Vous êtes sur la page 1sur 620

Sasakian Geometry

Charles P. Boyer and Krzysztof Galicki

Department of Mathematics and Statistics, University of New Mex-


ico, Albuquerque, N.M. 87131
Preface

This book is a result of a fifteen year long collaboration which has led us to
understand and appreciate the importance of Sasakian manifolds as an integral
part of Riemannian Geometry. In the early nineties neither of us was aware or
particularly interested in Sasakian structures. This rapidly changed in 1992 when,
together with Ben Mann, we realized that smooth 3-Sasakian manifolds, which are
automatically Einstein, are orbibundles over positive quaternionic Kähler orbifolds.
In the smooth case this bundle was first described by Konishi. As quaternionic
Kähler orbifolds are easily manufactured via quaternionic reduction, likewise we
were able to construct large families of smooth compact 3-Sasakian spaces with
relative ease. Searching through the literature afterwards we were surprised that a
folklore conjecture attributed to Tanno stated that any 3-Sasakian manifold should
be a spherical space form.
Soon after we began to understand that the 3-Sasakian geometry, though very
interesting, was too specialized a case. To a large extent our main motivation
had to do with Einstein metrics. We realized that the 3-Sasakian manifolds were
merely a special case of the more general theory of Sasaki-Einstein manifolds and
shifted our focus to that more general case. With time it became clearer and clearer
that Sasakian geometry is one of the richest sources of complete Einstein metrics of
positive scalar curvature. We began to appreciate that Sasakian geometry, naturally
“sandwiched” between two different Kähler geometries, was at least as interesting
and important as the latter.
It was Nigel Hitchin who first suggested to us that perhaps it is time to write
a modern book on the subject. We started thinking about the project in 2002. It
was a challenging endeavor: we aimed at writing a monograph which should give a
fairly complete account of our own contributions to the subject combined with an
advanced graduate level textbook describing the foundational material in a modern
language. Unlike the Kähler case, there are very few books of this sort discussing
Sasakian manifolds.
Another difficulty which we were to discover later was the rapid development of
the subject. Some new results, such as the construction of Sasaki-Einstein metrics
on exotic spheres with János Kollár, had to do with the dynamics of our own
research program. But many other important results were obtained by others.
To complicate things even more, Sasaki-Einstein manifolds appear to play a very
special rôle in the so-called AdS/CFT duality conjecture and they have received an
enormous amount of attention among physicists working in this area. Some of the
intriguing new results which we decided to include in the book, albeit very briefly,
indeed came quite unexpectedly from considerations in Superstring Theory.
In the end we know our book will not be as complete as we would have wanted
or hoped it to be. We finally had to stop writing while knowing all too well that the

iii
iv PREFACE

book is neither complete nor perfect. Few books of this sort are. We are pleased to
see so much renewed interest in the subject and hope that our work will be of help
to a number of mathematicians and physicists, researchers and graduate students
alike.
Over the years we have benefited immensely from many discussions with many
colleagues and collaborators about the mathematics contained in this book. Here
we are happy to take this opportunity to thank: B. Acharya, I. Agricola, D. Alek-
seevsky, V. Apostolov, F. Battaglia, C. Bär, H. Baum, F. Belgun, L. Bérard Bergery,
R. Bielawski, C. Böhm, A. Buium, D. Blair, O. Biquard, J.-P. Bourguignon, R.
Bryant, D. Calderbank, A. Čap, J. Cheeger, T. Colding, A. Dancer, O. Dearricott,
T. Draghici, Y. Eliashberg, J. Figueroa O’Farrill, M. Fernandez, T. Friedrich, E.
Gasparim, P. Gauduchon, H. Geiges, W. Goldman, G. Grantcharov, K. Grove, M.
Harada, T. Hausel, G. Hernandez, L. Hernandez, R. Herrera, O. Hijazi, N. Hitchin,
H. Hofer, T. Holm, J. Hurtubise, J. Isenberg, G. Jensen, J. Johnson, A. de Jong,
D. Joyce, L. Katzarkov, J. Konderak, M. Kontsevich, J. Kollár, H. B. Lawson,
C. LeBrun, B. Mann, P. Matzeu, S. Marchiafava, R. Mazzeo, J. Milgram, A. Mo-
roianu, P.-A. Nagy, M. Nakamaye, T. Nitta, P. Nurowski, L. Ornea, H. Pedersen,
P. Piccinni, M. Pilca, Y.-S. Poon, E. Rees, P. Rukimbira, S. Salamon, R. Schoen,
L. Schwachhöfer, U. Semmelmann, S. Simanca, M. Singer, J. Sparks, J. Starr, S.
Stolz, A. Swann, Ch. Thomas, G. Tian, M. Verbitsky, M. Wang, J. Wiśniewski, R.
Wolak, D. Wraith, S.-T. Yau, D. Zagier, and W. Ziller. It is probably inevitable
that we have missed the names of some friends and colleagues to whom we deeply
apologize.
We owe our deep gratitude to János Kollár. He is not just a collaborator on
one of the more important papers we wrote. His continuous help and involvement
throughout writing of this book was invaluable. Much of the material presented
in chapters 10 and 11 was written with his expert help and advice. We also wish
to thank Ilka Agricola, Elizabeth Gasparim, Eugene Lerman, Michael Nakamaye,
and Santiago Simanca and Thomas Friedrich for carefully reading certain parts
of our book and providing invaluable comments and corrections. We thank Evan
Thomas for helping with computations used to compile the various tables appearing
in the appendices. We also thank our graduate students: J. Cuadros, R. Gomez,
D. Grandini, J. Kania, and R. Sanchez-Silva for weeding out various mistakes while
sitting in several courses we taught at UNM using early versions of the book.
The second author would like to express special thanks to the Max-Planck-
Institut für Mathematik in Bonn for the generous support and hospitality. Several
chapters of this book were written during K.G.’s sabbatical visit at the MPIM
during the calendar year 2004 and also later during two shorter visits in 2005 and
2006. Both of us thank the National Science Foundation for continuous support of
our many projects, including this one.
Finally, we thank Jessica Churchman and Alison Jones from the Oxford Uni-
versity Press for their patience, continuing interest in our work on the project and
much help in the latter stages of preparing the manuscript for publication.
Last but not least we would like to thank our families, especially Margaret and
Rowan, for support and patience and for putting up with us while we were “working
on The Book” day and night.

Albuquerque, February 2007.


Contents

Preface iii

Introduction 1

Chapter 1. Structures on Manifolds 9


1.1. Sheaves and Sheaf Cohomology 9
1.2. Principal and Associated Bundles 14
1.3. Connections in Principal and Associated Vector Bundles 18
1.4. G-Structures 23
1.5. Pseudogroup Structures 36
1.6. Group Actions on Manifolds 38

Chapter 2. Foliations 51
2.1. Examples of Foliations 51
2.2. Haefliger Structures 52
2.3. Leaf Holonomy and the Holonomy Groupoid 54
2.4. Basic Cohomology 59
2.5. Transverse Geometry 60
2.6. Riemannian Flows 69

Chapter 3. Kähler Manifolds 75


3.1. Complex Manifolds and Kähler Metrics 76
3.2. Curvature of Kähler Manifolds 82
3.3. Hodge Theory on Kähler Manifolds 88
3.4. Complex Vector Bundles and Chern Classes 92
3.5. Line Bundles and Divisors 94
3.6. The Calabi Conjecture and the Calabi-Yau Theorem 102

Chapter 4. Fundamentals of Orbifolds 105


4.1. Basic Definitions 105
4.2. Orbisheaves and orbibundles 108
4.3. Groupoids, Orbifold Invariants and Classifying Spaces 115
4.4. Complex Orbifolds 123
4.5. Weighted Projective Spaces 133
4.6. Hypersurfaces in Weighted Projective Spaces 138
4.7. Seifert Bundles 144

Chapter 5. Kähler-Einstein Metrics 151


5.1. Some Elementary Considerations 152
5.2. The Monge-Ampère Problem and the Continuity Method 153
5.3. Obstructions in the Positive Case 160
v
vi CONTENTS

5.4. Kähler-Einstein Metrics on Hypersurfaces in CP(w) 162


5.5. Automorphisms and the Moduli Problem 173

Chapter 6. Almost Contact and Contact Geometry 179


6.1. Contact Structures 180
6.2. Almost Contact Structures 190
6.3. Almost Contact Metric Structures 195
6.4. Contact Metric Structures 198
6.5. Structures on Cones 201

Chapter 7. K-Contact and Sasakian Structures 207


7.1. Quasi-regularity and the Structure Theorems 207
7.2. The Transverse Geometry of the Characteristic Foliation 214
7.3. Curvature Properties of K-Contact and Sasakian Structures 219
7.4. Topology of K-Contact and Sasakian Manifolds 229
7.5. Sasakian Geometry and Algebraic Geometry 236
7.6. New Sasakian Structures from Old 250

Chapter 8. Symmetries and Sasakian Structures 257


8.1. Automorphisms of Sasakian Structures and Isometries 257
8.2. Deformation Classes of Sasakian Structures 265
8.3. Homogeneous Sasakian Manifolds 272
8.4. Symmetry Reduction and Moment Maps 276
8.5. Contact and Sasakian Reduction 290

Chapter 9. Links as Sasakian Manifolds 299


9.1. Preliminaries 299
9.2. Sasakian Structures and Weighted Homogeneous Polynomials 300
9.3. The Milnor Fibration and the Topology of Links 302
9.4. The Differential Topology of Links 312
9.5. Positive Sasakian Structures on Links 319
9.6. Links of Complete Intersections 326

Chapter 10. Sasakian Geometry in Dimensions Three and Five 329


10.1. Sasakian Geometry in Dimension Three 329
10.2. Sasakian Structures and the Topology of 5-Manifolds 335
10.3. Sasakian Links in Dimension Five 352
10.4. Regular Sasakian Structures on 5-Manifolds 360

Chapter 11. Sasaki-Einstein Geometry 369


11.1. Foundations of Sasaki-Einstein Geometry 370
11.2. Extremal Sasakian Metrics 378
11.3. Further Obstructions to Sasaki-Einstein Structures 382
11.4. Sasaki-Einstein Metrics in Dimensions Five 388
11.5. Sasaki-Einstein Metrics on Homotopy Spheres 403
11.6. The Sasaki-Einstein semi-group 407
11.7. Sasaki-Einstein Metrics in Dimensions Seven and Higher 409
11.8. Sasakian η-Einstein Metrics 417

Chapter 12. Quaternionic Kähler and Hyperkähler Manifolds 421


12.1. Quaternionic Geometry of Hn and HPn 422
CONTENTS vii

12.2. Quaternionic Kähler Metrics 428


12.3. Positive Quaternionic Kähler Manifolds and Symmetries 434
12.4. Quaternionic Kähler Reduction 438
12.5. Compact Quaternionic Kähler Orbifolds 443
12.6. Hypercomplex and Hyperhermitian Structures 453
12.7. Hyperkähler Manifolds 455
12.8. Hyperkähler Quotients 458
12.9. Toric Hyperkähler Metrics 461
12.10. ALE Spaces and Other Hyperkähler Quotients 465
Chapter 13. 3-Sasakian Manifolds 473
13.1. Almost Hypercontact Manifolds and 3-Sasakian Structures 474
13.2. Basic Properties 478
13.3. The Fundamental Foliations Fτ and FQ 480
13.4. Homogeneous 3-Sasakian Manifolds 491
13.5. 3-Sasakian Cohomology 494
13.6. Symmetry Reduction 499
13.7. Toric 3-Sasakian Manifolds 506
13.8. Cohomogeneity One 3-Sasakian 7-Manifolds 522
13.9. Non-Toric 3-Sasakian Manifolds in Dimension 11 and 15 525

Chapter 14. Sasakian Structures, Killing Spinors, and Supersymmetry 529


14.1. The Dirac Operator and Killing Spinors 529
14.2. Real Killing Spinors, Holonomy and Bär’s Correspondence 533
14.3. Geometries Associated with 3-Sasakian 7-manifolds 535
14.4. Geometries Associated with Sasaki-Einstein 5-manifolds 542
14.5. Geometric Structures on Manifolds and Supersymmetry 545
Appendix A 551
A.1. Preliminaries on Groupoids 551
A.2. The Classifying Space of a Topological Groupoid 555

Appendix B 559
B.1. Reid’s List of K3 Surfaces as hypersurfaces in CP4 (w) 559
B.2. Differential topology of 2k(S 3 × S 4 ) and 2k(S 5 × S 6 ) 560
B.3. Tables of Kähler-Einstein metrics on hypersurfaces CP(w) 561
B.4. Positive Breiskorn-Pham Links in Dimension 5 564
B.5. The Yau-Yu Links in Dimensions 5 567
Bibliography 569
Index 607
Introduction

In 1960 Shigeo Sasaki [Sas60] began the study of almost contact structures
in terms of certain tensor fields, but it wasn’t until [SH62] that what are now
called Sasakian manifolds first appeared under the name of “normal contact metric
structure”. By 1965 the terms “Sasakian structure” and “Sasakian manifold” be-
gan to be used more frequently replacing the original expressions. For a number of
years these manifolds were intensively studied by a group of Japanese geometers.
The subject did get some attention in the United States mainly due to the papers
of Goldberg and Blair. Nevertheless, the main interest in the field remained in-
side Japan, finding it hard to spread out and attract broader attention beyond its
birthplace either in the United States or in Europe.
Over a period of four years between 1965 and 1968 Sasaki wrote a three part
set of lecture notes which appeared as an internal publication of the Mathematical
Institute of the Tôhoku University under the title Almost contact manifolds, Part
I-III [Sas65, Sas67, Sas68]. Put together, the work amounts to almost 500 pages.
Even today, after 40 years, the breath, depth and the relative completeness of the
Sasaki lectures is truly quite remarkable. It is hard to understand why they did not
make it as a monograph in some prestigious Western book series; it is a pity. As it
is, the notes are not easily available and, consequently, not well-known1. Outside
Japan the first and important attempt to give a broader account of the subject was
given eight years later by Blair [Bla76a].
After 1968 Sasaki himself was less active although he continued to publish
until 1980. Yet he had already created a new subfield of Riemannian geometry
which slowly started to attract attention worldwide, not just in Japan. In 1966
Brieskorn wrote his famous paper describing a beautiful geometric model for all
homotopy spheres which bound parallelizable manifolds [Bri66]. In 1976 Sasaki
[ST76, SH76] realized that Brieskorn manifolds admit almost contact and con-
tact structures. (This very important fact was independently observed by several
other mathematicians: Abe-Erbacher [AE75], Lutz-Meckert [LM76], and Thomas
[Tho76].) Thirty years later the Brieskorn-Pham links as well as more general links
of weighted homogeneous polynomials, are the key players in several chapters of
our book. Yet again, Sasaki seemed to have had both the necessary intuition and
a broad vision in understanding what is and what is not of true importance2.
1We became aware of the Sasaki notes mainly because of our work on this book. We obtained
a copy of the lectures in 2003 form David Blair and we would like to thank him for sharing them
with us.
2We know little about Sasaki’s non-mathematical life. He was born in 1912 and what we do
know is from the volume of his selected papers edited in 1985 by Tachibana [Sas85]. There one
finds a short introduction by S. S. Chern and an essay by Sasaki in which he mostly discusses
his life as a working mathematician. He apparently died almost 20 years ago on August 12th,
1987. Sadly, his death passed without any notice, strangely forgotten. We could not find an
1
2 INTRODUCTION

Over the years Sasakian geometry has taken a back seat to other areas of
Riemannian geometry, most prominently to the study of Riemannian geometry
with reduced holonomy groups. Nevertheless, as we shall see, Sasakian geome-
try is closely related to all these other geometries. Although generally a Sasakian
manifold has the generic holonomy SO(n), the Riemannian cone over a Sasakian
manifold does have reduced holonomy making the study of Sasakian geometry quite
tractable. A quick perusal of Berger’s list of possible irreducible Riemannian holo-
nomy groups (see Table 1.4.1 below) shows that there are five infinite series of these
holonomy groups. All five are related to Sasakian geometry: generally a Sasakian
metric itself has generic holonomy SO(n), and its Riemannian cone has holonomy
U (n); the Riemannian cone of a general Sasaki-Einstein structure has holonomy
SU (n); whereas, the Riemannian cone of a general 3-Sasakian structure has holo-
nomy Sp(n). The remaining infinite series in Table 1.4.1 is the group Sp(n)Sp(1)
of quaternionic Kähler geometry which is closely related to 3-Sasakian geometry as
we discuss shortly. The two remaining irreducible Riemannian holonomy groups
G2 and Spin(7) are related to Sasakian geometry in a less direct way as we discuss
in Chapter 14. In this same spirit our book attempts to show Sasakian geometry
not as a separate subfield of Riemannian geometry but rather through its inter-
relation to other geometries. This is perhaps the most important feature of the
subject. The study of Sasakian manifolds brings together several different fields
of mathematics from differential and algebraic topology through complex algebraic
geometry to Riemannian manifolds with special holonomy.
The closest relative of Sasakian geometry is Kählerian geometry, the importance
of which is difficult to overestimate mainly because of its role in algebraic geometry.
But Sasakian geometry also has a very algebro-geometric flavor. In fact, there is an
‘algebraic structure’ on every Sasakian manifold; whereas, Voisin has shown recently
that there are Kähler manifolds that admit no algebraic structure whatsoever. To
better understand the relation between Sasakian and Kählerian geometries we begin
with the more familiar relation between contact and symplectic geometries. Let
(M, η, ξ) be a contact manifold where η is a contact form on M and ξ is its Reeb
vector field. It is easy to see that the cone (C(M ) = R+ × M, ω = d(tη)) is
symplectic. Likewise, the Reeb field defines a foliation of M and the transverse
space Z is also symplectic. When the foliation is regular the transverse space is a
smooth symplectic manifold giving a projection π called Boothby-Wang fibration,
and π ∗ Ω = dη relates the contact and the symplectic structures as indicated by

(C(M ), ω) ←- (M,η, ξ)


π
y
(Z, Ω).

English language mathematical obituary honoring his life and work and commemorating his death;
apparently, his passing was noted just in a short obituary of a local newspaper. After his death
Tokyo Science University where he briefly worked after retiring from Tôhoku took care of his
mathematical heritage with some of his manuscripts placed at the university library. We are
grateful to Professor Yoshinobu Kamishima for this information.
INTRODUCTION 3

We do not have any Riemannian structure yet. It is quite reasonable to ask


if there is a Riemannian metric g on M which “best fits” into the above diagram.
As the preferred metrics adapted to symplectic forms are Kähler metrics one could
ask for the Riemannian structure which would make the cone with the warped
product metric ḡ = dt2 + t2 g together with the symplectic form ω into a Kähler
manifold? Then ḡ and ω define a complex structure Φ̄. Alternatively, one could ask
for a Riemannian metric g on M which would define a Kähler metric h on Z via a
Riemannian submersion. Surprisingly, in both cases the answer to these questions
leads naturally and uniquely to Sasakian Geometry. Our diagram becomes

(C(M ), ω, ḡ, Φ̄) ←- (M, ξ,η, g, Φ)




π
y
(Z, Ω, h, J).

From this point of view it is quite clear that Kählerian and Sasakian geome-
tries are inseparable, Sasakian Geometry being naturally sandwiched between two
different types of Kählerian Geometry. Yet the two fared differently over the years.
Since Erich Kähler’s seminal article “several dominant figures of the mathematical
scene of the XXth century have, step after step along a 50 year period, transformed
the subject into a major area of Mathematics that has influenced the evolution of
the discipline much further than could have conceivably been anticipated by any-
one” writes Jean-Pierre Bourguignon in his tributary article The unabated vitality
of Kählerian geometry published in [Käh03]. Sasakian Geometry has not been as
lucky. There always has been interesting work in the area, but for unclear reasons
it has never attracted people with the same broad vision, people who would set
out to formulate and then work on fundamental problems. Yet, arguably Sasakian
manifolds are at least as interesting as Kählerian manifolds.
Our own research of the last decade has been an attempt to bring Sasakian
geometry back into the main stream. We believe that a modern book on Sasakian
manifolds is long overdue. Most of the early results are scattered, often buried
in old and hard to get journals. They are typically written in an old fashioned
language. Worse than that, articles with interesting results are drowned in a vast
sea of papers of little importance. There are very few graduate level texts on the
subject. There is the book Structures on manifolds written by Yano and Kon in 1984
[YK84], but this book is over 20 years old and treats both Sasakian and Kählerian
geometries as a subfield of Riemannian geometry. Recently, Blair substantially
updated his well-known Contact manifolds in Riemannian geometry in the Springer
Lecture Notes series [Bla76a] with Riemannian geometry of contact and symplectic
manifolds [Bla02]. Again the major emphasis as well as the techniques used are
Riemannian in nature. Our monograph naturally complements Blair’s as it employs
an entirely different philosophy and follows a different approach. First we develop
the important relations between Sasakian geometry and the algebraic geometry of
Kähler (actually projective algebraic) orbifolds. Secondly, our major motivation
to begin with was in proving the existence of Einstein metrics. So we have the
understanding of Sasaki-Einstein metrics as a main goal toward which to work, but
4 INTRODUCTION

we have also come to appreciate the beauty and richness of Sasakian geometry in
its own right.
Our book breaks more or less naturally into two parts. Chapters 1 through
9 provide an introduction to the modern study of Sasakian geometry. Starting
with Chapter 10 the book becomes more of a research monograph describing many
of our own results in the subject. However, the extensive introduction should
make it accessible to graduate students as well as non-expert researchers in related
fields. We have used parts or our monograph as a textbook for advanced graduate
courses. For example, assuming that the students have some basic knowledge of
Riemannian geometry, algebraic geometry, and some algebraic topology, a course
treating Sasakian geometry starting with Chapter 4 through Chapter 11 is possible,
drawing on the first three chapters as review. Many of the results in Chapters 4-11
are given with full proofs bringing the student to the forefront of research in the area.
As a guide we use the principle that proofs, or at least an outline of proofs, are given
for important results that are not found in another book. Our book also contains
many examples, for we believe that the learning process is substantially enhanced
by working through examples. We also have exercises scattered throughout the
text for the reader to sharpen her/his skills. Open problems of varying or unknown
difficulty are listed, some of which could be the basis of a dissertation. The text is
aimed at mathematicians, but we hope it will find many readers among physicists,
particularly those working in Superstring Theory.
We begin in Chapter 1 by introducing various geometries that play more or
less important roles in the way they relate to Sasakian structures. We espouse the
point of view that a geometric structure is best described as a G-structure which,
in addition, may or may not be (partially) integrable. As Sasakian manifolds are all
examples of Riemannian foliations with one-dimensional leaves, Chapter 2 takes the
reader into the world of foliations with a particular focus on the Riemannian case.
The literature is full of excellent books on the subject so we just select the topics
most relevant to us. Chapter 3 reviews some basic facts about Kähler manifolds.
Again, we are very selective choosing only what is needed later in describing the
two Kähler geometries of the “Kähler-Sasaki sandwich”. Of particular interest is
Yau’s famous proof of the Calabi conjecture.
A key tool that allows for connecting Sasakian structures to other geometric
structures is the theory of Riemannian orbifolds and orbifold bundles or ‘orbibun-
dles’. For that reason Chapter 4 is crucial in setting the stage for an in depth study
of Sasakian manifolds which begins later in Chapter 7. Orbifolds just as manifolds
have become a household name to the well trained geometer. Nevertheless, a lot
of important results are scattered throughout the literature, and orbifolds typically
appear within a specific context. There is a forthcoming book on orbifolds, Orb-
ifolds and Stringy Topology [ALR06] by Adem, Leida, and Ruan, but this has a
particularly topological bent. Hence, we take some time and effort to prepare the
reader introducing all the basic concepts from the point of view needed in subse-
quent chapters. By Chapter 5 we are ready for a second trip into the realm of
Kähler geometry. However, now the focus is on Kähler-Einstein metrics, in partic-
ular positive scalar curvature Kähler-Einstein metrics on compact Fano orbifolds.
We introduce some basic techniques that allow for proving various existence results.
We also briefly discuss obstructions. Chapter 6 presents the necessary foundational
INTRODUCTION 5

material on almost contact and contact geometry. This leads directly to the defi-
nition of a Sasakian structure introduced at the very end.
The study of Sasakian geometry finally begins with Chapter 7. We first present
the important structure theorems, and then gather all the results concerning the
geometry, topology, and curvature properties of both K-contact and Sasakian man-
ifolds. Most of the curvature results are standard and can be found in Blair’s book
[Bla02], but our main focus is different: we stress the relation between Sasakian
and algebraic geometry, as well as the basic cohomology associated with a Sasakian
structure. A main tool used in the text is the transverse Yau Theorem due to El
Kacimi-Alaoui. In the companion Chapter 8 we present known results concerning
symmetries of Sasakian structures. We introduce the Sasakian analogue of the bet-
ter known symplectic/contact reduction. Then we study toric contact and toric
Sasakian manifolds and prove several Delzant-type results.
Chapter 9 is devoted to the geometry of links of isolated hypersurface singulari-
ties as well as a review of the differential topology of homotopy spheres a la Kervaire
and Milnor [KM63]. A main reference for the study of such links is Milnor’s classic
text Singular points of complex hypersurfaces [Mil68], but also Dimca’s Singular-
ities and topology of hypersurfacs [Dim92] is used. The differential topology of
links is a beautiful piece of mathematics, and this chapter offers a hands-on “user’s
guide” approach with much emphasis on the famous work of Brieskorn [Bri66]. Of
importance for us is that when the singularities arise from weighted homogeneous
polynomials the links have a natural Sasakian structure with either definite (posi-
tive or negative) or null basic first Chern class. Emphasis is given to the positive
case which corresponds to having positive Ricci curvature. In Chapter 10 we dis-
cuss the Sasakian geometry in low dimensions. In dimension 3 there is a complete
classification. Dimension 5 is large enough to be interesting, yet small enough to
hope for some partial classification. We concentrate on the simply connected case
as there we can rely on the Smale-Barden classification. In terms of Sasakian struc-
tures our main focus is on the case of positive Sasakian structures. In considerable
detail we describe several remarkable theorems of Kollár which show how positivity
severely restricts the topology of a manifold which is to admit a positive Sasakian
structure.
Chapter 11 is central to the whole book and perhaps the main reason and jus-
tification for it. Much of this chapter is based on a new method for proving the ex-
istence of Einstein metrics on odd dimensional manifolds introduced by the authors
in 2001 [BG01b]. We realized there that links of isolated hypersurface singulari-
ties obtained from weighted homogeneous polynomials admit Sasakian structures.
Moreover, by using an orbifold version [BG00b] of an old result of Kobayashi, we
successfully tied the problem to making use of the continuity method for proving
the existence of Kähler-Einstein metrics on compact Kähler orbifolds. For the au-
thors this was the original “raison d’etre” for Chapters 4, 5, and 9 of the book.
In a series of papers the authors and their collaborators have successfully applied
this method to prove the existence of Sasaki-Einstein metrics on many 5-manifolds,
on odd dimensional homotopy spheres that bound parallelizible manifolds, as well
as on odd dimensional rational homology spheres. Furthermore, our method has
been substantially generalized by Kollár who has pushed our understanding much
further especially in dimension five. Although a complete classification is perhaps
6 INTRODUCTION

not within reach we now begin to have a really good grasp of Sasaki-Einstein ge-
ometry in dimension 5. In addition, we discuss toric Sasaki-Einstein geometry in
dimension 5 which began with the work of Gauntlett, Martelli, Sparks, and Wal-
dram [GMSW04b], and culminates with the very recent work of Cho, Futaki,
Ono, and Wang [FOW06, CFO07] which shows in arbitrary odd dimension that
any toric Sasakian manifold with positive anticanonical Sasakian structure admits
a compatible Sasaki-Einstein metric. We also discuss extremal Sasaki metrics de-
fined in analogy with the extremal Kähler metrics, and introduce the Sasaki-Futaki
invariant [BGS06]. In addition to lifting the well-known obstructions of positive
Kähler-Einstein metrics we also present some new results due to Gauntlett, Martelli,
Sparks, and Yau [GMSY06] involving two well known estimates, one due to Lich-
nerowicz, and the other to Bishop. We also present Sasaki-Einstein metrics obtained
via the join construction described earlier in Chapter 7. We end this long chapter
with a brief discussion of Sasakian-η-Einstein metrics.
Chapter 12 gives an extensive overview of various quaternionic geometries. The
main focus is on the positive quaternionic Kähler (QK) manifolds (orbifolds) and
on the hyperkëhler manifolds (orbifolds). The reason for such an extensive treat-
ment has to do with Chapter 13. The 3-Sasakian manifolds studied there cannot
be introduced without a deeper understanding of these two quaternionic geome-
tries, just as Sasakian and Sasaki-Einstein manifolds cannot be studied without
Kählerian and Kähler-Einstein geometry. The Sasaki-Einstein manifolds of Chap-
ter 13 have a completely different flavor than the ones that appeared in Chapter
11. It is not only that these occur only in dimensions 4m + 3, but also that they
have a somewhat richer geometric structure. In addition the method in which
the metrics are obtained is completely different. In Chapter 11, with some excep-
tions, we mostly get our existence results via the continuity method applied to the
Monge-Ampère equation. Very few metrics are known explicitly, though there are
exceptions. Most of the 3-Sasakian metrics we consider are obtained via symmetry
reduction similar to the hyperkähler and quaternionic Kähler reduction. Indeed the
three quotients are all related. So the manifolds and the metrics we get are quite
often explicit and can be studied as quotients. Again, there are some exceptions.
Finally, Chapter 14 gives a very brief overview of the rich theory of Killing spinors.
There we describe some other geometries and show how they relate to 3-Sasakian
7-manifolds and Sasaki-Einstein 5-manifolds. At the end we very briefly comment
on how Sasakian geometry naturally appears in various supersymmetric physical
theories. Both Sasaki-Einstein geometry and geometries with exceptional holonomy
have appeared in various models of supersymmetric String Theory fuelling vigorous
interest in them by mathematicians and physicists alike.
We also have added two appendices. The first appendix gives a very brief in-
troduction to groupoids and their classifying spaces which are employed in Chapter
4, while the second gives many tables listing links of hypersurface singularities that
are used throughout the book.
We have compiled a very extensive bibliography. There are various reasons for
its size. We should remark that in this day and age of easy internet access, with
MathSciNet and Google, it would make no sense to simply compile a bibliography
of every paper with the words ‘Sasaki’ or ‘Sasakian’ in the title. Anyone with an
access to MathSciNet can easily compile such a list of 809 papers3 so it would serve
3As we checked on February 19, 2007.
INTRODUCTION 7

no purpose to do it for this book. In a way, in spite of its size, we were very careful
and selective in choosing all bibliography items. Our book brings together so many
different areas of mathematics, that having good references becomes essential. In
some cases the proofs we give are only sketches and in such instances we wanted
to refer the reader to the best place he/she could find more details. That is quite
often the original source but not always. We refer to various books, monographs,
and lecture notes. We have tried to be both selective and accurate in attributing
various results with care. This can be at times a hard task. We suspect that we
did not always get the references and proper credits exactly right. This is almost
inevitable considering we are not experts in many of the areas of mathematics that
substantially enter as part of the book. In any case, we apologize for any omissions
and errors; these are certainly not intentional4.
Finishing this book was a challenging task. We had to deal with an increasing
number of new and interesting results appearing every few months. It is always a bit
dangerous to include material based on articles that have not yet been published.
To make the book as up-to-date as possible we took the risk to include some of
these new results without giving proofs. On the other hand we are happy to see
that the field is active and we very much hope that our book will become somewhat
outdated in a few years.

4There are many examples in the literature where incorrect attributions are made. A recent
case that we just uncovered is that of nearly Kähler manifolds which are usually attributed to
Gray [Gra69b, Gra70], yet they were discovered 10 years earlier by Tachibana [Tac59]. The
fact that any nearly Kähler 6-manifold is Einstein is also attributed to Gray in [Gra76], yet it
was proven earlier by Matsumoto [Mat72]. See Section 14.3.2.
CHAPTER 1

Structures on Manifolds

A unifying viewpoint for doing differential geometry involving different struc-


tures is that of a G-structure, where G can stand for ‘geometric’ or more appro-
priately a ‘Lie group’. A (first order) G-structure is just a reduction of the bundle
of frames of a manifold from the general linear group to a subgroup. For example,
from the point of view of G-structures a Riemannian metric on a manifold corre-
sponds to a reduction of the frame bundle with group GL(n, R) to the orthonormal
frame bundle with subgroup, the orthogonal group O(n, R). Many other examples
will be given below. It will be important to have at our disposal the general theory
of connections in principal and associated fibre bundles of which the G-structures
mentioned above are special cases. However, before embarking on the study of such
structures we give a short review of sheaves and their cohomology groups. Sheaves
are more general than bundles, but they are a bit too general for describing geomet-
ric structures on manifolds. They are mainly used to pass from local information
to global information.

1.1. Sheaves and Sheaf Cohomology


Sheaves, which were invented by Jean Leray as a prisoner of war during World
War II, have become an important tool in geometry. Here we give a very brief tour
of sheaf theory referring to the literature [Bre97, GR65, Hir66, Wel80] for more
detail and proofs. Our presentation follows [Wel80] fairly closely.
Definition 1.1.1: Let X be a topological space. A presheaf on X is an assignment
to each nonempty open set U ⊂ X a set F(U ) together with maps, called restriction
maps,

ρU
V : F(U )−
−−−→F(V )

for each pair of open sets U and V with V ⊂ U that satisfy the conditions ρU
U = idU
and ρUW = ρ U
V ◦ ρ V
W whenever, W ⊂ V ⊂ U.
Very often the sets F(U ) have some additional algebraic structure, such as a
group, a ring, or a module structure. In this case we assume that the restriction
maps preserve the algebraic structure. So a presheaf is a contravariant functor
from the category of open sets of a fixed topological space with inclusion maps as
morphisms to the category of groups, rings, or modules whose morphisms are the
homomorphisms of that category.
9
10 1. STRUCTURES ON MANIFOLDS

A morphism of presheaves f : F−−→G on X is a family of homomorphisms


fU : F(U )−−→G(U ) such that the diagram
hU
F(U −−−−−→
 ) G(U
)
 
 
(1.1.1) ρUV ρUV
 
y y
hV
F(V ) −−−−−→ G(V )
commutes, where V ⊂ U ⊂ X. If the homomorphisms fU : F(U )−−→G(U ) are
monomorphisms then F is a subpresheaf of G.
Given a presheaf F over X we can consider the direct limit
Fx = lim
−→
F(U )
x∈U

with respect to the restriction maps {ρU V }. Clearly, Fx inherits whatever algebraic
structure the sets F(U ) have. Fx is called the stalk of F at x, and for x ∈ U there
is a natural projection ρUx : F(U )− −→Fx by sending an element s ∈ F(U ) to its
equivalence class sx in the direct limit. sx is called the germ of s at x.
Definition 1.1.2: A presheaf F is called a sheaf if for every collection Ui of open
subsets of X with U = ∪Ui , the presheaf F satisfies the following two conditions
(i) If s, t ∈ F(U ) and ρU U
Ui (s) = ρUi (t) for all i, then s = t.
(ii) If si ∈ F(Ui ) and if for any j with Ui ∩ Uj 6= ∅ the equality ρU
Ui ∩Uj (si ) =
i

U
ρUji ∩Uj (sj ) holds for all i, then there is an s ∈ F(U ) such that ρUUi (s) = si
for all i.
A morphism of sheaves is a morphism of the underlying presheaves. In par-
ticular, an isomorphism of sheaves is a morphism of sheaves such that the all the
maps fU : F(U )−−→G(U ) are isomorphisms. Not every presheaf is a sheaf. For
example consider C and the presheaf that assigns to every open set U ⊂ C the
algebra of bounded holomorphic functions B(U ) on U. Since there are no bounded
holomorphic functions on all of C this presheaf violates condition (ii) of definition
1.1.2 above. Nevertheless, as we shall see shortly we can associate a sheaf to any
presheaf in a fairly natural way. First we give an important example of a presheaf
that is a sheaf.
Example 1.1.3: Let X and Y be topological spaces and consider the presheaf CX,Y
on X which associates to every open subset U of X the set CX,Y (U ) of all continuous
maps from U to Y. The restriction map ρU V is just the natural restriction, i.e. if
f ∈ CX,Y (U ) and V ⊂ U, then ρU V (f ) = f |V . It is easy to see that CX,Y satisfies
the two conditions of Definition 1.1.2 and thus defines a sheaf on X. An important
special case is obtained by taking Y to be either of the continuous fields F = R or
C. In this case we note that CX,F has the structure of an F-algebra. Two particular
cases of interest are when X is either a smooth real manifold or a complex manifold.
In these cases we are more interested in the subsheaf EX ⊂ CX,R of smooth functions
and the subsheaf OX ⊂ CX,C of holomorphic functions, respectively. Such sheaves
are called the structure sheaf of X.
Definition 1.1.4: An étale space over a topological space X is a topological space
Y together with a continuous map π : Y −−→X that is a local homeomorphism.
1.1. SHEAVES AND SHEAF COHOMOLOGY 11

For each open set U ⊂ X we can consider the set Γ(U, Y ) of continuous sections
of π, that is, the subset of f ∈ CX,Y (U ) that satisfy π◦f = idU . Then the assignment
that assigns to each open subset U of X the set of continuous sections Γ(U, Y ) forms
a subsheaf of CX,Y which we denote by SX,Y . We shall now associate to any presheaf
F an étale space F̃. Then by taking sections of F̃ we shall get a sheaf. Define F̃ by

(1.1.2) F̃ = tx∈X Fx .

Let π : F̃−−→X denote the natural projection that sends any sx ∈ Fx to x. We


define a topology on F̃ as follows: for each s ∈ F(U ) define a map s̃ : U −−→F̃ by
s̃(x) = sx . This satisfies π ◦ s̃ = idU . Then we let the sets

{s̃ | U is open in X, s ∈ F(U )}

be a basis for the topology on F̃. Then π and all functions s̃ are continuous, and it
is easy to check that π is a local homeomorphism, so that F̃ is an étale space over
X. But we have already seen that the set of sections of an étale space form a sheaf.
We denote the sheaf associated to the étale space F̃ by S̃X,Y . Thus, beginning
with a presheaf F we have associated a sheaf S̃X,Y . This sheaf is called the sheaf
associated to or generated by the presheaf F.
Exercise 1.1: Show that if one starts with a sheaf F then the sheaf associated to
F is isomorphic to F.
Definition 1.1.5: A ringed space is a pair (X, A) consisting of a topological
space X together with a sheaf of rings A on X, called the structure sheaf. (X, A)
is a locally ringed space if for each point x ∈ X, the stalk Ax is a local ring.
All ringed spaces considered in this book will be locally ringed spaces, so we
often omit the word ‘locally’ and just refer to a ringed space. We now consider the
structure sheaves EM and OM on a real or complex manifold M, respectively. We
denote by A the sheaf E on a real manifold or the sheaf O on a complex manifold.
Definition 1.1.6: Let (X, A) be a ringed space with structure sheaf A given by E
(or O), respectively. A sheaf F of A-modules is said to be locally free of rank
r if X can be covered by open sets U such that there is an isomorphism of sheaves
r times
F|U ≈ A|U ⊕ ··· ⊕ A|U . A locally free rank 1 sheaf is called an invertible sheaf.
Let π : E−−→M be an F-vector bundle over a smooth manifold M . As in the
case of an étale space the subset of f ∈ CM,E satisfying π ◦ f = idM defines a
subsheaf of CM,E , called the sheaf of germs of continuous sections of E. Actually,
we are more interested in the sheaf of germs of smooth sections of E which we
denote by E(E). On a complex manifold M a C-vector bundle can have a special
type of structure.
Definition 1.1.7: Let M be a complex manifold. A complex vector bundle π :
E−−→M on M is said to be holomorphic if E is a complex manifold with π
holomorphic, and the transition functions for E can be taken to be holomorphic
functions.
Of course, the transition functions being holomorphic is equivalent to the local
trivializations being holomorphic. Then we have
Proposition 1.1.8: Let M be a real or complex manifold. There is a one-to-one
correspondence between smooth vector bundles on M and locally free EM -sheaves on
12 1. STRUCTURES ON MANIFOLDS

M. Similarly, when M is a complex manifold, there is a one-to-one correspondence


between holomorphic vector bundles on M and locally free OM -sheaves on M.
The correspondence is given by associating to a holomorphic vector bundle
E the sheaf of germs of holomorphic sections of E. Because of Proposition 1.1.8
we shall often use interchangeably the concepts of vector bundles and locally free
sheaves. Of particular interest is the rank one case in the holomorphic category
which gives a one-to-one correspondence between holomorphic line bundles and
invertible OM -sheaves.
As mentioned previously the power of sheaf theory is as a tool in passing from
the local to the global. This is accomplished by the way of sheaf cohomology theory
which we now briefly describe.

1.1.1. Sheaf Cohomology. There are several approaches to sheaf cohomol-


ogy theory. There is the axiomatic approach used in [GR65], the derived functor
approach using the resolution by discontinuous sections due to Godement [God58]
and espoused in [Bre97, Wel80], and finally the Čech theoretic approach described
in [GH78b] and [Kod86]. We prefer to begin with the latter and then present the
so-called Abstract de Rham Theorem using resolutions as in [Wel80].
For simplicity we assume that X is paracompact and Hausdorff. Let F be a
sheaf X, and let U = {Ui }i be a locally finite cover by open sets. Define the set
C p (U, F) of p-chains of F by
Y
(1.1.3) C p (U, F) = F(Ui0 ∩ · · · ∩ Uip ),
i0 ,...,ip

where we assume that the indices ij are distinct. We denote elements of C p (U, F)
by σi0 ,...,ip . The coboundary operator δp : C p (U, F)−−→C p+1 (U, F) is defined by
p+1
X
(1.1.4) (δp σ)i0 ,...,ip+1 = (−1)j σi0 ,...,ibj ,...,ip+1 |Ui0 ∩···∩Uip ,
j=0

where as usual î means remove that index. This gives rise to a cochain complex
δ0 δ1 δp
C 0 (U, F) −−→ C 1 (U, F) −−→ · · · −−→C p (U, F) −−→ C p+1 (U, F)−−→ · · ·
and we define the cohomology groups H p (U, F) by
kerδp
H p (U, F) = .
Imδp−1
If V is a cover of X refining U, there are homomorphisms ψ : H p (U, F)−−→H p (V, F).
Thus, we can take the direct limit of the cohomology groups H p (U, F) as the cover
becomes finer and finer, so we define the cohomology group with coefficients in the
sheaf F by
(1.1.5) H p (X, F) = lim
−→
H p (U, F).
U
p
Note that generally the groups H (U, F) depend on the cover, but there are certain
special covers known as acyclic covers for which we have the following theorem of
Leray
Theorem 1.1.9: If the cover U is acyclic in the sense that H q (Ui0 ∩ · · · ∩ Uip , F)
vanishes for all q > 0 and all i0 , . . . , ip , then H ∗ (M, F) ≈ H ∗ (U, F).
1.1. SHEAVES AND SHEAF COHOMOLOGY 13

In practice, one always computes H ∗ (U, F) for an acyclic cover and then uses
the above theorem.
P3
Example 1.1.10: Consider the sphere S 2 = {(x1 , x2 , x3 ) ∈ R3 | i=1 (xi )2 = 1} and
the constant sheaf F = R, together with two covers. The first is the ‘stereographic
cover’ defined by two open sets U± = S 2 − {(0, 0, ±1)}. This cover is not acyclic
since U+ ∩ U− has the homotopy type of a circle. The second cover is a cover
consisting of a cover of the lower hemisphere by 3 open sets together with the
upper hemisphere. This gives an acyclic cover from which one can compute the
cohomology groups H i (S 2 , R) of S 2 . See Example 9.3 of [BT82].
The sheaf cohomology groups have some nice properties. For any sheaf F the
cohomology group H 0 (X, F) is the space of global sections F(X) of F, and a sheaf
morphism f : A−−→B induces a group homomorphism fq : H q (X, A)−−→H q (X, B).
Moreover, a short exact sequence of sheaves,

0−−−→A−−−→B−−−→C−−−→0

induces a long exact sequence in cohomology,


(1.1.6)
δ δ
0−→H 0 (M, A)−→H 0 (M, B)−→H 0 (M, C) −→ H 1 (M, A)−→ · · · −→ H q (M, A)−→ · · · ,

where δ is the well-known connecting homomorphism.


To proceed further we need some definitions.
Definition 1.1.11: A sheaf F is soft if for any closed subset S ⊂ X, the restriction
map ρX U : F(X)− −→F(S) is surjective. F is fine if for any locally finite open cover
{Ui } of X there is a partition
P of unity of F, i.e. there is a family of sheaf morphisms
υi : F−−→F such that υi = 1, and υi (Fx ) = 0 for all x in some neighborhood of
X − Ui .
The importance of soft sheaves is the following
Proposition 1.1.12: If F is a soft sheaf, then H q (X, F) = 0 for all q > 0.
Fine sheaves are special cases of soft sheaves, viz.
Proposition 1.1.13: Fine sheaves are soft.
Example 1.1.14: The structure sheaf EM of a real smooth manifold is fine, thus
soft. The structure sheaf OM of a complex manifold is neither fine nor soft, nor are
the constant sheaves.
Definition 1.1.15: Let F be a sheaf on X. A resolution of F is an exact sequence
of sheaves F i of the form

0−−→F−−→F 0 −−→F 1 −−→ · · · −−→F k −−→ · · · .

The resolution is acyclic if H q (X, F i ) = 0 for all q > 0 and all i ≥ 0.


A resolution of a sheaf F is conveniently written in the shorthand notation
0−−→F−−→F ∗ . By Proposition 1.1.12 if the sheaves F i of a resolution are soft
sheaves, then the resolution is acyclic. Thus, a resolution by soft or fine sheaves is
acyclic. A resolution of sheaves gives rise to a cochain complex C ∗ = C ∗ (F ∗ ) on
global sections
d0 d1 dk−1 dk
(1.1.7) 0−−→F(X)−−→F 0 (X) −−→ F 1 (X) −−→ · · · −−→ F k (X) −−→ · · · .
14 1. STRUCTURES ON MANIFOLDS

In general this sequence is exact only at F(X), but we do have dk+1 ◦ dk = 0. The
cohomology group
ker dk
H k (C ∗ ) =
im dk−1
of this complex is called the k th derived group of C ∗ . We are now ready for the
Abstract de Rham Theorem:
Theorem 1.1.16: Let F be a sheaf over a paracompact Hausdorff space X, and
let 0−−→F−−→F ∗ be an acyclic resolution of F. Then there is a natural isomorphism
ker dk
H q (X, F) ≈ H q (C ∗ ) = .
im dk−1
This is a very powerful theorem of which the usual de Rham theorem is a special
case. We give this as an example below. The Dolbeault Theorem is another special
case which will be given in Chapter 3.
Example 1.1.17: de Rham’s Theorem. Let M be a smooth manifold and let
E p denote the sheaf of germs of sections of the exterior bundles Λp M with E 0 = E.
Let F = R the constant sheaf on M. Then by the well-known Poincaré Lemma we
get a resolution of the constant sheaf R, viz
0−−→R−−→E 0 −−→E 1 −−→ · · · −−→E n−1 −−→E n −−→0.
Since the sheaves E p are fine this resolution is acyclic, so by the abstract de Rham
Theorem 1.1.16 we have an isomorphism
µ ¶
k k ker dk k
H (M, R) ≈ H = HdeRh (M, R).
im dk−1
Moreover, H k (M, R) can be identified with the k th singular cohomology group by
taking an acyclic resolution of the constant sheaf R by sheaves of singular cochains.
See [Wel80].

1.2. Principal and Associated Bundles


We begin by considering principal fibre bundles. The main reference here is
the classic text of Kobayashi and Nomizu[KN63, KN69].
Definition 1.2.1: A principal bundle P over M with Lie group G consists of:
(1) A pair of smooth manifolds P and M together with a smooth surjection π : P →
M.
(2) A smooth free action of the Lie group G on P , R : P × G → P given by the
“right” action (u, a) 7→ Ra u = ua such that the quotient is M , i.e., M = P/G.
(3) P is locally trivial. More precisely, every point p ∈ M has a neighborhood U
and a diffeomorphism χ : π −1 (U ) → U × G defined by χ(u) = (π(u), φ(u)), where
φ : π −1 (U ) → G satisfies the compatibility condition φ(ua) = φ(u)a.
As usual the manifold P is called the total space, the manifold M is called the base
space, and for p ∈ M , π −1 (p) is called the fibre at p. If we fix a point u0 ∈ π −1 (p)
the map sending a ∈ G to u0 a ∈ π −1 (p) identifies π −1 (p) with G. Moreover, one
easily sees that this map is a diffeomorphism from G to π −1 (p) whose inverse is φ.
When we want to emphasize the base space and the group we write P (M, G) for P ,
or we refer to P as a principal G-bundle over M. Notice that for any open subset
U ⊂ M, π −1 (U ) is a principal G-bundle over U. Another example is the “trivial
G-bundle” when P is just the product M × G.
1.2. PRINCIPAL AND ASSOCIATED BUNDLES 15

Now let {Uα } be an open cover of M such that for each α the map χα :
π −1 (Uα )−−→Uα × G is a diffeomorphism, and satisfies the compatibility condition
of Definition 1.2.1. Then if u ∈ π −1 (Uα ∩ Uβ ) we see that φβ (ua)(φα (ua))−1 =
φβ (u)(φα (u))−1 for all a ∈ G. Hence, we can define a smooth map ψβα : Uα ∩
Uβ −−→G by

(1.2.1) ψβα (π(u)) = φβ (u)(φα (u))−1 .


The smooth maps ψβα are called transition functions for the principal bundle
P (M, G), and one easily sees that for each p ∈ Uα ∩ Uβ ∩ Uγ they satisfy the
‘cocycle condition’
(1.2.2) ψγα (p) = ψγβ (p)ψβα (p).
The appellation ‘cocycle condition’ comes from the fact that equivalence classes of
transition functions are elements of the sheaf cohomology set H 1 (M, G), where G
denotes the sheaf of germs of maps from opens sets of M to G. Here two transition
0
functions ψβα and ψβα are equivalent if there are maps ψα : Uα −−→G such that
ψβα (p) = ψβ (p)ψβα (p)(ψα (p))−1 , and the inverse element in G is given by ψαβ (p) =
0

(ψβα (p))−1 . The reader is referred to Section 1.1 below for more details on sheaf
theory and its cohomology. We then have the well-known
Theorem 1.2.2: Let {Uα } be an open cover on the smooth manifold M and let
G be a Lie group. Suppose further that on each nonempty intersection Uα ∩ Uβ
there exist smooth maps ψβα : Uα ∩ Uβ −−→G satisfying (1.2.2). Then there is a
principal G-bundle P (M, G) with transition functions ψβα . Moreover, there is a
bijective correspondence between isomorphism classes of principal G-bundles and
elements of H 1 (M, G).
The first statement in the theorem says that the transition functions determine
the principal bundle, and the last statement says that the set H 1 (M, G) classifies
principal G-bundles over M. The definition of an isomorphism of principal bundles
will be given below. It is only in the case that G is Abelian, for example, G = S 1 ,
the circle, that H 1 (M, G) has itself the structure of an Abelian group, and more
standard techniques such as the exponential sequence can be used to relate this to
the integral cohomology of M. We will discuss this later in more detail.
Perhaps the most important example of a principal bundle is:
Example 1.2.3: The linear frame bundle L(M ). Recall that a frame u =
(X1 , . . . , Xn ) at a point p ∈ M is a basis of the tangent space Tp M. We let L(M )
denote the set of all frames at all points of M. Then π : L(M )−−→M is a principal
bundle on M with group GL(n, R). The action of GL(n, R) on L(M ) is just given
by matrix multiplication from the right, that is (u, A) 7→ uA for u ∈ L(M ) and
A ∈ GL(n, R). The local triviality can be seen by choosing a local coordinate chart
(U ; x1 , . . . , xn ) on M and writing the vectors Xi of the frame u in local coordinates
as
X j ∂
Xi = Xi ,
j
∂xj

where the Xij are the components of a non-singular matrix (Xij ) of smooth functions
on U. Then the diffeomorphism ψ : π −1 (U )−−→U × GL(n, R) is given by ψ(v) =
(π(v), (Xij (v))). We remark that a point u ∈ L(M ) can be viewed as a linear map
16 1. STRUCTURES ON MANIFOLDS

u : Rn −−→Tp M with p = π(u) by u(ei ) = Xi , where {e1 , . . . , en } denotes the


standard basis of Rn .
Exercise 1.2: Let G be a Lie group and H a closed Lie subgroup. Show that the
Lie group G can be viewed as a principal H-bundle G(G/H, H) over the homoge-
neous manifold G/H.
Now suppose we are given a principal G-bundle P (M, G) over M , and a prin-
cipal H-bundle Q(N, H) on N. A homomorphism of principal bundles consists of a
smooth map fˆ : Q−−→P together with a Lie group homomorphism h : H−−→G such
that for all u ∈ Q and a ∈ H, fˆ satisfies fˆ(ua) = fˆ(u)h(a). This condition implies
that fˆ maps fibres to fibres; hence, there is a smooth map f : N −−→M such that
the diagram

Q
 −−−−−→ P

 
(1.2.3)  
πQ πP
y y
f
N −−−−−→ M
commutes. If in addition fˆ is an embedding of smooth manifolds and h is a group
monomorphism then we say that fˆ : Q−−→P is an embedding of principal bundles.
This implies that the map f : N −−→M is also an embedding. In the case that
N = M and f is the identity map on M, then Q(M, H) is called a subbundle
or a reduction of P (M, G) to the group H. We also say that Q(M, H) is a reduced
subbundle of P (M, G). It may not be possible to reduce a principal bundle P (M, G)
to a given subgroup H. In general, there are topological obstructions to doing so. We
shall see many examples of this below. If also h is an isomorphism then Q(M, H)
and P (M, G) are isomorphic as principal bundles. We also say that P (M, G) is
trivial if it is isomorphic to the product M × G. Observe that the local triviality
conditions says that every point p ∈ M has a neighborhood U such that π −1 (U ) is
trivial. Concerning the transition functions we have
Proposition 1.2.4: A principal G-bundle P (M, G) can be reduced to a Lie sub-
group H if and only if there is an open cover {Uα } of M with transition functions
ψβα taking their values in H.
Proof. Suppose that P (M, G) can be reduced to a Lie subgroup H. Then there
is a principal H-bundle Q(M, H) together with a smooth bundle map f : Q−−→P
such that the diagram (1.2.3) commutes and f (ua) = f (u)a for all a ∈ H ⊂ G. On
−1
the open set πQ (Uα ) ⊂ Q(M, H) the maps φQ P Q P
α and φα are related by φα = φα ◦ f.
It follows that for any v ∈ P there exists an a ∈ G and u ∈ Q such that v = f (u)a
and φP Q P
α (v) = φα (u)a. Hence, the transition functions ψβα satisfy
P
(1.2.4) ψβα (v) = φP P
β (v)(φα (v))
−1
= φQ
β (u)a · a (φα (u))−1 = φQ
−1 Q Q
β (u)(φα (u))
−1
,
and thus, have their values in H.
P
Conversely, given transition functions ψβα : Uα ∩ Uβ −−→G which take their
P
values in the Lie subgroup H, a standard result says that ψβα is smooth as a
map into H. Thus, by Proposition 1.2.2 there is a principal H-bundle Q(M, H)
P
with transition functions ψβα . To construct the bundle map f we define maps
fα : πQ (Uα )−−→πP (Uα ) by putting fα = χ−1
−1 −1
P ◦ χQ . One easily sees that fα = fβ
1.2. PRINCIPAL AND ASSOCIATED BUNDLES 17

on Uα ∩ Uβ , and so defines a global bundle map f : Q−−→P with the requisite


properties. ¤

Let P (M, G) be a principal bundle and F a G-manifold, that is, a manifold


together with a smooth action of the Lie group G. We denote this action by ‘left
multiplication’ as x 7→ a−1 x, where x ∈ F and a ∈ G. Then we have the “product
action” on P × F defined by (u, x) 7→ (ua, a−1 x). The quotient space (P × F )/G
by this action is a smooth fibre bundle E(M, F, G, P ) called the bundle associated
to P (M, G) with fibre F. Note that the projection map πE is defined as follows: let
[u, x] denote equivalence class of the pair (u, x) ∈ P × F, where (u, x) is equivalent
to (u0 , x0 ) if there is an a ∈ G such that (u0 , x0 ) = (ua, a−1 x). Then define the
projection map πE : E−−→M by πE ([u, x]) = πP (u). One easily checks that this
is well defined. In the case that F is a vector space, say Rk , and G = GL(k, R)
the associated bundle E = E(M, Rk , GL(k, R), P ) is called a vector bundle (of rank
k) associated to the principal bundle P (M, GL(k, R)), or just a real vector bundle.
Similarly, if the fibre F is Ck with group G = GL(k, C), then E is called a complex
vector bundle. We shall often combine these an refer to a real or complex vector
bundle as an F-bundle, where F = R or C. A vector bundle of rank one is called a
line bundle.
Exercise 1.3: Use πE to define a differential structure on E(M, F, G, P ) that
makes it a smooth fibre bundle with projection map πE and fibre F.
Exercise 1.4: Show that the tangent bundle T M of M is a vector bundle associ-
ated to the principal bundle L(M ). More generally show that the tensor bundles
Tsr M, exterior bundles Λp M, and symmetric bundles S p M are vector bundles as-
sociated to the principal bundle L(M ).
One often describes structures on manifolds by tensor fields. Examples are
complex structures, symplectic structures, Riemannian structures. These arise as
smooth sections of certain bundles over M which are associated to the frame bundle
L(M ) or its subbundles. Recall that a section of a fibre bundle πE : E−−→M
is a (smooth) map s : M −−→E such that πE ◦ s = 1lM . Unless otherwise stated
sections will be smooth. Vector bundles always have sections (e.g., the zero section);
whereas, a principal bundle has a section if and only if it is trivial. This is easy to
verify directly, but also follows from Theorem 1.2.5 below. By a local section of E
−1
we shall mean a section of the bundle πE (U ) for some open set U in M.
Theorem 1.2.5: A principal G-bundle P (M, G) is reducible to a closed subgroup H
if and only if the associated bundle E(M, G/H, G, P ) admits a section s : M −−→E.
Moreover, there is a bijective correspondence between such sections s and subbundles
Q(M, H) of P (M, G).

Proof. Suppose P (M, G) is reducible to a closed subgroup H and let f :


Q−−→P be the subbundle. The associated bundle E can be identified with the
quotient space P/H by mapping the equivalence class [u, gH] ∈ E to the equivalence
class [ug] ∈ P/H. Let ρ : P −−→E = P/H denote the natural projection. Then it is
easy to see that ρ ◦ f is constant on the fibres of Q. Thus, we can define a section
s : M −−→E by setting s(x) = ρ(f (u)), where πP (f (u)) = x.
Conversely, let s : M −−→E = P/H be a section, and consider P as a principal
H-bundle over E = P/H with transition functions hβα : Vα ∩Vβ −−→H, where {Vα } is
an open cover of E. Then we can define transition functions for P (M, G) with values
18 1. STRUCTURES ON MANIFOLDS

in H by setting ψβα = hβα ◦ s. Thus, by Proposition 1.2.2 P (M, G) is reducible to


the subgroup H. The correspondence between sections and H-subbundles can be
seen to be 1-1. ¤
Let P (M, G) be a principal bundle on a smooth manifold M with G = GL(k, R),
and let E be an associated real vector bundle of rank k. Reductions of P (M, G) to
subgroups G ⊂ GL(k, R) correspond to adding certain structures to its associated
vector bundle E. In general, there are obstructions to be able to do this. For exam-
ple, a reduction to the group GL(k, R)+ corresponds to choosing an orientation on
E, which can be done if and only if the bundle E is orientable. Such obstructions are
often given in terms of so-called characteristic classes [MS74]. Obstruction theory
for general fibre bundles is expounded in the classic text of Steenrod [Ste51] and
also in [Hus66]. E is orientable if and only if its first Stiefel-Whitney class w1 (E)
vanishes. See Appendix A. Many more examples of such obstructions are discussed
for G-structures in Section 1.4. Another example of such a reduction is the reduc-
tion of P (M, GL(k, R)) to its maximal compact subgroup O(k, R). Since GL(k, R)
is homotopy equivalent to the orthogonal group O(k, R) there are no obstructions
to performing this reduction. This corresponds to choosing a Riemannian metric
on E. If E is also orientable then we can reduce further to the special orthogonal
group SO(k, R). The corresponding principal bundle is now P (M, SO(k, R)).
Another type of structure comes from lifting instead of reduction. Thus, since
π1 (SO(k, R)) = Z2 the group SO(k, R) has a two-sheeted universal covering group
called Spin(k). So given a P (M, SO(k, R)) or one of its associated vector bundles E
one can ask whether the bundle P (M, SO(k, R)) can be lifted to a covering bundle
P (M, Spin(k)) with group Spin(k)? When such a covering exists the bundle E is
said to have a spin structure. We refer to [LM89] for a full treatment of spin
structures. The exact sequence of Lie groups 0−−→Z2 −−→Spin(k)−−→SO(k)−−→0
induces the coboundary map w2 : H 1 (M, SO(k))−−−→H 2 (M, Z2 ) whose image is
just the second Stiefel-Whitney class w2 (E) described briefly in Appendix A. So
an oriented vector bundle E admits a spin structure if and only if w2 (E) = 0.
Moreover, the distinct spin structures on E are in one-to-one correspondence with
the elements of H 1 (M, Z2 ). We say that a smooth oriented manifold M is a spin
manifold if T M admits a spin structure. So M is spin if and only if w1 (M ) =
w2 (M ) = 0. Spin manifolds admit certain vector bundles that do not exist on non-
spin manifolds, namely, those whose fibres are representations of Spin(n) that are
not representations of SO(n) or GL(n, R). Such vector bundles are called spinor
bundles and its sections are called spinor fields. Evaluation at a point gives a
spinor. Notice that spinor bundles depend on a choice of Riemannian metric. A
study of these representations involves Clifford algebras which will not be treated
in any detail in this text, and will not appear until Chapter 14. See [LM89] for a
thorough treatment.

1.3. Connections in Principal and Associated Vector Bundles


In this section we briefly review the fundamentals of the theory of connections.
As the reader undoubtedly knows there are various settings and definitions for a
connection. We begin with the most general as well as most abstract, namely that
of connections in a principal fibre bundle. We then discuss connections in associated
vector bundles, and the relationship to the former. Both formulations have their
advantages and will be used in the sequel. This brief discussion will also allow us to
1.3. CONNECTIONS IN PRINCIPAL AND ASSOCIATED VECTOR BUNDLES 19

set our notation and terminology for the remainder of the text. Again the standard
reference for much of this material is the classic text of Kobayashi and Nomizu
[KN63, KN69].
Let P be a principal bundle over M with group G. Let u ∈ P and consider the
tangent space Tu P at u. Let Gu be the vertical subspace of Tu P consisting of all
vectors that are tangent to the fibre π −1 (π(u)). The differential of the restriction
of the map ψ to the fibre through u identifies Gu with the tangent space Te G at
the identity e ∈ G, and thus with the Lie algebra g of G.
Definition 1.3.1: A connection Γ in P is an assignment to each u ∈ P a com-
plimentary subspace Hu to Gu ⊂ Tu P , called the horizontal subspace, that for all
u ∈ P satisfies:
(i) Tu P = Gu ⊕ Hu .
(ii) Hua = (Ra )∗ Hu for all a ∈ G.
(iii) Hu depends smoothly on u.
In words a connection is a G-equivariant choice of compliment to the vertical
that varies smoothly with u. If X ∈ Tu P then vX and hX denotes the vertical
and horizontal components of X, respectively. We can describe a connection in
“fancier”terminology
F as follows. We define the vertical
F subbundle of T P by VP =
u∈P G u , and the horizontal subbundle by HP = u∈P Hu . Then we have an exact
sequence of G-modules
0 −→ VP −→ T P −→ Q −→ 0.
Then a connection in P is a splitting of this exact sequence as G-modules, and thus
gives a decomposition of the tangent bundle T P as G-modules, viz.
T P = VP ⊕ HP.
For a Lie group G acting smoothly on a manifold M there is well-known homo-
morphism from the Lie algebra g of G to the Lie algebra X (M ) of smooth vector
fields on M, and if G acts effectively (which we assume) this is a monomorphism.
Given A ∈ g we let A∗ denote its image in X (M ). If we specialize to the case of the
right action of G on a principal G-bundle P , the vector field A∗ is tangent to the
fibres π −1 (π(u)) at each point u ∈ P , and is called the fundamental vertical vector
field on P associated to A ∈ g. Evaluation of A∗ at a point u ∈ P gives a vector
space isomorphism of the Lie algebra g with the vertical tangent space Gu at u.
Definition 1.3.2: Given a connection Γ on P , we define the connection form
associated to Γ to be the 1-form ω on P with values in the Lie algebra g by setting
ω(X) equal to the unique A ∈ g such that (A∗ )u is the vertical component of X.
Clearly, ω satisfies ω(X) = 0 if X is horizontal. In the sequel we are particularly
interested in the case of a circle bundle. In this case the Lie algebra g ≈ R, and
it is common to take 1 as a generator of g. Actually a connection 1-form defines a
connection as is seen by the following
Proposition 1.3.3: The connection form ω satisfies
(i) ω(A∗ ) = A for all A ∈ g.
(ii) Ra∗ ω = ada−1 ω for all A ∈ g.
Conversely, given a g valued 1-form ω on P which satisfies conditions (i) and (ii),
there is a unique connection Γ in P whose connection form is ω.
20 1. STRUCTURES ON MANIFOLDS

We shall often refer to a connection 1-form ω as simply a connection. By


localizing and using partitions of unity one has
Theorem 1.3.4: Every principal bundle P (M, G) admits a connection.
Let P (M, G) be a principal bundle and V be a finite dimensional vector space
over F = R, C or the quaternions1 Q. Let Ωr (P, V ) denote the set of smooth r-forms
on P with values in V, that is the set of smooth sections of the bundle Λr (P )⊗V. This
set has some natural algebraic structures. It is an infinite dimensional vector space
over F as well as a C ∞ (P ) ⊗R F module. Consider a representation ρ : G → Aut V
on V . A pseudotensorial form of degree r and of type (ρ, V ) is r-form on P with
values in V such that
Ra∗ φ = ρ(a−1 ) · φ.
φ is called tensorial if φ is pseudotensorial and satisfies φ(X1 , . . . , Xr ) = 0 if at
least one of the vector fields X1 , . . . , Xr is vertical. In particular, a pseudotenso-
rial 0-form, which is automatically tensorial, is just a smooth G-equivariant map
from P to V. We denote by T r (P, V ) the closed subspace of tensorial r-forms. Now
suppose that Γ is a connection in P (M, G). Then the connection form ω of Γ is
pseudotensorial of type (ad, g), but not tensorial. Notice, however, Proposition
1.3.3 implies that the difference ω 0 − ω of any two connections is tensorial. Given
any pseudotensorial form φ we can define a tensorial form by taking its horizontal
projection, i.e., φh(X1 , . . . , Xr ) = φ(hX1 , . . . , hXr ). Generally dφ is only pseu-
dotensorial, even if φ is tensorial. However, the exterior covariant derivative Dφ
defined as the horizontal projection
(1.3.1) Dφ(X1 , . . . , Xr+1 ) = dφ(hX1 , . . . , hXr+1 )
is always tensorial. In particular, if ω is a connection 1-form, Dω is a tensorial 2-
form of type (ad, g) called the curvature 2-form of ω and usually denoted by Ω. It is
a smooth section of Λ2 (P ) ⊗ g, and satisfies the famous Cartan structure equations
(1.3.2) Ω = Dω = dω + 21 [ω, ω],
as well as the well-known Bianchi identities
(1.3.3) DΩ = 0.
The meaning of the bracket in the Equation (1.3.2) is [ω, ω](X, Y ) = [ω(X), ω(Y )].
Generally, the exterior covariant derivative of any tensorial form is
Dφ = dφ + [ω, φ].
A connection Γ or ω is said to be flat if its curvature Ω vanishes.
Tensorial forms can be described alternatively in terms of associated bundles.
Let E = P ×G V be the F-vector bundle on M with standard fibre V and associated
to the F-representation ρ of G. Then there is a 1-1 correspondence between tensorial
r-forms T r (P, V ) of type (ρ, V ) and smooth sections of the bundle Λr (M ) ⊗ E as
follows: By fixing a point u ∈ P the natural projection P × V −−→E gives an F-
linear isomorphism from the vector space V to the fibre Eπ(u) . Then if X1 , . . . Xr
are tangent vectors at π(u) ∈ M we define the section φ̃ of Λr (M ) ⊗ E by
φ̃(X1 , . . . , Xr ) = uφ(X1∗ , . . . , Xr∗ ),
1We shall need all three types of vector spaces in the sequel. Some care must be taken
when working with quaternionic vector spaces, due to the noncommutativity of Q. For example,
quaternionic vector spaces are considered by multiplication from the left.
1.3. CONNECTIONS IN PRINCIPAL AND ASSOCIATED VECTOR BUNDLES 21

where X ∗ denotes any vector on P that projects to X on M. One checks that φ̃


is independent of the choices made. In particular, smooth G-equivariant functions
P −−→V correspond to smooth sections of E.
Now a connection Γ in P induces a connection ∇ in the vector bundle E and,
more generally, in the vector bundles Λr (M ) ⊗ E. We denote the C ∞ (M )-module
of smooth sections of Λr (M ) ⊗ E by Ωr (M, E).
Definition 1.3.5: A (Koszul) connection ∇ on an F-vector bundle E is an F-
linear map on sections: ∇ : Γ(E) → Λ1 (M, E) satisfying the Leibnitz rule ∇(f s) =
f ∇s + df ⊗ s, where f ∈ C ∞ (M ) and s is a smooth section of E.
Given a connection Γ on P (M, G) and an associated vector bundle E = P ×G V,
we can define a connection ∇ on E by setting
∇ = uDu−1 .
Conversely, suppose we have an F-vector bundle E with associated principal bundle
P (M, G) and F-representation ρ of G on the standard fibre V so that E = P ×G V.
Then given a connection ∇ on E, we obtain a connection Γ on P as follows: Let
s = (s1 , . . . , sk ) be a G-frame of local sections of E. If X ∈ Tx M, then the subspace
Hu of Pu with π(u) = x defined by
Hu = {s∗ X | ∇X s = 0}
is a G-equivariant complement to the vertical subspace Gu . This defines the con-
nection Γ in P. The Koszul connection ∇ in the associated bundles is often referred
to as the covariant derivative, and a section s of E is said to be covariantly constant
if ∇s = 0.
As in Proposition 1.3.3 it is convenient to express the connection ∇ in terms of
a 1-form. However, this can only be done locally in terms of a local trivialization
of the vector bundle E. Given a local trivialization E|U ' U × V, we can write
(1.3.4) ∇|U = d + AU ,
where the exterior derivative d represents the flat connection on U and AU is a g-
valued 1-form on U. Given another such trivialization on the open set W ⊂ M, and
transition functions g : U ∩ W −−→G relating the two trivializations in the overlap,
one obtains the relation
(1.3.5) AW = g −1 AU g + g −1 dg.
Conversely, given a cover of M by open sets {Uα } with g-valued 1-forms AU on each
open set in the cover satisfying Equation (1.3.5) in the overlaps, one can reconstruct
the connection ∇ on E. This formulation is essentially Cartan’s definition of con-
nection, whereas, the formulation given in Proposition 1.3.3 is due to his student
Ehresmann.
There is a natural extension of Equation (1.3.4) to the bundles Λr (M ) ⊗ E. We
shall use the notation in [DK90] and write the exterior covariant derivative as
dA = d + A : Ωr (M, E)−−→Ωr+1 (M, E).
Here the symbol A denotes a family of g-valued 1-forms on open sets of M satisfying
Equation (1.3.5) which act linearly via the representation ρ∗ : g−−→End V on the
local sections of E obtained from a local trivialization of E. The curvature form
22 1. STRUCTURES ON MANIFOLDS

Ω = Dω on P corresponds by the isomorphism u to the smooth section F A = dA A


of Λ2 (M ) ⊗ g. The Cartan structure Equations (1.3.2) then take the form

(1.3.6) F A = dA A = dA + A ∧ A,

where here again we follow the convention in [DK90] using the wedge product in
lieu of the brackets to emphasize A as an endomorphism of V via local trivializa-
tions.
Exercise
P 1.5: Show that if we write P the g-valued connection 1-form as A =
Ai dxi and its curvature 2-form F A = Fij dxi ∧ dxj in a local coordinate chart
(U ; x), Equation (1.3.6) can be written as

∂Aj ∂Ai
Fij = − + [Ai , Aj ].
∂xi ∂xj

A connection Γ in P (M, G) allows us to define the notion of parallel translation


of a fibre of P along any piecewise smooth curve in M . This is done as follows: Let
x0 , x1 ∈ M and let γ : [0, 1]−−→M be a curve in M with γ(0) = x0 and γ(1) = x1 .
Now at each point u ∈ P there is a vector space isomorphism ι : Tπ(u) M −−→Hu .
So fixing u0 ∈ π −1 (x0 ) we can lift γ horizontally to a unique piecewise smooth
curve γ ∗ in P such that γ ∗ (0) = u0 and γ̇ ∗ (t) = ι(γ̇(t)) for all t ∈ [0, 1]. This gives
an isomorphism of fibres π −1 (γ(t)) ≈ π −1 (x0 ), called parallel translation along γ.
Now suppose that γ is a loop at x0 , then parallel translation gives automorphisms
of the fibre π −1 (x0 ). By composing loops and running the loop backwards, we see
that the set of such automorphisms form a group. Moreover, if we fix a point
u ∈ π −1 (x0 ) this group can be identified with a connected Lie subgroup of G,
called the holonomy group through u and denoted by Holu . If we restrict ourselves
to loops at π(u) that are null homotopic, then we obtain a normal subgroup Hol0u
of Holu known as the restricted holonomy group through u. The groups Hol0u and
Holu enjoy the following properties:
(i) If u, v ∈ P can be joined by a horizontal curve then Holu = Holv and
Hol0u = Hol0v ,
(ii) If v = ua for a ∈ G, then Holu = Ada Holv and H0 (u) = Ada Hol0u .
A fundamental result in the theory of holonomy groups is the so-called Reduc-
tion Theorem which we now state. Of course, we refer to [KN63] for its proof:
Theorem 1.3.6: Let M be a smooth connected manifold and P (M, G) be a prin-
cipal G-bundle with a connection Γ. Let u ∈ P be an arbitrary point and let P (u)
denote the subset of points in P that can be joined to u by a horizontal curve. Then
P (u) is a reduced subbundle of P with structure group Holu , and the connection Γ
restricts to a connection on P (u).
The subbundle P (u) is called the holonomy bundle through u, and we call such
a connection reducible. It has its values in the Lie algebra holu of Holu . A well-
known theorem of Ambrose and Singer [AS53] characterizes the Lie algebra holu
as precisely the Lie algebra spanned by the curvature Ωv (X, Y ), where v ∈ P (u)
and X, Y are arbitrary horizontal vectors at v. Conversely, if Q(M, H) ⊂ P (M G)
is principal subbundle corresponding to a Lie subgroup H ⊂ G, a connection Γ in
a principal bundle Q(M, H) can be extended to a connection in P (M, G).
1.4. G-STRUCTURES 23

1.4. G-Structures
In this section we describe structures on manifolds from the unifying viewpoint
of G-structures. There are several texts where this point of view is expounded
[Kob72, Mol77, Sal89, Ste83]. Here we consider only first order G-structures,
that is, reductions of the bundle of linear frames on M. Here is the precise definition.
Definition 1.4.1: Let G ⊂ GL(n, R) be a subgroup, then a G-structure on M is
a reduction of the frame bundle L(M ) to the subgroup G. The G-structure is said
to be integrable if every point of M has a local coordinate chart (U ; x) such that
the local section
³ ∂ ∂ ´
, . . . ,
∂x1 ∂xn
of L(M ) is a local section of the reduced bundle P (M, G). Such a coordinate chart
is called admissible.
Let us return to the bundle of linear frames L(M ) of a smooth manifold M.
On L(M ) there is a canonical Rn -valued 1-form θ defined as follows: as seen in
Example 1.2.3 we can view any point u ∈ L(M ) as a vector space isomorphism
u : R−−→Tπ(u) M. So θ can be defined by

(1.4.1) hθ, Xi = u−1 π∗ X,


where X ∈ Tu L(M ) and h·, ·i denotes the natural pairing between the tangent bun-
dle to the bundle of linear frames T L(M ) and its dual cotangent bundle T ∗ L(M ).
If P (M, G) is a G-structure on M, i.e., a subbundle of L(M ), then we can restrict θ
to P (M, G) giving a canonical 1-form on the G-structure. The canonical 1-form on
a G-structure P (M, G) behaves functorially under the action of the general linear
group GL(n, R) on L(M ). Indeed we have
Lemma 1.4.2: For any a ∈ G ⊂ GL(n, R) the transformation rule holds:
Ra∗ θ = a−1 θ.

Proof. For any vector field X on P (M, G), we have


hRa∗ θ, Xi = hθ, Ra∗ Xi = (ua)−1 π∗ Ra∗ X = a−1 u−1 π∗ X = ha−1 θ, Xi. ¤

Generally, there are topological obstructions to the existence of G-structures.


To see that such a reduction does not always exist let G = {e} the identity group.
Then an {e}-structure on M is nothing but a global frame or parallelism of M. But
it is well-known that global frames do not always exist, that is that M may not
be parallelizable. For example, the 2-sphere S 2 does not even have one nowhere
vanishing vector field let alone a global frame. Even if there is a G-structure
on M there may not be an integrable G-structure. For example, as seen below
every manifold M admits many O(n, R)-structures, but a compact M admits an
integrable O(n, R)-structure only if M is covered by a torus (see Example 1.4.7
below). On the other hand every GL(n, R) structure on M is integrable. The
following proposition is evident.
Proposition 1.4.3: A G-structure P (M, G) is integrable if and only if there is an
³ (α) ´
∂xi
atlas of coordinate charts {(Uα ; x(α) )}α∈I on M whose Jacobian matrices (β)
∂xj i,j
lie in G at all points of Uα ∩ Uβ .
24 1. STRUCTURES ON MANIFOLDS

Recall from Section 1.3 that associated to any connection Γ in a principal bundle
there is the fundamental curvature 2-form. So for any connection Γ in L(M ), usually
called a linear connection, we have its curvature Ω = Dω. However, since the linear
frame bundle L(M ) has a canonical 1-from θ associated to it, we have another
2-form associated to any linear connection Γ, namely Θ = Dθ, called the torsion
2-form. In the case of linear connections we can add to Cartan’s structure Equation
(1.3.2) the so-called First Structure Equation
(1.4.2) dθ + ω ∧ θ = Θ.
We can also add to the Bianchi identities (1.3.3), the First Bianchi identities
(1.4.3) DΘ = Ω ∧ θ.
Exercise 1.6: Show that for linear connections the usual expressions for the torsion
and curvature tensors
T (X, Y ) = ∇X Y − ∇Y X − [X, Y ], R(X, Y )Z = ∇X ∇Y Z − ∇Y ∇X Z − ∇[X,Y ] Z
are related to the corresponding Cartan expressions by
T (X, Y ) = u(2Θ(X ∗ , Y ∗ )), R(X, Y )Z = u(2Ω(X ∗ , Y ∗ ))(u−1 Z),
where X ∗ , Y ∗ are the horizontal lifts of the vectors X, Y ∈ Tπ(u) M , respectively,
and u ∈ L(M ) is any point.
This entire discussion holds for any G-structure P (M, G) ⊂ L(M ) with a con-
nection Γ. We shall often refer to a linear connection Γ in a G-structure P (M, G) as
a G-connection. A linear connection Γ with Θ = 0 is said to be torsion-free. Clearly
the notions of parallel translation and holonomy apply to the case of G-structures.
We now want to put Theorem 1.2.5 to work by seeing how certain natural
tensor fields define G-structures on M. Suppose that T0 is an element of the tensor
algebra T (Rn ) over Rn and that G is the largest closed Lie subgroup of GL(n, R)
that leaves T0 invariant. Viewing a point u ∈ L(M ) as a vector space isomorphism
u : Rn −−→Tπ(u) M, we obtain an induced isomorphism u∗ : T (Rn )−−→T (Tπ(u) M ) of
the tensor algebras. Then the image T = u∗ T0 is a section of the tensor algebra
bundle T (M ). If T0 is a tensor of type (r, s) then T is a section of the tensor bundle
Tsr M. In any case because of the invariance of T0 under G, the tensor field T defines
a section of the associated bundle L(M )/G. In this case we say that P (M, G) is a
G-structure defined by the tensor T0 .
Proposition 1.4.4: Let P (M, G) be a G-structure defined by the tensor T0 . Then
P (M, G) is integrable if and only if there exists an atlas of charts {(Uα ; x(α) )}α∈I
on M such that the corresponding tensor field T = u∗ T0 has constant components
on U.
Proof. (⇒) Let P (M, G) be integrable and let (U ; x) be a coordinate chart,

then the frame u = ( ∂x 1
, . . . , ∂x∂n ) belongs to P (M, G). Let {ei } denote the standard
n j ∂
basis for R and {e } its dual basis. Then ∂x i
= u(ei ), and dxj = u(ej ). So if
T0 = tji11···i j1
···js ei1 ⊗ · · · ⊗ eir ⊗ e ⊗ · · · ⊗ e
r js

is a tensor of type (r, s), we have


∂ ∂
T = u∗ T0 = tji11···i
⊗ ··· ⊗
···js
r
⊗ dxj1 ⊗ · · · ⊗ dxjs .
∂xi1 ∂xir
Hence, T has constant components on U.
1.4. G-STRUCTURES 25

(⇐) Conversely, suppose that (U ; x) is a coordinate chart such that T = u∗ T0


has constant components. So T is a constant section of the associated bundle
L(U )/G. Then there is a linear transformation of coordinates on U to new coor-
dinates (y1 , . . . , yn ) such that the frame u = ( ∂y∂ 1 , . . . , ∂y∂n ) belongs to P (M, G).
Thus, the G-structure P (M, G) is integrable. ¤
One easily sees that if a G-structure P (M, G) is defined by a tensor field T and Γ
is a G-connection, then T is covariantly constant with respect to Γ, or equivalently,
∇T = 0. Generally, the ‘integrability condition’ of Definition 1.4.1 can be very
restrictive, and we will discuss various ‘levels of integrability’. First, it is easy to
show that
Proposition 1.4.5: An integrable G structure admits a torsion-free connection.
It is now convenient to consider a much less restrictive definition of integrability.
Definition 1.4.6: A G-structure P (M, G) is said to be 1-integrable if it admits
a torsion-free G-connection.
So an integrable G-structure is automatically 1-integrable. The failure of the
existence of torsion-free G-connection can be seen as the first order obstruction to
integrability, hence, the name 1-integrable. The question of uniqueness of a torsion-
free connection, assuming one exists, is related to prolongations of G-structures
which we shall briefly treat below (See Definition 1.6.8). We now wish to consider
many examples of G-structures. Our first example is a good example where non-
integrable G-structures are of more interest than the integrable ones.
Example 1.4.7: Riemannian metrics. We consider a reduction of the frame
bundle L(M ) to the orthogonal group
O(n, R) = {A ∈ GL(n, R) | At A = 1ln } .
The reduced bundle O(M ) ⊂ L(M ) is called the orthonormal frame bundle of M.
Theorem 1.2.5 says that such a reduction is equivalent to a choice of section of the
associated bundle E(M, GL(n, R)/O(n, R), L(M )) = L(M )/O(n, R). We show that
such a section is just a Riemannian metric on M. As mentioned previously each
point u ∈ L(M ) gives an isomorphism of the standard vector space Rn with the
tangent space Tπ(u) M. Let h·, ·i denote the Euclidean inner product on Rn , and let
X, Y ∈ Tπ(u) M. Then
gu (X, Y ) = hu−1 X, u−1 Y i
defines an inner product on Tπ(u) M. Furthermore, if a ∈ O(n, R) we have
gua (X, Y ) = h(ua)−1 X, (ua)−1 Y i = ha−1 u−1 X, a−1 u−1 Y i
= hu−1 X, u−1 Y i = gu (X, Y ),
where the second to the last equality holds by the invariance of h·, ·i under O(n, R).
This shows that gu is constant along the fibres of O(M ), and thus, is a section of
L(M )/O(n, R). Thus, the associated bundle L(M )/O(n, R) can be identified with
a subbundle of the vector bundle Sym2 T ∗ M of symmetric covariant 2-tensors on
M. The choice of u modulo O(n, R) corresponds to a choice of Riemannian metric
on M. Since manifolds are paracompact the standard partition of unity argument
shows that such sections and hence, such reductions always exist.
When are O(n)-structures integrable? According to Proposition 1.4.4 this oc-
curs when the metric tensor g has constant components in some coordinate chart
26 1. STRUCTURES ON MANIFOLDS

(U ; x), and then by a change of coordinate, say (V ; y) the metric can be brought
to the form X
g= (dyi )2
i
in V. By the Second Fundamental Theorem of Riemannian geometry this happens
precisely when the Riemann curvature tensor vanishes. For example, if M is a com-
pact 2-dimensional manifold, then it is either a torus or a Klein bottle depending
on whether it is orientable or not. This is quite restrictive. On the other hand
the First Fundamental Theorem of Riemannian geometry says that there exists
a unique torsion-free Riemannian connection, denoted ∇g ≡ ∇, called the Levi-
Civita connection. So all Riemannian G-structures are 1-integrable. Similarly, a
reduction to the group SO(n, R) = O(n, R) ∩ GL+ (n, R) corresponds to oriented
Riemannian geometry. In particular, one can consider the parallel translation de-
fined by the Levi-Civita connection and its associated holonomy group which is a
subgroup of the structure group O(n, R) (SO(n, R) in the oriented case). Since this
connection ∇g is uniquely associated to the metric g, we denote it by Hol(g), and
refer to it as the Riemannian holonomy group or just the holonomy group when
the context is clear. Indeed, it is precisely this Riemannian holonomy that plays
an important role in this book. Now on a Riemannian manifold (M, g) there is a
canonical epimorphism π1 (M )−−−→Hol(g)/Hol0 (g), in particular, if π1 (M ) = 0 then
Hol(g) = Hol0 (g). In 1955 Berger proved the following theorem [Ber55] concerning
Riemannian holonomy:
Theorem 1.4.8: Let (M, g) be an oriented Riemannian manifold which is neither
locally a Riemannian product nor locally symmetric. Then the restricted holonomy
group Hol0 (g) is one of the following groups listed in Table 1.4.1.

Table 1.4.1: Berger’s Riemannian Holonomy Groups


Hol0 (g) dim(M ) Geometry of M Comments
SO(n) n orientable Riemannian generic Riemannian
U (n) 2n Kähler generic Kähler
SU (n) 2n Calabi-Yau Ricci-flat Kähler
Sp(n) · Sp(1) 4n quaternionic Kähler Einstein
Sp(n) 4n hyperkähler Ricci-flat
G2 7 G2 -manifold Ricci-flat
Spin(7) 8 Spin(7)-manifold Ricci-flat

We will encounter all the geometries listed in this table throughout this book.
Most of them will already be introduced in this chapter as G-structures. Orig-
inally, Berger’s list included Spin(9) but Alekseevsky proved that any manifold
with such holonomy group must be symmetric [Ale68]. In the same paper Berger
also claimed a classification of all holonomy groups of torsion-free affine (linear)
connections that act irreducibly. He produced a list of possible holonomy rep-
resentations up to what he claimed was a finite number of exceptions. But his
classification had some gaps discovered 35 years later by Bryant [Bry91]. An infi-
nite series of exotic holonomies was found in [CMS96] and finally the classification
in the non-Riemannian affine case was completed by Merkulov and Schwachhöfer
[MS99]. We refer the reader to [MS99] for the proof, references and the history of
the general affine case. In the Riemannian case a new geometric proof of Berger’s
1.4. G-STRUCTURES 27

Theorem is now available [Olm05]. An excellent review of the subject just prior
to the Merkulov and Schwachhöfer’s classification can be found in [Bry96]. We
should add that one of the first non-trivial results concerning manifolds with the
exceptional holonomy groups of the last two rows of Table 1.4.1 is due to Bonan
[Bon66] who established Ricci-flatness of manifolds with parallel spinors.
It is clear in the Riemannian case that non-integrable structures are much
more interesting than integrable ones, and there is a big gap between the integrable
structures and the 1-integrable structures. The obstructions to integrability occur
at order two, namely the Riemannian curvature. In order to study these further one
needs to study certain invariants of the G-structure which involves a detailed study
of the Riemannian curvature. For a general G-structure these invariants involve
Spencer cohomology to which we refer the reader to the literature [Fuj72, KS72,
SS65, Ste83]. In the case of O(n, R)-structures the relevant Spencer cohomology
group consists of the Riemann curvature tensor. In our next example the integrable
case is very rich.
Example 1.4.9: Almost complex structures. A complex structure on a real
vector space V is an endomorphism J of V that satisfies J 2 = −1l, where 1l denotes
the identity endomorphism of V. We can give V the structure of a complex vector
space by defining scalar multiplication with scalars in C by
(a + ib)v = av + bJv.
Conversely, if V is a complex vector space we define the endomorphism J on V by
Jv = iv.
Let {v1 , . . . , vn } be basis for V as a complex vector space, then it is easy to check
that {v1 , . . . , vn , Jv1 , . . . , Jvn } is a basis for V as a real vector space. In particular,
this implies that any real vector space with a complex structure has even real
dimension. For example, by writing zj = xj + iyj for j = 1, . . . , n, the complex
vector (z1 , . . . , zn ) in the complex vector space Cn is identified with the real vector
(x1 , . . . , xn , y1 , . . . , yn ) in the real vector space R2n . The canonical complex structure
J0 on V is given in block form with respect to the standard basis of R2n by
µ ¶
0 1ln
J0 = ,
−1ln 0
where 1ln denotes the n × n identity matrix. We can identify the complex linear
group GL(n, C) with the subgroup of matrices in GL(2n, R) that commute with J0 .
Thus, the complex n × n matrix A + iB is identified with the real 2n × 2n matrix
µ ¶
A B
.
−B A
On a manifold M an almost complex structure is an endomorphism Jx of the
tangent space Tx M that varies smoothly with x and satisfies Jx2 = −1lx at each
point x ∈ M, or in other words, a smooth section J of the endomorphism bundle
End T M satisfying J 2 = −1l. As in the Riemannian case such a section can be
obtained from the canonical complex structure J0 on R2n by identifying u ∈ L(M )
with a linear map u : R2n −−→Tπ(u) M. Thus, we define

Ju = uJ0 u−1 .
28 1. STRUCTURES ON MANIFOLDS

The invariance of J0 under the subgroup GL(n, C) ⊂ GL(2n, R) shows that Ju is


constant along the fibres of the bundle CL(M ) of complex linear frames on M and
thus, gives a section of the associated bundle E(M, GL(2n, R)/GL(n, C), L(M )) =
L(M )/GL(n, C) which is the subbundle of End T M consisting of endomorphisms
that satisfy J 2 = −1l. Unlike in the Riemannian case it is not always possible to find
such a section, or equivalently to obtain a reduction of the real frame bundle L(M )
to the complex linear group GL(n, C). In general, this is a fairly deep topological
question. The existence of an almost complex structure J on M is equivalent to
the splitting of the complexified tangent bundle T M ⊗ C = T 1,0 M ⊕ T 0,1 M. If
one has such a splitting, J can be defined to be multiplication by i on T 1,0 M
and multiplication by −1 on T 0,1 M. Conversely, given J one easily sees that it is
diagnonalizable over T M ⊗ C with eigenvalues ±i.
By Proposition 1.4.3 a GL(n, C)-structure is integrable when there is an atlas
(α) (α) (α) (α)
of coordinate charts {Uα ; x1 , . . . , xn , y1 , . . . , yn }α∈I whose Jacobian matrix
(α)
lies in GL(n, C) ⊂ GL(2n, R). This corresponds to the complex coordinates zj =
(α) (α) (β)
xj +iyj being independent of the complex conjugate coordinates z̄i , that is the
transition functions are holomorphic or equivalently satisfy the Cauchy-Riemann
equations. An integrable almost complex structure is called a complex structure,
and a smooth manifold M with a complex structure is called a complex manifold.
The structure tensor of the G-structure that measures the failure of integrability is
the Nijenhuis torsion tensor defined by
(1.4.4) NJ (X, Y ) = [JX, JY ] − [X, Y ] − J[X, JY ] − J[JX, Y ].
The vanishing of this tensor field is equivalent to the integrability of the almost com-
plex structure. If one assumes that M is real analytic this follows from a Frobenius
integrability argument (cf. [KN69], Appendix 8). However, in the C ∞ case this is
a much deeper result due to Newlander and Nirenberg [NN57]. Summarizing an
almost complex structure is integrable if and only if the Nijenhuis torsion tensor
NJ vanishes. This is indeed the torsion of an almost complex connection, so the
only obstruction to integrability occurs at level one, that is, in this case 1-integrable
implies integrable.
Following up on the remark at the end of the second paragraph of Example
1.4.9, one can ask why a partition of unity argument doesn’t work in the case of
almost complex structures. The point is that one cannot guarantee the condition
J 2 = −1l at every point of the manifold by using partitions of unity. A better under-
standing is obtained from the point of view of G-structures. An alternative proof
of the existence of Riemannian metrics on any manifold is as follows: by Theorem
1.2.5 giving a Riemannian metric is equivalent to giving a reduction of the frame
bundle to the orthogonal group O(n, R). This reduction can be accomplished (gen-
erally over any paracompact space) since as topological spaces the general linear
n(n+1)
group GL(n, R) decomposes as O(n, R) × R 2 , the so-called Iwasawa decompo-
sition. Thus, GL(n, R) is homotopy equivalent to O(n, R). Then paracompactness
allows one to construct this homotopy equivalence at each point of M giving the
reduction. However, in the case of reducing L(M ) to the complex linear group
GL(n, C) there is no such homotopy equivalence. Indeed, GL(2n, R) is homotopy
equivalent to O(2n, R), whereas, again the Iwasawa decomposition for GL(n, C)
says that GL(n, C) is homotopy equivalent to the unitary group U (n), that is, the
subgroup of GL(n, C) that leaves invariant the standard Hermitian form h·, ·i on
1.4. G-STRUCTURES 29

Cn . It is well-known that O(2n, R) and U (n) are not homotopy equivalent, e.g.
they have different homology groups. So, generally, there are obstructions to such
reductions. A general theory of obstructions is expounded in the classic text of
Steenrod [Ste51]. A well-known example of an even dimensional manifold which
admits no almost complex structure is the 4-sphere S 4 . This can be shown either by
obstruction theory or using the Atiyah-Singer index theorem. Another interesting
example is S 6 which is known to admit an almost complex structure, but it is still
an open question whether S 6 has an integrable almost complex structure, i.e., a
complex structure.
Exercise 1.7: Let G = GL+ (n, R) be the subgroup of matrices in GL(n, R) whose
determinant is positive. Show that M admits such a G-structure if and only if M
is orientable. Show that a GL+ (n, R)-structure on M coincides with a choice of
orientation of M. What about integrability?
Exercise 1.8: Show that an almost complex structure on M determines an orien-
tation on M, hence, any almost complex manifold is orientable.
Notice that an oriented Riemannian manifold gives a reduction of L(M ) to the
special orthogonal group SO(n) = {A ∈ O(n) | det(A) = 1}, and that SO(n) =
GL+ (n, R) ∩ O(n). We will frequently use G-structures that occur as a combination
of other G-structures or, alternatively, a G-structure defined by more than one
tensor. Another important example of this is:
Example 1.4.10: Almost Hermitian structures. Let G = U (n) the unitary
group defined by
U (n) = {A ∈ GL(n, C) | Āt A = 1ln }.
This is precisely the group that leaves the standard Hermitian form hu, vi = u¯1 v1 +
· · · + u¯n vn in Cn invariant. Since U (n) = GL(n, C) ∩ O(2n), a U (n)-structure or
almost Hermitian structure consists of an almost complex structure J together with
a Riemannian metric g on M satisfying the compatibility condition g(JX, JY ) =
g(X, Y ). The two tensor fields g and J give rise to a 2-form ω called the Kähler
form defined by ω(X, Y ) = g(X, JY ). The structure is called almost Kähler if ω is
closed, Hermitian if J is integrable, and it is called Kähler if both J is integrable
and ω is closed. A U (n)-structure is integrable if both the almost complex structure
and the Riemannian structures are integrable which is very restrictive. The only
compact examples are covered by a complex torus. A less restrictive notion is that
of being 1-integrable. A U (n)-structure is 1-integrable if and only if it is Kähler.
This is equivalent to the U (n) bundle coinciding with the holonomy bundle. We
shall discuss these structures in much more detail in Chapter 3.
These ideas can also be applied to indefinite or pseudo-Riemannian and pseudo-
Hermitian metrics2. For example,
Example 1.4.11: Pseudo-Riemannian structures. Let O(p, q) denote the sub-
group of GL(n, R) that leaves the quadratic form u21 + · · · + u2p − u2p+1 − · · · − u2p+q
invariant. Then an O(p, q)-structure on M is the same as a choice of pseudo-
Riemannian metric g on M of signature (p, q). If p or q equals 1, then this is called
a Lorentzian structure and g a Lorentzian metric. Similarly, one has almost pseudo-
Hermitian structures by considering the pseudo-unitary groups U (p, q) which is the

2One should be cautioned that the terminology pseudo-Hermitian structure is used in an


entirely different sense in [Web78]
30 1. STRUCTURES ON MANIFOLDS

group that leaves invariant an indefinite Hermitian form of signature (p, q). Clearly,
U (p, q) = GL(p + q, C) ∩ O(2p, 2q). The integrability question is similar to that of
Example 1.4.7.
Unlike Riemannian metrics (structures), pseudo-Riemannian metrics do not
always exist on M. See for example [HE73, O’N83] for the existence of Lorentzian
metrics on M. In fact, it is easy to see that M admits a Lorentzian metric if and
only if it admits a nowhere vanishing vector field.
Exercise 1.9: Define the group
CO(n) = {A ∈ GL(n, R) | At A = c1ln , c > 0} .
Show that a CO(n)-structure on M coincides with a conformal class of Riemannian
metrics on M. Recall that a conformal class is an equivalence class of Riemannian
metrics, where the Riemannian metrics g and g 0 are equivalent if there exists a
positive function λ on M such that g 0 = λg.
A CO(n)-structure on M is called a conformal structure on M ; however, CO(n)
is usually not called the conformal group, a name that is usually reserved for the
automorphism group of a conformal structure, and this group is generally larger
than CO(n). This phenomenon is related to prolongations of G-structures which is
briefly touched upon below. The reader is referred to [Kob72, Ste83] for complete
discussions of prolongations of G-structures. Conformal structures can be defined
similarly in the case of pseudo-Riemannian structures. Our next example will also
be of fundamental importance to us.
Example 1.4.12: Almost symplectic structures. Consider the non-degenerate
antisymmetric bilinear form
ω0 (u, v) = u1 vn+1 + · · · + un v2n − un+1 v1 − · · · − u2n vn
on the vector space R2n . Notice that ω0 can also be written in terms of the standard
Euclidean metric h·, ·i and canonical complex structure J0 as
ω0 (u, v) = hu, J0 vi.
The subgroup of GL(2n, R) that leaves ω0 invariant is called the real symplectic
group and denoted3 by Sp(n, R). It is easy to see that this group can be defined by
Sp(n, R) = {A ∈ GL(2n, R) | At J0 A = J0 }.
An almost symplectic structure on a manifold M is then given by transporting
the antisymmetric form ω0 to the manifold as before by the linear isomorphism
u : R2n −−→Tp M, viz. for X, Y ∈ Tp M we define
ω(X, Y ) = ω0 (u−1 X, u−1 Y ).
ω is a section of the bundle E(M, GL(2n, R)/Sp(n, R), L(M )) = L(M )/Sp(n, R)
which can be identified with the subbundle of the exterior bundle Λ2 M bundle con-
sisting of non-degenerate 2-forms. By Theorem 1.2.5 this corresponds to a reduction
of L(M ) to the real symplectic group Sp(n, R). The appearance of the tensor J0 sug-
gests that there may be a connection between almost complex structures and almost
symplectic structures. This is indeed the case for Sp(n, R) ∩ O(2n, R) = U (n) =
GL(n, C) ∩ O(2n, R) Since, an O(2n, R)-structure exists on any 2n-dimensional
3The notation for this group is by no means standard. It is sometimes written as Sp(2n, R),
and is written as Sp(2n) in [MS98]. We shall employ the notation Sp(n) for the ‘compact
symplectic group’ which is different.
1.4. G-STRUCTURES 31

manifold, we see that M 2n admits an almost symplectic structure if and only if it


admits an almost complex structure. Of course, the two structures are different.
By Proposition 1.4.4 an Sp(n, R)-structure is integrable if and only if the tensor
field ω has constant components with respect to some local coordinates. In this
case there are local coordinates (x1 , . . . , xn , y1 , . . . , yn ) on M such that ω takes the
form
X n
ω= dxi ∧ dyi .
i=1
A well-known theorem of Darboux says that an Sp(n, R)-structure is integrable if
and only if the tensor field ω is closed, that is dω = 0. An integrable Sp(n, R)-
structure on M is called a symplectic structure. In this case the pair (M, ω) is
called a symplectic manifold.
Let (M, ω) be a symplectic manifold and consider the group S(M, ω) of sym-
plectomorphisms on M, i.e., the subgroup of Diff M that leaves ω invariant. The
‘Lie algebra’ s(M, ω) of S(M, ω) is the subalgebra of vector fields X on M such that
£X ω = 0. This equation implies that the 1-form X ω is closed. If it is also exact
then the there is a smooth function H on M such that X ω = dH. In this case
the function H is called a Hamiltonian function and the X a Hamiltonian vector
field. Notice that H is only defined up to a constant and if H 1 (M, R) = 0 then
every X ∈ s(M, ω) is Hamiltonian. We refer to [LM87] for further development.
From the discussion above about Lie groups, it should be clear that an almost
Hermitian structure gives naturally an almost symplectic structure. So we recover
Example 1.4.10 as a particular case of an almost symplectic structure with an added
compatible Riemannian structure.
Exercise 1.10: Define the conformal symplectic group by
CSp(n, R) = {A ∈ GL(2n, R) | At J0 A = cJ0 , c ∈ R∗ } ,
where R∗ denotes R − {0}. Show that a CSp(n, R)-structure on M coincides with
a choice of equivalence class of non-degenerate 2-forms, where the 2-forms ω and
ω 0 are equivalent if there is a nowhere zero function λ on M such that ω 0 = λω.
Such a structure can be called a conformal almost symplectic structure. Show that
if the CSp(n, R)-structure is integrable, then the non-degenerate 2-forms are closed
and λ is a constant. The resulting structure is known as a conformal symplectic
structure.
Example 1.4.13: p-dimensional distribution A p-dimensional distribution on
an n-dimensional manifold M is a choice of p-dimensional subbundle E of the
tangent bundle T M. The group G that stabilizes E is given by
nµ ¶ o
A 0
GL(p, q; R) = | A ∈ GL(p, R), B ∈ GL(q, R), C ∈ Mq,p ,
C B
where Mp,q denotes the p by q real matrices, and p + q = n. An integrable
GL(p, q; R)-structure on M is just a foliation F of M, that is, every point of M
has a ‘foliated coordinate chart’ (U ; φ) with coordinates (x1 , . . . , xp ; y1 , . . . , yq ) such

that ( ∂x 1
, . . . , ∂x∂ p ) spans E|U, and the Jacobian matrix of the change of foliated
charts lies in GL(p, q; R). An atlas U = {(Uα ; φα )}α∈I of such foliated charts is
called a foliated atlas for M. Two foliated atlases U and V of the same codimen-
sion q correspond to the same foliation F if and only if they satisfy the coherence
property that the disjoint union U t V is a foliated atlas for F.
32 1. STRUCTURES ON MANIFOLDS

Globally, the foliation F gives a partition of M into disjoint p-dimensional


immersed submanifolds Lα , called the leaves of F. By the classical Frobenius The-
orem, a p-dimensional distribution is integrable if and only if it involutive, that is
the Lie bracket of any sections of the subbundle E is also a section of E. This has a
dual formulation as follows: let E 0 denote the subbundle of T ∗ M that annihilates
E. Then the Frobenius Theorem says that E is integrable if and only if the ideal
I(E 0 ) generated by the sections of E 0 is closed under exterior differentiation, i.e.,
I(E 0 ) is a differential ideal. Of particular interest in foliation theory is a study
of the transverse geometry and this is studied in more detail in Chapter 2. The
transverse geometry is the geometry of the space of leaves of the foliation defined
to be the quotient space defined by the equivalence relation that x ∼ y if x and
y lie on the same leaf. However, this can be very crudely defined as a topological
space, so one studies the transverse geometry directly on the foliated manifold.
Alternatively, a foliation can be described by patching together local submer-
sions. This is the point of view advanced by Haefliger [Hae70] and will be discussed
in Section 2.2. A (global) submersion (hence, a fibre bundle) is a special case of a
foliation. It is often called also a simple foliation. More development of foliation
theory is given below in Chapter 2.
There are various special cases of Example 1.4.13 of special interest.
Example 1.4.14: Almost product structures. This is Example 1.4.13 with
C = 0. In this case the tangent bundle to M splits as a direct sum of two subbun-
dles T M = D+ ⊕ D− . By Proposition 1.4.3 an integrable almost product structure
implies the integrability of the complementary subbundles. In particular, the uni-
versal covering M̃ is actually a product of two manifolds, one of dimension p and
the other of dimension q. One can easily see that an almost product structure is
equivalent to the existence of an endomorphism P ∈ End(T M ) such that P 2 = 1l.
Namely, given such P we can introduce the projection operators
1
P± = (1l ± P ), P+ ◦ P− = P− ◦ P+ = 0, P+ + P− = 1l.
2
We get the desired spitting T M = D+ ⊕ D− by taking D± = P± (T M ). As in
the almost complex case the Nijenhuis tensor NP (X, Y ) is the obstruction to the
integrability of the two distributions D± . A Riemannian almost product structure
is a further reduction of the structure group to O(p, R) × O(q, R) which is not
obstructed, i.e., if a manifold admits an almost product structure it also admits an
almost product Riemannian metric structure. It follows that a Riemannian almost
product structure is a triple (M, Π, g), where Π is an almost product structure and
g is a Riemannian metric with the property
g(P X, P Y ) = g(X, Y ).
A Riemannian almost product structure is said to be a locally Riemannian product
structure if ∇P = 0, i.e., Π is parallel with respect to the Levi-Civita connection of
g. The condition ∇P = 0 implies the integrability NP (X, Y ) = 0. Note also that if
a manifold M admits an almost product structure with group GL(p, R) × GL(q, R)
then it also admits an O(p, q) pseudo-Riemannian structure.
Exercise 1.11: Let (M n , P, g) be an almost product structure. Show that P is a
local O(n, R) matrix with all eigenvalues ±1 and Tr(P ) = 2q −n, where q = rk(D+ )
is the number of positive eigenvalues. Furthermore, show that one can choose a
local chart and a basis so that P, P± take the following simple forms
1.4. G-STRUCTURES 33

µ ¶ µ ¶ µ ¶
1lq 0 1lq 0 0 0
P = , P+ = , P− = .
0 −1ln−q 0 0 0 1ln−q

We are especially interested in almost product structures when p = 2m and


q = 1 which is related to almost contact structures treated in detail in Chapter 6.
Unfortunately, the terminology with regard to contact and almost contact struc-
tures is very misleading. For example, a contact structure is generally not an almost
contact structure in contrast to the usual use of this terminology; moreover both
of these structures are defined in [Gra59]! This was noted and fixed in [Sto74], at
least for topological purposes, but we would like consistency with geometric struc-
tures as well. We shall refer to our structure as an almost contact G-structure to
distinguish it from an almost contact structure whose definition is by now too well
ensconced in the literature.
Example 1.4.15: Almost contact G-structure. Consider the group G defined
as follows:
nµ ¶ o
A 0
G= | A ∈ CSp(m, R), b ∈ GL(1, R), c ∈ M1,2m .
c b
This G-structure picks out a codimension 1 subbundle with a conformal almost
symplectic structure on it. Here an integrable G-structure is related to what could
be called a conformal cosymplectic structure.
In the case of Example 1.4.15 we shall be more interested in the non-integrable
case. In Chapter 6 we shall describe its relation with almost contact and contact
structures.
Another structure that is related to both Examples 1.4.9 and 1.4.13 is a CR-
structure. Here the concept of an almost CR-structure is not standard. Indeed
the definition of a CR-structure varies as well. Some authors require that a CR-
structure be defined by a codimension one subbundle (cf. [Jac90]), the idea being
that of modelling a real hypersurface in Cn+1 . We shall be more general.
Example 1.4.16: Almost CR-structures. An almost CR-structure of codimen-
sion q is a G-structure with G given by
nµA 0 ¶ o
G= | A ∈ GL(m, C), B ∈ GL(q, R), C ∈ Mq,2m ,
C B
where GL(m, C) is realized as a subgroup of GL(2m, R) as described in Example
1.4.9. This G picks out a 2m-dimensional subbundle E of T M with an almost
complex structure, that is a smooth section J of the endomorphism bundle End E
satisfying J 2 = −1lE . We say that (E, J) defines a CR-structure on M if for any
smooth sections X, Y of E the vector field [JX, Y ] + [X, JY ] is a section of E and
the Nijenhuis tensor NJ , as defined in Example 1.4.9, vanishes. We are interested
in the case of codimension one, i.e., q = 1. If M is orientable then there exists a
1-form η on M such that E = ker η. The symmetric bilinear form L(X, Y ) on E
defined by L(X, Y ) = dη(X, JY ) is called the Levi form of the CR-structure. The
CR-structure is said to be strictly pseudoconvex if L is either positive or negative
definite. For an extensive treatment of CR-structures we refer the reader to the
recent book [DT06].
34 1. STRUCTURES ON MANIFOLDS

Example 1.4.17: f -structures. An f -structure of corank q is a special kind of


an almost product structure. It is a G-structure with G given by
nµA 0 ¶ o
G= | A ∈ GL(n, C), B ∈ GL(q, R) .
0 B
f -structures were first introduced and studied by Yano [Yan63]. One can easily see
that the reduction implies an existence of a non-vanishing section f of the bundle
End(T M ) with constant rank such that
f 3 + f = 0.
Given f we define two commuting projections P+ = −f 2 and P− = f 2 + 1l, which,
in turn, give the splitting T M = D+ ⊕ D− . Clearly, f 2 (f X) = −f (X) so that f
defines and almost complex structure on the subbundle D+ = Im(f ) ⊂ T M . In
particular, further reduction to the subgroup U (n) × O(q, R) can be achieved by
choosing a locally product Riemannian metric g on M which is compatible with f
in the following natural sense
g(f X, Y ) = −g(X, f Y ).
Clearly, such a metric always exists and the triple (M, f, g) is said to be a metric f -
structure. When q = 0, an f -structure is simply an almost complex structure, and
with a compatible metric it is an almost Hermitian structure of two earlier examples.
When q = 1 an f -structure is a special case of an almost contact G-structure. In
particular, when q = 1 and the structure group reduces further to GL(n, C) × 1 an
f -structure is a (strict) almost contact structure discussed in Chapter 6. It follows
that there exist a nowhere vanishing vector field ξ and its dual 1-form η such that
[Bla76a]
(1.4.5) f = −1l + η ⊗ ξ.
Note that in the latter case picking a compatible Riemannian metric, i.e., the
metric g which satisfies g(f X, f Y ) = g(X, Y ) − η(X)η(Y ) is always possible and
gives further reduction of the structure group to U (n) × 1 which is the standard
definition of an almost contact (metric) structure due to Gray [Gra59]. When q ≥ 2
one can consider the so-called globally framed (metric) f -structure as the further
reduction if the structure group to U (n) × 1lq . The definition implies that the
bundle D− = Ker(f ) is parallelizable, i.e., there exist global frame field {ξ1 , . . . , ξs }
spanning D− = Ker(f ). Such structures were introduced and studied by Goldberg
and Yano [GY71, GY70] and Blair [Bla70]. In some literature they are also
called f.pk-structure (see the recent book [FIP04] for more details and references).
Example 1.4.18: Quaternionic structures. Quaternionic G-structures ap-
peared early on in the work of Libermann [Lib54], Berger [Ber55], Obata [Oba56],
and Martinelli [Mar59], and continued in [Ale68, Bon67, Gra69a, Wol65] to
cite a few. Our brief treatment here follows that of Salamon [Sal86] with further
development in Chapter 12. Let Hn be a quaternionic vector space with n > 1.
By writing each quaternionic coordinate as ui = u0i + u1i i + u2i j + u3i k we natu-
rally identify Hn ' R4n . Let GL(n, H) ⊂ GL(4n, R) be a group of non-singular
quaternionic n × n matrices and consider the product GL(n, H) × GL(1, H). For
any (A, λ) we can define GL(n, H) × GL(1, H)-action on Hn by (u; (A, λ)) 7→ A · uλ.
The central R∗ acts trivially and we denote the quotient group by GL(n, H)Sp(1),
1.4. G-STRUCTURES 35

where Sp(1) ⊂ GL(1, H) is the subgroup of unit quaternions. In particular, it fol-


lows that G = GL(n, H)Sp(1) ¡ is a subgroup
¢ of GL(4n, R), and one easily sees
that GL(n, H)Sp(1) = GL(n, H) × Sp(1) /Z2 . An almost quaternionic struc-
ture on a manifold M is a reduction of the frame bundle L(M ) to the subgroup
GL(n, H)Sp(1). Theorem 1.2.5 gives a one-to-one correspondence between sections
of the associated vector bundle E(M, GL(4n, R)/G and reductions Q of the frame
bundle L(M ) to the group G. Moreover, in analogy with Example 1.4.9 we see that
sections of E(M, GL(4n, R)/G) correspond to the existence of a triple of local sec-
tions {I1 , I2 , I3 } of a rank 3 vector subbundle Q ⊂ End(T M ) satisfying the algebra
of the quaternions. Hence, one can consider either a reduction to the bundle Q
or the existence of the subbundle Q as an almost quaternionic manifold. In the
world of quaternionic geometry the usual notion of integrability is much too strong.
For example, the only simply connected compact manifold which admits a quater-
nionic atlas whose transition functions are in GL(n, H)Sp(1) is the quaternionic
projective space HP(n) [Mar70, Som75]. More interesting quaternionic geome-
tries emerge by considering 1-integrable (recall Definition 1.4.6) almost quaternionic
G-structures. A 1-integrable almost quaternionic structure is called a quaternionic
structure. In the case of quaternionic structures torsion-free connections are not
unique. Such a connection [Opr77, Opr84] is called an Oproiu connection in
[AM96b] or a quaternionic connection.
One can always equip M with a Riemannian metric g which preserves Q, i.e.,
g(JY, JY ) = g(X, Y ) for any local section J of Q. The triple (M, Q, g) is then
called almost quaternionic Hermitian manifold. This corresponds to a further re-
duction of the bundle Q to its maximal compact subgroup Sp(n)Sp(1). An almost
quaternionic (Hermitian) manifold with a torsion-free connection preserving Q is
called a quaternionic (Hermitian) manifold [Sal86]. In dimension greater than 4 an
almost quaternionic Hermitian structure is 1-integrable if and only if it is quater-
nionic Kähler [AM96b]. The usual definition of a quaternionic Kähler structure
(in dimension greater than 4) is that the Riemannian holonomy lies in Sp(n)Sp(1).
Quaternionic structures will be described in more detail in Chapter 12. In dimen-
sion 4 the definitions of quaternionic and quaternionic Kähler involve ‘duality’ of
Weyl’s conformal curvature and will be defined in Chapter 12.
Example 1.4.19: Hypercomplex structure. A GL(n, H)-structure on a mani-
fold M is called an almost hypercomplex structure. Thus, an almost hypercomplex
manifold is the special case of an almost quaternionic manifold, where the sub-
bundle Q ⊂ End(T M ) admits a trivialization by global sections I ≡ {I1 , I2 , I3 }
satisfying quaternion algebra. Notice that the quaternionic relations on Q allows
one to define a two-sphere’s worth of almost complex structures

(1.4.6) I(τ ) = τ1 I1 + τ2 I2 + τ3 I3 ,

where τ = (τ1 , τ2 , τ3 ) ∈ S 2 . So M together with I(τ ) is usually called an almost


hypercomplex manifold. A GL(n, H)-connection satisfies ∇I(τ ) = 0 for all τ ∈
S 2 . An almost hypercomplex structure that is 1-integrable is called hypercomplex.
In this case torsion-free GL(n, H)-connections are unique. Such a connection is
called the Obata connection . Compact four dimensional manifolds admitting a
hypercomplex structure were classified in [Boy88a].
One can always equip M with a Riemannian metric g which preserves I(τ ),
i.e., g(I(τ )Y, I(τ )Y ) = g(X, Y ) for any τ ∈ S 2 . The triple (M, I, h) is then called
36 1. STRUCTURES ON MANIFOLDS

an almost hyperhermitian manifold and provides further reduction of the structure


group to Sp(n). A hypercomplex structure with a compatible Riemannian metric
is called a hyperhermitian structure. An almost hyperhermitian structure is 1-
integrable if and only if it is hyperkähler. This is equivalent to the usual definition
of a hyperkähler structure as one in which the Riemannian holonomy is contained
in Sp(n). Both hypercomplex and hyperkähler structures are treated in more detail
in Chapter 12.
Example 1.4.20: Quaternionic f -structures. A G-structure with the group
G = GL(n, H)Sp(1) × 1lm ⊂ GL(4n + m, R) is called an almost quaternionic f -
structure of corank m. It can be seen that such a reduction is equivalent to the
existence of two subbundles Q ⊂ End(T M ) and E ⊂ T M of rank 3 and m respec-
tively, such that Q admits local sections {J1 , J2 , J3 } satisfying quaternionic algebra
relations, and E is parallelizable. One can choose global sections {ξ1 , . . . , ξm } of E
so that
Xm
Ji (ξα ) = 0, ηα ◦ Ji = 0, Ji2 = −1l + ηα ⊗ ξα ,
α=1
where α = {1, . . . , m}, i ∈ {1, 2, 3} and {η1 , . . . , ηm } is the dual frame in E ∗ . When
the sections {J1 , J2 , J3 } are global, i.e., the bundle Q is trivial we call (M, Q, E) a
hyper f -structure of corank m. Such structure is a G-structure with G = Sp(n)×1lm .
When m = 0 these definitions reduce to the usual notions of almost quaternionic
and hypercomplex geometry in the two previous examples. When m = 3 a hyper f -
structure is called an almost hypercontact structure introduced by Kuo [Kuo70] as
an almost contact (metric) 3-structures. The geometry of quaternionic and hyper
f -structures of corank 3 was studied by [Her96] and we will come back to some of
these examples in Chapter 13.

1.5. Pseudogroup Structures


We have seen that the notion of G-structure is a unifying concept for study-
ing geometric structures. Another unifying concept is the idea of a pseudogroup
of transformations. Roughly speaking a pseudogroup is a ‘group of transforma-
tions’, where there may be some problems with domains of definition. The concept
goes back to Lie and Cartan, but the precise notion was defined more recently by
Ehresmann [Ehr53]. The ideas were then developed much further by Kodaira and
Spencer among others, cf. [GS66, Kod60b, KS58, Spe62]. Here is the definition
following [Kob61]
Definition 1.5.1: A pseudogroup of transformations on Rn is a set Γ of local
diffeomorphisms which satisfy
(i) Each f ∈ Γ is a diffeomorphism of an open subset of Rn , called the
domain of f , onto an open subset of Rn , called the range of f.
(ii) If U = ∪i Ui , where each Ui is an open subset of Rn , then a diffeomor-
phism f with domain U belongs to Γ if and only if its restriction to each
Ui belongs to Γ.
(iii) For every open subset U of Rn , the identity transformation on U belongs
to Γ.
(iv) If f ∈ Γ then so is f −1 .
(v) If f : U −−→V and f 0 : U 0 −−→V 0 are elements of Γ and V ∩ U 0 6= ∅, then
the diffeomorphism f 0 ◦ f : f −1 (V ∩ U 0 )−−→f 0 (V ∩ U 0 ) belongs to Γ.
1.5. PSEUDOGROUP STRUCTURES 37

Γ is said to be transitive if for every pair of points p, q ∈ Rn there is an f ∈ Γ


such that f (p) = q.
More generally, one can replace Rn by a different ‘model space’ [Kob61], for
example a smooth manifold M. The importance of pseudogroups for our purpose
lies in its ability to describe different types of atlases of coordinate charts.
Definition 1.5.2: Let Γ be a transitive pseudogroup, and M a second countable,
Hausdorff topological space. A Γ-atlas for M is a family of pairs {(Uα , φα )}α∈I ,
called charts, such that
(i) The collection {Uα }α∈I of open sets covers M, i.e., each Uα is an open
subset of M, and ∪α Uα = M.
(ii) Each φα is a homeomorphism of Uα onto an open subset φα (Uα ) ⊂ Rn
such that whenever Uα ∩ Uβ 6= ∅, the map
φβ ◦ φ−1
α : φα (Uα ∩ Uβ ) : −
−→φβ (Uα ∩ Uβ )
belongs to Γ.
A Γ-atlas on M is said to be maximal if it is not properly contained in any
other Γ-atlas on M.
Clearly, if Γ is the pseudogroup of all local diffeomorphisms of Rn , this definition
coincides with the usual definition of an atlas of coordinate charts for a differentiable
manifold. As in the usual case, it is easily seen that every Γ-atlas is contained in
a unique maximal Γ-atlas. Thus, a given Γ-atlas defines uniquely a Γ-structure on
M. The relation between transitive pseudogroups and G-structure should now be
evident from Proposition 1.4.3.
Proposition 1.5.3: Let G be a Lie subgroup of GL(n, R), and let ΓG be the pseu-
dogroup of local diffeomorphisms of Rn whose Jacobian matrix lies in G at each
point of its domain. Then the ΓG -structures on M are in 1 − 1 correspondence with
the integrable G-structures on M.
It should be clear that the pseudogroup ΓG is transitive since it contains the
translations. Notice that not every Γ-structure is a ΓG -structure for some G ⊂
GL(n, R). We will see an important example of this later. Of course, as with
integrable G-structures, or just G-structures, for a given Γ there may not exist any
Γ-structure at all on M. That is, generally, there are topological obstructions to
the existence of a Γ-structure. Two important ΓG -structures are for G = GL(n, C)
and G = Sp( n2 , R). It is clear from the two examples 1.4.9 and 1.4.12 that the
ΓGL(n,C) -structures are just the complex structures, while the ΓSp( n2 ,R) -structures
are the symplectic structures. The example of the six sphere S 6 is interesting in
this context. As mentioned in the paragraph after Example 1.4.9 S 6 is known to
admit an almost complex structure, i.e., a GL(3, C)-structure, but it is not known
whether it has a complex structure, i.e., a ΓGL(3,C) -structure. From the symplectic
point of view, S 6 also admits an almost symplectic structure, i.e., an Sp(3, R)-
structure, since it admits an almost complex structure; however, it cannot admit
a symplectic structure, i.e., a ΓSp(3,R) -structure, since the closed 2-form ω would
provide a non-vanishing element of H 2 (S 6 , R) by Stokes Theorem.
Following [SS65] we relate pseudogroups to collections of vector fields. Let
{Xα } be a collection of smooth vector fields each defined on an open set Uα ⊂ M.
Each Xα generates a local1-parameter group φα (t) of local diffeomorphisms defined
for small t and for some open subset of Uα . The collection {φα } defines a family
38 1. STRUCTURES ON MANIFOLDS

of local diffeomorphisms and thus generates a pseudogroup P on M. Fix a point


x ∈ M and consider the subpseudogroup P0 of P generated by those vector fields
in the collection that vanish at x. Then any local 1-parameter group φα (t) ∈ P0
satisfies φα (t)(x) = x for all t in its domain. The differential φα (t)∗ of any such
local 1-parameter group is a linear map on the vector space Tx M. Running through
all such local 1-parameter groups in P0 , determines a Lie subalgebra g0 of gl(n, R).
We call g0 the linear isotropy algebra of the pseudogroup P. This definition differs a
bit from the one in [SS65], but they should be equivalent. The connected subgroup
G0 of GL(n, R) determined by the Lie algebra g0 is called the linear isotropy group
of P.
Let Γ be a pseudogroup of local transformations4 on a smooth manifold M. We
can consider germs of elements f ∈ Γ, that is two maps f, g : U −−→V in Γ sending
x ∈ M to y ∈ M are equivalent if there is an open set W ⊂ U such that f |W = g|W .
The equivalence class fx is the germ of the map f at x. Note that fx can only be
evaluated at x in which case fx (x) = y. Thus, we have a source map s(fx ) = x,
a target map t(fx ) = fx (x) = y, and a composition defined as follows: if fx is a
germ of a map from x ∈ M to y ∈ M, and gy is a germ of a map from y to z, then
the composition gy ◦ fx is well defined and is the germ at x of the map g ◦ f from
x to z. One easily sees that the set G(Γ) of germs of maps of the pseudogroup Γ
forms a groupoid (cf. Appendix A), called a groupoid of germs of the pseudogroup
Γ. Furthermore, we can give G(Γ) a topology that makes the source map s a local
homeomorphism. So s : G(Γ)−−→M becomes an étale space over M. Thus, G(Γ)
is an étale groupoid constructed from germs, or in the terminology of Haefliger
[Hae58], a faisceau de groupoids (sheaf of groupoids). Conversely given a groupoid
of germs G on a manifold M, we can reconstruct a pseudogroup as follows: let
σ : U −−→G be a local section of the source map of G. Composing σ with the target
projection t : G−−→M gives a local diffeomorphism of M, and it is easy to check that
the set of all such diffeomorphisms forms a pseudogroup of local transformations of
M. Summarizing we have
Proposition 1.5.4: Let M be a smooth manifold and Γ a pseudogroup of local
transformations on M. Then the set G(Γ) can be given the structure of a étale
groupoid. Conversely given a groupoid of germs G over M, the set
{t ◦ σ | t is the target projection, σ : U −−→G is a local section of the source map s}
is a pseudogroup of local transformations on M.
This proposition says that the notion of pseudogroups of transformation on M
and that of groupoids of germs on M are equivalent.
Exercise 1.12: Check the details of Proposition 1.5.4. That is, given a pseu-
dogroup of transformations on M, show that passing to germs one obtains an étale
groupoid, and conversely, given a groupoid of germs on M, the set defined above
forms a pseudogroup.

1.6. Group Actions on Manifolds


In this section we will review some basic properties of group actions on mani-
folds. This is a classical subject treated in several texts [MZ55, Bre72, Kob72,
4By a pseudogroup of local transformations on M we mean any subpseudogroup of the
pseudogroup of local diffeomorphisms of M. More generally one could take M to be a topological
space and Γ a subpseudogroup of the pseudogroup of local homeomorphisms of M .
1.6. GROUP ACTIONS ON MANIFOLDS 39

DK00] to name a few. We mainly follow the approach and the notation of Duis-
termaat and Kolk and refer the reader to their modern textbook for more details
[DK00]. Recall that a topological group is a group G endowed with a topology
such that the multiplication map µ : M × M −→M , µ(g1 , g2 ) = g1 g2 and the inverse
map ι : M −→M , ι(g) = g −1 are continuous. A discrete group is a topological group
with the discrete topology. For a topological manifold M we denote by Hom(M )
the group of homeomorphisms of M. If M is a smooth manifold the group of diffeo-
morphisms is denoted by Diff(M ). We give Diff(M ) the compact-open C ∞ topology.
This is a good topology when M is compact, but in the non-compact case it does
not control the behavior at infinity. Since we deal mainly with compact manifolds
we do not concern ourselves with this failure. We refer the reader to [Ban97] for
further discussion of diffeomorphism groups.
Definition 1.6.1: An action A of a topological group G on a topological manifold
M is a homomorphism A : G −→ Hom(M ) of G into the group of homeomorphisms
of M . We say that the action A is continuous if the mapping
(1.6.1) (g, x) 7→ g · x = A(g)(x) : G × M −−→ M,
also denoted by A : G × M −→ M, is continuous. We say that A is a proper
action if the associated map G × M −→ M × M given by
(1.6.2) (g, x) 7→ (x, g · x)
is proper, i.e, the inverse image of any compact set is compact. If M is a smooth
manifold and G is only a topological group we say that A is smooth if A(g) ∈
Diff(M ) for each g ∈ G. However, if G is a Lie group a smooth action A means
that the map (1.6.1) is smooth.
Let A be an action of G on M and let e ∈ G be the identity element. For any
x ∈ M the subgroup Gx = {g ∈ G | g · x = x} is called the isotropy subgroup of the
action at the point x ∈ M , or the stabilizer of x ∈ M under the action. Then A is
called effective if Ker(A) = {1l}. The action A is free if Gx = {e} for all x ∈ M
and locally free if Gx is a finite group for all x ∈ M .
For each x ∈ M the orbit through x is defined as
(1.6.3) G · x = {y ∈ M |y = g · x, g ∈ G} ⊂ M.
The orbit G · x through x is an immersed submanifold of M, and there is a natural
identification of G · x with the coset space G/Gx .
It is clear that the relation x ∼ y if y ∈ G · x is an equivalence relation
partitioning M into orbits. We shall denote the quotient space M/ ∼ by M/G,
and call it the space of orbits. The map π : M −→M/G sending x 7→ G · x is the
canonical projection. We equip M/G with the quotient topology, i.e., V is open in
M/G if and only if π −1 (V ) is open in M .
Exercise 1.13: Let M be a manifold with group action A : G × M −→ M . Show
that M/G is Hausdorff if and only if the set {(x, y) ∈ M × M | y ∈ G · x} is a
closed subset of M × M .
In this book we will mainly be concerned with actions which yield Hausdorff
quotients. The importance of proper group actions is realized through the following
Proposition 1.6.2: If the action A of G on M is proper then the quotient M/G
is Hausdorff.
40 1. STRUCTURES ON MANIFOLDS

Example 1.6.3: A typical example of a non-proper group action is the so-called


irrational flow on a torus. Let M = T 2 = SO(2)×SO(2) which is an Abelian group.
So for any α ∈ R we can consider a subgroup G(α) = {(A(t), A(αt), t ∈ R}, where
µ ¶
cos t − sin t
A(t) = .
− sin t cos t
Then G(α) acts on M by left (or right) multiplication. For α ∈ Q the orbits are
all closed in M . In this case G(α) is homeomorphic to S 1 and so is the quotient
M/G(α). However, for α ∈ R \ Q the orbits are not closed. They are immersed but
not embedded submanifolds homeomorphic to R.
The closure of any orbit equals to M and the quotient M/G(α) is non-Hausdorff.
There are also examples with non-Hausdorff quotients with all orbits closed.
Consider, for instance, G = R acting on R2 via
¡ ¢
t, (x1 , x2 ) 7→ (x1 + tx2 , x2 ), t ∈ R, (x1 , x2 ) ∈ R2 .
Here the orbits are indeed closed: they are either lines parallel to the x1 -axis, or
points on the x1 -axis. It is easy to see that the quotient is non-Hausdorff.
Definition 1.6.4: A G-manifold is a smooth manifold M together with a smooth
action A of a group G.
The importance of proper and free actions is immediately realized by the fol-
lowing classical result.
Theorem 1.6.5: Let G be a Lie group acting smoothly, freely, and properly on
M . Then the orbit space M/G is a smooth manifold of dimension dim(M ) −
dim(G) with a unique differentiable structure having the following property: if
π : M −→ M/G is the canonical projection then for every a ∈ M/G there exists an
open neighborhood U 3 a and a diffeomorphism
τ = (χ, s) : π −1 (U) −→ G × U
such that, for each x ∈ π −1 (U), g ∈ G we have
(1.6.4) π(x) = s(x), τ (g · x) = (gχ(x), s(x)).
In fact, it follows that the map π is a smooth submersion and M = P (M/G, G)
is a principal G-bundle over M/G, and the map τ is a local trivialization. If we add
a bit more structure the hypothesis on the action of Theorem 1.6.5 can be weakened
to be only locally free. This is of great importance to us in this book, and will be
discussed further in the forthcoming chapters. See, for example, Theorem 2.5.11
below.
A case of particular interest to us is when a group G leaves invariant some
geometric structure. Following [Kob72] we have
Definition 1.6.6: Let P be a G-structure over M. A diffeomorphism f : M −−→M
is an automorphism of the G-structure if the induced map f∗ : P −−→P is an
isomorphism.
The automorphisms of P form a subgroup of Diff(M ) and is denoted by Aut(P ).
In particular, if the G-structure P is defined by a tensor field T then f : M −−→M
is an automorphism of P if and only if it leaves T invariant. There are many
such examples of interest to us throughout this book. We shall often denote the
automorphism group by Aut(T ), but in some cases there is more common standard
terminology. For example, if (M, g) is a Riemannian manifold the automorphism
1.6. GROUP ACTIONS ON MANIFOLDS 41

group Aut(g) is called the isometry group and denoted by Isom(M, g). We shall
come across other examples as we proceed. Since Aut(P ) a subgroup of Diff(M ) it
is endowed with the compact-open topology, and we are interested in when it is a
Lie group. Recall [Kob72] that a Lie subalgebra g ⊂ gl(n, R) is said to be elliptic
if g contains no matrix of rank one. Then we have
Proposition 1.6.7: Let P be a G-structure on a compact manifold M with group
G and Lie algebra g. If g is elliptic then Aut(P ) is a Lie group with respect to the
compact-open C ∞ topology.
Examples of elliptic Lie algebras are the orthogonal Lie algebras o(n, R), the
complex Lie algebra gl(n, C) ⊂ gl(2n, R), and any of their subalgebras. Thus,
the automorphism groups of a compact Riemannian manifold, or of a compact
almost complex manifold are Lie groups. In general compactness is a necessary
hypothesis for Aut(P ) of an elliptic G-structure to be a Lie group. For example the
complex automorphism group Aut(Cn ) is not a Lie group. However, for Riemannian
and certain other G-structures compactness is not necessary. In order to develop
more we need the notion of prolongation. Since this is amply treated in several
books [Ste83, Kob72] we give only a very brief treatment here. Let V be an
n-dimensional vector space and g a Lie subalgebra of gl(n, R) = End(V ).
Definition 1.6.8: The first prolongation g(1) of g is defined by
g(1) = {T ∈ Hom(V, g) | T (u)v = T (v)u for all u, v ∈ V }.
(1)
The k th prolongation g(k) is defined inductively by g(k+1) = g(k) . A Lie algebra
g is said to be of finite type of order k if for some k, g(k) = 0, but g(k−1) 6= 0
otherwise g has infinite type.
Similarly one can define the prolongations of a Lie group G ⊂ GL(n, R). Such
prolongations give rise to higher order G-structures for which we have no real need.
We say the a G-structure is of finite type of order k if its Lie algebra g is of finite
type of order k. It is easy to see that if h ⊂ g and g(k) = 0 then h(k) = 0. It
is equally easy to show that o(n, R)(1) = 0, and that gl(n, R) and gl(n, C) are of
infinite type. However, gl(n, H) is of finite type, in fact gl(n, H)(1) = 0. Many of the
G-structures that are of importance in this book are of infinite type, in particular
those of Examples 1.4.9, 1.4.12 1.4.13 1.4.15, and 1.4.16. G-structures that are of
finite type, but with g(1) 6= 0 are also of interest. This occurs in two important
cases for us, the conformal G-structures with G = CO(n) and the quaternionic G-
structures with G = GL(n, H)Sp(1). These both have g(1) 6= 0, but g(2) = 0. When
a G-structure is 1-integrable, the vector space g(1) parameterizes the torsion-free
connections, and if g(1) = 0 the torsion-free connection is unique. We now have
Theorem 1.6.9: Let P (M, G) be a G-structure of finite type of order k, Then
Aut(P ) is a Lie group of dimension at most dim M + dim g + dim g(1) + · · · +
dim g(k−1) .
The proof of this general theorem is given in [Ste83, Kob72]. It, of course,
has many precursors the most famous of which is the somewhat stronger result
essentially due to Myers and Steenrod Theorem [MS39]:
Theorem 1.6.10: Let (M, g) be a Riemannian manifold with finitely many con-
nected components. Then the group of isometries Isom(M, g) is a finite-dimensional
Lie group. The action of Isom(M, g) on M is proper and its Lie group topology
coincides with the compact-open C ∞ topology as a subgroup of Diff(M ).
42 1. STRUCTURES ON MANIFOLDS

In particular, this theorem says that the isometry group Isom(M, g) of a com-
pact manifold is compact. We discuss many other examples of G-structures of
finite type throughout the book. The G-structures dealt with in this book for
which Aut(P ) is not a Lie group include the symplectic and contact structures.
Definition 1.6.11: Let G be a group acting on a manifold M . We say that the
action is transitive if for any x, y ∈ M there exists g ∈ G such that y = g · x. In
such a case we say that M is G-homogeneous. If Aut(P ) is the automorphism
group of a G-structure P (M, G) we say that M is a homogeneous G-structure
if Aut(P ) acts transitively on M. If G is a Lie group which acts transitively on M
we say that M is a homogeneous manifold.
When a manifold admits a Lie group G acting transitively and is a subgroup
of the Aut(P ) for some G-structure P (M, G), we say that M is a homogeneous (G-
structure) manifold, for example a homogeneous complex or homogeneous contact
manifold, etc.
Example 1.6.12: Let G be a Lie group and H any subgroup. The H acts on G
from the left and from the right:

(1.6.5) AL (h)g = hg,

(1.6.6) AR (h)g = gh−1 .

The corresponding quotient spaces are denoted by H\G and G/H. Both of these
actions are free but they may or may not be proper. If H is a closed subgroup of
G then the H-action is proper and the quotient G/H is a homogeneous smooth
G-manifold under the natural left G-action.
More generally, we will be interested in proper actions which are not necessarily
free. In such case the space of orbits is not always a manifold; nevertheless, it has
a very tractable structure, namely that of an orbifold which we discuss in Chapter
4.
Consider the differential (Ax )∗ : g −→ Tx M , where g ' Te G is identified with
the tangent space of G at the identity g = e.
Definition 1.6.13: Let A : G × M −→ M be a smooth action of a Lie group on a
manifold M . A smooth slice at x0 ∈ M for the action A is a smooth submanifold
S ⊂ M , x0 ∈ S such that
(i) Tx0 M = (Ax0 )∗ (g) ⊕ Tx0 S and Tx0 M = (Ax0 )∗ (g) + Tx0 S for all x ∈ S,
(ii) S is Gx0 -invariant,
(iii) if x ∈ S, g ∈ G, and A(g)(x) ∈ S, then g ∈ Gx0 .
It follows that the inclusion S ,→ M induces a bijective map Gx0 · x 7→ G · x
from the space S/Gx0 of Gx0 orbits in S onto an open neighborhood of G · x0 in
M/G.
Definition 1.6.14: The action A : G × M −→ M is said to be proper at x0 if
for any convergent sequence xi −→x0 in M and a sequence {gi }i∈N in G such that
gi · xi −→x0 , there exists a convergent subsequence of {gi }i∈N .
Naturally, if A : G × M −→ M is proper then it is proper at every point x ∈ M .
But the converse is in general not true. However, if A is proper at every point of
M and M/G is Hausdorff then A is proper.
1.6. GROUP ACTIONS ON MANIFOLDS 43

Lemma 1.6.15: The action A : G × M −→ M is proper at x0 ∈ M if and only


if there exists a neighborhood U 3 x0 such that {g ∈ G | A(g)(U) ∩ U 6= ∅} has a
compact closure in G. In particular Gx0 is a compact subgroup of G.
The following important theorem was proved by Palais [Pal61].
Theorem 1.6.16: Let A : G × M −→ M be a smooth action of a Lie group G on
a manifold M . Suppose that the action is proper at x0 ∈ M . Then there exists a
smooth slice S at x0 for the action A.
If M is a smooth manifold and G acts properly, the quotient is both Hausdorff
and paracompact. One can define a function on an open subset of M/G to be
smooth if and only if π ∗ f = f ◦ π is a smooth function on M . Then a standard
partition of unity and averaging argument can be used to construct G-invariant
Riemannian metric on M .
Theorem 1.6.17: Let A : G × M −→ M be a proper action of G on M . Then M
has a G-invariant Riemannian metric g.
Theorems 1.6.10 and 1.6.17 show that in the case of smooth manifolds with
a finite number of connected components the category of smooth proper effective
Lie group actions coincides with the category of closed subgroups of isometries for
Riemannian structures on M .
Definition 1.6.18: A smooth stratification of a manifold M is a partition Σ of
M into locally closed sets such that
(i) Any M ∈ Σ is a smooth, connected, finite-dimensional manifold without
boundary.
(ii) Σ is locally finite.
(iii) If M1 , M2 ∈ Σ and M1 ∩ M 2 6= ∅ then M1 ⊆ M 2 .
(iv) If M1 , M2 ∈ Σ, M1 6= M2 and M1 ⊆ M 2 then dim(M1 ) < dim(M2 ).
The third condition is called the frontier condition. Furthermore, Σ is said to be
Whitney regular if the following two conditions are met.
(a) For each M1 , M2 ∈ Σ, M2 ⊆ M 1 , M1 6= M2 and each sequence {xi }i∈N
such that
lim xi = x ∈ M2 , lim Txi M1 = L
i→∞ i→∞

we have Tx M1 ⊂ L.
(b) If {xi }i∈N is a sequence as in (a) and {yi }i∈N is a sequence in the limit
stratum M2 such that yi −→x and yi 6= xi for all i ∈ N then each limit of
one-dimensional subspaces R · λ(xi , yi ) of Txi M , for i−→∞ is contained
in L. Here λ is any diffeomorphism from an open neighborhood of the
diagonal in M × M to an open neighborhood of the zero section of the
tangent bundle T M and the set of limit lines does not depend on the
choice of λ.
We will now consider the a stratification induced by a proper effective action
A : G × M −→ M . As before we will follow closely the notation of [DK00].
Definition 1.6.19: We say that points x, y ∈ M and the orbits G · x, G · y ∈ M/G
are of the same type, with notation x ∼ y, G · x ∼ G · y respectively, if there
exists a G-equivariant bijection from G · x to G · y. We say that x dominates y and
G · x dominates G · y (y . x, G · y . G · x) if there exists a G-equivariant mapping
from G · x to G · y. The equivalence classes under ∼ are called orbit types in M
44 1. STRUCTURES ON MANIFOLDS

and M/G respectively and will be denoted by


Mx∼ = {y ∈ M | x ∼ y}, ∼
MG·x /G = {G · y ∈ M/G | G · y ∼ G · x}.
The relation . is a pre-order in M and we define
.
Mx. = {y ∈ M | x . y}, MG·x /G = {G · y ∈ M/G | G · y . G · x}.
The key to this definition is the following lemma describing the properties of
the above relations.
Lemma 1.6.20: We have the following
(i) x ∼ y ⇔ Gy = g −1 Gx g for some g ∈ G, i.e., Gx and Gy are conjugated
in G.
(ii) x . y ⇔ g −1 Gx g ⊂ Gy for some g ∈ G, i.e., Gx is conjugated in G to
a subgroup of Gy .
(iii) For any G-equivariant mapping Φ : G · x −→ G · y, Gx is a subgroup of
GΦ(x) and Φ is a smooth fibration with fiber GΦ(x) /Gx .
(iv) x . y and y . x ⇔ x ∼ y.
Definition 1.6.21: Let H be any subgroup of G. We define the set of fixed points
for H as
(1.6.7) M H = {y ∈ M | h · y = y, for all h ∈ H}.
We denote by (H) the conjugacy class of groups conjugated to H in G and by M(H)
the set of all points in M with stabilizers conjugated to H.
One can show that each connected component of M H is a closed submanifold
of M but different connected components can have different dimensions. Naturally,
the orbit types give a stratification of M and this stratification is typically called
stratification by orbit types. However, an orbit type can be a union of connected
components of different dimensions. That is why we consider a finer stratification
of each Mx∼ = M(Gx ) by the so-called local action type.
Definition 1.6.22: The points x, y ∈ M and the orbits G · x, G · y, respectively,
are said to be of the same local type (with the notation x ≈ y and G · x ≈ G · y) if
there exists a G-equivariant diffeomorphism from an open G-invariant neighborhood
U 3 x onto an open G-invariant neighborhood V 3 y in M .

The relation partitions further partitions Mx ∼ and MG·x /G into subclasses
called local action types and denoted by
(1.6.8) Mx≈ = {y ∈ M | y ≈ x}, ≈
MG·x /G = {G · y ∈ M/G | G · y ≈ G · y}.
The following theorem describes fundamental properties of the stratification by
local action types.
Theorem 1.6.23: Let A : G × M −→ M be a smooth and proper action.
(i) Each local action type is an open and closed subset of the corresponding
orbit type.
(ii) The set Mx≈ ∩ M Gx is open in M Gx and a locally closed smooth submani-
fold of M (that is all its connected components have the same dimension).
In addition it is N (Gx )-invariant, when N (Gx ) is the normalizer of Gx
in G.
(iii) The canonical projection π : M −→ M/G maps Mx≈ ∩M Gx onto MG·x ≈
/G,
its fibers in Mx ∩M are the orbits for the N (Gx )-action on Mx ∩M Gx .
≈ Gx ≈
1.6. GROUP ACTIONS ON MANIFOLDS 45


The action of N (Gx )/Gx is proper and free. This gives MG·x /G a unique
≈ Gx ≈
structure of a smooth manifold for which π : Mx ∩ M −→ MG·x /G is
the principal fiber bundle with structure group N (Gx )/Gx .
(iv) Mx≈ is a locally closed smooth submanifold in M . Furthermore, the pro-
jection π : Mx≈ −→ MG·x≈
is a smooth fibre bundle with fibre G/Gx .
(v) We have dim(Mx ) = dim(G) − dim(Gx ) + dim(Mx≈ ∩ M Gx ), and, hence,

dim(Mx≈ /G) = dim(N (Gx )) − dim(Gx ) + dim(Mx≈ ∩ M Gx ).


The local finiteness of the stratification follows now from the fact that in the
case of proper actions, locally there can only be finitely many orbit types. More
precisely
Proposition 1.6.24: For a proper smooth Lie group action G on M there are
locally only finitely many distinct orbit types with locally only finitely many con-
nected components. If M is compact there are only finitely many distinct connected
components of orbit types and the same is true if M is a finite-dimensional vector
space and G acts linearly.
Theorem 1.6.25: The connected components of the orbit types in M from a Whit-
ney stratification of M .
A part of the information about the orbit type stratification for the action
A : G × M −→ M can be encoded into the directed graph Γ = Γ(A, G, M ). The
vertices of Γ are the connected components of the orbit types in M/G. If Γ1 , Γ2 ∈ Γ
then we draw an arrow from Γ1 −→Γ2 if Γ1 ⊂ Γ2 . It follows that if Γ1 −→Γ2 then
Γ2 . Γ1 and dim(Γ1 ) > dim(Γ2 ) as long as Γ1 and Γ2 are distinct. Γ1 −→Γ2 defines
a partial ordering in Γ in which any chain has at most 1 + dim(M/G) elements.
Example 1.6.26: [Complex weighted projective plane] Let M = S 5 ⊂ C3 , be
the unit sphere in C3 , G = S 1 and consider the action Ap : S 1 × S 5 −→ S 5 given
by

Ap (λ, z) = (λp1 z1 , λp2 z2 , λp3 z3 ),

where z = (z1 , z2 , z3 ) are the linear complex coordinates on C3 , λ ∈ S 1 is a unit com-


plex number, and p = (p1 , p2 , p3 ) are positive integers such that gcd(p1 , p2 , p3 ) = 1,
dij = gcd(pi , pj ) > 1 for all i 6= j. Then the directed graph Γ has 7 vertices which
can be described as follows: Let β be any non-empty subset of {1, 2, 3} and denote
Γβ the S 1 -invariant subset in S 5 /S 1 corresponding to the non-vanishing of all com-
plex coordinates zj on S 5 ⊂ C3 with j ∈ β. With this notation Γ can be described
as follows.
46 1. STRUCTURES ON MANIFOLDS

Γ o ΓD 123
E 1Y44 ª
®®
® 44 ª 333
® 4 ªªª 33
®® 44
® 4 ªªª 33
®® 44 ª 33
® 44 ªª 33
® 4 ªª ª 3¼
®®
Γ313 o Γ1235 / Γ2
E
33
33 ­­­ 555 ®®®
­ 555 ®
33 ­­ ®®
33 ­ 55 ®
33 ­­ 55
®®
®
33 ­­­ 55
5½ ®® ®
¼ ¦­
Γ3 o Γ23

In the case any dij = gcd(pi , pj ) = 1 or/and some p0j s equal to 1, the directed
graph is obtained by deleting the corresponding vertices and the arrows that orig-
inate there and appropriately re-defining the set Γ123 . In the case p = (1, 1, 1) we
get just one vertex as the S 1 -action is free and there is only one orbit type.
Definition 1.6.27: Let A : G×M −→ M be a smooth proper action of a Lie group
G on M . Then the orbit G · x, x ∈ M is said to be a principal orbit if its local
action type Mx≈ is open in M , i.e., Mx≈ belongs to the maximal element of Γ. We
write M princ = {x ∈ M | G · x is a principal orbit}, and M princ /G for the set of
principal obits.
Definition 1.6.28: The orbit G · x is said to be regular if the dimension of the
orbits G · y is constant for all y sufficiently close to x. Points lying on regular
orbits are said to be regular points and their collection forms M reg . The points in
M reg \ M princ are called exceptional and any orbit through an exceptional point is
called an exceptional orbit. Finally, the points in M \ M reg are called singular
and any orbit through a singular point is called a singular orbit.
It follows that in every connected component of G/M a nearby non-principal
orbit is an exceptional orbit if and only if it has the same dimension as the principal
orbit. A nearby non-principal orbit is singular if and only if its dimension is strictly
smaller than the dimension of the a principal orbit. Note that in the Example
1.6.26 all orbits are regular so that M reg = M . Hence, non-principal orbits are all
exceptional.
Clearly, M princ ⊂ M reg ⊂ M and both are dense open subsets in M . Both of
these sets are also G-invariant so that the inclusions project to open dense subsets
M princ /G ⊂ M reg /G ⊂ M/G. The following theorem is typically called principal
orbit theorem.
Theorem 1.6.29: Let A : G × M −→ M be a smooth proper action of a Lie group
G on M . Them M \ M reg is equal to the union of local orbit types of codimension
≥ 2 in M . For every connected component M ◦ of M , the subset M reg ∩ M ◦ is
connected, open and dense in M ◦ . Each connected component of G/M contains
only one principal orbit type which is a connected, open, and dense subset of it. In
particular, if M is connected there is only one principal orbit type.
Without loss of generality, by restricting to subsets of M of the from G · M ◦ ,
one can assume that G/M is actually connected. In particular, this is always the
1.6. GROUP ACTIONS ON MANIFOLDS 47

case when M itself is connected. In this book such will most often be the case.
When both M/G and M are connected the general stratification picture simplifies
a lot. We have
Corollary 1.6.30: Let A : G × M −→ M be a smooth proper action of a Lie group
G on M such that M/G is connected. Let x ∈ M princ and let C be the connected
component of x in M princ ∩ M Gx . Then
(i) M princ = Mx≈ = Mx∼ .
(ii) The union of connected components of M Gx that meet M princ is a closed
smooth submanifold of M which contains N (Gx ) · C = M princ ∩ M Gx .
(iii) G(C) = {g ∈ G | A(g)(C) = C} induces a G-equivariant diffeomorphism
G/Gx ×G(C) /Gx C ' M princ .
We next consider some special cases in some more detail. One important class
of actions is when G is a compact Lie group acting locally freely on a compact
manifold M , in which case the quotient M/G can be identified with a compact
smooth orbifold. These will be discussed in chapter 4. In fact, most, though not
all compact orbifolds arise this way.
Definition 1.6.31: A smooth, proper, locally free action A : S 1 ×M −→ M is called
a Seifert fibered S 1 -structure and M is called Seifert fibered S 1 -manifold.
The actions Ap of the Example 1.6.26 give infinitely many distinct Seifert fibred
S 1 -structures on S 5 . In Chapter 4 we will give a more general definition of a Seifert
fibration of a manifold M in the language of orbifolds and orbibundles.
Definition 1.6.32: Let (M, g) be a Riemannian manifold and let G = Isom(M, g)
be its isometry group. We say that (M, g) is homogeneous if G acts transitively
on M . We say that (M, g) is of cohomogeneity 0 < k ≤ dim(M ) when the
principal orbit of the G-action has dimension dim(M ) − k.
The most tractable cases amenable to further analysis are the homogeneous, co-
homogeneity one, and cohomogeneity two, cases. Homogeneous Riemannian mani-
folds will be of special interest to us throughout this book, but they are trivial from
the point of view of the stratification discussed in this section. The next case is that
of cohomogeneity one. In this case the stratification is simple as there are not too
many candidates for M/G. We have the following theorem of Mostert [Mos57]:
Theorem 1.6.33: Let G be a compact connected Lie group acting on a connected
manifold M of dimension n such that there is at least one orbit of dimension n − 1.
Then the quotient space M/G must be one of the following
(i) M/G = R,
(ii) M/G = [−1, ∞),
(iii) M/G = S 1 ,
(iv) M/G = [−1, +1].
Furthermore, the non-principal orbits correspond to the boundary points.
The general description can be found in Section 8 of Chapter IV of [Bre72], in
particular, Theorem IV.8.2 and its proof where it is assumed that M is compact.
Compactness eliminates the first two cases. In the third case M ' (G/H) ×Z R is a
bundle over S 1 and has infinite fundamental group. We follow the presentation in
[GZ02]. However, we focus our attention on the last case where there are exactly
two non-principal orbits.
48 1. STRUCTURES ON MANIFOLDS

Consider any G-invariant metric g and a geodesic c : [−1, +1] → M perpendic-


ular to the orbits with endpoints x± = c(±1). The isotropy group at every point
c(t), −1 < t < 1 equals to Gc(t) = H. Let us denote by K± = Gx± the isotropy
groups at the end points of the geodesic. All orbits G/H through c(t), t 6= ±1
l +1
are principal. The two non-principal orbits are G/K± = G · x± . Let D±± be
the normal disc to the orbit G/K± at x± , where l± = dim(G) − dim(K± ) are the
dimensions of non-principal orbits. By the slice theorem M decomposes as
(1.6.9) M = G ×K− Dl+ +1 ∪G/H G ×K− Dl− +1
and we have the following group diagram
(1.6.10) = G aBB
j− ||| BB j+
|| BB
| BB
||
K− K+
aBB |=
BB |
BB ||
h− BB ||
|| h+
H
which is also written as H ⊂ {K− , K+ } ⊂ G. The non-principal orbits are denoted
by B± = G/K± and note that K± /H ' S l± must be spheres. The non-principal
orbit, say B− = G/K− , is exceptional if l− = 0. Otherwise, B− is singular. Hence,
a cohomogeneity one manifold M with two non-principal orbits determines the
diagram 1.6.10. Conversely, any such group diagram determines M via (1.6.9). We
remark that in the case there is only one non-principal orbit M = M ×K− Dl− +1 .
Remark 1.6.1: Observe that a group diagram as in (1.6.10), where we assume
that h± are embeddings, but j± are only homomorphisms with finite kernel and
j− ◦ h− = j+ ◦ h+ = j0 with K± /H = S l± , defines a cohomogeneity one orbifold
O: The regular orbits, being hypersurfaces, have no orbifold singularities, and we
can therefore assume that j0 is an embedding, although we still allow the action
of G to be ineffective otherwise. A neighborhood of a singular orbit is given by
D(B± ) = G×K± Dl± +1 , where K± acts on G via right multiplication: g·k = gj± (k)
and on Dl± +1 via the natural linear extension of the action of K± on S l± . This
then can be written as D(B± ) = G ×(K± /kerj± ) (Dl± +1 /kerj± ) and the singularity
normal to the smooth singular orbit G/j± (K± ) is S l± /kerj± .
Parker [Par86] began a classification of the smooth cohomogeneity one actions
of compact Lie groups on compact connected 4-manifolds although there are some
missing cases. In Section 12.5, we shall be interested in certain cohomogeneity one
SO(3)-actions. Here we present an example, shown to us by Wolfgang Ziller, for
the group Sp(1) that is relevant to the discussion in Section 12.5, in particular to
diagram 12.5.9.
Example 1.6.34: Let M be a compact 4-manifold, G = Sp(1) = SU (2), with quo-
tient of type (iv) above. We consider four cases where the non-principal orbits are
singular of dimension one. The first case we consider is M = S 4 . Here we represent
S 4 as the unit sphere in the five dimensional representation of SO(3) consisting
of trace-free symmetric matrices. (See Equations 12.5.7 and 12.5.8 below). This
action can be lifted to an non-effective action of the group Sp(1) of unit quaternions
which we treat here in order to facilitate the forthcoming discussion. We have the
1.6. GROUP ACTIONS ON MANIFOLDS 49

group diagram
Sp(1)
< cFF
xxx
j− FFj+
x FF
xxx FF
x
K− K ,
cFF ; +
FF www
FF ww
h− FFF ww h+
ww
Q
where Q is the 8-element quaternion group Q = {±1, ±i1 , ±i2 , ±i3 }, K− = hei1 θ , i2 i
and K− = hei2 θ , i1 i. Each group in this diagram can be factored by a central Z2
giving diagram 12.5.9. This is action 62 in [Par86].
All other possible diagrams are obtained from this one by the removal of certain
finite subgroups from K± which gives a new manifold together with a branch cover.
For example, consider
Sp(1)
: eKK
uuu
j− KK j+
u KK
uuu KK
K
uu
K−
dII hei2 θ i,
II s9
s
II ss
I
h− II sssh+
ss
Z4 = hi2 i

with K− as before. This gives M = CP2 as a well-known 2-fold branch cover


π : CP2 → S 4 . This action is missing in [Par86].
Further branching is obtained by taking
Sp(1)
sss9 eLLL
LLLj+
j−
sss LLL
ss L
ss
hei1 θ i hei1 θ i.
eKKK r9
KKK
KKK rrrrr
r
h− K rrr h+
Z2 = {±1}

This gives M = CP1 × CP1 as a 2-fold branched cover π : CP1 × CP1 → CP2
branched over a conic which is the image of the diagonal. In [Par86] this is action
50 with n = 2.
The final example corresponds to the group diagram
Sp(1)
w; cHH
j− ww w HH j+
w HH
ww HH
w w H
i1 θ
he i hei1 θ i.
dHH u:
HH u
HH uu
h− HHH u uuh+
H uu u
1
50 1. STRUCTURES ON MANIFOLDS

which corresponds to M = CP2 #CP2 : a blow-up of CP2 at one point. This is


action 45 in [Par86].
Exercise 1.14: Fill in the details for Example 1.6.34.
We finish this section with a short discussion of the cohomogeneity 2 case. In
general, when M/G is 2-dimensional the situation can be quite complicated even
when it is connected. Nevertheless, under some mild assumptions, the structure is
actually quite manageable in which case M/G is a disc with boundary, the boundary
corresponding to non-principal orbits. The following theorem, due to Montgomery,
Samelson, and Yang [MSY56], is Theorem 8.6 in Chapter IV of [Bre72] to which
we refer for the proof.
Theorem 1.6.35: Let M be a compact connected manifold with a cohomogeneity 2
action of a compact connected Lie group G. Suppose H1 (M, Z) = 0 and that there
exist a singular orbit in M/G. Then there are no exceptional orbits and M/G is a
2-disc with the boundary corresponding to the non-principal orbits.
By definition, any circle action on a 3-dimensional manifold is of cohomogene-
ity 2. The following example shows that the existence of a singular orbit in the
hypothesis of Theorem 1.6.35 is indeed necessary.
Example 1.6.36: [Football orbifold] Let M = S 3 ⊂ C2 , be the unit sphere in
C2 , G = S 1 and consider the action Ap : S 1 × S 3 −→ S 3 given by
Ap (λ, z) = (λp1 z1 , λp2 z2 ),
where z = (z1 , z2 ) are the linear complex coordinates on C2 , λ ∈ S 1 is a unit
complex number, and p = (p1 , p2 ) are positive integers such that gcd(p1 , p2 ) = 1.
There are three orbit types:
(a) principal orbits through (z1 , z2 ) ∈ S 2 when neither complex coordinate
vanishes and with G(z1 ,z2 ) = {e},
(b) two exceptional orbits: through (z1 , 0) ∈ S 3 with G(z1 ,0) = Zp2 , and
through (0, z2 ) ∈ S 3 with G(0,z2 ) = Zp1 .
The set M princ /S 1 can be identified with a punctured disc (cylinder) with the
boundary as two circles corresponding to the exceptional orbits. The quotient
space is the orbifold CP1 (p1 , p2 ) which is topologically a sphere with two isolated
orbifold points.
In Chapter 13 we will consider many examples of 3-Sasakian manifolds with
cohomogeneity 2 isometric actions for which Theorem 1.6.35 applies.
CHAPTER 2

Foliations

Foliations were defined in Example 1.4.13. It is the purpose of this section to


gather together those well-known facts concerning foliations that we shall need in
the sequel. The theory of foliations is very well developed, there being many books
on the subject [CC00, CC03, CL85, God91, HH86, Mol88, Rei83, Tam92,
Ton97]. The modern theory began with the work of Ehresmann and Reeb in the
1940’s (cf. [ER44]), but as with much in mathematics its roots go back to the 19th
century. The type of foliations that we are mainly concerned with in this book are
the so-called Riemannian foliations. This type of foliation has been developed in
the books of Reinhart [Rei83], Molino [Mol88] and Tondeur [Ton97].

2.1. Examples of Foliations

We give here several examples of foliations which should provide insight later
on. Only those examples which illustrate concepts pertinent to our development
are given.
Example 2.1.1: Submersions. As mentioned previously a submersion f : M −−→N,
where M and N are smooth manifolds, is a very special case of a foliation, called
a simple foliation. A fibre bundle is a special case of a submersion; however, not
every submersion satisfies the local triviality condition. If one removes a point from
a fibre of a fibre bundle, one still has a submersion, but local triviality fails. For
a more interesting examples of submersions that are not locally trivial see [CC00]
pages 6-7. The Ehresmann Fibration Theorem states that if f is proper then a
submersion is a fibre bundle.
The next example illustrates phenomena that will re-occur in our work.
Example 2.1.2: Linear flows on tori. For simplicity we consider the two dimen-
sional torus T 2 defined as the quotient space by the integer lattice, viz. R2 /(Z × Z).
A linear flow is the flow generated by the vector fields
∂ ∂
X = a1 + a2
∂x1 ∂x2
where ai ∈ R for i = 1, 2 are not both zero. For each such pair (a1 , a2 ) ∈ R2 −{(0, 0)}
the vector field X generates a subbundle E of T T 2 , and thus defines a 1-dimensional
foliation F on T 2 . (Frobenius integrability is automatic for 1-dimensional distribu-
tions). The leaf passing through the point (x1 (0), x2 (0)) ∈ T 2 is given by the image
of the flow φ : R−−→R2 defined by
φ(t) = (x1 (0) + a1 t, x2 (0) + a2 t).
We assume that a1 6= 0. There are two cases to consider:
51
52 2. FOLIATIONS

(i) The ratio aa21 is rational. In this case the vector field X is periodic and
the leaves of the foliation are circles. The foliation F is simple, and in
fact describes T 2 as an S 1 bundle over S 1 .
(ii) The ratio aa12 is irrational. In this case the leaves are diffeomorphic to
R, and each leaf is dense in T 2 . Hence, the topology of the quotient
space T 2 /F is not even a T1 space (referring to the separation axioms of
point-set topology).
Exercise 2.1: Prove the assertions (i) and (ii) of Example 2.1.2.
Our next example generalizes a well-known result in Lie theory.
Example 2.1.3: A locally free action of a Lie group. Let G be a Lie group
which acts locally freely on a smooth manifold M, that is there is a smooth map
G × M −−→M such that the isotropy subgroup at any point is a discrete subgroup
of G. Let g denote the Lie algebra of G. The action induces a monomorphism
ψ : g−−→Γ(T M ) of Lie algebras, and since this action is locally free the image ψ(g)
generates a subbundle E of T M of dimension (dim g). Furthermore, since ψ(g) is
subalgebra of Φ(T M ), the subbundle E is integrable and defines a foliation F on
M. In the next chapter we show that if the action of G is also proper and there
is a Riemannian metric on M with isometry group G, then the space of leaves
M/G is an orbifold such that the natural projection π : M −−→M/G is an orbifold
Riemannian submersion. This generalizes to the case of locally free action the well-
known theorem of Lie theory that says that if G acts properly and freely on M,
then the space of leaves of F, i.e., the orbit space M/G of the action is a smooth
manifold such that M is the total space of a principal G-bundle over M/G.

2.2. Haefliger Structures


(α) (α) (α) (α)
Let {(Uα , φα )} be a foliated atlas with coordinates φα = (x1 , . . . , xp ; y1 , . . . , yq ).
Then there are local submersions fα : Uα −−→Rq defined to be the composition of
φα : Uα −−→Rp+q = Rp × Rq with projection onto the second factor. The inverse
image f −1 (y) of a point y ∈ Rq defines a plaque of F, that is a component of
Uα ∩ Lα . In the overlap of foliated charts Uα ∩ Uβ the local submersions fα , fβ are
related by
(2.2.1) fβ = τβα fα ,
where the transition functions τβα : fα (Uα ∩Uβ )−−→fβ (Uα ∩Uβ ) are diffeomorphisms
that satisfy Haefliger’s cocycle conditions
−1
(2.2.2) ταβ = τβα on Uα ∩ Uβ , τγα = τγβ τβα on Uα ∩ Uβ ∩ Uγ .
This point of view has particular interest in regard to transverse G-structures which
are defined below in Definition 2.5.2.
Haefliger [Hae62] has generalized this notion in order to construct a correct
homotopy theory for foliations. It has also been generalized by using other pseu-
dogroups, cf. [Hae71].
Definition 2.2.1: Let Γ be a subpseudogroup of the pseudogroup ΓGL(q,R) of local
diffeomorphisms of Rq , and let X be a topological space. A Haefliger cocycle on
X is given by the following data:
(i) an open cover U = {Uα }α∈I of X,
2.2. HAEFLIGER STRUCTURES 53

(ii) continuous maps fα : Uα −−→Rq , called local projections,


(iii) for each α, β ∈ I and x ∈ Uα ∩ Uβ , a diffeomorphism τγβ ∈ Γq from
a neighborhood of fα (x) onto a neighborhood of fβ (x) where the germ
of the map fα at the point fα (x) varies continuously with x, such that
fβ = τβα fα , and for all x ∈ Uα ∩ Uβ ∩ Uγ one has τγα = τγβ τβα .
We remark that if the local projections fα are open maps, then the second
condition in (iii) follows from the first. It is also clear that if X is a smooth
manifold and the local projections fα are local submersions we recover Haefliger’s
description of a foliation given above. In this case we say that the Haefliger cocycle
is a Haefliger cocycle of a foliation, and in the case of a proper subpseudogroup
Γq ⊂ ΓGL(q,R) , it is a Haefliger cocycle of a foliation with an additional transverse
structure. We shall consider this case in more detail in section 2.5. Definition 2.2.1
also makes perfect sense if we replace Rq by an appropriate q-dimensional ‘model
space’ N, and Γq by any pseudogroup of transformations on N.
Next we give an alternative definition of a Haefliger cocycle in terms of groupoids.
We refer to Appendix A for a brief discussion of groupoids. Actually this second
definition is a bit more general since it holds for any topological groupoid G. When
G is a groupoid of germs this definition is completely equivalent to Definition 2.2.1
by Proposition 1.5.4.
Definition 2.2.2: Let G be a topological groupoid with X as its space of units. A
Haefliger cocycle on X consists of an open cover {Uα } together with a family
of continuous maps gαβ : Uα ∩ Uβ −−→G such that for every x ∈ Uα ∩ Uβ ∩ Uγ the
condition gαγ (x) = gαβ (x)gβγ (x) holds. Two such cocycles are equivalent if they
can be extended to a cocycle on the union of the covers.
Notice that local projection maps are implicit in this definition. They are
obtained by the composition fα = s ◦ gαα : Uα −−→X. We now take the groupoid G
to be a groupoid of germs G(Γq ) of a pseudogroup Γq . Haefliger’s cocycle condition
2.2.2 is a cocycle for the cohomology set H 1 (U, G(Γq )). In this set two cocycles gβα
0
and gβα are equivalent if there exist germs of diffeomorphisms ζα , ζβ ∈ Γq such that
0
on fα (Uα ∩ Uβ ) the relation gβα = ζβ ◦ gβα ◦ ζα−1 holds. The equivalence stated in
Definition 2.2.2 is the following: we say that the Haefliger cocycle gαβ corresponding
to the cover U is equivalent to the Haefliger cocycle gα0 0 β 0 corresponding to the cover
U0 if there is a Haefliger cocycle gα0000 β 00 on the disjoint union U t U0 that restricts
to gαβ on U and to gα0 0 β 0 on U0 . As usual by taking the inductive limit over finer
covers we obtain the cohomology set H 1 (X, G(Γq )) whose elements we refer to as
Γq -structures or Haefliger structures. Unfortunately, the classes in H 1 (X, G(Γq ))
are not invariant under homotopy, so we need to define the set Γq (X) of homotopy
classes of elements in H 1 (X, G(Γq )). We say the two cocycles g1 , g2 ∈ H 1 (X, G(Γq ))
are homotopic if there exists an element Υ ∈ H 1 (X × I, G(Γq )) such that ι∗j Υ = gj ,
where I = [0, 1], and ιj : X−−→X × I are the natural face maps.
Exercise 2.2: Show that the property of being homotopic defines an equivalence
relation on H 1 (X, G(Γq )).
Now we are interested in the case when the Haefliger structure comes from a
foliation F on a smooth manifold M. In this case we have
54 2. FOLIATIONS

Proposition 2.2.3: Let M be a smooth manifold with a foliation F of codimension


q. Then F determines a unique Haefliger structure (g) ∈ H 1 (M, G(Γq )). Further-
more, any two foliations that determine the same Haefliger structure are isomor-
phic.
Proof. Given a foliated atlas U = {(Uα ; φα )} we obtain local submersions
fα : Uα −−→Rq as above. This uniquely determines a cocycle gβα by Equation
x
2.2.1 which determines a unique germ gβα , and this germ determines a unique
1 q
Haefliger structure (g) ∈ H (U, G(Γ )). Suppose that V = {(Vα0 ; ψα0 )} is another
foliated atlas for F, and gα0 : Vα0 −−→Rq are the corresponding local submersions.
This determines a unique element (υ) ∈ H 1 (V, G(Γq )). Let W = {(Wα ; φα )} be a
common refinement of U and V. By abuse of notation we let fα , gα denote their
corresponding restrictions to Wα . Then for each α there is a diffeomorphism ζα ∈ Γq
such that gα = ζα ◦ fα . So by the uniqueness of the diffeomorphism in 2.2.1 and we
have υαβ = ζα ◦ gαβ ◦ ζβ−1 . Then by passing to germs this gives a unique class in
H 1 (W, G(Γq )), and hence, in H 1 (M, G(Γq )).
Suppose two foliations F and F 0 determine the same Haefliger structure, then
they determine the same element (g) ∈ H 1 (U, G(Γq )) for some common foliated
atlas U. ¤

2.3. Leaf Holonomy and the Holonomy Groupoid

In this section we describe an important invariant of foliations, the holonomy


groupoid. We begin by describing the leaf holonomy of a given leaf, a concept
which generalizes the classical ‘first return map’ of Poincaré from dynamical systems
theory. Let (M, F) be a foliated manifold with dimension of M equal to n, and
ι
the dimension of F equal to p. A submanifold N ,→ M is called a transversal or
transverse submanifold if at each point x ∈ N we have
Tι(x) M = T Lx + Tx N.
In order to construct leaf holonomy, we now describe, following [Mol88], a proce-
dure known as “sliding along the leaves”. Consider two points x0 and x00 lying on the
same leaf L0 . Choose transversals N and N 0 through x0 and x00 , respectively, and
a piecewise smooth path γ : [0, 1]−−→L0 such that γ(0) = x0 and γ(1) = x00 . Subdi-
vide the interval [0, 1] into subintervals [tα−1 , tα ] with 0 = t0 < t1 < · · · < tk = 1
such that each γ([tα−1 , tα ]) lies in a simple open set Uα of a foliated atlas. More-
over, since γ([tα−1 , tα ]) is connected it lies on a unique plaque of L0 ∩ Uα . Let Nα
be transversals at γ(tα ). Then since γ(tα−1 ) lies on a single plaque in Uα−1 ∩ Uα ,
for each α = 1, . . . , k there is a diffeomorphism φα from an open neighborhood
Vi−1 ⊂ Nα−1 about γ(tα−1 ) onto an open neighborhood Vα ⊂ Nα about γ(tα )
which is constructed from the local submersions fα : Uα −−→Rq by the commutative
diagram
φα
Vα−1 −−−−→ Vα ,
(2.3.1) & .
Rq
where the diagonal arrows are fα−1 |Vα−1 and fα |Vα , respectively. By looking at the
(α−1) (α−1) (α−1) (α−1)
local foliated coordinate charts (Uα−1 ; x1 , . . . , xp , y1 , . . . , yq ) and
2.3. LEAF HOLONOMY AND THE HOLONOMY GROUPOID 55

(α) (α) (α) (α)


(Uα ; x1 , . . . , xp , y1 , . . . , yq ) one can easily identify the coordinate represen-
tative of φα with the Haefliger cocycle g(α−1)α . It follows that the composition
φk1 = φk ◦ · · · ◦ φ1 : V0 −−→Vk is a diffeomorphism that depends on the subdivision
and the intermediate transversals only through their domains. It does, however,
depend on the choice of foliated atlas, and the transversals at x0 and x00 . The set of
all local diffeomorphisms obtained in this way form a pseudogroup ΓU ,N,N 0 , called
the holonomy pseudogroup of F. This is a bit of a misnomer; however, a different
choice of foliated atlas, and different choices of transversals give pseudogroups that
are locally conjugate, so it makes sense to speak of ‘the’ holonomy pseudogroup.
Since the diffeomorphism φk1 described above depends on the path, the subdi-
vision, and intermediate transversals through their domains, it makes sense to pass
to germs. We denote the germ of the diffeomorphism φk1 at x0 by hγ . This is now
independent of the subdivision and transversals. Notice also that if γ 0 is another
path with the same endpoints and Γ : [0, 1] × [0, 1]−−→ ∪α Uα is a homotopy between
γ and γ 0 , then hγ only depends on the homotopy class [γ] of γ (cf. [CC00]). Now
suppose that x00 = x0 in which case we can take N 0 = N. Then hγ defines the
germ of a local diffeomorphism of N which leaves x0 fixed. We let Gxx00 denote the
set of all such germs obtained by sliding along leaves. (The notation will be clear
shortly). This set forms a group under composition. Notice that if γ and γ 0 are
two loops at x0 , then hγ 0 ◦γ = hγ 0 ◦ hγ , so one gets a group epimorphism
(2.3.2) hx0 : π1 (L0 , x0 )−−−→Gxx00 ,→ Diffx0 ,
where Diffx0 denotes the group of germs of diffeomorphisms of N that fix x0 . It
can be identified with the group of germs of local diffeomorphisms of Rq which fix
the origin.
Definition 2.3.1: The group Gxx00 is called the holonomy group of the leaf L0 at
x0 .
If a different transversal N 0 is chosen at x0 , then one can identify the germ of
the transversal N 0 with the germ of the transversal N. So the leaf holonomy group
is well defined up to this identification. Furthermore, if x1 ∈ L0 is another point of
the same leaf, any path γ : [0, 1]−−→L0 satisfying γ(0) = x0 and γ(1) = x1 induces
an isomorphism
γ ∗ : Gxx00 −−−→Gxx11 ,

where γ (hx0 ) = hγ ◦ hx0 ◦ hγ −1 . Thus, it makes sense to talk about the holonomy
group Hol(L) of a leaf L.
Proposition 2.3.2: Let F be a simple foliation. Then Hol(L) is trivial for all
leaves L of F.

Proof. Since the foliation is simple the local submersions in diagram 2.3.1 are
the restrictions of a global submersion f . Thus, after identifying Vα−1 with Vα by
translation along the leaves, the map φα in 2.3.1 is the identity, so sliding along the
leaves produces the germ of the identity diffeomorphism. ¤
The following proposition says that for any foliation the “generic” leaf has
trivial holonomy.
Proposition 2.3.3: For any foliated manifold the leaves having a trivial holonomy
group form a residual (i.e., a countable intersection of dense open sets) set.
Proof. Cf. [CC00]. ¤
56 2. FOLIATIONS

The kernel of the homomorphism hx of Equation 2.3.2 is a normal subgroup


of the fundamental group of the leaf called the holonomy kernel, and denote by
Holker(L). A space (leaf) whose fundamental group is Holker(L) is called the ho-
lonomy covering. We consider several examples regarding holonomy. The first is a
continuation of Example 2.1.2.
Example 2.3.4: Linear flows on tori. Again for simplicity we consider only T 2 .
In both the irrational and rational cases the holonomy groups are all trivial, but
for different reasons. In the rational case the foliation is simple, so the holonomy is
trivial. In the irrational case the leaves are simply connected, so the holonomy is
again trivial. However, in this latter case the holonomy pseudogroup is nontrivial.
The next example is more interesting and plays an important role in the sequel.
Example 2.3.5: Linear flows on odd dimensional spheres. Consider the unit
sphere in Cn+1 defined by
S 2n+1 = {z = (z0 , . . . , zn ) ∈ Cn+1 | |z0 |2 + · · · + |zn |2 = 1}.
By linear flow we mean linear in the complex Cartesian coordinates z = (z0 , · · · , zn ).
Again for simplicity we consider the case n = 1. Let w0 and w1 be two non-zero
real numbers, and consider the real vector field on C2
µ ¶
∂ ∂
ξ = i w0 z0 + w1 z1 − c.c .
∂z0 ∂z1
Restricted to S 3 this vector field is everywhere tangent to S 3 , and defines a nowhere
vanishing vector field on S 3 . Thus, ξ generates a 1-dimensional foliation on S 3 whose
flow is given by
(z0 , z1 ) 7→ (e2πiw0 t z0 , e2πiw1 t z1 ).
w1
Again there are two cases to consider depending on whether the ratio w 0
is rational
or irrational. If the ratio is rational by a reparameterization and considering the
complex conjugate coordinates if necessary, we can take the pair (w0 , w1 ) to be
coprime positive integers. In this case all the leaves are circles. The leaf through
x0 = (1, 0) is a circle which returns to x0 at t = w10 , whereas the nearby leaves with
z1 6= 0 take w0 times as long to return. So the holonomy group Hol(Lx0 ) ≈ Zw0 .
Similarly we see for x1 = (0, 1) the holonomy through the leaf Lx1 is Hol(Lx1 ) ≈
Zw1 . We also see that Holker(Li ) ≈ wi Z. All other leaves have trivial holonomy.
Now consider the case when the ratio w w0 is irrational. Consider the leaf through
1

a point with z0 z1 6= 0, say the point ( √2 , √12 ). This leaf is non-compact; in fact,
1

it is the irrational flow on the torus T 2 in S 3 defined by |z0 |2 = |z1 |2 = 12 . So this


leaf is not closed in S 3 . Its closure is just the 2-torus T 2 . The same phenomenon
occurs for any point (z0 , z1 ) ∈ S 3 with z0 z1 6= 0. However, the two leaves through
x0 = (1, 0) and x1 = (0, 1) are circles for they return after a time t = w10 and t = w11 ,
respectively. Nevertheless, all leaves have trivial holonomy which can be seen from
(v) of Theorem 2.5.10 below, but as in Example 2.3.4 the holonomy pseudogroup
is non-trivial.
The holonomy groups attached to a foliation can be studied together by form-
ing a more global object, the holonomy groupoid. Although the ideas go back to
Ehresmann, Reeb, and Thom, it appears that its first precise formulation was given
by Winkelnkemper [Win83]. See also [Rei83].
2.3. LEAF HOLONOMY AND THE HOLONOMY GROUPOID 57

Proposition 2.3.6: Let (M, F) be a foliated manifold. The collection of all triples
(x, y, [γ]), where x, y ∈ M lie on the same leaf L, γ is a piecewise smooth path
from x to y, and [γ] denotes the holonomy equivalence class of γ, i.e., γ ∼ γ 0
if γ 0 γ −1 = e ∈ Gyy , is a Lie groupoid of dimension n + p called the holonomy
groupoid1 of (M, F) and denoted by G = G(M, F). Furthermore, if every holonomy
germ is conjugate to an analytic germ, G(M, F) is Hausdorff.
Proof. In Appendix A.1 we give a brief review of groupoids. Clearly, the set
of objects G0 is M and this is a smooth Hausdorff manifold, while G1 is the set
of triples {(x, y, [γ])}. We show that G1 is a smooth manifold by following Connes
[Rei83] and exhibiting a system of coordinates on it. We then check that the
structure maps are smooth. Write the coordinates as (v, w) and (v 0 , w) having the
same transverse coordinate w at the beginning and end of a path, respectively. Here
v and v 0 are coordinates along the leaf. Then (v 0 , v, w) is a coordinate system on
G1 , and a change of coordinates is of the form
F (v 0 , v, w) = (f 0 (v 0 , w), f (v, w), g(w)),
where (v, w) 7→ (f (v, w), g(w)) is a change of coordinates on M. This gives G1 a
locally Euclidean topology of dimension n+p that is not necessarily Hausdorff. This
topology is also 2nd countable (see [Rei83], pg 137). Moreover, G1 is Hausdorff if
and only if each holonomy germ at x ∈ M which is the identity on some open set
whose closure contains x is the identity germ at x [Rei83, Win83]. In particular,
if every holonomy germ is conjugate to an analytic germ, the groupoid G(M, F)
will be Hausdorff.
The canonical source and target maps s : G−−→M and t : G−−→M, are defined
by
(2.3.3) s(x, y, [γ]) = x, t(x, y, [γ]) = y,
while the unit is u(x) = (x, x, 0), where 0 denotes the constant loop. Further-
more, the associative multiplication on G is defined when the range of one element
coincides with the source of the other, specifically in our case we have
(2.3.4) (x, y, [γ]) · (x0 , y 0 , [γ 0 ]) = (x, y 0 , [γ ◦ γ 0 ]) if y = x0
The inverse map is defined by ι(x, y, [γ]) = (y, x, [γ −1 ]). To see that the structure
maps are smooth, we write in coordinates:
s(v 0 , v, w) = (v, w), t(v 0 , v, w) = (v 0 , w), u(v, w) = (v, v, w),
0 0
ι(v , v, w) = (v, v , w), (v , v , w) · (v , v, w) = (v 00 , v, w).
00 0 0

So the structure maps are smooth. Furthermore, both s and t are submersions.
This completes the proof. ¤
The fact that the holonomy groupoid G(M, F) of a foliation is, in general, not
Hausdorff is of very little consequence for us, since it known to be Hausdorff when
the foliation F is Riemannian (see Definition 2.5.4 and Theorem 2.5.10 below)
which is the case of interest to us. Notice that F induces a foliation2 of dimension
2p on G(M, F) by pulling back the foliation F to G(M, F) by the source map s. We
denote this foliation on G(M, F) by FG . If E denotes the integrable subbundle of
T M corresponding to the foliation F, then the integrable subbundle of T G(M, F)
1This is called the graph of the foliation in [Win83].
2A different 2p-dimensional foliation having trivial holonomy groups is described in [Win83].
58 2. FOLIATIONS

corresponding to FG is s∗ E ⊕ V, where V is the vertical bundle consisting of the


tangent vectors to the fibres of the submersion s. So the leaf L(x,y,[α]) of FG through
(x, y, [α]) ∈ G(M, F) projects under s and t to the leaf Lx of F. Furthermore,
these projections induce an isomorphism of the corresponding holonomy groups,
see [Rei83] Proposition IV.2.44. Notice that the both subbundles s∗ E and V of
T G(M, F) are separately integrable. Summarizing we have
Proposition 2.3.7: Let F be a p-dimensional foliation on a manifold M. Then
the foliation G(M, F) F pullsback under either s or t to a 2p-dimensional foliation
FG on the holonomy groupoid G(M, F) that is invariant under the involution ι. If
L(x,y,[α]) denotes the leaf of FG through (x, y, [α]) ∈ G(M, F) and Lx denotes the
leaf of F through x ∈ M, then there is an isomorphism Hol(L(x,y,[α]) ) ≈ Hol(Lx ).
More information about the holonomy groupoid G(M, F) can be ascertained
when F is a Riemannian foliation. This will be described later in this section. Now
we consider another important groupoid introduced by Haefliger [Hae84] which he
called the transverse holonomy groupoid. It is closely related to, but different from
G(M, F). Here is the definition.
Definition 2.3.8: Let N be a complete transversal. The transverse holonomy
N
groupoid associated to N is the full subgroupoid GN of G(M, F).
N
GN is defined in the Appendix by Equations A.1.2. Here complete means that
N intersects each leaf at least once. Complete transversals always exist since we
N
do not require them to be connected. Of course, GN depends on N, but Haefliger
N
[Hae84] shows that GN is independent of N up to Morita equivalence. In fact
Theorem 2.3.9: Let F be a p-dimensional foliation on a manifold of dimension
p+q, and let N be any complete transversal. Then the transverse holonomy groupoid
N
GN is an étale Lie groupoid of dimension q that is Morita equivalent to the holonomy
groupoid G(M, F).

Proof. G N = t−1 (N ) is a submanifold of G1 such that

codim(G N ,→ G1 ) = codim(N ,→ G0 ) = p.

Since G1 has dimension 2p + q we see that dim G N = p + q = dim G0 . Now consider


the source map sN = s|G N restricted to G N . If we show that this map sN : G N −−→G0
is an immersion, it will be a local diffeomorphism. Then GN N
= s−1
N (N ) will be
N N
a submanifold of G of dimension q. So the restriction of sN to GN will be a
N
local diffeomorphism. Thus, GN will be étale. To see that the differential dsn is
injective at a morphism (g : x−−→y) ∈ G N ⊂ G1 , we consider a tangent vector
v ∈ Tg G1 . It will be tangent to G N if (dt)g (v) ∈ Ty N. Since s : G1 −−→G0 is
a submersion, it defines a p-dimensional foliation F̃ on G1 whose leaves are the
connected components of the fibres of the submersion. Thus, if dsN (v) = 0 with v
tangent to G N , then v is both tangent to the fibres of the source submersion and
dtg (v) is tangent to N. But (dt)g maps the tangent space to the leaf L̃g of F̃ at g to
the tangent space to the leaf Ly of F at y. Thus, v = 0 which proves the injectivity
of dsN , or equivalently, that sN is a local diffeomorphism.
N
To prove the last statement we consider the smooth functor F : GN −−→G(M, F)
defined by F (x, y, [g]) = (ιN (x), ιN (y), [g]). We see from the fact that the transversal
N is complete that sN is surjective. This shows that the map s◦π2 : N ×G0 G1 −−→G0
on pairs (x, g) satisfying ιN (x) = t(g) is surjective. Moreover, one easily checks
2.4. BASIC COHOMOLOGY 59

N
that corresponding square is a pullback. Thus, F : GN −−→G(M, F) is an essential
N
equivalence, and hence GN and G(M, F) are Morita equivalent. ¤
N
In particular, Theorem 2.3.9 says that the étale groupoid GN is independent of
N up to Morita equivalence. We shall see later that Morita equivalent groupoids
N
have homotopy equivalent classifying spaces. We will consider GN further in the
Chapter 4.
We are most interested in the case where the foliation arises from a locally free
action of a Lie group (2.1.3) with finite isotropy groups. So we recall the action
groupoid from the Appendix (Definition A.1.5). For G a Lie group and M a smooth
manifold, we consider the action groupoid M o G. According to Reinhart ([Rei83],
pg. 138) the space of morphisms of the holonomy groupoid of a foliation coming
from a locally free action of a Lie group G is just M × G. So in this case these two
groupoids are isomorphic.

2.4. Basic Cohomology

Basic cohomology was introduced by Reinhart in [Rei59b]. We begin with a


brief review following [Ton97]. Let F be a foliation of a smooth manifold M. A
differential r-form ω on M is said to be basic if for all vector fields V on M that
are tangent to the leaves of F the following conditions hold:
(2.4.1) V ω = 0, £V ω = 0.
In a local foliated coordinate chart (U, φ) with φ = (x1 , . . . , xp ; y1 , . . . , yq ) a basic
r-form ω takes the form
X
(2.4.2) ω= ωi1 ...ir (y1 , . . . , yq )dyi1 ∧ · · · ∧ dyir ,
where the sum is taken over all repeated indices.
r
Let ΩB (F) denote sheaf of germs of basic r-forms on M, and ΩrB (F) the set of
its global sections. The direct sum ΩB (F) = ⊕r ΩrB (F) is closed under addition and
exterior multiplication, and so is a subalgebra of the algebra of exterior differential
forms on M. Furthermore, exterior derivation takes basic forms to basic forms, that
is,
(2.4.3) £V dω = d£V ω = 0, and V dω = £V ω − d(V ω) = 0.
Thus, the subalgebra ΩB (F) forms a subcomplex of the de Rham complex, and
∗ ∗
its cohomology ring HB (F) = HB (F, dB ) is called the basic cohomology ring of
F. Here dB denotes the restriction of the exterior derivative d to the basic forms
ΩB (F).
r
The groups HB (F) are defined for all 0 ≤ r ≤ q, where q is the codimension
of the foliation, but generally they may be infinite dimensional for r ≥ 2. However,
they are finite dimensional for Riemannian foliations on compact manifolds which
r
is mainly the case of interest to us. Let us look at the groups HB (F) for r = 0, 1.
0
The set ΩB (F) is just the set of functions that are constant along the leaves of
F, so the cocycles with respect to dB are just the constant functions. Thus, if
0 1 1
M is connected HB (F) ≈ R. Let us now consider HB (F). We claim that HB (F)
1
injects into H (M, R) induced by the natural inclusion ΩB (F)−−→Ω(M ). Let ω be
a 1-cocycle, i.e., dB ω = 0, where ω is a smooth section of Ω1B (F). So if ω = df for
some smooth function on M, then we have V f = V df = V dω = 0 for any
60 2. FOLIATIONS

vector field V tangent to the leaves of F. But this says that f ∈ Ω1B (F), so ω is a
coboundary with respect to dB . This proves the claim. We have arrived at
Proposition 2.4.1: For any foliation F on a connected manifold M,
0
(i) HB (F) ≈ H 0 (M, R) ≈ R.
(ii) HB (F) is a subgroup of H 1 (M, R).
1

The appropriate notion of basic for vector fields is


Definition 2.4.2: A vector field X on a foliated manifold (M, F) is said to be
foliate with respect to F if for every vector field V tangent to the leaves of F, the
Lie bracket [X, V ] is tangent to the leaves of F.
Let us set some notation. The Lie algebra of all smooth vector fields on M is
denoted by Γ(M ). We denote by ΞF (M ) the Lie subalgebra of Γ(M ) consisting of
those vector fields that are tangent to the leaves of the foliation F. These are just
the smooth sections of the integrable vector bundle E defining F. The set of foliate
vector fields also forms a Lie subalgebra of Γ(M ). It is denoted by fol(M, F). From
Definition 2.4.2 one sees that it is precisely the normalizer of ΞF (M ) in Γ(M ).
Equivalently, its local one parameter group preserves the foliation. We denote the
subgroup of the diffeomorphism group Diff(M ) that preserve the foliation F by
Fol(M, F), that is
(2.4.4) Fol(M, F) = {φ ∈ Diff(M ) | φ∗ E ⊂ E}.
Exercise 2.3: Show that in local foliated coordinates (x1 , . . . , xp ; y1 , . . . , yq ) a fo-
liate vector field X takes the form
Xq Xp
∂ ∂
(2.4.5) X= Aa (y1 , . . . , yq ) + B i (x1 , . . . , xp ; y1 , . . . , yq ) ,
a=1
∂y a i=1
∂x i

where Aa and B i are smooth functions of the indicated variables.

2.5. Transverse Geometry

We are interested in foliations with a ‘transverse structure’. These were intro-


duced by Conlon [Con74] and further developed by Molino [Mol75, Mol88] in
a slightly different guise. Here we follow [Mol88], but our notation is somewhat
different. Let F be a p-dimensional foliation on the manifold M with integrable
subbundle E. There is an exact sequence of vector bundles, viz.
πQ
(2.5.1) 0−−−→E−−−→T M −−−→ Q−−−→0.
The quotient bundle Q is called the normal bundle of the foliation, and it is often
denoted by ν(F) to emphasize the foliation. Notice that any foliate vector field X
projects to a section X̄ of ν(F) that is independent of the coordinates along the
leaves. This follows easily from the coordinate form 2.4.5 together with a change in
foliated coordinate charts. The set of all such sections forms a Lie algebra, denoted
by trans(M, F), and we have an exact sequence of Lie algebras
(2.5.2) 0−−−→ΞF (M )−−−→fol(M, F)−−−→trans(M, F)−−−→0.
Here the Lie bracket on trans(M, F) is defined by [X̄, Ȳ ] = [X, Y ]. We call elements
of trans(M, F) transverse vector fields. Note also that the Lie derivative of basic
tensor fields with respect to transverse vector fields is well defined [Mol88].
2.5. TRANSVERSE GEOMETRY 61

A transverse frame at the point x ∈ M is a basis (Y1 , . . . , Yq ) for the fibre Qx of


Q at x. (Yi is not a tangent vector at x, but an equivalence class of tangent vectors
modulo vectors in Ex .) The set LT (M, F) of all transverse frames at all points of
M forms a principal GL(q, R) bundle on M, called the transverse frame bundle on
(M, F). Let G ⊂ GL(q, R) be a Lie subgroup and let πT : PT (M, G, F)−−→M denote
the corresponding principal G subbundle of LT (M, F). As with the ordinary frame
bundle a point z ∈ LT (M, F) can be viewed as the linear map z : Rq −−→QπT (z)
that assigns to the standard basis of Rq the frame z = (Y1 , . . . , Yq ). Also as in the
usual case LT (M, F) and its subbundles PT (M, G, F) have a canonical Rq -valued
1-form θT defined by
(2.5.3) hθT , Xi = z −1 πQ πT ∗ X.
The foliation F on M lifts to a foliation FT on LT (M, F) as follows: we define a
foliation on LT (M, F) by the p-dimensional subbundle ET ⊂ T LT (M, F) generated
by the vectors in Tz LT (M, F) that satisfy
(2.5.4) ETz = {Xz ∈ Tz LT (M, F) | Xz θT = Xz dθT = 0 for all z ∈ LT (M, F)}.
We now have
Proposition 2.5.1: Let (M, F) be a foliated manifold and let LT (M, F) be its
transverse frame bundle. The subbundle ET defined by equation 2.5.4 is integrable,
and thus defines a foliation FT on LT (M, F). The leaves of this foliation project
under πT to the leaves of the foliation F.

Proof. Let X, Y be vector fields on LT (M, F) that are sections of the sub-
bundle ET . Then
0 = dθT (X, Y ) = XθT (Y ) − Y θT (X) − θT ([X, Y ]) = −θT ([X, Y ]),
so [X, Y ] θT = 0. But also we have for any section X of ET that
£X θT = d(X θT ) + X dθT = 0.
Thus,
0 = [£X , £Y ]θT = £[X,Y ] θT = d([X, Y ] θT ) + [X, Y ] dθT = [X, Y ] dθT
which implies that ET is integrable by Frobenius’ Theorem.
Now suppose that LT is a leaf of FT . Let U be a simple open set of M such
that the quotient map of the foliation restricted to U is a submersion fU : U −−→Rq .
Then on ET |U = πT−1 (U ), the canonical 1-form θT is just the pullback fU∗ θ, where
θ is the canonical 1-form on the frame bundle L(Rq ) on Rq . So in local coordinates
(x1 , . . . , xp ; y1 , . . . , yq ; A11 , . . . , Aq1 , . . . , Aqq ) on ET , equation 2.5.4 implies that the
tangent vectors to ET that are tangent to the leaves of the foliation FT are spanned
∂ ∂
by the ∂x j
and ∂A i . Thus, these vectors project under πT to vectors that are tangent
j
to the leaves of the foliation F on M. But since πT is surjective this implies that
the leaf LT of FT projects to a leaf of F. ¤

The foliation FT is called the lifted foliation. We are now ready for
Definition 2.5.2: Let G be a Lie subgroup of GL(q, R), and let PT (M, G, F) be
a principal G subbundle of LT (M, F). We say that PT (M, G, F) is a transverse
G-structure if at each point z ∈ PT (M, G, F) the tangent space Tz PT (M, G, F)
contains the subspace ETz tangent to the leaves of the lifted foliation FT .
62 2. FOLIATIONS

Important examples for us will be when G = O(q, R), G = GL(m, C), and
G = U (m), where q = 2m. These correspond to transverse Riemannian structures,
transverse almost complex structures, and transverse almost Hermitian structures,
respectively. The following result [Mol88] gives a characterization of transverse
G-structures in terms of the local submersions.
Proposition 2.5.3: Let PT (M, G, F) be a principal G subbundle of LT (M, F).
Then PT (M, G, F) is a transverse G-structure if and only if for every simple open
set U, the bundle PT (M, G, F)|U is the pullback by πT : LT (U, F|U )−−→L(Rq ) of a
G-structure P (Rq , G) on Rq .

Proof. The argument is purely local, and the details are left to the reader. ¤

2.5.1. Transverse Riemannian Geometry. The foliations that interest us


most are those with a transverse Riemannian structure and actually more, a trans-
verse Kähler structure. It turns out that allowing for a transverse Riemannian
structure adds quite a bit of tractability to the subject which was begun by Rein-
hart [Rei59a], and subsequently developed by Molino and others. See [Mol88]
and references therein.
Definition 2.5.4: A Riemannian foliation is a foliation with a transverse Rie-
mannian structure, i.e., a transverse G-structure with G = O(q, R).
A Riemannian foliation (M, F) defines a transverse Riemannian metric gT on
M by
(2.5.5) gT (X, Y ) = hz −1 πQ X, z −1 πQ Y i,
where X, Y ∈ T MπT (z) and πQ is the projection in the exact sequence 2.5.1. Notice
that gT defines a nonnegative symmetric bilinear form on T M, and viewed as a
section of Sym2 T M has kernel E ⊗ T M + T M ⊗ E, and thus defines a Riemannian
metric on the quotient bundle Q.
Proposition 2.5.5: Let F be a codimension q foliation on M. There is a 1-1 corre-
spondence between transverse O(q, R)-structures on M and transverse Riemannian
metrics on M.
The second definition is that of a bundle-like metric which is due to Reinhart
[Rei59a].
Definition 2.5.6: A Riemannian metric g is said to be bundle-like with respect
to a foliation F if for any foliate horizontal vector fields X, Y , the function g(X, Y )
is basic, i.e., for any vector field V along the leaves of F the equation V g(X, Y ) = 0.
The following is well-known [Mol88]
Proposition 2.5.7: If g is a bundle-like metric with respect to a foliation F, then
F is a Riemannian foliation. Conversely, if F is a Riemannian foliation with trans-
verse metric gT , then there exist bundle-like metrics g whose associated transverse
metric is gT .

Proof. Let g be a bundle-like metric on M. This metric splits the exact se-
quence (2.5.1) as
(2.5.6) T M = E ⊕ E⊥,
2.5. TRANSVERSE GEOMETRY 63

where E ⊥ denotes the subbundle of T M consisting of all vectors perpendicular to


the vectors of E with respect to the Riemannian metric g. Thus, any vector field
X on M can be decomposed as
X = X > + X ⊥,
where X > is a section of E and X ⊥ is a section of E ⊥ . Define a transverse met-
ric on (M, F) by gT (X, Y ) = g(X ⊥ , Y ⊥ ). Since g is bundle-like, this makes F a
Riemannian foliation.
Conversely, given a Riemannian foliation, we have a transverse metric gT de-
fined by Equation 2.5.5, and we can choose a Riemannian metric in the vector
bundle E to obtain a bundle-like metric on M. ¤

In terms of Haefliger structures of Section 2.2, the transverse metric gT can be


obtained by pulling back Riemannian metrics on Rq by the local submersions fα .
In this case the transition functions ταβ local isometries between local Riemannian
metrics. In fact one can give an equivalent definition of a Riemannian foliation in
terms Haefliger cocycles:
Proposition 2.5.8: Let (M, F) be a foliated manifold, and let {τ } be the associated
Haefliger cocyle. Then (M, F) is a Riemannian foliation if and only if there are

Riemannian metrics gα on each fα (Uα ) such that gα = τβα gβ .

Proof. Given a foliation F whose associated Haefliger structure satisfies gα =



τβα gβ with Riemannian metrics gα on each open set fα (Uα ) ⊂ Rq , we obtain a
transverse Riemannian metric gT,α on each open set Uα and the cocycle conditions
imply that these metrics agree on each overlap Uα ∩ Uβ giving a global transverse
Riemannian metric gT . Hence, the foliation is Riemannian. Conversely, if F is
Riemannian we have a transverse Riemannian metric gT which extends to a bundle-
like metric g on M. Since for any foliate vector fields X, Y the function g(X, Y ) is
basic, we can define Riemannian metrics gα on each fα (Uα ) by gα (fα∗ X, fα∗ Y ) =

g|Uα (X, Y ). It is then easy to check that gα = τβα gβ . ¤

For a Riemannian foliation on a compact manifold we have [Mol88, Rei83]


Proposition 2.5.9: Let (M, F) be a Riemannian foliation on a compact connected
manifold M. Then the leaves of F all have the same universal cover.
In fact a common covering of the leaves appears naturally in the holonomy
groupoid G(M, F) of a Riemannian foliation. Generally, for a Riemannian foliation
G(M, F) has some nice special properties. We say that a Riemannian submersion
is horizontally complete if any horizontal geodesic can be extended for all values of
its affine parameter.
Theorem 2.5.10: Let F be a p-dimensional Riemannian foliation on the manifold
M. Then
(i) The holonomy groupoid G(M, F) of F is a Hausdorff Lie groupoid.
(ii) There is a unique Riemannian metric gG on G(M, F) such that both s
and t are Riemannian submersions.
(iii) The inverse map ι is an isometry with respect to gG .
(iv) The 2p dimensional foliation FG on G(M, F) is a Riemannian foliation
such that each leaf is locally the Riemannian product of leaves of the
p-dimensional foliations.
64 2. FOLIATIONS

(v) If the source submersion s : G(M, F)−−→M is horizontally complete, then


it is a locally trivial fibre bundle whose fibres are precisely the holonomy
covers of the leaves of F.
(vi) If M is compact and all leaves are compact, then G(M, F) is compact.
Proof. (i): The holonomy groupoid is a Lie groupoid by Proposition 2.3.6.
Since the transverse geometry of F is orthogonal, so are the leaf holonomy groups.
Thus, the holonomy is linear, and hence, analytic. So the result follows by Propo-
sition 2.3.6. For the remainder see Proposition IV.4.23 of [Rei83]. ¤
Next we give a theorem of Molino [Mol88] that is of great importance to us in
the sequel. It says that a Riemannian foliation with compact leaves has a tractable
space of leaves, namely an orbifold which we discuss in detail in Chapter 4.
Theorem 2.5.11: Let F be a p-dimensional Riemannian foliation with compact
leaves on the manifold M. Then the space of leaves M/F admits the structure of
a p-dimensional orbifold such that the canonical projection π : M −−→M/F is an
orbifold Riemannian submersion.
Proof. The proof can be found in [Mol88], Proposition 3.7. ¤
Now we discuss the relation between the curvature of (M, g) and the curvature
of the transverse metric gT , where g is a bundle-like metric whose transverse com-
ponent is gT . Accordingly we split T M as in 2.5.6. This relationship is essentially
that of O’Neill for Riemannian submersions [O’N66, Rei83, Ton97]. We first
define the induced connection ∇T in the bundle E ⊥ .
(
πE ⊥ (∇X Y ) if X is a smooth section of E ⊥ ,
(2.5.7) ∇TX Y =
πE ⊥ [V, Y ] if X = V is a smooth section of E,
where Y is a smooth section of E. It is left as an exercise to mimic the standard
proof of the first fundamental theorem of Riemannian geometry that ∇T is the
unique torsion free connection such that the transverse metric gT is parallel, i.e.,
∇TX gT = 0.
Exercise 2.4: Show that the connection ∇T is the unique torsion-free metric con-
nection with respect to the transverse metric gT .
We let RT denote the Riemannian curvature tensor field with respect to gT .
Then if Z, W are any vector fields on M, and X is a vector field in E ⊥ , we have
(2.5.8) RT (Z, W )X = ∇TZ ∇TW X − ∇TW ∇TZ X − ∇T[Z,W ] .
Proposition 2.5.12: For any smooth section V of E the following hold:
(i) £V ∇T = 0.
(ii) V RT = 0.
(iii) £V RT = 0.
Proof. (i) follows as usual from the fact that ∇T is determined uniquely in
term the transverse metric gT and £V gT = 0. (ii): If W is a vector field on M and
X is a smooth section of E ⊥ we have
RT (V, W )X = ∇TV ∇TW X − ∇TW ∇TV X − ∇T[V,W ] X
= £V ∇TW X − ∇TW £V X − ∇T£V W X = (£V ∇T )W (X) = 0.
We leave the proof of (iii) as an exercise. ¤
2.5. TRANSVERSE GEOMETRY 65

Exercise 2.5: Prove (iii) of Proposition 2.5.12.


Let us fix some common notation [Bes87, Ton97]: we let Ei denote any vector
fields on M, while we let U, V, W denote vector fields in the integrable subbundle E,
and X, Y, Z be vector fields in E ⊥ . We also let π : T M −−→E, and π ⊥ : T M −−→E ⊥
denote the corresponding natural projections. Following O’Neill [O’N83] (see also
Gray [Gra67]) we introduce the type (2, 1) tensor field
(2.5.9) TE1 E2 = π(∇π(E1 ) π ⊥ (E2 )) + π ⊥ (∇π(E1 ) π(E2 )).
Exercise 2.6: Show that T defined by 2.5.9 is a tensor field of type (2, 1).
Clearly, T only depends of projection of E1 along the leaves, so
(2.5.10) TX U = TX Y = 0.
We also have
(2.5.11) TU V = π ⊥ (∇U V ), TU X = π(∇U X),
(2.5.12) TU V = TV U,
(2.5.13) T is anti-symmetric, i.e., g(TU V, X) = −g(V, TU X).
The symmetry property Equation 2.5.12 follows by taking U and V as coordinate
vector fields along the leaves of the foliation F (valid since T is a tensor field),
and using the fact that ∇ is torsion-free, while the anti-symmetry property 2.5.13
follows since ∇ is a metric connection. Notice that by Equation 2.5.13 T vanishes
if and only if TU V = 0 for all vector fields tangent to the leaves of F. Since any leaf
L is an immersed submanifold of M, we see from Equation 2.5.11 that L is totally
geodesic if and only if T vanishes along L. If TU V = 0 for all U and V then all
leaves of F are totally geodesic. In this case we say that the foliation F is a totally
geodesic foliation. Summarizing we have
Proposition 2.5.13: A foliation F is totally geodesic if and only if T = 0.
Reversing the roles of the subbundles E and E ⊥ in Equation 2.5.9 we obtain
(2.5.14) AE1 E2 = π(∇π⊥ (E1 ) π ⊥ (E2 )) + π ⊥ (∇π⊥ (E1 ) π(E2 )).
Exercise 2.7: Show that A defined by 2.5.14 is a tensor field of type (2, 1).
For O’Neill’s A tensor field, we have the following
(2.5.15) AU V = 0, AU X = 0,

(2.5.16) AX U = π (∇X U ), AX Y = π(∇X Y ),
(2.5.17) AX Y = −AY X,
(2.5.18) A is anti-symmetric, i.e., g(AX U, Y ) = −g(U, AX Y ).
Equations 2.5.15 and 2.5.16 are obvious from the definition. Equation 2.5.17 will
be proved in Proposition 2.5.14 below, while again Equation 2.5.18 follows since
∇g = 0.
Proposition 2.5.14: For any horizontal vector fields X, Y we have
1
AX Y = π([X, Y ]).
2
Proof. We compute
π([X, Y ]) = π(∇X Y − ∇Y X) = AX Y − AY X.
66 2. FOLIATIONS

The result will now follow if we establish Equation 2.5.17. It suffices to show that
AX X = 0. Without loss of generality we may take the horizontal vector field X to
be foliate. We have for arbitrary vertical U

g(U, AX X) = g(U, π(∇X X)) = g(U, ∇X X) = −g(∇X U, X)


1
= −g(∇U X, X) − g([X, U ], X) = − U g(X, X) = 0.
2

The last expression vanishes since g(X, X) is basic when X is foliate. ¤

Notice that Equation 2.5.18 implies that A vanishes precisely when AX Y van-
ishes for all horizontal vector fields X and Y, and Proposition 2.5.14 says that
AX Y vanishes for all X, Y if and only if the subbundle E ⊥ is integrable. Next let
us see how the covariant derivative on M decomposes. We let ∇ ˆ be the induced
connection along the leaves. Then we have Gauss’ formula for the leaves

(2.5.19) ˆ U V + TU V.
∇U V = ∇

We also have

(2.5.20) ∇U X = TU X + π ⊥ (∇U X),


(2.5.21) ∇X U = π(∇X U ) + AX U,
(2.5.22) ∇X Y = AX Y + ∇TX Y.

Concerning the covariant derivatives of A and T we have the following:


Lemma 2.5.15: The following hold:
(i) (∇U A)V = −ATU V , (∇X T )Y = −TAX Y ,
(ii) (∇X A)V = −AAX V , (∇V T )Y = −TTV Y ,
(iii) g((∇U A)X V, W ) = g(TU V, AX W ) − g(TU W, AX V ),
(iv) g((∇E1 A)X Y, V ) is anti-symmetric in X, Y,
(v) g((∇E1 T )U V, X) is anti-symmetric in U, V,
(vi) Sg((∇Z A)X Y, V ) = Sg(AX Y, TV Z), where S denotes the cyclic sum
taken over X, Y, Z.

Proof. We establish the first identity of (ii), and leave the rest as an exercise.
For any vector field F we have

(∇X A)V F = ∇X (AV F ) − A∇X V F − AV (∇X F ) = −AAX V F. ¤

Exercise 2.8: Finish the proof of Lemma 2.5.15.


Theorem 2.5.16: Let R, R̂, and RT denote the Riemann curvature associated to
ˆ and ∇T , respectively Then the following curvature identities
the connections ∇, ∇,
2.5. TRANSVERSE GEOMETRY 67

hold:
g(R(U, V )W, W 0 ) = g(R̂(U, V )W, W 0 ) + g(TU W, TV W 0 ) − g(TV W, TU W 0 );
g(R(U, V )W, X) = g((∇U TV )W, X) − g((∇V TU )W, X);
g(R(X, U )Y, V ) = g(TU X, TV Y ) − g((∇X TU )V, Y )
− g(AX U, AY V ) − g((∇U A)X Y, V );
g(R(U, V )X, Y ) = g((∇V A)X Y, U ) − g((∇U A)X Y, V )
+ g(AX V, AY U ) − g(AX U, AY V )
+ g(TU X, TV Y ) − g(TV X, TU Y );
g(R(X, Y )Z, U ) = g(AY Z, TU X) + g(AZ X, TU Y )
+ g((∇Z A)X Y, U ) − g(AX Y, TU Z);
g(R(X, Y )Z, Z ) = g(RT (X, Y )Z, Z 0 ) + 2g(AX Y, AZ Z 0 )
0

+ g(AX Z, AY Z 0 ) − g(AX Z 0 , AY Z)
Proof. The first equation is just Gauss’ Equation along the leaves viewed as
immersed submanifolds, while the second equation is the Codazzi equation along
the leaves. We leave the proof for the other equations to the reader. ¤
We now have an easy corollary for the various sectional curvatures.
Corollary 2.5.17: Let X, Y, U, V be linearly independent unit vector fields with
respect to the metric g, and let K, K T , K̂ denote the sectional curvatures with respect
to the metrics g, gT , g|L , respectively. Then one has
K(U, V ) = K̂(U, V ) + |TU V |2 − g(TU U, TV V ),
K(X, U ) = g((∇X T )U U, X) − |TU X|2 + |AX U |2 ,
K(X, Y ) = K T (X, Y ) − 3|AX Y |2 .
We would like expressions for the Ricci and scalar curvatures similar to those
of Theorem 2.5.16. First we need some preliminary development which we lift from
Besse [Bes87]. In what follows we let {Xi }i and {Ua }a denote local orthonormal
frames for the horizontal subspace E ⊥ and vertical subspace E, respectively. The
following identities hold:
X X
(2.5.23) g(AX , AY ) = g(AX Xi , AY Xi ) = g(AX Ua , AY Ua ),
i a
X X
(2.5.24) g(AX , TU ) = g(AX Xi , TU Xi ) = g(AX Ua , TU Ua ),
i a
X
(2.5.25) g(AU, AV ) = g(AXi U, AXi V ),
i
X
(2.5.26) g(T X, T Y ) = g(TUa X, TUa Y ).
a
Next for any tensor field T we define the projected divergence operators,
X
(2.5.27) δT T = − (∇Xi T )Xi ,
i
X
(2.5.28) δ̂T = − (∇Ua T )Ua ,
a
68 2. FOLIATIONS

which satisfy δ = δ̂ + δ T with respect to the ordinary divergence operator δ. We


define the mean curvature vector field N along the leaves of F by
X
(2.5.29) N= TUa Ua
a
which vanishes if and only if each leaf is a minimal submanifold of M. We are now
ready for the relations between the Ricci curvatures. We let Ric, RicT , and Ric d
denote the Ricci curvatures of g, gT , and the family of metrics g|L , respectively.
Theorem 2.5.18: The Ricci curvatures Ric, RicT , and Ric d satisfy
X
d
Ric(U, V ) = Ric(U, V ) − g(N, TU V ) + g(AU, AV ) + g((∇Xi T )U V, Xi ),
i

Ric(X, U ) = g((δ̂T )U, X) + g(∇U N, X) − g((δ T A)X, U ) − 2g(AX , TU ),


1
Ric(X, Y ) = RicT (X, Y ) − 2g(AX , AY ) − g(T X, T Y ) + (g(∇X N, Y ) + g(∇Y N, X)).
2
Defining the quantities
X X
(2.5.30) |A|2 = g(AXi , AXi ) = g(AUa , AUa ),
i a
X X
2
(2.5.31) |T | = g(T Xi , T Xi ) = g(TUa , TUa ),
i a
we can give the relation between the various scalar curvatures, viz.
Corollary 2.5.19: If s, sT , ŝ denote the scalar curvatures of the metrics g, gT , g|L ,
respectively, then
s = sT + ŝ − |A|2 − |T |2 − |N |2 − 2δ T N.

2.5.2. Transverse Complex Geometry. Here we describe only the basics


of transverse complex and almost complex geometry. We present much more detail
in the special case of interest to us, namely transverse almost Kähler and Kähler
geometry, in Sections 7.2 and 7.5 below. For related work on transverse Kähler
foliations see [NT88]. Let (M, F) be a foliated manifold, then
Definition 2.5.20: A transverse almost complex structure is a smooth sec-
tion J¯ of End ν(F) that satisfies J¯2 = −1lν(F ) .
This implies that the normal bundle ν(F) is even dimensional, say 2q. Then
alternatively, a transverse almost complex structure on a foliated manifold (M, F)
is a transverse G-structure with G = GL(q, C). We are particularly interested in
the case when the transverse almost complex structure is integrable. This is done
in terms of a Nijenhuis torsion tensor in the case of almost contact structures
in subsection 6.5.4 of Chapter 6. Here it seems both more conceptual and more
convenient to describe the integrable case in terms of Haefliger structures. It is
more convenient from the point of view of deformation theory adopted in Section
8.2. Let (M, F) be a foliated manifold of codimension 2q with local submersions
fα : Uα −−→R2q = Cq related by Equation 2.2.1 with transition functions τβα that
satisfy Haefliger’s cocycle conditions 2.2.2. Then
Definition 2.5.21: The foliated manifold (M, F) is said to have a transverse
holomorphic structure if the local diffeomorphisms τβα : fα (Uα ∩Uβ )−−→fβ (Uα ∩
Uβ ) are biholomorphisms for all α, β.
2.6. RIEMANNIAN FLOWS 69

It is easy to see that a foliated manifold (M, F) with a transverse holomorphic


structure defines a transverse almost complex structure J¯ on the normal bundle
ν(F). We can then decompose the complexification ν(F)C = ν(F) ⊗ C into the
eigenspaces of J,¯

(2.5.32) ν(F)C = ν(F)1,0 ⊕ ν(F)0,1 .


Similarly, we get a splitting of the co-normal bundle
(2.5.33) ν ∗ (F)C = ν ∗ (F)1,0 ⊕ ν ∗ (F)0,1 ,
and this induces a splitting of the exterior differential algebra over ν ∗ (F)C
M
Λr ν ∗ (F)C = Λp,q ν ∗ (F)C ,
p+q=r

where Λp,q ν ∗ (F)C = Λp ν ∗ (F)1,0 ⊗ Λq ν ∗ (F)0,1 . We let Ωp,q


B denote the set of basic
forms of type (p, q), that is the set of smooth basic sections of Λp,q ν ∗ (F)C . As in
the usual case there is a splitting
M p,q
(2.5.34) ΩrB (F) ⊗ C = ΩB ,
p+q=r

as well as the basic Dolbeault complex


∂¯ ∂¯
(2.5.35) 0−−→Ωp,0
B − −→ Ωp,1
B − −→ · · · −−→Ωp,n
B − −→0,
p,q
together with its basic Dolbeault cohomology groups HB (Fξ ). This will be devel-
oped much further in the context of K-contact and Sasakian geometry in Section
7.2.
We end this section with a discussion of transformation groups that preserve
the transverse holomorphic structure. Notice that any φ ∈ Fol(M, F) induces
a map φ̄∗ : ν(F) → ν(F). So we define the group of transversely holomorphic
¯ by
transformations HT (F, J)
(2.5.36) ¯ = {φ ∈ Fol(M, Fξ ) | φ̄∗ ◦ J¯ = J¯ ◦ φ̄∗ }.
HT (F, J)
Since a 1-parameter subgroup of any element of ΞF (M ) induces the identity on
¯ is infinite dimensional. We are mainly interested in the
ν(Fξ ), the group HT (F, J)
¯
infinitesimal version, that is in the ‘Lie algebra’ of HT (F, J).
Definition 2.5.22: Let (M, F) be a foliated manifold with a transverse holomor-
¯ We say that a vector field X ∈ fol(M, F) is transversally holo-
phic structure J.
morphic if given any section Ȳ of ν(F), we have
[X̄, J¯ Ȳ ] = J¯ [X, Y ] .
¯
The set of all such vector fields will be denoted by hT (F, J).
T ¯
Exercise 2.9: Show that h (F, J) is a Lie subalgebra of trans(M, F).

2.6. Riemannian Flows

We shall be interested in 1-dimensional Riemannian foliations.


Definition 2.6.1: A one dimensional oriented Riemannian foliation is called a
Riemannian flow.
70 2. FOLIATIONS

Before we begin our study of Riemannian flows, we briefly mention one dimen-
sional foliations that are never Riemannian, the so-called Anosov flows [Ano69].
Any oriented 1-dimensional foliation on a smooth manifold M is determined by a
nowhere vanishing vector field V. The flow φt determined by V is said to be Anosov
if the tangent bundle T M splits as T M = LV ⊕ E s ⊕ E u , where LV is the trivial
line bundle generated by V and E s (E u ) denote the stable (unstable) subbundles
defined as follows: there exist constants c ≥ 1 and a > 0 such that for all t ≥ 0 the
estimates hold:
(i) |(φt )∗ X| ≤ ce−at |X| for every smooth section X of E s ,
(ii) |(φt )∗ X| ≥ c−1 eat |X| for every smooth section X of E u .
Here the norm | · | is taken with respect to some Riemannian metric on M, but
when M is compact the condition of being Anosov is actually independent of the
metric. An important example of an Anosov flow is the geodesic flow on the unit
tangent bundle over a compact manifold of constant negative curvature [Ano69].
See also [Ton97].
For a given Riemannian foliation (M, F) one is interested in whether the leaves
are closed in the topology of M. This is of particular interest to us in the case of
Riemannian flows. In order to study this in more detail we consider certain types
of “singular foliations”. Here we follow Molino [Mol88].
Definition 2.6.2: Let M be a smooth manifold. A singular Riemannian folia-
tion SF is a partition of M into connected immersed submanifolds, called leaves
that satisfy:
(i) The Lie subalgebra of vector fields tangent to the leaves of SF is transitive
on each leaf.
(ii) There exists a Riemannian metric g on M such that every geodesic that
is perpendicular at one point to a leaf remains perpendicular to every leaf
it meets.
It is not difficult to see that any ordinary Riemannian foliation satisfies these
conditions, and thus is a special case of a singular Riemannian foliation. Further-
more, Molino shows a converse in the following sense. If F is a foliation on M (in
the usual sense) and the conditions of Definition 2.6.2 hold, then F is a Riemannian
foliation with respect to g. For us the importance of singular Riemannian foliations
is given by
Proposition 2.6.3 (Molino): Let (M, F) be a Riemannian foliation on a compact
connected manifold M. Then the closures F̄ of the leaves of F form a singular
Riemannian foliation for which the leaves are embedded submanifolds of M.

Proof. See [Mol88] Proposition 6.2 and Theorem 5.1. ¤

We are interested in the case when the Riemannian flow is an isometry, and then
many simplifications occur. Indeed the following theorem of Carrière [Car84b] (see
also Appendix A of [Mol88] by Carrière) becomes much easier to prove when the
flow is an isometry.
Theorem 2.6.4: Let F be Riemannian flow on a compact manifold M. Then the
leaf closure L̄ of a leaf L of F is diffeomorphic to a torus T k , and F restricted to
L̄ is conjugate to a linear flow.
2.6. RIEMANNIAN FLOWS 71

Proof. Here we give the proof assuming that the flow is an isometry, but the
theorem holds more generally. See [Car84b] for the proof of the general case. Let g
be a bundle-like Riemannian metric for (M, F) and let φ denote the one parameter
subgroup of the isometry group I(M, g) generated by F. Since M is compact, so is
I(M, g), and the closure φ̄ is a torus T ⊂ I(M, g). For any x ∈ M the closure φx
of the orbit φx coincides with the orbit φ̄x of the closure φ̄. So if φx denotes the
isotropy subgroup at x, we see that φ̄/φ¯x is a torus T k . Hence, the leaf closure L̄
is a torus. Furthermore, φ is a subgroup of the torus, so it is conjugate to a linear
flow. ¤
Note that the dimension k = k(L) of the torus in general depends on the leaf L
and is clearly bounded by the dimension of the manifold. This leads to
Definition 2.6.5: The toral rank or just rank rk(M, F) of a Riemannian flow
F on a compact manifold M is defined by
rk(M, F) = max k(L).
L∈M

It is clear that 1 ≤ rk(M, F) ≤ dim M. The rk(M, F) = 1 case will be of great


interest to us. In this case the leaves are all circles, and the space of leaves will be
a compact orbifold. See Chapter 4 for details.
Definition 2.6.6: A one dimensional foliation F on M is said to be isometric if
there is a Riemannian metric g on M such that the flow is an isometry of g.
We have
Proposition 2.6.7: An isometric one dimensional foliation is Riemannian.
Proof. Let V be a nowhere vanishing vector field on M generating a one
dimensional foliation F that is isometric with respect to a Riemannian metric g,
and let X, Y be any horizontal foliate vector fields on M. Then
(2.6.1) V g(X, Y ) = (£V g)(X, Y ) + g([V, X], Y ) + g(X, [V, Y ]).
The first term vanishes since V is an infinitesimal isometry (Killing vector field),
and the last two terms vanish since X, Y are both horizontal and foliate. Thus, g
is a bundle-like metric for (M, F). So F is Riemannian by Proposition 2.5.7. ¤
The converse is not true in general; however, see Proposition 2.6.11 below.
Carrière has related the isometry of a one dimensional foliation to the orbits being
geodesics. The following result is well-known [Ton97].
Lemma 2.6.8: Let F be a one dimensional foliation on a manifold M, and let V
be a nowhere vanishing vector field generating F. Then the following are equivalent:
(i) There exists a Riemannian metric g on M such that V has unit length
and the orbits of V are geodesics.
(ii) There exists a 1-form η on M such that η(V ) = 1 and £V η = 0.
(iii) There exists a 1-form η on M such that η(V ) = 1 and V dη = 0.
Proof. Choose a metric g on M such that V has unit length and define η to
be the dual of V , that is η = g(V, ·). Then the equivalence of (ii) and (iii) follows
immediately from the well-known formula £X = ιX ◦ d + d ◦ ιX . Now let ∇ denote
any affine connection on M. Then, the orbits of V are geodesics with respect to the
connection ∇ if and only if the mean curvature vector field (∇V V )⊥ vanishes, where
⊥ indicates the component perpendicular to V. Equivalently, this means that the
72 2. FOLIATIONS

dual mean curvature 1-form κ = g((∇V V )⊥ , ·) vanishes. But for any vector field
X perpendicular to V , we have
κ(X) = g(∇V V, X) = V g(V, X) − ∇V g(V, X) − g(V, ∇V X)
= (£V η)(X) + η([V, X]) − ∇V g(V, X) − η(∇V X)
= (£V η)(X) + η(∇V X) − η(∇X V ) + η(τ (V, X)) − ∇V g(V, X) − η(∇V X)
1
= (£V η)(X) + (∇X g)(V, V ) − (∇V g)(V, X) + η(τ (V, X)),
2
where τ denotes the torsion tensor of ∇. In particular, choosing ∇ to be the Levi-
Civita connection proves the result. ¤

Now we have the following result due to Carrière.


Proposition 2.6.9: A Riemannian flow whose orbits are geodesics is isometric.
Conversely, the orbits of a one dimensional isometric foliation are geodesics.

Proof. Let F be a Riemannian flow generated by the nowhere vanishing vector


field V whose orbits are geodesics, and let g be a bundle-like metric such that V
has unit length. Consider Equation 2.6.1 for horizontal vector fields X, Y. Since
this equation is tensorial we can take X, Y to be foliate as well. So the last two
terms on the right hand side vanish as does the left hand side since g is bundle-like.
Thus, (£V g)(X, Y ) = 0 for X, Y horizontal. On the other hand if η is the1-form
dual to V , then by (iii) of Lemma 2.6.8 we have for X horizontal
0 = dη(V, X) = V η(X) − Xη(V ) − η([V, X]) = −η([V, X]),
which implies that (£V g)(V, X) = 0. Also (£V g)(V, V ) = 0 since V has constant
length. Thus, V is a Killing vector field and the flow is isometric.
For the converse let F be a one dimensional isometric foliation. Then F is
generated by a nowhere vanishing vector field V, and there is Riemannian metric
g on M such that V is a Killing vector field. Let η be the 1-form dual to V with
respect to g. Then condition (ii) of Lemma 2.6.8 is satisfied. Thus, the orbits of V
are geodesics. ¤

The following is an immediate consequence of the proof of Theorem 2.6.4.


Proposition 2.6.10: Let (M, F) be an isometric Riemannian flow with toral rank
rk(M, F) = k. Then the isometry group I(M, g) contains a torus subgroup Tk of
dimension k. Moreover, the closure of any leaf is a subtorus Tj ⊂ Tk of dimension
1 ≤ j ≤ k.
Molino (See [Mol88] and references therein) has studied the leaf closures for
general Riemannian foliations F and has shown the existence of a sheaf of local
transverse Killing vector fields C(M, F) that he calls the commuting sheaf of the
foliation. Proposition 2.6.10 says that for isometric Riemannian flows the sheaf
C(M, F) admits precisely k global sections. In fact on a compact manifold M a
result of Molino and Sergiescu [MS85] says that the commuting sheaf admits a
global trivialization if and only if F is isometric. Furthermore, Proposition 5.5
of [Mol88] says that for a Riemannian foliation on a compact simply connected
manifold, C(M, F) admits a global trivialization. Thus,
Proposition 2.6.11: On a simply connected compact manifold every Riemannian
flow is isometric.
2.6. RIEMANNIAN FLOWS 73

We are now ready for a Theorem of Wadsley [Wad75] that will be of great
importance to us in the sequel.
Theorem 2.6.12: Let F be a one dimensional foliation on M such that all the
leaves of F are circles. Then there is a smooth action A : S 1 × M −−→M whose
orbits are precisely the leaves of F if and only if there exists a Riemannian metric
on M such that the leaves of F are geodesics.
Next we discuss an obstruction to the existence of Riemannian flows noticed by
Carrière. This comes from a seminal paper of Gromov [Gro81] in which he defines
two important invariants of smooth manifolds. The first, called the minimal volume
is defined by

(2.6.2) MinVol(M ) = inf Vol(M, g),


|K(g)|≤1

where Vol(M, g) denotes the volume of M with respect to the Riemannian metric g,
and the infimum is taken over all complete metrics whose sectional curvatures K(g)
are bounded in absolute value by 1. One easily sees that, for example, for compact
oriented 2-manifolds M 2 one has MinVol(M 2 ) = 2π|χ(M 2 )|, where χ denotes the
Euler characteristic. On the other hand Gromov noticed that if a smooth manifold
M admits a locally free S 1 action, then by scaling the metrics along the orbits of
the S 1 , one sees that MinVol(M ) = 0, and Carrière [Car84b] noticed that this
argument generalizes to arbitrary Riemannian flows.
The second invariant of Gromov is intuitively less geometrical, but computa-
tionally more useful. It is the so-called simplicial volume of M defined3 P as follows:
consider the singular chain complex C∗ of all finite combinations c = i ri σi of
singular simplices
P σi in M with real coefficients ri . Define the simplicial norm of
c by ||c|| = i |ri |. This gives a pseudo-norm on the singular homology H∗ (M, R)
by ||[z]|| = inf z ||[z]||, where the infimum is taken over all singular homology cycles
representing [z]. If M is an orientable manifold, its simplicial volume ||M || (also
known as the Gromov invariant of M ) is defined to be the simplicial norm of the
fundamental class of M. If M is not orientable, then ||M || = 21 ||M̃ ||, where M̃ is the
2-fold cover of M. The simplicial volume enjoys the following important properties:
(i) If f : M −→M 0 is a proper map of degree d, then ||M || ≥ d||M 0 ||, and if
f is a d-sheeted covering map, ||M || = d||M 0 ||.
(ii) If M1 is compact and M2 is arbitrary, then there is a positive constant
C depending only on the dimension of M1 × M2 such that
C||M1 || · ||M2 || ≥ ||M1 × M2 || ≥ C −1 ||M1 || · ||M2 ||.
(iii) If dim M1 = dim M1 ≥ 3, then ||M1 #M2 || = ||M1 || + ||M2 ||.
Furthermore, Gromov proves the following estimate for the simplicial and minimal
volumes
(2.6.3) ||M || ≤ (n − 1)n n!MinVol(M ).
We mention that if M is closed (compact without boundary), then ||M || is com-
pletely determined by the fundamental group. In particular, compact simply con-
nected manifolds have vanishing Gromov invariant. (More generally ||M || vanishes
for any compact manifold whose fundamental group is amenable).

3Here M can be any topological space.


74 2. FOLIATIONS

Summarizing we have the following theorem of Carrière.


Theorem 2.6.13: If a compact manifold M admits a Riemannian flow, then
MinVol(M ) = 0. In particular, ||M || = 0 and all the Pontrjagin numbers of M
vanish.
We are interested in seeing which manifolds M have non-vanishing Gromov in-
variant ||M ||. In [Gro81] Gromov proves a theorem which he attributes to Thurston
which says that a complete Riemannian manifold of finite volume whose sectional
curvatures are pinched between −k and −1 for some k > 1 has non-vanishing
Gromov invariant ||M ||. At about the same time Inoue and Yano [IY82] proved
that any compact Riemannian manifold whose sectional curvatures is bounded from
above by −δ for δ > 0 has non-vanishing Gromov invariant. We state the precise
results as:
Theorem 2.6.14: Let (M, g) be a n-dimensional Riemannian manifold.
(i) Suppose that (M, g) is complete with finite volume 0 < Vol(M, g) < ∞,
and that the sectional curvatures of g are pinched between −k and −1 for
some real number k > 1. Then the estimate
Vol(M, g) ≤ Cn ||M ||
holds, where Cn is a positive constant depending only on n.
(ii) Let (M, g) be compact without boundary whose sectional curvatures are
bounded from above by −δ for some δ > 0. Then there exists a positive
constant Cn such that
||M || ≥ Cn δ n/2 Vol(M, g).
Hence, in both cases ||M || 6= 0.
It follows immediately that no manifold satisfying either of the two conditions
of this theorem can admit a Riemannian flow.
CHAPTER 3

Kähler Manifolds

Kähler metrics1 were introduced and studied by Erich Kähler in 1933 [Käh33]
(recently reprinted in [Käh03]). But it was not until late the 40ties when the
importance of Kähler manifolds in both Riemannian and algebraic geometry was
finally realized. This was largely due to the fundamental work of Chern. Already
in 1946 Chern introduces the notion of his Chern classes for Hermitian manifolds
[Che46]. As a special case, it follows that the cohomology class of the Ricci 2-form
ρω on a Kähler manifold (M, J, g, ωg ) does not depend on the metric g but only on
the complex structure and it is a fixed multiple of the first Chern class of M
[ρg ] = 2πc1 (M ) ,
(cf. Proposition 3.6.1 of the last section of this chapter). All of a sudden Kähler
metrics dramatically gained in their importance. Throughout the 50ties they were
studied by such giants as Borel, Chern, Hodge, Kobayashi, Kodaira, Lichnerowicz,
Spencer, among many others, and culminating in Weil’s well known book [Wei58].
However, it was not until 10 years after the Chern’s article when Calabi grasped
the real significance of this remarkable relation and what it means in the context
of Kähler-Einstein metrics [Cal56, Cal57]. Calabi’s work directly or indirectly re-
sulted in several mesmerizing conjectures which captivated the mathematical world
for years to come and stimulated incredible interest in Kähler geometry that has
remained until today2. The relation
λ[ωg ] = [ρg ] = 2πc1 (M )
will be the focal point of Chapter 3 and even more so in Chapter 5. Since Kähler
geometry and algebraic geometry from a differential viewpoint play a crucial rôle
in this book, we take some care in setting things up in spite of the fact that there
are several comprehensive treatments of this subject in a book form. We mention
here the following books where excellent presentations of Kähler geometry can be
found: [KN69, LB70, GH78b, Wel80, Zhe00, Voi02, Bal06] to name a few.
We thus recall the basic properties of Kähler manifolds and Kähler metrics for the
most part without proofs. We focus on the curvature properties, Chern classes and
end with a discussion of the famous Calabi Conjecture. In Chapter 4 we describe

1Schouten considered Kähler manifolds 4 years earlier in the article Über unitäre Geometrie
but never got much credit for his work [Sch29, SvD30]. Curiously, neither Schouten nor Kähler
appeared very interested in their own inventions. The term Kähler manifolds became standard
after the WWII in the late 40ties. Interestingly, Schouten and von Dantzig arrive at the Kähler
condition while investigating parallel transport of the Levi-Civita connection associated to the a
Hermitian metric.
2We refer the reader interested in the historical development of Kähler geometry from its
birth to the volume devoted to the mathematical works of Kähler [Käh03] and to Jean-Pierre
Bourguignon’s excellent article in that volume.

75
76 3. KÄHLER MANIFOLDS

how Kähler and algebraic geometry generalize to the category of Kähler orbifolds.
Finally, in Chapter 5 we focus primarily on Kähler-Einstein metrics of positive
scalar curvature.

3.1. Complex Manifolds and Kähler Metrics


Let us start with a definition of a complex manifold. There are several fre-
quently used approaches to define such a structure. Perhaps the most natural one
is the one that imitates the usual definition of a smooth structure on a real manifold.
Definition 3.1.1: Let M be a real manifold of dimension 2n. A complex chart
on M is a pair (U, φ) such that U ⊂ M is open and φ : U → Cn is diffeomorphism
between U and an open set φ(U ) ⊂ Cn . We say that M is a complex manifold if
it admits and atlas A = {Uα , φα }α∈Ω of complex charts whose transition functions
φβ ◦ φ−1
α : φα (Uα ∩ Uβ )−−−→φβ (Uα ∩ Uβ )

are all biholomorphic.


Recall from Example 1.4.9 that if M is a real manifold than a smooth section J
of the bundle of endomorphisms End(T M ) such that J 2 = −1l is called an almost
complex structure. Note that we must have dim(M ) = 2n. We can extend J to
act on the complexified tangent bundle T M ⊗R C by C-linearity. Then J induces
a splitting T M ⊗R C = T 1,0 M ⊕ T 0,1 M, where T 1,0 M and T 0,1 M are eigenspaces
with eigenvalues ±i, respectively. Note that T M is naturally isomorphic to T 1,0 M
by the map X 7→ 21 (X − iJX).
An almost complex structure is said to be integrable if M admits an atlas of
complex charts with holomorphic transition functions such that J corresponds to
the induced complex multiplication on T M ⊗R C. Hence, a manifold with an inte-
grable almost complex structure is (by definition) a complex manifold. Conversely,
given a complex manifold M we can define J by
∂ ∂
J =i ,
∂zj ∂zj
where (z1 , . . . , zn ) is a local holomorphic coordinate system.
The theorem of Newlander and Nirenberg [NN57] asserts that integrability of
J is equivalent of the vanishing of the Nijenhuis tensor Equation 1.4.4 In view of
this we have the following
Theorem/Definition 3.1.2: Let M be a real manifold and J an almost complex
structure on M . The Nijenhuis tensor NJ ≡ 0 on M if and only if J is integrable
in which case we call J a complex structure and the pair (M, J) a complex
manifold.
The second part of this statement is very often taken as an alternative definition
of a complex manifold. As we will see it is perhaps the most geometric and indeed
we will use it most often. A third way of describing a complex manifold as a real
manifold with torsion-free GL(n, C)-structure was described earlier in Example
1.4.9. For maps between complex manifolds we have
Definition 3.1.3: Let (M1 , J1 ), (M2 , J2 ) be two complex manifolds. A map f :
M1 → M2 is called holomorphic if J2 (df (v)) = df (J1 (v)) for all v ∈ Γ(T M ). If,
in addition, f −1 exists and is holomorphic than f is called a biholomorphism
between M1 and M2 .
3.1. COMPLEX MANIFOLDS AND KÄHLER METRICS 77

If M2 = M1 = M and J2 = J1 = J the set of biholomorphisms from M to itself


form a group Aut(M, J). Generally it is not a Lie group; however, it is a Lie group
when M is compact by Proposition 1.6.7. In fact if M is compact it is a complex
Lie group. Explicitly, [Kob72] we have
Proposition 3.1.4: Let M be a compact complex manifold. Then Aut(M, J) is
a complex Lie transformation group and its Lie algebra aut(M, J) consists of the
holomorphic vector fields on M.
For a much more extended discussion of the complex automorphism group on
Kähler manifolds we refer the reader to Chapter III of [Kob72].
One can extend J to act on the complexified cotangent bundle T ∗ M ⊗R C which
then splits as T ∗(1,0) M ⊕ T ∗(0,1) M. This in turn defines the splitting of the bundle
of complex p-forms on M
p
M p
M
Λp T ∗ M ⊗R C = Λj T ∗(1,0) M ⊕ Λp−j T ∗(0,1) M = Λj,p−j M ,
j=0 j=0

where the last equality defines the space of forms of type (j, p − j). Note that such
decomposition holds even if J is only almost complex.
Proposition 3.1.5: Let (M, J) be an almost complex manifold. The following
conditions are equivalent:
(i) [V, W ] ∈ Γ(T 1,0 M ) for all V, W ∈ Γ(T 1,0 M ),
(ii) [V, W ] ∈ Γ(T 0,1 M ) for all V, W ∈ Γ(T 0,1 M ),
(iii) Im(d) |Γ(Λ1,0 M ) ⊂ Γ(Λ2,0 M ) ⊕ Γ(Λ1,1 M ),
Im(d) |Γ(Λ0,1 M ) ⊂ Γ(Λ0,2 M ) ⊕ Γ(Λ1,1 M ),
(iv) Im(d) |Γ(Λp,q M ) ⊂ Γ(Λp+1,q M ) ⊕ Γ(Λp,q+1 M ),
(v) NJ ≡ 0.

Proof. We will only show the equivalence of (i), (iii), and (v) as the rest is
evident. To establish equivalence of (i) and (iii) we use

2dω(X, Y ) = X(ω(Y )) − Y (ω(X)) − ω([X, Y ]) ,

which, if ω is a 1-form of type (1, 0) and X, Y are vector fields of type (0, 1), clearly
vanishes. Hence, dω does not have any (0, 2)-part. To establish equivalence of (i)
and (v) we consider B = [X − iJX, Y − iJY ] for any two real vector fields X, Y on
M. It is easy to see that B + iJB = −N (X, Y ) − iJ(N (X, Y )).. But B + iJB = 0
if and only if B is a vector field of type (1, 0) if and only if NJ ≡ 0. ¤

Thus, if (M, J) is a complex manifold, composing the exterior derivative acting


on (k, l)-forms with projections on Γ(Λk+1,l M ) and Γ(Λk,l+1 M ), defines natural
splitting of d into two parts. They are denoted by ∂ and ∂, ¯ where d = ∂ + ∂.
¯ Since
d2 = 0 we get ∂ 2 = ∂¯2 = 0 and ∂ ∂¯ = −∂∂. ¯ We also denote by E k,l the sheaf of
germs of smooth sections of the bundles Λk,l M, and use the same notation ∂, ∂, ¯ etc.
for these operators acting on local sections of E k,l . We also let Ωp denote the sheaf
of germs of holomorphic sections of the bundle Λp,0 M. Then, by the well-known
¯
∂-Poincaré lemma [GR65], we have a resolution of this sheaf, namely
∂¯ ∂¯ ∂¯
0−−→Ωp −−→E p,0 −−→ E p,1 −−→ · · · −−→E p,n−1 −−→ E p,n −−→0 .
78 3. KÄHLER MANIFOLDS

Moreover, since the sheaves E k,l are fine sheaves, this resolution is acyclic. So if we
define the Dolbeault cohomology groups [Dol53] of (M, J)
¡ ¢
k,l Ker ∂¯ : Γ(Λk,l ) → Γ(Λk,l+1 )
H∂¯ (M ) = ¡ ¢
Im ∂¯ : Γ(Λk,l−1 ) → Γ(Λk,l )
to be the derived groups of the cochain complex on global sections (1.1.7), the Ab-
stract de Rham Theorem 1.1.16 gives the well-known Dolbeault Theorem [Dol53]
Theorem 3.1.6: Let M be a complex manifold. Then there is an isomorphism
H q (M, Ωp ) ≈ H∂p,q
¯ (M ) .

In Section 3.4 we shall give a more general version of this theorem by twisting
with a vector bundle.
Definition 3.1.7: Let (M, J) be an almost complex manifold and let g be a Rie-
mannian metric on M such that
g(JX, JY ) = g(X, Y ), X, Y ∈ Γ(T M ) .
Then g is called an Hermitian metric and (M, J, g) is called an almost Hermit-
ian manifold. Furthermore, given a Hermitian metric we define its fundamental
2-form or Kähler form ωg of g by
ωg (X, Y ) = g(X, JY ), X, Y ∈ Γ(T M ) .
The triple (J, g, ωg ) is called an almost Hermitian structure on M .
Note that an almost complex manifold always admits a Hermitian metric.
An almost Hermitian structure (J, g, ωg ) with an integrable J is called Hermit-
ian. Given an Hermitian metric g we can extend it to a unique Hermitian scalar
product h on the complexified tangent bundle T M ⊗R C satisfying
(i) h(Z̄, W̄ ) = h(Z, W ) for all Z, W ∈ Γ(T M ⊗R C),
(ii) h(Z, Z̄) > 0 fora all Z 6= 0 ∈ T M ⊗R C,
(iii) h(Z, W̄ ) = 0 for Z ∈ Γ(T 1,0 M ) and W ∈ Γ(T 0,1 M ).
Locally, in a complex coordinate chart (U ; z1 , . . . , zn ) we can write the metric h as
n
X
(3.1.1) h= hij̄ dzi ⊗ dz̄j ,
i,j=1

where hij̄ (p) is a Hermitian n×n matrix. Then the Hermitian form h can be written
in terms of the Riemannian metric g and fundamental 2-form3 as h = g − 2iωg and
ωg can be written as
n
i X
(3.1.2) ωg = h dzi ∧ dz̄j .
2 i,j=1 ij̄

Also, note that the volume form satisfies


µ ¶n
i
ωgn = n! det(hij̄ )dz1 ∧ dz̄1 ∧ · · · ∧ dzn ∧ dz̄n
2
and, hence, ωgn = n!dvolg , where dvolg is the Riemannian volume form of (M, g).

3The factor -2 multiplying ω is chosen so that the volume form satisfies ω n = n!dvol .
g g g
3.1. COMPLEX MANIFOLDS AND KÄHLER METRICS 79

Definition 3.1.8: A Hermitian manifold (M, J, g, ωg ) is said to be Kähler if ωg


is a closed 2-form. We call g a Kähler metric, ωg its Kähler form, and the
triple (J, g, ωg ) a Kähler structure on M.
When we want to emphasize a Kähler form, we shall use the notation (M, ω)
to denote a Kähler manifold. A manifold with a Kähler metric is often called a
manifold of Kähler type.
From the previous discussion we see that a Kähler form is a real (1, 1)-form
which is both ∂-closed and ∂-closed. ¯ Hence [ωg ] defines a Dolbeault cohomology
1,1
class in H∂¯ (M ). Next we state several propositions and lemmas which describes
some basic local and global properties of Kähler manifolds.
Proposition 3.1.9: Let (M, J, g, ωg ) be an almost Hermitian manifold of real di-
mension 2n and let ωg be the fundamental 2-form associated to g. Let ∇ be the
Levi-Civita connection of g. Then the following conditions are equivalent:
(i) ∇J = 0,
(ii) ∇ωg = 0,
(iii) the holonomy group Hol(M, g) is contained in U (n),
(iv) (M, J, g, ωg ) is 1-integrable,
(v) (M, J, g, ωg ) is Kähler.
Yet another characterization of the Kähler condition is
Proposition 3.1.10: A Hermitian metric g on a complex manifold (M, J) is
Kähler if and only if for any point p ∈ M there exists a local holomorphic chart
(U ; z1 , . . . , zn ) such that hij̄ (p) = δij , dhij̄ (p) = 0.
Such a metric is said to osculate to order 2 to the Euclidean metric, and the
coordinate chart is said to be normal at p.
We consider some important example of complex Kähler manifolds [KN69].
Example 3.1.11: Let M = R2n and let (x1 , . . . , xn , y1 , . . . , yn ) be the global coor-
dinate chart on M. We can define
∂ ∂ ∂ ∂
J = , J =− .
∂xj ∂yj ∂yj ∂xj
Clearly, J 2 = −1l and NJ ≡ 0 as J is a constant matrix on M . We can define
z1 , . . . , zn by zj = xj + iyj for j = 1, . . . , n. We have
∂ ∂ ∂ ∂
J =i , J = −i
∂zj ∂zj ∂ z̄j ∂ z̄j
so that (z1 , . . . , zn ) ∈ Cn is a (global) holomorphic chart compatible with J. Now let
us write the Hermitian metric h and the fundamental 2-form in complex coordinates:
Let (z1 , . . . , zn ) be the global holomorphic chart with zi = xi + iyi . Since dzi =
dxi + idyi , dz̄i = dxi − idyi i, i = 1, . . . , n and the Hermitian metric h reads
Xn n
X
h= hij̄ dz i ⊗ dz̄ j = dzi ⊗ dz̄i .
i,j=1 i=1

As the fundamental 2-form ωg is simply minus the imaginary part of h we can write
n n
i X iX
ωg = hij̄ dzi ∧ dz̄j = dzi ∧ dz̄i
2 i,j=1 2 i=1

which is clearly closed.


80 3. KÄHLER MANIFOLDS

The fiducial example of a compact complex manifold is given by our next


example.
Example 3.1.12: Complex Projective Space CPn is defined as the set of lines
through the origin in Cn+1 . Let z = (z0 , . . . , zn ) be a point (vector) in Cn+1 \ {0}.
We say that two non-zero vectors z and z0 are equivalent if there is a λ ∈ C∗
such that z0 = λz. Then CPn is the quotient space (Cn+1 \ {0})/∼ . We let
π : Cn+1 \ {0}−−→CPn denote the natural projection. Local coordinate charts
are defined as follows. Let Ũi be the open subspace of Cn+1 such that zi 6= 0. Then
CPn is covered by open sets {Ui = π(Ũi )}ni=0 , together with homeomorphisms
φi : Ui −−→Cn = R2n defined by
µ ¶
z0 zi−1 zi+1 zn
φi ([z0 , . . . , zn ]) = ,..., , ,..., .
zi zi zi zi
The coordinates on the right hand side of this equation are called affine coordinates,
d
and it is convenient to set zi = 1 and write |z|2 = |z1 |2 + · · · + |z 2 2
i | + · · · + |zn | ,
where as usual the hat means remove that element. Then the Fubini-Study metric
in Ui ⊂ CPn is defined by
∂2
gkj̄ = ln(1 + |z|2 ) ,
∂zk ∂ z̄j
and the Kähler form is
ÃP P P !
i i dzk ∧ dz̄k ( k z̄k dzk ) ∧ ( j zj dz̄j )
(3.1.3) ωg = ∂ ∂¯ ln(1 + |z|2 ) = k
− .
2 2 1 + |z|2 (1 + |z|2 )2
Clearly, ωg is closed and one can check that both the Riemannian metric g and
Kähler form ωg are globally defined on CPn .
Notice that by restricting the map π to the unit sphere in Cn+1 gives the well-
known Hopf fibration S 1 −−→S 2n+1 −−→CPn which as should become apparent will
be considered as the fiducial example for this book.
This example has an interesting generalization, namely:
Example 3.1.13: Let Grp (Cq+p ) be the set of p-dimensional complex subspaces of
the complex vector space Cq+p . We will define complex structure on the complex
Grassmannian Grp (Cq+p ) by construction an atlas of holomorphic charts. Let
(z1 , . . . , zp+q ) be the natural coordinates on Cq+p . We can think of zα as the complex
linear mapping zα : Cq+p → C. Now, consider the partition of the set {1, . . . , p +
q} = α ∪ αc = {α1 , . . . , αp } ∪ {αp+1 , . . . , αp }, where both sets are ordered in the
increasing order. Let let Uα ⊂ Grp (Cq+p ) be the subset of p-planes W for which
zα1 |W , . . . , zαp |W are linearly independent. Since, for each W ∈ Uα , the mappings
zα1 |W , . . . , zαp |W form a basis for the dual space of W , we can write
p
X
zαp+k |W = wkj (zαj |W ) , k = 1, . . . , q.
j=1

The coefficients wkj uniquely define a p × q complex matrix which maybe thought
of as an element of Cpq . Setting
ψα (W ) = (wkj )
for each partition α defines maps ψα : Uα → Cpq . It is a simple exercise to
¡ check
¢ that
these maps are injective and onto. Hence, the family S = {Uα ; ψα } of p+1 p charts
3.1. COMPLEX MANIFOLDS AND KÄHLER METRICS 81

form a holomorphic atlas on Grp (Cq+p ). There is a generalization of Equation 3.1.3


which shows that Grp (Cq+p ) is Kähler [KN69]. When p = 1, q = n we recover
the standard atlas of the (n + 1) holomorphic charts on complex projective space
described in Example 3.1.12. In particular, we have Gr1 (Cn+1 ) = CPn .
One can describe further interesting generalizations by considering the set of
all nested subspaces Ck1 ⊂ Ck2 ⊂ · · · ⊂ Ckl ⊂ Cn with 0 < k1 < k2 < · · · < kl < n.
This gives rise to the so-called flag manifolds Flk1 ,...,kl (Cn ), cf. [Akh90, Akh95].
In the case l = 1 we recover the Grassmannian. Even further generalizations are
obtained by considering the generalized flag manifolds G/P, where G is a complex
semi-simple Lie group and P is a parabolic subgroup (the ordinary flag manifolds
correspond to taking G = SL(n, C)). These are all homogeneous Kähler manifolds.
In fact, a well-known theorem of Borel and Remmert [BR62] (See also [Akh95])
says that any compact homogeneous Kähler manifold M is of the form (G/P ) ×
A(M ), where A(M ) is a complex torus known as the Albanese of M.
¯ = 1 (∂ + ∂)(
Now, let f ∈ C ∞ (M ). As the i∂ ∂f ¯ ∂¯ − ∂)f is real we conclude that
2
¯
i∂ ∂f is a real closed 2-form of type (1, 1). We have the following two lemmas
¯
Lemma 3.1.14: [local i∂ ∂-lemma] Let ρ ∈ Γ(Λ1,1 U ) be a smooth, closed, real
(1, 1)-form on a unit disc U = DCn (1) ⊂ Cn . Then there exists f ∈ C ∞ (U ) such
¯ .
that ρ = i∂ ∂f
¯
Lemma 3.1.15: [global i∂ ∂-lemma] Let (M, J) be a compact complex manifold
and let γ be a real (1, 1)-form on M satisfying γ = dα for some real 1-form α. Then
there exists a smooth real function f such that γ = i∂ ∂f. ¯ In particular, if g1 , g2 are
two different Kähler metrics on M such that [ωg1 ] = [ωg2 ] ∈ H 2 (M, R). Then
¯ ,
ωg2 = ωg1 + i∂ ∂f
and f is unique up to a constant.
Lemma 3.1.14 says that on a Kähler manifold locally one can always find a
function which “generates” the Kähler form (and, hence, the Kähler metric) via
the simple formula ωg = i∂ ∂f.¯ Such a function is called a Kähler potential. How-
ever, if M is compact a Kähler potential cannot exists globally. This is a simple
consequence of the following
Lemma 3.1.16: Let (M, J, g, ωg ) be a compact Kähler manifold with the Kähler
form ωg . Then [ωgk ] ∈ H 2k (M, R) is non-trivial for all k = 0, . . . , n.

Proof. This is a simple consequence of the Stokes’ theorem and the fact that
ωgn = n!dvolg . For let ωgk = dα be exact. Then
Z Z Z Z
0 < n!vol(M ) = ωgn = ωgk ∧ ωgn−k = dα ∧ ωgn−k = d(α ∧ ωgn−k ) = 0 ,
M M M M

so we get a contradiction. In particular, when k = 1 the lemma implies that there


¯
is no globally defined function on M such that ωg = i∂ ∂f. ¤

Let (M, J) be a complex manifold and let N be a submanifold of M. Then, in


particular at each point p ∈ N , the tangent space Tp N ⊂ Tp M is a vector subspace.
We say that N is a complex submanifold of M if Tp N is a complex subspace for
each p ∈ N, i.e., J(Tp N ) = Tp N. Now, if N is a complex submanifold then the
restriction JN of J to T N is a complex structure on N and the inclusion map
82 3. KÄHLER MANIFOLDS

ι : N ,→ M has the property


J(dp ι(w)) = (dp ι)(JN (w)) , ∀w ∈ Tp N ,
that is ι is a holomorphic map between two complex manifolds (N, JN ) and (M, J).
Since we can pullback Kähler forms, we have a simple but important result.
Proposition 3.1.17: A complex submanifold of a Kähler manifold is Kähler.
Example 3.1.18: Consider the previous example with M = CPn and let the inclu-
sion map ι : N ,→ CPn be defined by setting
N = Nf = {m ∈ CPn |f (m) = 0} ,
where f = (f1 , . . . , fk ) and fk ∈ C[z0 , . . . , zn ] are complex homogeneous polynomials
in (n + 1) variables for each k = 0, 1, . . . , n. Any such set is called a projective
algebraic variety. If, in addition NJ is smooth then it is a complex submanifold in
the way described above; hence, it is a smooth projective algebraic variety which
by Proposition 3.1.17 is always Kähler. For instance, if k = 1 then N is called a
hypersurface. For example, let f = f, where f is
f ([z0 , . . . , zn ]) = z0p + · · · + zn+1
p
.
Then Nf is called a Fermat hypersurface. When p = 2 we get a complex quadric.
The concept of variety can vary somewhat depending on the context. We now
formalize this. An important point is that varieties allow for a certain type of
singular behavior.
Definition 3.1.19: We have
(i) An affine algebraic variety is the common zero locus of a collection of
polynomials in Cn .
(ii) A projective algebraic variety is the common zero locus in CPn of a
collection of homogeneous polynomials in Cn+1 .
(iii) An analytic (sub)variety V of a complex manifold M is a closed subset
that can written locally as the common zero locus of a finite collection of
local holomorphic functions in M . In particular, A closed subset N ⊂ M
of a complex manifold M is said to be a hypersurface if every point
p ∈ N has an open neighborhood U and a nonzero holomorphic function
f : U −−→C such that N ∩ U = {p ∈ U | f (u) = 0}.
By a variety we mean any of these three, and by an algebraic variety either of
the first two. When working with algebraic varieties it is common to use the Zariski
topology which is defined by the condition that its closed subsets are precisely the
common zero loci of polynomials. So by a subvariety of an algebraic variety V we
shall mean any Zariski closed subset of V.
Definition 3.1.20: A variety V is said to be irreducible if it cannot be written
as the union of two subvarieties V1 and V2 with Vi 6= V.

3.2. Curvature of Kähler Manifolds


Let (M, J, g, ωg ) be a Kähler manifold and ∇ the Levi-Civita connection. We
can extend ∇ in a C-linear way to Γ(T M ⊗R C). In a local chart (U ; z1 , . . . , zn ) we
have { ∂z∂ 1 , . . . , ∂z∂n } and { ∂∂z̄1 , . . . , ∂∂z̄n } as bases for T 1,0 M and T 0,1 M, respectively.
3.2. CURVATURE OF KÄHLER MANIFOLDS 83

Define the Christoffel symbols ΓC


AB as follows
n
X
∂ ∂
∇∂/∂zA = ΓC
AB , A, B, C ∈ {1, . . . , n, 1̄, . . . , n̄} .
∂zB ∂zC
C=1

It follows by C-linearity that



ΓC
AB = ΓĀB̄ , ΓC C
AB = ΓBA .

Lemma 3.2.1: The only non-vanishing Christoffel symbols are Γkij and Γk̄īj̄ , where
X ∂gj k̄
Γnij = g nk̄ .
∂zi

Proof. We have
∂ X ∂ X ∂
∇∂/∂zi = Γkij + Γk̄ij ,
∂zj ∂zk ∂ z̄k
k k̄

∂ X ∂ X ∂
∇∂/∂zi = Γkij̄ + Γk̄ij̄ .
∂ z̄j ∂zk ∂ z̄k
k k̄

Since, ∇J = 0 and J acts by multiplication by i on the basis of T 1,0 M and −i on


the basis of T 0,1 M , we get
µ ¶
∂ ∂
∇∂/∂zi J = J∇∂/∂zi .
∂zj ∂zj
By comparing two sides of this equation we get
µX X ¶ µX X ¶
∂ ∂ ∂ ∂
i Γkij + Γk̄ij =i Γkij − Γk̄ij ,
∂zk ∂ z̄k ∂zk ∂ z̄k
k k̄ k k̄

which means that Γk̄ij = 0. Similar argument applied to the second equation yields
Γk̄ij̄ = 0, while Γk̄ij = 0 follows by the symmetry two lower indices. Hence, the
only non-zero Christoffel symbols are Γkij and their complex conjugate Γk̄īj̄ . Since
X ·g(Y, Z) = g(∇X Y, Z) + g(Y, ∇X Z) we obtain
µ ¶ µX ¶ X
∂ ∂ ∂ ∂ l ∂ ∂
g = g , =g Γij , = Γlij glk̄ .
∂zi j k̄ ∂zi ∂zj ∂ z̄k ∂zl ∂ z̄k
l l

Contracting both sides of this equation with the inverse of the metric gives the
formula. ¤

In a local chart (U ; z1 , . . . , zn ) we can define the following matrix-valued (1, 0)-


form
X ∂g
θ = g −1 ∂g = g −1 dzi ,
i=1
∂zi
where g = (gij̄ ). The above lemma implies that
Proposition 3.2.2: Let (M, J, g, ωg ) be a Kähler manifold and let ∇ be the Levi-
Civita connection of g extended to T M ⊗R C ' T 1,0 M ⊕ T 0,1 M. Accordingly, ∇
decomposes ∇ = ∇1,0 + ∇0,1 , θ is the connection 1-form of ∇1,0 .
84 3. KÄHLER MANIFOLDS

Note that we have the following simple expression


X j X ∂gj k̄ ∂
Γij = g j k̄ = ln G ,
j
∂z i ∂z i
j,k̄

where G = det(g) = det(gij̄ ). The determinant G ∈ C ∞ (U) is a smooth real


function defined locally on U. The last equality follows from the well-known formula
det(A) = exp{Tr ln A} as we have
µ ¶ X
∂ ∂g ∂gkn̄
ln G = Tr g −1 = g n̄k .
∂zi ∂zi ∂zi
n̄,k

These simple formulas for the Christoffel symbols simplify the expression for the
full Riemann curvature tensor. Recall that the Riemannian curvature tensor R of
the metric g one M is defined by R(X, Y ) : Γ(T M )−→Γ(T M ), where
R(X, Y )Z = ∇X ∇Y Z − ∇Y ∇X Z − ∇[X,Y ] Z , X, Y, Z ∈ Γ(T M ) .
Point-wise R(X, Y ) is a skew-endomorphism of Tp M at any p ∈ U . Since also
R(X, Y ) = −R(Y, X) one can think of R as a 2-from on M with values in skew-
endomorphisms of T M. It is sometimes convenient to think of R as the curvature
operator, that is a sections of ¯2 Λ2 T M defined by
R(X ∧ Y, W ∧ Z)) = g(R(X, Y )Z, W ) , X, Y, Z, W ∈ Γ(T M ) .
Alternatively, one can introduce R as the section of ⊗4 (T ∗ M ), i.e., the four-linear
map R : Γ(T M ) × Γ(T M ) × Γ(T M ) × Γ(T M )−→C ∞ (M ), defined by
R(X, Y, V, Z) = g(R(X, Y )Z, V ) , X, Y, V, Z ∈ Γ(T M ) .
All these different pictures are useful.
If (M, J, g) is complex than one can extend all these curvature tensors by C-
linearity to T M ⊗R C. In addition, on a Kähler manifold we have ∇J = 0 so that
R(V, W )JZ = JR(V, W )Z , V, W, Z ∈ Γ(T M ⊗R C) .
The above property has many important consequences. Since g is Hermitian we
get R(U, V, JW, JZ) = R(U, V, W, Z). Hence, R(U, V, W, Z) = 0 unless W, Z are
of different type. In particular, in the Kähler case one can define the following
curvature tensors:
Definition 3.2.3: Let (M, J, g, ωg ) be a Kähler manifold and let R be the curvature
tensor of (M, g). Extending R to T M ⊗R C we define the following:
(i) Riemann curvature tensor R as section of ⊗4 (T ∗1,0 M ), i.e., the map
R : Γ(T 1,0 M ) × Γ(T 0,1 M ) × Γ(T 1,0 M ) × Γ(T 0,1 M )−→C ∞ (M, C), given
by
(X, Y, Z, W ) 7→ R(X, Ȳ , Z, W̄ ) ,
(ii) Riemann curvature tensor RX,Ȳ as a section of End(T 1,0 M ), i.e.,
the map RX,Ȳ : (Γ(T 1,0 M ))−→Γ(T 0,1 M ), X, Y ∈ Γ(T 1,0 M ) given by
RX,Ȳ (Z) = R(X, Ȳ )Z̄ ,
(iii) A real (1, 1)-form Ω with values in Γ(End(T 1,0 M )), i.e., a skew-Hermitian
map Ω : Γ(T 1,0 M ) × Γ(T 0,1 M )−→Γ(End(T 1,0 M ))
Ω(X, Ȳ ) = RX,Ȳ , X, Y ∈ Γ(T 1,0 M ) ,
called the Kähler-Riemann curvature form.
3.2. CURVATURE OF KÄHLER MANIFOLDS 85

(iv) Riemann curvature operator R as a section of ¯2 Γ(Λ1,1 M ), i.e., the


map R : Γ(Λ1,1 M ) × Γ(Λ1,1 M )−→C ∞ (M, C) given by
R(X ∧ Ȳ , V ∧ W̄ ) = R(X, Ȳ , Z, W̄ ) .
Proposition 3.2.4: Let (U ; z1 , . . . , zn ) be a local coordinate chart on a Kähler
manifold (M, J, g, ωg ). Let R ∈ Γ(⊗4 (T ∗1,0 M )) be the Riemann curvature tensor.
Then
∂ 2 gkl̄ X ∂gkr̄ ∂gpl̄
Rij̄kl̄ = − g pr̄ .
∂zi ∂ z̄j p,r̄,m ∂zi ∂ z̄j

Proof. It remains to do the local computations. We have


µ ¶ Ã µ ¶ !
∂ ∂ ∂ ∂ ∂ ∂ ∂ ∂
Rij̄kl̄ = R , , , =g R , , .
∂zi ∂ z̄j ∂zk ∂ z̄l ∂zi ∂ z̄j ∂ z̄l ∂zk

Since
µ ¶ µX ¶
∂ ∂ ∂ ∂ ∂ ∂
R , = ∇∂/∂zi ∇∂/∂ z̄j − ∇∂/∂ z̄j ∇∂/∂zi = ∇∂/∂zi Γn̄j̄ l̄ .
∂zi ∂ z̄j ∂ z̄l ∂ z̄l ∂ z̄l n̄
∂ z̄n

Hence
à µX ¶ ! à !
∂ ∂ ∂ X
n̄ ∂ ∂
Rij̄kl̄ = g ∇∂/∂zi Γn̄j̄ l̄ , = g Γj̄ l̄ , −

∂ z̄n ∂zk ∂zi n̄
∂ z̄n ∂zk

³X ∂ X m ∂ ´ ∂ ³X n̄ ´ X
g Γn̄j̄ l̄ , Γik = Γj̄ l̄ gn̄k − Γn̄j̄ l̄ Γm
ik gn̄m =

∂ z̄n p ∂zm ∂zi n̄ n̄,p

∂ ³X n̄p ∂gl̄p ´ X ∂gl̄p mr̄ ∂gkr̄ ∂ 2 gkl̄ X ∂gkr̄ ∂gpl̄


= g gn̄k − g n̄p g gn̄m = − g pr̄ .
∂zi n̄,p ∂ z̄j n̄,p,r̄,m
∂ z̄j ∂zi ∂zi ∂ z̄j p,r̄,m ∂zi ∂ z̄j

Definition 3.2.5: Let RX,Ȳ : Γ(T 1,0 M )−→Γ(T 1,0 M ), X, Y ∈ Γ(T 1,0 M ) be the
Riemann curvature tensor on a Kähler manifold (M, J, g, ωg ). The Ricci curva-
ture tensor is the map Ricω : Γ(T 1,0 M ) ⊗ Γ(T 1,0 )−→C ∞ (M, C) defined by the
trace
Ricω (X, Ȳ ) = Tr(RX,Ȳ ) .
A priori it is not clear that Ricω (X, Ȳ ) is the Ricci curvature tensor in the usual
sense. However, it is clear that Ric(X, Ȳ ) defines a Hermitian symmetric form on
Tp1,0 M . For with respect to any unitary basis {e1 , . . . , en } of Tp1,0 M we have
n
X
Ricω (X, Ȳ ) = R(X, Ȳ , ei , ēi ) ,
i=1

so that
n
X n
X
Ricω (X, Ȳ ) = R(X̄, Y, ēi , ei ) = R(Y, X̄, ei , ēi ) = Ricω (Y, X̄) .
i=1 i=1
86 3. KÄHLER MANIFOLDS

Proposition 3.2.6: The Ricci curvature tensor Ricω : T 1,0 M ⊗ T 1,0 M → C is a


Hermitian symmetric form on Tp1,0 M at each p ∈ M . In a local chart (U ; z1 , . . . , zn )
this form can be written as
Xn
Ricω = Rij̄ dzi ⊗ dz̄j ,
i,j̄=1

where
∂2
Rij̄ = − (ln G) .
∂zi ∂ z̄j
The real part of this Hermitian symmetric form is the Ricci curvature tensor of the
metric g while the imaginary part is a real (1, 1)-form on M called the Ricci form
and denoted by ρω ≡ ρg . With respect to the local chart (U ; z1 , . . . , zn ) we have
iX ¯ G) .
ρω = − Rij̄ dzi ∧ dz̄j = −∂ ∂(ln
2
i,j̄

Proof. By definition we have


X ∂2
Rij̄ = − g kl̄ Rij̄kl̄ = − (ln G) .
∂zi ∂ z̄j
kl̄

A priori it is not obvious that Ric(X, Ȳ ) coincides with the usual definition of the
Ricci curvature. A simple calculation shows that this is indeed the case. ¤
Note that the Kähler Riemann curvature 2-form Ω = dθ + θ ∧ θ is simple the
curvature 2-form of the ∇1,0 part of the Levi-Civita connection. In local coordinates
we can write X
Ωlk = Rij̄kp̄ g lp̄ dzi ∧ dz̄j
i,j̄,p̄
and, hence, the Ricci form ρω is a closed 2-form which is an invariant under the
complex linear group GL(n, C), namely,
(3.2.1) ρω = Tr(Ω) .
There are several more curvatures typically considered in the context of complex
and Kähler manifolds. The usual notion of sectional curvature is one of them.
Recall that for any X, Y ∈ Γ(T M ) we define the sectional curvature of the 2-plane
σ ⊂ Tp M spanned at point p ∈ M by Xp and Yp as
R(X, Y, Y, X)
K(σ, p) = K(X, Y ) = ,
|X ∧ Y |2
where |X ∧ Y |2 = |X|2 |Y |2 − [g(X, Y )]2 = g(X, X)g(Y, Y ) − [g(X, Y )]2 is the area
square of the parallelogram spanned by X, Y. Now, if (M, g) is complete, simply
connected and of constant sectional curvature then (depending on the sign) (M, g) is
isometric to S n , Rn , or the real hyperbolic ball BRn (1). Now, in complex dimension
greater than 1 the only Kähler manifolds of constant sectional curvature are flat.
To obtain a Kähler analogue of the real space forms we introduce the following
Definition 3.2.7: Let (M, J, g, ωg ) be a Kähler manifold and let X, Y ∈ Γ(T 1,0 M ).
Then
R(X, X̄, Y, Ȳ )
B(X, Y ) =
g(X, X̄)g(Y, Ȳ )
3.2. CURVATURE OF KÄHLER MANIFOLDS 87

is called the bisectional curvature of g in the direction of X, Y while


R(X, X̄, X, X̄)
H(X) = B(X, X) =
[g(X, X̄)]2
is called the holomorphic sectional curvature of g in the direction of X.
If we write X = √12 (U + iJU ), Y = √12 (V + iJV ) for some real vector fields
U, V ∈ Γ(T M ) we have
|JU ∧ V |2 |U ∧ V |2
B(X, Y ) = K(JU, V ) + K(U, V ) ,
|U |2 |V |2 |U |2 |V |2
and H(X) = K(U, JU ), where K the sectional curvature of g. In particular, if g
is of positive (negative, non-positive, nonnegative) sectional curvature then g is of
positive (negative, non-positive, nonnegative) bisectional curvature. Clearly, the
holomorphic sectional curvature is the curvature of all J-invariant planes. It is not
hard to show that on a Kähler manifold the holomorphic sectional curvature H
determines the Riemann curvature tensor completely.
Example 3.2.8: This is a continuation of the discussion of complex projective space
as described in Example 3.1.12 An easy computation gives
n!
ωgn = (i/2)2 dz1 ∧ dz̄1 ∧ · · · ∧ dzn ∧ dz̄n ,
(1 + |z|2 )n+1
which means that G = det(gij̄ ) = (1 + |z|2 )−(n+1) . Hence, The Ricci curvature
tensor
∂2
(3.2.2) Rij̄ = − ln G = (n + 1)gij̄ .
∂zi ∂ z̄j
Exercise 3.1: Show that the Fubini-Study metric is U (n + 1)-invariant and that
U (n + 1) acts on M transitively.
It now follows that in order to compute the full Riemann curvature tensor Rij̄kl̄
it is sufficient to compute it at one point [1, 0, . . . , 0]. Using local expressions for
Rij̄kl̄ and gij̄ one can easily see that at z = 0 we have Rij̄kl̄ = δij̄ δkl̄ + δil̄ δkj̄ . In
particular, as gij̄ |z=0 = δij̄ at any other point
Rij̄kl̄ = gij̄ gkl̄ + gil̄ gkj̄
and M is of constant holomorphic sectional (and bisectional) curvature c = +1.
Example 3.2.9: [Bergman Metric on the Complex Ball.] Consider M =
BCn (1) = {z ∈ Cn | |z|2 < 1} and let
i ¯
ωg = ∂ ∂ ln(1 − |z|2 ) .
2
Likewise, one can show that Bergman metric g is U (n, 1)-invariant and that U (n, 1)
acts on M = BCn (1) transitively. Repeating the calculations of the previous example
we can easily see that Rij̄ = −(n + 1)gij̄ and (M, g) is of constant holomorphic
sectional curvature c = −1.
It can be shown that any two simply-connected complete Kähler manifolds with
constant sectional curvature c are holomorphically isometric. Hence, the above
examples together with the flat metric on Cn show that
88 3. KÄHLER MANIFOLDS

Theorem 3.2.10: Let (M 2n , J, g, ωg ) be a complete Kähler manifold of constant


holomorphic sectional curvature c. Depending on the sign of c and up to scaling,
f is holomorphically isometric to CPn , Cn , or B n (1).
the universal cover M C

3.3. Hodge Theory on Kähler Manifolds


Let (M 2n , J, g, ωg ) be a compact Hermitian manifold. We define the Hodge star
operator to be the map ? : Γ(Λk,l M )−→Γ(Λn−k,n−l )M defined by
ωgn
(3.3.1) φ ∧ ?ψ = g(φ, ψ̄)dvolg = g(φ, ψ̄) , φ, ψ ∈ Γ(Λk,l M ) .
n!
Note that this map is linear over R but conjugate-linear over C. In addition, we
have ?2 φ = (−1)k+1 φ. The action of the Hodge star operator on any complex k-
form φ ∈ Γ(Λk M ⊗ C) is defined by the splitting of φ as a sum of forms of type
(k − j, j). Locally, in a holomorphic chart (U ; z1 , . . . , zn ), n = dimC (M ) one can
write ½ X ¾
Γ(Λk,l M ) = ωI J¯dzI ∧ dz̄J , ωI J¯ ∈ C ∞ (U, C) ,
|I|=k,|J|=l
where
I = (i1 , . . . , ik ), 1 ≤ i1 < · · · < ik ≤ n, dzI = dzi1 ∧ · · · ∧ dzik ,
J = (j1 , . . . , jl ), 1 ≤ j1 < · · · < jk ≤ n, dz̄J = dz̄k1 ∧ · · · ∧ dz̄jl .
Let us assume that (U ; z1 , . . . , zn ) are normal coordinates, i.e., {dz1 , . . . , dzn } is
a unitary frame of Γ(Λ1,0 M ). Denote by I, ˆ Jˆ the likewise ordered complements of
I, J in {1, . . . , n} and define σ(I) by
dzI ∧ dzIˆ = (−1)σ(I) dz1 ∧ · · · ∧ dzn .
Then we have an explicit formula
?(ωI J¯dzI ∧ dz̄J ) = 2k+l−m in ²IJ ωI J¯dzIˆ ∧ dz̄Jˆ ,
where the sign ²IJ = (−1)n(n−1)/2+(n−k)l+σ(I)+σ(J) . For compact M we can define
an L2 Hermitian inner product on each Γ(Λk,l M ) by
Z
(3.3.2) hφ, ψi = φ ∧ ?ψ, φ, ψ ∈ Γ(Λk,l M ) .
M
One can check that ? is an isometry with respect to this inner product, i.e.,
h?φ, ?ψi = hφ, ψi. We proceed to define
(3.3.3) ∂¯∗ = − ? ∂?
¯ : Γ(Λk,l M )−→Γ(Λk,l−1 M ) .
It follows that this is a formal adjoint of the ∂¯ operator as for any φ ∈ Γ(Λk,l−1 M )
and any ψ ∈ Γ(Λk,l M ) we have
¯ ∧ ?ψ − φ ∧ ?∂¯∗ = ∂φ
∂φ ¯ ∧ ?ψ + (−1)k+l−1 φ ∧ ∂¯ ? ψ = ∂(φ
¯ ∧ ?ψ) = d(φ ∧ ?ψ) ,
¯ ψi = hφ, ∂¯∗ ψi . Similarly, once can define
so that, by Stokes’ theorem, h∂φ,
∂ ∗ = − ? ∂? : Γ(Λk,l M )−→Γ(Λk−1,l M )
which is an adjoint of the ∂ operator. Hence, we have 6 different operators acting
¯ ∂¯∗ and 3 natural Laplacians
on forms d, d∗ , ∂, ∂ ∗ , ∂,
∆d = dd∗ + d∗ d , ∆∂ = ∂∂ ∗ + ∂ ∗ ∂ , ∆ ¯ = ∂¯∂¯∗ + ∂¯∗ ∂¯ .

3.3. HODGE THEORY ON KÄHLER MANIFOLDS 89

The Laplacian ∆∂¯ is a self-adjoint operator and


¯ ∂φi
h∆∂¯φ, φi = h∂φ, ¯ + h∂¯∗ φ, ∂¯∗ φi ≥ 0 .

Definition 3.3.1: The space Hk,l (M ) ⊂ Γ(Λk,l M )


Hk,l (M ) = {φ ∈ Γ(Λp,q M ) | ∆∂¯φ = 0} .
is called the space of harmonic (k, l)-forms. Furthermore, dim(Hk,l (M )) = hk,l (M )
are called the Hodge numbers of M .
We now have the famous Hodge Theorem.
Theorem 3.3.2: On a compact Hermitian manifold (M 2n , J, g, ωg ) the dimensions
dim(Hk,l (M )) < ∞ for all 0 ≤ k, l ≤ n. Furthermore, we have the decompositions
¯
Γ(Λk,l M ) = Hk,l (M ) ⊕ ∂(Γ(Λ k,l−1 ¯
M )) ⊕ ∂(Γ(Λ k,l−1
M )) .
In particular, any class in H∂k,l
¯ (M ) has a unique harmonic representative, i.e., we
have a complex vector space isomorphism
Hk,l (M ) ' H∂k,l
¯ (M ) .

Let us introduce the projections


π k,l : Γ(Λ∗ T M ⊗ C)−→Γ(Λk,l M ) .

π r = ⊕k+l=r π k,l : Γ(Λ∗ T M ⊗ C)−→Γ(Λr M ⊗ C) .


Since the metric g is Hermitian we have πk,l ? = ?πk,l . In general, given a complex
∆d -harmonic r-form ψ the projection πk,l ψ need not be ∆∂¯-harmonic. However,
on a Kähler manifold we have some additional structure.
To show this let us introduce the operator
L : Γ(Λk,l M )−→Γ(Λk+1,l+1 M )
defined by wedging with the Kähler form ωg , i.e., L(φ) = φ ∧ ωg . The adjoint of L
in the L2 -norm is given by
L∗ = (−1)p+q ? L? : Γ(Λk,l M )−→Γ(Λk−1,l−1 M ) .
We have the following
Lemma 3.3.3: [Kähler Identities] Let (M, J, g, ωg ) be a Kähler manifold. Then
[L, ∂ ∗ ] = i∂¯ , [L, ∂¯∗ ] = −i∂ ,
[L∗ , ∂] = i∂¯∗ , ¯ = −i∂ ∗ .
[L∗ , ∂]

Proof. Since both sides of all these inequalities are first order differential
operators it is enough to check the for the Euclidean space Cn with the standard
Hermitian metric. The general statement follows then from the fact that Kähler
condition is equivalent to the existence of normal coordinates. ¤

Exercise 3.2: Verify the identities of Lemma 3.3.3 in normal coordinates.


Proposition 3.3.4: Let (M, J, g, ωg ) be a compact Kähler manifold. The
(i) [L∗ , ∆d ] = [L∗ , ∆d ] = 0 ,
(ii) ∆d = 2∆∂ = 2∆∂¯ ,
(iii) [∆d , π p,q ] = 0 .
90 3. KÄHLER MANIFOLDS

Proof. Now, the proof of the proposition. We have


¯ ∗ + ∂¯∗ ) + (∂ ∗ + ∂¯∗ )(∂ + ∂)
∆d = (∂ + ∂)(∂ ¯ = ∆∂ + ∆ ¯ + (∂ ∂¯∗ + ∂¯∗ ∂) + (∂∂ ¯ .
¯ ∗ + ∂ ∗ ∂)

Using Kähler identities we can replace ∂¯∗ = −i[L∗ , ∂] and

(∂ ∂¯∗ + ∂¯∗ ∂) = −i(∂[L∗ , ∂] + [L∗ , ∂]∂) = 0 .


¯ = −i∂ ∗ , we have
On the other hand, using [L∗ , ∂]

−i∆∂ = ∂(−i∂ ∗ ) + (−i∂ ∗ )∂ = ∂(L∗ ∂¯ − ∂L


¯ ∗ ) + (L∗ ∂¯ − ∂L
¯ ∗ )∂ =

= ∂L∗ ∂¯ − L∗ ∂ ∂¯ + ∂∂L
¯ ∗ − ∂L
¯ ∗ ∂ = [∂, L∗ ]∂¯ + ∂[∂,
¯ L∗ ] = −i(∂¯∗ ∂¯ + ∂¯∂¯∗ ) = −i∆ ¯ .

There are two important consequences of the above propositions. The first
is the so-called Hodge decomposition. The second is the so-called hard Lefschetz
theorem. Let us first discuss the Hodge structure. We define

Zdk,l (M )
H k,l (M, C) = ,
dΓ(Λ∗ M ⊗ C) ∩ Zdk,l (M )
to be the space of d-closed forms of type (k, l) modulo d-exact forms and let Hdr (M )
be the space of ∆d -harmonic r-forms. Since the usual Laplacian ∆d is real, com-
¯
mutes with π k,l and equals 1/2 of the ∂-Laplacian we have

Hr (M ) = ⊕k+l=r Hk,l (M ) ,

H∂k,l k,l k,l


¯ (M ) = H (M ) = H (M ) = H
k,l
(M, C) .

Using the usual Hodge theorem HDR (M ) = H∗ (M ) for the Laplacian ∆d we get
Theorem 3.3.5: For a compact Kähler manifold we have

H r (M, C) = ⊕k+l=r H k,l (M, C) , H k,l (M, C) = H l,k (M, C) .

Since [?, ∆∂¯] = 0, by Kodaira-Serre duality4 the map

? : Hk,l (M )−→Hn−k,n−l (M ),

is an isomorphism. Summarizing, we have the following properties of the Hodge


numbers
Corollary 3.3.6: Let(M 2n , J,P
g, ωg ) be a compact Kähler manifold, hk,l (M ) be the
Hodge numbers, and br (M ) = k+l=r hk,l (M ) the Betti numbers of M . Then for
all 0 ≤ k, l ≤ n we have
(i) hk,l (M ) < ∞,
(ii) hk,k (M ) ≥ 1 and hn,n = h0,0 = 1,
(iii) hk,l (M ) = hl,k (M ) = hn−k,n−l (M ),
1
(iv) b2r+1 (M ) are even and h1,0 (M ) = h0,1 (M ) = 2 b1 (M ) is a topological
invariant.

4The general version of the Kodaira-Serre Duality Theorem is given in Theorem 3.4.9 below.
3.3. HODGE THEORY ON KÄHLER MANIFOLDS 91

For any compact Kähler manifold once can consider the Hodge diamond, that
is the arrangement of the Hodge numbers in a diamond-shape array. For instance,
when n = 3 we can have
h3,3

h3,2 h2,3

h3,1 h2,2 h1,3

h3,0 h2,1 h1,2 h0,3

h2,0 h1,1 h0,2

h1,0 h0,1

h0,0

The above corollary implies that every Hodge diamond has two symmetries: con-
jugation gives the symmetry through the central vertical axis while Hodge star
yields the symmetry through the center of the diamond. There are more relations
between Hodge numbers. They are due to the Lefschetz decomposition which we
will discuss below.
P2n
Exercise 3.3: Let us define an operator h = r=0 (n − r)π r . Show that we have
the following relations

[L∗ , L] = h, [h, L] = −2L, [h, L∗ ] = 2L∗

and, hence, {L, L∗ , h} are generators of the Lie algebra sl(2, R).
Definition 3.3.7: Let (M, J, g, ωg ) be a compact Kähler manifold. We define the
primitive cohomology groups of M as kernels of the L∗ -operator, i.e.,
¡
P r (M, C) = ker L∗ : H r (M, C)−→H r−2 (M, C)),
¡
P k,l (M, C) = ker L∗ : H k,l (M, C)−→H k−1,l−1 (M, C)).
The dimension of P r (M, C) is called primitive rth Betti number of M.
Since {L, L∗ , h} all commute with the Kähler Laplacian we have

P r (M, C) = ⊕k+l=r P k,l (M, C).

The generators {L, L∗ , h} give a finite-dimensional representation of the Lie algebra


sl(2, R) acting on H ∗ (M, C), where H r (M, C) is the eigenspace of the operator h
with eigenvalue (n − r). The next theorem follows directly from the theory of
finite-dimensional representations of sl(2, R):
Theorem 3.3.8: (Hard Lefschetz) On a compact Kähler manifold M the map
Lk : H n−k (M )−−→H n+k (M ) is an isomorphism for 1 ≤ k ≤ n. Furthermore,
M
H r (M ) = Lk P r−2k (M ) .
k≥0
92 3. KÄHLER MANIFOLDS

3.4. Complex Vector Bundles and Chern Classes


Let (M, J) be a smooth manifold and let E−→M be a complex vector bundle
of complex rank r over M . Let ∇ be a complex Koszul connection (cf. Definition
1.3.5) on E with curvature form Ω. Consider the space Mr×r (C) of complex r × r
matrices. For any A ∈ Mr×r (C) we define
det(A + λI) = fr (A) + λfr−1 (A) + · · · + λr−1 f1 (A) + λr .
Clearly, for any i = 1, . . . , r the function fi : Mr×r (C)−→C is a GL(r, C)-invariant,
complex homogeneous polynomial of deg(fi ) = i. Note that fi is the ith elementary
symmetric function of the eigenvalues of A. In particular, fr (A) = det(A) and
f1 (A) = Tr(A).
Definition 3.4.1: Let E−→M be rank r complex vector bundle over M, and let ∇
be a complex connection on E with curvature 2-form Ω. For each i = 1, . . . , r we
define the 2i-form
µ ¶
i
fi Ω = ci (E, ∇)

and call it the ith Chern form of E.
We have the following
Proposition/Definition 3.4.2: Given (E, ∇) and any 1 ≤ i ≤ r, the ith Chern
form ci (E, ∇) is closed. Furthermore, if ∇ e is another complex connection on E
e
the difference ci (E, ∇) − ci (E, ∇) is exact, i.e., the cohomology class [ci (E, ∇)] ∈
2i
HDR (M ) is independent of ∇. The resulting cohomology class is called the ith
Chern class of E and is denoted by ci (E).
When working
P∞ with Chern classes it is convenient to consider the total Chern
class c(E) = i=0 ci (E), where c0 (E) = 1 and the sum is always finite for a finite
rank vector bundle, so c(E) ∈ H ∗ (M, Z). For Whitney sums the total Chern class
satisfies c(E ⊕ E 0 ) = c(E)c(E 0 ). If Ē denotes the complex conjugate bundle to E,
then ci (Ē) = (−1)i ci (E) and since Ē is isomorphic to the dual complex vector
bundle ci (E ∗ ) = (−1)i ci (E). There are also certain “multiplicative sequences” of
interest [Hir66, MS74]. We define the Chern roots xi of a rank k complex vector
Qk
bundle by c(E) = 1 + c1 (E) + · · · + ck (E) = i=1 (1 + xi ). Then the Chern character
ch(E) and the Todd class Td(E) are defined by
k
X k
Y xi
(3.4.1) ch(E) = exi , Td(E) = .
i=1 i=1
1 − e−xi

Generally, these are elements in the rational cohomology ring H ∗ (M, Q). Here is
the celebrated Hirzebruch-Riemann-Roch Theorem:
Theorem 3.4.3: Let E be a holomorphic vector bundle on a compact complex
manifold M n . Then
χ(M, E) = t2n (Td(M )ch(E)) ,
Pn i
where χ(M, E) = i=1 h (M, O(E)) is the holomorphic Euler characteristic of E.
When E is the trivial bundle we get an invariant of the complex structure,
namely the holomorphic Euler characteristic χ(M, O) and Theorem 3.4.3 gives the
‘Todd-Hirzebruch formula’ χ(M, O) = t2n (Td(M )), the later being known as the
Todd genus.
3.4. COMPLEX VECTOR BUNDLES AND CHERN CLASSES 93

Example 3.4.4: The complexified tangent bundle. Consider the case where E =
T M ⊗ C is the complexified tangent bundle of a complex manifold which splits as
T M ⊗ C = T 1,0 M ⊕ T 0,1 M. Now T 1,0 M is a holomorphic vector bundle, called the
holomorphic tangent bundle, which is isomorphic as a real vector bundle to T M.
Now T 1,0 M has a complex connection ∇1,0 which can be used to compute its Chern
classes, ci (M ) = ci (T 1,0 M ), called the ith Chern class of the complex manifold M,
and denoted simply by ci when the complex manifold M is understood. These
Chern classes depend only on the homotopy class of the complex structure J on M .
Note that the top class cn (M ) = e(M ) is the Euler class. Given any partition I =
i1 , . . . , ir of the integer n, we can define the I th Chern number cI [M ] by evaluation
on the fundamental homology class [M ], that is, cI [M ] = hci1 · · · cir , [M ]i. Note
that the top P Chern number cn [M ] is just the Euler-Poincaré characteristic cn [M ] =
χ(M ) = i (−1)i bi (M ).
In low dimensions the Todd-Hirzebruch formula reduces to well-known classical
formulae. Recall that h0,1 = q is called the irregularity. For n = 1 (compact
Riemann surfaces) we get h0,1 = g, where g is the genus of the Riemann surface.
For n = 2 we get Noether’s formula for compact complex surfaces, χ(M, O) =
1
1 − q + pg = 12 (c21 [M ] + c2 [M ]).
Now suppose that (M, J) is a complex manifold and E is a complex vector
bundle on M. We can consider the tensor product bundle E ⊗ Λk,l M, and we let
E k,l (E) denote the sheaf of germs of smooth sections of E ⊗Λk,l M. Smooth sections
of this sheaf are (k, l)-forms with coefficients in E, the set of which we denote by
Ak,l (E). The connection ∇ in E induces a connection, also written as ∇, in Ak,l (E).
This connection splits as ∇ = ∇1,0 + ∇0,1 giving maps
∇1,0 : Ak,l (E)−→Ak+1,l (E), ∇0,1 : Ak,l (E)−→Ak,l+1 (E).
We are interested in the case when E admits a holomorphic structure. Then simi-
larly, we let Ωp (E) denote the sheaf of germs of holomorphic sections of the bundle
Λp,0 ⊗ E. So we shall analyze the structure of connections in holomorphic vector
bundles following [Kob87].
Theorem 3.4.5: A smooth complex vector bundle E over a complex manifold ad-
mits a holomorphic structure if and only if there is a connection ∇ in E such that
¯
∇0,1 = ∂.
The holomorphic structure in E is uniquely determined by the condition ∇0,1 =
¯
∂, and this condition says that the (0, 2) component ∇0,1 ◦ ∇0,1 of the curvature
of ∇ vanishes. It is straightforward to generalize Definition 3.1.7 to an arbitrary
complex vector bundle.
Definition 3.4.6: An Hermitian metric h on a complex vector bundle E is an
assignment of an Hermitian inner product to each fibre Ex of E that varies smoothly
with x. A connection ∇ in E is called an Hermitian connection if ∇h = 0.
A vector bundle equipped with a Hermitian metric is often called an Hermitian
vector bundle. Using partitions of unity one easily sees that Hermitian metrics
exists on complex vector bundles.
Proposition 3.4.7: Let E be a holomorphic vector bundle with an Hermitian met-
ric h. Then there exists a unique Hermitian connection ∇ such that ∇0,1 = ∂. ¯
Again for a proof see [Kob87]. The unique connection of Proposition 3.4.7 is
called the Hermitian connection. There is a version of Hodge theory tensored with
94 3. KÄHLER MANIFOLDS

holomorphic vector bundles with an Hermitian metric. Such a choice of Hermitian


metric gives an isomorphism τ between a holomorphic vector bundle E and its dual
E ∗ . Then we can define the Hodge star isomorphism
?E = ? ⊗ τ : Ak,l (E)−−→An−k,n−l (E ∗ )
or, alternatively, on the sheaf level
?E = ? ⊗ τ : E k,l (E)−−→E n−k,n−l (E ∗ ).
We can now extend the Hodge inner product 3.3.2 to an inner product on Ak,l (E)
by
Z
(3.4.2) hφ , ψi = φ ∧ ?E ψ , φ, ψ ∈ Ak,l (E) ,
M

where we write φ, ψ ∈ A (E) = Γ(Λk,l M ⊗ E) as φ = φ ⊗ e, and ψ = ψ ⊗ f,


k,l

respectively, in which case we have φ ∧ ?E ψ = (φ ∧ ?ψ)he, f iE , where he, f iE is the


pairing between E and E ∗ .
Now consider the cochain complex
∂¯
· · · −−→Ak,l (E) −−→ Ak,l+1 (E)−−→ · · ·
and let H∂p,q
¯ (M, E) denote the derived cohomology groups of this complex. The
operator ∂ has a (formal) adjoint ∂¯E
¯ ∗ ¯ E with respect to the Hodge inner
= − ?E ∂?
product 3.4.2, and we have complex Laplacian ¤̄E = ∂¯∂¯E

+ ∂¯E
∗ ¯
∂ acting on A∗,∗ (E).
k,l
As in Definition 3.3.1 we define the space H (M, E) of harmonic E-valued (k, l)-
forms by
(3.4.3) Hk,l (M, E) = {φ ∈ Ak,l (E) | ¤̄E φ = 0} .
Combining the resulting vector bundle version of the Hodge isomorphism The-
orem 3.3.2 with the Abstract de Rham Theorem 1.1.16, and using the acyclic ∂- ¯
resolution of the sheaf Ωp (E) and one gets Serre’s generalized Dolbeault Theorem
for holomorphic vector bundles [Ser55]:
Theorem 3.4.8: Let M be a complex manifold, and E a holomorphic vector bundle
on M. Then there are isomorphisms
H q (M, Ωp (E)) ≈ H∂p,q
¯ (M, E) ≈ H
p,q
(M, E) .
This theorem can be used to prove the celebrated Kodaira-Serre Duality The-
orem [Ser55]:
Theorem 3.4.9: Let M be a compact complex manifold of complex dimension n,
and let E be a holomorphic vector bundle over X. There there is a conjugate-linear
isomorphism
ψ : H q (M, Ωp (E))−−−→H n−q (M, Ωn−p (E ∗ )) .

3.5. Line Bundles and Divisors


The purpose of this section is to give a brief review of the fundamental concepts
employed in complex manifold theory, namely, line bundles and divisors as well as to
discuss their interrelationship. We refer to the literature [GH78b, Wel80, Voi02,
Laz04a] for complete treatments of these important subjects.
3.5. LINE BUNDLES AND DIVISORS 95

3.5.1. Line Bundles. We begin by considering the set of smooth complex


line bundles on a complex manifold M. These are determined by their transition
functions which have values on in the group GL(1, C) = C∗ . The isomorphism
classes of complex line bundles form a group under tensor product, and by Theorem
1.2.2 isomorphism classes of C∗ -bundles over M are in one-to-one correspondence
with elements of the sheaf cohomology group H 1 (M, E ∗ ). Thus, we consider the
exponential short exact sequence of sheaves on M
ι exp
(3.5.1) 0−−−→Z −−−→ E −−−→ E ∗ −−−→0 ,
where the map ι is ι(k) = 2πik and the exponential map sends the germ f of any
holomorphic function to exp(f ). Since E is a fine sheaf, Propositions 1.1.12 and
1.1.13 imply H q (M, E) = 0 for all q > 0. So the induced long exact cohomology
sequence (1.1.6) gives an isomorphism H 1 (M, E ∗ ) ≈ H 2 (M, Z) which says that
the topological invariant H 2 (M, Z) can be thought of as the group of complex line
bundles on M. This isomorphism is realized by associating to a complex line bundle
L its first Chern class c1 (L).
To study the holomorphic line bundles on M we consider the exact sequence
ι exp
(3.5.2) 0−−−→Z −−−→ O −−−→ O∗ −−−→0 .
This induces a long exact sequence in cohomology,
δ
(3.5.3) −−→H 1 (M, O)−−→H 1 (M, O∗ ) −−→ H 2 (M, Z)−−→H 2 (M, O)−−→ · · · .
The group H 1 (M, O∗ ) represents the group of holomorphic line bundles on M with
group multiplication being the tensor product, and the inverse bundle being the
dual bundle. This group is called the Picard group of M and often denoted by
Pic(M ). As seen above the connecting homomorphism δ takes a holomorphic line
bundle L to its first Chern class c1 (L), and the group H 2 (M, Z) is isomorphic to the
group of topological complex line bundles on M. So if H 2 (M, O) 6= 0 we see that
not every complex line bundle gives rise to a holomorphic line bundle. Similarly,
if H 1 (M, O) 6= 0, there can be inequivalent holomorphic bundles associated to the
same complex line bundle. The kernel of the map δ is denoted by Pic0 (M ) and
represents the subgroup of holomorphic line bundles that are trivial topologically.
The quotient group Pic(M )/Pic0 (M ) is known as the Neron-Severi group denoted
by N S(M ). The rank of N S(M ) is called the Picard number of M and denoted
by ρ(M ). When M is a smooth projective algebraic variety, we have 1 ≤ ρ(M ) ≤
b2 (M ), and it follows from the well-known Lefschetz Theorem on (1, 1) classes that
there is an isomorphism N S(M ) ≈ H 1,1 (M, C) ∩ H 2 (M, Z). So N S(M ) is a free
Abelian group of rank ρ(M ).
The notion of positivity is fundamental in the study of holomorphic line bun-
dles.
Definition 3.5.1: Let M be a complex manifold and L a holomorphic line bundle
on M. We say that L is positive (negative) if it carries an Hermitian met-
ric whose curvature form Ω with respect to the Hermitian connection is a positive
(negative) (1, 1)-form on M.
Since Chern classes can be computed with respect to any connection, we see
i
that the first Chern class c1 (L) can be represented by 2π Ω. Notice that when L is
i
positive, the form 2π Ω defines a Kähler metric ω on M with integral cohomology
class. This leads to
96 3. KÄHLER MANIFOLDS

Definition 3.5.2: The pair (M, L) consisting of a complex manifold M with a


positive line bundle L is called a polarized Kähler manifold.
Of course, the complex manifold M is necessarily a manifold of Kähler type, and
i
the Kähler structure is that determined by the curvature of L, namely, ω = 2π Ω. So
we often refer to the pair (M, [ω]) as a polarized Kähler manifold. If M is a Kähler
manifold with an integral Kähler class [ω] or equivalently a positive line bundle L,
we also say that M admits a polarization or is polarized by [ω] or L.
There is a complex line bundle canonically associated with every complex man-
ifold, called the canonical bundle.
Definition 3.5.3: Let M be a complex manifold of complex dimension n. The
nth exterior power Λn T ∗(1,0) M = Λn,0 M is a holomorphic line bundle, called the
−1
canonical bundle and denoted by KM . The dual or inverse line bundle KM is
called the anticanonical bundle.
When the underlying manifold M is understood we often write just K for KM .
It is easy to see that
Proposition 3.5.4: The first Chern class of KM satisfies
c1 (KM ) = −c1 (M ) .
There are important discrete invariants associated to KM and hence, to M. On
any complex manifold we define the plurigenera by
(3.5.4) Pm (M ) = dim H 0 (M, O(Km )) = h0 (M, O(Km )) .
Recall that P1 (M ) is called the geometric genus and is usually denoted by pg . Note
that by the well-known Dolbeault Theorem 3.4.8, pg = hn,0P . Another important
invariant is the holomorphic Euler characteristic χ(M, O) = i hi (M, O).
Consider the commutative ring
X
(3.5.5) R(M ) = C ⊕ H 0 (M, O(Km )) .
m≥1

R(M), called the canonical ring, has finite transcendence degree tr(R(M )) over C.
Definition 3.5.5: Let M be a compact complex manifold. The Kodaira dimension
of M is defined by
(
−∞ if R(M ) = C
Kod(M ) =
tr(R(M )) − 1 otherwise,
where tr(R) denotes the transcendence degree of the ring R.
More generally one can define (cf. [Laz04a]) the Iataka dimension κ(M, L) of
any line bundle L by simply replacing K by L in Equation 3.5.5. Then Kod(M ) =
κ(M, K). The Iataka dimension and, therefore, the Kodaira dimension can be de-
fined on any normal projective algebraic variety. We shall make use of the Iataka
dimension when discussing the ‘orbifold Kodaira dimension’ in Chapter 4. Recall
that the transcendence degree of the field of meromorphic functions on a compact
complex manifold is called the algebraic dimension and is denoted by a(M ). Re-
garding the Kodaira dimension we have Kod(M ) ≤ a(M ) ≤ n = dim M. The
Kodaira dimension gives a measure of the asymptotic growth of the plurigenera.
Theorem 3.5.6: Let M be compact complex manifold. Then
(i) Kod(M ) = −∞ if and only if Pm (M ) = 0 for all m;
3.5. LINE BUNDLES AND DIVISORS 97

(ii) Kod(M ) = 0 if and only if Pm (M ) = 0 or 1, but not 0 for all m;


(iii) Kod(M ) = k if and only if there exist constants α, β such that αmk ≤
Pm (M ) ≤ βmk .
The importance of the Kodaira dimension is that it is a birational invariant.
Recall [Har77] that a rational map is defined as follows: (it is not a map in the usual
sense) let X, Y be varieties (either both affine or both projective) and consider pairs
(U, φU ), where U is a nonempty open subset of X, and φU : U −−→Y is a holomorphic
map. Two such pairs (U, φU ) and (V, φV ) are equivalent if φU |U ∩V = φV |U ∩V . Then
a rational ‘map’ φ : X−−→Y is an equivalence class of such pairs. φ is said to be
dominant of if the image φU is dense in Y. A birational map is a rational map that
admits an inverse, namely, a rational map ψ : Y −−→X such that ψ ◦ φ = 1lX and
φ ◦ ψ = idY . For a thorough treatment of rational maps we recommend the recent
book [Har92].
The most important example of a rational map is the ‘blowing-up map’ which
we now describe.
Example 3.5.7: [Blowing-up] Let Br (0) be a ball of radius r centered at 0 in
Cn , where n ≥ 2, and let z = (z1 , . . . , zn ) be the standard coordinates in Cn . Then
blowing-up Br (0) at 0 is the complex manifold defined by

B̃ = {(z, y) ∈ B × CPn−1 | zi yj = zj yi , ∀ 1 ≤ i < j ≤ n} ,

where y = (y1 , . . . , yn ) are homogeneous coordinates for CPn−1 . There is a natural


surjective holomorphic map π : B̃−−→B defined as the restriction to B̃ of the
projection map B × CPn−1 onto the first factor. For z 6= 0 the fibre π −1 (z) is the
single point (z, [z]); whereas, at z = 0 we have π −1 (0) ≈ CPn−1 . So B̃ is B with the
origin 0 replaced by a projective space CPn−1 . The fibre π −1 (0) = CPn−1 is called
the exceptional divisor, and often denoted by E. Away from the exceptional divisor
π is a biholomorphism. Replacing B̃ by B is called blowing-down. One can show
n
that B̃ is diffeomorphic to B#CP , where # denotes the connected sum operation
n n
and CP is CP with the reverse orientation.
It is easy to transfer the blowing-up process to an arbitrary complex manifold
M. Let p ∈ M and V a neighborhood containing p that is biholomorphic to B with
p mapping to 0. Identifying V with B, we let M f be the manifold obtained from M
f
by replacing V = B by B̃. So M is a complex manifold that is diffeomorphic to
n
M #CP . Analytically, π is a birational map, so we have the following birational
invariants: a(M f) = a(M ), Pm (M f) = Pm (M ), and Kod(M f) = Kod(M ). We also
i i f ∗
have an isomorphism H (M, OM ) ≈ H (M , OM f) for all i ≥ 0, and KM
f = π KM ⊗
O((n−1)E), where O(D) is the line bundle associated to the divisor D as described
in subsection 3.5.2 below.
Blowing-up can be applied to singular points of algebraic varieties to obtain
smooth manifolds. This procedure has led to the celebrated Hironaka Resolution
of Singularities Theorem which says that a singular algebraic variety can be desin-
gularized after a finite sequence of blowing-ups.
Now for certain complex manifolds there is a “vanishing theorem” H q (M, O) =
0 for all q > 0. When this happens we get an isomorphism between the Picard
group H 1 (M, O∗ ) of holomorphic line bundles on M and the topological invariant
H 2 (M, Z). Such vanishing theorems are of great importance in complex geometry.
98 3. KÄHLER MANIFOLDS

Here we only give the well-known Kodaira-Nakano vanishing Theorem, and we refer
to Chapter 4 of [Laz04a] and [SS85] for further development.
Theorem 3.5.8: Let M be a compact complex manifold of complex dimension n,
and let L be a holomorphic line bundle on M.
(i) If L ⊗ K−1 is positive, then H q (M, O(L)) = 0 for all q > 0. In particular,
if the anticanonical bundle K−1 is positive, H q (M, O) = 0 for all q > 0.
(ii) If L is negative, then H q (M, Ωp (L)) = 0 for all p + q < n.
Compact manifolds with positive anticanonical bundles are called Fano man-
ifolds, and it follows immediately from the long exact sequence 3.5.3 and (i) of
Theorem 3.5.8 that
Corollary 3.5.9: For any Fano manifold M there are isomorphisms Pic(M ) ≈
N S(M ) ≈ H 2 (M, Z). So for Fano manifolds ρ(M ) = b2 (M ).
Example 3.5.10: A simple example of a Fano manifold is the complex projective
space CPn as discussed in Examples 3.1.12 and 3.2.8. It follows from Equation 3.2.2
and Proposition 3.6.1 that c1 (CPn ) is positive. It is known that H 2 (CPn , Z) =
Z, and it follows from Proposition 3.5.4 and Theorem 3.5.8 that Pic(CPn ) ≈
N S(CPn ) ≈ Z. The positive generator of Pic(CPn ) is known as the hyperplane
bundle and denoted by H or as we do using the invertible sheaf notation O(1). So
every holomorphic line bundle on CPn is O(n) for some n ∈ Z. The line bundle
O(−1) is called the tautological bundle since its total space is the Cn+1 from which
CPn is constructed. It is easy to compute the canonical bundle of CPn finding
KCPn = O(−n − 1).
Example 3.5.11: Low dimensional Fano manifolds. The smooth Fano varieties
have been classified for n = 1, 2, 3. There is a unique one dimensional smooth Fano
variety up to isomorphism; it is the complex projective plane CP1 . For n = 2 they
are classical and known as del Pezzo surfaces. A smooth del Pezzo surface is, up to
isomorphism, one of the following:
(i) CP2 ,
(ii) a smooth quadric Q = CP1 × CP1 ,
(iii) a double cover F1 of a quadric cone Q0 CP3 ramified along a smooth curve
of degree 6 not passing through the vertex of the cone,
(iv) a double cover F2 of CP2 ramified along a smooth curve of degree 4,
(v) a surface Fd ⊂ CPd , d = KF2 d , where 3 ≤ d ≤ 7,
(vi) a geometrically ruled surface F1 with the exceptional section s, s2 = −1.
One can show that surfaces Fd , d = 1, . . . , 7 can be obtained by blowing up of 9 − d
points on CP2 which are in general position, i.e., no two of these points lie on a
line nor any three lie on a conic. Also, F1 can be obtained by blowing CP2 at 1
point. Thus, as a smooth manifold a del Pezzo surface must be diffeomorphic to
2 2
CP2 , Q = CP1 × CP1 , F1 = CP2 #CP , or Fd = CP2 #(9 − d)CP , where 1 ≤ d ≤ 7.
The classification for n = 3 was begun by Fano and almost completed by
Iskovskikh in [Isk77, Isk78, Isk79]. However, an additional 3-fold was found
a bit later by Mukai and Umemura [MU83], and it was shown by Prokhorov
[Pro90] to complete the classification. We refer to [IP99] for a survey including
the complete list of Fano 3-folds. Generally, Fano manifolds and orbifolds are of
much importance for us in this book. For recent treatments of Fano varieties see
[Kol96, IP99].
3.5. LINE BUNDLES AND DIVISORS 99

More generally, any generalized flag manifold G/P is Fano. Fano manifolds
have the important property that they can always be embedded into a complex
projective space CPN for some N, i.e., they are examples of projective algebraic
varieties as discussed in Example 3.1.18. The beautiful Kodaira embedding theorem
gives precise conditions to have a projective algebraic variety.
Theorem 3.5.12: Let M be a compact complex manifold and L a holomorphic line
bundle on M. Then L is positive if and only if there is a holomorphic embedding
φ : M −−→CPN for some N such that φ∗ O(1) = Lm for some m > 0.
The condition of the holomorphic embedding gives rise to
Definition 3.5.13: A holomorphic line bundle L on a complex manifold M is
said to be ample if there is an embedding φ : M −−→CPN for some N such that
φ∗ O(1) = Lm for some m > 0. If we can take m = 1 then L is said to be very
ample.
The Kodaira Embedding Theorem can now be reformulated in two more equiv-
alent ways.
Theorem 3.5.14: Let M be a compact complex manifold. Then
(i) A holomorphic line bundle L on M is positive if and only if it is ample
in which case M is a projective algebraic variety.
(ii) M admits polarization if and only if it is projective algebraic.

So a Fano manifold could be defined by the condition that the anticanonical


−1
line bundle KM is ample. In this case the corresponding embedding is said to be
an anticanonical embedding. Similarly, if the canonical line bundle KM is ample,
then the manifold M is projective algebraic and the corresponding embedding is
called a canonical embedding.
A very important criterion for the ampleness of a line bundle was discovered
by Nakai in the case of complex surfaces and generalized to arbitrary smooth al-
gebraic varieties by Moishezon. It was generalized further to complete schemes by
Kleiman. Here we simply state the version for smooth algebraic varieties and refer
to Lazarsfeld’s recent book [Laz04a] for further discussion and proof of the scheme
theoretic version.
Theorem 3.5.15: Let M be a compact manifold of complex dimension n, and L
a holomorphic line bundle on M. Then L is ample if and only if
Z
c1 (L)k > 0
V
for every irreducible subvariety V ⊂ M of dimension k.
In the case of surfaces Nakai’s criterion can be stated as L is ample if and only
if c21 (L) > 0 and
Z
c1 (L) > 0
D
for every effective divisor D on M.
If the Kähler class [ω] is a rational class, that is it lies in H 2 (M, Q), then k[ω] is
an integral class for some positive integer k. So k[ω] defines a positive holomorphic
line bundle L on M, and Theorem 3.5.14 implies that (M, ω) is a projective algebraic
variety. A Kähler manifold (M, ω) with [ω] ∈ H 2 (M, Q) is called a Hodge manifold.
100 3. KÄHLER MANIFOLDS

Example 3.5.16: A complex torus is the quotient manifold T n = Cn /Λ, where Λ


is a lattice, that is a discrete subgroup of Cn of rank 2n. A flat Kähler structure on
Cn induces a Kähler structure ω on T n . However, for a generic complex structure,
the Kähler class [ω] ∈ H 1,1 (M, C) ∩ H 2 (M, R) is not a rational class. So generally
a complex torus T n is not a projective algebraic variety. When one can find a
compatible Kähler form ω such that [ω] is a rational class, T n is projective algebraic
by the Kodaira Embedding Theorem 3.5.12 in which case it is called an Abelian
variety. Alternatively, a complex torus T n which admits a positive line bundle L
is an Abelian variety. So the pair (T n , L) is called a polarized Abelian variety. For
a complete treatment of Abelian varieties we refer the reader to [LB92].

3.5.2. Divisors. There are two equivalent way to describe divisors on smooth
complex manifolds. Since they are not equivalent for singular algebraic varieties or
more generally for complex spaces, we discuss both of these here. The singular case
is treated in Section 4.4 below.
Definition 3.5.17: A Weil divisor D on a complex manifold M is a locally finite
formal linear combination of irreducible analytic hypersurfaces Vi
X
D= a i Vi , ai ∈ Z ,
i

where locally finite means that every point p ∈ M has a neighborhood intersecting
only finitely many of the Vi ’s. D is said to be effective if ai ≥ 0 for all i with not
all ai equal to zero.
Under the formal sum operation Weil divisors form a group, called the divisor
group and denoted by Div(M ). Now from Definition 3.1.19 a divisor is described
locally by the zero set of holomorphic functions.
Recall that a meromorphic function on an open set U ⊂ M is a ratio f /g of
relatively prime holomorphic f, g on U. We let M denote the sheaf of holomor-
phic functions and M∗ the subsheaf of not identically zero meromorphic functions.
Consider the short exact sequence
(3.5.6) 0−−→O∗ −−→M∗ −−→M∗ /O∗ −−→0 .
We have
Definition 3.5.18: Let X be a complex manifold or an algebraic variety. Elements
of the group
H 0 (X, M∗ /O∗ )
are called Cartier divisors on X. A Cartier divisor is principal if it is the divisor
of a global meromorphic function, that is, it is in the image of the natural quotient
map H 0 (X, M∗ )−−→H 0 (X, M∗ /O∗ ).
Since on a smooth manifold the local rings Ox are unique factorization domains
(UFD) Weil divisors and Cartier divisors coincide.
Theorem 3.5.19: Let M be a smooth complex manifold. Then there is an isomor-
phism
Div(M ) ≈ H 0 (M, M∗ /O∗ ) .
This isomorphism does not hold on singular complex spaces. For example the
complex spaces discussed in Section 4.4.1 below are not ‘locally factorial’. On
smooth complex manifolds we shall often identify Weil divisors and Cartier divisors
3.5. LINE BUNDLES AND DIVISORS 101

by just referring to a divisor. Two divisors D and D0 on M are said to be linearly


equivalent, written D0 ≡ D, if D0 = D + (f ), where (f ) denotes the divisor defined
by the global meromorphic function f as follows: write f locally as f = g/h as
(f ) = ord(g)Zg −ord(h)Zh . Here Zg denotes the zero set of the holomorphic function
g and ord(g) denotes the its order of vanishing. We denote by |D| the set of all
divisors on M that are linearly equivalent to D.TIt is called the linear system of
divisors defined by D. The common intersection D0 ∈|D| D0 is called the base locus
of linear system |D|.
We now describe the relationship between line bundles and divisors. From the
short exact sequence (3.5.6) one has
(3.5.7) 0−−−→H 0 (M, M∗ /O∗ )/H 0 (M, M∗ )−−−→H 1 (M, O∗ )−−−→H 1 (M, M∗ ) .
This says that every divisor D on M determines a holomorphic line bundle O(D),
and the line bundle O(D) is holomorphically trivial if and only if D is the di-
visor of a meromorphic function. The quotient H 0 (M, M∗ /O∗ )/H 0 (M, M∗ ) is
called the Cartier divisor class group and is denoted by CaCl(M ). Furthermore,
H 1 (M, M∗ ) = 0 if and only if every holomorphic line bundle on M has a global
meromorphic section. In this case we get an isomorphism between the divisor class
group and the Picard group. This happens, for example, for smooth projective
algebraic varieties, that is
Proposition 3.5.20: Let X be a smooth projective algebraic variety. Then Cl(X) ≈
Pic(X).
We now briefly describe the intersection theory for divisors. The general theory
is laid out in [Ful84], but our treatment follows closely the abbreviated version in
[Laz04a]. Here we work in the category of varieties (either affine or projective).
Let X be an irreducible variety. To each Cartier divisor D ∈ H 0 (X, M∗ /O∗ ) we
associate a line bundle O(D), and to this line bundle we can associate a Chern
class c1 (O(D)) ∈ H 2 (X, Z). Now let C ⊂ X be an irreducible curve, that is a one
dimensional irreducible complex subvariety of X, and denote by [C] its homology
class in H2 (X, Z). Then
Definition 3.5.21: The intersection number D · C ∈ Z of a Cartier divisor D
with an irreducible curve C is defined by
Z
D·C = D = hc1 (O(D)), [C]i ∈ Z ,
C
where h·, ·i denotes the Kronecker pairing.
This definition can easily be generalized as follows: let V be a k-dimensional
subvariety of X, and let D1 , . . . , Dk be Cartier divisors on X. Then we define the
intersection number by
(3.5.8) D1 · · · · · Dk · V = hc1 (O(D1 )) ∪ · · · ∪ c1 (O(Dk )) , [V ]i
Definition 3.5.22: We say that two Cartier divisors D, D0 are numerically equiv-
alent, denoted by D0 ≡n D if D0 ·C = D·C for all irreducible curves C. A divisor D
is numerically trivial if it is numerically equivalent to 0. We denote by Num(X)
the subgroup of H 0 (X, M∗ /O∗ ) consisting of numerically trivial Cartier divisors.
We now have another characterization of the Neron-Severi group, namely as
the quotient group of numerical classes of Cartier divisors,
N S(X) = H 0 (X, M∗ /O∗ )/Num(X) .
102 3. KÄHLER MANIFOLDS

We shall also make use of the following terminology [Laz04a].


Definition 3.5.23: Let X be a complete variety of dimension n. A (Cartier) divi-
sor D is said to be numerically effective or nef if for every irreducible curve C
we have D · C ≥ 0. The divisor D is big if the Iataka dimension κ(X, O(D)) = n.
An important criterion for the bigness of a nef divisor is: a nef divisor D is big
if and only if Dn > 0.

3.6. The Calabi Conjecture and the Calabi-Yau Theorem


The Ricci form ρω defined in Proposition 3.2.6 plays a special role in Kähler
geometry. Indeed, from Equation 3.2.1 and Definition/Proposition 3.4.2 we have
Proposition 3.6.1: Let (M, J, g, ωg ) be a Kähler manifold and ρω the Ricci form
of g. Then
2
[ρω ] = 2πc1 (M ) ∈ HDR (M ) ∩ H 1,1 (M, C).
This proposition says that on a Kähler manifold the Ricci curvature 2-form ρω
of any Kähler metric represents the cohomology class 2πc1 (M ). The well-known
Calabi Conjecture [Cal56, Cal57] is the question whether or not the converse is
also true, that is, the assertion that on a compact complex manifold any real closed
(1, 1)-form that represents 2πc1 (M ) is the Ricci form of a unique Kähler metric g on
M. This conjecture was proven in the celebrated work of Yau which was announced
in [Yau77] and the details of which appeared in [Yau78]. (Uniqueness was actually
proved by Calabi [Cal57] before the conjecture was settled). By now there are
several excellent treatments in book form [Aub82, Aub98, Joy00, Siu87, Tia00].
While these proofs all assume one is working with compact Kähler manifolds, the
proofs readily extend to the case of compact Kähler orbifolds since the computations
are local in nature and one needs only to consider local functions and sections on the
local uniformizing neighborhoods that are invariant under the local uniformizing
groups.
To be more specific we begin with a couple of definitions
Definition 3.6.2: Let (M, J, g, ωg ) be a compact Kähler manifold. The Kähler
cone of M
K(M ) = {[ω] ∈ H 1,1 (M, C) ∩ H 2 (M, R) | ω = ωh for some Kähler metric h} ,
i.e., it is the set of all possible Kähler classes on M .
It is easy to show that K(M ) is a convex open set in H 1,1 (M, C) ∩ H 2 (M, R).
Of particular interest to us is the Kähler lattice KL (M ) which is the intersection
of K(M ) with the Neron-Severi group, i.e., KL (M ) = K(M ) ∩ N S(M ). Beyond
surfaces, not much is known about the Kähler cone. However, recent progress has
been made by Demailly and Paun [DP04] who characterize K(M ) as a connected
component of the set of (1, 1) cohomology classes which are numerically positive on
all analytic cycles. This generalizes the well-known Nakai-Moishezon criterion for
ample line bundles.
Definition 3.6.3: Let (M, J, g, ωg ) be a compact Kähler manifold and K(M ) its
Kähler cone. For any fixed Kähler class [ω] ∈ K(M ) we define
K[ω] = {h ∈ Γ(¯2 (T M )) | h is a Kähler metric and [ω] = [ωh ]}
to be the space of all Kähler metrics in a given cohomology class.
3.6. THE CALABI CONJECTURE AND THE CALABI-YAU THEOREM 103

¯
The global i∂ ∂-lemma provides for a very simple description of the space of
Kähler metrics K[ω] . Suppose we have a Kähler metric g in a with Kähler class
[ωg ] = [ω] ∈ K(M ). If h ∈ K[ω] is another Kähler metric then, up to a constant,
¯ We could
there exists a global function φ ∈ C ∞ (M, R)R such that ωh − ωg = i∂ ∂φ.
fix constant by requiring, for example, that M φdvolg = 0. Hence, we have
Corollary 3.6.4: Let (M, J, g, ωg ) be a compact Kähler manifold with [ωg ] = [ω] ∈
K(M ). Then, relative to the metric g the space of all Kähler metric is K[ω] can be
described as
¯ > 0, R
K[ω] = {φ ∈ C ∞ (M, R) | ωh = ωg + i∂ ∂φ M
φdvolg = 0} ,
where the 2-form ωh > 0 means that ωh (X, JY ) is a Hermitian metric on M .
The following theorem is the statement of the famous conjecture made by Calabi
and proved by Yau in 1977-78 [Yau77, Yau78].
Theorem 3.6.5: Let (M, J, g, ωg ) be a compact Kähler manifold. Then any real
(1, 1)-form ρ on M which represents the cohomology class 2πc1 (M ) is the Ricci
form of a unique Kähler metric h such that [ωh ] = [ωg ].
¯
Let us reformulate the problem using the global i∂ ∂-lemma. We start with a
given Kähler metric g on M in Kähler class [ωg ] = [ω]. Since ρg also represents
2πc1 (M ) there exists a globally defined function f ∈ C ∞ (M, R) such that
¯ .
ρg − ρ = i∂ ∂f
Appropriately, f could be called a discrepancy potential
R function for the Calabi
problem and we could fix the constant by asking that M (ef − 1)dvolg = 0.
Now suppose that the desired solution of the problem is a metric h ∈ K[ω] . We
know that the Kähler form of h can be written as
¯ ,
ωh = ωg + i∂ ∂φ
for some smooth function φ ∈ C ∞ (M, R). We normalize φ as in previous corollary.
Combining these two equations we see that
¯ .
ρh − ρg = i∂ ∂f
If we define a smooth function F ∈ C ∞ (M, R) relating the volume forms of the two
metrics dvolh = eF dvolg then the left-hand side of the above equation takes the
following form
¯ = ρh − ρg = i∂ ∂φ
i∂ ∂F ¯ ,

and, ¯ − f ) = 0. Hence, F = f + c. But since we normalized


hence, simply i∂ ∂(F
R
M
(e − 1)dvolg = 0 we must have c = 0. Hence, F = f, or dvolh = ef dvolg . We
f

can now give two more equivalent formulations of the Calabi Problem.
Theorem 3.6.6: Let (M, J, g, ωg ) be a compact Kähler manifold, [ωg ] = ω ∈ K(M )
the corresponding Kähler class and ρg the Ricci form. ConsiderR any positive (1, 1)-
¯ with
form Ω on M such that [Ω] = 2πc1 (M ). Let ρg −Ω = i∂ ∂f, (ef −1)dvolg = 0.
M
(i) There exists a unique Kähler metric h ∈ K[ω] whose volume form satisfies
dvolh = ef dvolg .
(ii) Let (U ; z1 , . . . , zn ) be a local complex chart on M with respect to which the
metric g = (gij̄ ). Then, up to a constant, there exists a unique smooth
104 3. KÄHLER MANIFOLDS

function φ in K[ω] , which satisfies the following equation


³ 2
´
det gij̄ + ∂z∂i ∂φz̄j
= ef .
det(gij̄ )
The equation in (ii) is called the Monge-Ampère equation. Part (i) gives a
very simple geometric characterization of the Calabi-Yau theorem. On a compact
Kähler manifold one can always find a metric with arbitrarily prescribed volume
form. The uniqueness part of this theorem was already proved by Calabi. This
part involves only the Maximum Principle. The existence proof uses continuity
methods and it involves several difficult a priori estimates. These were found by
Yau [Yau78] in 1978. For the details we refer the reader to the books mentioned
previously. Here we only list some immediate consequences of this theorem:
Corollary 3.6.7: Let (M, J, g, ωg ) be a compact Kähler manifold with c1 (M ) = 0.
Then M admits a unique Kähler Ricci-flat metric.
Such a metric has holonomy group inside SU (n). Kähler manifolds with this
property are called Calabi-Yau manifolds.
Corollary 3.6.8: Let (M, J, g, ωg ) be a compact Kähler manifold with c1 (M ) > 0.
Then M admits a Kähler metric of positive Ricci curvature.
We can combine this corollary with an older result of Kobayashi [Kob61] to
give a beautiful proof of
Theorem 3.6.9: Any Fano manifold is simply connected.
−1
Proof. Let M be a Fano manifold. Then its anticanonical line bundle KM is
−1
positive, so c1 (M ) = c1 (KM ) > 0. Thus, by Corollary 3.6.8 M has a Kähler metric
of positive Ricci curvature. By Myers’ Theorem M has finite fundamental group.
Let M f be the universal cover of M. It is compact, Fano and a finite, say d-fold,
cover of M. By the Kodaira-Nakano Vanishing Theorem 3.5.8 H q (M f, O) = 0 for
all q > 0 which implies that the holomorphic Euler characteristic χ(Mf, O) = 1. On
f
the other hand, χ(M , O) = dχ(M, O) = d implying that d = 1. So M is simply
connected. ¤
We shall return to important consequences of Yau’s Theorem in Chapter 5.
CHAPTER 4

Fundamentals of Orbifolds

The notion of orbifold was introduced under the name V-manifold by Sa-
take [Sat56] in 1956, and subsequently he developed Riemannian geometry on
V-manifolds [Sat57] ending with a proof of the Gauss-Bonnet theorem for V-
manifolds. Contemporaneously, Baily introduced complex V-manifolds and gen-
eralized both the Hodge decomposition theorem [Bai56], and Kodaira’s projective
embedding theorem [Bai57] to V-manifolds. Somewhat later in the late 1970’s
and early 1980’s Kawasaki generalized various index theorems [Kaw79, Kaw78,
Kaw81] to the category of V-manifolds. It was about this time that Thurston
[Thu79] rediscovered the concept of V-manifold, under the name of orbifold, in his
study of the geometry of 3-manifolds, and defined the orbifold fundamental group
π1orb . By now orbifold has become the accepted term for these objects and we shall
follow suit. However, we do use the name V-bundle interchangeably with orbibun-
dle for fibre bundles in this category. At this stage we can also mention the recent
book [ALR06] which we refer to for certain foundational results, especially from
the groupoid point of view. Although we start with a very general setting, the real
focus of this chapter is on complex and specifically on Kähler orbifolds as they are
be of crucial importance in understanding the quasi-regular Sasakian structures of
Chapter 7. However, more general orbifold structures, such as quaternionic Kähler
orbifolds of Chapters 12 and 13, will appear and will be equally important.

4.1. Basic Definitions


Orbifolds arise naturally as spaces of leaves of Riemannian foliations with com-
pact leaves, and we are particularly interested in this point of view. Conversely,
every orbifold can be realized in this way. In fact, given an orbifold O, we can
construct on it the V-bundle of orthonormal frames whose total space P is a
smooth manifold with a locally free action of the orthogonal group O(n) such that
O = P/O(n). Thus, every orbifold can be realized as the quotient space by a lo-
cally free action of a compact Lie group. The definition given here is close to
Satake’s original definition [Sat56, Sat57] (cf. [Bai56]), but with more modern
terminology as in Moerdijk and Pronk [MP97]. Here F denotes either R or C.
Definition 4.1.1: Let X be a paracompact Hausdorff space. An orbifold chart
or local uniformizing system on X is a triple (Ũ , Γ, ϕ), where Ũ is connected
open subset of Fn containing the origin, Γ is a finite group acting effectively as
diffeomorphisms of Ũ , and ϕ : Ũ −−→U is a continuous map onto an open set U ⊂ X
such that ϕ ◦ γ = ϕ for all γ ∈ Γ and the induced natural map of Ũ /Γ onto U is a
homeomorphism. An injection or embedding between two such charts {Ũ , Γ, ϕ}
and {Ũ 0 , Γ0 , ϕ0 } is a smooth embedding λ : Ũ −−→Ũ 0 such that ϕ0 ◦λ = ϕ. An orbifold
atlas on X is a family U = {Ũi , Γi , ϕi } of orbifold charts such that
105
106 4. FUNDAMENTALS OF ORBIFOLDS

(i) X = ∪i ϕi (Ũi ),
(ii) given two charts (Ũi , Γi , ϕi ) and (Ũj , Γj , ϕj ) with Ui = ϕi (Ũi ) and Uj =
ϕj (Ũj ) and a point x ∈ Ui ∩Uj , there exist an open neighborhood Uk of x,
and a chart (Ũk , Γk , ϕk ) such that there are injections λik : (Ũk , Γk , ϕk )−→
(Ũi , Γi , ϕi ) and λjk : (Ũk , Γk , ϕk )−→(Ũj , Γj , ϕj ).
An atlas U is said to be a refinement of an atlas V if there exists an injection of every
chart of U into some chart of V. Two orbifold atlases are said to be equivalent
if they have a common refinement. A smooth orbifold (or V-manifold) is a
paracompact Hausdorff space X with an equivalence class of orbifold atlases. We
denote the orbifold by X = (X, U). If every finite group Γ consists of orientation
preserving diffeomorphisms and there is an atlas such that all the injections are
orientation preserving then the orbifold is orientable.
Remarks 4.1.1: 1) The finite groups Γi are called local uniformizing groups, and it
follows from the usual slice theorem that we can always take them to be finite sub-
groups O(n, F) ⊂ GL(n, F), where O(n, R) = O(n) is the usual orthogonal group,
and O(n, C) = U (n), the unitary group. Of course, in the orientable case the local
uniformizing groups lie in SO(n). The condition that the local uniformizing groups
act effectively is not always imposed in the literature, and there are occasions when
this is too restrictive. However, in our situation we are almost always concerned
with ‘effective orbifolds’1 so it is convenient to incorporate this as part of our defi-
nition. We shall refer to the more general situation when at least one of the local
uniformizing groups does not act effectively as a non-effective orbifold [HM04]. 2)
Every orbifold atlas is contained in a unique maximal atlas satisfying the required
properties. Thus, we shall often think of the orbifold X = (X, U), where U is the
unique maximal atlas. This justifies the notation. 3) Notice that any γ ∈ Γ satisfies
ϕ(γ · x) = ϕ(x) for any x ∈ Ũ . Thus, every γ defines an injection by x 7→ γ · x.
The following lemma was proved by Satake in [Sat57] under the added as-
sumption that the fixed point set has codimension at least two2. However, this
assumption was removed recently by Moerdijk an Pronk [MP97]. We refer to
Proposition 2.12 and Lemma 2.11 of [MM03] for the proof.
Lemma 4.1.2: Let λ1 , λ2 : (Ũ , Γ, ϕ)−→(Ũ 0 , Γ0 , ϕ0 ) be two injections. Then there
exists a unique γ 0 ∈ Γ0 such that λ2 = γ 0 ◦ λ1 .
Let (X, U) be an orbifold and choose a local uniformizing system {U, Γ, ϕ}.
Let x ∈ X be any point, and let p ∈ ϕ−1 (x), then up to conjugacy the isotropy
subgroup Γp ⊂ Γ depends only on x, and accordingly we shall denote this isotropy
subgroup by Γx . We now have
Definition 4.1.3: A point of X whose isotropy subgroup Γx 6= id is called a singu-
lar point. Those points with Γx = id are called regular points. The set of singular
points is called the orbifold singular locus or orbifold singular set, and is
denoted by Σorb (X).
The subset of regular points is an open dense subset of X. The isotropy groups
give a natural stratification of X by saying that two points lie in the same stratum
if their isotropy subgroups are conjugate. Thus, we get a decomposition of X as
orb
(4.1.1) X = Xreg tj Σorb
j (X) , Σorb (X) = tj Σorb
j (X) ,

1Effective orbifolds are sometimes called reduced orbifolds, cf. [CR04].


2The codimension two condition occurred as part of Satake’s original definition of V-manifold.
4.1. BASIC DEFINITIONS 107

where the union is taken over all conjugacy classes. The dense open subset of regular
orb
points Xreg forms the principal stratum corresponding to the trivial conjugacy
class. It is well-known that an orbifold singular point may or may not be an honest
singularity. For example as a variety X may have both smooth points and singular
points. The set of smooth or regular points Xreg is generally larger than the set
orb
of orbifold regular points, i.e., Xreg ⊂ Xreg . Likewise, the orbifold singular locus
Σ (X) is larger than the singular locus Σ(X) as a variety, Σ(X) ⊂ Σorb (X). This
orb

is discussed in Section 4.4 below.


One way to construct an orbifold is from a group Γ acting properly discontin-
uously on a manifold M. The quotient then has the structure of an orbifold in a
natural way [Sat57]. To see this consider a point p ∈ M. The isotropy subgroup
Γp ⊂ Γ is finite since the action is properly discontinuous. Moreover, there is a
neighborhood Ũp of each point p such that γ Ũp ∩ Ũp = ∅ if γ 6∈ Γp and γ Ũp = Ũp
if γ ∈ Γp . Then the natural projection ϕ : Ũp −−→Ũp /Γp = Up is a local homeo-
morphism satisfying ϕ ◦ γ = ϕ for γ ∈ Γp . Construct a cover {Up } for M/Γ and
local uniformizing systems of the form {Ũ , Γ, ϕ} by adding when necessary finite
intersections of the form Up1 ∩· · ·∩Upk to guarantee that condition (ii) of Definition
4.1.1 holds. The quotient M/Γ, which is always Hausdorff, is then an orbifold.
Definition 4.1.4: An orbifold X = (X, U) is said to be developable3 if X = M/Γ
for some discrete group Γ acting properly discontinuously on M, and the orbifold
atlas U is constructed as above.
Although not every orbifold is developable, it will be proven later in Proposition
4.2.17 that every orbifold can be realized as the quotient of a manifold by a compact
Lie group4 G. In this case G acts locally freely so that X is the space of leaves of
a foliation F, and the isotropy subgroup Γx is precisely the leaf holonomy group of
the leaf x. (See Example 2.1.3 below.) This identifies the local uniformizing groups
with the leaf holonomy groups of F. An orbifold X is a smooth manifold or in the
complex analytic category a complex manifold if and only if Γx = id for all x ∈ X.
In this case we can take Γ = id and ϕ = id, and the definition of an orbifold reduces
to the usual definition of a smooth manifold.
Definition 4.1.5: The order υ(X ) of an orbifold X = (X, U) is the least common
multiple of the orders of the isotropy groups when it exists. If there is no such least
common multiple, we say X has infinite order.
Example 4.1.6: Orbifold structures on S 2 . This example is taken from [BH99].
Consider S 2 = C ∪ {∞} as the Riemann sphere with two marked points, the north
pole ∞ and south pole 0, respectively. We construct an orbifold structure on S 2
using two uniformizing charts ϕ0 : C−−→V0 = S 2 \ {∞} and ϕ∞ : C−−→V∞ =
S 2 \ {0} defined by ϕ0 (z) = z m and ϕ∞ (w) = w−n . The injection λ satisfies
m
w = λ(z) = (1/z) n so that ϕ0 (z) = ϕ∞ ◦ λ(z) holds. The uniformizing groups
fixing ∞ and 0, respectively, are cyclic groups Cm , Cn of order m and n, respectively.

3The name good was used by Thurston in [Thu79], but we do not like this terminology,
since from our viewpoint it is precisely Thurston’s ‘bad orbifolds’ that are good in the sense that
they are important and useful to us. On the other hand Thurston’s good orbifolds are of less
importance to us. We prefer the appellation ‘developable’ that was introduced by Bridson and
Haefliger in [BH99].
4It is not known whether this holds for noneffective orbifolds. However, see [HM04] for
recent progress on this issue.
108 4. FUNDAMENTALS OF ORBIFOLDS

This orbifold is developable if and only if m = n in which case it is the quotient


space S 2 /Cm . Note that in all cases the order υ of this orbifold is lcm(m, n).
For orbifold structures on Riemann surfaces of genus greater than 1 we refer
to the constructions in Chapter IIIG of [BH99]. In this very interesting chapter
Bridson and Haefliger discuss orbifolds from the point of view of étale groupoids
(cf. Proposition 4.3.1 below). In particular, they generalize the following result
attributed to Gromov in [BH99].
Proposition 4.1.7: A complete Riemannian orbifold5 of non-positive sectional
curvature is developable.
The notions of structure preserving maps and bundle theory, or more gener-
ally sheaf theory, are fundamental to doing geometry on any object. The stan-
dard notions of smooth maps between orbifolds as well as V-bundles on orb-
ifolds (orbibundles) have been given in an analogous manner to manifolds (see
[Sat56, Sat57, Bai56, Bai57]). However, it wasn’t realized until recently (see
for example [CR04, MP97]) that certain problems arise, namely it does not in
general induce morphisms of sheaves or V-bundles. This led to the notion of “good
map” in [CR04]; however, we follow the development in [MP97] and [Moe02].
Definition 4.1.8: Let X = (X, U) and Y = (Y, V) be orbifolds. A map f : X−−→Y
is said to be smooth if for every point x ∈ X there are orbifold charts (Ũ , Γ, ϕ)
about x and (Ṽ , Φ, ψ) about f (x) such that f maps U = ϕ(Ũ ) into V = ψ(Ṽ )
and there exist local lifts f˜Ũ : Ũ −−→Ṽ satisfying ψ ◦ f˜Ũ = f ◦ ϕ. Furthermore,
f is said to be good if it is compatible with the injections in the sense that if
λij : (Ũj , Γj , ϕj )−−→(Ũi , Γi , ϕi ) is an injection on X , then there is an injection
µ(λij ) : (Ṽj , ϕj , ψj )−−→(Ṽi , ϕi , ψi ) on Y, such that
(i) f˜ ◦ λij = µ(λij ) ◦ f˜ .
Ũi Ũj
(ii) µ(λij ◦ λjk ) = µ(λij ) ◦ µ(λjk ).
We shall often write f : X −−→Y for a smooth map of orbifolds and refer to
this as an orbifold map. It is easy to see that the composition of smooth maps is
smooth, so the collection of orbifolds with smooth maps forms a category. Thus,
we also have
Definition 4.1.9: Two orbifolds X = (X, U ) and Y = (Y, V) are said to be dif-
feomorphic or equivalent if there exist smooth maps f : X−−→Y and g : Y −−→X
such that f ◦ g = 1lY and g ◦ f = 1lX , where 1lX , 1lY are the respective identity maps.
Notice that in particular a diffeomorphism between orbifolds gives a homeo-
morphism of the underlying topological spaces.

4.2. Orbisheaves and orbibundles


We now wish to describe sheaves and bundles over orbifolds. Following [MP97]
we have
Definition 4.2.1: Let X = (X, U) be an orbifold. A sheaf F on X or orbisheaf
on X consists of a sheaf FŨ over Ũ for each orbifold chart (Ũ , Γ, ϕ) such that
for each injection λ : (Ũ , Γ, ϕ)−−→(Ṽ , Φ, ψ) there is an isomorphism of sheaves
F(λ) : FŨ −−→λ∗ FṼ which is functorial in λ.

5For the definition of a Riemannian orbifold, see Definition 4.2.11 below.


4.2. ORBISHEAVES AND ORBIBUNDLES 109

Let F be an orbisheaf on X , and U = (Ũ , Γ, ϕ) an orbifold chart. By 3)


of Remarks 4.1.1 each γ ∈ Γ defines an injection λγ , so there is a sheaf map
F(γ) : FŨ −−→λ∗γ FŨ , and if s ∈ FŨ ,x is a point in the stalk over x ∈ Ũ , then
F(γ)(s) lies in the stalk over γ −1 x. This defines an action of Γ on the sheaf FŨ ,
which says that FŨ is a Γ-equivariant sheaf on Ũ . So every orbisheaf F is equivariant
under the local uniformizing groups Γi . We now have
Definition 4.2.2: The structure orbisheaf OX of an orbifold X is the orbisheaf
defined by the structure sheaf OŨ on each orbifold chart (Ũ , Γ, ϕ).
OX is well defined since each injection λ : (Ũ , Γ, ϕ)−−→(Ũ 0 , Γ0 , ϕ0 ) induces an
isomorphism OŨ ≈ λ∗ OŨ 0 by sending f ∈ OŨ ,x to f ◦ λ−1 ∈ (λ∗ OŨ 0 )x . The
structure sheaves that are of most interest to us are the smooth structure sheaf
when OŨ is the sheaf of C ∞ functions on Ũ for real orbifolds, the holomorphic
structure sheaf when OŨ is the sheaf of holomorphic functions on Ũ for complex
orbifolds, and the regular structure when OŨ is the sheaf of regular functions on
Ũ .
Definition 4.2.3: A morphism of orbisheaves α : F−−→F 0 is a family of sheaf
maps αŨ : FŨ −−→FŨ0 , one for each orbifold chart, that is compatible with the injec-
tions λ in the sense that for each injection the diagram
αŨ
FŨ −−−−−−−−−→ FŨ0
 
 
F (λ) F 0 (λ)
 
y y
λ∗ αṼ
λ∗ FṼ −−−−−−−−−→ λ∗ FṼ0
commutes.
So the collection of all sheaves of sets, Abelian groups, rings, etc. forms a cate-
gory6 X̃ that we call the category of orbisheaves on X . It is important to realize that
X̃ is quite different from the category of sheaves on X unless the orbifold structure
is trivial, i.e. all the local uniformizing groups Γi are the identity. Nevertheless,
one can identify sheaves with orbisheaves.
Lemma 4.2.4: Let X = (X, U) be an orbifold. Then a sheaf FX on X induces
an orbisheaf, denoted by FX , on X by pulling back the restrictions FUi to the local
uniformizing neighborhoods Ũi . Conversely, an orbisheaf FX on X induces a sheaf
FX on X by locally taking the direct image sheaves of the subsheaf of Γ-invariant
sections on the local uniformizing neighborhoods.
Proof. This follows directly from the definitions. Given an orbifold atlas
U = {Ũi , Γi , ϕi } and an injection λji : {Ũi , Γi , ϕi }−−→{Ũj , Γj , ϕj }, there are sheaves
FŨi = ϕ∗i FUi and FŨj = ϕ∗j FUj on X . We have Ui ⊂ Uj , so
λ∗ji FŨj = λ∗ji ϕ∗j FUj ≈ (ϕj ◦ λji )∗ (FUj |Ui ) = ϕ∗i FUi = FŨi .
Similarly, it is easy to check that this is functorial in λ.
To describe the converse consider an orbisheaf F on X . This gives a sheaf FŨ
on each local uniformizing neighborhood (Ũ , Γ, ϕ). Then we get a sheaf F on X
6Actually this is a special type of topos called a smooth étendue by Grothendieck and Verdier
[AGV72], but we shall not use this terminology.
110 4. FUNDAMENTALS OF ORBIFOLDS

by taking the direct image of the Γ-invariant sections, that is, F is the sheaf on X
associated to the presheaf U 7→ H 0 (ϕ−1 (U ), FŨ )Γ . ¤

Any map f˜ : X̃ −−→Ỹ of orbisheaves sends OX to OY , and thus induces a mor-


phism f : X −−→Y. However, the converse is not true. We need a good orbifold map
as defined by Definition 4.1.8 to induce a morphism of orbisheaves. Nevertheless,
the following does hold [MP97]:
Proposition 4.2.5: A diffeomorphism of orbifolds X ≈ Y induces an equivalence
of orbisheaves X̃ ≈ Ỹ.

Proof. See [MP97] for details. ¤

The structure sheaf OX of an orbifold X is a sheaf of commutative rings on each


local uniformizing neighborhood Ũi , and we can construct an orbisheaf M of OX
modules on X by giving a sheaf MŨ of OŨ modules on each orbifold chart (Ũ , Γ, ϕ).
r times
As usual for a sheaf FŨ on Ũ we denote by (FŨ )r the direct sum FŨ ⊕ · · · ⊕ FŨ .
Definition 4.2.6: An orbisheaf M of OX modules on X is said to be locally
V-free or just locally free if each point x ∈ X has an orbifold chart such that
MŨ ≈ (OŨ )r for some positive integer r, called the rank of M. An orbisheaf of
rank one is called invertible orbisheaf or a V-invertible sheaf.
We use the terms locally free and locally V-free interchangeably when referring
to orbisheaf. The terms V-free and V-invertible are used mainly for emphasis since
a sheaf induces an orbisheaf which is generally a different structure.
Many of the usual differential geometric concepts that hold for smooth or com-
plex analytic manifolds also hold in the orbifold category, in particular the impor-
tant notion of a fiber bundle.
Definition 4.2.7: A V-bundle or orbibundle over an orbifold X = (X, U) con-
sists of a fiber bundle BŨ over Ũ for each orbifold chart (Ũi , Γi , ϕi ) ∈ U with Lie
group G and fiber F a smooth G-manifold which is independent of Ũi together with
a homomorphism hŨi : Γi −−→G satisfying:
(i) If b lies in the fiber over x̃i ∈ Ũi then for each γ ∈ Γi , bhŨi (γ) lies in the
fiber over γ −1 x̃i .
(ii) If the map λji : Ũi −−→Ũj is an injection, then there is a bundle map
λ∗ji : BŨj |λji (Ũi ) −−→BŨi satisfying the condition that if γ ∈ Γi , and
γ 0 ∈ Γj is the unique element such that λji ◦ γ = γ 0 ◦ λji , then hŨi (γ) ◦
λ∗ji = λ∗ji ◦ hŨj (γ 0 ), and if λkj : Ũj −−→Ũk is another such injection then
(λkj ◦ λji )∗ = λ∗ji ◦ λ∗kj .
If the fiber F is a vector space of dimension r and G acts on F as linear transfor-
mations of F, then the V-bundle is called a vector V-bundle of rank r. Similarly,
if F is the Lie group G with its right action, then the V-bundle is called a principal
V-bundle.
Remark 4.2.1: Both the definitions of orbisheaves and orbibundles (V-bundles)
consist of a sequence sheaves or bundles defined on the disjoint union Ũ = ti Ũi of
the local uniformizing neighborhoods that satisfy certain compatibility conditions
with respect to the action of the local uniformizing groups and the injections.
The space Ũ is the space of objects of a certain étale groupoid, and this plays
4.2. ORBISHEAVES AND ORBIBUNDLES 111

an important role when describing invariants of an orbifold. We discuss this in


detail in the next section.
We now describe the total space E of a V-bundle over X, and see that it inherits
a (generally) non-effective orbifold structure. We construct an orbifold structure
on E with local uniformizing systems {BŨi , Γ∗i , ϕ∗i } by gluing together the sets
BŨj /Γ∗i . By choosing the local uniformizing neighborhoods of X small enough, we
can always take BŨi to be the product Ũi × F which we shall heretofore assume.
The action of the local uniformizing group Γi on Ũi extends to an action on Ũi × F
by sending (x̃i , b) ∈ Ũi × F to (γ −1 x̃i , bhŨi (γ)), so the local uniformizing groups Γ∗i
are defined to be subgroups of Γi that stabilize the point (xi , b). We are particularly
interested in the case of a principal bundle in which case the fibre is a Lie group G,
so the image hŨi (Γ∗i ) acts freely on F. Then we have
Lemma 4.2.8: Let G be a Lie group and P the total space of a principal V-bundle
over an orbifold X = (X, U ). Then P is a smooth manifold if and only if the maps
hŨi are injective for all i.
Remark 4.2.2: We shall often denote a V-bundle by the standard notation π :
P −−→X and think of this as an “orbifold fibration”. It must be understood, however,
that an orbifold fibration is not a fibration in the usual sense. Shortly, we shall
show that it is a fibration rationally. As in the case of orbisheaves a smooth map
between orbifolds does not in general induce a bundle map except in the case
of diffeomorphisms as in Proposition 4.2.5. For more general maps we need to
assume that the map is good as in Definition 4.1.8. However, isomorphism classes
of V-bundles can be described in the usual manner. We leave this description to
the reader. An absolute V-bundle resembles a bundle in the ordinary sense, and
corresponds to being able to take hŨ = id, for all local uniformizing neighborhoods
Ũ . In particular, the trivial V-bundle X × F is absolute. Another important notion
introduced by Kawasaki [Kaw78] is that of proper. A V-bundle E is said to be
proper if the local uniformizing groups Γ∗i of E act effectively on X when viewed
as subgroups of the local uniformizing groups Γi on X. Any V-bundle with smooth
total space is clearly proper. The Kawasaki index theorems such as his Riemann-
Roch Theorem used in section 3.2 require the V-bundles to be proper.
The notion of sections of bundles works just as well in the orbifold category.
Definition 4.2.9: Let E = (E, U ∗ ) be a V-bundle over an orbifold X . Then a
section σ of E over an open set W ⊂ X is a section σŨ of the bundle BŨ for each
orbifold chart (Ũ , Γ, ϕ) such that U ⊂ W and for any x̃ ∈ Ũ we have
(i) For each γ ∈ Γ, σŨ (γ −1 x̃) = hŨ (γ)σŨ (x̃).
(ii) If λ : (Ũ , Γ, ϕ)−−→(Ũ 0 , Γ0 , ϕ0 ) is an injection, then λ∗ σŨ 0 (λ(x̃)) = σŨ (x̃).
If each of the local sections σŨ is continuous, smooth, holomorphic, etc., we say
that σ is continuous, smooth, holomorphic, etc., respectively. Given local sections
σŨ of a vector V-bundle or orbisheaf we can always construct Γ-invariant local
sections by “averaging over the group”, i.e., we define an invariant local section by
I 1 X
(4.2.1) σŨ = σŨ ◦ γ .
|Γ|
γ∈Γ

Notice that this determines a well defined map from the underlying space, namely
σU : U = ϕ(Ũ )−−→BŨ . Using invariant local sections over each orbifold chart we
112 4. FUNDAMENTALS OF ORBIFOLDS

obtain global invariant sections. We often work with these invariant sections, and
view objects on an orbifold interchangeably as invariant objects on Ũ or objects on
U. Conversely, given a section σ of a sheaf FX over the regular subspace Xreg we
can extend it to a Γi invariant section over each local uniformizing neighborhood
Ũi . This gives a section of the corresponding orbisheaf FX .
The standard notions of tangent bundle, cotangent bundle, and all the associ-
ated tensor bundles all have V-bundle analogues [Bai56, Sat56, Sat57].
Example 4.2.10: [Tangent V-bundle of X ] Given a uniformizing chart (Ũi , Γi , ϕi )
in U , we take BŨi = T Ũi the tangent bundle of the open submanifold Ũi ⊂
Rn . Let (yi1 , . . . , yin ) denote the local coordinates of Ũi . The fibre F is Rn with
its coordinate basis {∂/∂yik } at each point p̃ ∈ Ũi , and for each injection λji :
(Ũi , Γi , ϕi )−−→(Ũj , Γj , ϕj ) we have the Jacobian matrix

³ ´
∂yjk ◦λji
Jλji = ∂yil

which satisfies Jλji (p̃) ∈ GL(n, R) for each p̃ ∈ Ũi , and its inverse is the transition
function for the bundle BŨi . So G = GL(n, R). Since, as mentioned previously,
each local uniformizing group Γi defines an injection, the homomorphisms hŨi are
injective, and satisfy condition i) of Definition 4.2.7. Furthermore, the injections
define a bundle map satisfying condition ii) of the definition. This defines the
tangent V-bundle T X = (T X, U ∗ ), where T X is the total space of the V-bundle
and the orbifold atlas U ∗ has orbifold charts of the form (Ũi × Rn , Γ∗i , ϕ∗i ), where
the local uniformizing group Γ∗i is Γi acting on Ũi ×Rn by (x̃, v) 7→ (γ −1 x̃, hŨi (γ)v),
and ϕ∗i : Ũi × Rn −−→(Ũi × Rn )/Γ∗i is the natural quotient projection. Thus, the
tangent V-bundle of an orbifold is always proper. A smooth invariant section of
T X is called a vector field on X.
The cotangent V-bundle and the tensor bundles are constructed similarly.
Thus, one easily constructs Riemannian metrics, symplectic 2-forms, connections,
etc. such that the injections are maps of the corresponding G-structure. In partic-
ular, the objects are invariant under the local uniformizing groups. For example,
Definition 4.2.11: A Riemannian metric g on an orbifold X = (X, U) is a Rie-
mannian metric gi on each local uniformizing neighborhood Ũi that is invariant un-
der the local uniformizing group Γi , and the injections λji : (Ũi , Γi , ϕi )−−→(Ũj , Γj , ϕj )
are isometries, i.e., λ∗ji (gj |λji (Ui ) ) = gi . Similarly, if X = (X, U) is a complex orb-
ifold, then an Hermitian metric h is a Γi -invariant Hermitian metric on each
neighborhood Ũi such that the injection maps are Hermitian isometries. An orb-
ifold with a Riemannian (Hermitian) metric is called a Riemannian (Hermitian)
orbifold.
Then a slight modification of usual partition of unity argument gives (cf.
[MM03] for details of the proof):
Proposition 4.2.12: Every orbifold admits a Riemannian metric, and every com-
plex orbifold admits an Hermitian metric.
Remark 4.2.3: If an orbifold structure is given on a smooth manifold M, then an
orbifold Riemannian metric is not necessarily a Riemannian metric on M.
4.2. ORBISHEAVES AND ORBIBUNDLES 113

Integration theory also goes through. In particular, if V is an open subset of


ϕ(Ũ ) then the integral of an n-form (measurable) σ is defined by
Z Z
1
(4.2.2) σ= σ .
V |Γ| ϕ−1 (V ) Ũ
So all of the standard integration techniques, such as Stokes’ theorem, hold on
orbifolds.
Example 4.2.13: [V-bundle or orbibundle L(X ) of linear frames of X ] Con-
tinuing with the notation of Example 4.2.10, we now put F = GL(n, R). The
local bundle BŨ is now just the frame bundle on the local uniformizing neighbor-
hood Ũ . The linear frame V-bundle L(X ) = (L(X), U ∗ ) has orbifold charts of the
form (Ũi × GL(n, R), Γ∗i , ϕ∗i ), where the local uniformizing group Γ∗i is Γi acting
on Ũi × GL(n, R) by (x̃, A) 7→ (γ −1 x̃, AhŨi (γ)), and ϕ∗i : Ũi × GL(n, R)−−→(Ũi ×
GL(n, R))/Γ∗i is the natural quotient projection. In this case Γ∗i acts freely on the
second factor, so the action of the local uniformizing groups Γ∗i of the linear frame
V-bundle is always free. Thus, the total space L(X) of the linear frame V-bundle
L(X ) of any orbifold X is a smooth manifold. The right action of GL(n, R) on
L(X) is only locally free with the uniformizing groups Γi as the isotropy group.
We wish to describe reduction of principal orbibundles, but first we need the
notion of an orbibundle map.
Definition 4.2.14: A V-bundle (orbibundle) map α : E−−→E 0 consists of a
family of bundle maps αŨ : BŨ −−→BŨ0 , one for each orbifold chart, that is compatible
with the injections λ : (Ũ , Γ, ϕ)−−→(Ṽ , Φ, ψ) and λ0 : (Ũ 0 , Γ0 , ϕ0 )−−→(Ṽ 0 , Φ0 , ψ 0 ) in
the sense that the diagram
αṼ
BṼ −−−−−−−−−→ BṼ0 0
 
 
λ∗ λ0∗
 
y y
αŨ
BŨ −−−−−−−−−→ BŨ0 0
commutes.
By refining the orbifold charts if necessary we can take Ũ such that BŨ ≈ Ũ ×F.
1
Then by identifying BŨ with Ũ × F and letting αŨ denote the map αŨ on the first
factor, we see that Definition 4.2.14 implies αṼ ◦ λ = λ0 ◦ αŨ
1 1
. To define a reduction
of a principal orbibundle to a subgroup, we require that the diagram (1.2.3) on
the uniformizing neighborhoods patch together to give an orbibundle map. Let us
describe this on the level of the principal frame orbibundle, as the generalization
to arbitrary principal orbibundles is straightforward. Let ι : G−−→GL(n, R) be a
subgroup and consider the corresponding subbundle QŨ of the linear frame bundle
LŨ on the uniformizing neighborhood Ũ of an orbifold X . Then diagram (1.2.3)
becomes
αŨ
QŨ −−−−−→ LŨ
 
 
(4.2.3) πQ πL
y y
id
Ũ −−−−−→ Ũ ,
114 4. FUNDAMENTALS OF ORBIFOLDS

where αŨ satisfies αŨ (ua) = αŨ ι(a). Combining this diagram with the one of
Definition 4.2.14 gives the commutative diagram:

αŨ
(4.2.4) 6 QŨ /6 LŨ
λ∗ mmmmm mmmmm
Q mm
m mmm
mmmmm mmmmmλ∗L
mmm α mmm
QṼ Ṽ
/ LṼ
πQ πL

πQ ² πL ²
m Ũ ________ ____id_____________/ Ũ
m mm
mmm mmm
mmmmm m
λ mmm
m
mm λ mm
² vmmmmm id
² vmmmmm
Ṽ / Ṽ

This then together with Example 4.2.13 allows us to define the notion of a G-
structure on an orbifold.
Definition 4.2.15: Let X be an orbifold and G be a subgroup of GL(n, R). Then a
G-structure on an orbifold X is a reduction of the linear frame orbibundle L(X )
to the subgroup G such that the natural homomorphism is an orbibundle map.
The notion of integrable G-structure works equally well on orbifolds X . It simply
means that the G-structure on each local uniformizing neighborhood Ũ is integrable.
Exercise 4.1: Formulate the reduction process for general principal orbibundles
and state and prove the orbifold analog of Theorem 1.2.5.
Example 4.2.16: [The orthonormal frame orbibundle] The discussion in Ex-
ample 1.4.7 carries over to the orbifold case with the obvious changes. As in that
case a reduction of the linear frame orbibundle to the orthogonal group is equiva-
lent to a choice of an orbifold Riemannian metric. That this can always be done
follows from a standard partition of unity argument. Since the local uniformizing
groups Γi can be taken to be subgroups of the orthogonal group O(n), the total
space LO(X) of the orthonormal frame V-bundle is also a smooth manifold.
This example immediately implies the following result of Satake.
Proposition 4.2.17: Every orbifold X = (X, U) can be presented as a quotient
space of a locally free action of a compact Lie group G on a manifold M.
This proposition implies that every orbifold can be realized as the space of leaves
of a foliation. Of course, generally the converse is not true; Molino’s Theorem 2.5.11
gives conditions under which a converse holds.
The usual relation between locally free sheaves and vector bundles also holds
for orbifolds. We consider both real and complex vector V-bundles and refer to
them as F-vector V-bundles, where F = R or C.
Proposition 4.2.18: There is a one-to-one correspondence between isomorphism
classes of F-vector V-bundles on X and isomorphism classes of locally V-free or-
bisheaves on X . In particular, F line V-bundles correspond to invertible orbisheaves.
4.3. GROUPOIDS, ORBIFOLD INVARIANTS AND CLASSIFYING SPACES 115

Proof. Given a rank r vector V-bundle E over an orbifold X we construct


the sheaf FŨ of local section of the local bundle BŨ . By the usual result FŨ is a
locally free sheaf of rank r. Then the bundle map λ∗ associated with each injection
λ : (Ũ , Γ, ϕ)−−→(Ṽ , Φ, ψ) gives an isomorphism F(λ) : FŨ −−→λ∗ FṼ that is functorial
in λ. This gives a locally V-free orbisheaf F.
Conversely, if F is a locally V-free orbisheaf then on each local uniformizing
neighborhood Ũ , the locally free sheaf FŨ corresponds to the sheaf of sections of
some vector bundle BŨ on Ũ . The transition functions gij : Ũi ∩ Ũj −−→GL(r, R) are
obtained from the isomorphisms gi : FŨi −−→OŨ r
by setting gij = gi ◦ gj−1 . ¤
i

4.3. Groupoids, Orbifold Invariants and Classifying Spaces


In this section we describe certain orbifold invariants which are important to us
since they give rise to certain invariants of deformation classes of Sasakian struc-
tures. For this description it is convenient to use the language of groupoids. It
turns out that orbifolds can be conveniently represented by étale groupoids. The
relation between orbifolds and groupoids was first elucidated by Haefliger in his
seminal paper [Hae84], and more recently developed by Moerdijk and collabora-
tors, [MP97, Moe02, MM03]. Again we refer the reader to the Appendix A.1
for basic facts about groupoids. Haefliger’s point of departure is his groupoid of
germs discussed in Proposition 1.5.4.
Consider an orbifold X = (X, U ) with a given orbifold atlas U . Let Ũ = ti Ũi
denote the disjoint union of the Ũi , and consider the pseudogroup PX of local
diffeomorphisms of Ũ generated by the injections λji and their inverses. The action
of PX on Ũ defines an equivalence relation on Ũ by saying that x ∼ y if they lie on
the same orbit, that is, if there exists an f ∈ PX such that y = f (x). The quotient
space Ũ /PX is identified with the underlying space X of the orbifold. To see this
we define the projection map ϕ : Ũ −−→X to be the union of the maps ϕi , and
notice that xi ∈ Ũi , and xj ∈ Ũj satisfies ϕ(xi ) = ϕ(xj ) if and only if there exists
an injection λji : (Ũi , Γi , ϕi )−−→(Ũj , Γj , ϕj ) such that ϕ ◦ λji = ϕ. So we let G(U)
denote the groupoid of germs of diffeomorphisms associated to PX as in Proposition
1.5.4. G(U) is a proper effective étale Lie groupoid with space of objects G0 = Ũ
and space of arrows G the set of germs of the injections. Thus, we have identified
X with the orbit space G0 /G. In this sense we say that the orbifold X = (X, U) is
represented by the groupoid G(U). In fact we have [MM03]
Proposition 4.3.1: Let X = (X, U) be an orbifold with a fixed atlas U, then G(U)
is a proper effective étale groupoid. Furthermore, if X 0 = (X 0 , U 0 ) is another such
orbifold, then G(U ) and G(U 0 ) are Morita equivalent if and only if the orbifolds X
and X 0 are isomorphic.
Conversely, given a proper effective étale Lie groupoid G, the isotropy subgroup
Gxx at any point x ∈ G0 is finite, and x has an open neighborhood Ux with an effective
action of Gxx which induces an isomorphism of étale groupoids [MM03]
G|U ≈ Ux o Gxx ,
where the right hand side is the action groupoid of Definition A.1.5. Now we can
construct a canonical orbifold structure on the space of orbits G0 /G as follows: we
can choose the neighborhood Ux of x so that it is diffeomorphic to an open subset
Ũx of Rn via a diffeomorphism φx : Ux −−→Ũx . Let π : G0 −−→G0 /G be the natural
116 4. FUNDAMENTALS OF ORBIFOLDS

projection. Then we obtain an atlas U of orbifold charts of the form (Ũx , φx ◦ Gxx ◦
φ−1 −1
x , π ◦ φx ). Now suppose that Vy is contained in Ux with ιy the natural inclusion.
Then the injection

λxy : (Ṽy , ψy ◦ Gyy ◦ ψy−1 , π ◦ ψy−1 )−−→(Ũx , φx ◦ Gxx ◦ φ−1 −1


x , π ◦ φx )

is given by λxy = φx ◦ ιy ◦ ψy−1 . Thus, (G0 /G, U) is an orbifold. We have arrived at


Proposition 4.3.2: Let G be a proper effective étale groupoid. Then G0 /G has a
canonical orbifold structure.
Thus, an orbifold structure on a paracompact Hausdorff space is completely
equivalent to a proper étale Lie groupoid. This gives credence to the name orbifold
groupoid of Definition A.1.6. The groupoid G also defines a Haefliger cocyle by
Definition 2.2.2 which in this case is just the cocycle defined by the cover U together
with the injection maps from double intersections viewed as maps from Ũi ∩ Ũj to
G. This cocycle defines an element {λji } ∈ H 1 (Ũ , G). These cocycles are related to
principal G bundles. We refer to Section 2.2 and [Hae84, GH06] for more detail.
Actually it is only the Morita equivalence class of a groupoid that is pertinent
to our discussion of orbifold invariants. The following theorem due to Moerdijk and
Pronk [MP97, Moe02, MM03] plays an important role relating the holonomy
groupoid of a foliation to orbifolds.
Theorem 4.3.3: Let G be a Lie groupoid. Then the following are equivalent:
(i) G is Morita equivalent to an effective proper étale groupoid, i.e., an orb-
ifold groupoid.
(ii) G is Morita equivalent to the holonomy groupoid of a foliation with com-
pact leaves and finite holonomy groups.
(iii) G is Morita equivalent to an action groupoid M oK, where K is a compact
Lie group acting on M such that each isotropy group Kx is finite and acts
effectively on the normal bundle.
The reader is invited to compare Theorem 4.3.3 dealing with categorical equiv-
alence of groupoids with Molino’s Theorem 2.5.11 about the space of leaves of a
foliation while keeping in mind the relation between proper étale groupoids and
orbifolds. From the sheaf theoretic viewpoint we have, along similar lines, the
following theorem of Moerdijk and Pronk [MP97]:
Theorem 4.3.4: The following are equivalent as categories (ringed topoi):
(i) The topos of orbisheaves on orbifolds.
(ii) The topos of equivariant sheaves on a G-manifold X with finite isotropy
groups and faithful slice representations.
(iii) The topos of sheaves on some effective proper étale groupoid.
We are now ready to describe some important orbifold invariants. As discussed
in Appendix A.2, Haefliger constructed a classifying space BG for any topological
groupoid G. On any orbifold we can choose a Riemannian metric by Proposition
4.2.12. In this case the space of arrows of the étale Lie groupoid G(U) is generated
by the pseudogroup of local isometries. Haefliger [Hae84] gives an explicit Borel-
type construction of the classifying space of an orbifold. Let us look at the linear
frame bundle of Example 4.2.13 of an orbifold X from the point of view of the étale
groupoid G(U ) representing it. We define the orthonormal frame bundle of the
space of units Ũ of G(U) by LO(Ũ ) = ti LO(Ũi ). It is a principal O(n)-bundle over
4.3. GROUPOIDS, ORBIFOLD INVARIANTS AND CLASSIFYING SPACES 117

Ũ . The differential of the groupoid action on Ũ induces an action of the groupoid


G on LO(Ũ ), and as in Example 4.2.13 this action is free. From the definitions one
easily sees that the quotient space LO(Ũ )/G is the orthonormal frame V-bundle
LO(X) of the orbifold which is a smooth manifold as discussed in Example 4.2.13.
Furthermore, LO(X) depends only on X , that is, on the equivalence class of orbifold
atlases, and not on the particular atlas U. Thus, we may also consider LO(Ũ ) to
be a principal G-bundle over LO(X).
Recall that the universal principal O(n)-bundle EO(n)−−→BO(n) whose total
space EO(n) is a contractible space on which O(n) acts freely [Hus66, Mil56a].
So we can construct the associated O(n)-bundle LO(Ũ )×O(n) EO(n). It is an O(n)-
bundle over Ũ with contractible fibres EO(n). Since the action of the groupoid G is
trivial on EO(n) and commutes with the O(n) action on LO(Ũ ), the groupoid G acts
on both the total space and base space of the bundle p̃ : LO(Ũ ) ×O(n) EO(n)−−→Ũ .
From the definitions and construction in Appendix 2.5.16, we see that this is just
the universal space EG, i.e.,
(4.3.1) EG = LO(Ũ ) ×O(n) EO(n).
Taking the quotient by G one gets a V-bundle p : LO(X) ×O(n) EO(n)−−−−→X
so the classifying space BG is LO(X) ×O(n) EO(n). We thus have a commutative
diagram
π
EG
 −−−−−−−−−→ BG
 
 
p̃ 
(4.3.2)  p
y y
φ
Ũ = G0 −−−−−−−−−→ G0 /G = X .
Note that since LO(X) is independent of a particular atlas representing the orbifold
X , so is BG. In fact from Theorem A.2.3, up to weak homotopy equivalence BG
depends only on the Morita equivalence class of G. Thus, it makes sense to define
Definition 4.3.5: We call BG the classifying space of the orbifold X , and denote
it by BX .
It follows also from Theorem 4.3.3 that the weak homotopy type of BG is an
invariant of any foliation whose holonomy groupoid is Morita equivalent to the
orbifold groupoid G. Note also that the map p is not a fibration, since the fibres are
not all homotopy equivalent. A generic fibre of p, that is a fibre over a regular point
x ∈ X is EO(n), a contractible space, whereas, a singular fibre is an Eilenberg-
MacLane space K(Γ, 1), where Γ is the local uniformizing group.
The above discussion allows us, following [Hae84, HS91] to make the impor-
tant
Definition 4.3.6: We define the orbifold cohomology, homology, and homo-
topy groups by
i
Horb (X , Z) = H i (BX , Z) , Hiorb (X , Z) = Hi (BX , Z) , πi (X )orb = πi (BX ) .
We caution that the orbifold cohomology groups defined here are not the orb-
ifold cohomology groups defined by Chen and Ruan [CR04] which are not used in
this book. In [Thu79] Thurston introduced the orbifold fundamental group π1orb
as a group of deck transformations. It was shown in [HD84] that the orbifold
118 4. FUNDAMENTALS OF ORBIFOLDS

fundamental group π1orb (X ) defined in Definition 4.3.6 is equivalent to Thurston’s


better known definition. See Theorem 4.3.19 below. We should mention if the orb-
ifold X is a smooth manifold these orbifold groups coincide with the usual groups.
It should be noted that generally these orbifold groups are not topological invari-
ants of X, but invariants of the orbifold structure. In order to better understand
the orbifold cohomology groups, we study the Leray spectral sequence of the map
p : BX −−→X. For any sheaf F on BX we let Rs p(F) denote the derived functor
sheaves, that is the sheaves associated to the presheaves U 7→ H s (p−1 (U ), F). Then
Leray’s theorem [Bre97] says that there is a spectral sequence Err,s with E2 term
given by
E2r,s = H r (X, Rs p(F))
converging to H r+s (X , F).
Lemma 4.3.7: Let X = (X, U) be a compact orbifold and A an Abelian group.
Then the derived functor sheaves Rs p(A) have their support contained in the sin-
gular set Σorb of X with stalks given by the group cohomology H s (Γi , A) at xi ∈
Ui ∩ Σorb .
Proof. Let (Ũi , Γi , ϕi ) be a local uniformizing system for the open set Ui =
ϕi (Ũi ) ⊂ X. On Ũi the action of the groupoid G is the action of Γi . So p−1 (Ui ) is
identified with Ũi ×Γi EO(n) which is an Eilenberg-MacLane space K(Γi , 1) = BΓi .
Thus, H q (p−1 (Ui ), A) is the group cohomology H q (Γi , A). So for q > 0 the stalks
vanish at regular points and is H q (Γi , A) at the singular point xi . ¤
Since the group cohomology H q (Γi , A) vanishes if the order |Γi | is invertible in
A, the spectral sequence collapses and we have [Hae84]
Corollary 4.3.8: Let X = (X, U) be an orbifold of order υ(X ) < ∞. Then if
q
υ(X ) is invertible in A, the groups Horb (X , A) and H q (X, A) are isomorphic. In
q q
particular, Horb (X , A) ' H (X, A) if A = Q, R, C, or Zp , where gcd(p, υ(X)) = 1.
There is a similar spectral sequence for the homology groups [CM00] such that
Hr (X, Rs p(F)) =⇒ Hr+s
orb
(X , F),
so Corollary 4.3.8 holds for the homology groups as well. We should also warn the
p
reader that the groups Horb (X , F) or Hporb (X , F) do not in general vanish above
the dimension of X.
We are especially interested in orbifolds whose local uniformizing groups are
finite subgroups of the circle group S 1 , so we give these a special name.
Definition 4.3.9: An orbifold X = (X, U) is said to be locally cyclic if it has an
orbifold atlas all of whose local uniformizing groups are cyclic groups.
In the case of cyclic groups the cohomology groups H q (Γi , A) are well under-
stood. Indeed, for Zm [Bro82]


Z for q = 0,
H q (Zm , Z) = Zm for q > 0 even,


0 for q odd.
We want to relate orbibundles on X to ordinary bundles on BX . To do so
we need the analogue of an “atlas of charts” on BX . This is done by passing to
the finite models of EO(n) as discussed in Appendix A. So a smooth map on EG
means smooth maps on the finite models EG(k) which are smooth manifolds. Let
4.3. GROUPOIDS, ORBIFOLD INVARIANTS AND CLASSIFYING SPACES 119

{Ũi , Γi , ϕi } be a cover of uniformizing charts for the orbifold X, and let L̃i = LO(Ũi )
denote the restriction of LO(Ũ ) to Ũi . The group O(n) acts locally freely on L̃i
with isotropy group Γi fixing the orthonormal frames over the center ai ∈ Ui . So
we have homeomorphisms L̃i /O(n) ≈ Ũi /Γi ≈ Ui . Thus, we can cover BX by
neighborhoods of the form Ũi ×Γi EO(n), where Γi . Now by refining the cover if
necessary we have injection maps λji : Ũi −−→Ũj and these induce the change of
“charts” maps
Gji : Ũi ×Γi EO(n)−−→Ũj ×Γj EO(n)
given by Gji ([x̃i , e]) = [λji (x̃i ), e]. This is well-defined since the unique element γj is
identified with γi under the identification of Γi as a subgroup of Γj . This will allow
us to construct local data on Ũi ×Γi EO(n) by considering smooth (or holomorphic)
data on the Ũi and continuous data on EO(n) which is smooth on the finite models
EO(n)(k) and invariant under the Γi action. Since the λji are diffeomorphisms (or
biholomorphisms) this local data will then patch to give global data on BX . For
example we denote by E the sheaf of germs of complex-valued functions on BX that
are smooth in Ũi and continuous in EO(n). We shall call global (local) sections of
the sheaf E smooth functions on BX . Similarly we refer to “smooth” maps from
BX . For example, if f : X1 = (X1 , U1 )−−→X2 = (X2 , U2 ) is a good map of orbifolds,
then this induces a smooth map F : BX1 −−→BX2 . However, by Proposition 4.2.5
we have an equivalence of orbisheaves, and hence an equivalence of orbibundles.
Thus,
Proposition 4.3.10: An orbifold diffeomorphism f : X1 −−→X2 induces a diffeo-
morphism F : BX1 −−→BX2 .
Similarly, because of the commutative diagram
Ũi × EO(n) −−−→ Ũi
 
 
 ϕi
y y
pi
Ũi ×Γi EO(n) −−−→ Ui .
the map p : BX −−→X is smooth, since its local covering maps pi are smooth. Now
we are ready for our description of V-bundles in terms of ordinary bundles.
Theorem 4.3.11: Let X = (X, U) be an orbifold. There is a one-to-one correspon-
dence between isomorphism classes of orbibundles on X with group G and generic
fiber F and isomorphism classes of bundles on BX with group G and fiber F.
Proof. A V-bundle on X is a bundle on Ũi for each local uniformizing neigh-
borhood Ũi together with a group homomorphism hŨi ∈ Hom(Γi , G) that satisfy
the compatibility conditions of Definition 4.2.7. This gives an action of Γi on BŨi .
Now cover BX by neighborhoods of the form Ũi ×Γi EO(n), where we make use of
the fact that the local uniformizing groups Γi can be taken as subgroups of O(n).
There is an action of Γi on Ũi ×F ×EO(n) given by (x̃i , u, e) 7→ (γ −1 x̃i , uhUi (γ), eγ),
and this gives a G-bundle over Ũi ×Γi EO(n) with fiber F for each i. Moreover, the
compatibility condition (ii) of Definition 4.2.7 guarantees that these bundles patch
together to give a G-bundle on BX with fiber F.
Conversely, given a G-bundle on BX with fiber F, restricting to Ũi ×Γi EO(n)
gives a G-bundle there. Since for each i Ũi ×Γi EO(n) is the Eilenberg-MacLane
space K(Γi , 1), there is a one-to-one correspondence between isomorphism classes
120 4. FUNDAMENTALS OF ORBIFOLDS

of G-bundles on Ũi ×Γi EO(n) and conjugacy classes of group homomorphisms


Hom(Γi , G). The fact that these G-bundles come from a global G-bundle on BX
implies that the compatibility conditions (ii) of Definition 4.2.7 are satisfied. The
correspondence can be seen to be bijective. ¤
It is well-known that isomorphism classes of principal G-bundles on a CW
complex X are in one-to-one correspondence with the homotopy classes [X, BG] of
maps from X to BG. Thus, we have
Corollary 4.3.12: Let G be a Lie group. There is a one-to-one correspondence be-
tween principal orbibundles on the orbifold X = (X, U) with group G and homotopy
classes of maps [BX , BG].
Another corollary of Theorem 4.3.11 whose proof is straightforward is:
Corollary 4.3.13: The isomorphism classes of orbibundles on X with group G
are in one-to-one correspondence with elements of the non-Abelian cohomology set
H 1 (BX , G) = Horb
1
(X , G), where G is the sheaf of germs of maps to the group G.
Recall the cohomology ring H ∗ (BG) for the classical groups.

H ∗ (BO(n), Z2 ) ≈ Z2 [w1 , w2 , . . . , wn ] ,
H ∗ (BU (n), Z) ≈ Z[c1 , c2 , . . . , cn ] ,
(4.3.3) H ∗ (BSp(n), Z) ≈ Z[p1 , p2 , . . . , pn ] ,
where wi are the Stiefel-Whitney classes, ci are the Chern classes, and pi the
Pontryagin classes. By Theorem 4.3.11 a principal G V-bundle P on X can be
represented uniquely up to homotopy by P ≈ ψ ∗ EG, where ψ : BX −−→BG
is a classifying map. So for any ring A the V-bundle P defines a map ψ ∗ :
H ∗ (BG, A)−−→Horb ∗
(X , A). Elements in the image of this map are the orbifold
characteristic classes of the orbibundles on X . For example the Chern classes of a
principal U (n)-orbibundle, or the Pontryagin classes of a principal Sp(n) V-bundle

are integral classes in Horb (X , Z).
Now Chern-Weil theory proceeds on orbifolds as on manifolds. As in the mani-
fold case every principal V-bundle P admits a connection, and this can be described
by an invariant connection 1-form on P with values in the Lie algebra g of G, that
is a smooth section of Λ1 P ⊗g that is invariant under the local uniformizing groups.
Let I(G) denote the algebra of polynomials on g that are invariant under the adjoint
action of G on g. The algebra I(G) is graded by degree,
X
I(G) = I k (G),
k

and to each polynomial f ∈ I k (G) one defines a 2k-form on the orbifold P by


1 X ¡ ¢
f (Ω)(X1 , . . . , X2k ) = ²σ f Ω(Xσ(1) , Xσ(2) ), . . . , Ω((Xσ(2k−1) , Ωσ(2k) ) ,
2k! σ
where Ω is the curvature 2-form of ω, X1 , . . . , X2k are smooth vector fields on P,
the sum is taken over all permutations σ, and ²σ denotes the sign of σ. This 2k-form
is basic and projects to a closed 2-form on the underlying space X of the orbifold.
Moreover, its de Rham cohomology class is independent of the choice of connection
ω. This gives the Weil homomorphism

w : I(G)−−−→HdRh (X, R) ≈ H ∗ (X, R) .
4.3. GROUPOIDS, ORBIFOLD INVARIANTS AND CLASSIFYING SPACES 121

When X is a smooth manifold then w(I(G)) lies in the image of the inclusion
H ∗ (X, Z)−−→H ∗ (X, R). This is not true for orbifolds. However, we shall see that
w(I(G)) does lie in the image of H ∗ (X, Q)−−→H ∗ (X, R). Moreover, there is a uni-
versal homomorphism I(G)−−→H ∗ (BG, R) which is an isomorphism when G is com-
pact. So in this case we get a homomorphism ψ ∗ : H ∗ (BG, R)−−→H ∗ (X, R). El-
ements of H ∗ (X, R) that are in the image of ψ are known as real characteristic
classes. We now relate this construction of real characteristic classes in H ∗ (X, R) to
the integral characteristic classes described above. First we see that Corollary 4.3.8
implies that p : BX −−→X gives a ring isomorphism p∗ : H ∗ (X, Q)−−→Horb ∗
(X , Q).
∗ ∗
So the integral classes in Horb (X , Z) ⊂ Horb (X , Q) map to rational classes in
H ∗ (X, Q) under (p∗ )−1 . Furthermore, since both constructions are universal the
characteristic classes constructed by the Chern-Weil theory pullback under p∗ to

the integral characteristic classes in Horb (X , Z). Of course, we have worked with
the principal bundles here, but everything works for the corresponding associated
vector bundles as well. We denote the orbifold rational Chern (Pontrjagin) classes
in H ∗ (X, Q) of a complex (real) V-bundle E on an orbifold X by corb orb
i (E)(pi (E)),
respectively. Summarizing we have
Theorem 4.3.14: Let G = U (n) or Sp(n) and X an orbifold. Then the orbifold
Chern classes corb orb
i (E) and orbifold Pontrjagin classes pi (E) are defined for any
V-bundle E on X with group G, respectively. Moreover, p∗ corb ∗ orb
i (E) and p pi (E)

are integral classes in Horb (X , Z).
Similarly for G = SO(2n) we can define the orbifold Euler class eorb ∈ H 2n (X, Q)
such that p∗ eorb ∈ H 2n (X , Z), and eorb vanishes for the odd special orthogonal
groups. Satake [Sat57] defines the orbifold Euler characteristic χorb (X) as
X 1
(4.3.4) χorb (X) = (−1)dim(Ci ) ,
|Γ(Ci )|
Ci

where the sum is taken over a cell decomposition of X that is compatible with the
stratification 4.1.1 in the sense that every cell lies in a single stratum. This can be
rewritten in the more invariant form as
X χ(S)
(4.3.5) χorb (X) = (−1)dim(S) ,
|Γ(S)|
S

where the sum is taken overall strata S of the stratification (4.1.1), and χ(S) is the
ordinary Euler class of the stratum S. It is emphasized that generally χorb (X) is a
rational number. Satake [Sat57] proved the orbifold version of the Gauss-Bonnet
Theorem. The Euler class eorb ∈ H 2n (X, Q) can be represented by the top invariant
curvature form Ω of the Riemannian curvature, so the Gauss-Bonnet formula takes
the form
Z
(4.3.6) χorb (X) = heorb , [X]i = Ω,
X

where [X] ∈ H2n (X, Q) denotes the fundamental class of the compact oriented
2n-dimensional orbifold X.
We are particularly interested in the case of circle V-bundles on an orbifold. In
this case the group G is the circle group U (1), and the relevant characteristic class
is corb
1 ∈ H 2 (X, Q), or equivalently its image p∗ corb
1
2
∈ Horb (X , Z). We have
122 4. FUNDAMENTALS OF ORBIFOLDS

Theorem 4.3.15: The isomorphism classes of circle V-bundles E over an orbifold


2
X = (X, U) are in one-to-one correspondence with elements of Horb (X , Z), and the
∗ orb
bijection is given by the orbifold first Chern class p c1 . Furthermore, the total
space of E is a smooth manifold if the local homomorphisms hŨi of the local uni-
formizing groups Γi mapping into the group U (1) of the bundle are monomorphisms
for all local uniformizing charts (Ũi , Γi , ϕi ).
Proof. Let E denote sheaf of germs of smooth complex-valued functions on
BX and E ∗ the sub-sheaf of nowhere vanishing complex valued smooth functions.
The isomorphism classes of complex line bundles on BX , hence complex line V-
bundles on X, are in one-to-one correspondence with the elements of the cohomology
1
group Horb (X , E ∗ ). Now BX is an infinite paracompact CW complex, and E is a
fine sheaf, so the exponential exact sequence
0−−−→Z−−−→E−−−→E ∗ −−−→0
gives an isomorphism
δ
1
Horb (X , E ∗ ) −−→ Horb
2
(X , Z) .
Now let S denote the sheaf of germs of maps from BX to U (1). We can identify S
as a subsheaf of E ∗ by identifying U (1) as the unit circle in C. There is a deformation
retraction of E ∗ onto S as étale spaces, so there is an isomorphism Horb 1
(X , E ∗ ) ≈
1 ∗
Horb (X , S ). Now Corollary 4.3.13 implies the first statement. In the case of circle
bundles E over a CW complex it is well-known that the connecting homomorphism
δ can be represented by the first Chern class of E. The last statement follows from
the discussion after Definition 4.2.7. This completes the proof. ¤
It is straightforward to generalize Theorem 4.3.15 in analogy with the manifold
case to
Theorem 4.3.16: The isomorphism classes of principal T n V-bundles E over an
2
orbifold X = (X, U ) are in one-to-one correspondence with elements of Horb (X , Zn ).
In [HS91] Haefliger and Salem study torus actions on orbifolds. There it is
2
shown that Horb (X , Zn ) classifies T n V-bundles up to local equivalence. This gives
a monomorphism H 2 (X, Zn )−−→Horb 2
(X , Zn ) which, of course, is an isomorphism
rationally. [HS91] also study the exact homotopy sequence associated with a locally
free action of T n on an orbifold. We shall need
Definition 4.3.17: A (smooth) action of a Lie group G on an orbifold X =
(X, U) is a continuous action A : G × X−−→X such that for each g ∈ G and
x ∈ X there are local uniformizing charts (Ũ , Γ, ϕ) and (Ũ 0 , Γ0 , ϕ0 ) with x ∈ ϕ(Ũ ),
A(g, x) ∈ ϕ0 (Ũ 0 ), together with an open neighborhood V of g ∈ G and a smooth
map à : V × Ũ −−→Ũ 0 such that ϕ0 (Ã(g, x̃)) = A(g, ϕ(x̃)) and for each g ∈ V the
map x̃ 7→ Ã(g, x̃) is a diffeomorphism.
The Haefliger-Salem theory applies to the following situation. Let G be a
compact Lie group acting locally freely on an orbifold Y with quotient orbifold
X . This gives rise to a fibration EO(n) × G−−→BY−−→BX , which induces the long
exact homotopy sequence. Thus, we have
Theorem 4.3.18: Let G be a compact Lie group that acts locally freely on an
orbifold Y = (Y, V) with quotient orbifold X = (X, U). Then the sequence
· · · −−→πi (G)−−→πiorb (Y)−−→πiorb (X )−−→πi−1 (G)−−→ · · ·
4.4. COMPLEX ORBIFOLDS 123

of homotopy groups is exact.


We are particularly interested in the orbifold fundamental group π1orb (X ) =
π1 (BX ). Thurston [Thu79] has given an alternative definition of π1orb (X ) as the
group of deck transformations. Here we state this as a theorem.
Theorem 4.3.19: Let X = (X, U) be an orbifold. Then there exists a universal
covering orbifold X̃ = (X̃, Ũ). The orbifold fundamental group π1orb (X ) is the group
of deck transformations of the universal covering map π : X̃−−→X.

Proof. The existence of a universal covering orbifold was given in Chapter


13 of [Thu79]. The proof that Haefliger’s definition and Thurston’s definition of
π1orb (X ) coincide was given in [HD84] to which we refer. ¤

4.4. Complex Orbifolds


On a complex orbifold there is a Γ-invariant tensor field J of type (1, 1) which
describes the complex structure on the tangent V-bundle T X. The almost complex
structure J gives rise in the usual way to the V-bundles Ar,s of differential forms
of type (r, s). The standard concepts of Hermitian and Kähler metrics hold equally
well on orbifolds, and all the special identities involving Kähler, Einstein, or Kähler-
Einstein geometry hold. In particular, the standard Weizenböck formulas hold.
This is described in more detail in Chapter 3.

4.4.1. The Underlying Space and Divisors. Let X = (X, U) be a complex


orbifold. Then the underlying complex space X is a paracompact Hausdorff space
with at worst quotient singularities. We can give X the structure of a locally C-
ringed space (X, O) as follows: The structure sheaf OX has stalks Ox for x ∈ U ⊂ X
that are isomorphic to the local ring OCΓn of germs of Γ-invariant holomorphic func-
tions, where Γ is the local uniformizing group of the uniformizing system (Ũ , Γ, ϕ).
Such rings are reduced, that is, they contain no nilpotent elements. (The reader is
referred to [GR65, GR84] for basic facts about locally ringed spaces.) Since the
orb
local uniformizing groups are the identity on the dense set Xreg of regular orbifold
points, the ring Ox is isomorphic to the ring OCn of convergent power series in
Cn when x ∈ X is regular. In particular, OCn is a unique factorization domain,
or UFD for short. However, at a singular orbifold point x ∈ Σorb (X), the ring
Ox is isomorphic to the ring OCΓnx , which for Γx 6= id is not necessarily a UFD
[Mum61, Bri68]. Nevertheless it follows from a standard result (cf. [Bou89], pg
323) that the rings OCΓnx are always integrally closed.
Recall that an element A ∈ GL(n, C) is called a reflection if A has eigenvalue
1 with multiplicity n − 1. Thus, a reflection fixes a hyperplane in Cn , hence, its
name. A finite subgroup Γ ⊂ GL(n, C) is a reflection group if it is generated by
reflections. A finite subgroup Γ ⊂ GL(n, C) is small if it contains no reflections.
We have the following well-known result [Pri67] (See also [Dim92]):
Theorem 4.4.1: Let Γ ⊂ GL(n, C) be a finite subgroup. Then
(i) The quotient space Cn /Γ is smooth if and only if Γ is a reflection group.
(ii) There is a small group Φ such that Cn /Γ and Cn /Φ are biholomorphic.
In the case n = 1 every finite subgroup of GL(1, C) = C∗ is a reflection group
and has the form Zn for some integer n ≥ 2, so we have
Corollary 4.4.2: For every finite subgroup Γ ⊂ C∗ , the quotient C/Γ is smooth.
124 4. FUNDAMENTALS OF ORBIFOLDS

It follows from Theorem 4.4.1 that if a local uniformizing group Γ of an orbifold


X = (X, U) contains a reflection, the orbifold singular set will be bigger than the
singular set of the underlying complex space. Thus we have Σ(X) ⊂ Σorb (X ), and
it follows from Theorem 4.4.1 that Σ(X) = Σorb (X ) if and only if none of the local
uniformizing groups Γi of the orbifold X = (X, U) contain a reflection. If some Γi
contains a reflection, then the reflection fixes a hyperplane in Ũi giving rise to a
ramification divisor on Ũi and a branch divisor on X. We discuss this more in detail
below.
Now combining the discussion of the last several paragraphs we arrive at
Proposition 4.4.3: Let OX be the structure sheaf of a complex orbifold X =
(X, U). Then
(i) If x ∈ Xreg then Ox is isomorphic to the ring OCn of convergent power
series.
(ii) If x ∈ Σ(X) with local uniformizing group Γ, then there is a nontrivial
small subgroup Φ of Γ such that Ox is isomorphic to the subring OCΦn of
Φ invariant convergent power series, so Ox is not a UFD.
(iii) For any x ∈ X the local ring Ox is integrally closed.
This Proposition together with well-known results (cf. [GR84]) have the fol-
lowing consequence:
Proposition 4.4.4: The locally ringed space (X, OX ) associated to a complex orb-
ifold has the following properties:
(i) (X, OX ) is a reduced normal complex space,
(ii) the singular locus Σ(X) is a closed reduced complex subspace of X and
has complex codimension at least two in X,
(iii) the smooth locus Xreg is a complex manifold and a dense open subset of
X.
If Σ(X) is compact it follows from Proposition 4.4.4 and Hartogs’ Theorem
that any holomorphic function on Xreg extends to a holomorphic function on all
of X. Furthermore, Weil divisors can be defined on (X, OX ), although there may
be none in the purely analytical setting. Similarly Cartier divisors can also be de-
fined. However, generally in contrast to Theorem 3.5.19, they are not the same as
we discuss below. But Xreg is a complex manifold, so Theorem 3.5.19 applies and
Weil divisors and Cartier divisors coincide on Xreg . Let us set some notation as
in subsection 3.5.2. For any normal variety X we let Div(X) denote the group of
Weil divisors on X, and H 0 (X, M∗ /O∗ ) the group of Cartier divisors. Similarly,
we denote by Cl(X) the divisor class group of Weil divisors modulo linear equiva-
lence, and by CaCl(X) the group of Cartier divisor classes (Cartier divisors modulo
principal divisors). Moreover, by (ii) of Proposition 1.1.5 we have
Lemma 4.4.5: Let X be the underlying complex space of a complex orbifold. Then
Cl(X) ≈ Cl(Xreg ) ≈ CaCl(Xreg ).
However, not every Cartier divisor on Xreg extends to a Cartier divisor on X.
If D is a Weil divisor with x ∈ D ∩ Σ(X) 6= ∅, then Ox is not a UFD, so D is
not locally principal near x. The obstruction for being locally principal near x lies
in the local divisor class group Cl(X, x) = Cl(Ũ /Φ). Thus, the Cartier divisors are
identified with the subgroup of Div(X) consisting of locally principal divisors. So
the group CaCl(X) is generally a subgroup of the divisor class group Cl(X). In our
4.4. COMPLEX ORBIFOLDS 125

singular situation in lieu of Theorem 3.5.19 we have only an exact sequence


(4.4.1) 0−−−→CaCl(X)−−−→Cl(X)−−−→Cl(X)/CaCl(X)−−−→0 .
If X is a projective algebraic variety (or scheme) [Har77], which is what is of
interest to us in this book, then there is an isomorphism CaCl(X) ≈ PicX, where
PicX denotes the Picard group of holomorphic line bundles (equivalently, invertible
sheaves) on X. Because of this we shall often use PicX to denote the group CaCl(X)
as well as the group of holomorphic line bundles or invertible sheaves. If D ∈
Div(X) the sheaf O(D) is not necessarily invertible. It is, however, reflexive of rank
1 [Kol04], and one can characterize Cl(X) as group of rank 1 reflexive sheaves. The
underlying space X of a complex orbifold enjoys another important property.
Definition 4.4.6: A normal reduced complex space is said to be Q-factorial if
every Weil divisor is Q-Cartier, i.e., some multiple of it is a Cartier divisor.
From the definition of orbifolds it is easy to see that
Proposition 4.4.7: The underlying space X of a complex orbifold is Q-factorial.
In order to work efficiently with complex orbifolds, it is convenient to consider
Q-divisors on X. These are just formal finite linear combinations of analytic hyper-
surfaces with rational coefficients, that is, Weil divisors with coefficients in Q. Then
generalizing Definition 4.4.6 we say that a Q-divisor is Q-Cartier if some multiple
of it is a Cartier divisor. The concepts of effective, nef, big, etc. all apply equally
well to Q-divisors, but now for example the intersection number is only a rational
number instead of an integer. We refer to the literature [Laz04a, Laz04b, Mat02]
for more details. It will be convenient for us to work in the so-called log category,
that is one considers pairs (X, ∆), where X is the underlying space of a complex
orbifold and ∆ is a Q-divisor on X.
Definition 4.4.8: The branch divisor ∆ of an orbifold X = (X, U) is a Q-divisor
on X of the form
X³ 1 ´
∆= 1− Dα ,
α

where the sum is taken over all Weil divisors Dα that lie in the orbifold singular
locus Σorb (X), and mα is the gcd of the orders of the local uniformizing groups
taken over all points of Dα and is called the ramification index of Dα .
Notice that if a Weil divisor D lies in Σorb (X), the local uniformizing group
of every point is nontrivial, and the gcd of the orders of all the local uniformizing
groups on D is its ramification index. As emphasized by Kollár [Kol04], pg. 15,
a complex orbifold X = (X, U ) is uniquely determined by the pair (X, ∆). The
local branched covers ϕi : Ũi −−→Ui are uniquely determined by the conditions
that ϕi is unramified over Ui \ (Σorb (X) ∩ Ui ), and the ramification index mα
of Dα is the largest integer dividing the ramification index of ϕi for all i with
Dα ∩ Ui 6= ∅. So one can think of the orbifold X as the pair (X, ∆). This point
of view is also espoused in [Cam04]. Then by applying Theorem 4.4.1 we see
that one can associate to every complex orbifold (X, ∆) another complex orbifold
(X, ∅) which we call the unramified orbifold associated with (X, ∆). Notice that if
Σ(X) = ∅ then (X, ∅) is a smooth complex manifold. We denote the order of the
unramified orbifold (X, ∅) by ῡ = ῡ(X), and more generally the order of the local
uniformizing group Γx at x ∈ Σ(X) by ῡ(x). If the unramified orbifold (X, ∅) is of
finite type, multiplication by ῡ defines a map Cl(X)−−→CaCl(X). More generally
126 4. FUNDAMENTALS OF ORBIFOLDS

if D ∈ Div(X) and ῡ(D) = lcmx∈D ῡ(x) then ῡ(D)D is Cartier. We illustrate the
discussion of this subsection with an example.
Example 4.4.9: Consider the weighted projective spaces CP(w) described in Sec-
tion 4.5 below. For simplicity we take n = 2 and w = (1, 4, 6). The orbifold struc-
ture on CP(1, 4, 6) is the standard orbifold structure described in Section 4.5. In
this case the Weil divisor D0 = (z0 = 0) gives a branch divisor 21 D0 with ramifica-
tion index 2. The order of the orbifold (CP(1, 4, 6), 12 D0 ) is υ = 12. The unramified
orbifold is CP(1, 2, 3) with its standard orbifold structure whose order is ῡ = 6.
Now CP(1, 2, 3) has two singular points, (0, 1, 0) with local uniformizing group the
cyclic group Z2 , and (0, 0, 1) with local uniformizing group Z3 . Both of these points
are contained in the Weil divisor D0 , and D0 is not a Cartier divisor. However, 6D0
is Cartier. Notice also that the divisors D1 = (z1 = 0) and D2 = (z2 = 0) are not
Cartier, but 3D1 and 2D2 are Cartier. Both CaCl(CP(1, 2, 3)) and Cl(CP(1, 2, 3))
are isomorphic to Z, and we have an exact sequence
0−−−→CaCl(CP(1, 2, 3))−−−→Cl(CP(1, 2, 3))−−−→Z6 −−−→0 .
So Pic(CP(1, 2, 3)) ≈ CaCl(CP(1, 2, 3)) ≈ 6Cl(CP(1, 2, 3)).
We end this subsection by recalling
Definition 4.4.10: A canonical divisor KX is any divisor on X such that its
restriction to Xreg is associated to the canonical bundle ΛnXreg .
Notice that KX is only defined up to linear equivalence.

4.4.2. Orbidivisors and Orbisheaves. As with bundles and sheaves the


notion of a divisor in the category of orbifolds consists of a sequence of divisors,
one on each local uniformizing neighborhood that patch together in the right way.
Definition 4.4.11: An orbidivisor or Baily divisor on a complex orbifold X is
a Cartier divisor DŨ on each local uniformizing system {Ũ , Γ, ϕ} that satisfies the
two conditions
(i) If for each x ∈ X and γ ∈ Γ, f ∈ Dγx then f ◦ γ ∈ Dx .
(ii) If λ : {Ũ , Γ, ϕ}−−→{Ũ 0 , Γ0 , ϕ0 } is an injection and f ∈ Dλ(x)
0
then f ◦ λ ∈
Dx . Here D denotes the divisor sheaf on Ũ .
A Baily divisor is called absolute if on each local uniformizing system {Ũ , Γ, ϕ}
the divisor DŨ can be written as DŨ = (f ), where f is the quotient of Γ-invariant
holomorphic functions on Ũ .
Thus, a Baily divisor is viewed as a divisor on the space of objects G0 = Ũ
of the associated étale groupoid that satisfy certain compatibility conditions with
respect to the morphisms of the groupoid.
There are several important orbisheaves on complex orbifolds that we shall
work with. First, there is the structure orbisheaf OX defined by Definition 4.2.2,
where now OŨ is the sheaf of holomorphic functions on Ũ . Similarly there is the
meromorphic orbisheaf MX consisting of meromorphic functions on each local uni-
formizing neighborhood Ũ . Finally, there is the canonical orbisheaf of a complex
orbifold. On a complex orbifold X of complex dimension n, we denote by Ωp the
orbisheaf of holomorphic differential p-forms on X . This is the orbisheaf constructed
from the usual sheaf ΩpŨ on each orbifold chart. Ωp is a locally free orbisheaf of
¡ ¢ i
rank np .
4.4. COMPLEX ORBIFOLDS 127

Definition 4.4.12: The canonical orbisheaf of a complex orbifold X of complex


dimension n is the orbisheaf Ωn .
Given a Weil divisor D on X we can lift its restriction DUi to Ui to a divisor
D̃Ũi on the uniformizing neighborhood Ũi by D̃Ũi = ϕ−1 i (D ∩Ui ). It is easy to check
that this defines a Baily divisor D̃ on the orbifold X = (X, U ). In fact, if m ∈ Z
divides the order of all the local uniformizing groups at all points of a Q-divisor
1
m D, then it lifts to a Baily divisor. In particular, since the ramification index of a
branch divisor ∆ is the gcd of the orders of the uniformizing groups at all point of
∆ we have arrived at
Proposition 4.4.13:
P bα The branch divisor ∆ or more generally any Q-divisor on
X of the form α m α
Dα , where mα is a ramification index and bα ∈ Z, lifts to a
Baily divisor R on X = (X, U ).
A Baily divisor R obtained as the lift of a branch divisor is called a ramification
divisor. The following is straightforward
Proposition 4.4.14: To each Baily divisor D on X there corresponds a complex
line V-bundle [D] or equivalently an invertible orbisheaf O(D). Absolute Baily di-
visors on a complex orbifold X correspond to Cartier divisors on the underlying
complex space X.
Remark 4.4.1: Let X = (X, U ) be a complex orbifold. Its canonical orbisheaf
by Proposition 4.2.18 determines an orbibundle which will be called the canonical
orb
orbibundle of X and denoted by KX . We will sometimes write KX , especially
in the presence of a non-trivial branch divisor ∆. The anticanonical orbibundle
−1 orb
will always be denoted by KX . One can interchangeably think of KX as the
canonical orbibundle, the orbifold canonical divisor (see below), as well as the
orbifold canonical class (rational orbifold first Chern class) in H 2 (X, Q). On the
other hand, when X = (X, ∆) one may consider a simpler complex orbifold: namely
(X, ∅). This orbifold’s canonical divisor coincides with the canonical divisor KX
defined in (4.4.10). When X is non-singular, KX traditionally also denotes the
canonical line bundle. When X has quotient singularities this is not defined. On
the other hand, there is a well-defined canonical orbibundle of the complex orbifold
−1
(X, ∅) and this orbibundle will be denoted by KX with its dual written as KX .
The most important Baily divisor on a complex orbifold X = (X, U) is the
orb
orbifold canonical divisor KX which is any Baily divisor associated to the canonical
orbibundle KX . In the presence of branch divisors Dα an orbifold canonical divisor
orb
KX is not the same (meaning not linearly equivalent) as the canonical divisor KX
of the underlying complex space X. In fact we have
orb
Proposition 4.4.15: The orbifold canonical divisor KX and canonical divisor
KX are related by
X³ 1 ´ ∗
orb
KX ≡ ϕ∗ KX + 1− ϕ Dα .
α

orb
In particular, KX ≡ ϕ∗ KX if and only if there are no branch divisors.
1
Proof. Let (1 − m )D be an irreducible branch divisor. Assume that D in-
tersects Ui nontrivially, where Ũi is a local uniformizing neighborhood. By defini-
tion the ramification index m divides the order of Γi . We can choose coordinates
(x1 , . . . , xn ) on Ũi so that at a general point of the ramification divisor Ri the map
128 4. FUNDAMENTALS OF ORBIFOLDS

ϕ takes the form


ϕi (x1 , . . . , xn ) = (xm
1 , x2 , . . . , xn ) = (z1 , . . . , zn ) .

A local section dx1 ∧ · · · ∧ dxn of the canonical bundle satisfies


ϕ∗i (dz1 ∧ · · · ∧ dzn ) = mx1m−1 dx1 ∧ dx2 ∧ · · · ∧ dxn ,
so that the ramification divisor Ri is represented by xm−1 1 = 0. This expresses a
local section of the canonical bundle KŨi as the sum of a local section of the pullback
of the canonical bundle KUi and a pole of order m−1 along the ramification divisor
which translates to a pole of order m−1 m along the pullback of the branch divisor
D, that is
m−1 ∗
KŨi ≡ ϕ∗i KUi + (m − 1)Ri = ϕ∗i KUi + ϕi D .
m
This holds in all uniformizing neighborhoods, and for all irreducible branch divisors.
This proves the result. ¤
In terms of the orbifold rational Chern class corb
1 (X ) defined in the paragraph
above Theorem 4.3.14, Proposition 4.4.15 implies
X³ 1 ´
(4.4.2) corb
1 (X ) = −c1 (KX ) = c1 (X) − 1− c1 (O(Dα )) ∈ H 2 (X, Q) .
α
m α

Here c1 (X) = −c1 (KX ) ∈ H 2 (X, Q) is the first Chern class of X. We will sometimes
write corb orb
1 (X ) = c1 (X, ∆) to indicate the branch divisor.
The orbifold Kodaira dimension can be defined analogously with the manifold
orb
case if one replaces the canonical bundle KX by the canonical orbibundle KX .
Definition 4.4.16: Let X = (X, U) be a compact complex orbifold. The orbifold
Kodaira dimension Kodorb (X ) is defined
(
orb −∞ if R(X ) = C
Kod (X ) =
tr(R(X )) − 1 otherwise,
where tr(R) denotes the transcendence degree of the ring R.
Here R(X ) is the orbifold canonical ring defined by
X
(4.4.3) R(X ) = C ⊕ H 0 (X , O(mKXorb
)).
m≥1

Note that by passing to invariant sections Kodorb (X ) is just the Itaka dimension of
the line bundle
µ ³ X³ ¶
orb ∗ 1 ´ ∗ ´
O(υ(KX )) = O υ ϕ KX + 1− ϕ Dα .
α

Of special interest to us is the case Kodorb (X ) = −∞.


Definition 4.4.17: We say that a compact complex orbifold X is a (log) Fano
orbifold if corb
1 (X ) > 0. Alternatively, a Fano orbifold is a complex orbifold whose
−1
anti-canonical orbibundle KX is ample. A Fano orbifold surface is also called a
del Pezzo orbifold.
Actually this is a special case of a somewhat more general definition used in
Mori theory which we will make use of in Chapter 10.
Definition 4.4.18: A log del Pezzo surface is a pair (X , ∆), where
4.4. COMPLEX ORBIFOLDS 129

(i) S is aPnormal, projective surface,


(ii) ∆ = ai Di is a linear combination of distinct irreducible divisors with
0 ≤ ai ≤ 1, ai ∈ Q, and
(iii) −(KS + ∆) is ample. (That is, a suitable multiple of it is an ample
Cartier divisor.)
The Q-divisor ∆ is called the boundary.
Thus a 2-dimensional log Fano orbifold is a log del Pezzo surface. In the orbifold
case the boundary satisfies ai = mm i −1
i
, where the mi are natural numbers. Such a
boundary is sometimes called standard.
In analogy with the manifold case we also say that a complex orbifold X =
(X, U) is a Calabi-Yau orbifold if corb 1 (X ) = 0. However, an important result of
Kollár [Kol05] given in Proposition 4.7.11 below says that under mild conditions
this is the same as c1 (X) = 0.
Remark 4.4.2: It is clear from (4.4.2) that in the presence of a non-trivial branch
divisor ∆, the complex space X may be a Fano variety in the usual algebraic sense
of the anti-canonical bundle of X being ample, while, depending on the divisor
∆, the complex orbifold X = (X, U ) may or may not be Fano in the orbifold
sense. We would like to highlight this important difference: our definition of a
Fano orbifold is not equivalent to X being Fano as an algebraic variety. We will
soon consider many examples of this situation. We warn the reader that when X
is not smooth other inequivalent notions of Fano or Q-Fano can be found in the
literature. Our extension of the Fano condition to the orbifold case is quite natural
in several different contexts. Most importantly, as we shall see later, a compact
Kähler Fano orbifold admits an orbifold Kähler metric of positive Ricci curvature
in the same Kähler class, this being one of the consequences of the orbifold version
of the Calabi-Yau theorem.
Example 4.4.19: Let (Σ; z1 , . . . , zk ) be a Riemann surface with k marked points.
We can give (Σ; z1 , . . . , zk ) an orbifold structure by defining local uniformizing
systems (Ũi , Cmi , ϕi ) centered at the point zi , where Cmi is the cyclic group of
order mi and ϕi : Ũi −−→Ui = Ũi /Cmi is the branched covering map ϕi (z) = z mi .
An orbifold canonical divisor is given by
X³ 1 ´
orb ∗
K(Σ;z 1 ,...,zk )
≡ ϕ K Σ + 1 − (zi ) ,
i
mi
where KΣ is an ordinary canonical divisor on the Riemann surface Σ. Thus, if g
denotes the genus of Σ the orbifold Chern number is
Xk
1
corb
1 (Σ; z1 , . . . , zk ) = 2 − 2g − k +
i=1
mi

which equals the orbifold Euler characteristic χorb (Σ; z1 , . . . , zk ). In order that
(Σ; z1 , . . . , zk ) be a Fano orbifold we need g = 0, and k < 4. When g = 0 we
have Σ = S 2 with k = 1, 2, 3 marked points, and the Chern number becomes
Xk
1
corb 2
1 (S ; z1 , . . . , zk ) = 2 − k + .
i=1
m i

For k = 1 we have corb 2


1 (S ; z1 ) = 1 + (1/m1 ), while for k = 2, we recapture Example
orb 2
4.1.6 with c1 (S ; z1 , z2 ) = (1/m1 ) + (1/m2 ). For k = 3 we obtain a Fano orbifold
130 4. FUNDAMENTALS OF ORBIFOLDS

only for (m1 , m2 , m3 ) = (p, 2, 2), (3, 3, 2), (4, 3, 2), and (5, 3, 2). These can all be
realized as hypersurfaces of Brieskorn-Pham polynomials in a weighted projective
space, see Example 4.6.15 below and subsection 10.1.1 of Chapter 10. Fano orbifold
structures on Riemann surfaces are usually called ‘spherical’ [Mil75]. As in the
usual case there is a uniformization theorem for Riemannian orbifold metrics with
Gaussian curvature equal to 1, 0, −1. This is no longer a topological invariant, but an
invariant of the orbifold structure according to the orbifold Gauss-Bonnet formula
4.3.6.
Exercise 4.2: Determine for which k and mi one obtains ‘Euclidean orbifolds’
(i.e., corb
1 = 0) and ‘hyperbolic orbifolds’ (corb 1 < 0) on S 2 .

4.4.3. The Orbifold Picard and Néron-Severi Groups. In [BG97b] the


authors introduced the set Picorb (X) of equivalence classes holomorphic line V-
bundles over a complex orbifold X = (X, U ). First we notice that complex line
V-bundles on X correspond to circle bundles on X which form a group by Corollary
1
4.3.13 namely Horb (X, E ∗ ). There are several approaches to deal with holomorphic
orb
objects such as Pic (X). One may: 1) work directly on the V-bundle using holo-
morphic data in Definition 4.2.7, or 2) work on BX using Theorem 4.3.11 in the
holomorphic context. These two approaches are actually very closely related since
holomorphic data on BX really means holomorphic data on the set of objects of the
étale groupoid G, and this is the same as holomorphic data on the V-bundles since
the “charts” of BX are Ũi ×Γi EU (n) with holomorphic data on Ũi and smooth
data on EU (n). In this case the group G is a complex Lie group and the fibre F is
a complex manifold. In fact, we are most interested in the case of holomorphic line
bundles or holomorphic C∗ bundles7. We have arrived at the holomorphic version
of Theorem 4.3.11:
Theorem 4.4.20: Let X = (X, U ) be an complex orbifold. There is a one-to-one
correspondence between isomorphism classes of holomorphic V-bundles on X with
group G and generic fiber F and isomorphism classes of holomorphic bundles on
BX with group G and fiber F. In particular, holomorphic line V-bundles on X
correspond to holomorphic line bundles on BX .
This leads to
Definition 4.4.21: The Abelian group of holomorphic line V-bundles on an orb-
ifold X is denoted by Picorb (X ).
The relationship between the orbifold Picard group Picorb (X ) and the ordinary
Picard group Pic(X) is given by:
Proposition 4.4.22: Let X = (X, U) be a complex orbifold of order υ. Then the
following hold:
(i) Pic(X) is naturally a subgroup of Picorb (X ),
(ii) Pic(X) ⊗Z Q ≈ Picorb (X ) ⊗Z Q,
(iii) If υ is finite, in particular if X is compact, the map L 7→ Lυ

Picorb (X )−−−−→Pic(X)
is a homomorphism of Abelian groups.

7A third approach using Grothendieck topologies was taken in [SW99].


4.4. COMPLEX ORBIFOLDS 131

Proof. (i): Ordinary holomorphic line bundles on X corresponds to absolute


line orbibundles on X , so there is a monomorphism Pic(X)−−→Picorb (X ). (ii):
This follows from Corollary 4.3.8 and Theorem 4.4.20. (iii): If υ is finite, the map
L 7→ Lυ , is a group homomorphism sending Picorb (X ) to Pic(X), since the absolute
orbibundle Lυ is identified with an element of Pic(X). ¤
We now study holomorphic line V-bundles on X by considering ordinary line
bundles on BX that are holomorphic in the complex coordinates on Ũ . Let X be
a compact complex orbifold. We consider the classifying space BX as covered by
“charts” of the form Ũi ×Γi EU (n). In the complex category the local uniformizing
neighborhoods Ũi are identified with Cn with the origin as the fixed point of Γi , and
the injections λji of Definition 4.1.1 are biholomorphisms on their image. We now
define the sheaf A of germs of holomorphic functions on BX to be the sub-sheaf
of E consisting of functions that are holomorphic in the Ũi . Similarly, A∗ denotes
the subsheaf of germs of nowhere vanishing functions. So the last statement of
Theorem 4.4.20 can be rephrased as
Theorem 4.4.23: On a complex orbifold there is an isomorphism
Picorb (X ) ' Horb
1
(X , A∗ ) .
The exponential exact sequence 0−−→Z−−→A−−→A∗ −−→0 on BX yields the long
exact sequence
p∗ corb
1
1
· · · −−→Horb 1
(X , Z)−−→Horb (X , A)−−→Picorb (X ) −−−→ Horb
2 2
(X , Z)−−→Horb (X , A)−−→ · · ·
which, in turn, gives rise to an exact sequence
p∗ corb
1
1
0−−−→Horb 1
(X , A)/Horb (X , Z)−−−→Picorb (X ) −−−−→ Horb
2
(X , Z)
and the image of the map p∗ corb
1 is the orbifold Neron-Severi group
N S orb (X ) = Picorb (X )/Horb
1
(X , A) ≈ H 1,1 (X , C) ∩ Horb
2
(X , Z) .
Following the usual convention the rank ρorb (X ) of N S orb (X ) is called the orbifold
Picard number.
Next we give slight generalization of the Fano index. We generalized the notion
of Fano index to orbifolds in [BG00b]. However, it is convenient to be more general
and consider a divisibility index as in [BGO07].
Definition 4.4.24: Let X be a complex orbifold such that corb 1 (X ) is definite. The
divisibility index or just index of X is the largest positive integer m such that
p∗ corb
1 (X ) 2
m is an element of the free part of Horb (X , Z). The index of X is denoted by
IX .
Now we have
Lemma 4.4.25: Let X be a complex orbifold with IX = m. Then there is a holo-
morphic line V-bundle L ∈ Picorb (X ) such that Lm = KX −1
.
Proof. The idea of the proof is simple, but since we are working on BX we
write out the details. First on BX we define the following sheaf E p,q of “differential
forms”: Let (z1 , . . . , zn ) be complex coordinates on Ũi . Then using the standard
multi-index notation, we construct the sheaf E p,q whose stalks are spanned by
elements of the form fIJ (z, e)dzI ∧ dz̄J , where I = i1 · · · ip and J = j1 · · · jq are
the usual multi-indices, and f is a smooth function on Ũi × EO(2n) satisfying
132 4. FUNDAMENTALS OF ORBIFOLDS

f (γ −1 z, eγ) = f (z, e) for γ ∈ Γi . We have, as usual, differential operators ∂¯ and the


Dolbeault complex
∂¯ ∂¯ ∂¯
· · · −−→ E p,q −−→ E p,q+1 −−→ · · · .
Notice that E 0,0 = E and ker(∂¯ : E 0,0 −−→E 0,1 ) = A.
−1
By Theorem 4.3.15 there is a complex line V-bundle L such that Lm = KX .
The transition function hij for L are nowhere vanishing local sections of E over
−1
Ũi ×Γi EO(2n) ∩ Ũj ×Γj EO(2n) and the transition functions for Lm = KX are
hm −1 ¯ m ¯
ij . But since KX is holomorphic ∂hij = 0; hence, ∂hij = 0 implying that L is
holomorphic. ¤

4.4.4. Kähler and Algebraic Orbifolds. It is usually straightforward to


generalize the geometry of Kähler manifolds as reviewed in Chapter 3 to the case
of orbifolds. However, as seen previously subtleties can and do occur. Here we
study the geometry of putting Kähler metrics on a complex orbifold. As discussed
in Section 4.2 a section of an orbibundle on an orbifold is just a sequence of sections
on the local uniformizing covers which patch together in the right way. We then
work with invariant sections by averaging over the local uniformizing groups as
in Equation 4.2.1. In this way the usual notions carry over easily to orbifolds.
For example, if X = (X, U) is a complex orbifold with an Hermitian metric g
and corresponding Hermitian 2-form ωg defined by ωg (X, Y ) = g(X, JY ). Then
X is a Kähler orbifold if ωg is a closed 2-form. We shall often denote a Kähler
orbifold by (X , ω) or by (X , g, ωg ) when we want to emphasize the orbifold metric
g. Of course, this definition is essentially the same as the manifold case, only the
interpretation of the objects is different. Here g, J and ωg are sequences of Γi -
invariant Hermitian metrics, almost complex structures, and a 2-forms, respectively,
on each local uniformizing neighborhood that are compatible with the injections
as discussed previously. With this in mind the results of Chapter 3 carry over
to orbifolds in a straightforward way. Now we are mainly interested in polarized
Kähler orbifolds whose definition is the orbifold analog of Definition 3.5.2, viz.
Definition 4.4.26: The pair (X , L) consisting of a complex orbifold X = (X, U)
with a positive line orbibundle L is called a polarized Kähler orbifold.
Exactly as in Definition 3.5.1, a positive line orbibundle is an orbibundle that
carries an Hermitian metric whose curvature form Ω with respect to the Hermitian
connection is positive. Analogous to the manifold case polarized Kähler orbifold are
algebraic. This is the content of the Kodaira-Baily Embedding Theorem [Bai57]:
Theorem 4.4.27: Let X = (X, U) be a compact complex orbifold and suppose that
X has a positive orbibundle L. Then X is a projective algebraic variety.
This theorem is proved in much that same way as the usual Kodaira embedding
theorem by noticing that if υ is the order of the orbifold then Lυ is a positive
line bundle on X in the usual sense. See [Bai57] for details. In the process one
also makes use of a vanishing theorem, that we call the Kodaira-Baily Vanishing
Theorem.
Theorem 4.4.28: Let X be a compact complex orbifold, KX the canonical orbi-
−1
bundle and L any holomorphic line orbibundle on X . If L ⊗ KX is a positive line
orbibundle, then
H q (X, O(L)) = 0 for q ≥ 1 .
4.5. WEIGHTED PROJECTIVE SPACES 133

This theorem is of much interest in its own right. In analogy with the manifold
case we have
Definition 4.4.29: A compact Kähler orbifold X = (X, ω) is called a Hodge
orbifold if [p∗ ω] lies in Horb
2
(Z, Z), where p : BZ−−→Z is the natural projection
defined in section 4.3.
With this definition a restatement of the Kodaira-Baily embedding theorem
[Bai57] is:
Theorem 4.4.30 (Baily): A Hodge orbifold is a polarized projective algebraic va-
riety.
Now Yau’s theorem and its corollaries have a direct translation to the orbifold
case.
Theorem 4.4.31: Let (X , g, ωg ) be a compact Kähler orbifold. Then any real
(1, 1)-form ρ on X which represents the cohomology class 2πcorb 1 (X ) is the Ricci
form of a unique Kähler orbifold metric h such that [ωh ] = [ωg ].
It follows from Theorem 4.3.4 that this theorem is a special case of the more
general foliation version of El Kacimi-Alaoui [EKA90] given in Theorem 7.5.19
below. The orbifold version of Theorem 3.6.6 and its corollaries follows directly
from Theorem 4.4.31. These will be used to much advantage in Chapter 5.

4.5. Weighted Projective Spaces


In this section we discuss weighted projective spaces both as singular (or non-
singular) algebraic varieties, and as complex orbifolds. It is important to regard
orbifolds as “added structure”, and distinguish the orbifold from its underlying
algebraic variety. In this sense an orbifold is a type of Deligne-Mumford stack.
We begin with the algebraic geometric treatment of weighted projective spaces.
References here are [Del75, Dol82, IF00].
Given a sequence w = (w0 , . . . , wn ) of positive integers one can form the
graded polynomial ring S(w) = C[z0 , . . . , zn ], where zi has weight wi . As an alge-
braic variety [Dol82] the weighted projective space with weights w = (w0 , . . . , wn )
is defined to be the projective scheme Proj(S(w)), and is denoted by CP(w) =
CP(w0 , . . . , wn ). Weighted projective spaces are examples of normal projective al-
gebraic varieties with only ‘mild’ quotient singularities. As in the case of ordinary
projective space the coordinates (z0 , . . . , zn ) are called homogeneous coordinates.
It is convenient to view the weights as the components of a vector w ∈ (Z+ )n+1 ,
and we shall usually assume that they are ordered w0 ≤ w1 ≤ · · · ≤ wn and
that gcd(w0 , . . . , wn ) = 18. Notice that these two conditions are not additional
assumptions since the ordering can be performed by a permutation of the affine
coordinates, and the gcd condition can be obtained by redefining the coordinate on
C∗ . It will at times be necessary to perform these operations.
In this book we are very interested in a canonical orbifold structure that comes
naturally by defining CP(w) as a certain orbit space. So our ‘official’ definition of
a weighted projective should reflect this.

8We will sometimes use the notation gcd(w , . . . , w ) ≡ (w , . . . , w ). Since (w , . . . , w )


0 n 0 n 0 n
also denotes a vector w, the latter notation will only be used when dropping the ‘gcd’ in front of
the brackets causes no ambiguity.
134 4. FUNDAMENTALS OF ORBIFOLDS

Definition 4.5.1: The weighted C∗ -action on the affine space Cn+1 , denoted by
C∗ (w), is defined by
(z0 , . . . , zn ) 7→ (λw0 z0 , . . . , λwn zn ) ,
where w = (w0 , . . . , wn ) is a sequence of positive integers.
Then we obtain the weighted projective space with a canonical orbifold struc-
ture, viz.
Definition 4.5.2: The weighted projective space CP(w) is defined as the quo-
tient space (Cn+1 \ {0})/C∗ (w).
Let us describe the canonically induced orbifold structure on CP(w). Set Ũi =
{(z1 , . . . , zm ) | zi = 1}, and let Γi ⊂ C∗ (w) be the subgroup of with roots of unity.
Note that Ũi is invariant under the action of Γi . Set Ui = Ũi /Γi . Note that the
C∗ (w)-orbits on (Cm \ {0}) \ (zi = 0) are in one-to-one correspondence with the
points of Ui . Let us explicitly construct an orbifold atlas on CP(w). We cover
CP(w) by open sets of the form
(4.5.1) Ui = {[z0 , . . . , zn ] ∈ CP(w) | zi 6= 0} .
For the local uniformizing system {Ũi , Γi , ϕi } we have open sets Ũi ≈ Cn with affine
coordinates (yi0 , . . . , ŷii , . . . , yin ) which satisfy

wi
zjwi
(4.5.2) yij = wj .
zi
Here ŷ means that element is removed. Let us use some vector notation and
write yi = (yi0 , . . . , ŷii , . . . , yin ) and yiwi = (yi0 wi wi
, . . . , ŷii wi
, . . . , yin ). The maps ϕi :
wi
Ũi −−→Ui are ϕi (yi ) = yi . The local uniformizing groups Γi are the cyclic groups
Zwi which act on Ũi by sending yi to ²yi , where ² ∈ Zwi . The origin is clearly
a fixed point of this action. The atlas U of uniformizing charts consists of the
{Ũi , Γi , ϕi } together with their nontrivial ‘intersections’ which are charts of the form
{Ũi0 ∩ · · · ∩ Ũik , Γi0 ···ik , ϕi0 ···ik }, where Γi0 ···ik is the cyclic group Zgcd(wi0 ,...,wik ) .
The map ϕi0 ···ik : Ũi0 ∩ · · · ∩ Ũik −−→Ui0 ∩ · · · ∩ Uik is given by ϕi0 ···ik (yi0 ···ik ) =
gcd(wi ,...,wik )
yi0 ···ik 0 . Here the notation yi0 ···ik designates an n-vector lying in the open
set Ũi0 ∩ · · · ∩ Ũik ⊂ Cn . We illustrate the injection maps in double overlaps. There
is an open set Ṽij ⊂ Ũi ∩ Ũj so that the injection map λij,i : Ṽij −−→Ũi is written as
t gcd(wi ,wj )
λij,i (yij ) = yij , where t = wi and we choose the principal branch. This sat-
isfies the condition ϕi ◦ λij,i = ϕij , and any other such injection map satisfying this
condition differs by a different choice of branch, that is, by an element of Γi = Zwi .
Keeping with our notation we denote this orbifold by P(w) = (CP(w), U), and refer
to it as the standard orbifold structure on CP(w). Unless otherwise stated when we
refer to an orbifold structure on CP(w) we shall always mean the standard orbifold
structure. Clearly, from Definition 4.3.9 and the above description we have
Proposition 4.5.3: The orbifolds P(w) = (CP(w), U) are locally cyclic.
Weighted projective spaces are also important examples of toric varieties [Ful93]
since they admit the action of a complex torus (C∗ )n . This torus action can be de-
scribed as follows. Let the torus (C∗ )n+1 act on Cn+1 by
(z0 , . . . , zn ) 7→ (λ1 z0 , . . . , λn zn ) .
4.5. WEIGHTED PROJECTIVE SPACES 135

Then the quotient group (C∗ )n+1 /C∗ (w) acts on CP(w) preserving the standard
orbifold structure. So P(w) = (CP(w), U) is a toric orbifold as well.
The orbifold singular locus Σorb (CP(w)) is determined by setting certain of the
homogeneous coordinates equal to zero. For any proper subset J ⊂ {w0 , . . . , wn }
let SJ denote the subset of CP(w) such that zj = 0 for all j ∈ J. Let J c denote
the compliment of J in {w0 , . . . , wn }. Then SJ lies in Σorb (CP(w)) if and only if
gcd(wi1 , . . . , wik ) > 1, where J c = {wi1 , . . . , wik }. The local uniformizing group
at generic points in SJ is then the cyclic group Zgcd(wi1 ,...,wik ) . It is clear that
Σorb (CP(w)) is nonempty as long as w 6= (1, . . . , 1), and that Σorb (CP(w)) has a
natural stratification according to the subsets SJ .
We now consider some important sheaves and orbisheaves on CP(w). For each
integer n we have the sheaves OCP(w) (n) defined as the sheaf associated to the
graded S(w)-module S(w)(m), where S(w)(m)j = S(w)m+j . But by Lemma 4.2.4
OCP(w) (n) also gives rise to an orbisheaf which we denote OP(w) (n). We are in-
terested in Picorb (P(w)). Here we make use of Proposition 4.2.18 and identify
holomorphic line V-bundles on P(w) with invertible orbisheaves. We have
Theorem 4.5.4: The Abelian group Picorb (P(w)) ≈ Z is generated by the invert-
ible orbisheaves OP(w) (±1). That is, every invertible orbisheaf (holomorphic line
V-bundle) on P(w) is of the form OP(w) (n) for some n ∈ Z.

Proof. First the isomorphism Picorb (P(w)) ≈ Z follows from Proposition


4.4.22. On the uniformizing neighborhood Ũi the orbisheaf OP(w) (1) is defined by
the sheaf ϕ∗i OUi (1) = ϕ−1
i OUi (1) ⊗ϕ−1 OU OŨi . As usual we let (z0 , . . . , zn ) denote
i i
the homogeneous coordinates on CP(w) with zi of weight wi . Then the with root

wi z is a well defined section of ϕ−1 O
i i Ui (1) ⊗ϕ−1 OU OŨi for each i = 0, . . . , n. Now
i i
any section s of ϕ−1
i OUi (1) ⊗ϕ−1 OŨi is a finite linear combination of sections
i OUi
of the form
m m
zi0 i0 · · · zin in
,
zimi
where the ik run over the n indices P different than j, and the mik and mi are
nonnegative integers satisfying ik 6=i mik wik = mi wi + 1. Moreover, by Equation
q
w √
(4.5.2), zj = yij wi zi j for all j 6= i. Thus, s = p(yi ) wi zi , where p is a polynomial

in the components of yi . Hence, wi zi is a basis for ϕ∗i OUi (1) for each p i = 0, . . . , n
which implies that the orbisheaf OP(w) (1) is invertible. Similarly, wi zim is a basis
for ϕ∗i OUi (m), so OP(w) (m) is a V-invertible orbisheaf. Next we show that there is
an isomorphism of orbisheaves

(4.5.3) OP(w) (k) ⊗ OP(w) (l) ≈ OP(w) (k + l)


p
for all k, l ∈ Z. We know that for each k ∈ Z, wi zik is a basis for the
q orbisheaf
p p

ϕi OUi (k). This gives an isomorphism of bases wi k
zi ⊗ wi l
zi ≈ wi
zik+l , and
proves the isomorphism of orbisheaves (4.5.3). This shows that OP(w) (1) gener-
ates Picorb (P(w)). Clearly a similar analysis holds for OP(w) (−1). ¤

Lemma 4.5.5: The ordinary sheaves OCP(w) (n) are not generally invertible.
136 4. FUNDAMENTALS OF ORBIFOLDS

Proof. To see that OCP(w) (1) is not generally invertible we consider CP(1, 1, 2)
([BR86], pg 130). Consider the sections of OU2 (1). It is minimally generated by
z0 and z1 , but it is not free, since for example
z0 z1 z2
z0 − 0 z1 = 0 . ¤
z2 z2
We now consider the properties of weighted projective spaces as algebraic va-
rieties. First we give an alternative description of “weighted projective spaces” as
complex spaces. Consider the finite group Gw = Zw0 × · · · × Zwn . Then CPn /Gw
is a developable compact complex orbifold. We have [Dol82]
Lemma 4.5.6: As algebraic varieties there is an isomorphism CP(w) ≈ CPn /Gw .
But no such isomorphism of orbifolds exists unless w = (1, . . . , 1).
Proof. One easily sees that the isomorphism is induced by the map zi 7→ ziwi .
However, there can be no isomorphism of orbifolds which can be seen by considering
the orbifold invariant π1orb . By the long exact homotopy sequence 4.3.18 one sees
that π1orb (CP(w)) is the identity; whereas, π1orb (CPn /Gw ) ≈ Gw . ¤
In the future by a weighted projective space we shall mean CP(w) with its
induced orbifold structure as a quotient of S 2n+1 by a locally free circle action.
There is another isomorphism of algebraic varieties between weighted projective
spaces that is more important to us in the sequel. This was first noticed by Delorme
[Del75]. First we give some notation.
Definition 4.5.7: For any weight vector w we define for each i = 0, . . . , n the
integers:
(i) di = gcd(w0 , . . . , wi−1 , wi+1 , . . . , wn ),
(ii) ei = lcm(d0 , . . . , di−1 , di+1 , . . . , dn ),
(iii) aw = lcm(d0 , . . . , dn ),
and a new weight vector w̄ = ( we00 , . . . , wenn ).
Note that (w0 , . . . , wn ) = 1 implies that (di , dj ) = 1 for all i, j, so in this case
the ei and aw are just the corresponding products. More generally the following is
easily verified:
Lemma 4.5.8: The following hold:
(i) di |wj for all j 6= i,
(ii) ei |wi for all i,
(iii) wi |υ(CP(w)) for all i,
(iv) di |υ(CP(w)) for all i,
(v) aw |υ(CP(w)).
Note that (ii) implies that the weight vector w̄ is integer-valued, and that
w = w̄ if and only if di = 1 for all i = 0, . . . , n. This leads to the notion of
well-formedness.
Definition 4.5.9: We say that the weight vector w = (w0 , . . . , wn ) or the weighted
projective space CP(w) with its standard orbifold structure P(w) is well-formed
if di = 1 for all i = 0, . . . , n.
Thus, w is well-formed if and only if w = w̄. We now have [Del75]
Proposition 4.5.10: There is an isomorphism CP(w) ≈ CP(w̄) of algebraic va-
rieties. But there is no orbifold isomorphism unless w̄ = w. In particular, as
4.5. WEIGHTED PROJECTIVE SPACES 137

algebraic varieties every weighted projective line CP(w0 , w1 ) is isomorphic to the


projective line CP1 .
Proof. Following [Dol82] define the graded subring S 0 of the graded ring
S(w) by
M
S0 = S(w)kaw .
k∈Z

Then writing S = C[y0 , . . . , yn ] we see that yi = zidi has weight aweiwi since ei di =
0

aw . Thus, by 2.4.7 of [Gro61] we have Proj(S(w̄)) ≈ Proj(S 0 ) ≈ Proj(S(w)) which


proves the first statement. For the second statement it is clear that unless w̄ = w
the weighted projective spaces CP(w̄) and CP(w) have different local uniformizing
groups. The last statement is clear from Definition 4.5.7. ¤
Remark 4.5.1: The isomorphism of Proposition 4.5.10 is due to the existence of
the reflection groups that leave hyperplanes invariant (See Theorem 4.4.1 and the
preceding paragraph). Elements of such groups are also called pseudoreflections
[Dol82] or quasireflections [IF00]. In the case of CP(w) these finite groups are the
cyclic groups Zdi .
Example 4.5.11: Here we give an explicit example of the isomorphism of Proposi-
tion 4.5.10 which relates to the famous Poincaré homology sphere as well as other
homology 3-spheres (cf. Example 10.1.10 below). The weighted projective space is
CP(6, 2(6k − 1), 3(6k − 1)), where k ≥ 1. Here we have w = (6, 2(6k − 1), 3(6k − 1))
so d0 = 6k − 1, d1 = 3, d2 = 2, aw = 6(6k − 1). Let us go through the isomorphism
in steps, first using d0 then d1 and finally d2 . We have
CP(6, 2(6k − 1), 3(6k − 1)) ∼ CP(6, 2, 3) ∼ CP(2, 2, 1) ∼ CP(1, 1, 1) = CP2 .
Thus, as algebraic varieties CP(6, 2(6k − 1), 3(6k − 1)) is isomorphic to the complex
projective plane CP2 , and similarly with the projective spaces occurring at the
intermediate stages.
The following is a result of Delorme [Del75].
Lemma 4.5.12: For any j ∈ Z there is a unique J ∈ Z and sheaf isomorphisms
OCP(w) (j) ≈ OCP(w) (Jaw ) ≈ OCP(w̄) (J) .
Proof. The isomorphisms can be described as follows (cf. [BR86]): recall
that OCP(w) (j) is the sheaf associated to the graded S(w)-module S(w)(j) de-
fined by S(w)(j)l = S(w)j+l . Now there is an equality of schemes Proj(S(w)) =
Proj(S 0 (w)), where S 0 (w) is the subring of S(w) defined by
M
S 0 (w) = S(w)kaw .
k∈Z

The isomorphisms are then induced by the equality of graded S 0 (w)-modules


M M b (j) ³ X ´
S(w)(j)kaw = z00 · · · znbn (j) S(w)(j − bi (j)wi ) ,
kaw
k∈Z k∈Z i

where bi (j) are the unique integers


¡ 0P≤ bi (j) < ¢di satisfying j = bi (j)wi + ci di for
some ci ∈ Z. So the integer J is j − i bi (j)wi /aw . ¤
Exercise 4.3: Show that if di > 1 for all i = 0, . . . , n, then there is a sheaf
isomorphism OCP(w) (|w|) ≈ OCP(w̄) .
138 4. FUNDAMENTALS OF ORBIFOLDS

It should be clear from the explicit construction of these isomorphisms that


they cannot be isomorphisms of orbisheaves. Here is an example showing that the
isomorphism of orbisheaves given by Equation 4.5.3 is not generally an isomorphism
of sheaves, that is the canonical homomorphism of sheaves
(4.5.4) OCP(w) (k) ⊗ OCP(w) (l)−−−→OCP(w) (k + l)
is not generally an isomorphism.
Example 4.5.13: [BR86] Consider the weighted projective space CP(2, 3). As al-
gebraic varieties CP(2, 3) ≈ CP1 . So by Lemma 4.5.12 and its proof there are
isomorphisms
OCP(2,3) (2) ≈ OCP1 , OCP(2,3) (4) ≈ OCP1 , OCP(2,3) (6) ≈ OCP1 (1).
It follow that the canonical homomorphism 4.5.4 cannot be an isomorphism.
Nevertheless, we do have
Lemma 4.5.14: For any integers k, l there is an isomorphism of sheaves
OCP(w) (kaw ) ⊗ OCP(w) (law )−−−→OCP(w) ((k + l)aw ).
orb
We now turn to the description of orbifold canonical divisors KP(w) of the
orbifolds P(w), and their rational Chern classes.
Lemma 4.5.15: The following hold:
P
orb
(i) KP(w) ≡ ϕ∗ KCP(w̄) + i (1 − d1i )[(zi )].
(ii) aw corb orb
1 (KP(w) ) = c1 (OP(w) (−|w|)).
In particular, any weighted projective space with its standard orbifold structure is
log Fano.
Proof. (i) follows easily from Proposition 4.4.15. To prove (ii) we notice
that the divisor (zi ) has degree w̄i , and we can clear the denominator in (i) by
multiplying by aw . This avoids working with “fractional sheaves”. So we see that
(i) and Equation 4.4.2 give
³ ¡ X 1 ¢´
aw corb orb
1 (KP(w) ) = c1 OP(w) aw (−|w̄| + (1 − )w̄i )
di
³ X 1 ´ ³ X w̄i ´
= c1 OP(w) (aw (1 − − 1)w̄i ) = c1 OP(w) (−aw )
di di
³ X wi ´
= c1 OP(w) (−aw ) = c1 (OP(w) (−|w|)) .
ei di
Here we have made use of Definition 4.5.7, Lemma 4.5.8, and aw = ei di . This proves
(ii). ¤
Remark 4.5.2: Lemma 4.5.15 is far from true generally. Example 4.6.15 below
illustrates non-standard orbifold structures on CPn that are not Fano.

4.6. Hypersurfaces in Weighted Projective Spaces

Next we are interested in certain hypersurfaces, or more generally, complete


intersections in CP(w). As we shall see throughout this text, these hypersurfaces
provide a wide range of interesting examples of Kähler orbifolds, and ultimately
Sasakian structures on odd dimensional manifolds. Although the end product al-
most exclusively involves the hypersurface case, the foundations are laid for the
4.6. HYPERSURFACES IN WEIGHTED PROJECTIVE SPACES 139

more general case of weighted homogeneous varieties. We are interested in both


their algebraic geometric as well as their orbifold properties.
Definition 4.6.1: A polynomial f ∈ C[z0 , . . . , zn ] is said to be a weighted ho-
mogeneous polynomial of degree d and weight w = (w0 , . . . , wn ) if for any
λ ∈ C∗
f (λw0 z0 , . . . , λwn zn ) = λd f (z0 , . . . , zn ) .
The space of weighted homogeneous polynomials of degree d is a λd -eigenspace
of the induced C∗ (w)-action on C[z0 , . . . , zn ] and in analogy with the classical case
will be denoted by H 0 (CP(w), O(d)). If V ⊂ Cn+1 is a variety defined by a collection
of weighted homogeneous polynomials f1 , . . . , fp with weights (w0 , . . . , wn ), then V
is invariant under the C∗ -action 4.5.1. This statement has a converse due to Orlik
and Wagreich [OW71b].
Proposition 4.6.2: Let V ⊂ Cn+1 be an irreducible analytic variety which is in-
variant under the C∗ -action defined by Equation 4.5.1. Then V is an algebraic
variety and the ideal of polynomials vanishing on V is generated by weighted homo-
geneous polynomials of Definition 4.6.1.

Proof. Let C{z1 , . . . , zn } denote the ring of convergent power series. Then
for f ∈ C{z0 , . . . , zn } there is a unique polynomial fi for each i such that

X
(4.6.1) f (λ · z) = λi fi (z) .
i=0

This series converges for λ small enough. P i If f vanishes on V near the origin 0
in Cn+1 , then z ∈ V implies that i λ fi (z) = 0 for λ small enough. Hence,
fi (z) = 0 for all i and all z ∈ V near o. Let I(V ) denote the ideal of functions in
(j)
C{z0 , . . . , zn } vanishing on V, and let J be the ideal generated by {fi }, where
(j)
the f ’s are generators of I(V ). Clearly, J ⊂ I(V ). Now the locus of zeroes of
J is V, since if z 6∈ V and is within the radius of convergence of f (j) for all j,
(j) (j)
there must be some fi with fi (z) 6= 0. Thus, by Hilbert’s Nullstellensatz the
(j)
radical of J is I(V ). Let J 0 √ be the ideal generated by {fi √} in the polynomial ring
0 0 0
C[z0 , . . . , zn ], and let I = J . Then I C[z0 , . . . , zn ] = J = I(V ). So I(V ) is
generated by polynomials, and the algebraic variety defined by I(V ) is V. Now if f
is polynomial in I(V ) then there is only a finite number of i’s in Equation (4.6.1)
with fi different from zero. Since the f (j) ’s are generators of I(V ), the weighted
(j)
homogeneous polynomials fi also generate I(V ). ¤

Remark 4.6.1: Notice that for a given (n + 1)-tuple of weights w = (w0 , . . . , wn )


and degree d there may be no such polynomial f. So the existence of a weighted
homogeneous polynomial f ∈ C[z0 , . . . , zn ] puts constraints on the set of the pair
(w; d).
In analogy with Definition 3.1.19 we have
Definition 4.6.3: A weighted (homogeneous) variety X is the common zero
locus in CP(w) of a collection of weighted homogeneous polynomials in C[z0 , . . . , zn ].
A weighted (homogeneous) hypersurface Xf is the zero locus in CP(w) of
a single weighted homogeneous polynomial f. A weighted variety X is called a
weighted complete intersection if the number of polynomials in the collection
equals the codimension of X.
140 4. FUNDAMENTALS OF ORBIFOLDS

Since each weighted homogeneous polynomial has a degree, we can associate


precisely c = codim(X) degrees (d1 , . . . , dc ) to a weighted complete intersection X.
We call d = (d1 , . . . , dk ) the multidegree of the complete intersection.
Notice that a weighted variety defines another affine variety as the zero locus in
Cn+1 of a collection of weighted homogeneous polynomials which is clearly invariant
under the C∗ (w)-action. This leads to
Definition 4.6.4: Let X ⊂ CP(w) be a weighted variety in CP(w) and π : Cn+1 \
{0} → CP(w) the canonical projection. The punctured affine cone C ∗ (X) over
X is defined by C ∗ (X) = π −1 (X) and the affine cone C(X) is the completion of
C ∗ (X) in Cn+1 .
Notice that C ∗ (X) cannot have isolated singularities. We shall assume that
when a weighted homogeneous polynomial f does exist that it is non-degenerate in
the sense that each variable occurs. In this treatise we shall only be interested in
the case where either f is regular at the origin o or when o is an isolated singular
point. Notice that f will be regular at o if and only if some variable, say zn , occurs
as a linear monomial in f, and this occurs if and only if wn = d. In this case we
have a linear cone and we can remove the variable zn and consider the reduced
weighted homogeneous polynomial g ∈ C[z0 , . . . , zn−1 ]. Below we give conditions
on (w; d) such that the hypersurface Xf has an isolated singularity at the origin.
Definition 4.6.5: We say that a weighted variety X ⊂ CPn (w) is quasi-smooth
of dimension m if its affine cone C(X) is smooth of dimension m + 1 outside the
origin 0, that is, C ∗ (X) is smooth.
When X ⊂ CP(w) is quasi-smooth all its singularities are quotient singularities.
In fact since a weighted homogeneous variety is invariant under the C∗ (w)-action
we have
Proposition 4.6.6: Let ι : X−−→CP(w) be a quasi-smooth weighted homogeneous
variety. Then the orbifold structure P(w) = (CP(w), U) on CP(w) naturally in-
duces a locally cyclic orbifold structure X on X.

Proof. Since X is quasi-smooth and invariant under the C∗ -action, the local
uniformizing groups of the orbifold X are just the isotropy subgroups of the weighted
C∗ (w)-action on the punctured affine cone C ∗ (X) ⊂ Cn+1 . If x ∈ X has local
uniformizing group Γx , then this is also the local uniformizing group of ι(x) ∈
CP(w). So the set of local uniformizing groups of the orbifold X is precisely the
subset of local uniformizing groups of P(w) taken over all image points of ι. Then
since P(w) is a cyclic orbifold, so is X . ¤

We will sometimes emphasize the degree d of a weighted variety and write X =


Xd ⊂ CP(w). Note that the pair (d, w) determines not just one f satisfying the non-
degeneracy requirement, but rather a family of polynomials in H 0 (CP(w), O(d))
and by abuse of notation we shall write interchangeably Xd for a specific member
or a sufficiently general member of this deformation family. The context should
make clear which is meant. Let us consider an important example which will play
a significant role in this book.
Example 4.6.7: Consider weighted homogeneous polynomials f of the form

(4.6.2) f (z0 , . . . , zn ) = z0a0 + · · · + znan ∈ H 0 (CP(w), O(d)) ,


4.6. HYPERSURFACES IN WEIGHTED PROJECTIVE SPACES 141

where ai > 1 and wi ai = d for all i = 0, . . . , n. Such polynomials were introduced


by Pham [Pha65] and used by Brieskorn [Bri66] in his famous study of exotic
spheres. They have become known as Brieskorn-Pham polynomials, and we shall
often employ the abbreviation BP polynomials. Note that since we always assume
gcd(w0 , . . . , wn ) = 1, we have d = lcm(a0 , . . . , an ). It is easy to see that the zero
locus Xf ⊂ CP(w) is quasi-smooth, so Xf is an orbifold with orbifold structure in-
duced by the standard orbifold structure of CP(w). This orbifold structure depends
on the pair (w; d) or equivalently on the exponent vector a = (a0 , . . . , an ). We denote
the induced orbifold structure by X (a). Now we can perturb the BP polynomial f
as follows. Let p ∈ H 0 (CP(w), O(d)) be any weighted homogeneous polynomial of
degree d and weight vector w. Let Xa,p = Xf +p denote the corresponding weighted
homogeneous hypersurface together with its affine cone Ca,p = Cf +p . We say that
Xa,p and/or Ca,p is a (weighted homogeneous) perturbation of the BP polynomial f
if the following genericity condition holds:
(GC) The intersections of Ca,p with any number of hyperplanes (zi = 0)
are all smooth outside the origin.
We shall often be interested in a ‘generic’ member of this family.
We can naturally extend Definition 4.6.3 to weighted complete intersections,
and as in the hypersurface case we will write Xd1 ,...,dk ⊂ CP(w) to denote a suffi-
ciently general member of the family defined by the pair (w; d).
We now extend the notion of well-formedness given in Definition 4.5.9 to sub-
varieties of weighted projective spaces. This seems to have first appeared in [IF00].
Definition 4.6.8: A subvariety X ⊂ CP(w) of codimension k is well-formed if
CP(w) is well-formed and X contains no codimension k + 1 singular stratum of
CP(w). It follows that codimX (X ∩ CP(w)sing ) ≥ 2.
It follows from Definition 4.6.8 and Theorem 4.4.1 that a quasi-smooth well-
formed variety X ⊂ CP(w) is a locally cyclic complex orbifold with no branch
divisors. Thus, we have
Proposition 4.6.9: Let X be a quasi-smooth subvariety of CP(w). Then X is
well-formed if and only if its orbifold singular locus and algebro-geometric singular
locus coincide, equivalently X has no branch divisors.
In the case of complete intersections the well-formedness condition can be ex-
pressed as a divisibility condition on the pair (w; d).
Corollary 4.6.10: Let Xd ⊂ CP(w) be a weighted complete intersection of multi-
degree d = (d1 , . . . , dk ). Then Xd is well-formed if and only if CP(w) is well-formed
and for all 1 ≤ m ≤ k the gcd of any (n − 1 − k + m) of {w0 , . . . , wn } must divide
at least m of {d1 , . . . , dk }. If Xd ⊂ CP(w) is a weighted homogeneous hypersurface,
then Xd is well-formed if and only if CP(w) is well-formed and for all 0 ≤ i < j ≤ n
we have gcd(w0 , . . . , ŵi , . . . , ŵj , . . . , wn )|d. In particular, if Xd = Xf , where f is a
Brieskorn-Pham polynomial, then Xd is well-formed if and only if CP(w) is well-
formed.
We continue the discussion of Example 4.6.7.
Example 4.6.11: For a BP polynomial f (or a perturbation thereof) of Equation
4.6.2, we see that the branch divisor of X (a) is
X³ 1´
∆= 1− Di ,
i
di
142 4. FUNDAMENTALS OF ORBIFOLDS

where di is defined in Definition 4.5.7 and Di are the hyperplane divisors (zi = 0).
Notice that if di = 1 there is no contribution of Di to the branch divisor, or
equivalently Di is not contained in the orbifold singular locus Σorb (X). Of course,
in the well-formed case di = 1 for all i, and there is no branch divisor. When
we want to emphasize the appearance of branch divisors we shall use the notation
X (a) = (X, ∆) with the a or equivalently the BP polynomial f understood. We can
obtain other orbifold structures on the same variety X by applying the procedure
of Section 4.5. Ultimately, we end up with a well-formed weight vector w̄ with the
orbifold structure (X, ∅). This will give rise to a new BP polynomial with a new
exponent vector b which we describe in more detail below.
It follows from Proposition 4.4.15 that in the non-well-formed case the orbifold
canonical divisor and canonical divisor are not the same. In this case it is the
orbifold canonical bundle and corresponding orbisheaf that is of interest to us. In
either case the adjunction formula holds as usual as long as one replaces the canon-
orb
ical bundle KX by the orbifold canonical bundle KX . Somewhat more generally
we have [BGK05]
Proposition 4.6.12: Let f ∈ H 0 (CPn (w), O(d)) and let Xf and Cf be as in
Definition 4.6.3. Assume also that Xf is quasi-smooth, n ≥ 2 and s(d − |w|) > 0.
Then the following three spaces are naturally isomorphic:
(i) Global sections of ⊗s KXorb
.
(ii) C (w)-invariant global sections of ⊗s KCf .

(iii) The space of weighted homogeneous polynomials of weight s(d − |w|),


modulo multiples of f .
Proof. Differential forms on Xf correspond to C∗ -invariant differential forms
on Cf \ {0}, so global sections of ⊗s KXorb
can be identified with C∗ (w)-invariant
s
global sections of ⊗ KCf \{0} . If n ≥ 2 then Cf is a hypersurface of dimension
≥ 2 with an isolated singularity at the origin, thus normal. Hence, by Hartogs
global sections of ⊗s KCf agree with global sections of ⊗s KCf \{0} . This shows the
equivalence of (i) and (ii).
The C∗ (w)-action on η has weight |w| − d, thus ⊗s KCf is the trivial bundle on
Cf , where the C∗ (w)-action has weight s(|w| − d). Its invariant global sections are
thus given by homogeneous polynomials of weight s(d − |w|) times the generator
η. ¤
Putting s = 1 in Proposition 4.6.12 gives the adjunction formula
orb
(4.6.3) KX = OX (d − |w|) .
The general adjunction formula can be proved as usual as long as one replaces the
canonical divisor by the orbifold canonical divisor.
Proposition 4.6.13 (Adjunction formula): Let D be a quasi-smooth orbidivisor
in the weighted projective orbifold P(w). Then the adjunction formula
orb orb orb
KD = O(KD ) = O(KP(w) + D)|D
holds.
It follows that
d·corb orb
1 (KX (a) ) = c1 (OP(w) (d − |w|))
and we get
4.6. HYPERSURFACES IN WEIGHTED PROJECTIVE SPACES 143

Corollary 4.6.14: Let Xd ⊂ CP(w) be a quasi-smooth weighted hypersurface of


degree d. The orbifold Xf = (Xf , ∆) is log Fano if and only if d < |w|.
One can easily generalize this statement to the case of complete intersections.
Let us consider an example of a family of quasi-smooth but not well-formed hy-
persurfaces which illustrates various concepts discussed earlier as well as being of
interest to us later.
Example 4.6.15: Consider the Brieskorn-Pham polynomials f of Examples 4.6.7
and 4.6.11 where now the (n+1)-tuple a = (a0 , . . . , an ) consists of pairwise relatively
prime integers for all i = 0, . . . , n. Then d = a0 · · · an and wi = a0 · · · aˆi · · · an , so
di = ai . Clearly, CP(w) is not well-formed and as an algebraic variety we have
CP(w) = CPn . Likewise, it is easy to see that Xd ⊂ CP(w) is simply CPn−1 .
However, the orbifold structure very much depends on a with
µ X³ ¶
1 ´
X (a) = (Xd , ∆) = CPn−1 , 1 − Di ,
ai
where Di ⊂ CPn−1 are the coordinate hyperplanes zi = 0. According to the Corol-
lary 4.6.14 the orbifold X (a) is log Fano if and only if
Xn
1
(4.6.4) 1< .
a
i=0 i

Thus Xd = CPn−1 is smooth and certainly Fano as an algebraic variety while the
orbifold X (a) = (CPn−1 , ∆) is log Fano only when the inequality (4.6.4) holds.
This can be seen by computing the orbifold first Chern class of X (a). Let [H] be
the hyperplane class. Then c1 (CPn−1 ) = n[H], [Di ] = [H] and, hence,
Xn µXn ¶
1
corb
1 (X (a)) = n[H] − (1 − 1/ai )[H] = − 1 [H] .
i=0
a
i=0 i

When n = 2 the orbifold X (a) is topologically just CP1 ' S 2 , but the orbifold
structure and the geometry of X (a) depend on a in a rather complicated way
[Mil75]. In particular, X (a) admits an orbifold Riemannian metric of constant
positive curvature if and only if it is log Fano. As we are considering the case that
a0 , a1 , a2 are pairwise relatively prime, this occurs in only one case, a = (2, 3, 5). If
P2
1 > i=0 a1i the orbifold geometry is hyperbolic, and one can show that X (a) is
actually a quotient of a smooth Riemann surface Σg of genus g = g(a) > 0 by a
discrete group of isometries. Hence, in such a case X (a) admits a negative constant
curvature orbifold metric. See Example 4.4.19 as well as Proposition 4.1.7. We
shall revisit this and similar examples in Chapters 5, 9, 10 and 11.
More generally, for the case of a Riemann surface Σw realized as hypersurfaces
of a weighted homogeneous polynomial in CP(w0 , w1 , w2 ), one has the genus formula
[OW71a, Orl72b]
1 ³ d2 X gcd(wi , wj ) X gcd(d, wi ) ´
(4.6.5) g(Σ(w;d) ) = −d + −1 .
2 w0 w1 w2 i<j
wi wj i
wi

Notice that this equation reduces to the well-known (cf. [GH78b]) genus formula
g = 21 (d − 1)(d − 2) for smooth curves of degree d in CP2 when all the weights are
1.
144 4. FUNDAMENTALS OF ORBIFOLDS

We finish this section with the quasi-smoothness criterion for hypersurfaces


Xd ⊂ CP(w). We first note that all linear cones are quasi-smooth. If a weighted
hypersurface is not a linear cone we get [IF00]
Theorem 4.6.16: Let Xd ⊂ CPn (w) be a general hypersurface which is not a
linear cone. Then Xd is quasi-smooth if and only if for every non-empty subset
I = {i0 , . . . , ik−1 } ⊂ {0, 1, . . . , n} either one of the two following conditions holds
m
(i) There exists a monomial zim0 0 · · · zik−1
k−1
∈ H 0 (CP(w), O(d)).
m0,µ mk−1,µ
(ii) For all µ = 1, . . . , k there exist zi0 · · · zik−1 zeµ ∈ H 0 (CP(w), O(d)),
where {eµ } are k distinct elements.
Exercise 4.4: [Quasi-smoothness criterion for surfaces] Show that a general
hypersurface Xd ⊂ CP(w0 , w1 , w2 , w3 ) is quasi-smooth if and only if all of the
following three conditions hold [IF00, JK01b]
(i) For each i = 0, . . . , 3 there is a j and a monomial zimi zj ∈ H 0 (CP(w), O(d)).
Here j = i is possible.
b
(ii) If gcd(wi , wj ) > 1 then there is a monomial zibi zj j ∈ H 0 (CP(w), O(d)).
b
(iii) For every i, j either there is a monomial zibi zj j ∈ H 0 (CP(w), O(d)),
c d
or there are monomials zici zj j zk and zidi zj j zl ∈ H 0 (CP(w), O(d)) with
{k, l} 6= {i, j}.

4.7. Seifert Bundles


Seifert fibred 3-manifolds were originally introduced and studied by Seifert in
1932 [Sei32, ST80]. Roughly speaking, these are 3-manifolds which are unions
of circles in a very particular way. More precisely, one could start with a trivial
fibred solid torus S 1 × D2 . Then a fibred solid torus is a solid torus which is finitely
covered by a trivial solid torus. This can be obtained by cutting the trivial solid
torus along a disc, rotating the disc through the angle 2πq/p and gluing it back
together. A fibred Klein bottle is a solid Klein bottle which is doubly covered by a
trivial solid torus. Up to fibred homeomorphism there is only one such fibred solid
Klein bottle.
Definition 4.7.1: A Seifert fibred 3-space is a 3-manifold M with a decomposition
of M into circles called fibres such that each circle in M has a neighborhood in M
which is a union of fibers and it is isomorphic to a fibred solid torus or Klein bottle.
Seifert original definition [Sei32] excludes the Klein bottle case. The Definition
4.7.1 is due to Scott [Sco83] and in the world of 3-manifolds it has some obvious
advantages. Epstein proved that a compact 3-manifolds is a Seifert fibred space
(in the above sense) if and only if it is fibred by circles [Eps72]. Hence, one
can introduce the transverse space X with the canonical projection π : M −→ X.
Clearly, X is an orbifold surface and M , in a certain sense, is an orbifold bundle
over X. However, this definition is perhaps too general from our point of view.
For example, it can be seen that if M is a Seifert fibred space then M admits an
S 1 -action leaving the foliation invariant if and only if one can coherently orient all
the fibers. Hence, in general a Seifert fibred 3-space is not a principal S 1 -orbifold
bundle over a compact 2-dimensional orbifold surface. We will be slightly less
general in defining Seifert fibred n-spaces for which both M n and X are (naturally)
oriented. On the other hand we want to be able to work with Seifert fibrations in
4.7. SEIFERT BUNDLES 145

arbitrary dimension. For the extensive discussion of the general theory of Seifert
fibred 3-spaces we refer the interested reader to [Sco83].
Higher dimensional Seifert fibred manifolds were investigated first Orlik and
Weigreich [OW75] who observed that in many cases of interest Seifert fibred mani-
folds correspond to holomorphic Seifert C∗ -bundles, and started to develop a general
theory of holomorphic Seifert G-bundles for any complex Lie group G. More re-
cently Kollár has generalized this construction and we closely follow his approach
[Kol04, Kol05]. First we give some preliminaries. Recall [GR84] that a complex
space X (all our complex spaces are reduced and second countable unless other-
wise stated) is holomorphically convex if for every infinite discrete closed subset S
of X there is a holomorphic function whose set of values on S is unbounded. A
holomorphically convex space X is Stein if every compact complex subspace of X
is finite.
Definition 4.7.2: Let X be a normal complex space. A Seifert C∗ -bundle over
X is a normal complex space Y together with a holomorphic map f : Y → X and
a C∗ -action on Y satisfying the following two conditions.
(i) The pre-image f −1 (U ) of any open Stein subset U of X is Stein and
C∗ -equivariant (with respect to the trivial action on X).
(ii) For every p ∈ X, the C∗ -action on the fiber Yp = f −1 (p), C∗ × Yp → Yp
is C∗ -equivariantly biholomorphic to the natural C∗ -action on C∗ /µm for
some m = m(p, Y /X), where µm ⊂ C∗ denotes the group of mth roots of
unity.
The number m(p, Y /X) is called the multiplicity of the Seifert fiber over p.
One can always assume that the C∗ -action is effective, that is, m(p, Y /X) =
1 for generic p ∈ X. Note that even when Y is smooth, X can have quotient
singularities.
A classification of Seifert C∗ -bundles for X a smooth manifold with H1 (X, Z)
torsion free is given in [OW75]. The most general case in arbitrary characteristic
is discussed in detail in [Kol04]. We begin with the following description of Seifert
C∗ -bundles Let f : Y → X be a Seifert C∗ -bundle. The set of points {p ∈ X :
m(p, Y /X) > 1} is a closed analytic subset of X. It can be written as the union
of Weil divisors ∪Di ⊂ X and of a subset of codimension at least 2 contained in
Sing(X). The multiplicity m(p, Y /X) is constant on a dense open subset of each
Di , this common value is called the multiplicity of the Seifert
P C∗ -bundle over Di ;
denote it by mi = m(Di ). We call the Q-divisor ∆ = (1 − m1i )Di the branch
divisor of f : Y → X. We will write f : Y → (X, ∆) to indicate the branch divisor.
Kollár shows that there is the following correspondence [Kol04]
Theorem 4.7.3: Let P X be a normal reduced complex space with at worst quotient
singularities and ∆ = (1 − m1i )Di a Q-divisor. There is a one-to-one correspon-
dence between Seifert C∗ -bundles f : Y → (X, ∆) and the following data:
(i) For each Di an integer 0 ≤ bi < mi , relatively prime to mi , and
(ii) a linear equivalence class of Weil divisors B ∈ Div(X).
We now wish to relate Seifert bundles to orbibundles on orbifolds. First follow-
ing [Kol05] we consider the case when X is smooth as in [OW75]. First let |∆|
denote the support of the Q-divisor ∆, and set m(∆) = lcm(mi ). Then Y \f −1 (|∆|)
is an m(∆)-sheeted
¡ P bi covering
¢ space of the C∗ -bundle over X\|∆| with Chern class
m(∆) [B] + i mi [Di ] . (Here for convenience we adopt the common convention of
146 4. FUNDAMENTALS OF ORBIFOLDS

identifying a line bundle [D] associated with a divisor with its Chern class c1 ([D]).)
This is a well-defined integral cohomology class since the mi divide m(∆) for all i.
More generally if X is singular but satisfies the hypothesis of Theorem 4.7.3, then
the Seifert bundle f : Y −−→(X, ∆) is determined uniquely by its restriction to the
smooth locus Xreg . Now we define the Chern class of the Seifert bundle as a rational
cohomology class of X which defines an integral class p∗ c1 (Y /X) ∈ Horb2
(X , Z).
Definition 4.7.4: The Chern class of the Seifert bundle f : Y → (X, ∆) is
defined as
X
bi 2
c1 (Y /X) = [B] + mi [Di ] ∈ H (X, Q) ,

where B is a Weil divisor as in Theorem 4.7.3.


The triple (Di , mi , bi ) are often called orbit invariants of the Seifert bundle.
We shall see in Theorem 4.7.10 below that under fairly mild conditions a Seifert
bundle is uniquely determined by its Chern class. We are now ready for
Theorem 4.7.5: Every C∗ -Seifert bundle f : Y → (X, ∆) with Y smooth is a
C∗ -orbibundle over the locally cyclic orbifold X = (X, ∆). Conversely, if (X, ∆) is
locally cyclic orbifold and f : Y −−→(X, ∆) is a C∗ -orbibundle over X whose local
uniformizing groups inject into the group C∗ of the orbibundle, then f : Y −−→(X, ∆)
is a C∗ -Seifert bundle with Y smooth.

Proof. Let f : Y → (X, ∆) be a Seifert C∗ -bundle with Y smooth. For p ∈ X


pick any y ∈ f −1 (p) and a µm -invariant smooth hypersurface Vp ⊂ Y transversal
to f −1 (p) for m = m(p, Y /X). Then {ϕp : Vp → Up = Vp /µm } gives an orbifold
structure on X = (X, U) on X. The orbifold branch divisor coincides with the
branch divisor of the Seifert bundle. Since Y is smooth each Up is a quotient by the
cyclic group µm . Hence X = (X, P ∆) is locally cyclic. Now consider the Weil divisor
bi
B ∈ Div(X) and the Q-divisor mi Di ∈ Div(X) ⊗ Q. The Weil divisor B lifts
to Baily divisor B̃ on P
X and since the mi are the ramification indices of a branch
bi
divisor, the Q-divisor Di also lifts to a Baily divisor R by Proposition 4.4.13.
Pm bi
i
So the Q-divisor B + mi i lifts to a unique Baily divisor which by Proposition
D
4.4.14 corresponds to an orbibundle over X . For the converse one can simply retrace
the steps and use Lemma 4.2.8. ¤

Our main interest in Seifert bundles involves the circle subgroup of C∗ . As a


real Lie group, C∗ ∼
= S 1 × R. Since a Seifert bundle involves subgroups of S 1 only,
we can write Y = M × R. Then restricting f to M we have
Definition 4.7.6: We call f : M → (X, ∆) a Seifert S 1 -bundle or simply a
Seifert bundle.
Clearly, Y and M have isomorphic homology and homotopy groups, and Y is
smooth if and only if M is smooth. So it follows that f : M → (X, ∆) is a principal
S 1 -orbibundle over the orbifold X = (X, ∆), and Theorem 4.7.5 becomes
Theorem 4.7.7: Every Seifert S 1 -bundle f : M → (X, ∆) with M smooth is a S 1 -
orbibundle over the locally cyclic orbifold X = (X, ∆). Conversely, if (X, ∆) is lo-
cally cyclic orbifold and f : M −−→(X, ∆) is a S 1 -orbibundle over X whose local uni-
formizing groups inject into the group S 1 of the orbibundle, then f : M −−→(X, ∆)
is a Seifert S 1 -bundle with M smooth.
4.7. SEIFERT BUNDLES 147

In general, the total space of an orbibundle or Seifert bundle over an orbifold


is an again an orbifold. So we want to characterize the smoothness of M or equiv-
alently Y in terms of algebro-geometric properties. For this we describe the divisor
class group of a quotient singularity. Let DimC X = n, and pick a point x ∈ X such
that Y is smooth along f −1 (x). Let {Ũ , Γ, ϕ} be a local uniformizing system for X
with x ∈ U = ϕ(Ũ ). The local uniformizing group is the cyclic group µm . We can
take Ũ to be the polydisc Dn ⊂ Cn in which case U = ϕ(Ũ ) is biholomorphic to
Dn /µm . By diagonalizing the cyclic group action, we see that locally the irreducible
components of ∆ passing through x are the quotients of (some of) the coordinate
hyperplanes in Dn . By an analysis similar to that in the proof of Lemma 4.5.12
we can obtain equivalent varieties by removing the reflection elements of µm which
amounts to writing m = r · m1 · · · mn , where the m1 , . . . , mn are pairwise relatively
prime and the multiplicities of the irreducible components of ∆ passing through x.
(We add the necessary number of 1’s when there are fewer than n such components.)
The quotient D̄n = Dn /µm1 ···mn is smooth, so as varieties Dn /µm ∼ = D̄n /µr and
the µr -action is fixed point free outside a codimension 2 subset. So the local divisor
class group Cl(X, x) at x is the cyclic group µr(x) ≈ Zr . Given a Weil divisor D
containing x, we can restrict it to the neighborhood U ∼ = Dn /µr(x) of x. This gives
a well defined map Rx : Cl(X)−−→Cl(X, x) ≈ Zr(x) whose kernel is precisely the
subgroup of classes of locally principal divisors at x. Following Kollár we introduce
some useful notation:
1. m(x, ∆) = lcm(mi | x ∈ Di ),
2. m(∆) = lcm(m(x, ∆) | x ∈ X) = lcm(mi ),
3. m(x, X) = r(x),
4. m(X) = lcm(m(x, X)) | x ∈ X),
5. m(x, X, ∆) = m(x),
6. m(X, ∆) = lcm(m(x, X, ∆) | x ∈ X) = lcm(m(x) | x ∈ X) = υ(X).
Note that m(x, X, ∆) = m(x, ∆)m(x, X) and m(X, ∆)|m(∆)m(X) but the latter
can be different. Also since mi divides m(∆) for each i, the class m(∆)c1 (Y /X)
gives a well defined element of Cl(X). Similarly, m(X, ∆)c1 (Y /X) gives a well
defined element of PicX. Now we define
X m(x, ∆)bi
(4.7.1) m(x, ∆) · c1 (Y /X) = m(x, ∆)[B] + [Di ]
mi
i|x∈Di

and note that it is a well defined element of Cl(X). Kollár then gives the following
simple smoothness criterion [Kol05]:
P bi
Proposition 4.7.8: Let Y = Y (B, mi Di ) be a Seifert bundle over the orbifold
P
(X, (1 − m1i )Di ). Then Y is smooth along f −1 (x) if and only if Rx (m(x, ∆) ·
c1 (Y /X)) is a generator of the local class group Cl(X, x).
P
Let M → (X, (1 − m1i )Di ) be the associated Seifert S 1 -bundle. This, as we
P
remarked is a principal circle orbibundle over the orbifold (X, (1 − m1i )Di ). It is
easy to see that this is equivalent to the requirement that the orbifold uniformizing
groups inject into the structure group S 1 of the principal orbibundle.
Some Pfairly general information can be obtained for a general Seifert bundle
M → (X, (1 − m1i )Di ) with a smooth total space. By Leray’s Theorem there is
148 4. FUNDAMENTALS OF ORBIFOLDS

a Leray spectral sequence (in cohomology) whose E2 term is


(4.7.2) E2i,j = H i (X, Rj f∗ ZM ) =⇒ H i+j (M, Z) .
Since every fibre is an S 1 , the only non-vanishing derived functor sheaf is R1 f∗ ZM .
As with Corollary 4.3.8 it is straightforward to show
P
Proposition 4.7.9: Let M → (X, (1 − m1i )Di ) be a Seifert S 1 -bundle, and let
A is any group in which υ(X) is invertible. Then
(i) there is an isomorphism R1 f∗ AM ≈ AX : hence, there is a Gysin sequence
ι∗ jp δ
· · · −−→H p (X, AX ) −−→ H p (M, AM ) −−→ H p−1 (X, AX ) −−→ H p+1 (X, AX )−−→ · · · ,
(ii) there is a natural injection τ : R1 f∗ ZM −−→ZX which is an isomorphism
on points where m(x) = 1,
(iii) the differential d2 : E 0,1 = H 0 (X, R1 f∗ AM )−−→H 2 (X, AX ) is identified
with cupping with c1 (Y /X).
Even so, in general, it is not easy to compute the cohomology ring of the bundle,
or even understand the fundamental group π1 (M ). However, assuming that M is
smooth, from Theorem 4.3.18 we get an exact sequence
(4.7.3) π2orb (X, ∆)−−→Z−−→π1 (M )−−→π1orb (X, ∆)−−→1
which often allows for an explicit computation of H1 (M, Z). In principle, the orb-
ifold invariants of (X, ∆) and the first Chern class of the Seifert bundle should be
sufficient to determine the cohomology of M at least when H1orb (X , Z) = 0. Indeed
using results in [OW75] Kollár [Kol04] proves
Theorem 4.7.10: Let X = (X, ∆) be a complex orbifold such that H1orb (X , Z) = 0.
Then a Seifert C∗ -bundle (hence, a Seifert S 1 -bundle) f : Y −−→(X, ∆) is uniquely
determined by its Chern class c1 (Y /X) ∈ H 2 (X, Q).
Another important result of [Kol05] says that when H1orb (X , Z) = 0 there are
no branch divisors on a Calabi-Yau orbifold.
P
Proposition 4.7.11: Let f : M → (X , (1 − m1i )Di ) be a Seifert bundle, M
smoothPand H1orb (X , Z) = 0. Suppose that X P is a Calabi-Yau orbifold, that is,
KX + (1 − m1i )Di is numerically trivial. Then (1− m1i )Di = 0 and KX orb
≡ KX .
So it is important to have conditions for the vanishing of H1orb (X , Z) or even
better for π1orb (X ). Necessary and sufficient conditions were worked out by Kollár
[Kol05] in the case of complex dimension 2, but before describing these we give a
result of importance to us.
Proposition 4.7.12: Let X = (X, U) be a quasi-smooth weighted complete inter-
section in CP(w) with dimX ≥ 2. Then π1orb (X ) = 0.

Proof. This is a direct consequence of Proposition 9.6.1 below and the long
exact homotopy sequence 4.3.18 applied to the Seifert S 1 -bundle. ¤

4.7.1. Seifert Bundles of Dimension Five. We now specialize to the case


of complex surfaces where much more is known, but first we need a general definition
concerning singularities. We recall [KM98, Laz04a] that a normal variety V has
rational singularities if there exists a resolution of singularities π : V 0 −−→V such
that the derived functor sheaves Ri π∗ OV 0 vanish for i > 0.
4.7. SEIFERT BUNDLES 149

Proposition 4.7.13: Let X be a normal projective surface with rational singu-


larities,
P and let X = (X, ∆) be an orbifold structure with branch divisor ∆ =
1 orb
i (1 − mi )Di . Then H1 (X , Z) = 0 if and only if the following conditions hold:
(i) H1 (Xreg , Z) = 0, and P
(ii) the map ψ : H 2 (X, Z)−−→ j Zmj defined by ψ(α) = α(Dj ) mod mj is
surjective.
If in addition H1 (X, Z) = 0 and H2 (X, Q) = Q, then H1orb (X , Z) = 0 if
and only if
(iii) Cl(X) ≈ Z,
(iv) the mj are pairwise relatively prime, and
(v) for every j the degree of Dj is relatively prime to mj .
Again one can say more for projective surfaces; Kollár [Kol05] shows that in
this case the spectral sequence degenerates at the E 3 term and one obtains:
P 1
Theorem 4.7.14: Let f : M 5 → (X, ∆ = i (1 − mi )Di ) be a smooth Seifert
bundle over a projective surface with rational singularities. Set s = b2 (X) and
assume that H1orb (X , Z) = 0. Then the cohomology groups H i (M 5 , Z) are
i 0 1 2 3 4 5
P
H i (M 5 , Z) Z 0 Zs−1 ⊕ Zd Zs−1 ⊕ i (Zmi )
2g(Di )
Zd Z
where d is the largest natural number such that m(∆)c1 (Y /X) ∈ Cl(X) is divisi-
ble by d, and g(D) is the genus of the Riemann surface D. (Here we employ the
convention that Z1 = {0}).
For details we refer to the literature [OW75, Kol05]. Theorem 4.7.14 says
that the torsion in H 3 (M 5 , Z) (or equivalently in H2 (M 5 , Z) when H1 (M 5 , Z) = 0)
is determined by the nonrational curves in the branch divisor ∆. Indeed one has
an immediate corollary of Theorem 4.7.14 and Proposition 4.6.9 which was first
obtained in [BG01b] by the authors using a result of Randell [Ran75].
Corollary 4.7.15: Let Lf be the link of an isolated hypersurface singularity de-
fined by a weighted homogeneous polynomial f in four complex variables, and let X
denoted the corresponding hypersurface in CP(w0 , w1 , w2 , w3 ). Suppose further that
X is well-formed, then Tor (H2 (Lf , Z)) = 0.
We end this section by considering an example that illustrates some of the ideas
presented in this chapter.
Example 4.7.16: We build on our discussion of Example 4.4.9 where X = CP(1, 2, 3),
but with a different branch divisor as in Example 6.6 of [Kol05]. Here the branch
divisor has two irreducible components, namely ∆ = 21 D0 + 10 0
11 D , where D =
0
6 3 2
(z0 + z1 + z2 = 0) and D0 is as in Example 4.4.9. Recall also that D0 is a generator
orb
of Cl(CP(1, 2, 3)) ≈ Z. Let us compute KX of X = (X, ∆). First, the divisor D0 is
0
a member of the linear system |O(6)|, so D ≡ 6D0 . Second, the canonical bundle
of CP(1, 2, 3) is O(−6), so from Proposition 4.4.15 we have
orb 1 10 1 60 1
KX ≡ −6D0 + D0 + 6D0 ≡ (−6 + + )D0 ≡ − D0 .
2 11 2 11 22
So X = (CP(1, 2, 3), 12 D0 + 10 0
11 D ) is a Fano orbifold.
1
We consider the orbibundle over (X, ∆) with Chern class c1 (Y /X) = 22 [D0 ].
5
To check the smoothness of Y or equivalently M we need to determine the orbit
invariants.
150 4. FUNDAMENTALS OF ORBIFOLDS

Since Cl(X) ≈ Z, we can write B ≡ kD0 for some integer k. Then using
Definition 4.7.4 we write
1 1 b
D0 ≡ kD0 + D0 + (6D0 )
22 2 11
for 0 < b < 11. We get k = −1 and b = 1. Now X = CP(1, 2, 3) has precisely
two singular points, x2 = [0, 1, 0], and x3 = [0, 0, 1] both of which lie in D0 but
not in D0 . So m(x2 , ∆) = m(x3 , ∆) = 2, and it follows from Equation 4.7.1 that
m(xi , ∆)·c1 (Y /X)) = −2D0 + 22 D0 = −D0 . But Rxi (D0 ) is a generator of Cl(X, xi )
for both i = 2, 3, and by Proposition 4.7.8 the total spaces M 5 and Y are smooth.
Using Theorem 4.7.14 we show that M 5 is a rational homology sphere with
H1 (M 5 , Z) = 0 and H2 (M 5 , Z) = Z11 ⊕ Z11 . Now m(∆) = lcmi (mi ) = 22, so
m(∆)c1 (Y /X) = [D0 ] implying that d = 1. Furthermore, from the genus formula
4.6.5 we see that g(D0 ) = 1, whereas, D0 has g(D0 ) = 0. So H2 (M 5 , Z) = Z11 ⊕Z11 .
CHAPTER 5

Kähler-Einstein Metrics

Recall that a Riemannian metric g on a manifold M is Einstein if Ricg = λg


for some constant λ. So there are three cases, positive, negative, and null, depend-
ing on whether λ is positive, negative, or zero, respectively. The reader is referred
to the ‘Bible’ of Einstein geometry, namely Besse [Bes87], and to [LW99] for
more recent developments. It is worth mentioning that although there has been
an enormous amount of progress in producing new examples of Einstein metrics,
most of the main problems in Section I of the Introduction of [Bes87] still re-
main. The one notable exception is the ten year old result of Catanese and LeBrun
[CL97] which shows that in dimension 4n with n ≥ 2 there are smooth compact
manifolds that admit both positive and negative Einstein metrics. However, an
important problem that still remains is that there is still no known topological
obstruction to the existence of Einstein metrics in dimensions greater than four
where one has the well-known Hitchin-Thorpe inequality and its generalizations
[LeB96, Kot98b, LeB99, LeB01]. Indeed there has been much progress in our
understanding of Einstein geometry on four manifolds. The reader is referred to
the excellent expository articles of LeBrun [LeB99, LeB03] for a discussion about
what was known about Einstein metrics in dimension four circa 2003. Of particular
interest is the relationship between the existence of Einstein metrics and differential
topology. This is developed further in Chapter 11 in the context of Sasaki-Einstein
geometry. However, in dimension greater than four very little is known about both
the existence and the moduli problem concerning Einstein metrics. At this stage
the best we can do is to produce examples. An important theme in Einstein ge-
ometry that is not touched on at all in this book is the notion of collapse and its
relation to the moduli space of Einstein metrics developed mainly by Anderson,
Cheeger, Colding, and Tian. See [Che03, CT06] and references therein.
The Einstein equations if Kähler geometry have immediate cohomological con-
sequences. We must have
λ[ωg ] = [ρg ] = 2πc1 (M )
so that the first Chern class of M cannot be indefinite. From the outset that
excludes most compact Kähler manifolds from considerations. As discussed in the
last section of Chapter 3 Yau’s proof of the Calabi Conjecture [Yau77, Yau78]
implies each compact complex manifold with c1 (M ) = 0 must have a Kähler Ricci-
flat metric. However, in the two other cases when c1 (M ) < 0 and c1 (M ) > 0 one
needs to consider a modified Monge-Ampère problem. The existence of a Kähler-
Einstein metric of negative scalar curvature on a compact manifold with c1 (M ) < 0
was proved contemporaneously and independently by and Aubin [Aub76, Aub78].
In the case of c1 (M ) > 0 there are known obstructions to solving the Monge-Ampère
problem. In the 50ties the only known obstructions had to do with the theorem of
151
152 5. KÄHLER-EINSTEIN METRICS

Matsushima which says that if (M, J, g, ωg ) is Kähler-Einstein then the Lie algebra
h(M ) of the holomorphic vector fields on M must be reductive [Mat57b]. This
led Calabi to conjecture that if h(M ) = 0 there should be no obstructions. If true
that would have indeed provided a very effective way of deciding on whether or not
a given Fano manifold admits a Kähler-Einstein metric. It took years of research
and truly deep insights of Tian to show that, on one hand, this conjecture is true
for surfaces but, on the other hand, it fails in higher dimensions.
So the problem is to give necessary and sufficient conditions for the class c1 (M )
to admit a Kähler-Einstein metric. For compact manifolds this was done in 1997
by Tian [Tia97] who gave necessary and sufficient analytic conditions in terms of
the properness of a certain functional. This is also related to a certain stability
criterion. Along the way Tian also exhibited the first example of a smooth Fano
3-fold with no holomorphic vector fields which admits no Kähler-Einstein metric, as
well as proving that del Pezzo surfaces admit positive Kähler-Einstein metrics if and
only if the Lie algebra of infinitesimal automorphisms is reductive [Tia90]. In spite
of all this, the problem of formulating necessary and sufficient algebraic conditions
under which a compact Fano manifold (orbifold) admits an Einstein metric still
remains open. It is precisely finding algebraic conditions that is of upmost interest
to us. We refer the reader to Tian’s book [Tia00] and review article [Tia99] as well
as his original articles for a complete treatment (see also Chapter 7 of [Aub98]).
More generally, today one is interested in the larger class of the so-called “ex-
tremal Kähler metrics” (cf. [Tia00, Sim04] and references therein) which general-
ize both Kähler-Einstein metrics and Kähler metrics of constant scalar curvature.
We will discuss extremal metrics briefly in Section 5.3. The subject of extremal
Kähler metrics and how they relate to various notions of geometric and algebraic
stability goes far beyond the scope of this book. Here we had to settle for a careful
selection of topics most relevant for the material of Chapter 11 with the focus on
the existence of Kähler-Einstein metrics on compact Fano manifolds and orbifolds.

5.1. Some Elementary Considerations


Here we concentrate on the case that g is both Kähler and Einstein in which
case we say that g is Kähler-Einstein. It follows from Proposition 3.6.1 that if
(M, J, g, ωg ) is Kähler-Einstein then the first Chern class must be proportional to
the cohomology class of the Kähler form, that is, c1 (M ) = λ[ωg ], so we have
Proposition 5.1.1: A necessary condition that a Kähler manifold admits a Kähler-
Einstein metric is that the first Chern class c1 (M ) is either definite or null.
One often refers to such metrics as the canonical Kähler metrics (cf. [Tia00]).
There is an analogous statement for Kähler orbifolds X = (X, U ), but here we must
use the orbifold first Chern class corb
1 as defined in Theorem 4.3.14 which in general
is different than c1 (X). Thus, we have
Proposition 5.1.2: A necessary condition that a Kähler orbifold X admits an
orbifold Kähler-Einstein metric is that the orbifold first Chern class corb
1 (X ) is
either definite or null.
There can exist Kähler-Einstein orbifold metrics on algebraic varieties X whose
first Chern class c1 (X) is indefinite. Of course, in this case corb
1 (X ) must be definite
or null.
5.2. THE MONGE-AMPÈRE PROBLEM AND THE CONTINUITY METHOD 153

We can consider the existence of Kähler-Einstein metrics with arbitrary Ein-


stein constant λ. By scaling we can assume that λ = 0, ±1. Specifically, let
(M, J, g, ωg ) be a compact Käher manifold. We would like to know if one can al-
ways find a Kähler-Einstein metric h ∈ K[ωg ] . Recall that on a Kähler-Einstein
manifold ρg = λωg . This implies that 2πc1 (M ) = λ[ωg ]. Now, if c1 (M ) > 0 we
must have λ = +1 because [ωg ] is the Kähler class, so that when M is compact
[ωg ] 6= 0. Similarly, when c1 (M ) < 0 the only allowable sign of a Kähler-Einstein
metric on M is λ = −1. Clearly, when c1 (M ) = 0 we must have λ = 0 as [ωg ] 6= 0.
As we have already pointed out the λ = 0 case follows from Yau’s solution to the
Calabi conjecture.
Remark 5.1.1: Note that when M is not compact one can have Kähler-Einstein
metrics on M of any sign. The simplest example here is M = Cn . First we can
take the flat metric which is Einstein with λ = 0. Then we can identify M with a
complex hyperbolic ball and put the Bergman metric on it with λ = −1. Finally,
we can take M = Cn ⊂ CP(n) with the Fubini-Study metric and λ = +1. The first
two metrics are complete, the third is not. One cannot have a complete metric of
positive Ricci curvature on a non-compact manifold.

5.2. The Monge-Ampère Problem and the Continuity Method


This section is a continuation of Sections 3.6 and 4.4.4. Let (X , g, ωg ) be
a Kähler orbifold with underlying complex space X and Kähler class [p∗ ωg ] ∈
2
Horb ¯
(X , Z). Let us reformulate the existence problem using the global i∂ ∂-lemma.
Suppose there exists an orbifold Einstein metric h ∈ K[ωg ] . Starting with the original
Kähler metric g on X we have a globally defined function f ∈ C ∞ (X , R)Γ that is
invariant under the local uniformizing groups such that
¯ .
ρg − λωg = i∂ ∂f
As before we will
R call f a discrepancy potential function. We also fix the constant
by asking that X (ef − 1)dvolg = 0. Let h ∈ K[ω] be an Einstein metric for which
¯
ρh = λωh . Using global i∂ ∂-lemma once again we have a globally defined invariant
¯ We shall fix the constant in ϕ
function ϕ ∈ C ∞ (X , R)Γ such that ωh − ωg = i∂ ∂ϕ.
later. Using these two equations we easily get
¯ − λϕ) .
ρg − ρh = i∂ ∂(f
Defining F so that ωhn = eF ωgn we can write this equation as
¯ = i∂ ∂(f
i∂ ∂F ¯ − λϕ) .

This implies that F = f − λϕ + c. We have already fixed the constant in f so


Rc depends
f −λϕ
only on the choice of ϕ. We can make c = 0 by choosing ϕ such that
X
(e − 1)dvolg = 0. Hence, using Theorem 4.4.31, Theorem 3.6.6 becomes
Proposition 5.2.1: Let (X , ω) be a compact Kähler orbifold with λcorb 1 > 0, where
λ = ±1, or with corb
1 = 0 and λ = 0. Let [ω] ∈ H 2
orb (X , Z) be a Kähler class and
g, h two Kähler metrics in K[ω] with Ricci forms ρg , ρh . Let f, ϕ ∈ C ∞ (X , R)Γ be
¯ ωh − λωg = i∂ ∂f.
defined by Rρg − λωg = i∂ ∂f, ¯ Fix the relative constant of f − λϕ
f −λϕ
by setting X (e − 1)dvolg = 0. Then the metric h is Einstein with Einstein
constant λ if and only if ϕ satisfies the following Monge-Ampère equation
ωhn = ef −λϕ ωgn ,
154 5. KÄHLER-EINSTEIN METRICS

which in a local uniformizing complex chart (U ; z1 , . . . , zn ) is written as


³ 2
´
det gij̄ + ∂z∂i ∂ϕz̄j
(5.2.1) = ef −λϕ .
det(gij̄ )
Note that by setting λ = 0 we get the Monge-Ampère equation for the original
Calabi problem described in Theorems 3.6.5 and 3.6.6. The character of the Monge-
Ampère equation above very much depends on the choice of λ. The case of λ < 0 is
actually the simplest, as the necessary a priori C 0 -estimates can be derived using
the Maximum Principle. This was done by Aubin [Aub76] and independently by
Yau [Yau77, Yau78]. As discussed previously the extension to the case of compact
orbifolds is straightforward, so we have
Theorem 5.2.2: Let (X , J, g, ωg ) be a compact Kähler orbifold with corb 1 < 0 or
orb
c1 = 0. Then there exists a unique orbifold Kähler-Einstein metric h with [ωh ]
such that ρh = λωh .
The case of λ > 0 is more difficult and it has been known for quite some time
that there are actually non-trivial obstructions to the existence of Kähler-Einstein
metrics which we discuss in the next section. For ease of discussion we first treat the
positive case for smooth manifolds. The first important step toward understanding
the c1 (M ) positive case was taken by Aubin [Aub84] who showed how to apply
the continuity method in this case. Rewrite equation (5.2.1) with a parameter
t ∈ [0, 1] :
³ ∂2ϕ ´
(5.2.2) det gij̄ + = det(gij̄ )ef −tϕ
∂zi ∂ z̄j
subject to the condition

∂2ϕ
(5.2.3) gi0j̄ = gij̄ + >0
∂zi ∂ z̄j
which defines a Kähler metric on M. Assuming that a solution ϕ(t) of Equation
(5.2.2) exists at t, we get a Kähler metric gt0 and its 2-form ωt0 . We set g00 = g, and
let ρ0t be the Ricci form of gt0 , then a calculation shows that
(5.2.4) ρ0t = (1 − t)ω + tωt0 .
Now we define the set
(5.2.5) S = {t ∈ [0, 1] | Equation (5.2.2) has a solution ϕ(t)} .
By Theorem 3.6.6 ϕ(0) is a solution, so S is non-empty. If S = [0, 1] then Equation
(5.2.4) gives us a Kähler-Einstein metric at t = 1. So to prove the existence of a
Kähler-Einstein metric it suffices to show that the set S is both open and closed in
[0, 1]. That S is open is an inverse function theorem argument [Aub84, Aub98].
The problem then is to find necessary and sufficient conditions that guarantee that
S is closed. Such necessary and sufficient analytic conditions were found by Tian
in 1997 [Tia97] in terms of the properness of a certain functional; however, such
conditions are often hard to implement. The applications we have in mind in this
book require algebraic conditions that are easy to give.
The first steps toward finding sufficient conditions for guaranteeing the exis-
tence of a Kähler-Einstein metric were taken by Tian [Tia87] and independently
5.2. THE MONGE-AMPÈRE PROBLEM AND THE CONTINUITY METHOD 155

by Siu [Siu88]. Let us define1


© R ª
P (M, g) = ϕ ∈ C 2 (M ) | positivity (5.2.3) holds, M
ϕ dvolg = 0 .
Then Tian [Tia87] shows that there exist positive constants α and C depending
only on M and g such that
Z
(5.2.6) e−αϕ dvolg ≤ C
M

for all functions ϕ ∈ P (M, g). This gives a uniform bound


© ª
(5.2.7) α(M ) = sup ∃ C > 0 such that (5.2.6) holds for all ϕ ∈ P (M, g) .
α>0

It is easy to see that α(M ) depends only on the cohomology class [ωg ], and is
invariant under biholomorphisms. Then we have [Tia87] (also [Siu88]):
Theorem 5.2.3: Let (M, J, g, ωg ) be a compact Kähler manifold of complex di-
n
mension n with c1 (M ) > 0. If ωg ∈ c1 (M ) and α(M ) > n+1 , then M admits a
Kähler-Einstein metric in the same Kähler class as ωg .
In certain cases Tian is able to obtain a lower bound for α(M ) as follows.
Proposition 5.2.4: Let (M, J, g, ωg ) be a compact Kähler manifold having N fami-
lies of curves {Cα1 }, . . . , {CαN }, where α ∈ CPn is the parameter, and N subvarieties
S1 , . . . , SN such that
(i) S1 ∩ · · · ∩ SN¡ = ∅, ¢
(ii) M \ Sj = ∪α Cαj ∩ (M \ Sj ) , Cαj ∩ Cβj ∩ (M \ Sj ) = ∅, and Cαj ∩ (M \ Sj )
is smooth for each α,
(iii) for all z ∈ M \ ∪i Si , {Tz Cαj i | Cαj j ∈ z} spans Tz M, and for all z ∈ Si ,
either {Tz Cαj j | z ∈ Cαj j ∩ (M \ Sj )} spans Tz M , or there exists Cαj j such
that z ∈ Cαj j ∩ (M \ Sj ) with Cαj j ∩ Si = {finite number of points},
(iv) for all i = 1, . . . , N and all α ∈ CPn , 4volg (Cαi ) ≤ β.

Then α(M ) ≥ β .
It is easy to generalize these results to orbifolds, and we give orbifold examples
later where this inequality holds. In order to apply Proposition 5.2.4 successfully
to prove the existence of Kähler-Einstein metrics one needs to invoke some symme-
tries. So one considers a subgroup G of the group of automorphisms of the Kähler
structure, and defines the subset PG (M, g) = {ϕ ∈ P (M, g) | ϕ is G invariant}.
Then we define
© ª
(5.2.8) αG (M ) = sup ∃ C > 0 such that (5.2.6) holds for all ϕ ∈ PG (M, g) .
α>0

Clearly, αG (M ) ≤ α(M ). Then Theorem 5.2.3 has a G-invariant version, namely


Theorem 5.2.5: Let (M, J, g, ωg ) be a compact Kähler manifold of complex di-
mension n with c1 (M ) > 0, and let G be a compact subgroup of the group of Kähler
n
automorphisms. If ωg ∈ c1 (M ) and αG (M ) > n+1 , then M admits a Kähler-
Einstein metric in the same Kähler class as ωg .
It is this G-invariant version that is most effective for proving the existence of
Kähler-Einstein metrics on smooth manifolds.

1Our normalization R ϕ dvol = 0 which follows [Aub98] is different from Tian’s.


M g
156 5. KÄHLER-EINSTEIN METRICS

Example 5.2.6: [Fermat hypersurfaces] Both Tian and Siu showed how to prove
the existence of Kähler-Einstein metrics on Fermat hypersurfaces. Here we follow
Tian [Tia87]. Consider the Fermat hypersurface Fn,p defined by
z0p + · · · + znp = 0
in CPn for n ≥ 3. First, by adjunction theory Fn,p is Fano if and only if n+1−p > 0.
The hypersurface Fn,p is invariant under the group G generated by the sym-
metric group Σn+1 acting as permutations of the n + 1 homogeneous coordinates
z0 , . . . , zn and n + 1 copies of the cyclic group Cp ≈ Zp acting by zi 7→ ζi zi , where
ζi ∈ Cp . Now in the presence of symmetries we can obtain an improved lower bound.
We need to replace item (iv) in Proposition 5.2.4 by
(iv)’ Let Gj ⊂ G be the subgroup that preserves the fibration of M \ Sj by the
curves Cαj ∩ (M \ Sj ), then Sj is invariant under Gj and the inequality
4volg (Cαi )
≤β,
ord(Gj )
|Gj |
where ord(Gj ) = minz∈(M \Sj ) |Stabz ⊂Gj | , where Stabz is the isotropy
subgroup at z in Gj .

Then αG (Fn,p ) ≥ β . This makes β smaller giving a better bound for αG (Fn,p ).
Tian (see [Tia87] for details) then obtains the estimate β ≤ 4π(n+1−p)
2 . So using
Theorem 5.2.5, we get the existence of Kähler-Einstein metrics if
n−1 4π 2
> αG (M ) ≥ ≥
n β (n + 1 − p)
holds. That is, we get Kähler-Einstein metrics for p = n − 1, n. Nadel [Nad90]
has improved this result by proving the existence of Kähler-Einstein metrics on
Fn,p for n2 ≤ p ≤ n using multiplier ideal sheaves. We should also mention that the
complex quadric Fn,2 is an Hermitian symmetric space (the oriented Grassmannian
G+2 (R
n+1
)) that is known to admit a Kähler-Einstein metric [KN69].
Let us briefly describe what happens if we fail to obtain the uniform bound
(5.2.6). Then there exists an increasing sequence {tk } with tk < 1 such that
Z
(5.2.9) e−αϕk dvolg −−−→∞ as k−−→∞

n
for all α ∈ ( n+1 , 1). We now introduce the concept of a multiplier ideal sheaf
[Nad89, Nad90].
Definition 5.2.7: Let I(αϕ) denote the sheaf associated to the presheaf defined by
associating to each open set U ⊂ M the set of sections
n Z o
f ∈ OM (U ) | |f |2 e−αϕ dvolg < ∞ .
U

We can now state Nadel’s Theorem [Nad90]:


Theorem 5.2.8: Let (M, J, g, ωg ) be a compact Kähler manifold of complex dimen-
sion n with c1 (M ) > 0, and let G be a compact group of automorphisms. If M does
n
not admit a G-invariant multiplier ideal sheaf I(αϕ) 6= OM for all α ∈ ( n+1 , 1),
then M admits a Kähler-Einstein metric.
5.2. THE MONGE-AMPÈRE PROBLEM AND THE CONTINUITY METHOD 157

The point is that if one can find a sequence satisfying Equation 5.2.9 then there
is a non-trivial multiplier ideal sheaf I(αϕ) 6= OM , otherwise I(αϕ) = OM , and the
set S of Equation 5.2.5 becomes the whole interval [0, 1] giving a Kähler-Einstein
metric at t = 1. We refer to the literature [Aub98, DK01, Nad90] for further
discussion and proofs.
Once one has some familiarity working with orbifolds, the generalization of
these ideas to orbifolds is straightforward. Here computations are done on the
local uniformizing covers (Ũ , Γ) working with Γ-invariant objects. So one works
with orbisheaves as discussed in Chapter 4. Let OX denote the structure orbisheaf
of Definition 4.2.2. One now writes down Equations (5.2.2)-(5.2.5) on the local
uniformizing neighborhoods with all objects being Γ-invariant. Then a multiplier
ideal orbisheaf is:
Definition 5.2.9: Let I(αϕ) denote the orbisheaf associated to the presheaf defined
by associating to each local uniformizing open set (Ũ , Γ) the set of sections
n Z o
f ∈ OX (Ũ )Γ | |f |2 e−αϕ dvolg < ∞ .

The orbifold version of Theorem 5.2.8 was given by Demailly and Kollár [DK01].
Theorem 5.2.10: Let X be a compact Kähler orbifold of complex dimension n
with corb
1 (X ) > 0, and let G be a compact group of automorphisms. If X does not
n
admit a G-invariant multiplier ideal orbisheaf I(αϕ) 6= OX for all α ∈ ( n+1 , 1) ,
then X admits a Kähler-Einstein orbifold metric.
Remark 5.2.1: Recall from Equation 4.4.2 that corb 1 (X ) and c1 (X) are different in
the presence of branch divisors. So positive Kähler-Einstein orbifold metrics can
exist on algebraic varieties that are not Fano. They are, however, orbifold Fano.
Multiplier ideal sheaves can also be described in terms of log resolutions of
singularities as we now explain [Dem96, Dem01, Kol97, KM98, Laz04b]. The
idea is that the uniform boundedness of the integrals (5.2.9) can be expressed in
terms of the mildness of the singularities of a pair (X, D), where X is a normal
variety and D is a Q-divisor on X. However, in order to define log resolutions, we
need to make sense out of pulling back divisors by a map. So we assume that X
is Q-factorial. Thus, Weil Q-divisors are Q-Cartier, so their pullback by a map is
well defined. This causes no problem since we are interested in algebraic varieties
X that are the underlying complex space of a complex orbifold and Propositions
1.1.5 and 4.4.7 guarantee that such an X is both normal and Q-factorial.
P
Definition 5.2.11: A divisor D = i Di is said to have simple normal cross-
ings if each Di is smooth and D is defined in a local neighborhood P chart (U ; z)
by equations of the form z1 · · · zk = 0 forPsome k ≤ n. A Q-divisor i ai Di has
simple normal crossings if the divisor i Di has simple normal crossings.
Definition 5.2.12: Let D be a Q-divisor on X. A log resolution of the pair
(X, D) is a projective
P birational map µ : X 0 −−→X with X 0 smooth such that the

divisor µ D + i Ei has simple normal crossings. Here Ei are the exceptional
divisors.
We first describe the multiplier ideal sheaves associated with the underlying
variety. Given a log resolution µ : X 0 −−→X of (X, D) we can define a multiplier
ideal sheaf I(D) by
¡ ¢
(5.2.10) I(D) = µ∗ OX 0 KX 0 − µ∗ bKX + Dc ⊆ OX ,
158 5. KÄHLER-EINSTEIN METRICS

where bDc denotes the integral part of D. It follows from a theorem of Esnault and
Viehweg (cf. [Laz04b], pg 156) that I(D) is independent of the log resolution.
In the case of an orbifold X = (X, U ) we replace the variety X by a local
uniformizing neighborhood φ : Ũ −−→U, and if the orbifold has a non-trivial branch
divisor ∆ then we must replace the canonical divisor φ∗ KU with the orbifold canon-
ical divisor KŨorb = φ∗ (KU + ∆). This gives rise to a multiplier ideal orbisheaf I(D)
defined as the sequence of sheaves associated to the presheaves which associates to
each local uniformizing neighborhood Ũ the sheaf
¡ ¢
(5.2.11) I(D)(Ũ ) = µ∗ OX 0 KX 0 − µ∗ (KŨorb + φ∗ D) ⊆ OŨ .
We now make the connection between the orbisheaf I(ϕ) P of Definition 5.2.9
and the orbisheaf defined by Equation (5.2.11). Let D = i ai Di be an effective
Γ-invariant Q-divisor on a local uniformizing neighborhood Ũ , and assume that
Di is principal, that is, determined
P by a Γ-invariant holomorphic defining function
gi . Then the function ϕD = i ai log |gi | is plurisubharmonic and locally defines a
multiplier ideal sheaf
n Z o
|f |2
I(ϕD )Ũ = f ∈ OX (Ũ )Γ | Q 2a
dvolg < ∞
Ũ |gi | i
on Ũ . Now the gi can be taken as the coordinate functions of a local coordinate
chart. So the condition that
Q the integral be locally finite is that the holomorphic
function f be divisible by |gi |mi , where mi is an integer satisfying mi − ai > −1
for each i. Note that since ai ∈ Q this means that mi ≥ bai c. By Theorem 9.3.42
of [Laz04b] the multiplier ideal sheaf I(ϕD )Ũ is equal to the analytic extension
I(D)an (Ũ ) of the algebraic multiplier ideal sheaf I(D)(Ũ ) on each uniformizing
neighborhood Ũ . Thus, the corresponding multiplier ideal orbisheaves I(ϕD ) and
I(D)an coincide. Hereafter, we shall use these two orbisheaves interchangeably and
denote the orbisheaf by I(X , D).
Definition 5.2.13: Let D be an effective Q-divisor on an orbifold X = (X, U).
Then
(i) The pair (X, D) is Kawamata log terminal ¡ or klt¢ if I(X , D) = OX .
(ii) The pair (X, D) is log canonical if I X , (1 − ²)D = OX for all 0 <
² < 1.
An alternative formulation of Definition 5.2.13 can be obtained by noticing that
the condition I(X , D)(Ũ ) = OŨ is equivalent to the condition
¡ ¢
(5.2.12) ordE KX 0 − µ∗ φ∗ bKU + ∆ + Dc > −1

for every exceptional divisor in the log resolution µ : X 0 −−→Ũ . This gives rise to the
definition used in [BGN03b] (and also in [KM98]), namely
Definition 5.2.14: The pair (X, D) is klt or Kawamata log-terminal) if for
each local uniformizing neighborhood Ũ there exists a log resolution of singularities
µ : X 0 → Ũ such that
X
KX 0 ≡n µ∗ (KŨorb + φ∗ D) + ai Ei

with ai > −1 for all µ-exceptional Ei . If ai ≥ −1 then the pair (X, ∆) is said to
have log-canonical singularities.
5.2. THE MONGE-AMPÈRE PROBLEM AND THE CONTINUITY METHOD 159

Clearly, a klt pair (X, D) is log-canonical. It is now a simple matter to refor-


mulate Theorems 5.2.8 and 5.2.10 in terms of the klt condition. For example, we
have
Theorem 5.2.15: Let (X , J, g, ω) be a compact Kähler orbifold of complex di-
mension n with corb 1 (X ) > 0, and let G be a compact group of automorphisms
n+²
on (X , J, g, ω). Assume that there is an ² > 0 such that (X, n+1 D) is klt for
orb
every G-invariant effective Q-divisor D ≡n −KX . Then X has an orbifold Kähler-
Einstein metric.
This theorem has given some rather startling results which we describe in Chap-
ter 11 [JK01b, BGN03b, BGK05] even for the case G = {id}. Let us rephrase
this more explicitly in terms of holomorphic sections for the case G = {id} [DK01].
Theorem 5.2.16: Let (X , J, g, ω) be a compact Kähler orbifold of complex dimen-
−1
sional n such that the orbifold anticanonical divisor KX is ample. The continuity
n
method produces a Kähler-Einstein orbifold metric h on X if there is a γ > n+1
0 orb −s
such that for every s ≥ 1 and for every holomorphic section τs ∈ H (X , O(KX ))
γ
we have |τs |− s ∈ L2 (X), that is
Z

|τs |− s dvolg < +∞ .

When applying this theorem we will need to check the L2 condition along
certain divisors Hi . This is accomplished by reducing the problem to an analogous
problem on Hi and using induction. In algebraic geometry, this method is called
inversion of adjunction. Conjectured by Shokurov, the following version is due
to Kollár [Kol92, 17.6]. It was observed by [Man93] that it can also be derived
from the L2 -extension theorem of Ohsawa and Takegoshi [OT87]. See [Kol97] or
[KM98] for more detailed expositions.
Theorem 5.2.17 (Inversion of adjunction): Let M be a smooth manifold, H ⊂ M
a smooth divisor with equation (h = 0) and g a holomorphic function on M . Let
gH denote the restriction of g to H and assume that it is not identically zero. The
following are equivalent:
(i) |g|−c |h|δ−1 is locally L2 near H for every δ > 0.
(ii) |gH |−c is locally L2 on H.
We end this section by briefly describing the situation for complex surfaces. In
this case Tian [Tia90] has given a classification of all smooth compact complex
surfaces that admit a positive Kähler-Einstein metric. Recall from Example 3.5.11
2
that the del Pezzo surfaces are precisely CP2 , CP1 × CP1 , or CP2 #kCP , where
1 ≤ k ≤ 8. Now there are well-known obstructions to the existence of positive
Kähler-Einstein metrics as discussed in Section 5.3 below. In particular, Example
2
5.3.2 says that CP2 #kCP with k = 1 or 2 cannot admit a Kähler-Einstein metric.
However, Tian’s theorem says that these are the only smooth compact complex
del Pezzo surfaces which do not admit a Kähler-Einstein metric. Specifically, Tian
[Tia90] proved
Theorem 5.2.18: A smooth compact complex surface M with c1 (M ) positive ad-
mits a Kähler-Einstein metric if and only if the Lie algebra of the automorphism
group of the complex structure is reductive.
160 5. KÄHLER-EINSTEIN METRICS

Somewhat earlier Tian and Yau [TY87] had shown that for each smooth man-
2
ifold CP2 #kCP with 3 ≤ k ≤ 8 there is a complex structure admitting a Kähler-
Einstein metric, and independently Siu [Siu88] proved the existence of a Kähler-
2
Einstein metric on CP2 #3CP . We mention that for 5 ≤ k ≤ 8 there are non-trivial
moduli of Kähler-Einstein metrics. See Exercise 5.3 below.

5.3. Obstructions in the Positive Case


The first types of obstructions that were found involved the Lie algebra of
holomorphic vector fields. Let h(M ) be the complex Lie algebra of all holo-
morphic vector fields on M. Matsushima [Mat57a] proved that on a compact
Kähler-Einstein manifold with c1 (M ) > 0, h(M ) must be reductive, i.e., h(M ) =
Z(h(M )) ⊕ [h(M ), h(M )]. This result was generalized by Lichnerowicz [Lic57] and
then later by Calabi [Cal85]. For convenience here we give the Lichnerowicz ver-
sion. The more general result of Calabi will be generalized further to certain trans-
verse Kähler structures in Theorem 11.3.1 below. A detailed proof of Calabi’s
theorem appears in [Sim04].
Theorem 5.3.1: Let (M, J, g, ωg ) be a compact Kähler manifold of constant scalar
curvature. Then h(M ) = a(M ) ⊕ h0 (M ), where a(M ) is the subalgebra of h(M )
consisting of all parallel vector fields and h0 (M ) is the complexification of the Lie
algebra of Killing vector fields. In particular, if (M, J) admits a Kähler-Einstein
metric, then h(M ) is reductive.
For further discussion and proof we refer the reader to [Kob72] and [Bes87].
Example 5.3.2: For an easy application of this theorem consider Mk = CPn blown
up at k points in general position, where 1 ≤ k ≤ n. It is easy to see that the
Lie algebra h(Mk ) is not reductive; hence, Mk cannot admit a Kähler-Einstein
metric. For details in the n = 2 case see, for example, [Fut88] or [Tia00]. The
generalization to n dimensions is straightforward.
Theorem 5.3.1 was generalized by Calabi [Cal85] to the case of extremal met-
rics. Extremal Kähler metrics are critical points of the ‘energy’ functional
Z
E(g) = sg2 dvolg ,
M

where sg is the scalar curvature of the Kähler metric g. Solving the Euler-Lagrange
equations for this functional shows that a Kähler metric g is extremal if and only
if the vector field gradg sg is holomorphic. Thus, Kähler metrics of constant scalar
curvature are extremal. In particular, Kähler-Einstein metrics are extremal. How-
ever, in the more general case of non constant scalar curvature extremal metrics,
the Lie algebra h0 (M ) is more complicated. See, for example, [Fut88] or [Sim04]
for details and proofs.
The next type of obstruction again involves the Lie algebra h(M ) and was
found by Futaki [Fut83]. We follow the presentation in [Sim04]. Given a Kähler
metric on M or equivalently 2-form ω, its Ricci form ρ represents the first Chern
class c1 (M ), up to a factor of 2π. But by Hodge theory ρ has a unique harmonic
representative ρH , and by Lemma 3.1.15 we can write

(5.3.1) ¯ ω
ρ = ρH + i∂ ∂ψ
5.3. OBSTRUCTIONS IN THE POSITIVE CASE 161

for some real function ψω , called the Ricci potential. Then the Futaki invariant or
character is a map F : h × K(M )−−→C defined by
Z
(5.3.2) F(X, [ω]) = X(ψω )dvolω .
M

One can check that F only depends on the cohomology class [ω], and not on the
specific Kähler metric or form ω. Furthermore, it is easy to see that for any two
holomorphic vector fields X, Y, we have F([X, Y ], [ω]) = 0. The main result on
Futaki invariant is
Theorem 5.3.3: Let (M, J, g, ωg ) be a compact Kähler manifold, and suppose that
the g is an extremal metric. Then g has constant scalar curvature if and only if
F(X, [ω]) = 0 for all holomorphic vector fields X.
We are particularly interested in the case that c1 (M ) is positive and the coho-
mology class [ω] the Kähler metric is proportional to c1 (M ). Then after an appro-
priate scaling, Equation (5.3.1) becomes

(5.3.3) ¯ ω.
ρ = ω + i∂ ∂ψ

In this case an extremal Kähler metric with F(X, [ω]) = 0 is a Kähler-Einstein


metric.
Example 5.3.4: [Fut88] Consider CPm1 × CPm2 and let Hi denote the hyperplane
bundle over CPmi . Let πi : CPm1 × CPm2 −−→CPmi be the natural projection, and
let Mm1 ,m2 be the total space of the projectivization CP(π1∗ H1 ⊕ π2∗ H2 ). Then
c1 (Mm1 ,m2 ) is positive, and the connected component of the automorphism group
of Mm1 ,m2 is the projectivized group P(GL(m1 , C)×GL(m2 , C)). So the Lie algebra
of holomorphic vector fields is reductive. Let X be an infinitesimal generator of
the kernel of the projection map P(GL(m1 , C) × GL(m2 , C))−−→P(GL(m1 , C)) ×
P(GL(m2 , C)). Then generally F(X, [ω]) 6= 0. For example Futaki [Fut88] shows
that for M1,2 , F(X, [ω]) = −90. Thus, M1,2 does not admit a Kähler-Einstein
metric.
In particular, the non-vanishing of F is an obstruction to the existence of Kähler-
Einstein metrics. Just as Matsushima theorem the Futaki invariant provides ob-
structions only in the presence of holomorphic vector fields. Another conjecture
of Calabi (formulated long before Futaki’s papers) speculated that Kähler-Einstein
metrics should exist on a Kähler manifold with c1 (M ) > 0 and no holomorphic
vector fields. This was shown, however, to be false, first in the category of orbifolds
by Ding and Tian [DT92], and finally Tian [Tia97] exhibited a smooth Fano 3-fold
with no holomorphic vector fields and no Kähler-Einstein metric (see also [Tia00]).
In order to understand the existence and obstructions in the absence of any holo-
morphic vector fields Tian constructs a generalized Futaki character [Tia97]. He
then formulates an alternative to the ‘false’ Calabi conjecture: vanishing of his
generalized obstruction should now be necessary and sufficient. All that quite nat-
urally leads to the study of the various notions of geometric/algebraic stability.
Simplifying a bit, in the case of the canonical Kähler metrics, in essence one deals
with the following fundamental questions:
Open Problem 5.3.1: Let (M, J, g, ωg ) be a compact Kähler manifold (orbifold)
with positive (orbifold) first Chern class. What are the necessary and sufficient
conditions for M to admit an (orbifold) Kähler-Einstein metric? In case M is a
162 5. KÄHLER-EINSTEIN METRICS

complex algebraic variety, can such conditions be understood/formulated exclu-


sively in terms of the algebraic-geometric properties of M ? Finally, given such M
can one effectively determine if M can admit a Kähler-Einstein metric or not?
Obstructions to the existence of Kähler-Einstein orbifold metrics on log Fano
orbifolds will be discussed later within the context of Sasakian geometry in Sections
11.3.2 and 11.3.3.

5.4. Kähler-Einstein Metrics on Hypersurfaces in CP(w)


The algebraic geometry of hypersurfaces and complete intersections in weighted
projective spaces CP(w) was discussed in Section 4.6 where they were first intro-
duced. In particular, they are all examples of compact Kähler orbifolds (manifolds).
In this section we shall examine some applications of the general existence results
discussed in Section 5.2 when applied to this specific case.

5.4.1. Smooth Hypersurfaces in CP(w). Let us begin by considering ex-


amples of smooth embedded hypersurfaces in CP(w). For convenience we mainly
treat the case dimension two, that is, complex surfaces. First we recall the well-
known work of Freedman (cf. [GS99]) which says that simply connected smooth
compact 4-manifolds M are determined up to homeomorphism by their intersection
form Q : H2 (M, Z) × H2 (M, Z)−−→Z. In fact, Freedman proved that for every uni-
modular symmetric bilinear form Q there exists a compact simply connected topo-
logical 4-manifold whose intersection form is Q. Unfortunately, in the case of even
intersection form it is not known exactly which correspond to smooth 4-manifolds.
There is an outstanding conjecture, known as the 11 8 -conjecture, that if true, would
solve the problem. The case of odd intersection form is relatively easy. All sim-
ply connected smooth 4-manifolds with odd intersection form are homeomorphic
to b+ 2 − 2 ±
2 CP #b2 CP , where b2 is the dimension of the maximal positive (negative)
subspace of Q on H2 (M, Z). Since the pioneering work of Donaldson and Freedman
in the early eighties, the study of the differential topology of 4-manifolds has been
a major theme in modern mathematics. We refer the reader to the monographs
[DK90, FM94, GS99, Sco05] for a detailed treatment of this very interesting
subject.
Example 5.4.1: We continue the discussion of the Fermat hypersurfaces Fn,p ⊂
CPn+1 of Example 5.2.6. As mentioned there Fn,p are Fano if n + 1 > p and are
known to admit a Kähler-Einstein metric if in addition p ≥ n2 . Furthermore, when
p ≥ n + 1 the hypersurfaces Fn,p also admit a Kähler-Einstein metric by the Aubin-
Yau Theorem 5.2.2. For simplicity we consider now the case of complex surfaces,
i.e., n = 3. Here there are two del Pezzo surfaces, p = 2 which is a complex quadric
biholomorphic to CP1 ×CP1 and it is well-known that this admits a Kähler-Einstein
metric. Of course, this isµdiffeomorphic
¶ to S 2 × S 2 and its intersection form Q is
0 1
given by the matrix H = . The case p = 3 is known to be CP2 blown up at
1 0
6 points, that is, F3,3 = CP2 #6CP2 which is also known to admit Kähler-Einstein
metrics by Tian and Yau [TY87]. Its intersection form Q = h1i ⊕ 6h−1i. The case
p = 4 gives the well-known K3 surface which admits a Ricci flat metric by Yau’s
result 5.2.2. Its intersection form is 2(−E8 ) ⊕ 3H, where E8 is the Cartan matrix
for the simple Lie algebra E8 . Generally for the Fermat surface F3,p the intersection
5.4. KÄHLER-EINSTEIN METRICS ON HYPERSURFACES IN CP(w) 163

form Q is lp h1i ⊕ mp h−1i for p odd and lp H ⊕ np (−E8 ) for p even, where
1 3 1 1
lp = (p − 6p2 + 11p − 3), mp = (p − 1)(2p2 − 4p + 3), np = p(p2 − 4) .
3 3 24
So, for example, the Fermat surface F3,5 is homeomorphic to 9CP2 #44CP2 , whereas,
F3,6 is homeomorphic to 4K3#9(S 2 × S 2 ). For p ≥ 5 the Fermat surfaces F3,p are
simply connected complex surfaces of general type all of which admit negative
Kähler-Einstein metrics. The general method will be discussed in more detail in
Example 5.4.2 below. See also [DK90, GS99].
Next we consider some complex algebraic manifolds that are hypersurfaces in
the more general weighted projective spaces CP(w). According to Proposition 9.3.22
any Brieskorn-Pham polynomial with pairwise relatively prime weights describes a
smooth projective algebraic variety in CP(w). Again for simplicity we consider the
surface case. The Table below gives the examples of all del Pezzo surfaces Xw that
can be written as smooth hypersurfaces in weighted projective spaces. All of these
are known to admit positive Kähler-Einstein metrics [TY87]. We only indicate a
particular polynomial, as the most general such hypersurface involves many such
monomials which for reasons of space we do not include. We shall discuss this
further in Section 5.5.

Table 1: del Pezzo surfaces as smooth hypersurfaces in CP3 (w)


|w| − d w fw d Xw
1 (1,1,1,1) z03 + z13 + z23 + z33 3 CP2 #6CP2

1 (1,1,1,2) z04 + z14 + z24 + z32 4 CP2 #7CP2

1 (1,1,2,3) z06 + z16 + z23 + z32 6 CP2 #8CP2

2 (1,1,1,1) z02 + z12 + z22 + z32 2 CP1 × CP1

3 (1,1,1,d) z0d + z1d + z2d + z3 d CP2

There are only two K3 surfaces that can be written as smooth hypersurfaces
in weighted projective spaces. Of course, these admit Ricci flat Kähler-Einstein
metrics. They are given in Table 2.

Table 2: K3 surfaces as smooth hypersurfaces in CP3 (w)


|w| − d w fw d Xw
0 (1,1,1,1) z04 + z14 + z24 + z34 4 K3

0 (1,1,1,3) z06 + z16 + z26 + z32 6 K3

For all other Brieskorn-Pham polynomials with relatively prime weights, the
index |w|−d is negative. There are infinitely many of these and they are all surfaces
of general type and admit negative Kähler-Einstein metrics. It is easy to see that
for any such surface the geometric genus is at least 2. Suffice it to present a few
examples.
164 5. KÄHLER-EINSTEIN METRICS

Example 5.4.2: These examples are complex surfaces Xw of general type with
c1 (X) negative, and unimodular indefinite intersection form Q that are smoothly
embedded in complex weighted projective space CP(w). By Proposition 9.3.22 these
can be obtained from Brieskorn polynomials with relatively prime weights.
By Freedman smooth simply connected compact 4-manifolds with indefinite
intersection form Q are determined up to homeomorphism by the parity, rank and
signature of Q. We have the following two cases according to the parity of Q.
(i) If Q is odd, Xw is homeomorphic to b+ 2 −
2 CP #b2 CP .
2

(ii) If Q is even, Xw is homeomorphic to mK3#n(S × S 2 ), where m and n


2

are given by
τ b2 − |τ | 3τ
m=− , n= + .
16 2 16
We mention here the well-known theorem of Rohlin (cf. [GS99]) which states
that for a smooth 4-manifold with Q even the signature is divisible by 16. To
determine the homeomorphism type of Xw it suffices to compute the numbers
b+ ± +
2 = b2 (Xw ), or alternatively b2 (Xw ) and b2 (Xw ). The former is easily computed
using the techniques discussed in Chapter 9. While for the latter we use the well-
known formula for complex surfaces with even second Betti number, b+ 2 = 1 + 2pg ,
and pg = h0,2 = h2,0 . So by Proposition 4.6.12 pg equals the number of monomials
of degree d − |w|. Then b± 2 and τ can easily be determined. We consider the case
of odd and even intersection form separately.
We give two tables of infinite sequences of smooth hypersurfaces in the weighted
projective space CP(w). Table 3 lists two infinite sequences with Q odd; whereas,
Table 4 gives one infinite sequence for the case Q even. Here both k and l are
integers greater than 1. Other sequences can be worked out in a straightforward
way.

Table 3: Hypersurfaces in CP3 (w) homeomorphic to b+ 2


2 CP #b2 CP
− 2

b−
2 w fw d b+
2

14k2 − 12k + 4 (1, 1, 1, k) z03k + z13k + z23k + z33 3k 4k2 − 6k + 3

12l2 − 6l + 1 (1,1,1,2l) z04l + z14l + z24l + z32 4l 4l2 − 6l + 3

The first two members of each of these sequences are 7CP2 #36CP2 , 21CP2 #94CP2 ,
and 7CP2 #37CP2 , 21CP2 #91CP2 . Let us analyze the first member of each sequence
a bit further illustrating an important phenomenon discovered by LeBrun [LeB96,
LeB01] and Kotschick [Kot98a].
Example 5.4.3: This example is taken from [LeB03]. The first member of each
sequence are the surfaces
X0 = {z06 + z16 + z26 + z33 = 0} , X2 = {z08 + z18 + z28 + z32 = 0} ,
and as mentioned they are complex surfaces of general type that are homeomorphic
to 7CP2 #36CP2 and 7CP2 #37CP2 , respectively. Now blow-up the first one X0 at
one point to obtain another complex surface X1 homeomorphic to 7CP2 #37CP2 .
Using Seiberg-Witten theory, LeBrun [LeB01] gives the estimate
2
(2χ + 3τ )(X1 ) > c21 (X0 )
3
5.4. KÄHLER-EINSTEIN METRICS ON HYPERSURFACES IN CP(w) 165

for X1 = X0 #CP2 to admit any Einstein metric whatsoever. Now c21 (X0 ) = 3
so (2χ + 3τ )(X1 ) = c21 (X1 ) = 2. Thus, X1 and X2 are both homeomorphic to
7CP2 #37CP2 with X2 admitting a Kähler-Einstein metric, but X1 admitting no
Einstein metric at all.
Exercise 5.1: Perform an analysis similar to Example 5.4.3 for the general se-
quences in Table 3.
A series giving surfaces of general type with even intersection form is:

Table 4: Hypersurfaces in CP3 (w) homeomorphic to mK3#n(S 2 × S 2 )


m w fw d n
l(l−1) 5l2 −9l+2
2 (1, 1, 1, 2l + 1) z04l+2 + z14l+2 + z24l+2 + z32 4l+2 2

The first two members of this sequence are K3#2(S 2 × S 2 ) and 3K3#10(S 2 × S 2 ).
As indicated in the previous examples, Kähler-Einstein metrics abound in the
negative case. So we concentrate on the positive and null cases. Before passing on
to the case of Fano hypersurfaces which is our main interest, we briefly discuss the
Calabi-Yau case.

5.4.2. Quasi-smooth Calabi-Yau Hypersurfaces in CP(w). Recall that


an orbifold X = (X, U) is Calabi-Yau if corb
1 (X ) = 0, and that this is equivalent to
the triviality of the ordinary canonical class when H1orb (X , Z) = 0 by Proposition
4.7.11. Combining this with Propositions 4.7.12, 4.6.13, and Theorem 4.4.31 gives
Theorem 5.4.4: Let X = (X, U) be a quasi-smooth weighted hypersurface in
CP(w) satisfying |w| = d. Then w is well-formed, and X admits a unique Ricci
flat orbifold metric.
When dimC X = 2 one can say more. Indeed, Kollár [Kol05] shows that
the minimal resolution of X is a smooth K3 surface as long as H1orb (X , Z) = 0.
Furthermore, if b2 (X) = 22 then X = K3. Thus using Theorem 4.7.14 we can
sharpen the results for the surface case (See Theorem 10.3.8 below for details).
Theorem 5.4.5: Let X = (X, U) be an orbifold K3 surface with H1orb (X , Z) = 0.
Then H2 (X, Z) = Zs , where 3 ≤ s ≤ 22 and s = 22 if and only if X = K3.
In 1979 Miles Reid [Rei80] produced a list of 95 weighted K3 surfaces as well-
formed hypersurfaces in the weighted projective space CP(w0 , w1 , w2 , w3 ). This list
is reproduced in [IF00] and includes the two cases in Table 2 above. In Appendix
B.1 we give Reid’s list together with the second Betti number of each surface X
which can be computed using the Milnor-Orlik procedure described below in Chap-
ter 9, Subsection 9.3.1. Explicitly, we have a computer program which computes
the second Betti number of the link (Seifert bundle over X) by a means of the
formula of Corollary 9.3.13 below. Interestingly, Reid’s list produces all possible b2
except b2 = 3 and 18.
Theorem 5.4.4 holds equally well for quasi-smooth
P weighted complete intersec-
tions as long as one replaces the degree d by |d| = i di . A list of 84 weighted K3
surfaces as well-formed weighted complete intersections was also given in [IF00].
In higher dimensions the physicists have generated many examples. In partic-
ular, Candelas et.al. [CLS90] produced a list of over 6000 Calabi-Yau orbifolds in
166 5. KÄHLER-EINSTEIN METRICS

CP(w0 , w1 , w2 , w3 , w4 ). This was one of the first papers that illustrated the inter-
esting new phenomenon known as mirror symmetry which has since taken on a life
of its own with important consequences for enumerative geometry. We refer the
reader to the recent books [CK99, HKK+ 03] and references therein.

5.4.3. Quasi-smooth Fano Hypersurfaces in CP(w). Suppose X ⊂ CP(w)


is a quasi-smooth irreducible hypersurface in weighted projective space with its in-
duced orbifold structure X . We also assume that X is Fano, that is the orbifold
−1
anticanonical bundle KX is ample. In the case that X is given as the zero set of
a weighted homogeneous polynomial, the Fano condition is IX = |w| − d > 0, by
Corollary 4.6.14. In terms of divisors we want to describe under what conditions
the log pair (Xd , D) is klt whenever D ≡ −γKX orb
, where γ > n−1 n and D is ef-
fective. As discussed in Section 5.2 this amounts to L2 conditions on the sections
−1
of powers of the orbifold anticanonical bundle O(KX ) = O(IX ). We emphasize
once more that the klt condition is only a sufficient condition for the existence of
a Kähler-Einstein orbifold metric, and that we give only sufficient conditions to be
klt. We have [BGK05] the following direct consequence of Theorem 5.2.16 applied
to weighted hypersurfaces.
Corollary 5.4.6: Let f ∈ H 0 (CP(w), O(d)) be quasi-smooth weighted homoge-
neous polynomial of degree df and weight vector w, and let Xdf denote the cor-
responding weighted homogeneous hypersurface with its induced orbifold structure
X . Assume further that X is a Fano orbifold, i.e., that IX = |w| − df > 0.
The continuity method produces a Kähler-Einstein orbifold metric on Xdf if there
is a γ > n−1 n such that for every nontrivial weighted homogeneous polynomial
g ∈ H (Xdf , O(sIX )) of degree sIX , the function |g|−γ/s is locally L2 on the punc-
0

tured affine cone Cf∗ .


Generally it is not easy to decide whether a given function |g|−c is locally L2
or not. However, for g defined as in Corollary 5.4.6, it is not difficult, though it
can be somewhat tedious, to estimate the multiplicities of its zeroes by intersection
theory. The following theorem was given in [BGK05].
Theorem 5.4.7: Let Xd ⊂ CP(w) be a weighted homogeneous hypersurface as in
Corollary 5.4.6, and assume further that the intersections of Cf with any number of
hyperplanes (zi = 0) are smooth outside the origin. If h is a weighted homogeneous
polynomial of degree
Q dh and for any c ∈ R the estimate cdf dh < mini,j {wi wj }
holds, then |h|−c i |zi |δi −1 is locally L2 on Cf∗ for any δi > 0.
This theorem is a generalization of Proposition 11 of [JK01b] as well as results
in [JK01b, BGK05] and in the Appendix of the preprint version of [BGN02a]
(math.DG/0108113). These all correspond to the case δi = 1 in Theorem 5.4.7,
and in this case one easily obtains the following
Corollary 5.4.8: Let Xd ⊂ CP(w) be a weighted homogeneous hypersurface as in
Corollary 5.4.6. Then Xd admits a Kähler-Einstein orbifold metric if the following
estimate holds:
n
dIX < min{wi wj } .
(n − 1) i,j
outline of proof. The proof of Theorem 5.4.7 is somewhat technical, so we
provide only an outline, and refer to the literature for details. The proof proceeds
by induction on dimension, reducing to the surface case. Suppose for a moment that
5.4. KÄHLER-EINSTEIN METRICS ON HYPERSURFACES IN CP(w) 167

X is a surface with at worst quotient singularities. Johnson and Kollár [JK01b]


gave sufficient conditions for the pair (X, D) to be klt, namely
Lemma 5.4.9: Let X be a normal surface with at worst quotient singularities and
D a Weil divisor on X. Then (X, D) is klt if the following three conditions hold:
Pr
(i) If D = i=1 αi Di then αi < 1 for all i,
(ii) for all smooth points x ∈ X, multx (D) ≤ 1,
(iii) if P ⊂ X is a singular point and π : Y → X a local finite cover re-
solving the singularity at P then π ∗ D has multiplicity at most one at
Q = π −1 (P ).
Condition (i) is also necessary, while neither (ii) nor (iii) is necessary. When
passing to the case of general dimension we cut X with the appropriate number of
hypersurfaces, and the condition of the hypothesis on the intersection of hyperplanes
allows us to apply inversion of adjunction (Theorem 5.2.17) to conclude that (X, D)
is klt. In general dimension the conditions (i) and (ii) remain essentially the same,
whereas, we must replace condition (iii) by
(iii’) Suppose x ∈ X is a singular point of X with local uniformizing group Γx
and suppose H1 , H2 , . . . , HdimX are hypersurfaces through x (including
the Di ), then we require the intersection number to satisfy
1
X · H1 · · · · · HdimX ≤ .
|Γx |
The details can be found in varying degree of detail in several places [JK01b,
JK01a, BGK05], but perhaps the most detail appears in the Appendix of the
preprint version (math.DG/0108113) of [BGN02a]. ¤

Let us now specialize to the case of a Brieskorn-Pham polynomial


Pof Examples
n
4.6.7 and 4.6.11, then the Fano condition |w| − d > 0 becomes 1 < i=0 a1i . So if
Xd ⊂ CP(w) is a Fano hypersurface arising from a BP polynomial, the estimate of
Corollary 5.4.8 takes the form
Xn
1 n n 1 o
(5.4.1) 1< <1+ min .
a
i=0 i
n − 1 i,j ai aj

These conditions are far from optimal, and we shall see how Theorem 5.4.7 gives
better estimates in the non-well-formed case. This comes about by reducing, via
the procedure of Section 4.5, to a well-formed orbifold. Let f be a BP polynomial
or a weighted homogeneous perturbation thereof.
Lemma 5.4.10: Let w = (w0 , . . . , wn ) be the weight vector and a = (a0 , . . . , an )
the exponent vector of a BP polynomial f of degree d, and let dj , ej and aw be given
by Definition 4.5.7. Then there is a BP polynomial g of degree d¯ = lcm(bi ) with a
well-formed weight vector w̄ = (w̄0 , . . . , w̄n ) and exponent vector b = (b0 , · · · , bn )
such that there is an isomorphism of algebraic varieties Xf ≈ Xg . Furthermore, the
following conditions hold
¯
(i) d = aw d,
(ii) d = C dj , where C j = lcm(ai |i 6= j),
j

(iii) bj = gcd(aj , C j ),
for all j = 0, . . . , n.
168 5. KÄHLER-EINSTEIN METRICS

Proof. We apply the procedure discussed in Section 4.5 (cf. Example 4.5.11).
We begin with the BP polynomial f = z0a0 + · · · + znan such that Xf ⊂ CP(w).
By Proposition 4.5.10 there is an isomorphism π : CP(w)−−→CP(w̄) of algebraic
d
varieties defined in local coordinates by π ∗ xj = zj j . Now π has degree aw = lcm(di )
and π ∗ xj has degree wj dj = dj ej w̄j = aw w̄j . Thus, xj has weight w̄j . Now d = wi ai
for all i = 0, . . . , n, and by Lemma 4.5.8 dj divides wi for all i 6= j, and is relatively
prime to wj . Thus, dj must divide aj for each j. So we define bi = adii and consider
Pn
the BP polynomial g = i=0 xbi i together with its corresponding weighted variety
Xg ∈ CP(w̄). Then we have
n
X n
X n
X ¡ ¢
f (z0 , . . . , zn ) = ziai = (zidi )bi = (π ∗ xi )bi = π ∗ g(x0 , . . . , xn ) .
i=0 i=0 i=0

Then Xg and Xf are isomorphic as algebraic varieties (but not of course as weighted
varieties or orbifolds). It follows from Lemma 4.5.8 that w̄ is well-formed. This
proves the first part.
For the three conditions, we first see that (i) follows from the fact that the de-
gree of π is aw . (ii) follows from the definition of C j and dj together with the equality
ai wi = d for all i. Then (iii) follows from dbj = C j dj bj = C j aj = d gcd(aj , C j ). ¤
We are now ready to prove our main estimate for Brieskorn-Pham polynomials
which was given in [BGK05]. Once again this estimate is not optimal at least in
certain cases as we shall see shortly.
Theorem 5.4.11: Let Xd be a weighted homogeneous hypersurface arising from a
Brieskorn-Pham polynomial or a weighted homogeneous perturbation thereof (i.e.,
condition (GC) of Example 4.6.7 holds). Then the orbifold Xd is Fano if and only
if
Xn
1
1<
a
i=0 i
and admits a Kähler-Einstein orbifold metric if
Xn
1 n n1 1 o
1< <1+ min , .
a
i=0 i
n − 1 i,j ai bi bj
In Theorem 5.4.11 the klt condition is the right hand inequality, and since
bi ≤ ai for all i it clearly is at least as good as the estimate of Equation (5.4.1).
We shall see that in a sense the less well-formed w is the better the estimate
becomes, whereas, in the well-formed case the estimate coincides with the estimate
of Equation (5.4.1). Notice also from Lemma 5.4.10 that if the ai ’s are pairwise
relatively prime the bi ’s are all 1, so the estimate is fairly good in this case, although
it can be improved upon as seen by Theorem 5.4.12 below.
Proof of Theorem 5.4.11. By Corollary 5.4.6 we need to show that for ev-
ery s > 0 and
P for every weighted homogeneous polynomial g of weighted degree
sIXd = sd( a−1 i − 1), the function |g|−γ/s
is locally L2 on the punctured affine

cone Cf +p of the perturbed BP polynomial f + p, where p is any weighted homo-
geneous polynomial of degree df such that the affine cone Cf +p satisfies condition
(GC) of Example 4.6.7. By Lemma 5.4.10 we reduce this to a problem on a per-
turbation of the simpler Brieskorn-Pham polynomial g to which we can apply the
sharper estimate given by Theorem 5.4.7. As in the proof of Lemma 5.4.10 we
5.4. KÄHLER-EINSTEIN METRICS ON HYPERSURFACES IN CP(w) 169

easily arrive at a perturbation q of g such that p(z0 , . . . , zn ) = q(x0 , . . . , xn ). This


then gives a branched covering map π : Cf∗+p −−→Cg+p ∗
of the punctured affine cones
of the perturbed polynomials.
More generally we consider an arbitrary weighted homogeneous polynomial
h with degree dh . If the variable zi has exponent ci in h then there is a unique
nonnegative integer ri such that zici = ziri zidi bi = ziri xbi i . So there is a polynomial
H(x0 , . . . , xn ) of degree dH such that
Y
(5.4.2) h(z0 , . . . , zn ) = ziri H(x0 , . . . , xn ).
i

Since the Jacobian of π has a (di − 1)-fold zero along (zi = 0), we see that
Q γe
− i + 1 −1
(?) |h|−γ/s is locally L2 on Cf∗+p if and only if |H|−γ/s · |xi | sdi di is
2 ∗
locally L on Cg+q .
We now apply Corollary 5.4.6 and Theorem 5.4.7 to the latter estimate with
h ∈ H 0 (CP(w), O(sIX )). First, each variable xi must have an exponent greater
than −1. This is Pequivalent to ei < γ −1 s. But weP know from above that ei wi ≤
dh = sIX = sd( ai − 1). So if we assume that ( a−1
−1 n
i − 1) < n−1 mini { ai } we
1

have
X n n1o n1o
ei ≤ ai s( a−1
i − 1) < ai s min < γ −1
a i s min < γ −1 s.
n − 1 i ai i ai
Next if we can satisfy the condition in the hypothesis of Theorem 5.4.7 for the
well-formed perturbed BP polynomial g + q, then we can use (?) together with
Corollary 5.4.6 to conclude the proof. Consider a weighted homogeneous polynomial
h ∈ H 0 (X, O(sIX )). We have using Equation (5.4.2)
d2g ³X 1 ´ n n d2 o n
g
dH dg ≤ dh = d2g s −1 ≤s min =s min{w̄i w̄j } . ¤
df i
a i n − 1 i,j b b
i j n − 1 i,j

For the special case of X = CPn−1 Ghigi and Kollár [GK05] have improved
the klt estimate of Theorem 5.4.11 by roughly a factor of n. Their idea is based
on the paper [AGP06] where the existence of Kähler-Einstein metrics on certain
manifolds of the form M/G, where G is a finite group, implies the existence on the
n−1
Galois cover M. In [GK05] this method is applied ¡ n−1to the
P¡identity
¢ map
¢ on CP .
1
Let us consider the orbifold X (a) = (Xd , ∆) = CP , 1 − ai Di , of Example
4.6.15, where the ai ’s are pairwise relatively prime, Di = {zi = 0} for i = 1, . . . , n
and D0 = {z1 + · · · + zn = 0}. Under these conditions the Ghigi-Kollár theorem is
Theorem 5.4.12: Let a = (a0 , . . . , an ) ¡be a sequence
P¡ of 1relatively
¢ ¢ prime integers.
Then the orbifold X (a) = (Xd , ∆) = CPn−1 , 1 − ai Di admits a Kähler-
Einstein metric if the following estimate holds:
Xn
1 n1o
0< − 1 < n min .
a
i=0 i
i ai
Surprisingly, it turns out that the Ghigi-Kollár estimate in this case is sharp,
i.e., the converse of this theorem is also true. We will discuss this in the proof
of Theorem 11.5.1 in Chapter 11 after we introduce obstructions to the existence
of Sasaki-Einstein metrics on links. In particular, we shall see that these new
obstructions give non-existence result for Kähler-Einstein metrics on many log del
Pezzo surfaces.
170 5. KÄHLER-EINSTEIN METRICS

We are interested in finding sequences of integers a = (a0 , . . . , an ) with n ≥ 3


that satisfy the inequalities of Theorem 5.4.11. For low values of n these inequalities
are easily solved on a computer. However, it is not only instructive to obtain some
solutions by hand, but also we can prove some interesting results by judicious
choices of sequences. We consider a very important sequence that does not satisfy
the inequalities, but marks the borderline of the left hand inequality. It is the
so-called extremal sequence or Sylvester’s sequence [GKP89] determined by the
recursion relation
(5.4.3) ck+1 = 1 + c0 · · · ck = c2k − ck + 1
beginning with c0 = 2. It starts as
2, 3, 7, 43, 1807, 3263443, 10650056950807, ...
The importance of this sequence is that it satisfies
Xn
1 1
(5.4.4) =1− .
c
i=0 i
c0 · · · cn

So we see that sequences of the form a = (a0 = c0 , . . . , an−1 = cn−1 , an ) satisfy the
left hand inequality as long as an < c0 · · · cn−1 . Furthermore, by construction the
first n elements of such sequences are all pairwise relatively prime. It remains to
analyze the right hand inequality.
Example 5.4.13: We continue the discussion of Example 4.6.15. Here recall that
a = (a0 , . . . , an ) consists of pairwise relatively prime integers, so d = a0 · · · an
and wi = a0 · · · aˆi · · · an . Thus, bj = gcd(aj , C j ) = 1 for all j in which case we
can use the sharper estimate of Theorem 5.4.12. There are many sequences that
satisfy this condition. Take a = (a0 = c0 , . . . , an−1 = cn−1 , an ) with cn−1 <
an < c0 · · · cn−1 with an relatively prime to c0 · · · cn−1 , then both inequalities are
satisfied. For example for n = 3 this gives the sequence (2, 3, 7, a3 ) with a3 =
11, 13, 17, 19, 23, 25, 29, 31, 37, 41. But we can also relax the condition an−1 = cn−1 ;
for example, we can take the sequence (2, 3, 5, a3 ) giving a larger range for a3 ,
namely, a3 = 7, 11, 13, 17, 19, 23, 29, 31, 37, 41, 43, 47, 49, 53, 59. Next, for n = 4 we
take a4 to be any integer between 43 and 1806 that is relatively prime to 1806, etc.
This example is just a special case and with the aid of a computer one can take
full advantage of the sharper estimates of both Theorem 5.4.11 and Theorem 5.4.12.
Our program, which is described in [BGKT05], generates a list of 68 sequences
for n = 3 and 8494 for n = 4. We shall make important use of these results later
in Chapter 11. In particular, we show that the number of sequences grow double
exponentially with dimension.
Now we discuss the more general case of weighted homogeneous polynomials,
which is based on the papers [JK01b, JK01a, Ara02, BGN03b, BGN02b].
These papers all assume that the weights are well-formed. This was done mainly for
convenience and can be dropped. However, for our discussion here we shall assume
that the weight vector w is well-formed, and restrict ourselves to the surface case,
i.e., n = 3. We first need to impose the conditions guaranteeing quasi-smoothness,
Exercise 4.4. From condition (i) of Exercise 4.4 we see that we need
(5.4.5) mi wi + wj(i) = d = w0 + w1 + w2 + w3 − IX ,
5.4. KÄHLER-EINSTEIN METRICS ON HYPERSURFACES IN CP(w) 171

where mi is a positive integer and both i and j(i) are integers ranging over 0, 1, 2, 3.
Note we can have j(i) = i. Since here we are only interested in Fano orbifold, we
impose the condition that the index be positive. In what follows we assume without
loss of generality that the weights are ordered w0 ≤ w1 ≤ w2 ≤ w3 . Then we have
Lemma 5.4.14: Assuming the quasi-smoothness conditions (i) and (iii) of Exer-
cise 4.4, and that IX > 0,, the following bounds hold
1. 1 ≤ m3 ≤ 2.
2. Either 2 ≤ m2 ≤ 4, or 2w0 ≤ IX .
3. Either 2 ≤ m1 ≤ 10, or 2IX = w0 + w1 , or 2IX ≥ 3w0 , or quasi-
smoothness condition(ii) of Exercise 4.4 is violated.
Proof. The proof is straightforward, though a bit tedious, so we refer the
reader to [BGN03b] for details. ¤
The solutions of the linear system (5.4.5) satisfying Lemma 5.4.14 split into two
cases depending on whether m0 is bounded or not. The case that m0 is unbounded,
gives series solutions depending on an integer k, and the case when m0 is bounded,
give the so-called sporadic solutions. The series solutions can be worked out by
hand, but the sporadic solutions require the aid of a computer. In [BGN03b]
we modified the computer program of [JK01b] to solve the system (5.4.5) for any
index I as well as to discard some solutions that are not quasi-smooth. Notice that
there are solutions with arbitrary m1 that are quasi-smooth, but our lemma implies
that these must satisfy 2I ≥ 3w0 or 2I = w0 + w1 . As we shall see, in both cases the
sufficient condition for the existence of a Kähler-Einstein metric fails. Examples of
such a series are: (1, 1, k, k) of degree 2k and index 2, (1, 2, 2k + 1, 2k + 1) of degree
2(2k + 1) and index 3, and (2, 2, 2k + 1, 2k + 1) of degree 2(2k + 1) and index 4, as
well as the general series (I − n, I + n, w, w + n) of degree 2w + n + I and index I.
There are also double series such as (1, 1, m, m + k) of degree 2m + k and index 2.
We now make use of the inequalities in Lemma 5.4.14 involving IX .
Lemma 5.4.15: If 2IX ≥ 3w0 or 2IX = ¡w0 + w1 then ¢ for every ² > 0 there exists
a divisor D ∈ | − KX | such that the pair X, 2+² 3 D is not klt.
Proof. The first inequality is easy for it implies that the divisor D0 = (z0 = 0)
satisfies rD0 ∈ |OX (−KX )| for r ≥ 32 . But this says that 2+² 3 D0 has multiplicity
2+²
greater than 1 at a smooth point, so (X, 3 D) cannot be klt by (ii) of Lemma
5.4.9.
Next assume the second equality holds. We notice in the proof (Lemma 5.1 in
[BGN03b]) of Lemma 5.4.14 that 2I = w0 + w1 occurs in precisely one case and
in this case we have w3 = w1 + w2 − I. But then w must have the form
w = (I − n, I + n, w, w + n)
+
for some w ∈ Z with w ≥ I + n, and some non-negative integer n < I and in
this case the degree d = 2w + n + I. Now suppose that z0a0 z1a1 z2a2 z3a3 is a monomial
occurring with nonzero coefficient in the polynomial defining X . We claim that if
a0 = 0 then a1 6= 0. Indeed, if a0 = a1 = 0 then
a2 w + a3 w = 2w + 2I + n .
But w ≥ I+n so the only possible solutions would require a2 +a3 = 3 and w = 2I+n
and one readily checks that these hypersurfaces are never quasi-smooth. Thus the
divisor D = {z0 = 0} ∩ X has at least two components, E and F, where E is the
172 5. KÄHLER-EINSTEIN METRICS

line z0 = z1 = 0 and F is defined, inside the weighted projective plane {z0 = 0},
by a polynomial f (z1 , z2 , z3 ). Moreover, f (z1 , z2 , z3 ) = z22 + z12 (g(z1 , z2 , z3 )). Note
that the point P = (0, 0, 0, 1) ∈ X since w + n does not divide the degree of X .
Thus if π : C2 −−→X is a local cover of the quotient singularity at P = (0, 0, 0, 1)
then π ∗ D = π ∗ E + π ∗ F has multiplicity at least 1 + mult0 (π ∗ F ) at the origin. To
compute the multiplicity of π ∗ F at the origin, let Y = Z(f (z1 , z2 , z3 )), the zero set
of f. Then Y ∩ X = F ∪ G, where G does not contain the point (0, 0, 0, 1). Thus
mult0 (π ∗ F ) = mult0 (π ∗ Z(f )) ≥ 2 .
Consequently, (X, 2+²3 D) is never klt as it always has multiplicity > 2 at 0. But
I
I−n D ∈ | − K X | so this completes the proof of the lemma. ¤
Now with the aid of our computer program we have
Theorem 5.4.16: Let X = (X, U) be a well-formed quasi-smooth log del Pezzo
surface with weight vector w, index 0 < IX ≤ 10 and degree d embedded in the
weighted projective space CP(w) = CP(w0 , w1 , w2 , w3 ). Suppose further that X sat-
isfies the klt condition on the pair (X, 2+²
3 D), where D ∈ | − KX |, then X must
belong to one of the cases listed in Tables B.2.1 and B.2.2 in Appendix B.
Outline of Proof. With IX > 0 the orbifold X is Fano, and the quasi-
smoothness and klt assumptions together with Lemmas 5.4.14 and 5.4.15 imply the
range for mi indicated in Lemma 5.4.14 for the computer program as well as the
inequalities 2I < 3w0 and 2I 6= w0 + w1 . The computer program then produces
tables B.2.1 and B.2.2 of Appendix B assuming the indicated range for the index
IX . The last column of these tables indicates whether the klt condition (which
implies that X admits a Kähler-Einstein metric) holds (Y) or is unknown (?). To
verify the klt condition for the sporadic case we use the following corollary which is
a sharpening of Corollary 5.4.8 for the surface case and whose proof can be found
in [BGN03b]:
Corollary 5.4.17: Let w = (w0 , w1 , w2 , w3 ) and X ⊂ CP(w) be a quasi-smooth
surface of degree d = w0 + w1 + w2 + w3 − IX . Then X admits a Kähler-Einstein
orbifold metric if 2IX d < 3w0 w1 . If the line {z0 = z1 = 0} 6⊂ X then 2IX d < 3w0 w2
is also sufficient. If the point (0, 0, 0, 1) 6∈ X then 2IX d < 3w0 w3 is also sufficient.
The proof of the klt condition in the series case is more delicate. Again full
proofs can be found in [BGN03b, BGN02b]. ¤
Remark 5.4.1: The computer program indicates that there are neither series solu-
tions nor sporadic solutions satisfying the hypothesis of Theorem 5.4.16 for IX > 10.
In fact, an easy argument shows that there are no such solutions for sufficiently large
IX . 2
A more systematic study of all (not just well-formed) log del Pezzo hypersur-
faces in CP(w0 , w1 , w2 , w3 ) can be carried out in a similar way. Here the starting
point has recently been established by Yau and Yu [YY05] who, based on the much
earlier work of Kouchnirenko [Kou76] and Orlik and Randell [OR77], presented a
table of 19 distinct types of quasi-smooth weighted homogeneous hypersurfaces Xf
in CP(w0 , w1 , w2 , w3 ). More precisely we have

2The code for the C program used to generate the tables of the Theorem 5.4.16 are available
at: http://www.math.unm.edu/˜galicki/papers/publications.html.
5.5. AUTOMORPHISMS AND MODULI 173

Theorem 5.4.18: Let f (z) ∈ C[z0 , z1 , z2 , z3 ] be such that Yf = {z ∈ C4 | f (z) = 0}


has only an isolated singularity at the origin. Then f (z) = h(z) + g(z), where
h(z) is one of the 19 classes of weighted homogeneous polynomials with isolated
singularities at the origin listed in Table B.5.1 of Appendix B and h(z) and g(z) have
no monomial in common. If f (z) is weighted homogeneous with rational weights
(d/w0 , d/w1 , d/w2 , d/w3 ) then so are h(z) and g(z). Furthermore, let Yh = {z ∈
C4 | h(z) = 0} and let Lf and Lh be the associated links. Then Lf is equivariantly
diffeomorphic to Lh .
Open Problem 5.4.1: All Xh ⊂ CP(w0 , w1 , w2 , w3 ), where h(z) is one of the
weighted homogeneous polynomials mentioned in Theorem 5.4.18, have definite or
zero first Chern class. One can partition the list depending on the sign of corb 1 (Xh )
and try to determine which of the log del Pezzo surfaces can possibly admit a
Kähler-Einstein metric using the methods of this chapter.
We end this section with a brief discussion of the higher dimensional case. The
ideas described here can clearly be applied to any dimension, although computer
searches become much more time consuming as the dimension increases. In this
regard the only further results known are in dimension 7, where Johnson and Kollár
[JK01a] have produced a list of 4442 Fano 3-folds that embed into weighted projec-
tive 4-spaces CP(w0 , w1 , w2 , w3 , w4 ). Out of this list they showed that 1936 of these
admit Kähler-Einstein metrics using the techniques described above. However, so
far little is known about the topology of the orbifolds on this list. To date the only
known results are those of [BGN02a] where 184 3-folds with b3 = 0 were isolated.
This was used to give Sasaki-Einstein metrics on rational homology 7-spheres, and
will be discussed further in Chapter 11.
Example 5.4.19: [Iano-Fletcher/Reid Fano 3-folds] Consider any orbifold K3
surface of the Appendix B.1. Suppose we take any Xd ⊂ CP(w0 , w1 , w2 , w3 ) repre-
sented by some Xf with some f = f (z0 , z1 , z2 , z3 ). Then g = wd + f (z0 , z1 , z2 , z3 )
is a weighted homogeneous polynomial of the same degree and with weights w =
(1, w0 , w1 , w2 , w3 ). The 3-fold Xg is Fano of index one as I = 1 + w0 + w1 + w2 +
w3 − d = 1. In [IF00] Iano-Fletcher shows that
Lemma 5.4.20: There is a one-to-one correspondence between the following two
classes
(i) the set of families of quasismooth, well-formed weighted hypersurfaces
Xd2 ⊂ CP(w0 , w1 , w2 , w3 ) with only canonical singularities and trivial orb-
orb
ifold canonical class KX = 0,
3
(ii) the set of families Xd ⊂ CP(1, w0 , w1 , w2 , w3 ) of quasismooth weighted
3-folds of index I = 1.
Hence, one gets 95 examples of anti-canonically embedded Fano 3-folds which
correspond to the Reid’s K3 orbifold surfaces. It was recently shown by Chelstov
that all but 5 of these admit Kähler-Einstein orbifold metrics [Che07].

5.5. Automorphisms and the Moduli Problem


In this section we are mainly concerned with describing the moduli space of
Kähler-Einstein metrics on weighted homogeneous hypersurfaces in CP(w). This
will then give moduli for Sasaki-Einstein and Sasaki-η-Einstein metrics which we
discuss in detail in Chapter 11.
174 5. KÄHLER-EINSTEIN METRICS

5.5.1. Automorphisms of Weighted Projective Spaces. We begin by


discussing the group Aut(CP(w)) of complex automorphisms of a weighted projec-
tive space CP(w). The group C∗ (w) of Definition 4.5.1 is a subgroup of the group
Aut(Cn+1 ) of biholomorphic maps from Cn+1 to itself, and we let Z(C∗ (w)) denote
the centralizer of C∗ (w) in Aut(Cn+1 ). Clearly C∗ (w) ⊂ Z(C∗ (w)). We have
Lemma 5.5.1: The automorphism group Aut(CP(w)) is isomorphic to the quo-
tient group Z(C∗ (w))/C∗ (w) and Z(C∗ (w)) can be characterized as the subgroup
of Aut(Cn+1 ) consisting of biholomorphic maps defined by sending zi to a linear
combination of monomials P of degree wi for all i = 0, . . . , n. Thus, Aut(CP(w)) is a
Lie group of dimension i h0 (CP(w), O(wi )) − 1.

Proof. First, Aut(CP(w)) is a Lie group by Proposition 1.6.7 which holds


equally well for compact orbifolds. Furthermore, since CP(w) = (Cn+1 \{0})/C∗ (w)
the elements of Aut(CP(w)) all lift to biholomorphic maps of Cn+1 \ {0} which
commute with C∗ (w). It follows by Hartogs’ Theorem that Aut(CP(w)) is a closed
subgroup of Aut(Cn+1 ). Moreover, if τ ∈ Aut(Cn+1 ) commutes with C∗ (w) then it
descends to element in Aut(CP(w)). So there is a group epimorphism
Z(C∗ (w))−−→Aut(CP(w))
whose kernel is C∗ (w).
Now let τ ∈ Z(C∗ (w)). Then τ (zi ) = gi (z0 , . . . , zn ), where gi is a holomorphic
function of the n + 1 variables z0 , . . . , zn . Since τ commutes with C∗ (w), gi satisfies
gi (λw0 z0 , . . . , λwn zn ) = λwi gi (z0 , . . . , zn ) ,
that is, gi is a weighted homogeneous polynomial of degree wi . This means that
gi consists of a linear combination of monomials of degree wi . Such monomials
span the finite dimensional vector space H 0 (CP(w), O(wi )) whose dimension is
h0 (CP(w), O(wi )). ¤

Now let Xf be a quasi-smooth weighted hypersurface in CP(w), and let Aut(Xf )


denote its complex automorphism group. We are interested in knowing when
Aut(Xf ) is a closed subgroup of Aut(CP(w)). Then
Lemma 5.5.2: If n ≥ 4 or if n = 3 and |w| 6= d, then Aut(Xf ) is the closed
subgroup of Aut(CP(w)) that leaves f invariant, that is, τ ∗ f = f. Hence, Aut(Xf )
is an algebraic group.

Proof. Let ιf : Xf −−→CP(w) denote the inclusion map. If n ≥ 4 then


Picorb (Xf ) = Z and the line orbibundle ι∗f O(1) generates Picorb (Xf ) = Z by
the Grothendieck-Lefschetz theorem [Gro68]. Thus, ι∗f O(1) is invariant under
Aut(Xf ), so it is a closed algebraic subgroup of Aut(CP(w)). This also occurs when
n = 3 and |w| 6= d since in this case either Korb or (Korb )−1 is ample, invari-
ant under Aut(Xf ), and the line orbibundle ι∗f O(1) is a unique root of Korb or
(Korb )−1 . ¤

Our next result says that Aut(Xf ) is a discrete group, or alternatively that the
Lie algebra aut(Xf ) vanishes, in most cases.
Lemma 5.5.3: If 2wi < d for all but at most one i then the group Aut(Xf ) is
discrete. If in addition n ≥ 4 or n = 3 and |w| 6= d, then the group Aut(Xf ) is
finite.
5.5. AUTOMORPHISMS AND MODULI 175

Proof. For the first statement it is enough to show that there are no one
parameter subgroups of Aut(Xf ) under the hypothesis of the lemma. Let τt be a
one parameter subgroup of Aut(Xf ) ⊂ Aut(CP(w)), and write
X
zi0 = τt (zi ) = zi + tj gij (z0 , . . . , zn ) .
j≥1

The condition of invariance is


f (τt (z0 ), . . . , τt (zn )) = f (z0 , . . . , zn ) .
Differentiating with respect to zi , expanding in a Taylor series in t and equating
the coefficients of tj for each j gives
X ∂f
gkj = 0.
∂zk
k
∂f ∂f
Since the degree of gkj is wk and the degree of ∂z k
is d − wk , and not all ∂z k
’s
vanish, this implies that either all gkj vanish or there is a cancellation for two
different values of k, say k and l. But the only way this can happen is that wk = wl
∂f ∂f
and that both ∂z k
and ∂z l
are linear in zk or zl . But this implies that wk = wl = d2 .
This proves the first statement. The second statement follows from Lemma 5.5.2
since an algebraic group has finitely many components. ¤

This result was stated incorrectly in [BGK05]. The hypothesis n ≥ 3 or n = 2


and |w| 6= d was left off. It was, however, used correctly since all cases treated in
[BGK05] are Fano. That the extra hypothesis is necessary in general is illustrated
by the classical example of Fano and Severi who exhibited a quartic surface in CP3 ,
where Aut(Xf ) is an infinite discrete group. This is discussed in detail in [MM64].
We mention also the related result of Kobayashi [Kob59a, Kob72] which says
that Aut(Xf ) is finite for complex manifolds with negative first Chern class.

5.5.2. The Moduli Problem for Kähler-Einstein metrics. It is not our


intention here to discuss moduli theory in any generality. This is a very important
part of modern mathematics which is treated in detail elsewhere, cf. [Kod86,
MFK94, SU02]. Rather, we are interested in the moduli problem for Kähler-
Einstein metrics, and more specifically mainly those arising from hypersurfaces
embedded in weighted projective spaces. Later in Section 8.2.1 we review Kodaira-
Spencer theory in the context of the transverse holomorphic geometry of a foliation,
but here we proceed in a much more pedestrian fashion by contenting ourselves with
a count of effective parameters for Kähler-Einstein orbifold metrics. For now the
only bit of information from complex deformation theory that we need explicitly
is the well-known fact [Kod86] that H 1 (X, Θ) is the ‘tangent space’ to the space
of deformations of complex structures on X, where Θ is the sheaf of germs of
holomorphic vector fields on X. First, we mention a general result
Theorem 5.5.4: Let X = (X, U) be a compact orbifold with a Kähler-Einstein
orbifold metric g. Then
(i) If c1 (X )orb < 0 then g is unique up to homothety.
(ii) If c1 (X )orb = 0 then g is unique up to homothety in each positive (1, 1)
cohomology class.
(iii) If c1 (X )orb > 0 then g is unique up to a complex automorphism.
176 5. KÄHLER-EINSTEIN METRICS

In the manifold case uniqueness in (i) and (ii) was proved originally by Calabi
and follows from Theorem 5.2.2. Case (iii) is due to Bando and Mabuchi [BM87].
A proof can also be found in [Aub98]. The proof for orbifolds is essentially the
same.
We mention that in case (iii) of Theorem 5.5.4, the Lie algebra of the complex
automorphism group is the complexification of the Lie algebra of the isometry group
by Theorem 5.3.1. Theorem 5.5.4 says that the moduli space of Kähler-Einstein
metrics is determined by the moduli space of complex structures. In case (ii) one
should also take into account the contribution from all the admissible (1, 1)-classes.
Thus, one needs to consider the moduli space of ‘marked pairs’ consisting of complex
manifold (or orbifold) X and Kähler class on it (cf. [BHPVdV04], pg 334-336,
and Chapter 12 of [Bes87]). For example, in the case of K3 surfaces this moduli
space has real dimension 60; therefore, the moduli space of Kähler-Einstein metrics
(up to homothety) on K3 has real dimension 59. However, the moduli space of
marked pairs is of much less interest to us in this book since we are more concerned
with the moduli space of a polarized Kähler orbifold.
The study of Einstein moduli spaces on complex manifolds is due to Koiso
[Koi83]. Here we present a very brief summary of his work which has been treated
in much more depth in Chapter 12 of [Bes87]. Following Besse we let g be an
orbifold Einstein metric on an orbifold3 M , and denote the space of infinitesimal
Einstein deformations by ε(g). This space is characterized by symmetric 2-tensors
h which satisfy certain equations due to Berger and Ebin [BE69] who show that
ε(g) is finite dimensional and should be thought of as the tangent space to the
‘premoduli space’ of Einstein structures near g. Here we are interested in the case
that X is a complex orbifold and g is a Kähler-Einstein metric on it. In this case
one gets a splitting ε(g) = ε(g)H ⊕ ε(g)A into Hermitian and skew-Hermitian parts.
Then for manifolds Koiso [Koi83] proves the following result (which also holds for
complex orbifolds as long as c1 (X) is replaced by corb1 (X)).
Theorem 5.5.5: Let (J, g) be a Kähler-Einstein structure on X with fixed volume
say, one, and assume that J belongs to a complete4 family of complex structures.
Then
(i) if corb orb
1 (X) < 0 or c1 (X) = 0, then every Einstein metric g in the
0

premoduli space is Kähler with respect to a deformed complex structure,


(ii) if corb
1 (X) > 0, ε(g)H = 0, and there are no nonzero holomorphic vector
fields on X, then every Einstein metric g 0 in the premoduli space is Kähler
with respect to a deformed complex structure.
Moreover, the following hold
(a) if corb 1
1 (X) < 0, then dim ε(g) = 2dimC H (X, Θ),
orb
(b) if c1 (X) = 0, then
dim ε(g) = 2dimC H 1 (X, Θ) − 2dimC H 2,0 (X, C) + dim H 1,1 (X, R) − 1 ,
(c) if corb 1
1 (X) > 0, then dim ε(g) ≥ 2dimC H (X, Θ).
This implies that the subset of Kähler-Einstein metrics is open in the premoduli
space of infinitesimal Einstein deformations.

3The treatment in [Koi83] as well as [Bes87] assumes that M is a manifold, but the discus-
sion carries through for orbifolds by working on the local uniformizing covers.
4See Koiso or Definition 1.7 of [KS58] for the definition.
5.5. AUTOMORPHISMS AND MODULI 177

Exercise 5.2: Show that Koiso’s proof of Theorem 5.5.5 holds for orbifolds when
c1 (X) is replaced by corb
1 (X).

5.5.3. The Moduli Problem for Weighted Hypersurfaces. We now turn


to the problem at hand, namely the description of the moduli space of complex
embeddings of a projective algebraic orbifold X as a weighted hypersurface into a
fixed weighted projective space CP(w).
Proposition 5.5.6: Let Xf be a quasi-smooth weighted hypersurface in CP(w)
corresponding to the weighted homogenous polynomial f of degree d and weight
vector w = (w0 , . . . , wn ). Let Aut(Xf ) denote the group of automorphisms of Xf .
Assume also that n ≥ 3. Then the complex orbifolds X = (Xf , U ) form a continuous
family of complex dimension
¡ ¢ X 0¡ ¢
h0 CP(w), O(d) − h CP(w), O(wi ) + dim Aut(Xf ) .
i

Proof. The space of ¡ weighted homogeneous


¢ polynomials
¡ of degree
¢ d is the
complex vector space H 0 CP(w), O(d) of dimension h0 CP(w), O(d) . The action
of Aut(CP(w))
¡ on CP(w)
¢ induces a representation of Aut(CP(w)) on the vector
space H 0 CP(w), O(d) . The hypersurfaces cut out by two polynomials
¡ ¢
f, g ∈ H 0 CP(w), O(d)
are isomorphic if they differ by a transformation in Aut(CP(w)). So we get a complex
family of orbifolds of complex dimension at least
¡ ¢ X 0¡ ¢
h0 CP(w), O(d) − h CP(w), O(wi ) .
i
¡ ¢
However, the isotropy subgroup of f ∈ H 0 CP(w), O(d) is Aut(Xf ), so any one
parameter subgroup of Aut(Xf ) cannot contribute to the equivalence between f
and g implying the result. ¤

We now combine Proposition 5.5.6 with Theorem 5.5.4 to obtain


Theorem 5.5.7: Let Xf be a quasi-smooth weighted hypersurface of degree d in
CP(w) and let I = |w| − d < 0. Then
£ ¡ ¢ P ¡ ¢¤
(i) If I < 0 then Xf admits a 2 h0 CP(w), O(d) − i h0 CP(w), O(wi )
dimensional family of Kähler-Einstein
£ ¡ metrics
¢ Pup to¡ homothety, ¢¤
(ii) If I = 0 Xf admits a 2 h0 CP(w), O(d) − i h0 CP(w), O(wi ) di-
mensional family of Kähler-Einstein metrics up to homothety in each
positive (1, 1) cohomology class.
(iii) If I > 0 and£ Xf¡ admits a Kähler-Einstein
¢ P 0¡ metric for¢a generic f then¤
0
it admits a 2 h CP(w), O(d) − i h CP(w), O(wi ) + dim Aut(Xf )
dimensional family of Kähler-Einstein metrics up to homothety.
We give an application of this theorem.
Example 5.5.8: Consider the log del Pezzo orbifold surface X210 of degree 210
obtained from the sequence a = (35, 7, 3, 2). This is represented by the zero set
of the weighted homogeneous polynomial f = f7 (z05 , z1 ) + z23 + z32 in the weighted
projective space CP(w) with weight vector w = (6, 30, 70, 105), where f7 is any
sufficiently general homogenous septic polynomial. The conditions of Theorem
178 5. KÄHLER-EINSTEIN METRICS

5.4.11 are easily seen to be satisfied. So X210 admits Kähler-Einstein metrics.


Aut(CP(w)) is given modulo the C∗ (w) action by the transformations
(z0 , z1 , z2 , z3 ) 7→ (α0 z0 , α1 z1 + β10 z05 , α2 z2 , α3 z3 ) .
P
So we have i h0 (CP(w), O(wi )) = 5, and by Lemma 5.5.3 dim Aut(Xf )) = 0.
Now it is easy to see that h0 (CP(w), O(210)) = 10. So we have a 2(10 − 5) = 10
parameter family of Kähler-Einstein metrics on X210 . In Chapter 11 we show that
this gives a 10 parameter family of Sasaki-Einstein metrics on the 5-sphere S 5 .
An important result of Kodaira and Spencer ([KS58], Theorem 14.1) says that
for smooth hypersurfaces Hn,d of degree d in the complex projective space CPn+1
the complex analytic family given by Proposition 5.5.6 is actually complete except
when n = 2 and d = 4. It would be interesting to obtain a similar result for weighted
hypersurfaces.
Open Problem 5.5.1: Discuss the completeness problem for weighted homoge-
neous hypersurfaces in CP(w).
It is also easy to work out the dimension of the moduli space of Kähler-Einstein
metrics on Xk equal to CPn blown-up at k points for certain k. For simplicity we
consider n = 2 and the del Pezzo surfaces Xk for 0 < k ≤ 8. As discussed previously
there are no Kähler-Einstein metrics on X1 or X2 .
Exercise 5.3: Assuming Tian’s Theorem 5.2.18 show that the moduli space of
Kähler-Einstein metrics on Xk for 4 ≤ k ≤ 8 has real dimension 4(k − 4).
Exercise 5.4: Compute the dimension of the parameter space of Kähler-Einstein
metrics for the first three entries of Table 1 in Subsection 5.4.1 using Theorem 5.5.7.
See that it agrees with the answer in Exercise 5.3.
Exercise 5.5: Compute the dimension of the parameter space of Kähler-Einstein
metrics for the surfaces in Tables 2, 3, and 4 of Subsection 5.4.1
CHAPTER 6

Almost Contact and Contact Geometry

In this chapter we describe the basics of contact and almost contact geometry.
Contact and symplectic geometries both lie at the very heart of the mathematical
foundations of Hamiltonian and Lagrangian mechanics. As such they have been
studied extensively for over three centuries. We refer the reader interested in the
geometric description of mechanics to the classic texts of Arnold [Arn78] or Abra-
ham and Marsden [AM78] on the subject. In this chapter we rather focus on a
slightly different aspect of almost contact and contact structures in the way they
naturally lead towards the development of Sasakian geometry. We also do not touch
on the interesting advances in contact topology, but refer the reader to the recent
review [Gei06].
Our goal is to show how Riemannian metrics can be naturally incorporated
into the fabric of contact geometry. Starting with a quasi-regular (strict) contact
manifold (M, η) we will consider it simultaneously with the associated symplectic
cone on one hand and the symplectic orbifold transverse space on the other. At
the beginning of this chapter we shall develop all the necessary non-Riemannian,
i.e., contact and symplectic ingredients. At the end we will arrive at the definition
of the Sasakian geometry together with the full description of how it relates to the
two Kähler geometries of the cone and the transverse space, respectively. We can
summarize the context and the structure of Chapter 6 by putting together the two
diagrams of the Introduction.

(C(M ), ω) ←- (M,η, ξ) (C(M ), ω, Φ̄, ḡ) ←- (M, ξ,η, Φ, g)


 
 
π =⇒ π
y y
(Z, Ω) (Z, Ω, J, g)

The geometric structures described in the second box will be the main subject of
Chapter 7. We remark yet again that, unfortunately, the terminology is not com-
pletely standard although preference has been given to defining a contact structure
as a codimension one subbundle of the tangent bundle that is as far as possible from
being integrable (see Proposition 6.1.9 below). However, what is worse, is that pref-
erence has also been given to defining an almost contact manifold [Gra59] in terms
of an equivalence class of reductions of the orthonormal frame bundle to the group
U (n) × 1, and with this convention a contact manifold is not necessarily an almost
contact manifold! Although this, in the words of Stong [Sto74], is “terrible”, it
seems to be too well ensconced in the literature to be rectifiable. Fortunately, for
179
180 6. ALMOST CONTACT AND CONTACT GEOMETRY

our purposes this problem of terminology is mute, since the confusion only arises
when the contact line bundle is not trivial, and the contact structures of interest to
us always have trivial contact line bundle. Nevertheless, we take some pains to set
things up correctly in the beginning. We shall then revert to the usual convention.

6.1. Contact Structures


6.1.1. Contact Transformations in R2n+1 . Contact transformations arose
in the theory of Analytical Mechanics developed in the 19th century by Hamilton,
Jacobi, Lagrange, and Legendre. But its first systematic treatment was given by
Sophus Lie [Lie77] in 1896 under the German name Berührungstransformation.
Consider R2n+1 with Cartesian coordinates (x1 , . . . , xn ; y1 , . . . , yn ; z), together with
a 1-form α given by
X
(6.1.1) α = dz − yi dxi .
i

It is easy to see that α satisfies


(6.1.2) α ∧ (dα)n 6= 0
on R2n+1 . A 1-form that satisfies this equation is called a contact form.
Remark 6.1.1: The choice of α in equation (6.1.1) is not unique. Other choices
are probably just as reasonable, for example,
n
0 1X
α = dz + (xi dyi − yi dxi ) .
2 i=0

This clearly satisfies (6.1.2), and differs from α by an exact form, namely, 12 d(x · y).
We shall mainly use the form α of equation (6.1.1) which we call the standard
1-form on R2n+1 . There is local converse to this, namely, the contact version of
Darboux’s theorem.
Theorem 6.1.1: Let α be a 1-form on R2n+1 that satisfies equation (6.1.2). Then
there is an open set U ⊂ R2n+1 and local coordinates (x1 , . . . , xn ; y1 , . . . , yn ; z) such
that α has the form (6.1.1) in U.
Darboux’s theorem says that locally all contact forms look the same. A proof
can be found in most books on Analytical Mechanics, e.g., [AM78, LM87]. The
modern proof is due to Moser [Mos65] and Weinstein [Wei77a]
Definition 6.1.2: A diffeomorphism φ : U −−→V where U, V are open sets of R2n+1
is called a contact transformation if there is a nowhere vanishing smooth func-
tion f such that
φ∗ α = f α .
If f ≡ 1 on U , then φ is called a strict contact transformation.
The collection ΓCon of all such contact transformations forms a pseudogroup,
called the contact pseudogroup; it is one of the transitive infinite, simple pseu-
dogroups of Lie and Cartan. In fact ΓCon is an example of a transitive pseudogroup
which does not come from an integrable G-structure. It is, however, related to a
non-integrable G-structure as we shall see.
Exercise 6.1: Show that the subset of strict contact transformations forms a sub-
pseudogroup ΓsCon of ΓCon .
6.1. CONTACT STRUCTURES 181

Next we describe the Lie algebra of infinitesimal contact transformations. By


taking φt to be a local 1-parameter group of transformations generated by a vector
field X on U , it is easy to see that the infinitesimal version of Definition 6.1.2 is:
Definition 6.1.3: An infinitesimal contact transformation is a local vector
field X defined on an open set U ⊂ R2n+1 that satisfies
£X α = f α ,
where f is a smooth function on U. If f vanishes on U, then X is called a infinites-
imal strict contact transformation. Let scon(U ) and con(U ) denote the subsets
of all vector fields on U consisting of infinitesimal strict contact transformations
and infinitesimal contact transformations, respectively.
Exercise 6.2: Show that con(U ) forms a Lie subalgebra of the Lie algebra of all
vector fields on U. Show that scon(U ) is a Lie subalgebra of con(U ).
It is easy to see that the vector space con(U ) of all infinitesimal contact trans-
formations on U forms a Lie algebra under the Lie bracket of vector fields. In order
to see that the contact pseudogroup ΓCon is transitive, we shall compute explic-
itly the Lie algebra of infinitesimal contact transformations. Using the well-known
identity £X α = d(X α) + X dα, one sees that X must take the form
µ Xn ¶ Xn Xn µ ¶
∂F ∂ ∂F ∂ ∂F ∂F ∂
(6.1.3) X= F− yi − + yi + ,
i=1
∂y i ∂z i=1
∂y i ∂x i i=1
∂z ∂x i ∂y i

where F is an arbitrary smooth function on U known as a Hamiltonian function


for the infinitesimal contact transformation X.
Example 6.1.4: Consider the subalgebra h2n+1 of scon(R2n+1 ) spanned by the
vector fields
∂ ∂ ∂ ∂
(6.1.4) Z= , Xi = , Yi = + xi
∂z ∂xi ∂yi ∂z
which satisfy the Lie bracket identities
[Xi , Yj ] = δij Z , [Xi , Z] = [Yi , Z] = 0 .
h2n+1 is known as the Heisenberg algebra, and the Lie group H2n+1 that it generates
is the Heisenberg group. Here we have realized H2n+1 as a Lie subgroup of the
pseudogroup ΓsCon of strict contact transformations on R2n+1 .
Exercise 6.3: Show that the Heisenberg group H2n+1 is realized as a subgroup of
ΓsCon by the transformations (using vector notation)
(x, y, z) 7→ (x + a, y + b, z + c + x · b) .
Notice that as manifolds H2n+1 = R2n+1 whose group law is given by (a, b, c) ·
(x, y, z) = the right hand side of the above equation. It is clear that H2n+1 acts
transitively on R2n+1 ; hence, we have arrived at
Proposition 6.1.5: ΓsCon and ΓCon are transitive pseudogroups.
6.1.2. Basic Definitions. We begin with the most general definition of a
contact manifold. We give a more restricted definition below, which is the one we
concentrate on in this monograph.
Definition 6.1.6: A 2n+1 dimensional manifold M with a ΓCon -structure is called
a contact manifold. If M has a ΓsCon -structure, then it is called a strict contact
manifold.
182 6. ALMOST CONTACT AND CONTACT GEOMETRY

The definition of contact manifold appears to be due to Gray [Gra59]. This


has also been given the somewhat cumbersome name contact structure in the wider
sense, but we shall not use this terminology. A contact structure on M gives rise to
a real line bundle L called the contact line bundle as follows. In the chart (Ui ; φi )
we have the contact form φ∗i α, and in the overlap Ui ∩ Uj there exists a nowhere
vanishing function ρij such that

(6.1.5) φ∗i α = ρij φ∗j α .

The {ρij } define a cocycle for the sheaf cohomology group H 1 (M, G`(1, R)) where
G`(1, R) is the sheaf of nowhere vanishing real functions on M, and thus a real
line bundle L. The cocycle {ρij } is trivial or a coboundary if for all i, j the ρij
can be taken to be positive, that is, local sections of the sheaf G`(1, R)+ of strictly
positive functions. In this case, there are positive functions fi such that for all i, j
we have fi φ∗i α = fj φ∗j α. Thus, we have a nowhere vanishing global section η of L
defined on each Ui by η|Ui = fi φ∗i α. This trivializes L in which case M admits a
global contact 1-form. In the sequel we shall deal almost exclusively with contact
manifolds with trivial contact line bundle L. In fact it is quite common to take this
as the definition of a contact manifold. We shall often consider a contact structure
ΓCon on a strict contact manifold M. In this case the maximal atlas defining the
strict contact structure is a subatlas of the corresponding maximal atlas defining
ΓCon . In this case the contact structure can be defined as an equivalence class of
globally defined contact 1-forms. Thus, we are led to our second slightly more
restrictive definition of a contact manifold.
Definition 6.1.7: A 2n + 1 dimensional manifold M is a (strict) contact man-
ifold if there exists a 1-form η, called a contact 1-form, on M such that

η ∧ (dη)n 6= 0

everywhere on M. A contact structure on M is an equivalence class of such


1-forms where η 0 ∼ η if there is a nowhere vanishing function f on M such that
η 0 = f η.
For purposes of this section only, we shall refer to a contact structure in the
sense of Definition 6.1.7 as a strict contact structure. Clearly for any real line
bundle L the tensor product with itself L2 is trivial. This gives rise to the following
result, attributed to Sasaki [Bla02]:
Proposition 6.1.8: Let M 2n+1 be a contact manifold which is not strict, then
its 2-fold cover is a strict contact manifold. In particular, every simply connected
contact manifold is a strict contact manifold.
A contact structure also gives rise to a codimension one subbundle of T M.
On Ui we define Di = kerφ∗i α, and on Ui ∩ Uj equation (6.1.5) holds implying that
Di = Dj . So we get a well defined subbundle D of T M, called the contact subbundle,
or contact distribution which is ‘as far from being integrable as possible’. This
terminology stems from the definition of integrability in the sense of the classical
Frobenius Theorem. Let D be a codimension one subbundle of T M so there is an
exact sequence of vector bundles

(6.1.6) 0−−−→D−−−→T M −−−→Q−−−→0 .


6.1. CONTACT STRUCTURES 183

Dualizing we see that Q∗ is just the contact line bundle L, so we have the exact
sequence
0−−−→L−−−→T ∗ M −−−→D∗ −−−→0 .
Now D is integrable if and only if there are local sections αi of L that satisfy
αi ∧ dαi = 0 for all i. Hence, we say that a subbundle D of T M defined as the
kernel of the local 1-forms αi is as far from being integrable as possible if the local
1-forms satisfy αi ∧ (dαi )n 6= 0. Thus, the contact subbundle actually characterizes
the contact structure, that is, we have
Proposition 6.1.9: Let M be a (2n + 1)-dimensional manifold. Then a codimen-
sion one subbundle D of T M defines a contact structure on M if and only if D is
as far from being integrable as possible.
Definition 6.1.10: If the vector bundle D is oriented we say that the contact
manifold M is co-oriented.
The following is a result of Stong [Sto74] which corrects an error in Proposition
2.2.1 and Theorem 2.2.2 of [Gra59]. This error was propagated in [Kob72].
Proposition 6.1.11: Let M be (2n+1)-dimensional contact manifold with contact
bundle D. Then
(i) If n is odd, then M is oriented.
(ii) If n is even, then M is co-oriented. Thus, in this case M has a strict
contact structure if and only if M is oriented.
Proof. There is a cover {Ui } of M together with 1-forms αi on Ui defining
the contact structure. In the overlap Ui ∩ Uj 6= ∅ there are non-vanishing functions
fi , fj such that fi αi = fj αj . Then on Ui ∩ Uj 6= ∅ we have
µ ¶n µ ¶n+1
n fj n n fj
(dαi ) |D = (dαj ) |D , αi ∧ (dαi ) = αj ∧ (dαj )n .
fi fi
Thus, if n is odd, we get a well-defined orientation; whereas, if n is even we get an
orientation on D. In the latter case M is orientable if and only if we can choose the
fi all positive, that is, if and only if the contact line bundle is trivial. ¤
Corollary 6.1.12: Let M be a strict contact manifold. Then M is both oriented
and co-oriented.
Remark 6.1.2: Recall that if (M 2n , J) is a complex manifold and M̄ denotes the
same smooth manifold but with reversed orientation than it is a non-trivial problem
to determine if M̄ admits any complex structure J¯ compatible with that orientation.
The simplest example is CP2 which is not a complex manifold. For surfaces the
complete answer can be found in [Kot97]. One can ask a similar question about
any strict contact manifold (M 2n+1 , η). That is, if we reverse the orientation on
M fixed by η, is there a compatible contact form η̄ on M̄ ? The answer is yes when
n = 1 due to Maritnet’s theorem (cf. Theorem 6.1.22). When n = 2 one can
always take −η. Hence, the first dimension when this question becomes of interest
is dimension 7. Unfortunately, very little is known in general about existence and
non-existence of contact structures in higher dimensions. In particular, other than
orientability there are very few known obstructions (cf. Theorem 6.2.7).
In this book almost all contact structures that we discuss are both oriented and
co-oriented. Notice that if (M, η) is a co-oriented contact manifold then (D, dη)
defines a symplectic subbundle of T M.
184 6. ALMOST CONTACT AND CONTACT GEOMETRY

Contact structures have been defined in the complex setting by Kobayashi


[Kob59b], and will play and important role in Chapters 12 and 13. Complex
contact transformations can be defined in analogy with Definition 6.1.2 by simply
replacing R2n+1 with C2n+1 . Then the coordinates and 1-form α of equation (6.1.1)
are holomorphic. One then defines the pseudogroup Γcon C of biholomorphic maps
from open sets C2n+1 to open sets of C2n+1 leaving α invariant. Then in complete
analogy with Definition 6.1.6 we have
Definition 6.1.13: A complex manifold M with a ΓCon C -structure is called a com-
plex contact manifold.
Such a manifold must have odd complex dimension, and in analogy with the real
case a complex contact structure is given by locally defined holomorphic 1-forms
(Ui , γi ) satisfying
(i) γi ∧ (dγi )n 6= 0 in Ui for each i,
(ii) if Ui ∩Uj 6= ∅ then there exists a nowhere vanishing holomorphic function
fij on Ui ∩ Uj which satisfies γi = fij γj for all distinct i, j.
The locally defined 1-forms patch together to define a holomorphic line bundle L,
so that the γi are 1-forms with values in L. Then the equation
n+1
γi ∧ (dγi )n = fij γj ∧ (dγj )n
1

gives a trivialization of the bundle KM ⊗ Ln+1 , hence, an isomorphism L ≈ KMn+1 .
This entire discussion works equally well for orbifolds. We only need to take the
open sets Ui to be the local uniformizing neighborhoods and the 1-forms γi and
transition functions fij to be invariant under the local uniformizing groups. Of
orb
course, KM must be replaced by the canonical orbibundle KX . Summarizing we
have
Proposition 6.1.14: Let (X , L) be a complex contact orbifold of dimension 2n+1.
1
orb − n+1
Then L ≈ (KX ) . In particular, the index IX is divisible by n+1. Conversely, if
X = (X, U) orbifold with a holomorphic line orbibundle L satisfying Ln+1 ⊗ KX orb

1l, then (X , L) is a complex contact orbifold.

6.1.3. Examples. Our first example is the most obvious.


Example 6.1.15: M = R2n+1 . The global 1-form given by equation (6.1.1) defines
a strict contact structure on R2n+1 . The contact subbundle D is spanned by the
vector fields ∂xi + yi ∂z and ∂yi . This is called the standard contact structure
on R2n+1 .
The next example will be fundamental in the sequel.
Example 6.1.16: M = S 2n+1 , the unit (2n + 1)-sphere. Consider the 1-form
2n+2
on
PnR defined in the standard Cartesian coordinates (x0 , . . . , xn , y0 , . . . , yn ) by
(x
i=0 i dy i − yi dxi ). Let η denote its restriction to the unit sphere. We claim that
η ∧ (dη)n 6= 0 everywhere on S 2n+1 . A straightforward computation shows that
³X
n
η ∧ (dη)n = 2n n! ci ∧ dyi ∧ · · · ∧ dxn ∧ dyn
xi dx0 ∧ dy0 ∧ · · · ∧ dx
i=0

n
X ´¯
− ci ∧ · · · ∧ dxn ∧ dyn ¯¯
yi dx0 ∧ dy0 ∧ · · · ∧ dxi ∧ dy ,
S 2n+1
i=0
6.1. CONTACT STRUCTURES 185

c means the dx is omitted. Now the 1-form β = Pn (xi dxi +yi dyi )
where as usual dx i=0
is normal to the sphere S 2n+1 and an easy computation gives
n
X ¡ ¢
β ∧ η ∧ (dη)n = 2n n! (xi )2 + (yi )2 dx0 ∧ dy0 ∧ · · · ∧ dxn ∧ dyn .
i=0

This shows that η ∧ (dη) vanishes nowhere on S 2n+1 . This defines the standard
n

contact structure on S 2n+1 . Notice also that η is invariant under the reflection
(xi , yi ) 7→ (−xi , −yi ) for all i = 0, . . . , n. Thus, we get an induced contact structure
on the real projective space RP2n+1 as well.
There is a straightforward generalization of the example above due to Gray
[Gra59] (see also [Bla02, YK84]) that shows that any immersed hypersurface in
R2n+2 such that no point of the tangent space to the hypersurface contains the
origin 0 ∈ R2n+2 admits a strict contact structure. Here is the construction:
Proposition 6.1.17: Let M be an immersed hypersurface in R2n+2 such that no
tangent space of M contains the origin of R2n+2 . Then M has a strict contact
structure.
Proof. Again consider the 1-form in Cartesian coordinates (x0 , . . . , x2n+1 ) of
R2n+2 defined by
Xn
α= (xi dxn+1+i − xn+1+i dxi ) .
i=0
Choose a point x0 ∈ M together with a basis (v1 , . . . , v2n+1 ) of Tx0 M. Consider
the vector w whose components are wi = (?dxi )(v1 , . . . , v2n+1 ), where ? denotes
the Hodge star operator with respect to the Euclidean metric on R2n+2 . Now w is
perpendicular to Tx0 M viewed as a subspace of R2n+2 . Regarding x0 as a vector,
a simple computation gives
α ∧ (dα)n (v1 , . . . , v2n+1 ) = x0 · w .
Thus, the right hand side of this equation vanishes if and only if the tangent space
Tx0 M contains the origin. So under the hypothesis of the proposition, the restriction
of α to M defines a strict contact structure on M. ¤
Another variation on the same theme, again due to Gray, uses the cotangent
bundle T ∗ N of a manifold N. On T ∗ N there is a tautological or canonical 1-form
γ defined as follows: if v ∈ Tα T ∗ N at the covector α ∈ T ∗ N, then γ is defined by
(6.1.7) γ(v) = hα, π∗ vi ,
where π : T ∗ N −−→N is the natural projection and h·, ·i denotes the natural pairing
between Tx N and Tx∗ N. It is not difficult to show that (dγ)n+1 6= 0. Thus, the
cotangent bundle of any manifold has a natural symplectic structure.
Proposition 6.1.18: Let M be an immersed submanifold of T ∗ N with its canonical
1-form γ. Then the restriction of γ to M defines a contact structure on M if and
only if the following conditions hold:
(i) dim (M ∩ π −1 (x)) = n for all x ∈ N.
(ii) M ∩ N = ∅, where N is identified with the zero section of T ∗ N.
(iii) No tangent space to M ∩ π −1 (x) in π −1 (x) contains the origin in π −1 (x).
Furthermore, if the conditions are satisfied the contact structure is strict.
186 6. ALMOST CONTACT AND CONTACT GEOMETRY

Proof. The proof is similar to that of Proposition 6.1.17. If (x0 , . . . , xn ) are


local coordinates on an open set U ⊂ N, and (y0 , . . . , yn ) are coordinates in the
y0 , . . . , yn ) are coordinates in π −1 (U ). In these coordinates
fibres, then (x0 , . . . , xn , P
n
γ takes the form γ = i=0 yi dxi . Clearly, γ vanishes along the zero section of

T N , so condition (ii) is necessary. As in the previous case let (v1 , . . . , v2n+1 ) be a
local frame for M ∩ π −1 (U ). Define a tangent vector w to the fibres π −1 (x) whose
components with respect to the coordinate frame in π −1 (x) are
ci ∧ dxi ∧ · · · ∧ dyn ∧ dxn (v1 , . . . , v2n+1 ) ,
wi = dy0 ∧ dx0 ∧ · · · ∧ dy
where only the dy’s are removed. Then a simple computation gives
n
X
(6.1.8) γ ∧ (dγ)n (v1 , . . . , v2n+1 ) = yi wi .
i=0

Now the proposition follows as before. ¤

As an application of this proposition we have immediately


Example 6.1.19: The unit sphere bundle T1∗ N of a manifold N with respect to
some Riemannian metric. One easily checks that T1∗ N satisfies the three conditions
of Proposition 6.1.18. Thus, T1∗ N has a (strict) contact structure.
Proposition 6.1.18 suggests the following
ι
Definition 6.1.20: Let (N, ω) be a symplectic manifold. A hypersurface M ,→ N
is said to be of contact type if there is a 1-form α on M such that ι∗ ω = dα.
Next is an example of a contact structure due to Gray [Gra59] which shows
(ironically) that (ii) of Proposition 2.2.1 as well as Theorem 2.2.2 of [Gra59] and
Theorem 7.3 of Chapter I of [Kob72] are incorrect. This was first noticed in
[Sto74].
Example 6.1.21: M = Rn+1 × RPn . Let xi for i = 0, . . . , n denote the Cartesian
coordinates on Rn+1 and ti for i = 0, . . . , n denote the homogeneous coordinates
on the real projective space RPn . Let Ui ⊂ Rn+1 × RPn be the affine neighborhood
defined by ti 6= 0. Clearly {Ui }ni=0 cover M. We define the contact structure by a
sequence of 1-forms αi defined in Ui by
X tj X tj
αi = dxi + dxj = dxj .
ti j
ti
j6=i

In the overlap Ui ∩ Uj we see that


ti
αj = αi ,
tj
and this defines the contact line bundle L which is non-trivial since it is induced
by the tautological line bundle on RPn . Hence, there is no globally defined contact
1-form on M which defines the contact structure, that is, the contact structure is
not strict. An easy calculation also shows that
µ ¶n+1
ti
αj ∧ (dαj )n = αi ∧ (dαi )n ,
tj
so M is orientable if and only if n is odd, and in this case M is not co-orientable.
Stong [Sto74] noticed the following generalization of this example:
6.1. CONTACT STRUCTURES 187

Exercise 6.4: Consider the projectivized cotangent bundle π : M = P(T ∗ N )−−→N,


where N has dimension n+1. Let U be the open set of the chart defined by the local
coordinates (x0 , . . . , xn ), and let (y0 , . . . , yn ) denote the homogeneous coordinates
in the fibres of π −1 (U ) ⊂ T ∗ N. Let Ui ⊂ M be theP open set of π −1 (U ) defined by
−1 n
setting yi 6= 0. On Ui define the 1-form αi = (yi ) j=0 yj dxj . Show that the αi ’s
patch together to define a contact structure on M.
For further examples see Blair’s recent book [Bla02]. Of particular interest is
the fundamental theorem of Martinet in dimension three [Mar71].
Theorem 6.1.22: Every orientable 3-manifold admits a contact structure.
Following Blair [Bla02] we give an explicit contact structure on the 3-torus T 3 .
Example 6.1.23: M = T 3 = R3 /Z3 . Let (x, y, z) denote standard Cartesian coor-
dinates in R3 , and consider the 1-form
η = sin ydx + cos ydz .
Then we find η ∧ dη = −dx ∧ dy ∧ dz. The contact subbundle D is spanned by the
vector fields
n∂ ∂ ∂ o
, cos y − sin y .
∂y ∂x ∂z
In particular, Blair discusses the so-called overtwisted contact structures in
dimension 3 of Eliashberg. These seem to have no analogues in higher dimensions.
Lutz [Lut79] proved the existence of contact structures on principal T 2 -bundles
over 3-manifolds. In particular, the 5-torus T 5 admits a contact structure. It has
recently been proven that every odd dimensional torus admits a contact structure
[Bou02].
Exercise 6.5: Let (x1 , . . . , x5 ) denote the standard coordinates on T 5 = R5 /Z5 .
Show that the 1-form
η = (sin x1 cos x3 − sin x2 sin x3 )dx4 + (sin x1 sin x3 + sin x2 cos x3 )dx5
+ sin x2 cos x2 dx1 − sin x1 cos x1 dx2 + cos x1 cos x2 dx3
defines a contact structure on T 5 .
A strict contact structure has an associated canonical vector field introduced
by Reeb [Ree52] which is of great importance in the sequel.
Lemma 6.1.24: On a strict contact manifold (M, η) there is a unique vector field
ξ, called the Reeb vector field, satisfying the two conditions
ξ η = 1, ξ dη = 0 .
Proof. The volume element η ∧ (dη)n gives an isomorphism between the
C (M )-module of vector fields on M and the C ∞ (M )-module of 2n-forms on

M. Thus, there is a unique vector field ξ that satisfies the equation ξ (η ∧ (dη)n ) =
(dη)n . Contracting again with respect to ξ gives ξ (dη)n = 0, and expanding gives
(dη)n = η(ξ)(dη)n − η ∧ (ξ (dη)n ) = η(ξ)(dη)n which implies η(ξ) = 1. But then
since dη is a 2-form 0 = ξ (dη)n = n(ξ dη) ∧ (dη)n−1 . But since the rank of the
2-form dη is n, this implies that ξ dη = 0. ¤
The Reeb vector field is often called the characteristic vector field and it
uniquely determines a 1-dimensional foliation Fξ on (M, η) called the character-
istic foliation. We let Lξ denote the trivial line bundle consisting of tangent vectors
188 6. ALMOST CONTACT AND CONTACT GEOMETRY

that are tangent to the leaves of Fξ . This splits the exact sequence 6.1.6 as
(6.1.9) T M = D ⊕ Lξ .
Definition 6.1.25: The characteristic foliation Fξ of a strict contact structure is
said to be quasi-regular if there is a positive integer k such that each point has a
foliated coordinate chart (U, x) such that each leaf of Fξ passes through U at most k
times. If k = 1 then the foliation is called regular. We use the terminology non-
regular to mean quasi-regular, but not regular; whereas, if Fξ is not quasi-regular,
it is said to be irregular.
The notion of quasi-regularity in Definition 6.1.25 is due to Thomas [Tho76]
who used the terminology almost regular instead of quasi-regular. He then proved
that the flow of the Reeb vector field for any almost regular contact structure is
equivalent to a locally free circle action. However, this follows from the earlier
theorem of Wadsley [Wad75] 2.6.12.
Remark 6.1.3: We will also say that the Reeb vector field ξ is regular (quasi-
regular, non-regular, irregular). Note that since the Reeb vector field is uniquely
determined by the contact form, it is perfectly acceptable to talk about a regular
(quasi-regular, non-regular, irregular) contact form η on M as well. However, in a
given contact structure different choices of the 1-form η produce different character-
istic foliations which can have drastically different behavior. We shall see explicit
examples of this in Chapter 7 (cf. Example 7.1.12). Hence, it makes no sense to
say, for example, that a contact structure on M is regular. Nevertheless, we can
ask
Question 6.1.1: Given a contact manifold (M, D) does there exist a regular 1-form
η in the contact structure D?
When the answer is yes, by abuse of language, we will sometimes say that a contact
manifold M is regular, meaning that M is a strict contact manifold (M, η) which
has been quipped with a regular contact from η.
In 1958 Boothby and Wang [BW58] (cf. Theorem 8.3.5) studied homogeneous
contact manifolds which they proved must be regular in the sense of Question 6.1.1.
This will be discussed in Chapter 8. In particular, they show that a regular strict
compact contact manifold has a canonical structure of a principal circle bundle over
a compact symplectic manifold.
Theorem 6.1.26: Let (M, η) be a regular compact strict contact manifold. Then
M is the total space of a principal circle bundle π : M −→Z over the space of leaves
Z = M/Fξ . Furthermore, Z is a compact symplectic manifold with symplectic
form Ω, [Ω] ∈ H 2 (Z, Z), and η is a connection form on the bundle with curvature
dη = π ∗ Ω.
Remark 6.1.4: This theorem does, in a sense, serve as the backbone of the material
of the next several chapters. It is quite clear the it can easily be extended to the
quasi-regular case where instead of circle bundle and a smooth symplectic manifold
we can talk about orbibundles and symplectic orbifolds. We will discuss and prove
various generalizations of this result in this and the next chapter (cf. Theorem 6.3.8
and Theorem 7.1.3).
Let (M, η) be a strict contact manifold. An immersed submanifold ι : N −−→M
is called an integral submanifold of the contact structure if ι∗ η = 0. An integral
submanifold N is said to be maximal if it is not contained as a proper subset of any
6.1. CONTACT STRUCTURES 189

other integral submanifold. Maximal integral submanifolds of a contact manifold


are often called Legendrian submanifolds.
Proposition 6.1.27: On a contact manifold (M, η) of dimension 2n+1 the Legren-
drian submanifolds have dimension n.
Proof. It is easy to see that integral submanifolds of dimension n exist. By
Darboux’s theorem there is a coordinate chart (U ; x1 , . . . , xn , y1 , . . . , yn , z) such
that η takes the form of equation (6.1.1). Then define the submanifold locally by
xi = ai , z = c for some constants ai , c. This shows that submanifolds of dimension
n exist. On the other hand suppose that N is an integral submanifold of dimension
k, and let X and Y be vector fields that are tangent to N. Then the equation
g(ΦX, Y ) = dη(X, Y ) = 0
says that ΦX is normal to N. Since ξ is clearly normal to N there can be at most
n linearly independent local vector fields tangent to N. ¤
6.1.4. Global Contact Transformations. We begin by describing some el-
ementary properties of the group of contact transformations. A good book for the
further study of this important group is that of Banyaga [Ban97].
Definition 6.1.28: Let M be a strict contact manifold, and let Con(M, D) denote
the group of contact transformations, i.e., the subgroup of the group Diff(M ) of
diffeomorphisms of M that leaves the contact distribution D invariant. Alternatively
fixing a contact form η such that D = ker η, then Con(M, D) can be characterized
as
Con(M, D) = {φ ∈ Diff(M ) | φ∗ η = f η for f ∈ C ∞ (M ) nowhere vanishing} .
With the 1-form η fixed we are also interested in the subgroup Con(M, η) of strict
contact transformations defined by the condition φ∗ η = η.
The uniqueness of the Reeb vector field implies
Lemma 6.1.29: If φ ∈ Con(M, η) then φ∗ ξ = ξ.
We denote the “Lie algebras”1 of Con(M, D) and Con(M, η) by con(M, D) and
con(M, η), respectively. They can be characterized as follows:
con(M, D) = {X ∈ Γ(M ) | £X η = gη for some g ∈ C ∞ (M )} ,

(6.1.10) con(M, η) = {X ∈ Γ(M ) | £X η = 0} ,


where Γ(M ) denotes the Lie algebra of smooth vector fields on M. Clearly, con(M, η)
is a Lie subalgebra of con(M, D). The infinitesimal analog of Lemma 6.1.29 is
Lemma 6.1.30: If X ∈ con(M, η) then [X, ξ] = 0, so ξ lies in the center of
con(M, η). In particular, every vector field in con(M, η) is foliate with respect to the
characteristic foliation Fξ , i.e., con(M, η) is a subalgebra of fol(M, Fξ ).
In contrast to the symplectic case, every infinitesimal contact transformation
is Hamiltonian. More explicitly,
Proposition 6.1.31: The contact 1-form η induces a Lie algebra isomorphism
Φ between the Lie algebra con(M, D) of infinitesimal contact transformations and

1The quotes occur since such Lie algebras are actually associated with the corresponding
pseudogroups, not groups of global transformations.
190 6. ALMOST CONTACT AND CONTACT GEOMETRY

the Lie algebra C ∞ (M ) of smooth functions with the Jacobi bracket {·, ·}. The
isomorphism is defined by
Φ(X) = η(X).
Furthermore, under this isomorphism the subalgebra con(M, η) is isomorphic to
the subalgebra C ∞ (M )ξ of functions in C ∞ (M ) that are invariant under the flow
generated by the Reeb vector field ξ, that is the subalgebra of functions that are basic
with respect to the characteristic foliation Fξ .
Recall [LM87] that the Jacobi bracket is defined by {f, g} = η([Xf , Xg ]). The
function η(X) is known as the contact Hamiltonian function of X. Conversely, for
any smooth function f the vector field Xf = Φ−1 (f ) is an infinitesimal contact
transformation. In particular, ξ = Φ−1 (1).
6.1.5. Deformation Theory of Contact Structures. The deformation the-
ory of contact structures was done by Gray [Gra59], the main result being Gray’s
Stability Theorem which essentially says that there are no local invariants in contact
geometry. It is the analog of Moser’s Stability Theorem of symplectic geometry,
cf. [MS98, Ban97]. Gray’s idea was to apply the Kodaira-Spencer deformation
theory to contact geometry; however, it is possible to give a proof that is similar to
the proof of Moser’s Theorem [MS98]. For us it suffices here to give a statement of
the result. Recall that an isotopy between two diffeomorphisms φ0 , φ1 ∈ Diff(M )
is a smooth map φ : M × [0, 1]−−→M such that
(i) φ(x, 0) = φ0 (x),
(ii) φ(x, 1) = φ1 (x),
(iii) φt (x) = φ(x, t) ∈ Diff(M ) for all t ∈ [0, 1].
Here is Gray’s Stability Theorem:
Theorem 6.1.32: For t ∈ [0, 1] let ηt be a smooth family of contact forms on
a compact manifold M. Then there exists an isotopy φt and a family of smooth
nowhere vanishing functions ft such that φ∗t ηt = ft η0 for each t ∈ [0, 1].
Let Dt = ker ηt with D0 = D. Then this theorem says that every contact
structure in the smooth family Dt of contact structures is contactomorphic to D,
so there are no local invariants for contact structures.

6.2. Almost Contact Structures


6.2.1. Definitions and Topological Properties. Arguably the most effi-
cient way to define almost structures is through G-structures which we shall employ.
Our first definition of an almost contact structure is given below in 6.2.2 is com-
pletely non-standard, but at least it is terminologically consistent. As mentioned
previously the standard definition of almost contact structures [Bla76a, Bla02,
YK84] is too restrictive. For example, the contact manifold M = Rn+1 × RPn of
Example 6.1.21 is not an almost contact manifold in the usual sense. Recall the
almost contact G-structure of Example 1.4.15, then we have
Theorem 6.2.1: A contact structure on a manifold M is a non-integrable almost
contact G-structure.
Proof. Let us look at the contact pseudogroup ΓCon at the level of the tangent
bundle. Let φ ∈ ΓCon , then its differential φ∗ leaves the contact subbundle D
invariant. Choose a local 1-form αi representing the contact structure. Then dαi |D
defines a symplectic structure on D over Ui . Fix a point x ∈ Ui , and let ΓCon
x denote
6.2. ALMOST CONTACT STRUCTURES 191

the subpseudogroup generated by the vector fields in con(Ui ) that vanish at x. The
differentials of the elements of ΓCon
x span the linear isotropy algebra of con which
determines the G-structure. Let φ ∈ ΓCon ∗
x , then φ leaves the symplectic form dαi |D
invariant up to a multiplicative factor. So the linear isotropy group is precisely the
group G of Example 1.4.15. Furthermore, since αi and the symplectic form dαi |D
satisfy equation (6.1.2), this G-structure is not integrable. ¤
Hence, we choose to define an almost contact structure in terms of G-structures
as follows:
Definition 6.2.2: An almost contact structure on M is a reduction of the
frame bundle to the group G defined in Example 1.4.15, that is, an almost contact
G-structure.
Thus, an almost contact structure picks out a codimension one subbundle D of
T M with a conformal symplectic structure on it. By reducing to maximal compact
subgroups we have
Theorem 6.2.3: A manifold M of dimension 2n + 1 admits an almost contact
structure if and only if its frame bundle can be reduced to the group U (n) × Z2 × Z2 ,
where the first Z2 corresponds to the sign of the conformal factor in CSp(n, R)
and the second Z2 corresponds to the sign of the determinant. Moreover, if M is
orientable, there is a reduction to the group U (n) × Z2 , which further reduces to the
group U (n) × 1 if and only if the subbundle D of T M is orientable. Furthermore, if
M is non-orientable, but D is orientable, then there is a reduction to the subgroup
U (n) × Z2 .
Proof. The group CSp(n, R) is homotopy equivalent to Z2 × Sp(n, R), where
the group Z2 is identified with the sign of the conformal factor, and Sp(n, R) is
homotopy equivalent to its maximal compact subgroup U (n). So the group G of
Example 1.4.15 is homotopy equivalent to the block matrices of the form
nµC 0¶ o
G= | C ∈ H, b ∈ Z2 ,
0 b
where H ≈ U (n) × Z2 is realized as the subgroup of O(2n, R) generated by matrices
of the form  
−1 0 · · · 0
µ ¶ 0 1 · · · 0
A B  
,  .. .. ,
−B A  . ··· . 0
0 ··· 0 1
where A and B are n by n block matrices that satisfy
Āt A + B̄ t B = 1ln , (Āt B)t = Āt B .
If M is orientable we can take det M = b, so the group reduces to U (n) × Z2 .
Moreover, the subbundle D is orientable if and only if we can take det C = 1, in
which case the group reduces to U (n) × 1. ¤
Combining Theorem 6.2.3 and Proposition 6.1.11 we have
Corollary 6.2.4: Let M be a 2n + 1-dimensional contact manifold, then the frame
bundle can be reduced to the subgroup U (n) × Z2 .
(i) If n is odd, then M can be reduced to the group U (n) × 1 if and only if
M is co-orientable.
192 6. ALMOST CONTACT AND CONTACT GEOMETRY

(ii) If n is even, then M can be reduced to the group U (n) × 1 if and only if
M is orientable.
(iii) M has a strict contact structure if and only if M can be reduced to the
group U (n) × 1.
We shall now discuss the standard definition of an almost contact structure
[Bla76a, Bla02, YK84], which is more closely related to strict contact structures
than contact structures. However, even in this regard, the definition includes ad-
ditional structure, namely a choice of vector field. Also from the point of view of
G-structures, it would be more consistent to define almost contact structures using
a ‘transverse symplectic form’ instead of a ‘transverse almost complex structure’.
Of course, in the presence of a metric the two are equivalent, so we shall stick with
what is commonly done. At least for the purposes of this section we shall often add
the word ‘strict’ to the definition for emphasis.
Definition 6.2.5: A (strict) almost contact structure on a differentiable man-
ifolds M is a triple (ξ, η, Φ), where Φ is a tensor field of type (1, 1) (i.e., an endo-
morphism of T M ), ξ is a vector field, and η is a 1-form which satisfy
η(ξ) = 1 and Φ ◦ Φ = −1l + ξ ⊗ η ,
where 1l is the identity endomorphism on T M. A smooth manifold with such a
structure is called an almost contact manifold.
Such an almost contact structure (ξ, η, Φ) has three other almost contact struc-
tures canonically associated to it, namely (ξ, η, −Φ), (−ξ, −η, Φ), and (−ξ, −η, −Φ).
Notice that an almost contact manifold is necessarily odd dimensional, and one eas-
ily verifies the identities [Bla76a, Bla02, YK84]
(6.2.1) Φ ◦ ξ = 0, η ◦ Φ = 0.
Thus, an almost contact manifold M comes canonically equipped with the following
structures: Since the vector field ξ is nowhere vanishing, it generates a 1-dimensional
subbundle Lξ of the tangent bundle T M ; hence, an almost contact manifold M has
associated to it a 1-dimensional foliation Fξ , called the characteristic foliation. The
1-form η is called the characteristic 1-form and it plays the role of the characteristic
form χ in foliation theory [Ton97], even though χ requires a Riemannian metric for
its definition. The codimension one subbundle D = ker η has an almost complex
structure defined by J = Φ|D . Thus, as in the case of a contact structure there is a
canonical splitting of the tangent bundle
(6.2.2) T M ' D ⊕ Lξ ,
and from the point of view of G-structures, a choice of a strict almost contact
structure on a 2n + 1 dimensional manifold M corresponds to a reduction of the
frame bundle to the subgroup GL(n, C) × 1 of GL(2n + 1, R). Thus, in particular,
an almost contact structure is naturally an almost product structure of type (2n, 1)
with a transverse almost complex structure. Hence, the subbundle D, which we
refer to as the horizontal subbundle of T M, together with the endomorphism J
provides M with an almost CR structure (D, J) of codimension one. Combining
this discussion with Theorem 1.2.5 we have arrived at
Theorem 6.2.6: Let M be a smooth manifold of dimension 2n + 1. There is a
one-to-one correspondence between the strict almost contact structures (ξ, η, Φ) on
M and the reductions of the frame bundle L(M ) to the group GL(n, C) × 1.
6.2. ALMOST CONTACT STRUCTURES 193

Notice that a choice of strict almost contact structure (ξ, η, Φ) on M chooses


both an orientation and a transverse orientation on (M, Fξ ). The almost contact
structure (−ξ, −η, Φ) chooses the opposite orientation, and (ξ, η, −Φ) chooses the
opposite orientation on D if and only if n is odd.
Corollary 6.2.4 together with standard topological obstruction theory give nec-
essary conditions for the existence of a strict almost contact structure. See Appen-
dix A for a brief discussion of characteristic classes and [MS74, LM89] for more
details. The characteristic classes on M are obtained by pulling back the universal
characteristic classes on the classifying space BU (n) by the classifying map. But it
is well-known that BU (n) has vanishing odd Stiefel-Whitney classes w2i+1 as well
as vanishing integral Stiefel-Whitney classes Wi . The former implies the vanishing
of all Stiefel-Whitney numbers, and by a famous theorem of Thom that M is the
boundary of a compact manifold. Summarizing, we have
Theorem 6.2.7: Let (M, ξ, η, Φ) be a strict almost contact manifold of dimension
2n + 1, then the odd Stiefel-Whitney classes w2i+1 (M ) vanish and all the integral
Stiefel-Whitney classes Wi (M ) vanish. Thus, all Stiefel-Whitney numbers vanish,
and M is the boundary of a compact manifold.
Actually, this result holds a bit more generally for all three cases of Corollary
6.2.4. Theorem 6.2.7 is essentially due to Gray [Gra59] who also proved that the
vanishing of W3 (M ) is sufficient in dimension 5.
Corollary 6.2.8: [Gra59] A 5-dimensional orientable manifold M admits an al-
most contact structure (ξ, η, Φ) if and only if the integral Stiefel-Whitney class
W3 (M ) vanishes.

6.2.2. Equivalence and Automorphisms. Two strict almost contact man-


ifolds (M, ξ, η, Φ) and (M 0 , ξ 0 , η 0 , Φ0 ) are isomorphic if there is a diffeomorphism
f : M −−→M 0 such that

(6.2.3) ξ 0 = f∗ ξ, η 0 = (f −1 )∗ η , Φ0 ◦ f∗ = f∗ Φ .

When (M, ξ, η, Φ) is a contact manifold such an isomorphism induces an isomor-


phism of the corresponding symplectic vector bundles (D, dη) and (D0 , dη 0 ). We
shall also be interested in certain automorphism groups. For any vector subbun-
dle E ⊂ T M we let Aut(E) denote the subgroup of diffeomorphisms that leave E
invariant, that is,

(6.2.4) Aut(E) = {φ ∈ Diff(M ) | φ∗ E ⊂ E} .

In the case that E = ker η for some 1-form η, this is equivalent to φ∗ η = f η for some
nowhere vanishing function f. Thus, when η is a contact form Aut(D) = Con(M, D).
Similarly, if E = Lξ the line bundle generated by a nowhere vanishing vector
field ξ, then Aut(Lξ ) = Fol(Fξ ). We also have subgroups Aut(η) ⊂ Aut(D) and
Aut(ξ) ⊂ Aut(Fξ ) defined by

(6.2.5) Aut(η) = {φ ∈ Diff(M ) | φ∗ η = η}

and

(6.2.6) Aut(ξ) = {φ ∈ Diff(M ) | φ∗ ξ = ξ} ,


194 6. ALMOST CONTACT AND CONTACT GEOMETRY

respectively. These groups are all infinite dimensional and have corresponding ‘Lie
algebras’ defined by
(6.2.7) aut(D) = {vector fields X | £X η = f η for some f ∈ C ∞ (M )} ,
(6.2.8) aut(η) = {vector fields X | £X η = 0} ,
(6.2.9) aut(ξ) = {vector fields X | [X, ξ] = 0} .
Notice that when ξ generates a foliation Fξ on M the Lie algebra aut(ξ) is a Lie
subalgebra of the Lie algebra fol(M, Fξ ) of foliate vector fields.
Let us now consider the case of an almost contact structure (ξ, η, Φ). In this
case we define the group leaving the tensor field Φ invariant by
(6.2.10) Aut(M, Φ) = {φ ∈ Diff(M ) | φ∗ Φ = Φφ∗ } .
Its ‘Lie algebra’ of infinitesimal transformations is given by
(6.2.11) aut(M, Φ) = {vector fields X | £X Φ = 0} .
It follows from Equations 6.2.1 that
(6.2.12) Aut(M, Φ) ⊂ Aut(D) ∩ Fol(M, Fξ ) ,
and similarly for their Lie algebras. It is also convenient to define the group of
CR-transformations of an almost contact manifold (M, ξ, η, Φ) by
(6.2.13) CR(M, D, J) = {φ ∈ Aut(D) | φ∗ J = Jφ∗ } .
In the case that the almost CR structure is integrable and M is compact, this group
was shown to be a Lie group in [CM74]. Its Lie algebra cr(M, D, J) is defined by
(6.2.14) cr(M, D, J) = {X ∈ aut(D) | £X J = 0} .
Notice that £X J makes sense even though J is not a tensor field on M . This
follows from the fact that the vector field X leaves D invariant. In [BGS06] the
following lemma is proved under a more restrictive hypothesis, but it is easy to see
that the proof given there holds in the present case.
Lemma 6.2.9: Let (M, ξ, η, Φ) be an almost contact manifold with underlying al-
most CR structure (D, J) with D = ker η and J = Φ|D . Then
Aut(M, Φ) = CR(M, D, J) ∩ Fol(M, Fξ ) .
We now define the automorphism group Aut(ξ, η, Φ) of an almost contact struc-
ture as
(6.2.15) Aut(ξ, η, Φ) = {f ∈ Diff M | f∗ ξ = ξ, f ∗ η = η, f∗ Φ = Φf∗ } .
Clearly, we have Aut(ξ, η, Φ) = Aut(η) ∩ Aut(ξ) ∩ Aut(M, Φ). We also define the ‘Lie
algebra’ of Aut(ξ, η, Φ) by
(6.2.16) aut(ξ, η, Φ) = {vector fields X | £X η = £X Φ = [X, ξ] = 0} .
However, unlike the strict contact case the vector field ξ does not necessarily lie in
aut(ξ, η, Φ). See for example Lemmas 6.3.3 and 6.3.4 below.
Notice that in the case that η defines a contact structure, Lemma 6.1.29 implies
that the group Aut(ξ, η, Φ) can be characterized as
(6.2.17) Aut(ξ, η, Φ) = {f ∈ Con(M, η) | f∗ Φ = Φf∗ } = Con(M, η) ∩ Aut(M, Φ) .
6.3. ALMOST CONTACT METRIC STRUCTURES 195

Exercise 6.6: Let (M, ξ, η, Φ) be an almost contact manifold. Show that the Lie
algebra cr(M, D, J) can be characterized as the set of all vector fields X on M such
that £X Φ = −Φ[X, ξ] ⊗ η.
The main theorem on automorphisms of strict almost contact structures is due
to Sasaki [Sas67].
Theorem 6.2.10: Let M be a compact manifold with a strict almost contact struc-
ture (ξ, η, Φ). Then the group of automorphisms Aut(ξ, η, Φ) is a Lie group.
Proof. By Theorem 6.2.6 the Lie algebra of the structure group is gl(n, C) ⊂
gl(2n, R), which is elliptic since it has no matrix of rank one. So the theorem follows
from Proposition 1.6.7. ¤

6.3. Almost Contact Metric Structures


In order to investigate further we need a compatible metric structure.
Definition 6.3.1: Let M be a (strict) almost contact manifold. A Riemannian
metric g on M is said to be compatible with the almost contact structure if for
any vector fields X, Y on M we have
g(ΦX, ΦY ) = g(X, Y ) − η(X)η(Y ) .
An almost contact structure with a compatible metric is called an almost contact
metric structure.
One immediately sees by setting Y = ξ in the equation above that g(ξ, X) =
η(X). It is easy to construct compatible Riemannian metrics on an almost contact
manifold. Compatible metrics are often called associated metrics [Bla02].
Exercise 6.7: Let g 0 be any Riemannian metric on a strict almost contact manifold
(M, ξ, η, Φ) and define
1¡ ¢
g = g 0 ◦ (Φ2 ⊗ Φ2 ) + g 0 ◦ (Φ ⊗ Φ) + η ⊗ η .
2
Show that g defines a Riemannian metric on M that is compatible with the almost
contact structure. Conversely, show that any compatible metric must take the form
g = gD + η ⊗ η, where gD is a metric on the vector bundle D.
Almost contact metric structures correspond to almost product structures of
type (2n, 1) with a transverse almost Hermitian structure. As with Theorem 6.2.6
we have
Proposition 6.3.2: There is a one-to-one correspondence between strict almost
contact metric structures (ξ, η, Φ, g) on M and reductions of the frame bundle to
the group U (n) × 1.
If M is a smooth manifold of dimension 2n + 1 with an almost contact metric
structure, then there is a natural 2-form Ψ on M of rank 2n defined by
(6.3.1) Ψ(X, Y ) = g(X, ΦY ) .
In this case it is easy to see that the 2n + 1 form η ∧ (Ψ)n 6= 0 everywhere on M,
and that Ψ is skew-symmetric. The following lemma is a reformulation of Lemma
2.6.8 in terms of almost contact manifolds.
Lemma 6.3.3: Let (M, ξ, η, Φ) be an almost contact manifold. Then the following
are equivalent:
196 6. ALMOST CONTACT AND CONTACT GEOMETRY

(i) There exists a compatible Riemannian metric on M such that the orbits
of ξ are geodesics.
(ii) £ξ η = 0.
(iii) ξ dη = 0.
We remark that for 1-dimensional foliations condition (i) in Lemma 6.3.3 is
equivalent to tautness [Ton97]. Thus, we can rephrase Lemma 6.3.3 as:
Lemma 6.3.4: Let (M, ξ, η, Φ) be an almost contact manifold. The associated
characteristic foliation Fξ is taut if and only the characteristic 1-form η is invariant
under the flow of ξ.
Next we wish to characterize those almost contact structures that have Rie-
mannian metrics that are ‘compatible’ with the characteristic foliation, that is that
the characteristic foliation be a Riemannian foliation 2.5.4. As shown in Proposition
2.5.7 this is equivalent to the condition that the compatible metric be bundle-like,
cf. 2.5.6. We show that this happens precisely when the transverse metric gD = g|D
is invariant under the flow of ξ, that is when £ξ gD = 0.
Proposition 6.3.5: Let (M, ξ, η, Φ) be an almost contact manifold and g be a
Riemannian metric that is compatible with the almost contact structure (ξ, η, Φ).
Then g is bundle-like and Fξ is taut if and only if ξ is a Killing field of g.

Proof. By Exercise 6.7 any compatible metric g takes the form


1¡ 0 ¢
g = gD ⊕ η ⊗ η, gD = g ◦ (Φ2 ⊗ Φ2 ) + g 0 ◦ (Φ ⊗ Φ) ,
2
where g 0 is an arbitrary Riemannian metric. The computation is local and there
is a local frame of foliate vector fields. Computing the Lie derivatives we find for
foliate horizontal vector fields X, Y
(£ξ g)(X, Y ) = (£ξ gD )(X, Y ) , (£ξ g)(X, ξ) = dη(ξ, X), (£ξ g)(ξ, ξ) = 0
from which the result easily follows by Lemma 6.3.3. ¤

It would be interesting to find examples of almost contact structures whose


characteristic foliation is not taut, but nevertheless admits a compatible bundle-
like metric. Next we give the orbifold version of a Theorem of Hatakeyama [Hat63]
who shows that the total space of a principal G1 bundle over an almost complex
manifold where G1 is a 1-dimensional Lie group admits the structure of an almost
contact manifold in a natural way. Now let (N, J) be an almost complex orbifold,
and π : M −−→N a principal G1 V-bundle over N. Generally, the total space M is
an orbifold with fibres diffeomorphic to G1 . There are two cases (we assume G1 is
connected): (1) G1 = R and (2) G1 = S 1 . But according to Definition 4.2.7 there
is a group homomorphism from the local uniformizing groups Γi to G1 . But for
case (1) G1 = R the only homomorphism from a finite group to the additive reals
is trivial. Hence, π : M −−→N is an absolute V-bundle, i.e. an ordinary real line
bundle, and these are classified by homotopy classes of maps [N, BZ2 ] ≈ H 1 (N, Z2 ).
In this case the singularities of N are not desingularized by the V-bundle, so if N
is a non-trivial orbifold so is M. Henceforth, for case (1) we restrict ourselves to
the smooth category where Hatakeyama’s theorem holds. Case (2) with G1 = S 1 is
more interesting since M can desingularize N. By Lemma 4.2.8 this occurs precisely
when the homomorphisms hi : Γi −−→S 1 are injective.
6.3. ALMOST CONTACT METRIC STRUCTURES 197

In either case Hatakeyama’s construction goes through with only slight modi-
fication. Let ξ denote the vector field on M generating the G1 action, and let η be
a connection 1-form on the V-bundle π : M −−→N. Then with respect to the basis 1
of the Lie algebra R of G1 , we have η(ξ) = 1, and the horizontal bundle D = ker η.
Now any point u ∈ M can be written as u = (x̃, a) ∈ Ũ × S 1 , where Ũ is local uni-
formizing neighborhood, and there is a vector space isomorphism π∗ : Du −−→Tx̃ Ũ .
Let qu : Tx̃ Ũ −−→Du denote the inverse map. Then we have a commutative diagram
Rh(γ)∗
Tx̃Ũ −−−−→ Tγ −1
x̃ Ũ
 
 
(6.3.2) qu quh(γ)
y y
Rh(γ)∗
Du −−−−→ Duh(γ) ,
where (Ũ , Γ, φ) is a local uniformizing system with γ ∈ Γ, and h is the homomor-
phism? We then define the tensor field Φ by
(6.3.3) Φu = qu ◦ Jx̃ ◦ π∗ .
This tensor field is invariant under the group of the bundle, namely, Φua = (Ra )∗ Φu .
Theorem 6.3.6: Let (N, J) be an almost complex orbifold and let π : M −−→N be
a principal G1 V-bundle over N. Suppose that ξ is a generator of the G1 action
that corresponds to 1 in the Lie algebra g1 ≈ R, and that η is a connection 1-form
on M. Then (ξ, η, Φ) with Φ defined by equation (6.3.3) defines an almost contact
structure on M. Furthermore, if h is a Hermitian metric on N compatible with J
then
g = π∗ h + η ⊗ η
is a bundle-like Riemannian metric on M compatible with the almost contact struc-
ture, and ξ is a Killing vector field for g.
Our next task is to formulate a converse to Theorem 6.3.6. Notice that the
ingredients are a bundle-like metric with ξ a Killing vector field, and a ξ-invariant
tensor field Φ. On an almost contact manifold (M, ξ, η, Φ) these conditions are
related mainly through Proposition 6.3.5; however, the invariance of Φ is not pinned
down completely.
Lemma 6.3.7: On an almost contact manifold (M, ξ, η, Φ) the following relations
hold:
£ξ Φ(ξ) = 0, η ◦ (£ξ Φ) ◦ Φ = £ξ η, (£ξ Φ)h ◦ Φ + Φ ◦ (£ξ Φ)h = 0 ,
where the superscript h means horizontal component. So if the tensor field Φ is
invariant under the flow of ξ then Fξ is taut.
Proof. The first equation follows from the first of equations (6.2.1). To prove
the remaining two equations, take the Lie derivative with respect to ξ of the second
equation in Definition 6.2.5 giving
(6.3.4) £ξ Φ ◦ Φ + Φ ◦ £ξ Φ = ξ ⊗ £ξ η .
Taking η of this implies that the second equation holds on D while the second
equation of (6.2.1) implies that it hold also on ξ. Taking the horizontal component
equation (6.3.4) gives the third equation. The last statement follows from the
second equation and Lemma 6.3.4. ¤
198 6. ALMOST CONTACT AND CONTACT GEOMETRY

First we shall make the simplifying assumption that the leaves of the charac-
teristic foliation Fξ are compact. In this case the leaves are all diffeomorphic to a
circle, however, the foliation need not be simple.
Theorem 6.3.8: Let (M, ξ, η, Φ) be an almost contact manifold such that the leaves
of the characteristic foliation are all compact. Suppose also that (M, ξ, η, Φ) admits
a compatible Riemannian metric g such that ξ is a Killing vector field which leaves
the tensor field Φ invariant. Then the space of leaves M/Fξ has the structure of
an almost Hermitian orbifold such that the canonical projection π : M −−→M/Fξ is
an orbifold Riemannian submersion, and a principal S 1 V-bundle over M/Fξ with
connection 1-form η.

Proof. Since ξ is a Killing vector field for the compatible Riemannian metric
g, Proposition 6.3.5 says that g is bundle-like and the leaves of Fξ are geodesics.
Thus, Fξ is a taut Riemannian foliation. Hence, since the leaves are all compact
Molino’s theorem 2.5.11 implies that the space of leaves is an orbifold, and the
projection π : M −−→M/Fξ is an orbifold Riemannian submersion. Moreover, by
Lemma 6.3.3 the leaves are geodesics, so Wadsley’s Theorem 2.6.12 implies that
the leaves are the orbits of an S 1 action on M. Since the holonomy groups of Fξ are
finite cyclic, this action is locally free, and thus π : M −−→M/Fξ is a principal S 1
V-bundle over M/Fξ . Again by Lemma 6.3.3 the horizontal subbundle D of T M
is invariant under the S 1 action and thus defines a connection on π : M −−→M/Fξ ,
and it is clear from Definition 1.3.2 that η is a connection 1-form of D. Finally,
for any vector field X on M/Fξ we can define an endomorphism I on M/Fξ by
setting (IX)h = ΦX h where the superscript h denotes the horizontal lift. This is
well defined by the invariance of the tensor field Φ under the circle action. So one
easily sees that I defines an almost complex structure on the orbifold M/Fξ . ¤

We end this section by mentioning in passing the classification of types of almost


contact metric structures, although we do not make use of them in this text. This
is somewhat involved, being accomplished by studying the representations of the
structure group U (n) × 1 on certain natural vector spaces [AG86, CG90, CM92].

6.4. Contact Metric Structures


The case of most interest to us is when the almost contact structure is a strict
contact structure. We now investigate this case further.
We now wish to relate strict contact structures to strict almost contact struc-
tures. While a strict contact structure uniquely determines the characteristic vector
field ξ, we need to add some more structure, namely the endomorphism Φ, in order
to obtain an underlying almost contact structure! Recall that on a contact man-
ifold a choice of 1-form η gives the contact bundle the structure of a symplectic
vector bundle, namely (D, dη). An almost complex structure J in D is said to be
compatible (cf. [MS98]2) with the symplectic form dη if dη(JX, JY ) = dη(X, Y )
for all vector fields X, Y , and dη(JX, X) > 0 for all X 6= 0 hold. We can extend
the endomorphism J to an endomorphism Φ on all of T M by setting Φξ = 0. The
conditions then become
(6.4.1) dη(ΦX, ΦY ) = dη(X, Y ) for all X, Y , dη(ΦX, X) > 0 for all X 6= 0,

2We use the opposite convention to McDuff and Salamon [MS98] who use ω(X, JX) > 0.
6.4. CONTACT METRIC STRUCTURES 199

and one easily checks that (ξ, η, Φ) defines an almost contact structure on M. Ac-
cordingly, we have
Definition 6.4.1: Let (M, η) be a contact manifold. Then an almost contact struc-
ture (ξ, η 0 , Φ) is said to be compatible with the contact structure if η 0 = η, ξ is its
Reeb vector field, and there is a smooth section Φ of the endomorphism bundle of
T M such that Φ2 = −1l + ξ ⊗ η and conditions (6.4.1) hold. We denote by AC(η)
the set of compatible almost contact structures on (M, η).
Every compatible almost contact structure is determined uniquely by the endo-
morphism Φ which in turn is determined uniquely by the almost complex structure
on D. So we can give the set AC(η) the subspace topology of the space of sections
Γ(End T M ) with the C ∞ compact open topology. Then we have
Proposition 6.4.2: Let (M, η) be a strict contact manifold. Then there is a
one-to-one correspondence between compatible almost complex structures J on the
symplectic vector bundle (D, dη) and compatible almost contact structures (ξ, η, Φ).
Moreover, the space AC(η) is contractible.

Proof. The first statement is clear from what was said above, and the second
statement follows from Proposition 2.61 of [MS98]. ¤

Notice that dη(ΦX, Y ) is symmetric in X and Y and is essentially the Levi form
Lη of the almost CR-structure defined by (D, Φ|D). So the last of conditions (6.4.1)
implies that the Levi form Lη is strictly pseudoconvex; hence, Lη = dη ◦ (Φ ⊗ 1l)
defines a Riemannian metric gD on the vector bundle D. We can then define a
metric g on M by
(6.4.2) g = gD + η ⊗ η = dη ◦ (Φ ⊗ 1l) + η ⊗ η
which, as one can easily check, is compatible with the almost contact structure
(ξ, η, Φ), i.e., (ξ, η, Φ, g) is an almost contact metric structure. Exercise 6.7 shows
that compatible Riemannian metrics always exist on a strict contact manifold.
Riemannian metrics of the form (6.4.2) are called associated Riemannian metrics,
and there is precisely one for every compatible almost contact structure. Thus,
Proposition 6.4.3: The set of associated Riemannian metrics are in one-to-one
correspondence with elements of AC(η).
Definition 6.4.4: A (strict) contact manifold (M, η) with a compatible (strict) al-
most contact metric structure (ξ, η, Φ, g) such that g(X, ΦY ) = dη(X, Y ) is called a
contact metric structure, and (M, ξ, η, Φ, g) is called a contact metric man-
ifold.
Remark 6.4.1: It is possible to have a contact manifold (M, η) with a compatible
almost contact structure (ξ, η, Φ) and a Riemannian metric g which is compatible
with the almost contact structure, but which is not the associated metric, i.e.,
g 6= dη ◦ (Φ ⊗ 1l). We refer to such a structure as a (strict) contact structure with a
compatible Riemannian metric. Of course, a contact metric structure is a contact
structure with a compatible Riemannian metric. Unless stated to the contrary, the
metric of a contact metric structure will mean an associated Riemannian metric.
An immediate corollary of Lemmas 6.3.3 and 6.1.24 is
Corollary 6.4.5: On a strict contact manifold (M, η) the characteristic foliation
Fξ is taut with respect to any compatible Riemannian metric.
200 6. ALMOST CONTACT AND CONTACT GEOMETRY

Lemma 6.4.6: On a strict contact metric manifold (M, η, g) the 1-form η is co-
closed, i.e., δη = 0.

Proof. Recall that δ = −?d?, where ? is the Hodge star operator with respect
to g. The Riemannian volume dvolg satisfies
1
dvolg = η ∧ (dη)n ,
n!
so we have
1
δη = − ? d ? η = − ? d(dη)n = 0 . ¤
n!

We are interested in the case when the characteristic foliation Fξ of a contact


metric structure is a Riemannian foliation. Although a Riemannian flow is not
necessarily isometric in general, it turns out that if the characteristic flow Fξ of a
contact metric structure is Riemannian if and only if it is isometric. First we have
Definition 6.4.7: A contact metric structure (ξ, η, Φ, g) is called K-contact if ξ
is a Killing vector field of g, i.e., if £ξ g = 0. A manifold with such a structure is
called a K-contact manifold.
Proposition 6.4.8: On a contact metric manifold (M, η, g) the following condi-
tions are equivalent:
(i)The characteristic foliation Fξ is a Riemannian foliation.
(ii)g is bundle-like.
(iii)The Reeb flow is an isometry.
(iv) The Reeb flow is a CR transformation, that is, it belongs to CR(M, D, J).
(v) The Reeb flow leaves the (1, 1) tensor field Φ invariant, that it belongs to
Aut(M, Φ).
(vi) The contact metric structure (M, η, g) is K-contact.

Proof. The equivalence of (i) and (ii) is Proposition 2.5.7. The two condi-
tions of Lemma 6.1.24 that characterize the Reeb vector field imply that both the
transverse symplectic form dη and the contact form η are invariant under the Reeb
flow. From Definition 2.5.6 g is bundle-like if and only if the transverse metric gD
is basic, that is, if and only if the Reeb flow leaves gD invariant as well. Hence,
the equivalence of (ii) and (iii). Since dη is invariant under the Reeb flow, gD is
invariant if and only if J is invariant, or equivalently if and only if Φ is invariant.
This implies the equivalence of (iii) with both (iv) and (v). Also (iii) and (vi) are
clearly equivalent. ¤

We are particularly interested in a certain type of CR-structure, namely when


the Levi form Lη = dη ◦ (Φ|D ⊗ 1l) is strictly pseudoconvex. The following definition
will prove convenient.
Definition 6.4.9: Let (D, J) be a strictly pseudoconvex almost CR structure on
M . We say that (D, J) is of Sasaki type if there exists a K-contact structure
S = (ξ, η, Φ, g) such that D = ker η and Φ|D = J.
Clearly, if the almost CR structure on M is integrable, then the K-contact
structure S will be Sasakian.
6.5. STRUCTURES ON CONES 201

6.5. Structures on Cones


We wish to describe certain types of conical geometry. Let M be a smooth
manifold, and consider the cone on M as C(M ) = M × R+ . We shall identify M
with M × {1}. We first describe the Riemannian situation.

6.5.1. Riemannian Cones. In this case the natural Riemannian structure


on C(M ) is not the product metric, but the so-called warped product [O’N83]
M ×r2 R+ , where r denotes the coordinate on R+ . More generally, we have
Definition 6.5.1: For any Riemannian metric gM on M, the warped product
metric on C(M ) = R+ × M is the Riemannian metric defined by
g = dr2 + φ2 (r)gM ,
where r ∈ R+ and φ = φ(r) is a smooth function, called the warping function. If
φ(r) = r then (C(M ), g) is simply called the Riemannian cone or metric cone
on M.
In this and the following chapters we will almost exclusively deal with the
metric cones. However, a more general cone metrics will also be discussed in Chap-
ter 14. It is clear that there is a one-to-one correspondence between Riemannian
metrics on M and cone metrics on C(M ). The cone metric admits a group of ho-
mothety transformations defined by (x, r) 7→ (x, λr). The infinitesimal generator of
the homothety group is the Liouville (or Euler) vector field defined by

(6.5.1) Ψ=r .
∂r
It is clear that the metric cone of Definition 6.5.1 is homogeneous of degree 2 and
satisfies
£Ψ g = 2g .

6.5.2. Almost Complex and Almost Hermitian Cones. We are inter-


ested in the geometry on C(M ) that corresponds to an almost contact structure on
M. Given an almost contact structure (ξ, η, Φ) on M we define a section I of the
endomorphism bundle of the tangent bundle T C(M ) of the cone by
(6.5.2) IY = ΦY + η(Y )Ψ , IΨ = −ξ .
It is easy to see that I defines an almost complex structure on C(M ) that is ho-
mogeneous of degree 0 in r. Alternatively, I is invariant under the flow of Ψ and
we
(6.5.3) £Ψ I = 0 .
Conversely, we begin with an almost complex structure I on C(M ) such that
(i) I is invariant under Ψ, i.e., equation (6.5.3) holds.
(ii) The vector field IΨ is tangent to M.
From this one easily sees that ξ = IΨ defines a nowhere vanishing vector field on M.
Letting Lξ denote the trivial line bundle generated by ξ, we have an exact sequence
0−−→Lξ −−→T M −−→Q−−→0 ,
and there is a one-to-one correspondence between the splittings of this exact se-
quence and 1-forms η on M that satisfy η(ξ) = 1. This correspondence is given by
202 6. ALMOST CONTACT AND CONTACT GEOMETRY

η 7→ ker η. One can then easily define an endomorphism Φ of T M by


(
I, on ker η
(6.5.4) Φ=
0 on Lξ .
It is easy to check that Φ2 = −1l + ξ ⊗ η, so (ξ, η, Φ) defines an almost contact
structure on M.
The choice of η in the paragraph above can be determined by giving a compat-
ible Riemannian structure. Explicitly,
Proposition 6.5.2: There is a one-to-one correspondence between almost contact
structures (ξ, η, Φ) on M and almost complex structures I on C(M ) satisfying the
two conditions above and equation (6.5.4). Furthermore, if g is a Riemannian
metric on M that is compatible with the almost contact structure, then the cone
metric h = dr2 +r2 g is Hermitian with respect to the almost complex structure I, and
the one-to-one correspondence above induces a one-to-one correspondence between
almost contact metric structures on M and almost Hermitian cone structures on
C(M ).
Proof. As described in the paragraph above an almost contact structure
(ξ, η, Φ) on M determines a unique almost complex structure I on C(M ) that
is invariant under Ψ, and vice-versa. Furthermore, if g is a metric on M satisfying
g(ΦX, ΦY ) = g(X, Y ) − η(X)η(Y ), then one checks (exercise) that the cone metric
h = dr2 + r2 g is Hermitian, that is that h satisfies h(IX, IY ) = h(X, Y ), where I
is given by equation (6.5.2). This correspondence is clearly injective. Finally, given
an almost Hermitian structure (I, h) where h is a cone metric and I is invariant
under Ψ, we define η by η(X) = r12 h(IΨ, X) and Φ by equation (6.5.4) and check
that g is compatible with the almost contact structure. ¤
Exercise 6.8: Fill in the details of the proof of Proposition 6.5.2 by showing that
the cone metric h = dr2 + r2 g is Hermitian if and only if g is compatible with the
almost contact structure.

6.5.3. Symplectic Cones. We begin with the following


Definition 6.5.3: A symplectic cone is a cone C(M ) = M × R+ with a sym-
plectic structure ω which has a one parameter group ρt of homotheties whose infin-
itesimal generator is a vector field on R+ .
Definition 6.5.3 is referred to as a symplectic Liouville structure in [LM87].
Notice that there is a freedom regarding the degree of homogeneity of the homoth-
eties. As we shall want the symplectic structure on C(M ) to be compatible with
the standard cone metric given in Definition 6.5.1, we ask that the symplectic form
ω be homogeneous3 of degree 2, that is the equation
(6.5.5) 2ω = £Ψ ω = d(Ψ ω)
holds. Notice that it follows from this equation that ω is an exact 2-form, so we
define the 1-form η̃ on C(M ) by η̃ = Ψ ω giving ω = dη̃. We can define a 1-form
η on M by η = η̃|M ×{1} . Then we see that η̃ = r2 η, and this gives ω as
(6.5.6) ω = 2rdr ∧ η + r2 dη .

3Note that in [BG00a] we took ω to be homogeneous of degree one.


6.5. STRUCTURES ON CONES 203

The condition that ω n+1 6= 0 implies that η gives M the structure of a strict
contact manifold with a fixed 1-form. Conversely, one can easily reverse this proce-
dure: given a strict contact manifold (M, η), defining ω on C(M ) by equation 6.5.6
gives the cone C(M ) a symplectic structure with homotheties. In [Eli98, MS98]
(C(M ), ω) is called a symplectization of M while it is called a symplectification of
M in [LM87]. The terminology is apparently due to Arnold [Arn78]. We have
arrived at the well-known result
Proposition 6.5.4: Let η be a 1-form on the manifold M. Then η defines a strict
contact structure on M if and only if the 2-form d(r2 η) defines a symplectic struc-
ture on the cone C(M ) = M × R+ .
Let (η, g) be a contact metric structure on M with g an associated metric. Then
by Proposition 6.4.3 there is a unique almost contact structure (ξ, η, Φ) satisfying
g(X, Y ) = dη(ΦX, Y ) and ξ dη = 0. The metric g = dη ◦ (Φ ⊗ 1l) + η ⊗ η on M
corresponds to the metric dr2 + r2 (dη ◦ (Φ ⊗ 1l) + η ⊗ η) on C(M ). We have arrived
at
Proposition 6.5.5: There is a one-to-one correspondence between the contact met-
ric structures (ξ, η, Φ, g) on M and almost Kähler structures (dr2 + r2 g, d(r2 η), I)
on C(M ).
So what does the K-contact condition on M correspond to? We have
Proposition 6.5.6: A contact metric structure (ξ, η, Φ, g) is K-contact if and only
if Ψ − iξ pseudo-holomorphic with respect to the almost Kähler structure (dr2 +
r2 g, d(r2 η), I) on C(M ).

Proof. By the properties of the Reeb vector field one easily sees that Ψ − iξ is
pseudo-holomorphic, i.e., an infinitesimal automorphism of I if and only if £ξ I =
£ξ Φ = 0. But by Proposition 6.4.8 this holds if and only if £ξ gD = 0. ¤

Remark 6.5.1: In the case that the complex vector field Ψ − iξ on C(M ) is not
pseudo-holomorphic, a quotient formed by dividing by the resulting C∗ action, or
equivalently, by the symplectic reduction of the S 1 action will lose both the almost
complex structure and the Riemannian structure.

6.5.4. Integrability and Normality. The main thrust of this monograph


will deal with the case in which the complex structure on the C(M ) is integrable.
This leads to [SH62]
Definition 6.5.7: An almost contact structure (ξ, η, Φ) is said to be normal if the
corresponding almost complex structure I on C(M ) is integrable.
Combining this definition with Proposition 6.5.2 immediately gives
Proposition 6.5.8: There is a one-to-one correspondence between the normal al-
most contact structures (ξ, η, Φ) on M and almost complex structures I on C(M )
that are integrable and satisfy the conditions of Proposition 6.5.2. Furthermore, if
g is a Riemannian metric on M that is compatible with the almost contact struc-
ture, then the cone metric h = dr2 + r2 g is Hermitian with respect to the complex
structure I.
As is well-known and discussed in Example 1.4.9 the integrability of I is equiv-
alent to the vanishing of its Nijenhuis torsion tensor NI given by equation (1.4.4).
More generally the Nijenhuis torsion tensor NΦ can be defined for any (1, 1)-tensor
204 6. ALMOST CONTACT AND CONTACT GEOMETRY

field Φ, viz.

(6.5.7) NΦ (X, Y ) = [ΦX, ΦY ] + Φ2 [X, Y ] − Φ[X, ΦY ] − Φ[ΦX, Y ] .

Note that the Nijenhuis torsion tensor of a (1, 1)-tensor field is a tensor field of type
(1, 2), and that NΦ (Y, X) = −NΦ (X, Y ). In particular we want to determine the
normality condition in terms of the Nijenhuis torsion tensor for the tensor field Φ
of an almost contact structure.
Theorem 6.5.9: An almost contact structure (ξ, η, Φ) on M is normal if and only
if NΦ = −2ξ ⊗ dη.

Proof. Since NI is a tensor field of type (1, 2) on C(M ), it suffices to compute


it on pairs of vector fields of the form X, Y and X, Ψ, where X and Y are vector
fields on M. Using equation (6.5.2)

NI (X, Y ) = [IX, IY ] − [X, Y ] − I[X, IY ] − I[IX, Y ]


= [ΦX + η(X)Ψ, ΦY + η(Y )Ψ] − [X, Y ]
− I[X, ΦY + η(Y )Ψ] − I[ΦX + η(X)Ψ, Y ]
= [ΦX, ΦY ] + Φ2 [X, Y ] − Φ[X, ΦY ] − Φ[ΦX, Y ] + [η(X)Ψ, ΦY ]
+ [ΦX, η(Y )Ψ] − η([X, Y ])ξ − Xη(Y )IΨ + Y η(X)IΨ
− η([ΦX, Y ] + [X, ΦY ])Ψ
= NΦ (X, Y ) − (ΦY η(X) − ΦXη(Y ) − η([ΦX, Y ] + [X, ΦY ]))Ψ
+ (Xη(Y ) − Y η(X) − η([X, Y ]))ξ
= NΦ (X, Y ) + 2dη(X, Y )ξ + 2(dη(ΦX, Y ) + dη(X, ΦY ))Ψ
¡ ¢
= NΦ (X, Y ) + 2dη(X, Y )ξ − 2 (£ΦY η)(X) − (£ΦX η)(Y ) Ψ .

Next we compute

NI (X, Ψ) = [IX, IΨ] − [X, Ψ] − I[X, IΨ] − I[IX, Ψ]


= −[ΦX + η(X)Ψ, ξ] + Φ[X, ξ] + η([X, ξ])Ψ
− Φ[ΦX + η(X)Ψ, Ψ] − η([ΦX + η(X)Ψ, Ψ])
¡ ¢
= −[ΦX, ξ] + Φ[X, ξ] + η([X, ξ]) + ξη(X) Ψ
= (£ξ Φ)(X) + (£ξ η)(X)Ψ .

At this stage it is traditional to define certain tensor fields, and we shall follow suit.

(6.5.8) N (1) (X, Y ) = NΦ (X, Y ) + 2dη(X, Y )ξ ,


(6.5.9) N (2) (X, Y ) = (£ΦX η)(Y ) − (£ΦY η)(X) ,
(6.5.10) N (3) (X) = (£ξ Φ)(X), N (4) (X) = (£ξ η)(X) .

Equating to zero separately the components of NI (X, Y ) that are tangent to M


and those that are proportional to Ψ, shows that (ξ, η, Φ) is normal if and only if
N (i) (X, Y ) = 0 for all i = 1, . . . , 4. However, the following lemma shows that the
vanishing of N (1) is both necessary and sufficient for normality.
Lemma 6.5.10: If N (1) vanishes so does N (i) for i = 2, 3, 4.
6.5. STRUCTURES ON CONES 205

Proof. For any vector field X on M, consider


N (1) (X, ξ) = NΦ (X, ξ) + 2dη(X, ξ)ξ
= Φ2 [X, ξ] − Φ[ΦX, ξ] − ξη(X)ξ − η([X, ξ])ξ
= −[X, ξ] + η([X, ξ])ξ − Φ[ΦX, ξ] − ξη(X)ξ − η([X, ξ])
= −[X, ξ] − Φ[ΦX, ξ] − ξη(X)ξ .
Applying η to this equation shows that the vanishing of N (1) implies η([X, ξ]) +
ξη(X) = 0 which is equivalent to £ξ η = ξ dη = 0. This shows that N (4) = 0.
Next we notice that replacing X by ΦX in the equation for N (1) (X, ξ), gives
(6.5.11) N (1) (ΦX, ξ) = [ξ, ΦX] − Φ[ξ, X] = (£ξ Φ)(X) = N (3) (X) .
So the vanishing of N (1) implies the vanishing of N (3) .
Finally, we apply η to N (1) (ΦX, Y ) which gives
η(N (1) (ΦX, Y )) = η(NΦ (ΦX, Y )) + 2dη(ΦX, Y )
= η([Φ2 X, ΦY ]) + 2dη(ΦX, Y )
= −η([X, ΦY ]) + η([η(X)ξ, ΦY ]) + 2dη(ΦX, Y )
= −η([X, ΦY ] − ΦY η(X) + η(X)η([ξ, ΦY ]) + 2dη(ΦX, Y )
= 2dη(ΦX, Y ) − 2dη(ΦY, X) − 2η(X)dη(ξ, ΦY )
¡ ¢
= 2 (£ΦX η)(Y ) − (£ΦY η)(X) − 2η(X)dη(ξ, ΦY )
= 2N (2) (X, Y ) − 2η(X)dη(ξ, ΦY ) .
But we saw in the argument for the vanishing of N (4) that the vanishing of N (1)
implies ξ dη = 0, hence, the last term on the right vanishes whenever N (1) vanishes,
so the vanishing of N (1) implies the vanishing of N (2) . ¤
This completes the proof of the theorem. ¤
Let us compare normality with the integrability of the underlying almost CR-
structure.
Corollary 6.5.11: An almost contact metric structure (ξ, η, Φ, g) on M is normal
if and only if
(i) The almost CR-structure (D, Φ|D ) is integrable, and
(ii) Φ is invariant under ξ, i.e., N (3) = 0.
Proof. We set J = Φ|D and compute the Nijenhuis tensor for X, Y sections
of D.
NJ (X, Y ) = [JX, JY ] − [X, Y ] − J([X, JY ] + [JX, Y ]) = NΦ (X, Y ) + 2dη([X, Y ])ξ .
Moreover, from equations (6.5.8) and (6.5.11) we have
(£ξ Φ)(X) = N (1) (ΦX, ξ) = NΦ (ΦX, ξ) + 2dη(ΦX, ξ)ξ .
Using these two equations the result follows from Theorem 6.5.9. ¤
In the case of a contact metric structure the four tensor fields defined above
simplify a bit.
Proposition 6.5.12: Let (ξ, η, Φ, g) be a contact metric structure. Then N (2) =
N (4) = 0. Furthermore, N (3) = 0 if and only if (ξ, η, Φ, g) is K-contact.
206 6. ALMOST CONTACT AND CONTACT GEOMETRY

Proof. First we have


N (4) = (£ξ η) = d(η(ξ)) + ξ dη = 0
when (ξ, η, Φ, g) is contact. Next we have
N (2) (X, Y ) = (£ΦX η)(Y ) − (£ΦY η)(X)
= ΦX(η)(Y )) − η([ΦX, Y ]) − ΦY (η)(X)) + η([ΦY, X])
= 2dη(ΦX, Y ) − 2dη(ΦY, X) = 2g(Y, X) − 2g(X, Y ) = 0 .
For N (3) we notice first that since (ξ, η, Φ, g) is contact, £ξ η = £ξ dη = 0. Since g
is the associated metric 6.4.2, we have
£ξ g = dη ◦ (£ξ Φ ⊗ 1l) .
(3)
So the vanishing of N implies that ξ is Killing. Conversely, if ξ is Killing, the
non-degeneracy of the bilinear form dη on D implies that N (3) (X) = 0 when X is
a section of D. But since Φξ = 0, N (3) vanishes identically on ξ. ¤
Our final definition of this subsection is the major object of study in this mono-
graph.
Definition 6.5.13: A normal contact metric structure S = (ξ, η, Φ, g) on M is
called a Sasakian structure. A pair (M, S) called a Sasakian manifold.
It follows immediately from Corollary 6.5.11 and Proposition 6.5.12 that
Proposition 6.5.14: A Sasakian structure S = (ξ, η, Φ, g) is a K-contact structure
whose almost CR-structure is integrable.
Actually as seen in Lemma 6.3.3 N (4) vanishes as soon as there is a compatible
metric of the almost contact structure such that integral curves of ξ are geodesics.
From the definition of normality 6.5.7 and Proposition 6.5.5, Definition 6.5.13 is
equivalent to:
Definition 6.5.15: A contact metric manifold (M, ξ, η, Φ, g) is Sasakian if its
metric cone (C(M ), dr2 + r2 g, d(r2 η), I) is Kähler.
CHAPTER 7

K-Contact and Sasakian Structures

In this chapter we begin a detailed study of our main subject: Sasakian Geome-
try. Since there are many properties that also hold for K-contact structures, we will
often consider both. The reason for imposing the K-contact condition and not work-
ing with the more general contact metric structures comes from Proposition 6.4.8
which says, among other things, that a contact metric structure is K-contact if and
only if the characteristic foliation Fξ is Riemannian. The latter condition is neces-
sary for much of the geometry that we develop. Indeed, early (circa 1960) in the de-
velopment of global contact geometry the ‘nice’ situation was that described by the
well-known Boothby-Wang fibration (Theorem 6.1.26 of Chapter 6). As mentioned
in Chapter 6 Sasakian structures were first studied by Sasaki and Hatakeyama in
[SH62] under the name normal contact metric structures (cf. Definition 6.5.13).
In 1963 Hatakeyama, Ogawa, and Tanno [HOT63] introduced the name K-contact
manifold, and proved the basic Theorem 7.3.7 found below. There was a systematic
development of the curvature properties of both K-contact and Sasakian geometries
throughout the sixties and seventies, see the books [Bla76a, Bla02, YK84]. The
present chapter summarizes all the basic geometric properties of such manifolds
(M, ξ, η, Φ, g) as well as adding some more recent developments of a topological
nature due mainly to Rukimbira and those of a algebro-geometric nature due to
the authors and their collaborators. Throughout this chapter we will learn how
most of these properties, starting with the key ‘structure’ Theorem 7.1.3, closely
evolve around the geometric objects depicted in the following diagram:

(C(M ), ω, Φ̄, ḡ) ←- (M, ξ,η, Φ, g)




π
y
(Z, Ω, J, h).

7.1. Quasi-regularity and the Structure Theorems


In this section we begin with a more detailed discussion of the regularity prop-
erties introduced in Definition 6.1.25. To best illustrate this definition consider
the linear flows on the sphere described by Example 2.3.5. It is not difficult to
show that these flows are the characteristic flows belonging to K-contact, in fact,
Sasakian structures on S 2n+1 . We shall see this explicitly shortly. Note that the
case in Example 2.3.5 when the ratio w0 /w1 is rational (irrational) corresponds to
a quasi-regular (irregular) contact form, respectively. For quasi-regular K-contact
or Sasakian structures we have:

207
208 7. K-CONTACT AND SASAKIAN STRUCTURES

Definition 7.1.1: The order of a quasi-regular K-contact structure S is the least


common multiple (assuming it exists) of the orders of the leaf holonomy groups of
the characteristic foliation, and is denoted by υ(S).
A K-contact manifold with υ < ∞ is said to be of finite order. In the quasi-
regular case the order υ(S) coincides with the order of the orbifold Z of Definition
4.1.5. Of course, any compact quasi-regular K-contact manifold is of finite or-
der, and the order of a regular K-contact structure is one. In the compact case
quasi-regularity implies K-contact. In fact, compactness of the leaves is all that is
necessary. We see that from the existence of the Reeb vector field 6.1.24, Lemma
6.3.3, and Wadsley’s Theorem 2.6.12 we obtain
Proposition 7.1.2: Let (ξ, η) be a quasi-regular strict contact structure on M such
that the leaves of the characteristic foliation Fξ are compact. Then the Reeb flow
generates a locally free circle action on M. Furthermore, there exists a Riemannian
metric g such that the induced structure (ξ, η, g) is K-contact.
Proof. Since the Reeb vector field ξ satisfies £ξ η = 0, Lemma 6.3.3 implies
that there exists a compatible Riemannian metric such that the leaves of Fξ are
geodesics. Thus, Wadsley’s Theorem implies that the leaves are the orbits of a
locally free circle action.
To prove the second statement we let φt denote the Reeb flow generating the
locally free circle action. Let g0 be any compatible metric. The orbits φt (x) are
geodesics with respect to this metric, but g0 is not necessarily invariant under the
circle action generated by φt . However, the metric g defined by
Z
g= φ∗t g0 dt
S1
1
is invariant under the S -action, and the Reeb flow φt is an isometry. Furthermore,
Lemma 6.3.3 implies that the orbits φt (x) are also geodesics of g. Hence, (M, η, g)
is K-contact by Proposition 6.4.8. ¤
The second statement of this proposition is due to Rukimbira [Ruk95a]. We
now have a special case of Theorem 6.3.8 for K-contact manifolds, namely our first
structure theorem:
Theorem 7.1.3: Let (M, ξ, η, Φ, g) be a quasi-regular K-contact manifold with
compact leaves. Then
(i) The space of leaves M/Fξ is an almost Kähler orbifold Z such that the
canonical projection π : M −−→M/Fξ is an orbifold Riemannian submer-
sion.
(ii) M is the total space of a principal S 1 orbibundle (V-bundle) over M/Fξ
with connection 1-form η whose curvature dη is the pullback by π of a
symplectic form ω on M/Fξ .
(iii) The symplectic form ω defines a non-trivial integral orbifold cohomology
class, that is, [p∗ ω] ∈ Horb
2
(M/Fξ , Z) where p is the natural projection
defined by equation (4.3.5).
(iv) The leaves of Fξ are all geodesics.
(v) If the characteristic foliation Fξ is regular, then the circle action is free
and M is the total space of a principal S 1 bundle over an almost Kähler,
hence, symplectic manifold defining an integral class [ω] ∈ H 2 (M/Fξ , Z).
(vi) (M, ξ, η, Φ, g) is Sasakian if and only if (M/Fξ , ω) is Kähler.
7.1. QUASI-REGULARITY AND THE STRUCTURE THEOREMS 209

Remark 7.1.1: Part (v) of this theorem is essentially a restatement of the Boothby-
Wang Theorem 6.1.26. In the regular K-contact (Sasakian) case it was proved
by Hatakeyama (cf. [Hat63], Theorem 4). Hence, Theorem 7.1.3 is the orb-
ifold/orbibundle generalization of the Boothby-Wang fibration. Actually the K-
contact condition can be dropped as seen by Proposition 7.1.2.
The reader should recall the relation between quasi-regular Sasakian manifolds
and Seifert S 1 -bundles as discussed in Section 4.7. Indeed, Theorem 4.7.7 says
that S 1 orbibundles over complex orbifolds and Seifert S 1 -bundles are one and
the same. Then Theorem 7.5.2 below says that a compact quasi-regular Sasakian
manifold is nothing but an S 1 -Seifert bundle over a normal projective algebraic
variety. Of course, dimension 3 is the classical case (cf. [Orl72b] and references
therein). This was generalized to n dimensions when the projective algebraic variety
is smooth in [OW75]. As presented in Section 4.7, Kollár has recently begun a
more general study of Seifert bundles [Kol07, Kol05, Kol04, Kol06a] which has
already produced substantial results in Sasakian geometry that we discuss later in
Chapter 10. Although Seifert bundles were originally defined in the case that M/S 1
is complex, it is straightforward to generalize this to the case of a locally free circle
action [Kol06a]. Here we are only interested in the case of quasi-regular K-contact
structures, and we think of these as orbibundles over symplectic orbifolds Z. As in
the complex case [OW75], if M is a compact quasi-regular K-contact manifold of
order υ, there is a commutative diagram
(7.1.1) MD
DD
DD
DD
D"
π Mυ ,
zz
zz
zz
² |zz
Z
where Mυ is the quotient M/Zυ . Generally, Mυ is a developable orbifold with
π1orb (Mυ ) ≈ Zυ . Now there is an induced action of S 1 /Zυ = S 1 on Mυ which is
free, so the southwest arrow is equivalent to the projection map of a principal S 1
bundle over Z. We can also consider the cone C(M ) which is a C∗ orbibundle over
Z in which case the commutative diagram 7.1.1 becomes
(7.1.2) C(M )
JJ
JJ
JJ
JJ
J$
π C(Mυ ) .
ss
sss
ss
² ysss
Z
When M is Sasakian, so is Mu , and the orbifold is projective algebraic. In the
case the southeast arrow is a branched cover. This shows that the natural orbifold
fibration associated to any compact quasi-regular Sasakian manifold factors in terms
of a branched cover and an S 1 -bundle map.
Theorems 7.1.3 and 4.3.18 give some immediate information about quasi-regular
K-contact manifolds.
210 7. K-CONTACT AND SASAKIAN STRUCTURES

Proposition 7.1.4: Let M be a quasi-regular K-contact manifold with orbifold


fibration π : M −−→Z. Then we have
(i) πi (M ) = πiorb (Z) for i ≥ 3.
(ii) The sequence
0−−→π2 (M )−−→π2orb (Z)−−→Z−−→π1 (M )−−→π1orb (Z)−−→{1}
is exact. In particular, if M is simply connected, then π1orb (Z) = {1} and
π2orb (Z) ≈ π2 (M ) ⊕ Z.
More information relating the homology and cohomology groups of M and Z
can be obtained by studying harmonic theory as well as the Leray spectral sequence
of the map π. This is treated in more detail in Section 7.4.
Example 7.1.5 (Sasakian structure on S 2n+1 [SH62]): We begin Pn with the
standard contact structure1 on S 2n+1 , namely the restriction of η0 = i=0 (yi dxi −
xi dyi ) to the unit sphere
n
X n
X ¡ ¢
S 2n+1 = {z = (x, y) ∈ Cn+1 | |zi |2 = (xi )2 + (yi )2 = 1} ,
i=0 i=0
where xi , yi denote the real and imaginary components of zi ∈ C. In complex
coordinates the 1-form η0 takes the form
n
iX
η0 = − (zj dz̄j − z̄j dzj ) .
2 j=0

The Reeb vector field is


n ³
∂ ´
n ³
X ∂ X ∂ ∂ ´
ξ0 = yi − xi = −i zj − z̄j
i=0
∂xi ∂yi j=0
∂zj ∂ z̄j

while the (1, 1)-tensor field Φ0 is obtained by restricting the standard complex
structure is
Xn ³ ´ Xn ³ ´
∂ ∂ ∂ ∂
I= ⊗ dxi − ⊗ dyi = i ⊗ dzj − ⊗ dz̄j
i=0
∂yi ∂xi j=0
∂zj ∂ z̄j

on R2n+2 to D ⊂ T S 2n+1 and extends trivially to the real line bundle generated by
ξ0 . Explicitly, we find
X³ ∂ ∂ ∂ ∂ ´
Φ0 = (xi xj −δij ) ⊗dyj −(yi yj −δij ) ⊗dxj +xj yi ⊗dyj −xi yj ⊗dxj .
i,j
∂xi ∂yi ∂yi ∂xi

The canonical metric g0 on S 2n+1 is the flat metric on R2n+1 restricted to the
S 2n+1 . It has constant sectional curvature 1 and satisfies
g0 = dη0 ◦ (Φ0 ⊗ 1l) + η0 ⊗ η0 .
It is easy to check that £ξ0 g0 = 0, so (ξ0 , η0 , Φ0 , g0 ) is a K-contact structure.
Moreover, since the I describes the standard flat complex structure on Cn+1 , this
1Our choice of 1-form η here is the negative of the one used in Example 6.1.16. The reason
0
for this is to keep the standard conventions of [Bla02, YK84] for Sasakian geometry, while at
the same time using the standard complex structure on Cn+1 and not the complex conjugate. Of
course there is no change in the underlying contact structure, but only a change of orientation or
co-orientation.
7.1. QUASI-REGULARITY AND THE STRUCTURE THEOREMS 211

structure is Sasakian. We refer to the structure S0 = (ξ0 , η0 , Φ0 , g0 ) as the stan-


dard Sasakian structure on S 2n+1 . We shall also refer to the conjugate struc-
ture S0c = (−ξ0 , −η0 , −Φ0 , g0 ) as a standard Sasakian structure when no confusion
should arise. (See Definition 7.5.11 below).
These are both regular Sasakian structures, and are nothing but the Hopf
fibrations as principal S 1 bundles over the complex projective space CPn . Now
Cn+1 with the standard complex structure and the positive oriented C∗ -action
(i.e., z 7→ reit z) is the total space of the tautological line bundle O(−1). Since
for the Sasakian structure S0 the circle action has the negative orientation, S0 is
associated with the positive line bundle O(1), as it must by Theorem 7.5.1. The
conjugate Sasakian structure S0c is associated with complex line bundle O(−1) but
with the complex conjugate structure, and as such it is also a positive line bundle.
Before turning to the irregular case, we give an inversion theorem to Theorem
7.1.3. In the regular K-contact (and also Sasakian) case the result was observed by
Hatakeyama [Hat63].
Theorem 7.1.6: Let (Z, ω, J) be an almost Kähler orbifold with [p∗ ω] ∈ Horb 2
(Z, Z),
and let M denote the total space of the circle V-bundle defined by the class [ω]. Then
the orbifold M admits a K-contact structure (ξ, η, Φ, g) such that dη = π ∗ ω where
π : M −−→Z is the natural orbifold projection map. Furthermore, if all the local
uniformizing groups of Z inject into the structure group S 1 , then M is a smooth
K-contact manifold.
Remark 7.1.2: Notice that the Sasakian structure on M is not uniquely deter-
mined, but is determined only up to a gauge transformation of the S 1 V-bundle.
Once a gauge is chosen the Sasakian structure on M is then uniquely determined by
the Kähler structure and S 1 V-bundle on Z. We discuss such gauge transformations
further in Section 7.5.
The following is immediate:
Corollary 7.1.7: Let M be a smooth manifold that admits a quasi-regular K-
contact structure (ξ, η, Φ, g), then the leaf holonomy groups, or equivalently, the
local uniformizing groups of the orbifold M/Fξ are all cyclic.
To study the irregular case we recall the toral rank of a Riemannian flow 2.6.5.
Here we specialize to a K-contact structure (ξ, η, Φ, g) on a smooth manifold M.
We denote the isometry group of (M, g) by Isom(M, g), and the group of automor-
phisms of the K-contact structure (M, η, g), by Aut(ξ, η, Φ, g). We let isom(M, g)
and aut(ξ, η, Φ, g) denote the corresponding Lie algebras. We study these Lie groups
and Lie algebras more thoroughly in Chapter 8. When M is compact the well-known
theorem of Myers and Steenrod says that Isom(M, g) is a compact Lie group.
Moreover, Aut(ξ, η, Φ, g) is a closed Lie subgroup of Isom(M, g) [Tan69a]. In the
K-contact case the Reeb flow belongs to the automorphism group Aut(ξ, η, Φ, g)
which is a compact Lie group. Thus, the closure T of the Reeb flow is a compact
commutative Lie group which lies in Aut(ξ, η, Φ, g). Now the Reeb flow is a strict
contact transformation lying in the center of the group of strict contact transforma-
tions [LM87]; hence, it lies in the center of Aut(ξ, η, Φ, g). It follows that T also lies
in the center of Aut(ξ, η, Φ, g). Combining this discussion with Carrière’s Theorem
2.6.4, we have
Proposition 7.1.8: Let (M, ξ, η, Φ, g) be a compact K-contact metric manifold.
Then the leaf closures of the Reeb flow are the orbits of a torus T lying in the
212 7. K-CONTACT AND SASAKIAN STRUCTURES

center of the Lie group Aut(ξ, η, Φ, g) of automorphisms of (ξ, η, Φ, g) and the Reeb
flow is conjugate to a linear flow on T.
The dimension of the maximal torus in Proposition 2.6.3 is the toral rank
rk(M, Fξ ) of (ξ, η, Φ, g), and the contact geometry places a smaller upper bound on
the toral rank [Ruk94].
Lemma 7.1.9: Let (M, ξ, η, Φ, g) be a compact K-contact metric manifold of di-
mension 2n+1. Then the rank rk(M, Fξ ) is independent of the metric g and satisfies
1 ≤ rk(M, Fξ ) ≤ n + 1. Moreover, rk(M, Fξ ) ≤ rk Aut(ξ, η, Φ, g) ≤ dim a(ξ, η, Φ, g).

Proof. Consider the Lie algebra t of T. It consists of the Reeb vector field ξ
together with the infinitesimal generators of the leaf closures. The projections of
these generators onto D are global sections of Molino’s commuting sheaf [Mol79,
Mol88] C(M, Fξ ). Thus, they give integral submanifolds of the subbundle D. But
from Proposition 6.1.27 the integral submanifolds of D of maximal dimension, that
is the Legendrian submanifolds of the contact structure, have dimension n. Hence,
these together with the Reeb vector field generate a torus of dimension at most
n + 1. Furthermore, Molino shows that the commuting sheaf is independent of the
transverse metric, so rk(M, Fξ ) is independent of g. ¤

Remark 7.1.3: The case rk(M, Fξ ) = 1 is the quasi-regular case, while the other
extreme rk(M, Fξ ) = n + 1 is the toric case studied in [BM93a, BM96, BG00a].
It might seem that the irregular case could prove to be intractable, but this
is not at all the case at least if M is compact thanks to the recent approximation
theorem of Rukimbira [Ruk95a]. This result is based on Carrière’s Theorem 2.6.4
on Riemannian flows. The theorem proves that any irregular K-contact structure
can be approximated by quasi-regular ones. We thus arrive at our second structure
theorem:
Theorem 7.1.10: Let M be compact. Any K-contact structure (ξ, η, Φ, g) on M
is either quasi-regular or there is a sequence of quasi-regular K-contact structures
{(ξj , ηj , Φj , gj )}j converging in the compact-open C ∞ -topology to (ξ, η, Φ, g). In par-
ticular, if M admits an irregular K-contact structure, it admits many locally free
circle actions.
Conversely, any quasi-regular strict contact structure η on a compact manifold
M admits a compatible K-contact structure (ξ, η, Φ, g).

Proof. Assume the K-contact structure is irregular. Since M is compact we


can apply Proposition 7.1.8 which says that the closure L̄ of any leaf L of the
characteristic foliation Fξ is a torus. Furthermore, by Proposition 2.6.3 L̄ is the
leaf of a singular Riemannian foliation F¯ξ which is the closure of Fξ . The vector
fields defining F¯ξ span a commutative Lie subalgebra t(ξ, η, Φ, g) of the Lie algebra
a(ξ, η, Φ, g) of infinitesimal automorphisms of the K-contact structure. Since by
Proposition 7.1.8 the flow of ξ is conjugate to a linear flow on L̄, there exist a
sequence of vector fields {ρj } ⊂ t(ξ, η, Φ, g) such that ξj = ξ + ρj is quasi-regular
for each j and limj→∞ ρj = 0. Now define a 1-form
η
ηj = .
1 + η(ρj )
We can choose the vector fields ρj small enough in the compact-open C ∞ -topology
so that ηj is a contact 1-form for each j.
7.1. QUASI-REGULARITY AND THE STRUCTURE THEOREMS 213

Next we define the (1, 1)-tensor fields and metrics by


1
(7.1.3) Φj = Φ − Φρj ⊗ η ,
1 + η(ρj )
and
(7.1.4) gj = dηj ◦ (Φj ⊗ 1l) + ηj ⊗ ηj ,
where D = ker η = ker ηj . Again one can choose the vector fields ρj small enough so
that (ξj , ηj , Φj , gj ) are K-contact structures converging to (ξ, η, Φ, g). This proves
the first statement. Since M is compact and the flow is quasi-regular, the leaves of
Fξ are compact. The second statement then follows from the second statement of
Proposition 7.1.2. ¤
An immediate corollary of Theorem 7.1.10 is
Corollary 7.1.11: Every K-contact manifold admits a fixed point free circle action.
Exercise 7.1: Fill in the details of the proof of Theorem 7.1.10 by showing that
the tensor fields (ξj , ηj , Φj , gj ) satisfy the conditions of Definition 6.4.1.
To consider non-regular, both quasi-regular and irregular examples we con-
struct deformation of the round sphere of Example 7.1.5. Such deformations will
be formalized later in section 8.2.
Example 7.1.12: Consider the standard Sasakian structure S0 = (ξ0 , η0 , Φ0 , g0 ) de-
scribed in Example 7.1.5. Now the isometry group I(S 2n+1 , g0 ) of the round sphere
is O(2n + 2, R), and the automorphism group Aut(ξ0 , η0 , Φ0 , g0 ) of the Sasakian
structure is the subgroup of O(2n + 2, R) that commutes with ξ0 , namely, U (n + 1).
The maximal torus Tn+1 of U (n + 1) has n + 1 dimensions. Let {ei }ni=0 denote the
standard basis for its Lie algebra tn+1 = Rn+1 . Under the action of Tn+1 on S 2n+1 ,
the standard basis vectors {ei }ni=0 correspond to the vector fields {Hi }ni=0 , where
∂ ∂
Hi = y i − xi ,
∂xi ∂yi
P
and the Reeb vector field ξ0 can be written as ξ0 = i Hi .
We now “deform” the standard Sasakian structure of Example 7.1.5 by con-
sidering any positive vector w = (w0 , . . . , wn ) ∈ tn+1 = Rn+1 with wi > 0 for all
i = 0, . . . , n. The vector field ξw corresponding to w is the one described in Exam-
ple 2.3.5 and the corresponding 1-form is denoted by ηw . In this way we obtain a
deformed Sasakian structure Sw = (ξw , ηw , Φw , gw ) defined by
n
X η
(7.1.5) ξw = wi Hi , ηw = Pn 2 2
,
i=0
w
i=0 i ((x i ) + (yi ) )

where Φw and gw are determined by equations (7.1.3) and (7.1.4), respectively with
j replaced by w. Notice that
n
X
η(ξw ) = 1 + η(%w ) = wi ((xi )2 + (yi )2 ) > 0 .
i=0

This deformation procedure was first introduced by Takahashi [Tak78] and will be
formalized in Section 8.2. It follows (cf. Section 8.2.3) that Sw is indeed a Sasakian
structure as long as all the components of w are positive. We shall refer to Sw as
the weighted Sasakian structure on S 2n+1 , and we denote the unit sphere with
214 7. K-CONTACT AND SASAKIAN STRUCTURES

2n+1
this Sasakian structure by Sw . In this case we have the ‘linear’ flow as described
in Example 2.3.5.
The deformed structures Sw , however, are not all distinct. The Weyl group
W(SU (n + 1)) = Nor(Tn+1 )/Tn+1 acts as outer automorphisms on the maximal
torus, and it is well-known that W(SU (n + 1)) = Σn+1 , the permutation group
on n + 1 letters. So for any σ ∈ Σn+1 the Sasakian structures Sw and Sσ(w) are
isomorphic. Thus, we shall frequently take the weights to be ordered such that
w0 ≤ w1 ≤ · · · ≤ wn . It is important to note that all the Sasakian structures Sw
belong to the same underlying contact structure on S 2n+1 , namely the standard
one. Later we shall discover Sasakian structures on S 2n+1 belonging to distinct,
non-standard contact structures.
If all the components of w are rational numbers, then the leaves characteristic
foliation Fξ are all circles and Fξ is equivalent to a locally free circle action by
Wadsley Theorem 2.6.12. In this case we can clear the denominators by multiplying
by their least common multiple and then redefine the circle action. So without loss
of generality we can assume that the wi are all integers satisfying gcd(w0 , . . . , wn ) =
1. Now the Sasakian structure Sw on S 2n+1 is quasi-regular and we have a natural
orbifold fibration π : S 2n+1 −−→P(w) where P(w) denotes the weighted projective
space CP(w) with its natural orbifold structure as discussed in Section 4.5.
Exercise 7.2: Show that for the structures Sw of Example 7.1.12 the conjugation
c
map Sw 7→ Sw is induced by an isometry of gw .

7.2. The Transverse Geometry of the Characteristic Foliation


7.2.1. Basic Cohomology of K-Contact Structures. Recall the basic de

Rham complex ΩB (F) of a foliation and its basic cohomology ring HB (F) from
Section 2.4. As seen previously a K-contact structure has a taut Riemannian flow,
and in this case the spectral sequence of the complex simplifies. Our treatment
follows [Ton97]. As seen in Proposition 2.6.3 the closure of the leaves of the
characteristic foliation Fξ is a torus T ⊂ Aut(ξ, η, Φ, g) ⊂ Isom(M, g). Let Ω(M )T
denote the subalgebra of Ω(M ) consisting of T-invariant forms. Since the orbit of ξ
is dense in T any basic form is T-invariant, so we have ΩB (F) ⊂ Ω(M )T . We have
Proposition 7.2.1: There is an exact sequence of complexes
ξ
0−−−→Ω∗B (Fξ )−−−→Ω∗ (M )T −−−→ Ω∗−1
B (Fξ )−
−−→0.
Proof. First we show that ξ α is basic for any T-invariant form α. This
means the two conditions ξ (ξ α) = 0 and ξ d(ξ α) are satisfied. The first is
an identity while the second follows from £ξ (ξ α) = 0, since ξ ∈ T. Furthermore,
up to sign ξ is a map of complexes. Next we show that ξ is surjective. Let
α ∈ Ωr−1
B (Fξ ), then ξ (η ∧ α) = α, and if T is an element of the Lie algebra t of
T, we have
£T (η ∧ α) = (£T η) ∧ α + η ∧ £T α = 0.
The first term vanishes since T ∈ t ⊂ a(M, ξ, η, Φ, g), and the second term vanishes
since basic forms are T-invariant. This proves the surjectivity. Finally, exactness
of the middle term follows from the definitions of ΩB (Fξ ) and Ω(M )T and the fact
that ξ ∈ t. ¤
Now the cohomology of the complex Ω(M )G for any compact connected Lie
group is isomorphic to the usual de Rham cohomology [GHV73], so we get a long
7.2. THE TRANSVERSE GEOMETRY OF THE CHARACTERISTIC FOLIATION 215

exact cohomology sequence


(7.2.1)
ι∗ jp δ
p p−1 p+1
· · · −−−→HB (Fξ ) −−−→ H p (M, R) −−−→ HB (Fξ ) −−−→ HB (Fξ )−−−→ · · · ,
where δ is the connecting homomorphism given by δ[α]B = [dη ∧ α]B = [dη]B ∪
[α]B , and jp is the composition of the map induced by ξ with the isomorphism
H r (Ω(M )T ) ≈ H r (M, R) mentioned above. This sequence is of fundamental im-
portance in the study of Sasakian geometry. It is the foliation analogue of the
rational or real Gysin sequence for the quasi-regular case such as in Proposition
4.7.9. In fact, we have
Proposition 7.2.2: Let (M, ξ, η, Φ, g) be a quasi-regular K-contact manifold with
π : M −−→M/Fξ an orbifold Riemannian submersion. Then π induces a ring iso-
morphism

π ∗ : H ∗ (M/Fξ , R) −−→ HB

(Fξ ) .
Exercise 7.3: Prove Proposition 7.2.2.
On compact K-contact manifolds the basic cohomology groups enjoy some
special properties. In particular, there is a transverse Hodge theory [EKAH86,
KT87, Ton97] which we shall briefly review. We begin by defining the transverse
Hodge star operator ¯?. It is defined in terms of the usual Hodge star by
(7.2.2) ¯?α = ?(η ∧ α) = (−1)r ξ ?α .
The adjoint δB : ΩrB (Fξ )−−→Ωr−1
B (Fξ ) of dB is defined by
(7.2.3) δB = −¯?dB ¯? .
The basic Laplacian ∆B is defined in terms of dB and its adjoint δB by
(7.2.4) ∆B = dB δB + δB dB .
r
The space HB (Fξ ) of basic harmonic r-forms is then defined to be the kernel of
∆B : ΩB (Fξ )−−→ΩrB (Fξ ), or alternatively HB
r r
(Fξ ) = ker dB ∩ kerδB . The transverse
Hodge theorem [EKAH86] then says that each basic cohomology class has a unique
r r
harmonic representative, i.e., HB (Fξ ) = HB (Fξ ). From transverse Hodge theory
we get
Proposition 7.2.3: Let (M, ξ, η, Φ, g) be a compact K-contact manifold of dimen-
sion 2n + 1. Then we have
r r
(i) The groups HB (Fξ ) are finite dimensional and HB (Fξ ) = 0 for r > 2n.
2n
(ii) HB (Fξ ) ≈ R.
2
(iii) The class [dη]B ∈ HB (Fξ ) is non-trivial.
2r
(iv) HB (Fξ ) > 0 for 0 ≤ r ≤ n.
(v) H 1 (M, R) ≈ HB 1
(Fξ ).
(vi) (transverse Poincaré duality) There is a non-degenerate pairing
Ψ : H r (Fξ ) ⊗ H 2n−r (Fξ )−−−→R .
Proof. (i) follows from transverse Hodge theory in the usual way. (ii): With
p = 2n + 1 the exact sequence 7.2.1 gives an isomorphism
j2n+1
(7.2.5) H 2n+1 (M ) 2n
≈ HB (Fξ ),

proving (ii). To prove (iii) we notice that the 2-form dη satisfies η ∧ (dη)n 6= 0,
and that the isomorphism j2n+1 in (7.2.5) (which is essentially ξ ) implies that dη
216 7. K-CONTACT AND SASAKIAN STRUCTURES

2
defines a non-trivial class in HB (Fξ ). Clearly, (iv) holds since (dη)n ∈ HB
2n
(Fξ ) is
non-trivial. To prove (v) we notice that the beginning of the exact sequence 7.2.1
is
j1 δ0
1
0−−→HB 2
(Fξ )−−→H 1 (M, R) −−→ R −−→ HB (Fξ ).
By (iii) δ0 is injective, so {0} = ker δ0 = im j1 ; hence, j1 is the zero map. (vi)
follows by transverse Hodge theory [KT87, Ton97]. The pairing Ψ is that induced
from the pairing
Ψ̃ : ΩrB (Fξ ) ⊗ Ω2n−r
B (Fξ )−−−→R
defined by
Z
(7.2.6) Ψ̃(α, β) = η ∧ (α ∧ β) .
M

This induces an isomorphism on the transverse harmonic forms, and defines a


non-degenerate inner product hα, βiB = Ψ̃(α, ¯?β). See [KT87, Ton97] for details.
¤

2
Remark 7.2.1: The class [dη]B ∈ HB (Fξ ) is called the fundamental basic class of
the K-contact structure (ξ, η, Φ, g), and if M is compact [dη]B 6= 0.
Next we consider some fundamental invariants of K-contact structures [BGN03a].
We define the basic Betti numbers

(7.2.7) bB r
r (Fξ ) = dim HB (Fξ ),

and the basic Euler characteristic of Fξ given by


2n
X
(7.2.8) χ(Fξ ) = (−1)r bB
r (Fξ ).
r=0

Notice that by (v) of Proposition 7.2.3 b1 (Fξ ) = b1 (M ), a topological invariant of


M. Using transverse Poincaré duality we have
Proposition 7.2.4: On a compact K-contact manifold M 2n+1 we have

bB B
2n−r (Fξ ) = br (Fξ ), χ(Fξ ) = 2 − 2bB B n B
1 (Fξ ) + 2b2 (Fξ ) − · · · + (−1) bn .

We end this section with a result from [EKAN93] which implies that the basic
cohomology ring is an invariant of a K-contact structure.
Theorem 7.2.5: Let M, M 0 be manifolds of dimension 2n + 1 with complete K-
contact structures S = (ξ, η, Φ, g) and S 0 = (ξ 0 , η 0 , Φ0 , g 0 ), respectively. Suppose
further that there exists a diffeomorphism f : M −−→M 0 sending S to S 0 , i.e., such
that
ξ 0 = f∗ ξ, η 0 = (f −1 )∗ η, Φ0 ◦ f∗ = f∗ Φ, g 0 = (f −1 )∗ g .
Then, f induces a ring isomorphism f ∗ : HB
∗ ∗
(Fξ0 )−−→HB (Fξ ).
Actually in Theorem 7.4.14 below we shall see that for any Sasakian structures
on diffeomorphic manifolds, the basic cohomology groups are isomorphic. However,
it says nothing about the ring structure.
7.2. THE TRANSVERSE GEOMETRY OF THE CHARACTERISTIC FOLIATION 217

7.2.2. Basic Cohomology of Sasakian Structures. The basic cohomology


p,q
groups HB (Fξ ) are fundamental invariants of a Sasakian structure which enjoy
many of the same properties as the ordinary Dolbeault cohomology of a Kähler
structure. Here we explore some of the elementary properties.
Theorem 7.2.6: Let (M, S) be a compact Sasakian manifold of dimension 2n + 1.
Then we have
n,n
(i) HB (Fξ ) ≈ R.
1,1
(ii) The class [dη]B 6= 0 lies in HB (Fξ ).
p,p
(iii) HB (Fξ ) > 0.
2p+1
(iv) HB (Fξ ) has even dimension for p < [n/2]. L
r p,q
(v) (Transverse Hodge Decomposition) HB (Fξ ) = p+q=r HB (Fξ ).
p,q
(vi) Complex conjugation induces an anti-linear isomorphism HB (Fξ ) ≈
q,p
HB (Fξ ).
(vii) (Transverse Serre Duality) There is an isomorphism
H p,q (Fξ ) ≈ H n−p,n−q (Fξ ) .
Proof. The proof is very similar to the standard case [EKA90]. As usual we
define the operator L : Ωr (Fξ )−−→Ωr+2 (Fξ ) by
(7.2.9) Lα = α ∧ dη ,
and its adjoint Λ : Ωr (Fξ )−−→Ωr−2 (Fξ ) by
(7.2.10) Λ = −¯?L¯? .
We also have the adjoint operators
∂¯∗ : Ωp,q (Fξ )−→Ωp,q−1 (Fξ )
and
∂ ∗ : Ωp,q (Fξ )−→Ωp−1,q (Fξ )
¯ respectively, defined by
of ∂ and ∂,
(7.2.11) ∂¯∗ = −¯?∂¯? , ¯? .
∂ ∗ = −¯?∂¯
As usual, we get the following
Lemma 7.2.7: On a Sasakian manifold the following hold:
(i) [Λ, ∂] = −i∂¯∗ ,
¯ = i∂ ∗ ,
(ii) [Λ, ∂]
(iii) ∂ ∂ + ∂¯∗ ∂ = 0, ∂ ∗ ∂¯ + ∂∂
¯∗ ¯ ∗ = 0,
∗ ∗ ¯ ¯∗
(iv) ∂∂ + ∂ ∂ = ∂ ∂ + ∂ ∂. ¯ ∗¯

If we define
(7.2.12) ∆B = ∂∂ ∗ + ∂ ∗ ∂ , ∆B ¯ ¯∗ ¯∗ ¯
¯ = ∂∂ + ∂ ∂ ,
∂ ∂
then
Lemma 7.2.8: On a Sasakian manifold one has
∆B = 2∆B
∂¯ .
The theorem easily follows from Lemmas 7.2.7 and 7.2.8. ¤
Remark 7.2.2: Item (vii) of Theorem 7.2.6 holds without the Kählerian hypothesis.
Thus, it holds for normal almost contact structures such that N (3) = 0.
218 7. K-CONTACT AND SASAKIAN STRUCTURES

Theorem 7.2.6 gives rise to fundamental invariants of the Sasakian structure.


Now we have the basic Hodge numbers
(7.2.13) hp,q p,q
B (Fξ ) = dim HB (Fξ )

which are related to the basic Betti numbers by


X
(7.2.14) bB
r (Fξ ) = hp,q (Fξ ) ,
p+q=r

and satisfy
(7.2.15) hp,q q,p
B (Fξ ) = hB (Fξ ) , hp,q n−p,n−q
B (Fξ ) = hB (Fξ ) .
In addition to the basic Euler characteristic on a Sasakian manifold we define the
basic holomorphic Euler characteristic by
n
X
(7.2.16) χhol (Fξ ) = (−1)q h0,q
B (Fξ ) .
q=0

There is also a transverse Lefschetz decomposition [EKA90]. Consider the


operator L : ΩrB (Fξ )−−→Ωr+2 r
B (Fξ ) of equation (7.2.9), and let P (Fξ ) denote the
so-called basic primitive cohomology group, that is, the kernel of the map Λ :
r r−2 r
HB (Fξ )−−→HB (Fξ ) defined by Equation 7.2.10. Notice that as a subset of HB (Fξ ),
each element of P r (Fξ ) has a unique harmonic representative. Thus, the transverse
Hodge Theorem [EKAH86] allows us to define the set P r (Fξ ) of primitive basic
harmonic forms. Now the Sasakian version of the transverse Hard Lefschetz The-
orem [EKA90] is:
Theorem 7.2.9: On a compact Sasakian manifold (M, S) of dimension 2n + 1 the
n−k n+k
map Lk : HB (Fξ )−−→HB (Fξ ) is an isomorphism for 1 ≤ k ≤ n. Furthermore,
M
r
HB (Fξ ) = Lk P r−2k (Fξ ) .
k≥0

We have the following useful corollary of Theorem 7.2.9:


Corollary 7.2.10: On a compact Sasakian manifold (M, S) of dimension 2n + 1
we have an isomorphism
r r−2
HB (Fξ ) ≈ P r (Fξ ) ⊕ LHB (Fξ )
for 0 ≤ r ≤ n. Moreover, since L is injective for this range of r we also have
bB B B B r
r = pr + br−2 , where pr = dim P (Fξ ).
If the dimension of M is 4n+1 we can also define the basic Hirzebruch signature
τB (Fξ ) to be the signature of the bilinear form defined by the cup product on the
2n
middle basic cohomology group HB (Fξ ). Using the transverse Lefschetz Theorem
7.2.9 [EKA90] one can obtain the usual formula:
X X
(7.2.17) τB (Fξ ) = (−1)p hp,q
B (Fξ ) = (−1)p hp,q
B (Fξ ) .
p,q p≡q(2)

As is usual in complex geometry we introduce the basic geometric genus pg (Fξ ) =


h0,n
B (Fξ ), the basic arithmetic genus pa (Fξ ) = (−1)n (χhol (Fξ ) − 1), and the basic
irregularity q = q(S) = h0,1 1 B 1
B . We remark that q = 2 b1 (Fξ ) = 2 b1 (M ) is actually a
topological invariant by (v) of Proposition 7.2.3. In the case n = 2, things simplify
7.3. CURVATURE PROPERTIES OF K-CONTACT AND SASAKIAN STRUCTURES 219

much more. All the basic Hodge numbers are given in terms of the three invariants
q, pg (Fξ ) and h1,1
B (Fξ ), and we have the relations

χB (Fξ ) = 2 + 2pg (Fξ ) − 4q + h1,1


B (Fξ ), τB (Fξ ) = 2 + 2pg (Fξ ) − h1,1
B (Fξ ),
χB (Fξ ) + τB (Fξ ) = 4χhol (Fξ ).
The fact that q is actually a topological invariant has some nice consequences for
n = 2.

7.3. Curvature Properties of K-Contact and Sasakian Structures


For any geometry it is important to study the curvature properties of a preferred
connection. In our case this connection is almost always the Levi-Civita connection,
but other important connections can and do occur. Nevertheless, our development
here concerns the curvature of the Levi-Civita connection.
7.3.1. Some Elementary Properties of Contact Metric Structures.
Here we develop some useful formulas for K-contact and Sasakian structures. We
begin by proving a general expression for the covariant derivative of Φ for an almost
contact metric structure.
Lemma 7.3.1: Let (ξ, η, Φ, g) be an almost contact metric structure on M and
define the 2-form Ω by Ω(X, Y ) = g(X, ΦY ). Then the following formula holds:
2g((∇X Φ)Y, Z) = 3dΩ(X, ΦY, ΦZ) − 3dΩ(X, Y, Z) + g(N (1) (Y, Z), ΦX)
+ N (2) (Y, Z)η(X) + 2dη(ΦY, X)η(Z) − 2dη(ΦZ, X)η(Y )
for any vector fields X, Y, Z on M.
Proof. The formula follows by a straightforward albeit tedious computation
from the well-known formulas for the covariant and exterior derivative, namely
2g(∇X Y, Z) = Xg(Y, Z) + Y g(X, Z) − Zg(X, Y )
+ g([X, Y ], Z) + g([Z, X], Y ) − g([Y, Z], X) ,
and
3dΩ(X, Y, Z) = XΩ(Y, Z) + Y Ω(Z, X) + ZΩ(X, Y )
− Ω([X, Y ], Z) + Ω([Z, X], Y ) − Ω([Y, Z], X) .
The details are left as an exercise. ¤
Lemma 7.3.1 easily simplifies in the case of a contact metric structure with its
associated metric.
Lemma 7.3.2: Let (ξ, η, Φ, g) be a contact metric structure on M. Then we have
2g((∇X Φ)Y, Z) = g(N (1) (Y, Z), ΦX) + 2dη(ΦY, X)η(Z) − 2dη(ΦZ, X)η(Y ) .
In particular, ∇ξ Φ = 0.
Lemma 7.3.3: For a contact metric structure we have
(i) N (3) is a symmetric endomorphism in the sense that
g(N (3) (X), Y ) = g(N (3) (Y ), X).
(ii) ∇ξ = −Φ − 12 ΦN (3) .
(iii) Φ ◦ N (3) + N (3) ◦ Φ = 0.
(iv) N (3) (ξ) = 0, η ◦ N (3) = 0.
220 7. K-CONTACT AND SASAKIAN STRUCTURES

(v) tr N (3) = 0.
Proof. To prove (i) we first compute
(3)
N (X) = (£ξ Φ)(X) = [ξ, ΦX] − Φ[ξ, X] = ∇ξ (ΦX) − ∇ΦX ξ − Φ(∇ξ X − ∇X ξ)
= (∇ξ Φ)(X) + Φ∇ξ X − ∇ΦX ξ − Φ∇ξ X + Φ∇X ξ = Φ∇X ξ − ∇ΦX ξ ,
where we have used the last statement of Lemma 7.3.2. Thus, N (3) (ξ) = Φ∇ξ ξ
which vanishes by Lemma 2.6.8. Next we claim that g(N (3) (X), ξ) = η(N (3) (X)) =
0. From above we have η(N (3) (X)) = −η(∇ΦX ξ) which vanishes since ∇ is a metric
connection. This proves (ii) when X or Y is ξ. Now assume that X and Y are
orthogonal to ξ. Continuing our computation above
g(N (3) (X), Y ) = g(Φ∇X ξ, Y ) − g(∇ΦX ξ, Y ) = −g(∇X ξ, ΦY ) + g(ξ, ∇ΦX Y )
= η(∇X ΦY ) + η(∇ΦX Y ) ,
and hence,
g(N (3) (Y ), X) = η(∇Y ΦX) + η(∇ΦY X) .
But it is not difficult to see that the equation N (2) (X, Y ) = 0 implies that these
two expressions are equal. This proves (i).
To prove (ii) we consider the equation of Lemma 7.3.2 with Y = ξ, and X and
Z orthogonal to ξ. We have
2g((∇X Φ)ξ, Z) = g(N (1) (ξ, Z), ΦX) − 2dη(ΦZ, X)
= g(Φ2 [ξ, Z] − Φ[ξ, ΦZ], ΦX) − 2g(ΦZ, ΦX)
= −g(Φ(£ξ Φ)Z, ΦX) − 2g(Z, X)
= −g((£ξ Φ)Z, X) + η((£ξ Φ)Z)η(X) − 2g(Z, X) .
Therefore,
−2g(Φ∇X ξ, Z) = −g(N (3) (Z), X) − 2g(X, Z) = −g(N (3) (X), Z) − 2g(X, Z) .
Since this holds for arbitrary Z we have 2Φ∇X ξ = N (3) (X) + 2X. Applying Φ to
this equation proves (ii).
Both (iii) and (iv) are a result of Lemma 6.3.7. Finally, (v) holds since by (iii)
if λ is an eigenvalue of the symmetric operator N (3) so is −λ. ¤
Combining Lemma 7.3.3 and Proposition 6.5.12 we have
Proposition 7.3.4: A contact metric structure (ξ, η, Φ, g) is K-contact if and only
if ∇ξ = −Φ.
For a K-contact structure (ξ, η, Φ, g) the characteristic foliation Fξ is Riemann-
ian, so we have the O’Neill tensors T, A at our disposal. First, one easily sees that
T = 0, and
1 1
(7.3.1) AX Y = π([X, Y ]) = η([X, Y ])ξ = −dη(X, Y )ξ .
2 2
Indeed,
Lemma 7.3.5: For a K-contact manifold the following identities hold:
(i) T = 0,
(ii) AX Y = −dη(X, Y )ξ,
(iii) AX ξ = −ΦX,
(iv) |A|2 = 2n.
7.3. CURVATURE PROPERTIES OF K-CONTACT AND SASAKIAN STRUCTURES 221

Proof. It only remains to prove (iii) and (iv). We have that

g(AX ξ, Y ) = −g(ξ, AX Y ) = −η(AX Y ) = dη(X, Y ) = −g(ΦX, Y )

holds for all horizontal Y. This proves (iii) since AX ξ is horizontal. To prove (iv)
we let {Ei }2n
i=1 denote a local orthonormal frame for D, and using (iii) we compute

2n
X 2n
X 2n
X
2
|A| = g(AEi ξ, AEi ξ) = g(ΦEi , ΦEi ) = g(Ei , Ei ) = 2n . ¤
i=1 i=1 i=1

We can write A in the succinct form

(7.3.2) AX = −ΦX ⊗ η − ξ ⊗ X dη .

Combining Lemmas 2.5.15 and 7.3.5 we have


Lemma 7.3.6: For a K-contact manifold the following identities hold:
(i) (∇ξ A)ξ = 0 ,
(ii) (∇X A)ξ = AΦX ,
(iii) η((∇ξ A)X ξ) = 0 .
We can obtain more information about the behavior of the covariant derivatives
of A from Theorem 2.5.16, but first we need to investigate curvature properties of
K-contact manifolds.

7.3.2. Curvature and K-Contact Structures. Here we collect some useful


results concerning the Riemannian curvature R(X, Y ) of K-contact manifolds. We
begin with the fundamental result of [HOT63]:
Theorem 7.3.7: A Riemannian manifold (M, g) is K-contact if and only if (M, g)
admits a Killing vector field ξ of unit length such that the sectional curvature of all
plane sections containing ξ equals one.
First we obtain a general expression for R(X, ξ)Y for a metric contact structure.
Lemma 7.3.8: For a contact metric structure (ξ, η, Φ, g) we have for any X

1 1
R(X, ξ)ξ = X − η(X)ξ − (N (3) )2 (X) + Φ(∇ξ N (3) )(X).
4 2
Moreover, if (ξ, η, Φ, g) is K-contact manifold then for any vector field X and for
Y is a section of D the following identities hold:
(i) R(X, ξ)ξ = X − η(X)ξ ,
(ii) R(X, ξ)Y = −(∇X Φ)(Y ) .

Proof. We compute

R(X, ξ)Y = ∇X ∇ξ Y − ∇ξ ∇X Y − ∇[X,ξ] Y


= ∇X (∇Y ξ + [ξ, Y ]) − ∇∇X Y ξ − [ξ, ∇X Y ] − ∇[X,ξ] Y
(7.3.3) = ∇X ∇Y ξ − ∇∇X Y ξ − (£ξ ∇)X Y .
222 7. K-CONTACT AND SASAKIAN STRUCTURES

Putting Y = ξ in this equation and using Lemmas 7.3.2 and 7.3.3 we have
¡ ¢
R(X, ξ)ξ = −∇∇X ξ ξ − [ξ, ∇X ξ] − ∇[ξ,X] ξ
1
= −∇−ΦX− 12 ΦN (3) (X) ξ − [ξ, −ΦX − ΦN (3) (X)] − Φ[ξ, X]
2
1
− ΦN (3) ([ξ, X]))
2
1 1 1
= −Φ(ΦX + ΦN (3) (X)) − ΦN (3) (ΦX + ΦN (3) (X)) + N (3) (X)
2 2 2
1 (3) 2 1 (3)
+ (N ) (X) + Φ(£ξ N )(X)
2 2
1 (3) 2 1
= X − η(X)ξ + (N ) (X) + N (3) (X) + Φ(£ξ N (3) )(X) .
4 2
Now it is easy to verify the formula
(7.3.4) ∇ξ N (3) = £ξ N (3) − 2ΦN (3) − Φ(N (3) )2 .
By applying Φ to this formula and substituting it into the expression for R(X, ξ)ξ
above proves the first statement. (i) follows immediately from the first statement
by Proposition 6.5.12. To prove (ii) we have from above
R(X, ξ)Y = ∇X ∇Y ξ − ∇∇X Y ξ − (£ξ ∇)X Y
which for a K-contact structure becomes
R(X, ξ)Y = −∇X ΦY + Φ∇X Y = −(∇X Φ)(Y )
and proves (ii). ¤
Exercise 7.4: Prove formula 7.3.4 above.
Proof of Theorem 7.3.7. Suppose first that (M, ξ, η, Φ, g) is K-contact. Then
ξ is a unit length Killing vector field, and using Lemma 7.3.8 we have for X or-
thogonal to ξ,
g(R(X, ξ)ξ, X) g(X, X)
K(X, ξ) = = = 1.
g(X, X) g(X, X)
Conversely, suppose that ξ is a unit length Killing vector field. We define η
and Φ by
η(X) = g(ξ, X), ΦX = −∇X ξ .
Then as in the proof of Lemma 2.6.8 we have Φξ = −∇ξ ξ = 0. Now since ξ is a
Killing vector field it leaves ∇ invariant, so we have from above
R(X, ξ)ξ = −∇∇X ξ ξ = ∇ΦX ξ = −Φ2 X .
This gives for any vector field X orthogonal to ξ,
g(R(X, ξ)ξ, X) g(Φ2 X, X)
1 = K(X, ξ) = =− .
g(X, X) g(X, X)
Since this must hold for all vector fields X orthogonal to ξ, we must have Φ2 X = −X
for all vector fields X orthogonal to ξ. Thus, Φ2 X = −X + η(X)ξ for arbitrary X.
Moreover, since ξ is a Killing and ∇ is metric and torsion-free we have
2dη(X, Y ) = Xη(Y ) − Y η(X) − η([X, Y ]) = Xg(ξ, Y ) − Y g(ξ, X) − g(ξ, [X, Y ])
= g(∇X ξ, Y ) + g(ξ, ∇X Y ) − g(∇Y ξ, X) − g(ξ, ∇Y X) − g(ξ, [X, Y ])
= g(∇X ξ, Y ) − g(∇Y ξ, X) = 2g(∇X ξ, Y ) = −2g(ΦX, Y ) = 2g(X, ΦY ) .
7.3. CURVATURE PROPERTIES OF K-CONTACT AND SASAKIAN STRUCTURES 223

This proves that the structure (ξ, η, Φ, g) is K-contact. ¤

Next we consider the Ricci curvature


X ¡ ¢ X ¡ ¢
(7.3.5) Ric(X, Y ) = g R(X, Ei )Ei , Y = g R(Ei , X)Y, Ei ,
i i

where Ei is a local orthonormal frame. We have a result of Blair [Bla77]:


Proposition 7.3.9: A contact metric structure (ξ, η, Φ, g) on a manifold of dimen-
sion 2n + 1 is K-contact if and only if Ric(ξ, ξ) = 2n.

Proof. Let {E0 = ξ, E1 , . . . , E2n } be a local orthonormal frame, then we have


from Lemma 7.3.8
2n
X ¡ ¢
Ric(ξ, ξ) = g R(Ei , ξ)ξ, Ei
i=1
2n ³
2n
X X ¡1 ¢ 1 ¡ ¢´
= g(Ei , Ei ) − g (N (3) )2 (Ei ), Ei − g Φ(∇ξ N (3) )(Ei ), Ei
i=1 i=1
4 2

1 X³ ¡ (3) ¢´
2n
¢ ¡
= 2n − g N (Ei ), N (3) (Ei ) − 2g Φ(∇ξ N (3) )(Ei ), Ei .
4 i=1
Now since on a contact metric manifold X and ΦX are orthogonal, we can take the
local orthonormal frame to be of the form {ξ, E1 , . . . , En , ΦE1 , . . . , ΦEn }. Then we
have
2n
1 X ¡ (3) ¢
Ric(ξ, ξ) = 2n − g N (Ei ), N (3) (Ei )
4 i=1
n n
1X ¡ ¢ 1X ¡ ¢
− g Φ(∇ξ N (3) )(Ei ), Ei − g Φ(∇ξ N (3) )(ΦEi ), ΦEi .
2 i=1 2 i=1
We claim that the last two terms cancel so that we have
2n
1X 1
(7.3.6) Ric(ξ, ξ) = 2n − (g(N (3) (Ei ), N (3) (Ei )) = 2n − tr (N (3) )2 .
4 i=1 4
Since each term under the sum is positive, this proves the result. ¤

Exercise 7.5: Prove the following formula for contact metric structures
1
ΦR(ΦX, ξ)ξ = R(X, ξ)ξ + 2Φ2 X + (N (3) )2 (X) ,
2
and use it to prove the claim above verifying equation (7.3.6).
Next we relate the curvature for K-contact manifolds to the O’Neill tensor field
A. First we notice that the first two equations and the fourth equation of Theorem
2.5.16 are vacuous in the K-contact case. In what follows X, Y, Z, Z 0 are always
sections of D.
Lemma 7.3.10: On a K-contact manifold (M, ξ, η, Φ, g) the following identities
hold:
¡ ¢
(i) η ¡(∇ξ A)X Y ¢ = 0, ¡ ¢
(ii) η (∇Z A)X Y = −g (∇Z Φ)(X), Y ,
224 7. K-CONTACT AND SASAKIAN STRUCTURES

¡ ¢ ¡ ¢
(iii) g R(X, Y )Z, Z 0 = g RT (X, Y )Z, Z 0 + 2dη(X, Y )dη(Z, Z 0 )
+ dη(X, Z)dη(Y, Z 0 ) − dη(X, Z 0 )dη(Y, Z) .
Proof. (i): The third equation of Theorem 2.5.16 gives the identity
¡ ¢ ¡ ¢ ¡ ¢
g R(X, ξ)Y, ξ = −g AX ξ, AY ξ − g (∇ξ A)X Y, ξ
¡ ¢ ¡ ¢ ¡ ¢
= −g ΦX, ΦX − g (∇ξ A)X Y, ξ = −g(X, Y ) − g (∇ξ A)X Y, ξ .
Using the well-known Riemann curvature identities and Lemma 7.3.8 the left hand
side of this equation reduces as
¡ ¢ ¡ ¢
g R(X, ξ)Y, ξ = −g R(X, ξ)ξ, Y = −g(X, Y ) .
This proves (i). To prove (ii) we notice that the fifth equation of Theorem 2.5.16
becomes ¡ ¢ ¡ ¢ ¡ ¢
g R(X, Y )Z, ξ = g (∇Z A)X Y, ξ = η (∇Z A)X Y .
Again by the Riemann curvature identities and Lemma 7.3.8, the left hand side is
¡ ¢ ¡ ¢ ¡ ¢
g R(X, Y )Z, ξ = g R(Z, ξ)X, Y = −g (∇Z Φ)(X), Y
which proves (ii). Then (iii) follows immediately by (vi) of Theorem 2.5.16 and (ii)
of Lemma 7.3.5. ¤
From Theorem 2.5.18 one obtains relations between the Ricci curvature of g
and the transverse Ricci curvature for K-contact manifolds. But first we need a
lemma.
Lemma 7.3.11: Let (ξ, η, Φ, g) be a contact metric structure on M, and consider
the local orthonormal frame {ξ, E1 , . . . , En , ΦE1 , . . . , ΦEn }. Then the following re-
lation holds:
2n
X ¡ ¢
g (∇Ei Φ)(Ei ), X = 2nη(X) ,
i=1
where En+i = ΦEi .
Proof. From Lemma 7.3.2 we get
¡ ¢ ¡ ¢
(7.3.7) 2g (∇Ei Φ)Ei , X = g (NΦ (Ei , X), ΦEi + 2g(Ei , Ei )η(X) .
Now it is easy to see that
NΦ (X, Ei+n ) = −ΦNΦ (X, Ei ) − η([Ei , ΦX])ξ .
Using this equation and summing on i in Equation 7.3.7 gives the result. ¤
We now have
Theorem 7.3.12: Let (M, ξ, η, Φ, g) be a K-contact manifold and let RicT denote
the Ricci curvature of the transverse metric gT . Then the following identities hold:
(i) Ric(X, ξ) = 2nη(X) for any vector field X,
(ii) Ric(X, Y ) = RicT (X, Y ) − 2g(X, Y ), where X, Y are sections of D.
Proof. To prove (i) we see that the second equation of Theorem 2.5.18 gives
¡ ¢
Ric(X, ξ) = −g (δ T A)X, ξ .
So choosing a local orthonormal basis {Ei }i of D as above, we compute using (ii)
of Lemma 7.3.10
X ¡ ¢ X ¡ ¢
g(δ T A)X, ξ) = g (∇Ei A)Ei )(X), ξ = − g (∇Ei Φ)(Ei ), X .
i i
7.3. CURVATURE PROPERTIES OF K-CONTACT AND SASAKIAN STRUCTURES 225

So applying Lemma 7.3.11 implies (i). For (ii) we only need to compute g(AX , AY )
using Equation 2.5.23
g(AX , AY ) = g(AX ξ, AY ξ) = g(ΦX, ΦY ) = g(X, Y ) . ¤
For sectional curvatures of a K-contact structure we have
Lemma 7.3.13: Let X, Y be smooth sections of D of unit length. Then the sec-
tional curvatures satisfy
(i) K(X, ξ) = 1,
(ii) K(X, Y ) = K T (X, Y ) − 3(dη(X, Y ))2 .
Proof. This follows immediately from Theorem 7.3.7, Corollary 2.5.17, and
Lemma 7.3.5. ¤
We are also interested in the so-called Φ-sectional curvatures which play a
role in Sasakian geometry analogous to the holomorphic sectional curvatures in
Kählerian geometry.
Definition 7.3.14: The Φ-sectional curvature H(X) of a K-contact manifold
is defined by H(X) = K(X, ΦX), where K is the ordinary sectional curvature.
Then an immediate consequence of Lemma 7.3.13 is:
Lemma 7.3.15: Let X be a section of D of unit length. Then
H(X) = K T (X, ΦX) − 3 .
In the next section we shall see that in the case of Sasakian manifolds, the
Φ-sectional curvature determines the full curvature tensor R.
7.3.3. Curvature Properties of Sasakian Structures. Now we give nec-
essary and sufficient condition on the covariant derivative of Φ for an almost contact
structure to be Sasakian.
Theorem 7.3.16: An almost contact metric manifold (M, ξ, η, Φ, g) is Sasakian if
and only if
∇X Φ = ξ ⊗ X g − X ⊗ η .
Proof. Let (M, g) be embedded in its cone (C(M ), dr2 +r2 g) as the hypersur-
face r = 1. Then the almost contact metric structure (ξ, η, Φ, g) is Sasakian if and
only if the cone metric ḡ = dr2 + r2 g is Kähler with respect to complex structure
given by equation (6.5.2). Furthermore, the complex structure I is Kähler with
respect to the metric ḡ if and only if it is parallel with respect to the Levi-Civita
connection ∇ ¯ of ḡ. Now for any vector fields X, Y on M we have, using Gauss’
¯
formula ∇X Y = ∇X Y + g(sX, Y )Ψ and Weingarten’s equation ∇ ¯ X Ψ = −sX,
0 = (∇¯ X I)(Y ) = ∇
¯ X IY − I ∇
¯ XY = ∇ ¯ X (ΦY + η(Y )Ψ) − (Φ + Ψ ⊗ η)∇ ¯ XY
= ∇X ΦY + g(sX, ΦY )Ψ + Xη(Y )Ψ − η(Y )sX
− Φ∇X Y − g(sX, Y )IΨ − η(∇X Y )Ψ
¡ ¢
= (∇X Φ)(Y ) − g(X, Y )ξ + η(Y )X + Xη(Y ) − g(X, ΦY ) − η(∇X Y ) Ψ .
Noting that the normal terms give an identity for any almost contact metric struc-
ture, one sees that the tangential term gives the result. Notice that we have used
the easily verified fact that the ‘shape operator’ s equals −1l for a hypersurface M
embedded metrically in the cone C(M ) as r = 1. ¤
226 7. K-CONTACT AND SASAKIAN STRUCTURES

Exercise 7.6: Show that the normal terms in the above equation vanish identically
on any almost contact metric manifold.
Next we have a characterization of Sasakian structures in terms of curvature.
Proposition 7.3.17: Let (ξ, η, Φ, g) be a contact metric structure. Then the fol-
lowing are equivalent:
(i) (ξ, η, Φ, g) is Sasakian.
(ii) R(X, Y )ξ = η(Y )X − η(X)Y.
(iii) R(X, ξ)Y = η(Y )X − g(X, Y )ξ.

Proof. (i) ⇒ (ii) and (iii): We have


R(X, Y )ξ = ∇X ∇Y ξ − ∇Y ∇X ξ − ∇[X,Y ] ξ
¡ 1 ¢ ¡ 1 ¢
= −∇X ΦY + ΦN (3) (Y ) + ∇Y ΦX + ΦN (3) (X)
2 2
1
+ Φ[X, Y ] + ΦN (3) ([X, Y ])
2
1¡ ¢
= −(∇X Φ)(Y ) + (∇Y Φ)(X) − (∇X ΦN (3) )((Y ) − (∇Y ΦN (3) )(X) .
2
If (ξ, η, Φ, g) is Sasakian then N (3) = 0, and Theorem 7.3.16 implies (ii). That (i)
implies (iii) is an immediate consequence of Lemma 7.3.8 and Theorem 7.3.16.
Now assume that either (ii) or (iii) holds. Then we have R(X, ξ)ξ = X for any
X orthogonal to ξ. So using Exercise 7.5 we have
1 (3) 2
(N ) (X) = ΦR(ΦX, ξ)ξ − R(X, ξ)ξ + 2X = Φ2 X + X = 0.
2
But by Lemma 7.3.3 N (3) is symmetric and annihilates ξ, so this implies that N (3) =
0. Thus, (ξ, η, Φ, g) is K-contact. But (ii) of Lemma 7.3.8 says that R(X, ξ)Y =
−(∇X Φ)(Y ), so (iii) implies (i) by Theorem 7.3.16. To see that (ii) also implies (i)
we use the symmetries of the Riemann curvature tensor to give
¡ ¢ ¡ ¢ ¡ ¢ ¡ ¢
g (∇X Φ)(Y ), Z = −g R(X, ξ)Y, Z = g R(Y, Z)ξ, X = g η(Z)Y − η(Y )Z, X
¡ ¢
= g g(X, Y )ξ − η(Y )X, Z .
This holds for arbitrary Z so (ii) implies (i) again by Theorem 7.3.16. ¤

The proof of Theorem 7.3.16 suggests that we should look for conditions that
a real hypersurface in a Kähler manifold be Sasakian. We have
Theorem 7.3.18: A real hypersurface M in a Kähler manifold (N, I, ḡ) is Sasakian
if the shape operator s satisfies s = −1l + βξ ⊗ η for some smooth function β, with
ξ = −IV , V a unit normal vector field of M ,→ N and η the dual 1-form of ξ.
Conversely, if (M, S) is a Sasakian manifold embedded as a real hypersurface in N
then the shape operator s satisfies the above condition.

Proof. Let M be a (2n + 1)-dimensional hypersurface in a Kähler manifold


(N, I, ḡ) of complex dimensional n+1 with complex structure tensor I. Let g denote
the restriction of ḡ to M and ∇ ¯ the Levi-Civita connection on N. Let ∇ be the
Levi-Civita connection on M. Then the Gauss and Weingarten equations are
(7.3.8) ¯ X Y = ∇X Y + g(sX, Y )V,
∇ ¯ X V = −sX ,

7.3. CURVATURE PROPERTIES OF K-CONTACT AND SASAKIAN STRUCTURES 227

where X, Y are vector fields on M, and V is a unit normal field to M. Define


ξ = −IV and Φ by IX = ΦX + η(X)V where η is the dual 1-form to ξ. It is easy
to check that (ξ, η, Φ, g) defines an almost contact metric structure on M.
Now since I is parallel with respect to ∇, ¯ we have ∇ ¯ X IY = I ∇
¯ X Y. The left
hand side gives using equations (7.3.8)
¯ X IY = ∇
∇ ¯ X (ΦY + η(Y )V ) = ∇X ΦY + g(sX, ΦY )V + Xη(Y )V − η(Y )sX ,
while the right hand sign equals
¯ X Y = I(∇X Y + g(sX, Y )V ) = Φ∇X Y + η(∇X Y )V − g(sX, Y )ξ .
I∇
Equating the components tangent to M gives
(7.3.9) (∇X Φ)(Y ) = η(Y )sX − g(sX, Y )ξ .
By Theorem 7.3.16 M is Sasakian if and only if
(∇X Φ)(Y ) = g(X, Y )ξ − η(Y )X .
Equating this with Equation 7.3.9 and using the symmetry of the second funda-
mental form gives the solution
sX = −X + βη(X)ξ
for some smooth function β, which proves the result. ¤
Theorem 7.3.18 says that the embedding of a Sasakian hypersurface into a
Kähler manifold is umbilic on the contact subbundle. The principal curvatures are
all equal there. Another approach to the problem of embedded hypersurfaces in
Kähler manifolds is given by
Exercise 7.7: Let M be a real hypersurface embedded in a Kähler manifold N.
Show that
NΦ (X, Y ) + 2dη(X, Y )ξ = η(X)[Φ, s]Y − η(Y )[Φ, s]X .
Thus, (M, ξ, η, Φ, g) is Sasakian if [Φ, s] = 0.
Next we give a brief discussion of Φ-sectional curvatures which were defined in
Definition 7.3.14. Only the main results are given here. The proofs as well as more
detail are given in several books [Bla76a, Bla02, YK84]. All of these results have
strong analogs in Kählerian geometry.
Theorem 7.3.19: On a Sasakian manifold M the Φ-sectional curvature H deter-
mines the curvature. Moreover, assuming dim M ≥ 5, if the Φ-sectional curvature
H at any point is independent of the choice of Φ-section at the point, then it is
constant, say c, on M and the Riemannian curvature is given by
c + 3³ ´
R(X, Y )Z = g(Y, Z)X − g(X, Z)Y
4
c − 1³
+ η(X)η(Z)Y − η(Y )η(Z)X + g(X, Z)η(Y )ξ − g(Y, Z)η(X)ξ
4 ´
+ dη(Z, Y )ΦX − dη(Z, X)ΦY + 2dη(X, Y )ΦZ .
Notice that the case c = 1 greatly simplifies giving the formula for a Riemannian
metric of constant sectional curvature equal to 1. Thus, c = 1 gives the standard
round sphere.
Unlike Kähler geometry Sasakian geometry is not scale invariant. That is, if
g is a Sasakian metric and λ is a non-zero constant, λg is not Sasakian. There is,
228 7. K-CONTACT AND SASAKIAN STRUCTURES

however, a transverse homothety deformation or D-homothety in [YK84] which we


now describe. The earliest reference seems to be the thesis of Moskal [Mos66]. It
was used subsequently by Sasaki in his lecture notes [Sas68] and by Tanno [Tan68].
Consider the deformation of a contact metric structure S = (ξ, η, Φ, g) defined by
(7.3.10) ξ 0 = a−1 ξ, η 0 = aη, Φ0 = Φ, g 0 = ag + (a2 − a)η ⊗ η
for a ∈ R+ . We call this deformation a transverse homothety deformation. We shall
see later that this also occurs as part of what we call a-homology classes described
in Definition 7.5.13 below. It is easy to see that
Proposition 7.3.20: If S = (ξ, η, Φ, g) is a contact metric, K-contact, or Sasakian
structure, then so is S 0 = (ξ 0 , η 0 , Φ0 , g 0 ) determined by equations (7.3.10). Moreover,
S 0 belongs to the same underlying CR structure as S.
Exercise 7.8: First prove Proposition 7.3.20. Then show that if S = (ξ, η, Φ, g) is
the standard round sphere S 2n+1 with sectional curvature 1, the Sasakian structure
S 0 = (ξ 0 , η 0 , Φ0 , g 0 ) obtained by a transverse homothety deformation has constant
Φ-sectional curvature c = a4 − 3.
We denote the sphere S 2n+1 with constant Φ-sectional curvature equal to c by
2n+1
S (c). Note that in this case the range of c is −3 < c < ∞.
Exercise 7.9: Consider the transverse homothety of a Sasakian structure defined
in equation (7.3.10). Show that if ∇0 and ∇ denote the Levi-Civita connections of
g 0 and g, respectively, then the difference tensor A = ∇0 − ∇ is given by
A = −(a − 1)(Φ ⊗ η + η ⊗ Φ).
Show that the Riemann curvature tensor R0 for g 0 is related to that for g by
R0 (X, Y )Z = R(X, Y )Z + (∇X A)(Y, Z) − (∇Y A)(X, Z)
¡ ¢ ¡ ¢
+ A X, A(Y, Z) − A Y, A(X, Z) ,
and determine an explicit expression in terms of the Sasakian structure (ξ, η, Φ, g).
Show directly that the Ricci tensor of g 0 is given by
(7.3.11) Ricg0 = Ricg − 2(a − 1)g + 2(a − 1)(an + n + 1)η ⊗ η .
Exercise 7.10: Prove relation (7.3.11) using Theorem 7.3.12 as well as the trans-
formation formula for the scalar curvature,
(7.3.12) sg0 = a−1 (sg + 2n) − 2n .
Next we describe another natural Sasakian structure.
Example 7.3.21: Recall the standard
P contact structure on R2n+1 defined by equa-

tion (6.1.1), namely η = dz − i yi dxi . The Reeb vector field is given by ξ = ∂z .
Next we define
Xµ³ ∂ ∂ ´ ∂

Φ= + yi ⊗ dyi − ⊗ dxi .
i
∂xi ∂z ∂yi
Then the associated metric 6.4.2 becomes
X¡ ¢
(7.3.13) g= (dxi )2 + (dyi )2 + η ⊗ η .
i

Notice from (7.3.13) that the transverse metric gT is flat, so it follows from Lemma
7.3.15 that the Φ-sectional curvature H(X) = −3 for all sections X of D of unit
7.4. TOPOLOGY OF K-CONTACT AND SASAKIAN MANIFOLDS 229

length. We denote this Riemannian structure by R2n+1 (−3). This example is related
to the (2n+1)-dimensional Heisenberg group as we discuss in Example 8.1.12 below.
Exercise 7.11: Prove that S = (ξ, η, Φ, g) is a Sasakian structure on R2n+1 . What
happens to (R2n+1 , S) under a transverse homothety?
Example 7.3.22: Recall the complex ball BCn (1) of Example 3.2.9 with the Bergman
metric g0 of constant sectional curvature −1. We can scale this metric to have con-
stant holomorphic sectional curvature k < 0. Let gT denote this metric, and ωT
the corresponding Kähler form. Since BCn (1) is simply connected there is a 1-form
α such that ωT = dα. We construct a Sasakian structure on BCn (1) × R by setting

η = dt+π ∗ α and ξ = ∂t , where π : BCn (1)×R−−→BCn (1) is the natural projection and
t is the coordinate on R. We define Φ to be the horizontal lift of the almost complex
structure tensor J and zero in the vertical direction. The metric is g = gT + η ⊗ η
as usual. Then S = (ξ, η, Φ, g) defines a Sasakian structure on BCn (1) × R which by
Lemma 7.3.15 has constant Φ-sectional curvature H = c = k − 3. We denote this
Riemannian structure by BCn (k) × R.
Exercise 7.12: Show that BCn (k) × R maps to BCn (1) × R under a transverse ho-
mothety.
The examples discussed above are known as Sasakian space forms and the
following theorem of Tanno [Tan69a] gives the classification of simply connected
Sasakian space forms. So there are precisely three Sasakian space forms up to
transverse homothety.
Theorem 7.3.23: Let M (c)be a complete simply connected Sasakian manifold of
constant Φ-sectional curvature c. Then M (c) is one of the following:
(i) If c > −3, M (c) is isomorphic to S 2n+1 (c).
(ii) If c = −3, M (c) is isomorphic to R2n+1 (−3).
(iii) If c < −3, M (c) is isomorphic to BCn (k) × R, where c = k − 3.
For Sasakian Φ-symmetric spaces the reader is referred to [Bla02] and refer-
ences therein.

7.4. Topology of K-Contact and Sasakian Manifolds


We had seen previously in Section 6.2 that a necessary condition for a manifold
to admit a strict almost contact structure is the vanishing of certain characteristic
classes, namely the odd Stiefel-Whitney classes. Here we discuss some additional
topological obstructions for a manifold to admit a K-contact or Sasakian struc-
ture. The first such obstruction was noticed by Rukimbira [Ruk93] in 1993, and
somewhat later by Itoh [Ito97]. Rukimbira’s result applies to what he called R-
contact structures, that is, contact metric structures whose Reeb vector field defines
a Riemannian flow. This condition is ostensively weaker than the K-contact con-
dition. (Clearly, any K-contact structure is R-contact.) However, later Rukimbira
[Ruk95a] proved that every R-contact structure admits a compatible K-contact
metric. Hence, we restrict ourselves to K-contact structures.
Recall that the cuplength of a topological space M, denoted by cup(M ) is the
largest integer k such that there exists classes α1 , . . . , αk ∈ H̃ ∗ (M, R) satisfying
α1 ∪ · · · ∪ αk 6= 0. Clearly if M is an orbifold or manifold, we have 0 ≤ cup(M ) ≤
dim(M ), and if in addition M is compact cup(M ) > 0. For example, the (2n + 1)-
torus T 2n+1 has maximal cuplength, namely cup(T 2n+1 ) = 2n + 1. Now we have
230 7. K-CONTACT AND SASAKIAN STRUCTURES

Theorem 7.4.1: Let M be a compact K-contact manifold of dimension 2n + 1.


Then,
1 ≤ cup(M ) ≤ 2n.
We have an immediate corollary of this theorem:
Corollary 7.4.2: The manifolds Σg1 × · · · × Σgl × S 1 do not admit a K-contact
structure, where Σg denotes a Riemann surface of genus g ≥ 1. In particular, a
torus cannot admit a K-contact structure.
On the other hand odd dimensional tori do admit contact structures [Bou02],
as evidenced by Example 6.1.23 and Exercise 6.5. The proof of Theorem 7.4.1 is
based on the following well-known result of Tachibana [Tac65].
Lemma 7.4.3: Let (M, ξ, η, g) be a compact K-contact manifold. Then every har-
monic 1-form α satisfies α(ξ) = 0.
Proof. Any closed 1-form α can be decomposed into α = β + f η where β
satisfies β(ξ) = 0. By Hodge theory harmonic forms are invariant under isometries,
and ξ is an infinitesimal isometry. Thus,
0 = £ξ α = d(α(ξ)) = df,
so f is constant. This implies that dβ + f dη = 0. So by Stokes theorem
Z Z
n−1
0= d(β ∧ η ∧ (dη) =− f η ∧ (dη)n ,
M M
which, since η ∧ (dη)n is a volume form, implies that f = 0. So α = β proving the
result. ¤
Proof. (of Theorem 7.4.1). By Hodge theory every cohomology class can be
represented uniquely by a harmonic form, and the wedge product on forms induces
the cup product on de Rham cohomology classes. So suppose that cup(M ) = 2n+1,
then the Riemannian volume dvol form can be written uniquely up to order as
dvol = α1 ∧ · · · ∧ α2n+1 ,
where αi are harmonic 1-forms. Since M admits a K-contact structure, the har-
monic forms αi can be given with respect to the K-contact metric g. Moreover, the
characteristic vector field ξ is a Killing vector field, and we can extend ξ to a local or-
thonormal basis {ξ, e1 , . . . , e2n }. But then Lemma 7.4.3 gives dvol(ξ, e1 , . . . , e2n ) =
0, which implies the vanishing of the cohomology class [dvol] giving a contradic-
tion. ¤
We mention that Example 3.5.16 below gives Sasakian, hence, K-contact man-
ifolds, with cup(M ) = 2n. Rukimbira [Ruk94] noticed that one can do better than
the upper bound given by Lemma 7.1.9, namely
Theorem 7.4.4: Let (M, ξ, η, Φ, g) be a compact K-contact manifold of dimension
2n + 1 with first Betti number b1 = b1 (M ). Then
1 ≤ rk(M, Fξ ) ≤ min(n + 1, 2n + 1 − b1 ) .
Proof. By Hodge theory there are precisely b1 independent harmonic 1-forms
α1 , . . . , αb1 . Let Z1 , . . . , Zb1 denote their dual harmonic vector fields, and let k =
rk(M, Fξ ). Then by Proposition 2.6.3 the leaf closures of the characteristic foliation
Fξ are tori T whose maximal dimension is k. So on a dense open set the leaf
closures are tori Tk of dimension precisely k. Furthermore, Tk lies in the center of
7.4. TOPOLOGY OF K-CONTACT AND SASAKIAN MANIFOLDS 231

the automorphism group Aut(ξ, η, Φ, g). So if V ∈ tk , the Lie algebra of Tk , it is a


Killing vector field. Thus, by Tachibana’s Lemma 7.4.3 we have
0 = αi (V ) = g(Zi , V )
for all i = 1, . . . , b1 and any V ∈ tk . It follows that k + b1 ≤ 2n + 1 which together
with Lemma 7.1.9 proves the result. ¤
Together with Theorem 7.1.3, Rukimbira’s Theorem 7.4.4 has an interesting
corollary.
Corollary 7.4.5: Let (M, ξ, η, Φ, g) be a compact K-contact metric manifold of
dimension 2n + 1 whose first Betti number b1 (M ) = 2n. Then M is the total space
of a principal S 1 V-bundle over a compact almost Kähler orbifold Z whose first
Betti number b1 (Z) = 2n.
Another consequence of Tachibana’s Lemma 7.4.3 is
Proposition 7.4.6: There are no parallel vector fields on a compact K-contact
manifold.
Proof. Parallel vector fields are harmonic so Lemma 7.4.3 implies that a par-
allel vector field is orthogonal to ξ. Thus, for ρ parallel and for X an arbitrary
vector field we have
0 = ∇X g(ξ, ρ) = g(∇X ξ, ρ) + g(ξ, ∇X ρ) = g(ΦX, ρ)
by Proposition 7.3.4. But ρ being parallel is a section of D, so it must vanish by
the non-degeneracy of Φ on D. ¤
A quasi-regular K-contact manifold clearly has an infinite number of closed
characteristics, that is, an infinite number of closed integral curves of the charac-
teristic vector field ξ. The famous Weinstein conjecture[Wei79] states that for any
compact hypersurface of contact type with vanishing first Betti number, the Reeb
vector field ξ has a periodic orbit, i.e., a closed integral curve. It is also widely
believed that the conjecture holds when the condition on the first Betti number is
dropped. Much progress has been made recently on this conjecture, but it still re-
mains open in full generality. We refer the reader to [Bla02, MS98] and references
therein for further discussion and results. For K-contact manifolds the conjecture
was proven by Banyaga [Ban90], and more recently Rukimbira [Ruk95b, Ruk99]
proved the following refinement:
Theorem 7.4.7: Let (M, ξ, η, Φ, g) be a compact K-contact manifold of dimension
2n + 1. Then ξ has at least n + 1 closed integral curves. Furthermore, if M is
simply connected and ξ has precisely n + 1 closed characteristic curves. Then M is
homeomorphic to the sphere S 2n+1 .
Proof. Let (M, ξ, η, Φ, g) be a compact irregular K-contact manifold of di-
mension 2n + 1. Then by Theorem 7.1.10 there exists a quasi-regular K-contact
structure (ξ1 , η1 , Φ1 , g1 ) on M close to (ξ, η, Φ, g). Furthermore, by Theorem 2.6.4,
the vector fields ξ and ξ1 are equivalent to linear flows on a torus; hence, they com-
mute and ξ1 is an infinitesimal automorphism of the K-contact structure (ξ, η, Φ, g).
Consider the function η(ξ1 ). It is invariant under the circle action generated by ξ1 .
If p ∈ M is a critical point of the function η(ξ1 ), then we have
0 = d(η(ξ1 ))(p) = (£ξ1 η)(p) − (ξ1 dη)(p) = −(ξ1 dη)(p).
232 7. K-CONTACT AND SASAKIAN STRUCTURES

It follows that the vectors ξ1 (p) and ξ(p) are proportional. Furthermore, since η(ξ1 )
is invariant under the circle action generated by ξ1 , it follows that ξ1 (p) and ξ(p)
are proportional at all points of the leaf L1 of Fξ1 through p. So the leaves of Fξ
and the leaves of Fξ1 coincide at the critical points of η(ξ1 ). Since the leaves of Fξ1
are all circles, the leaves of Fξ are circles at these critical points.
Now we can use Lusternik-Schnirelman theory to get a lower bound on the
number of critical curves. This was done by Weinstein [Wei77b] by showing that
the Lusternik-Schnirelman Theorem generalizes to compact orbifolds (Notice that
this proves Weinstein’s conjecture for K-contact manifolds assuming that critical
points exist). In fact Weinstein shows that there are at least Cat(M/Fξ1 ) criti-
cal curves of any circle invariant function on M, where Cat(M/Fξ1 ) denotes the
Lusternik-Schnirelman category2 of the orbifold M/Fξ1 . But on any topological
space S we have the inequality Cat(S) ≥ cup(S) + 1. Clearly, on a compact almost
Kähler orbifold of dimension 2n, we have cup ≥ n. Putting these estimates together
proves the first part.
The proof of the second statement uses Morse-Bott theory. We only give an
outline of the proof here. First using curvature computations Rukimbira shows that
the Hessian is non-degenerate, so Bott’s generalization [Bot54] of Morse theory
holds. Moreover, the fact that there is a transverse almost complex structure implies
that the Morse indices are all even. Then since there are precisely n + 1 closed
characteristics, one for each index 0, 2, . . . , 2n, the Morse inequalities imply that all
the Betti numbers are either 0 or 1. Then using the gradient flow for the function
η(ξ1 ) one obtains a stratification of M 2n+1 as follows:
M 1 ⊂ M 3 ⊂ · · · ⊂ M 2n−1 ⊂ M 2n+1 ,
where each M 2k+1 a is closed embedded submanifold with k + 1 non-degenerate
critical circles of the restriction of η(ξ1 ) to M 2k+1 . Rukimbira then shows using
the Mayer-Vietoris sequence that if H1 (M 2k+1 , Z) = 0 then H1 (M 2k−1 , Z) = 0,
and if in addition M 2k−1 is a homology sphere then so is M 2k+1 . By hypothesis
H1 (M 2n+1 , Z) = 0, and using Carrière’s classification [Car84a] of Riemannian
flows on 3-manifolds one sees that M 3 is diffeomorphic to S 3 . This then together
with Smale’s generalized Poincaré Theorem [Sma61] proves the result. ¤

Another related theorem of Rukimbira [Ruk95b] is:


Theorem 7.4.8: Let (M, ξ, η, Φ, g) be a compact K-contact manifold, and suppose
that ξ has only a finite number of closed integral curves. Then b1 (M ) = 0.
A corollary of this result is that any toric K-contact manifold M must have
b1 (M ) = 0. There is a stronger result for toric K-contact manifolds due to Lerman
[Ler04b] which says that the fundamental group of any K-contact toric manifold
is finite and determined combinatorially by the image of the contact moment map.
We will treat toric contact manifolds in some detail later.
Our next example of an obstruction to K-contact structures involves Gromov’s
invariants MinVol and ||M || discussed at the end of Chapter 2. Since a K-contact
structure is in particular a Riemannian flow, we have immediately from Theorem
2.6.13

2Recall that for a topological space Cat(S) is the minimal number of open sets needed to
cover S. This definition is not standard though. For example, the minimal number of such open
sets minus one is frequently used [CLOT03].
7.4. TOPOLOGY OF K-CONTACT AND SASAKIAN MANIFOLDS 233

Corollary 7.4.9: Let M admit a K-contact structure. Then both MinVol(M ) and
Gromov invariant ||M || vanish.
From Theorem 2.6.14, the estimate 2.6.3, and the properties of the simplicial
volume discussed in Chapter 1, we have
Corollary 7.4.10: If M decomposes as M = M1 # · · · #Ml where some Mi is
compact and admits a metric whose sectional curvature is bounded from above by
−δ for some δ > 0, then M does not admit a K-contact structure.
For obstructions to Sasakian manifolds we have the following theorem which ap-
pears to have been discovered independently by various geometers [BG67, Fuj66,
Tan67]. It is the analogue of a well-known theorem in Kähler geometry.
Theorem 7.4.11: Let M 2n+1 be a compact manifold admitting a Sasakian struc-
ture S. Then the pth Betti number bp (M ) is even for p odd with 1 ≤ p ≤ n, and
for p even with n < p ≤ 2n.
Proof. For p = 1 this follows immediately from (v) of Proposition 7.2.3, (vi)
of Theorem 7.2.6 and equation (7.2.14). For p > 1 it is much more involved. We
present the main technical result as a lemma and leave the proof as an exercise.
The first part is a generalization of Lemma 7.4.3 and the lemma itself is again due
to Tachibana [Tac65]:
Lemma 7.4.12: Let (M 2n+1 , S) be a compact Sasakian manifold. Then for 1 ≤
p ≤ n any harmonic p-form u is orthogonal to η, i.e., ξ u = 0. Moreover, for the
same range of p, the p-form Φu defined by
p
X
Φu(X1 , . . . , Xp ) = u(X1 , . . . , ΦXi , . . . , Xp )
i=1

is harmonic.
Assuming the lemma, we proceed with the proof of Theorem 7.4.11. Define the
p-form Iu by
Iu(X1 , . . . , Xp ) = u(ΦX1 , . . . , ΦXp ),
and for each k = 0, . . . , p consider the iterates Φk u. By Lemma 7.4.12 each Φk u
is harmonic. Furthermore, it should be clear that Iu can be written as a linear
combination of the harmonic forms {Φk u}pi=0 . Hence, Iu is itself harmonic. So I
is a map I : Hp −−→Hp on harmonic p-forms. Moreover, it satisfies I 2 = (−1)p 1l
on Hp . Thus, for p ≤ n odd, the real harmonic forms u and Iu are independent,
so the result follows for this range of p. The other range of p follows by Poincaré
duality. ¤
Exercise 7.13: Prove Lemma 7.4.12.
More can be said about harmonic theory on compact Sasakian manifolds.
Proposition 7.4.13: Let (M 2n+1 , S) be a compact Sasakian manifold, and let p
be an integer satisfying 1 ≤ p ≤ n. Then a p-form u is harmonic if and only if it is
primitive and basic harmonic.
Proof. Let u be a harmonic p-form on M. Then since u is closed and horizontal
by Lemma 7.4.12, u is basic. By the definition of ¯?, equation (7.2.2), we see that
?u = η ∧ ¯?u. Thus,
d(?u) = dη ∧ ¯?u − η ∧ d(¯?u) = L¯?u − η ∧ dB ¯?u .
234 7. K-CONTACT AND SASAKIAN STRUCTURES

The two terms on the right hand side of this equation are independent so the left
hand side vanishes if and only if both terms on the right vanish. The vanishing
of the first term on the right says that u is primitive, while the vanishing of the
second says that u is basic harmonic. ¤

Proposition 7.4.13 says that the basic primitive cohomology groups P r (Fξ ) of
a Sasakian manifold are actually topological invariants. This implies then that the
r
basic cohomology groups HB (Fξ ) are also topological invariants. More precisely,
Theorem 7.4.14: Let M and M 0 be homeomorphic compact Sasakian manifolds
with Sasakian structures S = (ξ, η, Φ, g) and S 0 = (ξ 0 , η 0 , Φ0 , g 0 ), respectively. Then
r r
their basic cohomology groups are isomorphic, i.e., HB (Fξ0 ) ≈ HB (Fξ ). Further-
more, the Betti numbers br and basic Betti numbers br are related by br = bB
B B
r −br−2
B
for 0 ≤ r ≤ n. In particular, if r is odd, br = br .

Proof. By Proposition 7.4.13 the basic primitive cohomology groups P r (Fξ )


are isomorphic to the de Rham cohomology of M which is a topological invariant.
So P r (Fξ ) ≈ P r (Fξ0 ). But the transverse hard Lefschetz Theorem 7.2.9 computes
r
HB (Fξ ) in terms of P r (Fξ ) by wedging with dη, and computes HB r
(Fξ0 ) in terms
r 0
of P (Fξ0 ) by wedging with dη , and these are isomorphic operations. This proves
the first statement. The second statement then follows from Corollary 7.2.10. ¤

r
Although the basic cohomology groups HB (Fξ ) cannot distinguish Sasakian
p,q
structures, the basic Hodge groups HB (Fξ ) can as we show by explicit exam-
ple later. Notice that Theorem 7.4.14 and (iv) of Theorem 7.2.6 yield Theorem
7.4.11 as an immediate corollary. Both proofs of Theorem 7.4.11, however, rely on
Tachibana’s Lemma 7.4.12. In the case of a quasi-regular K-contact or Sasakian
manifold M , Proposition 7.2.2 together with Theorem 7.4.14 relate the real coho-
mology of M with the real cohomology of the space of leaves Z = M/Fξ . However,
they do not give integral information. For integral information one needs to analyze
the Leray spectral sequence of the corresponding Seifert S 1 -bundle. In this regard
see Proposition 4.7.9.
Next we discuss without proof some rigidity results regarding Sasakian man-
ifolds with positive sectional curvature. This work appears to have begun with
Goldberg [Gol65, Gol67] under the added assumption of regularity, but reached
its fruition in the work of Tanno [Tan68] and the unpublished thesis of Moskal
[Mos66] which is described in detail in Sasaki [Sas68].
Theorem 7.4.15: Let (M, S) be a compact Sasakian manifold with positive sec-
tional curvature. Then b2 (M ) = 0. If in addition the scalar curvature is constant,
then (M, g) is isometric to the sphere S 2n+1 (c) of constant Φ-sectional curvature c
for some −3 < c < ∞.
The first statement is due to Tanno [Tan68] with a weaker version established a
few years earlier in [TO66], while the second statement is due to Moskal. Somewhat
stronger results are available in the low dimensions of three and five, as well as under
the added assumption of being Einstein. These are treated in detail in Chapters 10
and 11.
The first known example of a symplectic manifold M 4 that does not admit a
Kähler structure is the well-known primary Kodaira surface from complex surface
7.4. TOPOLOGY OF K-CONTACT AND SASAKIAN MANIFOLDS 235

theory [BPVdV84]. This was pointed out by Thurston3 [Thu76] who constructed
it as a non-trivial T 2 -bundle over T 2 with first homology group equal to Z3 . This
gives rise to a K-contact manifold M 5 that admits no Sasakian structure. Later M 4
was shown to admit an indefinite Kähler structure in [FGG88]. Further examples
of compact almost Kähler manifolds admitting no Kähler structure are given in
[CFdL85, CFG86]. Many of these examples, including Thurston’s, involve the
Heisenberg group. The geometry of Thurston’s example is discussed more fully in
[Abb84].
Example 7.4.16: A non-trivial T 3 -bundle over T 2 . Consider the real three-step
nilpotent group G5 (R) which as a set is R5 . Using vector notation the group action
is
(7.4.1) (c, a, b) · (z, x, y) = (x + a, Ay + b, z + c + a2 x1 + b2 y1 + a2 b2 y2 + a2 y22 ) ,
where µ ¶
1 a2
A= .
0 1
Let G5 (Z) denote G5 with integer values for its parameters, and let M 5 denote the
quotient G(R)/G(Z). Then π1 (M 5 ) ≈ G5 (Z) which as a set is Z5 . The commutator
subgroup is [G5 (Z), G5 (Z)] = Z2 . So H1 (M 5 , Z) ≈ Z3 . Moreover, by construction
M 5 is a S 1 -bundle over Thurston’s non-trivial T 2 -bundle over T 2 which is an almost
Kähler manifold. Thus, M 5 is K-contact but not Sasakian by Theorems 7.1.6 and
7.4.11. It does, however, admit a pseudo-Sasakian structure owing to the indefinite
Kähler structure on the base. This is a type of normal almost contact structure.
As with Kähler manifolds it is natural to ask to what extend K-contact mani-
folds are more general than Sasakian manifolds. In the symplectic case, the first sim-
ply connected examples of non-Kähler symplectic manifolds was given by McDuff
[McD84]. Later Gompf [Gom95] gave a systematic construction which produced
infinitely many examples of simply connected symplectic non-Kähler manifolds.
It seems likely that a similar result should be true in the K-contact case, but so
far there are no known simply connected K-contact manifolds which do not admit
Sasakian structures. Thus, we have
Open Problem 7.4.1: Find examples of simply connected K-contact manifolds
which cannot admit Sasakian structures.
The work of Kollár [Kol05, Kol06a] should prove important in attacking this
problem. Indeed Kollár has discovered torsion obstructions to the existence of
Sasakian structures in dimension five. We discuss this work in detail in Chapter
10. Perhaps it is worth mentioning here that the techniques of Gompf and Kollár
appear to be quite different. Another known obstruction to the existence of Kähler
structures is the existence of non-vanishing Massey products [DGMS75]. This
technique has been utilized in some of the references mentioned above, cf. [CFG93].
However, non-vanishing Massey products do not obstruct Sasakian structures as the
three dimensional Heisenberg group shows. Nevertheless, certain Massey products
do obstruct Sasakian structures, but they depend on the basic cohomology class

3Thurston gave his example as a counterexample to a result of Guggenheimer who claimed


that the odd Betti numbers of a symplectic manifold were even after P. Libermann had noticed
that Guggenheimer’s proof was incomplete.
236 7. K-CONTACT AND SASAKIAN STRUCTURES

of a K-contact structure. It would be very interesting to obtain a topological


characterization of such obstructions.
As with Kähler manifolds another interesting question concerns the possible
fundamental groups of compact Sasakian manifolds. In the case of compact Kähler
manifolds quite a bit is known concerning which groups can be the fundamental
group of a compact Kähler manifold [ABC+ 96], whereas, almost nothing is known
for compact Sasakian manifolds. As with Kähler manifolds it is clear that there are
restrictions on the fundamental group of a compact Sasakian manifold, for example
Theorem 7.4.11 says that the first Betti numbers of any compact Sasakian manifold
must be even. Moving away from the Kähler category to symplectic geometry, there
is again a result of Gompf [Gom95] saying that any finitely presented group can
be realized as the fundamental group of a compact symplectic manifold. Similarly
A’Campo and Kotschick [AK94] have shown that any finitely presented group
is the fundamental group of some compact contact manifold. It is not known
whether there are any restrictions on the fundamental group of a compact K-contact
manifold. As in the Kähler case we have
Definition 7.4.17: A group Γ is called a Sasaki group if it is the fundamental
group of some compact Sasakian manifold. The first Betti number of Γ b1 (Γ) is
the rank of the Abelianization Γ/[Γ, Γ].
As with Kähler metrics Sasakian metrics can be lifted to coverings, so subgroups
of finite index of a Sasaki group is Sasaki. Thus, we have the following negative
result:
Proposition 7.4.18: A group Γ which contains a finite index subgroup Γ0 with
b1 (Γ0 ) odd is not Sasaki. In particular any non-trivial free group cannot be Sasaki.
It appears as though the analogy with Kähler groups stops with this proposi-
tion. For example we have seen above that Sasaki groups unlike Kähler groups can
have non-vanishing Massey products.

7.5. Sasakian Geometry and Algebraic Geometry


We begin the study of Sasakian structures per se. First we translate Theorem
7.3.7 to the Sasakian situation. In the case of a compact quasi-regular Sasakian
manifold M we know that the space of leaves M/Fξ is a compact Riemannian
orbifold Z. But since the transverse geometry on M is Kähler, the orbifold must
be Kähler. But actually in the quasi-regular case more is true. It follows from
Theorem 7.3.7 that M is the total space of a V-bundle over Z, and the curvature
of the connection form η is precisely the pullback of the Kähler form on Z. Thus,
Z satisfies an orbifold integrality condition which we now elaborate. This integral-
ity condition ties Sasakian geometry on compact manifolds to projective algebraic
geometry.
We can now combine the results of this section with Theorem 6.3.8 or Theorem
7.1.3 to give one of the foundational theorems of our subject.
Theorem 7.5.1: Let S = (ξ, η, Φ, g) be a compact quasi-regular Sasakian structure
on a smooth manifold of dimension 2n + 1, and let Z denote the space of leaves of
the characteristic foliation (just as topological space). Then
(i) Z carries the structure of a Hodge orbifold Z = (Z, ∆) with an orbifold
Kähler metric h and Kähler form ω which defines an integral class [p∗ ω]
7.5. SASAKIAN GEOMETRY AND ALGEBRAIC GEOMETRY 237

2
in Horb (Z, Z) in such a way that π : (M, g)−−→(Z, h) is an orbifold Rie-
mannian submersion, and a principal S 1 orbibundle over Z. Furthermore,
ω satisfies π ∗ ω = dη.
(ii) The fibers of π are geodesics.
(iii) Z is also a Q-factorial, polarized, normal projective algebraic variety.
(iv) If S is regular then the orbifold structure is trivial and π : M −−→Z is a
principal circle bundle over a smooth projective algebraic variety.
Proof. (i) and (ii) follow from Theorem 7.1.3 and the fact that the transverse
almost contact structure is integrable. That Z is a projective algebraic variety
follows from the Kodaira-Baily embedding theorem 4.4.30, and since the only sin-
gularities are quotient singularities it is both normal and Q-factorial. Now ω defines
a rational class in H 2 (Z, Q) which pullback to the integral class [p∗ ω] ∈ Horb
2
(Z, Z)
and thus defines a holomorphic line V-bundle whose unit circle bundle is the prin-
cipal S 1 V-bundle. Thus, (Z, [ω]) is a polarization. ¤
Notice that a quasi-regular Sasakian structure on a compact manifold can be
thought of as a transverse projective algebraic structure. For compact Kähler mani-
folds there was a conjecture attributed to Kodaira that says that any Kähler struc-
ture can be approximated or equivalently is a deformation of an algebraic structure.
It was shown to be true in case of compact Kähler surfaces by Kodaira [Kod60a].
However, it has been recently shown by Voisin [Voi04] that this does not hold in
complex dimension four and higher. More explicitly Voisin has given examples of
simply connected compact Kähler manifolds whose cohomology ring is not isomor-
phic to the cohomology ring of a projective algebraic variety. Perhaps surprisingly
the analog in Sasakian geometry does hold. A corollary of the Structure Theorem
7.1.10 which will be stated precisely in the next section says that every Sasakian
structure is a deformation of a quasi-regular Sasakian structure. Thus, in a real
sense (no pun intended), Sasakian geometry is more algebraic than Kählerian ge-
ometry.
The Inversion Theorem 7.1.6 now becomes
Theorem 7.5.2: Let (Z, ω, J) be a Hodge orbifold with [p∗ ω] ∈ Horb 2
(Z, Z), and
let M denote the total space of the circle V-bundle defined by the class [ω]. Then
the orbifold M admits a quasi-regular Sasakian structure (ξ, η, Φ, g) such that dη =
π ∗ ω where π : M −−→Z is the natural orbifold projection map, or alternatively π :
M −−→Z is a Seifert S 1 -bundle. Furthermore, if all the local uniformizing groups of
Z inject into the structure group S 1 , then Z is a cyclic orbifold and M is a smooth
Sasakian manifold.
2n+1
Example 7.5.3: Consider the weighted sphere Sw as described in Example
7.1.12 with integral weights ordered as 0 < w0 ≤ · · · ≤ wn . The Sasakian structure
Sw = (ξw , ηw , Φw , gw ) is quasi-regular for every such weight vector w. We shall
assume that gcd(w0 , . . . , wn ) = 1 for otherwise we can perform a transverse homo-
thety 7.3.10 to obtain another Sasakian structure. This doesn’t change the orbifold
structure of the base space Z, but only resales the Kähler metric h. In our case
Zw = P(w) the weighted projective space of Definition 4.5.2. This is a Kähler orb-
ifold as well as a Q-factorial normal projective algebraic variety. The polarization is
defined by the unique Kähler class [ωw ] satisfying π ∗ [ωw ] = [dηw ]B . On the orbifold
Zw the canonical orbisheaf of Definition 4.4.12 is the orbisheaf OP(w) (−|w|). The
Sasakian structure Sw is determined by the orbifold structure of Zw . For example,
238 7. K-CONTACT AND SASAKIAN STRUCTURES

consider the weighted Sasakian structure Sw with w = (6, 2(6k − 1), 3(6k − 1)) with
order υ(S) = 6(6k − 1). The orbifold Zw is that given by Example 4.5.11, namely
P(6, 2(6k − 1), 3(6k − 1)). As an algebraic variety P(6, 2(6k − 1), 3(6k − 1)) is just
P2 , but not as orbifolds. The Sasakian structure with Z = P2 is that of the round
sphere of Example 7.1.5 and has order υ = 1.
Another similar example is:
Example 7.5.4: Consider as above the weighted projective space CP2 (w0 , w1 , w2 )
defined by the usual weighted C∗ action on C3 \ {0}, where now w0 , w1 , w2 are pair-
wise relatively prime integers. As an algebraic variety CP2 (w0 , w1 , w2 ) is equivalent
[Kol96] to CP2 /Zw0 × Zw1 × Zw2 . But as orbifolds these are distinct, since the for-
mer has π1orb = 0, whereas the latter has π1orb = Zw0 × Zw1 × Zw2 . The induced
metrics are also different. In the latter case the metric is just the Fubini-Study
metric pushed to the quotient which is Kähler-Einstein. The Sasakian structure on
the corresponding S 1 V-bundle is just the standard Sasaki-Einstein structure on the
lens space S 5 /Zw0 × Zw1 × Zw2 ; whereas, in the former case the Sasakian structure
is just the weighted sphere as described in Example 7.1.12 and the induced metric
is ellipsoidal and not Einstein. Likewise, the metric h on CP2 (w0 , w1 , w2 ) is not
Kähler-Einstein. In both cases the order is the same, namely υ = w0 w1 w2 .
Let Z be a Kähler orbifold and recall the Kähler cone K(Z) from Definition
3.6.2. In K(Z) we define the Kähler lattice KL (Z) of the complex orbifold Z to be
the lattice
(7.5.1) KL (Z) = {[ω] ∈ K(Z) | [p∗ ω] ∈ Horb
2
(Z, Z)} .
From the Inversion Theorem 7.5.2 we have
Theorem 7.5.5: Let (Z, ω) be a Hodge orbifold. Every [ω] ∈ KL (Z) determines
a principal S 1 V-bundle π : M[ω] −−→Z and choosing a connection η in M[ω] whose
curvature is π ∗ ω determines a Sasakian structure on M[ω] .
Recall that the class [ω] ∈ H 2 (Z, Q) is just the rational Chern class of the S 1
orbibundle, and generally [ω] does not uniquely determine a complex line orbibun-
dle. However, Kollár (Proposition 53 in [Kol04]) has recently proven that [ω] does
uniquely determine a complex orbibundle if H1orb (Z, Z) = 0 in which case the S 1
V-bundle π : M[ω] −−→Z with its CR structure is uniquely determined.

7.5.1. The Space of Sasakian Structures. Here we consider those Sasakian


structures that have the same Reeb vector field. Fix a Sasakian structure S =
(ξ, η, Φ, g) and define the set
(7.5.2) F(ξ) = {Sasakian structures (ξ 0 , η 0 , Φ0 , g 0 ) | ξ 0 = ξ} .
We give the set F(ξ) the C ∞ compact-open topology as sections of vector bundles,
and call it the space of Sasakian structures compatible with ξ or just the space of
Sasakian structures when the Reeb vector field is understood to be ξ. Define the real
1-form ζ = η 0 − η. Then ζ is basic, so [dη 0 ]B = [dη]B . Hence, all Sasakian structures
2
in F(ξ) correspond to the same basic cohomology class in HB (Fξ ). Furthermore, it
is clear that the order υ(S) does not depend on the Sasakian structure S ∈ F(ξ),
and thus is an invariant of the space F(ξ).
Let J¯ denote the underlying almost complex structure on ν(Fξ ), and πν :
T M −−→ν(Fξ ) the natural projection. We define the subspace F(ξ, J) ¯ ⊂ F(ξ) to be
7.5. SASAKIAN GEOMETRY AND ALGEBRAIC GEOMETRY 239

the subset of all Sasakian structures (ξ 0 , η 0 , Φ0 , g 0 ) ∈ F(ξ) such that the diagram
Φ0
T
M −−−−→ T
M
 
 
(7.5.3) πν πν
y y

ν(Fξ ) −−−−→ ν(Fξ ),

commutes. The elements of F(ξ, J) ¯ consist of all Sasakian structures S 0 ∈ F(ξ) with
the same transverse holomorphic structure.
We can now give an alternative description of F(ξ, J), ¯ but first we need the
transverse ∂ ∂¯ lemma due to El Kacimi-Alaoui [EKA90].
Lemma 7.5.6: Let (M, S) be a compact Sasakian manifold, and let ω, ω 0 be basic
real closed (1, 1)-forms such that [ω]B = [ω 0 ]B , then there exists a smooth basic
¯
function φ such that ω 0 = ω + i∂ ∂φ.

Proof. The proof is similar to the usual case [EKA90]. ¤

As in the Kähler case the basic 2-form dη can be written locally in terms of a
¯ So as with Kähler geometry, Sasakian
basic potential function φ, viz. dη = i∂ ∂φ.
geometry is locally determined by one potential function [GKN].
A real basic (1, 1)-form ω on a Sasakian manifold (M, ξ, η, Φ, g) is said to be
positive, written ω > 0, if ω(ΦX, X) > 0 for all smooth sections X of D. We can
obtain a characterization of the space of Sasakian metrics on M whose Reeb vector
field is ξ and whose transverse holomorphic structure is J.¯ We now define certain
subspaces of the space of all smooth basic functions on (M, Fξ ).
R
¯ > 0,
C01 (S) = {φ ∈ C ∞ (M )B | dη + i∂ ∂φ φ dvolg } ,
M
0 ∞
R
C0 (S) = {ψ ∈ C (M )B | M ψ dvolg = 0} .
Proposition 7.5.7: Let (M, ξ, η, Φ, g) be a compact Sasakian manifold with under-
¯ Then F(ξ, J)
lying transverse holomorphic structure J. ¯ = C 1 (S)×C 0 (S)×H 1 (M, Z).
0 0

¯ then as mentioned above dη and dη 0


Proof. Let S 0 = (ξ, η 0 , Φ0 , g 0 ) ∈ F(ξ, J),
1,1
represent the same basic cohomology class in HB (Fξ ). Thus, by Lemma 7.5.6
(7.5.4) ¯
dη 0 = dη + i∂ ∂φ
for some smooth basic function φ. Since S 0 is Sasakian we must have dη + i∂ ∂φ ¯ > 0.
However, the function φ does not uniquely determine the Sasakian structure. Define
dc = 2i (∂¯ − ∂) so that i∂ ∂φ
¯ = ddc φ. Then the 1-form ζ = η 0 − η − dc φ is closed and
basic; hence, it defines an element of H 1 (M, R) ≈ H 1 (Fξ ) which lies in the image of
the natural inclusion H 1 (M, Z)−−→H 1 (M, R). Let α denote the class in H 1 (M, Z)
defined by ζ. Then any other closed basic 1-form ζ 0 belonging to the same class α
differs from ζ by dψ where ψ is a smooth basic function on M. Since both φ and ψ
are only defined
R up to a constant, they can be chosen to satisfy the normalization
condition M f dvolg = 0. This then determines the pair (φ, ψ) uniquely. It is now
easy to see that the map S 0 7→ (φ, ψ, α) is one-to-one.
Conversely, with a fixed Sasakian structure S = (ξ, η, Φ, g) let (φ, ψ, α) ∈
C01 (S) × C00 (S) × H 1 (M, Z) be given. The class α ∈ H 1 (M, Z) defines a closed
240 7. K-CONTACT AND SASAKIAN STRUCTURES

1-form ζ up to a the differential of a smooth basic function. So the pair (ψ, α) de-
fines the 1-form ζ uniquely, the constant being fixed by the normalization condition
on ψ. Define η 0 by

(7.5.5) η 0 = η + dc φ + ζ .

Then η 0 is a contact form on M since if η 0 ∧ (dη + ddc φ)n were to vanish at some
point, taking the interior product with ξ shows that (dη+ddc φ)n would vanish which
violates the positivity condition. Now putting ζ 0 = dc φ + ζ we have η 0 = η + ζ 0 . We
define the endomorphism Φ0 and metric g 0 by

Φ0 = Φ − ξ ⊗ ζ 0 ◦ Φ,
(7.5.6) g 0 = dη 0 ◦ (Φ0 ⊗ 1l) + η 0 ⊗ η 0 .

Clearly, Φ0 and Φ project to the same transverse holomorphic structure J. ¯ To see


that Φ0 is uniquely determined by the triple (φ, ψ, α) we notice that if Φ0 and Φ
both project to J¯ then there is a smooth 1-form θ such that Φ0 = Φ + ξ ⊗ θ.
But then the condition η 0 ◦ Φ0 = 0 implies that θ = −ζ 0 ◦ Φ. Finally, we need to
check the positivity condition dη 0 ◦ (Φ0 ⊗ 1l) > 0, as well as the condition (Φ0 )2 =
−1l + ξ ⊗ η 0 . These are straightforward and left as an exercise. Furthermore, under
the above identification the topology of F(ξ, J)¯ coincides with the relative topology
as a subspace of C ∞ (M )B × C ∞ (M )B × H 1 (M, Z). ¤

The group H 1 (M, Z) labels the components F(ξ, J) ¯ α of F(ξ, J).


¯ We have
Corollary 7.5.8: Each component F(ξ, J) ¯ α is contractible.
¯
If we fix the component F(ξ, J)α , it is convenient to think of a transformation
of the form given by equation (7.5.5) as a deformation within a give component.
This point of view was adobted in [Bel01] and we formalize this as:
Definition 7.5.9: A transformation within F(ξ, J) ¯ α of the form η 7→ η + ζ for
some 1-form ζ is called a deformation of type II.
Deformations of type I will be defined in Definition 8.2.10 below.
Remark 7.5.1: It is interesting to note that in the quasi-regular case F(ξ, J) ¯ can
be viewed as a subspace of the space of connections in a U (1) V-bundle, and indeed
even in the irregular case as a subspace of the space of connections in a 1-dimensional
¯ is the subspace of connections with strictly positive
foliation. Explicitly, F(ξ, J)
curvature satisfying the normalization condition. The function ψ is then identified
with a null-homotopic gauge transformation.
Our next result says that of all the Sasakian structures in F(ξ) have the same
volume. (We always take the volume element dvolg to have positive orientation.)
Explicitly,
Proposition 7.5.10: Let M 2n+1 be a compact Sasakian manifold. Then for S, S 0 ∈
F(ξ) arbitrary, we have Vol(M, g) = Vol(M, g 0 ).

Proof. The infinitesimal volume element of a Sasakian metric is related to


the contact volume element, up to orientation, by
1
dvolg = η ∧ (dη)n .
n!
7.5. SASAKIAN GEOMETRY AND ALGEBRAIC GEOMETRY 241

Now by our discussion above there is a smooth basic function φ and closed basic
1-form ζ such that
1 0
dvolg0 = η ∧ (dη 0 )n = (η + dc φ + ζ) ∧ (dη + ddc φ)n = η ∧ (dη + ddc φ)n ,
n!
where the last equality holds since dc φ + ζ is basic. Now expand
Xn µ ¶
c n n n
(dη + dd φ) = (dη) + (dη)n−i ∧ (ddc φ)i ,
i=1
i
so that
Z Z
1
Vol(M, g 0 ) = dvolg0 = η ∧ (dη + ddc φ)n
M n! M
Z ³ Xn µ ¶ ´
1 n n
= η ∧ (dη) + (dη)n−i ∧ (ddc φ)i
n! M i=1
i
Z n µ ¶Z
1 1 X n ¡ ¢
= η ∧ (dη)n + η ∧ d (dη)n−i ∧ dc φ ∧ (ddc φ)i−1
n! M n! i=1 i M
Z n µ ¶ Z
1 1 X n ¡ ¢
= η ∧ (dη)n − d η ∧ (dη)n−i ∧ dc φ ∧ (ddc φ)i−1
n! M n! i=1 i M
Z
= dvolg = Vol(M, g) ,
M
where the last line follows by Stokes’ theorem. ¤
¯ when it is understood that the
Henceforth, we shall often write F(ξ) for F(ξ, J)
transverse complex structure is fixed. Later in Section 8.2.1 we discuss deformations
of the transverse holomorphic structure. It will be convenient to consider a set
of Sasakian structures that is slightly larger than F(ξ), namely those Sasakian
structures that have the same characteristic foliation. First we have
Definition 7.5.11: For any Sasakian structure S = (ξ, η, Φ, g) there is the con-
jugate Sasakian structure defined by S c = (ξ c , η c , Φc , g) = (−ξ, −η, −Φ, g) ∈
F(Fξ ) .
Clearly, this definition holds equally well for K-contact structures. We define
F(Fξ ) to be the set of all Sasakian structures whose characteristic foliation is Fξ .
Clearly, we have F(ξ) ⊂ F(Fξ ). Recall the transverse homothety 7.3.10. Then fixing
S we define
[
(7.5.7) F+ (Fξ ) = F(a−1 ξ) ,
a∈R+

and F− (Fξ ) to be the image of F+ (Fξ ) under conjugation. We now have


Lemma 7.5.12: The decomposition holds
F(Fξ ) = F+ (Fξ ) t F− (Fξ ) .

Proof. The one inclusion is obvious. Let S 0 = (ξ 0 , η 0 , Φ0 , g 0 ) ∈ F(Fξ ). Then


there is a nowhere vanishing smooth function f such that ξ 0 = f −1 ξ, and η 0 = f η.
Then the condition ξ 0 dη 0 = 0 implies that
(7.5.8) df = (ξf )η,
242 7. K-CONTACT AND SASAKIAN STRUCTURES

and this implies that Xf = 0 for any horizontal vector field X. So f only depends
on the leaf coordinates. Taking d of Equation 7.5.8 gives

0 = d(ξf ) ∧ η + (ξf )dη .

If we can show that the first term vanishes, then ξf = 0 so f will be constant
which will prove the lemma. Now the first term d(ξf ) ∧ η = dh (ξf ) ∧ η where
dh is the exterior derivative in the horizontal direction. So we compute Xξf for
X horizontal. We have, using the fact that f is independent of the horizontal
coordinates, Xξf = [X, ξ]f = 0, since [X, ξ] is horizontal. ¤

This discussion shows that if we fix a foliation Fξ and a class α ∈ H 1 (M, Z),
the subset of homology classes represented by compatible Sasakian structures forms
2
a line minus the origin in HB (Fξ ), and that conjugation interchanges the positive
and negative rays. We shall refer to F± (Fξ )α as an a-homology class of Sasakian
structures, or simply a Sasakian ray, and we shall always mean F+ (Fξ )α unless
otherwise stated. We need
Definition 7.5.13: Two Sasakian structures S = (ξ, η, Φ, g) and S 0 = (ξ 0 , η 0 , Φ0 , g 0 )
in F(Fξ )α on a smooth manifold M are said to be a-homologous if there is an
a ∈ R+ such that ξ 0 = a−1 ξ and [dη 0 ]B = a[dη]B .
The a-homology classes form a set of two elements that can be identified with
2
positive and negative rays in HB (Fξ ). So every Sasakian structure in F(Fξ )α is a-
homologous to S or its conjugate S c . For example, the standard Sasakian structure
S0 on S 2n+1 discussed in Example 7.1.5 is a representative in the standard a-
deformation class F+ (Fξ0 ), while S0c = (−ξ0 , −η0 , −Φ0 , g0 ) is a representative in
F− (Fξ0 ). Of course, the other representatives of F(Fξ0 ) do not, in general, have
constant sectional curvature. The diffeomorphism C : S 2n+1 −−→S 2n+1 induced by
complex conjugation on Cn+1 sends S0 to S0c . Note that C is an element of the
contactomorphism group Con(S 2n+1 , D0 = ker η0 ) (as well as the isometry group
Isom(M, g0 )) which preserves orientation if n is odd, and reverses orientation for
n even. For n odd C reverses the co-orientation. A natural question to ask is:
Question 7.5.1: Given a Sasakian manifold (M, S) when does there exist a C ∈
Con(M, D) ∩ Isom(M, g0 ) such that C(S) = S c ?
This question is discussed further in Chapter 8. Summarizing we have
Theorem 7.5.14: The space of Sasakian structures F(Fξ )α corresponding to a
class α ∈ H 1 (M, Z) consists of two contractible components F+ (Fξ )α and F− (Fξ )α
that are isomorphic under conjugation. Moreover, any two Sasakian structures in
F(Fξ )±
α are contactomorphic.
We are specially interested in the case that the Sasakian structure S is quasi-
regular. Then according to Theorem 7.5.2 S determines a Seifert S 1 -bundle. Since
quasi-regularity is a property of the foliation Fξ , every S ∈ F(Fξ )+ α determines
the same Seifert bundle, whereas, elements of F(Fξ )− α correspond to the ‘conjugate
Seifert S 1 -bundle’ with the orientation of the fibres reversed and with the complex
conjugate structure on Z. Thus, we can think of F(Fξ )+ , or by abuse of terminology
F(Fξ ), as a certain type of Seifert S 1 -bundle. This leads us to
Definition 7.5.15: If Fξ is quasi-regular we call F(Fξ )+ α (or even F(Fξ )α ) a
Sasaki-Seifert structure.
7.5. SASAKIAN GEOMETRY AND ALGEBRAIC GEOMETRY 243

Thus, a Sasaki-Seifert structure is just an a-deformation class of Sasakian struc-


tures under the additional assumption of quasi-regularity. Strictly speaking the
transverse complex structure J¯ is fixed for a given Sasaki-Seifert structure since
the complex structure of the orbifold X = (X, ∆) is fixed. However, as we shall
see Sasaki-Seifert structures often occur in continuous families (moduli), and we
will sometimes use the term Sasaki-Seifert structure to refer to an entire family of
such structures. The context should make this clear. Note that the Approximation
Theorem 7.1.10 says that a manifold which admits a Sasakian structure necessarily
admits a Sasaki-Seifert structure.
Open Problem 7.5.1: The following two problems are of paramount interest to
us in this book.
(1) Determine which manifolds admit Sasakian structures.
(2) Determine all Sasaki-Seifert structures on a given manifold M.
In full generality these problems seem quite hopeless to answer. However, we
shall see in Chapter 10 that under some added conditions, such as positivity to be
discussed in the next section, one can go quite far in certain cases.
7.5.2. Chern Classes and Basic Chern Classes. Consider the complex
vector bundle D on a Sasakian manifold (M, ξ, η, Φ, g). As such D has Chern
classes c1 (D), . . . , cn (D) which can be computed by choosing a connection ∇D in D
[Kob87]. We choose ∇D to be the transverse connection ∇T of 2.5.7 in which case
the curvature is just the transverse curvature RT given by equation (2.5.8). Let us
choose a local foliate unitary transverse frame (X1 , . . . , Xn ), and denote by ΩT the
transverse curvature 2-form with respect to this frame. We have
Proposition 7.5.16: ΩT is a basic (1, 1)-form.
Proof. The transition functions on the transverse coordinates for the foliation
Fξ are holomorphic and only involve the transverse coordinates. So the usual proof
carries over to this case. ¤
Since the curvature 2-form ΩT has type (1, 1) it follows as in ordinary Chern-
Weil theory that
Theorem/Definition 7.5.17: The kth Chern class ck (D) of the complex vector
bundle D is represented by the basic (k, k)-form γk determined by the formula
³ 1 ´
det 1ln − ΩT = 1 + γ1 + · · · + γk .
2πi
k,k
Since γk is a closed basic (k, k)-form it represents an element in HB (Fξ ) ⊂
2k th
HB (Fξ ) that is called the basic k Chern class and denoted by ck (Fξ ).
Notice that the Chern classes ck (D) are invariants of the underlying contact
structure.
Recall the transverse Ricci tensor RicT of gT . It is related to the Ricci tensor
Ricg of g by the second equation in Theorem 7.3.12. Now as usual define the Ricci
form ρg and transverse Ricci form ρTg by
(7.5.9) ρg (X, Y ) = Ricg (X, ΦY ), ρTg (X, Y ) = RicT (X, ΦY )
for smooth sections X, Y of D. It is easy to check that these are anti-symmetric of
type (1, 1) and Theorem 7.3.12 implies that they are related by
(7.5.10) ρTg = ρg + 2dη .
244 7. K-CONTACT AND SASAKIAN STRUCTURES

Thus, as in the usual case we have


1,1
Lemma 7.5.18: The basic class 2πc1 (Fξ ) ∈ HB (Fξ ) is represented by the trans-
T
verse Ricci form ρg .
More importantly there is a transverse version of Yau’s famous Theorem given
by El Kacimi-Alaoui [EKA90]:
Theorem 7.5.19: If 2πc1 (Fξ ) is represented by a real basic (1, 1) form ρT , then
it is the Ricci curvature form of a unique transverse Kähler form ω T in the same
basic cohomology class as dη.
This is translated into Sasakian language as
Theorem 7.5.20: Let (M, ξ, η, Φ, g) be a Sasakian manifold whose basic first Chern
1
class c1 (Fξ ) is represented by the real basic (1, 1) form 2π ρ, then there is a unique
Sasakian structure (ξ, η1 , Φ1 , g1 ) ∈ F(ξ) homologous to (ξ, η, Φ, g) such that ρg1 =
ρ − 2dη1 is the Ricci form of g1 , and η1 = η + ζ1 , with ζ1 = 12 dc φ. The metric g1
and endomorphism Φ1 are then given by equations (7.5.6).
We also have
Proposition 7.5.21: The basic Chern classes ck (Fξ ) are independent of the choice
of a Sasakian structure in F(Fξ ). In particular, if Fξ is quasi-regular ck (Fξ ) only
depends on the Sasaki-Seifert structure.
Proof. Since the γk is invariant under a GL(n, C) change of local frame, it
is also invariant under a homothety of the transverse metric. So the basic Chern
classes ck (Fξ ) are invariant under a transverse homothety, as well as invariant under
the addition of a (1, 1)-form in the image of ∂ ∂.¯ ¤
We denote by ck (D)R the image of the class ck (D) under the natural inclusion
H 2k (M, Z)−−→H 2k (M, R).
Lemma 7.5.22: The basic Chern class ck (Fξ ) maps to ck (D)R under the natural
2k
inclusion ι∗ . Moreover, an element [α]B ∈ HB (Fξ ) maps to ck (D)R if and only if
there exists a closed basic (2k − 2)-form β such that
[α]B = ck (Fξ ) + [Lβ]B .
In particular, ck (D) is a torsion class if and only if there exists a closed basic
(2k − 2)-form β such that ck (Fξ ) = [Lβ]B .
Proof. Since the connection in D is the transverse connection ∇T and its
curvature form is basic by Proposition 7.5.16, the Chern classes ck (D) are computed
by the γk ’s of Theorem/Definition 7.5.17. This proves the first statement.
From the exact sequence (7.2.1) we have the diagram of exact sequences
0



y
2k
(7.5.11) H (M,
 Z)


y
δ ι∗
2k−2 2k
· · · −→HB (Fξ ) −−→ HB (Fξ ) −−−→ H 2k (M, R) −−→ · · · .
So the second statement follows since δ([β]B ) = [dη]B ∪ [β]B = [dη ∧ β]B = [Lβ]B ,
and the last statement is clear from the diagram (7.5.11). ¤
7.5. SASAKIAN GEOMETRY AND ALGEBRAIC GEOMETRY 245

In case of a quasi-regular Sasakian structure, the basic Chern classes are related
to the real orbifold Chern classes in a rather obvious way, namely:
Proposition 7.5.23: Let (M, ξ, η, Φ, g) be a quasi-regular Sasakian manifold with
S 1 orbibundle π : M −−→Z. Then as real cohomology classes ck (Fξ ) = π ∗ corbk (Z).

Proof. This follows immediately from the fact that the transverse Ricci form
ρT is the pullback of the Ricci form ρ̂ on the orbifold Z. ¤
We now concentrate on the first Chern classes c1 (D) and c1 (Fξ ). We give an
immediate corollary of Lemma 7.5.22, but first a definition.
Definition 7.5.24: A Sasakian structure S = (ξ, η, Φ, g) ∈ F(Fξ ) or a Sasaki-
Seifert structure F(Fξ ) is said to be of positive (negative) type if c1 (Fξ ) can be
represented by a positive (negative) definite (1, 1)-form. If either of these two con-
ditions is satisfied F(Fξ ) is said to be of definite type. F(Fξ ) is said to be of null
type if c1 (Fξ ) = 0. A Sasakian structure (ξ, η, Φ, g) which is neither definite nor
null is said to be of indefinite type. A Sasakian structure S = (ξ, η, Φ, g) ∈ F(Fξ )
is called anticanonical (canonical) if corb 1 (Z) is a positive (negative) multiple of
[dη]B .
We apply the definition to F(Fξ ) whether Fξ is quasi-regular or not. We usually
drop the terminology ‘type’ by saying for example, ‘a positive Sasakian structure’
instead of ‘a Sasakian structure of positive type’. The following self-evident result
says that being anticanonical or canonical is a property of the deformation class
F(Fξ ).
Proposition 7.5.25: If S = (ξ, η, Φ, g) ∈ F(Fξ ) is anticanonical (or canonical)
then so is any other Sasakian structure S 0 = (ξ, η, Φ, g) ∈ F(Fξ ). An anticanonical
(canonical) Sasakian structure is positive (negative), respectively.
The converse of he last statement is certainly not true as seen from the Wang-
Ziller manifolds of Example 7.6.10 below.
Now we consider the case k = 1 in Lemma 7.5.22. Then diagram (7.5.11)
becomes
0



y
2
H (M,
 Z)


y
δ ι∗
0 2
0−→HB (Fξ ) −−→ HB (Fξ ) −−−→ H 2 (M, R) −−→ · · · ,
and Lemma 7.5.22 implies
Corollary 7.5.26: Let S = (ξ, η, Φ, g) be a Sasakian structure with underlying
contact bundle D. Then c1 (D) is a torsion class if and only if there exists a real
number a such that c1 (Fξ ) = a[dη]B . So
(i) if c1 (Fξ ) 6= 0 then c1 (D) is a torsion class if and only if the Sasakian
structure S is either anticanonical or canonical,
(ii) if S is a null Sasakian structure then c1 (D) is a torsion class.
Remark 7.5.2: Let M be a quasi-regular Sasakian manifold with b2 (M ) > 0 and
space of leaves Z = M/Fξ with its induced orbifold structure Z = (Z, U). Assume
also that corb
1 (Z) is either positive or negative. The condition of being anticanonical
246 7. K-CONTACT AND SASAKIAN STRUCTURES

or canonical chooses a ray in the Kähler cone K(Z). Thus, there are many positive
(negative) Sasakian structures that are not anticanonical or canonical corresponding
to choices of Seifert S 1 -bundle (orbibundle) whose first Chern class c1 (M/Z) is not
a multiple of c1 (Fξ ). The situation is different in the null case. If corb
1 (Z) = 0,
every (1, 1) class in K(Z) satisfying the correct positivity and integrality conditions
gives rise to a null Sasakian structure which stand more or less on equal footing.
For example as discussed later in Chapter 11, all null Sasakian structures admit
a null η-Einstein metric, whereas, only relatively few positive (negative) Sasakian
structures are anticanonical (canonical) which is a necessary condition to admit a
non-null η-Einstein metric.
In analogy with common terminology of smooth algebraic varieties we see that
a positive Sasakian structure is a transverse Fano structure, while a null Sasakian
structure is a transverse Calabi-Yau structure. The negative Sasakian case cor-
responds to the canonical bundle being ample, and could be called a transverse
canonical structure; however, this conflicts with the terminology introduced in Def-
inition 7.5.24, so we do not use it.
Note that c1 (D) is an invariant of the complex vector bundle D, but it is
related to a topological invariant, namely the second Stiefel-Whitney class w2 (M ).
It is well-known (cf. [LM89]) that a vector bundle E admits a spin structure if
and only if w2 (E) = 0, and that the distinct spin structures on E are in one-to-one
correspondence with the elements of H 1 (M, Z2 ). Furthermore, if E is a complex
vector bundle then w2 (E) is the mod 2 reduction of c1 (E). Indeed, let (M, η) be a
strict contact manifold with contact bundle D. Let ξ be the Reeb vector field, then
we have a natural splitting
T M = D ⊕ Lξ ,
where Lξ is the trivial real line bundle generated by ξ. Thus,

w2 (M ) = w2 (T M ) = w2 (D ⊕ Lξ ) = w2 (D) .

So if M admits a compatible Sasakian structure, the contact bundle D is a complex


vector bundle, so w2 (M ) is the mod 2 reduction of c1 (D) ∈ H 2 (M, Z). Hence,
Corollary 7.5.26 implies
Theorem 7.5.27: Let (M, ξ, η, Φ, g) be a Sasakian manifold with H1 (M, Z)tor = 0,
and suppose that c1 (Fξ ) = a[dη]B for some real number a. Then M is spin. In
particular, if M satisfies H1 (M, Z)tor = 0 and admits an anticanonical, canonical,
or null Sasakian structure, then M is spin.
Theorem 7.5.27 says that w2 (M ) is an obstruction for a simply connected man-
ifold M to admit a Sasakian structure S = (ξ, η, Φ, g) satisfying c1 (Fξ ) = a[dη]B ,
and the next example shows that some condition on H1 (M, Z) is necessary. (Note
that Proposition 2.6 of [BGN03a] is incorrect. This is illustrated by Example
7.6.10 below where a counterexample is given. The correct statement concerning
Sasakian manifolds and spin is Theorem 7.5.27).
Example 7.5.28: Consider the real projective spaces RP2n+1 which clearly admit
Sasakian structures (the induced metric from round sphere metric on S 2n+1 is
2
Sasakian). For any Sasakian structure HB (Fξ ) is one dimensional, so c1 (Fξ ) =
a[dη]B is automatically satisfied. Moreover [LM89],

w2 (RP2n+1 ) ≡ (n + 1) mod 2 .
7.5. SASAKIAN GEOMETRY AND ALGEBRAIC GEOMETRY 247

So RP2n+1 is not spin if n is even. This shows that a condition on H1 (M, Z) is


needed.
The type of Sasakian structures that can occur on rational homology spheres
are quite restrictive. We can give a characterization of type of Sasakian structures
that can occur on rational homology spheres in terms of the basic cohomology ring

HB (Fξ ).
Proposition 7.5.29: Let M 2n+1 be a manifold with a Sasakian structure S =
(ξ, η, Φ, g). Then M 2n+1 is a rational homology sphere if and only if the basic coho-

mology ring HB (Fξ ) is isomorphic to the real cohomology ring of complex projective
space CPn . Moreover, any Sasakian structure S on a rational homology sphere sat-
isfies c1 (Fξ ) = a[dη]B for some non-zero constant a, or equivalently, the Sasakian
structure S is either anticanonical or canonical. Hence, S is either positive or
negative.
Proof. To prove the first statement we analyze the basic long exact sequence
(7.2.1). If M is a rational homology sphere this implies the isomorphisms
p−1 p+1
HB (Fξ ) ≈ HB (Fξ )
1 2
for p = 1, . . . , 2n − 1, and since HB (Fξ ) = 0 and HB (Fξ ) is generated by [dη]B ,
this implies the ring isomorphism

£ ¤
(7.5.12) HB (Fξ ) ≈ R [dη]B /([dη]B )n+1 .
Conversely, the ring isomorphism (7.5.12) implies that the connecting homomor-
phism δ in (7.2.1) is an isomorphism. Putting p = 2r+1 shows that H 2r+1 (M, R) =
H 2r+2 (M, R) = 0 for r = 0, . . . , n − 1 which implies that M 2n+1 is a rational ho-
mology sphere. This proves the first statement.
For the second statement we first notice that the ring isomorphism 7.5.12 im-
plies that c1 (Fξ ) = a[dη]B for some a ∈ R. Moreover, a cannot vanish since this
would imply that the basic geometric genus pg (Fξ ) 6= 0 implying bn (M ) 6= 0 by the
transverse Hodge decomposition 7.2.6 which would give a contradiction. ¤
Notice that Proposition 7.5.29 and Theorem 7.5.27 have the following
Corollary 7.5.30: A rational homology sphere M with H1 (M, Z)tor = 0 which
admits a Sasakian structure is a spin manifold.
An example of a rational homology sphere satisfying H1 (M, Z)tor = 0 is the
simply connected homogeneous 5-manifold SU (3)/SO(3). This cannot admit a
Sasakian structure. In fact as discussed in Section 10.2 below SU (3)/SO(3) does
not admit a contact structure nor an almost contact structure. Other examples of
rational homology 5-spheres, due to Kollár [Kol05], which cannot admit a Sasakian
structure are discussed in Chapter 10. We shall see some non-spin Sasakian mani-
folds in Example 7.6.10 below.
Although not every Sasakian manifold is spin, every Sasakian manifold (in
fact, any K-contact manifold) admits a Spinc -structure since w2 (M ) is the mod
2 reduction of an integral class, namely c1 (D). In fact since any Sasakian mani-
fold M 2n+1 admits a reduction of the frame bundle to the group U (n) × 1, there
is a canonical Spinc -structure on M 2n+1 determined by the canonical homomor-
phism U (n)−−→Spinc2n . We refer to Appendix D of [LM89] for the details of Spinc -
structures. The observation that any Sasakian manifold is Spinc was first made in
[Mor97].
248 7. K-CONTACT AND SASAKIAN STRUCTURES

7.5.3. Positive Sasakian Structures. In this book a special role is played


by Sasakian structures with positive basic first Chern class c1 (Fξ ). Our main results
in this area appear in [BGN03a]. However, before embarking on this discussion we
give a very brief discussion of the well-known obstructions to the existence of pos-
itive scalar curvature metrics. For a more thorough treatment we refer the reader
to [LM89, RS01]. The first of these obstructions was discovered in 1963 by Lich-
nerowicz [Lic63a] in his study of harmonic spinors in dimension 4n. Later Hitchin
[Hit74] generalized this to an invariant α(M ) which lives in the real K-theory
group KOn which is non-vanishing only in dimensions 0 mod 4, or 1, 2 mod 8.
Then the Hitchin-Lichnerowicz Theorem says that if M is a compact spin manifold
with α(M ) 6= 0, then M does not admit a metric of positive scalar curvature. A
converse to this theorem was attacked by Gromov and Lawson [GL80] with much
progress toward its solution, but the full converse was left as a conjecture, known
as the Gromov-Lawson conjecture. The simply connected case was settled in 1992
by Stolz [Sto92], but the full conjecture still remains open. Stolz proved that a
simply connected manifold of dimension n ≥ 5 admits a metric of positive scalar
curvature if and only if α(M ) = 0. This is known to be false in dimension 4 [RS01].
For a complete statement of the Gromov-Lawson conjecture and its variants the
reader is referred to [RS01]. In the context of Sasakian geometry in dimension 1
mod 8 one can ask
Question 7.5.2: Are there Sasakian manifolds that do not admit metrics of posi-
tive scalar curvature?
From Stolz’s Theorem or the earlier result of Gromov and Lawson [GL80] any
simply connected manifold of dimension 5 or 7 admits a metric of positive scalar
curvature, so for an affirmative answer to the question in the simply connected case
one should start with dimension 9 and the Hitchin exotic spheres which we briefly
discuss later in Section 9.4.
We now turn to a general discussion of positive Sasakian structures. Before
doing so we mention that Kollár [Kol05] has recently found strong conditions on
the torsion subgroup of H2 (M, Z) for a simply connected five manifold M to admit
a positive Sasakian structure. We describe this in detail in Chapter 10.
Theorem 7.5.31: Let S = (ξ, η, Φ, g) be a positive Sasakian structure on a com-
pact manifold M of dimension 2n + 1. Then M admits a Sasakian structure S 0 =
(ξ 0 , η 0 , Φ0 , g 0 ) with positive Ricci curvature a-homologous to S.
Proof. c1 (Fξ ) can be represented by a positive definite basic (1, 1)-form ρ,
so by Theorem 7.5.20 there is a Sasakian structure S1 = (ξ, η1 , Φ1 , g1 ) ∈ F(ξ)
homologous to (ξ, η, Φ, g) such that ρg1 = ρ − 2dη1 is the Ricci form of g1 . Let g1T
denote the transverse Kähler metric of this Sasakian structure. Then by Theorem
7.3.12 the Ricci curvatures of g1T and g1 are related by
Ricg1 |D1 ×D1 = Ricg1T − 2g1T .
Next for any real number a > 0 we can perform a transverse homothety (cf. equa-
tion (7.3.10)) of the Sasakian structure by defining
1 1
g2T = g1T , η2 = η1 , ξ2 = aξ, Φ2 = Φ1 ,
a a
in which case S2 = (ξ2 , η2 , Φ2 , g2 ) is a Sasakian structure where g2 = g2T + η2 ⊗ η2
by Proposition 7.3.20. Notice also that S1 and S2 both have the same contact
7.5. SASAKIAN GEOMETRY AND ALGEBRAIC GEOMETRY 249

subbundles with the same underlying transverse complex structures, and the same
characteristic foliations. Now since the Ricci tensor is invariant under homothety
we find
2
Ricg2 |D2 ×D2 = Ricg2T − 2g2T = Ricg1T − g1T .
a
But since Ricg1T > 0 and M is compact there exists a0 ∈ R+ such that for all a > a0
we have Ricg2 |D2 ×D2 > 0. But also for any Sasakian metric we have Ricg2 (X, ξ2 ) =
2nη2 (X) which implies that the Ricci curvature of g2 is positive. But clearly S2 is
a-homologous to S which proves the result. ¤
Next we have a variation of Myers’ Theorem based on a result due to Hasegawa
and Seino [HS82].
Theorem 7.5.32: Let M be a manifold which admits a positive Sasakian structure
S = (ξ, η, Φ, g) with g complete and c1 (Fξ ) > δ for some δ > 0. Then M is compact
with finite fundamental group.
Proof. As in the Proof of Theorem 7.5.31 we have an a-homologous Sasakian
structure (aξ, η2 , Φ2 , g2 ) such that
2 2
Ricg2 |D2 ×D2 = Ricg2T − 2g2T = Ricg1T − g1T > δ −
a a
which is positive for a large enough. So the result follows by (i) of Theorem 7.3.12
and Myers’ Theorem (cf. [Pet98]). ¤
As in algebraic geometry positivity implies strong restrictions on the type of
transverse holomorphic structure allowed. So we can think about positive Sasakian
structures as transverse Fano structures as the following result [BGN03a] suggests.
Proposition 7.5.33: Let S = (ξ, η, Φ, g) be a quasi-regular positive Sasakian struc-
ture on a compact manifold M with S 1 orbibundle π : M −−→Z, then
(i) Ricg > −2 and the orbifold Z is Fano, i.e., corb 1 (Z) > 0.
(ii) hp,0 (Fξ ) = 0 for all p > 0, so χhol (Fξ ) = 1, and pg (Fξ ) = pa (Fξ ) = 0.
(iii) The underlying complex space Z of the orbifold Z is simply connected.
(iv) As an algebraic variety Z is uniruled and its Kodaira dimension is −∞.
Proof. (i) follows from equation (7.5.10). (ii): Any non-trivial element of
HBp,0
(Fξ ) is represented by a basic (p, 0)-form in the kernel of ∂,¯ i.e., by a trans-
versely holomorphic p-form α. Since S = (ξ, η, Φ, g) is quasi-regular, M is the total
space of an orbifold S 1 V-bundle over a compact Kähler orbifold Z. Moreover, since
α is basic, it descends to a non-trivial element α̌ ∈ H p,0 (Z). Now since the Sasakian
structure S is positive so is the Kähler structure on Z, i.e. corb1 (Z) > 0, and this
implies hp,0 (Z) = 0 by the Kodaira-Baily vanishing Theorem 4.4.28.
The proof of simple connectivity is essentially a translation of Kobayashi’s proof
in [Kob61] for the case of manifolds to the case of complex orbifolds. By positivity
of the transverse Ricci form ρT = π ∗ ρZ , the orbifold Z has at most a finite cover.
By (ii) the transverse holomorphic Euler number χ(Z, OZ ) = 1, and this must
also hold on any finite cover. But from the above Z is projective algebraic, hence
complete. So the version of the Hirzebruch-Riemann-Roch Theorem for singular
algebraic varieties due to Baum, Fulton, and Macpherson (cf. [Ful84],pg 354)
applies and we have
1 = χ(Z, OZ ) = hTd(Z), [Z]i ,
250 7. K-CONTACT AND SASAKIAN STRUCTURES

where Td(Z) is the Todd class. Moreover, this equation holds on any k-fold cover
Z̃ as well, and hTd(Z̃), [Z̃]i = khTd(Z), [Z]i . This forces k = 1. ¤
Similar results proving simple connectivity in a purely algebro-geometric set-
ting, that is without the orbifold hypothesis appear in [Tsu88] and [Tak00]. If we
restrict ourselves to the regular case we obtain a finiteness result as a corollary of the
well-known rational connectedness of smooth Fano varieties [Cam92, KMM92],
namely
Theorem 7.5.34: There is at most a finite number of deformation classes F(Fξ )
of regular positive Sasakian structures in any dimension.

7.6. New Sasakian Structures from Old


In this section we describe methods for obtaining new Sasakian structures from
old ones. The first is an invariance condition for submanifolds analogous to the
situation in Kähler geometry. The second is a join construction which applies
to quasi-regular Sasakian structures. This was originally constructed for Sasaki-
Einstein structures in [BG00b] and developed further recently in [BGO07].
7.6.1. Immersions and Invariant Submanifolds. We begin with a defini-
tion about submanifolds of Sasakian manifolds.
Definition 7.6.1: Let (M 2n+1 , ξ, η, Φ, g) be a K-contact manifold. An immersed
submanifold N of M is said to be an invariant submanifold if the two conditions
hold:
(i) ξ is tangent to N at all points of N,
(ii) ΦTp N ⊂ Tp N for all points p ∈ N.
This leads to the simple but important result of Okumura [Oku68] (see also
[YK84]).
Theorem 7.6.2: An invariant submanifold N of a Sasakian (K-contact) manifold
(M 2n+1 , ξ, η, Φ, g) is Sasakian (K-contact) whose structure is given by the restric-
tion of the tensor fields (ξ, η, Φ, g) to N.
Proof. The two conditions of Definition 7.6.1 guarantee that ξ and Φ re-
stricted to well-defined tensor fields on N. Of course, the restriction of the tensor
fields η and g to N are always well-defined. Hence, the tensor fields (ξ, η, Φ, g)
satisfy all the relations on N that they satisfy on M, in particular the normality
condition 6.5.9, NΦ = −2ξ ⊗ dη, is satisfied. Thus, N with the structure (ξ, η, Φ, g)
is Sasakian. ¤
Theorem 7.6.2 will be of much interest to us in Chapter 9. Many further
results about submanifolds of Sasakian manifolds can be found in [YK84] and
references therein. In view of Theorem 7.6.2 it is natural to ask whether there is a
CR or Sasakian version of the famous Kodaira Embedding Theorem 3.5.12. Such
a theorem was claimed by Ornea and Verbitsky in [OV05] as an offshoot of their
study of Vaisman immersion theorems. But they have recently shown [OV06b]
that their claim was mistaken and have provided a fix. We shall not delve into
the details of Vaisman manifolds. Suffice it to say that they are a special type
of locally conformally Kähler manifold and that if M is a Sasakian manifold then
M × S 1 is a Vaisman manifold. Thus, a Vaisman manifold is a kind of generalized
Hopf manifold which is actually the name that Vaisman gave to them [Vai82]. See
7.6. NEW SASAKIAN STRUCTURES FROM OLD 251

the book [DO98] of Dragomir and Ornea for details and references. Here is the
embedding theorem of [OV06b]:
Theorem 7.6.3: Let N be a compact Sasakian manifold. Then there is a positive
integer N and CR-embedding of N into the sphere S 2N +1 such that cone C(S 2N +1 )
is biholomorphic to CN \ {0}.
It is important to note that the induced CR structure on S 2N +1 is not generally
the standard CR structure on S 2N +1 . Another related result [MY06] gives a
CR embedding of not necessarily compact Sasakian manifolds into CN , but not
necessarily into a sphere.

7.6.2. The Join Construction. In this section we apply a construction due


to Wang and Ziller [WZ90] to define a multiplication on the set of quasi-regular
Sasakian orbifolds. This was done originally in [BG00b] in the case of Sasaki-
Einstein orbifolds which is perhaps of more interest, but there is an easy general-
ization to the strict Sasakian case [BGO07]. The idea is quite simple and is based
on the fact that product of Kähler orbifolds can be given a Kähler orbifold structure
in many natural ways.
Definition 7.6.4: We denote by SO the set of compact quasi-regular Sasakian
orbifolds, by SM the subset of SO that are smooth manifolds, and by R ⊂ SM the
subset of compact, simply connected, regular Sasakian manifolds. The set SO is
topologized with the C m,α topology, and the subsets are given the subspace topology.
The set SO is graded by dimension, that is,

M
(7.6.1) SO = SO2n+1 ,
n=1

and similarly for SM and R. For each pair of relatively prime positive integers
(k1 , k2 ) we define a graded multiplication
(7.6.2) ?k1 ,k2 : SO2n1 +1 × SO2n2 +1 −−→SO2(n1 +n2 )+1
as follows: Let M1 , M2 ∈ SO of dimension 2n1 + 1 and 2n2 + 1, respectively. Since
each orbifold (Mi , Si ) has a quasi-regular Sasakian structure Si , its Reeb vector
field generates a locally free circle action, and the quotient space by this action has
a natural orbifold structure Zi [Mol88]. Thus, there is a locally free action of the 2-
torus T 2 on the product orbifold M1 ×M2 , and the quotient orbifold is the product
of the orbifolds Zi . (Locally free torus actions on orbifolds have been studied in
[HS91]). Now the Sasakian structure on Mi determines a Kähler structure ωi
on the orbifold Zi , but in order to obtain an integral orbifold cohomology class
2
[ωi ] ∈ Horb (Zi , Z) we need to assure that the period of a generic orbit is one. By a
result of Wadsley [Wad75] the period function on a quasi-regular Sasakian orbifold
is lower semi-continuous and constant on the dense open set of regular orbits. This
is because on a Sasakian orbifold all Reeb orbits are geodesics. Thus, by a transverse
homothety we can normalize the period function to be the constant 1 on the dense
open set of regular orbits. In this case the Kähler forms ωi define integer orbifold
2
cohomology classes [ωi ] ∈ Horb (Zi , Z). If Zi denotes the underlying complex space

associated with the orbifold Zi , one should not confuse Horb (Zi , Z) with H ∗ (Zi , Z).
However, they are isomorphic rationally. Now each pair of positive integers k1 , k2
give a Kähler form k1 ω1 + k2 ω2 on the product. Furthermore, [k1 ω1 + k2 ω2 ] ∈
252 7. K-CONTACT AND SASAKIAN STRUCTURES

2
Horb (Z1 × Z2 , Z), and thus defines an S 1 V-bundle over the orbifold Z1 × Z2 whose
total space is an orbifold that we denote by M1 ?k1 ,k2 M2 . We have
Definition 7.6.5: The orbifold M1 ?k1 ,k2 M2 constructed above is called the (k1 , k2 )-
join of (M1 , S1 ) and (M2 , S2 ).
By construction M1 ?k1 ,k2 M2 admits a quasi-regular Sasakian structure by
choosing a connection 1-form on M1 ?k1 ,k2 M2 whose curvature is π ∗ (k1 ω1 + k2 ω2 ).
This Sasakian structure is unique up to a gauge transformation of the form η 7→
η + dψ where ψ is a smooth basic function and will be denoted by S1 ?k1 ,k2 S2 .
This defines the maps in (7.6.2). If Si are quasi-regular Sasakian structures on
the compact manifolds Mi , respectively, we shall use the notation S1 ?k1 ,k2 S2 and
M1 ?k1 ,k2 M2 interchangeably depending on whether we want to emphasize the
Sasakian or manifold nature of the join. Notice also that if gcd(k1 , k2 ) = m and
(k10 , k20 ) = ( km1 , km2 ), then gcd(k10 , k20 ) = 1 and M1 ?k1 ,k2 M2 ≈ (M1 ?k10 ,k20 M2 )/Zm . In
this case the cohomology class k10 ω1 + k20 ω2 is indivisible in Horb 2
(Z1 × Z2 , Z). Note
that M1 ?k1 ,k2 M2 can be realized as the quotient (M1 × M2 )/S 1 (k1 , k2 ), where the
S 1 action is given by the map
(7.6.3) (x, y) 7→ (eik2 θ x, e−ik1 θ y) .
Alternatively, we can think of the (k1 , k2 )-join as an operation on the set of Sasaki-
Seifert structures endowed with the quotient topology.
We are interested in restricting the map ?k1 ,k2 of (7.6.2) to the subset of smooth
Sasakian manifolds, that is in the map
(7.6.4) ?k1 ,k2 : SM2n1 +1 × SM2n2 +1 −−→SO2(n1 +n2 )+1 .
If M1 and M2 are quasi-regular Sasakian manifolds, we are interested under what
conditions the orbifold M1 ?k1 ,k2 M2 is a smooth manifold. Recall from Definition
7.1.1 the order υ of a quasi-regular Sasakian manifold. We have
Proposition 7.6.6: Let (M1 , S1 ) and (M2 , S2 ) be compact quasi-regular Sasakian
manifolds of orders υ1 and υ2 , respectively. Then for each pair of relatively prime
positive integers k1 , k2 , the orbifold (M1 ?k1 ,k2 M2 , S1 ?k1 ,k2 S2 ) is a smooth quasi-
regular Sasakian manifold if and only if gcd(υ1 k2 , υ2 k1 ) = 1. Furthermore, if M1
and M2 are simply connected, so is M1 ?k1 ,k2 M2 .
Proof. Since M1 × M2 is a 2-torus V-bundle over Z1 × Z2 , the S 1 V-bundle
M1 ?k1 ,k2 M2 can be realized as a quotient (M1 × M2 )/S 1 by choosing a certain ho-
momorphism S 1 −−→S 1 ×S 1 . For the V-bundle on Z1 ×Z2 with first Chern class given
by [k1 ω1 + k2 ω2 ] we have the action on M1 × M2 defined by (x, y) 7→ (τ k2 x, τ −k1 y).
The condition that M1 ?k1 ,k2 M2 is in SM is that there are no fixed points (x, y)
under the above action. If υx , υy denote the orders of the leaf holonomy groups
of x, y, respectively, then the condition that (x, y) be a fixed point under the S 1
action generated by the Reeb vector field is that gcd(υx k2 , υy k1 ) = g > 1. This con-
dition never holds if and only if gcd(υx k2 , υy k1 ) = 1 for all pairs (x, y). This proves
the first statement. The second statement follows from the long exact homotopy
sequence. ¤
We make note of some special cases. If M1 is regular, then M1 ?k1 ,k2 M2 is
smooth if and only if k2 is chosen relatively prime to the order υ2 . If Mi are both
regular Sasakian manifolds, then so is M1 ?k1 ,k2 M2 . The following proposition is
essentially given in [WZ90].
7.6. NEW SASAKIAN STRUCTURES FROM OLD 253

Proposition 7.6.7: Let (M1 , S1 ) and (M2 , S2 ) be compact quasi-regular Sasakian


manifolds and assume gcd(υ1 k2 , υ2 k1 ) = 1. Then M1 ?k1 ,k2 M2 is the associated S 1
orbibundle over Z1 with fibre M2 /Zk2 . In particular, if M1 is regular and k2 = 1 then
for each positive integer k, M1 ?k,1 M2 is an M2 -bundle over the Kähler manifold
Z1 .
Proof. Following [WZ90] we break up the S 1 action on M1 × M2 into stages.
First divide by the subgroup Zk2 of the circle group S 1 (k1 , k2 ) defined by equation
(7.6.3) giving M1 × (M2 /Zk2 ). Letting [y] ∈ M2 /Zk2 denote the equivalence class of
y ∈ M2 , we see that the quotient group S 1 /Zk2 acts on M1 × M2 /Zk2 by (x, [y]) 7→
k1
(eiθ x, [e−i k2 θ y]) which identifies M1 ?k1 ,k2 M2 as the orbibundle over Z1 with fibre
M2 /Zk2 associated to the principal S 1 orbibundle π1 : M1 −−→Z1 . This proves the
result. ¤
Notice that the type of the two Sasakian manifolds is compatible with the join
operation in the following sense.
Proposition 7.6.8: Let (M1 , S1 ) and (M2 , S2 ) be compact quasi-regular Sasakian
manifolds. Then the join M1 ?k1 ,k2 M2 is positive, negative, or null if and only if
both Sasakian manifolds M1 and M2 are positive, negative, or null, respectively.
Proof. It follows from Theorem 7.1.3 and Proposition 7.5.23 that for any
quasi-regular Sasakian structure on a compact manifold M we have an orbifold
submersion π : M −−→Z satisfying c1 (Fξ ) = π ∗ corb
1 (Z) as real cohomology classes.
So the sign or vanishing of c1 (Fξ ) and corb
1 (Z) coincide. Furthermore, for any pair
of integers (k1 , k2 ) we have
c1 (F1 ?k1 ,k2 F2 ) = π ∗ corb ∗ orb ∗ orb
1 (Z1 × Z2 ) = π c1 (Z1 ) + π c1 (Z2 ) = c1 (F1 ) + c1 (F2 ) ,

where F1 ?k1 ,k2 F2 is the foliation defined by the structure S1 ?k1 ,k2 S2 on M1 ?k1 ,k2 M2
and F1 , F2 are the characteristic foliations of S1 , S2 , respectively. Now suppose
that the Sasakian structures S1 and S2 are both definite of the same type. Then
c1 (F1 ) + c1 (F2 ) can be represented by either a positive definite or negative definite
basic (1, 1)-form. The null case is clear. ¤
We are interested in computing the topology of the join M1 ?k1 ,k2 M2 . We first
obtain some general information about the low Betti numbers of the (k1 , k2 )-join
of two orbifolds which follows easily from Theorem 7.4.14
Lemma 7.6.9: Let Mi ∈ SO2ni +1 and let M1 ? M2 be any (k1 , k2 )-join. Then
(i) b1 (M1 ? M2 ) = b1 (M1 ) + b1 (M2 ),
(ii) b2 (M1 ? M2 ) = b2 (M1 ) + b2 (M2 ) + 1 if ni ≥ 1,
(iii) b3 (M1 ? M2 ) = b3 (M1 ) + b3 (M2 ) if ni ≥ 3,
(iv) b4 (M1 ?M2 ) = b4 (M1 )+b4 (M2 )+b2 (M1 )b2 (M2 )+b2 (M1 )+b2 (M2 )+1
if ni ≥ 4.
Exercise 7.14: Prove Lemma 7.6.9
When ni is outside the indicated range, the formulae are slightly different, but
are easily worked out. For general br the formulae are increasingly more compli-
cated and are different depending on whether r is even or odd, or whether the
range conditions are satisfied or not. In order to determine the cohomology of
the (k1 , k2 )-join M1 ?k1 ,k2 M2 of two Sasakian manifolds one needs to analyze the
Leray-Serre spectral sequences in specific cases. Following [WZ90] we describe
254 7. K-CONTACT AND SASAKIAN STRUCTURES

the general setup. First we apply the classifying functor to an orbifold fibration
S 1 −−→M −−→Z giving the fibration M −−→BZ−−→BS 1 to which we can apply a
Leray-Serre spectral sequence. Next for a pair of Sasakian manifolds Mi ∈ SM,
we consider the commutative diagram of orbifold fibrations
1
M1 × M2 −−−−→ B(M1 ?k1 ,k2 M2 ) −−−−→ BS

  
  
  ψ
(7.6.5) = y 
y y
M1 × M2 −−−−→ BZ1 × BZ2 −−−−→ BS 1 × BS 1 .
The maps are all the obvious ones. In particular, ψ is determined by the S 1 action
ψ(τ ) = (τ k2 , τ −k1 ). Of course, if M1 ?k1 ,k2 M2 is a smooth manifold then the B
does not occur in the middle term of the top row. The important point in the
analysis of this diagram is that the differentials in the spectral sequence of the top
fibration are determined through naturality by the differentials in the sequence of
the bottom fibration. Wang and Ziller [WZ90] apply this method to compute the
integral cohomology ring of torus bundles over products of projective spaces.
p,q
Example 7.6.10: [The Wang-Ziller manifolds] The Wang-Ziller manifolds Mk,l
studied in [WZ90] are defined to be S 1 -bundle over CPp × CPq whose first Chern
class is kα + lβ where α and β are the positive generators of H ∗ (CPp , Z) and
H ∗ (CPq , Z), respectively and k, l ∈ Z+ . Thus, they are just the (k, l)-join S 2p+1 ?k,l
p,q
S 2q+1 = Mk,l . For simplicity we assume that gcd(k, l) = 1, so by Proposition
p,q
7.6.6 Mk,l is a simply connected manifold admitting natural Sasakian structures.
Moreover, since the base of the circle bundle is CPp × CPq with its product complex
structure, these Sasakian structures are all positive by Proposition 7.6.8. To analyze
p,q
the topology of the manifolds Mk,l we follow Wang and Ziller [WZ90] and consider
the free T 2 action on S 2p+1 ×S 2q+1 defined by (x, y) 7→ (eiθ1 x, eiθ2 y) where (x, y) ∈
p,q
Cp+1 × Cq+1 . The quotient space is CPp × CPq , and Mk,l can be identified with
2p+1 2q+1
the quotient of S ×S by the circle defined by (x, y) 7→ (eilθ x, e−ikθ y).
Applying the spectral sequence analysis described above Wang and Ziller determine
p,q
the integral cohomology ring of the manifolds Mk,l . Assuming that 1 ≤ p ≤ q they
show that
p,q ¡ ¢
(7.6.6) H ∗ (Mk,l , Z) = Z[x, y]/ (lx)p+1 , xq+1 , xp+1 y, y 2 ,
where x is a 2-dimensional class and y is a 2q + 1-dimensional class. Letting π :
p,q
Mk,l −−→CPp ×CPq denote the natural bundle projection, we see that the classes α, β
pull back as ι∗ π ∗ α = lx, ι∗ π ∗ β = −kx. Here, by abuse of notation, we let π ∗ α, π ∗ β
2
also denote the basic classes in HB (Fξ ). Furthermore, the basic first Chern class is
∗ ∗
c1 (Fξ ) = (p + 1)π α + (q + 1)π β, so we get
c1 (D) = ι∗ c1 (Fξ ) = (p + 1)ι∗ π ∗ α + (q + 1)ι∗ π ∗ β = (l(p + 1) − k(q + 1))x .
Thus, we have
p,q
(7.6.7) w2 (Mk,l ) = (l(p + 1) + k(q + 1))x mod 2.
In certain cases one can determine the manifold completely [WZ90]. For example,
1,q
consider p = l = 1 in which case Mk,1 is an S 2q+1 -bundle over S 2 . The S k -bundles
2
over S are classified by π1 (SO(k + 1)) ≈ Z2 [Ste51]. So there are precisely two
S 2q+1 -bundles over S 2 , and they are distinguished by w2 . From Eequation (7.6.7)
7.6. NEW SASAKIAN STRUCTURES FROM OLD 255

1,q
we get w2 (Mk,1 ) = k(q + 1)x mod 2. Thus, if k is even or q is odd we get the trivial
bundle S × S 2q+1 ; whereas, if q is even and k is odd, we get the unique non-trivial
2

S 2q+1 -bundle over S 2 . This gives an infinite number of distinct deformation classes
F(Fξ ) of positive Sasakian structures on these manifolds. Furthermore, Wang and
Ziller show using the foundational work of Sullivan [Sul77] that for fixed q and
1,q
l the manifolds Mk,l the manifolds have only a finite number of diffeomorphism
types. In dimension five (q = 1) we can do somewhat better. In fact for any pair of
1,1
relatively prime positive integers (k, l) Mk,l is diffeomorphic to S 2 × S 3 ; whereas
later in Chapter 10 we construct a positive Sasakian structure on the non-trivial
S 3 -bundle over S 2 as well as a family of indefinite Sasakian structures. Notice for
1,1
Mk,l , c1 (D) = 2(k − l)x. Other examples can be found in [WZ90].
Generally, given two known quasi-regular Sasakian manifolds M1 and M2 , it can
be quite difficult to compute the diffeomorphism type or even the homeomorphism
type of M1 ?k1 ,k2 M2 . However, see Theorem 11.7.8 below. For further discussion
of Sasakian structures it is convenient to introduce a definition.
Definition 7.6.11: We say that (M, S) ∈ SO is S-reducible if there are positive
integers (k1 , k2 ) such that M can be written as M = M1 ?k1 ,k2 M2 for some
(M1 , S1 ), (M2 , S2 ) ∈ SO. M is S-irreducible if it is not S-reducible.
It is clear from the construction that S-irreducibility corresponds to Riemann-
ian irreducibility on the space of leaves. S-reducibility first occurs in dimension
5. Every simply connected S-reducible regular positive Sasakian 5-manifold is a
1,1
Wang-Ziller manifold Mk,l = S 3 ?k,l S 3 which is diffeomorphic to S 2 × S 3 for every
pair of relatively prime integers (k, l). Clearly (ii) of Lemma 7.6.9 implies
Corollary 7.6.12: Any Sasakian structure on a rational homology sphere is S-
irreducible.
CHAPTER 8

Symmetries and Sasakian Structures

In this chapter we will undertake a detailed study of the symmetries of con-


tact and Sasakian structures. Automorphisms of almost contact and Sasakian
structures have been investigated early on in series the papers of Tanno [Tan62,
Tan63a, Tan63b, Tan64, Tan69a, Tan70, Tan71] (cf. Vol. II of Sasaki’s
lecture notes [Sas68]). Here, in a systematic way, we begin by exploring the
relation between the isometry group and its subgroup of transformations which
also preserve the Sasakian structure. We then introduce a deformation theory for
Sasakian structures. The first specific application of the background material of
the first two section is the theory of compact homogeneous Sasakian manifolds.
This is a Sasakian version of a similar effort to understand homogeneous com-
pact complex and contact manifolds undertaken in the 50ties by Boothby and
Wang [Wan54, BW58, Boo80]. In fact, as we show, every compact homoge-
neous contact manifold admits a compatible homogeneous Sasakian structure (cf.
Theorem 8.3.6). The second half of this chapter is devoted to the study of contact
(Sasakian) moment maps, convexity, and the Delzant type theorems for toric con-
tact (Sasakian) manifolds. Contact/Sasakian reduction is indeed a natural analogue
of better known symplectic/Kähler quotient of Marsden-Weinstein1 [MW74].

8.1. Automorphisms of Sasakian Structures and Isometries


Given a Riemannian manifold (M, g) we let Isom(M, g) denote the isometry
group of g, that is, Isom(M, g) is the subgroup of Diff(M ) that leaves g invariant.
The well-known theorem of Myers and Steenrod (cf. [Kob72]) says that Isom(M, g)
is a finite dimensional Lie group and if M is compact so is Isom(M, g). We let
isom(M, g) denote the Lie algebra of Isom(M, g). We now wish to consider contact
metric structures with the metric given by equation (6.4.2). In Section 6.2.2 we
discussed several important groups associated with an almost contact structure,
namely, the automorphism group Aut(ξ, η, Φ), the group of CR transformations
CR(M, D, J), the automorphism group Aut(Φ) of the tensor field Φ, as well as the
group Con(M, D) of contact transformations in the case of a contact structure. Here
we relate these groups to the isometry group Isom(M, g) in the case of a contact
metric structure (M, ξ, η, Φ, g). First we notice that for contact structures we have
a natural string of subgroup inclusions
(8.1.1) Aut(ξ, η, Φ) ⊂ Aut(M, Φ) ⊂ CR(M, D, J) ⊂ Con(M, D) .

1Sir Michael Atiyah once remarked that the correct original reference regarding the sym-
plecting momentum mapping ought to be the works of Archimedes, c. 287-212 B.C. Indeed, the
height function on the 2-sphere is the symplectic moment map and quite clearly Archimedes was
investigating this function in his work on areas and volumes (cf. famous treatises On Conoids
and Spheroids and On the Sphere and the Cylinder).

257
258 8. SYMMETRIES AND SASAKIAN STRUCTURES

The relation with isometries can be seen by restricting the contact transformations
to the subgroup Con(M, η) of strict contact transformations.
Proposition 8.1.1: Let (M, ξ, η, Φ, g) be a contact metric manifold. Then
Aut(ξ, η, Φ) = Con(M, η) ∩ Aut(M, Φ) = Con(M, η) ∩ CR(M, D, J)
= Con(M, η) ∩ Isom(M, g) .
Furthermore, Aut(ξ, η, Φ) is a closed Lie subgroup of Isom(M, g), hence, it is a Lie
group. If M is compact so is Aut(ξ, η, Φ).

Proof. By equation (8.1.1) to prove the equalities on the first line we need
only show that if φ ∈ Con(M, η) ∩ CR(M, D, J) then φ ∈ Aut(ξ, η, Φ). But this
follows easily from Lemma 6.1.29 and equation (6.2.1). Furthermore, it is clear
from equation (6.4.2) that this implies f ∗ g = g. This proves the containment
Aut(ξ, η, Φ) ⊂ Con(M, η) ∩ Isom(M, g). For the converse, suppose f ∗ η = η and
f ∗ g = g, then using equation (6.4.2) we have
dη(ΦX, Y ) = f ∗ (dη ◦ Φ ⊗ 1l)(X, Y ) = dη(Φf∗ X, f∗ Y ) = dη(Φf∗ X, f∗ Y )
= dη(f∗ f∗−1 Φf∗ X, f∗ Y ) = f ∗ dη(f∗−1 Φf∗ X, Y ) = dη(f∗−1 Φf∗ X, Y )
which holds for all vector fields X, Y. But then f∗−1 Φf∗ = Φ by the non-degeneracy
of dη on D. This proves the equality as sets. But since all the groups involved
are defined as subgroups of Diff(M ) this is an equality as groups. This equality
characterizes Aut(ξ, η, Φ) as a subgroup of the Lie group Isom(M, g) that leaves
the 1-form η invariant. Thus, it is a closed Lie subgroup which is compact if M is
compact. ¤

Definition 8.1.2: Let (M, ξ, η, Φ, g) be a contact metric manifold. We define


the automorphism group of (M, ξ, η, Φ, g) to be the automorphism group of
the underlying almost contact structure and denote it by Aut(M, ξ, η, Φ, g). We let
aut(M, ξ, η, Φ, g) denote the Lie algebra of Aut(M, ξ, η, Φ, g).
We know by Proposition 6.4.8 that
Proposition 8.1.3: A contact metric structure S = (ξ, η, Φ, g) with underlying
almost CR structure (D, J) is K-contact if and only if ξ ∈ cr(D, J) in which case ξ
actually lies in the center of aut(ξ, η, Φ, g).
We now describe the relationship between Aut(M, ξ, η, Φ, g) and CR(M, D, J).
The study of the group CR(M, D, J) has a long and varied history [Web77, Lee96,
Sch95]. We state the most general result given in [Sch95]. This result holds for
a general strictly pseudoconvex CR manifold M , that is an integrable almost CR
manifold with a positive definite Levi form.
Theorem 8.1.4: Let M be a 2n + 1 dimensional manifold with a strictly pseudo-
convex CR structure (D, J), and CR automorphism group CR(D, J). If CR(D, J)
does not act properly on M , then:
(i) If M is non-compact, then M is CR diffeomorphic to the Heisenberg
group with its standard CR structure.
(ii) If M is compact, then M is CR diffeomorphic to the sphere S 2n+1 with
its standard CR structure.
In particular, if M is a closed manifold not CR diffeomorphic to the sphere, the
automorphism group of its CR structure is compact.
8.1. AUTOMORPHISMS OF SASAKIAN STRUCTURES AND ISOMETRIES 259

As an immediate corollary we have


Corollary 8.1.5: Let M be a 2n + 1 dimensional manifold with a strictly pseudo-
convex CR structure (D, J) on which CR(D, J) does not act properly. Then (D, J)
is of Sasaki type.
Theorem 8.1.6: Let (D, J) be a strictly pseudoconvex CR structure of Sasaki type
on a compact manifold M. Then there exists a Sasakian structure S = (ξ, η, Φ, g),
with underlying CR structure (D, J), whose automorphism group Aut(S) is a max-
imal compact subgroup of CR(D, J). In fact, except for the case when (M, D, J)
is CR diffeomorphic to the sphere S 2n+1 with its standard CR structure, the auto-
morphisms group Aut(S) of S is equal to CR(D, J). For the standard CR structure
(D, J) on the sphere, CR(S 2n+1 , D, J) = SU (n + 1, 1).
Proof. This is accomplished by averaging over the maximal compact subgroup
of CR(D, J). See [BGS06] for details. The last statement was apparently given
first in [Web77]. ¤
Now we focus on the characteristic foliation Fξ rather than the CR structure.
For a K-contact manifold we have an exact sequence
(8.1.2) 0−−→{ξ}−−→aut(ξ, η, Φ, g)−−→q(ξ, η, Φ, g)−−→0
and we wish to study the quotient algebra q(ξ, η, Φ, g).
Now on a K-contact manifold we have a transverse almost Hermitian structure
defined by a transverse almost complex structure J¯ and a compatible Hermitian
metric gT . Recall from Section 2.5 the Lie algebra trans(M, Fξ ) of transverse vector
fields which are actually equivalences classes of vector fields on M. Since taking Lie
derivatives of basic geometric objects with respect to transverse vector fields is well
¯ gT ) denote the Lie algebra of infinitesimal automorphisms of
defined, we let aut(J,
the transverse almost Hermitian structure, that is,
(8.1.3) ¯ gT ) = {X̄ ∈ trans(M, Fξ ) | £X̄ J¯ = 0,
aut(J, £X̄ gT = 0} .
¯ gT ), however, as
We would like to identify the Lie algebras q(ξ, η, Φ, g) and aut(J,
the following shows this is not always possible. First we have
Lemma 8.1.7: Let X ∈ aut(ξ, η, Φ, g) and let X̄ denote its projection onto q(ξ, η, Φ, g).
Then X̄ ∈ aut(J,¯ gT ).

Proof. This is easy. For example,


[X̄, J¯Ȳ ] = [X̄, JY
¯ ] = [X, ΦY ] = Φ[X, Y ] = J[
¯ X̄, Ȳ ]

with a similar computation to show that £X̄ gT = 0.


¤
Conversely, given an element X̃ ∈ aut(J, ¯ gT ), we would like to extend X̄ to a
vector field in X ∈ aut(ξ, η, Φ, g) by choosing as a representative of X̄ the section X̃
of D, writing X = X̃ + aξ and demanding that £X η = 0. This implies the existence
of a smooth basic function a such that da = −X̃ dη. The obstruction for solving
1
this equation lies in HB (Fξ ). In fact we have an exact sequence
δ
(8.1.4) ¯ gT ) −−→ H 1 (Fξ ) ,
0−−→{ξ}−−→aut(ξ, η, Φ, g)−−→aut(J, B

where δ(X̄) = [X̃ dη]B . We have arrived at


260 8. SYMMETRIES AND SASAKIAN STRUCTURES

Theorem 8.1.8: Let S = (ξ, η, Φ, g) be a K-contact structure on a smooth manifold


¯ gT )
M with characteristic foliation Fξ . Then a transverse vector field X̄ ∈ aut(J,
extends to an infinitesimal automorphism X ∈ aut(ξ, η, Φ, g) if and only the basic
1
cohomology class [X̃ dη]B ∈ HB (Fξ ) vanishes.
Exercise 8.1: Prove that if the obstructions [X̃ dη]B vanish for any transverse
¯ gT ), then q(ξ, η, Φ, g) and aut(J,
Killing field X̄ ∈ aut(J, ¯ gT ) are isomorphic as Lie
algebras.
1
In the quasi-regular case the vanishing of [X̃ dη]B ∈ HB (Fξ ) on M translates
to the vanishing of the class [X̌ ω] ∈ H (Z, R) on Z, where π ∗ ω = dη. This is just
1

the condition that the infinitesimal automorphism X̌ ∈ aut(J, ω) be Hamiltonian.


Thus, we can rephrase Theorem 8.1.8 as
Corollary 8.1.9: Let (Z, J, ω) be an almost Kähler orbifold with polarization de-
2
fined by [ω] ∈ Horb (Z, Z). Then an infinitesimal automorphism X̌ ∈ aut(J, ω) lifts
to an infinitesimal automorphism X ∈ aut(ξ, η, Φ, g) of the induced K-contact struc-
ture on the total space M of the circle V-bundle whose orbifold first Chern class is
[ω] (cf. Theorem 7.5.2) if and only if X̌ is Hamiltonian.
Generally in the quasi-regular case the Lie algebra aut(J, ¯ gT ) is isomorphic to
the Lie algebra of infinitesimal automorphisms of the almost Kähler orbifold M/Fξ .
Exercise 8.2: Prove this statement.
Notice that the minimal possible dimension of aut(ξ, η, Φ, g) is one. In [Tan69b]
Tanno proved that the maximal dimension of the group of automorphisms of an
almost contact metric structure is (n + 1)2 . This maximum is realized in Example
8.1.12.
Proposition 8.1.10: Let M 2n+1 be a 2n+1-dimensional manifold with a K-contact
structure (ξ, η, Φ, g) of rank k. Then
k ≤ dim aut(ξ, η, Φ, g) ≤ (n + 1)2 .

Proof. The lower bound follows from Proposition 7.1.8. By the exact se-
quence (8.1.2), it suffices to work locally on a local transversal Cn with its almost
Hermitian structure. Now any transverse Killing vector field at a point p ∈ Cn
is determined by Xpi and (∇j X i )p . From the invariance of gT and J¯ above, the
later lies in the Lie algebra u(n) of the unitary group, and the former is an arbi-
trary complex vector in Cn . Thus, the maximal dimension is n2 + 2n. Adding the
characteristic vector field ξ on M to this gives the result. ¤

Next we give some examples. The first two examples show that the maximal
dimension in Proposition 8.1.10 is realized.
Example 8.1.11: We consider the standard Sasakian structure (ξ0 , η0 , Φ0 , g0 ) on
the round sphere S 2n+1 discussed in Example 7.1.5. Since this is just the usual
Hopf fibration over CPn with its standard Fubini-Study metric, the automorphism
group Aut(ξ0 , η0 , Φ0 , g0 ) is U (n + 1) with the central S 1 being generated by the
Reeb vector field. Here Aut(ξ0 , η0 , Φ0 , g0 ) acts transitively on S 2n+1 , and realizes
the maximal dimension of Proposition 8.1.10.
Our next example shows that the exact sequence of Lie algebras (8.1.2) does
not always split, and that on a non-compact manifold aut(ξ, η, Φ, g) need not be
reductive.
8.1. AUTOMORPHISMS OF SASAKIAN STRUCTURES AND ISOMETRIES 261

Example 8.1.12: This is a continuation of Example 7.3.21. One easily sees that
the Sasakian structure S = (ξ, η, Φ, g) in 7.3.21 has a transitive group of Sasakian
automorphisms. In fact R2n+1 is the total space of the 2n + 1-dimensional Heisen-
berg group H2n+1 defined in Exercise 6.3. The Heisenberg Lie algebra h2n+1 is
generated by the left invariant vector fields
∂ ³ ∂ ∂ ´ ∂
ξ= , Si = + xi , Ri =
∂z ∂yi ∂z ∂xi
which satisfy the Lie bracket relations
(8.1.5) [Ri , Sj ] = δij ξ, [Ri , ξ] = [Si , ξ] = 0 .
Thus, h2n+1 is a non-trivial central extension of the 2n-dimensional Abelian Lie
algebra. Moreover, it is easy to check that h2n+1 ⊂ aut(ξ, η, Φ, g), however, there
are many more symmetries. The automorphism group of the flat Kähler structure
on R2n = Cn is the semidirect product of the unitary group U (n) with the 2n-
dimensional translation group, or in terms of Lie algebras we have a non-trivial
extension
0−−→t2n −−→aut(J,¯ gT )−−→u(n)−−→0 .
¯
We can lift elements of aut(J, gT ) to vector fields on H by using formula (6.1.3).
From all of this we get a commutative diagram of exact sequences
0 0
 
 
y y
=
{ξ} −−−−→ {ξ}
 
 
y y
(8.1.6) 0 −−−−→ h2n+1 −−−−→ aut(ξ, η, Φ, g) −−−−→ u(n) −−−−→ 0
  
  '
y y y
0 −−−−→ t2n −−−−→ ¯ gT )
aut(J, −−−−→ u(n) −−−−→ 0
 
 
y y
0 0
Thus, we obtain the Lie algebra of infinitesimal automorphisms aut(ξ, η, Φ, g) of
the Sasakian manifold (H2n+1 , ξ, η, Φ, g) as a non-trivial extension of u(n) by the
Heisenberg algebra h2n+1 . In dimension 3 (n = 1) this is one of Thurston’s eight
model geometries [Thu97].
The following is an example where none of the infinitesimal automorphisms of
(Z, ω, J) lift to infinitesimal automorphisms of (M, ξ, η, Φ, g).
Example 8.1.13: Continuing from Example 3.5.16, we let (V, ω) be an Abelian
variety. Consider a polarization defined by the nonzero class [ω] ∈ H 2 (V, Z). For
simplicity we consider the principal polarization defined by the standard Kähler
structure defined by
iX
ω= dzj ∧ dz¯j
2 j
with local coordinates zj = xj + iyj . The vector fields ∂xj and ∂yj are in aut(J, ω),
but none are Hamiltonian since ∂xj ω is not exact, with the same for ∂yj . Thus,
262 8. SYMMETRIES AND SASAKIAN STRUCTURES

the nontrivial circle bundle S 1 −−→M −−→T 2n whose first Chern class is [ω] admits a
Sasakian structure S = (ξ, η, Φ, g) by Theorem 7.5.2. In this case aut(ξ, η, Φ, g) is
the one dimensional Lie algebra generated by ξ. This is consistent with the well-
known fact [LB92] that the automorphism group of a polarized Abelian variety
is finite. Notice that here we have cup(M ) = 2n, and that c1 (Fξ ) = 0. In this
case the manifold M is the nil-manifold N2n+1 = H2n+1 (R)/H2n+1 (Z) which is a
homogeneous manifold for the Heisenberg group H2n+1 (R) and a Sasakian manifold,
but it is not homogeneous Sasakian.
This example is indicative of a more general result.
Theorem 8.1.14: Let M be a compact manifold with a K-contact structure S =
(ξ, η, Φ, g).
(i) If c1 (Fξ ) < 0, then aut(ξ, η, Φ, g) = {ξ}. Hence, any K-contact structure
with c1 (Fξ ) < 0 is quasi-regular.
(ii) If S is a quasi-regular K-contact structure with c1 (Fξ ) ≤ 0 and b1 (M ) =
0, then aut(ξ, η, Φ, g) = {ξ}. Hence, when b1 (M ) = 0 any K-contact
structure with c1 (Fξ ) = 0 is quasi-regular.
(iii) If S is a Sasakian structure with c1 (Fξ ) ≤ 0, then aut(ξ, η, Φ, g)
= {ξ}. Hence, any Sasakian structure with c1 (Fξ ) ≤ 0 is quasi-regular.
Proof. Let X ∈ aut(ξ, η, Φ, g) − {ξ}. We claim that X must vanish. Assume
not, then X and ξ span a 2-dimensional commutative subalgebra t2 of aut(ξ, η, Φ, g).
First assume that S is quasi-regular and that c1 (Fξ ) ≤ 0. Now by the Trans-
verse Yau Theorem 7.5.19, or equivalently the Sasakian version, Theorem 7.5.20,
we can assume that the transverse Ricci curvature RicgT ≤ 0. Then we have an
orbifold Riemannian submersion π : M −−→Z with almost Kähler orbifold struc-
ture (Z, h) whose Ricci curvature satisfies Rich ≤ 0. Furthermore, by Lemma
8.1.7 any infinitesimal automorphism X of S projects to an infinitesimal auto-
morphism of the almost Kähler structure on Z. So to prove (ii) we notice that
b1 (Z) = bB 1 (M ) = b1 (M ) = 0, so the orbifold (Z, h) has no parallel vector fields.
But 0 ≥ c1 (Z)orb implies Rich ≤ 0, so Bochner’s well-known theorem (cf. [Pet98]),
which holds equally well for compact orbifolds, implies that any Killing field on
(Z, h) is parallel. This implies that X = 0 and proves (ii).
For (iii) we first assume that S is quasi-regular and Sasakian. Again X projects
to a parallel vector field X̃. But now J X̃ is also parallel since h is Kähler. So
X̃ ω = J X̃ h is harmonic, and thus represents a non-trivial element of H 1 (Z, R).
So by Corollary 8.1.9 X̃ cannot be the projection of a vector field X ∈ aut(ξ, η, Φ, g).
This is a contradiction unless X = 0 proving (iii) under the added assumption that
S is quasi-regular. But since every quasi-regular Sasakian structure with c1 (Fξ ) ≤ 0
has a one dimensional automorphism group, it follows from Proposition 7.1.8 that
no irregular Sasakian structure can exist in this case. This proves (iii).
To prove (i) we first assume that S is quasi-regular, so as above M is an orbifold
Riemannian submersion over an almost Kähler orbifold (Z, h) with Rich < 0. Thus,
by Bochner there are no Killing vector fields on (Z, h). But by Lemma 8.1.7 every
X ∈ aut(ξ, η, Φ, g) projects to an automorphism of (Z, h) implying that X ∈ {ξ}.
This proves (i) for the quasi-regular case. But then as in the proof of (iii) no
irregular K-contact structure can exist. ¤
The maximal dimension of Isom(M 2n+1 , g) is (2n + 1)(n + 1). We want to
understand the difference between isom(M, g) and aut(ξ, η, Φ, g) or more globally
8.1. AUTOMORPHISMS OF SASAKIAN STRUCTURES AND ISOMETRIES 263

between Isom(M, g) and Aut(ξ, η, Φ, g) on a K-contact or Sasakian manifold. First


we notice that any nonzero Killing vector field in isom(M, g) − aut(ξ, η, Φ, g) cannot
commute with ξ. Explicitly,
Lemma 8.1.15: Let S = (ξ, η, Φ, g) be a K-contact structure, and let ρ be a Killing
vector field that commutes with ξ. Then ρ ∈ aut(ξ, η, Φ, g). Moreover, the following
hold:
ξ(η(ρ)) = 0, ρ(η(ρ)) = 0 .

Proof. For the first statement we need to show that £ρ η and £ρ Φ both
vanish. Since, £ρ g = 0, the vector field ρ is affine, that is, £ρ ∇ = 0, where ∇ is
the Levi-Civita connection. Thus,
−£ρ Φ = £ρ (∇ξ) = (£ρ ∇)(ξ) + ∇[ρ, ξ] = 0 .
To show that £ρ η vanishes, we have
£ρ η = £ρ (ξ g) = [ρ, ξ] g+ξ £ρ g = 0 .
To prove the two equations we notice that
ξ(η(ρ)) = (£ξ η)(ρ) + η([ξ, ρ]) = 0 ,
which proves the first equation, while for the second we have
ρ(η(ρ)) = (£ρ η)(ρ) = 0 . ¤

To understand better the difference between isometries and automorphisms, we


develop two more lemmas the first of which is due to Tachibana and Yu [TY70],
and the second essentially due to Tanno [Tan70].
Lemma 8.1.16: Let (ξ, η, Φ, g) and (ξ 0 , η 0 , Φ0 , g) be two Sasakian structures sharing
the same Riemannian metric g. Then either g(ξ, ξ 0 ) is constant or (M, g) is a space
form with sectional curvature equal to 1.

Proof. Put f = g(ξ, ξ 0 ) and consider the vector field ρ = grad f. It is easiest
∂f
to work in local coordinates (U ; xi ). Let ρi = dxi
. Then using Proposition 7.3.4 and
Theorem 7.3.16 one obtains the differential equation
∇i ∇k ρj + 2ρi gkj + ρk gij + ρj gik = 0.
Then a Theorem of Obata [Oba66] says that the only nonzero solution for ρi occurs
when g is a metric of constant sectional curvature equal to 1. ¤

Our second lemma addresses the case when f = g(ξ, ξ 0 ) is constant.


Lemma 8.1.17: Let S = (ξ, η, Φ, g) and S 0 = (ξ 0 , η 0 , Φ0 , g) be two Sasakian struc-
tures sharing the same Riemannian metric g, and suppose that (M, g) is not a space
form with sectional curvature equal to 1. Then either
(i) S 0 = S,
(ii) S 0 = S c the conjugate Sasakian structure, or
(iii) S and S 0 are part of a 3-Sasakian structure (cf. Proposition 13.2.1 below).

Proof. Since (M, g) is not a space form with sectional curvature equal to 1,
Lemma 8.1.16 says that g(ξ, ξ 0 ) is a constant, say a. By the Schwarz inequality
264 8. SYMMETRIES AND SASAKIAN STRUCTURES

|a| ≤ 1, and a = 1 gives case (i) while a = −1 gives case (ii). Thus, we can assume
that |a| < 1. Following Tanno [Tan70] we define
00 ξ 0 − aξ
ξ =√ ,
1 − a2
and it is easy to see that
00 00 00
g(ξ, ξ ) = 0, g(ξ , ξ ) = 1 .
00 00 00 00
Now we define η = g(ξ , ·), and Φ = −∇ξ . So by linearity Proposition 7.3.17
00 00 00 00
implies that S = (ξ , η , Φ , g) s a Sasakian structure orthogonal to S. For con-
00
venience we set ξ 1 = ξ, ξ 2 = ξ , etc, and define ξ 3 = Φ1 ξ 2 . Then using
0 = X(g(ξ 1 , ξ 2 )) = g(Φ1 X, ξ 2 ) + g(Φ2 X, ξ 1 )
together with Definition 6.2.5 we see that Φ1 ξ 2 = −Φ2 ξ 1 . So from the torsion
freeness of ∇ one gets [ξ 1 , ξ 2 ] = 2Φ1 ξ 2 = 2ξ 3 . Similar relations are obtained by
cyclic permutation of the superscripts and Proposition 13.2.1 implies the result. ¤

We now have
Theorem 8.1.18: Let (M, ξ, η, Φ, g) be a complete Sasakian manifold that is not
isometrically covered by the round sphere, nor part of a 3-Sasakian structure. Then
either Aut(ξ, η, Φ, g) = Isom(M, g) or Isom(M, g) is a Z2 extension of Aut(ξ, η, Φ, g).
In particular, if (M, ξ, η, Φ, g) is a complete non-compact Sasakian manifold, then
Aut(ξ, η, Φ, g) = Isom(M, g) or Isom(M, g) is a Z2 extension of Aut(ξ, η, Φ, g).

Proof. Assuming the hypothesis of Theorem 8.1.18, let ψ ∈ Isom(M, g), and
consider the Sasakian structure S ∗ = (ψ∗ ξ, ψ∗−1 η, ψ∗ Φψ∗−1 , g). By Lemma 8.1.17
the only possibilities are S ∗ = S, or S ∗ = S c . In the former case Aut(ξ, η, Φ, g) =
Isom(M, g), whereas, the latter gives Isom(M, g) as a Z2 extension of Aut(ξ, η, Φ, g).
The last statement follows since both spheres and complete 3-Sasakian manifolds
are compact, cf. Corollary 13.2.3 below. ¤

Corollary 8.1.19: Under the hypothesis of Theorem 8.1.18, we have


(i) isom(M, g) = aut(ξ, η, Φ, g).
(ii) 1 ≤ dim Isom(M, g) ≤ (n + 1)2 .
Since every complete Sasakian structure has at least a one dimensional group of
automorphisms, in the case of quasi-regular structures we can always form nontrivial
smooth quotients. Evidently we have
Proposition 8.1.20: Let (M, ξ, η, Φ, g) be a compact quasi-regular Sasakian man-
ifold, and let G be a finite subgroup of Aut(ξ, η, Φ, g). Then the quotient space M/G
is an orbifold which inherits the Sasakian structure S = (ξ, η, Φ, g) from M. More-
over, if G acts freely on M, then M/G is a Sasakian manifold. In particular, every
compact quasi-regular Sasakian manifold admits quotients by cyclic subgroups Cm of
the circle group generated by ξ. Moreover, if gcd(m, υ(S)) = 1, the quotient M/Cm
is a Sasakian manifold.
One can use this proposition to construct Sasakian structures on lens spaces
from the Sasakian structure on the weighted sphere.
Example 8.1.21: [Lens spaces] Let Sw denote the weighted Sasakian structure
of weight vector w ∈ (R+ )n+1 and S0 the standard Sasakian structure on S 2n+1 .
8.2. DEFORMATION CLASSES OF SASAKIAN STRUCTURES 265

Also let Cp denote the cyclic subgroup of the maximal torus Tn+1 generated by the
action
(z0 , . . . , zn ) 7→ (ζz0 , ζ q1 z1 , . . . , ζ qn zn ) ,
where ζ is a primitive pth root of unity, and is relatively prime to the qi ’s. The
quotient of S 2n+1 by this action is a smooth manifold known as the lens space
L(p; q1 , . . . , qn ). Since Tn+1 ⊂ Aut(Sw ), the Sasakian structures Sw as well as S0
all pass to the quotient manifold L(p; q1 , . . . , qn ). Notice that if we choose w =
(1, q1 , . . . , qn ) the cyclic subgroup Cp ⊂ Tn+1 is actually a subgroup of the circle
1
group Sw generated by the Reeb vector field ξw . Thus, in this case the Sasakian
structure Sw on L(p; q1 , . . . , qn ) is the one obtained from Proposition 8.1.20. As
in Example 7.1.12 all the Sasakian structures Sw belong to the same underlying
contact structure on L(p; q1 , . . . , qn ).
Much is known about the topology of lens spaces especially in dimension 3
(i.e., n = 1). For example it is well-known (cf. [Bre93]) that L(p, q) is homotopy
equivalent to L(p, q 0 ) if and only if either qq 0 or −qq 0 is a quadratic residue mod p.
There is a theorem of Brody (cf. [Bre93]) that is perhaps less well-known. It says
that L(p, q) and L(p, q 0 ) are homeomorphic if and only if either q 0 ≡ ±q mod p or
qq 0 ≡ ±1 mod p.
Notice that one can define quotient manifolds that are “locally Sasakian” as
follows. Let M be a Sasakian manifold and σ ∈ Isom(M, g)\Aut(ξ, η, Φ, g). Then
by Theorem 8.1.18 σ is an involution. Suppose further that this involution acts
freely on M. Then M/{σ} is not Sasakian, but it has a “locally Sasakian” structure
in the sense that its 2-fold cover is Sasakian and the Riemannian metric is just the
Sasakian metric of M pushed to the quotient. However, generally such manifolds
may not even be orientable. We shall give explicit examples of this phenomenon in
Chapters 9 and 14. See Examples 9.3.23 and 14.2.5 below.

8.2. Deformation Classes of Sasakian Structures


8.2.1. Deformations of Transverse Holomorphic Structures. Here we
briefly sketch the deformation theory of transverse holomorphic structures due to
Duchamp and Kalka [DK79, DK80, DK84], Girbau, Haefliger, and Sundarara-
man [GHS83], and Gomez-Mont [GM80]. In [DK79, DK80] an extra assump-
tion was made that is not convenient for our purposes. So we shall mainly follow
[GHS83], but adapted to our purpose. For the basics of deformation theory we re-
fer to [Kod86] and [Kur71]. Here we want to think of M as a Sasakian manifold of
dimension 2n + 1, but we shall be more general in the beginning. We are interested
for now only in the transverse holomorphic structure. So in this subsection F will
denote a transverse holomorphic foliation on M. Let (S, 0) denote a germ at 0 of
an analytic space, that is, a space S with a local structure ring O0 (S) ≈ O0 (Ck )/J,
where O0 (Ck ) denotes the algebra of germs of holomorphic on Ck at the origin, and
J is some ideal. A smooth function g on (S, 0) is an equivalence class of smooth
functions on a neighborhood of 0 ∈ Ck , where the two functions are equivalent if
their difference lies in the ideal J.
A germ of a deformation of a transversely holomorphic foliation F on M pa-
rameterized by (S, 0) is given by an open cover {Uα }α∈I of M and a family of local
submersions fαs : Uα −−→Cn parameterized by (S, 0) that are holomorphic in s ∈ S
266 8. SYMMETRIES AND SASAKIAN STRUCTURES

for each x ∈ Uα , and satisfy the condition


(8.2.1) fαs = ταβ
s
◦ fβs on Uα ∩ Uβ ,
s
where ταβ : fβs (Uα ∩Uβ )−−→fαs (Uα ∩Uβ ) is a holomorphic family of biholomorphisms.
s
Thus, for each s ∈ S the map ταβ belongs to the pseudogroup ΓAutut(I) consisting
of local automorphisms of the standard complex structure I on Cn . Here we denote
the germ of the transverse holomorphic foliation by F s , and the original foliation
F = F 0 is defined by the local submersions fα0 . This induces a foliation F S on
(S, 0) × M given by the local submersions (s, x) 7→ (s, fαs (x)) of (S, 0) × Uα to
(S, 0)×Cn . The transverse structure of this foliation is locally isomorphic to (S, 0)×
Cn . Consider two deformations of F parameterized by (S, 0). These can be given
by two families of local submersions {fαs : Uα −−→Cn }α and {fγ0s : Uγ0 −−→Cn }γ .
Two such deformations are isomorphic if there is a smooth family hs : M −−→M
of diffeomorphisms parameterized by (S, 0) such that on Uα ∩ Uγ0 there exists a
s
holomorphic family ταγ of biholomorphisms such that
fαs = ταγ
s
◦ fγ0s ◦ hs .
Infinitesimal deformations are obtained by differentiating equation (8.2.1) with
respect to s and evaluating at s = 0. This gives rise to the Kodaira-Spencer map
which we now describe. Let Θ̂F denote the sheaf of germs of vector fields on M
that are infinitesimal automorphisms of the transversely holomorphic foliation F,
and let ΘF be the quotient sheaf by the sheaf of smooth vector fields tangent to
the leaves of F. Let Cn have its standard complex structure I and let ΘCn denote
the sheaf of germs of vector fields preserving I. Then ΘF |Uα ≈ fα∗ ΘCn , and upon
restricting further to Uα ∩ Uβ we have ΘF |Uα ∩Uβ ≈ fα∗ ΘCn |Uα ∩Uβ . Now writing
s (β) (β)
the Haefliger cocycle ταβ in the transverse complex coordinates (z1 , . . . , zn ), we
define a section of the sheaf ΘF |Uα ∩Uβ by
n ³
X s i ´
∂(ταβ ) (β) ∂
θαβ = (z1 , . . . , zn(β) ) .
∂s s=0 ∂z (α)
i=1 i

From Haefliger’s cocycle condition (2.2.2) we see that θαβ satisfies the cocycle con-
dition
(8.2.2) θαβ + θβγ + θγα = 0
in the triple overlap Uα ∩ Uβ ∩ Uγ 6= ∅, and thus defines an element [θαβ ] ∈
H 1 (M, ΘF ). The Kodaira-Spencer map is the linear map ρ : T0 S−−→H 1 (M, ΘF )
d
that assigns to the tangent vector ( ds )s=0 ∈ T0 S the cohomology class [θαβ ] de-
fined by the vector field θαβ . More generally, one can consider the full cohomology
ring H ∗ (M, ΘF ) which has the structure of graded Lie algebra induced by the
Lie bracket on ΘF . The basic fact about these cohomology groups is that they
are finite dimensional for compact manifolds. This follows from the ellipticity of
the transverse structure and was proven independently in [DK79] and [GM80].
Summarizing we have
Proposition 8.2.1: Let M be a compact manifold with a transverse holomorphic
foliation F. Then the cohomology groups H r (M, ΘF ) are finite dimensional.
The major result of any deformation theory is that the deformation space (S, 0)
be versal, that is in the case at hand, that any germ of a transverse holomorphic
8.2. DEFORMATION CLASSES OF SASAKIAN STRUCTURES 267

deformation that is ‘close’ to F is isomorphic to a germ of a transverse holomorphic


foliation F s for some s ∈ (S, 0). Here is the result [GHS83]:
Theorem 8.2.2: Let F be a transverse holomorphic foliation on a compact man-
ifold M, and let ΘF denote the sheaf of germs of transversely holomorphic vector
fields defined above. Then
(i) There is a germ (S, 0) of an analytic space parameterizing a germ of
0
a deformation F s of F such that if F s is any germ of a deformation
0
parameterizing F by the germ (S , 0), there is a holomorphic map φ :
0
(S 0 , 0)−−→(S, 0) so that the deformation F φ(s) is isomorphic to F s .
(ii) The Kodaira-Spencer map ρ : T0 S−−→H 1 (M, ΘF ) is an isomorphism.
(iii) There is an open neighborhood U ⊂ H 1 (M, ΘF ) and a holomorphic map
Ψ : U −−→H 2 (M, ΘF ) such that (S, 0) is the germ at 0 of Ψ−1 (0). The
2-jet of Ψ satisfies j 2 Ψ(u) = 12 [u, u].
The element [u, u] ∈ H 2 (M, ΘF ) is the primary obstruction to performing the
deformation, and (i) of Theorem 8.2.2 says that the analytic space (S, 0) is versal.
It is known as the Kuranishi space of the deformation. Theorem 8.2.2 has the
following important corollaries.
Corollary 8.2.3: If H 1 (M, ΘF ) = 0, then any infinitesimal deformation of F is
trivial, that is, every F s close enough to F is isomorphic to F.
Corollary 8.2.4: If H 2 (M, ΘF ) = 0, then the Kuranishi space of deformations of
F is isomorphic to an open neighborhood of 0 in H 1 (M, ΘF ).
Notice that (ii) and (iii) of the Theorem 8.2.2 imply that if H 2 (M, ΘF ) does
not vanish then the Kuranishi space S is singular.

8.2.2. Deformations of Transverse Holomorphic Sasakian Structures.


We now apply the ideas of the previous subsection to the case where M is a compact
Sasakian manifold. First, we have the evident
Proposition 8.2.5: The characteristic foliation Fξ of a Sasakian structure S is a
transverse holomorphic foliation.
Since by Theorem 7.1.10 any Sasakian structure on M can be approximated by
quasi-regular Sasakian structures, we restrict our discussion of deformation theory
to the latter. Here we denote by ΘFξ the sheaf of transversely holomorphic vector
fields on M, and by ΘZ the sheaf of germs of holomorphic vector fields on the
complex orbifold Z.
Proposition 8.2.6: Let S be a quasi-regular Sasakian structure on a compact man-
ifold M with characteristic foliation Fξ , and let π : M −−→Z denote the quotient
projection. Then there is an exact sequence

0−−→H 1 (Z, ΘZ )−−→H 1 (M, ΘFξ )−−→H 0 (Z, ΘZ )−−→H 2 (Z, ΘZ ) .

Proof. The Leray sheaf H∗ (π, ΘFξ ) of the map π is the sheaf associated to
the presheaf U 7→ H ∗ (π −1 (U ), ΘFξ |π−1 (U ) ). By Leray’s Theorem there is a spectral
sequence for which

E2p,q = H p (Z, Hq (π, ΘFξ )) =⇒ H p+q (M, ΘFξ ) .


268 8. SYMMETRIES AND SASAKIAN STRUCTURES

Furthermore, the stalk Hq (π, ΘFξ )y at a point y ∈ Z is isomorphic to


(
(ΘZ )y , if q = 0, 1;
H q (π −1 (y), ΘFξ |π−1 (y) ) ≈ H q (S 1 , (ΘZ )y ) ≈
0, otherwise.

Thus we have
(
H p (Z, ΘZ ), if q = 0, 1;
E2p,q =
0, otherwise.
So the spectral sequence collapses at the third term and the exact sequence follows
as for the Gysin sequence. ¤

There are two immediate corollaries to this proposition.


Corollary 8.2.7: If H i (Z, ΘZ ) = 0 for i = 1, 2, then H 1 (M, ΘFξ ) ≈ H 0 (Z, ΘZ ).
Corollary 8.2.8: If H 0 (Z, ΘZ ) = 0, then H 1 (M, ΘFξ ) ≈ H 1 (Z, ΘZ ).
These are the two extreme cases for which one can think of a deformation
of a transverse holomorphic structure as arising form either a deformation of the
complex structure on Z or a deformation of the vector field ξ. The first corollary
applies when there are no nontrivial deformations of the complex structure on Z,
in which case the nontrivial deformations of the transverse holomorphic structure
comes from deformations of the vector field ξ which arise from the infinitesimal
complex automorphisms of Z. An example of this type is M = S 2n+1 with its
standard Sasakian structure. In this case the base space Z = CPn which has no
deformation space. So H 1 (S 2n+1 , ΘFξ ) ≈ H 0 (CPn , ΘZ ). Deformations of this type
include the weighted spheres of Example 7.1.12 which correspond to elements lying
in the maximal torus of H 0 (CPn , ΘZ ).
On the other hand, the second corollary says that if there are no infinitesimal
complex automorphisms, then the all deformations of the transverse holomorphic
structure comes from the deformations of the complex structure on Z. This is the
case in Example 8.1.13 for the nil-manifold N2n+1 , where the base space Z is a
polarized smooth Abelian variety An of complex dimension n. In this case we have
H 1 (N2n+1 , ΘFξ ) ≈ H 1 (An , ΘZ ).
In order to treat deformations through Sasakian structures more constraints
are needed. This should be contrasted with the Kählerian case where the well-
known Kodaira-Spencer stability theorem [KS60] implies that locally infinitesimal
complex deformations of Kählerian structures remain Kählerian. The point is that
transversely holomorphic deformations which deform a Riemannian foliation need
not remain Riemannian. Thus, we wish to deform through Riemannian, or more
s
precisely Kählerian foliations in which case the Haefliger cocyles ταβ are taken from
the subpseudogroup ΓAut(I,g) ⊂ ΓAut(I) of isometries of some locally defined Kähler
metric g on Cn . Thus, infinitesimally the corresponding vector fields on M are also
local Killing vector fields for the transverse metric gT , where gT = fα∗ g for all α.
We denote the subsheaf of ΘFξ of such vector fields by ΘFξ ,a . Now let Fξ be the
characteristic foliation of a quasi-regular Sasakian structure S = (ξ, η, Φ, g) on M.
Then by Theorem 7.5.1 there is an orbifold submersion π : M −−→Z onto a compact
Kähler orbifold Z whose Kähler metric h satisfies gT = π ∗ h. So we let ΘZ,a denote
the subsheaf of ΘZ that leaves invariant the Kähler metric h on Z. Then exactly
as in Proposition 8.2.6 we have
8.2. DEFORMATION CLASSES OF SASAKIAN STRUCTURES 269

Proposition 8.2.9: Let S be a quasi-regular Sasakian structure on a compact man-


ifold M with characteristic foliation Fξ , and let π : M −−→Z denote the quotient
projection. Then there is an exact sequence:
0−−→H 1 (Z, ΘZ,a )−−→H 1 (M, ΘFξ ,a )−−→H 0 (Z, ΘZ,a )−−→H 2 (Z, ΘZ,a ).
We mention that in the case of certain deformations through Hermitian folia-
tions, Duchamp and Kalka [DK80]have given a splitting theorem.

8.2.3. Deformations of the Characteristic Foliation. We consider de-


formations of the characteristic foliation Fξ with a fixed transverse holomorphic
structure, that is infinitesimal deformations such that the Kodaira-Spencer map
ρ lies in H 1 (M, ΘFξ ,a )/H 0 (Z, ΘZ,a ). Rather than studying the Kodaira-Spencer
map, we proceed directly.
This type of deformation was first considered by Takahashi [Tak78] who con-
sidered deformations by adding an infinitesimal automorphism ρ ∈ aut(ξ, η, Φ, g).
However, here we consider somewhat more general deformations first introduced
by Gauduchon and Ornea [GO98] in the 3-dimensional case, and then used by
Belgun in his classification of 3-dimensional Sasakian structures [Bel00, Bel01].
Explicitly, we consider deformations which preserve the underlying contact struc-
ture as well as the underlying almost CR structure, but deform the contact 1-form
by sending η 7→ f η for some positive function f. We then write
(8.2.3) η̃ = f η, ξ˜ = ξ + ρ ,
where ξ˜ is the Reeb vector field of η̃. The conditions
˜ = 1,
η̃(ξ) ξ˜ dη̃ = 0
then imply the relation
1 1
(8.2.4) f= = .
˜
η(ξ) 1 + η(ρ)
It is clear that η̃ is a contact 1-form as long as η(ξ) ˜ > 0, and that the contact
subbundle D = ker η = ker η̃ remains unchanged. The deformations also preserve
the CR structure, that is, Φ̃|D = Φ|D , which extends to an endomorphism by setting
Φ̃ = Φ − Φξ˜ ⊗ η̃. Moreover, we can easily check that
Φ̃ξ˜ = 0 Φ̃2 = −1l + ξ˜ ⊗ η̃
˜ η̃, Φ̃) is an underlying almost contact structure of the contact struc-
hold. Thus, (ξ,
ture. Next we define a Riemannian metric g̃ by
g̃ = dη̃ ◦ (Φ̃ ⊗ 1l) ⊕ η̃ ⊗ η̃ .
This metric is compatible with the contact structure, so (ξ, ˜ η̃, Φ̃, g̃) is a contact
metric structure. But according to Corollary 6.5.11 the structure (ξ, ˜ η̃, Φ̃, g̃) is
normal if and only if the almost CR structure Φ̃|D is integrable and Φ̃ is invariant
˜ The former condition holds since the CR structure hasn’t changed. So if
under ξ.
we begin with a Sasakian structure a deformation through K-contact structures is
automatically a deformation through Sasakian structures.
Definition 8.2.10: Let (D, J) be a strictly pseudoconvex almost CR structure of
Sasaki type on a manifold M. We denote by F(D, J) the set of all K-contact struc-
tures having (D, J) as its underlying CR structure. A deformation within F(D, J)
270 8. SYMMETRIES AND SASAKIAN STRUCTURES

given by a transformation of the form of equation (8.2.3) is called a deformation


of type I.
We now investigate the properties of the space F(D, J). We shall assume that
(D, J) is of Sasaki type so that F(D, J) is nonempty. We have immediately from
Definition 6.4.9 and Proposition 6.4.8 that
Proposition 8.2.11: A contact metric structure S = (ξ, η, Φ, g) is in F(D, J) if
and only if ξ ∈ cr(D, J).
Next we identify F(D, J) with a convex subset of cr(D, J). Let us fix a K-contact
structure S0 = (ξ0 , η0 , Φ0 , g0 ) ∈ F(D, J), and let cr+ (D, J) be the subset of cr(D, J)
such that η0 (ξ) > 0. There is a natural map
(8.2.5) ι : F(D, J)−−−−→cr+ (D, J)
defined by ι(S) = ξ, and one easily sees that this map is injective. We claim that
it is also surjective. Let ξ ∈ cr+ (D, J) and define η by
η0
η= .
η0 (ξ)
Clearly η(ξ) = 1, and since ξ ∈ cr+ (D, J), it leaves D invariant. This implies that
ξ dη = £ξ η = 0. So ξ is the Reeb vector field of η. We then define Φ = Φ0 −Φ0 ξ ⊗η,
and g = dη ◦ (Φ ⊗ 1l) + η ⊗ η. So S = (ξ, η, Φ, g) ∈ F(D, J) by Proposition 8.2.11.
Hereafter, we shall identify F(D, J) with cr+ (D, J) without further ado.
Proposition 8.2.12: The subset cr+ (D, J) is an open convex cone in cr(D, J) that
is invariant under the adjoint action of CR(D, J).

Proof. That cr+ (D, J) is an open subset of cr(D, J) follows immediately from
its definition. Convexity is clear since if ξ1 , ξ2 ∈ cr+ (D, J), then η0 ((1−t)ξ1 +tξ2 ) =
(1 − t)η0 (ξ1 ) + tη0 (ξ2 ) > 0 for all 0 ≤ t ≤ 1. The conic property is also clear. For
groups of transformations the adjoint action is that induced by the differential. So
for φ ∈ CR(D, J), we have η0 (φ∗ ξ) = (φ∗ η0 )(ξ) = fφ η0 (ξ) > 0 for some positive
function fφ , implying that cr+ (D, J) is invariant. ¤

The tangent space TS (cr+ (D, J)) at a point S ∈ cr+ (D, J) is identified with the
Lie algebra cr(D, J). Let us now assume that M is compact and that the almost CR
structure is integrable, so by Theorem 8.1.6 the Lie algebra cr(D, J) is reductive.
We have [BGS06]
Theorem 8.2.13: Let M be a compact manifold of dimension 2n + 1 with a
CR structure (D, J) of Sasaki type. Then the Lie algebra cr(D, J) decomposes as
cr(D, J) = tk + p, where tk is the Lie algebra of a maximal torus Tk of dimension k
with 1 ≤ k ≤ n + 1 and p is a completely reducible Tk module. Furthermore, every
X ∈ cr+ (D, J) is conjugate to a positive element in the Lie algebra tk .

Proof. First assume that (D, J) is not the standard CR structure on the
sphere. Then according to Proposition 8.1.6 there is a Sasakian structure S0 ∈
S(D, J) such that CR(D, J) = Aut(S0 ) which is a compact Lie group. Then by a
well-known result of Lie theory every element in the Lie algebra aut(S0 ) is conjugate
under the adjoint action of the group Aut(S0 ) to one on tk . Moreover, by Proposition
8.2.12 positivity is preserved under this action. The range for the dimension of the
maximal torus follows by a well-known result of Sasakian geometry.
8.2. DEFORMATION CLASSES OF SASAKIAN STRUCTURES 271

In the case that (D, J) is the standard CR structure on the sphere, we know
[Web77] that CR(D, J) = SU (n + 1, 1), in which case there are several maximal
Abelian subalgebras of cr(D, J) = su(n + 1, 1). However, it can be seen from a case
by case analysis that the only Abelian subgroup where the positivity condition can
be satisfied is in the case of a maximal torus. This can be ascertained, for example,
by looking at Theorem 6 of [Dav05]. ¤

We now define t+ + + +
k = tk (D, J) = tk ∩ cr (D, J). As with cr (D, J) this is a
convex open cone in the Lie algebra tk , and it can be viewed as the moduli space of
Sasakian structures that are compatible with the CR structure (D, J). The space
t+
k (D, J) was introduced and studied in [BGS06]. It is called the Sasaki cone.
Explicitly,
Corollary 8.2.14: The following identifications hold:
t+ + +
k = tk (D, J) = cr (D, J)/CR(D, J) = F(D, J)/CR(D, J) .

Let us consider the case k = 1 which is somewhat special. Since the Reeb
vector field is central k = 1 implies that dim aut(S) = dim cr(D, J) = 1. In this
case F(D, J) = cr+ (D, J) = t+ +
1 = R and F(D, J) consists of a 1-parameter family
of Sasakian structures given by Sa = (ξa , ηa , Φa , ga ), where a > 0 and
ξa = a−1 ξ, ηa = aη , Φa = Φ , ga = ag + (a2 − a)η ⊗ η .
So the deformation is a transverse homothety. By Theorem 8.1.14 this case arises
when c1 (Fξ ) ≤ 0, that is:
Proposition 8.2.15: Let S = (ξ, η, Φ, g) be a Sasakian structure with c1 (Fξ ) ≤ 0.
Then the Sasaki cone has dimension one.
From another point of view one can formulate the condition of deforming
through Sasakian structure as a condition on the function f. This was done in
[GO98] in terms of the Hessian. We define the Hessian Hf (X, Y ) of a function f
by Hf (X, Y ) = (∇X df )(Y ).
Proposition 8.2.16: The deformed contact metric structure (ξ,˜ η̃, Φ̃, g̃) is Sasakian
if and only if
Hf −1 (ΦX, ΦY ) = Hf −1 (X, Y )
for all sections X, Y of D.
The proof of Proposition 8.2.16 is left as a series of exercises the first of which
is to prove the following
Lemma 8.2.17: Let (ξ, η, Φ, g) be a contact metric structure on M, and consider
a deformation of the contact 1-form defined by equations (8.2.3) and (8.2.4). Then
the vector field ρ takes the form
³1 − f ´ 1
ρ= ξ − Φgrad f −1 .
f 2
Exercise 8.3: Prove Lemma 8.2.17.
Exercise 8.4: Prove that
1 1
g(∇X Q, Y ) = − Hf −1 (X, ΦY ) + df −1 (ξ)g(X, Y ) ,
2 2
where 2Q = −Φgrad f −1 , and use this to prove Proposition 8.2.16.
272 8. SYMMETRIES AND SASAKIAN STRUCTURES

Exercise 8.5: Show that if ρ is a Killing vector field such that ρ = hξ for some
function h, then h must be constant.
An example of this type of deformation where ρ is a Killing vector field is the
deformation away from the standard sphere to the weighted sphere as described
in Example 7.1.12. These are obtained by taking ρ to lie on a maximal torus
in the automorphism group of the standard Sasakian round sphere. More gener-
ally, such deformations have an important application, namely in the proof of the
Sasakian analogue of Kodaira’s conjecture mentioned earlier. The Sasakian case,
however, manifests itself somewhat differently than the complex case in that type
I deformations do not change the underlying CR structure. This is a special case
of Rukimbira’s Theorem 7.1.10, namely
Theorem 8.2.18: On a compact manifold every Sasakian structure S = (ξ, η, Φ, g)
is a deformation of a quasi-regular (algebraic) Sasakian structure.

8.3. Homogeneous Sasakian Manifolds


Recall from Definition 1.6.11 that a homogeneous manifold is a smooth man-
ifold M on which there is a smooth transitive action of a Lie group G. Given a
homogeneous manifold M one is often interested in homogeneous structures, that
is, geometric structures that are invariant under the action of G. For example, a
homogeneous Riemannian structure is a homogeneous manifold M together with a
G-invariant Riemannian metric. This can be phrased by saying that the Lie group
Isom(M, g) of isometries of (M, g) acts transitively on M. However, the group of
automorphisms of a G-structure is not generally a Lie group. When it is, the G-
structure is said to be of finite type. Contact and symplectic structures are not of
finite type. So we have
Definition 8.3.1: A contact structure D on a manifold M is a homogeneous
contact structure if there is a Lie subgroup G ⊂ Con(M, D) that acts transitively
on M. Similarly, a strict contact structure defined by a global 1-form η is homoge-
neous is if there is a Lie subgroup G ⊂ Con(M, η) that acts transitively on M. In
this case we refer to the triple (M, η, G) as a homogeneous contact manifold.
Since this book mainly describes certain contact metric structures whose group
of automorphisms is a finite dimensional Lie group by we focus our attention on
finite dimensional Lie subgroups of Con(M, η). However, before embarking on the
description of the automorphism group of K-contact structures, we discuss briefly
a more general result of Lichnerowicz [Lic63b] that we make of use of later. Recall
that a Lie algebra g is reductive if its maximal solvable ideal (radical) coincides with
its center, alternatively, the derived Lie algebra [g, g] is semi-simple. Then in the
compact case, we have the following important theorem of Lichnerowicz [Lic63b]:
Theorem 8.3.2: Let (M, η) be a compact strict contact manifold, and let g be a
finite dimensional subalgebra of con(M, η). Then g is reductive.
This theorem will be important in Section 8.3 where we prove a general state-
ment about compact homogeneous contact manifolds due to Boothby and Wang
[BW58]. In the case of a K-contact structure much of the structure of its auto-
morphism group Aut(ξ, η, Φ, g) can be deduced directly from the well-known Myers-
Steenrod Theorem.
8.3. HOMOGENEOUS SASAKIAN MANIFOLDS 273

Corollary 8.3.3: Let M be a smooth manifold with a contact metric structure


S = (ξ, η, Φ, g). Then its automorphism group Aut(ξ, η, Φ, g) is a closed finite di-
mensional Lie subgroup of the isometry group Isom(M, g). Furthermore, if M is
compact the Lie algebra of infinitesimal automorphisms aut(ξ, η, Φ, g) is reductive
with non-trivial center. In this case the corresponding Lie group Aut(ξ, η, Φ, g) is
up to covering isomorphic to G × T, where G is a compact semi-simple Lie group
and T is a torus.
Without loss of generality we assume that the Lie group G is connected. A
particular case of interest to us concerns the addition of a compatible Riemannian
metric in which case we do obtain a G-structure of finite type. Hence, we make the
definition below. However, as we shall show the fundamental results of Boothby and
Wang [BW58, Boo80] go a long way toward showing that they are all equivalent.
Definition 8.3.4: A metric contact structure (η, g) on a manifold M is said to be
a homogeneous metric contact structure if the group Aut(M, η, g) acts transi-
tively on M. Similarly, a K-contact structure or Sasakian structure S = (ξ, η, Φ, g)
on a manifold M is said to be a homogeneous K-contact or Sasakian struc-
ture if the group Aut(S) acts transitively on M.
We have the following important theorem of Boothby and Wang [BW58].
Theorem 8.3.5: Let (M, η, G) be a homogeneous contact manifold. Then the con-
tact form η is regular. Moreover, M = G/K is a fiber bundle over G/H0 K with
fiber H0 K/K, where H0 is the connected component of a one dimensional Lie group
H, and H0 K/K is either diffeomorphic to S 1 or to R.

Proof. Standard Lie theory together with the foundational work of Palais
[Pal57] applied to the flow of the Reeb vector field proves regularity. Since the
fibres are one dimensional, the result follows. ¤

We mention that in [Ale90] homogeneous contact manifolds are related to co-


adjoint orbits. Let us recall the following well-known terminology. Let Gc be a
complex semi-simple Lie group. A maximal solvable complex subgroup B is called
a Borel subgroup, and B is unique up to conjugacy. Any complex subgroup P that
contains B is called a parabolic subgroup. Then the homogeneous space Gc /P is
called a generalized flag manifold. A well-known result of Wang [Wan54] (see also
[Akh90]) says that every simply-connected compact homogeneous Kähler manifold
is a generalized flag manifold Gc /P . Thus, we can easily construct homogeneous
Sasakian manifolds by considering circle bundles over generalized flag manifolds.
An additional result of Borel and Remmert [BR62] says that a compact homoge-
neous Kähler manifold X is the product of its Albaneese A(X) and a generalized
flag manifold; however, the next theorem shows that the Albaneese factor cannot
contribute to compact homogeneous contact manifolds. We now have the main
result of homogeneous contact manifolds.
Theorem 8.3.6: Let (M, η, G) be a compact homogeneous contact manifold. Then
(i) M admits a homogeneous Sasakian structure with contact form η.
(ii) M is a non-trivial circle bundle over a generalized flag manifold.
(iii) M has finite fundamental group, and the universal cover M f of M is
compact with a homogeneous Sasakian structure.
274 8. SYMMETRIES AND SASAKIAN STRUCTURES

Proof. We sketch the proof. Let G be a Lie group acting transitively on


M preserving the contact form η. Lichnerowicz’s Theorem 8.3.2 says that the Lie
algebra g of G is reductive. Furthermore, since the action is transitive a corollary
[Lic63b] of Theorem 8.3.2 says that g is either semi-simple or has a one dimensional
center. But we know that the group of contactomorphisms of M has the one
dimensional subgroup generated by the Reeb vector field in its center. So we can
assume that G is simply connected with a one dimensional center Z. We write M =
G/K for some closed subgroup K of G. Then the Boothby-Wang Theorem 8.3.5
implies that M = G/K is a circle bundle over the compact homogeneous symplectic
manifold B = G/H. The group G does not act effectively on B since H contains the
center Z. But B = G/H = (G/Z)/(H/Z), so writing G = G1 Z we see that G1 acts
transitively on B with isotropy subgroup H1 = H ∩G1 . Now G1 is simply connected
and semi-simple, and acts effectively up to a discrete normal subgroup. Thus, a
theorem of Borel [Bor60] implies that G1 is compact and B is simply connected.
So G1 is compact semi-simple, then according to Borel [Bor54] and Lichnerowicz
[Lic53] H is a centralizer of a torus, and G/H is homogeneous Kähler2. But the
Kähler form ω defines a positive integral cohomology class [BW58], so B = G/H
is projective algebraic by the Kodaira embedding theorem. So lifting the Kähler
structure to M gives a homogeneous Sasakian structure with the same contact form
η. This proves (i).
By a result of Wang [Wan54] B = G/H is a generalized flag manifold, that is
it can be written in the form B = Gc /P, where Gc is a complex semi-simple Lie
group and P a parabolic subgroup. (In fact, G is a maximal compact subgroup of
Gc .) Since the Kähler form ω on B transgresses to dη on M, the circle bundle is
non-trivial, proving (ii). The first part of (iii) now follows since the last statement
implies that b1 (M ) = 0. Furthermore, the group of the deck transformations of
the cover p : Mf−−→M is a cyclic subgroup of the central S 1 , so the homogeneous
Sasakian structure on M lifts to M f. ¤

A weaker version of this theorem was given originally by Boothby and Wang
[BW58] under the more restrictive assumption that M = G/H is not only com-
pact, but also simply connected. However, somewhat later Boothby [Boo80] (in a
memorial paper dedicated to the memory of H.C. Wang) was able to remove sim-
ply connectedness from the hypothesis (see also [DMR82]). In Boothby’s paper
[Boo80] he pulls back the contact form η to a 1-form on G, and argues that the
semi-simple factor G1 of G must act transitively. (Actually he incorrectly concludes
that the central element does not occur).
Theorem 8.3.6 says that every homogeneous contact structure on a compact
manifold is homogeneous Sasakian, and that every homogeneous Sasakian structure
on a compact manifold can be realized as a circle bundle over a generalized flag
manifold. However, we do not know whether the homogeneous Sasakian structure
is determined uniquely by the homogeneous contact 1-form. Borel and Hirzebruch
[BH58] give an example of two inequivalent homogeneous complex structures on
the flag manifold B = U (4)/(U (2)×U (1)×U (1)). If both of these complex structures
were compatible with the same Kähler form ω, one could construct two inequivalent

2There seems to be some folklore around this issue. Boothby and Wang [BW58] attribute
the result to Lichnerowicz citing mimeograph notes from Princeton.
8.3. HOMOGENEOUS SASAKIAN MANIFOLDS 275

homogeneous Sasakian structures with the same contact form on the circle bundle
over B determined by [ω].
Since a compact simply connected Lie group is semi-simple, we have
Corollary 8.3.7: Let (M, η, G) be a compact homogeneous contact manifold, then
G can be taken to be a compact semi-simple Lie group.
Remark 8.3.1: Notice that if a K-contact or Sasakian structure S = (ξ, η, Φ, g) is
homogeneous (under the Lie group G), then S 0 obtained from S by a transverse
homothety 7.3.10 is also a homogeneous K-contact or Sasakian structure. More
generally, if S 0 ∈ F(Fξ ) and ζ = η 0 − η is a G invariant basic 1-form, then S 0 is also
homogeneous.
p,q
Example 8.3.8: The Wang-Ziller manifolds Mk,l described in Example 7.6.10 are
examples of homogeneous Sasakian manifolds, since the base manifold CPp ×CPq is
a generalized flag manifold Gc /P. Furthermore, they are the (k, l)-join of the homo-
geneous Sasakian manifolds S 2p+1 and S 2p+1 with Gc = S(GL(p, C) × GL(q, C)).
This example exhibits some more general properties of the join construction
applied to homogeneous Sasakian manifolds. First we note that (k1 , k2 )-join of
two simply connected homogeneous Sasakian manifolds is also a simply connected
homogeneous Sasakian manifold as long as k1 and k2 are relatively prime, which
we assume. Thus, the set of simply connected homogeneous Sasakian structures H
forms a submonoid Moreover, a well-known theorem of Borel [Bor54] states that a
simply connected homogeneous Kähler manifold corresponding to a complex semi-
simple Lie group G is Riemannian irreducible if and only if G is simple. Thus, we
have
Theorem 8.3.9: The subset H of compact simply connected homogeneous Sasakian
manifolds is closed under the (k1 , k2 )-join operation (with gcd(k1 , k2 ) = 1). Fur-
thermore, a simply connected homogeneous Sasakian manifold corresponding to a
semi-simple Lie group G is S-irreducible if and only if G is simple.
In Example 8.1.12 we have studied the Heisenberg group H2n+1 as an example
of a Lie group that admits not only a contact structure, but a Sasakian structure.
In fact, it is easy to see that H2n+1 admits a homogeneous Sasakian structure.
However, there are very few semi-simple Lie groups that admit a homogeneous
contact structure as the following theorem Boothby and Wang [BW58] shows.
Theorem 8.3.10: Let (M = G, η, G) be a homogeneous contact manifold with G
semi-simple, then G is locally isomorphic to either SO(3) or SL(2, R).
Of course, the locally isomorphic compact Lie groups SU (2) and SO(3) admit
homogeneous Sasakian structures, but according to Corollary 8.3.7 and Theorem
8.3.10 these are the only compact Lie groups that admit a homogeneous contact
structure.
Example 8.3.11: The real Heisenberg group H2n+1 = H2n+1 (R) is a nilpotent Lie
group, and as mentioned above it carries a homogeneous Sasakian structure. H2n+1
276 8. SYMMETRIES AND SASAKIAN STRUCTURES

can be realized in terms of certain n + 2 by n + 2 nilpotent matrices, namely


 
1 x1 · · · xn z
0 1 0 · · · y1 
 
 .. . .. .. 
(8.3.1) . ··· .
 
0 · · · 0 1 yn 
0 ··· 0 0 1
In terms of these matrices the group multiplication on H2n+1 defined in Exercise 6.3
is matrix multiplication from the right. As discussed previously in Example 7.3.21
the contact 1-form α of equation (6.1.1) is invariant under this action, and has a
compatible Sasakian structure with constant Φ-sectional curvature −3. Now con-
sider the nilmanifold N2n+1 = H2n+1 (R)/H2n+1 (Z) of left cosets, where H2n+1 (Z)
is the subgroup of matrices of the form of equation (8.3.1) with integer entrees.
This is the nontrivial circle bundle over an Abelian variety discussed in Example
3.5.16. The contact structure D defined by α is invariant under the right action,
and defines a contact structure, also denoted by D, on the quotient N2n+1 . In fact
the entire Sasakian structure on H2n+1 passes to the quotient. However, D is not
invariant under the group action from the left. Thus, N2n+1 is a homogeneous
manifold and a contact (actually Sasakian) manifold, but it is not a homogeneous
contact manifold.

8.4. Symmetry Reduction and Moment Maps


In this section we discuss moment maps and symmetry reduction in various
contexts. We begin with the well-known case of symplectic reduction. Since this
has been treated extensively in many books and review articles (cf. [AM78, GS84,
MS98, CdS01, CdS03, OR04]), we only give a brief description. Historically
Souriau’s book [Sou70] is very important. Apparently it is where the term ‘moment
map’ had its origin. Furthermore, it has recently been translated into English
[Sou97].
8.4.1. A Brief Review of the Moment Map and Symplectic Reduc-
tion. Let (M, ω) be a symplectic manifold and let G be a Lie group acting effec-
tively as symplectomorphisms of M, i.e., there is a group monomorphism A : G →
S(M, ω). We call such an action a symplectic action. Recall from Example 1.4.12
the definition of a Hamiltonian vector field.
Definition 8.4.1: A symplectic action A : G → S(M, ω) is said to be Hamilton-
ian if for each element τ ∈ g the associated vector field X τ is Hamiltonian, i.e.,
X τ ω = dH for some function H ∈ C ∞ (M ).
So the obstruction to being “Hamiltonian” lies in the de Rham cohomology
group H 1 (M, R). In particular, if this group vanishes then every vector field X
corresponding to an element in the Lie algebra g is Hamiltonian. The definition of
a Hamiltonian action given by Definition 8.4.1 is that of [LM87], but it is not stan-
dard; for example this is called weakly Hamiltonian in [MS98]. We let ham(M, ω)
denote the Lie algebra of Hamiltonian vector fields on M. Then a Hamiltonian
action A induces a monomorphism of Lie algebras, viz. A∗ : g−−→ham(M, ω). As
mentioned in Example 1.4.12 the Hamiltonian function is defined only up to a
constant, so we have an exact sequence of Lie algebras
0−−−→R−−−→C ∞ (M )−−−→ham(M, ω)−−−→0 .
8.4. SYMMETRY REDUCTION AND MOMENT MAPS 277

Recall that on a symplectic manifold C ∞ (M ) has a natural Lie algebra structure


given by the Poisson bracket {·, ·} which is defined by

(8.4.1) {f, g} = ω(X f , X g ) ,


where X f denotes the Hamiltonian vector field defined by the smooth function f. So
a natural question is: when does the Lie algebra homomorphism A∗ : g− →ham(M, ω)
lift to a Lie algebra homomorphism φ : g−−→C ∞ (M )? This is determined by the
vanishing of a certain cocycle in the Lie algebra cohomology group H 2 (g, R), and
there is always a choice of constants that makes this cocycle vanish. From now
on we assume that the action of G on M is Hamiltonian and this choice has been
made. This stronger notion is what is called a Hamiltonian action in [MS98] and
it always gives rise to an equivariant moment map or momentum map µ : M −−→g∗
defined by
(8.4.2) dhµ, τ i = −X τ ω,
where τ ∈ g and X τ = A∗ τ. Furthermore, the map µ is Ad∗ -equivariant in the
sense that
µ(Ag (x)) = Ad∗g µ(x) .
The moment map allows one to construct new symplectic manifolds from old ones
by the procedure of symmetry reduction. The main result is the famous Marsden-
Weinstein-Meyer reduction theorem [Mey73, MW74]. The version given here is
from [Ler03a].
Theorem 8.4.2: Let (M, ω) be a symplectic manifold with a Hamiltonian action
of a Lie group G, and let µ : M −−→g∗ denote the corresponding moment map.
Suppose further that α ∈ g∗ is a regular value of µ, and that the action of the
isotropy subgroup Gα ⊂ G is proper on µ−1 (α). Then Gα acts locally freely on
µ−1 (α), and the quotient Mα = µ−1 (α)/Gα is naturally a symplectic orbifold. If
in addition the action of Gα is free on µ−1 (α), then the quotient Mα is a smooth
symplectic manifold. Furthermore, if (M, ω) has a compatible Kähler structure and
G acts as Kähler automorphisms, then the quotient has a natural Kähler structure.
For a proof of this and related results see [GS84]. There has been much work
on adapting the reduction procedure to various settings, cf. [OR04]. The case
of most interest is when G is Abelian. In this case the isotropy subgroup Gα for
any α ∈ g∗ is the whole group G, and any choice of constant gives an equivariant
moment map. Moreover, we can employ the so-called ‘shift trick’. If µ : M −−→g∗ is
a moment map for an Abelian group G, and α ∈ g∗ is a regular value, then 0 is a
regular value of the ‘shifted moment map’ µ0 = µ − α.
If α is not a regular value of µ the situation is more complicated. Neverthe-
less, it was shown by Sjamaar and Lerman [SL91] that in this general case the
reduced space is a stratified symplectic space satisfying Goretsky-MacPherson ax-
ioms [GM88]. More precisely,
Theorem 8.4.3: Let (M, ω) be a Hamiltonian G-manifold with the moment map
µ : M −−→g∗ . Let M(H) denote the stratum consisting of orbits of type H < G. Then
NH = µ−1 (0) ∩ M(H) is a manifold and the orbit space

µ−1 (0) ∩ M(H)


M̂H =
G
278 8. SYMMETRIES AND SASAKIAN STRUCTURES

has a natural symplectic structure ω̂H whose pullback to NH coincides with the
restriction of ω to NH ⊂ M . Consequently, the stratification of M by orbit types
induces the decomposition of the reduced space M̂ = µ−1 (0)/G into a disjoint union
of symplectic manifolds [
M̂ = M̂H .
H<G
An important tool in understanding the relation between the symplectic (Kähler)
and the algebraic quotient is the norm of the moment map f = ||µ||2 defined rel-
ative to any G-invariant inner product on the Lie algebra g. Kirwan proved that
the following [Kir84]
Theorem 8.4.4: Let G be a compact Lie group and (M, ω) a G-Hamiltonian space
with the moment map µ : M −−→g∗ . For any G-invariant inner product on g the
function f = ||µ||2 is an equivariantly perfect Morse function over Q.
Suppose (M, ω) is a compact G-Hamiltonian space such that the 0 ∈ g∗ is
a regular value of the moment map µ : M −−→g∗ . Equivalently, the compact Lie
group G acts on µ−1 (0) only with finite isotropy points and the quotient (M̂ , ω̂) is
a compact symplectic orbifold. We have a natural isomorphism
(8.4.3) H ∗ (M̂ , Q) ' HG

(µ−1 (0), Q) ,

where HG (µ−1 (0), Q) is the G-equivariant cohomology of the zero level set. The
inclusion µ−1 (0) ,→ M induces a map
∗ ∗
(8.4.4) κ : HG (M, Q)−−→HG (µ−1 (0), Q) ' H ∗ (M̂ , Q)
called the Kirwan map. Under the above assumption we have [Kir84]

Theorem 8.4.5: The Kirwan map κ : HG (M, Q)−−→H ∗ (M̂ , Q) is surjective.
Kirwan’s surjectivity is the key tool in computing the rational cohomology of
the reduced space. For example, it can be effectively used to determine H ∗ (M̂ , Q)
when G is a torus in which case the kernel of the Kirwan map can be computed
[Gol02]. Tolman and Weitsman give an explicit formula for the cohomology ring
of the symplectic reduction M̂ in terms of the cohomology ring of M and fixed
point data. Under certain conditions, their formula also holds for the integral
cohomology ring, and can be used to show that the cohomology of the reduced
space is torsion-free [TW03].
Recall that a complex Lie group K is reductive if and only if K = GC is the
complexification of any maximal compact subgroup G. Consider any Kähler man-
ifold (M, I, ω, g) with a holomorphic action of K = GC , where G ⊂ Aut(M, I, ω, g)
and M is G-Hamiltonian. Let µ : M −→g∗ be the associated moment map.
Definition 8.4.6: The point x ∈ M is analytically semi-stable if the closure of
GC -orbit through x intersects µ−1 (0). The set of semi-stable points is denoted by
M ss ←- µ−1 (0). The point x ∈ M is called analytically stable if the closure of
GC -orbit through x intersects µ−1 (0) at a point where dµ is surjective.
Note that the notion of analytic (semi-)stability depends on the Kähler metric
and the momentum map. Sjamaar proves [Sja95]
Theorem 8.4.7: The inclusion µ−1 (0) ,→ M ss induces a homeomorphism
M̂ = µ−1 (0)/G −−→ M ss /GC ,
as a map of stratified spaces.
8.4. SYMMETRY REDUCTION AND MOMENT MAPS 279

The connection between the two quotients in the above theorem is established
via the Morse function f = ||µ||2 . The set of semi-stable points M ss can be seen as
the minimum Morse startum M min , i.e., set of all points whose paths of steepest
descent under f have limit points in µ−1 (0) [Kir84, Sja95].
Theorem 8.4.8: Let (M, J, ω, g) be a connected Kähler manifold and let
G ⊂ Aut(M, J, ω, g) a compact Lie group. Let complexification GC act holomorphi-
cally on M and let M be a Hamiltonian G-space with the moment map µ : M −→g∗ .
Then the set M ss coincides with the minimum Morse stratum M min and induced
by µ−1 (0) ,→ M min map
M̂ = µ−1 (0)/G −−→ M min /GC
is a homeomorphism of stratified Kähler spaces.
Example 8.4.9: One important example is when M is a projective algebraic va-
riety. Suppose K acts on a nonsingular complex projective variety M ⊂ CPn via
a representation ρ : K−→GL(n + 1, C). We can choose coordinates so that ρ re-
stricts to the unitary representation of G ⊂ U (n + 1) acting on M . Suppose M
is G-Hamiltonian and let µ : M −→g∗ be the associated moment map. Mumford’s
geometric invariant theory associates to such an action a projective quotient variety
denoted by M//K. The quotient is obtained by restricting the so-called semi-stable
points M ss ⊂ M . Mumford’s notion of (semi-)stability is purely algebraic. How-
ever, one can show that in the integral case the analytic stability is the same as
the algebraic stability so that we get the connection between the geometric and the
algebraic quotients [MFK94, Sja95].
Theorem 8.4.10: Let M ∈ CPn and let K = GC be a reductive complex Lie
group acting on M via a representation ρ : K−→GL(n + 1, C). Suppose M is a
G-Hamiltonian space. Then the notions of analytic and algebraic semi-stability
coincide. Hence, the inclusion µ−1 (0) ,→ M ss induces a homeomorphism M̂ =
µ−1 (0)/G −−→ M//K, as a map of stratified spaces.
We end this brief discussion with a remark that many of these results still hold
in the more general case when M and G are not compact as long as the moment
map µ is assumed to be proper.
Next we give the well-known convexity theorem of Atiyah, Guillemin, and
Sternberg [Ati82, GS82a].
Theorem 8.4.11: Let (M, ω) be a compact 2n-dimensional symplectic manifold
with an effective Hamiltonian action of a torus T of dimension k. Let µ : M −−→t∗
be the corresponding Ad∗ -equivariant moment map. Then the level sets of µ are
connected and the image µ(M ) is a convex polytope in t∗ .
By definition a convex polytope is the convex hull of a finite set of points {pi },
called vertices, in t∗ . It can be represented as
nX
N X o
µ(M ) = ri pi | ri = 1, ri ≥ 0 .
i=1 i

The proof of Theorem 8.4.11 uses Bott-Morse theory and is given for example in
[MS98]. The inverse image under µ of the vertices is precisely the fixed point set
of the T-action. More generally, the inverse image of the k-dimensional faces is
the subset fixed by an (n − k)-dimensional subtorus of T. We also mention that
codimension one faces are called facets.
280 8. SYMMETRIES AND SASAKIAN STRUCTURES

We are particularly interested in the case k = n. We have


Definition 8.4.12: A toric symplectic manifold is a quadruple (M, ω, A, µ),
where M is a 2n dimensional manifold, ω is a symplectic 2-form on M, A is an
effective Hamiltonian action of an n-torus T with moment map µ : M −−→t∗ .
Notice that choosing a particular component H of the moment map gives in
this case what is often called a completely integrable Hamiltonian system. A re-
markable theorem of Delzant [Del88] says that compact toric symplectic manifolds
are completely described by the image of their corresponding moment maps.
Definition 8.4.13: A Delzant polytope ∆ ⊂ Rn is a polytope satisfying the
following three conditions:

(i) There are n edges meeting at each vertex (simplicity).


(ii) If p is a vertex then each edge meeting at p has the form p + tui , where
t ≥ 0 and ui ∈ Zn (rationality).
(iii) For each vertex the corresponding u1 , . . . , un can be chosen as a Z-basis
for Zn (smoothness).
Now Delzant’s theorem is:
Theorem 8.4.14: There is a one-to-one correspondence between compact toric
symplectic manifolds and Delzant polytopes.
The proof of Delzant’s theorem has an important corollary, viz.
Corollary 8.4.15: A compact toric symplectic manifold admits a compatible Kähler
structure.
There is an orbifold version of all of this due to Lerman and Tolman [LT97]
which we now describe. First we see that by using the Definition 4.3.17 of a Lie
group action on an orbifold that Definition 8.4.12 easily generalizes to orbifolds.
Definition 8.4.16: A toric symplectic orbifold is a quadruple (X , ω, A, µ),
where X is a 2n dimensional orbifold, ω is a symplectic 2-form on X , A is an
effective Hamiltonian action of an n-torus T with moment map µ : X −−→t∗ .
The Convexity Theorem 8.4.11 holds for symplectic orbifolds with an effective
torus action. However, one obtains an ‘orbifold Delzant polytope’ given by Defi-
nition 8.4.13 with the smoothness condition (iii) dropped. In addition to classify
orbifolds one needs to assign labels to the facets of the polytope. We have
Definition 8.4.17: A Lerman-Tolman polytope or lt-polytope for short is a
convex rational simple polytope ∆ ⊂ t∗ with dim ∆ = dim t∗ and a positive integer
attached to each open facet.
Let X = (X, U) be a symplectic orbifold with a closed non-degenerate 2-form
ω. Since the local uniformizing groups preserve ω, the dimensions of the orbifold
strata differ by an even integer. In particular, the orbifold singular set Σorb (X)
has codimension at least two. The inverse image µ−1 (F ) may or may not lie in
Σorb (X), but if it µ−1 (F ) ∩ Σorb (X) 6= ∅ then µ−1 (F ) ⊂ Σorb (X). This follows from
Theorem 6.4 of [LT97]. Moreover, if F is an open facet with label mF ∈ Z+ , the
inverse image µ−1 (F ) is fixed by an S 1 and lies in the stratum of X defined by
ZmF as a local uniformizing group. So for any open facet F with label mF = 1 the
inverse image µ−1 (F ) lies in the regular stratum of the orbifold. Thus, we use the
convention that a facet with label 1 is equivalent to no label. We see that facets
with mF > 1 correspond to orbifold branch divisors, cf. Definition 4.4.8. Generally,
8.4. SYMMETRY REDUCTION AND MOMENT MAPS 281

the local uniformizing group at a point x ∈ X can be described as follows: let F(x)
denote the set of open facets whose closure contains µ(x), and let vF denote the
primitive outward normal to F ∈ F(x). Let Tx denote the isotropy subgroup of
T at x, and let ` ⊂ tx denote its integral lattice. The set {mF vF } generates a
sublattice `ˆ of `; the local uniformizing group at x is then isomorphic to `/`. ˆ
We now have the Delzant-Lerman-Tolman theorem:
Theorem 8.4.18: There is a one-to-one correspondence between compact toric
symplectic orbifolds and lt-polytopes. Moreover, the toric symplectic orbifold is a
symplectic reduction of Cn+1 by a closed Abelian subgroup of U (n + 1).
As before the proof of convexity uses Bott-Morse theory, while the construction
of a toric symplectic orbifold from an lt-polytope is similar to Delzant’s original
proof, see Theorem 8.1 in [LT97]. As in the smooth manifold case we have
Corollary 8.4.19: A compact toric symplectic orbifold X admits a compatible in-
variant complex structure, hence, a Kähler structure. In fact its underlying complex
space X is a projective algebraic variety.
Next we give an example similar to Example 6.8 in [LT97].
Example 8.4.20: Weighted projective space CP(2, 3, 6) with its standard Kähler
structure. This is a quotient of S 5 ⊂ C3 by the weighted Sw 1
-action with weights
w = (2, 3, 6), or equivalently the weighted C -quotient of C3 \ {0}. It can also

be constructed by symplectic reduction of the standard symplectic structure on


C3 . However, in this case 0 is not a regular value of the ‘unshifted’ moment map.
We take the shifted moment map for this circle action to be µ1 (z) = 2|z0 |2 +
3|z1 |2 + 6|z2 |2 − 6, where z = (z0 , z1 , z2 ) are the standard coordinates for C3 . The
choice of the shift vector 6 is a matter of convenience. Generally, for any weighted
projective space CP(w), a convenient choice is lcm(w0 , . . . , wn ). Now we have the
identification CP(2, 3, 6) = µ−1 1 1
1 (0)/Sw . The circle Sw can be viewed as the kernel
3 2
of the map à : T −−→T induced by the matrix
µ ¶
3 0 −1
0 2 −1

viewed as a map A : Z3 −−→Z2 .


We want to compute the T2 -moment map µ2 : CP(2, 3, 6)−−→t∗2 . This can be
viewed as a “quotient moment map”. The full T3 -moment map µ3 : C3 −−→t∗3 is
µ3 (z) = (|z0 |2 , |z1 |2 , |z2 |2 ). Now the restriction µ3 |µ−1 (0) descends to a map µ̄3 :
1
CP(2, 3, 6)−−→t∗3 . To find µ2 we compose this with a map t∗3 −−→t∗2 obtained by
identifying t∗2 with the hyperplane defined by y 7→ AT y + (0, 0, 1). The image
µ̄3 (CP(2, 3, 6)) lying in this hyperplane is the polytope defined by

(8.4.5) µ̄3 (CP(2, 3, 6)) = {(x0 , x1 , x2 ) ∈ t∗3 | xi ≥ 0, 2x0 + 3x1 + 6x2 = 6} .

To obtain the corresponding lt-polytope in t∗2 corresponding to this polytope, we


notice that the vertices map as follows: (0, 0, 1) 7→ (0, 0), (0, 2, 0) 7→ (0, 1), and
(3, 0, 0) 7→ (1, 0). Thus, lt-polytope is the standard simplex in t∗2 = R2 ; however,
there are labels 2 on the horizonal line from (0, 0) to (1, 0), and 3 on the vertical line
from (0, 0) to (0, 1). The diagonal line from (0, 1) to (1, 0) has a 1 or equivalently
no label.
282 8. SYMMETRIES AND SASAKIAN STRUCTURES

Recall from Proposition 4.5.10 that as algebraic varieties there are isomor-
phisms
CP(2, 3, 6) ≈ CP(1, 3, 3) ≈ CP(2, 1, 2) ≈ CP(1, 1, 1) = CP2 .
All four of these orbifolds correspond to the same simplex, namely the standard
one. They are distinguished as orbifolds by their labels. Of course, the standard
simplex with no labels corresponds to CP2 . The simplex of the orbifold CP(2, 1, 2)
has horizontal and vertical labels, 2 and 1, respectively; while the simplex of the
orbifold CP(1, 3, 3) has horizontal and vertical labels, 1 and 3, respectively.
Exercise 8.6: Consider the toric symplectic orbifold CP2 /Γ with the symplectic
form descending from the standard Fubini-Study symplectic form on CP2 , and the
action of Γ ≈ Z2 × Z3 × Z6 on CP2 is given by [z0 , z1 , z2 ] 7→ [ζ0 z0 , ζ1 z1 , ζ2 z2 ] with
ζ02 = 1, ζ13 = 1, ζ26 = 1. Show that the lt-polytope for this toric symplectic orbifold
is the standard simplex as in Example 8.4.20 but now with label 6 on each face.
While polytopes are the combinatorial data involved in symplectic geometry,
the combinatorial data of algebraic geometry are known as fans. We refer to the
books [Oda88] and [Ful93] and reference therein for details. Example 8.4.20 illus-
trates a more general phenomenon.
Proposition 8.4.21: Let X1 = (X1 , U1 , ω1 ) and X1 = (X2 , U2 , ω2 ) be compact
toric symplectic orbifolds.
(i) Then their lt-polytopes ∆1 and ∆2 are isomorphic as labelled polytopes if
and only if X1 and X2 are isomorphic as toric symplectic orbifolds.
(ii) ∆1 and ∆2 are isomorphic polytopes, but with possibly different labels,
if and only if X1 and X2 are isomorphic as toric projective algebraic
varieties.

Proof. Part (i) is proven in [LT97] to which we refer the reader. To prove (ii)
we recall that a convex rational polytope determines a fan, and the fan completely
determines [Oda88] the underlying space as a toric algebraic variety. So isomorphic
polytopes give isomorphic toric algebraic varieties and vice-versa. ¤

It follows from the Morse theory on orbifolds developed in [LT97] that the odd
dimensional cohomology groups on toric symplectic orbifolds vanish. See Proposi-
tion III.47 [Ler03a] of for a proof.
Proposition 8.4.22: Let (Z, ω) be a compact toric symplectic orbifold with un-
derlying topological space Z. Then H 2k+1 (Z, R) = 0.

8.4.2. The Contact Moment Map and Convexity. In the modern liter-
ature the relation between contact structures and symplectic cones seems to have
first appeared in Arnold’s book [Arn78]. This connection was also used in the
context of group representation theory in [GS82b].
Let (M, η) be a strict contact manifold of dimension 2n + 1 with Reeb vector
field ξ, and C(M ) its symplectification. On C(M ) we let S(C(M ), ω) denote the
group of symplectomorphisms of (C(M ), ω), and S0 (C(M ), ω) the subgroup of
S(C(M ), ω) that commutes with homotheties, i.e., the automorphism group of
the symplectic Liouville structure. The corresponding Lie algebras are denoted by
s(C(M ), ω) and s0 (C(M ), ω), respectively. Given a vector field X ∈ s0 (C(M ), ω),
the fact that X commutes with the Liouville vector field Ψ implies that X = XM ,
8.4. SYMMETRY REDUCTION AND MOMENT MAPS 283

where XM is a vector field on M lying in con(M, η). One easily sees that ( cf.
[LM87] pg. 314)
Proposition 8.4.23: There is an isomorphism S0 (C(M ), ω) ≈ Con(M, η) of topo-
logical groups induced by the natural inclusion M −−→C(M ). Infinitesimally, there
are Lie algebra isomorphisms s0 (C(M ), ω) ≈ con(M, η) ≈ C ∞ (M )ξ induced by the
maps X 7→ XM 7→ η(XM ). Furthermore, ξ is in the center of con(M, η).
We now consider the moment map construction for C(M ) and M. Let G be a Lie
group acting effectively on the symplectic cone (C(M ), ω) which leaves invariant the
symplectic form ω and which commutes with the homothety group. In particular,
G gives rise to a group homomorphism A : G−−→S0 (C(M ), ω), and since the action
is effective A(G) is a subgroup of S0 (C(M ), ω). Recall from equation (6.5.5) that
on the symplectic cone C(M ), the 2-form ω is exact. Hence, every action A :
G−−→S0 (C(M ), ω) is Hamiltonian. As usual from this we obtain a moment map
µ̃ : C(M )−−→g∗ ,
and as before we shall assume that it is equivariant. Explicitly µ̃ is defined by
dhµ̃, τ i = −X τ ω,
where X τ denotes the vector field on C(M ) corresponding to τ ∈ g. For simplicity
we denote the function hµ̃, τ i by µ̃τ . As in Section 6.5.3 we take the symplectic
2-form ω to be homogeneous of degree 2 and write ω = dη̃ = d(r2 η). Here η is the
contact 1-form on M, where we identify M with the subset M × {1} ⊂ C(M ). An
easy computation then shows that up to an additive constant
(8.4.6) µ̃τ = η̃(X τ ).
With this choice we see that under homotheties r 7→ et r we have
(8.4.7) ω 7→ e2t ω, η̃ 7→ e2t η̃, µ̃ 7→ e2t µ̃ .
We now specialize to the case when the Lie group G is a torus T with Lie
algebra t. In this case the kernel of the exponential map exp : t−−→T forms a lattice
called the integral lattice of T and denoted by ZT . This lattice represents the set of
circle subgroups of T. There is a convexity theorem for symplectic cones [dMT97],
but first we have
Definition 8.4.24: A subset C ⊂ t∗ is called a polyhedral cone if it can be
represented by
\
C = {y ∈ t∗ | hy, vi i ≥ 0}
for some finite set of vectors {vi }. It is rational if in addition all the vectors vi lie
in the integral lattice ZT .
The vectors vi are the ‘inward normal vectors’ of the polyhedral cone. Without
loss of generality we can assume that each vi is primitive in the sense that svi 6∈ ZT
for 0 < s < 1, and that the set {vi } is minimal, that is, for any index j, C 6=
∩i6=j {y ∈ t∗ | hy, vi i ≥ 0}.
Theorem 8.4.25: Let (C(M ), ω) be a symplectic cone with M compact, and let
A(T) be a torus subgroup of S0 (C(M ), ω). Assume further that there is an element
τ ∈ t such that hµ̃, τ i > 0. Then the image µ̃(C(M )) is a convex polyhedral cone.
The positivity condition hµ̃, τ i > 0 is quite strong. For example, it is easy to
see that it implies that 0 is not in the image of the moment map µ̃.
284 8. SYMMETRIES AND SASAKIAN STRUCTURES

We wish to relate the symplectic moment map on C(M ) to a contact moment


map on M. This is easy to do since G commutes with homotheties, and we get
an induced action of G on M. By Proposition 8.4.23 we can identify G with a Lie
subgroup of Con(M, η) which by abuse of notation we continue to denote by G.
This gives a moment map by restriction, viz.

µη : M −−→g∗ ,

(8.4.8) µη = µ̃|M ×{1} , µτη = η(X τ ) .

Exercise 8.7: Show that the contact moment map µη of equation (8.4.8) can be
defined intrinsically without reference to the symplectification C(M ).
Notice that if f is a G-invariant function on M, and η 0 = f η is another contact 1-
form in the same contact structure, then we get a different symplectic form ω 0 which
is again G-invariant, and this gives rise to a different moment maps, namely µ̃0 = f µ̃,
and µ0η0 = f µη . It is convenient to have a moment map that only depends on the
underlying contact structure and not on the choice of 1-form. In order to define such
a moment map, we follow Lerman [Ler02a, Ler03a] and consider the annihilator
Do of D in T ∗ M. This is just the contact line bundle L defined by equation (6.1.5),
which in the current case is trivial. Choosing a contact 1-form η (nowhere vanishing
section of Do ) trivializes the bundle Do ≈ M × R, as well as chooses an orientation
of Do , so we can split Do \ {0} as D+ o
∪ D− o
in which case sections of D+ o
are
identified with {f η} with f > 0. Now the idea is to represent the cone C(M ) by
o
D+ ≈ M × R+ . Consider the group Con(M, D) of contactomorphisms of (M, D). By
pullback Con(M, D) lifts to an action on T ∗ M which leaves Do invariant. Con(M, D)
has a subgroup Con(M, D)+ defined to be the subset of elements in Con(M, D) that
o
leave D+ invariant. Notice that for any choice of contact 1-form η there is a sequence
of subgroups Con(M, η) ⊂ Con(M, D)+ ⊂ Con(M, D).
Definition 8.4.26: Let G be a Lie group acting effectively as contactomorphisms
that preserves the co-orientation, that is A(G) ⊂ Con(M, D)+ . We define a moment
o
map Υ : D+ −−→g∗ by
hΥ(x, α), τ i = hα, X τ (x)i.
If the action of G preserves a contact 1-form η of the contact structure (i.e.,
A(G) ⊂ Con(M, η)), then we have a commutative diagram
o
(8.4.9) D+
O @@
@@
@@
@@ Υ
η @@
@@
@@
@@
µη Ã
M / g∗

or in symbols Υ ◦ η = µη . An observation of Lerman [Ler02a] says that for proper


group actions the condition that G preserves a contact 1-form is not a restriction.
Lemma 8.4.27: Let G be a Lie group acting properly and effectively on a co-
oriented contact manifold M such that A(G) ⊂ Con(M, D)+ Then there exists a
o
G-invariant section of D+ , that is a G-invariant contact form η such that D = ker η.
8.4. SYMMETRY REDUCTION AND MOMENT MAPS 285

Proof. Choose a contact 1-form η 0 . If G is compact we define an invariant


contact form η by averaging, viz.
Z
1
η= g ∗ η 0 dg .
Vol(G) G
If G is non-compact we can use a slice theorem [Pal61] as done in Proposition 2.5.2
of [DK00], or in [Loo01]. ¤
Thus, for proper effective actions we can assume that A(G) ⊂ Con(M, η) for
o
some contact 1-form η. In this case the identification of D+ with the cone C(M )
holds as G-manifolds, and the moment map Υ coincides with µ̃. The advantage of
working with Υ over µη is obvious; the latter depends on the choice of 1-form η,
whereas, the former depends only on the contact structure D up to a choice of co-
orientation. Nevertheless, since this book mainly deals with Sasakian or K-contact
geometry where a 1-form is fixed working with the moment map µη will often be
sufficient.
However, when working with G-invariant contact structures, it is convenient to
define the moment cone
(8.4.10) C(µη ) = {tγ ∈ g∗ | γ ∈ µη (M ), t ∈ [0, ∞)}
which is an invariant of the contact structure with its G action. Clearly, we have
o o
C(µη ) = Υ(D+ ) ∪ {0}, and C(µη ) is independent of the choice of 1-form η in D+ .
Accordingly, we denote C(µη ) by C(Υ).
Definition 8.4.28: We say that a Lie group action A : G−−→Con(M, D)+ is of
Reeb type if there are a contact 1-form η of the contact structure D and an
element ς ∈ g such that X ς is the Reeb vector field of η.
Notice that a vector ς ∈ g such that X ς is the Reeb vector field satisfies
hΥ, ςi = 1. We call ς the characteristic vector of g, and the equation hΥ, ςi = 1 the
characteristic hyperplane in g∗ . (This hyperplane is called the Reeb hyperplane in
[BM93a, BM96]). Now choosing the standard Cartesian coordinates (r0 , . . . , rn )
on g∗ and expanding the characteristic vector in the standard basis {ei }i for g, this
hyperplane can be written as
Xn
(8.4.11) ri wi = 1 .
i=0
When the coefficients wi are integers it will often be convenient to rescale the
coordinates ri so that the characteristic hyperplane takes the form
Xn
(8.4.12) ri wi = lcm(w0 , . . . , wn ) .
i=0
Then the hyperplane arising in Example 8.4.20 is just the characteristic hyperplane
of the associated contact structure. Concerning actions of Reeb type we have
Proposition 8.4.29: The action A : G−−→Con(M, D)+ is of Reeb type if and only
if there is an element τ in the center z of g such that hΥ, τ i > 0.
Proof. The only if part is clear. So if τ ∈ g satisfies hΥ, τ i > 0, we have
η(X τ ) > 0 for any section η : M −−→D+
o
. We then define the contact form
1
η0 = η.
η(X τ )
286 8. SYMMETRIES AND SASAKIAN STRUCTURES

Since τ is in the center of g, the 1-form η 0 is G-invariant. But then X τ is the Reeb
vector field of η 0 . ¤

Proper actions of Reeb type have a particularly nice property, namely


Proposition 8.4.30: Let (M, D) be a co-oriented contact manifold, and let A :
G−−→Con(M, D)+ be a proper effective action of Reeb type of a Lie group G. Then
there exists a G-invariant K-contact structure S = (ξ, η, Φ, g) belonging to the con-
tact structure defined by D.

Proof. By Lemma 8.4.27 there exists a G invariant contact form η with D =


ker η. Choose a compatible Riemannian metric. Then as in the proof of Proposition
2.5.2 of [DK00] we obtain a G-invariant Riemannian metric g compatible with the
contact structure. Then the Reeb vector field ξ of η is an infinitesimal isometry
of g. Moreover, Φ is defined to be the endomorphism obtained from dη using the
metric g. The structure (ξ, η, Φ, g) is then K-contact. ¤

Proposition 8.4.30 together with Theorem 7.1.10 have an important corollary.


Corollary 8.4.31: Let (M, D) be a co-oriented compact contact manifold with an
effective k-torus action A : Tk −−→Con(M, D)+ . Suppose also that this action is of
Reeb type. Then there is a compatible quasi-regular Tk -invariant K-contact struc-
ture S = (ξ, η, Φ, g) so that M is an orbifold fibration π : M −−→Z over a compact
symplectic orbifold (Z, ω) with π ∗ ω = dη. Furthermore, the quotient T/S 1 of T
by the S 1 generated by the Reeb vector field acts effectively on Z preserving the
symplectic structure.
Actually the base orbifold Z has a Tk−1 invariant almost Kähler structure.
Indeed we are mainly concerned with the case when G is a torus T in which case
one has a convexity theorem for certain contact manifolds. It was first proved
by Banyaga and Molino [BM93a, BM96] in the case of compact toric contact
manifolds, that is when the torus has maximal dimension. It was then generalized
by Lerman [Ler02b] for any torus of dimension bigger than two under the condition
that the image of the moment map does not contain the origin. Before describing
this we have
Definition 8.4.32: A toric contact manifold is a quadruple (M, D, T, Υ), where
(M, D) is an oriented and co-oriented contact manifold of dimension 2n + 1 with
an effective action A : T−−→Con(M, D)+ of a (n + 1)-torus T and moment map
Υ : D0+ −−→t∗n+1 .
Lerman [Ler02a] classified all compact toric contact manifolds. In particular,
he shows that most of these are toric contact manifolds of Reeb type. The ones
that are not of Reeb type are completely described and are of less interest to us
in this book. We state Lerman’s theorem below without proof. Another result in
[Ler02a] which is obvious for toric contact manifolds of Reeb type is:
Proposition 8.4.33: For toric contact manifolds 0 is not in the image of the
moment map Υ.
The convexity theorem is:
Theorem 8.4.34: Let (M, D) be a compact co-oriented contact manifold of dimen-
sion 2n + 1 and let A : T−−→Con(M, D)+ be an effective action of a torus on M
with moment map Υ. Assume also that dim T > 2, and that 0 is not in the image
8.4. SYMMETRY REDUCTION AND MOMENT MAPS 287

of Υ. Then the fibres of Υ are connected and the moment cone C(Υ) is a convex
rational polyhedral cone.
When the action A is of Reeb type a slightly weaker convexity theorem follows
from Theorem 8.4.25 and Proposition 8.4.29. Since A of Reeb type implies that 0 is
not in the image of the moment map, the convexity result is subsumed by Theorem
8.4.34 at least in the case that dim T > 2. We now choose a contact 1-form η, i.e.,
o
a smooth section of D+ , together with its Reeb vector field. This determines the
characteristic hyperplane hΥ, ςi = 1 which intersects the polyhedral cone C(Υ) in
a polytope. By applying Corollary 8.4.31 and using the Lerman-Tolman Theorem
8.4.18 it is easy to see [BG00a] that this polytope is rational if and only if the
characteristic vector lies in the integral lattice ZT . Summarizing we have
Theorem 8.4.35: Let A : T−−→Con(M, D)+ be an effective action of a torus of
Reeb type. Then the moment cone C(Υ) ⊂ t∗ is a convex rational polyhedral cone.
o
Furthermore, a choice of T-invariant contact form η in D+ together with its Reeb
ς ∗
vector field X intersects C(Υ) ⊂ t with the characteristic hyperplane hΥ, ςi =
1 giving the image µη (M ) as a convex polytope in the characteristic hyperplane.
Moreover, the polytope is rational if and only if the characteristic vector ς lies in
the integral lattice ZT .
Actually the convex rational polyhedral cones of Theorem 8.4.35 must obey
other conditions. The following definition is due to Lerman [Ler02a].
Definition 8.4.36: A rational polyhedral cone C is said to be good if it satisfies
the following condition: if
k
\ © ª
y ∈ t∗ | hy, vij i ≥ 0
j=1

is a k-face of C then the vij are independent over Z and the equality
nX
k o nX
k o
aj vij | zj ∈ R ∩ ZT = mj vij | mj ∈ Z
j=1 j=1

holds.
Identifying t with its double dual t∗∗ , this condition says that the annihilator
of a k face of C is precisely the Lie algebra of a subtorus H of T.
We are now ready for Lerman’s classification theorem [Ler02a].
Theorem 8.4.37: Let (M, D, T, Υ) be a compact toric contact manifold. Then
(i) If dim(M ) = 3 and the action of T2 is free, then M is diffeomorphic to
the 3-torus T3 , and there is a section η of D0+ having the form cos nt dθ1 +
sin nt dθ2 for some positive integer n.
(ii) If dim(M ) = 3 and the action of T2 is not free, then M is diffeomorphic
to a lens space or S 1 × S 2 and (M, D, T2 , Υ) is classified by two rational
numbers r, q with 0 ≤ r < 1, r < q.
(iii) If dim(M ) = 2n+1 > 3 and the action of Tn+1 is free, then M is a prin-
cipal Tn+1 -bundle over S n . Moreover, each principal Tn+1 -bundle over
S n has a unique Tn+1 -invariant contact structure making it a contact
toric manifold.
(iv) If dim(M ) = 2n + 1 > 3 and the action of Tn+1 is not free, then the
moment cone C(Υ) is a good rational polyhedral cone. Conversely, if
288 8. SYMMETRIES AND SASAKIAN STRUCTURES

C ⊂ t∗ is a good rational polyhedral cone, there is a unique toric contact


manifold (M, D, Tn+1 , Υ) whose moment cone C(Υ) is C.
We are mainly interested in case (iv) of this theorem. We have
Proposition 8.4.38: Let (M, D, T, Υ) be a compact toric contact manifold with
dim(M ) = 2n + 1 > 3. Then the action of T on M is of Reeb type if and only if it
is not free and the moment cone C(Υ) contains no nonzero linear subspace.

Proof. Suppose that the action of Tn+1 is of Reeb type, then by Proposition
8.4.30 there is a compatible K-contact structure, and by perturbing the Reeb vector
field if necessary it is quasi-regular. Thus, the space of leaves of the characteristic
foliation is a compact symplectic orbifold X . Moreover, the quotient n-torus Tn =
Tn+1 /S 1 , where S 1 is the subgroup of Tn+1 generated by the Reeb vector field,
acts on the orbifold X preserving its symplectic structure. Such an action is never
free if n > 1 [LT97].
Conversely suppose n > 1 and the action of Tn+1 is not free, then by (iv) of
Lerman’s Theorem 8.4.37 C(Υ) is a good rational polyhedral cone containing no
nonzero linear subspace. This implies that there is an element τ ∈ t such that
hΥ, τ i > 0. Thus, by Proposition 8.4.29 the action is of Reeb type. ¤

There is a counterexample to this proposition when dim(M ) = 3, namely S 1 ×


2
S , where the action is neither free nor of Reeb type. We give some examples of
toric contact manifolds of Reeb type.
Example 8.4.39: We consider the standard sphere S 2n+1 with the standard contact
structure 7.1.5. This is the contact structure induced from the standard symplectic
structure on R2n+2 . In this case the standard contact form η0 and its Reeb vector
field ξ0 are those given in Example 7.1.5. The maximal torus Tn+1 is generated
by the vector fields Hi of that example, and this provides the sphere S 2n+1 with
a toric contact structure. Let {ei }ni=0 denote the standard basis for tn+1 = Rn+1
corresponding to the vector fields Hi , that is, X ei = Hi , and let {e∗i }i denote the
dual basis of t∗n+1 . From equation (8.4.8) the moment map becomes
n
X n
X
µη0 (x, y) = η0 (Hi )e∗i = |zi |2 e∗i .
i=0 i=0

So the moment cone C(Υ) = C(µη0 ) is the union of {0} with the positive 2n+1 -
tant. If we let (q, p) denote the fibre coordinates of T ∗ M corresponding to (x, y)
respectively, we see that
n
X
Υ(x, y; q, p) = (qi yi − pi xi )e∗i .
i=0

Letting r0 , . . . , rn denote the coordinates for t∗n+1 we see that the characteristic
hyperplane is just
r0 + · · · + rn = 1 .
Thus, the image µη0 (S 2n+1 ) is just the standard n-simplex in the characteristic
hyperplane with ri ’s as barycentric coordinates. We want to identify this with a
Delzant polytope. For simplicity we put n = 2 and consider the exact sequence
ι $
0−−−→t1 −−−→ t3 −−−→ t2 −−−→0 ,
8.4. SYMMETRY REDUCTION AND MOMENT MAPS 289

where t1 is the one dimensional Lie algebra defined by the characteristic vector ς.
The map $ can be represented by the matrix
µ ¶
1 0 −1
A= .
0 1 −1

The characteristic hyperplane is then represented as AT y + (0, 0, 1). Thus we


see that the vertices (0, 0, 1), (1, 0, 0), (0, 1, 0) are mapped by $ to the vertices
(0, 0), (1, 0), (0, 1), respectively, giving the Delzant polytope as the standard sim-
plex in t∗2 = R2 . This corresponds to the toric symplectic manifold (CP2 , ω), where
ω is the standard symplectic form given by the Fubini-Study metric.
Choosing a weighted contact form ηw and equivalently a weighted Reeb vector
field ξw as in Example 7.1.12, determines a different characteristic hyperplane,
namely that of equation (8.4.12). The toric CR contact structure is represented
by the moment cone C(Υ), whereas, choosing a particular Reeb vector field in the
Sasaki cone t+ n+1 cuts C(Υ) by the corresponding characteristic hyperplane.
Example 8.4.40: We revisit Example 8.4.20 in the context of contact geometry.
There we described the weighted projective space as a shifted symplectic reduction
of the cone C3 −{0} by the weighted circle action Sw 1
with weights w = (2, 3, 6). The
−1 5
0 level set µ (0) of the shifted moment map is the weighted sphere Sw represented
2 2 2
as 2|z0 | + 3|z1 | + 6|z2 | = 6 which differs from the standard representation as in
5
Example 8.4.39. In this case the contact 1-form on Sw is just a rescaled standard
contact form, namely 61 η with Reeb vector field ξw = 2H0 +3H1 +6H2 . The rescaling
of the contact form corresponds to the rescaling of the characteristic hyperplane,
viz. 2r0 + 3r1 + 6r2 = 6 which coincides with the hyperplane described in Example
8.4.20. Now the rational polyhedral cone corresponding to the standard contact
structure is just the positive octant with the inward normal vectors (1, 0, 0), (0, 1, 0)
and (0, 0, 1). Intersecting the rescaled characteristic hyperplane with this cone gives
the rational polytope 8.4.5. Notice that on the level of contact geometry there is no
need for labels, the branch divisors are encoded in the equation for the characteristic
hyperplane. This information is lost, however, when transforming to polytopes in
t∗2 , hence, the need for labels at this level.
Example 8.4.41: Consider the lens spaces L(p, q1 , q2 ) with p relatively prime to
qi as discussed in Example 8.1.21. The contact structure on L(p, q1 , q2 ) is that
induced from the standard contact structure on S 5 . We follow [GL02] (Note that
L(p, q) should be L(q, p) in [GL02]) and consider the surjective map φ : T3 −−→T3
defined by φ(ζ0 , ζ1 , ζ2 ) = (ζ0p , ζ0−q1 ζ1 , ζ0−q2 ζ2 ). The kernel of φ is the subgroup Γ
of T3 defined by {ζ0p = 1, ζ1 = ζ0q1 , ζ2 = ζ0q2 } which is isomorphic to Zp . The lens
space L(p, q1 , q2 ) is the quotient of S 5 by the action (z0 , z1 , z2 ) 7→ (ζz0 , ζ q1 z1 , ζ q2 z2 ),
where ζ is a pth root of unity. There is an action of T3 /Γ on L(p, q1 , q2 ) induced
by the standard action of T3 on S 5 . The image of the induced moment map is the
polyhedral cone

Cp,q1 ,q2 = {r0 (1, 0, 0) + r1 (q1 , p, 0) + r2 (q2 , 0, p) | r0 , r1 , r2 ≥ 0} .

Consider the sublattice ` of the integral lattice T3Z generated by the inward nor-
mal vectors to the facets of Cp,q1 ,q2 . It is easy to see that ` is generated by
(0, 0, 1), (0, 1, 0) and (p, 0, 0). Thus, we verify item (ii) of Theorem 8.4.42 below
that π1 (L(p, q1 , q2 )) = T3Z /` ≈ Zp .
290 8. SYMMETRIES AND SASAKIAN STRUCTURES

Let us choose the standard contact form on S 5 and Reeb vector field ξw =
5
H0 + q1 H1 + q2 H2 with w = (1, q1 , q2 ). We represent the weighted Sw as |z0 |2 +
2 2
q1 |z1 | +q2 |z2 | = q1 q2 , where for simplicity we take q1 and q2 to be relatively prime.
The rescaled characteristic hyperplane is r0 + q1 r1 + q2 r2 = q1 q2 and intersecting
this with the cone Cp,q1 ,q2 gives the polytope in the x = q1 q2 plane with vertices
(q1 q2 , 0, 0), (q1 q2 , q2 p, 0), (q1 q2 , 0, q1 p). Identifying this plane with t∗2 = R2 gives the
lt-polytope with vertices (0, 0), (q2 p, 0), (0, q1 p) with no labels. This is the poly-
tope corresponding to the orbifold weighted projective space P(1, q1 , q2 ) with the
2
symplectic form pω, where [ω] ∈ Horb (P(1, q1 , q2 ), Z) is a generator. In particular,
the special case q1 = q2 = 1 gives CP2 with symplectic form pω, where ω is the
Fubini-Study 2-form.
Exercise 8.8: Discuss the case in Example 8.4.41 when q1 and q2 are not relatively
prime.
Concerning the topology of compact toric contact manifolds we have the fol-
lowing theorem due to Lerman. Items (ii) and (iii) below are given in [Ler04b]
and (i) follows from Proposition 8.4.22 and Corollary 8.4.31 by a Gysin or spectral
sequence argument.
Theorem 8.4.42: Let (M, D, T, Υ) be a compact toric contact manifold of Reeb
type of dimension 2n+1 > 3, and let ` denote the sublattice of the integral lattice ZT
generated by the normal vectors to the facets of the polyhedral cone C(Υ) associated
to M. Then
(i) The Betti numbers b2i+1 (M ) vanish for i = 0, . . . n − 2.
(ii) π1 (M ) is isomorphic to the finite Abelian group ZT /`.
(iii) π2 (M ) is a free Abelian group of rank N − n − 1, where N is the number
of facets of C(Υ).

8.5. Contact and Sasakian Reduction


The contact analogue of the Marsden-Weinstein-Meyer reduction theorem was
first treated systematically by Albert [Alb89]. However, it appeared earlier in
[GS82a] in the guise of co-isotropic reduction of symplectic cones which is closely
related to a well-known generalization of reduction due to Kazhdan, Kostant, and
Sternberg [KKS78]. A related 3-Sasakian version of contact reduction appeared
later in [BGM94a]. At this stage we were unaware of both [Alb89] and the
connection with [GS82a]. This also appears to be the case of Geiges [Gei97a],
and even more recently in [Loo01] where contact reduction in the more general
case when the contact structure is not necessarily co-oriented is considered. In this
book we are only concerned with ‘contact reduction at 0’, and Theorem 8.5.1 below
assumes both that the contact structure is co-oriented and that 0 is a regular value
of the moment map. The singular case when 0 is not a regular value of the moment
map is treated in [LW01]. As in the symplectic case this gives rise to stratified
quotients. A version of reduction at nonzero values of the contact moment map
was described in Albert [Alb89] and is implicit in [GS82a]. A somewhat different
approach to reduction at non-zero values of the contact moment map has recently
been developed in [Wil02], and a Sasakian version in [DO06]. Contact reduction
of cosphere bundles both at 0 and not are treated in [DOR02]. The important
consideration about reduction at 0 is that it is independent of the choice of contact
form.
8.5. CONTACT AND SASAKIAN REDUCTION 291

Theorem 8.5.1: Let (M, D) be an oriented and co-oriented contact manifold of


dimension 2n+1, and let G be a compact Lie group with an effective action A(G) ⊂
Con(M, D)+ . Let η be a G-invariant contact form with ker η = D, and let µη :
M −−→g∗ denote its moment map. Then if 0 is a regular value of µη , the quotient
M0 = µ−1
η (0)/G is a contact orbifold of dimension 2(n − dim g) + 1 whose contact
1-form η0 descends from η. Furthermore, the quotient M0 depends only on the
underlying contact structure D and not on the 1-form. If the action of G is free on
µ−1
η (0), then the quotient (M0 , D0 ) is a smooth contact manifold with D0 = ker η0 .

Proof. Since 0 is a regular value of µη , the differential dµη has rank equal
to the dimension of G at all points of the zero level set µ−1 η (0) which is a smooth
embedded G-invariant submanifold of M of codimension dim G. Moreover, the
differential dµη (x) : Tx M −−→Tµη (x) g∗ ≈ g∗ is the transpose of the evaluation map
τ 7→ X τ (x). So G acts locally freely on µ−1 η (0). Since G is compact there exists
a G-invariant Riemannian metric compatible with the contact form η such that
the locally free action of G defines a Riemannian foliation of M with compact
leaves. Thus, by Molino’s Theorem 2.5.11, the space of leaves M0 = µ−1 η (0)/G is
an orbifold, and the natural map π : µ−1 η (0)−−→µ −1
η (0)/G is an orbifold Riemannian
submersion. Furthermore, if the action is free, then all the local uniformizing groups
are trivial so the quotient M0 is a smooth manifold.
Now from Definition 8.4.26
(8.5.1) µ−1 τ
η (0) = ∩τ ∈g {η(X ) = 0} ,

so for all τ ∈ g the vector fields X τ restrict to µ−1


η (0) and lie in the restriction of
the contact subbundle D to µ−1 η (0). Then it follows from invariance that η|µ−1
η (0)
is
basic. So it descends to a contact 1-form η0 on M0 . Defining D0 = ker η0 implies
that (M0 , D0 ) is a contact orbifold. Moreover, from invariance (M0 , D0 ) depends
only on the contact structure D and not on the choice of 1-form. ¤
Remark 8.5.1: The hypothesis of Theorem 8.5.1 can be weakened to a proper
effective action of any Lie group preserving the contact structure. In this case
Molino’s Theorem does not apply. However, one can use a slice theorem such as
Theorem 2.4.1 of [DK00] to construct an orbifold chart on the quotient. This
applies equally well to Theorems 8.5.2 and 8.5.3 below.
Now let ξ denote the Reeb vector field of a contact manifold (M, η). Let A(G) ⊂
Con(M, η) be a proper effective action of a Lie group G. Notice that if ξ corresponds
to an element in the Lie algebra g, then 0 in not in the image of the moment map
µη . So we cannot perform reduction at 0 by G in this case. Thus when doing contact
reduction by a Lie group G we always assume that the action A(G) ⊂ Con(M, η) is
not of Reeb type. However, by Proposition 6.1.31 ξ lies in the center of con(M, η),
so ξ commutes with A∗ g. It follows that for all τ ∈ g we have ξη(X τ ) = 0. Thus,
from equation (8.5.1), at points of µ−1 η (0) the Reeb vector field ξ is tangent to
µ−1
η (0), and since ξ commutes with A ∗ g, it descends to a vector field ξ0 on the
quotient M0 . It is easy to check that ξ0 is the Reeb vector field of η0 .
Given the contact manifold (M, η) we choose a compatible almost contact struc-
ture (ξ, η, Φ) satisfying equations (6.4.1). We then choose the associated Riemann-
ian metric, i.e., g = dη ◦ (Φ ⊗ 1l) + η ⊗ η to obtain a contact metric structure
(ξ, η, Φ, g). Now let A(G) ⊂ Con(M, D)+ be an effective action of a compact Lie
group as in Theorem 8.5.1, but suppose also that A(G) preserves the contact metric
292 8. SYMMETRIES AND SASAKIAN STRUCTURES

structure, that is, A(G) ⊂ Aut(ξ, η, Φ, g). Then for any φ ∈ Aut(ξ, η, Φ, g) we have
φ∗ ξ = ξ , φ∗ η = η , φ∗ Φφ−1
∗ = Φ, φ∗ g = g .
Theorem 8.5.2: Assuming the hypothesis of Theorem 8.5.1, we let (ξ, η, Φ, g) be
a contact metric structure associated to the underlying contact structure. Suppose
in addition that A(G) is a subgroup of Aut(ξ, η, Φ, g). Then there exists a unique
contact metric structure (ξ0 , η0 , Φ0 , g0 ) on the quotient M0 such that ι∗ g = π ∗ g0
and ι∗ η = π ∗ η0 , where ι : µ−1 η (0)−−→M denotes the natural inclusion and π :
µ−1
η (0)−− →M 0 is an orbifold submersion. Furthermore, if (ξ, η, Φ, g) is K-contact
so is (ξ0 , η0 , Φ0 , g0 ).

Proof. We already have constructed (ξ0 , η0 ) on the quotient space M0 which


under the hypothesis of Theorem 8.5.1 is a contact orbifold. We need to construct
Φ0 and g0 . The latter is easy since by invariance the induced metric ι∗ g descends
to a unique metric g0 on the complement to the trivial line orbibundle on M0
generated by ξ0 . To construct Φ0 we notice that for any τ ∈ g the vector field X τ
restricted to the zero set µ−1 −1
η (0) is tangent to µη (0), whereas, the vector fields
τ −1
ΦX are sections of the normal bundle N µη (0). To see this we notice that for any
τ ∈ g, and for any vector Y tangent to µ−1η (0) this implies

0 = (£X τ η)(Y ) = dη(X τ , Y ) + d(η(X τ ))(Y ) = dη(X τ , Y ) .


So for all τ ∈ g and all Y tangent to µ−1 τ τ
η (0), we have g(ΦX , Y ) = −dη(X , Y ) = 0.
Now since the action A of G is locally free, the vector fields {X τ }τ ∈g gener-
ate a trivial integrable subbundle E of T M. Furthermore, the non-degeneracy of
Φ on the complement to the trivial line bundle Lξ gives an isomorphism Φ|E :
E−−→N µ−1 η (0). But we also have the orthogonal decomposition

T M |µ−1
η (0)
= F ⊕ E ⊕ N µ−1
η (0) ,

where F is the orthogonal complement to E in T µ−1 η (0). So the non-degeneracy


of Φ implies that Φ|F is an endomorphism of F satisfying (Φ|F )2 = −1lF + ξ ⊗ η.
Now F is isomorphic to the tangent bundle of M0 , and Φ|F is invariant under the
group action, so ΦF descends to a unique endomorphism Φ0 on T M0 satisfying
Φ20 = −1l + ξ0 ⊗ η0 . This constructs the contact metric structure on M0 .
If £ξ g = 0 it is easy to trace through the identifications to prove that £ξ0 g0 = 0
as well. Thus, (ξ0 , η0 , Φ0 , g0 ) is a K-contact structure if (ξ, η, Φ, g) is K-contact. ¤

The reduction which is of most interest to us is given by the next theorem


due to Grantcharov and Ornea [GO01]. There are several ways to prove this
theorem. The proof in [GO01] uses the curvature conditions of Proposition 7.3.17
for Sasakian manifolds. Our proof uses Theorem 8.5.2 and Theorem 7.3.16.
Theorem 8.5.3: Let (M, ξ, η, Φ, g) be a Sasakian manifold of dimension 2n+1 and
let G be a compact Lie acting as automorphisms of the Sasakian structure. Assume
that 0 is a regular value of the moment map µη : M −−→g∗ . Then the quotient
orbifold M0 with its induced structure S0 = (ξ0 , η0 , Φ0 , g0 ) is Sasakian.

Proof. Consider the (2, 1)-tensor field ∇Φ on M. Since (M, ξ, η, Φ, g) is Sasakian


∇Φ satisfies ∇Φ = ξ ⊗ g − 1l ⊗ η by Theorem 7.3.16. But from the proof of Theorem
8.5.2 this equation descends to ∇0 Φ0 = ξ0 ⊗ g0 − 1l ⊗ η0 , where ∇0 is the Levi-Civita
connection of g0 . Thus, (ξ0 , η0 , Φ0 , g0 ) is Sasakian by Theorem 7.3.16. ¤
8.5. CONTACT AND SASAKIAN REDUCTION 293

There is a simple relationship between contact reduction at 0 and symplectic


reduction of a symplectic cone. Let us apply Theorem 8.4.2 with α = 0 to a
symplectic cone C(M ). First notice that the symplectic 2-form ω given by equation
(6.5.5) is exact, so every symplectic action is Hamiltonian. By Proposition 8.4.23
the groups S0 (C(M ), ω) and Con(M, η) are identified, and their corresponding
moment maps are related by equation (8.4.8). We are ready for
Theorem 8.5.4: Let (M, η) be a contact manifold and (C(M ), ω) its symplectic
cone with ω = d(r2 η). Let G be a compact Lie group with an effective action A(G) ⊂
Con(M, η) = S0 (C(M ), ω), and let µη and µ̃ be the moment maps on M and C(M ),
respectively. Suppose that 0 ∈ g∗ is a regular value of the moment map µη̃ . Then
the symplectic reduction (C(M )0 , ω0 ) of (C(M ), ω) is the symplectic cone on the
contact reduction (M0 , η0 ) of (M, η).
Proof. The reduced space C(M )0 is µ̃−1 (0)/A(G), where by abuse of notation
A denotes the action of G on both M and C(M ). It follows from the transformations
(8.4.7) and the fact that A(G) commutes with homotheties that C(M )0 is a cone.
Moreover, equations (8.4.8) imply that µ̃−1 (0) = µ−1 +
η (0) × R . So C(M )0 is a cone
over M0 . Letting r denote the coordinate on R+ one easily sees by invariance under
A(G) that ω0 = d(r2 η0 ), so C(M )0 is the symplectic cone on M0 . ¤
Now suppose that M has a contact metric structure (ξ, η, Φ, g), then by Propo-
sition 6.5.5 this determines a unique almost Kähler structure (dr2 + r2 g, d(r2 η), I)
on the symplectic cone C(M ), where these structures are related by equations
(6.5.2). Since every element g ∈ Aut(ξ, η, Φ, g) lifts to an element in g ∈ Aut(dr2 +
r2 g, d(r2 η), I), Theorems 8.5.2 and 8.5.4 imply that the corresponding reduced
structures (ξ0 , η0 , Φ0 , g0 ) and (dr2 + r2 g0 , d(r2 η0 ), I0 ) on M0 and C(M0 ), respec-
tively are also related by equations (6.5.2). By Proposition 6.5.6 M is K-contact
if and only if Ψ − iξ is pseudo-holomorphic, so again by Theorems 8.5.2 and 8.5.4
the reduced structure (ξ0 , η0 , Φ0 , g0 ) will be K-contact with Ψ − iξ0 is pseudo-
holomorphic. The reductions are also clearly compatible having Sasakian structures
on M and M0 and the corresponding Kähler structures on C(M ) and C(M0 ). Let
us denote the reduction procedure, either symplectic or contact, by a double arrow
⇓ . Then we can summarize our results of reduction by a compact Lie group G with
the follow commutative diagram
M ←→ C(M )
(8.5.2) ⇓ ⇓
M0 ←→ C(M0 ),
where the two-sided arrow ←→ can be interpreted as either a quotient projection
M ←− C(M ) by R+ , or an inclusion M −→ C(M ) putting r = 1. In the case
that the contact metric structure (ξ, η, Φ, g) on M is quasi-regular, it is naturally
K-contact by Theorem 7.1.10 and we have a natural orbifold fibration π : M −−→Z.
Thus, we can extend our diagram to the commutative diagram
Z ←− M ←− C(M )
(8.5.3) ⇓ ⇓ ⇓
Z0 ←− M0 ←− C(M0 ).
One can think of the reduction of diagram 8.5.3 as a reduction of Morita equiv-
alent Lie groupoids. The left hand reduction is that of proper étale Lie groupoids
while the reduction of M and its cone C(M ) corresponds to the holonomy groupoids
294 8. SYMMETRIES AND SASAKIAN STRUCTURES

associated to the respective foliation. For reduction of groupoids see [MW88]. Let
us consider an example of the Sasakian reduction procedure given in [GO01] ap-
plied to the diagram 8.5.3.
Example 8.5.5: We begin by considering a particular case of the fibrations of the
top row of diagram 8.5.3, namely CP3 ←− S 7 ←− (C4 \ {0}). Writing C4 = C2 × C2
with coordinates x and y on the C2 factors. Consider the circle action Sk,l 1
of
−ilθ ikθ
Example 7.6.10 given by (x, y) 7→ (e x, e y), where k and l are relatively
prime integers. The moment map µ : (C4 \ {0})−−→R corresponding to this action
is
µ = −l|x|2 + k|y|2
and the zero set is the cone C(S 3 × S 3 ). Restricting to S 7 (or equivalently quoti-
enting by R+ ) gives µ−1 3 3 −1 1
S 7 (0) = S × S and the quotient µS 7 (0)/Sk,l is identified
1,1
with the Wang-Ziller manifolds Mk,l which as discussed in Example 7.6.10 are dif-
feomorphic to S 2 × S 3 for all relatively prime integers (k, l). Moreover, the quotient
of µ−1
S 7 (0) by the 2-torus defined by (x, y) 7→ (e
iθ1
x, eiθ2 y) is CP1 × CP1 , so our
reduction diagram 8.5.3 becomes
CP3 ←− S7 ←− C4 \ {0}
(8.5.4) ⇓ ⇓ ⇓
1,1 1,1
CP1 × CP1 ←− Mk,l ←− C(Mk,l ).
1,1
Here the manifolds Mk,l are inequivalent toric contact manifolds for each pair of
relatively prime integers (k, l). This shows that S 2 × S 3 admits a countably infinite
number of inequivalent toric contact structures. In fact we shall see that there are
many more toric contact structures on S 2 × S 3 .
1,1
Exercise 8.9: Construct the polyhedral cones of Mk,l and the Delzant polytopes
2 2
of the toric symplectic manifolds (S × S , ωk,l ), and discuss their relationship.
In [BG00a] the authors proved
Theorem 8.5.6: Let (M, D, T, Υ) be a compact toric contact manifold of Reeb type
with a fixed quasi-regular contact form η. Then (M, D, T, Υ) is isomorphic to the
reduction of a sphere S 2N −1 with its standard contact structure and with a fixed
1-form ηa by a subtorus of T.

Proof. Now M is compact of dimension, say, 2n + 1, and since η is quasi-


regular, the space of leaves Z of the characteristic foliation F is a compact sym-
plectic orbifold of dimension 2n. Furthermore, since M is toric, so is Z, that is,
there is an n dimensional torus Tn preserving the symplectic structure ω on Z.
2
Furthermore, by Theorem 7.1.3 ω represents an integral class in Horb (Z, Z) and M
π
1
is the total space of the principal S V-bundle M −−→ Z whose first Chern class is
represented by dη = π ∗ ω.
Now by the Lerman and Tolman Theorem 8.4.18 (Z, ω) is isomorphic to the
symplectic reduction (CN , ω0 ) with the standard symplectic structure by a torus
TN −n of dimension N − n. If µN −n : CN −−−→t∗N −n denotes the moment map for the
TN −n action, then (Z, ω) is isomorphic to (µ−1 N −n (λ)/T
N −n
, ω̂), where λ is a regular
value of µN −n and ω̂ is the unique symplectic 2-form induced by reduction. Let φ̂
denote the above isomorphism. It follows that the cohomology class of ω̂ = (φ̂−1 )∗ ω
2
is integral in Horb (µ−1
N −n (λ)/T
N −n
, Z). By the orbifold version of the Boothby-Wang
8.5. CONTACT AND SASAKIAN REDUCTION 295

theorem there is an S 1 V-bundle π : P −−→µ−1N −n (λ)/T


N −n
, a connection form η̂ on
∗ 1
P such that dη̂ = π̂ ω̂. Thus there is an S -equivariant V-bundle map

φ
M /P

π π̂

² φ̂
²
Z / µ−1 (λ)/TN −n
N −n

such that φ∗ dη̂ = dη. Thus, φ∗ η̂ and η differ by a closed 1-form. But the space
of closed 1-forms on M is path-connected, so one can find a one parameter family
of connections having the same curvature that connect φ∗ η̂ to η. So by Gray’s
Stability Theorem 6.1.32 φ∗ η̂ and η define the same contact structure. Thus, we
can choose φ∗ η̂ = η. Moreover, by equivariance the characteristic vector ς of the
contact manifold (P, η̂) lies in the Lie algebra tN −n , so we can split off the circle
that it generates and write TN −n = Sς1 × TN −n−1 , where TN −n−1 is an (N − n − 1)-
dimensional torus. It follows that P = µ−1 N −n (λ)/T
N −n−1
. Hereafter, we identify
−1 N −n−1 −1
(M, η) with (µN −n (λ)/T , η̂). Now µN −n (λ) is a torus bundle over a compact
manifold, so it is a compact manifold which by construction is an intersection of
N − n real quadrics in CN . ItPfollows that there is a component of the moment map
µN −n which takes the form i ai |zi |2 with ai > 0 for all i. Let a denoteP the vector in
RN whose ith component is ai , and consider the ellipsoid Σa = { i ai |zi |2 = 1} ∼ =
S 2N −1 . Then there is a TN −n -moment map νa : Σa −−→t∗N −n such that νa−1 (0) =
P
µ−1
N −n (λ). Furthermore, letting η0 = i (xi dyi − yi dxi ) we see that dη0 = ω0 on C
N

and that η0 |Σa = ηa |Σa , where ηa is the deformed 1-form of Example 8.4.39. Thus,
letting p : µ−1
N −n (λ)−−→M denote the natural submersion, and ι : µ−1 N −n (λ)−−→Σa
the natural inclusion, we see that p∗ η = ι∗ ηa , so (M, η) is obtained from (Σa , ηa )
by contact reduction. ¤

Now Corollary 8.4.19 says that every symplectic toric orbifold possesses an
invariant complex structure which is compatible with its symplectic form giving it
a compatible Kähler structure. Combining this with Theorem 8.5.6 gives
Corollary 8.5.7: Every compact toric contact manifold of Reeb type admits a
compatible invariant Sasakian structure.
We now give a brief discussion of a construction due to Lerman [Ler04a] that
allows one to construct K-contact structures on the total space of a fibre bundle
whose fibres are K-contact. This is a generalization of the join construction of
Section 7.6.2 as well as a certain fibre join construction due to Yamazaki [Yam99].
We then apply Lerman’s construction to toric Sasakian geometry. Actually one
can work within the pure contact setting, and it is the contact analog of symplectic
fibre bundles described in [GLS96]. Here is the definition.
π
Definition 8.5.8: A fibre bundle F −−→M −−→ B is called a contact fibre bundle
if
(i) F is a co-oriented contact manifold with contact bundle D.
296 8. SYMMETRIES AND SASAKIAN STRUCTURES

(ii) There exist an open cover {Ui }i∈I of B and local trivializations
φi : π −1 (Ui )−−→Ui ×F such that for every point p ∈ Ui ∩Uj the transition
functions φj ◦ φ−1i |{p}×F are elements of Con(F, D) .
+

We need the notion of fatness of a bundle due to Weinstein [Wei80]. Let α


be a connection 1-form in a principal bundle P (M, G) with Lie group G, and let
Ω = Dα denote its curvature 2-form. Let S ⊂ g∗ be any subset in the dual g∗ of
the Lie algebra g of G. We say that the connection α is fat on S if the bilinear map

(8.5.5) µ ◦ Ω : HP × HP : −−−−→R

is non-degenerate for all µ ∈ S.


We first recall the main lines of the construction, not in full generality, but
adapted to our needs. Let π : P → B be a principal G-bundle endowed with a
connection A (we don’t distinguish between the connection and its 1-form). Let F
be a K-contact manifold, with fixed contact form ηF and Reeb field ξF . Suppose
G ⊂ Aut(F, η), i.e., it acts (from the left) on F by strong contactomorphisms and
denote by Ψ : F → g∗ the associated momentum map. Then Lerman [Ler04a]
proves
Theorem 8.5.9: In the above setting, if the connection A is fat at all the points
of the image of the momentum map Ψ, then the total space M of the associated
bundle P ×G F admits a K-contact structure.
We recall that fatness at a point ζ ∈ g∗ means that the bilinear map ζ ◦ ΩA :
HA ×HA → R is non-degenerate, where HA and ΩA are, respectively, the horizontal
distribution and the curvature form defined by the connection A. In particular, if
G is a torus, the bundle π : P → B is identified, up to a gauge transformation,
by a connection form A such that dA = π ∗ ω with [ω] ∈ H 2 (B, Z). Then, if ω is a
symplectic form, A is certainly fat.
Now combining Theorem 8.5.9 with Corollary 8.5.7 one can prove [BGO07]
Theorem 8.5.10: Let F 2n+1 be a compact toric contact manifold of Reeb type and
with torus T n+1 ⊂ Con(F, η). Let π : P → B be a principal T n+1 bundle over a
toric compact symplectic manifold B. Then P ×T n+1 F is a toric Sasakian manifold.
It is easy to see [BGO07] how this generalizes the join construction. Suppose
that G = S 1 , F is a regular Sasakian manifold over a polarized Kähler base man-
ifold BF , and B is a polarized Kähler manifold, then taking the product complex
structure on B × BF reproduces the join construction.
In the paper [Gui94a] and the book [Gui94b], Guillemin constructs ‘canoni-
cal’ Kähler metrics on compact toric symplectic manifolds from the combinatorial
data of the Delzant polytopes. See also [CDG02, BG04] for further development.
Furthermore, this construction has been generalized to compact toric symplectic
orbifolds by Abreu [Abr01] using [LT97]. This gives rise to ‘special’ Sasakian
metrics on compact toric contact manifolds of Reeb type. However, such special
structures are far from unique. First, one can deform the characteristic foliation
within a fixed contact and CR structure as described in Section 8.2.3. When apply-
ing these deformations to toric geometry we also need to preserve the toric structure.
Thus, this type of deformation is performed by perturbing the contact 1-form as in
equation (8.2.3) in a Tn+1 invariant way, that is, the function f is invariant under
Tn+1 . We also take the vector field ρ in equation (8.2.3) to lie in the image A∗ tn+1 .
Such deformations are equivalent to deforming the characteristic hyperplane in t∗ .
8.5. CONTACT AND SASAKIAN REDUCTION 297

Second, one can add to the contact form η a Tn+1 -invariant basic 1-form ζ. This
gives another toric Sasakian structure. In the quasi-regular case this is equivalent
¯ to the symplectic form ω on the orbifold base, where f is a smooth
to adding i∂ ∂f
n
T -invariant function.
We now give a brief review of the Guillemin-Abreu construction. Let ∆ ⊂ t∗
be an lt-polytope and let ∆◦ denote its interior. Now ∆ can be represented by
inequalities of the form
(8.5.6) hx, ui i ≥ λi , i = 1, . . . , d = # of facets of ∆ ,
where ui is the inward normal to the ith facet and is a primitive element in the
lattice ZT . We define affine functions li : t∗ −−→R by
(8.5.7) li (x) = hx, mi ui i − mi λi ,
where mi is the label of the ith facet Fi , so li (x) > 0 if and only if x ∈ ∆◦ and
li (x) = 0 if and only if x ∈ Fi . Now according to Theorem 8.4.18 to each lt-
polytope ∆ there corresponds a compact toric symplectic orbifold (Z, ω, T, µ) such
that µ(Z) = ∆. The idea is to construct the Kähler structure on the orbifold Z
from a potential function on the polytope ∆. Define the Guillemin potential p∆ by
d
1X
(8.5.8) p∆ = li (x) log li (x)
2 i=1

which is a smooth function on the interior ∆◦ . The dense open subset µ−1 (∆◦ ) ⊂ Z
is diffeomorphic to ∆◦ × Tn on which we have the so-called action-angle coordinates
(x1 , . . . , xn ; θ1 , . . . , θn ). The potential p∆ gives rise to Guillemin metrics that are
constructed from the Hessian of a function G. Define
∂2G
gij = .
∂xi ∂xj
Then the Guillemin metric is defined by
X¡ ¢
(8.5.9) g∆ = gij dxi dxj + g ij dθi dθj ,

where g ij are the components of the inverse of gij and G = p∆ . This Kähler
structure coincides up to isometry with the Kähler structure on the dense open set
µ−1 (∆◦ ) induced from the reduction by Corollary 8.4.19. This can be checked by
choosing a local slice of the Tn action [Abr01]. Thus, one obtains a smooth Kähler
orbifold metric on all of Z = µ−1 (∆). Furthermore, one has the freedom to add to
p∆ the Hessian of a smooth function h on the whole polytope ∆ giving more general
toric Kähler orbifold metrics. In fact, while the canonical Guillemin metric enjoys a
certain combinatorial simplicity, it does not enjoy other important properties. For
example, it is not generally extremal3. See [Abr01] where the extremal metrics for
weighted projective spaces CP(w) are computed in terms of the combinatorial data.
We refer to the metric of equation (8.5.9) with G = p∆ as the canonical Guillemin
metric. This construction gives
Theorem 8.5.11: Let (M, D, T, Υ) be a compact toric contact manifold of Reeb
type with a fixed quasi-regular contact form η, and let π : M −−→Z denote the

3Nevertheless, a potential similar to the Guillemin potential had been introduced earlier by
Simanca [Sim91, Sim92] to construct extremal Kähler metrics.
298 8. SYMMETRIES AND SASAKIAN STRUCTURES

induced orbifold bundle over the compact Kähler orbifold (Z, ω) satisfying dη = π ∗ ω.
Then the Riemannian metric
g = π ∗ g∆ + η ⊗ η ,
where g∆ is the canonical Guillemin metric on the orbifold Z, is a toric Sasakian
metric.
We refer to the metric g of Theorem 8.5.11 as the Guillemin-Sasakian metric.
Of course, as with the canonical Guillemin metric we can add a Tn+1 invariant basic
1-form to η to obtain other toric Sasakian metrics. Unlike the Guillemin metric on
Z, the toric Guillemin-Sasakian metrics in Theorem 8.5.11 are unique only up to a
gauge transformation in the choice of η satisfying dη = π ∗ ω.
CHAPTER 9

Links as Sasakian Manifolds

Links arose early on in the study of knots in three dimensions (cf. [Rol76]),
but it is the higher-dimensional case that is of interest to us. We are interested
in links arising from isolated hypersurface singularities, although we also treat
briefly the case of complete intersection singularities. So good references here are
[Mil68, Dim92]. As we shall see in Section 9.2 links of isolated hypersurface sin-
gularities defined by weighted homogeneous polynomials admit Sasakian structures
in a natural way [Abe77, Tak78]. In fact it was realized in the mid seventies that
both Brieskorn manifolds as well as the more general links of isolated hypersurface
singularities arising from weighted homogeneous polynomials admitted both contact
[AE75, LM76, SH76] and almost contact [Abe76, ST76] structures. However,
it was Abe [Abe77] who first developed the connection with Sasakian geometry
proceeded by Takahashi [Tak78] who viewed the links as invariant submanifolds of
the weighted spheres of Example 7.1.5. See Proposition 9.2.2 below.

9.1. Preliminaries
We begin with the main definitions following Milnor’s beautiful classical text
[Mil68]. Let (z0 , . . . , zn ) denote the standard complex coordinates in Cn+1 , and
let f ∈ C[z0 , . . . , zn ] be a polynomial that is non-constant with respect to each of
∂f
the variables zi , i.e., the partial derivative ∂z i
is not everywhere zero for any i =
0, . . . , n. We are interested in the zero locus Vf of f which is a complex hypersurface
in Cn+1 . Recall that a point z ∈ Vf is called a singular point or critical point of Vf
if its differential df vanishes at z. Otherwise, z is called a regular or a simple point.
A critical point z is said to be isolated if there is a neighborhood U of z such that
all the points of U \ {z} are regular.
Definition 9.1.1: Let z0 be any point z0 ∈ Vf . We consider a sphere S²2n+1 (z0 ) of
radius ² centered about z0 . Then the intersection Lf (²) = Vf ∩ S²2n+1 (z0 ) is called
a link of Vf .
If z0 is a regular point it follows from the well-known Morse Lemma [Mil63]
that for ² sufficiently small, Lf (²) is diffeomorphic to the standard “unknotted big
sphere” inside the round sphere S²2n+1 . For that reason we are interested in the case
when z0 is singular. It should be clear that if z0 is an isolated singular point then for
² sufficiently small Lf (²) is independent of ² in the sense that Lf (²) and Lf (²0 ) are
diffeomorphic for 0 < ²0 < ². Thus, we shall write Lf (²) as Lf . When z0 is singular
Lf is often referred to as the link of a hypersurface singularity. We shall assume
that all variables z0 , . . . , zn occur in f for if not we would have a polynomial in one
less variable. We also assume that there is no constant term so that the origin lies
in Vf , and we want the origin to be a singular point of f which means that there
are no linear terms in f. Consider df : Cn+1 −−→Cn+1 the differential of f, and let
299
300 9. LINKS AS SASAKIAN MANIFOLDS

θ(f ) be the Jacobi ideal in C[z0 , . . . , zn ] generated by df. The multiplicity µf (z0 )
of f at an isolated singular point z0 ∈ Vf is defined to be the degree of the Gauss
map G : S²2n+1 −−→S 2n+1 defined by
df (z)
(9.1.1) G(z) = .
||df (z)||
If z0 is a regular point then µf (z0 ) is defined to be 0. Let z0 be the origin in Cn+1
and define the Milnor algebra as the quotient algebra
C[z0 , . . . , zn ]
(9.1.2) M (f ) = .
θ(f )
Then it can be shown [Mil68] that
(9.1.3) µf = µf (0) = dimC M (f ) .
µf is called the Milnor number of the isolated hypersurface singularity. We shall
see later how M (f ) encodes much of the geometry of the link Lf .

9.2. Sasakian Structures and Weighted Homogeneous Polynomials


In this section we consider the links of a hypersurface singularity defined by
the weighted homogeneous polynomials of Definition 4.6.1, and show that such
links have a natural Sasakian structure associated with them. According to the
Structure Theorem 7.1.10, in order that a link Lf admits a Sasakian structure it
is necessary that it also admits a locally free circle action. Thus, at least locally
the hypersurface Vf admits a locally free C∗ action. This leads us naturally to the
weighted homogeneous polynomials of Definition 4.6.1. Consider the affine space
Cn+1 together with the weighted C∗ (w)-action1 as in Definition 4.5.1.
The zero set Vf of a non-degenerate weighted homogeneous polynomial f, is
just the weighted affine cone Cf of Definition 4.6.3. If the origin is an isolated
singularity in Cf , then it is the only singularity of Cf . In this case we can take
² = 1 so that S²2n+1 (z0 ) becomes the unit sphere S 2n+1 centered at the origin with
its weighted Sasakian structure as in Example 7.1.12. In this case our link
(9.2.1) Lf = Cf ∩ S 2n+1 .
is a smooth manifold of dimension 2n − 1 which by the Milnor Fibration Theorem
9.3.1 is (n − 2)-connected. Compact (n − 2)-connected manifolds M of dimension
2n − 1 have been called highly connected by Wall [Wal67b] since their only non-
vanishing homology groups, Hi (M, Z) occur for i = 0, n−1, n, 2n−1. Wall [Wal67b]
and his students [Bar65, Wil72] have given a rough classification of these highly
connected manifolds which will be of much interest to us later. In particular, for
n = 3 there is a complete classification [Bar65] which we discuss in detail in
Chapter 10.
When the origin is an isolated singularity of a weighted homogeneous polyno-
mial, the link Lf is a smooth manifold. This leads to

1Our convention of calling the positive integers w weights is not entirely standard. In the
j
literature on links it is common to refer to the rational numbers d/wi as weights (cf. [MO70,
OW71b]). Our convention is clearly more natural in this book, since it relates to the C∗ (w)-action
defined earlier. However, rational weights will be introduced later in Definition 9.3.8.
9.2. SASAKIAN STRUCTURES AND WEIGHTED HOMOGENEOUS POLYNOMIALS 301

Definition 9.2.1: We denote by W HP2n−1 the set of diffeomorphism classes of


smooth closed (2n − 1)-dimensional manifolds Lf that can be realized as the link
of an isolated hypersurface singularity of a weighted homogeneous polynomial f in
n + 1 complex variables, and we let W HP2n−1 (d) denote the subset of W HP2n−1
that are realized by weighted homogeneous polynomials of degree d.
We exclude the case d = 1 as this does not have a singularity. So we have the
equality
[
(9.2.2) W HP2n−1 = W HP2n−1 (d) .
d≥2

There may be polynomials of degree d with different weight vectors which give rise to
different elements of W HP2n−1 (d), and there may be smooth manifolds represented
by links of different degrees. When we fix a weight vector w and a degree d we are
interested in all possible monomials such that Theorem 4.6.16 holds.
The circle subgroup S 1 ⊂ C∗ acts on the affine space Cn+1 by taking λ ∈
1
S in Definition 4.5.1. Recall from Examples 7.1.5, and 7.1.12 that the weighted
Sasakian structure (ξw , ηw , Φw , gw ) on the unit sphere S 2n+1 comes from the circle
action with the negative orientation generated by the vector field ξw of equation
(7.1.5). Thus, our convention here is λ = e−iθ with θ ∈ [0, 2π]. We now have
[BG01b, Tak78]
Proposition 9.2.2: Let Lf be the link of a non-degenerate weighted homogeneous
polynomial of weight w and degree d. The Sasakian structure Sw = (ξw , ηw , Φw , gw )
on S 2n+1 induces by restriction a Sasakian structure, also denoted by Sw , on the
link Lf .
Proof. According to Theorem 7.6.2 we only need to check that for all points
p ∈ Lf the two conditions hold:
(i) ξw is tangent to Lf .
(ii) Φw Tp Lf ⊂ Tp Lf .
To see that (i) holds, we differentiate the defining equation in Definition 4.6.1 with
respect to λ,
df
ξw f = λ = dλd f .

So ξw f vanishes at points of Cf , hence, at points of Lf . This proves (i).
To prove (ii) we first notice that since Cf is defined by a holomorphic equation
we have IT Cf ⊂ T Cf , where I is the standard almost complex structure on Cn+1 .
Moreover, we have
Φw |DLf = Φ|DLf = I|DLf ,
where DLf = ker ηLf and to avoid confusion we let ηLf denote the restriction of
ηw to Lf . Now at a point p ∈ Lf we have (DLf )p = Dp ∩ Tp Cf . Since Φw ξw = 0,
we need only check that Φw DLf ⊂ DLf . But Φw leaves D invariant, and equals I
on DLf which leaves T Cf invariant, so it leaves DLf invariant. ¤
2n+1
Recall from Example 7.1.12 the unit sphere Sw with the weighted Sasakian
structure Sw = (ξw , ηw , Φw , gw ), but now with w ∈ (Z+ )n+1 ⊂ (R+ )n+1 . Thus,
2n+1
Proposition 9.2.2 says that the natural inclusion ι : Lf −−→Sw is a Sasakian em-
bedding. It is also clear since the vector w lies in the positive integer lattice in Rn+1
2n+1
that the Sasakian structures Sw on both Lf and Sw are quasi-regular. Thus,
by Theorem 7.5.1 the spaces of leaves in both cases are compact polarized Hodge
302 9. LINKS AS SASAKIAN MANIFOLDS

orbifolds. We denote the space of leaves of Lf by Zf , and almost by definition the


2n+1
space of leaves of Sw is the weighted projective space CP(w) = CP(w0 , . . . , wn ).
Thus, we have a commutative diagram
2n+1
Lf −−−−→ Sw
 
 
(9.2.3) π 
y y
Zf −−−−→ CP(w) ,
where the horizontal arrows are Sasakian and Kählerian embeddings, respectively,
and the vertical arrows are principal S 1 -orbibundles and orbifold Riemannian sub-
mersions. Of course, the complex orbifold Zf can also be represented as the quo-
tient (Cf \ {0})/C∗w . Note that a choice of f determines (w; d). However, given
(w; d) typically there is no quasi-smooth polynomial f matching such data, and
if there is one, there can be a finite-dimensional family as indicated by Theorem
5.5.7. Moreover, for each fixed values of the parameters, there is an a-deformation
class of Sasakian structures adapted to the link. We often think of all quasi-smooth
members of such a family. It is convenient to fix some notation.
Definition 9.2.3: We denote by L∗ (w; d) any quasi-smooth link Lf (or the whole
family of such links) given by a weighted homogeneous polynomial f with weight
vector w and degree d. We let F(Fξ (w; d)) denote the a-deformation class of the
family.
We use the star to distinguish such general links from the Brieskorn-Pham type
links discussed later. Alternatively, one could think of it as the quasi-smoothness
restriction. We have the following immediate consequence of the orbifold adjunction
formula equation (4.6.3):
Proposition 9.2.4: Let Lf ∈ L∗ (w; d) and let Sw denote its induced Sasakian
structure. Then Sw is quasi-regular, satisfies c1 (Fξw ) = a[dηw ]B , and is
(i) positive if and only if |w| − d > 0 ,
(ii) null if and only if |w| − d = 0 ,
(iii) negative if and only if |w| − d < 0 .
In particular, the induced Sasakian structure of all such links Lf is either anti-
canonical, canonical, or null.

9.3. The Milnor Fibration and the Topology of Links


The topology of links of isolated hypersurface singularities is best understood
using the well-known Fibration Theorem of Milnor [Mil68] which we now discuss.
Theorem 9.3.1: Let z0 ∈ Vf be any point of the complex hypersurface Vf , and let
² > 0 be sufficiently small. Then the map φ : S²2n+1 (z0 ) − Lf −−→S 1 defined by
f (z)
φ(z) =
|f (z)|
is the projection map of a smooth fibre bundle with a smooth parallelizable fibre.
If z0 is an isolated singular point of f, then each fibre F has the homotopy type
of a bouquet of n-spheres S n ∨ · · · ∨ S n , and is homotopy equivalent to its closure
F̄ which is a compact manifold with boundary where the common boundary ∂ F̄ is
precisely Lf . Furthermore, Lf is a smooth (n − 2)-connected manifold of dimension
2n − 1.
9.3. THE MILNOR FIBRATION AND THE TOPOLOGY OF LINKS 303

The proof of this theorem uses Morse Theory [Mil63] and can be found in
[Mil68]. The number of S n ’s in the bouquet described in Theorem 9.3.1 is an
important invariant of the link and it is the Milnor number µ. If the point z0 is
regular, it follows from the Morse Lemma [Mil63] that µ = 0, and the link Lf is
diffeomorphic to the big (2n−1)-sphere in S 2n+1 . In view of Theorem 9.3.1 we have
Theorem 9.3.2: Let M ∈ W HP2n−1 . Then
(i) M is highly connected, that is, it is (n − 2)-connected.
(ii) M is the boundary of a compact (n − 1)-connected parallelizable manifold
V of dimension 2n.
(iii) If n is even the Betti number bn−1 (M ) is even.
(iv) If n ≥ 3 then M is a spin manifold.
Proof. Parts (i) and (ii) follow from the Milnor Fibration Theorem 9.3.1. By
Proposition 9.2.2 M admits a Sasakian structure, so its odd Betti numbers are even
up to the middle dimension by Theorem 7.4.11. This proves (iii), and (iv) follows
from Proposition 9.2.4 and Theorem 7.5.27 as soon as n ≥ 3. ¤
Using the Wang sequence of the Milnor fibration together with Alexander-
Poincaré duality gives the exact sequence [Mil68]
1l−h∗
(9.3.1) 0−−→Hn (Lf , Z)−−→Hn (F, Z) −−→ Hn (F, Z)−−→Hn−1 (Lf , Z)−−→0 ,
where h∗ , known as the monodromy map (or characteristic map), comes from the
action of a generator of π1 (S 1 ) on the homology of the fibre. From this we see that
Hn (Lf , Z) = ker(1l − h∗ ) is a free Abelian group, and Hn−1 (Lf , Z) = coker(1l − h∗ )
which in general has torsion, but whose free part equals ker(1l−h∗ ). So the topology
of Lf is encoded in the monodromy map h∗ . The characteristic polynomial ∆(t) of
the monodromy map is another important link invariant. It is a generalization of
the Alexander polynomial of knot theory, and has also been called the Alexander
polynomial [Hir71, HZ74] of Lf . We can get some immediate information from
the exact sequence (9.3.1). If ∆(1) 6= 0 then 1l − h∗ is non-singular which implies
Hn (Lf , Z) = 0, that is Lf is a rational homology sphere. Furthermore, if |∆(1)| = 1,
then 1l − h∗ is an isomorphism, so Lf is a homology sphere, and if n > 2 it must
be a homotopy sphere by the Hurewicz Theorem. Thus, we have arrived at:
Proposition 9.3.3: The following hold:
(i) Lf is a rational homology sphere if and only if ∆(1) 6= 0.
(ii) Lf is a homology sphere if and only if |∆(1)| = 1.
(iii) If Lf is a rational homology sphere, then the order of Hn−1 (Lf , Z) equals
|∆(1)|.
We now briefly consider the case n = 2. Let us take z0 to be the origin in
C3 . If the origin is at worst an isolated singularity of f , all that we know is that
Lf is a smooth connected 3-manifold. But the well-known Theorem of Mumford
[Mum61, Dim92] implies more.
Theorem 9.3.4: If the origin is at worst a normal singularity of f then Lf is
simply connected if and only if Lf is diffeomorphic to S 3 .
In fact in the simply connected case the origin is a regular point of f. Thus,
hypersurface singularities can say nothing about the Poincaré conjecture. Never-
theless, they play an important role in understanding homology spheres as we shall
see in Section 9.5.
304 9. LINKS AS SASAKIAN MANIFOLDS

9.3.1. Computing the Alexander Polynomial of Weighted Homoge-


neous Polynomials. In the case that f is a weighted homogeneous polynomial
(see Section 9.2 below), there is a well-known algorithm due to Milnor and Orlik
[MO70] for computing the free part of Hn−1 (Lf , Z) in terms of the characteris-
tic polynomial ∆(t) = det(t1l − h∗ ), namely the Betti number bn (Lf ) = bn−1 (Lf )
equals the number of factors of (t − 1) in ∆(t).
Now we shall assume that f is a quasi-smooth weighted homogeneous polyno-
mial. In this case the Milnor number µf can be calculated explicitly in terms of
the degree d and weights w = (w0 , . . . , wn ). This result is due to Milnor and Orlik
[MO70].
Proposition 9.3.5: Let Lf ∈ L∗ (w; d). Then
Yn ³ ´
d
(9.3.2) µ = µ(Lf ) = −1 .
i=0
wi

Proof. Consider the map g : Cn+1 −−→Cn+1 defined by


g(z0 , . . . , zn ) = (z0w0 , . . . , znwn ) .
Then f ◦ g is homogeneous of degree d and the ith component of the differential
df ◦ g is homogeneous of degree d − wi . So the degree of df ◦ g at the origin is
(d − w0 ) · · · (d − wn ). But the degree of df at the origin, which by definition is the
multiplicity µf , is obtained by dividing by the degree of g at the origin, giving the
formula. ¤

Remark 9.3.1: It is remarkable that for isolated hypersurface singularities the


product in formula (9.3.2) is always an integer although the ratios d/wi may not
be integers. This is not necessarily the case if the singularity is not isolated.
The closure F̄ of F is a manifold with boundary that is homotopy equivalent to
F, and whose boundary is precisely the link Lf . The topology of Lf is determined
1
by the monodromy map induced by the Sw -action. In the case of isolated hyper-
surface singularities determined by weighted homogeneous polynomials this has a
particularly nice form.
Proposition 9.3.6: Let f be a weighted homogeneous polynomial of an isolated
hypersurface singularity. Then the Milnor fibre F is diffeomorphic to the non-
singular hypersurface
{z ∈ Cn+1 | f (z) = 1} ,
and the monodromy map h∗ is induced by the map h : Cn+1 −−→Cn+1 defined by
¡ 2πw0 i 2πwn i ¢
h(z0 , . . . , zn ) = e d z0 , . . . , e d zn .
Notice that h is periodic of period d. Thus, the monodromy map h∗ is diago-
nalizable over C with eigenvalues d roots of unity. Identifying F with f −1 (1) we
let hj : F −−→F denote the j-fold iterate of h, and F j denote the fixed point subset
of F under hj . The F j ’s are determined by
© 2πijwl ª
(9.3.3) F j = z ∈ f −1 (1) | e d zl = zl for all l = 0, . . . , n .
By computing the Euler number χj of the fixed point sets F j , Milnor [Mil68]
was able to compute the Alexander polynomial ∆(t) of the link Lf using the Weil
9.3. THE MILNOR FIBRATION AND THE TOPOLOGY OF LINKS 305

zeta function
³X

χ ´
j j
(9.3.4) ζ(t) = exp t .
j=1
j

The zeta function (9.3.4) and the Euler numbers χj are related by integers ri
through
X Y
(9.3.5) χj = lrl , ζ(t) = (1 − tj )−rj .
l|j j|d

Then we have
Theorem 9.3.7: The Alexander polynomial ∆(t) of the link Lf of an isolated
singularity of a weighted homogeneous polynomial f is determined by
n+1 Y
∆(t)(−1) = (t − 1) (tj − 1)−rj .
j|d

See [Mil68] for the proof of this theorem. Milnor and Orlik [MO70] use these
facts to give an algorithm for computing the Alexander polynomial ∆(t) of the
monodromy map in terms of the degree d and weight vector w. The procedure is
this. Associate to any monic polynomial f with roots α1 , . . . , αk ∈ C∗ its divisor
(9.3.6) div f = hα1 i + · · · + hαk i
as an element of the integral ring Z[C∗ ] and let Λn = div(tn − 1). Then the Λ’s
satisfy the multiplication rule
(9.3.7) Λa Λb = gcd(a, b)Λlcm(a,b) .
Definition 9.3.8: Given Lf ∈ L∗ (w; d) we define rational weights
(9.3.8) (d/w0 , . . . , d/wn ) ≡ (u0 /v0 , . . . , un /vn ),
where (u; v) = (u0 , . . . , un ; v0 , . . . , vn ) is given by
d wi
(9.3.9) ui = , vi = .
gcd(d, wi ) gcd(d, wi )
As already remarked the rational weights were introduced in [MO70]. In the
literature on links they are quite often denoted by wi = ui /vi so, once again, we
alert the reader that our weights w are different. We prefer to think of rational
weights as a (u; v)-data on a link Lf ∈ L∗ (w; d). In particular, one can always
translate between the (w; d)-data and the (u; v)-data. In a certain sense they are
‘dual’ to one another. We are now ready for
Theorem 9.3.9: The divisor of the Alexander polynomial ∆(t) is given by
Y³ Λu ´ ¡ X ¢
div ∆ = i
− 1 = (−1)n+1 1 − rj Λj .
i
vi

Proof. From Theorem 9.3.7 we have


¡ X ¢
div ∆ = (−1)n+1 1 − rj Λj .

Since the rj ’s are obtained from the Euler numbers χj of the fixed point sets F j , we
need to express χj in terms of d and the wi ’s, or alternatively in terms of the ui ’s
306 9. LINKS AS SASAKIAN MANIFOLDS

and vi ’s. By equation (9.3.3) F j is given by the vanishing of certain coordinates zi .


Then by permuting the variables zi if necessary we see that if2
(9.3.10) j ≡ 0(ui ) for i ≤ k, j 6≡ 0(ui ) for i > k
then
F j = {z ∈ Cn+1 | f (z0 , . . . , zk , 0, . . . , 0) = 1} .
The polynomial fk (z0 , . . . , zk ) = f (z0 , . . . , zk , 0, . . . , 0) is weighted homogeneous
with an isolated singularity at the origin in Ck+1 . Thus, F j is the Milnor fibre of
fk . So its only non-vanishing reduced homology group is H̃k (F j , Z) ≈ Z µj , where
by Proposition 9.3.5
Yk ³ ´
d
µj = −1 .
i=0
wi
Thus, the Euler number of F j is
k ³

k ³
k
Y Y ui ´
χj = 1 + (−1) µj = 1 − 1− =1− 1− .
i=0
wi i=0
vi
Notice from equation (9.3.10) that we can rewrite this equation as
Y³ ui ´
(9.3.11) 1 − χj = 1− .
vi
ui |j

This determines the χj , and hence, the rj in terms of the wi ’s. To prove the second
equality we define for any integer m between 0 and n,
³ X ´
(9.3.12) δ(w0 , . . . , wm ) = (−1)m+1 1 − rj (w0 , . . . , wm )Λj .
Viewing this as an element of the rational group ring Q[C∗ ], we have
Lemma 9.3.10: The identity holds:
δ(w0 , . . . , wp ) = δ(w0 , . . . , wm )δ(wm+1 , . . . , wp ) .
Proof. Set sl = lrl , and let χl , sl be the sequences associated to the weights
w0 , . . . , wm through equations (9.3.5) and (9.3.11). Likewise, let χ0l , s0l be associated
to the weights wm+1 , . . . , wp , and χ00l , s00l be associated to the composite p-tuple
w0 , . . . , wp . Then from equation (9.3.11) we find
χ00j = χj + χ0j − χj χ0j .
P
Substituting χj = i|j si and similar expresses for the primed and double primed
variables gives X X X XX
s00l = sl + s0l − sk s0l .
l|j l|j l|j k|j l|j
Using this equation an induction argument gives
X
(9.3.13) s00j = sj + s0j − sk s0l .
lcm(k,l)=j

Now we write the sj ’s in terms of the rj ’s, and multiply equation (9.3.13) by Λj
and sum over j. This gives
X X X ³X ´³X ´
(9.3.14) rj00 Λj = rj Λj + rj0 Λj − rk Λk rl0 Λl .

2We use the notation n ≡ m(k) or n = m mod k for arithmetics mod k.


9.3. THE MILNOR FIBRATION AND THE TOPOLOGY OF LINKS 307

Here we have used equation (9.3.7). The lemma now follows easily from equation
(9.3.14). ¤
We need one more lemma to finish the proof of the theorem. Namely,
Lemma 9.3.11:
Yn ³ ´
Λ ui
δ(w0 , . . . , wn ) = −1 .
i=0
vi

Proof. The proof is by induction on n. For n = 0 we notice from equation


(9.3.11) that
(
u0
, if u0 |j;
χj (w0 ) = v0
0, otherwise.
Thus, sj satisfies
(
u0
sj (w0 ) = v0 , if j = u0 ;
0, if j =
6 u0 .
This shows that
³ Λ u0 ´ Λ u0
δ(w0 ) = −(1 − rj (w0 )Λj ) = − 1 − = − 1,
v0 v0
which proves the result for n = 0. Now an easy induction using Lemma 9.3.10
proves the result. ¤
This completes the proof of Theorem 9.3.9. ¤
The Milnor-Orlik algorithm computes the numbers rj in terms of (w; d)-data
or, alternatively, in terms of (u; v)-data. In fact, we have
Corollary 9.3.12:
X ui1 · · · uis
div ∆ = (−1)n+1−s Λlcm(ui1 ,...,uis ) ,
vi1 · · · vis lcm(ui1 , . . . , uis )
where the sum is taken over all the 2n+1 subsets {i1 , . . . , is } of {0, . . . , n}.
According to the exact sequence (9.3.1) the (n − 1)st and nth Betti numbers
bn (Lf ) = bn−1 (Lf ) of the link equals the number of factors of (t−1) in the Alexander
polynomial ∆(t), or equivalently the order of vanishing of ∆(t) at t = 1. Thus, using
Corollary 9.3.12 we get
Corollary 9.3.13: Let Lf ∈ L∗ (w; d) and let ∆(t) be its Alexander polynomial.
Then
X ui1 · · · uis
(9.3.15) bn−1 (Lf ) = (−1)n+1−s ,
vi1 · · · vis lcm(ui1 , . . . , uis )
where the sum is taken over all the 2n+1 subsets {i1 , . . . , is } of {0, . . . , n}.
Working with the Alexander polynomial, the characteristic polynomial of the
monodromy map, entails working over C, that is studying the map
(t1l − h∗ ) : Hn (F, C)−−→Hn (F, C) .
An alternative procedure to the one taken above would be to diagonalize (t1l − h∗ ),
that is find unimodular matrices U (t) and V (t) such that
¡ ¢
(9.3.16) U (t)(t1l − h∗ )V (t) = diag m1 (t), . . . , mµ (t) .
308 9. LINKS AS SASAKIAN MANIFOLDS

The minimal polynomial mµ (t) contains the same information as the characteristic
polynomial ∆(t). In order to handle torsion, however, one needs to work with the
integral polynomial ring Z[t] which is not a principal ideal domain. This poses
several difficulties which have still not been entirely overcome. Nevertheless, there
is a conjecture due to Orlik [Orl72a] saying that the matrices U (t) and V (t) can
be taken with values in the ring Z[t]. This implies
Conjecture 9.3.14: Hn−1 (Lf , Z) = Zm1 (1) ⊕ · · · ⊕ Zmµ (1) , where Z1 and Z0 are
the trivial and infinite cyclic groups, respectively.
Given an index set {i1 , . . . , is } we will denote by I all its 2s subsets and by
J all its proper subsets. For each ordered subset {i1 , . . . , is } ⊂ {0, 1, . . . , n}; i1 <
i2 < · · · < is one defines inductively the set of 2s positive integers, starting with
c∅ = gcd(u0 , . . . , un ) :
gcd(u0 , . . . , ûi1 , . . . , ûis , . . . , un )
(9.3.17) ci1 ,...,is = Q .
J cj1 ,...jt
In addition, starting with k∅ = ²n+1 , one defines
(9.3.18)
X uj1 · · · ujt
ki1 ,...,is ≡ ²n−s+1 κi1 ,...,is = ²n−s+1 (−1)s−t ,
vj1 · · · vjt lcm(uj1 , . . . , ujt )
I

where
(
0, if n − s + 1 is even;
²n−s+1 =
1, if n − s + 1 is odd.
For any 1 ≤ j ≤ r = bmax{ki1 ,...,is }c we set

Y
(9.3.19) dj = ci1 ,...,is .
ki1 ,...,is ≥j

The Conjecture 9.3.14 can now be stated more precisely


Conjecture 9.3.15: Hn−1 (Lf , Z)tor = Zd1 ⊕ · · · ⊕ Zdr .
We refer to the above computation of Hn−1 (Lf , Z)tor as the Orlik Algorithm.
Remark 9.3.2: It is useful to think of the Orlik Algorithm as follows: Consider
the indexing set Ω = {0, . . . , n}. For each subset A ⊂ Ω (including the empty
set ∅ and Ω itself) Orlik defines 2|Ω| = 2n+1 pairs of numbers: (cA , κA ). The
numbers {cA }A⊂Ω are all positive integers and they are defined inductively from
c∅ = gcd(u0 , . . . , un ). (We set cΩ = 1 though this does not matter.) The numbers
{κA }A⊂Ω are, in general, fractions. Starting with κΩ = b2 (Lf ) one sets some κA ’s
equal to zero depending on the cardinality |A| defining
³ ´
kA = 1 − (−1)|Ω\A| κA .

The numbers cA are then used to define the orders of the cyclic groups that enter
as torsion and the numbers kA determine the powers in the prime number de-
composition of the torsion. It is equally important to observe that for the torsion
computation only the pairs (cA , kA ) with kA ≥ 1 matter, so in practice starting
with 2n+2 numbers we end up with very few that are needed for the calculation of
the relevant homology group.
9.3. THE MILNOR FIBRATION AND THE TOPOLOGY OF LINKS 309

The formula (9.3.19) gives a simple way of computing the relevant torsion group
in all cases that Orlik’s conjecture is known to be valid. For example, we have the
following [Ran75, OR77]
Proposition 9.3.16: Conjecture 9.3.15 is true in the following cases:
(i) In dimensions 3, i.e., for n = 2.
(ii) For f (z) = z0a0 +· · ·+znan , i.e., the so-called Brieskorn-Pham polynomials.
(iii) For f (z) = z0a0 + z0 z1a1 + · · · + zn−1 znan .
(iv) For certain complete intersections of Brieskorn varieties given by, what
Randell calls, generalized Brieskorn polynomials [Ran75].
Exercise 9.1: [See Section 9.3.2 for the discussion of Brieskorn-Pham links and
notation conventions.] Use Orlik’s algorithm to show that if Lf is the Brieskorn-
Pham link Lf = L(21, 52, 55, 900) then H2 (Lf , Z) = Z24 ⊕ (Z7 )12 ⊕ (Z11 )6 ⊕ (Z13 )8 .
Even more interesting is the BP link Lg = L(78, 170, 627, 2261). Show that Orlik’s
algorithm gives H2 (Lg , Z) = Z1558 ⊕ (Z5 )36 ⊕ (Z7 )2 ⊕ (Z11 )16 ⊕ (Z13 )288 . In spite of
what one might think we did not come up with these links by a random computer
search. Can you explain how we chose them and come up with your own examples?
You will need to consider the Brieskorn graph G(a) for these two links (which will
shortly be defined in Section 9.3.2) to answer this question which eventually leads
to Theorem 10.3.5. Can one get links (BP or otherwise) which have more than the
‘⊕’ of four different cyclic groups to some even powers as torsion in H2 (Lf , Z)?
Exercise 9.2: Suppose that n = 2 and f ∈ L∗ (w0 , w1 , w2 ; d). Show that Or-
lik Algorithm reduces to the genus formula (4.6.5), i.e., we have g(Σ(w;d) ) =
b1 (Lf ) = κ012 . Consider next the n = 3 case and f ∈ L∗ (w0 , w1 , w2 , w3 ; d).
Show that Orlik’s formula (9.3.19) for the torsion in H2 (Lf , Z) depends only on
the numbers {k012 , k013 , k023 , k123 } and {c012 , c013 , c023 , c123 }. Further show that
kijk = 2g(Σ(w̃;d)˜ ), where

(wi , wj , wk ) d
w̃ = , d˜ = ,
gcd(wj , wj , wk ) gcd(wj , wj , wk )
0 ≤ i < j < k ≤ 3, and g(Σ(w̃;d) ˜ ), is calculated with the formula (4.6.5). Could
d˜ be fractional? Show by example that it does not follow that all kijk ’s in Orlik
Algorithm are integers. Explain how (4.6.5) can give a non-integer result.
Open Problem 9.3.1: Prove or disprove the Orlik Conjecture 9.3.15. For n = 3
the conjecture basically follows from Kollár’s Theorem 4.7.14 and Exercise 9.2 (cf.
Theorem 10.3.5).
To end this section we consider branched covers constructed as the link Lg of
the polynomial
(9.3.20) g = z0k + f (z1 , . . . , zm ) ,
where f is a weighted homogeneous polynomial in n variables with an isolated
singularity at the origin. Then Lg is a k-fold branched cover of S 2m+1 branched
over the link Lf . The degree of Lg is dg = lcm(k, df ), and the weight vector is
¡ df k
¢
wg = gcd(k,d ,
f ) gcd(k,df )
wf . We shall always assume that k ≥ 2, since the linear
case k = 1 is a hyperplane in a weighted projective space. The following theorem,
which is Theorem 7.1 of [BGN03c], is a reformulation of results of Savel’ev [Sav79].
It is a generalization of part (i) of the well-known Brieskorn Graph Theorem 9.3.18
below.
310 9. LINKS AS SASAKIAN MANIFOLDS

Theorem 9.3.17: Let f (z1 , . . . , zm ) be a weighted homogeneous polynomial of de-


gree d and weights w = (w1 , . . . , wm ) in Cm with an isolated singularity at the
origin. Let k ∈ Z+ and consider the link Lg of the equation
g = z0k + f (z1 , . . . , zm ) = 0 .
d ui
Write the numbers wi in irreducible form vi , and suppose that gcd(k, ui ) = 1
(d,kw)
for each i = 1, . . . , m. Then the link Lg has weights gcd(k,d) and degree lcm(k, d).
Furthermore, Lg is a rational homology sphere with the order |Hm−1 (Lg , Z)| =
k bm−2 (Lf ) , where bm−2 (Lf ) is the (m − 2)nd Betti number of Lf .
9.3.2. Links of Brieskorn-Pham Type. Here we discuss in more detail
links of polynomials of Brieskorn-Pham type (9.3.16) (abbreviated BP type) which
are of particular interest in the case of homology spheres. We also discuss when BP
type links admit regular Sasakian structures. In the case of BP polynomials it is
more convenient to describe the links in terms of the Brieskorn exponents written
as the components of vector, viz. a = (a0 , . . . , an ) ∈ Zn+1 + . We then write f (z) =
z0a0 + · · · + znan and denote the associated link by Lf = L(a)3. The relation between
the exponents and the weights is now simply that d = ai wi for all i = 0, . . . , n,
where d is the degree of f. Hence, on a BP link Lf ∈ L∗ (w; d) the exponents a are
the non-trivial part of the (u; v)-data, with u = a and v = (1, . . . , 1).
Let us recall the famous Brieskorn Graph Theorem4 [Bri66, Dim92]. To
the vector a one associates a graph G(a) whose n + 1 vertices are labelled by
a0 , . . . , an . Two vertices ai and aj are connected if and only if gcd(ai , aj ) > 1. Let
Gev (a) ⊂ G(a) denote the connected component of G(a) determined by the even
integers. Note that all even vertices belong to Gev (a), but Gev (a) may contain odd
vertices as well. Then we have
Theorem 9.3.18: The following hold:
(i) The link L(a) is a rational homology sphere if and only if either G(a)
contains at least one isolated point, or Gev (a) has an odd number of
vertices and for any distinct ai , aj ∈ Gev (a), gcd(ai , aj ) = 2.
(ii) The link L(a) is an integral homology sphere if and only if either G(a)
contains at least two isolated points, or G(a) contains one isolated point
and Gev (a) has an odd number of vertices and for any distinct ai , aj ∈
Gev (a), gcd(ai , aj ) = 2.
Proof. The proof of (i) is given in detail in [Dim92], and that of (ii) in
[Bri66]. ¤
In either case a very simple property of the graph G(a) conveys a lot of infor-
mation about the topology of L(a). According to the Proposition 9.2.4 the sign of
the Sasakian structure on any BP link L(a) is determined as follows
Corollary 9.3.19: The link L(a) admits a Sasakian structure that is
P 1
(i) positive if and only if > 1,
P 1 ai
(ii) null if and only if ai = 1 ,

3Note that each L(a) is a member of a family L∗ (w; d), where the pair (w; d) is uniquely
determined by a. That is why we use the asterisk to avoid the confusion.
4The fascinating history of this theorem has recently come to light in an interesting historical
account by Brieskorn [Bri00]. The graph theorem for homotopy spheres was conjectured by Milnor
in a letter (published in [Bri00]) to John Nash who showed the letter to Brieskorn.
9.3. THE MILNOR FIBRATION AND THE TOPOLOGY OF LINKS 311

P 1
(iii) negative if and only if ai < 1.
In particular, BP links provide a rich source of all three types of Sasakian
geometry. Let us consider several important examples of BP links.
Example 9.3.20: Let a = (a0 , . . . , an ) = (2, . . . , 2). Here we have the degree d of
f (z) is 2, and weights are all 1, i.e., w0 = · · · = wn = 1. Thus, by Proposition 9.3.5
the Milnor number µ = 1. According to Theorem 9.3.18 L(a) is a rational homology
sphere if and only if n is even. Let us compute the Alexander polynomial ∆(t) for
this link from the Milnor-Orlik procedure described above. First we note that
(Λ2 − 1)2 = (2 − 2)Λ2 + 1 = 1 .
Thus, we have (
Λ2 − 1, if n is even;
div ∆ =
1, if n is odd.
So the Alexander polynomial is
(
t + 1, if n is even;
∆(t) =
t − 1, if n is odd.
This gives (
Z2 , if n is even;
Hn−1 (L(a), Z) =
Z, if n is odd.
In this case the link L(a) = SO(n + 1)/SO(n − 1) is the Stiefel manifold
V2 (Rn+1 ) of oriented 2-frames in Rn+1 . It is a homogeneous space and it carries
SO(n + 1)-homogeneous Sasaki-Einstein structure discussed in Theorem 11.4.1.

Example 9.3.21: Consider a = (a0 , . . . , an ) = (k, 2, . . . , 2) with n ≥ 2. The cor-


responding BP link L(a) is a k-fold cyclic branched cover of S 2n−1 branched over
the link of the quadric z12 + · · · + zn2 . Now we have µ = k − 1. The divisor of the
Alexander polynomial is


Λk − 1, if n is even;
div ∆ = Λ2k − Λk − Λ2 + 1, if n and k are odd;


Λk − Λ2 + 1, if n is odd and k is even.
Hence,

k−1

t + tk−2 + · · · + t + 1, if n is even;
k−1
∆(t) = t − tk−2 + · · · − t + 1, if n and k are odd;


(t − 1)(tk−2 + tk−4 + · · · + t2 + 1), if n is odd and k is even.
An additional calculation of the torsion group with 9.3.15 yields the following an-
swer 

Zk , if n is even;
Hn−1 (L(a), Z) = 0, if n and k are odd;


Z, if n is odd and k is even.
In the first case L(a) is a rational homology sphere whose homotopy type changes
with k. The second case is more interesting. Here L(a) is always a homology and,
hence, a homotopy sphere. But as we shall soon see the smooth structure does
depend on k. The third case is perhaps the most interesting as both the topology
312 9. LINKS AS SASAKIAN MANIFOLDS

and the smooth structure depend on k. We shall discuss the differential topology
of these examples at the end of Section 9.4.
Another case of interest for Brieskorn-Pham links is when the weights are pair-
wise relatively prime. There are many such examples. For instance, for any rela-
tively prime pair of positive integers p, q one can take a = (p, q, pq, . . . , pq). Then
the weight vector is w = (q, p, 1, . . . , 1). Clearly, Zf for this BP link is a smooth
manifold. More generally, we have
Proposition 9.3.22: Let L(a) be a BP link with pairwise relatively prime weights
w = (w0 , . . . , wn ). Then L(a) is the total space of a circle bundle over a smooth
projective algebraic variety Za and the induced Sasakian structure is regular.
We end this section with example of a construction of “locally Sasakian” struc-
tures.
Example 9.3.23: Let L(a) be a link of a BP polynomial f (z) = z0a0 + · · · + znan .
Consider the involution on Cn+1 defined by complex conjugation z 7→ z̄. This
involution restricts to an involution σ on L(a) which sends the Sasakian structure
S defined by Proposition 9.2.2 to its conjugate S c . Moreover, if all the ai ’s are
even then since all the coefficients of f (z) are positive, the involution acts freely
on L(a). So the quotient L(a)/{σ} is a manifold which is not Sasakian, but it is
“locally Sasakian”. Notice that L(a)/{σ} is orientable if and only if n is even. The
simplest examples are the k-fold branched covers of quadrics of Example 9.3.20
with k even. In particular, considering k = 2 we see that L(a) is just the Stiefel
manifold V2 (Rn+1 ) with its homogeneous Sasaki-Einstein metric. So V2 (Rn+1 )/{σ}
admits an Einstein metric that is locally Sasakian, but not Sasakian.

9.4. The Differential Topology of Links


The Milnor Fibration Theorem 9.3.1 implies that links of isolated hypersurface
singularities are highly connected manifolds which are the boundaries of a paral-
lelizable manifold. Hence, they are (2n − 1)-dimensional closed oriented manifolds
whose only non-vanishing homology groups occur in dimensions 0, n − 1, n, 2n − 1.
The most obvious examples of such manifolds are the homotopy spheres, so we
shall begin there. In 1956 Milnor [Mil56b] created the field of differential topol-
ogy by constructing exotic differential structures on S 7 . Later the seminal work of
Kervaire and Milnor [KM63] (see also [Hir67]) showed that associated with each
sphere S n with n ≥ 5 there is an Abelian group Θn consisting of equivalence classes
of oriented homotopy spheres S n that are equivalent under oriented h-cobordism
which by Smale’s h-cobordism theorem [Sma62b, Mil65] implies equivalence un-
der oriented diffeomorphism. More explicitly Kervaire and Milnor proved
Theorem 9.4.1: For n ≥ 5 the set Θn of equivalence classes of oriented homotopy
spheres of dimension n-form a semi-group under the connected sum operation whose
Grothendieck group completion forms a finite Abelian group, also denoted Θn .
Kervaire and Milnor also defined the subgroup bPn+1 of Θn which is of great
importance to us.
Theorem/Definition 9.4.2: The subset bPn+1 of Θn consisting of those homo-
topy spheres which bound parallelizable manifolds form a subgroup of Θn of finite
index, and the short exact sequence
0−−→bPn+1 −−→Θn −−→Θn /bPn+1 −−→0
9.4. THE DIFFERENTIAL TOPOLOGY OF LINKS 313

splits. Furthermore, the quotient group Θn /bPn+1 injects into Πn (S)/J(SO), and
is an isomorphism if n 6= 2j − 2 for some j.
Here Πn (S) is the nth stable homotopy group of spheres and J is the well-known
Hopf-Whitehead J homomorphism, cf. [KM63, Rav86].
Kervaire and Milnor [KM63] also were able to determine bPn+1 in most cases.
They proved that bP2m+1 = 0 for m ≥ 1, that bP4m+2 is either 0 or Z2 , and that
for m ≥ 2, bP4m is cyclic of order
µ ¶
2m−2 2m−1 4Bm
|bP4m | = 2 (2 − 1) numerator ,
m
where Bm is the mth Bernoulli number defined by the power series
z z B1 2 B2 4
=1− + z − z + ··· .
ez − 1 2 2! 4!
The numerator of the rational number 4B m is tabulated in [MS74] up to m = 12.
m

For bP4m+2 the situation is still not entirely understood. It entails computing
the Kervaire invariant, which is hard. It is known (see the recent review paper
[Lan00] and references therein) that bP4m+2 = Z2 if 4m + 2 6= 2i − 2 for any i ≥ 3.
Furthermore, bP4m+2 vanishes for m = 1, 3, 7, and 15. The table below lists what
is known about the two groups in first few dimensions.

n Θn bPn+1
5 0 0
6 0 0
7 Z28 Z28
8 Z2 0
9 Z2 × Z2 × Z2 Z2
10 Z6 0
11 Z992 Z992
12 0 0
13 Z3 0
14 Z2 0
15 Z2 × Z8128 Z8128
16 Z2 0
17 Z2 × Z2 × Z2 × Z2 Z2
18 Z2 × Z8 0
19 ? Z130816

Let us make several remarks about exotic structures on spheres and the above
table. First, homeomorphism implies diffeomorphism in dimensions 1, 2, 3, and so
Perelman’s recently confirmed proof of the Poincaré conjecture implies Θ3 = 0.
However, in four dimensions the problem remains completely open. The so-called
smooth Poincaré conjecture asserts that there are no exotic 4-spheres5. In higher

5Apparently there is no consensus whether the conjecture is true or false. Four is the only
dimension where infinitely many exotic structures can and do exist on the same topological man-
ifold. Uncountably many are known to exist on R4 ! See the recent book [Sco05] for a discussion
of what is known about ‘the wild world of 4-manifolds’.
314 9. LINKS AS SASAKIAN MANIFOLDS

dimensions the computation of Θn depends on the higher homotopy groups of


spheres and those, in general, are not known.
Orientation is typically not a part of the definition of a smooth structure on a
manifold. Taking the group inverse in Θn corresponds to reversing the orientation.
Hence, to a differential geometer, in dimension 7 there are only 15 smooth structures
on spheres. The standard S 7 and one exotic 7-sphere admit orientation reversing
diffeomorphisms. The other 13 do not admit such a diffeomorphism and, hence, they
are counted “twice” in Θ7 . In dimension 9, however, there are indeed 8 different
smooth structures. Let Σn and −Σn be two distinct elements in Θn . Whether or
not one needs to distinguish the two depends on the context. For instance, if one
wants to establish existence (or non-existence) of some special Riemannian metric
on Σn any such result would automatically extends to −Σn . Then it is natural to
forget the orientation. On the other hand, on a Sasakian manifold orientation is
fixed, hence a part of the Sasakian structure. A construction of a Sasakian structure
on Σn does not automatically give a Sasakian structure on −Σn (cf. Remark 6.1.2).
In dimensions n = 8k+1 and n = 8k+2 there is another interesting subgroup of
Θn : the group of homotopy n-spheres which bound spin manifolds Θspin n ⊂ Θn . The
subgroup Θspin
n has index 2. The coset of this group, i.e., all homotopy spheres which
are not boundaries of spin manifolds is quite remarkable. In 1974 Hitchin [Hit74]
introduced the so-called α-invariant, showing that α(M ) 6= 0 is an obstruction to
M admitting any metric of positive scalar curvature. He observed that his invariant
does not vanish for half of the exotic spheres in dimensions 1 and 2 mod 8. These
exotic spheres are exactly the spheres which do not bound a spin manifold. For
example, in dimension 9 we get Θspin9 = Z2 × Z2 so that 4 exotic 9-spheres do not
admit any metric of positive scalar curvature. In dimension 10 one has three such
exotic spheres; two forgetting the orientation.
The computation of Θ9 and Θ17 is based on a paper of Brumfiel [Bru70], where
S
he shows that Θ4m+1 = Z2 ⊕(π4m+1 /im J) if 4m+2 6= 2j −2. Here J : πn (SO) → πnS
is the Adams’ J-homomorphism. As pointed out by Linus Kramer the relevant
stable homotopy groups of spheres are known and it follows that π9S /im J = Z2 ×Z2
S
and π17 /im J = Z2 × Z2 × Z2 .
Ten years after Milnor’s landmark paper, Brieskorn [Bri66] showed how all
odd-dimensional exotic spheres in the bPn+1 subgroup can be obtained as links of
isolated hypersurface singularities. In fact, all of such spheres can be realized as
very simple BP links. Let us recall Brieskorn’s result. We begin with an arbitrary
BP link L(a) of dimension at least 5. The Brieskorn Graph Theorem 9.3.18 tells
us when L(a) is a topological sphere. Suppose L(a) is homeomorphic to S 2n−1 . It
follows that L(a) ∈ bP2n is a boundary of a parallelizable manifold which is the
Milnor fiber F = Va2n . The determination of the smooth structure of L(a) = ∂Va2n
depends on the parity of n. We being with the simpler case of n = 2m + 1 odd.
Then is bP4m+2 either 0 or Z2 . When bP4m+2 = 0, without any further effort, one
concludes that all such BP links L(a) ∈ bP4m+2 are standard spheres. Such is the
case, for example, when n = 3, 7 (m = 1, 3). Brieskorn provides a complete answer
when bP4m+2 = Z2 and a partial answer when bP4m+2 is not known.
Theorem 9.4.3: Let L(a) be a topological sphere and n = 2m + 1 ≥ 5. The link
L(a) is the so-called Kervaire sphere Σ4m+1 if and only if the graph G(a) consists
of one isolated point ak and another connected component and ak ≡ ±3(8). In all
9.4. THE DIFFERENTIAL TOPOLOGY OF LINKS 315

other cases L(a) is diffeomorphic to the standard sphere. Furthermore, the Kervaire
sphere Σ4m+1 is exotic when bP4m+2 = Z2 .
The above theorem separates topological spheres among BP links L(a) ∈
bP2m+2 into the standard and exotic every time bP4m+2 = Z2 . But even when
bP4m+2 is not known Theorem 9.4.3 still separates standard spheres from the “po-
tentially exotic” Kervaire spheres.
Definition 9.4.4: Consider the polynomials
(9.4.1) f (z) = z06k−1 +z13 +z22 +· · ·+z2m
2
, g(z) = z02k+1 +z12 +z22 +· · ·+z2m+1
2
,
with k > 0. We denote the BP links associated with the above polynomials by
L(6k − 1, 3, 2, . . . , 2) = Σ4m−1
k , L(2k + 1, 2, . . . , 2) = Σ4m+1
k .
Example 9.4.5: L(3, 2, 2, 2) = Σ53 is a 5-sphere and it is a Kervaire sphere. But
there are no exotic 5-spheres so that L(3, 2, 2, 2) ' S 5 as is any other topological
5-sphere. L(3, 2, 2, 2, 2, 2) is Kervaire and Theorem 9.4.3 implies that it is the non-
trivial element in bP10 = Z2 . In particular, we get
(
4m+1 S 4m+1 , if 2k + 1 = ±1 mod8 ;
Σk = L(2k + 1, 2, . . . , 2) ≈
Σ4m+1 , if 2k + 1 = ±3 mod8 .
and the Kervaire spheres Σ4m+1 are always exotic when bP4m+2 6= 0.
To analyze Σk4m−1 we need more information. Suppose L(a) is a topological
(4m − 1)-sphere and an element in bP4m .
Theorem 9.4.6: Consider two BP links L(a) = ∂Va4m , L(b) = ∂Vb4m . There
exists an orientation preserving diffeomorphism L(a) ' L(b) if and only if
τ (Va4m ) = τ (Vb4m ) mod |bP4m | ,
where τ (Va4m ) is the Hirzebruch signature. In particular, L(a) is diffeomorphic to
the standard sphere when τ (Va4m ) = 0 mod |bP4m |.
Thus we only need to compute the Hirzebruch signature of the Milnor fiber
F = Va4m. Brieskorn derives a simple combinatorial index formula for
(9.4.2) τ (Va4m ) = τ+ (a) − τ− (a),
where
n 2m
X xi o
τ+ (a) = # x ∈ Z2m+1 | 0 < xi < ai and 0 < <1 mod 2 ,
a
i=0 i

n 2m
X xi o
τ− (a) = # x ∈ Z2m+1 | 0 < xi < ai and 1 < <2 mod 2 .
a
i=0 i
In principle, calculating the signature amounts to counting points in a finite subset
of an integral lattice whose size depends on a. For arbitrary very large exponents
this can be a challenging task. We remark that using a formula of Eisenstein, Zagier
(cf. [Hir71]) has rewritten this as:
N −1
(−1)m X π(2j + 1) π(2j + 1) π(2j + 1)
(9.4.3) τ (V 4m (a)) = cot cot · · · cot ,
N j=0
2N 2a0 2a2m

where N is any common multiple of the ai ’s.


316 9. LINKS AS SASAKIAN MANIFOLDS

Example 9.4.7: Consider the BP links Σk4m−1 = L(6k ± 1, 3, 2, . . . , 2) ∈ bP4m with


m > 1. For any k ≥ 1 the Brieskorn Graph Theorem 9.3.18 says this is a topological
sphere as G(a) has two isolated points. Let L(6k ± 1, 3, 2, . . . , 2) = ∂Vk4m > An
elementary calculation with (9.4.2) shows that

τ (Vk4m ) = (−1)m 8k .

Both L(5, 3, 2, . . . , 2) and L(7, 3, 2, . . . , 2) represent the so-called Milnor generator


in bP4m . In particular, L(6k ± 1, 3, 2, . . . , 2) ≈ #k Σ4m−1 1 realizes all elements in
bP4m by simply taking 1 ≤ k ≤ |bP4m |.
Summarizing Examples 9.4.5 and 9.4.7 gives the following result Brieskorn
[Bri66]:
Theorem 9.4.8: For n ≥ 3 every homotopy sphere Σ ∈ bP2n can be realized as
(i) Σ4m−1
k , k = 1, . . . , |bP4m |, when n = 2m
(ii) Σ4m+1
k , k = 1, 3, when n = 2m + 1.
In particular, Σ4m−1 4m+1
|bP4m | and Σ3 are diffeomorphic to the standard spheres.
In analogy with the Kervaire-Milnor group bP2n , Durfee [Dur77] defined the
group BP2n . Actually, he first defined this as a semigroup in [Dur71], and later in
[Dur77] with the same notation denoted the corresponding Grothendieck group.
Thus, we have
Definition 9.4.9: For n ≥ 3 let SBP2n denote the semigroup of diffeomorphism
classes of closed oriented (n − 2)-connected (2n − 1)-manifolds that bound paralleliz-
able manifolds, and let BP2n denote its Grothendieck completion.
As with bP2n multiplication in SBP2n is the connected sum operation, and the
standard sphere is a two-sided identity. Thus, SBP2n is a monoid. Furthermore,
for n ≥ 3 the Kervaire and Milnor group bP2n of Definition/Theorem 9.4.2 becomes
a subgroup of BP2n by Smale’s famous h-cobordism theorem [Sma62b, Mil65].
Unless otherwise stated we shall heretofore assume that n ≥ 3. We are mainly
interested in those highly connected manifolds that can be realized as links of
isolated hypersurface singularities defined by weighted homogeneous polynomials,
that is by the set W HP2n−1 of Definition 9.2.1. By (i) and (ii) of Theorem 9.3.2
one sees that there is a map

Φ : W HP2n−1 −−−→BP2n

which is the composition of the inclusion W HP2n−1 ,→ SBP2n with the natural
semigroup homomorphism SBP2n −−→BP2n . The image Φ(W HP2n−1 ) is a subset
of BP2n , and by (iii) of Theorem 9.3.2 it is a proper subset at least when n is
even. This can be contrasted with bP2n which according to Theorem 9.4.8 satisfies
bP2n ∩ Φ(W HP2n−1 ) = bP2n if n ≥ 3. Notice, however, that Φ(W HP2n−1 ) is not
generally a submonoid.
We now discuss invariants that distinguish elements of BP2n . First, by Poincaré
duality the only non-vanishing homology groups occur in dimension 0, n − 1, n and
2n − 1. Moreover, Hn (K, Z) is free and rank Hn (K, Z) = rank Hn−1 (K, Z). Thus,
our first invariant is the rank of Hn−1 (K, Z), so we define

(9.4.4) BP2n (k) = {K ∈ BP2n | rank Hn−1 (K, Z) = k} .


9.4. THE DIFFERENTIAL TOPOLOGY OF LINKS 317

Now BP2n is graded by its rank, so we have


M
(9.4.5) BP2n = BP2n (k) ,
k

and the monoid multiplication defines a map


(9.4.6) BP2n (k1 ) × BP2n (k2 )−−−−→BP2n (k1 + k2 ) .
Note that BP2n (0) is the submonoid of highly connected rational homology spheres.
The remaining known invariants [Wal67b, Dur71, Dur77] are a linking form
on the torsion subgroup of Hn−1 (K) and a quadratic invariant on the 2n-manifold
whose boundary is K. The precise nature of these invariants depends on whether
n is even or odd. For the case n even Durfee [Dur77] shows that for n ≥ 3 and
n 6= 4, 8 there is an exact sequence
Ψ
(9.4.7) 0−−−→bP2n −−−→BP2n −−−→ Z ⊕ KQ(Z)−−−→0 ,
where KQ(Z) denotes the Grothendieck group of regular bilinear form modules over
Z. Let us describe the map Ψ. The projection onto the first factor is just the rank of
Hn−1 (K) while the projection onto the second factor is Wall’s quadratic form [Wa2]
which is essentially the classical linking form b on the torsion subgroup of Hn−1 (K).
Any two manifolds K1 , K2 ∈ BP2n such that Ψ(K1 ) = Ψ(K2 ) differ by a homotopy
sphere, i.e., there is Σ ∈ bP2n such that K2 ≈ K1 #Σ, where ≈ means diffeomorphic.
But by Theorem 9.4.6 the elements Σ ∈ bP2n for n even are determined by the
signature of V. This completes the diffeomorphism classification for n 6= 4, 8 even.
The cases n = 4, 8 are more complicated [Wil72, Cro01]. Now, in addition to
the group Hn−1 (K) and the linking form b, there is an obstruction cocycle β̂ ∈
H n (K, πn−1 (SO)) ≈ H n (K, Z). The tangent bundle of K restricted to the (n − 1)-
skeleton is trivial and β̂ gives the obstruction to triviality on the n-skeleton. If the
torsion subgroup of Hn−1 (K) has odd order, then up to decomposability these are
all the invariants. However, if the torsion subgroup of Hn−1 (K) has even order,
things are even more complicated, and the analysis in [Wil72] was not complete. It
was recently completed in [Cro01]. The important point for us is that if the torsion
subgroup of Hn−1 (K) vanishes, K is determined completely up to homeomorphism
by the rank of Hn−1 (K). Summarizing we have
Theorem 9.4.10: Let M be a highly connected manifold in BP4n such that
H2n−1 (M, Z) = Zk . Then M is diffeomorphic to k#(S 2n−1 ×S 2n )#Σ4n−1 for some
Σ4n−1 ∈ bP4n .
Notice that by Theorem 7.4.11 k#(S 2n−1 × S 2n )#Σ4n−1 can admit a Sasakian
structure only if k is even. So by (iii) of Theorem 9.3.2 BP4n (2k + 1) ∩ W HP4n−1 =
∅, whereas we shall see that BP4n (2k) ∩ W HP4n−1 6= ∅ for all k.
For the case n odd the diffeomorphism classification was obtained by Wall
[Wal67b], but for our purposes, the presentation in [Dur71] is more convenient.
Let K ∈ BP2n with K = ∂V, where V can be taken as (n − 1)-connected and
parallelizable. In this case the key invariant is a Z2 -quadratic form
ψ : Hn (V, Z)/2Hn (V, Z)−−→Z2
defined by Wall [Wal62] as follows: Let X be an embedded n-sphere in V that
represents a non-trivial homology class in Hn (V, Z), and let [X] denote its image
318 9. LINKS AS SASAKIAN MANIFOLDS

in Hn (V, Z)/2Hn (V, Z). Then ψ([X]) is the characteristic class in


¡ ¢
ker πn−1 (SO(n))−−→πn−1 (SO) ≈ Z2
of the normal bundle of X. Let rad ψ denote the radical of ψ, i.e., the subspace of
the Z2 -vector space Hn (V, Z)/2Hn (V, Z), where ψ is singular. Then Durfee [Dur71]
(see also [DK75]) proves
Theorem 9.4.11: Let Ki ∈ BP2n for i = 1, 2 with n ≥ 3 odd be boundaries of
parallelizable (n−1)-connected 2n manifolds Vi with Z2 quadratic forms ψi . Suppose
that Hn−1 (K1 , Z) ≈ Hn−1 (K2 , Z), then
(i) if n = 3 or 7, then K1 and K2 are diffeomorphic;
(ii) if the torsion subgroups of Hn−1 (Ki , Z) have odd order and ψi |rad ψi ≡ 0
for i = 1, 2, then K1 ≈ K2 #(c(ψ1 )+c(ψ2 ))Σ, where c is the Arf invariant
and Σ is the Kervaire sphere, i.e., the generator of bP2n ;
(iii) if the torsion subgroups of Hn−1 (Ki , Z) have odd order and ψi |rad ψi 6≡ 0
for i = 1, 2, then K1 ≈ K2 ≈ K2 #Σ.
We end this section by discussing some results of Durfee and Kaufmann [DK75]
concerning the periodicity of branched covers. Let K ⊂ S 2n+1 be a simple fibered
knot or link (n ≥ 1), by which we mean an (n − 2)-connected (2n − 1)-embedded
submanifold of S 2n+1 for which the Milnor fibration theorem holds. If F is the
Milnor fiber of the fibration φ : S 2n+1 − K−−→S 1 then the monodromy map h :
Hn (F )−−→Hn (F ) is a fundamental invariant of the link K. Let Kk be a k-fold cyclic
branched cover of S 2n+1 branched along K. Then Durfee and Kauffman [DK75]
show that there is an exact sequence
1l+h+···+hk−1
(9.4.8) Hn (F ) −−−−−−−−→ Hn (F )−−→Hn (Kk )−−→0 .
So homologically Kk is determined by the cokernel of the map 1l + h + · · · + hk−1 .
Now suppose that K is a rational homology sphere and that the monodromy map
h of K has period d. Then since 1l − h is invertible, 1l + h + · · · + hd−1 is the zero
map in (9.4.8), and this determines the homology of Kd . The following lemma is
due to Durfee and Kauffman [DK75]
Lemma 9.4.12: Let K be a fibered knot in S 2n+1 which is a rational homology
sphere such that the monodromy map has period d. Suppose further that Kk is a
k-fold cyclic cover of S 2n+1 branched along K. Then
(i) Hn (Kd ) ≈ Hn (F ) ≈ Zµ , where µ is the Milnor number of K.
(ii) H∗ (Kk+d ) ≈ H∗ (Kk ) for all k > 0.
(iii) H∗ (Kd−k ) ≈ H∗ (Kk ) for all 0 < k < d.
Notice that (i) determines a large class of (n − 1)-connected (2n + 1)-manifolds
whose middle homology group Hn is free, and in certain cases this determines the
manifold up to homeomorphism. Items (ii) and (iii) give a homological periodicity.
Durfee and Kauffman also show that there are both homeomorphism and diffeo-
morphism periodicities in the case that n is odd and n 6= 1, 3, 7. In particular in
this case, when the link K is a rational homology sphere, Kk+d is homeomorphic
to Kk . To obtain the diffeomorphism periodicity let σk denote the signature of
the intersection form on the Milnor fiber Fk . Again assuming that K is a rational
homology sphere and h has periodicity d, one finds that Kk+d is diffeomorphic to
σd+1 σd+1 σd+1
8 Σ#Kk , where 8 Σ denotes 8 copies of the Milnor sphere Σ. Here we state
the slightly more general Theorem 6.4 of Durfee [Dur77]:
9.5. POSITIVE SASAKIAN STRUCTURES ON LINKS 319

Theorem 9.4.13: For even n 6= 2, 4, 8 let Ki be (n − 2)-manifolds that bound


parallelizable manifolds Vi , with i = 1, 2. Suppose that the quadratic forms of Ki
are isomorphic and Hn−1 (K1 , Z) ≈ Hn−1 (K2 , Z). Then σ(V2 )−σ(V1 ) is divisible by
8, and K2 is diffeomorphic to K1 # 18 (σ(V2 )−σ(V1 ))Σ, where σ(V ) is the Hirzebruch
signature of V.
Remark 9.4.1: Theorem 6.4 of [Dur77] as well as Theorem 5.3 of [DK75] ex-
clude the cases n = 4 and 8. However, it follows from [Wil72] and [Cro01] that
the diffeomorphism classification still holds in these cases since the links we are
considering here have no element of even order in the torsion subgroup of Hn−1 (In
fact the torsion subgroup vanishes in the case above). This remark also pertains to
the discussion for Theorem 3 below.

9.5. Positive Sasakian Structures on Links


The importance of positive Sasakian structures comes from Theorem 7.5.31
which says that they give rise to Sasakian metrics with positive Ricci curvature.
The existence of metrics of positive Ricci curvature on certain manifolds has been
an area of some interest. Most of the results obtained in this direction have used
the techniques of surgery theory. We mention here two such results. First there is
the work of Sha and Yang [SY91] who proved using surgery theory the existence
of metrics of positive Ricci curvature on the connected sum of products of spheres.
There is also the work of Wraith [Wra97] who proved that all homotopy spheres
that are boundaries of parallelizable manifolds admit a metric of positive Ricci
curvature. A proof of this result using the techniques elaborated in this book
was given in [BGN03c] and presented as Theorem 9.5.6 below. In [BGN03a]
a special case of the Sha-Yang result was established by showing the existence of
Sasakian metrics with positive Ricci curvature on the connected sums k(S 2 × S 3 ).
This and other examples on simply connected 5-manifolds will be discussed further
in Chapter 10. Of course, the existence of Sasaki-Einstein metrics as presented in
Chapter 11 also provides examples of Sasakian metrics of positive Ricci curvature.
We start with some applications of Theorem 7.5.31. First we can combine
Proposition 9.2.4 with Theorem 7.5.31 to obtain
Theorem 9.5.1: Let Lf be the link of an isolated hypersurface singularity of a
weighted homogeneous polynomial f, and suppose that |w| − d > 0. Then Lf admits
a Sasakian metric with positive Ricci curvature.
Next we have
Theorem 9.5.2: Let Lf 0 be the link of a w.h.p. f 0 (z2 , . . . , zn ) ∈ L∗ (w0 , d0 ). As-
sume that the origin in Cn−1 is the only singularity so that Lf 0 is smooth. Consider
the weighted homogeneous polynomial
f = z02 + z12 + f 0
of degree d = lcm(2, d0 ). Then the link Lf admits a Sasakian structure with positive
Ricci curvature and bn−1 (Lf ) = bn−3 (Lf 0 ).
Proof. There are two cases. If d0 is odd then the weight vector of f is w =
0 0
(d , d0 , 2w0 ), whereas, if d0 is even, then w = ( d2 , d2 , w0 ). In the first case we have
0

|w| − d = d0 + d0 + 2|w0 | − 2d0 = 2|w0 | > 0, while in the second case |w| − d =
d0 d0 0 0 0
2 + 2 +|w |−d = |w | > 0. In either case Lf admits a Sasakian metric with positive
Ricci curvature by Theorem 9.5.1. The Betti number equality is actually well-known
320 9. LINKS AS SASAKIAN MANIFOLDS

and follows from the Sebastiani-Thom Theorem [ST71, KN77, Dim92], but for
completeness we give an independent derivation suited to our context. We consider
the divisor div ∆f of the Alexander polynomial ∆f of f. By Theorem 9.3.9 we have
Y³ Λu ´ Y³ Λu0 ´ Y³ Λ u0 ´
div ∆f = i
− 1 = (Λ2 − 1)2 0
i
−1 = 0
i
− 1 = div ∆f 0 ,
i
vi 0
vi 0
vi
i i
0
and for i = 2, . . . , n we have wdi = wd 0 . So ui = u0i and vi = vi0 for the same range of
i
i. Thus, the Betti number equality follows from Corollary 9.3.12. ¤

9.5.1. Positive Sasakian Structures on Homology Spheres. We begin


by studying homology spheres, both rational homology spheres and homotopy
spheres. An immediate corollary of Theorem 9.5.2 is
Corollary 9.5.3: Let Lf 0 be the link of a weighted homogeneous polynomial that
is a rational homology sphere. Then

f = z02 + z12 + f 0

is the link Lf of a rational homology sphere that admits a Sasakian structure with
positive Ricci curvature.
Example 9.5.4: Consider a BP link L(a) with a = (k, p, . . . , p) for any relatively
prime positive integers k, p ∈ Z+ . Brieskorn Graph Theorem 9.3.18 implies that this
is always a rational homology sphere as the graph G(a) has one isolated vertex. In
addition, it is never a homology sphere unless p = 2 and n odd which was already
discussed in the example 9.3.21 of the previous section. Using corollary 9.3.19 we
see that L(a) is positive if and only if p(k−1) < nk and negative when p(k−1) > nk.
In either case one gets infinitely many examples. As expected, we cannot get null
links as p(k − 1) 6= nk when gcd(k, p) = 1. To see that one indeed gets infinitely
many, one can take the simplest case of rational homology 5-spheres and compute
the torsion group
H2 (L(a), Z) = (Zk )(p−1)(p−2) .
The only infinite series of positive links we get is L(k, 3, 3, 3), with k > 3 prime to
3. L(k, 3, 3, 3) is diffeomorphic to the Barden prime manifold Mk5 . These links will
be discussed further in Chapters 10 and 11 together with many other examples of
five- and seven-dimensional links. Note that any L(k, p, p, p) with p ≥ and k > 1
prime to p is negative with the exception of L(3, 4, 4, 4) whose torsion group equals
(Z3 )6 . Furthermore, L(k, p, p, p) is a 12 (p − 1)(p − 2)-fold connected sum of Mk5 .
Theorem 9.5.5: Every (2n − 2)-connected (4n − 1)-manifold that is the boundary
of a parallelizable manifold whose homology group H2n−1 (K, Z) is isomorphic to Z3
admits Sasakian metrics with positive Ricci curvature. There are precisely 2|bP4n |
such smooth oriented manifolds.

Proof. This is essentially a corollary of Proposition 7.2 of [Dur77] where


Durfee considers the link Kk of the Brieskorn-Pham polynomial z0k + z13 + z22 +
· · · + zn2 for even n ≥ 4. He shows that Hn (K2 , Z) ≈ Hn (K4 , Z) ≈ Z3 , but that K2
and K4 have inequivalent linking forms. Furthermore, K6l+2 is diffeomorphic to
n n
K2 #(−1) 2 lΣ4n−1 and K6l+4 is diffeomorphic to K4 #(−1) 2 lΣ4n−1 , where Σ4n−1 =
K5 is the Milnor generator. ¤
9.5. POSITIVE SASAKIAN STRUCTURES ON LINKS 321

There are two distinct non-homeomorphic topological manifolds in Theorem


9.5.5 and they are distinguished by their linking form in H2n−1 (K, Z) ≈ Z3 . Each
topological manifold is comprised of |bP4n | distinct diffeomorphism types.
For homotopy spheres we have the following result [BGN03c]:
Theorem 9.5.6: For n ≥ 3 let Σ2n−1 be a homotopy sphere which can be realized
as the boundary of a parallelizable manifold. Then Σ2n−1 admits Sasakian metrics
with positive Ricci curvature.

Proof. By Theorem 9.4.8 every homotopy sphere that is the boundary of a


parallelizable manifold can be realized as the link Lf of the Example 9.4.5. By
Theorem 9.5.2 any all such links admits Sasakian metrics with positive Ricci cur-
vature. ¤

For more examples of positive Sasakian structures on homotopy spheres see


[BGN03c].
López de Medrano [LdM71] has shown that the quotient space of any fixed
point free involution on a homotopy sphere is homotopy equivalent to a real pro-
jective space RPn . Here we are interested in fixed point free involutions on Σ2n+1
that also preserve the Sasakian structure. In particular, we consider the well-known
involution T on Σ4m+1
p defined by (z0 , z1 , . . . , z2m+1 ) 7→ (z0 , −z1 , . . . , −z2m+1 ). One
easily sees that the group hT i generated by T acts freely on Σ4m+1 , so by López
de Medrano’s theorem the quotient manifold Σ4m+1 p /hT i is homotopy equivalent
to RP4m+1 . Furthermore, it is clear that hT i is a subgroup of the weighted circle
1
group Sw with weights w = (2, 2p + 1, . . . , 2p + 1). Thus, the deformation class of
Sasakian structures passes to the quotient Σ4m+1 p /hT i to give a deformation class
of Sasakian structures together with their characteristic foliation Fξ . Furthermore,
by Theorem 9.5.2 the Sasakian structures on Σ4m+1 p /hT i are positive and admit
Sasakian metrics of positive Ricci curvature. We have
Theorem 9.5.7: On each of the known 22m oriented diffeomorphism types of ho-
motopy projective spaces RP4m+1 there exist deformation types of positive Sasakian
structures, and each deformation class contains Sasakian metrics of positive Ricci
curvature.

Proof. It is known [AB68, Bro73, Gif69a, Gif69b] that there are at least
22m diffeomorphism types on a homotopy real projective space of dimension 4m+1.
Atiyah and Bott [AB68] and Browder [Bro73] obtained the bound 22m−1 which
was then extended to 22m by Giffen [Gif69a, Gif69b]. Furthermore, for 0 ≤ p <
p0 < 22m the homotopy real projective spaces Σ4m+1 p0 /hT i and Σ4m+1
p /hT i are not
4m+1 2m
diffeomorphic. Thus, the Σp /hT i realize the 2 diffeomorphism types distin-
guished by Giffen and from the discussion above each of these admit deformation
classes of positive Sasakian structures. In dimension 4 + 1 = 5, the bound 22·1 = 4
is sharp. This proves the theorem. ¤

9.5.2. Further Examples of Positive Sasakian Structures. Here we de-


scribe some results concerning the existence of positive Sasakian structures on man-
ifolds other than rational homology spheres. The next two theorems were described
in [BG06c].
Theorem 9.5.8: Let n ≥ 2 be an integer, then for each positive integer k there
exist Sasakian metrics with positive Ricci curvature in Dn (k) of the |bP4n | oriented
322 9. LINKS AS SASAKIAN MANIFOLDS

diffeomorphism classes of the (4n − 1)-manifolds 2k#(S 2n−1 × S 2n ), where Dn (k)


is determined by an explicit formula given below.
Theorem 9.5.9: For each pair of positive integers integer (n, k) there exists a
(2n − 1)-connected (4n + 1)-manifold K with H2n (K, Z) free of rank k which admits
a Sasakian metric of positive Ricci curvature. Furthermore, K is diffeomorphic to
one of manifolds
#k(S 2n × S 2n+1 ) , #(k − 1)(S 2n × S 2n+1 )#T , #k(S 2n × S 2n+1 )#Σ4n+1 ,
where T = T1 (S 2n+1 ) is the unit tangent bundle of S 2n+1 , and Σ4n+1 is the Kervaire
sphere. For k = 1 the manifolds
S 2n × S 2n+1 , (S 2n × S 2n+1 )#Σ4n+1 , T
all admit Sasakian metrics with positive Ricci curvature. If n = 1 or n = 3 then
#k(S 2n × S 2n+1 ) admits a Sasakian metric with positive Ricci curvature for all k.
Proof. The links that we need to prove Theorems 9.5.8 and 9.5.9 involve
Brieskorn polynomials of the form
(9.5.1) fp,q = z0p + z1q + z22 + · · · + zn2 .
The link associated with fp,q is
−1
Lp,q = fp,q (0) ∩ S 2n+1 .
By Theorem 9.5.2 all such links admit Sasakian metrics with positive Ricci curva-
ture. One can view Lp,q as a p-fold branched cover of S 2n−1 branched over the link
Lq defined by the polynomial
fq = z1q + z22 + · · · + zn2 .
For Theorem 9.5.8 we need the link L2(2k+1),2k+1 , i.e., p = 2(2k +1), q = 2k +1,
with n even. In this case the degree of L2(2k+1),2k+1 is d = 2(2k + 1) which is the
period of the monodromy map of the link L2k+1 . Furthermore, L2k+1 is a homology
sphere which can be seen by computing its Alexander polynomial,
t2k+1 + 1
∆(t) = .
t+1
Now the link L2(2k+1),2k+1 is a 2(2k + 1) branched cover of S 2n−1 branched over
L2k+1 , so by (i) of Lemma 9.4.12, we have
(9.5.2) Hn−1 (L2(2k+1),2k+1 , Z) ≈ Hn (L2(2k+1),2k+1 , Z) ≈ Zµ = Z2k .
Here µ is the Milnor number of the link L2k+1 which is easily computed by the
formula for Brieskorn polynomials, namely
Yn
µ= (ai − 1) = (2k + 1 − 1) · 1 · · · 1 = 2k .
i=0
It now follows from Theorem 9.4.10 that the link L2(2k+1),2k+1 is diffeomorphic to
2k#(S n−1 ×S n )#Σ4n−1 for some Σ4n−1 ∈ bP4n . We now use the periodicity results
of Durfee and Kauffman to determine the diffeomorphism type. First we notice that
Theorem 9.4.10 together with Theorem 4.5 of [DK75] imply that for every positive
integer i and every positive integer k, the link L2i(2k+1),2k+1 is homeomorphic to
the connected sum 2k#(S n−1 × S n ). The diffeomorphism types are determined by
Theorem 9.4.13 (Theorem 6.4 of [Dur77], see also Theorem 5.3 of [DK75]) together
9.5. POSITIVE SASAKIAN STRUCTURES ON LINKS 323

with Remark 9.4.1. Let Fi,k denote the Milnor fibre of the link L2i(2k+1),2k+1 and
σ(Fi,k ) its Hirzebruch signature. Then Theorem 9.4.13 says that for each pair of
positive integers i, j there is a diffeomorphism
³ σ(F ) − σ(F ) ´
i,k j,k
(9.5.3) L2i(2k+1),2k+1 ≈ Σ #L2j(2k+1),2k+1 .
8
Here lΣ denotes the connected sum of l copies of the Milnor sphere, and a minus
sign corresponds to reversing orientation. It follows from Durfee’s theorem that
the difference in signatures is always divisible by 8, so this expression makes sense.
Equation (9.5.3) can be iterated; so it is enough to consider the case i = 2 and
j = 1. In order to determine how many distinct diffeomorphism types occur in
(9.5.3), we need to compute the signature of the Milnor fibres. To do so we adapt
formula (9.4.3) to the case of equation (9.5.3) with N = 2(2k + 1), namely, a =
(2(2k +1), 2k +1, 2 . . . , 2). Notice that we can always take the N in Zagier’s formula
9.4.3 to be this in which we shall denote τ (V 4m (a)) by td since the degree d =
2(2k + 1) is the periodicity as well. Likewise, we denote by t2d the signature t(a)
with a = (4(2k + 1), 2k + 1, 2 . . . , 2). We find
n 4k+1
(−1) 2 X π(2j + 1) π(2j + 1)
td = (−1)j cot2 cot ,
4k + 2 j=0 8k + 4 4k + 2

and
n 8k+3
(−1) 2 X π(2j + 1) π(2j + 1)
t2d = (−1)j cot2 cot .
8k + 4 j=0 16k + 8 4k + 2
|t2d −td |
We want to compute τk = 8 . After some algebra we find

π(2j + 1) ³ π(2j + 1) ´
8k+3
X π(2j + 1) π(2j + 1)
τk = γ (−1)j cot cot − cot cot ,
j=0
16k + 8 16k + 8 8k + 4 4k + 2
1
where γ = 64k+32 . Now τk is always an integer that is independent of n. We now
define
|bP4n |
Dn (k) = .
gcd(τk , |bP4n |)
which by equation (9.5.3) represents the number of distinct diffeomorphism types
that can be represented by our construction. Using MAPLE we give two tables
consisting of a list of τk and Dn (k) together with the ratio
D2 (k) 1
=
|bP8 | gcd(τk , |bP4n |)
for both the 7-manifolds #2k(S 3 × S 4 ) and the 11-manifolds #2k(S 5 × S 6 ) for
various values of k. These two tables are given in Appendix B.2. This proves
Theorem 9.5.8.
Remark 9.5.1: Notice that the prime factorization of |bP4n | consists of high powers
of two together with odd primes coming from the Bernoulli numbers. Since τk is
independent of n, this gives rise to a bit of a pattern for the ratios D n (k)
|bP4n | . It is
obvious that for k = 1 all possible diffeomorphism types occur, but this seems also
to hold for k = 2. It is of course true whenever |bP4n | is relatively prime to 3. If we
look at the next case namely, bP16 , we see that |bP16 | = 8128 = 26 · 127. Comparing
324 9. LINKS AS SASAKIAN MANIFOLDS

this with |bP12 | = 992 = 25 · 31, we see that the same ratios will occur for the case
#2k(S 7 × S 8 ) as for #2k(S 5 × S 6 ) for k = 1, . . . , 30.
To prove Theorem 9.5.9 we have n odd and there are several cases. First we
take p = 2(2k + 1), q = 2k + 1 as in the proof of Theorem 9.5.8. Again this leads to
the link L2(2k+1),2k+1 with free homology satisfying equation (9.5.2) except now n
is odd. Next we consider q = 2k in equation (9.5.1). The link L2k of the Brieskorn-
Pham polynomial f2k = z12k + z22 + · · · + zn2 is a rational homology sphere by the
Brieskorn Graph Theorem. Furthermore, its monodromy map has period 2k. Then
choosing p = 2k in equation (9.5.1) the link L2k,2k is 2k-fold branched cover over
S 2n+1 branched over the rational homology sphere L2k , so by item (i) of Lemma
9.4.12, we have
Hn−1 (L2k,2k , Z) ≈ Hn (L2k,2k , Z) ≈ Zµ = Z2k−1 .
These two cases now give links whose middle homology groups are free of arbitrary
positive rank. However, unlike the case for n even this does not determine the
homeomorphism type unless n = 3, 7 in which case there is a unique diffeomorphism
class. Indeed Theorem 9.4.11 implies we need to compute the quadratic form ψ,
and this appears to be quite difficult in all but the simplest case. From Theorem
9.4.11 one can conclude [Dur71] that if M ∈ BP4n+2 with H2n (M, Z) free of
rank one, then it is homeomorphic to S 2n × S 2n+1 or the unit tangent bundle
T = T1 (S 2n+1 ). So the diffeomorphism types at most differ by an exotic Kervaire
sphere Σ4n+1 . Furthermore, S 2n × S 2n+1 , T and (S 2n × S 2n+1 )#Σ4n+1 generate
the torsion-free submonoid of BP4n+2 , there being relations in the monoid, namely,
T #T = #2(S 2n × S 2n+1 ) and T #Σ4n+1 = T (Some further relations may exist
depending on n such as T1 (S 3 ) ≈ S 2 × S 3 ). This proves the first statement in
Theorem 9.5.9.
To prove the second statement we follow Durfee and Kauffman and consider
a slightly different Brieskorn-Pham polynomial, namely z02k + z12 + · · · + zn2 . For
k = 1 we get as before a link L2,2 whose middle homology group is free of rank one.
Thus, it is diffeomorphic to one of the three generators above by (i) of Lemma 9.4.12.
Now as k varies we have a homological periodicity by (ii) and (iii) of Lemma 9.4.12.
Durfee and Kauffman show that there is an 8-fold diffeomorphism periodicity, and
they compute the ψ invariant to show that
L2,2 ≈ T , L4,2 ≈ (S 2n × S 2n+1 )#Σ4n+1 ,
L6,2 ≈ T #Σ4n+1 ≈ T , L8,2 ≈ S 2n × S 2n+1 .
This proves Theorem 9.5.9. ¤
9.5.3. Exotic Contact Structures. The material discussed here is well-
known and taken from Sato [Sat77] and Morita [Mor75]. Given an oriented almost
contact structure on a compact manifold M 2n−1 , we let A(M ) be the set of homo-
topy classes of almost contact structures on M. This is in one-to-one correspondence
with the set of homotopy classes of almost complex structures on M × R [Sat77],
and when M = Σ2n−1 is a homotopy sphere, the latter is known [Mor75]:


 Z ⊕ Z2 if n ≡ 0 mod 4;


Z if n ≡ 1 mod 4;
(n−1)!
(9.5.4) A(Σ2n−1 ) = π2n−1 (SO(2n)/U (n)) ∼ =

 Z if n ≡ 2 mod 4;


Z (n−1)! if n ≡ 3 mod 4.
2
9.5. POSITIVE SASAKIAN STRUCTURES ON LINKS 325

Theorem 9.5.10: On each odd homotopy sphere Σ2n−1 ∈ bP2n , n > 2, there
exist countably infinitely many deformation classes of positive Sasakian structures
belonging to non-isomorphic underlying contact structures. Hence, the moduli space
of deformation classes of Sasakian metrics with positive Ricci curvature on Σ2n−1
has infinitely many positive components.
Proof. The proof for the homotopy spheres Σ4m−1 and Σ4m+1 is somewhat
different. The case Σ4m−1 uses the proof of Theorem 4.1(i) of Morita [Mor75]. We
represent each homotopy sphere Σ4m−1 ∈ bP4m by the link of a BP polynomial.
The underlying almost contact structures can be distinguished by an invariant ∂,
Morita’s invariant, which is defined in terms of the Hirzebruch signature τ of the
Milnor fibre, the Milnor number µ of the link, and the Bernoulli numbers. Explicitly,
(−1)m τ (V4m ) 1
(9.5.5) ∂(L(a)) = − µ(L(a)) ,
4Sm 2
where
22m (22m−1 − 1)
Sm = Bm .
m!
It is more convenient to work with the link L(6k + 1, 3, 2, . . . , 2) rather than the
standard Brieskorn link with leading term 6k − 1. This is not a problem since the
two series both represent all homotopy spheres in bP4m . A simple computation gives
³ 2 ´
∂(L(6k + 1, 3, 2, . . . , 2)) = −6 k.
Sm
Moreover, there are an infinite number of k that correspond to the same homotopy
sphere Σ4m−1 ∈ bP4m . Thus, we will have a countably infinite number of distinct
underlying almost contact structures on the same homotopy sphere as long as Sm 6=
1
3 . We show that this always holds. Recall the well-known estimate for the Bernoulli
2(2m)!
numbers Bm > (2π) 2m (cf. Appendix B of [MS74]). We then have

22m (22m−1 − 1) 2(22m−1 − 1)(2m)! ³ m + 1 ´m


2m−1
Sm = Bm > > 2(2 − 1)
m! π 2m m! π2
which is clearly greater than 1 for all m ≥ 9. One can easily check, for example
using MAPLE6, that Sm > 1 for all 2 ≤ m < 9. This completes the proof for this
case.
Now every homotopy sphere of the form Σ4m+1 can be represented as a BP link
4m+1
Σk of Definition 9.4.4. Moreover, we know that Σ4m+1 k is the standard sphere
if 2k + 1 ≡ ±1(8), and the Kervaire sphere if 2k + 1 ≡ ±3(8) which is exotic when
4m + 2 6= 2i − 2. In this case the Morita invariant is simply given in terms of the
Milnor number µ of the link, viz.
1
∂(Σ4m+1
k ) = µ(Σ4m+1 k ) = k.
2
Thus, as 2k + 1 varies through ±1(8) or ±3(8), the invariant ∂(Σ4m+1 k ) varies
through 0(4) and 3(4), or 1(4) and 2(4), respectively. Let a(m) denote the order of
the group A(Σ4m+1
k ) defined by Eequation (9.5.4). Since for m > 1 a(m) ≡ 0(4) we
can choose k such that ∂(Σ4m+1k ) is any number mod a(m). Thus, there are a(m)
distinct underlying almost contact structures which is finite in this case. To obtain
6The notation for Bernoulli numbers is not standard. We use the notation used by topologists,
whereas, Maple uses the notation used by applied mathematicians. If we denote this Bernoulli
number by bm , then our Bm equals (−1)m−1 b2m .
326 9. LINKS AS SASAKIAN MANIFOLDS

an infinite number of inequivalent contact structures we use the recent results of


Ustilovsky [Ust99] who showed by explicitly computing the contact homology of
Eliashberg, Givental, and Hofer (cf. [Eli98]) that for the standard sphere S 4m+1
distinct k’s satisfying 2k + 1 ≡ ±1(8) give inequivalent contact structures. How-
ever, Ustilovsky’s proof works equally well for any homotopy sphere Σ4m+1 . This
completes the proof. ¤

Remarks 9.5.1: (1) It is interesting to note that for the 5-sphere S 5 (m = 1


above), Ustilovsky’s results distinguish infinitely many inequivalent contact struc-
tures, whereas, there is a unique underlying almost contact structure. (2) It is
tempting to adapt the computations in [Ust99] to the case of the homotopy pro-
jective spaces RP4m+1 . However, in this case complications arise, not the least of
which is the fact that the simplification of the contact homology arising from the
fact that all the Conley-Zehnder indices are even no longer holds. (3) An algorithm
for computing the contact homology for general Brieskorn manifolds has been given
recently in [vK04].

9.6. Links of Complete Intersections


Here we consider complete intersection singularities defined by p weighted ho-
mogeneous polynomials f1 , . . . , fp in Cn+1 of degrees d1 , . . . , dp , respectively. Our
main references are [Dim92, Dol82, IF00, Loo84]. We require that each of the
polynomials fi for i = 1, . . . , p is a weighted homogeneous polynomial of degree
di with weight vector w. We use vector notation and write f = (f1 , . . . , fp ), and
d = (d1 , . . . , dp ); the latter is called the multidegree of f . We are interested in the
weighted affine cone Cf defined by
Cf = {(z0 , . . . , zn ) ∈ Cn+1 | f (z0 , . . . , zn ) = 0} .
We shall assume that dimC Cf = n + 1 − p and that the origin in Cn+1 is an isolated
singularity, in fact the only singularity, of f . Then the link Lf defined by
(9.6.1) Lf = Cf ∩ S 2n+1 ,
is a smooth manifold of dimension 2(n − p) + 1. Furthermore, it is well-known
[Loo84] that
Proposition 9.6.1: Let f = (f1 , . . . , fp ) be p weighted homogeneous polynomials of
multidegree d = (d1 , . . . , dp ) with weight vector w = (w0 , . . . , wn ). Let Lf be the link
of an isolated complete intersection defined by F. Then Lf is (n − p − 1)-connected.
The C∗ (w) action on Cn+1 restricts to an action on Cf , and the associated S 1
action restricts to an action on both S 2n+1 and Lf . It follows that ξw is tangent to
the submanifold Lf and, by abuse of notation, we shall denote by ξw , ηw , Φw , gw
2n+1
the corresponding tensor fields on both Sw and Lf . Now Φw coincides with Φ
2n+1
on the contact subbundle D on S which defines an integrable almost complex
structure on D. Moreover, since f is a holomorphic function on Cn+1 the Cauchy-
Riemann equations imply that for any smooth section X of D we have Φw X(f ) = 0.
Thus, Lf is an invariant submanifold of S 2n+1 with its weighted Sasakian structure,
2n+1
or alternatively the inclusion Lf ,→ Sw is a Sasakian embedding. This gives an
easy generalization of a previous result of the first two authors [BG01b], viz.
Theorem 9.6.2: The quadruple (ξw , ηw , Φw , gw ) gives Lf a quasi-regular Sasakian
2n+1
structure such that Lf ,→ Sw is a Sasakian embedding.
9.6. LINKS OF COMPLETE INTERSECTIONS 327

Since the S 1 (w)-action is locally free on both Sw 2n+1


and Lf , we have well-
defined quotient spaces which by a standard theorem are compact Kähler orbifolds.
2n+1
The quotient of Sw is the well-known weighted projective space CP(w), and the
quotient of Lf is an algebraic variety Zf . It is also a suborbifold of CP(w) whose
singularities are quotient singularities induced by those of CP(w). So there is a
commutative diagram
2n+1
Lf −−−−→ Sw
 
 
(9.6.2) π 
y y
Zf −−−−→ CP(w) ,
where the horizontal arrows are Sasakian and Kählerian embeddings, respectively,
and the vertical arrows are orbifold Riemannian submersions. The right vertical
map is nothing but the weighted version of the well-known Hopf fibration. The
algebraic variety Zf is a complete intersection in the weighted projective space
CP(w), and the condition that the affine cone Cf be smooth away from the origin
meansPthat Zf is quasi-smooth [Dol82]. As with the weight vector w we define
|d| = di . Then we have the following obvious generalization of Proposition 9.2.4:
Proposition 9.6.3: The link Lf is a
(i) canonical Sasakian structure if and only if |d| − |w| > 0,
(ii) null Sasakian structure if and only if |d| − |w| = 0,
(iii) anticanonical Sasakian structure if and only if |d| − |w| < 0.
In particular all links of complete intersections are either positive definite, negative
definite, or null.
CHAPTER 10

Sasakian Geometry in Dimensions Three and Five

In this chapter we shall discuss the geometry of Sasakian manifolds in first two
non-trivial dimensions. The topology and geometry of compact three-dimensional
manifolds is extremely rich and varied. There has been enormous progress in the last
twenty years beginning with foundational work of Thurston [Thu97] and Hamilton
[Ham82], and culminating in the amazing papers of Perelman [Per02, Per03b,
Per03a]. Both the Poincaré Conjecture and the Thurston Uniformization Con-
jecture have finally succumbed to the powers of the modern Geometric Analysis.
See [KL06, MT06, CZ06] for full details of Perelman’s work. In contrast, the
Sasakian geometry of three-dimensional manifolds is much simpler owing to the fact
that any such manifold must be positive, negative or null which essentially reduces
their classification to the uniformization of the compact two-dimensional surfaces.
Furthermore, up to a cover there exists only one compact Sasaki-Einstein manifold
and it is the unit 3-sphere. In dimension five the situation is far more compli-
cated. There is, however, a classification (see Theorem 10.2.3 below) of closed
simply connected five manifolds up to diffeomorphism due to Smale [Sma62a] and
Barden [Bar65]. As we shall see in this chapter, in many ways dimension five is
large enough to accommodate many interesting Sasakian structures and yet small
enough to hope for some kind of classification, at least in the simply-connected
case.

10.1. Sasakian Geometry in Dimension Three


By Martinet’s Theorem 6.1.22 every orientable 3-manifold admits a contact
structure. The study of contact structures on 3-manifolds is quite rich and leads
to many yet unsolved problems. They occur in two types: tight contact structures,
and overtwisted contact structures. The latter have been classified by Eliashberg
[Eli89], leaving the tight contact structures as an important area of investigation.
See [Eli92, Bla02, Hon06, Gei06] for details and references. In contrast, as shown
by Belgun, the Sasakian structures are much more rigid and can be completely
classified [Bel00, Bel01]. Before discussing Belgun’s theorem, we briefly consider
regular Sasakian structures in dimension 3. In the compact case this follows easily
from the classification of compact Riemann surfaces.
Proposition 10.1.1: Every regular compact Sasakian 3-manifold M is a circle
bundle over a compact Riemann surface Σg of genus g.
Geiges [Gei97b] showed that a compact 3-manifold admits a Sasakian structure
if and only if it is diffeomorphic to one of the following:
(i) S 3 /Γ with Γ ⊂ I0 (S 3 ) = SO(4).
f
(ii) SL(2, f
R)/Γ, where is universal cover of SL(2, R) and Γ ⊂ I0 (SL(2, R)).
3 3
(iii) Nil /Γ with Γ ⊂ I0 (Nil ).
329
330 10. SASAKIAN GEOMETRY IN DIMENSIONS THREE AND FIVE

Here Γ is a discrete subgroup of the connected component I0 of the corresponding


isometry group with respect to a ‘natural metric’, and Nil3 denotes the 3 by 3
nilpotent real matrices, otherwise known as the Heisenberg group. These are three
of the eight model geometries of Thurston [Thu97], and correspond precisely to
the compact Seifert bundles with non-zero Euler characteristic [Sco83]. For further
discussion of the isometry groups we refer to [Sco83]. We refer to the three model
geometries above as spherical, SLg2 -type, and nil geometry, respectively.
The purpose of this section then is to describe some examples represented as
links of isolated hypersurface singularities. First note that in dimension 3 the basic
cohomology group H 2 (Fξ ) is 1-dimensional so we cannot get indefinite forms other
than the zero class. Hence
Proposition 10.1.2: Let M be a compact 3-dimensional Sasakian manifold. Then
M is either canonical (negative), anticanonical (positive) or null.
These three types of Sasakian structures correspond precisely to the three model
geometries above. Belgun’s Theorem uniformizes these three cases. We prefer to
rephrase Belgun’s theorem in terms of Tanno’s Theorem 7.3.23 on constant Φ-
sectional curvature. In dimension three the Φ-sectional curvature is determined by
one function, namely H(X) = K(X, ΦX) (cf. [Bel01], Theorem 1).
Theorem 10.1.3: Let M be a 3-dimensional compact manifold admitting a Sasakian
structure S = (ξ, η, Φ, g). Then
(i) If S is positive, M is spherical, and there is a Sasakian metric of constant
Φ-sectional curvature 1 in the same deformation class as g.
(ii) If S is negative, M is of SL g2 type, and there is a Sasakian metric of
constant Φ-sectional curvature −4 in the same deformation class as g.
(iii) If S is null, M is nil, and there is a Sasakian metric of constant Φ-
sectional curvature −3 in the same deformation class as g.
Actually, when S is either null or negative than M is up to finite quotient a
regular Sasakian 3-manifold, i.e., a circle bundle over a Riemann surface of positive
genus. In this case the Sasaki cone is one-dimensional (cf. Proposition 8.2.15)
so the only non-trivial deformations are deformations of the Kähler structure on
the underlying (orbifold) Riemann surface. In the positive case, M is covered by
S 3 and its Sasakian structure is a deformation of a standard Sasakian structure
S0 . In particular, Theorem 10.1.3 states that any Sasakian structure on S 3 is a
deformation of the standard one. As we have seen in the previous chapter (cf.
Theorem 9.5.10) this is far from true for higher-dimensional spheres. Generally
speaking not much is known about the Sasakian structures on S 2n+1 , n > 1.
Exercise 10.1: Prove Theorem 10.1.3.
It is convenient when possible to represent Sasakian 3-manifolds as links of
quasi-smooth weighted homogeneous polynomials. In the case of positive Sasakian
structures this is always possible at least up to deformations, while in the null and
negative cases it is not.

10.1.1. Three-Dimensional Links. Since by Proposition 9.2.2 any link of a


quasi-smooth weighted homogeneous polynomial admits a class of Sasakian struc-
tures, one can obtain fairly explicit constructions of Sasakian structures in dimen-
sion three. Furthermore, it is easy to classify weighted homogeneous polynomials
in 3 variables [OW71a].
10.1. SASAKIAN GEOMETRY IN DIMENSION THREE 331

Proposition 10.1.4: Let h(z0 , z1 , z2 ) be a weighted homogeneous polynomial in


three complex variables having an isolated singularity at the origin. Then there
are a permutation σ on the set {0, 1, 2}, non-zero complex numbers α0 , α1 , α2 , and
weighted homogeneous polynomials f and g with no monomials in common such
that h can be written as h = f + g, where f (α0 zσ(0) , α1 zσ(1) , α2 zσ(2) ) is equal to1:
(I) z0a0 + z1a1 + z2a2 ,
(II) z0a0 + z1a1 + z2a2 z1 with a1 > 1,
(III) z0a0 + z1a1 z2 + z2a2 z1 with a1 > 1, a2 > 1,
(IV) z0a0 + z1a1 z0 + z2a2 z1 with a0 > 1,
(V) z0a0 z1 + z1a1 z2 + z2a2 z0 ,
(VI) z0a0 + z0 (z2a2 + z1a1 ) + z2b2 z1b1 with (a0 − 1)(a2 b1 + a1 b2 ) = a0 a1 a2 ,
(VII) z0a0 z1 + z0 (z2a2 + z1a1 ) + z2b2 z1b1 with (a0 − 1)(a2 b1 + a1 b2 ) = (a0 a1 − 1)a2 .
We shall use Orlik’s notation K(a0 , a1 , a2 ; I), . . . , K(a0 , a1 , a2 , b1 , b2 ; VII) when
referring to any link associated to these classes of polynomials. Note that the link
K(a0 , a1 , a2 ; I) = L(a0 , a1 , a2 ) is of Brieskorn-Pham type. All the other types (II-
VII) can be easily written as L∗ (w; d) links in the notation of Chapter 9. For
example, we have
µ ¶
d d d(a1 − 1)
K(a0 , a1 , a2 ; II) ∈ L∗ , , ;d ,
a0 a1 a1 a2
µ ¶
d d(1 − a2 ) d(1 − a1 )
K(a0 , a1 , a2 ; III) ∈ L∗ , , ;d .
a0 1 − a1 a2 1 − a1 a2
The 5-dimensional analogue of the link K(a0 , a1 , a2 ; II) will be revisited later in
this chapter (cf. Example 10.3.7).
Exercise 10.2: Compute L∗ (w; d) form of each link of Proposition 10.1.4. In each
case find I = |w| − d and determine all the links with a fixed sign of I, i.e., separate
out positive, null, and negative links. Note that there is some duplication in the
list of Proposition 10.1.4. For instance, since K(2, 2, 2; III) ∈ L∗ (3, 2, 2; 6) we easily
conclude that K(2, 2, 2; III) is equivariantly equivalent to the BP link K(2, 3, 3, I) ∈
L∗ (3, 2, 2; 6). Using the first part of this exercise find all such equivalences in the
list of Proposition 10.1.4.
The following conjecture of Milnor [Mil68] was proven by Orlik [Orl70]
Theorem 10.1.5: Let Lf ∈ L∗ (w0 , w)1, w2 ; d). Then
1. The fundamental group π1 (L) is infinite and the universal cover of L is
an open 3-cell if and only if d ≥ |w|.
2. π1 (L) is infinite nilpotent if and only if d = |w|.
Proof. To prove the first statement Orlik shows that π1 (L) is finite if and
only if |w| > d. If π1 (L) is finite, then the geometry of L is elliptic and L is covered
by S 3 . In this case it is well-known that π1 (L) is one of the finite subgroups Γ of
SU (2), namely:
(i) Γ = Zp the cyclic group of order p,
(ii) Γ = D∗m a binary dihedral group of order 4(m − 2), where m > 3,
(iii) Γ = T∗ the binary tetrahedral group,
(iv) Γ = O∗ the binary octahedral group,
1Actually in [OW71a] Orlik and Wagreich missed the last two cases. The correction appears
in [OW77].
332 10. SASAKIAN GEOMETRY IN DIMENSIONS THREE AND FIVE

(v) Γ = I∗ the binary icosahedral group.


Orlik [Orl70] shows that all of these geometries are realized by the links occurring
in Proposition 10.1.4 and that this exhausts all cases with |w| > d (cf. Exercise
10.2 for the I > 0 links). These are:
(i) Γ = Zp gives the lens spaces L(p, 1) = K(2, 2, p; I).
(ii) The binary dihedral group Γ = D∗m gives K(2, m − 1, 2; II).
(iii) The binary tetrahedral group Γ = T∗ gives K(2, 3, 4; I).
(iv) The binary octahedral group Γ = O∗ gives K(2, 3, 3; II).
(v) The binary icosahedral group Γ = I∗ gives K(2, 3, 5; I).
This proves the first part of (1), and the second statement of (1) follows from the
work of Waldhausen [Orl70, Wal67a]. It should be noted that only the lens spaces
of the form L(p, 1) occur. Proof of (2) is equally simple. The condition |w| = d
is very restrictive and any such link turns out to be equivariantly diffeomorphic to
the three BP links K(3, 3, 3; I), K(2, 3, 6; I), K(2, 4, 4; I) (cf. Exercise 10.2 for the
null links). One can easily calculate the fundamental groups for these three BP
links to check it is nilpotent [Orl70]. Actually, a few years later Milnor calculated
the fundamental group of all BP links L(a0 , a1 , a2 ) [Mil75]. ¤

Remark 10.1.1: It is well-known that the five classes given in the proof are the
five classes of finite subgroups Γ ⊂ SU (2) the universal cover of the rotation group
SO(3) and that they are related to the hypersurface singularities as given. The
links are all diffeomorphic to S 3 /Γ and they are not simply Sasakian but they
exhaust all possible compact 3-Sasakian manifolds in dimension 3 (see Chapter 13
for the definition and the classification). In particular, the cones on these links can
be identified with C2 /Γ ⊂ C3 , where the embedding is given by the vanishing of
the associated polynomial. These are precisely the classical Kleinian singularities.
Even more remarkable is that these are related to the Dynkin diagrams of certain
simple Lie algebras. The Kleinian singularities, their resolutions, their relation to
the Dynkin diagrams of type A-D-E and the representation theory of the discrete
subgroups of SU (2) will be discussed in Section 12.10. Here, we summarize all this
in the following table:
Diagram Subgroup |Γ| Polynomial
Ap−1 Zp p z0p + z12 + z22
Dm D∗m 4(m − 2) z02 z1 + z1m−1 + z22
E6 T∗ 24 z04 + z13 + z22
E7 O∗ 48 z03 + z13 z0 + z22
E8 I∗ 120 z05 + z13 + z22

Let us now specialize to the case of links L(a) = L(a0 , a1 , a2 ) of Brieskorn-Pham


polynomials which, after Milnor [Mil75], are usually called Brieskorn manifolds.
There are precisely four classes of elliptical (i.e., |w| > d) Brieskorn 3-manifolds,
namely
L(p, 2, 2), L(3, 3, 2), L(4, 3, 2), L(5, 3, 2).
Looking at the above table we see that only three of these explicitly appear; the
fourth one L(3, 3, 2) ≡ K(2, 2, 2; III) is equivalent to one on the list of Proposi-
tion 10.1.4 as pointed out in Exercise 10.2. There are precisely three Brieskorn 3-
manifolds in the Euclidean case (|w| = d), namely L(6, 3, 2), L(4, 4, 2), and L(3, 3, 3).
10.1. SASAKIAN GEOMETRY IN DIMENSION THREE 333

These are realized by arbitrary weighted homogeneous polynomials of degrees 6, 4


and 3 respectively, and they exhaust all possibilities that satisfy |w| = d. The table
below gives the weights, degrees, and number of monomials of the three polynomi-
als.

manifold weight vector degree # of monomials


M1 (1, 2, 3) 6 7
M2 (1, 1, 2) 4 9
M3 (1, 1, 1) 3 10

The remaining Brieskorn 3-manifolds are hyperbolic. Which links of an isolated


hypersurface singularities defined by weighted homogeneous polynomials in three
variables are homology 3-spheres have been characterized by Saeki [Sae87]. The
following easily follows from the Brieskorn’s graph Theorem 9.3.18.
Proposition 10.1.6: Let Lf denote the link of an isolated hypersurface singularity
defined by a weighted homogeneous polynomial f in three variables. Suppose further
that Lf represents a homology sphere, then Lf has the same knot type as the link
of a Brieskorn-Pham polynomial.
The problem of finding which Brieskorn manifolds are homology spheres can
be determined from Brieskorn’s Graph Theorem 9.3.18.
Proposition 10.1.7: The link L(a) is a homology sphere if and only if gcd(ai , aj ) =
1 for all i 6= j = 0, 1, 2.
It is clear from Theorem 10.1.5 and Propositions 10.1.6 and 10.1.7 that the only
non-trivial homology 3-sphere with finite fundamental group that can be realized as
a hypersurface link of a weighted homogeneous polynomial is the case K(2, 3, 5; I)
which is the famous Poincaré homology sphere. Furthermore, it is known that I∗
is the only non-trivial finite group that can occur as the fundamental group of an
integral homology 3-sphere [Sav02].
Example 10.1.8: [Poincaré Homology Sphere] This example has a very inter-
esting history. When Poincaré first made his famous conjecture he first conjectured
that any homology 3-sphere is homeomorphic to S 3 . He then constructed a coun-
terexample S 3 /I∗ using Felix Klein’s binary icosahedral group I∗ which is a perfect
group (its commutator subgroup [I∗ , I∗ ] is all of I∗ ) of order 120. Thus, π1 (S 3 /I∗ ) =
I∗ , whereas its Abelianization H1 (S 3 /I∗ , Z) ≈ π1 (S 3 /I∗ )/[π1 (S 3 /I∗ ), π1 (S 3 /I∗ )] =
0. He then rephrased his conjecture as “any homotopy 3-sphere is homeomorphic
to S 3 ”. The icosahedral group I is the group of isometries of a regular icosahedron
(dodecahedron), and the binary icosahedral group I∗ is the lift of I ⊂ SO(3) to its
universal cover SU (2). It has a presentation

I∗ = {(a, b) | (ab)2 = a3 = b5 } .

It is well-known that there are many 3-dimensional homology spheres, and all
but one, the Poincaré homology sphere have infinite fundamental group. According
Saeki’s theorem, Proposition 10.1.6, any link of an isolated hypersurface singularity
in three variables which represents a homology sphere is a Brieskorn link L(a).
Furthermore, Proposition 10.1.7 says that L(a) is a homology sphere if and only if
the components of a are relatively prime. Let us consider a simple example.
334 10. SASAKIAN GEOMETRY IN DIMENSIONS THREE AND FIVE

Example 10.1.9: Consider the Brieskorn polynomial


f = z07 + z13 + z22
whose zero locus Zf is a hypersurface of degree d = 42 in the weighted projective
space CP(6, 14, 21). We wish to study the Brieskorn manifold L(7, 3, 2). It has in-
finite fundamental group by Orlik’s Theorem 10.1.5, and is a homology 3-sphere
which is an orbifold fibration over CP1 with a nontrivial orbifold structure. L(7, 3, 2)
f
is the quotient of the universal cover SL(2, R) of SL(2, R) by a co-compact discrete
f
subgroup Γ ⊂ SL(2, R). There is a covering of L(7, 3, 2) by the total space of a
non-trivial circle bundle over a Riemann surface of genus g = 3.
Example 10.1.10: Consider the Brieskorn polynomial
f = z06k−1 + z13 + z22
whose zero locus Zf is a hypersurface of degree d = 6(6k − 1) in the weighted
projective space CP(6, 2(6k −1), 3(6k −1)). We wish to study the link L(6k−1, 3, 2).
By Orlik’s Theorem 10.1.5 π1 (L(6k − 1, 3, 2)) is infinite if and only if d − |w| =
6k − 7 ≥ 0. So there is one case with finite fundamental group, namely the link
L(5, 3, 2) = K(2, 3, 5; I) which represents the Poincaré homology sphere S 3 /I ∗ .
Thus, by Proposition 10.1.7, for all k > 1 the link L(6k − 1, 3, 2) represents a
homology 3-sphere with infinite π1 . This homology 3-sphere is an orbifold fibration
over CP1 with a non-trivial orbifold structure. By Milnor’s Theorem [Mil75] the
f
link L(6k − 1, 3, 2), for (k > 1), is the quotient of the universal cover SL(2, R) of
f
SL(2, R) by a co-compact discrete subgroup Γ ⊂ SL(2, R). Furthermore, L(6k −
1, 3, 2) has a finite covering by a manifold that is diffeomorphic to a non-trivial
circle bundle over a Riemann surface of some genus g > 1.
Next we give a brief discussion of how one can distinguish 3-dimensional inte-
gral homology spheres. This can be done by the Casson invariant λ which roughly
speaking counts the number of irreducible representations of π1 in SU (2). More
explicitly for a Brieskorn homology sphere L(a0 , a1 , a2 ) Fintushel and Stern proved
that λ(L(a0 , a1 , a2 )) equals − 21 times the number of conjugacy classes of irre-
ducible representations of π1 (L(a0 , a1 , a2 )) into SU (2). Furthermore, they showed
how λ(L(a0 , a1 , a2 )) can be computed from the Hirzebruch signature τ (V ) of the
parallelizable manifold V (a0 , a1 , a2 ) whose boundary is the link L(a0 , a1 , a2 ). Ex-
plicitly,
τ (V (a0 , a1 , a2 ))
λ(L(a0 , a1 , a2 )) = .
8
One can easily compute the signature of V (6k − 1, 3, 2) using formula (9.4.2) to
find that it equals −8k, giving λ(L(6k − 1, 3, 2)) = −k. Thus, L(6k − 1, 3, 2) give
an infinite sequence of inequivalent homology spheres. Moreover, it is known that
λ(L(7, 3, 2)) = −1, and since k = 1 from the previous sequence is the Poincaré
homology sphere, which is clearly distinct, we see that L(7, 3, 2) is also distinct
from the others. For further discussion and references we refer the reader to the
recent book [Sav02].
Exercise 10.3: One can easily see that L(6k + 1, 3, 2) is also an infinite sequence
of homology spheres. Note that from the Hirzebruch signature computations the
Casson invariants λ(L(6k − 1, 3, 2)) = λ(L(6k + 1, 3, 2)) = −k. Yet, as remarked
L(5, 3, 2) and L(7, 3, 2) are completely different though they are both homology
spheres. What can you say about L(6k − 1, 3, 2) and L(6k + 1, 3, 2), for k > 1?
10.2. SASAKIAN STRUCTURES AND THE TOPOLOGY OF 5-MANIFOLDS 335

All regular null Sasakian structures on a compact manifold can be determined


from the work of Milnor [Mil75]. These are realized as quotients of the three-
dimensional Heisenberg group H3 (R) over R by the three-dimensional integral
Heisenberg group H3 (Z), cf. Example 8.1.12. Up to covering these are all non-
trivial circle bundles over a complex torus, and it is known [Mil75] that all such
bundles can be constructed from the three-dimensional Heisenberg group H3 (R)
which be represented as a subgroup of GL(3, R) by considering matrices of the
form
 
1 x z
(10.1.1) A =  0 1 y  ; x, y, z ∈ R .
0 0 1
We also define subgroups H3 (k, Z) ⊂ H3 (Z) of matrices for which x, y, z are integers
divisible by some positive integer k. It is easy to see that the quotient manifold
Mk = H/Hk is a circle bundle over a torus with Chern number ±k. The fibration
π : Mk −→T 2 is explicitly given by
(10.1.2) π(A) = (x mod k, z mod k) .
Now from example 7.3.21 the Sasakian structure S = (ξ, η, Φ, g) on H3 (R) is defined
by the formulas
½ ∂
η = dz − y dx, ξ = ∂z , g = dx2 + dy 2 + (dz − y dx)2
(10.1.3) ∂ ∂ ∂
Φ = ( ∂x + y ∂z ) ⊗ dy − ∂y ⊗ dx.
Now one can easily check that the tensor fields defining S are invariant under the
action of H3 (R) on itself. Hence, S defines a Sasakian structure on the compact
quotient manifolds Mk . Milnor [Mil75] observed that the integral homology can
easily be computed, viz.
H1 (Mk , Z) = Hk /[Hk , Hk ] = Z ⊕ Z ⊕ Zk .
However, as indicated by the table 2 above only the cases k = 1, 2, 3 can be realized
as links of isolated hypersurface singularities of weighted homogeneous polynomials.

10.2. Sasakian Structures and the Topology of 5-Manifolds


To begin our study of Sasakian structures on 5-manifolds, we would like to
review basic results concerning smooth simply connected 5-manifolds. Dimension
5 provides a quite useful venue to study and understand the different classes of
Sasakian manifolds. For the most part, although not entirely, we restrict our con-
siderations to simply connected 5-manifolds. The main reason for doing so is an
important theorem of Markov [Mar58, Mar60] which states
Theorem 10.2.1: The class of compact smooth orientable n-manifolds, n > 4, is
not classifiable under diffeomorphism, combinatorial equivalence, homeomorphism
or homotopy type.
Every such manifold M is given by a triangulation. A classification is under-
stood to mean finding all classes of pairwise inequivalent manifolds and then giving
a finite algorithm which, from the triangulation of M , would determine the class
M belongs to. It can be shown that even the question of simple connectivity of M
cannot be decided by a finite algorithm. Surprisingly perhaps, if we restrict our
attention to simply connected 5-manifolds the full classification is not only possible,
336 10. SASAKIAN GEOMETRY IN DIMENSIONS THREE AND FIVE

but also it is very simple: the complete set of invariants consist of H2 (M, Z) and
one other invariant i(M ) introduced in [Bar65] and defined below.
Theorem/Definition 10.2.2: Let M be a compact, smooth, oriented, 1-connected
5-manifold (or Smale-Barden manifold throughout this section). Let us write
H2 (M, Z) as a direct sum of cyclic groups of prime power order
M¡ ¢c(pi )
(10.2.1) H2 (M, Z) = Zk ⊕ Zpi ,
p,i

where k = b2 (M ), and c(p ) = c(p , M ). The non-negative integers k, c(pi ) are


i i

determined by H2 (M, Z) but the subgroups Zpi ⊂ H2 (M, Z) are not unique. One
can choose the decomposition (10.2.1) such that the second Stiefel-Whitney class
map
w2 : H2 (M, Z) → Z2
is zero on all but one summand Z2j . The value j is unique, denoted by i(M ),
and called the Barden invariant of M . It can take on any value j for which
c(2j ) 6= 0, besides 0 and ∞. Alternatively, i(M ) is the smallest j such that there
is an α ∈ H2 (M, Z) such that w2 (α) 6= 0 and α has order 2j .
The following theorem was proved by Smale [Sma62a] in the spin case and later
generalized by Barden [Bar65] without the spin assumption. We shall formulate
it using Barden’s notation and refer to it as the Smale-Barden Theorem.
Theorem 10.2.3: The class B of simply connected, closed, oriented, smooth, 5-
manifolds is classifiable under diffeomorphism. Furthermore, any such M is diffeo-
morphic to one of the spaces
(10.2.2) Mj;k1 ,...,ks = Xj #Mk1 # · · · #Mks ,
where −1 ≤ j ≤ ∞, s ≥ 0, 1 < k1 and ki divides ki+1 or ki+1 = ∞. A
complete set of invariants is provided by H2 (M, Z) and i(M ) and the manifolds
X−1 , X0 , Xj , X∞ , Mj , M∞ are characterized as follows
M H2 (M, Z) i(M )
X−1 = SU (3)/SO(3) Z2 1
M1 = X0 = S 5 0 0
Xj , 0 < j < ∞ Z2 ⊕ Z2j
j j
X∞ Z ∞
Mk , 1 < k < ∞ Zk ⊕ Zk 0
M∞ = S 2 × S 3 Z 0

We refer to B as the class of Smale-Barden manifolds or B-manifolds for short,


and define the subclass B0 to be the subclass of M ∈ B such that i(M ) = 0.
Elements of B 0 are called Smale manifolds, and this is precisely the subclass in B
that admit a spin structure, i.e., when w2 (M ) = 0. It is understood that s = 0
means that no Mki occurs. Note that X∞ is the non-trivial S 3 bundle over S 2 . With
the exception of X1 = X−1 #X−1 , the manifolds Xj ’s are indecomposable under
the connected sum operation. On the other hand M0,k1 ,...,ks in the decomposition
10.2.2 is given in terms of invariant factors k1 |k2 | · · · |ks ; equivalently, this can be
written in terms of elementary divisors pi ’s such as in the decomposition 10.2.1. For
example, there are four inequivalent Smale manifolds with ord(H2 (M, Z)) = 362 .
They can be written in terms of either invariant factors or elementary divisors.
10.2. SASAKIAN STRUCTURES AND THE TOPOLOGY OF 5-MANIFOLDS 337

The equalities of the two types are M0,36 = M36 = M4 #M9 with s = 1 and as
M0,2,18 = M2 #M18 = 2M2 #M9 , M0,3,12 = M3 #M12 = 2M3 #M4 , and M0,6,6 =
M6 #M6 = 2M2 #2M3 with s = 2. As usual, we write nM to denote the n-fold
connected sum of M with itself for n ≥ 1. It is also convenient to use the convention
0M = S 5 . Note that for 1 < j < ∞, the rational homology 5-spheres M2j and Xj
have the same H2 (M, Z) but i(M2j ) = 0 while i(Xj ) = j. Likewise, M∞ and X∞
(as S 3 -bundles over S 2 ) are distinguished by their Barden invariants. We will be
interested in the following further special classes of Smale-Barden manifolds:
Definition 10.2.4: Let B be the class of all Smale-Barden manifolds. We define
the following subclasses of B with obvious further inclusions
(i) SFBR ⊆ SFB ⊆ B: B-manifolds admitting a free smooth S 1 -action; (a
Seifert fibered S 1 -structure),
(ii) ACBQR ⊆ ACB ⊆ B: B-manifolds admitting a quasi-regular almost contact
structure (an almost contact structure),
(iii) CBR ⊆ CBQR ⊆ ACBQR ⊆ CB ⊆ B: B-manifolds admitting a regular contact
structure (a quasi-regular contact structure, a contact structure),
(iv) SBR ⊆ SBQR ⊆ SB ⊆ KB ⊆ CBQR ⊆ B: B-manifolds admitting regular
Sasakian structure; (a quasi-regular Sasakian structure, a Sasakian struc-
ture, a K-contact structure),
(v) SB+R ⊆ SB+QR ⊆ SB+ ⊆ SB ⊆ B: B-manifolds admitting a (regular, quasi-
regular) positive Sasakian structure; ( SB−R ⊆ SB−QR ⊆ SB− ⊆ SB ⊆ B,
SBN R ⊆ SBN QR ⊆ SBN ⊆ SB ⊆ B for negative and null Sasakian struc-
tures),
(vi) SE R ⊆ SE QR ⊆ SEB ⊆ SB+ ⊆ B: B-manifolds admitting a (regular,
quasi-regular) Sasaki-Einstein structure.
For all of these subclasses of B we denote the corresponding subclass of Smale
manifolds by a superscript 0, for example the class of Seifert fibered Smale manifolds
is denoted by SFB0 . and the regular Seifert fibered Smale manifolds by (SFBR )0 .
We begin by mentioning some equalities that follow directly from previous
results.
Proposition 10.2.5: The following equalities hold:
KB = CBQR , SBQR = SB , SB+QR = SB+ , SB−QR = SB− , SBN QR = SBN .

Proof. The first equality follows from Proposition 7.1.2 while the remaining
4 are a direct consequence of the Approximation Theorem 7.1.10. ¤

Recall from Theorem 8.1.14 that in the last two cases a much stronger result
holds, namely that any such Sasakian structure is quasi-regular.
The inclusions SFBR SFB B which are completely understood. First
concerning the inclusion SFBR SFB , it is easy to see that a Smale-Barden manifold
M admits a free circle action if and only if H2 (M, Z) is torsion free (cf. [Gei91,
Prop.10]), and we discuss the regular case further in Section 10.4 below.
The inclusion SFB B is much more subtle. Recall from the previous section
that an existence of a Seifert fibration on a compact oriented 3-manifold imposes
a very strong topological restriction: Only six models of the eight Thurston Ge-
ometries are Seifert fibered spaces and, as we have seen, three are Sasakian. Like-
wise, in the case of the Smale-Barden manifolds, the existence of a Seifert fibered
338 10. SASAKIAN GEOMETRY IN DIMENSIONS THREE AND FIVE

S 1 -structure must impose severe restrictions on the pair {H2 (M, Z), i(M )}. The
Barden invariant i(M ) is forced to assume only three possible values. In addition,
the size of k = b2 (M ) imposes restriction on the structure of the torsion group, the
lower the b2 (M ) the more restrictive the condition. More precisely, Kollár [Kol06a]
proves
Theorem 10.2.6: Let b2 (M ) = k and write H2 (M, Z) as in equation (10.2.1).
Then M ∈ SFB if and only if
(i) i(M ) ∈ {0, 1, ∞},
(ii) for every prime p #{i : c(pi ) > 0} ≤ k + 1,
(iii) i(M ) = ∞ ⇒ #{i : c(2i ) > 0} ≤ k.
We will sometimes refer to conditions (i-iii) of Theorem 10.2.6 as “Condition K”.
Actually, Kollár observes that the decomposition 10.2.1 and the Barden invariant
i(M ) can be considered also for certain non-simply connected 5-manifolds. He
obtains a slightly more general result: a compact oriented 5-manifold M with
H1 (M, Z) = 0 admits a Seifert fibered S 1 -structure only if the conditions of the
Theorem 10.2.6 are satisfied. Under the π1 (M ) = 0 assumption this is also a
sufficient condition. Theorem 10.2.6 is not a classification of all Seifert fibered S 1 -
structures on Smale-Barden manifold. For rational homology spheres k = b2 (M ) =
0 and we get
Corollary 10.2.7: Let b2 (M ) = 0. Then M ∈ SFB if and only if i(M ) = 0, 1 and
for every prime p #{i : c(pi ) > 0} ≤ 1.
Thus, for example, both M3 and M9 admit Seifert fibered S 1 -structures while
M3 #M9 does not. It should be noted that Kollár’s proof does not give a classi-
fication of all Seifert fibered S 1 -structures on any given Smale-Barden 5-manifold.
In fact, one can exhibit infinitely many topologically distinct fixed point free S 1 -
actions on every M as in (10.2.6). In principle, the classification of all S 1 -actions
on 5-manifolds is reduced to a question on 4-dimensional cyclic orbifolds, but the
4-dimensional problem is rather delicate.
The proof of Theorem 10.2.6 is quite involved and it can be found in [Kol06a].
The main ideas follow [Sei32, OW75] as one considers any compact smooth ori-
ented 5-manifold with a fixed point free circle action and takes the quotient space
Z =P M/S 1 . Z is a compact orbifold with an orbifold structure (Z, ∆), where
∆ = (1 − m1i )Di is a formal sum of codimension 2 closed subspaces Di ⊂ Z.
When (Z, ∆) is an algebraic orbifold surface, ∆ is the branch divisor defined in
Definition 4.4.8, but here one must work without any complex analytic assump-
tions on Z. The main technical aspect of the proof is relating the invariants of
M to the invariants of the topological Seifert bundle M −→(Z, ∆). In particular,
Kollár first shows that the conditions (i-iii) of Theorem 10.2.6 are necessary. After
that, for each i(M ) P= 0, 1, ∞ separately he gives an explicit construction of a Seifert
bundle M −→(CP2 , (1 − m1i )Di ) with the prescribed allowable homology group.
Remark 10.2.1: The classification of circle actions on 4-manifolds was carried out
by Fintushel [Fin77, Fin78]. In this case the quotient is a 3-manifold (with bound-
ary corresponding to the fixed points) endowed with additional data involving links
and certain weights. Foundational questions concerning circle and torus actions on
orbifolds were also considered in [HS91].
We next turn our attention to the question when a Seifert fibered S 1 -structure
on a Smale-Barden manifold is associated to an (almost) contact structure. First
10.2. SASAKIAN STRUCTURES AND THE TOPOLOGY OF 5-MANIFOLDS 339

note, that in the realm of Smale-Barden manifolds Gray’s result, Corollary 6.2.8
translates as follows
Theorem 10.2.8: M ∈ ACB if and only if i(M ) = 0, ∞.
This theorem completely explains the inclusion ACB B. The simplest exam-
ple of a manifold which has a locally free S 1 -action but no almost contact structure
is the symmetric space X−1 = SU (3)/SO(3), also known as the Wu manifold
[Bar65]. As we are ultimately interested in those Smale-Barden manifolds ad-
mitting various Sasakian structures, we need to know, in particular, which such
manifolds admit a contact structure. Geiges showed that there are no further ob-
structions beyond the U (2)-reduction obstruction of Gray [Gei91], i.e., ACB = CB .
More precisely
Theorem 10.2.9: M ∈ ACB if and only if M 5 = Xj #Mk1 # · · · #Mks , where
j = 0, ∞, s ≥ 0, 1 < k1 and ki divides ki+1 or ki+1 = ∞. Furthermore, each such
M ∈ ACB admits a contact structure in every homotopy class of almost contact
structures.
Comparing the theorems of Geiges and Kollár we note that infinitely many
Smale-Barden manifolds cannot support any quasi-regular contact structure. The
question of which Seifert fibered S 1 -structures are compatible with some quasi-
regular (almost) contact structure is more difficult. In such cases the transverse
geometry is that of a compact symplectic 4-orbifold.
Definition 10.2.10: Let M ∈ B with b2 (M ) = k and write H2 (M, Z) as in 10.2.1.
We say that M satisfies condition G-K if the pair (H2 (M, Z}, i(M )) fulfills all of
the following
(i) i(M ) ∈ {0, ∞},
(ii) for every prime p #{i : c(pi ) > 0} ≤ k + 1,
(iii) i(M ) = ∞ ⇒ #{i : c(2i ) > 0} ≤ k.
Alternatively, M ∈ B satisfies condition G-K if and only if M ∈ CB ∩ SFB .
Note that condition G-K is necessary for M ∈ ACBQR . Unfortunately, we do not
know whether it is also sufficient. Hence, we arrive at the following two fundamental
questions
Question 10.2.1: Suppose M ∈ B satisfies the condition G-K.
(i) Does M admit a quasi-regular almost contact structure, i.e., what is the
inclusion ACBQR ⊆ CB ∩ SFB ?
(ii) Does M admit a quasi-regular contact structure, i.e., what is the inclusion
CBQR ⊆ ACBQR ⊆ CB ∩ SFB ?
Note that these are indeed two separate problems: if M admits a quasi-regular
almost contact structure the deformed contact structure (which must exist) does
not itself need be quasi-regular. Alternatively, we are interested in those elements of
CB ∩SFB such that the contact structure and Seifert fibered structure are compatible
in the sense that the S 1 action of the Seifert fibration is a Reeb vector field for a
1-form representing the contact structure.
The importance of the above question has to do with the following simple
observation: Any quasi-regular contact manifold admits a compatible K-contact
structure by Proposition 7.1.2. Also, any K-contact structure can be approximated
by a sequence of quasi-regular K-contact structures by Theorem 7.1.10. Restricting
to Smale-Barden manifolds gave us the first equality in Proposition 10.2.5, KB =
340 10. SASAKIAN GEOMETRY IN DIMENSIONS THREE AND FIVE

CBQR . To sum up, we get the following sequence of inclusions


(10.2.3) SB ⊆ KB = CBQR ⊆ ACBQR ⊆ CB ∩ SFB
which highlights the importance of condition G-K as a necessary condition for the
existence of any Sasakian structures.
Corollary 10.2.11: Let M ∈ B. The condition G-K is necessary for the existence
of any K-contact and, hence, a Sasakian structure on M .
It is disappointing that very little is known about all the inclusions in (10.2.3).
In particular, we do not know if SB ⊆ KB is proper. (However, recall Example
7.4.16 which gives a non-simply connected K-contact 5-manifold that admits no
Sasakian structure). This gives rise to
Open Problem 10.2.1: Find a Smale-Barden manifold that admits a K-contact
structure but does not admit a Sasakian structure.
Kollár offers an important glimpse into the question of which Smale-Barden
manifolds admit Sasakian structures. He shows that the existence of a Sasakian
structure can be further obstructed [Kol06a], i.e., the inclusion SB ⊆ CB ∩ SFB is
proper. Perhaps this should not be surprising for in the quasi-regular Sasakian case
the transverse geometry is algebraic, i.e., the space of leaves Z is a Hodge orbifold.
In particular, Di ⊂ Z must themselves be complex algebraic curves. The extra
condition leads to finer holomorphic obstructions as illustrated by the following
example [Kol06a].
P
Proposition 10.2.12: Let M → (Z, (1 − m1i )Di ) be a Seifert bundle over an
algebraic orbifold Z. Assume that H1 (M, Z) = 0 and
M¡ ¢c(pi )
H2 (M, Z) = Zpi ,
p,i

i.e., M is a rational homology sphere. Then there is a degree 2 polynomial q with


integer coefficients such that the set q(Z) contains all but 10√elements of the set
{c(pi )}. In particular, the set {c(pi )} contains at most 12 + 2 N elements of any
interval of length N.
Proof. If Z is an algebraic surface with quotient singularities, the Bogomolov-
Miyaoka-Yau-type inequalities of [Kol92, 10.8, 10.14] and [KM99, 9.2] imply that
P ¡ 1
¢
x 1 − rx < e(Z),
where rx is the order of the fundamental group of the link of a singular point x ∈ Z
and e(Z) is the topological Euler number. In our case e(Z) = 3 and so Z has at
most 5 singular points. Kollár shows that there are at most 10 complex algebraic
curves Di which pass through a singular point. Once Di is contained in the smooth
locus of Z, its genus is computed by the adjunction formula
2g(Di ) = (Di · (Di + KZ )) + 2 .
Since Z has Picard number 1, we conclude that there is a degree 2 polynomial with
integer coefficients q(t) such that 2g(Di ) ∈ q(Z). The statement follows now from
the following lemma:
Lemma 10.2.13: Let q(t) = at2 + bt + c withp a > 0. Then the set q(Z) intersects
every interval of length N in at most 2 + 2 N/a elements.
¤
10.2. SASAKIAN STRUCTURES AND THE TOPOLOGY OF 5-MANIFOLDS 341

The conditions of Proposition 10.2.12 are very restrictive as can be seen by the
following example.
P
Example 10.2.14: Let the pi be distinct primes and M → (Z, (1 − m1i )Di ) a
Seifert bundle over an algebraic orbifold such that H1 (M, Z) = 0 and
k
M ¡ ¢2i
H2 (M, Z) = Zpi .
i=1

Then (10.2.12) implies that k ≤ 23. On the other hand, the conditions of (10.2.6)
are satisfied for any k, and so there are such Seifert bundles for any k and pi (even
over Z = CP2 ). The surfaces Di , however, can not be chosen complex algebraic for
k > 23.
Corollary 10.2.15: There exists infinitely many distinct rational homology spheres
in CB ∩ SFB which do not admit any Sasakian structure.
Other than the above example, generally we cannot say much about the inclu-
sions SB ⊆ KB ⊆ CB ∩ SFB . On the other hand it follows that either the first, or
the second, or both inclusions are proper.

10.2.1. Positive Sasakian Structures. We now focus our attention on pos-


itive Sasakian manifolds SB+ CB ∩ SFB , the importance of which stems from
Theorem 9.5.1 which implies that any M ∈ SB+ admits a Sasakian metric with
positive Ricci curvature. In the positive case one can actually get much more infor-
mation. The point is that, by definition, when M ∈ SB+ we get Seifert S 1 -structures
M −→(Z, ∆) with corb 1 (Z, ∆) > 0. Hence, in such cases the base (Z, ∆) is a compact
del Pezzo orbifold surface with only cyclic orbifold singularities. The positivity as-
sumption, although not as strong as in the smooth case, still imposes restrictions on
the type of Seifert fibrations that can occur. These results are again due to Kollár
[Kol05, Kol06b] and the techniques used to prove them are ensconced deeply in
the minimal model program. Since the authors are not experts in this area, we
shall state the relevant results, and in many cases only sketch some ideas of the
proofs, referring to the literature for details.
We begin by sketching the main idea in applying the minimal model program.
One begins with a log del Pezzo surface (S, ∆) and obtains a minimal resolution S 0
by a sequence of blow ups. Then one obtains a minimal model S min by a sequence of
blow downs. S min is a minimal rational surface, and there are a limited number of
possibilities, namely it is either CP2 , CP1 ×CP1 , or one of the Hirzebruch surfaces Sn
with n ≥ 2. At each intermediate step one obtains a surface T whose anticanonical
divisor satisfies −KT ≡ ∆T + HT , where HT is big and nef (see Definition 3.5.23)
and ∆T is the proper transform of ∆. From Theorem 4.7.14 we know that the
torsion in H2 is determined by the non-rational curves in the branch divisor ∆ of
S, and it is not difficult to see that there is at most one such curve. So we can
1
write ∆ = (1 − m )D0 + ∆0 , where D0 is non-rational, and the remainder ∆0 of the
branch divisor consists of at most rational curves. A case by case analysis shows
that there are constraints between the ramification index m and the genus g of the
non-rational curve D0 . Let D00 denote the birational transform of D0 , E the section
of Sn of self-intersection −n, and F the fibre of Sn . Then the following possibilities
occur [Kol06b]:
342 10. SASAKIAN GEOMETRY IN DIMENSIONS THREE AND FIVE

(1) If m ≥ 6 and g ≥ 1 then S min is CP2 , CP1 × CP1 , or S2 and g = 1 with


D00 ∈ | − KS min |.
(2) If m ≥ 5 and g ≥ 2 then m = 5, S min is S3 and g = 2 with D00 ∈ |2E+6F |.
(3) If m ≥ 3 and g ≥ 4 then m = 3, S min is S5 and g = 4 with D00 ∈
|2E + 10F |.
(4) If m = 4 and g ≥ 2 then either S min is S3 and g = 2 with D00 ∈ |2E + 6F |
or S min is S2 and g = 2 with D00 ∈ |2E + 5F |.
Using this analysis it is straightforward to prove the following lemma which we
shall make use of in several places below. Details can be found in (6.8) and (6.13)
of [Kol05].
Pn
Lemma 10.2.16: Let (S, ∆ = i=0 ai Di ) be a log del Pezzo surface. Assume that
D0 has genus ≥ 1 and a0 ≥ 1/2. Then Di is rational for i ≥ 1 and
(i) g(D0 ) = 1 if a0 ≥ 5/6,
(ii) g(D0 ) ≤ 2 if a0 ≥ 3/4,
(iii) g(D0 ) ≤ 4 if a0 ≥ 2/3.
Assume in addition that a0 ≥ 11/12 and ai ≥ 1/2 for every i. Then (S, ∆) is a log
del Pezzo surface with Du Val singularities and ∆ = a0 D0 . The Picard number of
S is at most 9.
We recall the following definition from [KM98]. Let (0, X) be a normal surface
singularity with a minimal resolution f : Y −−→X. Then (0, X) is a Du Val Singu-
larity if for every exceptional curve Ei ⊂ Y we have KY · Ei = 0. A normal surface
X with Du Val singularities is characterized by the fact that KX is Cartier.
The first remarkable theorem of Kollár [Kol05] says that only very few groups
can occur as torsion subgroups of H2 (M, Z) of a simply connected 5-manifold that
admits a positive Sasakian structure. Recall from Theorem 4.7.14 that torsion
in H2 (M, Z) is a consequence of the existence of a non-rational component of the
branch divisor. Then the point of Lemma 10.2.16 is that the positivity of corb
1 forces
the existence of at most one non-rational component such that both the genus and
ramification index cannot be large, severely restricting the possibilities. We call
such torsion groups admissible. Explicitly,
Theorem 10.2.17: Suppose M ∈ SB+ . Then the torsion subgroup of H2 (M, Z) is
one of the following:
(i) (Zm )2 for any m ∈ Z+ ,
(ii) (Z5 )4 ,
(iii) (Z4 )4 ,
(iv) (Z3 )4 , (Z3 )6 or (Z3 )8 ,
(v) (Z2 )2n for any n ∈ Z+ .
Conversely, for each finite group G in the list above, there is an M ∈ (SB+ )0 with
H2 (M, Z)tors = G. (Note that m = 1 corresponds to zero torsion).

Proof. By the Approximation Theorem 7.1.10, M admits P ¡ a Seifert


¢ bundle
structure over a log del Pezzo orbifold (S, ∆), where ∆ = k 1 − m1i Di , and by
Theorem 4.7.14 if the torsion of H2 (L, Z) is non-trivial at least one of the curves
Di must be non-rational. Then the first part of Lemma 10.2.16 says that all the
other curves Di for i 6= 0 must be rational, and the list of possible torsion groups
follows from (i)-(iii) of the lemma.
10.2. SASAKIAN STRUCTURES AND THE TOPOLOGY OF 5-MANIFOLDS 343

The converse can be proved easily by example. For instance, the Brieskorn-
Pham link L(3, 3, 3, 3k) has homology H2 (L, Z) = Z6 ⊕ Zk ⊕ Zk for any k which
gives examples for (i) (including the case of no torsion which is k = 1). All other
torsion groups in (ii)-(v) can be realized as H2 (M, Z) of a rational homology sphere
in (SB+ )0 and they are given in the discussion of Theorem 10.2.19 below. ¤
Remark 10.2.2: It is quite remarkable that all possible torsion groups for simply
connected 5-manifolds admitting positive Sasakian structures can be realized by
links of weighted homogeneous polynomials. Notice that we do not claim that
every simply connected 5-manifold admitting a positive Sasakian structure can be
realized as the link of weighted homogeneous polynomials. This is clearly false as
X∞ shows. See also the table in Theorem 10.2.25 below.
One can be more precise about possible torsion when M ∈ B is a rational ho-
mology sphere. When M ∈ SB+ , in addition we have that π1 (M ) = π1orb (Z) = 0,
and corb
1 (Z) > 0. The first results concerning the existence of positive Sasakian
structures on rational homology 5-spheres were obtained by the authors in [BG02]
(with erratum [BG06b]) where it was shown that infinitely many rational ho-
mology 5-spheres admit positive Sasakian structures. However, since then Kollár
[Kol05] has obtained a complete classification of all rational homology spheres in
SB+ together with a complete classification of Seifert fibered structures in certain
special cases [Kol05]. First we need
Theorem 10.2.18: There is a one-to-one correspondence between
(i) Seifert bundle structures f : M → Z = (S, ∆) on 5-dimensional, compact
rational homology spheres with H1 (M, Z) = 0, and
(ii) compact, complex, 2-dimensional, cyclic orbifolds Z = (S, ∆) with H2 (S, Q) =
Q and H1orb (Z) = 0.
Under this correspondence, π1 (M ) = π1orb (Z).
The proof of this theorem is fairly straightforward and left to the reader, who
may consult [Kol05] for details. We are now ready for our main classification
theorem for rational homology spheres.
Theorem 10.2.19: Let M ∈ SB+ and b2 (M ) = 0. Then M ∈ (SB+ )0 , and
(i) if H2 (M, Z) contains a torsion element of order at least 12 then H2 (M, Z) =
(Zm )2 for some m not divisible by 30. Furthermore, the number of Seifert
bundle structures varies between 1 and 4, depending on m mod 30,
(ii) 2M5 has a unique family of positive Sasaki-Seifert structures,
(iii) 4M3 has a unique family of positive Sasakian Seifert structures,
(iv) 2M4 has exactly 2 families of positive Sasaki-Seifert structures,
(v) nM2 admits infinitely many families of positive Sasaki-Seifert structures
for each n ≥ 0,
(vi) the manifolds 2M3 , 3M3 , and Mm , with 2 < m < 12 all admit families of
positive Sasaki-Seifert structure. Moreover, M3 , M5 and 2M3 all admit
infinitely many positive Sasaki-Seifert structures.
In particular, with the exception of (Z30n )2 , all torsion groups of Theorem 10.2.17
appear as H2 (M, Z) for some M ∈ (SB+ )0 with b2 (M ) = 0.
Proof. First, we notice that M ∈ (SB+ )0 by Theorem 10.2.9, and second we
notice that H2 (M, Z) must be one of the groups listed in Theorem 10.2.17. Again
from the Approximation Theorem 7.1.10 we know that M must have a Seifert
344 10. SASAKIAN GEOMETRY IN DIMENSIONS THREE AND FIVE

S 1 -bundle structure M −−→(S, ∆), and by Theorem 10.2.18 there is a one-to-one


correspondence between these and their cyclic orbifolds. Using this we now proceed
to check each of the six assertions. For (i) we apply the second part of Lemma
10.2.16 to conclude that S is a log del Pezzo surface with Du Val singularities.
But such surfaces with Picard number 1 have been classified [Fur86, MZ88], and
described in Proposition 6.5 in [Kol05]. Under the m ≥ 12 restriction there are
basically only four possibilities for (Z, ∆):
1
1. Z = (S, ∆) = (CP2 , (1− m )D), where D is a smooth cubic and gcd(m, 3) =
1.
1
2. Z = (S, ∆) = (Q, (1 − m )D), where Q ⊂ CP3 is a quadric cone and
D is its intersection with a quadric, and gcd(m, 2) = 1. As an orbifold,
Q = CP(1, 1, 2).
1
3. Z = (S, ∆) = (CP(1, 2, 3), (1 − m )D), where D ∈ |O(6)| is smooth. [D]
is 6 times the generator in Cl(Z) and gcd(m, 6) = 1.
1
4. Z = (S, ∆) = (S5 , (1 − m )D) with an embedding S5 ⊂ CP5 as a degree
5 surface and D is a smooth hyperplane section. [D] is 5 times the
generator in Cl(Z) and gcd(m, 5) = 1.
In all four cases the curve D is a smooth member, of genus one, of the linear
system | − KZ |, and the Seifert bundles can be represented by links of weighted
homogeneous polynomials. For each such polynomial one needs to check that the
branch divisor has the correct ramification index, and then compute the genus of
the corresponding curve using the genus formula 4.6.5. In the first three cases
they are BP Links listed in the tables of Example 10.3.3 below: L(3, 3, 3, k) with
(k, 3) = 1, L(2, 4, 4, k) with (k, 2) = 1, and L(2, 3, 6, k) with (k, 6) = 1. In the last
case the link is given in Example 10.3.7 and is the link of f (z) = z06 + z13 + z23 + z0 z3k
with (k, 5) = 1. This is a weighted homogeneous polynomial of degree d = 6k with
weights w = (k, 3k, 2k, 5) and index I = 5. So the associated link Lf is positive
and one can show it is Mk . The list of four cases above shows that m cannot be
divisible by 30 and proves the statement in (i).
The manifold 2M5 is the Brieskorn-Pham link L(2, 5, 6, 6) and the uniqueness
of the Seifert fibration is shown in [Kol05]. This proves (ii). To establish (iii)
we observe that the manifold 4M3 is the Brieskorn-Pham link L(2, 3, 10, 10) and
the uniqueness of the Seifert fibration can be argued similarly [Kol06b]. The
case (S, ∆) = (S5 , (1 − 61 )D) gives M6 and it can be realized by the link of the
weighted homogeneous polynomial f (z) = z06 + z12 + z23 + z0 z36 . The manifolds Mm
for 2 < m < 12 with m 6= 6 can be all obtained as Brieskorn-Pham links (see
the discussion of positive BP links in Section 10.3 and Table B.4.1 in Appendix
B). Likewise, 2M3 = L(2, 3, 7, 14), 3M3 = L(2, 3, 8, 8). This proves (vi). To show
(v) consider f (z) = z02 z1 + z12n+1 + z22n+1 + z32 . This is a weighted homogeneous
polynomial of degree d = 4n + 4 with weights w = (2n, 2, 2, 2n + 1) and of index
I = 1. Hence, Z4n+2 ⊂ CP(2n, 2, 2, 2n + 1) is log del Pezzo with the associated
positive link Lf . One can show that H2 (Lf , Z) = (Z2 )2n . Lastly, to establish (iv),
take f (z) = z05 + z15 + z0 z22 + z34 . This is a homogeneous polynomial of degree d = 20
with weights w = (4, 4, 8, 5) and index I = 1. Thus the associated link Lf is positive
and one can show that H2 (Lf , Z) = (Z4 )4 . The second Seifert fibration comes
from the polynomial f (z) = z010 + z15 + z22 + z0 z34 . This is a weighted homogeneous
polynomial of degree d = 40 with weights w = (4, 8, 9, 20) with index I = 1. Kollár
shows there are no other positive Seifert fibered structures in this case [Kol06b].
10.2. SASAKIAN STRUCTURES AND THE TOPOLOGY OF 5-MANIFOLDS 345

On the other hand infinitely many positive Seifert fibrations on M2 , M3 , M5


and 2M2 can be realized by BP links of Example 10.3.3. In the case of nM2
k(2n+1)
for n > 0 one can take the link of f (z) = z0 + z12n+1 + z1 z22 + z32 with
gcd(k, 2n(2n + 1)) = 1 to obtain an infinite number of positive Sasakian structures
on nM2 . For the n = 0 case, i.e., S 5 , the link L(2, 2, p, q) gives infinitely many
positive Sasakian structures. ¤
Theorem 10.2.19 says that the problem of determining which simply connected
rational homology 5-spheres admit positive Sasakian structures is completely un-
derstood. We succinctly summarize this as
Corollary 10.2.20: Let M be a simply connected rational homology 5-sphere.
Then M admits a positive Sasakian structure if and only if M is one of the following
M = Mm , nM2 , 2M3 , 3M3 , 4M3 , 2M4 , 2M5 ,
where n, m are positive integers and m 6= 30k for some positive integer k.
Here it is understood that M1 = S 5 . The problem of determining all positive
Sasaki-Seifert structures on these manifolds is still very much open. This seems
somewhat intractable for M = S 5 , but in other cases it is quite manageable.
Open Problem 10.2.2: Classify all positive Sasaki-Seifert structures on M4 , 3M3
and Mm for 6 ≤ m ≤ 11.
With the exception of M4 it is probably true that the number of such structures
is finite in each case. These “sporadic” rational homology spheres already appear
in the classification of positive Brieskorn links 10.3.3. Certainly, in each case we
have at least one example. In some cases there are many.
We now consider the case of a general Smale-Barden manifold. The rank of
H2 (M, Z) and its torsion subgroup are not independent of each other, but the
classification problem becomes much more complicated. One observation that can
be made by looking at Theorems 10.2.17 and 10.2.19 is that the possible torsion
groups occur in two series each depending on a positive integer plus a handful of
‘exceptional’ cases.
Kollár [Kol05, Kol06b] obtains only partial results when H2 (M, Z)tor = Zm ⊕
Zm , but he is able to get a complete classification in certain exceptional cases. In
particular, there are definitive results in three of the five exceptional cases listed in
Theorem 10.2.17.
Theorem 10.2.21: Let M ∈ SB+ .
(i) If H2 (M, Z) = Zk ⊕ (Z5 )4 , then k = 0 and there is a unique family of
positive Sasaki-Seifert structures;
(ii) If H2 (M, Z) = Zk ⊕ (Z3 )8 , then k = 0 and there is a unique family of
positive Sasaki-Seifert structures;
(iii) If H2 (M, Z) = Zk ⊕ (Z4 )4 then k = 0, 1 and there are exactly 3 families
of positive Sasaki-Seifert structures: 2 for k = 0 and 1 for k = 1.
Proof. This follows by a detailed analysis of the four conditions above Lemma
10.2.16. We refer to [Kol05, Kol06b] for details. ¤
Remark 10.2.3: It is quite remarkable that all the 5-manifolds M ∈ SB+ of Theorem
10.2.21 can be realized as links of weighted hypersurface singularities. Cases (i) and
(ii) are realized by the BP links L(2, 5, 6, 6) and L(2, 3, 10, 10), respectively, which
are listed in Table B.4.1 of appendix B. For case (iii) two Sasaki-Seifert structures
346 10. SASAKIAN GEOMETRY IN DIMENSIONS THREE AND FIVE

occur on 2M4 and one on M∞ #2M4 , and these are realized as links of the non-BP
weighted homogeneous polynomials
f1 (z) = z05 + z15 + z24 + z0 z32 , f2 (z) = z010 + z15 + z0 z24 + z32 ,
for k = 0 (the corresponding hypersurfaces are Z20 ⊂ CP(4, 4, 5, 8) and Z40 ⊂
CP(4, 8, 9, 20), respectively) or
f3 (z) = z06 + z16 + z0 z24 + z32 ,
which is Z24 ⊂ CP(4, 4, 5, 12) for k = 1 [Kol06b]. We give a list of weighted
homogeneous hypersurfaces of this type in Example 10.3.7 below.
In [Kol06b] Kollár also obtains some results for non-simply connected 5-
manifolds. First, as in Proposition 8.1.20, there are the obvious quotients obtained
from the cyclic subgroups of the circle group generated by the Reeb vector field.
But there are other more interesting quotients. For example Kollár shows (See The-
orem 7 of [Kol06b]) that there is a free involution on the manifolds 5M∞ #2M2 and
4M∞ #2M2 giving 5-manifolds M with π1 (M ) = Z2 and H2 (M, Z) = Z2 ⊕ (Z4 )4 .
In order to proceed further we need to discuss the notion of a weighted blow up.
The general definition is Definition 4.56 in [KM98]. However, as in [Kol05] only a
special case is needed so we give the definition employed there. Let p be a point on
a smooth curve C ⊂ S and blow up p repeatedly n times. This gives one (−1)-curve
and an (n − 1)-chain of (−2)-curves. We can contract the chain of (−2)-curves to
obtain a singular neighborhood of type An−1 which can be represented locally by
an equation of the form xy − z n = 0. The resulting surface T is called the blow up
of S of type n at p. If S is a del Pezzo surface with cyclic Du Val singularities and
C ⊂ S is a smooth elliptic curve, then we denote by Bn1 ,...,nk S any surface which is
obtained as above by performing blow ups of type nP i at pi ∈ C. The anticanonical
class of Bn1 ,...,nk S will be nef and big if and only if ni < (KS2 ), and if this holds
then Bn1 ,...,nk S is a del Pezzo surface with cyclic Du Val singularities for a general
choice of the points pi . Applying the minimal program we obtain for S min either
one of the 4 log del Pezzo surfaces with Picard number one CP2 , CP(1, 2, 3), Q, S5
listed in the proof of Theorem 10.2.19 or CP1 × CP1 . Then Kollár proves [Kol05]
Proposition 10.2.22: There are 93 deformation types of del Pezzo surfaces with
cyclic Du Val singularities satisfying H1orb (Z) = 0. These are
P
(i) Bn1 ,...,nk CP(1, 2,
P3) for ni < 6, ni ≥ 2,
(ii) Bn1 ,...,nk Q for P ni < 8, ni ≥ 2,
(iii) Bn1 ,...,nk CP2 for ni <P 9, ni ≥ 2,
(iv) Bn1 ,...,nk CP1 × CP1 for ni < 8,
(v) S5 , B3 S5 , B4 S5 and B1 CP2 .
All these satisfy π1orb (Z) = 0.
Remark 10.2.4: The count of the 93 deformation types and the ranges for the ni
in Proposition 10.2.22 takes into account the following isomorphisms
B1 CP(1, 2, 3) ≈ B3 Q, B1 Q ≈ B2 CP2 , Bm,1 CP2 ≈ Bm (CP1 ×CP1 ), B1 S5 ≈ B5 CP2
as well as isomorphisms derived from them.
In [Kol06b] Kollár has loosened the restriction that π1orb (Z) = 0. In all he
has found 132 families of log del Pezzo surfaces with cyclic Du Val singularities
up to deformations giving precisely the 93 deformation types listed in Proposi-
tion 10.2.22 when π1orb (Z) = 0. We are interested in constructing orbibundles or
10.2. SASAKIAN STRUCTURES AND THE TOPOLOGY OF 5-MANIFOLDS 347

equivalently Seifert S 1 -bundles over Z with smooth total space2. Let us set some
notation. We let Z = (Z, ∆) denote one of 93 log del Pezzo surfaces (orbifolds)
given by Proposition 10.2.22 with Z = Bn1 ,...,nk S, where S is one of the 5 surfaces
CP2 , CP(1, 2, 3), Q, S5 , CP1 × CP1 . Let H be a positive generator of Cl(S), and let
Ei with i = 1, . . . , k denote the exceptional divisors. Then the divisor class group
Cl(Z) is freely generated by π ∗ H, E1 , . . . , Ek except in the case S = CP1 × CP1 ,
where we add a second P positive generator E0 . The anticanonical divisor of Z is
−KZ ≡ IS π ∗ H − i ni Ei , where IS is the Fano index of S. We choose a Seifert
S 1 -bundle f : M −−→(Z, ∆) with first Chern class
X X bi
(10.2.4) c1 (M/Z) = a[π ∗ H] + ci [Ei ] + [Di ] ,
i i
mi

where 0 ≤ bi < mi , gcd(bi , mi ) = 1. In order to have a Sasakian structure on M


we need to choose c1 (M/Z) > 0. This can always be done by choosing a large
enough. The condition for assuring the smoothness of M , given by Proposition
4.7.8, is that the local restriction of m(x, ∆) · c1 (M/Z) generates the local class
group Cl(Z, x) at each point x ∈ Z. But the singular points pi of the weighted blow
ups lie on their respective exceptional divisor Ei which generates the local class
group Cl(Z, pi ) = Zni . So M will be smooth if gcd(ni , ci ) = 1 for all i = 1, . . . , k,
and when S is singular gcd(IS , a) = 1. Furthermore, M will be simply connected if
in addition m(∆)c1 (M/Z) is a generator of Cl(Z).
Now the orbifold first Chern class of Z = (Z, ∆) is
(10.2.5)
³ X¡ 1 ¢ ´ X X³ 1 ´
corb
1 (Z) = c1 −(KZ + 1− Di ) = IS [π ∗ H] − ni [Ei ] − 1− [Di ] .
i
mi i i
mi

In order that M admit a positive Sasakian structure it is necessary that corb 1 (Z) >
0. Let us now specialize to the case at hand, namely that Z is one of the 93
log del Pezzo surfaces with cyclic Du Val singularities described by Proposition
10.2.22. In this case by Lemma 10.2.16 there is only one component, say D, in the
branch divisor ∆, and we let m denote its ramification index. Let us define d(Z) =
gcd(n1 , . . . , nk , IS ). Now the condition that π1orb (Z) = 0 implies by Proposition
4.7.13 that α 7→ α(D) mod m surjects onto Zm . Moreover, it is easy to see that
the restriction map π ∗ : H 2 (Z, Z)−−→H 2 (D, Z) is multiplication by d(Z). It follows
that we must have gcd(d(Z), m) = 1. Of course, in many cases d(Z) = 1 so there
is no condition. Summarizing we have
1
Proposition 10.2.23: Let Z = (Z, ∆ = (1 − m )D) denote one of 93 log del Pezzo
orbifold surfaces given by Proposition 10.2.22 with corb 1 (Z) > 0, and let M −
−→Z be
the Seifert S 1 -bundle whose first Chern class c1 (M/Z) is given by equation (10.2.4).
Then M is a smooth simply connected 5-manifold with a positive Sasakian structure,
that is M ∈ SB+ if and only if the following conditions hold:
(i) c1 (M/Z) > 0 and gcd(ni , ci ) = 1 for all i = 1, . . . , k, and when S is
singular gcd(IS , a) = 1.
(ii) gcd(d(Z), m) = 1.
(iii) m(∆)c1 (M/Z) is a generator of Cl(Z).

2Although here we are discussing the case of del Pezzo surfaces with cyclic Du Val singular-
ities, the analysis that follows is slightly more general.
348 10. SASAKIAN GEOMETRY IN DIMENSIONS THREE AND FIVE

We can now combine Propositions 10.2.22 and 10.2.23 with Theorems 10.2.17
and 10.2.21.
Theorem 10.2.24: Let M 5 ∈ SB+ with k = b2 (M ). Then M 5 is kM∞ #N or
X∞ #(k − 1)M∞ #N, where N is Mm , nM2 , 2M3 , 3M3 , 4M3 , 2M4 , 2M5 . Further-
more,
(i) If N = 4M3 or 2M5 then k = 0, i.e., M 5 = N = 4M3 or 2M5 .
(ii) If N = 2M4 then k = 0, 1 so M 5 = 2M4 , M∞ #2M4 , or X∞ #2M4 .
(iii) If N = Mm with m ≥ 12 then k ≤ 8, so M 5 = Mm , kM∞ #Mm , or
X∞ #(k − 1)M∞ #Mm with 1 ≤ k ≤ 8, and the number of families of
Sasaki-Seifert structures is finite.

Proof. The only part of this theorem that is not a consequence of the results
mentioned is (iii) and this follows directly from the second statement of Lemma
10.2.16. The finiteness is a consequence of the fact that the branch divisor ∆
contains only one irreducible component again by the second statement of Lemma
10.2.16. ¤

Concerning the existence of positive Sasaki-Seifert structures much more is


known in the case of Smale manifolds. Corollary 10.2.20 says that this problem is
completely understood in the case of simply connected rational homology 5-spheres.
Furthermore, Table B.4.1 lists all rational homology 5-spheres admitting a positive
Sasakian structure that can be realized as the link of a BP polynomial. Remarkably
these together with the rational homology spheres in the table of Example 10.3.7
show that all admissible rational homology 5-spheres with a positive Sasakian struc-
ture can be realized by the link of a weighted homogeneous polynomial. However, it
is probably not true that all positive Sasaki-Seifert structures on admissible simply
connected rational homology 5-spheres can be realized by links of weighted homo-
geneous polynomials. As indicated by Open Problem 10.2.2 and as discussed earlier
the classification of all deformation types of positive Sasakian Seifert structures on
admissible rational homology 5-spheres is still open and perhaps even untractable.
Let us now turn to the more general case of Smale manifolds kM∞ #N, where
N is a rational homology 5-sphere with an admissible torsion group and k 6= 0. As
remarked earlier even though M30n does not admit a positive Sasakian structure,
kM∞ #M30n does for certain k as evidenced by Table B.4.2. Nevertheless, the
problem as to which M ∈ SB0 admit positive Sasakian structures is still very much
open. Much less is known about simply connected 5-manifolds in SB+ \ (SB+ )0 . Some
regular examples will be given in Section 10.4.
Open Problem 10.2.3: Determine the sets (SB+ )0 and SB+ precisely.
A special case of particular interest in the general case is when corb
1 (Z) is a ra-
tional multiple of c1 (M/Z), or equivalently when the Sasaki-Seifert structure is anti-
canonical, recall Definition 7.5.24. This condition automatically holds when M is a
rational homology sphere by Proposition 7.5.29, or for any link of a weighted homo-
geneous polynomial by (or more generally for a weighted complete intersection) by
Proposition 9.2.4. The terminology ‘preSE’ was introduced by Kollár in [Kol06b]
for the anticanonical case. The next theorem due to Kollár [Kol05, Kol06b] re-
duces the 93 cases of Proposition 10.2.23 to 19 in the case of anticanonical Sasakian
Seifert structures.
10.2. SASAKIAN STRUCTURES AND THE TOPOLOGY OF 5-MANIFOLDS 349

Theorem 10.2.25: Let M ∈ SB+ with H2 (M, Z) = Zk ⊕ (Zm )2 for some m ≥


12 and k = b2 (M ) > 0, and suppose that the Sasaki-Seifert structure F(Fξ ) is
anticanonical. Then F(Fξ ) is one of the cases listed in the table below. More
precisely, we list all such anticanonical Sasaki-Seifert structures together with the
corresponding 5-manifold and its representation as a link of weighted homogeneous
polynomial when it exists.
M Z Link
2
8M∞ #Mm B1,1,1,1,1,1,1,1 (CP ) L(2, 3, 6, 6m)
7M∞ #Mm B1,1,1,1,1,1,1 (CP2 ) L(2, 4, 4, 4m)
6M∞ #Mm B1,1,1,1,1,1 (CP2 ) L(3, 3, 3, 3m)
5M∞ #Mm B1,1,1,1,1 (CP2 )
4M∞ #Mm B1,1,1,1 (CP2 )
4M∞ #Mm B2,2,2 (CP × CP1 ), gcd(m, 2) = 1
1
L(2, 3, 6, 3m)
3M∞ #Mm B1,1,1 (CP2 )
3M∞ #Mm B2,2 (CP × CP1 ), gcd(m, 2) = 1
1
L(2, 4, 4, 2m)
2M∞ #Mm B1,1 (CP2 )
2M∞ #Mm B3,3 (CP2 ), gcd(m, 3) = 1 L(2, 3, 6, 2m)
2M∞ #Mm B2 (CP1 × CP1 ), gcd(m, 2) = 1 L∗ (m, m, m, 2; 3m)
M∞ #Mm B1 (CP2 )
M∞ #Mm CP × CP1 , gcd(m, 2) = 1
1

M∞ #Mm B4 (Q), gcd(m, 2) = 1 L∗ (m, 2m, 3m, 4; 6m)


M∞ #Mm B3 (CP2 ), gcd(m, 3) = 1 L∗ (m, m, 2k, 3; 4m)

Proof. Since m ≥ 12 (iii) of Theorem 10.2.24 says that any log del Pezzo sur-
face which occurs must have b2 ≤ 9 and be on the list given in Proposition 10.2.22.
We now add the condition that the Sasaki-Seifert structure be anticanonical. This
imposes stringent conditions on the Seifert S 1 -bundle. Since there is only one com-
ponent D in the branch divisor and D ∈ | − KZ | we find using equations (10.2.5)
and (10.2.4) that our condition becomes
X r X
a[π ∗ H] + ci Ei = (IS [π ∗ H] − ni Ei )
d(Z) i
k

for some integer r. But from Proposition 10.2.23 smoothness implies gcd(ci , ni ) = 1
for each i, and we also have gcd(d(Z), m) = 1. Thus, in the anticanonical case we
must have n1 = n2 = · · · = nk = d(Z), and gcd(a, IS ) = 1. Furthermore, when S
is singular n1 = · · · = nk = d(Z) = IS . These conditions cut down considerably
the possibilities from the list in Proposition 10.2.22. A straightforward argument
then shows that we obtain the 15 cases in the list above when b2 (M ) > 0. (The
4 remaining cases CP2 , Q, S 5 , and CP(1, 2, 3) have b2 (M ) = 0 giving 19 cases in
all). One can then easily compute the corresponding second Betti number by the
methods of Chapter 9.
The association with the corresponding link comes from our analysis in Section
10.3 below. The cases in the table where no link is given means that one cannot
realize M as the link of a weighted homogeneous polynomial when m ≥ 12 and
the Sasakian structure is positive. This was shown by Kollár in [Kol06b], and we
reproduce his proof: Assume the contrary, that is Z = (Z, ∆) is a hypersurface of
degree d in a weighted projective space CP(w). Then by the adjunction formula
350 10. SASAKIAN GEOMETRY IN DIMENSIONS THREE AND FIVE

(Proposition 4.6.13) we have


orb 1 |w| − d
−KZ ≡ −KZ − (1 − )D ≡ H|Z ,
m m
where H is the hyperplane class of the weighted projective space. Since D ∈
|−KZ |, we conclude that KZ is proportional to H|Z . By the Grothendieck-Lefschetz
theorem [Gro68], this implies that −KZ = rH|Z for some r ∈ Z. In the cases
Z = B1 CP2 , . . . , B11111 CP2 we have r = 1, so −KZ ≡ H|Z . Now
h0 (Z, OZ (−KZ )) = h0 (CP(w), OCP(w) (1)) ≤ 4
since this dimension is the number of times that 1 occurs among w = (w0 , w1 , w2 , w3 ).
But for the four cases for Z above we have h0 (Z, OZ (−KZ ) ≥ 5, giving a contra-
diction. A similar argument works for CP1 × CP1 . ¤

We make note, however, that 5M∞ #Mm can be written as a rather simple
link, namely L∗ (2m, 3m, 4m, 7; 14m) with gcd(m, 7) = 1, except that it is negative
for m > 1.
Remark 10.2.5: Note that from the table of Theorem 10.2.25 we can make a count
of the number of families of positive Sasakian Seifert structures in each case. In
all cases m > 11. We see that the positive Sasaki-Seifert structure on kM∞ #Mm
for k = 5, 6, 7, 8 is unique up to deformations. There are precisely two families of
positive Sasaki-Seifert structures on 4M∞ #Mm and 3M∞ #Mm when m is odd,
and one such family when m is even. There are three families of positive Sasakian
Seifert structures on 2M∞ #Mm when m is relatively prime to 6, two such families
when m is even and relatively prime to 3, two such families when m is an odd and
a multiple of 3, and only one such family when m is a multiple of 6. There are
four families of positive Sasaki-Seifert structures on M∞ #Mm when m is relatively
prime to 6, three such families when m is odd and a multiple of 3, two such families
when m is even and relatively prime to 3, and only one such family when m is a
multiple of 6.
Finally we briefly discuss the Smale manifolds with no torsion. These are
S 5 , kM∞ and they will be treated in much more detail in Chapter 11. Positive
Sasaki-Seifert structures are ubiquitous on S 5 and we discuss them further below
and in Chapter 11. For example in Section 10.3.1 we show that infinitely many
such structures arise already in the classification of positive BP links (cf. Table
B.4.3 of Appendix B).The first examples of positive Sasakian structures on the k-
fold connected sum kM∞ = k(S 2 × S 3 ) for all k were given in [BGN03a]. This is
given by the non-BP link appearing as the first entry of the table in Example 10.3.7
below. Many other examples for particular values of k were given in [BGN03b,
BGN02b] and will be discussed in the context of Sasaki-Einstein geometry in
Chapter 11. Again, just as in the sphere case, each kM∞ has infinitely many BP
link representatives (cf. Table B.4.4).

10.2.2. Smale-Barden Manifolds with Symmetries. So far we have fo-


cused on SFB ⊂ B as the existence of a locally free circle action is the necessary
pre condition for the Sasakian geometry. It is interesting to examine what happens
when M ∈ B is a G-manifold with dim(G) > 1. Not surprisingly, the larger the
symmetry group the more restrictions on the topology of M . Let us summarize all
of the known results here. Some of them were already mentioned and others will
10.2. SASAKIAN STRUCTURES AND THE TOPOLOGY OF 5-MANIFOLDS 351

be discussed in the following section. However, we feel it is interesting to put all


them together.
Proposition 10.2.26: Let M ∈ B.
(i) If M ∈ SFB then H2 (M, Z) satisfies condition K. Conversely, any such
manifold admits infinitely many inequivalent locally free circle actions.
(ii) If M admits a free circle action then H2 (M, Z)tor = 0. Conversely, any
such M admits a free circle action.
(iii) If M is homogeneous than M is X−1 , S 5 or M∞ .
(iv) If M is a cohomogeneity one manifold than M is X−1 , S 5 , M∞ , or X∞
and there are infinitely many cohomogeneity one structures in each case.
(v) If M admits an effective T 3 -action then H2 (M, Z)tor = 0. Conversely,
any such M admits infinitely many inequivalent T 3 -actions.
As discussed previously case (i) is due to Kollár [Kol06a]. Cases (ii) and
(v) are due to Geiges [Gei91] and Oh [Oh83], respectively and will be discussed
in the Section 10.4. All compact, homogeneous, 1-connected, 5-manifolds were
classified in [Gor80]. It is clear that X−1 , S 5 , M∞ and X∞ have infinitely many
cohomogeneity one structures. In the case of S 5 , M∞ and X∞ there are even
infinitely many cohomogeneity one Einstein metrics [Böh98, GHY04]. But the
fact that there are no other such manifolds and a complete classification of all
inequivalent cohomogeneity one diagrams has been carried out only recently by
Hoelcher [Hoe07]. We finish this chapter with a couple of open problems.
Open Problem 10.2.4: What can one say about M ∈ B admitting an effective
T 2 -action which does not extend to an effective T 3 -action? In particular, what can
one say about H2 (M, Z)tor in this case?
First, note that X−1 = SU (3)/SO(3) admits an affective T 2 -action and there-
fore such a condition no longer implies that H2 (M, Z)tor = 0. Secondly, we do get
many examples of Sasakian manifolds with T 2 -actions which are not toric. One
such example is the BP link L(2, 2, 2, k) with the naturally induced Sasakian struc-
ture which is SO(3) × U (1)-invariant and it is of cohomogeneity 1. More generally
all the links L(2, 2, p, q) ' (gcd(p, q) − 1)(S 2 × S 3 ) have Sasakian structures with
T 2 -symmetry.
These examples suggest the partial answer to the problem which was recently
communicated to us by Kollár. viz.
Proposition 10.2.27: Let M ∈ SB+ , and assume that the dimension of the isome-
try group of the Sasakian structure is greater than one. Then M = k(S 2 × S 3 ) for
some non-negative integer k (again k = 0 means M = S 5 ).

Proof. Let M have Sasakian structure S = (ξ, η, Φ, g). If the dimension of


Isom(M, g) > 1, then by Theorem 8.1.18 either M is S 5 with its round sphere
metric or dim Aut(S) = dim Isom(M, g). In both cases we have dim Aut(S) ≥ 2.
By the Approximation Theorem 7.1.10 we can assume that S is qusi-regular, and
that the additional element in aut(S) generates a circle. Since π1 (M ) = 0 the
exact sequence 8.1.4 implies that this circle action projects to a circle action on
the algebraic variety X = M/Fξ preserving the induced Kähler structure. Thus, it
extends to a C∗ action on the orbifold (X, ∆). Now sinceP S is positive, −(KX + ∆)
is also positive, so X is a rational surface. Writing ∆ = (1 − m1i )Di it suffices to
show by Theorem 4.7.14 that all the Di are rational curves.
352 10. SASAKIAN GEOMETRY IN DIMENSIONS THREE AND FIVE

So suppose that g(Dj ) > 0 for some j. Since each Di is invariant under the
C∗ action and a curve of genus g > 0 has no non-trivial C∗ actions, Di must lie
in the fixed point set of the action. Thus, we can assume that a general point of
Dj is weakly attracting fixed point. Then Theorem 4.3 of [BB73] implies that X
contains a Zariski open subset Uj containing Dj such that the C∗ action retracts
Uj to Dj . That is, there is a C∗ -equivariant morphism Uj −−→Dj whose fibres are
isomorphic to C. Thus, there is a non-constant rational map X 99K Dj , implying
that Dj is rational which gives a contradiction. ¤
Open Problem 10.2.5: Out of the cohomogeneity one Barden-Smale manifolds
we know that X−1 cannot be Sasakian. After a complete classification of all coho-
mogeneity one diagrams for S 5 , M∞ and X∞ is accomplished it is natural to ask
which of them admit an invariant Sasakian structure? Are there cohomogeneity
one diagrams with no compatible Sasakian structure at all?

10.3. Sasakian Links in Dimension Five


In this section we present a more systematic discussion of the Smale manifolds
that can be realized as links of weighted homogeneous polynomials in C4 . The set
of such manifolds was denoted by W HP5 in Chapter 9, and as seen in Propositions
9.2.2 and 9.2.4 such links admit a natural anticanonical, canonical, or null Sasakian
structure, and they are spin by (iv) of Theorem 9.3.2. Summarizing we have
Proposition 10.3.1: W HP5 ⊂ SB0 . Furthermore, every Sasakian structure realized
as a link of a weighted homogeneous polynomial is either anticanonical, canonical,
or null.
Just as in dimension three (see Proposition 10.1.4) there is a 5-dimensional
version of the classification problem of links of quasi-smooth weighted homogeneous
polynomials. As mentioned in Chapter 5 such a list of 19 types of polynomials was
given recently by Yau and Yu [YY05] and their table is reproduced as Table B.5.1
in Appendix B. In fact, we have a corollary of Theorem 5.4.18:
Theorem 10.3.2: Let f ∈ C[z0 , z1 , z2 , z3 ] be a weighted homogeneous polynomial
with weights and degree (w; d) and with only an isolated singularity at 0. Then
the associated link Lf ∈ L∗ (w; d) is equivariantly diffeomorphic to one of the links
given by the polynomials listed in Table B.5.1. of the Appendix B.
Note that there is a lot of duplication in Table B.5.1. On the other hand, it
follows from Theorem 10.3.2 that Table B.5.1. properly subsumes other tables of
Appendix B: namely, Reid’s examples in Table B.1.1, the well-formed examples
of Tables B.3.1-2 and the positive BP links classified in Tables B.4.1-4. The first
entry in Table B.5.1 is the just the set of Brieskorn-Pham polynomials and as seen
above it is quite remarkable how much can be obtained by working with these
alone. Some important non-BP Sasakian links are presented in the context of
Sasaki-Einstein geometry in Chapter 11.

10.3.1. Positive Sasakian Links. It is a straightforward albeit tedious task


to list all the positive links that occur in Table B.5.1. This is precisely what was
done in the preprint version of [YY05] (cf. math.AG/0303302). Indeed the authors
give quite an extensive list of all rational isolated hypersurface singularities defined
by weighted homogenous polynomials in four complex variables. The positivity of
the Sasakian structure on the link is equivalent to the rationality of the singularity
10.3. SASAKIAN LINKS IN DIMENSION FIVE 353

at the origin. They do not, generally, seem to be interested in the topology of the
links or the orbifold structure of the underlying log del Pezzo surfaces. However,
in an earlier paper Yau and Yu classify all links of Table B.5.1 which are positive
and diffeomorphic to the standard 5-sphere [YY02].
Next we determine those manifolds in SB+ which can be realized as Brieskorn-
Pham links L(a). It is easy to get a complete classification. Let us assume that the
exponents in a = (a0 , a1 , a2 , a3 ) are in non-decreasing order. We have to consider
P2
two separate situations: First let 0 a1i ≥ 1. Then the last exponent a3 is arbitrary
so that in each instance we get an infinite series of positive links. Apart from
(2, 2, p, ?), there are only six such cases: (2, 3, 3, k), (2, 3, 4, k), (2, 3, 5, k), (2, 3, 6, k),
(2, 4, 4, k), (3, 3, 3, k). One can show that L(2, 2, p, k) has no torsion in H2 (M, Z). It
follows that L(2, 2, p, q) ' (n − 1)M∞ , where n = gcd(p, q). For the six remaining
cases we consider subsequences dependingP on the modality of the fourth exponent
2
relative to the first three. Secondly, when 0 a1i < 1 we get only a finite list of 112
“sporadic” examples. These are L(3, 4, 4, 4), L(3, 4, 4, 5) and L(a), where a is one
of the following

(2, 3, 7, k) 6 < k < 42 (2, 3, 8, k) 7 < k < 24 (2, 3, 9, k) 8 < k < 18


(2, 3, 10, k) 9 < k < 15 (2, 3, 11, k) 10 < k < 14 (2, 4, 5, k) 4 < k < 20
(2, 4, 6, k) 5 < k < 12 (2, 4, 7, k) 6 < k < 10 (2, 5, 5, k) 4 < k < 10
(2, 5, 6, k) 5 < k < 8 (3, 3, 4, k) 3 < k < 12 (3, 3, 5, k) 4 < k < 8

In the four classification tables in Appendix B.4, we separate links according to the
following properties of the second homology group:
(i) H2 (L(a), Z) 6= 0 is pure torsion,
(ii) H2 (L(a), Z) has both non-trivial free part and non-trivial torsion piece,
(iii) H2 (L(a), Z) = 0,
(iv) H2 (L(a), Z) 6= 0 has no torsion subgroup.
In addition, in the case of vanishing torsion we separate out the sporadic links.
Out of the 112 sporadic examples most are actually spheres, more precisely, in 79
cases H2 (L(a), Z)tor = 0 of which 63 are diffeomorphic to S 5 and 16 are diffeomor-
phic to some nM∞ . These are split between in Tables B.4.3-4. The 33 sporadic links
with non-trivial torsion are rational homology spheres in 25 cases and are included
in Table B.4.1. The remaining 8 sporadic links with b2 > 0 and non-trivial torsion
can be found in Table B.4.2. The computation of H2 (L(a), Z) gives the following
classification:
Theorem 10.3.3: Let L(a) be a Brieskorn-Pham link and suppose L(a) ∈ SB+ . If
H2 (L(a), Z)tor 6= 0 then L(a) is either one of the rational homology spheres listed
in Table B.4.1 of Appendix B, or b2 (L) > 0 and L(a) is one of the Smale manifolds
listed in Table B.4.2. If H2 (L(a), Z)tor = 0 then L(a) is either S 5 with its positive
Seifert fibered structure listed in Table B.4.3 or b2 (L) > 0 and L(a) is one of the
Smale manifolds listed in Table B.4.4. In each case when H2 (L(a), Z)tor 6= 0 we
give the number of distinct positive Sasakian Seifert structures NSF (if finite, or
else NSF = ∞).
As mentioned above all positive links of isolated hypersurface singularities from
weighted homogeneous polynomials was given in [YY02]. So our Table B.4.3 con-
tains only the subclass of type I links (i.e., BP links) given in the 10 page list in
354 10. SASAKIAN GEOMETRY IN DIMENSIONS THREE AND FIVE

[YY02]. The computations for rational homology 5-spheres realized by Brieskorn-


Pham links L(a) are very easy due to a simplification of Orlik’s formulae (9.3.17)-
(9.3.19) in which case Orlik’s Torsion Conjecture 9.3.15 is known to hold.
Theorem 10.3.4: Let a = (a0 , a1 , a2 , a3 ) and L(a) be is a rational homology 5-
sphere. Without loss of generality assume that a0 = k is the only isolated vertex of
G(a). Then
H2 (L(a), Z) = (Zk )2g ,
where
X gcd(a1 , a2 ) gcd(a1 , a3 ) gcd(a2 , a3 )
2g = 2 − gcd(ai , aj ) + .
i<j
gcd(a1 , a2 , a3 )

Proof. Let L(a) be a BP link which is a rational homology 5-sphere. By the


Brieskorn Graph Theorem 9.3.18 we must have at least one isolated point in G(a),
say a0 = k. If there is more than one isolated vertex the torsion vanishes so that
we can assume that there is exactly one such vertex. The statement follows by
Proposition 9.3.16 from the calculation of the torsion group with Orlik’s algorithm.
We compute
c∅ = 1, c0 = gcd(a1 , a2 , a3 ), c1 = c2 = c3 = 1 .
gcd(aj , ak )
c0,i = , ci,j = 1 ,
gcd(a1 , a2 , a3 )
gcd(a1 , a2 , a3 )ai
c1,2,3 = k, c0,j,k = .
gcd(aj , ai )gcd(ai , ak )
However, the only non-vanishing ki1 ,...,is is k1,2,3 = 2g. ¤
Note that any k can occur and which g occur depends on the choice of k.
The Theorem 10.3.4 can be conveniently restated as follows: let L(a) be a rational
homology 5-sphere with non-trivial torsion. The graph G(a) consists of an isolated
point and a triangle. Let us label the triangle with numbers: the edges with
gcd(ai , aj ) and the face with gcd(a1 , a2 , a3 ). Then 2g − 2 equals the product of
the “edge numbers” divided by the “face number” minus the sum of the “edge
numbers”. The general case when b2 (L(a)) 6= 0 is only slightly more complicated.
Exercise 10.4: Let Lf = L(a1 , a2 , a3 ) be a 3-dimensional sublink of L(a0 , a1 , a2 , a3 ).
Show that 2g of Theorem 10.3.4 equals to the first Betti number of Lf which, in
turn, is twice the genus of the orbifold Riemann surface Xf which is the transverse
space of Lf .
Computing the homology of more general links is only slightly more involved.
Let L(a0 , a1 , a2 , a3 ) be a BP link and let G(a) be its Brieskorn graph. We can
easlily construct a labelled polytope, also denoted by G(a), where in addition all
six edges and all four faces should be labelled with the corresponding gcd’s. This
give a tetrahedron
/.-,
()*+
a0 C ()*+
/.-,
a1
CC {{{
CC{
G(a)= C
{{{{ CC
/.-,
()*+
a 2 ()*+
/.-,
a3
a0 a1 a2 a3
where we also label the interior of the tetrahedron with t = lcm(a 0 ,a1 ,a2 ,a3 )
. The re-
lation with the usual Brieskorn graph introduced in Chapter 9 is simple enough: we
just need to remove all edges of the tetrahedron labelled by 1 to give the Brieskorn
10.3. SASAKIAN LINKS IN DIMENSION FIVE 355

graph. Whether one uses a labelled polytope or a Brieskorn graph is a choice dic-
tated by convenience. In the situation of Theorem 10.3.4, for instance, it is more
convenient to use the Brieskorn graph which corresponds to removing the tree edges
meeting at a0 in the polytope. However, in the more general case, it is advantageous
to keep all edges and faces labelled by 1’s. We shall refer to such polytopes G(a) as
labelled Brieskorn polytopes. We get the following generalization of Theorem 10.3.4:
Theorem 10.3.5: Let L(a) = L(a0 , a1 , a2 , a3 ) be a BP link with labelled Brieskorn
polytope G(a). Then
H2 (L(a), Z) = Zκ ⊕ (Zm0 )2g0 ⊕ (Zm1 )2g1 ⊕ (Zm2 )2g2 ⊕ (Zm3 )2g3 ,
where
(i) m = (m0 , m1 , m2 , m3 ) are the reduced indices of the vertices a, i.e.,
mi is a part of ai which is not a part of any aj in the remaining 3 vertices.
(ii) g = (g0 , g1 , g2 , g3 ), where the integer 2gi − 2 is computed from the face
opposite to the vertex ai and it is the product of the “edge numbers”
divided by the “face number” minus the sum of the “edge numbers”.
(iii) κ + 3 − t is computed as the sum of the six “edge numbers” minus the
sum of four products of “edge numbers” divided by the “face numbers”
(one such term for each face).
The ‘decoding’ of the labelled Brieskorn polytope is just calculating the triple
(m, g, κ) which yields H2 (L(a), Z). Hence, Theorem 10.3.5 splits the calculation
with Orlik’s algorithm into two stages: labelling and decoding.
a =⇒ G(a) =⇒ H2 (L(a), Z) .
Note that mi ’s are always positive integers and, by construction, they must be
pairwise relatively prime. If some mi = 1 then Zmi is trivial. Likewise, 2gi is
a non-negative integer and if some gi = 0 then Z0mi is trivial regardless of the
corresponding mi .
Exercise 10.5: Prove Theorem 10.3.5.
Remark 10.3.1: In the case of more general links one can organize the information
coming from Orlik’s algorithm the same way it is presented in Theorem 10.3.5.
The combinatorial information is slightly more complicated as it depends on the
(u, v)-data. It can, however, be organized in a labelled rational Brieskorn polytope
G(u; v) in such a way that H2 (Lf , Z) is computed in almost the same fashion, i.e.,
the second arrow in (u; v) =⇒ G(u, v) =⇒ H2 (Lf , Z) is essentially the same as in
the BP case.
Exercise 10.6: In analogy with the BP case define G(u; v) and describe Orlik’s
Algorithm as labelling and decoding: (u; v) =⇒ G(u, v) =⇒ H2 (Lf , Z).
Example 10.3.6: We illustrate how to decode the labelled Brieskorn polytope by
considering a = (a0 , a1 , a2 , a3 ) = (27, 9, 3, 3). Then the reduced indices are m =
(m0 , m1 , m2 , m3 ) = (3, 1, 1, 1). There is one edge labelled by 9 and five others
labelled by 3. All four faces are labelled by 3 and the interior has t = 81. Now, one
can easily see that g = (g0 , g1 , g2 , g3 ) is given by 2g − (2, 2, 2, 2) = (0, 0, 12, 12).
Hence, g = (1, 1, 7, 7). Finally, κ + 3 = (9 + 5 · 3) − 2(27 + 9) + 81 = 30. Thus
H2 (L, Z) = Z30 ⊕ Z23 . Let us revisit one of the links of Example 9.1 but write a
using the prime number decomposition a = (2 · 3 · 13, 2 · 5 · 17, 3 · 11 · 19, 7 · 17 · 19).
Then the reduced indices m = (13, 5, 11, 7). There are four vertices labelled by
356 10. SASAKIAN GEOMETRY IN DIMENSIONS THREE AND FIVE

2, 3, 19, 17 and two labelled by 1. All four faces are labelled by 1 and t = 1938. Our
theorem easily gives g = (144, 18, 8, 1) while
κ + 3 = (2 + 3 + 17 + 19 + 1 + 1) − (2 · 3 + 17 · 19 + 3 · 19 + 2 · 17) + 1938 = 1561.
Generally speaking, the polytopes with more edges labelled by 1 lead to a
simpler recipe for the second homology. For example, when
/.-,
()*+
a0 ()*+
/.-,
a1

G(a)=

/.-,
()*+
a2 ()*+
/.-,
a3

then the formulas of Theorem 10.3.5 easily reduce to H2 (L(a), Z) = Zκ , where


κ = (gcd(a1 , a3 ) − 1)(gcd(a0 , a2 ) − 1). The special examples corresponding to this
type are the positive BP links L(2, 2, p, q) of Table B.4.4.
A similar analysis can be done with any of the Types in Table B.5.1. Since a
complete picture is quite involved we content ourselves by considering only a few
more examples of Type II in Table B.5.1.
Example 10.3.7: For a 4-tuple a = (a0 , a1 , a2 , a3 ) of positive integers consider the
weighted homogeneous polynomial
f (z) = z0a0 + z1a1 + z2a2 + z0 z3a3 .
This is a simplest modification of a BP type which is listed in Table B.5.1 as Type
II. One can easily see that
w = (d/a0 , d/a1 , d/a2 , (a0 − 1)d/a3 a0 ).
Hence, the link Lf is Fano if and only if
1 1 1 1 1
+ + + − > 1.
a0 a1 a2 a3 a0 a3
Just as in the BP case one can enumerate all the examples. However, many of
them are merely deformations of the BP links given in the tables in Appendix B.4
(the duplications we have already mentioned). Below are some interesting examples
which are not of BP type
M a w d
nM∞ (n + 1, n + 1, n + 1, 1) (1, 1, 1, n) n+1
2M∞ #Mk (3, 3, 3, k), (k, 2) = 1 (k, k, k, 2) 3k
M∞ #Mk (3, 6, 2, k), (k, 2) = 1 (2k, k, 3k, 4) 6k
Mk (6, 2, 3, k), (k, 5) = 1 (k, 3k, 2k, 5) 6k
M∞ #Mk (4, 2, 4, k), (k, 3) = 1 (k, 2k, k, 3) 4k
2M4 (10, 5, 2, 4) (4, 8, 20, 9) 40
2M4 (5, 5, 4, 2) (4, 4, 8, 5) 20
M∞ #2M4 (6, 6, 2, 4) (4, 4, 12, 5) 24
3M3 (4, 4, 3, 2) (3, 3, 4, 3) 12
nM2 (2n + 1, 2n + 1, 2, 2) (2, 2, (2n + 1), 2n) 2(2n + 1)
One could consider an arbitrary link Lf of some weighted homogenous polyno-
mial f of degree d and with weights w = (w0 , w1 , w2 , w3 ) of Table B.5.1.
Open Problem 10.3.1: Enumerate all positive non-BP links of Table B.5.1, i.e.,
all links of Type II-XIX with |w|−d > 0, and compute H2 (Lf , Z) for each such link,
thus determining the corresponding Smale manifold. Note that we could exclude all
10.3. SASAKIAN LINKS IN DIMENSION FIVE 357

the well-formed links as these have no torsion in H2 . (A classification of well-formed


links, under some additional assumptions, was carried out in [BGN03b]).
10.3.2. Null Sasakian Links. In the null case there appears to be only a fi-
nite number of compact simply connected manifolds admitting null Sasakian struc-
tures in each dimension. The key here is Proposition 4.7.11 of Kollár which states
that a Seifert bundle over a Calabi-Yau orbifold necessarily has no branch divisor.
Before discussing null Sasakian links outright, we give two general results concern-
ing null Sasakian structures in dimension five. The first of these results shows that
finiteness indeed holds in dimension five. This result is closely related to Theorem
5.4.5, but here we give a proof that is independent of that theorem. Both Theorem
10.3.8 and Corollary 10.3.11 were first given in [BGM06].
Theorem 10.3.8: Let S be a null Sasakian structure on a compact simply con-
nected 5-manifold M. Then 2 ≤ b2 (M ) ≤ 21 and H2 (M, Z) is torsion-free, and M is
diffeomorphic to kM∞ = k(S 2 ×S 3 ), where k = b2 (M ). Furthermore, if b2 (M ) = 21
then S is regular and M is diffeomorphic to 21(S 2 × S 3 ), and X = M/Fξ is a K3
surface.
Proof. A null Sasakian structure must be quasi-regular by (iii) of Theorem
8.1.14. So M is a S 1 -orbibundle over a orbifold K3 surface. By Proposition 4.7.11
orb
there are no branch divisors and KX ≡ KX is trivial. So the geometric genus
pg (X) and the transverse geometric genus pg (Fξ ) both equal 1. This implies that
b2 (X) is at least 3, so b2 (M ) is at least 2. This give the lower bound. The upper
bound comes from the fact that a smooth K3 is a minimal resolution of the singular
K3 surface X as shown in Corollary 10.4 of [Kol05]. This corollary also shows that
if b2 (M ) = 21, then X must be a smooth K3 surface, which in the absence of branch
divisors implies that the Sasakian structure is regular. That H2 (M, Z) is torsion
free follows from Proposition 4.7.11 and Theorem 4.7.14. Since c1 (Fξ ) = 0, the
manifold M is spin by Theorem 7.5.27, and so must be of the form given by the
Smale-Barden classification Theorem 10.2.3. ¤
We mention that the lower bound in Theorem 10.3.8 holds with the weaker
condition H1 (M, R) = 0. We now have
Corollary 10.3.9: S 5 and S 2 × S 3 or any quotient by a finite group do not admit
null Sasakian structures.
Compact 5-manifolds M with null Sasakian structures can be easily obtained
from the list of 95 orbifold K3 surfaces given in the Table B.1 in Appendix B.
Interestingly, we see from this table that there are gaps in the range of possible
null Sasakian 5-manifolds given by Theorem 10.3.8, namely b2 (M ) = 2, 17 do not
occur.
Example 10.3.10: As discussed in Section 5.4.2, in 1979 Reid [Rei80] produced
a list of 95 (possibly singular) K3 surfaces X given as well-formed hypersurfaces
in weighted projective space CP(w0 , w1 , w2 , w3 ). This list is given in Table B.1
of Appendix B. With each of Reid’s examples one can consider the associated 5-
dimensional link Lf which by Theorem 10.3.8 are all diffeomorphic to some k-fold
connected sum of S 2 × S 3 . The integer k = b2 (Lf ) can be computed in each case by
the Milnor-Orlik procedure described in Section 9.3.1, and Table B.1 includes this
information giving b2 (X) which equals k +1. For instance L(4, 4, 4, 4) ' 21(S 2 ×S 3 )
is regular and it is a circle bundle over the smooth K3-surface X4 ⊂ CP3 . The only
358 10. SASAKIAN GEOMETRY IN DIMENSIONS THREE AND FIVE

other regular circle bundle over a non-singular K3 surface on Reid’s list is also a
BP link, namely L(6, 6, 6, 2) ' 21(S 2 × S 3 ), which is a regular circle bundle over
X6 ⊂ CP(1, 1, 1, 3) given by the vanishing of f (z) = z06 + z16 + z26 + z32 . However, X4
and X6 are not the same as complex algebraic varieties, and the induced Sasakian
structures on L(4, 4, 4, 4) and L(6, 6, 6, 2) belong to inequivalent deformation classes
(of Sasakian structures).
As one sees from the Table B.1 of Appendix B all of Reid’s examples have
3 ≤ b2 (Lf ) ≤ 21. Hence, for instance, the link Lf with f (z) = z08 +z16 +z23 z0 +z33 must
be diffeomorphic to 3(S 2 × S 3 ), whereas, the BP link L(2, 3, 12, 12) is diffeomorphic
to 20(S 2 ×S 3 ). It is interesting that neither b2 (L) = 2 nor 17 occur on Reid’s list, so
we would need to consider possible singular K3 surfaces that are not hypersurfaces
in CP(w0 , w1 , w2 , w3 ) to fill these gaps. For instance, Fletcher gives a list of 84
examples of codimension 2 complete intersections [IF00], but here he assumes only
canonical singularities and that |w| ≤ 100, so there could be even more examples.
Exercise 10.7: Show that all the null links of Table B.5.1 (i.e., the ones with
|w| = d) are equivalent to Reid’s K3 surfaces of Table B.1.
Summarizing the results for the computation of the second Betti number of all
95 orbifold K3 surfaces given in Table B.1 of Appendix B we have
Corollary 10.3.11: k(S 2 × S 3 ) admit null Sasakian structures for 3 ≤ k ≤ 21 and
k 6= 17. Furthermore, the count of the number of null Sasaki-Seifert structures up
to deformation given by Reid’s list can be read off from Table B.1.
The question which of the simply connected compact spin 5-manifolds on
Smale’s list admit null Sasakian structures is related to classifying all orbifold K3-
surfaces which is still open. The more modest interesting question is whether k = 2
or 17 can occur. This leads us to:
Open Problem 10.3.2: Find examples of null Sasakian structures on 2(S 2 × S 3 )
and 17(S 2 × S 3 ), or prove that none can exist.
In higher dimensions there are many, though probably finite, null Sasakian
structures. Interesting examples include the over 6000 Calabi-Yau orbifolds in
complex dimension 3 [CLS90].

10.3.3. Negative Sasakian Links. We finish this section with a discussion


of negative Sasakian structure. These appear to be the least constrained in terms
of what kind of H2 (M, Z) one should expect when M is a Smale manifold. On the
other hand, other than examples we do not really have any classification result.
The Brieskorn-Pham links produce already very many examples. The simplest
ones are circle bundles over the Fermat hypersurfaces of degree d in P3 which are
described in Example 5.4.1. For d ≥ 5 we get regular negative Sasakian structures
on k(S 2 × S 3 ) for k = (d − 2)(d2 − 2d + 2) + 1. These begin with k = 52 and grow
rapidly.
Example 10.3.12: Consider L(k, k, k + 1, p). Such a link is negative as long as
k, p > 3 or with k = 3 and p > 12. Moreover, L(k, k, k + 1, p) is a 5-sphere as long
as p is prime to both k and k + 1 which can easily be arranged.
Proposition 10.3.13: Every negative Sasakian structure on S 5 is non-regular and
there exists infinitely many inequivalent negative Sasakian structures.
10.3. SASAKIAN LINKS IN DIMENSION FIVE 359

Proof. The existence of infinitely many non-regular negative Sasakian struc-


tures on S 5 follows immediately from Example 10.3.12. To prove the first statement
we assume that S 5 has a regular negative Sasakian structure. Then we have a circle
bundle S 5 −−→X with c1 (X) negative, π1 (X) = 0, and b2 (X) = 1. This is impossible
by a Theorem of Yau, cf. Theorem V.1.1 of [BPVdV84]. ¤

More generally one should be able to examine the existence of negative Sasakian
structures on Smale-Barden manifolds, and in particular, on rational homology
spheres. Unlike the positive case where we have Kollár’s classification Theorems
10.2.19 and 10.2.25, not much is known in general. However, we have
Theorem 10.3.14: Let M ∈ SB be a rational homology sphere. Then M is either
positive and the torsion in H2 (M, Z) is restricted by theorem 10.2.19 or M is neg-
ative. There exists infinitely many such manifolds which admit negative Sasakian
structures but no positive Sasakian structure. There exists infinitely many positive
rational homology spheres which also admit negative Sasakian structures.

Proof. To show that there are negative rational homology 5-spheres which
cannot admit any positive Seifert fibration one can consider the BP link L(p, p, p, k)
with (p, k) = 1 and 1/3p + 1/k < 1. By Orlik’s formula this has H2 (M, Z)tors =
(Zk )d , where d = p(p − 3) + 2. When p > 3 and k > 1 this produces torsion
disallowed by Kollár’s Theorem 10.2.19.
To get examples of rational homology 5-spheres which admit both positive and
negative Sasaki-Seifert structures we consider the links of [BG02]. One can easily
see, for example, that Mk admits both positive and negative Sasakian structures
for every prime k. ¤

Open Problem 10.3.3: Determine which simply connected rational homology 5-


spheres admit negative Sasakian structures.
Open Problem 10.3.4: Classify the negative Sasaki-Seifert structures on simply
connected rational homology 5-spheres.
These problems are very difficult in general, and Problem 10.3.4 is clearly more
difficult than Problem 10.3.3. Nevertheless, it is rather easy to produce examples
of negative Sasakian structures on k(S 2 × S 3 ); however, we haven’t been able to
find one nice sequence which does this, or even a sequence that works for odd k.
The examples below give the essential ideas.
Example 10.3.15: This is example 10.3.12 with p = k + 1. Now the BP link is
L(k, k, k + 1, k + 1) and it is easy to see that the induced Sasakian structures are
negative as long as k ≥ 4. Here the Milnor-Orlik method gives b2 (L(k, k, k + 1, k +
1)) = k(k − 1). This gives negative Sasakian η-Einstein structures on k(S 2 × S 3 )
for infinitely many k, but with larger and larger gaps as k grows. For low values of
k, for example, k = 4 and 5, we get negative Sasakian structures on 12(S 2 × S 3 ),
and 20(S 2 × S 3 ), respectively.
Example 10.3.16: Consider the BP links L(p, q, r, pqr) such that p, q, r ∈ Z+ are
pairwise relatively prime. The transverse space is a surface Xpqr ⊂ P(pr, qr, pq, 1)
and it is automatically well-formed, i.e., it has only isolated orbifold singularities.
Hence, L(p, q, r, pqr) is diffeomorphic to k(S 2 × S 3 ), where

k = b2 (L(p, q, r, pqr)) = (pqr − pq − pr − qr − 1) + p + q + r .


360 10. SASAKIAN GEOMETRY IN DIMENSIONS THREE AND FIVE

Now, L(p, q, r, pqr) is negative when the term in parenthesis is positive, i.e., when
pq + pr + qr + 1 < pqr. To simplify we can consider L(2, 3, r, 6r) with r prime to
both 3 and 2. L(2, 3, 7, 42) is one of the examples on Reid’s list. But from r > 7 we
get infinitely many examples with second Betti number b2 = 2(r − 1). For example
for r = 11, we see that L(2, 3, 11, 66) is diffeomorphic to 20(S 2 × S 3 ).
Notice that both Examples 10.3.15 and 10.3.16 produce negative Sasakian
structures on k(S 2 × S 3 ) for even k only. We haven’t yet found a nice series that
gives metrics for odd values of k. However, we can obtain negative Sasakian struc-
tures for odd k if we consider the more general weighted homogeneous polynomials.
Example 10.3.17: Here we give a few examples only to illustrate the method.
A much more extensive list can be generated with the aid of a computer a la
[JK01a, BGN03b]. We begin with two inequivalent ‘twin’ negative Sasakian
structures on S 2 × S 3 . Consider the weighted homogeneous polynomials
z021 z1 + z15 z2 + z23 z0 + z32 , z021 z1 + z15 z0 + z23 z1 + z32 .
Both have degree 316 and both give links diffeomorphic to S 2 × S 3 . The weight
vectors are w = (13, 43, 101, 158) and w = (11, 61, 85, 158), respectively. So |w| −
d = −1 giving negative Sasakian structures on S 2 × S 3 .
Similarly, the polynomials
z020 + z13 z3 + z23 z1 + z32 z0 z010 + z14 z2 + z22 z3 + z32 z0
have degree 40 and 159, respectively. They give negative Sasakian structures on
7(S 2 × S 3 ) and 2(S 2 × S 3 ).
We have
Corollary 10.3.18: 7(S 2 × S 3 ), 12(S 2 × S 3 ) and 20(S 2 × S 3 ) all admit positive,
null, and negative Sasakian structures.
Open Problem 10.3.5: Show that k(S 2 × S 3 ) admit negative Sasakian structures
for all k, or show that this is not true, and determine precisely for which k it is
true.
Open Problem 10.3.6: Determine which torsion groups correspond to Smale-
Barden manifolds admitting negative Sasakian structures.

10.4. Regular Sasakian Structures on 5-Manifolds


We begin with the homogenous case. Every compact Sasakian homogeneous
manifold is regular, in particular. Furthermore, the base of the circle bundle has
to be a compact homogenous Kähler surface. Using Theorem 8.3.6 and Example
7.6.10 we easily get
Theorem 10.4.1: Let M ∈ SB be a homogenous Sasakian space. Then M is one
of the following
(i) S 5 with the standard Sasakian structure fibering over CP2 ,
1,1
(ii) one of the Wang-Ziller manifolds Mk,l of Example 7.6.10 fibering over
CP × CP and diffeomorphic to S × S 3 .
1 1 2

Since the classification of compact complex surfaces uses the concept of mini-
mal models we do the same for regular Sasakian structures on compact 5-manifolds.
Let S = (ξ, η, Φ, g) be a regular Sasakian structure on a compact 5-manifold M 5 .
10.4. REGULAR SASAKIAN STRUCTURES ON 5-MANIFOLDS 361

Then by Theorem 7.5.1 M 5 is the total space of an S 1 -bundle over a smooth com-
pact polarized algebraic surface X 2 . The polarization is determined by a choice of
integral Kähler class. The regular Sasakian structures on compact 5-manifolds are
classified by picking out the algebraic surfaces in the Enriques-Kodaira classification
[BPVdV84, BHPVdV04] in complex surface theory. We have
Theorem 10.4.2: Let S = (ξ, η, Φ, g) be a regular Sasakian structure on a com-
pact 5-manifold M 5 . Then M 5 is the total space of an S 1 -bundle over a projective
algebraic surface X 2 which is obtained by blowing-up k = 0, 1, . . .-times one of the
following minimal surfaces:
(i) a minimal rational surface: CP2 or a Hirzebruch surface Sn with n ≥
0, n 6= 1.
(ii) a ruled surface of genus g ≥ 1.
(iii) an algebraic K3 surface.
(iv) an Enriques surface.
(v) an algebraic 2-torus (Abelian variety of dimension 2).
(vi) a hyperelliptic surface.
(vii) an algebraic minimal properly elliptic surface.
(viii) a minimal surface of general type.
Recall that a compact complex surface is minimal if it does not contain a (−1)-
curve. Thus, we can say that a regular Sasakian structure S on M 5 is minimal
if X 2 is minimal. In order to describe regular Sasakian structures on simply con-
nected 5-manifolds, we first discuss which algebraic complex surfaces are simply
connected. First we briefly discuss elliptic fibrations following [GS99]. The com-
plex surface CP2 #9CP2 admits elliptic fibrations of the form π : CP2 #9CP2 −−→CP1
with generic fibres an elliptic curve, and as such it is denoted by E(1). Given el-
liptic fibrations πi : Si −−→Ci we can form the fibre sum S 1 #f S2 as follows: Let ti
be points of Ci with generic fibres Fi = πi−1 (ti ), and take regular neighborhoods
Ni of these fibres in Si of the form D2 × T 2 . Glue the manifolds Si − Ni together
with an orientation reversing diffeomorphism on the boundary 3-tori Ti3 . This gives
an elliptic fibration π : S1 #f S2 −−→C1 #C2 . If one of the original elliptic fibrations
contains a cusp fibre, then the resulting 4-manifold is independent of the gluing dif-
feomorphism. Now it is known that E(1) admits elliptic fibrations containing cusp
fibres, so we obtain well defined 4-manifolds by defining E(n) = E(n − 1)#f E(1)
inductively. Note that E(2) = K3. A properly elliptic surface is an elliptic surface
with Kodaira dimension one.
We are interested in which Smale-Barden manifolds admit regular Sasakian
structures. The main point is a result of Geiges [Gei91] which says that any
simply connected 5-manifold that admits a free circle action has no torsion.
Theorem 10.4.3: Let M ∈ B. Then
(i) M ∈ SFBR if and only if M = S 5 , k(S 2 × S 3 ), or X∞ #(k − 1)(S 2 × S 3 ),
where k = b2 (M ) ≥ 1,
(ii) SFBR = SFBR ∩ AC = SFBR ∩ C = S R ,
(iii) (SFBR )0 = (S R )0 .
Proof. If M admits a smooth free circle action then it is the total space of a
principal S 1 -bundle over a smooth base Z. By the long exact homotopy sequence
Z is simply connected, and by Poincaré duality H 3 (Z, Z) = 0. Then the Gysin
sequence implies that H 3 (M, Z) is torsion-free, and this is the only possible torsion
362 10. SASAKIAN GEOMETRY IN DIMENSIONS THREE AND FIVE

in a simply connected 5-manifold. The converse will be proved below by proving


that the manifolds of Theorem 10.2.3 without torsion all admit regular Sasakian
structures. This follows immediately from Proposition 10.4.4 below. ¤
We shall give some examples below of regular Sasakian structures. Recall from
Definition 7.5.24 that there are four types of Sasakian structures, positive, negative,
null, and indefinite. We have
Proposition 10.4.4: Let S = (ξ, η, Φ, g) be a regular Sasakian structure on a
Smale-Barden manifold M. Then
(i) if S is null Sasakian then M = 21(S 2 × S 3 ),
(ii) if S is positive then M = S 5 , k(S 2 × S 3 ), or M = X∞ #(k − 1)(S 2 × S 3 ),
where 1 ≤ k ≤ 8,
(iii) if S is indefinite then M = k(S 2 × S 3 ), or M = X∞ #(k − 1)(S 2 × S 3 )
for k ≥ 1.
Conversely, regular null Sasakian structures exist on 21(S 2 × S 3 ); regular posi-
tive Sasakian structures exist on all the Smale-Barden manifolds S 5 , k(S 2 ×S 3 ), and
X∞ #(k − 1)(S 2 × S 3 ), where 1 ≤ k ≤ 8; and regular indefinite Sasakian structures
exist on k(S 2 × S 3 ), and X∞ #(k − 1)(S 2 × S 3 ) for all k ≥ 1.
Proof. Any regular Sasakian manifold must be one listed in Theorem 10.4.3.
Regular null Sasakian structures occur only on 21(S 2 × S 3 ) since the only smooth
algebraic surface with c1 = 0 is a K3 surface. Regular positive Sasakian structures
are all circle bundles over smooth del Pezzo surfaces which are completely classified
(see Example 3.5.11). It is straightforward to see that positive Sasakian structures
exist on all the manifolds indicated in the proposition by choosing the S 1 -bundle
appropriately as in Example 10.4.5. As for regular indefinite Sasakian structures it
is clear that they occur on S 1 -bundles over the equivariant blow-up of Hirzebruch
surfaces. It is also clear that the Picard group of the base must have rank at least
two, so S 5 is excluded. The existence of regular Sasakian structures of the type
indicated and on the indicated manifolds will then follow from the construction of
circle bundles over Hirzebruch surfaces and their blow-ups. This construction is
given in Examples 10.4.5 and Theorem 10.4.12 below. ¤
Example 10.4.5: [Circle Bundles over Hirzebruch Surfaces] Our discussion
here follows that in [BGO07]. We construct infinite families of deformation classes
of toric Sasakian structures on circle bundles over Hirzebruch surfaces. The total
space M of such circle bundles have b1 (M ) = 0 and b2 (M ) = 1. It is well-known
[Bar65] that there are precisely two simply connected closed 5-manifolds with
b2 = 1. They are S 2 × S 3 and the non-trivial S 3 -bundle over S 2 , denoted X∞ by
Barden. They are distinguished by their second Stiefel-Whitney class w2 , that is,
by whether they are spin or not, or equivalently by their Barden invariant i(M ).
Recall the Hirzebruch surfaces Sn [GH78b] are realized as the projectivizations
of the sum of two line bundles over CP1 , namely
¡ ¢
Sn = P O ⊕ O(n) .
They are diffeomorphic to CP1 × CP1 if n is even, and to the blow-up of CP2 at
2
f , if n is odd. For n = 0 and 1 we get CP1 × CP1
one point, which we denote as CP
2
f , respectively. Now Pic(Sn ) ≈ Z ⊕ Z, and we can take the Poincaré duals
and CP
of a section of O(−n) and the homology class of the fibre as its generators. The
10.4. REGULAR SASAKIAN STRUCTURES ON 5-MANIFOLDS 363

corresponding divisors can be represented by rational curves which we denote by


C and F, respectively satisfying
C · C = n, F · F = 0, C · F = 1.
Let α1 and α2 denote the Poincaré duals of C and F , respectively. The classes α1
and α2 can be represented by (1, 1) forms ω1 and ω2 , respectively, so that the (1, 1)
form ωl1 ,l2 = l1 ω1 + l2 ω2 determines a circle bundle over Sn whose first Chern class
is [ωl1 ,l2 ]. Since we want the total space our bundle to be simply connected, we
need to choose gcd(l1 , l2 ) = 1. We thus have circle bundles depending on a triple of
integers (l1 , l2 , n), with n non-negative,
π
S 1 −−→Ml1 ,l2 ,n −−→ Sn .
Now in order that Ml1 ,l2 ,n admit a Sasakian structure it is necessary that ωl1 ,l2 be
a positive (1, 1)-form, that is, ωl1 ,l2 must lie within the Kähler cone K(Sn ). The
conditions for positivity are by Nakai’s criterion Theorem 3.5.15,
(i) ωl21 ,l2 > 0,
R
(ii) D ωl1 ,l2 > 0 for all holomorphic curves D,
which in our case give l1 , l2 > 0.
Next we determine the diffeomorphism type of Ml1 ,l2 ,n . Since the Kähler class
[ωl1 ,l2 ] transgresses to the derivative of the contact form, dηl1 ,l2 , we see that π ∗ α1 =
−l2 γ and π ∗ α2 = l1 γ, where γ is a generator of H 2 (Ml1 ,l2 ,n , Z) ≈ Z. Now the first
Chern class of Sn is [GH78b]
(10.4.1) c1 (Sn ) = 2α1 − (n − 2)α2 ,
which pulls back to the basic first Chern class c1 (Fξ ) on Ml1 ,l2 ,n . So the first Chern
class of the contact line bundle D is given by
(10.4.2) c1 (D) = −[2l2 − l1 (2 − n)]γ .
Since Ml1 ,l2 ,n has a regular contact structure, there is no torsion in H2 (Ml1 ,l2 ,n , Z).
Thus, by Theorem 10.2.3 Ml1 ,l2 ,n is either M∞ or X∞ , depending on whether
Ml1 ,l2 ,n is spin or not. But we have w2 (Ml1 ,l2 ,n ) ≡ nl1 mod 2, so Ml1 ,l2 ,n is
diffeomorphic to M∞ = S 2 × S 3 if nl1 is even, and to X∞ if nl1 is odd. Choosing
n ≥ 2 gives countably infinite sequences of regular indefinite Sasakian structures
on S 2 × S 3 and X∞ . It is easy to see [Ful93, BGO07] that for every such triple
(l1 , l2 , n) of positive integers Ml1 ,l2 ,n has a toric Sasakian structure, that is a toric
contact manifold with a compatible Sasakian structure; see Sections 8.4 and 8.5.
The case of regular negative Sasakian structures is notably missing from Propo-
sition 10.4.4. The classification problem for regular negative Sasakian structures
appears to be quite difficult owing to the fact that very little is known about the
existence of simply connected algebraic surfaces of general type with low Picard
number. So we have
Open Problem 10.4.1: Classify the Smale-Barden manifolds that admit regular
negative Sasakian structures.
Even though the classification problem seems to be intractable, it is quite easy
to generate examples.
Example 10.4.6: Regular Negative Sasakian Structures. These are all circle bun-
dles over surfaces of general type. As indicated in Section 5.4.1 there are many
of these, but it appears to be difficult to obtain regular sequences that give all.
364 10. SASAKIAN GEOMETRY IN DIMENSIONS THREE AND FIVE

By constructing the circle bundles over the complex surfaces exhibited in Ta-
bles 3 and 4 we obtain regular negative Sasakian structures on k(S 2 × S 3 ) with
k = 18l(l − 1) + 6, 4l(4l − 3) + 3, and k = 12 l(27l − 31) for all l ≥ 2. Many other
sequences can be obtained by considering BP links with relatively prime weights.
The simplest of these sequences gives the Fermat hypersurfaces of degree greater
than 4 discussed in Example 5.4.1, but the series begins with k = 52. It appears
quite difficult to find examples for low values of k. The lowest k known to the au-
thors admitting a regular negative Sasakian structure is k = 8. This comes from
circle bundles over a Barlow surface [Bar85] which is a complex surface of general
type that is homeomorphic to CP2 blown-up at 8 points.
Example 10.4.7: [Circle Bundle over Fake CP2 ’s] A fake projective plane is a
smooth compact complex surface X 2 having the same Betti numbers as CP2 , but
not CP2 . It is known that such a surface is a projective algebraic surface of general
type and a quotient of the open unit ball in C2 by a cocompact torsion-free discrete
subgroup of the projectivization P U (2, 1) of SU (2, 1). In 1979 Mumford [Mum79]
(see also [BPVdV84], pg. 136) gave the first example of such a surface using p-adic
analysis, and recently there has been a classification by Prasad and Yeung [PY05].
Now π1 (X 2 ) is infinite, and c1 (X 2 ) < 0. Choose the Kähler class of X 2 to be the
positive generator in H 2 (X 2 , Z) ≈ Z, and let π : M 5 −−→X 2 denote the circle bundle
whose first Chern class is this class. By the long exact homotopy sequence, π1 (M 5 )
is also infinite, and by the rational Gysin sequence M 5 is a rational homology
sphere. Since c1 (X 2 ) pulls back to c1 (Fξ ), M 5 has a negative canonical Sasakian
structure.
This example proves the existence of regular negative Sasakian structures on
rational homology 5-spheres M 5 with an infinite fundamental group. Our next
result is another consequence Yau’s celebrated theorem [Yau77] (see also Theorem
V.1.1 of [BPVdV84]). It implies that if such a structure exists on a rational
homology 5-sphere, it must have infinite fundamental group.
Theorem 10.4.8: Let S be a regular Sasakian structure on a rational homology
5-sphere M 5 . Then either
(i) S is positive, and M 5 is S 5 or a lens space L(p, 1, 1) = S 5 /Zp with S in
the standard a-deformation class.
(ii) or S is negative in which case π1 (M 5 ) is infinite.

Proof. First by Proposition 7.5.29 the Sasakian structure is either positive or


negative. Since S is regular, M 5 is the total space of an S 1 -bundle over a compact
complex surface X 2 which by the rational Gysin sequence must have b1 = 0 and
b2 = 1. By Theorem V.1.1 of [BPVdV84] X 2 is either isomorphic to CP2 or
π1 (X 2 ) is infinite. If X 2 ≈ CP2 then M 5 is the total space of the circle bundle
over CP2 with Chern class equal pa, where p is a positive integer and a is the
positive generator of H 2 (CP2 , Z). In this case M 5 is S 5 if p = 1 and the lens
space L(p, 1, 1) for p = 2, 3, . . . . Since any two Kähler forms on CP2 define the
same cohomology class up to a constant, one obtains the standard a-deformation
class on M 5 . Furthermore, the Sasakian structure S is positive since c1 (Fξ ) is the
pullback of c1 (CP2 ) which is positive.
If S is negative then c1 (Fξ ) < 0, and this is the pullback of c1 (X 2 ). Thus,
c1 (X 2 ) is negative so again by Theorem V.1.1 of [BPVdV84] π1 (X 2 ) must be
infinite. But then π1 (M 5 ) is also infinite by the long exact homotopy sequence. ¤
10.4. REGULAR SASAKIAN STRUCTURES ON 5-MANIFOLDS 365

We have an immediate
Corollary 10.4.9: There is a unique regular Sasaki-Seifert structure on S 5 and
the lens spaces L(p, 1, 1), and it contains the standard round metric.

10.4.1. Toric Sasakian Structures on 5-manifolds. Smale-Barden mani-


folds admitting an effective action of a three-dimensional torus T 3 have been studied
by McGavran [McG77], who incorrectly ruled out the non-spin case, and by Oh
[Oh83] who obtained the complete classification. This was based on the earlier clas-
sification of torus actions on 4-manifolds by Orlik and Raymond [OR70, OR74].
The main result in Oh [Oh83] is
Theorem 10.4.10: Let M ∈ B have an effective T 3 -action. Then M is diffeomor-
phic to S 5 , kM∞ or X∞ #(k − 1)M∞ , where k = b2 (M ) ≥ 1. Conversely, all these
manifolds admit effective T 3 actions.
By using an S 1 -equivariant K-contact surgery, Yamazaki [Yam01] proved that
all Oh’s toric 5-manifolds also admit compatible K-contact structures. Then by
invoking Corollary 8.5.7 we obtain
Theorem 10.4.11: Let M be a closed simply connected 5-manifold with an effec-
tive T 3 -action. Then M admits toric Sasakian structures and is diffeomorphic to
S 5 , kM∞ or X∞ #(k − 1)M∞ , where k = b2 (M ) ≥ 1.
In [BGO07] we gave a construction of toric Sasakian structures on Smale-
Barden manifolds by constructing circle bundles over the T 2 -equivariant blow-ups
of Hirzebruch surfaces arriving at
Theorem 10.4.12: Let Sn,k be the equivariant k-fold blow-up of the Hirzebruch
surface Sn . Let πn,k : Mn,k −−→Sn,k be the circle bundle defined by integral Kähler
form ω̃l2 . Then for each positive integer l2 satisfying 2l2 + n > k, the manifold Mn,k
admits a toric regular Sasakian structure, Mn,k is diffeomorphic to kM∞ if n is
even, and if n is odd it is diffeomorphic to X∞ #(k − 1)M∞ . Thus, every regular
contact 5-manifold admits toric regular indefinite Sasakian structures.

Proof. It is well-known [Ful93] that the smooth toric complex surfaces are
all obtained by blowing-up Hirzebruch surfaces at the fixed points of the T 2 action.
Begin with a Hirzebruch surface Sn and blow-up Sn at one of the 4 fixed points
of the T 2 action. This gives a smooth toric surface Sn,1 which can be represented
by a Delzant polytope with 5 vertices. Repeat this procedure inductively to obtain
smooth algebraic toric surfaces Sn,k whose Delzant polytope has k + 4 vertices.
Choose a Kähler class [ω] lying on the Neron-Severi lattice, and construct the circle
bundle πn,k : Mn,k −−→Sn,k whose Euler class is [ω]. The Kähler form can be chosen
to be invariant under the T 2 action, and we can choose a T 2 invariant connection

η in πn,k : Mn,k −−→Sn,k whose curvature form satisfies dη = πn,k ω. This gives a
regular Sasakian structure S = (ξ, η, Φ, g) on Mn,k , and by Theorem 10.4.3 Mn,k is
diffeomorphic to either (k + 1)M∞ or X∞ #kM∞ . Since H 1 (Sn,k , Z) = 0 the torus
action T 2 is Hamiltonian, and thus lifts to a T 2 action in the automorphism group
of the Sasakian structure S by Corollary 8.1.9. Together with the circle group
generated by the Reeb field ξ this makes S a regular toric Sasakian structure.
It remains to show that all of the manifolds in B with no torsion in H2 occur.
For this we need to compute the second Stiefel-Whitney class w2 (Mn,k ) which is
the mod 2 reduction of the first Chern class of contact bundle Dn,k of Mn,k . First
366 10. SASAKIAN GEOMETRY IN DIMENSIONS THREE AND FIVE

we give a Kähler form of the complex manifolds Sn,k constructed in the paragraph
above. Let ω̃l1 ,l2 , ω̃1 , ω̃2 denote the proper transform of ωl1 ,l2 , ω1 , ω2 respectively,
where ωl1 ,l2 is the Kähler form of Example 10.4.5. Then the Kähler form on Sn,k
can be written as
X
(10.4.3) ω̃l1 ,...,lk+2 = li ω̃i ,
i

where ω̃i+2 is the (1, 1)-form representing the Poincaré duals αi+2
˜ of the exceptional
Pk
divisors Ei . Then writing the Kähler class as i=1 li α̃i , the positivity condition
becomes
k
X k
X k
X
2
(10.4.4) 0< li α̃i ∪ li α̃i = l1 (2l2 + nl1 ) − li+1 .
i=1 i=1 i=1

So ω̃l1 ,...,lk+2 defines a Kähler metric if this inequality is satisfied. Here we have used
the fact the exceptional divisor is in the kernel of the corresponding blow-up map.
Define the integer valued k + 2-vector l = (l1 , . . . , lk+2 ). It is convenient to choose
l = (1, l2 , 1, . . . , 1) in which case the positivity condition becomes 2l2 + n > k. For
simplicity we denote the corresponding Kähler form by ω̃l2 .

Next we need to compute the first Chern class of Dn,k . This is πn,k c1 (Sn,k )
modulo the transgression of the Kähler class on Sn,k , that is, modulo the relation

Pk+2
πn,k li α̃i = 0. Let β1 , . . . , βk+1 be a basis for H 2 (Mn,k , Z), and write πn,k

α̃i =
Pk+1 i=1
j=1 m β
ij j . We need to choose the k + 2 by k + 1 matrix (m ij ) such that
k+2
X
(10.4.5) li mij = 0.
i=1

Now using equation (10.4.1) we have


k
X
(10.4.6) c1 (Sn,k ) = 2α̃1 − (n − 2)α̃2 − α̃k+2
i=1

which gives
X X k+2
XX

(10.4.7) πn,k c1 (Sn,k ) = 2 m1j βj − (n − 2) m2j βj − mij βj .
j j i=3 j

We now make a judicious choice of the matrix (mij ).


 
−l2 2 2 ··· 2
 1 0 0 ··· 0
 
 . .. .. ..
 0 −2 .  .
 
(10.4.8) (mij ) =  . . 
 .. 0 .. 0 0
 
 . . 
 .. .. · · · −2 0
0 0 ··· 0 −2
The orthogonality condition (10.4.5) is satisfied and equation (10.4.6) becomes

πn,k c1 (Sn,k ) = 2l2 β1 − 6(β2 + · · · + βk+1 ) − (n − 2)β1 .
It follows that w2 (Mn,k ) ≡ n mod 2. ¤
10.4. REGULAR SASAKIAN STRUCTURES ON 5-MANIFOLDS 367

Such toric contact structures are very numerous. Here for purposes of illustra-
tion we consider only the case of circle bundles over Hirzebruch surfaces. A similar
construction should be possible for the general case, but it is much more tedious.
Let Fl1 ,l2 ,n denote the deformation class of Sasakian structures defined by the above
bundle construction. Then from Example 10.4.5 it is easy to show [BGO07]:
Theorem 10.4.13: For each triple of positive integers (l1 , l2 , m) satisfying gcd(l1 , l2 ) =
1, the manifold S 2 × S 3 admits the deformation classes Fl1 ,l2 ,2m and F2l1 ,l2 ,2m+1 of
regular toric Sasakian structures. For each triple of positive integers (l1 , l2 , m) satis-
fying gcd(l1 , l2 ) = 1, the manifold X∞ admits the deformation classes F2l1 −1,l2 ,2m−1
of regular toric Sasakian structures.
In fact each contact structure can have inequivalent toric contact structures cor-
responding to the number of non-conjugate maximal tori in the contactomorphism
group Con(S 2 × S 3 , η) or Con(X∞ , η) [Kar03, Ler03b, BGO07]. For simplicity
we consider the case n = 2m so S2m is diffeomorphic to S 2 × S 2 . For i = 1, 2 we
let σi denote the classes in H 2 (S 2 × S 2 , Z) given by pulling back the volume form
on the ith factor. Writing the Kähler class [ω] = a1 σ1 + a2 σ2 in terms of the this
basis, we see that
a1 = l1 m + l2 , a2 = l1 ,
and the positivity condition becomes a1 > ma2 > 0. We denote the corresponding
deformation classes of toric Sasakian structures on S 2 × S 3 by F(a1 , a2 , m). The
integers a1 , a2 are written as a, b in [Kar03] and [Ler03b]. In terms of the ai the
first Chern class 10.4.2 simplifies to c1 (D) = 2(a1 − a2 )γ. Thus, F(a1 , a2 , m) and
F(a01 , a02 , m0 ) belong to non-isomorphic contact structures if a01 − a02 6= a1 − a2 . The
following theorem is due to Lerman [Ler03b].
Theorem 10.4.14: For every pair of relatively prime integers (a1 , a2 ) there are
d aa12 e inequivalent regular toric Sasakian structures on S 2 × S 3 having the same
contact form ηa1 ,a2 . However, for each integer m = 0, . . . , d aa21 e the structures
F(a1 , a2 , m) are inequivalent as toric contact structures.
CHAPTER 11

Sasaki-Einstein Geometry

In some sense this chapter is the heart of our book, hence, its length. The
question of the existence of Einstein metrics on compact odd-dimensional mani-
folds has been and remains our main motivation in learning and further developing
Sasakian geometry. Since the time we began our study of Sasaki-Einstein manifolds
in the early nineties, there has been an enormous progress in proving the existence
of such geometries. In 1987 Besse asked1: ‘Are manifolds admitting an Einstein
metric rather scarce or numerous’ ? The first interesting odd (Sasakian) dimension
is dimension five, since in dimension three an Einstein manifold must be of constant
curvature. At the time the question was posed the only Einstein metrics known to
exist on compact simply connected 5-manifolds were: the round sphere metric on
S 5 , the homogeneous Kobayashi-Tanno metric on S 2 × S 3 , the product metric on
S 2 × S 3 , and the symmetric metric on SU (3)/SO(3). The subject of Einstein man-
ifolds has over the years been so dynamic that new important results appear almost
continuously. In order to cope with that phenomenon Besse’s book was appended
by a series of addenda which listed several new results, most of them not even
published at the time. One of them (see [Bes87], Addendum A) describes a result
of Wang and Ziller [WZ86] where it is shown, that using the Kobayashi bundle
construction [Kob63], one can obtain infinitely many inequivalent Einstein metrics
exist on S 2 × S 2m+1 for each m, as well as on the non-trivial S 2m+1 -bundle over S 2 .
So the floodgates were opened and it was soon realized that Einstein metrics tend
to exist in great profusion. About the same time the existence of Kähler-Einstein
metrics on certain smooth del Pezzo surfaces [TY87, Tia87, Siu88] was given
with the complete picture being understood somewhat later; all del Pezzo surfaces
save the one and two point blow-ups of CP2 admit Kähler-Einstein metrics [Tia90].
Then using the Kobayashi bundle construction this result automatically gives rise
to new families of Sasaki-Einstein metrics on 5-manifolds [FK89]. They naturally
generalize the Kobayashi-Tanno metric on S 2 × S 3 . Apart from the manifolds al-
ready mentioned, Tian’s theorem added six more members to the club of compact
5-manifolds admitting Einstein metric of positive scalar curvature. Throughout the
nineties this club remained very exclusive with just a total of merely nine members.
The current chapter of our book describes most advances that have been made in
the last few several years. Like Besse we are facing a difficult problem: new results
seem to appear with increasing frequency. Some of the material we have included
here has not yet been published. Today, 20 years after the publication [Bes87],
Einstein metrics are known to exist on all k-fold connected sums kM∞ = k(S 2 ×S 3 ),
on infinitely many rational homology 5-spheres Mp , as well as on infinitely many
connected sums of these. Most of these are actually Sasaki-Einstein. For many

1
See [Bes87], Chapter 0.I: The Main Problems Today.

369
370 11. SASAKI-EINSTEIN GEOMETRY

manifolds it appears that infinite sequences of Einstein metrics occur and with
large moduli. After 20 years the answer to the Besse’s question appears to be: Ein-
stein metrics in each dimension are numerous. It is quite conceivable that every
compact simply connected 5-manifold admits an Einstein metric. We should also
mention here that at least part of the reason for this explosion of interest in Sasaki-
Einstein metrics comes from Physics. As we discuss in Chapter 14 Sasaki-Einstein
metrics admit Killing spinors, something of great interest to physicists working in
conformal quantum field theory.

11.1. Foundations of Sasaki-Einstein Geometry


Although we are mainly interested in Sasaki-Einstein geometry, there is a nat-
ural generalization of the Einstein condition in the category of Sasakian geometry,
the so-called η-Einstein metrics. It is convenient to present the foundations in this
more general setting. Sasaki-η-Einstein structures are treated in more detail in
Section 11.8.
Definition 11.1.1: A contact metric structure S = (ξ, η, Φ, g) on M is said to be
η-Einstein if there are constants a, b such that Ricg = ag + bη ⊗ η.
Clearly, g is Einstein when b = 0. Little seems to be known concerning Einstein
metrics (or more generally η-Einstein metrics) in the general contact metric case
(see, however, Proposition 11.1.6 below), and as we shall see by Theorem 11.1.7 be-
low that K-contact Einstein implies Sasakian, so we restrict ourselves momentarily
to the Sasakian case. Our first result gives necessary conditions for the existence
of Sasaki-η-Einstein metrics on a Sasakian manifold M. There are restrictions on
both the type of contact structure D as well as the topology of the manifold M in
order for it to admit a Sasaki-η-Einstein structure.
Theorem 11.1.2: Let S = (ξ, η, Φ, g) be a Sasakian structure on M. Then a nec-
essary condition that there is a type II deformation of S to a Sasaki-η-Einstein
structure in F(Fξ ) is that there exists a real number a such that c1 (Fξ ) = a[dη]B .
The underlying contact bundle D must have vanishing real first Chern class, i.e.,
c1 (D) is a torsion class. In particular, a simply connected manifold admitting a
Sasaki-η-Einstein structure must be spin.
Proof. This follows from the discussion in Sections 7.5.1 and 7.5.2. The de-
tails are left as an exercise. ¤
Exercise 11.1: Prove Theorem 11.1.2.
Remarks 11.1.1: The fact that a simply connected Sasaki-Einstein manifold is spin
was first noticed by Moroianu [Mor97]. When F(Fξ ) is positive or zero Theorem
11.1.2 holds under the weaker condition that S = (ξ, η, Φ, g) is K-contact, but
Theorem 11.1.7 below shows that this is a mute point. The type II deformation
(7.5.9) always exists when F(Fξ ) is negative or zero by the transverse Yau Theorem
7.5.19 of El Kacimi-Alaoui.
We shall make use of the following important theorem which for the regular case
is due to Kobayashi [Kob63]. The general quasi-regular case appeared in [BG00b].
Recall from Theorem 7.1.3 that if M is a quasi-regular K-contact manifold with
compact leaves, the space of leaves M/Fξ is an almost Kähler orbifold (Z, h).
Theorem 11.1.3: Let M be a compact manifold of dimension 2n + 1 with a quasi-
regular K-contact structure (ξ, η, Φ, g). Then
11.1. FOUNDATIONS OF SASAKI-EINSTEIN GEOMETRY 371

(i) (ξ, η, Φ, g) is η-Einstein if and only if the almost Kähler orbifold (Z, h)
is Einstein.
(ii) g is Einstein if and only if h is Einstein with scalar curvature 4n(n + 1).
In particular, g is Sasaki-Einstein if and only if h is Kähler-Einstein with
scalar curvature 4n(n + 1).

Proof. An immediate consequence of Theorem 7.3.12 is: a K-contact struc-


ture (ξ, η, Φ, g) is η-Einstein if and only if the transverse metric gT is Einstein,
and g is Einstein if and only if gT is Einstein with Einstein constant 2n + 2. From
Theorem 7.1.3 there is an S 1 orbibundle π : M −−→Z such that the almost Kähler
orbifold (Z, h) satisfies π ∗ h = gT . Furthermore, the Ricci curvature Rich of the
orbifold metric h pulls back to RicT , so (i) follows. Moreover, g will be Einstein if
and only if gT is Einstein with Einstein constant 2n + 2. But this means that the
scalar curvature is 2n(2n + 2). ¤

An immediate corollary of Proposition 7.3.9 or Theorem 7.1.3 and the well-


known Myers’ Theorem is
Corollary 11.1.4: If a K-contact structure (ξ, η, Φ, g) on M 2n+1 is Einstein, then
the Einstein constant equals 2n. Hence, if M 2n+1 is complete, then it is compact
with diameter less than or equal to π and with finite fundamental group.
Recall our discussion of both Riemannian and symplectic cones in Section 6.5.
In particular Proposition 6.5.5 gives a one to one correspondence between contact
metric structures on a manifold M and almost Kähler structures on the cone C(M ).
Furthermore, by Proposition 6.5.6 a contact metric structure (ξ, η, Φ, g) is K-contact
if and only if the vector field Ψ − iξ pseudo-holomorphic with respect to the almost
complex structure on C(M ). We are interested in the case when the cone metric
ḡ = dr2 + r2 g satisfies the Einstein condition Ricḡ = λḡ. The following important
lemma seems to be well-known as folklore.
Lemma 11.1.5: Let (M, g) be a Riemannian manifold of dimension n, and con-
sider (C(M ) = M × R+ , ḡ) the cone on M with metric ḡ = dr2 + r2 g. Then if
ḡ is Einstein, it is Ricci-flat, and ḡ is Ricci-flat if and only if g is Einstein with
Einstein constant n − 1.

Proof. Let {θi }ni=1 be a local orthonormal coframe for (M, g), then we obtain
a local orthonormal coframe {φµ } for the cone metric ḡ on M × R+ by setting
φi = rθi , φ0 = dr ,
where 0 ≤ µ ≤ n. We employ the convention that the Greek indices range from 0 to
n, whereas, the Latin indices range from 1 to n, and we identify M with M × {1}.
Consider the Cartan structure equations (1.3.2)
X
dφµ + ωνµ ∧ φν = 0 ,
ν
X X
dωνµ + ωλµ ∧ ωνλ = µ
R̄νλρ φλ ∧ φρ .
λ λ<ρ

Here ωνµ denotes the connection 1-forms with respect to the Levi-Civita connection
µ
on M ×R+ which satisfies ωνµ +ωµν = 0, and R̄νλρ denotes the Riemannian curvature.
The first Cartan equations for the cone metric gives ω0i = θi , and that ωji are the
372 11. SASAKI-EINSTEIN GEOMETRY

connection 1-forms for the Levi-Civita connection with respect to g on M. The


second equations then imply
i i 1 i
R̄0µν =0 R̄jkl = 2 Rjkl − δ k[i δ j]l ,
r
where δ ik denotes the Kronecker delta symbol, and the brackets mean antisym-
metrize over i and j. Forming the Ricci tensor one now easily proves the result. ¤
Notice that if M has dimension 2n + 1 and (C(M ), ḡ) is Einstein then (M, g) is
Einstein with Einstein constant 2n. Thus, if g belongs to a contact metric structure
S = (ξ, η, Φ, g) and (C(M ), ḡ) is Einstein, then S = (ξ, η, Φ, g) is actually K-contact
by Proposition 7.3.9. Summarizing,
Proposition 11.1.6: If the metric cone (C(M ), ḡ) on a contact metric manifold
is Einstein, then the contact metric structure S = (ξ, η, Φ, g) is K-contact and
Einstein.
The question now arises concerning the consequences of adding to the hypothe-
sis of Theorem 11.1.3 the condition that S be Sasakian. Remarkably such an added
hypothesis is not needed when F(Fξ ) is positive or zero. The following theorem
proved by the authors in [BG01a] can be thought of as an odd-dimensional ver-
sion of the well-known Goldberg Conjecture [Gol69]. It says that if a K-contact
structure is Einstein it is Sasakian, that is the underlying almost CR-structure is
necessarily integrable.
Theorem 11.1.7: Let (M, ξ, η, Φ, g) be a compact K-contact manifold. Then if g
is Einstein (M, ξ, η, Φ, g) is Sasakian.
Proof. We first prove the theorem under the assumption that η is quasi-
regular. By Theorem 7.1.3 M is the total space of a principal S 1 orbibundle over
a compact almost Kähler orbifold Z. Furthermore, by Theorem 11.1.3 the induced
metric h on Z is almost Kähler-Einstein which has positive scalar curvature, since
g has positive scalar curvature. Now since Sekigawa’s [Sek1, Sek2] proof of the
Goldberg conjecture in the case of nonnegative scalar curvature only involves lo-
cal curvature computations together with a Bochner type argument using Stokes
Theorem, it carries over to the case of a compact orbifold. So the almost complex
structure on Z is integrable, and (Z, h) is Kähler-Einstein. It then follows from
Theorem 7.1.3 that (M, ξ, η, Φ, g) is normal, hence, Sasaki-Einstein. This proves
the result under the assumption of quasi-regularity.
Now assume that (M, ξ, η, Φ, g) is K-contact and Einstein, but not quasi-regular.
Then by Proposition 2.6.3 the Reeb vector field ξ lies in the commutative Lie subal-
gebra t(M, Fξ ) ⊂ a(M, g) which has dimension k > 1. Thus, there exists a sequence
of quasi-regular contact forms ηj and Reeb vector fields ξj ∈ t(M, Fξ ) that ap-
proximate (η, ξ) in the compact-open C ∞ topology. (In what follows we use this
topology on the space of smooth sections of all tensor bundles.) Explicitly, there is
a monotonically decreasing sequence {²j }∞ 1 with lim ²j = 0 such that
j→∞

1
(11.1.1) ηj = f (²j )η, ξ j = ξ + ρj , f (²j ) = ,
1 + η(ρj )
where f (²j ) are positive functions in C ∞ (M ) that satisfy lim f (²j ) = 1. Clearly
j→∞
ρj ∈ t(M, Fξ ) and lim ρj = 0. Moreover, ker ηj = ker η = D, so we have the same
j→∞
11.1. FOUNDATIONS OF SASAKI-EINSTEIN GEOMETRY 373

underlying contact structure. We also have the following easily verified relations
for the induced contact endomorphisms Φj :
1
(11.1.2) Φj = Φ − Φρj ⊗ η = Φ − f (²j )Φρj ⊗ η .
1 + η(ρj )
This implies that Φj ξj = 0 and that the almost complex structure J on D remains
unchanged. However, the induced metrics become
gj = f (²j )gD ⊕ f (²j )2 η ⊗ η + df (²j ) ◦ Φ ∧ η − f (²j )dη ◦ (Φρj ⊗ η) ⊗ 1l .
For ²j small enough gj are well defined Riemannian metrics on M which converge
to g, and can easily be seen to satisfy the compatibility conditions
gj (Φj X, Φj Y ) = gj (X, Y ) − ηj (X)ηj (Y ) .
Moreover, since ξj ∈ t ⊂ a(M, η), it follows that the functions f (²j ) ∈ C ∞ (M )T ,
where C ∞ (M )T denotes the subalgebra of C ∞ (M ) invariant under the action of
the torus T. Thus, from equation (11.1.2) we have
Lξj Φj = 0 .
Hence, (M, ηj , ξj , Φj , gj ) is a sequence of quasi-regular K-contact structures on
M whose limit with respect to the compact-open C ∞ topology is the original K-
contact Einstein structure (M, η, ξ, Φ, g). Now the metrics gj are not Einstein, but
their Ricci tensor can be seen to satisfy
Ricgj = λj gj + A(²j , ρj , g) ,
where A(²j , ρj , g) is a traceless symmetric 2-tensor field depending on ²j , ρj , g that
tends to 0 with ²j , and λj ∈ C ∞ (M ) satisfy lim λj = 2n.
j→∞
Now there is a sequence of orbifold Riemannian submersions πj : M −−→Zj ,
where (Zj , hj ) are a sequence of compact almost Kähler orbifolds satisfying πj∗ hj =
f (²j )gD . Moreover, it follows from the above limits that the scalar curvatures of
the hj are all positive. Notice that in Sekigawa’s proof [Sek2] of the positive scalar
curvature Goldberg conjecture, the Einstein condition is not used until Section 4
of [Sek2]. Following [Sek2] and making the necessary adjustments to our situation,
we find that there are nonnegative numbers δj and nonnegative smooth functions
Fj such that
Z
¡ šj ˇ ˇ 2 1 ˇ ˇ 4 ¢
(11.1.3) Fj + ||∇ j Jj ||Zj + ||∇j Jj ||Zj σj ≤ δj ,
Zj n 2n
where ∇ ˇ j , Jˇj , šj , σj and || · ||Z are the Levi-Civita connection, almost complex
j
structure, scalar curvature, volume element, and Riemannian norm, respectively on
(Zj , hj ). Now since the metrics g, gj are bundle-like the leaves of the characteristic
foliation are geodesics and the O’Neill tensors T and N vanish [Ton]. Moreover, for
any K-contact manifold of dimension 2n + 1 the O’Neill tensor A satisfies ||A||2 =
g(Aξ, Aξ) = 2n. Thus, we have the relation between the functions λj on M and the
scalar curvatures šj on Zj :
šj = (2n + 1)λj + 2n .
So that lim šj = 2n+(2n+1) lim λj = 4n(n+1) . Furthermore, we have lim δj = 0.
j→∞ j→∞ j→∞
Thus, since Fj (see [Sek2]) and šj are nonnegative for each j, the estimate 11.1.3
374 11. SASAKI-EINSTEIN GEOMETRY

implies the estimate


ˇ j Jˇj ||Z ≤ δj0 ,
||∇ j

where δj0 are nonnegative numbers satisfying lim δj0 = 0. Now for each j the hori-
j→∞
ˇ Jˇj is the horizontal projection (∇j Jj )h = (∇j Φj )h , where ∇j , Jj , Φj
zontal lift of ∇
are the corresponding Levi-Civita connection and tensor fields with respect to the
metrics on M. But on M Jj = J for all j and we have
||(∇J)h || = lim ||(∇j J)h ||j ≤ lim δj0 = 0 ,
j→∞ j→∞

where || · ||j is the Riemannian norm with respect to gj . So the almost CR structure
on D is integrable which by Proposition 6.5.14 implies that (M, ξ, η, Φ, g) is Sasaki-
Einstein. ¤
Exercise 11.2: Prove the following result of Okumura [Oku62]: If the Ricci tensor
of a Sasakian manifold (M, ξ, η, Φ, g) is parallel, then g is Einstein.
An alternative proof of Theorem 11.1.7 has been given recently by Apostolov,
Draghici, and Moroianu [ADM06]. Instead of using the transverse geometry as
in [BG01a], they work on the metric cone C(M ) by using a Weitzenbock-type
formula to show that if (C(M ), dr2 + r2 g) is almost Kähler and Einstein it must
be Kähler. They also notice that Theorem 11.1.7 cannot be improved in dimension
3 by assuming the weaker hypothesis that (M, ξ, η, Φ, g) is only a contact metric
structure. A simple counterexample is given by the 3-torus Example 6.1.23 with the
flat metric. However, this counterexample does not generalize to higher dimension,
for Blair [Bla76b] has shown that a contact metric structure in dimension 5 or
higher cannot be flat. Thus, it is still an open question as to whether the hypothesis
of Theorem 11.1.7 can be weakened to contact metric structures in dimension 5 and
greater.
There is another important characterization of the Sasaki-Einstein condition
which can be formulated as a corollary of Lemma 11.1.5:
Corollary 11.1.8: Let (M, g) be a Riemannian manifold. Then the metric g
is Sasaki-Einstein if and only if the cone metric ḡ is Calabi-Yau, i.e., (C(S), ḡ)
is Kähler Ricci-flat. In particular, it follows that the restricted holonomy group
Hol0 (ḡ) ⊂ SU (n + 1).
Now Hol0 (ḡ) is the normal subgroup of the full holonomy group Hol(ḡ) that is
the component connected to the identity. There is a canonical epimorphism
π1 (M ) = π1 (C(M )) −→ Hol(ḡ)/Hol0 (ḡ) ,
so if M is simply connected its structure group reduces to 1 × SU (n) and it will
admit a spin structure. So for Sasaki-Einstein manifolds we have a generalization
of the last statement of Theorem 11.1.2.
Corollary 11.1.9: Let M be a Sasaki-Einstein manifold such that the full holo-
nomy group of the cone metric Hol(ḡ) is contained in SU (n + 1). Then M admits
a spin structure.
We give some examples that illustrate the complications in the presence of
fundamental group. The hypothesis of this corollary is not necessary as the second
example shows.
Example 11.1.10: The real projective space M = RP2n+1 with its canonical metric
is Sasaki-Einstein, and the cone C(M ) = (Cn+1 \ {0})/Z2 with the usual antipodal
11.1. FOUNDATIONS OF SASAKI-EINSTEIN GEOMETRY 375

identification. We have Hol(ḡ) ' π1 (M ) ' Z2 . When n is odd the antipodal


map τ is in SU (n + 1), so M = RP2n+1 admits a spin structure. But when n
is even the antipodal map τ does not lie in SU (n + 1), which obstructs a further
reduction of the structure group. In this case it is well-known that M = RP2n+1
does not admit a spin structure. In fact the generator of Hol(ḡ) ' Z2 is the
obstruction. There are many other similar examples. An example that shows that
the hypothesis in Corollary 11.1.9 is not necessary is the following: Consider the
lens space L(p; , q1 , . . . , qn ) ' S 2n+1 /Zp where the qi ’s are relatively prime to p.
The action on Cn+1 \ {0} is generated by (z0 , z1 , . . . , zn ) 7→ (ηz0 , η q1 z1 , . . . , η qn zn ),
where η is a primitive pth root of unity. It is known P [Fra87] that if p is odd,
L(p; q1 , . . . , qn ) admits a spin structure. However, if i qi + 1 is not divisible by p,
the holonomy group Hol(ḡ) ' Zp does not lie in SU (n + 1).
Now combining the results from Theorems 7.5.1 and 11.1.7 we have
Theorem 11.1.11: Let S = (ξ, η, Φ, g) be a quasi-regular K-contact structure on
a compact manifold M of dimension 2n + 1, and let Z = (Z, ∆) denote the orbifold
structure on the space of leaves Z of the characteristic foliation. If g is Einstein,
then S is Sasakian and
(i) Z is an orbifold log Fano variety with an orbifold Kähler-Einstein metric
h with scalar curvature 4n(n + 1) and the projection π : M −−→Z is a
Seifert S 1 -bundle (equivalently and S 1 -orbibundle) with geodesic fibres.
(ii) If S is regular then π : M −−→Z is a principal circle bundle over the smooth
Fano manifold Z with a Kähler-Einstein metric.
The Inversion Theorem 7.5.2 in the Sasaki-Einstein setting gives
Theorem 11.1.12: Let Z be a compact Fano orbifold with π1orb (Z) = 0. Let π :
corb
1 (Z)
M −−→Z be the S 1 -orbibundle whose first Chern class is . Then π1orb (M ) = 0
Ind(Z)
and M is smooth if the local uniformizing groups of Z inject into S 1 . Furthermore,
there is a metric g on the total space M such that M is a compact simply connected
Sasakian manifold with Ricg > −2. If Z has an orbifold Kähler-Einstein metric h
then there is a unique metric g on M given by g = π ∗ h + η ⊗ η which is Sasaki-
Einstein.

11.1.1. Homogeneous Sasaki-Einstein Manifolds. There is one special


case where the hypothesis of Theorem 11.1.7 can be weakened from a K-contact
structure to an arbitrary metric contact structure, namely for compact homoge-
neous contact manifolds (M, η). This is a consequence of Theorem 8.3.6. From the
proof of that theorem we may assume that (M, η) is homogeneous under a compact
Lie group G. We then let Gc denote its complexification. We have
Theorem 11.1.13: Let (M, η, G) be a compact homogeneous contact manifold with
Reeb vector field ξ. Suppose also that (M, η, G) admits a compatible homogeneous
Einstein metric g. Then (ξ, η, g, G) is a homogeneous Sasaki-Einstein structure and
is an S 1 -bundle over a generalized flag manifold Gc /P. Conversely, given any gen-
eralized flag manifold Gc /P there is a circle bundle π : M −−→Gc /P whose total
space M is a simply connected homogeneous Sasaki-Einstein manifold, and all ho-
mogeneous Sasaki-Einstein manifolds are of the form M/Zk for some integer k ≥ 1.

Proof. The first statement is an immediate corollary of Theorem 8.3.6. To


prove the converse we notice that a theorem of Matsushima [Bes87] says every
376 11. SASAKI-EINSTEIN GEOMETRY

Gc /P admits a G invariant positive Kähler-Einstein metric with Kähler form ω.


Moreover, there is a subgroup U ⊂ G such that Gc /P = G/U. One then chooses
any circle bundle given by an integral class lying on the positive ray in the Kähler
cone defined by ω. One then obtains a Sasaki-Einstein metric on the total space M
of this circle bundle by Theorem 11.1.3. By construction one easily sees that this
metric is homogeneous. For each Gc /P and each k ∈ Z+ we get a homogeneous
S 1 -bundle over Gc /P with fundamental group π1 = Zk . It can be obtained as a
discrete Zk -quotient of the unique simply connected model. ¤

Note that the Stiefel manifold V2 (Rn+1 ) of Example 9.3.20 is a homogeneous


Sasaki-Einstein manifold that can be realized as the link of a quadric.
A somewhat weaker version of Theorem 11.1.13 was first given in [BG00b].
The following corollary lists all the possible G/P in the first three dimensions
[BG99, BG00b]:
Corollary 11.1.14: Let (M, S) be a compact homogeneous Sasaki-Einstein mani-
fold of dimension 2n + 1. Then M is a circle bundle over
(i) CP1 when n = 1, implying M = S 3 ;
(ii) CP2 or CP1 × CP1 when n = 2, implying M = S 5 or M = S 2 × S 3 ;
(iii) CP3 , CP1 × CP2 , CP1 × CP1 × CP1 , the complex flag F3,2,1 = SU (3)/T 2 ,
or the real Grassmannian Gr2 (R5 ) when n = 3, implying M = S 7 , the
1,2
Wang-Ziller manifold M2,3 , an S 2 × S 3 bundle over S 2 , the 3-Sasakian
homogeneous manifold SU (3)/S 1 , or the real Stiefel manifold V2 (R5 ),
respectively.
Exercise 11.3: Using the classification of low-dimensional homogeneous Einstein
manifolds [BK06] give the list of all homogeneous Sasaki-Einstein manifolds in
dimensions 9 and 11.
To end this subsection we mention a Sasaki-Einstein version of a sphere theorem
Theorem 11.1.15: Let M be a compact simply connected Sasaki-Einstein manifold
of positive sectional curvature. Then M is isometric to the round unit sphere.
It appears that the theorem was established for Sasaki-Einstein 5-manifolds
with additional assumption on sectional curvature κ ≥ 1/3 and in its full generality
by Moskal [Mos66]. However, Moskal’s thesis has never been published. On the
other hand, the proof of this theorem can be found in the Sasaki lecture notes (cf.
[Sas68], Theorem 45.1), which are perhaps better known, but have never really
been published either.

11.1.2. Moduli of Sasaki-Einstein Structures. Recall that Proposition


8.2.6 affirms that the infinitesimal deformations of the transverse holomorphic struc-
ture of the characteristic foliation of a quasi-regular Sasakian structure can be de-
scribed in terms of two pieces, the deformations of the underlying complex structure
which are elements of H 1 (Z, ΘZ ), and the deformation of the characteristic foliation
itself which can be characterized by the Sasaki cone t+ +
k (D, J) = cr (D, J)/CR(D, J)
as described by Corollary 8.2.14. We believe that a Sasaki-Einstein structure, if it
exists, will be unique within t+ k (D, J) with a fixed J. This is formulated as Conjec-
ture 11.1.18 below. So the deformations that we consider here come from deforming
the complex structure of the base orbifold. These deformations were discussed in
Section 5.5. We shall make use of
11.1. FOUNDATIONS OF SASAKI-EINSTEIN GEOMETRY 377

Proposition 11.1.16: Let (Z1 , h1 ) and (Z2 , h2 ) be two Kähler-Einstein Fano orb-
ifolds and M1 and M2 the corresponding Sasaki-Einstein manifolds. Assume also
that M1 is not a metric of constant sectional curvature one. Let φ : M1 → M2 be
an isometry. Then there is an isometry φ̄ : Z1 → Z2 which is either holomorphic
or anti-holomorphic, such that the following diagram commutes:
φ
M
1 −−−−→ M
2
 
 
π1 π2
y y
φ̄
Z1 −−−−→ Z2 .
1
Moreover, φ̄ determines φ up to the S -action induced by the Reeb vector field.

Proof. Since φ is an isometry we can assume that M2 = M1 = M and g2 =


g1 = g. Since g does not have constant curvature, Lemma 8.1.17 implies that either
φ∗ S2 = S1 , φ∗ S2 = S1c the conjugate Sasakian structure, S1c = (−ξ1 , −η1 , −Φ1 , g1 ),
or φ∗ S2 and S1 form part of a 3-Sasakian structure defined by Definition 13.1.9
below. In the first case φ ∈ Aut(S1 ), and in the second case it induces complex
conjugation on the base, that is Z2 = Z̄1 with the same orbifold metric. Finally
when φ∗ S2 and S1 form part of a 3-Sasakian structure, Z1 and Z2 are isomorphic
as complex Kähler-Einstein contact orbifolds by Theorem 13.3.1 below.
Conversely, any such biholomorphism or anti-biholomorphism φ̄ lifts to an or-
bibundle map φ : M1 −−→M2 uniquely up to the S 1 -action given by the Reeb vector
field. ¤

It is clear that our procedure produces inequivalent Sasaki-Einstein structures,


but Proposition 11.1.16 implies a much stronger result, namely
Theorem 11.1.17: Let M be a smooth compact manifold and let S = (ξ, η, Φt , gt )
be a family of quasi-regular Sasaki-Einstein structures on M induced by a continu-
ous family of inequivalent complex orbifolds Zt with Kähler-Einstein metrics. Then
the metrics gt are inequivalent as Einstein metrics.

Proof. The constant sectional curvature metric is rigid as an Einstein metric


[Bes87] and is covered by an S 1 -bundle over complex projective space. There are
no families of inequivalent complex structures on projective space, so none of the gt
can have constant sectional curvature. So Proposition 11.1.16 applies to conclude
that the gt are all inequivalent as Einstein metrics. ¤

Theorems 11.1.17 and 5.5.7 can be used to give a lower bound on the dimension
of the moduli space of Einstein metrics on many manifolds. We give many such
examples later in the chapter. Notice also that Theorem 11.1.17 only addresses
the moduli problem for quasi-regular structures. Nothing appears to be known
about the moduli problem for Sasaki-Einstein structures in the irregular case. As
we shall see irregular Sasaki-Einstein metrics do exist by Theorem 11.4.5 below. As
mentioned previously an interesting question is whether one can deform through
Sasaki-Einstein metrics by deforming the Reeb vector field within the Sasaki cone
t+
k (D, J). We think not, but we really have no good reason other than the lack of
any example.
Conjecture 11.1.18: Sasaki-Einstein metrics occur for at most one ξ ∈ t+
k (D, J).
378 11. SASAKI-EINSTEIN GEOMETRY

This is certainly true when k = 1 by Theorem 11.8.2 below.

11.2. Extremal Sasakian Metrics


In this section we define extremal metrics in the context of Sasakian geometry.
These2 were first introduced in [BGS06] and they naturally generalized the notion
of both Sasaki-Einstein metrics and Sasaki-η-Einstein metrics. Just as in the Kähler
case the concept is very useful in studying the obstruction theory which will be
discussed in further sections. We begin by introducing a Riemannian functional
whose critical points will describe an important subset of Sasakian structures in
¯ Our presentation follows [BGS06] closely.
F(ξ, J).
We denote by M(ξ, J) ¯ the set of all compatible Riemannian Sasakian metrics
¯ and define the ‘energy’ functional E : M(ξ, J)
arising from structures in F(ξ, J), ¯ →
R by
Z
(11.2.1) E(g) = s2g dvolg
M
2
which is just the L -norm of the scalar curvature of g.
The variation of a metric in M(ξ, J) ¯ depends upon the two functions ϕ and ψ
of Proposition 7.5.7. However, the transverse Kähler metric varies as a function of
ϕ only, and does so within a fixed basic cohomology class. The critical point of
(11.2.1), should it exist, will allow us to fix the parameter ϕ, which is determined
by the condition that dη̃ = dη + i∂ ∂ϕ ¯ be an extremal transverse Kähler metric
[Cal82]. The remaining gauge function parameter ψ represents nothing more than
a change of coordinates in the representation of the form η̃ of the Sasaki structure
in question. Thus, finding a critical point of E produces an extremal representative
¯
of F(ξ, J).

11.2.1. Variational Formulae and the Euler-Lagrange Equations. In


order to derive the Euler-Lagrange equation of (11.2.1), we describe the infinitesimal
variations of the volume form, Ricci tensor, and scalar curvature, as we deform the
¯
metric in M(ξ, J).
¯ that starts at (ξ, η, Φ, g)
Proposition 11.2.1: Let (ξ, ηt , Φt , gt ) be a path in F(ξ, J)
when t = 0, and is such that dηt = dη + ti∂ ∂ϕ ¯ for certain basic function ϕ, and for
t sufficiently small. Then we have the expansions
dvolt = (1 − 2t ∆B¡ϕ))dvol + O(t2
¢ ), 2
ρt ¯ 1
= ρ − ti∂ ∂ 2 ∆¡ B ϕ + ϕ + O(t ) , ¢
st ¯
= sT − 2n − t 12 ∆2B ϕ + 2(ρT , i∂ ∂ϕ) + O(t2 ) ,
for the volume form, Ricci form, and scalar curvature of gt , respectively. Here, the
geometric terms without sub-index are those corresponding to the starting metric g.

Proof. The proof in [BGS06] is a straightforward adaption of [Sim04] to


the Sasaki case. ¤

2In [BGS06] such metrics were called canonical Sasakian metrics. But we believe that this
terminology can be misleading, suggesting that it refers to Sasakian structures that are canonical
or anticanonical as given by Definition 7.5.24. So we prefer to call such metrics extremal (cf.
Definition 11.2.4 below) which is also more in agreement with the standard usage.
11.2. EXTREMAL SASAKIAN METRICS 379

¯ we introduce a
Associated to any Sasakian structure (ξ, η, Φ, g) in F(ξ, J),
B
basic differential operator Lg of order 4 whose kernel consists of basic functions
with transverse holomorphic gradient. Given a basic function ϕ : M 7→ C, we
consider the vector field ∂ # ϕ defined by the identity
(11.2.2) ¯ .
g(∂ # ϕ, · ) = ∂ϕ
Thus, we obtain the (1, 0)-component of the gradient of ϕ, a vector field that,
generally speaking, is not transversally holomorphic. In order to ensure that, we
¯ # ϕ = 0, that is equivalent to the fourth-order
would need to impose the condition ∂∂
equation
(11.2.3) (∂∂ ¯ #ϕ = 0 ,
¯ # )∗ ∂∂
¯ # )∗ ∂∂
because hϕ, (∂∂ ¯ # ϕiL2 = k∂∂¯ # ϕk2 2 .
L
We have that
¯ # ∗¯ # 1 2 ¯
(11.2.4) LB T T
g ϕ = (∂∂ ) ∂∂ ϕ = (∆B ϕ + 4(ρ , i∂ ∂ϕ) + 2(∂s ) ∂ ϕ) .
#
4
The functions on M that are transversally constant are always in the kernel of LB g .
The only functions of this type that are basic are the constants. Thus, the kernel
of LB
g has dimension at least 1.
Proposition 11.2.2: The first derivative of E at g ∈ M(ξ, J) ¯ in the direction of
the deformation defined by (ϕ, ψ) is given by
Z
d ¡ ¢
E(gt ) |t=0 = −4 ¯ # )∗ ∂∂
(sT − 2n) (∂∂ ¯ # ϕ dvol .
dt M

Proof. This result follows readily from the fact that s = sT − 2n, the varia-
tional formulae of Proposition 11.2.1, and identity (11.2.4). ¤
Theorem 11.2.3: A Sasaki metric g ∈ M(ξ, J) ¯ is a critical point of the energy
functional E of (11.2.1) if and only if the basic vector field ∂g# sTg = ∂g# sg is transver-
sally holomorphic.
So as in the Kähler case the Euler-Lagrange equations boil down to holomor-
phicity of the gradient of the scalar curvature. We are thus led to our fundamental
definition:
Definition 11.2.4: We say that S = (ξ, η, Φ, g) is an extremal Sasakian struc-
ture in F(ξ, J)¯ or that g is an extremal Sasakian metric if g satisfies the
condition of Theorem 11.2.3, that is if and only if g is transversally extremal.
11.2.2. Sasaki-Futaki Invariants. In this section we adapt the construction
of the well-known Futaki invariant to the Sasakian category. Again our presentation
¯ the
follows [BGS06] closely. Given any Sasakian structure (ξ, η, Φ, g) ∈ F(ξ, J),
¯ T
metric g is an element of M(ξ, J) whose transverse Ricci form ρ is basic. We
define the Ricci potential ψg as the function in the basic Hodge decomposition of
ρT given by
¯ g,
ρT = ρTh + i∂ ∂ψ
2
where ρTh is the harmonic representative of the basic cohomology class in HB (Fξ )
T T
represented by ρ . Notice that if Gg is the Green’s operator of the transverse
metric, we have that
ψg = −GTg (sTg ) = −GTg (sTg − 2n) = −GTg (sg ) = −Gg (s − sg,0 ) ,
380 11. SASAKI-EINSTEIN GEOMETRY

where sg and Gg are the scalar curvature and Green’s operator of g, and sg,0 is the
projection of sg onto the constants. The sequence of equalities above follows from
the fact that sg = sTg − 2n is a basic function. Thus, the Ricci potential ψg is itself
a basic function.
On a Sasakian manifold we have a transverse holomorphic structure, and thus
¯ of transversally holomorphic vector
from Definition 2.5.22 the Lie algebra hT (Fξ , J)
fields. So in analogy to the Kähler case equation (5.3.2) we have
Definition 11.2.5: With F(ξ, J) ¯ fixed we define a map SFξ,J¯ : hT (Fξ , J)−
¯ −→C by
Z
SF(X)ξ,J¯ = X(ψg )dvolg .

We call SF = SFξ,J¯ the Sasaki-Futaki invariant.


Since ψg is basic, the integrand in this expression can be fully written in terms
of the transversally holomorphic realization ΞX = 21 (XD − iJ(XD )) of X. Our next
result shows that SF is well-defined as a Lie algebra character.
Proposition 11.2.6: The map SFξ,J¯ only depends on the basic cohomology class
[dη]B (equivalently only on the space F(ξ, J))¯ and not on the particular Sasakian
¯
structure in F(ξ, J).
¯ starting at g for which the transverse
Proof. We take a path gt in M(ξ, J)
Kähler form is of the form
¯ ,
dηt = dη + ti∂ ∂ϕ
with the affine parameter ϕ a basic function. From the identity ∆B ψg = sTg,0 − sT ,
we see that the variation ψ̇g of ψg satisfies the relation

2(i∂ ∂ϕ, ¯ g )g + ∆B ψ̇g = −s˙T = 1 ∆2 ϕ + 2(ρT , i∂ ∂ϕ)


¯ i∂ ∂ψ ¯ g.
2 B
Hence,
Z
1 1 ¯ g,
ψ̇g − ψ̇g dvolg = ∆B ϕ + 2GTg (ρTh , i∂ ∂ϕ)
v 2
where v is the volume of M in the metric.
Since ρTh is harmonic, the last summand in the right side can be written as
−2GTg (∂ ∗ (∂¯∗ (ϕρTh ))). For convenience, let us set β = ∂¯∗ (ϕρTh ). Hence,
Z Z µ ¶
d 1 1
X(ψt )dvolgt = X ∆B ϕ − 2GTg (∂ ∗ β)) − ψ∆B ϕ dvolg .
dt 2 2
By the Ricci identity for the transverse metric, we have that
1
(∆B ϕ)α = −ϕ,γαγ + ϕ,γ (ψ,αγ + (rhT )αγ ) = −ϕ,γαγ + ϕ,γ ψ,αγ + βα ,
2
and so, after minor simplifications, we conclude that
Z Z
d γ
X(ψt )dvolgt = Ξα
X (ϕ,γ ψα − ϕ,γα ) dvolg +
dt Z
Ξα ∗
X (βα − 2(Gg ∂ β),α ) dvolg ,

where ΞX is the (1, 0)-component of XD .


11.2. EXTREMAL SASAKIAN METRICS 381

The first summand on the right above is zero because ΞX is holomorphic. This
is just a consequence of Stokes’ theorem. The second summand is also zero since
we have
Z Z
ΞαX (βα − 2(Gg ∂ ∗
β),α ) dvolg = (β − ∆B GTg β, Ξ[X )dvolg +
Z
(2∂ ∗ ∂GTg β, Ξ[X )dvolg ,

and β − ∆B GTg η = 0 while ∂Ξ[X = 0. Here, of course, Ξ[X is the (0, 1)-basic form
corresponding the (1, 0)-vector field ΞX . ¤

Proposition 11.2.7: The following hold


(i) SF(X) = 0 for any section X of the trivial line bundle Lξ .
¯
(ii) SF([X, Y ]) = 0 for any pair of vector fields X, Y ∈ hT (Fξ , J).
We leave the proof as
Exercise 11.4: Prove Proposition 11.2.7.
Our next proposition extends to extremal Sasakian metrics a now well-known
result in Kähler geometry originally due to Futaki [Fut83]. The Sasaki version
here is analogous to the expanded version of Futaki’s result presented by Calabi
[Cal85].
¯
Theorem 11.2.8: Let S = (ξ, η, Φ, g) be an extremal Sasakian structure in F(ξ, J).
Then, the metric g has constant scalar curvature if and only if SF( · ) = 0.

Proof. In one direction the statement is obvious: a constant scalar curvature


Sasakian metric has trivial Ricci potential function, so the Sasaki-Futaki invariant
¯
SF vanishes on every X ∈ hT (Fξ , J).
In order to prove the converse, we first observe that if X ∈ hT (Fξ , J) ¯ is a
#
transversally holomorphic vector field of the form X = ∂g f for some basic function
f, then
Z Z
SF(X) = − ∂g# f (Gg sg )dvolg = −2 (∂f, ¯ ∂G¯ T sg )g dvolg
g
Z
= −2 f (∂¯g∗ ∂G¯ T sg )dvolg ,
g

because the scalar curvature sg is a basic function also. Since 2∂¯∗ ∂¯ = ∆B , we


conclude that Z
#
SF(∂g f ) = − f (sg − sg,0 )dvolg .

Now, if the Sasakian metric g is a critical point of the energy function E in (11.2.1),
then ∂g# s = ∂ # sT = ∂g# sg is a transversally holomorphic vector field, and we
conclude that Z
SF(∂g# sT ) = − (sg − sg,0 )2 dvolg .

Thus, if SF( · ) = 0, then sg must be constant. ¤

Corollary 11.2.9: Let (ξ, η, Φ, g) be an extremal Sasakian structure in F(ξ, J). ¯


¯
Then the metric g ∈ M(ξ, J) is Sasaki-Einstein if and only if the basic first Chern
class satisfies 2πc1 (ν(Fξ )) = 2n[dη]B , and the Sasaki-Futaki invariant SFξ,J¯ van-
ishes.
382 11. SASAKI-EINSTEIN GEOMETRY

Proof. If g is Sasaki-Einstein, then g must have positive Einstein constant 2n,


and all the implications follow readily. Conversely, if g is extremal in F(ξ, J) ¯ and
SFξ,J¯ vanishes, its scalar curvature is constant. It follows that the scalar curvature
of the transverse metric is constant also, and this implies that ρg + 2dη = ρT is
transversally harmonic. As this form represents 2πc1 (Fξ ), which is also represented
by a constant multiple of dη, the uniqueness of the harmonic representative of a
class implies that ρT = λdη for some λ > 0. It then follows from this that the Ricci
tensor Ricg is a constant multiple of g. Thus, g is Einstein. ¤

Example 11.2.10: The only example so far where one has an explicit computation
of the Sasaki-Futaki invariant is for the weighted Sasakian structures in the standard
CR structure on the sphere S 2n+1 , and even for this case the computations are
somewhat involved. We refer to Examples 7.1.12 and 8.4.39 for a discussion of
these weighted spheres and their toric geometry. Indeed, in [BGS06] the Sasaki-
Futaki invariant, SFξw ,J¯w , is computed for an arbitrary weight vector w ∈ (R+ )n+1 ,
and it is shown that for the standard CR structure on S 2n+1 the entire Sasaki
cone t+n+1 (D, J) of Corollary 8.2.14 is represented by extremal Sasakian metrics.
Moreover, an explicit expression for SFξw ,J¯w is given which shows that it vanishes
if and only if w = w(1, . . . , 1) for some real number w. Moreover, the only Sasaki-
Einstein metric is when w = 1, proving the veracity of Conjecture 11.1.18 for
this case. Note that identifying the Lie algebra tn+1 with its dual t∗n+1 gives an
identification of the Sasaki cone t+ n+1 with the moment cone minus the cone point,
C(Υ) − {0}.

11.3. Further Obstructions to Sasaki-Einstein Structures


We have already discussed several obstructions to the existence of a Sasaki-
Einstein structure (in fact, of a Sasaki-η-Einstein structure). First there is Theo-
rem 11.1.2 which says in order for a Sasakian manifold (M, S) to admit a Sasaki-
Einstein structure, the first Chern class c1 (D) of the contact bundle D of S must
be a torsion class in H 2 (M, Z). Moreover, the basic first Chern class c1 (Fξ ) must
be positive. Both of these are conditions on the underlying contact CR structure.
But the first actually gives a topological obstruction, namely if π1 (M ) = {id}, then
w2 (M ) must vanish for M to admit a Sasaki-Einstein structure. If these obstruc-
tions vanish then Corollary 11.2.9 above says that the non-vanishing of the Sasaki-
Futaki invariant is another obstruction to the existence of a Sasaki-Einstein metric
¯ We now give three more obstructions the first of which is
within the class F(ξ, J).
a straightforward generalization of Theorem 5.3.1 as well as Calabi’s generalization
[Cal85]. The other two obstructions are quite new, arising from recent observations
of Gauntlett, Martelli, Sparks, and Yau [GMSY06].

11.3.1. Calabi-Matsushima-Lichnerowicz Obstruction. First notice that


¯ is infinite-dimensional, since the sections Γ(Lξ ) form a sub-
the Lie algebra hT (ξ, J)
algebra. However, owing to (i) of Proposition 11.2.7 it is really only the quotient
¯
algebra hT (ξ, J)/Γ(L ξ ) that is of interest. We are now ready to extend to the
Sasakian context a result of Calabi [Cal85] on the structure of the algebra of holo-
morphic vector fields of a Kähler manifold with an extremal metric. This result is
an extension of work of Lichnerowicz [Lic57] on constant scalar curvature metrics,
itself an extension of a result of Matsushima [Mat57b] in the Kähler-Einstein case.
11.3. FURTHER OBSTRUCTIONS TO SASAKI-EINSTEIN STRUCTURES 383

We also refer the reader to the theorem for harmonic Kähler foliations in [NT88],
which is relevant in this context.
¯ Let HgB be the space of
Consider a Sasaki structure S = (ξ, η, Φ, g) in F(ξ, J).
B
basic functions in the kernel of the operator Lg in (11.2.4), and consider the map
(11.3.1) ¯
∂g# : HgB → hT (ξ, J)/Γ(L ξ) ,

where ∂g# is the operator defined in (11.2.2). We use the Sasaki metric g to identify
the quotient space in the right side above with the holomorphic sections of (D, J),
which we shall refer to from here on as h(D, J). We also define the operator L̄Bg on
B B B
Hg by L̄g ϕ = Lg ϕ̄. It follows that
(11.3.2) (L̄B B #
g − Lg )ϕ = ∂g sg ∂ϕ − ∂g# ϕ ∂sg ,
where sg is the scalar curvature of g, a basic function.
The image h0 ∼= HgB /C of the map (11.3.1) is an ideal in h(D, J), and can be
identified with the space of holomorphic fields that have non-empty zero set. The
quotient algebra h(D, J)/h0 is Abelian.
Suppose now that (ξ, η, Φ, g) is an extremal representative of F(ξ, J),¯ so that g
#
is an extremal Sasakian metric. Let z0 be the image under ∂g of the set of purely
imaginary functions in HgB . This is just the space of Killing fields for the transverse
metric g T that are of the form J∇gT ϕ, ϕ ∈ HgB . Furthermore, by (11.3.2) we see
¯ 0 coincides with the commutator of ∂g# sg :
that the complexification z0 ⊕ Jz
¯ 0 = {X ∈ h(D, J) : [X, ∂g# sg ] = 0} .
z0 ⊕ Jz
Notice that we are implicitly using the fact here that ∂g# sg defines a section of
h(D, J), as sg is a basic function. We come to the main result of this section which
first appeared in [BGS06].
Theorem 11.3.1: Suppose that there exists an extremal Sasakian structure (ξ, η, Φ, g)
¯ Let HgB be the space of basic functions in the kernel of the operator LB
in F(ξ, J). g in
(11.2.4), and let h0 be the image of the map (11.3.1). Then we have the orthogonal
decomposition
¯
hT (ξ, J)/Γ(L ∼
ξ ) = h(D, J) = a ⊕ h0 ,
where a is the algebra of parallel vector fields relative to the transverse metric g T .
Furthermore,
h0 = z0 ⊕ Jz¯ 0 ⊕ (⊕λ>0 hλ ) ,
where z0 is the image of the purely imaginary elements of HgB under ∂g# , and hλ =
¯
{X ∈ hT (ξ, J)/Γ(L # ¯
ξ ) : [X, ∂g sg ] = λX}. In particular, the Lie algebra aut(J, gT )
T
of Killing vector fields for the transverse metric g is equal to
aut(J,¯ g T ) = a ⊕ z0 ∼= a ⊕ aut(ξ, η, Φ, g)/{ξ} .
The presence of the algebra a above does not contradict the absence of non-
trivial parallel vector fields on a closed Sasakian manifold. A vector field can be
parallel with respect to g T without being parallel with respect to g.
Proof. We sketch the main points, as the argument is an adaptation to our
situation of that in [Cal85, Sim04]. Given a section X in h(D, J), we look at the
Hodge decomposition of the (0, 1)-form that corresponds to it via the metric g T . It
¯
is ∂-closed, and both, its harmonic and ∂¯ components, are the dual of holomorphic
fields. The vector field dual to the harmonic component is g T -parallel. Since g T is
384 11. SASAKI-EINSTEIN GEOMETRY

an extremal metric, the operators LB B B


g and L̄g commute. We then restrict L̄g to Hg ,
B

and use the resulting eigenspace decomposition together with the identity (11.3.2) to
prove the first part of the theorem. The identification of z0 with aut(ξ, η, Φ, g)/{ξ}
is essentially a consequence of the exact sequence 8.1.4. ¤

Remark 11.3.1: This result obstructs the existence of special extremal Sasakian
structures in the same way it does in the Kählerian case. For instance, let (N, ω)
be the one-point or two-point blow-up of CP2 , and consider the circle bundle
π : M → N with Euler class [ω]. Now M has a natural Sasakian structure
S = (ξ, η, Φ, g) satisfying π ∗ ω = dη, then F(ξ, J) ¯ contains no Sasakian structure
(ξ, η̃, Φ̃, g̃) with g̃ a metric of constant scalar curvature. By Theorem 11.3.1 the
¯
structure of hT (Fξ , J)/Γ(L ξ ) obstructs it.

11.3.2. Bishop’s Obstruction. In the following two subsections we will briefly


discuss two obstructions recently discovered by Gauntlett, Martelli, Sparks, and
Yau [GMSY06]. It appears that the new obstructions are non-trivial only in the
case when the fundamental foliation is not regular. For this reason they have gone
unnoticed as obstructions to the existence of both positive Kähler-Einstein metrics
and Sasaki-Einstein metrics. The first obstruction is quite interesting arising from
a classical inequality due to Bishop which first appeared in a short note of Bishop
in the Notices of the American Mathematical Society in 1963, and then appeared in
the 1964 book of Bishop and Crittenden which has been recently reprinted by AMS
Chelsea [BC01]. An extended version due to Gromov [Gro81, Gro99] appears in
most recent texts in Riemannian geometry, cf. [Pet98, Ber03]. We only need the
original Bishop’s inequality.
Theorem 11.3.2: Let (M n , g) be a compact Einstein manifold of positive Einstein
constant normalized so that the metric cone (C(M ) = R+ × M, ḡ = dr2 + r2 g) is
Ricci-flat, so Ric(g) = (n − 1)g by Lemma 11.1.5. Then
vol(M, g) ≤ vol(S n , gcan ) ,
where gcan is the unit round sphere metric on S n .
Now, let M 2n+1 be a compact manifold with Sasakian structure S = (ξ, η, Φ, g).
Recall that by Proposition 7.5.10 the volume of any Sasakian metric in F(ξ) is the
same. On the other hand, if one includes transverse homotheties and consider
F(Fξ ), one can scale the volume of a Sasakian metric to an arbitrary chosen value.
However, it follows from Theorem 11.8.2 below that if a Sasaki-Einstein metric in
F(Fξ ) exists, it is unique within the transverse homothety class.
Definition 11.3.3: Let M 2n+1 be a compact manifold with a Sasakian structure
S = (ξ, η, Φ, g). The ratio
vol(M 2n+1 , g)
N (ξ) =
vol(S 2n+1 , gcan )
is called the normalized volume of the Sasakian structure S.
Let us rephrase Theorem 11.3.2 for the special case of Sasakian structures.
Corollary 11.3.4: Let M 2n+1 be a compact manifold with Sasakian structure S =
(ξ, η, Φ, g). If there is a Sasaki-Einstein metric in F(Fξ ), for some ξ 0 = cξ then
N (ξ 0 ) ≤ 1.
11.3. FURTHER OBSTRUCTIONS TO SASAKI-EINSTEIN STRUCTURES 385

Of course, a priori it is not clear that the inequality of Corollary 11.3.4 should
lead to any obstruction. However, we know from Proposition 7.5.10 that the volume
of any Sasakian structure (ξ, η, Φ, g) ∈ F(ξ) is the same. Suppose that one could
compute the volume of any, hence all structures in F(ξ). If such a computation
were to yield a normalized volume greater than one, Corollary 11.3.4 would imply
that none of the Sasakian structures in F(ξ) could be Sasaki-Einstein. Indeed, this
was recently observed to be the case in [GMSY06]. We stress here that one needs
the techniques of [MSY06a, MSY06b] to actually compute the volumes in many
cases. In the theory surrounding Maldacena’s AdS/CFT conjecture the normalized
volume of the Sasaki-Einstein horizon metric is related to the coupling constants
of the dual theory. This will be discussed in the last chapter. In [MSY06a]
Martelli, Sparks and Yau describe the methods of computing the volumes of Sasaki-
Einstein metrics in some special situations.
Example 11.3.5: Let L∗ (w; d) be the link of an isolated hypersurface singularity
with weight vector w = (w0 , . . . , wn ) and degree d. Since the singularity is isolated
the surface Xd ⊂ CPn (w) is quasi-smooth. The physicists have computed the
normalized volume N (w; d, n) of such a link and have shown that
d(|w| − d)n
(11.3.3) N (w; d, n) = Qn .
nn i=0 wi
This formula was first derived by Bergman and Herzog [BH02] under the assump-
tion that the link is well-formed, but this assumption is unnecessary as was noticed
in [MSY06b]. Hence, we have [GMSY06]
Theorem 11.3.6: Let L∗ (w; d) be a smooth link with the Sasakian structure de-
fined by diagram (9.2.3). If
n
Y
d(|w| − d)n > nn wi
i=0

then L∗ (w; d) cannot admit any Sasaki-Einstein structure in F(Fξ (w; d)).
We shall see explicitly in one of the next sections that this obstruction rules
out Sasaki-Einstein metrics in case of many links. Here we give two examples.
Example 11.3.7: Consider the link L(2, 2, . . . , 2, p) ⊂ S 2m+1 , where n = m > 2.
When m and p are both odd that this is Kervaire sphere. Assume p > 2 is odd.
Then we have d = 2p and I = pm − 2p + 2. Hence, the Bishop obstruction enters
when
(pm − 2p + 2)m > mm pm−1 .
The same inequality is obtained for p even. For any fixed m at p = 2 we always have
(2m − 2)m < mm 2m−1 . However, at some p = p0 (m) the left hand side dominates.
For instance, when m = 3, we have (p + 2)3 > 27p2 which happens when p > 20.
Another way of saying this is that L(2, 2, 2, p) has its normalized volume bigger
than 1 for p > 20. Bishop obstruction works even better for larger m. When m = 4
we have (p + 1)4 > 16p3 which is satisfied from p > 11. When m = 5 we get p > 8
and so on.
Example 11.3.8: Another interesting application of the Theorem 11.3.6 is the case
of the weighted projective space CP(w). Take f (z) = z0 , for example. This is a
homogeneous polynomial of degree d = w0 and its vanishing defines a hyperplane
in CPn (w) which is again a weighted projective space. In particular, L∗ (w; w0 )
386 11. SASAKI-EINSTEIN GEOMETRY

is simply the weighted Sasakian sphere with weights ŵ = (w1 , . . . , wn ) fibering


over the weighted projective space CPn−1 (ŵ) of one lower dimension. The normal-
ized volume formula of a hypothetical Sasaki-Einstein metric gives the following
beautiful expression
µ 1 Pn ¶n
2n−1 n i=1 wi
N (ξŵ ) = vol(Sŵ ) = p Qn .
i=1 wi
n

As the arithmetic average is greater than or equal to the geometric average we get,
Proposition 11.3.9: Let w = (w0 , . . . , wn ) and Sw be the weighted Sasakian struc-
ture on S 2n+1 . Then the nth root of the normalized volume N (Fξw ) of a hypothetical
Sasaki-Einstein metric equals the ratio of the arithmetic to the geometric averages
of all the weights. In particular, Bishop’s inequality obstructs the existence of a
Sasaki-Einstein metric in F(Fξw ) unless all the weights are equal, in which case we
get the standard Sasakian structure on the unit sphere.

11.3.3. Lichnerowicz’s Obstruction. One of the first results concerning


eigenvalues of the Laplace operator on a compact Riemannian manifold was proved
by Lichnerowicz. It is buried on page 135 of his famous book [Lic57]. The second
part of the statement is due to Obata [Oba62]. For further discussion see [Ber03].
Theorem 11.3.10: Let (M n , g) be a compact Riemannian manifold with Ricci
curvature Ric(g) ≥ n − 1. Then the first non-zero eigenvalue λ1 of the Laplacian
satisfies λ1 ≥ n. Furthermore, the equality λ1 = n happens only for manifolds
isometric to the sphere.
It easily follows from the above theorem that
Corollary 11.3.11: Let M 2n+1 be a compact manifold with a Sasaki-Einstein struc-
ture S = (ξ, η, Φ, g). Then the first non-zero eigenvalue λ1 of the Laplace operator
∆g is bounded
λ1 ≥ 2n + 1
and λ1 = 2n+1 if and only if S is the standard Sasaki-Einstein structure on S 2n+1 .
Following [GMSY06] we shall see that Lichnerowicz bound naturally leads to
an obstruction to the existence of Sasaki-Einstein metrics on certain links. Let M
be a Sasakian manifold of dimension 2n − 1 with Sasakian structure S = (ξ, η, Φ, g)
and let Y = C(M ) be the associated cone with the induced Kähler structure. Let
f be a holomorphic function on Y with

(11.3.4) £ξ f = cif ,
where c > 0 is a real constant called charge of f with respect to ξ. Since f is
holomorphic, this immediately implies that f = rc f˜, where f˜ is homogeneous of
degree zero under the Euler field Ψ = r∂r , In particular, f˜ is the pull-back to Y of
a function on the base of the cone M . Moreover, since (Y, ḡ) is Kähler, ∆Y f = 0,
where ∆Y = −∇2Y is the Laplacian on (Y, ḡ). For a metric cone, this is related to
the Laplacian on the base (M, g) at r = 1 by
µ ¶
1 1 ∂ ∂
(11.3.5) ∆Y = ∆M − 2n−1 r2n−1 .
r2 r ∂r ∂r
From this, one sees that
11.3. FURTHER OBSTRUCTIONS TO SASAKI-EINSTEIN STRUCTURES 387

(11.3.6) ∆M f˜ = λf˜ ,
where λ = c[c + (2n − 2)]. Thus any holomorphic function f of positive charge w.r.t
ξ, or, equivalently, degree zero under Ψ, corresponds to an eigenfunction of the
Laplacian on the base M. The charge c is then related simply to the eigenvalue λ
by the above formula.
Let us now assume that, (Y, ḡ) is Ricci-flat Kähler, or equivalently that (M, g)
is Sasaki-Einstein. The first non-zero eigenvalue λ1 of ∆M is bounded from below
as λ1 ≥ 2n − 1. This is of course Lichnerowicz’s λ1 -theorem. In terms of the charge
c the Lichnerowicz’s bound becomes c ≥ 1. This leads to a simple new holomorphic
obstruction to the existence of Sasaki-Einstein metrics on many links as well as an
obstruction to the existence of Kähler Ricci-flat Calabi-Yau cones.
Theorem 11.3.12: Let M be a (2n − 1)-dimensional manifold with a positive
Sasakian structure S = (ξ, η, Φ, g), and let Y be the cone with base M and its
induced Kähler structure. Suppose also that the complex structure induced by S on
the cone Y has c1 (Y ) = 0, and there exists a holomorphic function f on Y of charge
0 < c ≤ 1 with respect to the Reeb vector field ξ. Then Y admits no Ricci-flat cone
metric compatible with the Kähler structure and Reeb vector field ξ unless c = 1, in
which case the cone Y is smooth and biholomorphic to Cn . In particular, the space
of Sasakian structures F(Fξ ) on M does not have any Sasaki-Einstein structure
unless c = 1 and (M, g) is the standard round sphere.
It was shown in [GMSY06] by using a well-known result of Kobayashi and
Ochiai [KO73] that this new obstruction is non-trivial only in the non-regular
case. Again, we consider some simple examples of links.
Example 11.3.13: Let L∗ (w; d) be a link of isolated hypersurface singularities with
weight w = (w0 , . . . , wn ) and of degree d. Let us further assume that the link is a
smooth manifold, i.e., the surface Xd ⊂ CPn (w) is quasi-smooth. Further assume
that w0 ≤ · · · ≤ wn , It is easy to see that there exists a holomorphic function on
the cone with charge at least

nw0
(11.3.7) c= .
|w| − d
Hence, we get
Theorem 11.3.14: Let L∗ (w; d) be a smooth link with the Sasakian structure de-
fined by diagram (9.2.3). If the index
I = |w| − d > nw0
then L∗ (w; d) cannot admit any Sasaki-Einstein structure in F(Fξ(w;d) ).
We shall see explicitly in one of the next sections that this obstruction rules
out Sasaki-Einstein metrics in the case of many links. Here we just show that some
classical Brieskorn and Kervaire spheres are obstructed.
Example 11.3.15: Consider the Brieskorn spheres of dimension 2m + 1 given by
BP links L(2, . . . , 2, 3, 6k + 6) ⊂ S 2m+3 , where n = m + 1 and m is odd. Since
d = 6(6k+5), w = (6, 2(6k+5), 3(6k+5), . . . , 3(6k+5)) and I = (6k+5)(3m−4)+6.
Hence, the Lichnerowicz obstruction excludes links with I > 6(m + 1) which holds
388 11. SASAKI-EINSTEIN GEOMETRY

for any k ≥ 0 and any m ≥ 3. Note the when m is even the corresponding link is
diffeomorphic to the standard sphere S 2m+1 and it is still obstructed.
The second is a generalization of the example considered in [GMSY06]. Con-
sider the link L(2, 2, . . . , 2, p) ⊂ S 2m+1 , where n = m > 2. When m and p are both
odd that this is Kervaire sphere. Assume first p > 2 is odd (the inequality one gets
at the end does not depend on this assumption). Then d = 2p and w = (2, p, . . . , p)
and I = p(m − 2) + 2. The Lichnerowicz obstruction excludes links with I > 2m.
This gives
m−1
2<p<2
m−2
which can possibly be Sasaki-Einstein. When m = 3 this leaves L(2, 2, 2, 3) and
L(2, 2, 2, 4). As p = 4 realizes the equality it would have to be S 5 and it is not
so that it can also be excluded. This leaves L(2, 2, 2, 3). When m = 4 the only
possibility is L(2, 2, 2, 2, 3) in which case the equality holds so we can exclude this
link because it is a rational homology sphere but not S 7 . Hence, we get
Proposition 11.3.16: The positive BP links L(2, . . . , 2, 3, 6k+5) cannot admit any
Sasaki-Einstein structure in F(ξ(w; d)). If p > 2 the positive BP links L(2, . . . , 2, p)
cannot admit any Sasaki-Einstein structure in F(ξ(w; d)) with the possible exception
of L(2, 2, 2, 3) ' S 5 .
Remark 11.3.2: Note, that just as it is the case with the Bishop Obstruction the
weighted Sasakian structure Sw cannot admit a Sasaki-Einstein metric in F(ξw )
by Lichnerowicz obstruction, unless all weights are equal. In particular, we get
three different obstructions to the existence of a Kähler-Einstein metric on weighted
projective spaces.

11.4. Sasaki-Einstein Metrics in Dimensions Five


The only compact Sasaki-Einstein manifolds in dimension 3 are spherical space
forms. Dimension 5 is thus the first interesting case from the point of view of the
existence or non-existence of any Sasaki-Einstein metrics. Indeed, as we shall see,
one encounters a true cornucopia of interesting examples even on very simple man-
ifolds such as S 5 or S 2 × S 3 which is M∞ in the Smale-Barden notation employed
in Chapter 10. At the same time, at least in some special cases, one would hope for
some partial classification results. The question of which Smale manifolds admit
a quasi-regular Sasaki-Einstein structure is a question about the nature of the in-
clusions (11.4.1). As discussed in chapter 10 the set of positive Sasakian manifolds
SB+ is already known to be a relatively small subset. A complete answer would
require a classification of cyclic log del Pezzo surfaces which at the orbibundle level
should describe SB+QR B 0 . In addition one needs an orbifold version of the Tian’s
theorem which would answer the questions about the nature of the first inclusion
SE QR ⊆ SB+QR . The two problems are quite different in nature: the first has to do
with positivity in the presence of quotient singularities, the second with stability.
Both tasks appear to be quite hard. So a complete description of all Smale mani-
folds which admit some Sasaki-Einstein structure is probably out of reach at this
time. The problem of describing the moduli space of all Sasaki-Einstein structures
on a given Smale manifold M appears even more challenging. In the notation of
Chapter 10, ultimately we would like to understand all the inclusions
(11.4.1) SE R ⊆ SE QR ⊆ (SB+ )0 ⊆ B 0 .
11.4. SASAKI-EINSTEIN METRICS IN DIMENSIONS FIVE 389

We devote this section to a brief discussion of all that is known about them. For
organizational purposes we divide our discussion according to whether a Smale
manifold M has torsion in H2 (M, Z) or not. Moreover, since Sasaki-Einstein struc-
tures are ubiquitous on k(S 2 × S 3 ) we consider separately the case of toric Sasaki-
Einstein structures from the non-toric case. We treat the latter first.

11.4.1. Non-toric Sasaki-Einstein Structures on k(S 2 × S 3 ). We begin


with the case of regular Sasaki-Einstein manifolds, i.e., the inclusion SE R ⊆ B 0 .
By (ii) of Theorem 10.2.26 any M ∈ SBR has no torsion in H2 (M, Z). Furthermore,
the Einstein condition eliminates the appearance of X∞ , so we are left with S 5 and
k(S 2 × S 3 ), and these must fibre over a smooth del Pezzo surface with a Kähler-
Einstein metric. But as discussed at the end of Section 5.2 [Tia90] determined
precisely which compact complex surfaces admit positive Kähler-Einstein metrics,
so we can easily obtain a complete classification of regular Sasaki-Einstein struc-
tures. Furthermore, we actually get a complete description of the moduli of regular
Sasaki-Einstein structures in each case due to Tian’s Theorem 5.2.18. More pre-
cisely [FK89, BFGK91, BG00b]
Theorem 11.4.1: Let M be a simply connected five manifold with a regular Sasaki-
Einstein structure. Then M is diffeomorphic to kM∞ = k(S 2 × S 3 ), k ≥ 0 (recall
that k = 0 means S 5 ) with 0 ≤ k ≤ 8. Moreover,
(i) for each k = 0, 1, 3, 4, there is precisely one regular Sasaki-Einstein struc-
ture on k(S 2 × S 3 ) up to isometries,
(ii) for each 5 ≤ l ≤ 8 there is a 4(k − 4) parameter family of inequivalent
regular Sasaki-Einstein structures on k(S 2 × S 3 ).
The regular Sasaki-Einstein structures on S 5 and S 2 ×S 3 are homogeneous and
by Corollary 11.1.14 (ii) these are the only homogeneous simply connected Sasaki-
Einstein 5-manifolds. The Sasaki-Einstein metric on M∞ = T1 (S 3 ) was discovered
independently by Kobayashi [Kob63] and Tanno [Tan79] and we will call it the
Kobayashi-Tanno metric. It follows from the classification of toric smooth del Pezzo
surfaces that these two and 3(S 2 × S 3 ) are the only regular toric simply connected
Sasaki-Einstein manifolds in dimension 5.
Turning to the quasi-regular case, although we can find huge numbers of families
of Sasaki-Einstein structures on k(S 2 × S 3 ), there is no complete classification since
there is no classification of log del Pezzo orbifold surfaces admitting Kähler-Einstein
orbifold metrics even in the case of no branch divisors. With no branch divisors
present it follows from Theorem 4.7.14 that the total space of an S 1 -orbibundle is
diffeomorphic to some k(S 2 × S 3 ) for some k = 0, 1, . . . (again k = 0 means S 5 ).
There are actually three distinct constructions of Sasaki-Einstein structures on
these manifolds. One method uses the examples of log del Pezzo surfaces obtained
as quasi-smooth well-formed hypersurfaces Xd ⊂ CP3 (w) of degree d and index
I = |w| − d in weighted projective spaces. The techniques developed in Chapter 5
then can be used to show that some of these admit orbifold Kähler-Einstein metrics.
By Corollary 4.7.15 the well-formedness condition implies that the total space M
of the orbibundle has no torsion in H2 (M, Z). The first examples of quasi-regular
but not regular metrics on S 2 × S 3 and 2(S 2 × S 3 ) were obtained via this method in
[BG01b]. Later Johnson and Kollár classified all anti-canonically embedded (index
I = 1) log del Pezzo hypersurfaces in CP3 (w) which can possibly be shown to admit
an orbifold Kähler-Einstein metric using these methods [JK01b]. Their analysis
390 11. SASAKI-EINSTEIN GEOMETRY

produced an extensive list of available examples and an even longer list of possible
candidates. We further extended this classification to a higher index I < 10 in
[BGN02a, BGN03b]. The combined results of [Ara02, BG03] give two families
of quasi-regular non-regular Sasaki-Einstein structures on 8(S 2 × S 3 ). All the rele-
vant log del Pezzo surfaces are described in section 5.4 and in the tables of Appendix
B.2. Put together they give families of quasi-regular Sasaki-Einstein structures on
k(S 2 × S 3 ) for 0 < k < 9. The construction of [BGK05] with slightly improved es-
timates of [GK05] gives finite number of additional examples in the k = 0 (sphere)
case. Finally, in [Kol07] Kollár uses the more general approach of Seifert bundles
to prove that for every integer k ≥ 6 there are infinitely many 2(k − 1)-dimensional
families of Einstein metrics on k(S 2 × S 3 ). Let us summarize all these results in
the following theorem. For each k we give the minimum number of known distinct
Sasaki-Einstein structures as well as the (real) dimension of the space of effective
parameters of the largest known family. This gives a lower bound on the dimension
of the premoduli space of Einstein metrics on these manifolds. Summarizing, we
have
Theorem 11.4.2: Consider the Smale manifolds kM∞ = k(S 2 × S 3 ). We have
kM∞ ∈ SE QR for all k ≥ 0. Moreover
(i) S 5 admits at least 80 distinct non-regular quasi-regular Sasaki-Einstein
Seifert structures with the largest known family of dimension 10.
(ii) S 2 × S 3 admits at least 14 distinct non-regular quasi-regular Sasaki-
Einstein Seifert structures.
(iii) With the exception of k = 2, 8 the manifolds kM∞ = k(S 2 × S 3 ) admit
infinitely many distinct families of inequivalent non-regular quasi-regular
Sasaki-Einstein Seifert structures. The largest family depends on 2(k −1)
parameters.
(iv) 2(S 2 × S 3 ) admits at least two distinct 2 parameter families plus 21 in-
equivalent non-regular quasi-regular Sasaki-Einstein structures.
(v) 8(S 2 × S 3 ) admits infinitely many inequivalent non-regular quasi-regular
Sasaki-Einstein structures. The largest family has 16 parameters.

Proof. The proof of this theorem combines Theorems 5.4.16 and 11.1.12 to-
gether with the analysis in Section 5.5 and subsection 11.1.2 as well as Kollár’s
results in [Kol07] for k > 9. The details are left to the reader. ¤

It is very likely that 2(S 2 × S 3 ) admits infinitely many inequivalent Sasaki-


Einstein structures and, if so, could be combined with the statement for general
k. There are good candidates of three countable families of examples given by the
following sequences of log del Pezzo surfaces:
X12k+11 ⊂ CP2 (8, 4k + 1, 4k + 3, 4k + 5) , X12k+11 ⊂ CP2 (9, 3k + 2, 3k + 5, 6k + 1) ,

X36k+24 ⊂ CP2 (6, 6k + 5, 12k + 8, 18k + 9) .


They can be found in the first table of Appendix B.2. We were not able to prove
the existence of Kähler-Einstein metrics in either case but we believe that these
examples can be shown to admit such metrics via a more subtle analysis similar to
the one introduced in [BGN03b].
On the other hand 8(S 2 × S 3 ) appears to be somewhat special in that the
largest known family has dimension 16 and not 14 as predicted by (iii) of Theorem
11.4. SASAKI-EINSTEIN METRICS IN DIMENSIONS FIVE 391

11.4.2. Indeed, the examples of Kollár in [Kol07] give distinct infinite families de-
pending on 14 parameters. But the two log del Pezzo surfaces X10 ⊂ CP2 (1, 2, 3, 5)
and X15 ⊂ CP2 (1, 3, 5, 7) of index 1 are actually known to be Kähler-Einstein and
they give two distinct 16-dimensional families of non-regular Sasaki-Einstein struc-
tures. Notice in addition that the regular Sasakian structure on 8(S 2 × S 3 ) also
has dimension 16.

11.4.2. Toric Sasaki-Einstein Geometry. It follows from Oh’s Classifica-


tion Theorem 10.4.10 and Theorem 11.1.2 that toric Sasaki-Einstein structures can
only occur on the Smale manifolds kM∞ = k(S 2 × S 3 ). In addition, there are infin-
itely many distinct Sasaki-Einstein structures on S 2 × S 3 but unlike the examples
used to establish Theorem 11.4.2 they have to do with Sasakian toric geometry
and arise via a completely different construction. Considering just the well-formed
quasi-smooth hypersurfaces in CP3 (w) produces only the 14 examples of Theorem
11.4.2 with 4 more likely candidates all listed in Table B.3.2. The first toric exam-
ples were discovered by Gauntlett, Martelli, Sparks, and Waldram. It follows that
S 2 × S 3 admits infinitely many distinct quasi-regular toric Sasaki-Einstein struc-
tures [GMSW04b]. Toric geometry of these examples was further explored in
[MS05, MSY06a, MSY06b].
Consider the symplectic reduction of Cn (or equivalently the Sasakian reduction
of S 2n−1 ) as discussed in Section 8.5 by a k-dimensional torus T k . Every complex
representation of a T k on Cn can be described by an exact sequence
fΩ
0−−→T k −−→ T n −−→T n−k −−→0 .

The monomorphism fΩ can be represented by the diagonal matrix


à k k
!
Y ai Y a i
fΩ (τ1 , . . . , τk ) = diag τi 1 , . . . , τi n ,
i=1 i=1

where (τ1 , .., τk ) ∈ S 1 × · · · × S 1 = T k are the complex coordinates on T k , and


aiα ∈ Z are the coefficients of a k × n integral weight matrix Ω ∈ Mk,n (Z). The
following proposition follows easily from Theorem 8.5.3 or 8.5.4.
Proposition 11.4.3: Let X(Ω) = (Cn \ 0)//T k (Ω) denote the Kähler quotient of
the standard flat Kähler structure on (Cn \0) by the weighted Hamiltonian T k -action
with an integer weight matrix Ω. Consider the Kähler moment map
n
X
(11.4.2) µiΩ (z) = aiα |zα |2 , i = 1, . . . , k .
α=1

If all minor k × k determinants of Ω are non-zero then X(Ω) = C(Y (Ω)) is a


cone on a compact Sasakian orbifold Y (Ω) of dimension 2(n − k) − 1 which is the
Sasakian reduction of the standard Sasakian structure on S 2n−1 . In addition, the
projectivization of X(Ω) defined by Z(Ω) = X(Ω)/C∗ is a Kähler reduction of the
complex projective space CPn−1 by a Hamiltonian T k -action defined by Ω and it is
the transverse space of the Sasakian structure on Y (Ω) induced by the quotient. If
X
(11.4.3) aiα = 0, ∀ i = 1, . . . , k
α
392 11. SASAKI-EINSTEIN GEOMETRY

then c1 (X(Ω)) = c1 (D) = 0. In particular, the orbibundle Y (Ω)−→Z(Ω) is anti-


canonical. Moreover, the cone C(Y (Ω)), its Sasakian base Y (Ω), and the transverse
space Z(Ω) are all toric orbifolds.
Remark 11.4.1: The conditions on the matrix Ω that assure that Y (Ω) is a smooth
manifold are straightforward to work out. They involve gcd conditions on certain
minor determinants of Ω.
This proposition is nicely summarized by the reduction Diagram 8.5.3 which
in the current context becomes
CPn−1 ←− S 2n−1 ←− Cn \ (0)
(11.4.4) ⇓ ⇓ ⇓
Z(Ω) ←− Y (Ω) ←− C(Y (Ω)).
Both the toric geometry and the topology of Y (Ω) depend on Ω. Furthermore,
Y (Ω) comes equipped with a family of Sasakian structures. When n − k = 3,
assuming that Y (Ω) is simply connected (which is an additional condition on Ω),
we must have m(S 2 × S 3 ) for some m ≤ k. We will be mostly interested in the case
when m = k.
Example 11.4.4: This is a continuation of Example 8.5.5. Let us rephrase this
in terms of the notation of Proposition 11.4.3. So we consider Ω = (k, k, −l, −l).
1,1
The quotient Y (Ω) is the Wang-Ziller manifold Mk,l ' S 2 × S 3 which fibres over
1 1
Z(Ω) ' CP × CP . From Example 10.4.5, the first Chern class of the contact
1,1
bundle D is c1 (D) = 2(k − l)γ where γ is a generator of H 2 (Mk,l , Z). Thus, we see
that the only way to get an anticanonical Sasakian structure is to have k = l. This
gives anticanonical structures on S 2 × S 3 and its Zk quotients (S 2 × S 3 )/Zk . These
are known to admit Sasaki-Einstein metrics. In fact, the Sasaki-Einstein metric on
S 2 × S 3 is just the homogeneous Kobayashi-Tanno metric of Theorem 11.4.1.
Let us generalize this construction to an arbitrary circle reduction. We take
matrix Ω of Proposition 11.4.3 to be Ω = (p, −q) = (p1 , p2 , −q1 , −q2 ), where
p1 , p2 , q1 , q2 ∈ Z+ . We consider complex coordinates (x1 , x2 , y1 , y2 ) ∈ C4 with the
corresponding circle action (x1 , x2 , y1 , y2 ) 7→ (eip1 θ x1 , eip2 θ x2 , e−iq1 θ y1 , e−iq2 θ y2 ).
The moment map in these coordinates corresponding to this circle action is
µ = p1 |x1 |2 + p2 |x2 |2 − q1 |y1 |2 − q2 |y2 |2 .
So µ−1 (0) is again diffeomorphic to S 3 ×S 3 . If we denote the infinitesimal generators
of the circle action in the coordinates xi by Hi and those for yi by Li , then the
infinitesimal generator of this circle action is p1 H1 + p2 H2 − q1 L1 − q2 L2 , and
this circle will act freely on S 3 × S 3 as long as gcd(pi , qj ) = 1 for all i, j = 1, 2.
The Reeb vector field for the induced Sasakian structure can be taken as R =
p1 H1 + p2 H2 + q1 L1 + q2 L2 . So the induced Sasakian structure on Mp,q ≈ S 2 × S 3
is an orbibundle over CP(p1 , p2 ) × CP(q1 , q2 ). If we let α and β denote generators
2 2
for Horb (CP(p1 , p2 ), Z) and Horb (CP(q1 , q2 ), Z), respectively, we see that the first
Chern class of CP(p1 , p2 ) × CP(q1 , q2 ) is c1 = (p1 + p2 )α + (q1 + q2 )β. So pulling
back to Mp,q we find c1 (D) = π ∗ c1 = (p1 + p2 − q1 − q2 )γ, where γ is a generator for
H 2 (S 2 ×S 3 , Z) ≈ Z. So the Sasakian structure is anticanonical if and only if p1 +p2 =
q1 + q2 in agreement with Proposition 11.4.3. Recall from our previous discussion,
cf. Example 11.2.10, that weighted projective spaces with unequal weights do not
admit Sasaki-Einstein metrics. Thus, the Sasakian structure associated with this
orbibundle does not admit a compatible Sasaki-Einstein metric.
11.4. SASAKI-EINSTEIN METRICS IN DIMENSIONS FIVE 393

Nevertheless, Gauntlett, Martelli, Sparks, and Waldram [GMSW04b] gave


an explicit construction of a Sasaki-Einstein metric that is related to these toric
examples, but with a different Reeb vector field. In fact, they gave the first examples
of irregular Sasaki-Einstein metrics. They proved the following remarkable result
in the special case when Ω = (p, −q) = (p, p, −p + q, −p − q), where p and q are
relatively prime nonnegative integers with p > q. (The general case for k = 1 was
treated later in [CLPP05, MS05], see Remark 11.4.1 below). Let us simplify
notation a bit by writing X(Ω) = X p,q and Y (Ω) = Yp,q . Then we get:
Theorem 11.4.5: The 5-manifolds Yp,q are diffeomorphic to S 2 ×S 3 . Furthermore,
Yp,q admits a Sasaki-Einstein structure Sp,q = (ξp,q , ηp,q , φp,q , gp,q ) for any co-prime
p, q such that q < p. With the exception of (p, q) = (1, 0) all the metrics gp,q are
of cohomogeneity one. The isometry groups equal SO(3) × U (1)2 when both p, q
are odd, and U (2) × U (1) when one of them is even, respectively. The Sasaki-
Einstein structure Sp,q is quasi-regular if and only if
(11.4.5) 4p2 − 3q 2 = n2
which happens for infinitely many p, q. Otherwise, the Sasaki-Einstein structure is
irregular of rank 2. The normalized volume of (Yp,q , gp,q ) is given by the following
formula
q 2 [2p + (4p2 − 3q 2 )1/2 ]
(11.4.6) N (Yp,q ) =
3p2 [3q 2 − 2p2 + p(4p2 − 3q 2 )1/2 ]
and it is rational if and only if the Sasaki-Einstein structure is quasi-regular. Lastly,
g1,0 is the homogeneous Kobayashi-Tanno metric on S 2 × S 3 .
Proof. The original proof of this theorem can be found in [GMSW04b] and
the toric aspect of the metric was later described in [MS06]. The argument relies
on the explicit local expression for these metrics. Let us begin with the following
two-parameter family of local metrics
1 − cy 1 q(y)
ds2 = (dθ2 + sin2 θdφ2 ) + dy 2 + (dψ − cos θdφ)2
6 w(y)q(y) 9
2
(11.4.7) + w(y) [dα + f (y)(dψ − cos θdφ)] ≡ ds2B + w(y)[dα + A]2 ,
where
2(a − y 2 )
w(y) = ,
1 − cy
a − 3y 2 + 2cy 3
q(y) = ,
a − y2
ac − 2y + y 2 c
(11.4.8) f (y) = ,
6(a − y 2 )
and ds2B is a “transverse metric”. We stress that ds2B is not the transverse Kähler-
Einstein metric of ds2 but rather some transverse metric. This will be discussed
later. When c = 0 one can show that this has the local form of the Kobayashi-Tanno
metric on S 2 × S 3 . In such a case, it can also be seen that (p, q) = (1, 0) and the
metric is simply obtained via the Sasakian reduction of S 7 with zero momentum
level. When c 6= 0 one can scale the constant so that c = 1, which we shall assume
from now on. One can easily show that (11.4.7) is Einstein with Ric(g) = 4g for all
values of the constants a, c. To show that the metric is Sasaki-Einstein one needs
394 11. SASAKI-EINSTEIN GEOMETRY

to exhibit the Reeb vector field which will be done later. At this stage we have
a one-parameter family of metrics. We need to examine for what values of the
constant a one can get a complete metric on a compact manifold.
The compactification argument is based on the properties of the solutions of
the cubic polynomial Q(y) = a − 3y 2 + 2y 3 . The analysis of [GMSW04b] first
showed that the transverse space B can be made into a smooth complete compact
manifold with appropriate choices for the ranges of the coordinates. In particular,
for a ∈ (0, 1) one can take the ranges of the coordinates (θ, φ, y, ψ) to be 0 ≤ θ ≤ π,
0 ≤ φ ≤ 2π, y1 ≤ y ≤ y2 , 0 ≤ ψ ≤ 2π so that the base space B is an axially
squashed S 2 -bundle over the round 2-sphere. The latter is parameterized by θ, φ,
with ψ being an azimuthal coordinate on the axially squashed S 2 fibre. This bundle
is topologically trivial, i.e., B ∼
= S 2 × S 2 . The range of y is fixed so that 1 − y > 0,
2
a − y > 0, w(y) > 0, q(y) ≥ 0. Specifically, yi are two zeroes of q(y), i.e., are two
roots of the cubic Q(y). We have
Proposition 11.4.6: The metric extends to a complete Sasaki-Einstein metric
on S 2 × S 3 if and only if 0 < a < 1 and the polynomial Q(y) has three roots
y1 < 0 < y2 ≤ y3 such that y1 − y2 ∈ Q.
Proof. If 0 < a < 1 there are three real roots of Q(y), one negative y1 and
two positive y2 ≤ y3 . The values y = y1 , y2 then correspond to the south and north
poles of the axially squashed S 2 fibre. One can now choose the period of α so to
describe a principle S 1 -bundle over B. For this to happen the periods must be
commensurate

(11.4.9) P1 = `p, P2 = `q
with the periods Pi , i = 1, 2, given by
Z
1
(11.4.10) Pi = dA ,
2π Ci

where C1 and C2 give the standard basis for H2 (B, Z) = Z ⊕ Z. In this case, one
may take 0 ≤ α ≤ 2π` and the five-dimensional space is then the total space of an
S 1 fibration over B, with Chern numbers p and q over the two 2-cycles. An explicit
calculation shows that
3q
(11.4.11) y2 − y1 =
2p
and
q
(11.4.12) `= .
3q 2 − 2p2 + p(4p2 − 3q 2 )1/2
Moreover, as a function of the parameter (y2 − y1 )(a) is a monotonically increasing,
taking the range (0, 3/2). Thus for any p, q ∈ N such that 0 < q/p < 1 there
exists a ∈ (0, 1) for which (11.4.11) is satisfied. Furthermore, for any such p and
q which are relatively prime, the total space of the circle bundle over B is simply
connected. We denote this manifold by Mp,q . A standard argument shows that
H2 (Mp,q , Z) = Z and, hence, Mp,q must be diffeomorphic to S 2 × S 3 . ¤
Clearly, the Lie algebra isom(Mp,q , gp,q ) = su(2) ⊗ u(1) ⊗ u(1) but the isometry
group depends on the parity of p and q. For any p, q there are three commuting
11.4. SASAKI-EINSTEIN METRICS IN DIMENSIONS FIVE 395

∂ ∂ ∂
Killing vectors ∂φ , ∂ψ , and ∂γ , where α ≡ `γ so that the all three generators have
period 2π. In particular, the Sasaki-Einstein manifolds Mp,q are toric. One can
identify the toric structure of Mp,q with that of the quotient Yp,q by seeing that
their Delzant polytopes are equivalent [MS06]. Of course the metric gp,q is not
the Sasakian quotient metric unless (p, q) = (1, 0). When (p, q) 6= (1, 0), one can
easily see that the isometry group is SO(3) × U (1)2 when both p, q are odd and
U (2) × U (1) otherwise. In either case G = Isom(Yp,q , gp,q ) makes Yp,q = S 2 × S 3
into a cohomogeneity one manifold. In fact, one can see that there is a map f :
Yp,q −−→[y1 , y2 ] such that in the local chart (φ, θ, ψ, α; y) of equation (11.4.7) f is
projection onto the y-coordinate. Now for y ∈ (y1 , y2 ) we have f −1 (y) is either
SO(3) × U (1) or SU (2) × U (1) depending on the pair (p, q). These are the principal
orbits. Whereas, y = y1 and y = y2 corresponds to the singular orbits.
Let us change the local coordinates so that
(11.4.13) (α, ψ) = (−β/6 − ψ 0 /6, ψ 0 ) .
In this new chart the metric (11.4.7) can be expressed as
1−y dy 2 1
ds2 = (dθ2 + sin2 θdφ2 ) + + w(y)q(y)(dβ + cos θdφ)2
(11.4.14) 6 w(y)q(y) 36
1 1
+ [dψ 0 − cos θdφ + y(dβ + cos θdφ)]2 = ds2KE + ( dψ 0 + σ)2 ,
9 3
where now ds2KE is the transverse Kähler-Einstein metric with Kähler form

1−y 1
(11.4.15) dσ = sin θdθ ∧ dφ + dy ∧ (dβ + cos θdφ) .
3 3
Hence, ( 13 dψ 0 + σ)2 = η ⊗ η, where 3η = dψ 0 − cos θdφ + y(dβ + cos θdφ) is the dual
1-form to the Reeb vector field
µ ¶
∂ ∂ 1 ∂ ∂ 1 ∂
(11.4.16) ξ=3 0 =3 − =3 − .
∂ψ ∂ψ 6 ∂α ∂ψ 2` ∂γ
In particular, ξ = ξp,q has closed orbits if and only if ` is rational which is equivalent
to the Diophantine equation (11.4.5). If ` is irrational the generic orbits do not close
and their closure is the 2-torus generated by [∂/∂ψ, ∂/∂γ]. Hence, rank of these
metrics is equal to 2. The normalized volume N (Yp,q ) was calculated in [MS06]. ¤

As mentioned earlier, Y1,0 is just the homogeneous Kobayashi-Tanno metric on


S 2 × S 3 which is both toric and regular. The next simplest example is Y2,1 which,
as a toric contact (Sasakian) manifold, is a circle bundle over the blow up of CP2
at one point F1 = CP2 #CP2 [MS06]. As F1 cannot admit any Kähler-Einstein
metric, Kobayashi’s bundle construction cannot give a compatible Sasaki-Einstein
structure. But there is a choice of a Reeb vector field in the torus which makes
it possible to give Y2,1 a Sasaki-Einstein metric. The Sasaki-Einstein structure on
Y2,1 is not quasi-regular and this is the first such example in the literature. Hence,
S 2 × S 3 admits infinitely many toric quasi-regular Sasaki-Einstein structures and
infinitely many toric irregular Sasaki-Einstein structures of rank 2. We have the
following generalization of Theorem 11.4.5 due to [FOW06, CFO07]:
Theorem 11.4.7: Let Y (Ω) be as in Proposition 11.4.3. Then Y (Ω) admits a
toric Sasaki-Einstein structure which is unique up to a transverse biholomorphism.
396 11. SASAKI-EINSTEIN GEOMETRY

This existence of a Sasaki-Einstein metric is proved in [FOW06] although the


authors do not draw all the conclusions regarding possible toric Sasaki-Einstein man-
ifolds that can be obtained. They give one interesting example of an irregular
Sasaki-Einstein structure which generalizes the Y2,1 example of [MS05] in the fol-
lowing sense: One considers a regular positive Sasakian structure on the anticanoni-
cal circle bundle over the del Pezzo surface CP2 #2CP2 which gives a toric Sasakian
structure on 2(S 2 × S 3 ). The regular Sasakian structure on 2(S 2 × S 3 ) cannot
have any Sasaki-Einstein metric as its existence is obstructed by both the Sasaki-
Futaki invariant, Corollary 11.2.9, and by the Sasakian version of the Lichnerowicz-
Matsushima Theorem 11.3.1. However, as it is with Y2,1 Futaki, Ono and Wang
[FOW06] show that one can deform the regular structure to a unique irregu-
lar Sasaki-Einstein structure. A slightly different version of the Theorem 11.4.7 is
proved in [CFO07] where uniqueness is also established. Cho, Futaki and Ono work
with toric diagrams rather than with Kähler (Sasakian) quotients which amounts
to the same thing by Delzant’s construction. We should add that the results of
[CFO07] apply to the toric Sasaki-Einstein manifolds in general dimension and
not just in dimension 5.
Corollary 11.4.8: The manifolds k(S 2 × S 3 ) admit an infinite family of toric
Sasaki-Einstein structures for each k ≥ 1.
As in the k = 1 case one would expect infinitely many quasi-regular and in-
finitely many irregular such Sasaki-Einstein structures for each Ω satisfying all the
condition.
Remarks 11.4.1: The general anticanonical circle reduction was considered inde-
pendently in two recent papers, [CLPP05, MS05]. There it was shown that for
Ω = p = (p1 , p2 , −q1 , −q2 ), with pi , qi ∈ Z+ , p1 + p2 = q1 + q2 , and gcd(pi , qj ) = 1
for all i, j = 1, 2, the 5-manifold Y (Ω) ≈ S 2 × S 3 admits a Sasaki-Einstein structure
which coincides with that on Yp,q when p1 = p2 = p and q1 = p − q, q2 = p + q.
In [CLPP05] this family is denoted by L5 (a, b, c), where p = (a, b, −c, −a − b + c)
and they write the metric explicitly. However, in this case it appears to be harder
(though, in principle, possible) to write down the condition under which the Sasaki-
Einstein Reeb vector field ξ = ξ(a, b, c) is quasi-regular. A priori, it is not even clear
whether the quasi-regularity condition has any additional solutions beyond those
obtained for the subfamily Yp,q . Moreover, it follows from [CFO07] that the met-
rics of [CLPP05, MS05] describe all possible toric Sasaki-Einstein structures on
S2 × S3.
There have been similar constructions of a two-parameter family Xp,q of toric
Sasaki-Einstein metrics on 2(S 2 × S 3 ) [HKW05], and another two-parameter fam-
ily, called Zp,q , on 3(S 2 ×S 3 ) [OY06c]. All these examples, and many more, can be
obtained as special cases of Theorem 11.4.7 as they are all Y (Ω) for some choice of Ω.
The Yp,q , L5 (a, b, c), Xp,q and Zp,q metrics have received a lot of attention because
of the role such Sasaki-Einstein manifolds play in the AdS/CFT Duality Conjecture.
They created an avalanche of papers studying the properties of these metrics from
the physics perspective [ABCC06, OY06c, OY06b, OY06a, KSY05, HEK05,
BZ05, BBC05, BFZ05, SZ05, HKW05, BLMPZ05, BHK05, BFH+ 05,
Pal05, HSY04]. The AdS/CFT duality will be briefly discussed in the last chap-
ter of this book.
11.4. SASAKI-EINSTEIN METRICS IN DIMENSIONS FIVE 397

A recent construction of van Coevering shows that there many more examples
of quasi-regular toric Sasaki-Einstein structures on some Smale manifolds k(S 2 ×S 3 )
[vC06]. His construction is quite different from the one previously discussed. It
involves finding toric Sasakian submanifolds of the toric (in the quaternionic sense)
3-Sasakian 7-manifolds S(Ω) discussed in Section 13.7.1. Van Coevering proves
Theorem 11.4.9: k(S 2 ×S 3 ) admits infinite families of distinct toric quasi-regular
Sasaki-Einstein structures for each odd k > 1.

Proof. We only sketch the construction referring the reader to [vC06]. Con-
sider any toric QK orbifold O (see Section 12.5 for the classification and definition)
with its twistor space Z. Now, Z is a complex 3-fold with a complex contact struc-
ture and Kähler-Einstein metric. Note that Z is not toric as a complex variety.
However, it can be shown that one can always choose a hypersurface X ⊂ Z which
is a symmetric toric log del Pezzo surface. It follows from an orbifold generalization
of the result of Batyrev and Selivanova [BS99]that if Z is a symmetric toric log del
Pezzo surface, the Futaki invariant must vanish implying that Z admits an orbifold
Kähler-Einstein metric. The theorem then follows from the explicit construction
of all toric QK orbifolds S = S(Ω) as QK quotients. For any admissible matrix Ω
(see Definition 13.7.3) we get the following diagram:

(11.4.17) M 5 (Ω) / S 7 (Ω)

² ²
X 2 (Ω) / Z 3 (Ω)

²
O4 (Ω) ,

where the vertical maps are orbifold Riemannian submersions discussed in Chap-
ters 7 and 13, while the horizontal maps are Sasakian and Kählerian embeddings,
respectively. They depend on the weight matrix Ω and are explicitly obtained by
exhibiting X(Ω) as a hypersurface in the twistor space Z(Ω). However, the Kähler-
Einstein metric on the toric del Pezzo surface X(Ω) does not come from the explicit
Kähler-Einstein metrics on Z(Ω) and the Sasaki-Einstein metric on the orbibundle
M (Ω) does not come from the explicit 3-Sasakian metric on S(Ω). ¤

Example 11.4.10: We describe two elementary examples of this construction. We


begin with the case when O = S 4 with the standard round sphere metric which
is self-dual and Einstein. Its twistor space is Z = CP3 and we get the following
special case of the diagram 11.4.17
398 11. SASAKI-EINSTEIN GEOMETRY

(11.4.18) M = S2 × S3 / S7

² ²
X = CP1 × CP1 / CP3

²
S4 ,
where the second horizontal map is the standard realization of del Pezzo surface
CP1 × CP1 . Actually, the Sasakian embedding S 2 × S 3 ,→ S 7 has been studied
in [OP99] and it produces a Sasakian η-Einstein metric on S 2 × S 3 discovered
by Tanno. This metric can then be deformed to the Kobayashi-Tanno metric via
a transverse homothety. Of course, this metric is Sasaki-Einstein as mentioned
previously.
The second example is more general and better illustrates the power of the
method. Consider the family of Galicki-Lawson orbifolds O(p) for a triple of posi-
tive, pairwise prime integers p = (p1 , p2 , p3 ). The diagram 11.4.17 becomes then
(11.4.19) M (p) / S(p)

² ²
X(p) / Z(p)

²
O(p),
where the second horizontal map is described explicitly in [vC06]. The case of
p = (1, 1, 1) is special as then O(p) = CP2 with the Fubini-Study metric and all
the five spaces involved are smooth and compact Einstein manifolds. We have
(11.4.20) S(3) = 3(S 2 × S 3 ) / SU (3)/U (1)

² ²
X = CP2 #3CP2 / F2,1 = SU (3)/T 2

²
CP2 ,

where the del Pezzo surface X = CP2 #3CP2 is realized as hypersurface in a flag
manifold F2,1 . Alternatively, combining this with the twistor quotient description
of the homogeneous flag, one can describe X = CP2 #3CP2 as the usual Kähler
quotient of CP1 × CP1 × CP1 by a circle action (see, for instance, Example 7.3.11
in [Fut88]). As pointed out by Futaki the quotient metric is not Einstein. On the
other hand the existence of a unique Kähler-Einstein metric in this case was shown
by Tian and Yau [TY87] and by Siu [Siu88]. More generally, X(p) is a log del
Pezzo surface with b2 = 4 which can be considered as weighted deformation of the
smooth surface X = CP2 #3CP2 just as the twistor space Z(p) is a weighted flag
11.4. SASAKI-EINSTEIN METRICS IN DIMENSIONS FIVE 399

manifold [BGM94a, BG97b]. The 5-manifolds M (p), by Smale’s classification,


must all be diffeomorphic to 3(S 2 × S 3 ), but the toric structure and the Seifert
fibration M (p)−→X(p) depend on p and we get an infinite family of toric quasi-
regular Sasaki-Einstein structures on 3(S 2 × S 3 ). Taking an admissible Ω ∈ Ak,k+2
gives infinite families of such structures on (2k + 1)(S 2 × S 3 ).
It should be noted that van Coevering method fails to produce any toric Sasaki-
Einstein metrics on 2k(S 2 × S 3 ) which are also toric as Sasakian and contact man-
ifolds and they admit toric Sasaki-Einstein metrics for each k ≥ 1.
Open Problem 11.4.1: The Sasaki-Einstein manifolds Yp,q are of cohomogeneity
one. A very natural problem is to classify all Sasaki-Einstein 5-manifolds of cohomo-
geneity one. Any such space, if simply connected, must be diffeomorphic to either
S 2 × S 3 or S 5 . It is likely that all cohomogeneity one Sasaki-Einstein structures
on S 2 × S 3 are indeed give by the manifolds Yp,q . However, the question regarding
existence of cohomogeneity one Sasaki-Einstein metrics on S 5 is more subtle. Con-
sider L(2, 2, 2, k), for example. For any k > 1 odd this gives a cohomogeneity one
Sasakian structure on S 5 . However, for all but L(2, 2, 2, 3), Sasaki-Einstein struc-
tures are obstructed by Proposition 11.3.16. The case of L(2, 2, 2, 3) was studied in
[GMSY06] but the existence problem remains open.
Exercise 11.5: Compute the group diagram H ⊂ {K− , K+ } ⊂ G for each of the
manifolds Yp,q .
Exercise 11.6: Let k be odd. The isometry group of the Sasakian structure on the
link L(2, 2, 2, k) ' S 5 is SO(3) × U (1). Give the cohomogeneity one group diagram
for this action on S 5 . The transverse space X2k ⊂ CP(2, k, k, k) of this structure
is a cohomogeneity one orbifold Fano surface under the action of SO(3). Compute
the cohomogeneity one diagram of the SO(3) action on X2k and compare it with
the diagrams of the Exercise 1.6.34.

11.4.3. Sasaki-Einstein Structures on General Smale Manifolds. We


consider the problem of determining which Smale manifolds with non-vanishing
torsion in H2 (M, Z) admit Sasaki-Einstein metrics. Clearly, a necessary condition
is that M must admit an anticanonical Sasakian Seifert structure so that all the
restrictions of Theorems 10.2.19 and 10.2.25 apply. We begin by considering the
case of Sasaki-Einstein metrics on rational homology spheres. The first examples of
Sasaki-Einstein metrics occurring on simply connected rational homology spheres
appeared in [BG06a] (the preprint version of this paper appeared in November,
2003), but the definitive work is that of Kollár [Kol05, Kol06b]. It is quite
remarkable that almost all rational homology 5-spheres with a positive Sasakian
structure are known to admit a Sasaki-Einstein structure.
Theorem 11.4.11: Let M ∈ (SB+ )0 . When H2 (M, Z) = (Z5 )4 , (Z4 )2 , (Z3 )8 or
(Zm )2 , m 6= 30k, m ≥ 12 the moduli space of quasi-regular Sasaki-Einstein struc-
tures on M is non-empty and can be completely described in each case. In addition,
there are examples of Sasaki-Einstein structures on other positive rational homology
spheres of Theorem 10.2.19. More precisely, we have
(i) If M = Mk , k ≥ 12 for some k not divisible by 30 then M admits between
0 and 4 positive Sasakian Seifert structures, depending on k mod 30
which are given by the following links
• L(3, 3, 3, k), (k, 3, ) = 1,
400 11. SASAKI-EINSTEIN GEOMETRY

• L(2, 3, 6, k), (k, 6) = 1,


• L(2, 4, 4, k), (k, 2, ) = 1,
• L∗ (k, 2k, 3k, 5; 6k), (k, 5) = 1.
Each such Seifert structure yields a unique 2-parameter family of Sasaki-
Einstein metrics naturally parameterized by the moduli space of genus 1
curves.
(ii) 2M5 has a unique positive Sasakian Seifert structure given by the BP link
L(2, 5, 6, 6) which yields a 6-dimensional family of Sasaki-Einstein struc-
tures naturally parameterized by the moduli of genus 2 curves.
(iii) 4M3 has a unique positive Sasakian Seifert structure given by the BP link
L(2, 3, 10, 10) which yields a 14-dimensional family of Sasaki-Einstein
structures naturally parameterized by the moduli of hyperelliptic genus 4
curves.
(iv) 2M4 has exactly 2 distinct positive Sasakian Seifert structures given by
the links L∗ (4, 8, 9, 20; 40) and L∗ (4, 4, 5, 8; 20). In both cases the family
of Sasaki-Einstein metrics is naturally parameterized by the moduli space
of pairs (C, p), where π : C−→CP1 is a genus 2 curves and p ∈ C is not
a branch point of π.
(v) The manifolds 2M3 , 3M3 , and Mk , with 1 < k < 12 all admit Sasaki-
Einstein structures.

Proof. To prove (i) in the first three cases it is enough to apply Theorem
5.4.11. This was already done in [BG06a, BG05]. In particular, also M7 , M8 , M10 ,
and M11 admit Sasaki-Einstein structure which proves part of (v). In [Kol05]
Kollár improves it slightly to show that M5 also admits a Sasaki-Einstein struc-
ture. In the case of L∗ (k, 2k, 3k, 5; 6k), (k, 5) = 1, Kollár shows that it admits a
Sasaki-Einstein metric for k > 3. That gives the fourth series in (i) and M4 , M6 , M9
in (v).
Existence in part (ii) follows from the link representation and Theorem 5.4.11.
Kollár shows in [Kol05] the completeness of the construction.
Existence in part (iii) is proved by Kollár in [Kol05]. Let Fn ⊂ CPn+1 be the
cone over the degree n rational normal curve and C ⊂ Fn a smooth intersection
of Fn with a quadric. Kollár shows that one gets a Kähler-Einstein metric in
the following cases: (Z, ∆) = (F3 , 45 C), (F3 , 34 C), (F4 , 23 C) and (F5 , 23 C). Now,
M −→(F5 , 32 C) can be easily identified with L(2, 3, 10, 10) and the uniqueness and
completeness of the construction is established in [Kol06b].
Existence in part (iv) is first proved by Kollár in [Kol05] for L∗ (4, 4, 5, 8; 20)
by realizing that this link is a Seifert fibration over (Q, 43 C), where Q is the quadric
cone and a smooth degree 5 curve in Q. The existence of a Sasaki-Einstein metric
on L∗ (4, 4, 5, 8; 20) as well as the uniqueness and completeness of the construction
is argued in [Kol06b].
The existence for 2M3 was established in [Kol05] by the following example:
Let Sn be the Hirzebruch surface with a negative section E ⊂ Sn with E 2 = −n
and fiber F. 2M3 can be obtained as a Seifert bundle over (S1 , 32 D + 21 E), where
D ∈ |2E +4F |. In the 3M3 case, the existence was established in [BG06a] by taking
the BP link L(3, 4, 4, 4) and using Theorem 5.4.11. Alternatively Kollár shows that
3M3 is a Seifert bundle over (F4 , 23 C).
11.4. SASAKI-EINSTEIN METRICS IN DIMENSIONS FIVE 401

To finish (v) for M3 take the link L∗ (3, 3, 4, 6; 12). The estimates in [BG06a]
show that this has Sasaki-Einstein structure. For M2 see the remark in [Kol05],
Proof 9.6.
¤

Hence, with the exception of nM2 , n ≥ 2, each rational homology sphere in


(SB+ )0 is known to admit a Sasaki-Einstein structure.
Open Problem 11.4.2: The methods used by Kollár in [Kol05] should make it
possible to produce Sasaki-Einstein metrics on nM2 for small values of n > 1. The
question whether Sasaki-Einstein metrics can be found for any n is probably much
harder.
In the case of Theorem 10.2.25 when both b2 (M ) > 0 and the H2 (M, Z)tor are
non-trivial we have the following existence result
Theorem 11.4.12: Let M ∈ (SB+ )0 and k = b2 (M ) > 0.
(i) If M −→(Z, ∆) is one of the Seifert fibrations listed in the table of The-
orem 10.2.25 (i) then M admits at least a two-dimensional family of
compatible Sasaki-Einstein structures.
(ii) M∞ #2M4 has only one positive Seifert fibration giving rise to a fam-
ily of Sasaki-Einstein structures. This structure is realized by the link
L∗ (4, 4, 5, 12; 24). The family is naturally parameterized by the moduli
space of pairs (C, p), where π : C−→CP1 is a genus 2 curve and p ∈ C is
not a branch point of π.
(iii) M∞ #nM2 admits a Sasaki-Einstein structure for all n ≥ 1.

Proof. For the BP links of the Theorem 10.2.25 (i) one can use simple esti-
mates of Theorem 5.4.11 and this partial result was obtained in [BG05]. This gives
the Sasaki-Einstein metrics only in 6 out of the 15 cases. In all the cases the exis-
tence is established in [Kol05]. Observe, however, that some BP links give Sasaki-
Einstein structures also for m < 12. For example, L(3, 3, 3, 3m) and L(2, 4, 4, 4m)
give Sasaki-Einstein structures on 6M∞ #Mm and 7M∞ #Mm for any m > 2. The
link L(2, 3, 6, 6m) gives a Sasaki-Einstein structure on 8M∞ #Mm for any m > 4,
while the link L(2, 4, 4, 2m) gives a Sasaki-Einstein structure on 3M∞ #Mm , for
any odd m > 5 (see Example 74 of [BG05]).
For part (ii) one easily sees that the link is realized by the polynomial f6 (z0 , z1 )+
f1 (z0 , z1 )z24 + z32 , where fi is a polynomial of degree i and gives M∞ #2M4 . The
remainder of the proof can be found in [Kol06b].
Part (iii) is proved in [Kol05], where Kollár gives the following example: Let
Sn be the minimal ruled surface with a negative section E ⊂ Sn with E 2 = −n and
fiber F. Take a smooth curve C ⊂ |2E + (2n + 3)F | which is transverse to E. Then
g(C) = n + 2 and (Sn , (1 − 21 )C + (1 − m 1
)E) is a log Del Pezzo surface for m > 2n.
These give examples with H2 (M, Z) = Z ⊕ (Z/2)2n for any n ≥ 2. ¤

Corollary 11.4.13: The table below summarizes existence of Sasaki-Einstein met-


rics on Smale manifolds. The first column lists all cases when a Sasaki-Einstein met-
ric might exist. That is, any Smale manifold not listed there does not admit any
Sasaki-Einstein structure. The second column lists the cases when the existence is
known.
402 11. SASAKI-EINSTEIN GEOMETRY

M ∈ (B+ )0 S-E
kM∞ , k ≥ 0 any k
8M∞ #Mm , m > 2 m>4
7M∞ #Mm , m > 2 m>2
6M∞ #Mm , m > 2 m>2
5M∞ #Mm , m > 2 m > 11
4M∞ #Mm , m > 2 m>4
3M∞ #Mm , m > 2 m = 7, 9 or m > 10
2M∞ #Mm , m > 2 m > 11
M∞ #Mm , m > 2 m > 11
Mm , m > 2 m>2
2M5 , 2M4 , 4M3 , M∞ #2M4 yes
kM∞ #2M3 k=0
kM∞ #3M3 , k ≥ 0 k=0
kM∞ #nM2 , k ≥ 0, n > 0 (k, n) = (0, 1) or (1, n), n > 0
kM∞ #Mm , k > 8, 2 < m < 12

Question 11.4.1: The case of H2 (M, Z)tor = (Z5 )4 , (Z4 )2 , (Z3 )8 and (Zm )2 , m ≥
12 is completely understood in terms of the existence and the moduli of quasi-
regular Sasaki-Einstein structures. Let us therefore summarize some outstand-
ing questions/conjectures which describe what we do not yet know about Sasaki-
Einstein geometry of Smale manifolds. Let M ∈ B 0 .

(i) Suppose k = b2 (M ) > 9. We conjecture that M admits a Sasaki-Einstein


structure if and only if H2 (M, Z)tor = 0, i.e., M is diffeomorphic to
k(S 2 × S 3 ) where existence is known.
(ii) Suppose M admits a Sasaki-Einstein structure with dim(aut(ξ, η, Φ)) > 1.
We conjecture that in such a case H2 (M, Z)tor = 0. Recall the Corollary
11.4.8 stating that k(S 2 × S 3 ) admits toric Sasaki-Einstein structures for
any k ≥ 0.
(iii) We conjecture that kM∞ #Mm admit Sasaki-Einstein structures for all
0 ≤ k ≤ 8 and for all m > 2 giving 31 missing cases of Corollary 11.4.13.
(iv) What can one say about Sasaki-Einstein structures on kM∞ #2M3 and
kM∞ #3M3 ? Is k > 0 possible and, if so, which are the possible values it
takes?
(v) What can one say about Sasaki-Einstein structures on kM∞ #mM2 and
kM∞ #mM2 ? Is k 6= 1 possible and, if so, which are the possible values
it takes?
Remark 11.4.2: Some Smale-Barden manifolds admit other Einstein metrics. This
is known in four cases: X−1 = SU (3)/SO(3), S 5 , X∞ and M∞ . X−1 is a symmetric
space and the metric is Einstein. But X−1 is not even almost contact. S 2 × S 3 is
well-known to have infinitely many inequivalent homogeneous Einstein metrics dis-
covered by Wang and Ziller [WZ86, WZ90]. S 5 and S 2 × S 3 have infinitely many
inequivalent Einstein metrics of cohomogeneity one discovered by Böhm [Böh98].
Finally, we mention that both M∞ = S 2 × S 3 and as well as X∞ , the non-trivial
S 3 -bundle over S 2 , admit infinitely many Einstein but non Sasaki-Einstein metrics
that were recently constructed by the physicists [GHY04, HSY05, LPP04].
11.5. SASAKI-EINSTEIN METRICS ON HOMOTOPY SPHERES 403

11.5. Sasaki-Einstein Metrics on Homotopy Spheres


Let us recall that until very recently Einstein metrics on spheres appeared
to be very rare. Any sphere S n admits a metric of constant sectional curvature,
known as the standard metric (actually a homothety class of metrics) on S n . These
metrics are homogeneous and Einstein, that is the Ricci curvature is a constant
multiple of the metric. The spheres S 4m+3 , m > 1 are known to have another
Sp(m + 1)-homogeneous Einstein metric discovered by Jensen [Jen73]. In addi-
tion, S 15 has a third Spin(9)-invariant homogeneous Einstein metric discovered by
Bourguignon and Karcher [BK78]. In 1982 Ziller proved that these are the only
homogeneous Einstein metrics on spheres [Zil82]. No other Einstein metrics on
spheres were known until 1998 when Böhm constructed infinite sequences of non-
isometric Einstein metrics, of positive scalar curvature, on S 5 , S 6 , S 7 , S 8 , and S 9
[Böh98]. Böhm’s metrics are of cohomogeneity one and they are not only the first
inhomogeneous Einstein metrics on spheres but also the first non-standard Einstein
metrics on even-dimensional spheres. Then in September of 2003 the authors in
joint work with János Kollár (Evan Thomas helped us with the computer programs)
[BGK05, BGKT05] announced the existence of a plethora of Einstein metrics on
spheres in all odd dimensions greater than three, including exotic sphere which
bound parallelizable manifolds. A short review and history of our work appears in
[BG06d]. This section consists of a synopsis of our work. The main result is
Theorem 11.5.1: Let L(a) = L(a0 , . . . , an ) be a Brieskorn-Pham link.
(i) L(a) has a Sasaki-Einstein metric if
Xn
1 n n1 1 o
1< <1+ min , ,
a
i=0 i
n − 1 i,j ai bi bj

where the bi ≤ ai are defined before Theorem 5.4.11. L(a) is homeo-


morphic to S 2n−1 if and only if the conditions of Theorem (9.3.18) are
satisfied. The diffeomorphism type can be determined as in Section 9.3.2.
(ii) If a = (a0 , . . . , an ) are pairwise relatively prime. Then L(a) is home-
omorphic to S 2n−1 and it admits a Sasaki-Einstein metric if and only
if
Xn
1 n1o
1< < 1 + n min ,
a
i=0 i
i ai
(iii) Given two sequences a and a0 satisfying condition (i), the manifolds L(a)
and L(a0 ) are isometric if and only if a is a permutation of a0 .

Proof. Parts (i) follows from Theorems 11.1.3 and 5.4.11. The existence in
part (ii) follows from the improved Ghigi-Kollár estimate of Theorem 5.4.12. The
fact that this estimate is sharp3 follows from the Lichnerowicz obstruction of The-
orem 11.3.14 applied to this specific BP link. Part (iii) will follow by Proposition
11.1.16 and Theorem 13.3.1 below if one can show that the corresponding Kähler-
Einstein orbifolds Xd of Theorem 5.4.11 admit no holomorphic contact structure.
This is guaranteed by the following lemma whose proof is essentially found in
[BGK05]

3We thank János Kollár for pointing this out to us.


404 11. SASAKI-EINSTEIN GEOMETRY

Lemma 11.5.2: Let L(a) be a BP link and suppose the following estimate holds:
Xn
1 n n1o
< 1 + min .
a
i=0 i
2 i ai
Then the induced Sasakian Seifert structure on L(a) is not part of a 3-Sasakian
structure.
¤
Remark 11.5.1: It is interesting that the estimate of Lemma 11.5.2 lies between
the two estimates in Theorem 11.5.1. This allows for the intriguing possibility that
3-Sasakian structures could reside on some links of type (ii). We also remark that
for any link L(a) to admit a 3-Sasakian structure n must be even. Thus, (iii) of
Theorem 11.5.1 holds for the sequences of case (ii) if n is odd, or if n is even and
a satisfies the estimate of Lemma 11.5.2.
Using the theory described in Section 9.3.2 one can use Theorem 11.5.1 to
enumerate all m-tuples a in every dimension which give homotopy spheres and
then, depending on m determine the diffeomorphism type of the sphere L(a). This
was done in [BGK05, BGKT05] for case (i) of the theorem. Case (ii) of the
theorem provides additional examples [GK05].
Corollary 11.5.3: On S 5 we obtain 80 families of Sasaki-Einstein metrics that
are inequivalent as Einstein structures. Some of these admit non-trivial continuous
Sasaki-Einstein deformations, the largest of which depends on 10 real parameters.
Proof. We get 68 examples of BP links which satisfy the first inequality of the
Theorem 11.5.1. They are all positive BP links of Table B.4.3. Possible moduli can
be determined using the results of Sections 5.5 and 11.1.2. The largest family comes
from Example 5.5.8 which has (real) dimension 10. Out of the 63 sporadic spheres
in Table B.4.3 all but L(2, 3, 7, 7) and L(2, 3, 11, 11) can be shown to have a Sasaki-
Einstein structure using Theorem 11.5.1 (i). The remaining 7 examples are the
BP links L(2, 3, 4, 5) and L(2, 3, 5, k), k = 7, 8, 9, 11, 13, 14. The second inequality
gives an additional 12 examples. These are the BP links L(2, 3, 5, k) which admit
Sasaki-Einstein structure for k = 17, 19, 23, 29, 31, 37, 41, 43, 47, 49, 53, 59. None of
these 12, however, admit moduli. ¤
In the next odd dimension the situation becomes much more interesting. A
computer search [BGKT05] found 8, 610 distinct families of Sasaki-Einstein struc-
tures on the standard S 7 = Σ28 and exotic 7-spheres Σi which satisfy the inequality
(i) of Theorem 11.5.1. By the results of Kervaire and Milnor [KM63] as discussed
in Section 9.4, there are 28 oriented diffeomorphism types of topological 7-spheres.
(15 types if we ignore orientation.) Then Theorem 9.4.6 together with Brieskorn’s
signature formula 9.4.2 allows one to decide which L(a) corresponds to which ho-
motopy sphere by computing the Hirzebruch signature of the 8-manifold Va8 . This
computation was performed by a computer in [BGKT05] for the 8, 610 cases show-
ing that each oriented diffeomorphism type contains several hundred distinct fami-
lies of Sasaki-Einstein metrics. There, the number of distinct families ranged from
230 on Σ25 to 452 on Σ22 . However, the results of [GK05] increase these 8, 610
cases by more than 1000, so these numbers increase somewhat. We have
Corollary 11.5.4: Each of the 28 oriented diffeomorphism classes on S 7 admit
several hundred inequivalent families of Sasaki-Einstein structures, some of them
11.5. SASAKI-EINSTEIN METRICS ON HOMOTOPY SPHERES 405

(in each diffeomorphism class) depending on moduli. The largest of these families
depends on 82 real parameters and occurs on the standard sphere.

Proof. One can shows that already a = (2, 3, 7, 41, p), 43 ≤ p ≤ 256 real-
izes all 28 diffeomorphism types. This is an easy signature computation. In fact,
there are finitely many 5-tuples which are homology 7-spheres and satisfy either
of the inequalities of Theorem 11.5.1 giving hundreds of examples in each diffeo-
morphism class. The example with the biggest moduli is given by the sequence
a = (2, 3, 7, 43, 43 · 31) which is the standard 7-sphere with a 2(43 − 2) = 82-
dimensional family of Sasaki-Einstein metrics. The Sasaki-Einstein metrics arising
from case (ii) of Theorem 11.5.1 do not contribute moduli. ¤

Remark 11.5.2: Let us mention here that any orientation reversing diffeomorphism
takes a Sasaki-Einstein metric into an Einstein metric, but not necessarily a Sasaki-
Einstein metric, since the Sasakian structure fixes the orientation.
Since Milnor’s discovery of exotic spheres [Mil56b] the study of special Rie-
mannian metrics on them has always attracted a lot of attention. Perhaps the
most intriguing question is whether exotic spheres admit metrics of positive sec-
tional curvature. This problem remains open. In 1974 Gromoll and Meyer wrote
down a metric of non-negative sectional curvature on one of the Milnor spheres
[GM74]. More recently it has been observed by Grove and Ziller that all exotic
7-spheres which are S 3 bundles over S 4 admit cohomogeneity one metrics of non-
negative sectional curvature [GZ00]. But it is not known if any of these metrics
can be deformed to a metric of strictly positive curvature.
Lastly, we should add that although heretofore it was unknown whether Ein-
stein metrics existed on exotic spheres, Einstein metrics have been known to exist on
manifolds which are homeomorphic but not diffeomorphic. In dimension 7 there are
even examples of homogeneous Einstein metrics with this property [KS88]. Kreck
and Stolz find that there are 7-dimensional manifolds with the maximal number of
28 smooth structures, each of which admits an Einstein metric with positive scalar
curvature. Corollary 11.5.4 establishes the same result for 7-spheres. In dimension
4n + 1 we easily get
Theorem 11.5.5: For n ≥ 2 both the standard and Kervaire spheres of dimen-
sion (4n + 1) admit continuous families of inequivalent Sasaki-Einstein metrics.
The number of inequivalent families as well as the number of moduli grows doubly
exponentially with dimension.

Proof. Let {ci } be the Sylvester sequence of equation (5.4.3), consider se-
quences of the form
a = (a0 = 2c0 , . . . , am−2 = 2cm−2 , am−1 = 2, am ) ,
where am is relatively prime to all the other ai s. By an easy computation, condition
(i) of Theorem 11.5.1 is satisfied if 2cm−2 < am < 2cm−1 − 2. The relatively prime
condition is harder to pin down, but it certainly holds if in addition am is a prime
number. By the prime number theorem, the number of primes in the interval
[cm−1 , 2cm−1 ] is about
m−2
cm−1 (1.264)2 m−1
≥ m−1 ≥ (1.264)2 −4(m−1)
,
log cm−1 2 log 1.264
406 11. SASAKI-EINSTEIN GEOMETRY

so it is still doubly exponential in m. By Theorem 9.4.3, for even m, L(a) the


standard sphere if am ≡ ±1 mod 8 and the Kervaire sphere if am ≡ ±3 mod 8.
It is easy to check for all values of m that we get at least one solution of both
types. ¤

A partial computer search yielded more than 3 · 106 cases for S 9 and more
than 109 cases for S 13 , including a 21300113901610-dimensional family, see Ex-
ample 11.5.10. The only Einstein metric on S 13 known before was the standard
one.
In the remaining case of n ≡ 3 mod 4 the situation is more complicated. Recall
that for these values of n the group bPn+1 is quite large and we do not know how to
show that every member of it admits a Sasaki-Einstein structure, since our methods
do not apply to the examples given in [Bri66]. We believe, however, that this is
true:
Conjecture 11.5.6: All odd-dimensional homotopy spheres which bound paralleliz-
able manifolds admit Sasaki-Einstein metrics.
Theorem 11.5.7: The above conjecture is true in odd dimensions 1(4) and it is
true in dimension 7, 11, and 15.

Proof. In all dimensions 1(4) it follows from Proposition 11.5.5. In dimension


7 it is the Corollary 11.5.4. In dimensions 11 and 15 the proof follows a computer
search of links satisfying the estimates of Theorem 11.5.1(i). This was described in
[BGKT05]. ¤

Although we do not know the distribution of solutions between different dif-


feomorphism types, just as in the case of dimensions 1(4) in the remaining odd
dimensions Theorem 11.5.1(i) likewise yields doubly exponential growth.
Example 11.5.8: Consider sequences of L(a) of the form
a = (a1 = c1 , . . . , am−1 = cm−1 , am ) .
The troublesome part of the inequalities of Theorem 11.5.1(i) is the computation
of the bi . However bi ≤ ai thus it is sufficient to satisfy the following stronger
restriction:
Xm
1 m−1 n 1 o m−1 1
1< <1+ min =1+ · .
i=1
ai m − 2 i,j ai a j m − 2 am−1 am

By direct computation this is satisfied if cm − cm−1 < am < cm . At least a third of


these numbers are relatively prime to a1 = 2 and to a2 = 3, thus we conclude:
Proposition 11.5.9: The Theorem 11.5.1(i) yields at least
1 1 m−1
(cm−1 − 1) ≥ (1.264)2 − 0.5
3 3
inequivalent families of Sasaki-Einstein metrics on (standard and exotic) (2m − 3)-
spheres.
If 2m − 3 ≡ 1 mod 4 then by Theorem 9.4.3, all these metrics are on the
standard sphere. If 2m − 3 ≡ 3 mod 4 then all these metrics are on both standard
and exotic spheres but we cannot say anything in general about their distribution.
Finally, let us consider examples with the largest moduli.
11.6. THE SASAKI-EINSTEIN SEMI-GROUP 407

Example 11.5.10: We consider the sequences


a = (a1 = c1 , . . . , am−1 = cm−1 , am = (cm−1 − 2)cm−1 ) .
Any two of them are relatively prime, except for gcd(am−1 , am ) = cm−1 , and the
inequalities of Theorem 5.4.11 are satisfied. The Brieskorn-Pham polynomial has
weighted homogeneous perturbations of the form
a
xa1 1 + · · · + xm−2
m−2
+ G(xm−1 , xcmm−1 −2 ) ,
where G is any homogeneous polynomial of degree cm−1 . Up to coordinate changes,
these form a family of complex dimension cm−1 − 2. This example proves
m−1
Proposition 11.5.11: Our methods yield an at least 2(cm−1 −2) ≥ 2((1.264)2 −
2.5)-dimensional (real) family of pairwise inequivalent Sasaki-Einstein metrics on
some (standard or exotic) (2m − 3)-sphere.
As before, if 2m − 3 ≡ 1(4) then these metrics are on the standard sphere.

11.6. The Sasaki-Einstein semi-group


In this section we apply the join construction described in Section 7.6.2 to
the set of quasi-regular Sasaki-Einstein orbifolds. It is based on the toral bundle
construction of Wang and Ziller [WZ86, WZ90] for Einstein manifolds. Here we
see that the Einstein condition pins down the pair of integers (k1 , k2 ) of equation
(7.6.2).
Definition 11.6.1: We denote by SE the subset of SO with π1orb = 0 and whose
Sasakian structures are also Einstein, by SE s the subset of SE that are smooth
manifolds, and by RE ⊂ SE s the subset of regular Sasaki-Einstein manifolds. All
these subsets are given the subspace topology.
The condition π1orb = {1} is made to avoid complications. Now since every
element S ∈ SE is an S 1 -orbibundle over a compact Fano orbifold, Definition
4.4.24 of Fano index can be applied to SE. We have
Definition 11.6.2: The (Fano) index of an element S ∈ SE, denoted by Ind(S),
is defined to be the Fano index of the orbifold Z = S/S 1 . Furthermore, given two
elements S1 , S2 ∈ SE, the relative indices li of the pair (S1 , S2 ) by
Ind(Si )
li =
gcd(Ind(S1 ), Ind(S2 ))
for i = 1, 2.
The natural grading of SO given by equation (7.6.1) is clearly inherited by the
subset SE, that is,

M
SE = SE2n+1 ,
n=1
and similarly for SE s and RE. For the case n = 1 it follows from a theorem of
Hamilton [Bes87] that SE3s = RE3 = {pt}, namely S 3 with its standard Sasaki-
Einstein structure.
We now want to restrict our join map ?k1 ,k2 of equation (7.6.2) to SE2n1 +1 ×
SE2n2 +1 and ask when can we choose the pair (k1 , k2 ) of relatively prime integers
such that the image of ?k1 ,k2 lies in SE2(n1 +n2 )+1 ? Given S1 , S2 ∈ SE of dimension
2n1 + 1 and 2n2 + 1 respectively, we know that their respective spaces of leaves Z1
and Z2 are Kähler-Einstein Fano orbifolds of complex dimension n1 , n2 , respectively.
408 11. SASAKI-EINSTEIN GEOMETRY

The index of Z1 × Z2 is just gcd(Ind(Z1 ), Ind(Z2 )). Thus, by Theorem 11.1.12 the
S 1 -orbibundle on Z1 × Z2 whose first Chern class is
c1 (Z1 × Z2 ) c1 (Z1 ) + c1 (Z2 )
(11.6.1) =
Ind(Z1 × Z2 ) gcd(Ind(Z1 ), Ind(Z2 ))

is simply connected in the orbifold sense. So the appropriate S 1 -orbibundle over


Z1 × Z2 is the one whose orbifold first Chern class is given by equation (11.6.1).
2
Thus, if ωi denotes a Kähler form on Zi whose class [ωi ] ∈ Horb (Z, Z) is indivisible,
we see that

c1 (Z1 ×Z2 ) = c1 (Z1 )+c1 (Z1 ) = Ind(Z1 )[ω1 ]+Ind(Z2 )[ω2 ] = Ind(Z1 ×Z2 )[l1 ω1 +l2 ω2 ] .

Hence, we choose the pair (k1 , k2 ) to be the pair of relative indices (l1 , l2 ). In terms
of the Ricci form this translates to

ρl1 ω1 +l2 ω2 = 2πInd(Z1 × Z2 )(l1 ω1 + l2 ω2 ) ,

so l1 ω1 + l2 ω2 is Kähler-Einstein. We now only need to rescale the Kähler-Einstein


metric to obtain one whose scalar curvature is 4(n1 + n2 )(n1 + n2 + 1). Then by
Theorem 11.1.12 we get a Sasaki-Einstein metric on the total space S1 ?l1 ,l2 S2 of
the bundle whose orbifold first Chern class is [l1 ω1 + l2 ω2 ]. We have constructed
a multiplication on SE, defined by ? = ?l1 ,l2 , and refer to it as the join. It is
clear from the construction that if S1 , S2 ∈ SE then also S1 ? S2 ∈ SE with the
Sasaki-Einstein structure as defined above, and that this procedure can be iterated
in such a way that ? is associative. It is also commutative up to isomorphism.
Furthermore, ? is continuous in both factors. We have arrived at:
Theorem 11.6.3: The join operation ? = ?l1 ,l2 gives SE the structure of a com-
mutative associative topological semi-group.
Actually we are interested in the subset SE s of smooth Sasaki-Einstein mani-
folds. However, as in Proposition 7.6.6 it is not true that SE s is a subsemi-group.
From the discussion above and Proposition 7.6.6 it is easy to see that
Proposition 11.6.4: Let Si ∈ SE s with orders υi , and relative indices li , respec-
tively. Then
(i)Ind(S1 ? S2 ) = gcd(Ind(S1 ), Ind(S2 )).
(ii)S1 ? S2 ∈ SE s if and only if gcd(υ1 l2 , υ2 l1 ) = 1.
(iii)RE is a subsemi-group.
(iv) If S1 is regular (i.e., υ1 = 1) and Ind(S2 ) divides Ind(S1 ), then S1 ? S2 ∈
SE s independently of υ2 .
Before further discussion, we adapt Definition 7.6.11 to the current situation.
Definition 11.6.5: We say that S ∈ SE is SE-reducible if it can be written as
S = S1 ? S2 for some S1 , S2 ∈ SE. S is SE-irreducible if it is not SE-reducible.
Since a Sasaki-Einstein structure determines a unique Sasaki-Seifert structure
we have
Proposition 11.6.6: An element S ∈ SE is SE-reducible if and only if it is S-
reducible.
Using this it is easy to see that in terms of Sasaki-Einstein geometry Theorem
8.3.9 becomes
11.7. SASAKI-EINSTEIN METRICS IN DIMENSIONS SEVEN AND HIGHER 409

Theorem 11.6.7: The subset HE of simply connected homogeneous Sasaki-Einstein


manifolds forms a subsemi-group of RE. Furthermore, a simply connected Sasaki-
Einstein homogeneous manifold corresponding to a semi-simple Lie group G is SE-
irreducible if and only if G is simple.
It is clear that up to diffeomorphism there is precisely one regular SE-reducible
simply connected 5-manifold, namely the homogeneous Stiefel manifold V2 (R4 ) ≈
S 2 × S 3 ≈ S 3 ? S 3 with the Kobayashi-Tanno metric. When M is not simply
connected any reducible Sasaki-Einstein manifold would have to be a quotient of
S 2 × S 3 with the Kobayashi-Tanno metric. Proposition 11.6.4 will allow us to
construct smooth examples of quasi-regular Sasaki-Einstein manifolds in all odd
dimensions greater than or equal to 7. The next section is devoted to describing
such examples.

11.7. Sasaki-Einstein Metrics in Dimensions Seven and Higher


There are several different methods for proving the existence of Sasaki-Einstein
metrics in dimension 7 and higher. First we mention that in the homogeneous case
Theorem 11.1.13 gives us a complete classification. Next if one knows the existence
of a Kähler-Einstein metric on Fano manifold, one can construct the appropriate
circle bundle. But this is more or less hit or miss in higher dimensions, and not
much is known generally about the existence of Kähler-Einstein metric on Fano
manifolds of high dimension. Nevertheless, the physicists [GMSW04a] have been
able to construct explicit examples in some cases. Another method involving so-
called 3-Sasakian manifolds will be discussed separately Chapter 13.
The two somewhat systematic methods which have enjoyed some success in
producing Sasaki-Einstein metrics in high dimension are: (1) the join method de-
scribed in the previous section, and (2) links of isolated hypersurface singularities.
As far as (2) is concerned we have already dedicated an entire section to proving
the existence of Sasaki-Einstein metrics on homotopy spheres in odd dimensions.
These two methods are quite complimentary in the sense that (2) only gives ex-
amples on highly connected ((n − 1)-connected (2n + 1)-dimensional) manifolds;
whereas, method (1) always generates 2-dimensional cohomology classes.
Our join construction or more explicitly Theorem 11.6.3 allows us to construct
new Sasaki-Einstein structures from old ones. The price that we pay is that we
always increase dimension. For example if M is an odd-dimensional manifold then
Theorem 11.6.3 allows to propagate any Sasaki-Einstein metric on M to the next
odd dimension by taking a join of M with S 3 with its standard round Sasakian
metric. Of course not every such join will be a smooth manifold but it will nev-
ertheless have an orbifold Sasaki-Einstein structure. In many cases, however, the
join produces smooth examples as indicated by Proposition 11.6.4 and as we shall
soon see. We begin the discussion with the regular case.
The classification of regular Sasaki-Einstein 5-manifolds in Section 11.4 imme-
diately gives a classification of the regular SE-reducible 7-manifolds.
Proposition 11.7.1: Any simply connected regular SE-reducible 7-manifold is one
of the following:
S 3 ? S 3 ? S 3 , S 3 ? S 5 , S 3 ? Sk
for 3 ≤ k ≤ 8, where Sk is one of the Sasaki-Einstein circle bundles over the del
Pezzo surface discussed in Theorem 11.4.1.
410 11. SASAKI-EINSTEIN GEOMETRY

These examples have already been noted in [BFGK91]. Non simply connected
examples are obtained by dividing by a cyclic subgroup of the circle generated by
the characteristic vector field. Notice also that for 5 ≤ k ≤ 8 the 7-manifolds S 3 ?Sk
have continuous families of Sasaki-Einstein structures on them. We shall discuss
these shortly, but first we mention that in the list given in Proposition 11.7.1 only
the first two are homogeneous as seen in (iii) of Corollary 11.1.14.
The above theorem brings into the picture 3 additional Sasaki-Einstein spaces
which are SE-irreducible. One more such example occurs as a toric Sasaki-Einstein
manifold [Mab87]
Proposition 11.7.2: The simply connected regular toric Sasaki-Einstein 7-manifolds
are precisely one of the following:
(i) the reducible Sasaki-Einstein manifolds S 3 ? S 3 ? S 3 , S 3 ? S 5 or S 3 ? S3 ,
(ii) the Kobayashi bundle over the Fano 3-fold P(OCP1 ×CP1 ⊕OCP1 ×CP1 (1, −1)).
Other regular SE-irreducible 7-manifolds can be constructed as links (Fermat
hypersurfaces, cf. Example 11.7.15 below) and we will consider them in the fol-
lowing subsection. So far a complete classification of regular Sasaki-Einstein 7-
manifolds is lacking. Although the classification of Fano 3-folds is complete, it is
still not known precisely which ones admit Kähler-Einstein metrics. Actually the
classification of Fano 3-folds has an interesting history. This classification was be-
gun by Fano, but he missed a degree 22 3-fold V22 in CP13 . It was then found by
Iskovskikh [Isk79] in his study of Fano’s work, but a mistake was made and not all
were found. Mukai and Umemura [MU83] (See also [IP99]) produced a V22 that is
an equivariant compactification of SL(2, C)/I that was missed by Iskovskikh. Here I
is the icosahedral group. Later Prokhorov (see Proposition 4.3.11 of [IP99]) showed
that the Mukai-Umemura V22 completes the Fano-Iskovskikh classification of Fano
3-folds. Interestingly Tian [Tia97, Tia00] showed that there are deformations of
the Mukai-Umemura V22 which do not admit a Kähler-Einstein structure, giving
a counterexample to the folklore conjecture that every compact Kähler manifold
with no holomorphic vector fields admits a compatible Kähler-Einstein metric.
Thus, we have a blueprint for a complete classification of regular Sasaki-Einstein
7-manifolds. Determine which Fano 3-folds admit Kähler-Einstein metrics and con-
struct the corresponding anti-canonical circle bundle. Here we content ourselves by
quoting some general structure results which follow from the classification together
with results of Mori and Mukai [MM82].
Proposition 11.7.3: Let S be a compact regular Sasaki-Einstein 7-manifold. Then
(i) b2 (S) ≤ 9.
(ii) If b2 (S) ≥ 5 then S is reducible. Explicitly, S ≈ S 3 ?Sk , where 4 ≤ k ≤ 8.
Notice that the upper bound on the second Betti number for regular Sasaki-
Einstein 7-manifolds is realized by S 3 ?S8 , and there is a continuous 16-dimensional
family of Sasaki-Einstein structures on these manifolds. It is interesting to contem-
plate whether it is generally true that the bound on b2 is realized by a SE-reducible
element in RE. If this were true then the regular Sasaki-Einstein 9-manifold with
the largest second Betti number would be S8 ? S8 with b2 = 17. The answer proba-
bly lies in Mori theory. We end this section with a finiteness theorem that follows
from Theorem 7.5.34.
Theorem 11.7.4: There are at most a finite number of deformation types of reg-
ular Sasaki-Einstein manifolds in any given dimension.
11.7. SASAKI-EINSTEIN METRICS IN DIMENSIONS SEVEN AND HIGHER 411

11.7.1. Reducible Sasaki-Einstein Manifolds. We now give a brief gen-


eral discussion of the join construction. We know that Theorem 11.6.3 constructs
Sasaki-Einstein metrics on a new orbifold, and one can impose the condition of
Proposition 11.6.4 to assure that we obtain a smooth manifold. Once this is ac-
complished the more difficult part is to determine the homeomorphism or even
diffeomorphism type of the new manifold. Here we outline a procedure for deter-
mining the cohomology groups by studying the spectral sequences in the commu-
tative diagram 7.6.5 where the pair (k1 , k2 ) is now the relative index pair (l1 , l2 ),
so our diagram becomes
1
M1 × M2 −−−−→ M1 ? M2 −−−−→ BS

  
  
  ψ
(11.7.1) = y 
y y
M1 × M2 −−−−→ BZ1 × BZ2 −−−−→ BS 1 × BS 1 .
We reiterate that the important point in the analysis of this diagram is that the
differentials in the spectral sequence of the top fibration are determined through
naturality by the differentials in the sequence of the bottom fibration. If we have
complete information about the fibrations in the lower row, we can get completely
cohomological information about the top row. Here we present one such result
using this technique and refer the reader to [BG00b] for more examples.
Theorem 11.7.5: Let Sk denote the circle bundles over CP2 #kCP2 determined by
the positive generator on the ray containing first Chern class c1 (CP2 #kCP2 ). Then
the integral cohomology ring of the Sasaki-Einstein 7-manifolds S 3 ? Sk is given by


Z if q = 0, 7;

Zk+1 if q = 2, 5;
H q (S 3 ? Sk , Z) ≈

Zk2 if q = 4;


0 if otherwise,
with the ring relations determined by αi ∪ αj = 0, s2 = 0, 2αi ∪ s = 0, where αi , s
are the k + 1 two classes with i = 1, . . . , k. Moreover, if 3 ≤ k ≤ 8 the manifolds
S 3 ? Sk admit Sasaki-Einstein metrics.

Proof. We know that Sk is diffeomorphic to k(S 2 × S 3 ). By the Kobayashi-


Ochiai Theorem [KO73] we have Ind(Sk ) = 1. Thus, we have (l1 , l2 ) = (2, 1). By
considering the E2 term of the spectral sequence of the fibration in the lower row
of Diagram 11.7.1, we can compute the relevant differentials by naturality. Letting
u denote orientation class of S 3 , and βi the 3-classes in Sk , we find d4 (u) = s2 , and
d2 (βi ) = −2s ⊗ αi . This gives rise to torsion classes E32,2 ≈ Zk2 which survive to
E∞ . The remainder now follows by naturality and Poincaré duality. See [BG00b]
for more details. ¤

It is interesting to replace the manifolds Sk by one of their weighted counter-


parts in which case there are many more examples with Sasaki-Einstein metrics.
For example, we can consider the join S 3 ?2,1 k(S 2 × S 3 )w or S 3 ?1,1 k(S 2 × S 3 )w ,
where we write Sk (w) = k(S 2 × S 3 )w for any of the Sasaki-Einstein manifolds
consider in [BGN03b, BGN02b, BG03, Kol07] with Fano index I = 1 in the
first case, and I = 2 in the second. This gives Sasaki-Einstein metrics on simply
412 11. SASAKI-EINSTEIN GEOMETRY

connected 7-manifolds S 3 ? Sk (w) whose rational cohomology can be determined


as in the proof of Theorem 11.7.5.
Corollary 11.7.6: There are infinitely many deformation classes of Sasaki-Einstein
structures on simply connected 7-manifolds S 3 ?Sk (w) with the rational cohomology
of
(k + 1)(S 2 × S 5 ) .
This process can clearly be iterated as discussed in [BG00b]. For example,
by the iterated join of (n + 1) copies of S 3 one gets simply connected manifolds
of dimension 2n + 7 admitting Sasaki-Einstein structures and having the rational
cohomology of
n times
z }| {
S 2 × · · · × S 2 ×(k + 1)(S 2 × S 5 ) .
Together with our previous results in dimension five this gives
Corollary 11.7.7: There exist Sasaki-Einstein structures on simply connected man-
ifolds having arbitrarily large second Betti number in all odd dimensions greater
three.
Let us now consider the case S 3 ? Sw 5 5
where Sw is one of the 80 weighted 5-
spheres that admit Sasaki-Einstein metrics determined by Corollary 11.5.3. Rather
than using the method outlined above, we can apply Proposition 7.6.7 directly
giving sphere (or lens space) bundles over S 2 . This will allow us to determine the
homeomorphism type in many cases. Let L5 (l2 ) denote the lens space S 5 /Cl2 where
1
Cl2 ≈ Zl2 is the cyclic subgroup of the circle group Sw generated by the Reeb vector
field of the corresponding Sasakian structure, and l1 is the reduced Fano index of
5
Sw with respect to S 3 . It is straightforward to compute l2 = l2 (w) as a function of
w. Then we have
Theorem 11.7.8: If gcd(l2 , υ2 ) = 1 then S 3 ?l1 ,l2 Sw 5
is the total space of the fibre
2 5
bundle over S with fibre the lens space L (l2 ), and it admits Sasaki-Einstein met-
rics. In particular, for 16 different weight vectors w, the manifold S 3 ?2,1 Sw 5
is
2 5
homeomorphic to S × S and admits Sasaki-Einstein metrics including one 10-
dimensional family. Moreover, for three different weight vectors w, the manifold
S 3 ?1,1 Sw
5
is homeomorphic to S 2 × S 5 and admits Sasaki-Einstein metrics.

Proof. As mentioned above this is constructed using Proposition 7.6.7 with


M1 = S 3 with its standard round sphere Sasaki-Einstein structure, and M2 one
of the Sasaki-Einstein structures on S 5 mentioned above. So the first statement
follows. To prove the second statement we need to compute the relative indices.
Since I(S 3 ) = 2 for the standard Sasakian structure on S 3 , we need to consider
5
two cases, namely when I(Sw ) = 1, or 2. In both cases the relative index l2 = 1,
so we have S ?l1 ,1 Sw which is an S 5 bundle over S 2 . These are classified by their
3 5

second Stiefel-Whitney class w2 . But by construction the orbifold first Chern class
of CP1 × Z is proportional to the first Chern class of the S 1 -orbibundle defining
S 3 ?l1 ,1 Sw
5
. This implies that w2 (S 3 ?l1 ,1 Sw
5
) vanishes by Theorem 11.1.2. Now
5
all of the Sasakian structures on S can be represented as links of Brieskorn-Pham
polynomials of the form f = z0a0 + z1a1 + z2a2 + z3a3 or deformations thereof. So
we need to compute the Fano index I for the 80 cases in Corollary 11.5.3. As
mentioned in the proof of that corollary 68 were found in [BGK05], and 12 found
11.7. SASAKI-EINSTEIN METRICS IN DIMENSIONS SEVEN AND HIGHER 413

in [GK05]. For Brieskorn-Pham links the Fano index takes the form
³X3
1 ´
I = |w| − d = lcm(a0 , a1 , a2 , a3 ) −1 .
a
i=0 i

It is easy to write a Maple program to determine the Fano index of the 80 cases.
There are 16 with I = 1 and 3 with I = 2 giving 19 in all. For example the 10
parameter family given by Example 5.5.8 has a = (2, 3, 7, 35), and one easily sees
that I = 1. ¤
5
Remarks 11.7.1: We remark that only the Sw found in [BGK05] give rise to
2 5
Sasaki-Einstein metrics on S × S ; the ones found in [GK05] all have Fano index
greater than 2 and give non-trivial lens spaces. In fact the largest l2 obtained is
89. Generally, assuming gcd(l2 , υ2 ) = 1, an easy spectral sequence argument shows
that the manifolds S 3 ?l1 ,l2 Sw5
are simply connected with the rational homology
type of S × S , but with H (S 3 ?l1 ,l2 Sw
2 5 4 5
, Z) ≈ Zl2 .

11.7.2. Irreducible Sasaki-Einstein Manifolds: Rational Homology


Spheres. Here we discuss the problem of existence of Sasaki-Einstein metrics that
are SE-irreducible, so we need to rely on methods other than the join construction.
The most common method is that of construction Sasakian structures on links of
hypersurface singularities of weighted homogeneous polynomials. This is similar to
the 5-dimensional links discussed in the previous section. We begin with the case
of rational homology spheres which appeared in [BG06a]. By Corollary 7.6.12
rational homology spheres are always SE-irreducible.
Let f = f (z1 , . . . , zm ) be a quasi-smooth weighted homogeneous polynomial of
degree df in m complex variables, and let Lf denote its link. Let wf = (w1 , . . . , wm )
be the corresponding weight vector. We consider branched covers given by equa-
tion (9.3.20). Then Lg is a k-fold branched cover of S 2m+1 branched over the
link Lf . The degree of Lg is dg = lcm(k, df ), and the weight vector is wg =
¡ df k
¢
gcd(k,df ) , gcd(k,df ) wf . We shall always assume that k ≥ 2, since the linear case
k = 1 is a hyperplane in a weighted projective space. Notice that if gcd(k, d) > 1 and
gcd(k, ui ) = 1 for each i = 1, . . . , m, then any common factor of k and d must divide
wi for each i, and so will be an overall common factor which gives an equivalent link.
So hereafter we shall assume that gcd(k, d) = 1. Theorem 9.3.17 says that the link
Lg is a rational homology sphere such that the order |Hm−1 (Lg , Z)| = k bm−2 (Lf ) ,
where bm−2 (Lf ) is the (m − 2)nd Betti number of Lf . Let X = Xk,d,w denote the
quotient orbifold of Lg by the natural S 1 action. It is easy to see that Xk,d,w is
Fano if and only if
(11.7.2) k(|wf | − df ) + df > 0 .
It is easy to see that the klt condition (5.2.13) cannot be satisfied when |wf |−df > 0.
First we consider the case wf − d = 0. Then the klt condition (5.2.13) becomes
(11.7.3) (m − 1)d < mk min{wi } .
i
©1ª
This is clearly satisfied for k large enough, namely for k > m−1m d max wi . We
are particularly interested in the case of perturbations of Brieskorn-Pham (BP)
singularities where we can establish the existence results via Theorem 5.4.11. We
414 11. SASAKI-EINSTEIN GEOMETRY

will apply it to the weighted homogeneous polynomials of the form


m
X
P (z0 , . . . , zm ) = ziai + tp(z0 , . . . , zm ) ,
i=0

where t ∈ C, and w(p) = lcm(a0 , . . . , am ) which is the degree of the polynomial.


We impose a condition on the zero set Y (a, p) = P −1 (0), namely the genericity
condition (GC) of Example 4.6.7.
Example 11.7.9: [Branched Covers of Calabi-Yau hypersurfaces] First we consider
k-fold branched covers of Fermat-Calabi-Yau hypersurfaces
(11.7.4) z0k + z1m + · · · zm
m
=0
with gcd(k, m) = 1. For m ≥ 3 the link Mk2m−1 is a simply connected nontrivial
rational homology sphere. The (m − 2)-nd Betti number of the Calabi-Yau link is
³ (1 − m)m − 1 ´
bm−2 = (−1)m 1 + .
m
So by Theorem 9.3.17 |Hm−1 (Mk2m−1 , Z)| = k bm−2 . For example, in dimension 7
(m = 4) we get |H2 (Mk7 , Z)| = k 21 whereas, in dimension 9 we have |H2 (Mk9 , Z)| =
k 204 .
It is easy to see that the klt condition in Theorem 5.4.11 is satisfied if k >
m(m − 1) in which case the rational homology spheres Mk2m−1 admits a family of
Sasaki-Einstein metrics. From Theorem 5.5.7 one can easily determine the number
of effective real parameters µ, namely
µ ¶
2m − 1
µ=2 − 2m2 .
m
By Sterling’s formula one sees that µ grows exponentially with m. In dimension 7 we
get 38 real parameters, while in dimension 9 we have 202 effective real parameters.
In the orbifold category there are many Calabi-Yau hypersurfaces in weighted
projective spaces P(w). For example, one can consider w = (1, . . . , 1, m − 1) of
degree d = 2(m − 1), or w = (1, . . . , 1, m − 2, m − 2) of degree d = 3(m − 2), etc.
For rational homology spheres of dimension 7, we obtain examples from branched
covers of Reid’s list (cf. [IF00]) of 95 log K3 surfaces, and in dimension 9 from
branched covers of the over 6000 Calabi-Yau orbifolds in complex dimension 3
[CLS90].
Example 11.7.10: [Branched Covers in the Canonical Case] We consider
links LF of branched covers of a canonical Fermat hypersurface of the form
F = z0k + z1l + · · · + zm
l
=0
with l ≥ m + 1 and gcd(k, l) = 1. Here the quotient X will be Fano if and only if
l
k< .
l−m
Also for any branched cover we are only interested in k ≥ 2. Combining this with
the Fano condition gives an upper bound on l for fixed m, namely l < 2m. Thus,
we get the range for l
(11.7.5) m + 1 ≤ l ≤ 2m − 1 .
11.7. SASAKI-EINSTEIN METRICS IN DIMENSIONS SEVEN AND HIGHER 415

The link LF = Mk2m−1 is a rational homology sphere with |Hm−1 (Mk2m−1 , Z)| =
k bm−2 , where
³ (1 − l)m − 1 ´
(11.7.6) bm−2 = (−1)m 1 + .
l
We now look at the klt condition, that is, the right hand inequality of Theorem
5.4.11. We see that LF will have Sasaki-Einstein metrics if we choose k to satisfy
(m − 1)l2 l
(11.7.7) <k< .
(m − 1)l(l − m) + m l−m
As above we can compute the number of effective parameters µ to be
µ ¶
m+l−1
(11.7.8) µ=2 − 2m2 .
l
As a special case we consider l = m + 1. In this case the only solution to the
inequality 11.7.7 is k = m. This
¡ 2m ¢ gives Sasaki-Einstein metrics on rational homology
spheres depending on 2( m+1 − m2 ) effective real parameters.
Other solutions can be worked out, for example, the singularity
2m
z0k + z12m + · · · + zm−1 2
+ zm = 0.
This satisfies the inequalities of Theorem 5.4.11 if we choose k = 2m − 1.
Theorem 11.7.11: There exists continuous parameter families of Sasaki-Einstein
metrics on infinitely many simply connected rational homology spheres in every
odd dimension greater than 3. For some of these families the number of effective
parameters grows exponentially with dimension.
Example 11.7.12: In [BGN02a] we studied “sporadic” rational homology 7-spheres
which admit Sasaki-Einstein metric that were extracted from a classification of well-
formed orbifold Fano hypersurfaces anti-canonically embedded in CP4 (w) which
can be shown to admit Kähler-Einstein metrics using the estimate of Corollary
5.4.8. This classification was carried out by Johnson and Kollár in [JK01a] via
a computer search. We constructed a table of these sporadic rational homology
7-spheres which is part of an earlier preprint version of this article available as
math.DG/0108113.
Our table lists the weights w = (w0 , w1 , w2 , w3 , w4 ), degree d, Milnor number
7 7
µ, and the order of H3 (Mw,d , Z) of the rational homology 7-sphere Mw,d . It lists
7
Mw,d ’s in increasing order of their weights beginning with w0 . The first entry has
weights (17, 34, 75, 125, 175) while the last has weights (357, 388, 2231, 2975, 5593).
In the first entry the order of H3 is huge, 1712 a number over 500 trillion, while the
lowest order of H3 is 132 = 169. In general the orders of H3 tend to be quite large.
From a quick perusal of the table, it is easy to notice the existence of twins.
These are rational homology 7-spheres with the same degree d, Milnor number µ
and order of H3 . Twins often occur as adjacent listings with the same w0 , but this
is not always the case as with twins d = |H3 | = 10881, µ = 10880 and w0 = 101 and
109, and the twins d = |H3 | = 7777 with w0 = 141 and w0 = 167. Twins may also be
members of a larger set, such as the septuplets with d = |H3 | = 5761 and µ = 5760.
These have w0 = 157, 157, 185, 205, 214, 253, 271, respectively. Since twins have
the same Milnor number, it is tempting to conjecture that twins correspond to
homeomorphic or even diffeomorphic links, but we have no proof yet.
416 11. SASAKI-EINSTEIN GEOMETRY

11.7.3. Other Sasaki-Einstein Links. Our method of using links of hy-


persurface singularities to produce examples of Sasaki-Einstein metrics on homo-
topy spheres and rational homology spheres can also be used for general higher-
dimensional manifolds. The more difficult problem is determining the diffeomor-
phism or even homeomorphism type of the manifold. We have written a computer
program that computes the relevant Betti number of the link as well as the torsion
using Orlik’s Conjecture 9.3.15. So it can be applied to Brieskorn-Pham links or
the other type of link where Orlik’s algorithm is known to hold, cf. Proposition
9.3.16. The list of 4442 well-formed quasismooth Q-Fano 3-folds of index one found
by Johnson and Kollár [JK01a] contains 1936 3-folds where the klt estimates of
Corollary 5.4.8 are satisfied, and so admit a Kähler-Einstein metric. Then The-
orem 11.1.12 implies that the corresponding link L∗ (w0 , w1 , w2 , w3 , w4 ; d) admits
Sasaki-Einstein metrics. More generally, we have
Theorem 11.7.13: Let L∗ (w0 , . . . , wn ; d) be the link of an isolated hypersurface
singularity of degree d, with weight vector w = (w0 , . . . , wn ), and Fano index I.
Suppose also that the estimate
n
dI < min{wi wj }
(n − 1) i,j

holds. Then L∗ (w0 , . . . , wn ; d) admits a Sasaki-Einstein metric.


As usual the above estimate is far from optimal. For example as we saw the
estimate in Theorem 5.4.11 works better in the case of BP polynomials. Given
this to be the case we conclude our discussion by giving some examples. For each
example a different klt estimate is used.
Example 11.7.14: Here is a pair of twins taken from the end of the Johnson-
Kollár list. The klt estimate used here is the one of Theorem 11.7.13. The two
links are L∗ (355, 631, 5279, 12528; 37584) and L∗ (407, 547, 5311, 12528; 37584). Our
computer program shows that they both have b3 = 2, and Orlik’s algorithm suggests
that there is no torsion. However, we do not know whether Orlik’s Conjecture 9.3.15
holds in this case. If it does then these two links are homeomorphic to 2(S 3 × S 4 )
by Theorem 9.4.10. It would be interesting to compute their diffeomorphism type.
They would be diffeomorphic to 2(S 3 × S 4 )#Σ7 (w) where Σ7 (w) is a homotopy
7-sphere in bP8 which in principal depends on the weight vector w.
Next we consider examples of regular Sasaki-Einstein manifolds. These are cir-
cle bundles over the well-known Fermat hypersurfaces described in Example 5.2.6.
Example 11.7.15: [Links of Fermat Hypersurfaces]. First there are the well-
known Stiefel manifolds V2 (Rn+1 ) of 2-frames in Rn+1 . These are circle bundles
over the complex quadrics Fn,2 , and they admit Sasaki-Einstein metrics since their
base spaces Fn,2 = G2 (Rn+1 ) admit Kähler-Einstein metrics. Furthermore, they
are all homogeneous Sasaki-Einstein, and G2 (Rn+1 ) is a generalized flag manifold
whose parabolic subgroup P is maximal. The manifolds V2 (R2m+1 ) are rational
homology spheres with H 2m (V2 (R2m+1 ), Z) = Z2 , whereas, the manifolds V2 (R2m )
have H 2m−2 (V2 (R2m ), Z) = Z.
It was seen in Example 5.2.6 the Fn,p are Fano when p < n + 1 and that
the Nadel-Sui-Tian klt conditions hold for all Fermat hypersurfaces satisfying n2 ≤
p ≤ n. The BP link Sp,n = L(p, . . . , p) then admits regular Sasaki-Einstein metrics
for each p in the range n2 ≤ p ≤ n. The topology can be determined from the
11.8. SASAKIAN η-EINSTEIN METRICS 417

well-known equation for the nth Betti number


³ (1 − p)n+1 − 1 ´
bn = bn (p) = (−1)n−1 1 + .
p
The torsion can be computed by Orlik’s algorithm by Proposition 9.3.16. This
suggests the general result
(
n Zbn ⊕ Zp if n is even;
H (Sp,n , Z) =
Zbn if n is odd.

For example in dimension 7 our computer program gives H 4 (S4,4 , Z) = Z60 ⊕Z4
and H 4 (S3,4 , Z) = Z10 ⊕ Z3 , whereas, in dimension 9 we have H 5 (S5,5 , Z) = Z820
and H 5 (S4,5 , Z) = Z183 .
Our last example is a branched cover of a Fermat hypersurface.
Example 11.7.16: [Branched Covers of Fermat Hypersurfaces]. We consider the
link L(m, n) = L(mn, n, . . . , n) with BP polynomial f = z0mn + z1n + · · · znn of degree
mn. It is easy to see that the conditions of Theorem 5.4.11 are satisfied when m ≥ n.
The formula for the Betti number is straightforward, but somewhat messy and left
as an exercise. Furthermore, Orlik’s Conjecture holds in this case by Proposition
9.3.16. In dimension 7 our computer program suggests H 4 (L(m, 4), Z) = Z60 ⊕
Z4m ⊕ 20Zm . We have checked that this formula is correct for about 10 different
values of m, and leave its direct verification as an exercise. Similarly in dimension
9 our program suggests H 5 (L(m, 5), Z) = Z820 ⊕ 204Zm .
Many other similar examples can be worked out.
Exercise 11.7: Discuss the Sasaki-Einstein moduli problem for the links of Ex-
ample 11.7.15.
Exercise 11.8: Discuss the Sasaki-Einstein moduli problem for the links of Ex-
ample 11.7.16.

11.8. Sasakian η-Einstein Metrics


Here we discuss in some detail the less well-known case of η-Einstein metrics
that were defined in Definition 11.1.1. These metrics were introduced and studied
by Okumura [Oku62] and then named by Sasaki [Sas65] in his lecture notes in
1965. A more recent study of such metrics can be found in [BGM06].
For η-Einstein metrics we are mainly interested in K-contact and Sasakian
structures. More generally we could have defined an η-Einstein structure as in Def-
inition 11.1.1, but where a and b are smooth functions on M. However, analogously
to the Einstein case, we have
Proposition 11.8.1: Let (ξ, η, Φ, g) be a K-contact structure on M, and suppose
also that g satisfies
Ricg = ag + bη ⊗ η
for smooth functions a, b. Then if n > 1 the functions a and b are constants.

Proof. For an η-Einstein metric we have s = (2n + 1)a + b; whereas, since g


is also K-contact we have a + b = 2n. Thus, s = 2n(a + 1) and b = 2n − a, and
it suffices to show that s is constant. Now the contracted Bianchi identities are
418 11. SASAKI-EINSTEIN GEOMETRY

ds = 2δRicg where s is the scalar curvature and δ is the codifferential (cf.[Bes87]).


So
¡ ¢
ds = 2δRicg = 2δ ag + (2n − a)η ⊗ η = 2dB a .
Here we have used the fact that on tensor fields r the divergence operator δ satisfies
X
δr(X1 , . . . , Xr ) = (∇Ei r)(Ei , X1 , . . . , Xr ) ,
i

where {Ei } is a local orthonormal frame. But since ds = 2nda the result follows. ¤

Exercise 11.9: Fill in the missing details in the proof of Proposition 11.8.1 by
showing that δ((2n − a)η ⊗ η) = −(ξa)η and (δag) = (ξa)η + dB a.
Proposition 11.8.1 was first proved by Okumura [Oku62] when S = (ξ, η, Φ, g)
is Sasakian. Tanno [Tan79] also showed that for a Sasakian η-Einstein metric g
with a > −2, there is a transverse homothety transformation 7.3.10 such that the
resulting metric is Sasaki-Einstein. However, owing to Theorem 11.1.7 we can do
better [BG01a].
Theorem 11.8.2: Let (ξ, η, Φ, g) be an η-Einstein K-contact structure with a ≥
−2. Then (ξ, η, Φ, g) is Sasakian, and if a > −2 there exists a unique transverse
homothety whose resulting structure is Sasaki-Einstein.

Proof. Since g is η-Einstein it follows from (ii) of Theorem 7.3.12 that the
transverse metric gT is Einstein with Einstein constant a + 2. Since this is nonneg-
ative it follows from Theorem 11.1.7 that the almost CR structure is integrable4.
Thus, g is Sasakian. Now applying a transverse homothety 7.3.10 we get
(ξ 0 , η 0 , Φ0 , g 0 ) = (α−1 ξ, αη, Φ, αg + (α2 − α)η ⊗ η)
which is Sasakian by Proposition 7.3.20 is also Sasakian. Again using (ii) of Theo-
rem 7.3.12 we have for X, Y sections of D
Ricg0 (X, Y ) = RicgT0 (X, Y ) − 2g 0 (X, Y ) = RicgT (X, Y ) − 2αgT (X, Y )
= Ricg (X, Y ) + 2g(X, Y ) − 2αg(X, Y ) = (a + 2 − 2α)gT (X, Y )
a + 2 − 2α 0
= g (X, Y ) ,
α
where the second equality holds since the Ricci curvature of gT is invariant under a
homothety. Since g 0 is Sasakian Ricg0 (X, ξ 0 ) = 2ng 0 (X, ξ 0 ) by Theorem 7.3.12. So
a+2
g 0 is Sasaki-Einstein if and only if a > −2 and α = 2n+2 . ¤

Notice that in the 3-dimensional case (n = 1) Proposition 11.8.1 doesn’t


hold. Moreover, any K-contact manifold is automatically Sasakian in dimension
3, since two-dimensional almost complex structures are automatically integrable.
A Sasakian 3-manifold M is always η-Einstein in the wider sense of the definition
where a, b are functions on M . Hence, in this book we define a Sasakian 3-manifold
to be η-Einstein as in (11.1.1). On the other hand, when M is a metric contact
manifold one can study metrics for which Ricg = ag + bη ⊗ η, with a, b ∈ C ∞ (M )
(see [Bla02]). In dimension 3 a Sasakian η-Einstein structure is quite restrictive.

4In Theorem 11.1.7 this is proved for the case a > −2 only, but it follows easily that the
a = −2 case also holds, since Sekigawa’s proof holds in the Ricci flat case as well (cf. [BG01a]
for details).
11.8. SASAKIAN η-EINSTEIN METRICS 419

Proposition 11.8.3: A three-dimensional K-contact manifold is η-Einstein if and


only if it is Sasakian with constant Φ-sectional curvature.

Proof. The if part, which holds in all dimensions, follows easily from Tanno’s
classification theorem [Tan69b]. To prove the only if part we choose a local or-
thonormal bases X, ΦX, ξ for the Sasakian structure S = (ξ, η, Φ, g). Now in three
dimensions the Ricci curvature and the sectional curvature are related by [Pet98]
Ricg (X, X) = K(X, ΦX) + K(X, ξ) = K(X, ΦX) + 1 ,
Ricg (ΦX, ΦX) = K(X, ΦX) + K(ΦX, ξ) = K(X, ΦX) + 1 ,
Ricg (ξ, ξ) = K(X, ξ) + K(ΦX, ξ) = 2.
So if g is η-Einstein we have
Ricg = ag + (2 − a)η ⊗ η ,
and this gives K(X, ΦX) = a − 1. ¤

Thus, up to transverse homothety the η-Einstein metrics on a three-dimensional


manifold are the uniformizing metrics given in Theorem 10.1.3.
Now if g is a Sasakian-η-Einstein metric on M, then ρT = aωD for some constant
a, so [ρT ] = a[dη]. Thus, as a real cohomology class c1 (D) = ι∗ [ρT ] = 0, and from
Corollary 7.5.26 we have
Theorem 11.8.4: Let (M, g, ξ, η) be a compact Sasakian manifold with basic first
Chern class c1 (Fξ ) proportional to [dη]B . Suppose further that M has a Sasakian-
η-Einstein metric in the same deformation class F(Fξ ) as c1 (Fξ ). Then, c1 (D) is
a torsion class. In particular, if M is simply connected then c1 (D) = 0.
Hence, a non-torsion c1 (D) is the obstruction to the existence of a Sasakian-η-
Einstein metric. In particular in the case when a > −2 there is a canonical variation
to a Sasaki-Einstein metric[Tan, BG2], so c1 (D) is an obstruction to the existence
of a Sasaki-Einstein metric on M in the given deformation class.
Thus, Theorems 7.5.27 and 11.8.4 give
Corollary 11.8.5: Let (M, g, ξ, η) be a compact Sasakian-η-Einstein manifold with
H1 (M, Z)tor = 0. Then M is spin. In particular, a compact simply connected
Sasakian-η-Einstein manifold is spin.
We are interested in the transverse Calabi problem and its solution, Theorem
7.5.19 due to El Kacimi-Alaoui. From this theorem it is straightforward to obtain
the transverse version of the well-known Aubin-Yau result Theorem 5.2.2, namely
2
Theorem 11.8.6: If the basic first Chern class c1 (Fξ ) ∈ HB (Fξ ) is zero or can
be represented by a negative definite (1, 1)-form, then there exists a Sasakian-η-
Einstein metric g on M in the given ξ-deformation class with a = −2 in the first
case and a < −2 in the second. In both cases c1 (D) is a torsion class.
This theorem says that all null Sasakian and negative Sasakian structures ad-
mit the corresponding Sasakian-η-Einstein metrics. This applies to Reid’s list of 95
K3 surfaces, as well as all the examples of both null and negative Sasakian struc-
tures discussed in Section 10.3. Indeed, it is easy to construct examples of negative
Sasakian η-Einstein metrics on many manifolds in arbitrary odd dimensions, al-
though it appears much harder to obtain a systematic procedure as in the positive
case. Here we mention some results concerning homotopy spheres [BGM06].
420 11. SASAKI-EINSTEIN GEOMETRY

Example 11.8.7: Let the integers ri , i = 1, . . . , 2m be any pairwise prime positive


integers. Then the (4m + 1)-dimensional link L(2, 2r1 , . . . , 2r2m , a) is diffeomorphic
to the standard sphere if a ≡ P±1(8) and to the Kervaire sphere if a ≡ ±3(8).
The link will be negative when i r1i < a−2a which is easily satisfied for ri ’s large
enough. This gives
Theorem 11.8.8: In any dimension 4m + 1 both the standard and the Kervaire
spheres admit infinitely many inequivalent negative Sasakian η-Einstein structures.
In dimension 4m+3 one can construct examples of negative Sasakian structures
should exists on all homotopy spheres which bound parallelizable manifolds. That is
to say, all BP links that by Brieskorn Graph Theorem are homeomorphic to a sphere
of dimension 4m + 3 and are negative should easily contain all possible oriented
diffeomorphism types. We have checked it in dimension 7 with the following
Theorem 11.8.9: All 28 oriented diffeomorphism types of homotopy 7-spheres
admit negative Sasakian η-Einstein structures. Hence, they all admit Lorentzian
Sasaki-Einstein structures.
Proof. Consider L(k, k, k, k + 1, p), where p is prime to both k and k + 1. The
link L is a homotopy 7-sphere and it is negative for k and p large enough. Using
the computer codes of [BGKT05], it is easy to check that such links realize all 28
oriented diffeomorphism types. ¤
Similarly, one should be able to show that the homotopy spheres in dimensions
11 and 15 that bound parallelizable manifolds admit negative Sasakian structures.
11.8.1. Lorentzian-Sasakian Geometry. There is a canonical way to asso-
ciate a Lorentzian signature ‘metric’ to any Sasakian manifold. Let S = (ξ, η, Φ, g)
be a Sasakian structure on a manifold M. Then one easily sees that
(11.8.1) gL = g − 2η ⊗ η = dη ◦ (Φ × 1l) − η ⊗ η
defines a pseudo-Riemannian metric with signature (+, . . . , +, −) on M, and it is
clearly canonically associated to the Sasakian structure. In fact every Sasakian
structure has both a Riemannian and a Lorentzian metric associated to it in a
natural way. Notice that as with Riemannian metrics the same Lorentzian metric
is associated to both S and its conjugate S c . Recall [Bau00, Boh03] the following
Definition 11.8.10: Let (M, g) be a Lorentzian manifold of dimension 2n + 1 and
let ξ be a time-like Killing vector field such that g(ξ, ξ) = −1. We say that M is
Sasakian if Φ(X) = ∇X ξ satisfies the condition (∇X Φ)(Y ) = g(X, ξ)Y + g(X, Y )ξ
and Sasaki-Einstein if, in addition, g is Einstein.
Equivalently, we can require the metric cone C(M ) = (R+ ×M, −dt2 +r2 g, d(r2 η))
to be pseudo-Kähler of signature (2, 2n). In the Sasaki-Einstein case the cone met-
ric is pseudo-Calabi-Yau and the Einstein constant must equal to −2n. This is in
complete analogy with the Riemannian case.
Recall from Example 1.4.11 that a manifold admits a Lorentzian metric if and
only if it has a nowhere vanishing vector field.
CHAPTER 12

Quaternionic Kähler and Hyperkähler Manifolds

Quaternions were first described by the Sir William Rowan Hamilton1 in 1843.
Hamilton believed that his invention, like complex numbers, should play a funda-
mental rôle in mathematics as well as in physics. The jury is perhaps still out
on what, if any, importance should quaternions have in describing our physical
world. But there is little or no doubt that they have earned an important place in
Riemannian and algebraic geometry. Following Hitchin [Hit92] we would like to
argue that today’s rich theory of quaternionic manifolds, in some sense, vindicates
Hamilton’s conviction.
In this chapter we will recall some basic results concerning various quaternionic
geometries which were introduced briefly from the point of view of G-structures in
Example 1.4.18. Our main focus will be on positive quaternionic Kähler (QK) and
hyeprkähler (HK) manifolds, as these two geometries are of special importance in
the description and understanding of 3-Sasakian structures, the main topic of our
next chapter. It would be impossible here, in a single chapter, to give a complete
account of what is currently known about QK and HK spaces. Each case would re-
quire a separate monograph. Our goal is to describe some of the properties of such
manifolds relevant to Sasakian Geometry. Quaternionic Kähler geometry is tradi-
tionally defined by the reduction of the holonomy group Hol(M, g) to a subgroup of
Sp(n)Sp(1) ⊂ SO(4n, R). Observe that Sp(1) · Sp(1) ' SO(4) so any oriented Rie-
mannian 4-manifold has this property. It is generally accepted and, as we shall see
later, quite natural, to extend this definition in dimension 4 via an additional cur-
vature condition: an oriented Riemannian manifold (M 4 , g) is said to be QK if the
metric g is self-dual or anti-self-dual and Einstein. Interest in QK manifolds and this
holonomy definition dates back to the celebrated Berger Theorem 1.4.8. The Lie
group Sp(n)Sp(1) appears on Berger’s list of possible restricted holonomy groups of
an oriented Riemannian manifold (M, g) which is neither locally a product nor lo-
cally symmetric. In particular, the holonomy reduction implies that QK manifolds
are always Einstein [Ber66], though their geometric nature very much depends on
the sign of the scalar curvature. The model example of a QK manifold with positive
scalar curvature (positive QK manifold) is that of the quaternionic projective space
HPn . The model example of a QK manifold with negative scalar curvature (negative
QK manifold) is that of the quaternionic hyperbolic ball HHn . The first attempts
to study QK manifolds span over a decade and date back to the works of Bonan
[Bon64], Kraines [Kra65, Kra66], Wolf [Wol65], Alekseevsky [Ale68, Ale75],
Gray [Gra69a], Ishihara and Konishi [IK72, Ish73, Ish74, Kon75]. At this
early stage the departing point was the holonomy reduction and efforts to under-
stand what kind of geometric structures on the manifold would naturally lead to
1As a tribute, in this chapter, as in most literature on quaternionic structures, H stands for
Sir William Hamilton.
421
422 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

such a holonomy reduction. For example, it appears that, independently, Bonan


and Kraines were the first to consider the fundamental 4-form Ω of the quaternionic
structure and deduced some topological information. Wolf and Alekseevsky studied
and classified symmetric and some homogeneous examples, respectively. Ishihara
and Konishi explored some special geometric properties of such manifolds and their
relation to the 3-Sasakian spaces. The true revolution, however, came in the early
eighties. Salamon [Sal82] and, independently, Bérard Bergery [BB82] realized
that QK manifolds can be studied in the language of algebraic and holomorphic
geometry. Their twistor correspondence was a generalization of the beautiful Pen-
rose twistor space construction in dimension 4 to the case of QK manifold of any
quaternionic dimension. The power of the twistor correspondence which allows for
applying algebraic geometry when dealing with problems involving positive QK
manifolds will be illustrated by many results described in this chapter.
When the scalar curvature vanishes a QK manifold is necessarily locally hy-
perkähler. In the language of holonomy the hyperkähler manifolds are character-
ized by the reduction of the holonomy group Hol(M, g) to a subgroup of Sp(n) ⊂
SO(4n, R). In this sense HK geometry is a special case of QK geometry and, just
as in the previous case, we find the Lie group Sp(n) on Berger’s list. The model
example of HK geometry is that of a quaternionic vector space Hn with the flat
metric. The first study of hyperkähler manifolds appears to be that of Wakakuwa
[Wak58] who gave an example in local coordinates, but the name hyperkähler,
as well the name hypercomplex, is due to Calabi [Cal79] who constructed com-
plete hyperkähler metrics on the cotangent bundle T ∗ CPn . Of course, Yau’s famous
proof [Yau77] of the Calabi conjecture provides the K3 surface with a hyperkähler
structure. Hyperkähler manifolds are special cases of Calabi-Yau manifolds, and
so compact examples are important to mirror symmetry. There are several recent
books [VK99, GHJ03, NW04] treating hyperkähler manifolds. In this book we
are more interested in non-compact hyperkähler manifolds, especially hyperkähler
cones.

12.1. Quaternionic Geometry of Hn and HPn


The purpose of this section is to describe quaternionic geometries of some model
examples of quaternionic manifolds. We will do it in considerable detail using terms
which, in greater generality, will only be defined later. The quaternions H are the
associative, non-commutative real algebra
H = {u | u = u0 + u1 i1 + u2 i2 + u3 i3 , ua ∈ R} ' R4 .
The imaginary units are often denoted by {i1 , i2 , i3 } = {i, j, k}. The imaginary
quaternions Im(H) = span(i1 , i2 , i3 ) ' R3 and the multiplication rules are given by
the formula
3
X
(12.1.1) ia ib = −δab + ²abc ic ,
c=1

We define the quaternionic conjugate q̄ and the norm |u| by


3
X 3
X
ū = u0 − ua ia , and |u|2 = (ua )2 .
a=1 a=0
12.1. QUATERNIONIC GEOMETRY OF Hn AND HPn 423

The non-zero quaternions H \ {0} = H∗ = GL(1, H) from a group isomorphic to


R+ × Sp(1), where Sp(1) is the subgroup of unit quaternions and the isomorphism
is given explicitly by the map u 7→ (|u|, u/|u|). The group of unit quaternions
(12.1.2) Sp(1) = {σ ∈ H∗ ' GL(1, H) | σ̄σ = 1} ,
as a manifold, is just the unit 3-sphere in R4 . Furthermore, we have the group
isomorphism f : Sp(1) → SU (2) explicitly given by
µ 0 ¶
σ + σ 1 i1 σ 2 + σ 3 i1
(12.1.3) f (σ) = .
−σ 2 + σ 3 i1 σ 0 − σ 1 i1
It is known that Spin(4) = Sp(1) × Sp(1) and SO(4) ' Sp(1)Sp(1), where cus-
tomarily Sp(1)Sp(1) denotes the quotient of Sp(1) × Sp(1) by the diagonal Z2 .
This is yet another group isomorphism between classical groups which can be
explained using the quaternionic geometry of H ' R4 . Consider the action of
G = Sp(1)+ × Sp(1)− on H given by
(12.1.4) ϕ(σ,λ) (u) = σuλ̄ .
We assume the convention that the Sp(1)+ factor acts by the left quaternionic
multiplication while the Sp(1)− factor acts from the right. Clearly, the two actions
commute and the Z2 subgroup generated by (−1, −1) acts trivially. The quotient
acts on R4 preserving the Euclidean metric and orientation. This is the special
orthogonal group SO(4). It is worthwhile to write this action on R4 . The Sp(1)+
part is given by the following group homomorphism A+ : Sp(1) → SO(4):
 0 
σ −σ 1 −σ 2 −σ 3
σ 1 σ 0 −σ 3 σ 2 
(12.1.5) A+ (σ) = 
σ 2 σ 3
 = σ 0 1l4 + σ 1 I + + σ 2 I + + σ 3 I + ,
σ 0 −σ 1  1 2 3

σ 3 −σ 2 σ 1 σ0

where the matrices Ii+ = A+ (ei )


(12.1.6)     
0 −1 0 0 0 0 −1 0 0 0 0 −1
 1 0 0 0   0 1  0
I1+ =   + 0 0  , I + = 0 0 −1 ,
0 0 0 −1 , I2 = 1 0 0 0 3 0 1 0 0
0 0 1 0 0 −1 0 0 1 0 0 0

give a globally defined hypercomplex structure I + = {I1+ , I2+ , I3+ } on R4 . For a


purely imaginary τ = −τ̄ in Sp(1) one sets I + (τ ) = A+ (τ ) and gets the whole
S 2 -family of complex structures. We obtain the left hyperkähler structure on H
by further setting g0 − ω + = g0 − i1 ω1+ − i2 ω2+ − i3 ω3+ = du ⊗ dū, where the
multiplication in H is used to interpret the left hand side as an H-valued tensor.
This gives the standard Euclidean metric g0 and the three symplectic forms
(12.1.7) ωa+ = dub ∧ duc + du0 ∧ dua ,
where (a, b, c) is any cyclic permutation of (1, 2, 3). We can also introduce an H-
valued differential 2-form

1
(12.1.8) ω + = i1 ω1+ + i2 ω2+ + i3 ω3+ = Im(du ⊗ dū) = du ∧ dū .
2
424 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

The 2-from du ∧ dū is purely imaginary as α ∧ β = (−1)pq β̄ ∧ ᾱ, where p, q are the
respective degrees. The Sp(1)− part is given by A− : Sp(1) → SO(4) with
 0 
λ λ1 λ2 λ3
−λ1 λ0 −λ3 λ2 
(12.1.9) A− (λ) =  −λ2 λ3
 = λ0 1l4 + λ1 I − + λ2 I − + λ3 I − .
λ0 −λ1  1 2 3
3 2 1 0
−λ −λ λ λ
The matrices Ii− = A− (ei )
(12.1.10)
     
0 1 0 0 0 0 1 0 0 0 0 1
−1 0 0 0  0 1  0 −1 0

I1 =   , I− =  0 0  , I− =  0 ,
0 0 0 −1 2 −1 0 0 0 3 0 1 0 0
0 0 1 0 0 −1 0 0 −1 0 0 0
give a globally defined hypercomplex structure I − = {I1− , I2− , I3− }. Furthermore,
with the Euclidean metric one gets the right hyperkähler structure on H by setting
g0 + ω − = g0 + iω1− + i2 ω2− + i3 ω3− = dū ⊗ du. This gives
1
(12.1.11) ω − = i1 ω1− + i2 ω2− + i3 ω3− = Im(dū ⊗ du) = dū ∧ du ,
2
where, as before, we get the three symplectic forms
(12.1.12) ωa− = dub ∧ duc − du0 ∧ dua ,
for each cyclic permutation of (1, 2, 3). These are clearly fundamental 2-forms
associated to the complex structures {Ia− }. Note that by construction, for any (σ, λ)
one has [A+ (σ), A− (λ)] = 0 and the product A+ (σ)A− (λ) ∈ SO(4). In particular,
the two hypercomplex structures I + and I − commute. The hyperkähler structure
(g0 , Ia− , ωa− ) is preserved by Sp(1)+ (hypekähler isometry) while Sp(1)− acts by
rotating the complex structures on S 2 . The role of Sp(1)+ and Sp(1)− reverses
for (g0 , Ia+ , ωa+ ). With only little extra effort one can “compactify” this example
to see that another Lie group U (2) is a compact manifold with two commuting
hypercomplex structures, though U (2) admits no hyperkähler metric.
Remark 12.1.1: Consider the group of integers Z acting on H by translations of
the real axis. The action preserves both hypercomplex structures and the metric,
hence, the quotient H/Z ' S 1 × R3 is also a flat hyperkähler manifold with infinite
fundamental group π1 = Z. To indicate the difference, we will write the flat metric
in this case as g0 = dθ2 + dx · dx replacing x0 with the angle coordinate θ.
Example 12.1.1: [Quaternionic vector spaces]. Much of the above discussion
extends to Hn = {u = (u1 , . . . , un ) | uj = u0j +u1j i1 +u2j i2 +u3j i3 ∈ H, j = 1, . . . , n}.
Here and from now on we will choose to work with the left hyperkähler structure
on Hn , i.e., with the symplectic 2-forms given by
n
X
(12.1.13) g0 − ω = dūj ⊗ duj
j=1

so that
n X
X 3 n
X ¡ b ¢
(12.1.14) g0 = (dxaj )2 , ωa = dxj ∧ dxcj − dx0j ∧ dxaj
j=1 a=0 j=1
12.1. QUATERNIONIC GEOMETRY OF Hn AND HPn 425

for any cyclic permutation (a, b, c) of (1, 2, 3). The corresponding hypercomplex
structure is then given by left multiplication by {ī1 , ī2 , ī3 } = {−i, −j, −k} with the
standard basis as in 12.1.10, where 0, 1 are now matrices of size n × n.
We associate to g0 a quaternionic Hermitian inner product
n
X
(12.1.15) F (u, v) = ūj vj
j=1

and define
(12.1.16) Sp(n) = {A ∈ GL(n, H) |hAu, Avi = hu, vi} .
Now, Sp(n) × Sp(1) acts on Hn by
(12.1.17) ϕ(A,λ) (u) = A · uλ̄
with Sp(n)Sp(1) acting effectively. Clearly, Sp(n)Sp(1) is now a subgroup of
SO(4n). The group Sp(n) assumes the role of Sp(1)− and it acts by hyperkähler
isometries, while Sp(1) is the previous Sp(1)− and rotates the complex structures.
We will also work with complex coordinates (z, w) on Hn ' C2n writing
(12.1.18) u = z̄ + j w̄ = (x0 + ix1 ) + j(x2 − ix3 ) .
With such conventions we obtain
1X
(12.1.19) ω = iω1 + ω+ j = iω1 + (ω2 + iω3 )j = − du¯j ∧ duj ,
2 j

where
i X³ ´
n n
X
(12.1.20) ω1 = dzj ∧ dz̄j + dwj ∧ dw̄j , ω+ = dwj ∧ dzj .
2 j=1 j=1

Comparing with Example 3.1.11 we recognize (g0 , ω1 ) as the standard Hermitian


metric and Hermitian form on C2n . In addition, the (2, 0)-from ω+ is a complex
symplectic form so that
1 n
ω = dw1 ∧ · · · ∧ dwn ∧ dz1 ∧ · · · ∧ dzn
n! +
is the standard holomorphic volume form on C2n .
Example 12.1.2: Quaternionic projective space. We now use the left H∗ action
on Hn to introduce another model space of quaternionic geometry.
Definition 12.1.3: The quaternionic collectivization
HPn = PH (Hn+1 ) = (Hn+1 \ {0})/H∗
defined with respect to the left action of H∗ on Hn+1 is called the quaternionic
projective n-space.
Let S 4n+3 = {u ∈ Hn+1 | F (u, u) = 1} be the unit sphere in Hn+1 . The group
Sp(n + 1) acts on S 4n+3 transitively with the isotropy at every point Sp(n). Hence,
Sp(n + 1)
(12.1.21) S 4n+3 =
Sp(n)
426 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

is a homogeneous space and the induced metric is of constant sectional curvature


1. Note, that the Sp(1) subgroup of H∗ acts on the sphere and we get the natural
identification
Sp(n + 1
(12.1.22) HPn = S 4n+3 /Sp(1) ' ,
Sp(n) × Sp(1)
so we observe that HPn is actually a compact rank one symmetric spaces. If, in
addition, we make a choice {±1} ⊂ R∗ ⊂ C∗ ⊂ H∗ we can also define three more
projective spaces associated to HPn .
Definition 12.1.4: Let Hn+1 be the quaternionic vector space and HPn the asso-
ciated quaternionic projective space. We define
(i) Z = PC (Hn+1 ) = (Hn+1 \ {0})/C∗ ,
(ii) S = PR (Hn+1 ) = (Hn+1 \ {0})/R∗ ,
(iii) U = PZ2 (Hn+1 ) = (Hn+1 \ {0})/Z2 .
The spaces Z, S, U are called the twistor space, the Konishi bundle, and the
Swann bundle of HPn , respectively.
As homogeneous spaces we have
Sp(n + 1) Sp(n + 1)
Z = CP2n+1 ' , S = RP4n+3 ' .
Sp(n) × U (1) Sp(n) × Z2
Proposition 12.1.5: Let HPn be the quaternionic projective space. We have the
following natural fibre bundles defined by {±1} ⊂ R∗ ⊂ C∗ ⊂ H∗
(i) H∗ /C∗ = S 2 → Z → HPn ,
(ii) H∗ /R∗ = SO(3) → S → HPn ,
(iii) H∗ /Z2 → U → HPn ,
(iv) C∗ /R∗ = S 1 → S → Z ,
(v) C∗ /Z2 → U → Z ,
(vi) R∗ /Z2 = R+ → U → S .
The six bundles of this proposition are the six arrows in the following diagram

(12.1.23) (Hn \ {0})/Z2


o OOO
oo ooo OOO
OOO
oo o OO'
woo
2n−1 o
CP OOO RP4n−1 .
OOO nnn
n
OOO nnn
OO' ² wnnnnn
HPn−1
We shall see later in this chapter that all these bundles exist in a more general
setting. However, the following is a very special property of HPn and has to do
with vanishing of a certain secondary characteristic class, the Marchiafava-Romani
class ² defined in Definition 12.2.1 below. This class clearly vanishes for HPn since
H 2 (HPn , Z) = 0.
Proposition 12.1.6: With the exception of the first one, all the bundles of the
previous proposition admit global Z2 -lifting.
The existence of the bundle Sp(1) → S 4n+3 → HPn means that the structure
group of HPn is can be lifted from Sp(n)Sp(1) to Sp(n) × Sp(1). We will now
12.1. QUATERNIONIC GEOMETRY OF Hn AND HPn 427

construct an atlas on HPn . Consider homogeneous coordinates [u0 , . . . , un ] ∈ HPn .


These are defined in analogy with homogeneous charts on a complex projective
space by the equivalence of non-zero vectors in Hn+1 , with u ' u0 meaning u = u0 λ,
for some λ ∈ H∗ . Let
(12.1.24) Uj = {[u0 , . .¸. , un ] ∈ HPn | uj 6= 0}
and consider the maps φj : Uj → Hn defined by
φj ([u0 , · · · , un ]) = (u0 u−1 −1 −1 −1
j , . . . , uj−1 uj , uj+1 uj , . . . , un uj ) , j = 0, . . . , n.
Now A = {Uj , φj }j=0,...,n is clearly an atlas on HPn giving it a structure of differ-
entiable manifold. Consider the inhomogeneous quaternionic coordinates
(j)
xi = ui u−1
j , i 6= j, j = 0, . . . , n .
on Uj and H-valued 1-forms
(j) (j)
(12.1.25) dxi = (dui − xi duj )u−1
j , i 6= j, j = 0, . . . , n .
(j)
At each x ∈ HPn the forms define an isomorphism Tx HPn ' Hn of quater-
dxi
nionic vector spaces, and thus a local section η (j) ∈ Γ(Uj , L∗ (HPn )) of the principal
coframe bundle L∗ (HPn ) → HPn . Let η (k) ∈ Γ(Uk , L∗ (HPn )) be another such local
section and consider Uk ∩ Uj . An easy computation shows that at any x ∈ Uk ∩ Uj
(k) (j) (k) (j) (j)
(12.1.26) dxi = (dxi − xi dxk )[xk ]−1 , i 6= k .
(j) (j)
Note that by convention xj = 1, dxj = 0. The equations 12.1.26 imply that
pointwise in Uk ∩ Uj
(12.1.27) η (k) = Aη (j) q ,
where q has its values in GL(1, H) and A in GL(n, H) ⊂ GL(4n, R). The group
GL(n, H) × GL(1, H) does not act effectively, but there is an effective action of the
quotient group (GL(n, H) × GL(1, H))/R∗ = GL(n, H)Sp(1). Thus the structure
group of HPn reduces to GL(n, H)Sp(1). We are now ready to give HPn a Rie-
mannian metric which is induced by the flat metric on Hn+1 \ {0}. We can write
the quaternionic Hermitian form in homogeneous coordinates as
4 X j 4 X j j
(12.1.28) g0 − ω1 i1 − ω2 i2 − ω3 i3 = 2
du ⊗ duj − 4 (u du ) ⊗ (duk uk ) .
|u| j |u|
j,k
n
Note that the above equation defines the metric on HP as well as the three local
2-forms {ω1 , ω2 , ω3 } which are local sections of a 3-dimensional vector subbundle
Q ⊂ Λ2 T ∗ HPn . Using the language of H-valued forms we can introduce ω
4 X α 4 X α α
(12.1.29) −ω = 2
du ∧ duα − (u du ) ∧ (duβ uβ ) ,
|u| α |u|4
αβ

with ω = −ω, so that ω is purely imaginary. The constant c is equal to the so-
called quaternionic sectional curvature which generalizes the notion of holomorphic
sectional curvature in complex geometry. The quaternionic Kähler 4-form Ω is then
given by
(12.1.30) Ω = Ω = ω∧ω.
It is real and closed. We have the following
428 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

Theorem 12.1.7: The 4-form Ω is parallel. When n > 1 the holonomy group
Hol(g0 ) ⊂ Sp(n)Sp(1). When n = 1 HP1 ' S 4 and the metric g0 is simply the
metric of constant sectional curvature on S 4 which is self-dual and Einstein.

12.2. Quaternionic Kähler Metrics


Let M be a smooth 4n-dimensional manifold (n ≥ 1). Recall from Example
1.4.18 that M is almost quaternionic if there is a 3-dimensional subbundle Q ⊂
End(T M ) with the property that at each point x ∈ M there is a basis of local
sections {I1 , I2 , I3 } of Q satisfying the quaternion algebra, that is
(12.2.1) Ii ◦ Ij = −δij 1l + ²ijk Ik .
This definition is equivalent to M admitting a G-structure with G = GL(n, H)Sp(1).
Note that any oriented 4-manifold admits such a structure, but in higher dimen-
sions there are obstructions to admitting an almost quaternionic structure as we
now describe.
Suppose now that M carries a Riemannian metric g adapted to the quaternion
structure in the sense that each point on M has a neighborhood such that any local
section I of Q is a local isometry, i.e.,
(12.2.2) g(IX, IY ) = g(X, Y ) ,
for any local vector fields X, Y. Adapted metrics always exists and the resulting
triple (M, Q, g) is called an almost quaternionic Hermitian manifold, giving a fur-
ther reduction of the structure group to Sp(n)Sp(1). Given an adapted metric we
obtain a subbundle Q∗ ⊂ Λ2 T ∗ M which associates to each local section I of Q the
local non-degenerate 2-form ω defined by
(12.2.3) ωI (V, W ) = g(IX, Y ), X, Y ∈ Tx M .
The Sp(n)Sp(1)-structure is a principal Sp(n)Sp(1)-bundle P over M and as
such it can be regarded as an element of the cohomology group H 1 (M, Sp(n)Sp(1))
with coefficients in the sheaf Sp(n)Sp(1) of smooth Sp(n)Sp(1)-valued functions.
The short exact sequence
(12.2.4) 0 → Z2 → Sp(n) × Sp(1) → Sp(n)Sp(1) → 0
gives rise to the homomorphism
(12.2.5) δ : H 1 (M, Sp(n)Sp(1)) → H 2 (M, Z2 ) .
We have
Definition 12.2.1: Let ² = δ(P ) ∈ H 2 (M, Z2 ). Then ² is called the Marchiafava-
Romani class of (M, Q, g).
The Marchiafava-Romani class was introduced in [MR75] and it is the obstruc-
tion to lifting P to the Sp(n) × Sp(1) bundle. When n = 1 the sequence (12.2.4)
becomes
(12.2.6) 0 → Z2 → Spin(4) → SO(4) → 0
and it follows that ² equals the second Stiefel-Whitney class w2 (M ). For n > 1 we
can identify ² = w2 (Q) with the second Stiefel-Whitney class of the vector bundle
Q. Furthermore, we get [Sal82, MR75]
Proposition 12.2.2: Let (M 4n , Q, g) be an almost quaternionic Hermitian mani-
fold. Then w2 (M ) ≡ n²(2).
12.2. QUATERNIONIC KÄHLER METRICS 429

In particular, the Marchiafava-Romani class ² is the second Stiefel-Whitney


class of M if its dimension is 4(8). In complementary dimensions we get
Corollary 12.2.3: Any almost quaternionic manifold M of dimension 0(8) is spin.
Example 12.2.4: Consider the complex projective spaces CP4n of real dimension
8n. The first Chern class c1 (CP4n ) = (4n + 1)γ where γ is a positive generator of
H 2 (CP4n , Z). Since w2 is the mod 2 reduction of c1 , the manifold CP4n is not spin,
and so by Corollary 12.2.3 CP4n cannot admit an almost quaternionic structure.
The full obstruction theory for Sp(n)Sp(1)-structures (even in the 8-dimensional
case) is subtle and not completely understood. See, for example, the article by
Čadek and Vanžura [ČV98] where they prove
Theorem 12.2.5: Let M be a compact oriented 8-manifold. If the conditions

w2 (M ) = w6 (M ) = 0, 4p2 (M ) − p21 (M ) − 8e(M ) = 0, p21 (M ) + 4e(M ) ≡ 0(16)

hold, then M admits an almost quaternionic structure.


Not all these conditions are necessary, however. We know that the vanishing
of w2 is necessary, and Čadek and Vanžura remarked that the middle condition is
necessary. However, they also noticed that Borel and Hirzebruch [BH58] had com-
puted the mod 2 cohomology ring of the quaternionic Kähler manifold G2 /SO(4)
showing that w6 6= 0 which implies that the vanishing of w6 is not necessary.
Suppose {I1 , I2 , I3 } are locally defined smooth sections of Q which satisfy 12.2.1
at each point. Then these form a local orthonormal frame for Q with respect to the
1
standard metric hA, Bi = 2n Tr(At B) on End(T M ). Let {ωi }i=1,2,3 be the basis of
2-forms corresponding under 12.2.3. The associated exterior 4-form
3
X
(12.2.7) Ω= ωi ∧ ωi
i=1

is invariant under a change of frame and thus globally defined on M. It is non-


degenerate in the sense that Ωn is nowhere vanishing on M . The group Sp(n)Sp(1)
is precisely the stabilizer of the form Ω in GL(4n, R). The form Ω is called the
fundamental 4-form of the almost quaternion Hermitian structure (M, Q, Ω, g).
Definition 12.2.6: An almost quaternionic structure (M 4n , Q) with n > 1 is 1-
integrable if M admits a torsion-free connection ∇Q preserving the quaternionic
structure Q. In such a case (M 4n , Q) is called a quaternionic structure on M 4n ,
and if it has an adapted Riemannian metric g, the triple (M 4n , Q, g) is called a
quaternionic Hermitian manifold.
The case of real dimension 4, that is n = 1, is given in Definition 12.2.12 below.
Recall that the connection ∇Q is not unique. The obstruction to 1-integrability
has been studied by Salamon in [Sal86]. In the 4-dimensional case, there is no
obstruction as G = GL(1, H)Sp(1) = R+ × SO(4) so that G-structure is equiva-
lent to a choice of orientation and conformal class. In particular, the Levi-Civita
connection of any compatible metric preserves the G-structure and has no torsion.
But in higher dimensions, there are non-trivial obstructions.
Here we shall be interested in a very special class of quaternionic Hermitian
manifolds, namely the case where a torsion-free quaternionic connection ∇Q is also
a metric connection. In this case it must be the Levi-Civita connection.
430 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

Definition 12.2.7: An almost quaternionic Hermitian manifold (M 4n , Q, Ω, g) of


quaternionic dimension n > 1 is called quaternionic Kähler (QK) if ∇Q coin-
cides with the Levi-Civita connection, or alternatively, if the holonomy group Hol(g)
lies in Sp(n)Sp(1).
We also refer to such a manifold as one with a quaternionic Kähler structure.
We can easily see the holonomy definition to be equivalent to the following
Proposition 12.2.8: An almost quaternionic Hermitian manifold (M 4n , Q, Ω, g),
n > 1, is quaternionic Kähler if ∇Ω = 0, where ∇ denotes the Levi-Civita connec-
tion of g. In particular, an almost quaternionic Hermitian manifold (M 4n , Q, Ω, g),
n > 1, is quaternionic Kähler if it admits a parallel 4-form which is in the same
GL(4n, R)-orbit as Ω at each point x ∈ M.
The hypothesis ∇Ω = 0 clearly implies that dΩ = 0. Surprisingly, the following
theorem was proved by Swann [Swa89]:
Theorem 12.2.9: An almost quaternionic Hermitian (M 4n , Q, Ω, g) of quater-
nionic dimension n > 2 whose fundamental 4-form Ω is closed is quaternionic
Kähler.
The geometry of almost quaternionic Hermitian 8-manifolds is somewhat richer
as there are examples of such spaces for which the fundamental 4-form Ω is closed
but not parallel. Swann showed [Swa91] that
Theorem 12.2.10: An almost quaternionic Hermitian 8-manifold is quaternionic
Kähler if and only if the fundamental 4-form Ω is closed and the algebraic ideal
generated by the subbundle Q∗ ⊂ Λ2 T ∗ M is a differential ideal.
We now investigate some curvature properties of quaternionic Kähler manifolds.
Let {ω1 , ω2 , ω3 } be a local orthonormal frame field for Q∗ ⊂ Λ2 T ∗ M. If Ω is parallel
we get
X3
(∇ωi ) ∧ ωi = 0
i=1
from which it follows that
3
X
(12.2.8) ∇ωi = αij ⊗ ωj ,
j=1

where the αij are 1-forms which satisfy


(12.2.9) αij ≡ −αji ∀ i, j = 1, 2, 3 .
This means in particular that the subspace Γ(Q∗ ) ⊂ Γ(Λ2 T ∗ M ) is preserved by
the Levi-Civita connection. The equations (12.2.8) were considered by Ishihara
[Ish74]. The matrix
 
0 a12 a13
(12.2.10) A = −a12 0 a23 
−a13 −a23 0
is the connection 1-form with respect to the local frame field {ω1 , ω2 , ω3 }. The cur-
vature of this induced connection represents a component of the Riemann curvature
tensor R and is given by
F = dA − A ∧ A .
12.2. QUATERNIONIC KÄHLER METRICS 431

P3
Using the facts that dωi = αij ∧ ωj and d2 ωi = 0, one deduces that
j=1
X X
(12.2.11) Fij = dαij − αil ∧ αlj = λ ²ijk ωk
l k

for some constant λ. Since Q is an oriented 3-dimensional bundle, there is a canon-
ical identification SkewEnd(Q∗ ) ∼
= Q∗ via the cross-product. Using this identifica-
tion we can consider F as a map F : Λ2 T M −→ SkewEnd(Q∗ ) ∼ = Q∗ ⊂ Λ2 T M
and, as such, equation 12.2.11 simply states that F = λπQ∗ , where πQ∗ denotes
pointwise orthogonal projection πQ∗ : Λ2 T M → Q∗ . The full Riemann curvature
tensor R of QK manifold viewed as a symmetric endomorphism R : Λ2 T M −→
Λ2 T M (curvature operator), has the property that
(12.2.12) R |Q∗ = λIdQ∗ ,
where λ is a positive multiple of the scalar curvature s on M .
We will now use Equation 12.2.12 to extend our definition of quaternionic
Kähler manifolds to 4-dimensional spaces. Recall that the problem in dimension
four is that the structure group Sp(1)Sp(1) is isomorphic to the orthogonal group
SO(4) which just describes generic four dimensional oriented Riemannian geometry.
So the problem is caused by a certain low dimensional isomorphism of Lie groups.
Remarkably this same isomorphism of Lie groups provides us with the solution as
well. Now the Lie algebra so(n) and Λ2 (Rn ) are isomorphic as SO(n) modules. So
in dimension four we have a splitting so(4) = su(2) ⊕ su(2) giving rise to a splitting
(12.2.13) Λ2 T M = Λ2+ ⊕ Λ2− ,
where Λ2± are precisely the ± eigenspaces of the Hodge star operator ?. The bundles
Λ2+ and Λ2− are known as the bundles of self-dual and anti-self-dual 2-forms, respec-
tively. Reversing orientation interchanges the self-dual and anti-self-dual 2-forms.
(Note also that in dimension 4, the condition that ∇Ω = 0 is trivially satisfied since
Ω is the volume form). Fixing an orientation and identifying Q∗ with Λ2− we have
Definition 12.2.11: An oriented Riemannian 4-manifold (M, g) is called quater-
nionic Kähler if condition 12.2.12 holds.
With the reverse orientation Q∗ is identified with Λ2+ . Relative to the decom-
position (12.2.13), the curvature operator R can be represented by the matrix
 s

W+ + 12 1l Ric0
(12.2.14) R= ,
s
Ric0 W− + 12 1l
where W± are the self-dual and anti-self-dual Weyl curvatures, Ric0 is the trace-free
part of the Ricci curvature, and s is the scalar curvature. So Equation (12.2.12)
implies that Definition 12.2.11 is equivalent to
Definition 12.2.12: A 4-dimensional oriented Riemannian manifold (M, g) is
quaternionic Kähler if and only if it is self-dual (i.e., W− = 0) or anti-self-
dual (i.e., W+ = 0) and Einstein (i.e., Ric0 = 0). More generally, and oriented
4-manifold (M, g) is quaternionic if W− = 0 or W+ = 0.
Remark 12.2.1: It follows from Definition 4.2.15 that Definitions 12.2.6, 12.2.7
and 12.2.12 work equally well in the case of orbifolds. Thus, it makes perfect sense
to talk about quaternionic or quaternionic Kähler orbifolds. These will play an
important role in Sections 12.4 and 12.5 as well as Chapter 13.
432 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

Remark 12.2.2: Recall that changing the orientation of a quaternionic 4-manifold


interchanges W+ and W− . We shall stick with the more usual convention by saying
that a QK 4-manifold or orbifold with non-zero scalar curvature is self-dual and
Einstein, thus fixing the orientation. This becomes particularly important when
one adds a complex structure, since a complex structure fixes the orientation. The
complex manifold CP2 is self-dual and Einstein, not anti-self-dual; whereas, a K3
surface is anti-self-dual and Einstein. Neither of these manifolds are complex if one
reverses the orientation. Indeed, there are very few compact complex surfaces that
are complex with respect to the reverse orientation [Kot97].
There are at least two more justifications for adopting the Definition 12.2.12.
One is the theory of quaternionic submanifolds of QK manifolds. N ⊂ M is called
a quaternionic submanifold if for each x ∈ N, Tx N is an H∗ -submodule of Tx M .
Marchiafava observed that a 4-dimensional submanifold of a QK manifold is neces-
sarily self-dual and Einstein. The other justification comes in the theory of quater-
nionic Kähler reduction which will be discussed in a later section.
We now return to the general case. Following Salamon [Sal82] we decompose
the Riemannian curvature on a QK manifold in terms of its irreducible pieces
under the group Sp(n)Sp(1). We can write the cotangent bundle as T ∗ M = E ⊗ H
where E and H are locally defined vector bundles on M that transform as the
standard representations of Sp(n) and Sp(1), respectively. Although E and H
are not globally defined bundles their symmetric products S 2 (E), S 2 (H) and anti-
symmetric products Λ2 E, Λ2 H are. Note that S 2 (E) and S 2 (H) transform as
the adjoint representations of Sp(n) and Sp(1), respectively, so the vector bundle
A = S 2 (E) + S 2 (H) transforms as the adjoint representation of Sp(n)Sp(1). So the
(4, 0) Riemannian curvature tensor R can be thought of as a section of
S 2 (A) = S 2 (S 2 (E)) + S(S 2 (E) ⊗ S 2 (H)) + S 2 (S 2 (H))
satisfying the first Bianchi identities.
We already mentioned that the quaternionic projective space is quite special as
it is the only example of a compact quaternionic Kähler manifold which admits an
integrable Sp(n)Sp(1)-structure. The curvature tensor of the canonical symmetric
metric on HPn plays a key role in the more general setting. The following result
which is due to Alekseevsky [Ale68] is presented in the form of Salamon [Sal82]
to which we refer for a proof. This proof can also be found in [Bes87].
Theorem 12.2.13: Let (M 4n , Q, Ω, g) be a quaternionic Kähler manifold. The
Riemann curvature tensor can be written as
R = sR1 + R0 ,
where s is the scalar curvature, R1 is the curvature tensor of quaternionic projective
space HPn , and R0 is a section of S 2 (S 2 (E)).
The section R0 behaves like the curvature tensor of a hyperkähler manifold. In
particular, R0 has zero Ricci curvature and R1 has a traceless Ricci curvature, so
we get the following result due to Berger [Ber66]
Corollary 12.2.14: Any QK manifold is Einstein. A QK manifold with vanish-
ing scalar curvature s is locally hyperkähler, i.e., the restricted holonomy group
Hol0 (g) ⊂ Sp(n).
Theorem 12.2.13 and Corollary 12.2.14 are of fundamental importance to any
further study of quaternionic Kähler manifolds. They imply that QK geometry
12.2. QUATERNIONIC KÄHLER METRICS 433

splits into three cases, positive, negative, and null or the hyperkähler case. We will
discuss some basic properties of hyperkähler metrics in the last four sections of this
chapter, so until then we assume that scalar curvature is not zero. Furthermore, we
almost exclusively discuss the positive QK case, as it is this case that has a strong
connection with the main theme of this book. We end this section by defining three
important bundles that will play an important role in the next chapter.
Definition 12.2.15: Let (M, Q) be an almost quaternionic manifold. Let S(M )
be the SO(3)-principal bundle associated to Q. This principal bundle is called the
Konishi bundle of M . We define the following associated bundles S ×SO(3) F
(i) U(M ) = S(M ) ×SO(3) F, where F = H∗ /Z2 ,
(ii) Z(M ) = S(M ) ×SO(3) F, where F = S 2 is the unit sphere in Q.
The bundles U(M ), Z(M ) are called the Swann bundle, and the twistor space
of M, respectively.
The bundle S(M ) was first described in Konishi [Kon75] for quaternionic
Kähler manifolds, and the bundle U(M ) by Swann [Swa91], again for QK mani-
folds. The twistor space Z(M ) takes its name from Penrose’s twistor theory, cf. the
two volume set [PR87, PR88] and references therein. The twistor space construc-
tion used here has its origins in Penrose’s ‘non-linear graviton’ [Pen76]. It plays
an important role in understanding the geometry of both quaternionic and quater-
nionic Kähler manifolds. Here is why. Since the bundle Z is just the unit sphere in
Q, each point z ∈ Z represents an almost complex structure I(τ ) = τ1 I1 +τ2 I2 +τ3 I3
as in Section 12.1. Thus, a smooth section s of Z over an open set U ⊂ M is an
almost complex structure on U. So we can think of the twistor space as a bundle
of almost complex structures on M. Global sections do not exist generally, so M is
not almost complex. However, if V denotes the vertical subbundle of T Z consisting
of tangent vectors to the fibres F = S 2 , a choice of quaternionic connection ∇Q
determines an equivariant splitting T Z = V ⊕ H. So we obtain an almost com-
plex structure on Z by adding the standard complex structure I0 on S 2 to I(τ )
at each point z = (π(z), τ ) making Z an almost complex manifold. Moreover, the
antipodal map on the fibres induces an anti-holomorphic involution σ : Z−−→Z.
Then the main result concerning twistor spaces is the following theorem due to
Atiyah, Hitchin, and Singer [AHS78] in quaternionic dimension one, and Salamon
[Sal84] for quaternionic dimension greater than one which encodes the quaternionic
geometry of M in the complex geometry of Z.
Theorem 12.2.16: Let M be a quaternionic manifold (orbifold). Then the twistor
space Z(M ) is a complex manifold (orbifold). Moreover, the fibres of π : Z−−→M
are rational curves whose normal bundle is 2nO(1) and Z has a free anti-holomorphic
involution that is the antipodal map on the fibres.
For four dimensional manifolds (quaternionic dimension one) the converse is
true, that is, if the induced almost complex structure on Z is integrable, then
the conformal structure is self-dual (W− = 0). However, in higher dimension the
integrability of Z(M ) only implies the vanishing of a piece of the torsion of ∇Q .
We end this section with a brief discussion of some quaternionic manifolds
in dimension 4. The manifolds S 4 = HP1 , CP2 , K3, T 4 are all well-known to be
quaternionic. In fact, S 4 is self-dual Einstein with one orientation and anti-self-dual
Einstein with the other, CP2 is self-dual Einstein, and K3 is anti-self-dual Einstein
with the standard orientation induced by the complex structure. Of course T 4 ,
434 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

being flat, is self-dual, anti-self-dual and Einstein with either orientation. There
has been much work on self-dual and anti-self-dual structures on 4-manifolds over
the years, and it is not our purpose here to describe what is known. Indeed many
examples of such manifolds as well as orbifolds will make their appearance either
explicitly or implicitly in the present monograph. Usually they occur enjoying some
other property, such as being Einstein, or Kähler. We mention here only some
results of a more general nature. First, there is the existence of self-dual structures
on the connected sums kCP2 for k > 1 [Poo86, DF89, Flo91, LeB91b, PP95,
Joy95] as well as other simply connected 4-manifolds that are neither Einstein nor
complex. Second, Taubes [Tau92] has proven a type of stability theorem that says
that given any smooth oriented 4-manifold M then M #kCP2 admits a self-dual
conformal structure for k large enough.

12.3. Positive Quaternionic Kähler Manifolds and Symmetries


All known complete positive quaternionic Kähler manifolds are symmetric spaces
(see Conjecture 12.3.7 below), and Salamon [Sal82] showed that through dimension
16 any positive QK manifold must have a fairly large isometry group. Moreover,
a bit earlier Alekseevsky proved that all homogeneous positive QK manifolds must
be symmetric [Ale75]. These spaces had been classified by Wolf [Wol65] and they
are often called Wolf spaces. There is precisely one for each simple Lie algebra and
we have
Theorem 12.3.1: Let M be a compact homogeneous positive QK manifold Then
M = G/H is precisely one of the following:
Sp(n + 1) SU (m) SO(k)
, ¡ ¢, ,
Sp(n) × Sp(1) S U (m − 2)×U (2) SO(k − 4)×SO(4)

G2 F4 E6 E7 E8
, , , , .
SO(4) Sp(3)Sp(1) SU (6)Sp(1) Spin(12)Sp(1) E7 Sp(1)
Here n ≥ 0, Sp(0) denotes the trivial group, m ≥ 3, and k ≥ 7. In particular, each
such M is a symmetric space and, there is one-to-one correspondence between the
simple Lie algebras and positive homogeneous QK manifolds.
We remark that the integral cohomology group H 2 (M, Z) vanishes for M =
HP and it is Z for the complex Grassmannian M = Gr2 (Cn+2 ). In all other cases
n

we have H 2 (M, Z) = Z2 . The main results concerning positive QK manifolds are


due to LeBrun and Salamon. Before embarking into a description of their work we
give an infinitesimal rigidity theorem due to LeBrun [LeB88].
Theorem 12.3.2: Let (M, Q, Ω, g) be a compact positive QK manifold. If gt is a
family of positive QK metrics of fixed volume depending smoothly on R such that
g0 = g. Then there exists a family of diffeomorphisms ft : M → M depending
smoothly on t such that gt = ft∗ g.
This result is far from true for hyperkähler or negative QK manifolds. Mod-
uli in the hyperkähler case is well-known, and LeBrun has shown that the moduli
space of complete negative QK metrics on Hn is infinite-dimensional [LeB91c].
LeBrun and Salamon [LS94, LeB93] have strengthen LeBrun’s infinitesimal rigid-
ity Theorem 12.3.2 to a strong rigidity result. They give two theorems, the first is
a finiteness theorem, and second severely restricts the topological type of positive
12.3. POSITIVE QUATERNIONIC KÄHLER MANIFOLDS AND SYMMETRIES 435

QK manifolds. We should also mention that the rigidity results given below break
down entirely in the case of compact positive QK orbifolds.
Theorem 12.3.3: For each positive integer n there are up to isometries and rescal-
ings only finitely many compact positive quaternionic Kähler manifolds of dimen-
sion 4n.
outline of Proof. The proof of this theorem relies heavily on Mori theory
applied to the twistor space Z(M ). In the positive QK case Theorem 12.2.16 was
strengthened in [Sal82]. We state this as a lemma together with another result
from [LS94].
Lemma 12.3.4: Let M 4n+3 be a positive quaternionic Kähler manifold. Then its
twistor space Z(M ) is a Fano manifold with a positive Kähler-Einstein metric and
a complex contact structure. Moreover, two positive quaternionic Kähler structures
are homothetic if and only if their twistor spaces are biholomorphic.
Since Z(M ) has a complex contact structure c1 (M ) is divisible by n + 1. Then by
contracting extremal rays Wiśniewski [Wiś91] shows that b2 (Z(M )) = 1 with the
exception of three cases. Only one of these cases admits a complex contact struc-
ture, namely the flag variety P(T ∗ CPn+1 ) which is the twistor space of the complex
Grassmannian Gr2 (Cn+2 ). So we can conclude from Lemma 12.3.4 that b2 (M ) = 0
unless M = Gr2 (Cn+2 ) with its symmetric space metric. In the remaining cases we
are dealing with Fano manifolds with Picard number one. We can make use of the
rational connectedness theorem which imply that there are only a finite number of
deformation types of smooth Fano varieties. This was proven for Picard number one
in [Cam91, Nad91], and more generally without the Picard number restriction
in [Cam92, KMM92]. Then in line with Theorem 12.3.2 LeBrun and Salamon
show that there is only a finite number of Fano contact manifolds up to biholo-
morphism. We refer to [LS94, LeB93] for details. Then the second statement of
Lemma 12.3.4 implies that there are only a finite number of QK manifolds up to
homotheties. ¤
This proof says a lot about the topology of compact positive QK manifolds.
Here we collect the results of [LS94] together with what was known earlier about
the topology of positive QK manifolds.
Theorem 12.3.5: Let M be a compact positive QK manifold. Then
(i) π1 (M ) = 0;


0 if M is isometric to HPn ,
(ii) π2 (M ) = Z if M is isometric to Gr2 (Cn+2 ),


finite containing Z2 otherwise;
(iii) b2k+1 (M ) = 0 for all k ≥ 0;
(iv) b4i (M ) > 0 for 0 ≤ i ≤ n;
(v) b2i (M ) − b2i−4 (M ) ≥ 0 for 2 ≤ i ≤ n;
Pn−1 1
(vi) r=0 [6r(n − 1 − r) − (n − 1)(n − 3)]b2r (M ) = 2 n(n − 1)b2n (M ).

Proof. To prove (i) we see that by Lemma 12.3.4 Z(M ) is Fano so by Theorem
3.6.9 it is simply connected. Then (i) follows by the long exact homotopy sequence
applied to the fibration S 2 −−→Z(M )−−→M. For (ii) we recall in the proof of Theo-
rem 12.3.3 b2 (M ) = 0 unless M = Gr2 (Cn+2 ) in which case π2 (M ) = Z. But this to-
gether with (i) and universal coefficients imply that if b2 (M ) = 0 then H 2 (M, Z2 ) is
436 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

the 2-torsion of H2 (M, Z), and this is non-vanishing when the Marchiafava-Romani
class ² is non-vanishing. So the remainder of (ii) then follows from the following
result of Salamon [Sal82]:
Lemma 12.3.6: Let (M, Q, g) be a compact positive QK manifold with vanishing
Marchiafava-Romani class ². Then (M, g) ' (HPn , g0 ) with its canonical symmetric
metric g0 .
Part (iii) which is due to Salamon [Sal82] is a consequence of the fact that the
twistor space Z(M ) of a QK manifold has only (p, p)-type cohomology which in
turn is a consequence of the Kodaira-Nakano Vanishing Theorem 3.5.8 and a gen-
eralization due to Akizuki and Nagano [AN54]. Part (iv) follows immediately from
the non-degeneracy of the closed 4-form Ω of Equation (12.2.7). Part (v) follows
from the quaternionic version of the Lefschetz theory [Kra65, Bon82] as described
for example in Section 3.3. (vi) follows from index theory computations which we
refer to Section 5 of [LS94]. ¤

Remarks 12.3.1: Parts (iv) and (v) hold for any compact QK manifold, not only
positive ones. However, in the positive case we shall see in Proposition 13.5.5
below that the numbers β2i = b2i − b2i−4 are precisely the even Betti numbers of
the principal SO(3)-bundle associated to the quaternionic bundle Q.
In the absence of any counterexamples, Theorems 12.3.3 and 12.3.5 strongly
points towards
Conjecture 12.3.7: All compact positive QK manifolds are symmetric.
The above conjecture was first formulated in [LS94] and we will refer to it
as the LeBrun-Salamon Conjecture. Beyond Theorems 12.3.3 and 12.3.5 there are
several other results showing the conjecture to be true in some special cases. We
shall collect all these results in the following
Theorem 12.3.8: Let (M 4n , Q, Ω, g) be a compact positive QK manifold. Then
M is a symmetric space if
(i) n ≤ 3,
(ii) n = 4 and b4 = 1.

Proof. The statement in (i) dates back to the Hitchin’s proof that all compact
self-dual and Einstein manifolds of positive scalar curvature must be isometric to
either S 4 with the standard constant curvature metric or CP2 with the Fubini-
Study metric [Hit81] (see also [FK82, Bes87]). For n = 2 the result was proved
by Poon and Salamon [PS91]. The proof was greatly simplified in [LS94] using
the rigidity results of Theorem 12.3.5. The n = 3 case is a recent result of Herrera
and Herrera [HH02a, HH02b]. Their proof uses an old result which estimates
the size of the isometry group of M in lower dimension [Sal82]. The dimension
of the isometry group of a positive QK 12-manifold must be at least 6 and the
dimension of the isometry group of a positive QK 16-manifold must be at least 8.
In particular, when n = 3 the manifold M admits an isometric circle action. Using
some deep results concerning Â(M ) genus of non-spin manifolds with finite π2 (M )
and smooth circle actions Herrera and Herrera prove that
Lemma 12.3.9: Let M be a positive QK 12-manifold which is not Gr2 (C5 ). Then
Â(M ) = 0.
12.3. POSITIVE QUATERNIONIC KÄHLER MANIFOLDS AND SYMMETRIES 437

The result is then a consequence of the vanishing of Â(M ) and Theorem 12.3.5.
The result in (ii) follows from the estimate on the dimension of the isometry group
in this case and Betti number constraints of Theorem 12.3.5 [GS96]. ¤

Remark 12.3.1: The argument of [HH02a] does not work in 16-dimensional case
because all QK manifolds of quaternionic dimension 4 are automatically spin. Nev-
ertheless, the estimate on the size of the isometry group together with all the known
results can most likely be used to construct a proof of the LeBrun-Salamon Conjec-
ture in this case. However, as pointed out by Salamon in [Sal99], the biggest gap
in any potential geometric proof of this conjecture is the conundrum of whether a
QK manifold of quaternionic dimension n > 4 has any non-trivial Killing vector
fields.
There is another approach to the LeBrun-Salamon Conjecture which proceeds
via the algebraic geometry of the twistor space Z(M ) and uses Lemma 12.3.4.
The following, apparently stronger, conjecture was suggested by Beauville [Bea05,
Bea98]
Conjecture 12.3.10: Any compact Fano manifold with a complex contact struc-
ture is homogeneous.
This, of course implies the LeBrun-Salamon Conjecture. Several years ago there
were some attempts to use algebraic geometry to prove this result. Wiśniewski
even briefly claimed the proof of the conjecture but later Campana found a gap in
Wiśniewski’s argument. Campana briefly claimed to have bridged that gap but later
also withdrew the claim. Hence, as of the time of writing this monograph, both the
Beauville Conjecture 12.3.10 and LeBrun-Salamon Conjecture 12.3.7 remain open.
Let rk(M ) be the symmetry rank of M defined as the rank of its isometry group
Isom(M, g), i.e., the dimension of the maximal Abelian subgroup in Isom(M, g).
Bielawski [Bie99] proved that a positive QK manifold of quaternionic dimension
n with rk(M ) ≥ n + 1 is isometric to HPn or to the Grassmannian Gr2 (Cn+2 ).
Recently Fang proved several rigidity theorems for positive quaternionic Kähler
manifolds in terms its symmetry rank [Fan04]. Fang’s result slightly enhances
Bielawski’s theorem.
Theorem 12.3.11: Let (M 4n , Q, Ω, g) be a compact positive quaternionic Kähler
manifold. Then the isometry group Isom(M, g) has rank at most (n + 1), and M
is isometric to HPn or Gr2 (Cn+2 ) if rk(M ) ≥ n − 2 and n ≥ 10.
This theorem is quite interesting and apparently rather deep. It follows from
several different results. First recall that a quaternionic submanifold is one that pre-
serves the quaternionic structure. It is a well-known result of Gray that [Gra69a]
Proposition 12.3.12: Any quaternionic submanifold in a quaternionic Kähler
manifold is totally geodesic and quaternionic Kähler.
In [Fan04] Fang proves the following rigidity results for positive QK manifold
Theorem 12.3.13: Let (M 4n , Q, Ω, g) be a positive QK manifold. Assume f =
(f1 , f2 ) : N → M × M, where N = N1 × N2 and fi : Ni → M are quaternionic
immersions of compact quaternionic Kähler manifolds of dimensions 4ni , i = 1, 2.
Let ∆ be the diagonal of M × M and set m = n1 + n2 . Then
(i) If m ≥ n, then f −1 (∆) is non-empty.
(ii) If m ≥ n + 1, then f −1 (∆) is connected.
438 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

(iii) If f is an embedding, then for i ≤ m − n there is a natural isomorphism,


πi (N1 , N1 ∩ N2 ) → πi (M, N2 ) and a surjection for i = m − n + 1.
Remark 12.3.2: The study of homogeneous negative QK manifold is more deli-
cate. There are of course the non-compact duals of the Wolf spaces. Alekseevsky
showed that there are also non-symmetric homogeneous examples. He obtained
a classification of such spaces under the assumption that the symmetry group is
completely solvable[Ale75]. We will not describe these spaces here referring the
interested reader to an extensive review on this subject by Cortes [Cor00]. All
these are typically called Alekseevskian spaces. Much later de Wit and Van Proyen
discovered a gap in Alekseevsky’s classification [dWVP92] while considering some
supersymmetric σ-model coupled to supergravity. They filled in the gap and also
claimed that there should be no other homogeneous examples. Inspired by this
work Cortes [Cor96] provided a Lie algebraic proof filling the gap in Alekseevsky’s
original paper. A proof that all negative QK manifolds are the known Alekseevskian
spaces is, however, still lacking.

12.4. Quaternionic Kähler Reduction


In this section we describe an of the symmetry reduction method described in
Section 8.4 for QK manifolds. Just as in the case of hyperkähler quotients which
are introduced later in Section 12.8, the ideas originated in physics of supersym-
metric field theories. In 1983 Witten and Bagger observed that matter coupled
to 4-dimensional supergravity theory with N = 2 supersymmetries requires the
scalar fields of the coupling to be local coordinates on a negative scalar curvature
QK manifold [BW83]. Later more general Lagrangians of such theories were con-
structed and studied. In particular, some elements of the symmetry reduction can
be found in [dWLP+ 84, dWLVP85]. However, mathematical formulation of the
theory of quaternionic Kähler quotients and its application to the case of positive
QK manifolds was developed later in [Gal87a, GL88].
To begin we consider the spaces Γp (Q∗ ) ≡ Γ(Λp T ∗ M ⊗ Q∗ ) of smooth exterior
p-forms on M with values in the bundle Q∗ . The connection given on Q∗ induces
a “de Rham” sequence
d∇ =∇ d∇ d∇
(12.4.1) Γ0 (Q∗ ) −→ Γ1 (Q∗ ) −→ Γ2 (Q∗ ) −→ · · ·
such that
(12.4.2) d∇ ◦ d∇ (f ) = R(f )
for f ∈ Γ0 (Q∗ ).
Consider now the Lie group
Aut(M, Q, Ω, g) ≡ {g ∈ Isom(M, g) : g ∗ Ω = Ω}
and its Lie algebra
aut(M, Q, Ω, g) ≡ {V ∈ isom(M, g) : £V Ω = 0}
which is a Lie subalgebra of the Lie algebra isom(M, g) of Killing vector fields on
M. We have the following immediate consequence
Proposition 12.4.1: Let (M, Q, Ω, g) be a QK manifold of non-zero scalar cur-
vature. Then aut(M 4n , Q, Ω, g) = isom(M 4n , g). It follows that any one parameter
subgroup H ⊂ Isom(M 4n , g) is also a subgroup of Aut(M 4n , Q, Ω, g).
12.4. QUATERNIONIC KÄHLER REDUCTION 439

Proof. When M is symmetric all statements follow by inspection. When M


is not locally symmetric the holonomy Lie algebra hol = sp(n) ⊕ sp(1). Since M is
irreducible, by a theorem of Kostant [Kos55] any Killing vector field normalizes the
holonomy algebra and in particular the sp(1)-factor which defines the quaternionic
structure Q. Hence isom(M, g) normalizes Q and therefore any Killing vector field
V preserves Ω. The rest follows from the fact that both groups are compact Lie
groups. ¤

The full isometry group may contain discrete isometries which do not lie on
any one-parameter subgroup and these may not preserve the quaternionic 4-from
Ω. To each V ∈ isom(M, g) we associate the Q∗ -valued 1-form
ΘV ∈ Γ1 (Q∗ )
defined in terms of a local frame ω1 , ω2 , ω3 by
X
(12.4.3) ΘV ≡ (V ωi ) ⊗ ωi .
i

Clearly ΘV remains invariant under local change of frame field (i.e., under local
gauge transformations). We have [GL88]
Theorem 12.4.2: Assume that the scalar curvature of (M, Q, Ω, g) is not zero.
Then to each V ∈ isom(M, g) there corresponds a unique section µ ∈ Γ0 (Q∗ ) such
that
(12.4.4) ∇µ = ΘV .
In fact, under the canonical bundle isometry σ : SkewEnd(Q∗ ) −→ Q∗ , µ is given
explicitly by the formula
1
(12.4.5) µ= σ(£V − ∇V ) ,
λ
where λ is the constant positive multiple of the scalar curvature defined by 12.2.12.
We observe now that by the uniqueness in Theorem 12.4.2, the map V −→
µ transforms naturally under the group of automorphisms, specifically for g ∈
Aut(M, Q, Ω, g) and V ∈ isom(M, g) we have
(12.4.6) µg∗ (V ) = g∗ (µ) ,
¡ ¢ ¡ ¢
where g∗ µ (x) = g̃ µ(g −1 (x)) and where g̃ denotes the map induced by g on the
bundle Q∗ ⊂ Λ2 T M . Note also that g∗ V = Adg (V ). Hence, 12.4.6 means that the
diagram

Adg
isom(M,
 g) −−−−→ isom(M,
 g)
 
(12.4.7)  
µ µ
y y
g∗
Γ0 (Q∗ ) −−−−→ Γ0 (Q∗ ) ,
commutes.
Suppose now that G ⊂ Isom(M, g) is a compact connected Lie subgroup with
corresponding Lie algebra g.
440 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

Definition 12.4.3: The moment map associated to G is the section µ of the


bundle g∗ ⊗ Q∗ ∼ = Hom(g, Q∗ ) whose value at a point x is the homomorphism
V −→µ(x).
From the equivariance above we see immediately that the moment map is G-
equivariant. Since the action of G in the bundle g∗ ⊗ Q∗ is linear on the fibers, it
preserves the zero section. Consequently the set

(12.4.8) N = µ−1 (0) = {x ∈ M : µ(x) = 0}

is G-invariant. We have the following reduction theorem due to Galicki and Lawson
[GL88].
Theorem 12.4.4: Let (M, Q, Ω, g) be a quaternionic Kähler manifold with non-
zero scalar curvature. Let G ⊂ Isom(M, g) be a compact connected subgroup with
moment map µ. Let N0 denote the G-invariant subset of N = {x ∈ M : µ(x) = 0},
where µ intersects the zero section transversally and where G acts locally freely.
Then M̂ = N0 /G is a quaternionic Kähler orbifold.
The statement here is a slight generalization of the main theorem in [GL88].
The proof of the theorem proceeds along the lines discussed in [GL88] with the
exception that the locally free action allows us to use Molino’s Theorem 2.5.11 to
conclude that the quotient has an orbifold structure. We rephrase an important
special case of Theorem 12.4.4 as:
Corollary 12.4.5: Let M be as above and suppose G = ∼ T k ⊂ Isom(M, g) is a
k-torus subgroup generated by a vector fields Vk ∈ isom(M, g). If V1 ∧ . . . ∧ Vk is
a k-plane field at all points x ∈ N, then N/G is a compact quaternionic Kähler
orbifold.
Example 12.4.6: Consider the S 1 -action defined on HPn in homogeneous coordi-
nates as follows
ϕλ ([u0 , . . . , un ]) = [λu0 , . . . , λun ] ,

where λ is a complex unit. Recall that the projectivization is by right multiplication


u 7→ uq as described in Section 12.1. In a local trivialization we can identify the
local frame ω1 , ω2 , ω3 with the imaginary quaternions i, j, k, respectively, in which
case the moment map µ for the S 1 -action can be written as
n
X
(12.4.9) µ = ūiu = ūj iuj .
j=0

So the zero set µ−1 (0) is invariant under right multiplication by H∗ and so cuts out a
real codimension 3 subvariety of HPn . Moreover, one easily checks that it is a smooth
embedded submanifold that is invariant under the left action of U (n+1) by u 7→ Au.
The isotropy subgroup of the point [u] = [1, j, 0, · · · , 0] ∈ µ−1 (0)
¡ is identified with
¢
SU (2)×U (n−1). So µ−1 (0) is the homogeneous space U (n+1)/ SU (2)×U (n−1) ,
which identifies the quotient µ−1 (0)/S 1 with the complex Grassmannian M̂ =
Gr2 (Cn+1 ).
This can easily be generalized to the case of a weighted circle action

ϕλ ([u0 , . . . , un ]) = [λp0 u0 , . . . , λpn un ] ,


12.4. QUATERNIONIC KÄHLER REDUCTION 441

in which case the moment map becomes


n
X
(12.4.10) µ= pj ūj iuj .
j=0

In this case the quotients µ−1 (0)/S 1 are quaternionic Kähler orbifolds. The orbifold
stratification is analyzed for the special case p0 = q and p1 = · · · pn = p in [GL88].
Theorem 12.4.4 can be used to obtain many examples of compact QK orbifolds.
When the reduced space is 4-dimensional it is automatically self-dual and Einstein.
The only complete positive QK manifolds in dimension 4 are HP1 ' S 4 or CP2
with their standard symmetric space metrics. In this context the more interesting
quotients are those with orbifold singularities. They will be discussed in the next
section.
Just as in the symplectic case one can study more singular QK quotients with-
out assuming that the action of the quotient group G on N = µ−1 (0) is locally
free. A detailed study of this more general situation was done by Dancer and Swann
[DS97a]. Let (M, Q, Ω, g) be a QK manifold and let G be a connected Lie group
acting smoothly and properly on M preserving QK structure with moment map µ.
For any subgroup H ⊂ G we denote by M H the set of points in M fixed by H and
MH the set of points whose isotropy subgroups are exactly H. Further we write
M(H) for the set of points whose isotropy subgroups are conjugate to H in G. Now,
if MH is not empty then H must be compact. It follows that both MH ⊂ M H ⊂ M
are smooth manifolds. If N (H) is the normalizer of H in G then L = N (H)/H acts
freely and properly on MH with the quotient MH /L = M(H) /G a smooth manifold.
Thus M decomposes into the union of M(H) /G, where (H) runs over all conjugacy
classes of stabilizers. We define
µ−1 (0) ∩ M(H) µ−1 (0) ∩ MH
(12.4.11) M̂H = = .
G L
One can first show that the stratification of M by orbit types induces the stratifi-
cation of M̂ into a union of smooth manifolds, i.e.,
Theorem 12.4.7: Let H ⊂ G be a subgroup so that MH is not-empty. Then M̂H
is a smooth manifold.
However, not all the pieces M̂H have QK structures. Their geometry depends
on the way H acts on Q. To be more precise, let x ∈ MH and consider the differential
action of H on Tx M. Since H acts preserving the quaternionic structure we have the
representation H−→Sp(n)Sp(1) which induces the representation φ : H−→SO(3).
The group H acts on Q ' R3 via the composition of φ with the standard three-
dimensional representation. If x, y ∈ MH are on the same path-component then
parallel transport along any path joining x to y defines an H-equivariant isomor-
phism Tx M ' Ty M. It follows that the representation φ : H−→SO(3) is equivalent
at all points on a path component of MH . Hence, the image φ(H) < SO(3), up to
isomorphism, is the same on L-orbits. There are four possibilities [DS97a]:
• φ(H) is trivial ⇒ M̂H is a QK manifold,
• φ(H) = Zk , k > 1 ⇒ each path component of M̂H is covered by a Kähler
manifold,
• φ(H) is finite but not cyclic ⇒ MH is totally real in M (M̂H is “real”),
• φ(H) = SO(2) or SO(3) ⇒ M̂H is empty.
442 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

Even simple examples show that the stratification of the quotient by orbit type can
include all of the three possible pieces. However, there is a coarser stratification of
M̂ in which all pieces are in fact QK. Let M[H] be the set of points in M, where the
identity component of the stabilizer equals H and M([H]) the set of points whose
stabilizer has identity component conjugate to H in G.
Theorem 12.4.8: The union
[ (µ−1 (0) ∩ M([H]) )
M̂ =
G
H⊂G

taken over all compact connected subgroups H ⊂ G induces a canonical decomposi-


tion of M̂ into a union of QK orbifolds each of which is a finite quotient of a QK
manifold.

Proof. The key observation is that M[H] is an open submanifold in M H ,


where M H is a smooth QK submanifold of M. Hence, itself M[H] is a smooth QK
manifold. The restriction of µ to M[H] is the moment map for the locally free
action of L, hence (µ−1 (0) ∩ M[H] )/L = (µ−1 (0) ∩ M([H]) )/G is a QK orbifold by
Theorem 12.4.4. ¤

We finish this chapter with a brief discussion of Morse theory on QK manifolds.


The idea to consider f = ||µ||2 as the Morse function is quite natural as suggested in
analogy with Kirwan’s work on symplectic quotients [Kir84, Kir98]. The function
f was first introduced by Battaglia in [Bat96b, Bat99] and more recently also in
[ACDVP03]. The motivation behind [ACDVP03] was the fact that the so-called
BPS sates in 5-dimensional supergravity theory correspond to gradient flows on a
product M × N, where M is a negative QK manifold and N is a special Kähler
space. Such flows are generated by certain “energy function” f which is nothing
but the square of the moment map f = ||µ||2 . Battaglia was interested mostly in
the positive QK case and, it appears the authors of [ACDVP03] were not aware
of her work. We now describe some of the Battaglia’s results. Recall that a Morse
function f is called equivariantly perfect over Q if the equivariant Morse equalities
hold, that is if
X
P̂t (M ) = P̂t (µ−1 (0)) + tλF P̂t (F ) ,
where the sum ranges over the set of connected components of the fixed point
set, λF is the index of F, and P̂t is the equivariant Poincaré polynomial for the
equivariant cohomology with coefficients in Q. Battaglia proves
Theorem 12.4.9: Let (M 4n , Q, Ω, g) be a positive QK manifold acted on isomet-
rically by S 1 . Then the non-degenerate Morse function f = kµk2 is equivariantly
perfect over Q. The critical set of f is the union of the zero set f −1 (0) = µ−1 (0)
and the fixed point set of the circle action.
Moreover, the zero set µ−1 (0) is connected, and a fixed point component is
either contained in µ−1 (0) or does not intersect with µ−1 (0).
Proposition 12.4.10: Let M 4n be a positive QK manifold acted on isometrically
by S 1 . Then every connected component of the fixed point set, not contained in
µ−1 (0), is a Kähler submanifold of M \ µ−1 (0) of real dimension less than or equal
to 2n whose Morse index is at least 2n, with respect to the function f.
12.5. COMPACT QUATERNIONIC KÄHLER ORBIFOLDS 443

In [Bat99] Battaglia uses Morse theory to improve the results obtained earlier
in [Bat96b]. She shows that the quotient in Example 12.4.6 is unique in the
following sense.
Theorem 12.4.11: Let (M 4n , Q, Ω, g) be a positive QK manifold acted on isomet-
rically by S 1 . Suppose S 1 acts freely on N = µ−1 (0). Then M 4n is homotopic to
HPn with the quotient M̂ = Gr2 (Cn+1 ).

12.5. Compact Quaternionic Kähler Orbifolds


As already indicated the method of QK reduction enjoys much success if one
allows the quotient M of be a QK orbifold. The price that is paid is the loss of the
rigidity described in Section 12.3. Perhaps a more interesting observation about
such a generalization is that when M has orbifold singularities the total space of
the orbifold Konishi bundle S(M ), which is a principal orbibundle with structure
group G = SO(3) or G = Sp(1), may actually be smooth. That should not come
as a surprise to the reader familiar with earlier chapters of our book. In fact,
this happens exactly (just as it does in the case of orbifold circle V -bundles) when
the orbifold uniformizing groups are subgroups of the structure group G. In this
section we introduce some examples of compact positive QK orbifolds and discuss
some obvious classification problems.
The first examples of positive QK orbifolds were introduced in 1987 by Galicki
and Lawson [Gal87a, GL88]. We will briefly describe the construction slightly
generalizing the original example. The key to the construction is Corollary 12.4.5
of the previous section. Consider (M, g) = (HPn , gcan ) and an arbitrary reduction
of M be a k-dimensional Abelian subgroup of the isometry group. Such a reduction
is associated to a choice
(12.5.1) H = T k ⊂ Tmax = T n+1 ⊂ U (n + 1) ⊂ Sp(n + 1) ,
where Sp(n + 1) = Isom(HPn , gcan ) and Tmax = T n+1 is the maximal torus sub-
group. One can always choose T n+1 to be the set of diagonal matrices in the
unitary group U (n + 1). Any rational subtorus H is then determined by a collec-
tion of a non-zero integer vectors {θ 1 , . . . , θ n+1 } generating Rn+1−k . These can be
put together as a matrix Θ ∈ Mn+1−k,n+1 (Z). Dually, we can consider a matrix
Ω ∈ Mk,n+1 (Z) whose column vectors {ω 1 , . . . , ω n+1 } generate Rk . This gives the
exact sequence of Lie algebras
ω θ
(12.5.2) 0 −−→ h −−→ Rn+1 −−→ Rn+1−k −−→ 0 ,
and its dual
θ∗ Γ∗
(12.5.3) 0 −−→ Rn+1−k −−→ Rn+1 −−→ h∗ −−→ 0 ,
where θ(ei ) = Θei = θ i ∈ Rn+1−k and Γ∗ (ei ) = Ωei = ω i ∈ Rk with {e1 , . . . , en+1 }
being the standard basis in Rn+1 . There is corresponding exact sequence at the
group level 1−→H− →T n+1 − →T n+1−k −
→1 and the subtorus H is identified with the
image of the homomorphism fΩ : T k → T n+1
³Y
k k
Y aj
´
aj
(12.5.4) fΩ (τ1 , . . . , τk ) = diag τj 1 , . . . , τj n+1 ,
j=1 j=1
444 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

where (aji ) = Ω. Now, with each Ω we associate H acting on HPn , the QK moment
map µΩ , and the zero level set N (Ω) = µ−1 n
Ω (0) ⊂ HP . It is elementary to check
when the condition of Corollary 12.4.5 is satisfied. We have arrived at
Theorem 12.5.1: Suppose all k × k minor determinants of Ω do not vanish, i.e.,
any collection of k column vectors of Ω are linearly independent. Then the reduced
space O(Ω) = M̂ (Ω) is a compact positive QK orbifold. Furthermore, the Lie
algebra isom(O(Ω), ĝ(Ω)) contains (n + 1 − k) commuting Killing vector fields. In
particular, when k = n − 1, O(Ω) is a compact self-dual Einstein orbifold of positive
scalar curvature and 2 commuting Killing vector fields.
The rational cohomology of these orbifolds were computed independently in
[BGMR98] and [Bie97]. In particular, we have
Corollary 12.5.2: There exist compact toric positive self-dual Einstein orbifolds
with arbitrary second Betti number.
Remark 12.5.1: The case originally considered in [Gal87a, GL88] corresponds
to k = 1 and Ω = (q, p, . . . , p) with Ω = (1, . . . , 1) being the canonical quotient of
Example 12.4.6. More general cases were analyzed only much later in [BGM94a]
(Ω = (p1 , . . . , pn+1 )) and [BGMR98] (an arbitrary Ω), where it was realized that
the Konishi orbibundle of such orbifolds can often be a smooth manifold carrying
a natural Einstein metric. We shall return to a detailed analysis of these examples
in the next chapter after we define 3-Sasakian manifolds.
We now specialize to the case of compact 4-dimensional QK orbifolds (M, g),
i.e., 4-orbifolds with self-dual conformal structure with an Einstein metric g of
positive scalar curvature. Recall that when M is smooth it must be isomorphic to
S 4 or CP2 . On the other hand, Theorem 12.5.1 alone provides plenty of examples
of such spaces [GL88]. As orbifolds, some of them are the familiar examples of
weighted projective spaces introduced in Chapter 4.
Proposition 12.5.3: Let O(p) be O(Ω) of Theorem 12.5.1 with Ω = (p1 , p2 , p3 ) =
p, i.e., O(p) is a QK reduction of HP2 by the isometric circle action with weights
p. In addition assume that all pi ’s are positive integers such that gcd(p1 , p2 , p3 ) = 1.
Then
(i) there is smooth orbifold equivalence
(
CP2p1 +p2 p2 +p3 p3 +p1 , when pi is odd for all i,
O(p) ' 2 , 2 , 2
CP2p1 +p2 ,p2 +p3 ,p3 +p1 , otherwise;
(ii) the metrics g(p) defined by QK reduction are inhomogeneous unless p =
(1, 1, 1) in which case we get the Fubini-Study metric on CP2 ;
(iii) the QK metrics g(p) are Hermitian with respect to the standard complex
structure on the corresponding weighted projective space.
Proof. To proof (i) one needs to identify the level set of the moment map
with S 5 which is easily done. We refer to Section 13.7.4, where it is shown that
the level set of the 3-Sasakian moment map is diffeomorphic to the Stiefel manifold
V2 (C3 ) = U (3)/U (1). Now, in terms of the QK quotient, the level set must be
U (3)/U (2) ' S 5 . The result follows by observing that the circle action on v = z̄×w
(which can be though as a coordinate on the 5-sphere) has weights (p2 + p3 , p3 +
p1 , p1 + p2 ). (ii) can be proved by using the relation between the local form of any
positive toric metric given in Theorem 12.5.5 and QK toric quotients, This relation
12.5. COMPACT QUATERNIONIC KÄHLER ORBIFOLDS 445

was established in [CP02]. In particular, it easily follows that the metric is of


cohomogeneity two for all distinct weights, and of cohomogeneity one if exactly two
weights are equal. The case of equal weight gives the Fubini-Study metric which
is symmetric. Finally (iii) follows from the result of Apostolov and Gauduchon
[AG02].
¤
The above orbifolds are also quite interesting for another reason. In addition to
being self-dual and Einstein the metric g(p) is often of positive sectional curvature.
This curvature property of O(p) was discovered by Dearricott [Dea04, Dea05] and
by Blažić and Vukmirović [BV04]. First we have the following result Dearricott
Theorem 12.5.4: Let O(p) be the Galicki-Lawson orbifold with p1 ≤ p2 ≤ p3 .
Then the self-dual Einstein metric g(p) is of positive sectional curvature if and
only if
(12.5.5) σ3 (p1 + p2 + p3 , −p1 − p2 + p3 , −p1 + p2 − p3 , p1 − p2 − p3 ) > 4p33 ,
where σ3 denotes the third symmetric polynomial in 4 variables.
The paper of Blažić and Vukmirović uses quite different methods. In fact
their main theorem is a generalization of the Galicki-Lawson examples to the case
of pseudo-Riemannian metrics of split signature (+, +, −, −), where the quotient
construction involves paraquaternions. However, the curvature calculations apply
to the Riemannian case as well. For the U (2)-symmetric orbifolds O(p) = O(p, q, q)
case they calculate the pinching constants and get the following result.
Theorem 12.5.5: The self-dual √ Einstein√ metric on O(p) = O(p, q, q) has positive
sectional curvature if p2 < q 2 2 < 2p2 2 and then at every point x ∈ O(p, q, q)
the sectional curvature is k-pinched with 0 < k < 1 and
1 3 q 2 − p2 1 3 q 4 − p4
− ≤ k ≤ + , p ≤ q,
4 4 q 2 + p2 4 4 q 4 + p4
1 3 p4 − q 4 1 3 p2 − q 2
− ≤ k ≤ − , q ≤ p.
4 4 p4 + q 4 4 4 p2 + q 2
Clearly, k → 41 when p → q. Furthermore, k = 14 if and only if p = q = 1 in which
case O ' CP2 is symmetric.
When Ω ∈ Mn−1,n+1 (Z) the orbifold structure of O(Ω) is more involved. How-
ever, as each O(Ω) has two commuting Killing vectors, locally these metrics are
described by the following results of Calderbank and Pedersen [CP02]
Theorem 12.5.6: Let F (ρ, η) be a solution of the linear differential equation
3F
Fρρ + Fηη = 2

on some open subset of the half-space ρ > 0, and consider the metric g(ρ, η, φ, ψ)
given by
F 2 − 4ρ2 (Fρ2 + Fη2 ) dρ2 + dη 2
g=
4F 2 ρ2
(12.5.6) ¡ ¢2 ¡ ¢2
(F − 2ρFρ )α − 2ρFη β + −2ρFη α + (F + 2ρFρ )β
+ ¡ ¢ ,
F 2 F 2 − 4ρ2 (Fρ2 + Fη2 )
√ √
where α = ρ dφ and β = (dψ + η dφ)/ ρ. Then
446 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

(i) On the open set where F 2 > 4ρ2 (Fρ2 + Fη2 ), g is a self-dual Einstein
metric of positive scalar curvature, whereas on the open set where 0 <
F 2 < 4ρ2 (Fρ2 + Fη2 ), −g is a self-dual Einstein metric of negative scalar
curvature.
(ii) Any self-dual Einstein metric of non-zero scalar curvature with two lin-
early independent commuting Killing fields arises locally in this way (i.e.,
in a neighborhood of any point, it is of the form (12.5.6) up to a constant
multiple).
This theorem, together with the explicit construction of Theorem 12.5.1 leads
quite naturally to the question: Are all compact positive QK orbifolds admitting
two commuting Killing vectors obtained via some QK reduction O(Ω) of HPn ? A
partial answer to this question in the case when the Konishi bundle of O(Ω) is
smooth (which is an extra condition on Ω) was provided by Bielawski in [Bie99].
We shall discuss his result later in the context of smooth “toric” 3-Sasakian mani-
fold. More recently, via a more careful analysis of orbifold singularities Calderbank
and Singer proved the following [CS06a]
Theorem 12.5.7: Let (O, g) be a compact self-dual Einstein 4-orbifold of positive
scalar curvature whose isometry group contains a 2-torus. Then, up to orbifold
coverings, (O, g) is isometric to a quaternionic Kähler quotient of quaternionic
projective space HPn , for some n ≥ 1, by a (n−1)-dimensional subtorus of Sp(n+1).
There is yet another family of orbifold metrics due to Hitchin [Hit95a, Hit96,
Hit95b]. These metrics come from solutions of the Painlevé VI equation and as
such were also introduced by Tod [Tod94]. We describe these metrics in some
details here and come back to them once more in the next chapter. Consider the
space V defined by
(12.5.7) V = {B ∈ M3,3 (R) | BT = B, Tr(B) = 0}
of traceless symmetric 3 × 3 matrices with inner product hB1 , B2 i = tr(B1 B2 ).
Clearly, V ' R5 and SO(3) acts on V by conjugation B 7→ g −1 Bg, g ∈ SO(3) and
the unit sphere M = S 4 in V can be described as matrices in V whose eigenvalues
{λ1 , λ2 , λ3 } satisfy
3
X 3
X
(12.5.8) λi = 0 = 1 − λ2i .
i=1 i=1

This action has cohomogeneity one and is the Z2 -quotient of the first case described
in Example 1.6.34. The associated group diagram has the structure
(12.5.9) SO(3) .
j− nn nn7 gPPP +
PP j
n PPP
nnn PPP
nnn
K− = O(2) K+ = O(2)
gPPP 7
PPP nnnnn
PPP nn
PPP nn
h−
nnn h+
D = Z2 × Z2
with the generic orbit SO(3)/D, where D = Z2 × Z2 ⊂ SO(3) is the subgroup of
diagonal matrices which is the stabilizer of any generic point. The two degenerate
orbits B± = SO(3)/K± ' RP2 are both Veronese surfaces in S 4 that correspond
12.5. COMPACT QUATERNIONIC KÄHLER ORBIFOLDS 447

to the subset of matrices with two equal eigenvalues. If two eigenvalues are equal,
then they must be equal to ± √16 , and the two signs correspond to the two orbits B±
and the subgroups K± = O(2). The diagram 12.5.9 gives rise to the decomposition
¡ ¢
(12.5.10) S 4 = RP2 ∪ (0, +∞) × (SO(3)/D) ∪ RP2 .
Explicitly, we can parameterize the conic 12.5.8 by observing that (λ1 − λ2 )2 +
3(λ1 + λ2 )2 = 2 so that
(12.5.11) √ √
2 1 2 1
λ1 (t) = α(t) + √ β(t) , λ2 (t) = − α(t) + √ β(t) , λ3 (t) = −2β(t) ,
2 6 2 6
where α2 + β 2 = 1. Choose the standard rational parameterization
2t 1 − t2
α(t) = , β(t) = .
t2 +1 t2 + 1
Note that t = 0 gives λ1 = λ2 = √16 and t = +∞ gives the other degenerate
orbit λ2 = λ2 = − √16 so that we have an explicit diffeomorphism S 4 \ {B− , B+ } '
(0, +∞) × SO(3)/D given by (t, g) 7→ g −1 ∆(t)g, where g ∈ SO(3) and ∆(t) =
diag(λ1 (t), λ2 (t), λ3 (t)).
Any SO(3)-invariant metric on S 4 \ {B+ , B− } defines an invariant metric on
each orbit SO(3)/D. It follows that any such metric must be of the form
(12.5.12) g = f (t)dt2 + [T1 (t)]2 σ12 + [T2 (t)]2 σ22 + [T3 (t)]2 σ32 ,
where {σ1 , σ2 , σ3 } is the basis of Maurer-Cartan invariant one-forms dual to the
standard basis of the Lie algebra so(3). The equations for the most general self-
dual Einstein metric with non-zero scalar curvature Λ and in the diagonal form
(12.5.12) has been derived by Tod [Tod94]. It follows that
dx2 σ2 (1 − x)σ22 xσ 2
(12.5.13) Fg = + 12 + 2 + 23 ,
x(x − 1) F1 F2 F3
where {F1 (x), F2 (x), F3 (x)} satisfy the following first order system of ODEs
dF1 F2 F3 dF2 F3 F3 dF3 F1 F2
(12.5.14) =− , =− , =−
dx x(1 − x) dx x dx x(1 − x)
and the conformal factor
(12.5.15)
8xF12 F22 F32 + 2F1 F2 F3 (x(F12 + F22 ) − (1 − 4F32 )(F22 − (1 − x)F12 ))
−4ΛF = .
(xF1 F2 + 2F3 (F22 − (1 − x)F12 ))2
The expression for the conformal factor is algebraic in x, F1 , F2 , F3 so that the
problem reduces to solving the system (12.5.14). It turns out that this system can
be reduced to a single second order ODE: the Painlevé VI equation
d2 y 1 ³ 1 1 1 ´³ dy ´2 ³ 1 1 1 ´ dy
= + + − + + +
dx2 2 y y − 1 y − x dx x x − 1 y − x dx
(12.5.16)
y(y − 1)(y − x) ³ x x−1 x(x − 1) ´
+ 2 2
α+β 2 +γ 2
+δ ,
x (x − 1) y (y − 1) (y − x)2
where (α, β, γ, δ) = (1/8, −1/8, 1/8, 3/8). One can define an auxiliary variable z by
dy y(y − 1)(y − x) ³ 1 1 1 ´
(12.5.17) = 2z − + +
dx x(x − 1) 2y 2(y − 1) 2(y − x)
448 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

which then allows one to express the original functions {F1 , F2 , F3 } in terms of any
solution y = y(x) of the equation 12.5.16:
y(y − 1)(y − x)2 ³ 1 ´³ 1´
F12 = z− z− ,
x(1 − x) 2(y − 1) 2y
y 2 (y − 1)(y − x) ³ 1 ´³ 1 ´
(12.5.18) F12 = z− z− ,
x 2(y − x) 2(y − 1)
y(y − 1)2 (y − x) ³ 1 ´³ 1 ´
F12 = z− z− .
1−x 2y 2(y − x)
In a series of papers Hitchin analyzed the Painlevé VI equation giving an alge-
braic geometry description of the solutions in terms of isomonodromic deforma-
tions [Hit95a, Hit96, Hit95b]. In particular, any such solution can be described
in terms of a meromorphic function on an elliptic curve C̃ with a zero of order k
at a chosen point P and a pole of order k at a point −P . Hitchin’s description
gives explicit formulas for the coefficients of the metric {F1 , F2 , F3 } in terms of the
elliptic functions. In particular, Hitchin shows
Theorem 12.5.8: Choose an integer k ≥ 3 and consider the SO(3)-invariant met-
ric gk defined on (1, ∞) × SO(3)/D by the formula (12.5.13) via the corresponding
solution of the Painlevé VI equation with the metric coefficients Fi = Fi (x, k),
i = 1, 2, 3.
(i) The metric gk is a positive definite self-dual Einstein of positive scalar
curvature for all 1 < x < ∞.
(ii) The metric gk extends smoothly over at x = 1 over B− = RP2 and

as x → ∞ gk acquires an orbifold singularity with angle k−2 around
2
B+ = RP .
Hence, for any¡ integer k ≥ 3 the metric
¢ gk can be interpreted as an orbifold metric
on Ok = B− ∪ (1, ∞)×SO(3)/D ∪B+ ' S 4 , where (Ok , gk ) is a compact self-dual
Einstein orbifold of positive scalar curvature.
We shall return to these cohomogeneity one orbifold metrics in the next chapter
when we consider the Konishi bundle over Ok and the twistor space Z(Ok ). Here
we discuss the metric for some lower values of k. To each k one associates a solution
of the Painlevé VI equation y = y(s) and we write the metric gk as in 12.5.12
(12.5.19) gk = f (s; k)ds2 + [T1 (s; k)]2 σ12 + [T2 (s; k)]2 σ22 + [T2 (s; k)]2 σ32 .
Explicit computation shows that for k = 3 the metric is given by the following
solution of the Painlevé VI equation
s2 (2s2 + 5s + 2) s3 (s + 2)
y= 2
, with x = .
(2s + 1)(s + s + 1) 2s + 1
The components of g3 can easily be calculated with f (s; 3) = 3(1 + s + s2 )−2 and
3(1 + 2s)2 3(1 − s2 )2 3s2 (2 + s)2
T12 (s; 3) = , T22 (s; 3) = , T32 (s; 3) = .
(1 + s + s2 )2 (1 + s + s2 )2 (1 + s + s2 )2
In the arc length coordinates this metric can be easily transformed to
g3 = dt2 + 4 sin2 t σ12 + 4 sin2 (2π/3 − t) σ22 + 4 sin2 (t + 2π/3) σ32 ,
which shows that g3 is the standard metric on S 4 written in triaxial form. Hence,
the orbifold (O3 , g3 ) is actually non-singular and the metric is the standard one
12.5. COMPACT QUATERNIONIC KÄHLER ORBIFOLDS 449

on S 4 . For k = 4 the metric comes from the following solution of the Painlevé VI
equation
y = s, with x = s2 ,
so that
1 1 (1 − s)2 s
f (s; 4) = , T12 (s; 4) = , T22 (s; 4) = , T32 (s; 4) = .
4s(1 + s) 1+s (1 + s)2 1+s
In an arc length parameterization g4 becomes
(12.5.20) g4 = dt2 + sin2 t σ12 + cos2 2t σ22 + cos2 t σ32 ,
which is indeed locally the Fubini-Study metric on CP2 . The orbifold (O4 , gk ) has
π1orb = Z2 and its universal cover is (CP2 , gF S ). This corresponds to the second case
of Example 1.6.34.
Just to illustrate how complicated the metric coefficients get for larger values
of k, following Hitchin [Hit96] we also give explicit formulas for k = 6, 8. For k = 6
(the orbifold singularity at angle π/2) one gets
s(s2 + s + 1) s3 (2s + 2)
y= , with x = .
(2s + 1) s+1
This yields
(1 + s + s2 )
f (s; 6) = ,
s (s + 2)2 (2s + 1)2
(1 + s + s2 )
[T1 (s; 6)]2 = ,
(s + 2) (2s + 1)2
(s2 − 1)2
[T2 (s; 6)]2 = ,
(1 + s + s2 ) (s + 2) (2s + 1)
s (1 + s + s2 )
[T3 (s; 6)]2 = .
(s + 2)2 (2s + 1)
This gives the metric g6 of Equation 12.5.19 for the range 1 < s < ∞.
For k = 8 (the orbifold singularity at angle π/3) we get
4s(3s2 − 2s + 1) ³ 2s ´4
y= , with x = .
(s + 1)(1 − s)3 (s2 + 2s + 3) 1 − s2
This yields
(1 + s)(s2 + 2s − 1)(3 − 2s + s2 )(1 − 2s + 3s2 )(1 + 2s + 3s2 )
f (s; 8) = ,
(1 − s) s (1 + s2 )(s2 − 2s − 1) (3 + 2s + s2 )2 (s4 − 6s2 + 1)
(1 + s2 )(3 − 2s + s2 )(1 − 2s − s2 )2 (1 + 2s + 3s2 )
[T1 (s; 8)]2 = ,
(1 + 2s − s2 )2 (3 + 2s + s2 )2 (1 − 2s + 3s2 )
(1 − s)(1 + s)3 (3 − 2s + s2 )(1 − 2s + 3s2 )
[T2 (s; 8)]2 = ,
(1 + 2s − s2 )(3 + 2s + s2 )2 (1 + 2s + 3s2 )
4s(s + 1)2 (1 − 2s + 3s2 )(1 + 2s + 3s2 )
[T3 (s; 8)]2 = .
(1 + 2s − s2 )(3 − 2s + s2 )(3 + 2s + s2 )2

This metric g8 of Equation 12.5.19 becomes positive definite for the range 2 − 1 <
s < 1.
The Theorems of Hitchin [Hit95b], Calderbank and Pedersen [CP02], and
Calderbank and Singer [CS06a] are milestones in the broader problem of the clas-
sification of all compact self-dual Einstein 4-orbifolds. It seems plausible that these
450 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

are the only cohomogeneity one compact positive self-dual Einstein 4-orbifolds, but
a proof is lacking so far.
Open Problem 12.5.1: Classify all compact positive self-dual Einstein 4-orbifolds
with a cohomogeneity one action of a Lie group.
If one adds to this classifying the cohomogeneity two actions of SU (2) one
arrives at
Open Problem 12.5.2: Classify all compact positive self-dual Einstein 4-orbifolds
with at least a 2-dimensional isometry group.
The problem of finding examples of compact positive self-dual Einstein 4-
orbifolds without any assumption about symmetries seems quite intractable. No
compact orbifolds without any Killing vector fields are known at this time. How-
ever, there are two examples of positive QK 4-orbifolds with one Killing vector,
both obtained via symmetry reduction.
Example 12.5.9: [QK Extension of Kronheimer Quotients] The first orbifold
examples of positive QK orbifold metrics which are not toric were constructed in
[GN92]. The construction is a quaternionic Kähler modification of the Kronheimer
construction of hyperkähler ALE spaces discussed later in Section 12.10. We use
the notation there to explain the result. Let Γ ⊂ Sp(1) be a discrete subgroup.
Consider the quaternionic projective space PH (H|Γ| × H). The Kronheimer group
K(Γ) acts on H|Γ| as in 12.10.3. Let us consider a homomorphism b : K(Γ)−→Sp(1).
Such a homomorphism extends the action of K(Γ) to PH (H|Γ| × H) via

(12.5.21) g · [u, u0 ] = [g · u, b(g)u0 ], u ∈ H|Γ| , u0 ∈ H, g ∈ K(Γ) ,

where u 7→ g · u is the Kronheimer’s action discussed in 12.10.3. Consider the


map db : k−→sp(1) of Lie algebras. We can think of db ∈ k∗ ⊗ sp(1) by setting
hdb, Xi = db(X), X ∈ k. Let us denote the new action by K(Γ; b) and the QK
reduction of PH (H|Γ| × H) by O(Γ, b). We have the following [GN92]
Theorem 12.5.10: The QK reduced space O(Γ, b) is a compact positive self-dual
and Einstein orbifold if db ∈ k∗ ⊗sp(1) is in the Kronheimer’s good set of Definition
12.10.4.
Note that in the case of Γ = Zn the construction and Theorem 12.5.10 merely
give subfamilies of toric examples discussed earlier. However, in all cases when Γ
is non-Abelian we get families of positive self-dual Einstein orbifold metrics which
are not toric.
Example 12.5.11: [Abelian Quotients of Real Grassmannians] Consider the
positive QK structure on the Grasmannian of oriented 4-planes in Rn . The isometry
group SO(n) of the symmetric space Gr+ n
4 (R ) contains a torus and one can examine
+ n
possible QK reductions of Gr4 (R ). These are described in detail at the level of the
Konishi bundle in Section 13.9. In particular, there are only two possible Abelian
quotients of Gr+ n 3
4 (R ) for which the reduced space is 4-dimensional: (i) T -reduction
+ 8 2 + 7
of Gr4 (R ) and (ii) T -reduction of Gr4 (R ) (see Proposition 13.9.2). Both lead
to non-trivial examples of positive self-dual Einstein orbifold metrics which are not
toric and they were first introduced in [BGP02] and later studied in [Bis07]. In
particular, with the notation of Definition 13.9.1 and Proposition 13.9.2 we have
the following
12.5. COMPACT QUATERNIONIC KÄHLER ORBIFOLDS 451

Theorem 12.5.12: Let the weight matrices Θ12,3 ∈ M2,3 (Z) and Θ03,4 ∈ M3,4 (Z)
describe the choices of T 2 ⊂ T 3 ⊂ SO(7) and T 3 ⊂ T 4 ⊂ SO(8), respectively.
Let O(Θ12,3 ) and O(Θ03,4 ) denote the corresponding QK reductions of Gr+ 7
4 (R ) and
Gr+ 8
4 (R ). We have
(i) If all three 2 × 2 minor determinants of Θ12,3 are non-zero then O(Θ12,3 )
is a compact 4-orbifold.
(ii) If all four 3 × 3 minor determinants of Θ03,4 are non-zero then O(Θ03,4 )
is a compact 4-orbifold.
In both case we get compact orbifold families of positive self-dual Einstein metrics
with a one-dimensional isometry group.
With all the available examples one can naturally begin asking questions about
geometric properties of such metrics. It turns out that a pivotal role in understand-
ing such metrics is played by the non-linear PDE
(12.5.22) uxx + uyy + (eu )zz = 0 .
This equation was first described in [BF82] as providing solutions to the self-dual
Einstein equations with zero scalar curvature and one Killing vector field of ‘rota-
tional type’. In [BF82] it was shown that the zero scalar curvature (or vacuum)
self-dual Einstein equations admitting one Killing vector field amounts to solving ei-
ther the well-known three dimensional Laplace equation or Equation 12.5.22. Those
Killing fields that led to the three dimensional Laplace equation were called trans-
lational Killing fields, whereas, those leading to Equation 12.5.22 were called rota-
tional. The translational Killing fields have self-dual covariant derivative and are
well understood [TW79]. For example, they give rise to the well-known Gibbons-
Hawking Ansatz [GH78a]. On the other hand Equation 12.5.22 has proven to be
very resistent in offering up explicit solutions [Fin01]. Nevertheless, it has appeared
in a variety of settings, for example LeBrun [LeB91b] used it in his construction
of self-dual metrics on the connected sums of CP2 . (See also [PP98b, AG02]
for further development in terms of Hermitian-Einstein geometry). Moreover, it
can be viewed as an infinite dimensional version of the better known Toda lat-
tice equation associated with the Lie algebra of type An , and so it has become
known as the SU (∞)-Toda field equation [War90, Sav89] or alternatively the
Boyer-Finley equation [FKS02]. Its importance for us at this stage lies in the re-
markable observation made by Tod [Tod97] that finding solutions to the self-dual
Einstein equations with non-zero scalar curvature can be reduced to solving Equa-
tion (12.5.22). This equation as with the full self-dual or anti-self-dual Einstein
equations is related to integrability questions, infinite sequences of conservations
laws, and twistor theory, which we briefly discuss in Section 12.7. Actually there
are several cases where there are known implicit solutions to Equation (4.6.2). It
would be interesting to see if one could turn implicit solutions of Theorem 12.5.12
and Theorem 12.5.12 into explicit solutions of Equation (4.6.2).
Most 4-dimensional Einstein metrics appear as Riemannian metrics adapted to
some other geometric structure, self-dual (or anti-self-dual) metrics, and/or Kähler
metrics are perhaps the best known examples. Here is an interesting Venn dia-
gram taken from Tod [Tod97], where the special intersecting regions deserve some
comment.
Let us now discuss the overlapping areas of Figure 1. All of the labelled areas
can be related to Sasakian geometry by taking an appropriate S 1 or SO(3) bundle or
452 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

W+ = 0 (or W− = 0)
...................................................
......... .......
....... ......
...... ......
...... .....
..... .....
.
....... .....
... ...
.. ...
....
. ...
...
.
...
. ...
...
.
.... ...
.... ...
...
.... ...
.... . ... ...... ... . . .. . .. .. ................ .
. . ... . .. . . . . ........ .
... ............. ...
.
... .................. ........ ...................
..
........ ...
.. ... .. . ........
.............
. ...................
. .. ......
...
.. .
..... .... B .........
. .....
..... D ... ..........
.
... ... .... ..... . .....
.... ... .... ... ... ....
... ... ... .. ...
Einstein Kähler
.
. ... .
. . ... .
. ...
. ... ... ..
... ... ...
... ...
...
... ...
... A ... ...
...
.... ..... .. .
... ..... ...
....
.....
..... .... ... ........ ...
...
...... .. ... ........
.... ........ .. .. ...
.....
.
..... ................... . ... .. ........... ...
...
... ... ...................................... ..
. .
... ... .. ..
... ... .... ...
... ... ... ....
...
...
...
... C .... ....
... ... .. ..
... ... ... ...
... ... ... ...
... ... ... ...
... ... ... ...
..... ..... ...
...
...
....
..... ..... . .
..... ..... ..... .....
...... ...... .......... .....
....... ........... ......
........ .. .......
........... ........ ................... ........
......................................... .........................................

Figure 1. Tod’s Venn diagram of special metrics in dimension four.

orbibundle. In the area labelled C are the generic Kähler-Einstein 4-manifolds that
were treated in Chapter 5. In the regions labelled A and D the complex structure
chooses an orientation which breaks the equivalence between being self-dual or anti-
self-dual. Thus, in both region A and region D one must consider the two cases
separately. The two cases for region A consists of self-dual Kähler-Einstein, and
anti-self-dual Kähler Einstein. In the category of compact manifolds the former
consists of CP2 with the Fubini-Study metric and compact quotients of Hermitian
hyperbolic space with the Bergman metric [ADM96, Boy88b, Kod87, KS93b],
whereas, the later consists only of flat tori, K3 and Enriques surfaces, which are
all locally hyperkähler and have zero scalar curvature. Many more examples of
both self-dual Kähler-Einstein and anti-self-dual Kähler-Einstein appear when one
allows orbifold singularities [Bry01, ACG06], for example in the hyperkähler case
Reid’s list of 95 singular K3 given in Appendix B.1 and discussed in Section 5.4.2
occur. Likewise, complete metrics on non-compact 4-manifolds are plentiful in
all regions of the above diagram. We discuss hyperkähler geometry more fully
in Sections 12.7 through 12.10 with many 4-dimensional examples. In region D
many researchers [Bou81, Bry01, Che78, Der83, Ito84] have obtained results
concerning self-dual Kähler manifolds. Anti-self-dual Kähler metrics automatically
have zero scalar curvature [Der83]. Moreover, which possible compact complex
surfaces can admit such metrics have been delineated [Boy86]. There has been
much work [LeB91d, LS93, KP95, Tod95, Dan96, KLP97, RS05, DF06]
within the last fifteen years or so in proving the existence of scalar flat Kähler
metrics on 4-manifolds. Finally region B are the self-dual metrics which are not
Kähler. Here we are interested mainly in the case of positive scalar curvature, since
only these admit SO(3) or SU (2) orbibundles whose total space has a 3-Sasakian
structure. These have been discussed in detail above. The world of negative scalar
curvature self-dual Einstein metrics is fascinating with spectacular abundance of
complete metrics on non-compact manifolds [Ped86, Gal87b, Gal91, LeB91c,
Hit95b, Biq00, Biq02, CP02, CS04, BCGP05, Duc06]. The subject is worthy
of a separate book. We concentrate on the positive QK manifolds here as they are
bases of Konishi orbibundles of 3-Sasakian spaces considered in the next chapter.
However, negative QK manifolds are also related to 3-Sasakian (and not just semi-
Riemannian 3-Sasakian geometry) as first observed by Biquard [Biq99].
12.6. HYPERCOMPLEX AND HYPERHERMITIAN STRUCTURES 453

12.6. Hypercomplex and Hyperhermitian Structures


Recall from Example 1.4.19 the following
Definition 12.6.1: A smooth manifold M is said to be almost hypercomplex
if it admits a GL(n, H)-structure.
Alternatively, an almost hypercomplex structure is an almost quaternionic
structure such that the subbundle Q ⊂ End(T M ) is trivial. Thus, Q has a global
orthonormal frame {I1 , I2 , I3 } whose elements satisfy Equation (12.2.1). Such an
orthonormal frame can be viewed as a map I : R3 −−−→Q satisfying I(ea ) = Ia ,
where {ea }3a=1 is the standard basis for R3 . So given any two points τ , τ ∈ S 2 we
can write Equation (12.2.1) in terms of arbitrary frames of Q as
(12.6.1) I(τ ) ◦ I(τ 0 ) = −hτ , τ 0 i1l + I(τ × τ 0 ) ,
where hτ , τ 0 i is the standard inner product in R3 and τ × τ 0 is the cross-product.
So hypercomplex structure provides M with an S 2 ’s worth of complex structures.
We denote the family of complex structures satisfying Equation (12.6.1) by I and
refer to it as a hypercomplex structure.
Let g be a metric on M such that
(12.6.2) g(I(τ )X, I(τ )Y ) = g(X, Y ) ,
2
for any τ ∈ S nd X, Y ∈ Γ(T M ). Such a metric is said to be adapted to the
hypercomplex structure I and the pair (I, g) is called a hyperhermitian structure.
It is easy to see that such a metric always exists.
Definition 12.6.2: An almost hypercomplex manifold (M, I, g) with an adapted
metric g is called an almost hyperhermitian manifold.
As discussed in Chapter 1 an almost hyperhermitian structure is equivalent to
a reduction of the GL(n, H)-bundle to the subgroup Sp(n). Obata showed that
every almost hypercomplex manifold M 4n admits a canonical GL(n, H)-invariant
connection [Oba66], called the Obata connection.
Definition 12.6.3: An almost hypercomplex manifold (M, I) is called hypercom-
plex if all complex structures I(τ ), τ ∈ S 2 are integrable. A hypercomplex manifold
with an adapted metric is called hyperhermitian.
Actually, if any two orthogonal almost complex structures in the almost hyper-
complex structure are integrable then all the complex structures I(τ ) are integrable
[Oba66, Sal89]. In the hypercomplex case integrability can be expressed in sev-
eral different ways. For example, the Obata connection in general has non-trivial
torsion. But on a hypercomplex manifold this unique connection is torsion-free.
So an alternative definition of a hypercomplex structure is that it is an almost hy-
percomplex structure such that the Obata connection is torsion-free. In the lowest
dimension compact hyperhermitian 4-manifolds were classified by Boyer [Boy88a]
who proved
Theorem 12.6.4: Let (M, I, g) be a compact hyperhermitian 4-manifold. Then
(M, I, g) is conformally equivalent to one of the following
(i) a 4-torus with its flat metric,
(ii) a K3 surface with a Kähler Ricci-flat metric,
(iii) a coordinate quaternionic Hopf surface with its standard locally confor-
mally flat metric.
454 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

In higher dimensions there are many examples of hypercomplex structures, but


no classification results so far. For example, it is known which Lie groups admit such
structures [SSTVP88, Joy92, BDM96]. The simplest example here is G = U (2)
which as a compact complex surface is a Hopf surface, and it actually admits two
commuting hypercomplex structures. More generally [Joy92]
Theorem 12.6.5: Let G be a compact Lie group. Then there exist an integer
0 ≤ k ≤ max{3, rk(G)} such that U (1)k × G has a homogeneous hypercomplex
structure.
There is a natural construction of hypercomplex structures on the total space of
circle bundles over any 3-Sasakian manifold M [BGM98a] (See the next chapter for
a description of 3-Sasakian structures). For example, any trivial bundle S 1 ×M then
admits locally conformally hyperkähler structures that are automatically hypercom-
plex. However, non-trivial circle bundles give more interesting results. For example,
large families of hypercomplex structures were shown [BGM94b, BGM96a] to
exist on the complex Stiefel manifolds VC n
n,2 of 2-frames in C . (See also [Bat96a]).
There is also a good deformation theory for hypercomplex structures [PP98a].
Further discussion of hypercomplex structures and their relation to quaternionic
geometry can be found in [Joy92, AM96a, AM96b, PPS98].
Here we recall the quotient construction of Joyce [Joy91] which we shall use
later. Let (M, I) be a hypercomplex manifold. We define the automorphism group
Aut(M, I) of (M, I) by

(12.6.3) Aut(M, I) = {φ ∈ Diff(M ) | φ∗ I(τ ) = I(τ )φ∗ for all τ ∈ S 2 } .

Since a hypercomplex structure is a G-structure of finite type, it follows that the


group Aut(M, I) is a Lie group. Let H be a Lie subgroup of H ⊂ Aut(M, I).
Then H acts on M as complex automorphisms with respect to any of the complex
structures in {I(τ )}τ ∈S 2 .
Definition 12.6.6: Let (M, I) be a hypercomplex manifold. Given a compact Lie
subgroup H ⊂ Aut(M, I) a hypercomplex moment map is any H-equivariant
map µ = i1 µ1 + i2 µ2 + i3 µ3 : M → h∗ ⊗ sp(1) satisfying both of the following
conditions
(i) I1 dµ1 = I2 dµ2 = I3 dµ3 , where Ia acts on sections Γ(T ∗ M ⊗ h∗ ).
(ii) For any non-zero element ζ ∈ h and the induced vector field Xζ ∈ Γ(T M )
I1 dµ1 (Xζ ) 6= 0 on M.
Note that the condition (i) of 12.6.6 is equivalent to requiring that the complex
valued function µa + iµb on a complex manifold (M, Ic ) be holomorphic function
with respect to the complex structure Ic for any cyclic permutation (a, b, c) of
(1, 2, 3). Joyce proves the following [Joy91]
Theorem 12.6.7: Let (M, I) be a hypercomplex manifold H a any compact Lie
group in Aut(M, I). Choose any hypercomplex moment map µ and let ζ = ζ1 i1 +
ζ2 i2 + ζ3 i3 ∈ h∗ ⊗ sp(1), where all three ζi are in the center of h∗ . Suppose the
H-action on Nζ = µ−1 (ζ) has only finite isotropy groups and M̂ (ζ) = Nζ /H is an
orbifold. Then M̂ (ζ) has a naturally induced hypercomplex structure.
The hypercomplex quotient construction can be used to build many examples
of hypercomplex manifolds as we shall see in Chapter 13. The main point here
is that unlike in the case of hyperkähler reduction which will be defined in the
12.7. HYPERKÄHLER MANIFOLDS 455

following sections the hypercomplex reduction is much more flexible in the way one
chooses the associated moment map.
Since a hypercomplex manifold M is quaternionic, it has a twistor space Z(M )
which satisfies all the properties of Theorem 12.2.16. But also in the hypercomplex
case the trivialization of Q gives a trivialization of the twistor space Z(M ) = S 2 ×M
as smooth manifolds, but not as complex manifolds. Nevertheless, the projection p
onto the first factor is holomorphic, and we have a double fibration
Z(M )
p
(12.6.4) . &
CP1 Ã M
which gives a correspondence: points τ ∈ S 2 ' CP1 correspond to complex struc-
tures I(τ ) on M in the given hypercomplex structure I; points x ∈ M correspond
to rational curves in Z(M ) with normal bundle 2nO(1), called twistor lines.

12.7. Hyperkähler Manifolds


Given a an almost hyperhermitian manifold (M, I, g) we can use the metric to
define the 2-forms
(12.7.1) ω(τ )(X, Y ) = g(I(τ )X, Y ), X, Y ∈ Γ(T M ) .
In particular, given a hypercomplex structure I and choosing the basis {I1 , I2 , I3 }
we get the three fundamental 2-forms {ω1 , ω2 , ω3 } which trivializes the subbundle
Q∗ . By analogy with the almost Kähler case consider
Definition 12.7.1: An almost hyperhermitian manifold (M, I, g) is called almost
hyperkähler if the associated fundamental 2-forms are closed and it is called hy-
perkähler (HK) if the associated 2-forms are parallel with respect to the Levi-Civita
connection of g.
Unlike the Kähler case an almost HK manifold must automatically be HK
[Hit87a]. In fact we have the following equivalent characterization of hyperkähler
manifolds
Theorem 12.7.2: Let (M 4n , I, g) be an almost hyperhermitian manifold with the
fundamental 2-forms ωa (X, Y ) = g(Ia X, Y ), a = 1, 2, 3. Then the following condi-
tions are equivalent
(i) (M 4n , I, g) is hyperkähler,
(ii) (M 4n , I, g) is almost hyperkähler,
(iii) (M 4n , I, g) is 1-integrable.
(iv) ∇I1 = ∇I2 = ∇I3 = 0,
(v) Hol(g) ⊂ Sp(n).
In particular, an HK manifold is Kähler with respect to any choice of complex
structure in I(τ ), and the holonomy reduction
P implies that HK manifolds must
be Ricci-flat. Of course the 4-form Ω = a ωa ∧ ωa is parallel so that any HK
manifold is also QK, only the quaternionic bundle Q on M is trivial and the scalar
curvature vanishes. The following diagrams describes how HK geometry relates to
other quaternionic geometries discussed in previous sections
In dimension 4 the situation is special. Since Sp(1) ' SU (2) the HK condition
is equivalent to asking that M 4 be Kähler and Ricci-flat. From the decomposition
456 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

Quaternionic Kähler Quaternionic


Sp(n)Sp(1) −−−−→ GL(n, H)H∗
∇Ω = 0 Oproiu connection ∇
x x
 

 n>1 

Hyperkähler Hypercomplex
Sp(n) −−−−→ GL(n, H)
∇ω1 =∇ω2 =∇ω3 =0 Obata connection ∇

Figure 2. Quaternionic geometries in dimension ≥ 8.

of the Riemann curvature in (12.2.14) we see that the only non-zero component is
W− and such manifolds are also sometimes called half-flat.

W+ = 0 or W− = 0, W+ = 0 or W− = 0,
Einstein −−−−→ Conformal
Sp(1)Sp(1) R∗ × SO(4)
x x
 

 n=1 

Hyperkähler Hypercomplex
−−−−→
Sp(1) H∗

Figure 3. Quaternionic 4-manifolds.

When M 4 is a compact HK manifold then, up to cover, it must be either a K3


surface or a flat torus. On the other hand, if we do not insist on compactness the
question of the classification of such metrics remains wide open. Only partial clas-
sification results are known. All these metrics are important in General Relativity
Theory as they are vacuum solutions (Ricci-flat) of the Euclidean Einstein equa-
tions. Such solutions are called gravitational instantons. They will all be described
as certain quotients in the next section.
Proposition 12.7.3: Let (M 4n , I, g) be an HK manifold and consider the Kähler
structure (I1 , g, ω1 ). The complex 2-form ω+ = ω2 + iω3 is of type (2, 0) and holo-
morphic, i.e., it is a holomorphic symplectic form on M.
Proof. Let (U ; z1 , . . . , z2n ) be a holomorphic local chart on M with respect to
the complex structure I1 . Consider the 2-form ω+ (X, Z) = g(I2 X, Y ) + ig(I3 X, Y )
and extend it by linearity to the complexified tangent bundle T M ⊗ C. Setting
X = ∂∂z̄j we compute for any vector field Y
³ ∂ ´ ³ ∂ ´ ³ ∂ ´ ³ ∂ ´
ω+ (∂z̄j , Y ) = g I2 , Y + ig I3 , Y = ig I2 I1 , Y + ig I3 ,Y = 0,
∂ z̄j ∂ z̄j ∂ z̄j ∂ z̄j
12.7. HYPERKÄHLER MANIFOLDS 457

since I2 I1 = −I3 which implies that ω+ is a (2, 0)-form. It is holomorphic since it


is closed. ¤

Note that the (0, 2)-form conjugate under I1 to ω+ is ω− = ω2 − iω3 . Proposi-


tion 12.7.3 can easily be generalized to an arbitrary oriented orthonormal 3-frame
{τ1 , τ2 , τ3 }, namely
Proposition 12.7.4: Let (M 4n , I, g) be an HK manifold and consider the Kähler
structure (I(τ1 ), g, ω(τ1 )). The complex 2-form ω+ (τ1 ) = ω(τ2 ) + iω(τ3 ) is a holo-
morphic symplectic form on M with respect to the complex structure I(τ1 ).
We now consider briefly the twistor space Z(M ) of a hyperkähler manifold
M. For some references here see [HKLR87, Joy00, Sal86]. Since a hyperkähler
structure is hypercomplex, the twistor space Z(M ) of a hyperkähler manifold sat-
isfies all the properties of Theorem 12.2.16 as well as the correspondence given by
Diagram 12.6.4. But when M is hyperkähler the complex manifold (M, I(τ1 )),
which is the fibre over τ1 ∈ S 2 ' CP1 of the holomorphic fibration p, can be viewed
as a divisor in Z(M ) with a holomorphic symplectic form ω+ (τ1 ). This gives a
twisted holomorphic 2-form $ on Z(M ) that is a section of p∗ O(2) ⊗ Q̃∗ , where
Q̃∗ is the bundle Q∗ on M pulled back to Z(M ) under the right hand projection in
diagram 12.6.4. To express this holomorphic data explicitly P we choose the standard
basis {e1 , e2 , e3 } of R3 , and write a point τ ∈ S 2 as τ = a τa ea . These standard
coordinates are related to the complex affine coordinate t ∈ C ⊂ CP1 ' S 2 we have
[HKLR87]
µ ¶
1 − |t|2 t + t̄ t − t̄
(τ1 , τ2 , τ3 ) = , , i .
1 + |t|2 1 + |t|2 1 + |t|2
Then the complex structure IZ on Z(M ) as an endomorphism of the tangent space
T(x,t) Z(M ) = Tx M ⊕ Tt S 2 becomes

1 − |t|2 t + t̄ t − t̄
(12.7.2) IZ = I(τ ) + I0 = 2
I1 + 2
I2 + i I3 + I0 ,
1 + |t| 1 + |t| 1 + |t|2
where I0 denotes the standard complex structure on the tangent space Tt S 2 given
by multiplication by i. Of course, there is a similar expression for the other chart
with affine coordinate s = 1t centered about (−1, 0, 0). Now the twisted holomorphic
2-form $ on Z(M ) is written as

(12.7.3) $ = ω+ + 2tω1 − t2 ω− .
There is a converse to this twistor space construction [HKLR87], but first some
notation. Given a complex manifold Z with a holomorphic fibration p : Z → CP1 ,
we let TF denote the vertical subbundle of T Z, that is, the kernel of the differential
p∗ : T Z−−−→T CP1 .
Theorem 12.7.5: Let (Z, J) be a complex manifold of complex dimension (2n + 1)
equipped with the following data
(i) a holomorphic projection p : Z → CP1 ,
(ii) a holomorphic section $ of p∗ O(2) ⊗ Λ2 TF∗ which restricts to a holomor-
phic symplectic form on the fibres of p.
(iii) a free antiholomorphic involution σ : Z → Z that satisfies σ ∗ ($) = $,
and p ◦ σ = å ◦ p, where å is the antipodal map on S 2 .
458 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

Let M be the set of all rational curves C in Z with normal bundle 2nO(1) and
σ(C) = C. Then M is a hypercomplex manifold with a natural pseudo-HK metric
g. If g is positive definite then M is HK.

12.8. Hyperkähler Quotients


In this section we review the generalization of the Marsden-Weinstein construc-
tion described in Section 8.4 to HK manifolds with hyperholomorphic isometries.
Such reductions were first considered by Lindström and Roček as early as in 1983
[LR83], while investigating the so-called 4-dimensional N = 2 globally supersym-
metric σ-model theories. It is known that target manifolds of such σ-models are HK.
Lindström and Roček observed that one can “gauge away” hyperholomorphic sym-
metries. In the process one introduces auxiliary gauge fields without kinetic terms
in the Lagrangian, i.e., hyperholomorphic Killing vectors. The Euler-Lagrange
equation for such fields are algebraic (moment map equations) and eliminating
these fields leads to a new N = 2 supersymmetric σ model theory, hence, a new
HK metric. A few years later Hitchin gave the rigorous mathematical descrip-
tion of what is now known as hyperkähler reduction [HKLR87]. We will describe
this construction and some of the basic examples as it provides a blueprint for
much of the material of the next chapter. Let (M, I, g) be an HK manifold and
G ⊂ Aut(M, I, g) ⊂ Isom(M, g) be a Lie group acting smoothly and properly on M
by preserving the hyperkähler structure. Then G acts by symplectomorphism pre-
serving symplectic forms ωa , a = 1, 2, 3. Suppose the G-action is Hamiltonian with
respect to each symplectic form ωa . We will call such an action hyperhamiltonian.
Definition 12.8.1: A hyperkähler manifold (M, I, g) together with an effective
hyperhamiltonian G-action is called a hyperhamiltonian G-manifold.
As discussed in section 8.4.1 such an action gives rise to three G-equivariant
symplectic moment maps µa : M → g∗ . We can assemble these maps together to
get
Definition 12.8.2: Let (M, I, g) be a hyperkähler G-manifold. The map µ =
(µ1 , µ2 , µ3 ) = i1 µ1 + i2 µ2 + i3 µ3
(12.8.1) µ = M → g∗ ⊗ sp(1)
is called the hyperkähler moment map for the action of G.
We have the following natural generalization of the Marsden-Weinstein sym-
plectic reduction theorem [HKLR87]
Theorem 12.8.3: Let (M, I, g) be HK and G be a hyperhamiltonian action on M
with the HK moment map µ : M → g∗ ⊗ sp(1). Let λ = (λ1 , λ2 λ3 ) ∈ g∗ ⊗ sp(1) be
any element fixed by the co-adjoint action of G on its Lie co-algebra g∗ . Suppose λ
is a regular value of µ so that N = µ−1 (λ) ⊂ M is a manifold. Suppose further
that the orbit space M̂ (λ) = µ−1 (λ)/G is a manifold (orbifold). Then M̂ (λ) is an
HK manifold (orbifold) with the HK structure induced from M via inclusion and
projection maps.

Proof. We only sketch proof here. The manifold (M, I1 , ω1 , g) is a Kähler


manifold. The HK reduction can be seen as a two step process: First, we consider
the function
(12.8.2) µ+ = µ2 + i1 µ3 ,
12.8. HYPERKÄHLER QUOTIENTS 459

which is easily seen to be holomorphic on (M, I1 , g). Thus the set N+ = µ−1+ (0) is a
complex subspace of (M, I1 , g), in particular it must be Kähler. Note that N+ need
not be a smooth manifold, it is sufficient that it be smooth in some H-invariant
0 0
open neighborhood N+ such that N ⊂ N+ ⊂ N+ . The action of G restricts to
N+ with the Kähler moment map µ1 : N+ → g∗ . Hence, the reduced space M̂
is nothing but a Kähler reduction of N+ by the action of G. In particular, M̂ is
Kähler with the complex structure Iˆ1 induced from M by the quotient construction.
Now, the result follows by observing that the same argument applies to I2 and I3 ,
and M̂ is therefore Kähler with respect to all three complex structures {Iˆ1 , Iˆ2 , Iˆ3 }.
One can easily check that the induced complex structures satisfy the quaternionic
relations. ¤

We remark that, just as in the symplectic case, one can consider more general
“singular” quotients. This was done by Dancer and Swann [DS97a, Swa97] who
showed
Theorem 12.8.4: Let (M, I(τ ), g) be a hyperhamiltonian G-manifold with the
moment map µ : M −−→g∗ ⊗ sp(1). Furthermore, suppose G acts smoothly and
properly on M. Let M(H) denote the stratum consisting of orbits of type H < G.
Then NH = µ−1 (0) ∩ M(H) is a manifold and the orbit space
µ−1 (0) ∩ M(H)
M̂H =
G
has a natural HK structure. Consequently, the reduced space M̂ = µ−1 (0)/G is a
disjoint union of HK manifolds
[
M̂ = M̂H .
H<G

The proof is a corollary of the Sjamaar-Lerman Theorem 8.4.3. In particular,


each stratum is a smooth HK manifold but unlike the symplectic case it is not clear
whether the quotient is decomposable in the sense of Goretsky-MacPherson.
The rest of this section is devoted to introducing some basic examples of HK
reduction of the flat model Hn . We will often use the pair µ = (µ1 , µ+ ) to describe
the moment map in complex coordinates on (M, I1 , ω1 , g) relative to I1 . Since we
single out I1 we will also use {i, j, k} for the quaternions {i1 , i2 , i3 }.
Example 12.8.5: [Calabi metrics on T ∗ CPn ] Let u = z̄ + wj ∈ Hn ' Cn × Cn
and consider the diagonal action of G = S 1 ⊂ Sp(n) given by left multiplication
with g(t) = eit . The moment map for this action is
n
X
(12.8.3) µ(u) = iµ1 + µ+ j = ūk iuk .
k=1

In complex charts we get


n
X X
(12.8.4) µ1 (w, z) = −i (|wj |2 − |zj |2 ) , µ+ (w, z) = 2i wj zj ,
j=1 j

and the circle action reads (w, z) 7→ (eit w, e−it z). One could consider an arbitrary
level set of the moment map. However, as Sp(1)+ is a symmetry of the flat HK
metric, one can use it to choose the value of µ to be a constant multiple of i.
460 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

Further scaling the metric shows that it is sufficient to consider N = µ−1 (−i) and
N+ = µ−1
+ (−i) ⊂ N which are described as
n n
X X o
(12.8.5) N = (w, z) ∈ Cn × Cn | (|wj |2 − |zj |2 ) = 1, wj zj = 0 ,
j=1 j

n X o
(12.8.6) N+ = (w, z) ∈ Cn × Cn | wj zj = 0 .
j

Let M̂ = µ−1 (−i)/S 1 . We first want to identify M̂ with the Kähler reduction of N+
0 0
(or a G-invariant open set N+ ⊂ N+ ). For an appropriate choice of N+ its Kähler
0
reduction N+ will be an algebraic quotient of N+ ⊂ N+ by the complexification C∗
of G = S 1 . Note, however, that N+ 0
is not compact so we cannot rely on Kirwan’s
2
theorems in this setting[Kir84] . Nevertheless, we get the following identification
(12.8.7) M̂ = µ−1 (−i)/S 1 ' N+ 0
/C∗ ,
0
P
where N+ = {(w, z) ∈ Cn × Cn | j wj zj = 0, w 6= 0}, and C∗ acts by
(w, z) 7→ (λw, λ̄z). Thus, M̂ is the holomorphic cotangent bundle T ∗ CPn−1 . It
turns out that the HK metric obtained on T ∗ CPn−1 via this reduction is isometric
to the Calabi metric [Cal79], the first non-flat example of a complete HK manifold.
An N = 2 supersymmetric σ-model description of the metric is due to Lindström
and Roček [LR83] and, in the above language it appears in [HKLR87]. An explicit
expression for this metric in dimension 4 was discovered by Eguchi and Hanson
[EH79] and it is called the Eguchi-Hanson gravitational instanton.
Example 12.8.6: This example involves a non-compact hyperhamiltonian group
action of G = R on H × Hn ' Hn+1 defined for any p = (p1 , . . . , pn ) ∈ Rn by
(12.8.8) φp (u0 , u1 , . . . , un ) = (u0 − t, ep1 it u1 , . . . , epn it un ) ,
with the moment map
n
X
(12.8.9) µp (u) = iµ1 + µ+ j = 2Im(u0 ) + ipk ūk iuk .
k=1

We can always shift to the zero-level set and then the moment map equations can
be “solved” by writing
n
X
(12.8.10) Im(u0 ) = − pk ūk iuk .
k=1

This action is free and proper on H × Hn . In particular it is free and proper on


µ−1 −1
p (0) = Np . Denote the quotient manifold by M (p) = µp (0)/R. First, note
that M (p) is diffeomorphic to Hn . This follows from the observation that the set
S = {u ∈ µ−1 p (0) | Re(u0 ) = 0} is a global slice for this action. The induced
HK metric ĝ(p) can easily be calculated and g(0) = g0 is the flat metric. In
dimension 4, this metric ĝ(Λ) depends on one parameter Λ and when Λ = 0 we get
M (0) isomorphic to H ' C2 with the standard flat metric. Hence, (M (Λ), g(Λ))
is a smooth 1-parameter family of HK deformations of the Euclidean metric. The

2This is not a special feature of this example. On the contrary, this is what typically happens
with HK reductions of Hn by compact hyperhamiltonian G-actions.
12.9. TORIC HYPERKÄHLER METRICS 461

metric g(Λ) is called the Taub-NUT3 gravitational instanton and it has interesting
history. Just as the famous Schwarzschild metric, or Kerr solution, the metric
appears first in the Lorenzian signature. One can always perform the so-called
“Wick rotation” changing t 7→ it which locally gives a Riemannian metric with
similar properties. However, there is no reason for the Riemannian metric to extend
globally to a complete metric on some manifold. This is fairly rare and happens
only in special situations. It was Hawking who observed that this indeed is the
case for the Lorenzian Taub-NUT solution, giving rise to a complete Ricci flat
metric on R4 [Haw77]. For some time thereafter the metric was not really fully
understood, there being claims in the literature is that this metric was not Kähler.
We should point out that this is the only known complete Ricci-flat Kähler metric
on C2 apart from the standard one [LeB91a]. If one imposes a Euclidean volume
growth condition the only known example of such complete Ricci-flat Kähler (or
just Ricci-flat) metric on C2 is the Euclidean metric.

12.9. Toric Hyperkähler Metrics


It is easy to see that previous two examples fall into a special category of com-
plete hyperkähler manifolds: they admit n commuting hyperholomorphic Killing
vector fields, where n is the quaternionic dimension. Following Bielawski and
Dancer we consider [BD00, Bie99]
Definition 12.9.1: An HK manifold (M 4n , I, g) is locally toric if it admits
n commuting hyperholomorphic Killing vector fields, linearly independent at each
point x ∈ M 4n , i.e., locally M admits a free action of Rn by hyperholomorphic
isometries. Furthermore, (M 4n , I, g) is said to be a toric HK manifold if it is a
hyperhamiltonian T n -space.
We emphasize that ‘toric’ is understood here to be in the quaternionic sense.
A local description of such metrics in dimension 4 is due to Gibbons and Hawking
[GH78a] and in arbitrary dimension 4n it has been generalized by Lindström and
Roček [LR83]. The so-called Legendre transform method developed by Lindström
and Roček associates a 4n-dimensional hyperkähler metric with n commuting hy-
perholomorphic Killing vectors to every real-valued function F on an open subset
U ⊂ R3 ⊗ Rn which is harmonic on any affine 3-dimensional subspace L of the
form R3 ⊗ Rv, v ∈ Rn (such functions are sometimes called polyharmonic). The
construction proceeds as follows: Let us identify R3 ⊗ Rn with Rn × Cn and let
(x, z) ∈ Rn × Cn be coordinates on U . Given any polyharmonic function F (z, z̄) on
U we consider a real-valued function
n
X
(12.9.1) K(u, ū, z, z̄) = F (x, z, z̄) − 2 (ui + ūi )xi ,
i=1

where the xi are determined by


∂F
(12.9.2) = 2(ui + ūi ) .
∂xi
It is an elementary exercise to check that polyharmonicity of F turns K into a
Kähler potential of an HK metric. Furthermore, if we set y = i(ū − u) then

3The acronym NUT stands for Newmann-Unti-Tamburino and has become standard termi-
nology for describing a certain type coordinate singularity of the Einstein equations.
462 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

Xi = ∂/∂yi , i = 1, . . . , n yield n commuting hyperholomorphic Killing vector


fields with the corresponding hyperkähler moment maps
(12.9.3) µi = (µi1 , µi+ ) = (xi , zi ) , i = 1, . . . , n.
One can show that, relative to local coordinates (y, x, z), the HK metric takes
the form [PP88]
X³ ´
(12.9.4) g= Φij (dxi dxj + dzi dz̄j ) + (Φ−1 )ij (dyi + Ai )(dyj + Aj ) ,
i,j
√ P
where Φij = 41 Fxi xj and Aj = 2−1 l (Fxj z̄l dz̄l − Fxj zl dzl ). The functions Φij
are also polyharmonic. The n × n matrix [Φij ] locally determines the hyperkähler
and hyperhamiltonian structure. When n = 1 this is well-known Gibbons-Hawking
Ansatz. Renaming (y1 , x1 , z1 ) = (t, x1 , x2 + ix3 ) = (t, x), A1 = α · dx, Φ11 = V we
can write the metric in a more familiar form
(12.9.5) g = V (dx · dx) + V −1 (dt + α · dx)2 ,
where gradV = curlα. In particular, V is a solution of the Laplace equation so that
we can write
Xk
mi
(12.9.6) V (x1 , x2 , x3 ) = δ + .
i=1
|x − ai |
When δ = 0 these metrics are called k-center gravitational multi-instantons. The
first two values k = 1, 2 correspond to the Euclidean and Eguchi-Hanson metrics,
respectively. For larger values of k it is not easy to determine when the metric
is actually complete and even harder to see what the manifold Mk on which it is
defined is. When δ = 1 we get the so-called k-center Taub-NUT gravitational multi-
instantons with k = 1 corresponding to the Euclidean Taub-NUT metric discussed
in Example 12.8.6.
The two basic examples of this construction are flat S 1 -invariant metrics on
S × R3 and on H. In the first case we have
1

(12.9.7) F (x, z, z̄) = 2x2 − z z̄


and, consequently, Φ ≡ 1, while in the second case
(12.9.8) F (x, z, z̄) = x ln(x + r) − r ,
2 2
where r = x + z z̄ with Φ = 1/4r. More general forms are given in [BD00]. In
the latter, the functions F and the metrics for hyperkähler quotients of flat vector
spaces are computed. They are essentially obtained by taking linear combina-
tions and compositions with linear maps of the solution (12.9.8). Bielawski shows
[Bie99] that, in the case of a complete metric, the only other possibility is adding
a linear combination of (12.9.7), which corresponds to a Taub-NUT deformation of
Definition 12.9.4.
For an HK metric of the form (12.5.6) taking hyperkähler quotients by subtori is
simple. The moment map equations are now linear (in xi , zi ), and the hyperkähler
quotient corresponds to restricting the function F to an appropriate affine subspace
of R3 ⊗ Rn . In fact, the requirement that F be polyharmonic is a consequence of
the fact that we must be able to take hyperkähler quotients by any subtorus.
An explanation of this construction in terms of twistors was given by Hitchin,
Karlhede, Lindström and Roček [HKLR87]. In particular, they have shown that
12.9. TORIC HYPERKÄHLER METRICS 463

any hyperkähler 4n-manifold with a free hyperhamiltonian Rn -action which extends


to a Cn -action with respect to each complex structure and such that the moment
map is surjective is given by the Legendre transform. In fact, one can show [Bie99]
that the Legendre transform provides a complete local description of such metrics,
i.e.,
Proposition 12.9.2: Let (M 4n , g) be a locally toric HK manifold. Then g is locally
given by Equation 12.9.4.
The key to the understanding of the global properties of such metrics is the
HK quotient construction. The relevant spaces were first introduced in [HKLR87],
but the global properties of such metrics were studied only much later by Bielawski
and Dancer [BD00] culminating in a complete classification result by Bielawski
[Bie99].
We will discuss this classification here as it will be important in the next chap-
ter. First, we would like to identify hyperhamiltonian G-manifolds which are the
same in the sense of the following definition.
Definition 12.9.3: Let M, M 0 be two hyperhamiltonian G-manifolds and let µ, µ0
be the chosen moment maps. We say that M and M 0 are isomorphic as hy-
perkähler G-manifolds, if there is a hyperholomorphic G-equivariant isometry f :
M → M 0 such that µ = µ0 ◦ f.
Secondly, there is a natural relation between the flat metric and the Taub-NUT
metric on H. This is, however, only an example of a more general correspondence.
The construction of the Example 12.8.6 suggest the following definition.
Definition 12.9.4: Let M 4n be a connected complete HK manifold of finite topo-
logical type with an effective hyperhamiltonian action of G = Rp × T n−p . A Taub-
NUT deformation (of order m) of M is the hyperkähler quotient of M × Hm by
Rm , where Rm acts on M via an injective linear map ρ : Rm → Lie (T n ) = Rn .
Note that such a deformation M 0 is canonically T n -equivariantly diffeomorphic
to M by a diffeomorphism f which respects the hyperkähler moment maps µ, µ0 ,
i.e., µ = µ0 ◦ f. Bielawski shows that up to G-equivariant isometry and Taub-NUT
deformations one can restrict attention to the following HK quotients studied in
detail in [BD00]. We consider a T k ⊂ T m+1 ⊂ Sp(m + 1) action on Hm+1 defined
ω θ
via exact sequence 0 −−→ h −−→ Rm+1 −−→ Rm+1−k −−→ 0, and its dual, as in
(12.5.2, 12.5.3), and with m + 1 − k = n. This is a hyperhamiltonian T k -action for
any choice of the weight matrix Ω ∈ Mk,m+1 (Z) or, alternatively, Θ ∈ Mn,m+1 (Z).
We can consider the HK moment map for this action µΩ : Hm+1 −−→R3 × Rk

m+1
X
(12.9.9) µΩ (u) = (ūα iuα )ω α + c ,
α=1

where
m+1
X
(12.9.10) c= λα ω α
α=1

and λα = λ1α i + λ2α j + λ3α k are purely imaginary quaternions (or vectors in R3 ) for
α = 1, . . . , m + 1. In particular, c ∈ R3 ⊗ Rk is simply an arbitrary choice of the
constant in the definition of the HK moment map. Let us denote by λ the “moment
464 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

level set” data, i.e.,


λ ≡ λ1 i + λ2 j + λ3 k = (λ1 , . . . , λm+1 ) ∈ R3 ⊗ Rm+1 .
Consider now the HK quotient space, i.e.,
(12.9.11) M (Ω, c) ≡ M (Θ, λ) = µ−1 (0)/T k .
Clearly, the HK quotient M (Θ, λ) locally inherits a hyperhamiltonian T n -
action. However, it can be singular and the stratification depends on the choice of
the quotient data. Note that either choosing (Θ, λ) or (Ω, c) completely determines
both the T k action and the associated HK quotient. In fact, these two descriptions
are dual to one another and they are both useful. Consider the following codimen-
sion three affine subspaces in R3 ⊗ Rn :

(12.9.12) Hα = {(x1 , x2 , x3 ) ∈ R3 ⊗ Rn ; hxi , θ α i = λiα , i = 1, 2, 3} ,


α = 1, . . . , m + 1 and θ α ∈ Rn is the αth column of Θ.
Bielawski and Dancer prove the following theorem [BD00]
Theorem 12.9.5: Suppose the column vectors θ α of Θ ∈ Mn,m+1 (Z) are primitive
integer vectors spanning Rn . Suppose λ ∈ R3 ⊗Rm+1 is such that Hα are all distinct.
Then the HK quotient M (Θ, λ) is smooth if and only if
(i) every collection of n + 1 of the Hα have empty intersection, and
(ii) whenever some n of the Hk1 , . . . , Hkn have non-empty intersection, then
the set {θ k1 , . . . , θ kn } is a Z-basis for Zn .
The first condition is sufficient for M (Θ, λ) to be an HK orbifold with at worst
Abelian quotient singularities.
If condition (i) of Theorem 12.9.5 holds the orbifolds M (Θ, λ) and M (Θ, λ0 ) are
homeomorphic, and if λ0 = (λ1 , 0, 0), they are diffeomorphic. In particular, one can
always set λ2 = λ3 = 0 while investigating the topology of such quotients. Bielawski
and Dancer have given a formula for the Betti numbers of M (Θ, λ) in terms of
arrangements of certain hyperplanes. Consider the collection of hyperplanes
Hα1 = {y ∈ Rn ; hx, θ α i = λ1α , i = 1, 2, 3} ,
i.e., Hα1 in Rn defined by restricting Hα to the first coordinate in R3 factor. These
hyperplanes divide Rn into a finite family of closed convex polyhedra. Let A be
the polyhedral complex consisting of all faces of all dimensions of these polyhedra
and let C be the polyhedral complex consisting of all bounded polyhedra in A. We
have the following
Theorem 12.9.6: Let M = M (Θ, λ) be a toric HK orbifold of dimension 4n.
j
(i) Then M is simply connected ¡ ¢ and H (M, Q) = 0 for j odd;
Pn i−p i
(ii) b2p (M ) = i=p (−1) p di , where the integer di denotes the number of
i-dimensional elements of the complex C;
(iii) if M = M (Θ, λ) is a smooth manifold, there is a ring isomorphism
H ∗ (M, Z) ≈ Z[u1 , . . . , uN ], where ui are the first Chern classes of certain
complex line bundles on M.
Parts (i) and (ii) of this theorem are due to Bielawski and Dancer [BD00]
while (iii) is due to Konno [Kon00]. In [Bie99] it is shown that the quotients
M = M (Θ, λ) are essential part of the classification of locally toric HK manifolds.
On can show that
12.10. ALE SPACES AND OTHER HYPERKÄHLER QUOTIENTS 465

Theorem 12.9.7: Let M 4n be a connected complete hyperkähler manifold of finite


topological type with an effective hyperhamiltonian action of G = Rp × T n−p . Then
(i) If M is simply connected and p = 0, then M is isomorphic, as a hyper-
hamiltonian hyperkähler T n -manifold, to a hyperkähler quotient of some
flat Hd × Hm , m ≤ n, by T d−n × Rm .
(ii) If M is simply connected and p > 0, then M is isomorphic, as a hyper-
hamiltonian hyperkähler G-manifold, to the product of a flat Hp and a
4(n − p)-dimensional manifold described in part (i).
¢l
(iii) If M is not simply connected, then M is the product of a flat (S 1 × R3 ,
1 ≤ l ≤ n, and a 4(n − l)-dimensional manifold described in part (ii).
In the case of dimension four we can be more specific.
Corollary 12.9.8: Let M be a simply connected 4-dimensional complete hyperkähler
manifold with a non-trivial hyperhamiltonian vector field. If b2 (M ) = k > 0, then
M is isometric either to an ALE-space of type Ak (i.e., a multi-Eguchi-Hanson
space) or to its Taub-NUT-like deformation (i.e., to the hyperkähler quotient by R
of the product of such a space with H). If b2 (M ) = 0, then M is either the flat H
or it is the Taub-NUT metric on R4 .
In particular, Bielawski concludes the following HK analogue of the Delzant’s
theorem for complex toric manifolds:
Theorem 12.9.9: Complete connected HK hyperhamiltonian T n -manifolds of fi-
nite topological type and dimension 4n are classified, up to Taub-NUT deforma-
tions, by arrangements of codimension 3 affine subspaces Hα in R3 ⊗ Rn defined as
in (12.9.12) by
Hα = {(x1 , x2 , x3 ) ∈ R3 ⊗ Rn ; hxi , θ α i = λiα , i = 1, 2, 3}
for some finite collection of vectors θ α in Rn and scalars λiα , i = 1, 2, 3, such that,
for any p ∈ R3 ⊗ Rn , the set {θ α ; p ∈ Hα } is part of a Z-basis of Zn .
We end this section by briefly mentioning some fascinating recent work by
Hausel, Nakajima, Strurmfels and others that relates hyperkähler geometry, and
toric hyperkähler structures in particular to combinatorics and representation the-
ory [HS02, Nak98, Nak99] as well as to the famous ADHM construction of in-
stanton moduli spaces [BM93b, Nak99, Hau06] and to number theory [Hau05,
Hau06].

12.10. ALE Spaces and Other Hyperkähler Quotients


In this last section we shall describe some other examples of hyperkähler met-
rics with particular focus on examples relevant to 3-Sasakian geometry. Our first
goal is to describe quotient construction of ALE spaces discovered by Kronheimer
[Kro89a, Kro89b].

12.10.1. Classical McKay Correspondence. We begin by recalling some


elementary facts about discrete subgroups of SU (2), their representation, and the
classical McKay correspondence. These were already discussed in the proof of
Theorem 10.1.5, where they were denoted by: Zn , D∗n , T∗ , O∗ , I∗ . Let H = C2 be
the standard complex 2-dimensional representation of SU (2). In particular, H gives
a representation of each Γ ⊂ SU (2). Let {ρ0 , ρ1 , . . . , ρr } be the set of irreducible
466 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

representation of Γ with ρ0 the trivial representation. Then


Xr
(12.10.1) ρi ⊗ H = aij ρj .
i=0
McKay observed that there is a relation between the matrix A = (aij ), whose
entries are all either 0 or 1, is the adjacency matrix of the extended and Dynkin
diagram of type A-D-E. If C̃ is the Cartan matrix of the extended Dynkin diagram
then A + C̃ = 21lr+1 . Let {θ1 , . . . , θr } be the simple roots of the root system of the
associated Lie algebra. Let θ0 be the negative of the highest root. McKay further
noticed that
X r
θ0 = − ni θ i ,
1
where ni is the dimension of the representation ρi [McK80, McK81]. The regular
representation ρ of Γ decomposes as
M
ρ= Cni ⊗ ρi .
i
Separately for each Γ, all of this information can now be encoded in a “labelled”
extended simply-laced Dynkin graph.

'&%$
Ã!"#
1@A '&%$
Ã!"#
1 · · ·'&%$
Ã!"#
1 BC'&%$
Ã!"#
1
Zn+1 e n, n ≥ 0 :
A

'&%$
Ã!"#
1 PP '&%$
Ã!"#
nnn
1
P'&%$
D∗n e n, n ≥ 4 :
D
nnnÃ!"#
2 '&%$
Ã!"#
2 · · ·'&%$
Ã!"#
2 '&%$
Ã!"#
2 PP
P'&%$
Ã!"#
'&%$
1 Ã!"#
1

'&%$
Ã!"#
1

T∗ e6 :
E '&%$
Ã!"#
2

'&%$
Ã!"#
1 Ã!"#
'&%$
2 '&%$
Ã!"#
3 '&%$
Ã!"#
2 '&%$
Ã!"#
1

'&%$
Ã!"#
2
O∗ e7 :
E
'&%$
Ã!"#
1 Ã!"#
'&%$
2 '&%$
Ã!"#
3 '&%$
Ã!"#
4 '&%$
Ã!"#
3 '&%$
Ã!"#
2 '&%$
Ã!"#
1

'&%$
Ã!"#
3
I∗ e8 :
E
'&%$
Ã!"#
2 Ã!"#
'&%$
4 '&%$
Ã!"#
6 '&%$
Ã!"#
5 '&%$
Ã!"#
4 '&%$
Ã!"#
3 '&%$
Ã!"#
2 '&%$
Ã!"#
1

The vertices of the diagram correspond to the irreducible representations ρi with the

numbers in each vertex giving the dimension of that representation ni . The usual
Dynkin diagram is obtained from the extended one by removing one vertex which
corresponds to the trivial representation ρ0 . In particular, McKay’s observations
show that
Xr X
n2i = |Γ|, 2aij ni nj = 4|Γ| .
i=0 ij
12.10. ALE SPACES AND OTHER HYPERKÄHLER QUOTIENTS 467

Example 12.10.1: Consider the example of Γ = Zn+1 . Let τ n+1 = 1. Any irre-
ducible representation of Zn+1 is one-dimensional and ρi (x) = τ i x, x ∈ C. Now, the
two-dimensional representation H gives H(τ )·(x, y) = (τ x, τ̄ y). Clearly, H = ρ1 ⊕ρn
so that ρi ⊗ H = ρi+1 ⊕ ρi−1 . The vertex ρi of the McKay graph is joined by an
edge to the vertices ρi+1 and ρi−1 . This defines the extended Dynkin diagram A e n.

Exercise 12.1: Consider the example of Γ = Dn . As a subgroup of SU (2) the
binary dihedral group4 is generated by two matrices
µ ¶ µ ¶
0 −1 τ 0
(12.10.2) , , τ 2(n−2) = 1 .
1 0 0 τ̄
Note that for n = 3 we simply get the cyclic group Z4 . This gives the represen-
tation H. In particular, D∗4 is the group of quaternions Q = {±1, ±i, ±j, ±k}.
D∗4 has three non-trivial one-dimensional irreducible representation {ρ1 , ρ2 , ρ3 } and
one two-dimensional representation ρ4 . “Derive” the McKay’s graph for D∗4 as in
Example 12.10.1. Repeat this for any binary dihedral group D∗n .
12.10.2. Geometric McKay Correspondence and Kleinian Singulari-
ties. McKay’s observation is closely related to the algebraic geometry of Kleinian
singularities which relation we shall explain next. This is often referred to in the
literature as the geometric McKay correspondence.
Definition 12.10.2: For Γ ⊂ Sp(1) a finite subgroup, the quotient variety X =
C2 /Γ = SpecC[x, y]Γ is called a Kleinian singularity (also known as a simple
surface singularity, or a rational double point, or A-D-E type singularity).
The quotient can be embedded as a hypersurface X ⊂ C3 with an isolated
singularity at the origin with the defining equation f (z0 , z1 , z2 ) = 0 determined by
the conjugacy class of Γ. These polynomials have already appeared in the table of
Remark 10.1.1. Suppose now π : M −→X = C2 /Γ is a crepant resolution. Then the
divisor ∆ = π −1 (0) is the dual of the associated Dynkin diagram in the following
sense: the vertices of the Dynkin diagram correspond to rational curves Di with
self-intersection −2. Two curves intersect transversally at one point if and only if
the corresponding vertices are joined by an edge in the Dynkin diagram. Otherwise
they do not intersect. The collection of these curves {D1 , . . . , Dr } forms a basis for
H2 (M, Z). The intersection form with respect to this basis is the negative of the
Cartan matrix C.
12.10.3. Kronheimer-McKay Correspondence and Hyperkähler ALE
Spaces. Hitchin observed that in the case of Γ = Zn the crepant resolution
π : M −→C2 /Zn admits a family of complete hyperkähler metrics [Hit79]. In
fact, locally these metrics are produced via the Gibbons-Hawking Ansatz. Using
twistor methods Hitchin showed that Gibbons-Hawking gravitational instantons,
as hyperkähler ALE spaces, are the minimal resolution of singularity C2 /Zn . In
particular, the minimal resolution of the singularity C2 /Z2 is the cotangent bun-
dle T ∗ CP1 and the hyperkähler metric is the Eguchi-Hansom metric. Hitchin then
conjectured that such metrics should exist for all other spaces π : M −→C2 /Γ. It
was only after the discovery of the hyperkähler reduction and description of its
mathematical foundations in [HKLR87] that the conjecture was finally proved by
Kronheimer [Kro89a, Kro89b]. Kronheimer generalizes the quotient construction
4Recall that our notation is not completely standard. ∗ =
Our binary dihedral group Dn
Z2(n−2) o Z2 has order 4(n − 2) and not 4n.
468 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

described in [HKLR87] for Γ = Zn to the arbitrary group Γ ⊂ Sp(1). Again, the


absolute key is the McKay correspondence. Not surprisingly, the quotient can be
completely described by the extended Dynkin diagram associated to Γ. In a way,
Kronheimer’s quotient construction is “the third McKay’s correspondence”. To
make our statement more precise late us begin with a precise definition of an ALE
space.
Definition 12.10.3: Let Γ ⊂ Sp(1) and let (H, I + , g0 ) be the standard flat (left)
hyperkähler structure on H defined in Section 12.1. Let r : H/Γ−−→R≥0 be the
radius function on H/Γ. We say that an HK manifold (M, I, g) is Asymptotically
Locally Euclidean (ALE), and asymptotic to H/Γ, if there exists a compact
subset X ⊂ M and a map π : M \ X−→H/Γ that is a diffeomorphism between
M \ X and {x ∈ H/Γ|r(x) > R} for some R > 0 such that
∇k (π∗ (g) − g0 ) = O(r−4−k ) , ∇k (π∗ (I(τ )) − I0− (τ )) = O(r−4−k )
as r → ∞ and k ≥ 0, where ∇ is the Levi-Civita connection of the flat metric g0 .
Consider any extended Dynkin diagram ∆(Γ). ˜ With each vertex of ∆(Γ) ˜ we
associate the unitary group U (ni ) and with each edge the vector space Hni nj =
Hom(Cni , Cnj ) ⊕ Hom(Cnj , Cni ). One can think of each edge as the vector space
Mni ,nj (H) of quaternionic matrices. For each Γ, we define
M
(12.10.3) H|Γ| = aij Hom(Cni , Cnj ) , G(Γ) = U (n0 ) × · · · × U (nr ) .
i,j

The group G(Γ) acts naturally on H|Γ| and the action is hyperhamiltonian
with appropriately defined flat hyperkähler structure on each “edge”. However, the
action is not effective, hence, we take the quotient K(Γ) = G(Γ)/T, where T is
the central U (1) ⊂ G(Γ). The action of K(Γ) is then effective and it defines the
hyperkähler moment map

(12.10.4) µΓ : H|Γ| −−→k∗ ⊗ sp(1) .


Definition 12.10.4: We say that K(Γ)-invariant element ξ ∈ k∗ ⊗ sp(1) is in a
good set if the K(Γ)-action on the ξ-level set µ−1 Γ (ξ) of the moment map is free.
The notion of the good set is generic. The set Z of K(Γ)-invariant elements in
k∗ can be identified with the dual of the center. Kronheimer shows that ξ is not
in the good set if ξ ∈ Dθ ⊗ sp(1) ⊂ Z ⊗ sp(1), where Dθ are the walls of the Weyl
chamber. The reduction gives the following theorem [Kro89a, Kro89b]
Theorem 12.10.5: Let ξ ∈ k∗ ⊗ sp(1) be a G(Γ)-invariant element. Let M (Γ, ξ)
be the hyperkähler reduction of H|Γ| by the action of K(Γ) and with the momentum
level set ξ. Then M (Γ, 0) ' C2 /Γ and M (Γ, ξ) is a hyperkähler orbifold for any ξ.
In addition, when ξ is in the good set µ−1Γ (ξ) the smooth manifold M (Γ, ξ) gives a
family of complete ALE hyperkähler metrics on the crepant resolution of singularity
C2 /Γ.
When ξ is not in the good set, the HK orbifold M (Γ, ξ)−→C2 /Γ is a partial
resolution of the quotient singularity. In [Kro89b] Kronheimer shows that his
construction is also complete. That is any hyperkähler ALE space of Definition
12.10.3 is obtained as such quotient. The quotient metrics are known in local charts
for Γ = Zn (the Gibbons-Hawking Ansatz, see (12.9.5)) as they always have non-
trivial isometries. For the non-Abelian Γ the ALE metrics have no Killing vectors.
12.10. ALE SPACES AND OTHER HYPERKÄHLER QUOTIENTS 469

Recently, Cherkis and Hitchin gave explicit formulas for the ALE gravitational
instantons in the binary dihedral case [CH05].
Remark 12.10.1: The Kleinian singularities and discrete groups of SU (2) also give
one-to-one correspondence with all compact 3-Sasakian manifolds in dimension 3.
We shall discuss this in the next chapter.

12.10.4. Other Hyperkähler Metrics. In the last 25 years hyperkähler ge-


ometry has become an important field of Riemannian geometry. Already over a
decade ago, in a Séminaire Bourbaki review article Hitchin points out that the
richness of the theory of hyperkähler manifolds, in some sense, vindicates Hamil-
ton’s conviction that quaternions should play a fundamental role in mathematics
and physics [Hit92]. As it happens, many new ideas in this field have came from
mathematical physics. In this chapter we have only covered a small number of
selected topics most relevant to the material of Chapter 13. We would like to end
it with a brief discussion of several subjects we were unable to introduce. We re-
fer the interested reader to several books and review articles about the subject
[Hit87b, AH88, Hit95a, Dan99, VK99].
• [Instantons, Monopoles, and Stable Pairs] Many hyperkähler metrics emerge
in the description of the geometry of various moduli spaces. In 1983 Atiyah and
Bott made a fundamental observation that the moduli space of self-dual Yang-Mills
connections over a Riemann surface can be described as an infinite dimensional
Kähler quotient [AB83]. In such picture one first equips the space of all connec-
tions with a structure of an infinite dimensional Kähler manifold. On it acts an
infinite dimensional group of gauge transformations and the moment map for this
action is precisely the self-duality equation. Hence, the Kähler quotient, which
in turns out to be finite-dimensional, is the space of self-dual Yang-Mils connec-
tions modulo the gauge equivalence. When a Riemann surface is replaced by a
four-dimensional manifold (such as R4 or S 4 , for example) the space of connections
can be given a structure of an infinite-dimensional hyperkähler manifold. The HK
moment map for the gauge group action produces the self-duality equations and
the HK quotient can be naturally identified with the moduli space of instantons.
This picture explains why the k-instanton moduli spaces over S 4 carry a natu-
ral HK structure [AHDM78]. Yang-Mills connections on other 4-manifolds were
studied since. For instance, Kronheimer and Nakajima considered instantons over
the ALE spaces discussed earlier [KN90, Nak90]. The infinite dimensional HK
quotient picture is inherently present in a variety of different moduli problems. One
important case is the moduli space of solutions of the Bogomolny equations on a
3-manifold. These are known as monopoles. They can be viewed as Yang-Mills con-
nections which a translational symmetry. Many complete hyperkähler metrics have
been constructed as solutions of the Bogomolny equations or the related Nahm’s
equations [Nah82, AH88, Dan93, Dan94]. For instance, in the SU (2) case,
the universal cover M̃k0 of the moduli space of charge k monopoles with a fixed
center is a complete hyperkähler manifold of dimension 4(k − 1) [AH88]. The four-
dimensional case of M̃20 gives the famous SO(3)-invariant Atiyah-Hitchin monopole
metric [AH85]. Like in the Taub-NUT case the SO(3)-action acts by rotations on
the 2-sphere of complex structures I but the metric is quite different as it is not
toric: it has no hyperhamiltonian Killing vectors. Finally, we mention the mod-
uli space of R2 -invariant Yang-Mills equation. Such reduction naturally leads to
470 12. QUATERNIONIC KÄHLER AND HYPERKÄHLER MANIFOLDS

the moduli space of the so-called stable pairs or Yang-Mills-Higgs fields over an
arbitrary Riemann surface and was considered by Hitchin in [Hit87b].
• [Hyperkähler Manifolds of Type A∞ and D∞ ] The first example of a com-
plete hyperkähler manifold of infinite topological type was obtained by Anderson,
Kronheimer and LeBrun [AKL89]. They showed that one can take k → ∞ limit
in the Gibbons-Hawking Ansatz (12.9.5) under an appropriate assumption about
the distribution of the mass centers. Later these metrics were considered from an
algebraic viewpoint by Goto who also showed that a similar limit can be taken in
the D∗n ALE case [Got98].
• [Hyperkähler Deformations of ALE Spaces] Taub-NUT deformations of the
ALE gravitational instantons corresponding to Γ = Zn were discussed in Section
12.9. These metrics are no longer ALE but the are asymptotically locally flat.
Other ALE spaces do not admit any Taub-NUT deformations in the sense of Def-
inition 12.9.4 as they have no hyperholomorphic isometries. But the ALE spaces
of the binary dihedral group admit deformations similar to Taub-NUT deforma-
tions. Recall, that we can view the Taub-NUT deformation of the Zn ALE space as
follows: consider the space HΓ and replace one H-factor with the flat hyperkähler
manifold S 1 × R3 modifying the action of K(Γ) to be the translation on S 1 . Con-
sider a non-Abelian Γ and suppose we could replace the Euclidean metric on some
edge Hni nj with a non-Euclidean hyperkähler metric on Hni nj which, however, ad-
mits hyperhamiltonian action of U (ni ) × U (nj ). Dancer observed that there is one
such case: a complete U (2) × U (2)-invariant hyperkähler metric on H4 considered
as T ∗ G(2, C) [Dan93]. The metric is obtained as a monopole moduli space by
solving Nahm’s equations. Dancer shows that replacing the flat H4 with one copy
of M = T ∗ G(2, C) together with its monopole metric and then performing the HK
quotient of Kronheimer gives non-trivial deformations of the ALE metrics for each
D∗n , n ≥ 4. It is not clear if similar deformations exist for any of the ALE spaces of
the E-series.
• [Hyperkähler Metric on Coadjoint Orbits] Kronheimer showed that there
is a natural hyperkähler metric on a regular semisimple coadjoint orbits of a com-
plex group Lie group GC [Kro90a]. Kronheimer also proved that nilpotent orbits
admit hyperkähler structure [Kro90b]. Later Biquard [Biq96] and independently
Kovalev [Kov96] showed that there is a hyperkähler structure on any coadjoint
orbit of GC .
• [Hyperkähler Metrics on Cotangent Bundles] The first explicit non-trivial
example of a hyperkähler metric is the Calabi metric on the cotangent bundle
T ∗ CPn . Already in 1983 Lindström and Roček constructed a hyperkähler met-
ric on the cotangent bundle T ∗ Grn (Cn+m ) which is realized as a reduction of the
flat space Hm(n+m) by a hyperhamiltonian action of U (m) [LR83]. In fact, Lind-
ström and Roček derived an explicit formula for the Kähler potential of this metric
generalizing the formula given by Calabi [Cal79]. It is not surprising that Calabi-
Lindström-Roček metrics are only a special case. The hyperkähler metrics on coad-
joint orbits are complete if and only if the orbit is semisimple. These orbits are
then diffeomorphic to the cotangent bundle of flag manifolds for G. In such cases
one can write down the metric and Kähler potentials explicitly [BG97a, DS97b].
• [Compact Hyperkähler Manifolds] Much work has been done on the geometry
of compact hyperkähler manifolds with many new examples in higher dimensions.
12.10. ALE SPACES AND OTHER HYPERKÄHLER QUOTIENTS 471

We refer the reader to several extensive reviews on the subject and references therein
[GHJ03, VK99, NW04].
• [Quaternionic Geometries with Torsion] These structures arose from the
attempts of physicists to incorporate the so-called bosonic Wess-Zumino-Witten
term [WZ71, Wit83] in σ-models with (extended) supersymmetry. The idea of
considering torsion connections in such σ-models dates back to the early 80ties
(see [HS84] (Kähler with torsion) and [GHR84] (hyperkähler with torsion)). The
manifolds involved are not Kähler or hyperkähler, but Hermitian and hyperher-
mitian. In 1996 Howe and Papadopoulos introduced a formal definition of the
so-called HKT geometry (hyperkähler with torsion) and studied twistor spaces of
such manifolds [HP96]. This sparked a considerable interest in such models also
among mathematicians (cf. [GP00, GGP03]). Today physicists and mathemati-
cians alike continue studying KT (Kähler with torsion), CYT (Calabi-Yau with
torsion), HKT, and even QKT (quaternionic Kähler with torsion) geometries. This
subject, although very interesting, goes far beyond the scope and the main focus
of our book. We refer the reader to a couple of extensive review articles on the
mathematical foundations, the history, and the bibliography [Gra04, Agr06].
CHAPTER 13

3-Sasakian Manifolds

In this chapter we consider a special kind of Sasaki-Einstein manifolds which


we shall call manifolds (several different names have been introduced and used in
the earlier literature on this subject). In 1970 Kuo [Kuo70] refined the notion
of a Sasakian structure and introduced manifolds with Sasakian 3-structures (see
also [KT70], [TY70]). Independently, the same concept was invented by Udrişte
[Udr69]. Between 1970 and 1975 this new kind of geometry was investigated al-
most exclusively by a group of Japanese geometers, including Ishihara, Kashiwada,
Konishi, Kuo, Tachibana, Tanno, and Yu. Already in [Kuo70] we learn that the
3-Sasakian geometry has some interesting topological implications. Using earlier re-
sults of Tachibana about the harmonic forms on compact Sasakian spaces [Tac65],
Kuo showed that odd Betti numbers up to the middle dimension must be divisible
by 4. In 1971 Kashiwada observed that every 3-Sasakian manifold is Einstein with
a positive Einstein constant [Kas71]. In the same year Tanno proved an interest-
ing theorem about the structure of the isometry group of every 3-Sasakian space
[Tan70]. In a related paper he studied a natural 3-dimensional foliation on such
spaces showing that, if the foliation is regular, then the space of leaves is an Ein-
stein manifold of positive scalar curvature [Tan71]. Tanno clearly points to the
importance of the analogy with the quaternionic Hopf fibration S 3 → S 7 → S 4 ,
but does not go any further. In fact, Kashiwada’s paper mentions a conjecture
speculating that every 3-Sasakian manifold is of constant curvature [Kas71]. She
attributed this conjecture to Tanno and, at the time, these were the only known
examples.
Very soon after, however, it became clear that such a conjecture could not
possibly be true. This is due to a couple of papers by Ishihara and Konishi [IK72],
[Ish73]. They made a fundamental observation that the space of leaves of the
natural 3-dimensional foliations mentioned above has a “quaternionic structure”,
part of which is the Einstein metric discovered by Tanno. This led Ishihara to
an independent study of this “sister geometry”: quaternionic Kähler manifolds
[Ish74]. His paper is very well-known and is almost always cited as the source of
the explicit coordinate description of quaternionic Kähler geometry. Among other
results Ishihara showed that his definition implies that the holonomy group of the
metric is a subgroup of Sp(n)·Sp(1), thus providing an important connection with
the earlier studies of such manifolds by Alekseevsky [Ale68], Bonan [Bon67], Gray
[Gra69a], Kraines [Kra65], and Wolf [Wol65]. In 1975 Konishi [Kon75] proved
the existence of a Sasakian 3-structure on a natural principal SO(3)-bundle over any
quaternionic Kähler manifold of positive scalar curvature. This, with the symmetric
examples of Wolf, gives precisely all of the homogeneous 3-Sasakian spaces. Yet,
at the time they did not appear explicitly and escaped any systematic study until
much later.

473
474 13. 3-SASAKIAN MANIFOLDS

In fact, 1975 seems to be the year when 3-Sasakian manifolds are relegated to an
almost complete obscurity which lasted for about 15 years. From that point on the
two “sisters” fair very differently. The extent of this can be best illustrated by the
famous book on Einstein manifolds by Besse [Bes87]. The book appeared in 1987
and provided the reader with an excellent, up-to-date, and very complete account
of what was known about Einstein manifolds 10 years ago. But one is left in the
dark when trying to find references to any of the papers on 3-Sasakian manifolds we
have cited; 3-Sasakian manifolds are never mentioned in Besse. The other “sister”,
on the contrary, received a lot of space in a separate chapter. Actually Einstein
metrics on Konishi’s bundle do appear in Besse (see [Bes87] 14.85, 14.86) precisely
in the context of the SO(3)-bundles over positive quaternionic Kähler manifolds
as a consequence of a theorem of Bérard-Bergery ([Bes87], 9.73). Obviously, the
absence of 3-Sasakian spaces in Besse’s book was the result rather than the cause
of this obscurity. One could even say it was justified by the lack of any interest-
ing examples. The authors have puzzled over this phenomenon without any sound
explanation. One can only speculate that it is the holonomy reduction that made
quaternionic Kähler manifolds so much more attractive an object. Significantly, the
holonomy group of a 3-Sasakian manifold never reduces to a proper subgroup of
the special orthogonal group. And when in 1981 Salamon [Sal82, Sal86], indepen-
dently with Bérard-Bergery [Bes87], generalized Penrose’s twistor construction for
self-dual 4-manifolds introducing the twistor space over an arbitrary quaternionic
Kähler manifold, the research on quaternionic Kähler geometry flourished, fuelled
by powerful tools from complex algebraic geometry.

13.1. Almost Hypercontact Manifolds and 3-Sasakian Structures


We first describe some special kinds of almost contact and contact structures
which are closely related to the quaternionic geometries discussed in the previous
chapter.
Definition 13.1.1: Let M be a manifold with a family of almost contact struc-
tures {ξ(τ ), η(τ ), Φ(τ )} parameterized by points τ = (τ1 , τ2 , τ3 ) ∈ S 2 ⊂ R3 on
a unit sphere. We say that {ξ(τ ), η(τ ), Φ(τ )}τ ∈S 2 is an almost hypercontact
structure on M if
Φ(τ ) ◦ Φ(τ 0 ) − η(τ ) ⊗ ξ(τ 0 ) = −Φ(τ × τ 0 ) − (τ · τ 0 )1l ,

Φ(τ )ξ(τ 0 ) = −ξ(τ × τ 0 ), η(τ ) ◦ Φ(τ 0 ) = −η(τ × τ 0 )


for any two τ , τ 0 ∈ S 2 . Furthermore, a Riemannian metric g on M is said to be
compatible with (or associated to) the almost hypercontact structure if
g(Φ(τ )X, Φ(τ )Y ) = g(X, Y ) − η(τ )(X)η(τ )(Y )
2
for all τ ∈ S . In such a case {ξ(τ ), η(τ ), Φ(τ ), g}τ ∈S 2 is called an almost hy-
percontact metric structure on M.
The above definition is a natural extension of the standard notion of the almost
contact metric structure given in Definition 6.3.1. Just as before one can easily show
that compatible metrics always exist. We can recover the standard definition of the
almost contact (metric) 3-structure due to [Kuo70] via a choice of an arbitrary
orthonormal basis {e1 , e2 , e3 } in R3 and by setting Φ(ea ) = Φa , ξ(ea ) = ξa , and
η(ea ) = ηa . Hence,
13.1. ALMOST HYPERCONTACT MANIFOLDS AND 3-SASAKIAN STRUCTURES 475

Definition 13.1.2: An almost contact (metric) 3-structure {ξa , ηa , Φa , (g)}3a=1


on M is an almost hypercontact (metric) structure together with a choice of an or-
thonormal frame on R3 .
We should note here that our definition allows for two different sign convention
which depend on the orientation of the chosen frame {e1 , e2 , e3 }. Both conventions
can be found in the literature. The original definition in [Kuo70] corresponds to
choosing left-handed frame. In this chapter we will always choose the opposite: the
standard right-handed basis.
One can easily extend the results of Section 6.5 by considering the cone C(M ) =
R+ × M together with
(13.1.1) I(τ )Y = Φ(τ )Y + η(τ )(Y )Ψ , I(τ )Ψ = −ξ(τ ) , ∀τ ∈ S 2 .
This clearly defines an almost hypercomplex structure I on C(M ) which is invariant
under Ψ and such that the vector field I(τ )Ψ is tangent to M for each τ ∈ S 2 .
In particular, a manifold with an almost hypercontact structure must be of
dimension 4n + 3. Conversely, if C(M ) is a cone with an almost hypercomplex
structure which is Ψ-invariant and such that I(τ )Ψ is tangent to M for all τ ∈ S 2
we can define an almost hypercontact structure on M by taking ξ(τ ) = I(τ )Ψ. For
each τ we then let Lξ(τ ) denote the trivial line bundle generated by ξ(τ ), we have
an exact sequence
0−−→Lξ(τ ) −−→T M −−→Q−−→0 ,
and there is a one-to-one correspondence between the splittings of this exact se-
quence and 1-forms η(τ ) on M that satisfy ξ(τ ) η(τ ) = 1. This correspondence
is given by η(τ ) 7→ ker η(τ ). Then the tensors Φ(τ ) are given by

(
I(τ ), on ker η(τ ) ,
(13.1.2) Φ(τ ) =
0. on Lξ(τ ) .

One arrives at the following straightforward generalization of Proposition 6.5.2


Proposition 13.1.3: There is a one-to-one correspondence between almost hyper-
contact structures {ξ(τ ), η(τ ), Φ(τ ), g}τ ∈S 2 on M and Ψ-invariant almost hyper-
complex structures I on C(M ) for which the vector field I(τ )Ψ is tangent to M
and Equation 13.1.2 holds for each τ ∈ S 2 . Furthermore, if g is a Riemannian
metric on M that is compatible with the almost hypercontact structure, then the
cone metric h = dr2 + r2 g is hyperhermitian with respect to the almost hypercom-
plex structure I, and the one-to-one correspondence above induces a one-to-one
correspondence between almost hypercontact metric structures on M and almost
hyperhermitian cone structures on C(M ).
The following proposition is in complete analogy with 6.3.2 and it was already
observed by Kuo.
Proposition 13.1.4: There is a one-to-one correspondence between almost hy-
percontact metric structures {ξ(τ ), η(τ ), Φ(τ ), g}τ ∈S 2 on M and reductions of the
frame bundle to the group Sp(n) × 1l3 .
In particular, a 3-manifold admits such a structure if and only if it is paralleliz-
able. It is known that every compact orientable 3-manifold is parallelizable and,
hence,
476 13. 3-SASAKIAN MANIFOLDS

Corollary 13.1.5: Every compact orientable 3-manifold admits an almost contact


3-structure.
In higher dimensions there are other obstructions. For example, almost hyper-
complex manifolds are strict almost contact manifolds, so Theorem 6.2.7 implies
that the odd Stiefel-Whitney classes an all integral Stiefel-Whitney classes must
vanish, but the reduction to Sp(n) implies more.
Proposition 13.1.6: Let M 4n+3 be a compact manifold with an almost contact 3-
structure. Then the only possible non-vanishing Stiefel-Whitney classes are w4i (M ).
In particular, M is a spin manifold. Furthermore, all the integral Stiefel-Whitney
classes Wi (M ) must vanish.

Proof. Because Sp(n) ⊂ SU (2n) ⊂ U (2n) Theorem 6.2.7 implies every-


thing but the vanishing of w4i+2 . For this we notice that T M splits as Da ⊕ 1l,
where Da = ker ηa for any a = 1, 2, 3. The group of the vector bundle Da is
Sp(n) × 1l2 . So its characteristic classes are pulled back from the universal classes
in si ∈ H 4i (BSp(n), Z). See Appendix A. By the Universal Coefficients Theo-
rem H ∗ (BSp(n), Z2 ) is also generated by classes in dimension 4i. In particular the
classes w4i+2 must vanish. ¤

The reduction of the structure group to Sp(n)×1l3 is equivalent to the existence


of a section of the associated bundle to L(M ) with fiber F4n+3 = SO(4n+3)/Sp(n).
In particular, one needs to compute certain cohomology groups H i (M, πi−1 (F4n+3 )).
This can be done, at least in principle. In dimension 7 the situation is somewhat
special due to the following theorem [FKMS97]
Theorem 13.1.7: Let M be a compact smooth 7-manifold and let SU (2) ⊂ SU (3) ⊂
G2 ⊂ Spin(7), where the three subgroups of Spin(7) are the stability subgroups fixing
one, two, or three spinors, respectively. Then the following conditions are equiva-
lent:
(i) M admits reduction of the structure group to SU (2),
(ii) M admits reduction of the structure group to SU (3),
(iii) M admits reduction of the structure group to G2 ,
(iv) w1 (M ) = w2 (M ) = 0, i.e., M admits a spin structure.
In particular, every compact oriented spin 7-manifold admits an almost hypercontact
structure.
The above theorem follows from the fact that any compact orientable 7-manifold
admits two independent vector fields [Tho69] which, on a spin manifold, can be
used to define three linearly independent spinors at every point of m ∈ M 7 . Hence,
M 7 must admit three orthogonal spinor fields which can be used to define SU (2)-
structure [FKMS97]. It is perhaps worth remarking that we get the following
interesting corollary.
Corollary 13.1.8: Let M be a compact oriented 7-manifold. If M is spin then the
integral third Stiefel-Whitney class W3 (M ) must vanish.
The above, by inspection, is also a property of compact smooth orientable 5-
manifolds which follows from the Barden-Smale Classification Theorem 10.2.3. But
the result here is much stronger as we do not need to assume simple connectivity.
From the point of view of Proposition 13.1.3 there several special cases one can
consider. We have
13.1. ALMOST HYPERCONTACT MANIFOLDS AND 3-SASAKIAN STRUCTURES 477

Definition 13.1.9: Let {ξ(τ ), η(τ ), Φ(τ ), g}τ ∈S 2 be an almost hypercontact metric
structure on M with the corresponding Ψ-invariant almost hyperhermitian structure
I on C(M ). We say that S = {ξ(τ ), η(τ ), Φ(τ ), g}τ ∈S 2 is a
(i) metric contact 3-structure if (C(M ), I, g) is almost hyperkähler,
(ii) 3-Sasakian if (C(M ), I, g) is hyperkähler.
We denote by S a 3-Sasakian structure on M . Alternatively, relative to a
choice of a right-handed orthonormal basis {e1 , e2 , e3 } in R3 we set Sa = S(ea )
and S = {S1 , S2 , S3 } and also refer to this triple of Sasakian structures as the
3-Sasakian structure.
Remark 13.1.1: Any almost HK structure is automatically HK (see Theorem
12.7.2) so it is elementary that any metric contact 3-structure is automatically
3-Sasakian. This was first observed by Kashiwada [Kas01] and it eliminates the
need for the term “hypercontact” (as in “almost contact” vs. “contact”) altogether.
In particular, also the notion of the so-called 3-K-contact structure exists only as
part of a 3-Sasakian structure which was proved by Tanno [Tan96] in dimension
7, and by Jelonek [Jel01] in dimensions ≥ 15.
Remarks 13.1.1: We must warn the reader that there several other inequivalent
definitions of hypercontact, or quaternionic contact structures. For example, Geiges
and Thomas [GT95] define a hypercontact structure as a triple of contact forms
{α1 , α2 , α3 } and a Riemannian metric g compatible with an almost contact 3-
structure so that the 2-forms Ψa (X, Y ) = g(X, Φa Y ) = dαa (X, Y ) for all a = 1, 2, 3
and all X, Y ∈ Γ(T M ). This definition coincides with ours under the additional
requirement that the contact forms αa = ηa . But without that assumption the
notion is much weaker. In particular, the Geiges-Thomas hypercontact property
is preserved by taking connected sums. In contrast, connected sums of 3-Sasakian
manifolds are typically not 3-Sasakian even for crude topological reasons (Betti
number constraints). The Geiges-Thomas hypercontact structures were also inves-
tigated by Banyaga [Ban96, Ban97].
Another generalization which has led to important discoveries that are beyond
the scope of this book is due to Biquard [Biq99, Biq00] who observed that the
conformal infinity of a complete QK metric on a 4n-dimensional manifold with
boundary has a codimension 3 distribution on the boundary. Biquard used this to
define what he called a quaternionic contact structure showing that in dimensions
greater than 7 his quaternionic contact structure is always the conformal infinity of
some QK metric. In dimension 7, however, this is no longer automatic. Duchemin
solved the problem by proving a criterion for quaternionic contact structures to
be the conformal infinity of an 8-dimensional QK metric[Duc06]. The 3-Sasakian
structures described in the present text appear as special cases of Biquard’s more
general theory. More recently similar structures were also studied by Alekseevsky
and Kamishima in [AK04], where they are called (para-)quaternionic CR struc-
tures.
Remark 13.1.2: Even though a hypercontact manifold is automatically 3-Sasakian
there are several meaningful weakenings of the 3-Sasakian condition. Starting with
an almost hypercontact metric structure one can consider the following conditions:
(i) The integrability conditions for the associated almost hypercomplex struc-
ture on the cone C(M ) were studied by Kehowski [Keh04] and in such
a case an almost hypercontact metric structure is called hypernormal.
478 13. 3-SASAKIAN MANIFOLDS

(ii) When dΨa = dηa = 0, such a structure on M is said to be almost


cosymplectic 3-structure, and if the almost hypercontact structure is hy-
pernormal then M is said to be cosymplectic 3-structure. It is easy to
see that an almost cosymplectic 3-structure is necessarily cosymplectic
[FIP04].
(iii) When dΨa = 0 and the almost hypercontact structure is hypernormal
then M is said to be quasi-3-Sasakian.
(iv) Another possible generalization is by localizing [MO00]. That is, one
considers each of the tensors fields in the 3-structure {ξa , ηa , Φa } to be
sections of a 3-dimensional vector bundle.
Remark 13.1.3: As a 3-Sasakian manifold has not just 3 but an S 2 ’s worth of
Sasakian structures, in analogy with the term hyperkähler, the name hypersasakian
would be more consistent. However, most of the existing literature (including all
papers the two authors have written on the subject) uses the name Sasakian 3-
structure or 3-Sasakian structure. We feel it is too late at this stage to start using
anything else in its place.

13.2. Basic Properties


Let M be a 3-Sasakian manifold so that C(M ) is hyperkähler with a hypercom-
plex structure I(τ ). As in the Sasakian case we define ξ(τ ) = −I(τ )(r∂r ) for each
τ ∈ S 2 and show that these can be used to reconstruct the 3-structure S on M.
One important consequence of the quaternionic structure on C(M ) is the relations
it implies for the 2-sphere of the characteristic vector fields ξ(τ ) :
Proposition 13.2.1: If S = {ξ(τ ), η(τ ), Φ(τ ), g}τ ∈S 2 is a 3-Sasakian structure
on M then g(ξ(τ ), ξ(τ 0 )) = τ · τ 0 , [ξ(τ ), ξ(τ )] = 2ξ(τ × τ 0 ) and Φ(τ ) = −∇ξ(τ ).
Conversely, if M has three Sasakian structures {S1 , S2 , S3 } with characteristic vec-
tor fields {ξ1 , ξ2 , ξ3 } such that g(ξa , ξb ) = δab and [ξa , ξb ] = 2²abc ξc , then S =
{S1 , S2 , S3 } is 3-Sasakian.
Exercise 13.1: Prove Proposition 13.2.1.
In particular, Proposition 13.2.1 implies that the characteristic vector fields
{ξ1 , ξ2 , ξ3 } generate a local isometric Sp(1)-action on M . Also combining Proposi-
tion 13.2.1 together with Lemma 8.1.17 implies
Corollary 13.2.2: Suppose (M, g) admits two Sasakian structures S1 , S2 with
characteristic vector fields ξ1 , ξ2 , respectively. Suppose g(ξ1 , ξ2 ) is a constant dif-
ferent from ±1. Then M admits a 3-Sasakian structure.
Since a hyperkähler manifold is Ricci-flat, Proposition 11.1.5 implies
Corollary 13.2.3: Every 3-Sasakian manifold (M, S) of dimension 4n + 3 is Ein-
stein with Einstein constant λ = 2(2n+1). Moreover, if M is complete it is compact
with finite fundamental group.
The important result that every 3-Sasakian manifold is Einstein was first ob-
tained by Kashiwada [Kas71] using tensorial methods.
If (M, S) is compact the characteristic vector fields {ξ1 , ξ2 , ξ3 } are complete and
define a 3-dimensional foliation FQ on M. The leaves of this foliation are necessarily
compact as {ξ1 , ξ2 , ξ3 } defines a locally free Sp(1) action on M . Hence, the foliation
FQ is automatically quasi-regular and the space of leaves is a compact orbifold. We
shall denote it by O. In addition, for any τ ∈ S 2 we can consider the characteristic
13.2. BASIC PROPERTIES 479

vector field ξ(τ ) associated with the direction τ . Since ξ(τ ) generates a 1-parameter
subgroup of Sτ1 ⊂ Sp(1), it defines a 1-dimensional foliation Fτ ⊂ FQ ⊂ M with
compact leaves. Hence, ξ(τ ) is quasi-regular and defines a locally free circle action
on M.
Definition 13.2.4: Let (M, S) be a compact 3-Sasakian manifold with 3-Sasakian
structure S. We say that S is regular if the Sasakian structures S(τ ) are regular
for all τ ∈ S 2 .
Remark 13.2.1: When dim(M ) > 3 the above definition is equivalent to regularity
of the foliation FQ . When dim(M ) = 3 the leaf space of the foliation FQ is a single
point so that FQ is always regular. However, S is not always regular.
We have the following observation of Tanno [Tan71] concerning regularity
properties of the foliations Fτ ⊂ FQ which is essentially a corollary of Theorem
13.3.1 below.
Proposition 13.2.5: Let (M, S) be a compact 3-Sasakian manifold. Every Sasakian
structure S(τ ) ∈ S is quasi-regular. Moreover, if S(τ ) is regular for some τ = τ 0 ∈
S 2 then it is regular for all τ . Furthermore, if S is regular then either all the leaves
are diffeomorphic to SO(3) or all the leaves are diffeomorphic to S 3 .
Actually in the regular case it follows from a deeper result of Salamon [Sal82]
that all leaves are diffeomorphic to S 3 in precisely one case, namely when M =
S 4n+3 .
Remark 13.2.2: Note that every Sasaki-Einstein 3-manifold must also have a 3-
Sasakian structure. This is because in dimension four Ricci-flat and Kähler is
equivalent to hyperkähler. Every compact 3-Sasakian 3-manifold must be a space
of constant curvature 1. Hence, M is covered by a unit round 3-sphere and, in
fact, it is always the homogeneous spherical space form S 3 /Γ, where Γ is a discrete
subgroup of Sp(1) [Sas72]. The homogeneous spherical space forms in dimension 3
are well-known. They are Sp(1)/Γ, where Γ is one of the finite subgroups of Sp(1)
discussed in detail in Sections 10.1.1 and 12.10.1. The only regular 3-Sasakian
manifolds in dimension 3 are S 3 and SO(3). Compare this with the classification
by Belgun and Geiges of all Sasakian 3-manifolds given in Section 10.1.
A Sasaki-Einstein structure on a 3-Sasakian manifold does not have to be a
part of the 3-Sasakian structure. Several hundred Sasaki-Einstein metrics exist on
S 7 by Corollary 11.5.4 none of which are part of a 3-Sasakian structure which can
be proved using Exercise 13.2 below. It is worth observing that the only examples
that we are aware so far involve spheres and their quotients. It is possible that some
of the 3-Sasakian rational homology 7-spheres of Theorem 13.8.3 are diffeomorphic
to some of the Sasaki-Einstein rational homology 7-spheres discussed in Example
11.7.12. For example if we take l in Theorem 13.8.3 to be one of the primes listed in
the table referred to in Example 11.7.12, then the two rational homology 7-spheres
will have the same cohomology ring, but we have no way of knowing at this time
whether they are even homeomorphic. However, Corollary 13.3.2 below implies
that the Sasaki-Einstein metrics of Example 11.7.12 cannot be part of a 3-Sasakian
structure since they all have orbifold Fano index one. Of course, the previously
mentioned examples refer to distinct Riemannian metrics. Here is an example of
distinct Sasakian structures sharing the same Riemannian metric.
480 13. 3-SASAKIAN MANIFOLDS

Example 13.2.6: Recall that Lemma 8.1.17 says that the only Riemannian met-
ric that can share two distinct Sasakian structures that are neither conjugates nor
part of the same 3-Sasakian structure are metrics of constant sectional curvature
one. In fact it is easy to see that the constant sectional curvature one metric g0
on S 4n+3 admits two distinct 3-Sasakian structures defined by whether the Sp(1)-
action as unit quaternions is taken from the right or left. Now consider the lens
space Zk \S 4n+3 obtained by quotienting by the cyclic subgroup Zk of the diag-
onal ∆(Sp(1) ⊂ Sp(n). This can represented by multiplication from the left by
a diagonal matrix with diagonal entries ρ ∈ S 1 ⊂ Sp(1) such that ρk = 1. This
Zk commutes with the right Sp(1)-action, so the “right” 3-Sasakian structure S
on S 4n+3 descends to a 3-Sasakian structure on the quotient Zk \S 4n+3 . However,
the left 3-Sasakian structure on S 4n+3 does not descend to Zk \S 4n+3 since the
centralizer of Zk in ∆(Sp(1)) is a diagonal S 1 ⊂ Aut(Sl ). Nevertheless, the “left”
Sasakian structure determined by this S 1 does descend to a “left” Sasakian struc-
ture Sl on Zk \S 4n+3 . This left Sasakian structure is actually regular while none
of the Sasakian structures of the right 3-Sasakian structure can be regular unless
k = 1, 2.

13.3. The Fundamental Foliations Fτ and FQ


In this section we discuss the two fundamental foliations associated to 3-Sasakian
manifolds and describe their consequences. We begin with

13.3.1. The 1-dimensional Foliation Fτ . Fixing a Sasakian structure, say


S1 = (ξ1 , Φ1 , η1 ) in the 3-Sasakian structure S, the space of leaves of the corre-
sponding characteristic foliation F1 is a compact complex orbifold (Z1 , I1 ) which
by Corollary 13.2.3 and Theorem 11.1.3 admits a compatible Kähler-Einstein met-
ric. The subbundle D1 = ker η1 of T M together with Φ1 |D1 define the CR structure
on M , and it is related to the complex structure I1 on Z1 by I1 X = (Φ|D1 )X̂, where
X̂ denotes the horizontal lift of the vector field X on Z1 . Actually a 3-Sasakian
structure gives a special kind of CR structure, namely, a CR structure with a
compatible holomorphic contact structure. Notice that the complex valued one
form on M defined by η + = η2 + iη3 has type (1, 0) on M. Moreover, one checks
that η + is holomorphic with respect to the CR structure. Although the 1-form
η + is not invariant under the circle action generated by ξ1 , the trivial complex
line bundle L+ generated by η + is invariant. Thus, the complex line bundle L+
pushes down to a nontrivial complex orbibundle L on Z1 . Let V denote the one
dimensional complex vector space generated by L+ . Writing the circle action as
exp (iφξ1 ) shows that V is the representation with character e−2iφ , and since M
is a principal S 1 -orbibundle over Z1 , the twisted product L ' M ×S 1 V is a holo-
morphic line orbibundle on the complex orbifold (Z1 , I1 ). Now we can define a map
of orbibundles θ : T (1,0) Z1 −→ L by θ(X) = η + (X̂). Notice that θ(X) is not a
function on Z1 but a section of L. Now a straightforward computation shows that
η + ∧ (dη + )n is a nowhere vanishing section of Λ(2n+1,0) D1 on M, and thus θ ∧ (dθ)n
orb
is a nowhere vanishing section of KZ 1
⊗Ln+1 , implying the relation Ln+1 ⊗KZ
orb
1
= 1l
orb
in Pic (Z). This gives Z1 a complex contact structure by Proposition 6.1.14. It
is straightforward to perform this construction for an arbitrary choice of oriented
orthonormal 3-frame (τ 1 , τ 2 , τ 3 ). Choosing the complex structure I(τ 1 ) we define
η + (τ 1 ) = η(τ 2 ) + iη(τ 3 ). Thus, we obtain a complex contact structure θτ 1 on
13.3. THE FUNDAMENTAL FOLIATIONS Fτ AND FQ 481

the complex orbifold (Zτ 1 , I(τ 1 )) with a compatible orbifold Kähler-Einstein met-
ric hτ 1 . It is easy to see that there is a φ ∈ Sp(1) such that Sτ 1 = φ∗ S1 , and
that φ induces an isomorphism of the corresponding complex orbifolds with their
induced structures, viz. (Zτ 1 , I(τ 1 ), θτ 1 , hτ 1 ) ≈ (Z1 , I1 , θ, h). We have arrived at
[BGM93, BG97b]:
Theorem 13.3.1: Let (M, S) be a complete 3-Sasakian manifold, choose a direc-
tion τ ∈ S 2 , and let Zτ denote the space of leaves of the corresponding foliation Fτ .
Then Zτ is a compact locally cyclic Fano contact orbifold with a Kähler-Einstein
metric hτ of scalar curvature 8(2n + 1)(n + 1) such that the natural projection
π : M −→ Zτ is an orbifold Riemannian submersion with respect to the Riemann-
ian metrics g on M and hτ on Zτ . Furthermore, different τ , τ 0 ∈ S 2 give an
isomorphism (Zτ , I(τ ), θτ , hτ ) ≈ (Zτ 0 , I(τ 0 ), θτ 0 , hτ 0 ).
The space Zτ turns out to be the twistor space defined in Definition 12.2.15,
and in [BG97b] we named it, perhaps erroneously, the twistor space of the 3-
Sasakian manifold (M, S). Actually, there is another object that could merit the
name the twistor space of M, namely the trivial 2-sphere bundle S 2 × M with the
structure induced from the twistor space S 2 × C(M ) of the hyperkähler cone.
There is a nice corollary that follows from the proof of Theorem 13.3.1.
Corollary 13.3.2: Let Z be the twistor space of a complete 3-Sasakian manifold
M 4n+3 . Then its orbifold Fano index is divisible by n + 1.
Exercise 13.2: Let Xf be a weighted homogeneous hypersurface in CP(w) of
complex dimension 2n + 1. Show that if IXf = |w| − d < (n + 1)mini {wi } then Xf
can not admit a complex contact structure.
Example 13.3.3: As observed in Example 13.2.6 the only manifold that has two
distinct 3-Sasakian structures sharing the same Riemannian metric is the sphere
S 4n+3 with its constant sectional curvature one metric. A related result was proven
by LeBrun in [LeB95] where he shows that the only positive Kähler-Einstein man-
ifold that admits two distinct complex contact structures is CP2n+1 . Recall that
the manifold Zk \S 4n+3 of Example 13.2.6 has a non-regular 3-Sasakian structure.
Since the right and left circle actions commute we see that its twistor space is the
orbifold Zk \CP2n+1 with π1orb = Zk . Note also that the quotient of Zk \S 4n+3 by
the left circle action is just CP2n+1 .
Recall from Theorem 12.2.16 that the twistor space of a quaternionic manifold
is ruled by rational curves. The same is true in our case as long as one allows for
singularities. We have
Proposition 13.3.4: The twistor space Z of a 3-Sasakian manifold is ruled by a
real family of rational curves C with possible singularities on the singular locus of
Z. All the curves C are simply-connected, but π1orb (C) can be a non-trivial cyclic
group.
For any line orbibundle L we let L̂ denote L minus its zero section. Recall
Kawasaki’s notion [Kaw78] of proper from Remark 4.2.2.
Proposition 13.3.5: Let Z be the twistor space of a 3-Sasakian manifold M of
dimension 4n + 3, and assume that π1orb (Z) = 0. If the contact line orbibundle L
(or equivalently its dual L−1 ) has a root in Picorb (Z), then it must be a square
1 1
root, namely L 2 . Moreover, in this case if both L̂ and L̂ 2 are proper in the sense
482 13. 3-SASAKIAN MANIFOLDS

of Kawasaki, then we must have Z = P2n+1 . In particular, this holds if the total
space of L̂ is smooth.

Proof. Since Picorb (Z) is torsion free. By Proposition 13.3.4 Z is ruled by


rational curves C which on the singular locus take the form Γ\P1 . Now the restric-
tion L−1 |C is O(−2) which is an orbibundle if C is singular. In either case it has
only a square root namely the tautological orbibundle O(−1). Since these curves
C cover Z this proves the first statement. The second statement follows from a
modification of an argument due to Kobayashi and Ochiai [KO73] and used by
1
Salamon [Sal82]. The main point is that since L̂, L̂ 2 are proper and it follows
that we can apply Kawasaki’s Riemann-Roch Theorem [Kaw79] together with the
Kodaira-Baily Vanishing Theorem 4.4.28 to arbitrary powers of the line orbibundle
1
L 2 to give (n + 1)(2n + 3) infinitesimal automorphisms of the complex contact
structure on Z. Since π1orb (Z) = 0, these integrate to global automorphisms on Z
and the result follows. See [BG97b] for details. ¤

Remark 13.3.1: There is an error in the statement of Proposition 4.3 of [BG97b].


The error is in leaving out the assumptions that π1orb (Z) is trivial and that the
contact line bundle is proper. Example 13.3.6 below shows that the conclusion in
Proposition 13.3.5 does not necessarily hold if the hypothesis π1orb (Z) = 0 is omit-
ted. Likewise, Example 13.3.7 below gives a counterexample when the condition
that L be proper is omitted.
Example 13.3.6: Consider the 3-Sasakian lens space L(p; q) = Zp \S 7 constructed
as follows: S 7 is the unit sphere in the quaternionic vector space H2 with quater-
nionic coordinates u1 , u2 . The action of Zp is the left action defined by (u1 , u2 ) 7→
(τ u1 , τ q u2 ), where τ p = 1 and p and q are relatively prime positive integers. If
p = 2m for some integer m then −id is an element of Z2m , so the 3-Sasakian mani-
folds L(2m; q) and L(m; q) both have the same twistor space, namely Z = Zm \P3 ,
and π1orb (Z) ' Zm . There are clearly many similar examples in all dimensions equal
to 3 mod 4.
Example 13.3.7: Consider the 3-Sasakian 7-manifolds S(p1 , p2 , p3 ) described in
Section 13.7.4 below, where the pi ’s are pairwise relatively prime, and precisely one
of the pi ’s is even, say p1 . S(p1 , p2 , p3 ) is simply-connected and its twistor space
Z(p1 , p2 , p3 ) has π1orb (Z(p1 , p2 , p3 )) = 0. Now there is a Z2 acting on S(p1 , p2 , p3 ),
but not freely, which acts as the identity on Z(p1 , p2 , p3 ). Thus, Z2 \S(p1 , p2 , p3 )
has Z(p1 , p2 , p3 ) as its twistor space, and as an orbibundle Z2 \S(p1 , p2 , p3 ) −→
Z(p1 , p2 , p3 ) is not proper in the sense of Kawasaki. Thus, the orbibundle L̂ is not
proper, and Kawasaki’s Riemann-Roch theorem [Kaw79] cannot be applied.
We now wish to formulate a converse to Theorem 13.3.1.
Definition 13.3.8: A compact Fano contact orbifold Z is said to be good if the to-
tal space of the principal circle bundle M associated with the contact line orbibundle
L is a smooth compact manifold.
The following theorem given in [BG97b] can be thought of as an orbifold
version of a theorem of LeBrun [LeB95]:
Theorem 13.3.9: A good compact Fano contact orbifold Z is the twistor space
associated to a compact 3-Sasakian manifold if and only if it admits a compatible
Kähler-Einstein metric h.
13.3. THE FUNDAMENTAL FOLIATIONS Fτ AND FQ 483

Proof. The only if part is Theorem 13.3.1, so we assume that Z is a good


Fano contact orbifold with a compatible Kähler-Einstein orbifold metric h. Choose
the scale of h so that the scalar curvature is 8(2n + 1)(n + 1). Let π : M → Z
denote the principal orbifold circle bundle associated to L. It is a smooth compact
submanifold embedded in the dual of the contact line orbibundle L−1 . The Kähler-
Einstein metric h has Ricci form ρ = 4(n + 1)ω, where ω is the Kähler form on
−1
Z, and ρ represents the orbifold first Chern class of KZ . Let η1 be the connection

in π : M → Z with curvature form 2π ω. Then the Riemannian metric gM on M
can be defined by gM = π ∗ h + η12 . It follows from the Inversion Theorem 11.1.12
that gM is Sasaki-Einstein. As in Proposition 2.2.4 of [Swa91] the orbibundle
L ⊗ Λ(1,0) Z has a section θ such that the Kähler-Einstein metric h decomposes as
h = |θ|2 + hD , where hD is a metric in the codimension one orbibundle D = ker θ.
Let us write π ∗ θ = η + . Since M is a circle bundle in L−1 , the contact bundle L
trivializes when pulled back to M. This together with the condition that θ ∧ (dθ)n
is nowhere vanishing on Z implies that η+ is a nowhere vanishing complex valued
1-form on M. So the metric gM on M can be written as
gM = η12 + |η+ |2 + π ∗ hD .
We claim that this metric is 3-Sasakian. To see this consider the total space of the
dual of the contact line orbibundle minus its 0 section which is M × R+ = C(M ).
Put the cone metric dr2 + r2 g on C(M ). The natural C∗ action on C(M ) induces
homotheties of this metric. Now using a standard Weitzenböck argument, LeBrun
[LeB95] shows that C(M ) has a parallel holomorphic symplectic structure and his
argument works just as well in our case. Let ϑ denote the pullback of the contact
form θ to C(M ) which is a holomorphic 1-form on M that is homogeneous of degree
1 with respect to the C∗ action. Thus Υ = dϑ is a holomorphic symplectic form
on C(M ) which is parallel with respect to the Levi-Civita connection of the cone
metric. Hence, (C(M ), dr2 + r2 g) is hyperkähler. Furthermore, if {Ia }3a=1 denote
hyperkähler endomorphisms on C(M ), ϑ2 , ϑ3 are the real and imaginary parts of
ϑ, and ϑ1 is the pullback of η1 to C(M ), then LeBrun shows that
ϑ1 I1 = ϑ2 I2 = ϑ3 I3 .
It then follows from our previous work [BGM93] that g is 3-Sasakian. But by
construction Z is the space of leaves of the foliation generated by ξ1 , so Z must be
the twistor space of the compact 3-Sasakian manifold M. ¤
13.3.2. The 3-dimensional Foliation FQ . We now discuss the geometric
properties of the 3-dimensional foliation FQ generated by the characteristic vector
fields {ξ1 , ξ2 , ξ3 }. Basic references are [BGM94a, BG99]. First we make note of
some of the properties of the O’Neill tensor fields A and T discussed in Section
2.5.1. Since Fξa ⊂ FQ for each a many of the properties satisfied by the Sasakian
characteristic foliation Fξa hold for the 3-Sasakian foliation FQ . In particular, the
O’Neill tensor T Q vanishes for FQ . The only potentially non-vanishing components
satisfy TξQa ξb = (∇ξa ξb )h = −(Φb ξa )h = 0, and TξQa X = (∇ξa X)v = 0. For the A
tensor of O’Neill we notice that Equations (2.5.15)-(2.5.18) imply a direct analogy
with the Fξa foliation. Using the convention that X, Y are sections of H, we have
the 3-Sasakian version of Lemma 7.3.5:
Lemma 13.3.10: For a 3-Sasakian manifold the following identities hold:
(i) T Q = 0 ,
484 13. 3-SASAKIAN MANIFOLDS

P3
(ii) AQXY = − a=1 dηa (X, Y )ξa ,
(iii) AQ ξ
X a = −Φ aX ,
(iv) |AQ |2 = 12n .
Exercise 13.3: Prove Lemma 13.3.10 and give the relation between O’Neill’s ten-
sor Aa for the foliation Fa , and O’Neill tensor AQ for FQ .
We summarize some of the basic results in
Proposition 13.3.11: Let (M, S) be a 3-Sasakian manifold such that the char-
acteristic vector fields {ξ1 , ξ2 , ξ3 } are complete. Let FQ denote the canonical three
dimensional foliation on M. Then
(i) The foliation FQ is Riemannian, or equivalently the metric g is bundle-
like.
(ii) The leaves of FQ are totally geodesic spherical space forms Γ\S 3 of con-
stant curvature one, where Γ ⊂ Sp(1) = SU (2) is a finite subgroup.
(iii) The 3-Sasakian structure on M restricts to a 3-Sasakian structure on
each leaf.
(iv) The generic leaves are either SU (2) or SO(3).
Proof. The proof of (i), (ii), and (iii) follow from the basic relations for 3-
Sasakian manifolds and is left to the reader. To prove (iv) we notice that the
foliation FQ is regular restricted to the generic stratum M0 . By (ii) and regularity
there is a finite subgroup Γ ⊂ SU (2) such that the leaves of this restricted foliation
are all diffeomorphic to Γ\S 3 , which is 3-Sasakian by (iii). Now the regularity of
FQ on M0 implies that its leaves must all be regular with respect to the foliation
generated by ξ1 . But a result of Tanno [Tan70] says that the only regular 3-
Sasakian 3-manifolds have Γ = id or Z2 , in which case (iv) follows. ¤
Example 13.3.12: Consider the 3-Sasakian lens space L(p; q) = Zp \S 7 of Example
13.3.6. If p is odd then −id is not an element of Zp so the generic leaf of the
foliation FQ is S 3 . The singular stratum consists of two leaves both of the form
Zp \S 3 with leaf holonomy group Zp . These two leaves are described by u2 = 0 and
u1 = 0, respectively. If p is even then −id is an element of Zp , so the generic leaf
is SU (2)/Z2 = SO(3), and the leaf holonomy of the two singular leaves is Z p2 .
The next theorem was first proved by Ishihara [Ish74] in the regular case. Its
general version described here was proved in [BGM94a].
Theorem 13.3.13: Let (M, S) be a 3-Sasakian manifold of dimension 4n + 3 such
that the characteristic vector fields {ξ1 , ξ2 , ξ3 } are complete. Then the space of leaves
M/FQ has the structure of a quaternionic Kähler orbifold (O, gO ) of dimension
4n such that the natural projection π : M −→ O is a principal orbibundle with
group SU (2) or SO(3) and a Riemannian orbifold submersion such that the scalar
curvature of gO is 16n(n + 2). Furthermore, η = (η1 , η2 , η3 ) is a connection in this
principal orbibundle.
Proof. We can split T M = V3 ⊕ H, where V3 is the subbundle spanned by
the Lie algebra su(2) of characteristic vector fields {ξ1 , ξ2 , ξ3 } and the “horizontal”
bundle is the orthogonal complement H = V3⊥ . Let hΦa = Φa |H be the restriction
of characteristic endomorphisms. One can easily see that
X
hΦa ◦ hΦb = −δab 1l + ²abc hΦc .
c
13.3. THE FUNDAMENTAL FOLIATIONS Fτ AND FQ 485

It follows that H is pointwise a quaternionic vector space and O is a compact


quaternionic orbifold. We must show that the metric gO obtained from g by the
orbifold Riemannian submersion π : M → O has its holonomy group reduced to a
subgroup of Sp(n)·Sp(1). This can be done by constructing a parallel 4-form on
O. Consider φa = dηa and
X
Φ̄a = φa + ²abc ηb ∧ ηc .
b,c
P
It is easy to see that the 4-form Ω = a Φ̄a ∧ Φ̄a is horizontal and Sp(1)-invariant.
It follows that there is a unique 4-form Ω̂ on the orbifold O invariant under the
action of the local uniformizing groups such that π ∗ Ω̂ = Ω. One can show that
Ω̂ is parallel on O using standard tensor computation with O’Neill formulas (see
[BGM93, GL88] for details). In the case n = 1 the parallelism of the 4-form
does not further restrict Riemannian geometry of O. However, one can show that
(O, gO ) is a compact self-dual Einstein orbifold. Self-duality follows easily from
the fact that O is quaternionic. The fact that the metric is Einstein is a simple
computation and in the regular case can be found in [Tan71]. ¤

There is an important inversion theorem of Theorem 13.3.13 originally in the


regular case due to Konishi [Kon75]. By now there are several proofs of this,
all of them related. Given a quaternionic Kähler orbifold O one can construct
the Salamon twistor space Z and then get M from the Inversion Theorem 13.3.9
[BG97b]. Another approach would be to construct the orbifold version of Swann’s
bundle [Swa91] on O and then use the results of [BGM93] to obtain M. Here our
proof is essentially that of Konishi’s, only slightly modified to handle the orbifold
situation.
Theorem 13.3.14: Let (O, gO ) be a quaternionic Kähler orbifold of dimension 4n
with positive scalar curvature 16n(n+2). Then there is a principal SO(3) orbibundle
over O whose total space M admits a 3-Sasakian structure with scalar curvature
2(2n + 1)(4n + 3).

Proof. Let Q denote the orbisubbundle of End T O describing the quater-


nionic structure. Let {Ũi } be local uniformizing neighborhoods that cover O and
(i)
Ia a local framing of Q on Ũi that satisfies
(i)
Ia(i) ◦ Ib = −δab 1l + ²abc Ic(i) .
(i)
Since O is quaternionic Kähler there are 1-forms τa on each Ũi such that
(i)
∇Ia(i) = ²abc τb ⊗ Ic(i) .

Now the structure group of the orbibundle End T O is Sp(n)·Sp(1), and that of the
orbisubbundle Q is SO(3). Let π : M −→ O denote the principal SO(3) orbibundle
(i)
associated to Q. The local 1-forms τa are the components of an so(3) connection
P 3 (i)
τ (i) = a=1 τa ea , where {ea } denotes the standard basis of so(3) which satisfies
the Lie bracket relations [ea , eb ] = 2²abc ec . The local connection forms satisfy the
well-known relations
−1
τ (i) = adgij τ (j) + gij dgij
486 13. 3-SASAKIAN MANIFOLDS

in Ũi ∩ Ũj for some smooth map gij : Ũi ∩ Ũj −→ SO(3). Furthermore, from [Ish74]
one checks that the curvature forms
(i)
(13.3.1) ωa(i) = dτa(i) + ²abc τb ∧ τc(i)
(i) (i)
satisfy the relation 2gO (Ia X, Y ) = ωa (X, Y ). Now on each Ũi there exists a
smooth local section σi : Ũi −→ M and on M there is a global 1-form ηa such that
(i)
τa = σi∗ ηa . On M we define a Riemannian metric by
3
X
g = π ∗ gO + ηa ⊗ ηa .
a=1

By construction the vector fields ξa generating the SO(3) action on M are dual to
the forms ηa with respect to this metric, viz. gM (X, ξa ) = ηa (X) for any vector field
X on M. Also by construction the vector fields ξa are Killing fields with respect to
the metric g. Now define the (1, 1) -tensor field Φa = ∇ξa . Since the ξa are mutually
orthogonal vector fields of unit length on M, one easily checks that Φa ξb = −²abc ξc .
Thus Φa splits as
Φa = hΦa + ²abc ξb ⊗ ηc .
One then checks using 13.3.1 that on each open set π −1 (Ui ), hΦa equals the hori-
(i)
zontal component of (σi )∗ Ia . From this one then shows that
Φa ◦ Φb − ξa ⊗ ηb = −²abc Φc − δab 1l . ¤
If the homomorphisms ψi : Γi −→ SO(3) are injective the total space M will be
a smooth 3-Sasakian manifold. We summarize our discussion of the fundamental
foliations associated to a 3-Sasakian structure with the following
Definition 13.3.15: With every 3-Sasakian manifold (orbifold) (M, S) we asso-
ciated the diamond diagram ♦(M, S) of M and three more Einstein orbifold
spaces canonically associated to M :

(13.3.2) C(M )bF


y FFFF
yy FFFF
yy FFFF
y FFF1 Q
|yy "
o
Z FF M
FF w
FF www
FF ww
F# ² {www
O
of orbifold fibrations, where
(i) M ,→ C(M ) is the inclusion of M into the hyperkähler cone C(M ) ,
(ii) C(M )−−→O , C(M )−−→Z , C(M )−−→M are the Swann orbibundles,
(iii) M −−→O is the Konishi orbibundle,
(iv) M −−→Z is the twistor fibration.
Konishi’s construction gives an SO(3) orbibundle over O. In the case that O is a
smooth manifold there is a well-known obstruction to lifting this bundle to an Sp(1)
bundle, the Marchiafava-Romani class ² of Definition 12.2.1. This obstruction also
occurs when O is an orbifold as long as one uses Haefliger’s orbifold cohomology.
The class ² is the image of the connecting homomorphism
1 2
(13.3.3) δ : Horb (O, G) −→ Horb (O, Z2 ),
13.3. THE FUNDAMENTAL FOLIATIONS Fτ AND FQ 487

where G is the sheaf of germs of smooth orbifold maps from open sets of O to the
group Sp(n)·Sp(1). Since every 3-Sasakian structure S has canonically associated to
it a unique quaternionic Kähler orbifold, we can consider the Marchiafava-Romani
class ² to be an invariant of the 3-Sasakian structure S in which case we shall
sometimes write ²(S).
If following Salamon [Sal82] we write T O ⊗ C ' E ⊗ H, then ² is the second
Stiefel-Whitney class w2 of the orbibundle S 2 (H) over O. We have
Proposition 13.3.16: The principal SO(3)-orbibundle of Theorem 13.3.14 lifts to
2
a principal Sp(1) orbibundle if and only if ² ∈ Horb (O, Z2 ) vanishes. Moreover,
when ² = 0 the 3-Sasakian structure on the total space M of the SO(3) orbibundle
lifts to the total space M 0 of the Sp(1) orbibundle.
Thus, in the case that ² = 0 there are precisely two 3-Sasakian orbifolds M, M 0
corresponding to the quaternionic Kähler orbifold O. Let Z denote the twistor
space of the orbifold O. Then likewise, since S 2 ' SO(3)/S 1 ' Sp(1)/S 1 the two
3-Sasakian orbifolds M and M 0 have the same twistor space Z. When O is a smooth
manifold a result of Salamon [Sal82] says that ² = 0 if and only if the quaternionic
Kähler manifold O is quaternionic projective space. If we impose the condition
that the orbifolds M and M 0 are smooth manifolds, there is a similar result.
Theorem 13.3.17: If two 3-Sasakian manifolds (M, S) and (M 0 , S 0 ) are associ-
ated to the same quaternionic Kähler orbifold O or, equivalently, the same twistor
space Z, then both M and M 0 have the same universal covering space M̃ ' M̃ 0
which is the standard 3-Sasakian sphere.

Proof. We work with the twistor space Z. Now M and M 0 are unit circle
bundles in the line orbibundles L−1 and (L0 )−1 respectively. Moreover, since M 0
is a double cover of M, it follows that L = (L0 )2 . Consider the universal orbifold
cover Z̃ of Z with π1orb (Z̃) = 0. Pull back the orbibundles L and L0 to orbibundles
˜1
L̃ and L̃0 on Z̃ respectively. These bundles satisfy L̃0 = L 2 . By construction and
naturality of the covering maps L̃ is the contact line bundle on Z̃. Moreover, M̃
and M̃ 0 which are the total spaces of the pullbacks of M and M 0 to Z̃ are both
ˆ
smooth manifolds since they cover smooth manifolds. Thus, the orbibundles L̃0 and
ˆ
L̃ are proper, so we can apply Proposition 13.3.5 giving Z̃ ' P2n+1 . It follows that
M̃ 0 ' S 4n+3 . ¤

Remark 13.3.2: Konishi also considers the case when the quaternionic Kähler
manifold has negative scalar curvature. This gives a Sasakian 3-structure on M
with indefinite signature (3, 4n).

13.3.3. The Second Einstein Metric. Of course, by an Einstein metric we


actually mean a homothety class of Einstein metrics. In this section we shall show by
using a theorem of Bérard-Bergery [Bes87] that every 3-Sasakian manifold, as well
as its twistor space, has at least two distinct homothety classes of Einstein metrics.
The method involves the canonical variation [Bes87] associated with Riemannian
submersions. Due to the local nature of the calculations involved this construction
holds equally well for orbifold Riemannian submersions. The canonical variation is
constructed as follows [Bes87]:
Let π : M −→ B be an orbifold Riemannian submersion with g the Riemannian
metric on M. Let V3 and H denote the vertical and horizontal subbundles of the
488 13. 3-SASAKIAN MANIFOLDS

tangent bundle T M. For each real number t > 0 we construct a one parameter
family gt of Riemannian metrics on M by defining
(13.3.4) gt |V3 = tg|V3 , gt |H = g|H , gt (V3 , H) = 0 .
So for each t > 0 we have an orbifold Riemannian submersion with the same base
space. Furthermore, if the fibers of g are totally geodesic, so are the fibers of gt . We
apply the canonical variation to the orbifold Riemannian submersion π : M −→ O.
The metric as well as other objects on O will be denoted with a check such as ǧ.
Theorem 13.3.18: Every 3-Sasakian manifold M admits a second Einstein metric
of positive scalar curvature. Furthermore, the twistor space Z also admits a second
orbifold Einstein metric which is Hermitian-Einstein, but not Kähler-Einstein.
Proof. We apply the canonical variation to the orbifold Riemannian submer-
sion π : M −→ O. According to the Bérard-Bergery Theorem [Bes87], 9.73 there
are several conditions to check. First, the connection H must be a Yang-Mills
connection. By (i) of Lemma 13.3.10, the O’Neill tensor T Q vanishes. So the
Yang-Mills condition is [Bes87]
X
(13.3.5) g((∇Ei AQ )Ei X, ξa ) = 0
i

for each a = 1, 2, 3 and where Ei is a local orthonormal frame of H, X is any


horizontal vector field, and AQ is O’Neill’s tensor for FQ . Actually in our case
of a Sasakian metric we can prove the stronger condition g((∇Ei Aa )Ei X, ξa ) = 0
for all i = 1, . . . , n and all a = 1, 2, 3 by using (ii) of Lemma 7.3.10 and Theorem
7.3.16. Then using the relation between AQ and Aa Equation 13.3.5 follows. Second
|AQ |2 must be constant which is (iv) of Lemma 13.3.10. The final condition to be
satisfied is 9.73e of [Bes87]. Using the easily verified relations g(AXi , AXj ) = 3δij
and g(Aξa , Aξb ) = 4nδab this condition becomes (λ̌)2 − λ̂(12n + 18) > 0, where λ̌
and λ̂ are the Einstein constants for O and the fibers, respectively. Since in our
case λ̌ = 4(n + 2) and λ̂ = 2, we see that the inequality is satisfied.
The last statement, due also to Bérard-Bergery, follows also by applying the
canonical variation to the twistor space orbifold fibration π 0 : Z−−−→O. See 14.84
of [Bes87]. ¤
The scalar curvature of any metric gt in the canonical variation of the 3-
Sasakian metric g is given by the formula st = 16n(n + 2) + 6/t − 12nt [Bes87].
λ̂ 1
Moreover, the value of t that gives the second Einstein metric is t0 = λ̌− λ̂
= 2n+3 .
A similar computation can be done for the second Einstein metric on Z. We shall
return to these second Einstein metrics in Chapter 14.

13.3.4. 3-Sasakian Metrics and Positive Sectional Curvature. The only


3-Sasakian metrics of positive sectional curvature are the standard metrics on
spheres. On the other hand, Dearricott has shown that there is a relation between
certain 3-Sasakian metrics and metrics of positive sectional curvature [Dea04,
Dea05]. More precisely he has proven
Theorem 13.3.19: Let (M, S) be a 3-Sasakian manifold and let (O, gO ) be the
associated quaternionic Kähler orbifold. If gO is of positive sectional curvature
then the canonical variation gt of the 3-Sasakian metric g also has positive sectional
curvature for sufficiently small t.
13.3. THE FUNDAMENTAL FOLIATIONS Fτ AND FQ 489

The proof of this theorem can be found in [Dea04] and it easily follows from
the relation between sectional curvatures of the original metric and its canonical
variation. The condition on gO being itself of positive sectional curvature is very
strong and allows few examples. All of the known examples are the weighted pro-
jective spaces of Theorem 12.5.4 in dimension 4. These do lead to a new geometric
construction of positive curvature metrics on some Eschenburg spaces which will
be discussed later. Unfortunately, in higher dimensions, the only known example
of quaternionic Kähler spaces of positive sectional curvature are HPn .

13.3.5. Invariants and the Classification of 3-Sasakian Structures. We


consider the question of equivalence of 3-Sasakian manifolds. Note that a 3-Sasakian
structure S = {ξa , ηa , Φa }3a=1 on a manifold M is determined completely by the
metric g and the characteristic vector fields {ξ1 , ξ2 , ξ3 }.
Definition 13.3.20: Two 3-Sasakian manifolds (M, S) and (M 0 , S 0 ) are said to
be isomorphic if there exist a diffeomorphism F : M −→ M 0 and an φ ∈ Sp(1)
such that F ∗ g 0 = g and ξa0 = (Adφ )∗ F∗ ξa , where Ad denotes the adjoint action of
Sp(1) on its Lie algebra sp(1).
In practice we shall always choose a basis ξa0 of the 3-Sasakian structure on
M so that F∗ ξa = ξa0 . Now given such a diffeomorphism F : M −→ M 0 it is
0

clear that the corresponding foliations are F -related, that is that F∗ F1 = F10 and
0
F∗ FQ = FQ . This induces a commutative diagram of orbifold diffeomorphisms

F
0
M
 −→ M

 
 
y y
F1
(13.3.6) Z −→ Z0

 
 
y y
FQ
O −→ O0 .

This implies that if Lx is the leaf of FQ at x ∈ O, then F (Lx ) is the leaf at


FQ (x) ∈ O0 , that is, L0FQ (x) = F (Lx ). Let G(L) denote the leaf holonomy group of
the leaf L. Then we have G(L0FQ (x) ) ≈ G(Lx ). More generally let G(M ) denote the
holonomy groupoid of the foliation FQ . By Theorem 2.5.10 G(M ) is a Lie groupoid
and a smooth Hausdorff manifold of dimension 4n + 6. We have
Proposition 13.3.21: Let F : M −→ M 0 be an isomorphism of 3-Sasakian man-
ifolds. Then F induces an isomorphism F∗ : G(M ) −→ G(M 0 ) of Lie groupoids.
We are interested in the unordered list (Γ1 , Γ2 , . . .) of holonomy groups in G(M )
up to abstract isomorphism. This list is finite if M is complete and it provides
important invariants of a 3-Sasakian manifold. Since the leaves of the foliation FQ
are all spherical space forms, the groups Γi are all either subgroups of Sp(1) or all
subgroups of SO(3), depending on whether the Marchiafava-Romani class ² of the
quaternionic Kähler orbifold O is 0 or 1, respectively. Notice that it follows from
its definition and 13.3.6 above that the class ² is an invariant of the 3-Sasakian
structure on the manifold M. Indeed, ² can be identified with a certain secondary
characteristic class of the foliation FQ . Thus, the Marchiafava-Romani class splits
490 13. 3-SASAKIAN MANIFOLDS

the isomorphism classes S of 3-Sasakian manifolds into the disjoint union S0 + S1


depending on whether ² is 0 of 1. A further rough classification scheme is given by
Definition 13.3.22: M is said to be:
(i) regular if all the Γi are the identity.
(ii) of cyclic type if all the Γi are cyclic.
(iii) of dihedral type if all the Γi are either cyclic or dihedral or binary
dihedral with at least one Γi non-Abelian.
(iv) of polyhedral type if at least one of the Γi is one of the polyhedral
groups, tetrahedral, octahedral, or icosahedral (or the corresponding bi-
nary double covers) groups.
At this writing the only known examples of 3-Sasakian manifolds of polyhedral
or dihedral type are the spherical space forms Γ\S 4n+3 and Γ\RP4n+3 , where Γ is a
binary polyhedral or a binary dihedral group in the first case and a polyhedral or a
dihedral group in the second. The action is that induced by the diagonal action of
Γ on the quaternionic vector space Hn+1 from the left. All the known non-regular
3-Sasakian manifolds are of cyclic type and are discussed in detail in Section 13.7.
Open Problem 13.3.1: Prove the existence of 3-Sasakian structures of polyhedral
or dihedral type that are not space forms.
Theorem 13.3.23: Let M be a complete regular 3-Sasakian manifold with ² = 0.
Then M ' S 4n+3 or RP4n+3 .
For more results about regular 3-Sasakian manifolds see Section 13.4. Next we
consider an important infinitesimal rigidity result. In the regular case this rigidity
is a simple consequence of the results of LeBrun [LeB88] and Nagatomo [Nag92]
(see [GS96]). In the general case it was proved by Pedersen and Poon [PP99].
Theorem 13.3.24: Complete 3-Sasakian manifolds are infinitesimally rigid.

Proof. The deformation theory of 3-Sasakian manifolds is tied to the defor-


mation theory of hypercomplex manifolds studied previously in [PP98a]. Let M
be a complete 3-Sasakian manifold. Then the compact manifold S 1 × M has a
natural hypercomplex structure [BGM94a]. Thus, its twistor space W is compact
and fibers holomorphically over CP1 . Moreover, there is a holomorphic foliation on
W whose leaves are elliptic Hopf surfaces, and whose space of leaves is the twistor
space Z associated to M. The geometry of the corresponding deformation theory
is as follows. Deformations (Mt , gt ) of the 3-Sasakian structure (M0 , g0 ) on M
correspond to deformations of the hypercomplex structure on S 1 × M of the form
S 1 ×Mt . In turn these deformations correspond to deformations of the holomorphic
fibration p : W −→ CP1 . Thus, there are natural projections:

(13.3.7) WB
z BB
p zz BBΦ
zzz BB
z} z BÃ
CP1 Z,

where each fiber of p is a divisor in W diffeomorphic to S 1 × M and Φ is an


orbifold submersion whose leaves are elliptic Hopf surfaces. Now the product map
p × Φ : W −→ CP1 × Z is an orbifold submersion whose leaves are elliptic curves.
13.4. HOMOGENEOUS 3-SASAKIAN MANIFOLDS 491

The differential of p × Φ induces the exact sequence of sheaves


0 −→ OW −→ ΘW −→ Φ∗ ΘZ ⊕ p∗ ΘCP1 −→ 0 ,
where OW denotes the structure sheaf of W and Θ denotes the holomorphic tangent
sheaf. Then using standard techniques together with the Kodaira-Baily vanishing
theorem and the orbifold version of the Akizuki-Nagano vanishing theorem, Peder-
sen and Poon show that the virtual parameter space for 3-Sasakian deformations
lies in
(13.3.8) H 0 (Z, OZ ) ⊗ H 1 (F, OF ) ⊕ H 0 (Z, ΘZ ) ⊗ H 1 (F, OF )
⊕H 1 (Z, ΘZ ) ⊗ H 0 (F, OF ) ⊕ H 1 (W, p∗ ΘCP1 ) ,
where F is the generic elliptic Hopf surface S 1 × Sp(1). One then analyzes each
summand of 13.3.8 to show that there are no 3-Sasakian deformations. For exam-
ple, possible deformations lying in the last summand vanish by results of Horikawa,
while 3-Sasakian deformations lying in the second and third summands must pre-
serve the complex contact structure on Z. There are no such deformations in the
third summand by the Kodaira-Baily vanishing theorem. Elements in the second
summand correspond to complex contact transformations that are invariant under
the U (1) × U (1) action coming from a discrete quotient of the C∗ principal action
on L, and there are no such elements. Finally, elements of the first summand cor-
respond to scale changes in the S 1 factor of S 1 × Sp(1) and these hypercomplex
deformations do not come from 3-Sasakian ones. ¤
While this theorem says that there is no “infinitesimal moduli”, there can be
discrete moduli of 3-Sasakian structures. The two known examples of 3-Sasakian
manifolds, where the moduli consists at least of two structures is dues to [CEZ07]
and will be discussed later.

13.4. Homogeneous 3-Sasakian Manifolds


Since a 3-Sasakian structure consists of a 2-sphere’s worth of Sasakian struc-
tures or alternatively 3 orthogonal Sasakian structures, and each Sasakian structure
has an isometry group of at least one dimension, it follows that every 3-Sasakian
manifold (M, S) has a nontrivial isometry group Isom(M, g) of dimension at least
three. We wish to study the automorphism group of a 3-Sasakian structure and its
relation to Isom(M, g). The most natural definition is:
Definition 13.4.1: Let (M, S) be a 3-Sasakian manifold. We define the auto-
morphism group Aut(M, S) of the 3-Sasakian structure by
\
Aut(M, S) = Aut(S(τ )) .
τ ∈S 2

(M, S) is said to be a 3-Sasakian homogeneous space if Aut(M, S) acts tran-


sitively on M.
We now wish to give the 3-Sasakian analogue of Theorem 8.1.18. We begin
analyzing the structure of Isom(M, g) of a 3-Sasakian manifold with two lemmas.
Lemma 13.4.2: Let (M, S) be a 3-Sasakian manifold. Then
(i) there is a chain Aut(M, S) ⊂ Aut(S(τ )) ⊂ Isom(M, g) of closed Lie
subgroups for any τ ∈ S 2 , and
(ii) Aut(M, S) = Aut(S(τ1 )) ∩ Aut(S(τ2 )) if τ1 6= ±τ2 ,
492 13. 3-SASAKIAN MANIFOLDS

(iii) the subgroup


[
G3 = G(τ ) ⊂ Isom(M, g) ,
τ ∈S 2

where G(τ ) is the one parameter subgroup generated by ξ(τ ), is iso-


morphic to either Sp(1) or SO(3), and commutes with the subgroup
Aut(M, S).

Proof. The subgroup property in (i) is clear, and Proposition 8.1.1 says that
Aut(S(τ )) is a closed Lie subgroup of Isom(M, g) and the remainder follows from
(ii). By Corollary 13.2.2 it is enough to consider any two Sasakian structures
S(τ1 ), S(τ2 ) with τ1 6= ±τ2 implying (ii). It follows from Proposition 13.2.1 that
the Lie algebra of G(τ ) is sp(1); hence, G(τ ) is isomorphic to either Sp(1) or SO(3).
Since the one parameter groups G(τ ) lie in the center of Aut(S(τ )) for each τ , the
groups G3 and Aut(M, S) commute. ¤

The second lemma gives a characterization of Aut(M, S).


Lemma 13.4.3: Let (M, S) be a 3-Sasakian manifold, and let φ ∈ Isom(M, g).
Then the following are equivalent
(i) φ∗ ξa = ξa for all a = 1, 2, 3,
(ii) φ∗ ηa = ηa for all a = 1, 2, 3,
(iii) φ∗ ◦ Φa = Φa ◦ φ∗ for all a = 1, 2, 3,
(iv) φ ∈ Aut(M, S).

Proof. That (ii) is equivalent to (iv) follows from Proposition 8.1.1 and (ii)
of Lemma 13.4.2. (i) and (ii) are equivalent by an easy duality argument. But (i)
and (iii) are equivalent since Φa = −∇ξa for all a = 1, 2, 3 and φ preserves ∇. ¤

Exercise 13.4: Prove directly the infinitesimal version of Lemma 13.4.3, namely.
that for any Killing vector field X the following conditions are equivalent:
(i) £X Φa = 0, for a = 1, 2, 3,
(ii) £X ηa = 0, for a = 1, 2, 3,
(iii) £X ξa = 0, for a = 1, 2, 3,
(iv) X ∈ aut(M, S).
We are now ready for our main theorem concerning the structure of Isom(M, g).
Theorem 13.4.4: Let (M, S) be a complete 3-Sasakian manifold. If (M, S) is not
of constant sectional curvature, then
(i) isom(M, g) = aut(M, S) ⊕ sp(1),
(ii) either Isom(M, g) = Aut(M, S) × Sp(1) or Isom(M, g) = Aut(M, S) ×
SO(3);
whereas if (M, g) does have constant sectional curvature then either Aut(M, S) ×
Sp(1) or Aut(M, S) × SO(3) is a proper subgroup of Isom(M, g).

Proof. Since (M, g) is complete M is compact, so the Lie algebra isom(M, g)


is reductive. By Lemmas 8.1.17 and 13.4.2, we know that dim Isom(M, g) =
dim Aut(M, S) + 3. But then using the commutativity discussed above together
with the fact that isom(M, g) is reductive we obtain (i). But then this implies that
¡ ¢
Isom(M, g) = Aut(M, S) × Sp(1) /Γ ,
13.4. HOMOGENEOUS 3-SASAKIAN MANIFOLDS 493

where Γ is a finite normal subgroup. By Lemma 13.4.2 Aut(M, S) ∩ Sp(1) = {id},


and Aut(M, S) is a subgroup of Isom(M, g), so Γ must be a normal subgroup of
Sp(1). But the only normal subgroups of Sp(1) are the central Z2 and the identity
which implies (ii). ¤
Proposition 13.4.5: Let (M, S) be a 3-Sasakian homogeneous space. Then all
leaves are diffeomorphic and O = M/FQ is a quaternionic Kähler manifold, where
the natural projection π : M −
→O is a locally trivial Riemannian fibration. Further-
more, Aut(M, S) passes to the quotient and acts transitively on the space of leaves
O.
Proof. Let ψ : Aut(M, S) × M − →M denote the action map so that, for each
a ∈ Aut(M, S), ψa = ψ(a, ·) is a diffeomorphism of M to itself. Theorem 13.4.4
implies that the isometry group Isom(M, g) contains Aut(M, S) × G3 , where G3 is
either Sp(1) or SO(3). Since the Killing vector fields ξa for a = 1, 2, 3 are both the
infinitesimal generators of the group G3 and a basis for the vertical distribution V,
it follows that G3 acts transitively on each leaf with isotropy subgroup of a point
some finite subgroup Γ ⊂ G3 . Now let p1 and p2 be any two points of M and let L1
and L2 denote the corresponding leaves (not necessarily distinct) through p1 and p2 ,
respectively. Since Aut(M, S) acts transitively on M , there exists an a ∈ Aut(M, S)
such that ψa (p1 ) = p2 . Now ψa restricted to L1 maps L1 diffeomorphically onto
its image, and, since the Sp(1) factor acts transitively on each leaf and commutes
with Aut(M, S), the image of ψa lies in L2 . But the same holds for the inverse map
ψa−1 with L1 and L2 interchanged, so the leaves must be diffeomorphic. Thus, the
leaf holonomy is trivial and π : M − →M/FQ = O is a locally trivial Riemannian
fibration. The fact that the space of leaves O is a quaternionic Kähler manifold
now follows from Ishihara’s Theorem 13.3.13. Since Aut(M, S) commutes with G3 ,
it follows that Aut(M, S) passes to the quotient and acts transitively on O. ¤
The following classification theorem now follows from Proposition 13.4.5, the
result of Alekseevsky which states that all homogeneous quaternionic Kähler man-
ifolds of positive scalar curvature are symmetric [Ale75], and from Wolf’s classifi-
cation of homogeneous QK manifolds [Wol65] or from Boothby’s classification of
complex homogeneous contact manifolds [Boo61]
Theorem 13.4.6: Let (M, S) be a 3-Sasakian homogeneous space. Then M =
G/H is precisely one of the following:
Sp(n + 1) Sp(n + 1) SU (m) SO(k)
, , ¡ ¢, ,
Sp(n) Sp(n)×Z2 S U (m − 2)×U (1) SO(k − 4)×Sp(1)
G2 F4 E6 E7 E8
, , , , .
Sp(1) Sp(3) SU (6) Spin(12) E7
Here n ≥ 0, Sp(0) denotes the trivial group, m ≥ 3, and k ≥ 7. Hence, there is
one-to-one correspondence between the simple Lie algebras and the simply-connected
3-Sasakian homogeneous manifolds.
The dimension of the 3-Sasakian homogeneous spaces are as follows: the first
two have dimension 4n + 3, while the second has dimension 4m − 5 with m ≥ 3,
and the third 4k − 13 with k ≥ 7; the dimension of the 3-Sasakian homogeneous
spaces corresponding to the exceptional Lie groups are in the order presented
11, 31, 43, 67, 115, respectively.
494 13. 3-SASAKIAN MANIFOLDS

Below we give the fundamental diagram ♦(G/H) ≡ ♦(G) for each 3-Sasakian
homogeneous space of Theorem 13.4.6

R+ × G/H
.  &

(13.4.1) G/(H ·U (1)) −−−−
←−−− G/H,
y
& .
G/H ·Sp(1)
where G/(H ·U (1)) are precisely the complex homogeneous contact manifolds clas-
sified by Boothby [Boo61] and G/(H ·Sp(1)) are the Wolf spaces [Wol65].
Remark 13.4.1: Note that a homogeneous 3-Sasakian manifold is necessarily simply-
connected with the exception of the real projective space. This is in sharp con-
trast with the Sasaki-Einstein case. Also notice that a 3-Sasakian manifold can
be Riemannian homogeneous (i.e., the full isometry group acts transitively) but
not 3-Sasakian homogeneous. This is true for the lens spaces Γ\S 3 , with |Γ| > 2.
For example, Zk \S 3 , k > 2, is a homogeneous Sasaki-Einstein manifold but not
3-Sasakian homogeneous.
Theorem 13.4.6 does not specify what the 3-Sasakian metric on the coset G/H
is. In Section 13.6.2 we will describe a quotient construction of the 3-Sasakian
homogeneous spaces with G = SU (n + 1) and G = SO(n + 1). Here we quote a
theorem of Bielawski [Bie96], which gives an explicit description of these metrics
in all cases.
Theorem 13.4.7: Let M = G/H be one of the spaces in Theorem 13.4.6. and let
g = h ⊕ m be the corresponding decomposition of the Lie algebras. Then there is a
natural decomposition m = sp(1) ⊕ m0 and the metric g on M is given in terms of
the scalar product on m
1
||m||2 = −hσ, σi − hm0 , m0 i ,
2
where σ ∈ sp(1), m0 ∈ m0 , and h·, ·i is the Killing form on g. In particular, the
metric g is not naturally reductive with respect to the homogeneous structure on M.
Remark 13.4.2: In the case when M is of constant curvature the canonical metric
on S 4n+3 (or RP4n+3 ) is not the standard homogeneous metric on the homogeneous
space Sp(n + 1)/Sp(n) (or Sp(n + 1)/Sp(n) × Z2 ) with respect to the reductive
decomposition sp(n + 1) ' sp(n) + m. It is, of course, the standard homogeneous
metric with respect to the naturally reductive decomposition o(4n + 4) ' o(4n +
3) + m. This is quite special to the sphere and orthogonal group. In general the
3-Sasakian homogeneous metrics are not naturally reductive with respect to any
reductive decomposition.

13.5. 3-Sasakian Cohomology


In this section we describe some cohomological properties of 3-Sasakian man-
ifolds M. We discuss a vanishing theorem and then derive a relation between the
Betti numbers of M and the Betti numbers of the QK orbifolds O. We conclude
with various implications of these relations in the regular case.
13.5. 3-SASAKIAN COHOMOLOGY 495

13.5.1. A Vanishing Theorem. Let (M, S) be a compact 3-Sasakian man-


ifold of dimension 4n + 3 and 3-Sasakian structure S. Throughout this section we
shall suppose that p ≤ 2n + 1. Referring to the splitting of the tangent bundle of M
into T M = V3 ⊕ H, we shall say that a p-form u ∈ Ωp (M ) has bidegree (i, p − i) if it
Vp ∗ Vi Vp−i
is a section of the subbundle of T M isomorphic to the dual of V3 ⊗ H.
In particular, u is called 3-horizontal if it has bidegree (0, p), or equivalently if
ξa cu = 0 for a = 1, 2, 3. An element ω ∈ Ωp (M ) is called invariant if h∗ ω = ω for
all h ∈ Sp(1). In the regular case, there is a principal Sp(1)-bundle π : M → O,
and ω is both 3-horizontal and invariant if and only if it is the pullback π ∗ ω̂ of a
form ω̂ on the QK base O. Now the curvature forms Φ̄a defined in the proof of
Theorem 13.3.13 are horizontal with respect to the foliation FQ . The Killing fields
ξa transform according to the adjoint representation of G3 , and the same is true of
the associated triples ηa , dηa , and Φa . If h ∈ G3 , we may write
X
h∗ Φa = hab Φb , a = 1, 2, 3,
b
ab
where h are components of the image of h in G3 . The 3-forms
X X
(13.5.1) Υ = η1 ∧ η2 ∧ η3 , Θ= ηa ∧ Φ̄a = ηa ∧ dηa + 6Υ
a a

have respective bidegrees (3, 0), (1, 2), and are clearly invariant. Their exterior
derivatives are
(13.5.2) dΥ = η1 ∧ η2 ∧ Φ̄3 + η2 ∧ η3 ∧ Φ̄1 + η3 ∧ η1 ∧ Φ̄2 , dΘ = Ω + 2dΥ ,
where the 4-form Ω is defined in 12.2.7. In fact, Ω is the canonical 4-form determined
by the quaternionic structure of Theorem 13.3.13 of the subbundle H, and is the
pullback of the fundamental 4-form Ω̂ on the quaternionic Kähler orbifold O.
Proposition 7.4.13 implies that any harmonic p-form with p ≤ 2n + 1 on the
compact 3-Sasakian manifold M is 3-horizontal. We can apply the construction in
Lemma 7.4.12 so as to obtain Φa : Hp (M ) → Hp (M ), a = 1, 2, 3, p ≤ 2n + 1,
and
(Ia u)(X1 , X2 , . . . , Xp ) = u(Φa X1 , Φa X2 , . . . , Φa Xp ) .
As we showed in the proof of Theorem 7.4.11 each Ia defines a complex structure on
the space of odd degree harmonic p-forms. Now, using the basic identities satisfied
by S we can prove the following “quaternionic” extension of this property to get
the following result originally due to Kuo [Kuo70]:
Proposition 13.5.1: Let Ia : Hp (M ) → Hp (M ), a = 1, 2, 3, and p ≤ 2n + 1. Then
X
Ib ◦ Ia = (−δab )p I + (²abc )p Ic .
c

In particular, when p is odd, {I1 , I2 , I3 } defines an almost quaternionic structure


on the vector space Hp (M ).
We are now ready to prove the main theorem of this section.
Theorem 13.5.2: Let (M, S) be a compact 3-Sasakian manifold of dimension 4n+
3, and let u ∈ Hp (M ) with p ≤ 2n + 1.
(i) If p is odd then u = 0, that is Hp (M ) = 0.
(ii) If p is even then Ia u = u for a = 1, 2, 3.
496 13. 3-SASAKIAN MANIFOLDS

Proof. Let u ∈ Hp (M ). We shall in fact show that I1 u = I2 u irrespective of


whether p is even or odd; the result then follows from the identities in Proposition
13.5.1 and symmetry between the indices 1, 2, 3. We may choose an isometry h ∈
Sp(1) so that h∗ Φ1 = Φ2 . Both u and I1 u are harmonic, so h∗ u = u and
(I1 u)(X1 , . . . , Xp ) = (h∗ (I1 u))(X1 , . . . , Xp ) = u((h∗ Φ1 )(X1 ), . . . , (h∗ Φ1 )(Xp )) =

= u(Φ2 X1 , . . . , Φ2 Xp ) = (I2 u)(X1 , . . . , Xp ) . ¤

Corollary 13.5.3: Let (M, S) be a compact 3-Sasakian manifold of dimension


4n + 3. Then the odd Betti numbers b2k+1 of M are all zero for 0 ≤ k ≤ n.
Obviously Corollary 13.5.3 does not apply to compact Sasakian or even Sasaki-
Einstein manifolds as we saw in Chapters 10 and 11. In contrast to the Sasaki-
Einstein case, for example, when M is highly connected we get
Corollary 13.5.4: Let (M, S) be a 3-Sasakian manifold that is highly connected.
Then M is a rational homology sphere. In particular, a 3-Sasakian manifold that
is a the link of a complete intersection must be a rational homology sphere.
The only known example to date of a 3-Sasakian manifold that is the link of an
isolated hypersurface singularity or a complete intersection are certain space forms
in dimension 4n + 3.

13.5.2. 3-Sasakian Cohomology as Primitive Cohomology. We are go-


ing to consider the connection between the cohomology of M and that of O by
applying the Vanishing Theorem 13.5.2 and a rational Gysin sequence argument to
the diagram of orbifold bundles of ♦(M ):
M −−−→ Z.


 .
(13.5.3) y

The horizontal arrow is equivalent to the study of the transverse Kähler geom-
etry treated in Chapter 7. In particular, Theorem 7.4.14 gives the relation between
the Betti numbers bp (M ) of M and the basic Betti numbers bB p (Fξ ). In the quasi-
regular case the basic Betti numbers are just the Betti numbers of Z by Proposition
7.2.2. So we have bp (M ) = bp (Z) − bp−2 (Z) for p ≤ 2n + 1. The analogue of this
for the vertical arrow is
Proposition 13.5.5: Let (M, S) be a compact 3-Sasakian manifold of dimension
4n + 3 and let O = M/F3 . Then b2p (M ) = b2p (O) − b2p−4 (O) for p ≤ 2n + 1.

Proof. The result follows by applying the rational Gysin sequence to the
orbifold fibration L → M → O using Corollary 4.3.8. We have
δ
· · · → H i (M, Q) → H i−3 (O, Q)→H i+1 (O, Q) → H i+1 (M, Q) → H i−2 (M, Q) → · · ·
and the statement of the proposition follows easily from the vanishing of the odd
Betti numbers of M. ¤
13.5. 3-SASAKIAN COHOMOLOGY 497

Recall the vector space of basic primitive harmonic p-forms P p (Fξ ) defined in
Section 7.2.2. In the quasi-regular case P p (Fξ ) is isomorphic to vector space of
primitive harmonic p-forms H0p (Z, Q) of the orbifold Z = M/Fξ which in turn is
isomorphic to the cokernel of the injective mapping LZ : Hp−2 (Z) ,→ Hp (Z), p ≤
2n defined by wedging with the Kähler 2-form. We define the primitive Betti
numbers b0p (Z) of Z as the dimension of H0p (Z). Then as usual the primitive Betti
numbers of Z are just the Betti numbers of M and it follows from the fact that
for, 0 ≤ r ≤ 2n + 1, an r-form on M is harmonic if and only if it is the lift of
a primitive harmonic form on Z [BG97b]. Similarly, for p ≤ 2n + 2 the vector
space of primitive harmonic p-forms H0p (O, Q) of the orbifold O is isomorphic to
the cokernel of the injective mapping LO : Hp−4 (O) ,→ Hp (O) defined by wedging
with the quaternionic Kähler 4-form Ω. The injectivity of this mapping is well-
known in the smooth case [Bon67, Fuj87, Kra65] and it extends to the orbifold
case. We define the primitive Betti numbers b0p (O) of O as the dimension of H0p (O).
Proposition 13.5.5 says that the primitive Betti numbers of O are the usual Betti
numbers of M. Again, Proposition 13.5.5 is a consequence of the fact that an r-form
on M is harmonic if and only if it is the lift of a primitive harmonic form on O,
0 ≤ r ≤ 2n + 1.
13.5.3. Regular 3-Sasakian Cohomology, Finiteness, and Rigidity. In
this Section we shall assume that M is regular and, hence, both Z and O are smooth.
In this case, using the results of the previous section, one can easily translate all
the results about strong rigidity of positive quaternionic Kähler manifolds given
in Section 12.3 to compact regular 3-Sasakian manifolds. In particular, Theorem
12.3.3 and the first part of Theorem 12.3.5 translate to
Proposition 13.5.6: Let (M, S) be a compact regular 3-Sasakian manifold of di-
mension 4n + 3. Then π1 (M ) = 0 unless M = RP4n+3 and
½
Z iff M = SU (n + 2)/S(U (n) × U (1)),
π2 (M ) =
f inite otherwise.
Furthermore, up to isometries, for each n ≥ 1 there are only finitely many regular
3-Sasakian manifolds M.
As with quaternionic Kähler manifolds, there are very interesting relations
among the Betti numbers of regular 3-Sasakian manifolds.
Proposition 13.5.7: The Betti numbers of a regular compact 3-Sasakian manifold
(M, S) of dimension 4n + 3 satisfy
(i) b2 (M ) ≤ 1, with equality if and only if M = SU (l + 2)/S(U (l) × U (1)),
Xn
(ii) k(n + 1 − k)(n + 1 − 2k)b2k (M ) = 0.
k=1

Proof. (i) follows from Proposition 13.5.6 and (ii) for Salamon’s relation on
Betti numbers of O in Theorem 12.3.5 (iv) using Proposition 13.5.5. ¤
We give below a table for the coefficients in item (ii) of Proposition 13.5.7.
Notice that the coefficient in (ii) changes sign under the map k 7→ n − k + 1.
So there is an interesting symmetry among the coefficients. Table 13.1 gives the
relations for 2 ≤ n ≤ 9 as well as the 3 remaining dimensions of the 3-Sasakian
homogeneous manifolds corresponding to the exceptional simple Lie algebras. In
this case to save space we just write the coefficients of the left hand side of the
498 13. 3-SASAKIAN MANIFOLDS

equality as the relation can be determined from those. We remark that these Betti
number relations do not hold in general for non-regular 3-Sasakian manifolds as
seen in Proposition 13.7.34 below.

n relation on Betti numbers or coefficients thereof


2 b2 = b4
3 b2 = b6
4 2b2 + b4 = b6 + 2b8
5 5b2 + 4b4 = 4b8 + 5b10
6 5b2 + 5b4 + 2b6 = 2b8 + 5b10 + 5b12
7 7b2 + 8b4 + 5b6 = 5b10 + 8b12 + 7b14
8 28b2 + 35b4 + 27b6 + 10b8 = 10b10 + 27b12 + 35b14 + 28b16
9 12b2 + 16b4 + 14b6 + 8b8 = 8b12 + 14b14 + 16b16 + 12b18
10 15, 21, 20, 14, 5
16 40, 65, 77, 78, 70, 55, 35, 12
28 126, 225, 299, 350, 380, 391, 385, 364, 330, 285, 231, 170, 104, 35
Table 13.1

Notice that one can satisfy (ii) of Proposition 13.5.7 by the ‘duality condition’
b2k (M ) = b2n−2k+2 (M ). Such 3-Sasakian manifolds were called balanced in [GS96].
Not all the 3-Sasakian homogeneous manifolds are balanced as can be seen by
looking at the Poincaré polynomials given in Table 13.2 below.
Proposition 13.5.8: The Poincaré polynomials of the homogeneous 3-Sasakian
manifolds are as given in Table 13.2.
G H P (G/H, t)
Pn
SU (n + 2) SU (n) ×Zn T 1 (t2i + t4n+3−2i )
Pi=0
k−1 4i 8k−1−4i
SO(2k + 3) SO(2k − 1) × SU (2) i=0 (t + t )
4n+3
Sp(n + 1) Sp(n) 1+t
Pl
SO(2l + 4) SO(2l) × SU (2) t2l + t6l+3 + i=0 (t4i + t8l+3−4i )
E6 SU (6) 1 + t6 + t8 + t12 + t14 + t20 + · · ·
E7 Spin(12) 1 + t8 + t12 + t16 + t20 + t24 + t32 + · · ·
E8 E7 1 + t12 + t20 + t24 + t32 + t36 + t44 + · · ·
F4 Sp(3) 1 + t8 + t23 + t31
G2 SU (2) 1 + t11
Table 13.2

Recall the LeBrun-Salamon Conjecture 12.3.7 which says that any positive
quaternionic Kähler manifold must be symmetric. This translates to
Conjecture 13.5.9: Every regular 3-Sasakian manifold is 3-Sasakian homoge-
neous.
The two well-known classification results for positive quaternionic Kähler man-
ifolds (see Theorem 12.3.8) easily translate into:
Theorem 13.5.10: Let (M, S) be a compact regular 3-Sasakian manifold of di-
mension 4n + 3. If n < 4 then M = G/H is homogeneous, and hence one of the
spaces listed in Theorem 13.4.6.
13.6. SYMMETRY REDUCTION 499

The n = 0 case is trivial and it was an observation made by Tanno [Tan71].


The n = 1 case is based on [Hit81, FK82] and it was first observed in [FK90,
BFGK91]. The n = 2 case is based on [PS91] and was stated in [BGM93]. Fi-
nally, the n = 3 follows from more recent classification of compact positive quater-
nion Kähler manifolds [HH02a, HH02b]. One can compute the Poincaré poly-
nomials of all the 3-Sasakian homogeneous manifolds of Theorem 13.4.6. This was
done in [GS96] to which we refer for the proof.
We end this section by giving the 3-Sasakian version of a theorem of Salamon
[GS96] the proof of which involves index theory. A condensed proof for the n = 3
case can be obtained using the result of [HH02a, HH02b] together with Table
13.2, but the proof of the Herrera-Herrera part of Theorem 12.3.8 itself uses index
theory.
Proposition 13.5.11: Let (M, S) be a regular compact 3-Sasakian manifold of
dimension 4n + 3. If n = 3, 4 and b4 = 0, then M is either a sphere S 4n+3 or a real
projective space RP4n+3 .

13.6. Symmetry Reduction


In this section we give a general 3-Sasakian reduction procedure which con-
structs new 3-Sasakian manifolds from a given 3-Sasakian manifold (M, S) with a
non-trivial 3-Sasakian isometry group Aut(M, S) [BGM94a]. Actually, this is a
reduction that is associated with a quadruple of spaces of the fundamental diagram
♦(M, S). At the level of the hyperkähler cone C(M ) the reduction was discov-
ered by Lindström and Roček [LR83] in the context of supersymmetric σ-model
and later rigorously described by Hitchin et al. in [HKLR87]. In the case of the
quaternionic Kähler base O the reduction was discovered by the second author and
Lawson [Gal87a, GL88]. In this section we restrict ourselves to describing the
general procedure of reduction together with the homogeneous case arising from
reduction by a circle group, as well as a brief description of the singular case. The
large class of 3-Sasakian toric manifolds obtained by reduction is relegated to a
separate section, namely Section 13.7. It should also be understood that every
3-Sasakian reduction gives as well a reduction procedure for each of the spaces of
the fundamental diagram ♦(M, S).
13.6.1. The 3-Sasakian Moment Map. Let (M, S) be a 3-Sasakian mani-
fold with a nontrivial group Aut(M, S) of 3-Sasakian isometries. By the Definition
13.1.9, C(M ) = M × R+ is a hyperkähler manifold with respect to the cone metric
ḡ. The isometry group Aut(M, S) extends to a group Aut(C(M ), I, ḡ) ∼ = Aut(M, S)
of isometries on C(M ) by defining each element to act trivially on R+ . Further-
more, it follows easily from the definition of the complex structures I a that these
isometries Aut(C(M ), I, ḡ) are hyperkähler; that is, they preserve the hyperkähler
structure on C(M ). Any subgroup G ⊂ Aut(C(M ), I, ḡ) gives rise to a hyperkähler
moment map µ : M − →g∗ ⊗ R3 , where g denotes the Lie algebra of G and g∗ is its
dual (see the discussion of Section 12.8). Thus, we can define a 3-Sasakian moment
map
(13.6.1) µS : M −−→ g∗ ⊗ R3
by restriction µS = µ |S . We denote the components of µS with respect to the
standard basis of R3 , which we have identified with the imaginary quaternions,
by µaS . Recall that ordinarily moment maps determined by Abelian group actions
500 13. 3-SASAKIAN MANIFOLDS

(in particular, those associated to 1-parameter groups) are only specified up to an


arbitrary constant. This is not the case for 3-Sasakian moment maps since we
require that the group Sp(1) generated by the Sasakian vector fields ξ a acts on the
level sets of µS . However, we shall see that 3-Sasakian moment maps are given by
a particularly simple expression.
Proposition 13.6.1: Let (M, S) be a 3-Sasakian manifold with a connected com-
pact Lie group G acting on M by 3-Sasakian isometries. Let τ be an element of
the Lie algebra g of G and let X τ denote the corresponding infinitesimal isometry.
Then there is a unique 3-Sasakian moment map µS such that the zero set µ−1 S (0)
is invariant under the group Sp(1) generated by the vector fields ξa . This moment
map is given by
1
hµaS , τ i = ηa (X τ ) .
2
Furthermore, the zero set µ−1
S (0) is G invariant.

Proof. Using the Definition 13.1.9 we can define the 2-forms ωSa on M as
the restriction of the hyperkähler 2-forms ω a . Then any 3-Sasakian moment map
µaS (τ ) determined by τ ∈ g satisfies 2dµaS (τ ) = 2X τ ωSa = −X τ dηa . As X τ
is a 3-Sasakian infinitesimal isometry, Exercise 13.4 implies that 2hµaS , τ i differs
from ηa (X τ ) by a constant depending on a and τ. One then uses the invariance of
the zero set µ−1
S (0) to show that these constants must vanish. See [BGM94a] for
details. ¤

Henceforth by the 3-Sasakian moment map, we shall mean the moment map
µS determined in Proposition 13.6.1. Hence, the Definition 13.1.9 and Proposition
13.6.1 imply
Theorem 13.6.2: Let (M, S) be a 3-Sasakian manifold with a connected com-
pact Lie group G ⊂ Aut(M, S) acting on M smoothly and properly. Let µS be
the corresponding 3-Sasakian moment map and assume both that 0 is a regular
value of µS and that G acts freely on the submanifold µ−1 S (0). Furthermore, let
ι : µ−1
S (0) ,→ M and π : µ−1
S (0)−−→µ −1
S (0)/G denote the corresponding embedding
and submersion. Then (M///G = µ−1 S (0)/G, ǧ) is a smooth 3-Sasakian manifold
of dimension 4(n − dim g) + 3 with metric ǧ and characteristic vector fields ξˇa
(hence, the reduced 3-Sasakian structure Š) determined uniquely by the two condi-
tions ι∗ g = π ∗ ǧ and π∗ (ξa | µ−1 ˇ
S (0)) = ξa .
We conclude this part with the following fact concerning 3-Sasakian isometries
whose proof can be found in [BGM94a].
Proposition 13.6.3: Assume that the hypothesis of Theorem 13.6.2 holds. In ad-
dition assume that (M, S) is complete and hence compact. Let C(G) ⊂ Aut(M, S)
denote the centralizer of G in Aut(M, S) and let C0 (G) denote the subgroup of
C(G) given by the connected component of the identity. Then C0 (G) acts on the
submanifold µ−1 ∗
S (0) as isometries with respect to the restricted metric ι g and the
3-Sasakian isometry group Aut(M///G, Š) of the quotient (M///G, Š) determined in
Theorem 13.6.2 contains an isomorphic copy of C0 (G). Furthermore, if C0 (G) acts
transitively on M///G, then M///G is a 3-Sasakian homogeneous space.
It should be mentioned that it is not required that the 3-Sasakian isometry
group Aut(M///G, Š) of the reduced space acts effectively.
13.6. SYMMETRY REDUCTION 501

13.6.2. Regular Quotients And Classical Homogeneous Metrics. We


now apply the reduction procedure given in Theorem 13.6.2 to the round unit sphere
S 4n+3 to explicitly construct the Riemannian metrics for the 3-Sasakian homoge-
neous manifolds arising from the simple classical Lie algebras. These metrics are
precisely the ones associated to the three infinite families appearing in Theorem
13.4.6. The quotient construction applied to ♦(S 4n+3 ) explicitly describes all met-
rics in the fundamental diagrams ♦(G/H), where G is either the special unitary
SU (n + 1) or the orthogonal group SO(n + 1). To carry out this reduction we
must set some conventions. We describe the unit sphere S 4n+3 by its embedding
in flat space and we represent an element u = (u1 , . . . , un+1 ) ∈ Hn+1 as a column
vector. The quaternionic components of this vector are denoted by u0 for the real
component and by ua for the three imaginary components so that we can write
u = u0 + iu1 + ju2 + ku3 using the quaternionic units {i, j, k}. We also define
quaternionic conjugate ū = u0 − iu1 − ju2 − ku3 .
Now, the infinitesimal generators of the subgroup Sp(1) ⊂ H∗ acting by the
right multiplication on u are the defining vector fields ξ a for the Sasakian 3-
structure. These vector fields are given by
∂ ∂ ∂
ξar = u0 · − ua · − ²abc ub · ,
∂ua ∂u0 ∂uc
where the dot indicates sum over the vector components ui and the subscript r
means that these vector fields are the generators of the right action.
We will first consider G = U (1) acting on the sphere S 4n+3 as follows
ϕt (u) = τ u, τ = e2πit , u ∈ S 4n+3 .
Note that this action on the cone C(S 4n+3 ) = Hn+1 \ {0} has been already con-
sidered in the Example 12.8.5, where the corresponding hyperkähler moment map
was computed in (12.8.3). The 3-Sasakian moment map µS : S 4n+3 −−→ R ⊗ sp(1)
is the restriction
n+1
X
µS (u) = − ūα iuα .
α=1

Let N be the zero locus of µS . Using the complex vectors (x, y) ∈ Cn+1 ×Cn+1 as in
(12.8.5) one can then easily identify the zero-level set of the moment map with the
Stiefel manifold of complex 2-frames in Cn+1 , i.e., N ' Vn+1,2
C
= U (n+1)/U (n−1).
The following proposition is an immediate consequence of the reduction theorem
Proposition 13.6.4: Let N = µ−1 S (0) and ι : N ,→ S
4n+3
be the inclusion. Then

ι is an embedding and (N, ι gcan ) is the complex homogeneous Stiefel manifold
C
V2,n+1 = SU (n+1)/SU (n−1) of 2-frames in Cn+1 . Hence, the 3-Sasakian quotient
S 4n+3 ///U (1) = V2,n+1
C
/U (1) = SU (n + 1)/S(U (n − 1) × U (1)) with the 3-Sasakian
metric ǧ given by inclusion ι and submersion π : N −−→ N/U (1), i.e., ι∗ gcan = π ∗ ǧ.
Example 13.6.5: A similar construction can be carried out for the Sp(1)-action
on S 4n+3 defined by the left multiplication of u by a unit quaternion σ, i.e.,
ϕσ (u) = σu, σσ̄ = 1, u ∈ S 4n+3 .
This action is free on S 4n+3 and the zero-level set of the corresponding moment
R
map can be identified with the real Stiefel manifold V4,n+1 ' SO(n + 1)/SO(n − 3)
n+1
of oriented 4-frames in R with n ≥ 4. Hence, the reduced space S 4n+3 ///Sp(1) =
502 13. 3-SASAKIAN MANIFOLDS

SO(n+1)
SO(n−3)×Sp(1) . For the more detailed and uniform description of the geometry of
these two quotients see [BGM94a].
13.6.3. The Structure of Singular Quotients. In this section we will de-
scribe a more general situation, when the zero-level set of the 3-Sasakian moment
map (13.6.1) is not necessarily smooth and the group action on the level set is not
necessarily locally free.
Let G be a Lie group acting smoothly and properly on a manifold S and let
H ⊂ G be a subgroup. Using standard notation we will denote by SH ⊂ S the set
of points in S, where the stability group is exactly equal to H and by S(H) ⊂ S the
set of points with stabilizer conjugate to H in G. It follows than that the normalizer
N (H) of H in G acts freely on S(H) . Then we have the following theorem due to
Dancer and Swann [DS97a]:
Theorem 13.6.6: Let (S, g) be a 3-Sasakian manifold with a connected compact
Lie group G acting on S smoothly and properly by 3-Sasakian isometries. Let µS
be the corresponding 3-Sasakian moment map. Then the quotient µ−1 S (0)/G is a
union of the smooth, 3-Sasakian manifolds (S(H) ∩ µ−1 S (0))/G, where (H) runs over
the conjugacy classes of stabilizers of points in S.
Quite often S(H) does not meet the zero locus of the moment map. Then the
stratum (S(H) ∩ µ−1 S (0))/G is empty.
Example 13.6.7: We start with the 3-Sasakian sphere S 4n+3 in the notation of the
previous section. But now we consider a different U (1)-action
ϕp,q;m
t (u) = (τ p u1 , . . . , τ p um , τ q um+1 , . . . , τ q un+1 ), τ = e2πit , u ∈ S 4n+3 ,
where 0 ≤ m ≤ n + 1 and p, q ∈ Z. Let S(p, q; m) = S 4n+3 ///U (1) be the quotient.
(i) First, let p, q be relatively prime positive integers bigger than 1 and 2 ≤ m ≤
n − 1. Then, the stratified manifold S(p, q; m) consists of 3 strata. The stratum of
the highest dimension corresponding to H = {id} is an open incomplete 3-Sasakian
manifold. The two strata of lower dimension are easily seen to be the homogeneous
spaces: one with H = Zq is the homogeneous 3-Sasakian space of SU (n + 1 − m)
and the one with H = Zp is the homogeneous 3-Sasakian space of SU (m). In this
case, S(p, q; m) is actually a compact 3-Sasakian orbifold and the stratification of
Theorem 7.3.1 coincides with the orbifold stratification.
(ii) Consider S(0, p; m), where p > 1 and 2 ≤ m ≤ n − 1. There are two strata
now: the stratum of the highest dimension corresponds to H = Zp and the second
stratum is just the sphere S 4m−1 with H = U (1). The space S(0, p; m) is not an
orbifold but, as pointed out in [DS97a], it does have a length space structure.
4n+3
(iii) Consider S(0, 1; n). Here H is either U (1) or trivial but the set Sid does
not meet the zero locus of the moment map. Hence, there is only one stratum and
S(0, 1; n) = S 4n−1 .
(iv) Finally, consider S(0, 1; n − 1). The stability group H is either U (1) or trivial.
The stratum corresponding to H = U (1) is the sphere S 4n−5 . We leave it as an
exercise to the reader to show that S(0, 1; n − 1) is an orbifold and that it can be
identified with S 4n−1 /Z2 , where (w1 , . . . , wn ) ∈ S 4n−1 , where Z2 acts on the last
quaternionic coordinate by multiplication by ±1.
Example 13.6.8: [Kobak-Swann Quotient] This is a less trivial example of a
singular quotient which we examine in considerable detail. Recall that G2 /SO(4)
is one of the three possible models of positive quaternionic Kähler manifolds in
13.6. SYMMETRY REDUCTION 503

dimension 8. It turns out that this space, as well as the associated homogeneous
3-Sasakian manifold can be obtained via symmetry reduction. Before we describe
the geometry of this quotient construction let us briefly discuss some of the facts
about the two homogeneous Sasaki-Einstein geometries that are naturally associ-
ated with the exceptional Lie group G2 . They both come from the classical Lie
group isomorphism between SO(4) ⊂ G2 and Sp(1)− · Sp(1)+ ⊂ G2 . The two
Sp(1)± subgroups are very different. One of them has index 1 in G2 and the
other one has index 3. Consequently, the quotients are not of the same homo-
topy type as can be seen from the exact sequence in homotopy for the fibration
Sp(1)± −→ G2 −→ G2 /Sp(1)± . In particular, the two spaces can be distin-
guished by their third homotopy groups being trivial in one case and Z3 in the
other. One of these quotients, which we shall denote by G2 /Sp(1)− is diffeomor-
phic to the real Stiefel manifold V7,2 (R) = SO(7)/SO(5) of 2-frames in R7 [HL82].
As V7,2 (R) is 4-connected π3 (G2 /Sp(1)− ) = 0. The other quotient denoted here
by G2 /Sp(1)+ is one of the 11-dimensional 3-Sasakian homogeneous spaces and
π3 (G2 /Sp(1)+ ) = Z3 . G2 /Sp(1)+ fibers as a circle bundle over a generalized flag
Z = G2 /U (2)+ , which in turn is well-known to be the twistor space of the ex-
ceptional 8-dimensional Wolf space G2 /SO(4). The second homogeneous Sasaki-
Einstein manifold is a circle bundle over the complex flag G2 /U (2)− which can be
identified with the complex quadric in the 6-dimensional complex projective space
CP6 or, equivalently, the real Grassmannian Gr2 (R7 ) = SO(7)/SO(2) × SO(5) of
oriented 2-planes in R7 . We have the following diagram of Riemannian submersions:

G2
. &
G2 G2
Sp(1)+ Sp(1)− ' V7,2 (R)

(13.6.2) ↓ ↓

G2 G2
Z= U (2)+ U (2)− ' Gr2 (R7 )
& .
G2
SO(4)

Kobak and Swann [KS93a] proved the following:


Theorem 13.6.9: The quaternionic Kähler manifold Gr4 (R7 ) admits a U (1)-action
such that the quaternionic Kähler quotient is a compact quaternionic Kähler orb-
ifold O = Or ∪ CP(2) = G2 /(SO(4) × Z3 ).
One can easily reinterpret the Kobak-Swann construction in the language of
3-Sasakian geometry of the Konishi SO(3)-orbibundle over O (or, equivalently, the
hyperkähler geometry of the regular nilpotent orbit of sl(3, C)). In particular, we
have the following:
Theorem 13.6.10: The 3-Sasakian homogeneous manifold SO(7)/SO(3) × Sp(1)
admits a U (1)-action such that the 3-Sasakian quotient is a compact 3-Sasakian
orbifold M = Mr ∪ SU (3)/U (1) = Z3 \G2 /Sp(1)+ .
Theorem 13.6.10 is a straightforward translation of Theorem 13.6.9 into the
language of 3-Sasakian geometry and we could leave it at that. However, we will
504 13. 3-SASAKIAN MANIFOLDS

outline a constructive proof of this result as we will need it to describe an interesting


modification of this particular quotient [BGP02].
Using the Example 13.6.5 one can think of the homogeneous 3-Sasakian man-
ifold SO(7)/SO(3) × Sp(1) as the 3-Sasakian reduction S 4n−1 ///Sp(1) : Let u =
(u1 , . . . , u7 ) ∈ S 27 . Consider the Sp(1) action given by multiplication by unit
quaternion λ ∈ Sp(1) on the left that is
(13.6.3) ϕλ (u) = λu .
In the {i, j, k} basis the 3-Sasakian moment maps for this action read:
7
X 7
X 7
X
(13.6.4) µi (u) = uα iuα , µj (u) = uα juα , µk (u) = uα kuα .
α=1 α=1 α=1

Then, the common zero-locus of the moment maps


(13.6.5) N = {u ∈ S 4n−1 | µi (u) = µj (u) = µk (u) = 0}
is the Stiefel manifold N ' SO(7)/SO(3) = V4 (R7 ) of the orthonormal 4-frames
in R7 and the corresponding 3-Sasakian quotient S = N/Sp(1) is Konishi’s RP3 -
bundle over the real Grassmannian of oriented 4-planes in R7 . We can combine
Theorem 13.6.10 with this description to get
Corollary 13.6.11: The 3-Sasakian sphere S 27 admits an action of U (1) × Sp(1)
such that the 3-Sasakian quotient is a compact 3-Sasakian orbifold M = Mr ∪
SU (3)/U (1) = Z3 \G2 /Sp(1)+ .
We now turn to the explicit description of the U (1) quotient. Consider the
following subgroups of the group of 3-Sasakian isometries of the 27-sphere:
(13.6.6) Sp(7) ⊃ SO(7) ⊃ 1 × SO(6) ⊃ 1 × U (3),
where U (1) ⊂ U (3) is the central subgroup. Explicitly, f : [0, 2π) → SO(7)
 
1 0 0 0 µ ¶
0 A(t) 0 0  cos t sin t

f (t) =   , A(t) = ∈ SO(2).
0 0 A(t) 0  − sin t cos t
0 0 0 A(t)
The homomorphism f (t) yields a U (1)-action on S 27 and, after performing Sp(1)
reduction first, on the homogeneous 3-Sasakian manifold SO(7)/SO(3) × Sp(1) via
left multiplication f (t)u and the associated 3-Sasakian moment map can be written
as
X
(13.6.7) ν(u) = (u2α u2α+1 − u2α+1 u2α ) .
α=1,2,3

Note here that ν(u) does not depend on the u1 quaternionic coordinate.
Definition 13.6.12: Let us define the zero level set of this new moment map in-
tersected with N , that is Nν ≡ N ∩ ν −1 (0).
First we observe, following Kobak and Swann [KS93a] that
Lemma 13.6.13: The manifold Nν can be identified with U (1)·G2 = (S 1 ×G2 )/Z3 ,
where U (1) ∩ G2 = Z3 .
Proof. The argument is similar here to the one used by Kobak and Swann in
[KS93a] and it is based on the Proposition 1.10 of [HL82]. First, using the basis
13.6. SYMMETRY REDUCTION 505

{i, j, k} of unit imaginary quaternions, we write uα = u0α + iu1α + ju2α + ku3α and
introduce the 4 × 7 real matrix
 0   0
u1 u02 u03 u04 u05 u06 u07 f
u11 u12 u13 u14 u15 u16 u17  f 1 
(13.6.8) A=   
u21 u22 u23 u24 u25 u26 u27  ≡ f 2  ,
u31 u32 u33 u34 u35 u36 u37 f3

where to make the connection with the notation in [KS93a] we also think of the
rows of A as purely imaginary octonions Im(O). In the standard basis of Im(O)
we write f a = ua1 i + ua2 j + ua3 k + ua4 e + ua5 ie + ua6 je + ua7 ke. Let φ(a, b, c) = hab, ci
denote the 3-form defining the associative calibration [HL82] on Im(O), where
h·, ·i denotes the standard Euclidean inner product. Then writing ν = ν1 i + ν2 j +
ν3 k a straightforward computation shows that for a = 1, 2, 3 and ²abc the totally
antisymmetric tensor satisfying ²123 = 1

(13.6.9) νa = 2hf 0 f a + ²abc f b f c , ii ,

where the summation convention on repeated indices is used. Now, u ∈ N if and


only if the rows {f 0 , f 1 , f 2 , f 3 } of A form an orthonormal frame in R7 ' Im(O), and
one can identify G2 with a special kind of oriented orthonormal 4-frame, namely
those which are co-associative. This means that the 3-plane that is orthogonal to the
4-plane defined by the frame {f 0 , f 1 , f 2 , f 3 } is spanned by an associative subalgebra
of Im(O). Then one shows that these special 4-frames satisfy the U (1)-moment map
equation ν(u) = 0 and, hence, U (1) · G2 ⊂ N ∩ ν −1 (0). As, U (1) ∩ G2 = Z3 it
is enough to show that by acting with U (1) one gets the whole N ∩ ν −1 (0). The
argument is similar to the one presented in [KS93a]. (See [KS93a] Lemma 5.1
and the discussion that follows.) ¤

Now, Theorems 13.6.9 and 13.6.10 and Corollary 13.6.11 all follow from the
above lemma as we get the quotient
Nν U (1) · G2
(13.6.10) M= ' ' Z3 \G2 /Sp(1).
U (1) × Sp(1) U (1) × Sp(1)
We will examine an interesting generalization of this quotient construction later in
Section 13.9.1.
Remark 13.6.1: The U (1) × Sp(1) action on the level set Nν is not locally free. If
we divide by Sp(1) first and consider the U (1)-action on the orbit space Nν /Sp(1)
this circle action is quasi-free. This means that there are only two kinds orbits:
regular orbits with the trivial isotropy group and singular orbits (points) where
the isotropy group is the whole U (1). In such cases the quotient space is often an
orbifold (or even a smooth manifold). The stratification of the Theorem 13.6.10 is
precisely with respect to the orbit types as will be seen in the next section.
Question 13.6.1: It has been pointed out by Kobak and Swann in [KS98] that
up to quotients by a finite group the 3-Sasakian manifold F4 /Sp(3) (or, equiva-
lently, the Wolf space F4 /Sp(3) · Sp(1)) can be also obtained via symmetry reduc-
tion, but the question remains open for the three homogeneous spaces of E-series
E6 /SU (6), E7 /Spin(12), and E7 /E8 . Can these be obtained via symmetry reduc-
tion of a sphere (up to a quotient by a finite group)?
506 13. 3-SASAKIAN MANIFOLDS

13.7. Toric 3-Sasakian Manifolds


In this section we shall describe the quotient construction of large families of
3-Sasakian manifolds (and orbifolds) S(Ω) with a 3-Sasakian structure SΩ which
are Konishi orbibundles over the toric orbifolds O(Ω) of Theorem 12.5.1. These
are characterized by the property that Aut(S(Ω), SΩ ) ⊃ T m , where dim(S(Ω)) =
4m−1. In particular, the hyperkähler cone C(S(Ω)) is toric in the sense of Definition
12.9.1 so that, following the ideas of [BD00], we shall call such 3-Sasakian manifolds
toric. Let us point out, however, that a toric 3-Sasakian manifold is not a toric
contact (Sasakian) manifold as defined in 8.4.32, but it is toric in the quaternionic
sense. Much of the treatment in this section is taken from [BGMR98, BGM98b]
and [BGM94a].
13.7.1. Toral Reductions of Spheres. Consider the torus action defined
at the beginning of Section 12.5, only lifted to Hn+1 and then restricted to the
embedded unit (4n + 3)-dimensional sphere S 4n+3 embedded in the quaternionic
vector space Hn+1 . Every quaternionic representation of a k-torus T k acting on
Hn+1 as a subgroup of Sp(n + 1) can be described by an exact sequence (12.5.2)
or its dual sequence (12.5.3). The monomorphism fΩ : T k −−→Sp(n + 1) can be
represented by the diagonal matrix as in Equation (12.5.4) with the weight matrix
Ω = (alα )l=1,...,k
α=1,...,n+1 ∈ Mk,n+1 (Z).
Let {el }kl=1 denote the standard basis for t∗k ' Rk . Then the 3-Sasakian moment
4n+3
map
P lµΩ : S −−→t∗k ⊗ R3 of the k-torus action defined by fΩ is given by µΩ =
l µΩ el with X
µlΩ (u) = ūα ialα uα .
α
Let us further denote the triple (T , fΩ , ϕ(τ1 ,...,τk ) ) by T k (Ω).
k

Definition 13.7.1: N (Ω) = µ−1


Ω (0) and S(Ω) = S
4n+3
///T k (Ω) = N (Ω)/T k (Ω).
4n+3
Let SH denote all the points on the sphere, where the stability group H ⊂ T k
4n+3 4n+3
is exactly H. Because T k is Abelian SH = S(H) . Furthermore, let KH = T k /H
4n+3
and denote by S(Ω; H) = SH ∩ N (Ω)/KH . Following Theorem 13.6.6 we have
S
Proposition 13.7.2: The quotient S(Ω) = H S(Ω; H) is a disjoint union of 3-
Sasakian manifolds, where each stratum S(Ω; H) is smooth.
We will be interested in the case when S(Ω) is a compact orbifold (all stability
4n+3
groups H for which SH ∩N (Ω) are non-empty are discrete) or a compact smooth
manifold (there is only one stratum). Necessary and sufficient conditions for this to
happen can be expressed in terms of properties of the matrix Ω. First observe that,
without loss of generality, we can assume that the rank of Ω equals k. Otherwise,
one simply has an action of a torus of lower dimension and the whole problem
reduces to considering another weight matrix Ω with fewer¡ rows.
¢
We introduce the following terminology: consider the nk minor determinants
 1 
aα1 . . . a1αk
 .. 
(13.7.1) ∆α1 ...αk = det  ... . 
akα1 ... akαk
obtained by deleting n + 1 − k columns of Ω.
Definition 13.7.3: Let Ω ∈ Mk,n+1 (Z) be the weight matrix.
13.7. TORIC 3-SASAKIAN MANIFOLDS 507

(i) If ∆α1 ······αk 6= 0, ∀ 1 ≤ α1 < · · · < αk ≤ n + 1, then we say that Ω is


non-degenerate.
(ii) Suppose Ω is non-degenerate and let g be the k th determinantal divisor,
i.e., the gcd of all the k by k minor determinants ∆α1 ······αk . Then Ω is
said to be admissible if in addition we have
gcd(∆α2 ···αk+1 , ..., ∆α1 ···α̂s ···αk+1 , ..., ∆α1 ···αk ) = g
for all sequences of length (k+1) such that 1 ≤ α1 < · · · < αs < · · · <
αk+1 ≤ n + 1.

13.7.2. Equivalence Problem and Admissibility. Before we show how


these properties of the matrix Ω impact on the geometry of the quotient S(Ω)
we need to discuss the notion of the equivalence of T k -actions on S 4n+3 and ob-
tain a normal form for admissible weight matrices. We are free to change bases
of the Lie algebra tk . This can be done by the group of unimodular matrices
GL(k, Z). Moreover, if we fix a maximal torus T n+1 of Sp(n + 1), its normal-
izer, the Weyl group W(Sp(n + 1)) ' Σn+1 ./ (Z2 )n+1 , preserves the 3-Sasakian
structure on S 4n+3 and intertwines the T k actions. Thus, there is an induced action
of GL(k, Z) × W(Sp(n + 1)) on the set of weight matrices Mk,n+1 (Z). The group
GL(k, Z) acts on Mk,n+1 (Z) by matrix multiplication from the left, and the Weyl
group W(Sp(n + 1)) acts by permutation and overall sign changes of the columns.
Actually we want a slightly stronger notion of equivalence than that described
above. If the ith row of Ω has a gcd di greater than one, then by reparameterizing
aj i
the one-parameter subgroup τi0 = τidi we obtain τi α = (τi0 )bα , where gcd{biα }α = 1.
So the action obtained by using the matrix whose ith row is divided by its gcd di
is the same as the original action. The integers di all divide the kth determinantal
divisor g. We say that a non-degenerate matrix Ω is in reduced form (or simply re-
duced) if g = 1. The following easy lemma says that among non-degenerate matrices
it is sufficient to consider matrices in a reduced form.
Lemma 13.7.4: Every non-degenerate weight matrix Ω is equivalent to a matrix
in a reduced form.
Henceforth, we shall only consider matrices in a reduced form.
Definition 13.7.5: Let Ak,n+1 (Z) ⊂ Mk,n+1 (Z) denote the subset of reduced ad-
missible matrices. This subset is invariant under the GL(k, Z) × W(Sp(n + 1))-
action so the set Ak,n+1 (Z)/GL(k, Z) × W(Sp(n + 1)) of equivalence classes [Ω] is
well-defined. We let Ek,n+1 (Z) ⊂ Ak,n+1 (Z) denote a fundamental domain for the
action.
Our interest in Ak,n+1 (Z) is the following:
Theorem 13.7.6: Let S(Ω) be the quotient space of definition 13.7.1. Then
(i) if Ω is non-degenerate, S(Ω) is an orbifold.
(ii) If Ω is degenerate, then either S(Ω) is a singular stratified space which
is not an orbifold or it is an orbifold obtained by reduction of a lower
dimensional sphere S 4n−4r−1 by a torus T k−r (Ω0 ) or a finite quotient
of such, where 1 ≤ r ≤ k and Ω0 is non-degenerate. (When r = k the
quotient is the sphere S 4n−4k−1 ).
(iii) Assuming that Ω is non-degenerate S(Ω) is a smooth manifold if and only
if Ω is admissible.
508 13. 3-SASAKIAN MANIFOLDS

One can easily see that the non-degeneracy of Ω is not necessary for the quotient
space S(Ω) to be smooth or a compact orbifold (see Example 13.6.7(iv)). However,
Theorem 13.7.6(ii) shows that then we can reformulate the whole problem in terms
of another quotient and a new non-degenerate weight matrix Ω0 and can be found in
[BGM98b]. Theorem 13.7.6(iii) shows then the importance of admissible matrices
in the construction and it easily follows from the fact that non-degeneracy implies
that at most n − k quaternionic coordinates uj can simultaneously vanish on N (Ω)
[BGMR98].
Remark 13.7.1: Our discussion shows clearly that, if Ω, Ω0 ∈ Ak,n+1 (Z) such that
[Ω] = [Ω0 ] then the quotients S(Ω) ' S(Ω0 ) are equivalent as 3-Sasakian manifolds.
We believe that the converse of this is also true, though we will establish it later
only in certain cases.

13.7.3. Combinatorics and Admissibility. In general Theorem 13.7.6(iii)


is not yet an existence theorem, since Ak,n+1 (Z) could be empty. Indeed, for many
pairs (k, n) this is the case and we shall demonstrate this next.
Let Ω ∈ Ak,n+1 (Z). Since Ω is reduced there is a k by k minor determinant
that is odd. By permuting columns if necessary this minor can be taken to be the
first k columns. Now consider the mod 2 reduction Mk,n+1 (Z)−−→Mk,n+1 (Z2 ). We
have the following commutative diagram
GL(k, Z) ×Mk,n+1 (Z) −−→ Mk,n+1  (Z)
 
(13.7.2)  
y y
GL(k, Z2 ) × Mk,n+1 (Z2 ) −−→ Mk,n+1 (Z2 ) .

Let Ω̃ ∈ Ak,n+1 (Z2 ) denote the mod 2 reduction of Ω ∈ Ak,n+1 (Z). Since the first
k by k minor determinant of Ω is odd, the mod 2 reduction of this minor in Ω̃ is
invertible. Thus, we can use the GL(k, Z2 ) action to put Ω̃ in the form

 
1 0 ... 0 a1k+1 ... a1n+1
0 1 . . . 0 a2k+1 ... a2n+1 
 
(13.7.3) Ω̃ =  . . . .. .. .. 
 .. .. .. . . ··· . 
0 0 ··· 1 akk+1 ... akn+1

with aij ∈ Z2 .
Lemma 13.7.7: The set Ak,n+1 (Z) is empty for n > k + 1 and k > 4.

Proof. The second admissibility condition is equivalent to the condition that


every k by k + 1 submatrix of Ω̃ has rank k. By considering k − 1 of the first k
columns and 2 of last n + 1 − k columns, this condition implies (ajl , ajm ) 6= (0, 0) for
all j = 1, . . . , k, and k + 1 ≤ l < m ≤ n + 1. Similarly, by considering k − 2 of the
first k columns and 3 of last n + 1 − k columns, (13.7.3) implies
µ i ¶ µ ¶ µ i ¶ µ ¶
al aim air 1 1 1 al aim air 0 1 1
(13.7.4) 6= , 6= ,
ajl ajm ajr 1 1 1 ajl ajm ajr 0 1 1
where the last inequality is understood to be up to column permutation. Hence, it
follows that, up to column and row permutations, that any four triples of the last
13.7. TORIC 3-SASAKIAN MANIFOLDS 509

n − k columns of an admissible Ω̃ must have the form


 
1 1 1
0 1 1
(13.7.5)  .
1 0 1
1 1 0
So we see that we cannot add another row without violating the above condi-
tions. It follows that k ≤ 4. ¤

Similar analysis shows that


Lemma 13.7.8: The set Ak,n+1 (Z) is empty if k > 1 and n − k ≥ 4.
Remark 13.7.2: In view of the above lemmas and the fact that in the remainder
of this section we will be interested only in the smooth and compact quotients we
are left with the following possibilities:
(i) Trivial case of n = k. Then there are many admissible matrices Ω but
dim(S(Ω)) = 3 and it follows that S(Ω) = S 3 /Zp , where p = p(Ω)
depends on Ω. This case is of little interest.
(ii) Bi-quotient geometry with k = 1 and n > 1 arbitrary. Here Ω is just
a row vector p. The admissibility condition means that the entries are
non-zero and pairwise relatively prime. The quotient S(p) turns out to
be a bi-quotient of the unitary group U (n + 1) and we shall discuss its
geometry and topology in the next subsection.
(iii) The most interesting, 7-dimensional case of k = n−1. Here one easily sees
that there are many admissible matrices and we analyze the geometry
and topology of the quotients in a separate subsection. In particular,
when k = 1 we will completely determine the diffeomorphism type of the
reduced space S(p1 , p2 , p3 ) showing that there are examples of 3-Sasakian
7-manifolds with more than one 3-Sasakian structure S.
(iv) “Special” quotients with: (k, n) = {(2, 4), (2, 5), (3, 5), (3, 6), (4, 6), (4, 7)}.
These quotients are 11- or 15-dimensional and we give examples of ad-
missible weight matrices in each case. We shall show also that they
provide counterexamples to certain Betti number relations of Theorem
13.5.7 which are always satisfied in the regular case.

13.7.4. 3-Sasakian Structures as Bi-Quotients. When k = 1 we have


Ω = p = (p1 , . . . , pn+1 ) and we shall write S(Ω) = S(p), N (Ω) = N (p), fΩ = fp ,
µΩ = µp , and τ1 = τ. The quotients S(p) are generalizations of the homogeneous
examples discussed in Section 13.6.2. We get
A1,n+1 (Z) = {p ∈ (Z)n+1 | pi 6= 0 ∀i = 1, . . . , n+1 and gcd(pi , pj ) = 1 ∀i 6= j},

E1,n+1 (Z) = {p ∈ Zn+1 | 0 < p1 ≤ · · · ≤ pn+1 and gcd(pi , pj ) = 1 ∀i 6= j} .


Note that E1,n+1 (Z) can be identified with a certain integral lattice in the positive
Weyl chamber in t∗n+1 .
First, by studying the geometry of the foliations in the diagram ♦(S(p))
[BGM96a] one can solve the equivalence problem in this case. We get [BGM96b]:
Proposition 13.7.9: Let n ≥ 2 and p, q ∈ A1,n+1 (Z) so the quotients S(p) and
S(q) are smooth manifolds. Then S(p) ' S(q) are 3-Sasakian equivalent if and
only if [p] = [q].
510 13. 3-SASAKIAN MANIFOLDS

It is easy to see that for p ∈ A1,n+1 (Z) the zero locus of the moment map N (p)
C
is always diffeomorphic to the Stiefel manifold V2,n+1 of complex 2-frames in Cn+1 .
More generally,
Proposition 13.7.10: For each p ∈ (Z∗ )n the zero locus N (p)is diffeomorphic to
C
the complex Stiefel manifold Vn,2 .
C
Hence, we have the quotient S(p) = V2,n+1 /S 1 . We first observe that one
C
can identify V2,n+1 with the homogeneous space U (n + 1)/U (n − 1). Using this
identification we have
Proposition 13.7.11: For each p ∈ E1,n+1 (Z), there is an equivalence S(p) '
U (1)p \U (n + 1)/U (n − 1) as smooth U (1)p × U (n − 1)-spaces, where the action of
U (1)p × U (n − 1) ⊂ U (n + 1)L × U (n + 1)R is given by the formula
µ ¶
1l2 O
ϕpτ,B (W) = f p (τ )W .
O B

Here W ∈ U (n + 1) and (τ, B) ∈ S 1 × U (n − 1).


Note that the identification S(p) ' U (1)p \U (n + 1)/U (n − 1) is only true after
assuming that all the weights are positive, as the right-hand side is not invariant
under such sign changes. Proposition 13.7.11 shows that, in a way, the quotients
S(p) can be though of as a discrete “bi-quotient deformation” of the homogeneous
model S(1). Now let ιp : N (p) ,→ S 4n+3 be the inclusion and πp : N (p) →
S(p) be the Riemannian submersion of the moment map. Then the metric g(p) is
the unique metric on S(p) that satisfies ι∗p gcan = πp∗ g(p). Similarly we define the
other structure tensors of the 3-Sasakian structure Sp . Using the geometry of the
inclusion ιp one can show the following [BGM96b, BGM96a]
Theorem 13.7.12: Let¡ Aut(S(p), Sp ) be the group of 3-Sasakian isometries of
the manifold (orbifold) S(p), Sp ) and let k be the number of 1’s in p. Then the
connected component of Aut(S(p), Sp ) is S(U (k) × U (1)n+1−k ), where we define
U (0) = {id}. Thus, the connected component of the isometry group is the product
S(U (k) × U (1)n+1−k ) × SO(3) if the sums pi + pj are even for all 1 ≤ i, j ≤ n + 1,
and S(U (k) × U (1)n+1−k ) × Sp(1) otherwise.
In the case that p has no repeated 1’s, the cohomogeneity can easily be deter-
mined, viz. [BGM96b]
Corollary 13.7.13: If the number of 1’s in p is 0 or 1 then the dimension of the
principal orbit in S(p) equals n + 3 and the cohomogeneity of g(p) is 3n − 4. In par-
ticular, the 7-dimensional family (S(p), Sp ) contains metrics g(p) of cohomogeneity
0, 1, and 2.
Combining Proposition 13.7.11 with techniques developed by Eschenburg [Esc82,
Esc92] in the study of certain 7-dimensional bi-quotients of SU (3) one can compute
the integral cohomology ring of S(p) [BGM94a]:
Theorem 13.7.14: Let n ≥ 2 p ∈ E1,n+1 (Z). Then, as rings,
à !
¡ ¢ Z[b2 ]

H S(p), Z = ∼ ⊗ E[f2n+1 ] /R(p) .
[bn+1
2 = 0]

Here the subscripts on b2 and f2n+1 denote the cohomological dimension of each
generator. Furthermore, the relations R(p) are generated by σn (p)bn2 = 0 and
13.7. TORIC 3-SASAKIAN MANIFOLDS 511

Pn+1
f2n+1 bn2 = 0, where σn (p) = j=1 p1 · · · pˆj · · · pn+1 is the nth elementary symmet-
ric polynomial in the entries of p.
Notice that Theorem 13.7.14 shows that H 2n (S(p); Z) = Zσn (p) and hence
has the following corollary.
Corollary 13.7.15: The quotients (S(p), g(p)) give infinitely many homotopy in-
equivalent simply-connected compact inhomogeneous 3-Sasakian manifolds in di-
mension 4n − 1 for every n ≥ 2. In fact, there are infinite families that are not
homotopy equivalent to any homogeneous space.
Remark 13.7.3: Corollary 13.7.15 shows that the finiteness results for regular 3-
Sasakian manifolds discussed in Section 13.5.3 fail for non-regular 3-Sasakian man-
ifolds. Moreover, combining our results with a well-known finiteness theorem of
Anderson [And90] we have
Corollary 13.7.16: For each n ≥ 2 there are infinitely many 3-Sasakian (4n − 1)-
manifolds with arbitrarily small injectivity radii.

13.7.5. S(p1 , p2 , p3 ), Eschenburg Spaces, and Diffeomorphism Type.


We will now examine the n = 2 case in some detail. In particular we will show
how to determine the smooth structure. It can be seen that the spaces S(p) =
S(p1 , p2 , p3 ) are a special subfamily of more general bi-quotients of SU (3). We
have
Definition 13.7.17: Let k = (k1 , k2 , k3 ) ∈ Z3 , l = (l1 , l2 , l3 ) ∈ Z3 be such that
1
k1 + k2 + k3 = l1 + l2 + l3 . Consider the circle action Sk,l defined on SU (3) by

ϕk,l k1 k2 k3 k1 k2 k3
τ (A) = diag(τ , τ , τ ) · A · diag(τ̄ , τ̄ , τ̄ ) ,

where A ∈ SU (3) and |τ |2 = 1. Suppose (k, l) are chosen so that the associated
1
circle action Sk,l is free. Then the quotient
1
Mk,l = SU (3)/Sk,l ,
is a smooth compact manifold and it is called Eschenburg bi-quotient.
The action is free if and only if diag(τ k1 , τ k2 , τ k3 ) is not conjugate in SU (3) to
diag(τ l1 , τ l2 , τ l3 ). This translates into the following conditions, which must all be
satisfied:
gcd(k1 − l1 , k2 − l2 ) = 1, gcd(k1 − l2 , k2 − l1 ) = 1,
(13.7.6) gcd(k1 − l1 , k2 − l3 ) = 1, gcd(k1 − l2 , k2 − l3 ) = 1,
gcd(k1 − l3 , k2 − l1 ) = 1, gcd(k1 − l3 , k2 − l2 ) = 1.
It is easy to see compute the cohomology ring of Mk,l . In particular, the non-
trivial cohomology groups are H i (Mk,l , Z) = Z for i = 0, 2, 5, 7 and H 4 (Mk,l , Z) =
Z|r| , where r = r(k, l) = σ2 (k) − σ2 (l). These spaces were introduced by Eschen-
burg in 1982 showing that some of them admit positive sectional curvature metrics
[Esc82]. At that time these were the first examples of inhomogeneous Riemannian
metrics with this property. More precisely, later in [Esc84], Eschenburg proved
that
Proposition 13.7.18: A bi-quotient metric on Mk,l has positive sectional curva-
ture if and only if either li 6∈ [min{k1 , k2 , k3 }, max{k1 , k2 , k3 }] for each i = 1, 2, 3
or ki 6∈ [min{l1 , l2 , l3 }, max{l1 , l2 , l3 }] for each i = 1, 2, 3.
512 13. 3-SASAKIAN MANIFOLDS

Let us note that a given Eschenburg space can be represented as a bi-quotient


in many different ways. For example, we can change the order of the ki ’s in k or
the li0 s in l or switch all k with l. We can also simultaneously shift both k, l by
a = (a, a, a) or replace (k, l) by (−k, −l). All such operations do not change the bi-
quotient manifold. However, there is always a canonical choice. More importantly
after removing this redundance there are only finitely many Eschenburg bi-quotients
for each given r = r(k, l). More precisely
Proposition 13.7.19: Any positively curved Eschenburg space Mk,l has a unique
representative of the following form (k|l) = (a, b, c + d − a − b|0, c, d), where a ≥
b > c ≥ d ≥ 0. Furthermore, for each r = r(k, l) = σ2 (k) − σ2 (l) there are finitely
many such representatives.
This large collection of spaces contains several interesting subfamilies. One of
them is the homogeneous family (k|l) = (m, n, −n − m|0, 0, 0) of the so-called Aloff-
7
Wallach spaces Mmn [AW75] which are of much interest since they admit Einstein
metrics of positive sectional curvature [Wan82]. The importance of these spaces
to 3-Sasakian geometry follows from the following simple observation [BGM94a]
Proposition 13.7.20: S(p) = Mk,l with (k|l) = (p1 , p2 , p3 |0, 0, p1 + p2 + p3 ). In
particular, all 7-manifolds S(p) are Eschenburg bi-quotients and they have metrics
of positive sectional curvature.
Exercise 13.5: Prove Proposition 13.7.20 by writing an explicit identification of
S(p) and the bi-quotient M (p1 , p2 , p3 |0, 0, p1 + p2 + p3 ).
Note, however, that the positive sectional curvature bi-quotient metrics of Es-
chenburg has essentially nothing to do with the 3-Sasakian metric on S(p) which
necessarily always has negatively curved sections.
Because of their curvature properties the Eschenburg spaces M (k|l) have been
studied in great detail. Homotopy invariants for Eschenburg space have been
worked out independently by Kruggel [Kru98, Kru97] and Milgram [Mil00]. The
homeomorphism and diffeomorphism classification was first done for a certain sub-
class of Eschenburg spaces which include some of the S(p) spaces by Astey, Micha,
and Pastor [AMP97]. Later Kruggel [Kru05] obtained the diffeomorphism and
homeomorphism classification of a more general class of Eschenburg’s bi-quotients
by computing the Kreck-Stolz invariants [KS88]. In his construction one needs a
condition on (k, l), namely that the matrix aij = ki − lj contains at least one row
or column with pairwise relatively prime entries. Kruggel calls it condition (C) and
admits that he does not know if there are any Eschenburg spaces which satisfy it.
More recently Chinburg, Escher and Ziller found that there are many Eschenburg
spaces which satisfy this condition [CEZ07]. Note, however, that this condition is
automatically satisfied for the 3-Sasakian family S(p). Let us describe this classi-
fication and some of its important consequences. We closely follow the notation of
[CEZ07].
First, besides r(k, l) we can compute the first Portraying class and the linking
number of Mk,l . For the linking form we get

(13.7.7) Lk(k, l) = −s/r ∈ Q/Z ,


13.7. TORIC 3-SASAKIAN MANIFOLDS 513

where s = s(k, l) = σ3 (k) − σ3 (l). The Portraying class p1 (Mk,l ) = p1 (k, l) mod r
in Z|r| is
3
X 3
X
(13.7.8) p1 (k, l) = (ki − l1 )2 + (ki − l2 )2 − (l1 − l2 )2 .
i=1 i=1

Next one needs to compute two more Q/Z-invariants introduced by Kreck and Stolz
[KS88]. Let us assume that condition (C) holds for the j th column, there are at
1
most three exceptional orbits for the Sk,l action on the cobordism of SU (3) with
isotropy groups Z|k1 −lj | , Z|k2 −lj | and Z|k3 −lj | . After removing small equivariant
neighborhoods of these orbits the action becomes free and the quotient is a smooth
eight dimensional manifold Wk,l with boundary ∂(Wk,l ) = Mk,l ∪ L1 ∪ L2 ∪ L3 ,
where the Li are the following seven dimensional lens spaces.

L1 = L(k1 − lj ; k2 − lj , k3 − lj , k2 − l[j+1]2 , k3 − l[j+1]2 ) ,


(13.7.9) L2 = L(k2 − lj ; k1 − lj , k3 − lj , k1 − l[j+1]2 , k3 − l[j+1]2 ) ,
L3 = L(k3 − lj ; k1 − lj , k2 − lj , k1 − l[j+1]2 , k2 − l[j+1]2 ) ,
where we used the notation [n]p = m if n = λ · p + m for m = 1, . . . , p, for the
residue class [n] modulo p.
Since the invariants are additive with respect to unions, we obtain si (Wk,l ) =
si (Mk,l ) + si (L1 ) + si (L2 ) + si (L3 ) ∈ Q/Z , i = 1, 2 . Calculating si (Wk,l ), yields
the following expressions for the Kreck-Stolz invariants, which hold in the case
condition (C) is satisfied for the j th column.

(13.7.10)
4 | r(k, l) (k1 − lj ) (k2 − lj ) (k3 − lj ) | − q(k, l)2
s1 (Mk,l ) =
27 · 7 · r(k, l) (k1 − lj ) (k2 − lj ) (k3 − lj )
3
X
− s1 (k[i]3 − lj ; k[i+1]3 − lj , k[i+2]3 − lj , k[i+1]3 − l[j+1]2 , k[i+2]3 − l[j+1]2 )
i=1
q(k, l) − 2
s2 (Ek,l ) =
24 · 3 · r(k, l) (k1 − lj ) (k2 − lj ) (k3 − lj )
3
X
− s2 (k[i]3 − lj ; k[i+1]3 − lj , k[i+2]3 − lj , k[i+1]3 − l[j+1]2 , k[i+2]3 − l[j+1]2 ) ,
i=1

where

q(k, l) = (k1 − lj )2 + (k2 − lj )2 + (k3 − lj )2 +


+ (k1 − l[j+1]2 )2 + (k2 − l[j+1]2 )2 + (k3 − l[j+1]2 )2 − (lj − l[j+1]2 )2
and si (p; p1 , p2 , p3 , p4 ) = si (Lp (p1 , p2 , p3 , p4 )) ∈ Q/Z , i = 1, 2 are the Kreck-Stolz
invariants of the lens space Lp (p1 , p2 , p3 , p4 ) = S 7 /Zp . The freeness condition of
the action (13.7.6) implies that ki − lj 6= 0 for i, j = 1, 2, 3, and, hence, the above
expressions for s1 and s2 are well-defined.
By the Atiyah-Patodi-Singer index theorem the Kreck-Stolz invariants can
also be expressed as linear combinations of eta-invariants. Calculating these eta-
invariants for the lens spaces, one obtains:
514 13. 3-SASAKIAN MANIFOLDS

(13.7.11)
|d|−1 4
X Y µ ¶ |d|−1 4 µ ¶
1 kπdj 1 X Y kπdj
s1 (d; d1 , d2 , d3 , d4 ) = 5 cot + 4 csc ;
2 ·7·d j=1
d 2 ·d j=1
d
k=1 k=1
|d|−1
X 2πık 4
Y µ ¶
1 kπdj
s2 (d; d1 , d2 , d3 , d4 ) = (e |d| − 1) csc .
24·d j=1
d
k=1

For d = ±1 these expressions are interpreted to be 0. These formulas only hold


in the case that condition (C) is satisfied for the j th column. Similar formulas can be
derived in the case where the j th row consists of relatively prime entries [CEZ07].
Using these invariants we now state the classification theorems by Kruggel for the
Eschenburg spaces [Kru05, CEZ07].
Theorem 13.7.21: Assume the Eschenburg spaces Mk,l and Mk0 ,l0 both satisfy
condition (C). Then
1. Mk,l and Mk0 ,l0 are (orientation preserving) homeomorphic if and only if

(a) |r(k, l)| = |r(k0 , l0 )| ∈ Z ,


(b) Lk(k, l) ≡ Lk(k0 , l0 ) ∈ Q/Z ,
(c) p1 (k, l) ≡ p1 (k0 , l0 ) ∈ Z|r(k,l)| ,
(d) s2 (Mk,l ) ≡ s2 (Mk0 ,l0 ) ∈ Q/Z .

2. Mk,l and Mk0 ,l0 are (orientation preserving) diffeomorphic if and only if
in addition
s1 (Mk,l ) ≡ s1 (Mk0 ,l0 ) ∈ Q/Z .

3. Mk,l and Mk0 ,l0 are (orientation preserving) homotopy equivalent if and
only if

(a) |r(k, l)| = |r(k0 , l0 )| ∈ Z ,


(b) Lk(k, l) ≡ Lk(k0 , l) ∈ Q/Z ,
(c) s22 (Mbf k,l ) ≡ s22 (Mk0 ,l0 ) ∈ Q/Z ,

where s22 (Mk,l ) = 2 |r(k, l)| s2 (Mk,l ).

For the corresponding theorem in the orientation reversing case the linking
number and the Kreck-Stolz invariants change signs. Recall that in this theorem,
r(k, l) = σ2 (k) − σ2 (l), p1 (Mk,l ) = [2 σ1 (k)2 − 6 σ2 (k)] · u2 ∈ Z|r| , and the equality
of the linking forms can replaced by the equality of the numbers s(k, l) = σ3 (k) −
σ3 (l) ∈ Z|r(k,l)| .
Remark 13.7.4: In [KS88] Kreck-Stolz used another invariant s3 in the homeomor-
phism classification, and showed that r(k, l), s2 and s3 determine the homeomor-
phism type. Following [Kru05], the formula for the invariant s3 for the Eschenburg
spaces is easily seen to be:
13.7. TORIC 3-SASAKIAN MANIFOLDS 515

q(k, l) − 8
s3 (Ek,l ) =
22 · 3 · r(k, l) (k1 − lj ) (k2 − lj ) (k3 − lj )
3
X
− s3 (k[i]3 − lj ; k[i+1]3 − lj , k[i+2]3 − lj , k[i+1]3 − l[j+1]2 , k[i+2]3 − l[j+1]2 ),
i=1

where
|p|−1
X 4πık Y4 µ ¶
1 kπpj
s3 (p; p1 , p2 , p3 , p4 ) = (e |p| − 1) csc .
24 · p j=1
p
k=1

Remark 13.7.5: In the case of 3-Sasakian manifold S(p) all relevant invariants
other than Kreck-Stolz invariants can be written in terms of the elementary sym-
metric functions σi (p). We have r = r(p) = σ2 (p), s = s(p) = 2σ12 (p) − 4σ2 (p).
Theorem 13.7.21 can be used to obtain two important results concerning the
differential topology of 3-Sasakian manifolds. Like in the Sasaki-Einstein case there
exists 3-Sasakian manifolds which are homeomorphic but not diffeomorphic. Fur-
thermore, in spite of the local rigidity, there are non-isometric 3-Sasakian structures
on the same smooth manifold. Both results are expected, however, the examples
are relatively hard to find, and, somewhat curiously the known ones occur in pairs.
Theorem 13.7.22: For r = σ2 (p) ≤ 107 , there are exactly 138 pairs of 3-Sasakian
spaces S(p), S(p0 ) which are homeomorphic to each other, but not diffeomorphic.
The first few pairs, together with their relevant invariants, are listed in the
table below.

p r(p) p1 (p) s(p) s2 (p) s1 (p)


2393 82869
(171, 164, 1) 28379 27139 −335 − 56758 − 3178448
2393
(223, 60, 53) 28379 27139 −335 − 56758 − 1104513
3178448
80901 69409
(362, 291, 37) 129503 12564 45679 − 259006 14504336
80901 5767541
(423, 169, 98) 129503 12564 45679 − 259006 14504336
370663 393315
(717, 362, 13) 273581 91230 196280 1094324 − 1094324
370663 310179
(761, 241, 90) 273581 91230 196280 1094324 1094324
294993 74669
(891, 368, 43) 382025 −35741 334208 − 1528100 − 436600
294993 1442017
(928, 191, 183) 382025 −35741 334208 − 1528100 3056200
115166
(1265, 347, 2) 442179 −6448 346023 1326537 − 173889
611408
115166 21037
(1274, 311, 29) 442179 −6448 346023 1326537 − 611408

The search in [CEZ07] was performed with help of a computer code which
originally had some errors. L. Florit further ran corrected C codes of [CEZ07]
to find more examples with r = σ2 (p) ≤ 1.5 · 107 . He found an additional 83
homeomorphic pairs and one new pair that is diffeomorphic. In particular, we have
516 13. 3-SASAKIAN MANIFOLDS

Theorem 13.7.23: For r = σ2 (p) ≤ 107 , there are two pairs of 3-Sasakian spaces
S(p), S(p0 ) which are diffeomorphic to each other, but not isometric. They are
given by
(i) p = (2279, 1603, 384) and p0 = (2528, 939, 799) with r = 5143925.
(ii) p = (4219, 2657, 217) and p0 = (4637, 1669, 787) with r = 12701975.

13.7.6. 7-dimensional Toric 3-Sasakian Manifolds. First, let us observe


that as a consequence of Calderbank-Singer classification of all compact toric pos-
itive self-dual Einstein orbifolds in [CS06a] we have the following 3-Sasakian ana-
logue of Theorem 12.5.7.
Theorem 13.7.24: Any toric 7-dimensional compact 3-Sasakian orbifold is iso-
morphic to one of the quotients S(Ω), where Ω ∈ Mk,k+2 (Z) is some non-singular
weight matrix.
In this case we can easily see that there are infinitely many examples of admis-
sible weight matrices Ω for which S(Ω) is smooth. The simplest family of examples
is given by matrices of the form
 
1 0 . . . 0 a1 b1
 
0 1 . . . 0 a2 b2 
 
(13.7.12) Ω = . . . ..  ,
 .. .. . . ... ... .
 
0 0 · · · 1 ak bk
for which we have
Proposition 13.7.25: Let k be a positive integer, and let Ω ∈ Mk,k+2 (Z) be as
in 13.7.12. Then Ω ∈ Ak,k+2 (Z) if and only if (a, b) ∈ (Z∗ )k ⊕ (Z∗ )k and the
components (ai , bi ) are pairs of relatively prime integers for i = 1, · · · , k such that
if for some pair i, j ai = ±aj or bi = ±bj then we must have bi 6= ±bj or ai 6= ±aj ,
respectively.
Proposition 13.7.25 shows that Ak,k+2 (Z) is never empty and we have many
examples of compact smooth 7-dimensional quotients S(Ω) for arbitrary k > 1.
Some of these examples were first mentioned in [BGM93] and the idea of the
quotient is based on the result of [GN92]. As we shall not present here the complete
solution to the equivalence problem, we shall further assume that Ω ∈ Ak,k+2 (Z) is
arbitrary and shall determine some important topological properties of the quotients
S(Ω). More explicitly [BGMR98, BGM98a],
Theorem 13.7.26: Let Ω ∈ Ak,k+2 (Z). Then π1 (S(Ω)) = 0 and π2 (S(Ω)) = Zk .

Proof. First note that the groups T k+2 ×Sp(1) and T 2 ×Sp(1) act as isometry
groups on N (Ω) and S(Ω), respectively. Let us define the following quotient spaces:
Q(Ω) = N (Ω)/T k+2 × Sp(1), B(Ω) = N (Ω)/Sp(1) .
We have the following commutative diagram

N (Ω) −−−−−→ B(Ω)


 
 
(13.7.13)  
y y
S(Ω) −−−−−→ Q(Ω) .
13.7. TORIC 3-SASAKIAN MANIFOLDS 517

The top horizontal arrow and the left vertical arrow are principal bundles with
fibers Sp(1) and T k , respectively. The remaining arrows are not fibrations. The
right vertical arrow has generic fibers T k+2 , while the lower horizontal arrow has
generic fibers T 2 ·Sp(1) homeomorphic either to T 2 × RP3 or T 2 × S 3 depending on
Ω. The dimension of the orbit space Q(Ω) is 2.
The difficulty is in proving that both N (Ω) and B(Ω) are 2-connected. Once
this is accomplished the result follows by applying the long exact homotopy se-
quence to the left vertical arrow in diagram 13.7.13.
Lemma 13.7.27: Both N (Ω) and B(Ω) are 2-connected.
To prove this lemma we construct a stratification giving a Leray spectral se-
quence whose differentials can be analyzed. Let us define the following subsets of
N (Ω) : (Recall that, in this case, at most one quaternionic coordinate can vanish.)
N0 (Ω) = {u ∈ N (Ω)| uα = 0 for some α = 1, . . . , k + 2},
N1 (Ω) = {u ∈ N (Ω)| for all α = 1, . . . , k + 2, uα 6= 0 and there is a pair (uα , uβ )
that lies on the same complex line in H},
N2 (Ω) = {u ∈ N (Ω)| for all α = 1, . . . , k + 2, uα 6= 0 and no pair (uα , uβ )
lies on the same complex line in H} .
Clearly, N (Ω) = N0 (Ω)tN1 (Ω)tN2 (Ω) and N2 (Ω) is a dense open submanifold
of N (Ω). This stratification is compatible with the diagram 13.7.13 and induces
corresponding stratifications
(13.7.14) B(Ω) = B0 (Ω) t B1 (Ω) t B2 (Ω), Q(Ω) = Q0 (Ω) t Q1 (Ω) t Q2 (Ω).
The Bi (Ω) fiber over the Qi (Ω) whose fibers are tori T k+i . The strata are labelled
by the dimension of the cells in the resulting CW decomposition of Q(Ω). Using
known results about cohomogeneity 2 actions [Bre93] one can easily prove:
Lemma 13.7.28: The following are true
(i) The orbit space Q(Ω) is homeomorphic to the closed disc D̄2 , and the
subset of singular orbits Q1 (Ω) t Q0 (Ω) is homeomorphic to the boundary
∂ D̄2 ' S 1 .
(ii) Q2 (Ω) is homeomorphic to the open disc D2 .
(iii) Q1 (Ω) is homeomorphic to the disjoint union of k + 2 copies of the open
unit interval.
(iv) Q0 (Ω) is a set of k + 2 points.
After making certain choices Lemma 13.7.28 can be used to represent Q(Ω)
topologically as a polygon as illustrated in Figure 1.
Next one can easily show that π1 (B(Ω)) is Abelian; hence, π1 (B(Ω)) = H1 (B(Ω)).
Now we claim that H1 (B(Ω)) = H2 (B(Ω)) = 0, and since the fibers of the top hor-
izontal arrow of 13.7.13 are S 3 ’s this together with Hurewicz will prove Lemma
13.7.27. To prove this claim we define Y0 = Q0 (Ω), Y1 = Q0 (Ω) ∪ Q1 (Ω), and
Y2 = Q(Ω). Then, we filter B(Ω) by Xi = π −1 (Yi ) to obtain the increasing filtra-
tion
X0 = B0 (Ω), X1 = B0 (Ω) ∪ B1 (Ω), and X2 = B(Ω) .
The Leray spectral sequence associated to this filtration has E 1 term given by
1
Es,t ∼
= Hs+t (Xt , Xt−1 ; Z)
518 13. 3-SASAKIAN MANIFOLDS

vk
.......
.... •
....... .............
.......
.......
.........
.
.......
.......
.. .......
.
....... .......
....... k
.......
e
.......
.......
.......
.......
.....
.....
...
• vk+2
...
...
...
...
...
.. ...
...
. 2 σ ...
...
ek+2
...
.... ...
... ...
... ...
...
... ...
... ...
... ...
v2 • ......
.......
....... .
.. .......... • vk+1
....... .....
.....
.......
....... ..
.........
...
....... .......
e ..
1 .................. .......
.......
......
k+1 e
.......
....... ...
.........
....... ..........
.. •
v1

Figure 1. The quotient space Q(Ω).

with differential d1 : Hs+t (Xt , Xt−1 ; Z) −−→ Hs+t−1 (Xt−1 , Xt−2 ; Z), where we use
the convention that X−1 = ∅.
To compute these E 1 terms notice that all the pairs (Xt , Xt−1 ) are relative
manifolds so that one can apply the Alexander-Poincaré duality theorem. Hence,
by 13.7.14

(13.7.15) Hs (X0 ; Z) ∼= Hs (tk+2 T k ; Z) ,


(13.7.16) Hs (X1 , X0 ; Z) ∼
= H k+2−s (tk+2 T k+1 ; Z) ,
(13.7.17) Hs (X2 , X1 ; Z) ∼
= H k+4−s (T k+2 ; Z) ,
where tj T l means the disjoint union of j copies of T l . The Es,t
1
term of the spec-
tral sequence is described by the diagram of Figure 2. The computation of the
differentials is fairly tedious and we refer the reader to [BGM98a] for details.

s
...
...
...
...
...
...
...
• ...
...
• •
...
...
...
... k k+1 k+2
• Z( ) 2 (k+2)
• Z( ) • Z( )
...
... 2 (k+2) 2
...
...
...
...
...
... k(k+2) (k+1)(k+2) k+2
•Z ...
...
...
•Z •Z
...
...
...
k+2 k+2 Z
Z ...
... Z
...........................................................................................................................................................................................................................................................
• • • t

1
Figure 2. The diagram for the Es,t term.

¤
13.7. TORIC 3-SASAKIAN MANIFOLDS 519

Theorem 13.7.26 together with the well-known Hurewicz Theorem determines


the integral cohomology groups of S(Ω) except for the torsion in H 4 (S(Ω), Z).
Corollary 13.7.29: Ω ∈ Ak,k+2 (Z). Then H i (S(Ω), Z) are given as follows

H0 H1 H2 H3 H4 H5 H6 H7
Z 0 Zk 0 G(Ω) Zk 0 Z

where G(Ω) is a finite torsion group.


Recently Hepworth [Hep05a, Hep05b] determined the group G(Ω). Rather
than using the Leray spectral sequence as in [BGMR98], he considered the Serre
spectral sequence of the associated Borel construction of the toral action. Let us
set some notation. Changing slightly the notation of Equation 13.7.1 we let Ω ∈
Ak,k+2 (Z), i, j ∈ {1, . . . , k + 2} and ∆ij be the determinant of the k × k minor ob-
tained from Ω by deleting ith and j th columns. Set ²ij = (−1)i+j sign(i−j)sign(∆ij ).
By construction N (Ω)−−−→S(Ω) is a T k -bundle, so we let xp ∈ H 2 (S(Ω), Z) ≈ Zk
be the first Chern class associated to the ith factor of the T k -action. Then the xp
freely generate H 2 (S(Ω), Z). By abuse of notation we also let xp denote the gener-
ators of the cohomology ring ofP the classifying space BT k so that H ∗ (BT k , Z) ≈
Z[x1 , . . . , xk ]. We also put vi = p xp Ωpi . Then Hepworth proves
Theorem 13.7.30: G(Ω) is generated by the degree four symmetric products {xp xq }
subject only to the relations

²ij vi vj + ²jk vj vk + ²ki vk vi = 0

for i, j, k distinct. Furthermore, G(Ω) is a finite group of order


X
|∆i1 j1 | · · · |∆ik+1 jk+1 | ,

where the summand indexed by i1 , j1 , . . . , ik+1 , jk+1 is included in the sum if and
only if the graph on the vertices {1, . . . , k + 2} with edges {ir , jr } is a tree.
Since for an admissible Ω the determinants |∆ij | are nonzero, the sum above is
at least as large as the number of trees on the vertex set {1, . . . , k + 2}. A theorem
of Cayley states that the number of trees on a vertex set A is precisely |A||A|−2 .
We therefore have:
Corollary 13.7.31: |H 4 (S(Ω), Z)| > (b2 (S(Ω)) + 2)b2 (S(Ω)) .
Let Ω ∈ Ak,k+2 (Z) be as in (13.7.12). Theorem 13.7.30 implies that H 4 (S(Ω), Z)
will have size at least as large as each of the |ai |, |bj |. Therefore for each k, one
can construct an infinite sequence of Ω for which the groups H 4 (S(Ω), Z) are of
unbounded size. Hence, just as in the k = 1 case we get
Corollary 13.7.32: There are infinitely many homotopy types of S(Ω) for each
fixed second Betti number.
In addition, Hepworth computes the stable tangent bundle and, hence, the
Portraying classes of the S(Ω).
Theorem 13.7.33:
k+2
M
T S(Ω) ⊕ ²4k+1 ∼
=2 lp ,
R
p=1
520 13. 3-SASAKIAN MANIFOLDS

where, for 1 6 p 6 (k + 2), lp is the complex line bundle over S(Ω) with first Chern
class vp . Consequently
k+2
X
p1 (S(Ω)) = 2 vp2 .
p=1
In addition, in his thesis [Hep05a] Hepworth shows that the manifolds S(Ω) are
cobordant to much simpler spaces whose second Betti number equals 1. This suggest
that, just as in the case of S(p), one could obtain a diffeomorphism classification
of S(Ω) by computing their generalized Kreck-Stolz invariants.
Proposition 13.7.34: Let Ω ∈ Ak,k+2 (Z) so that S(Ω) is a smooth manifold. Let
Z(Ω) and O(Ω) be the associated twistor space and quaternionic Kähler orbifold,
respectively. Then we have b2 (S(Ω)) = b2 (O(Ω)) = b2 (Z(Ω)) − 1 = k.
This shows that inequality b2 ≤ 1 in Proposition 13.5.7 does not hold for non-
regular 3-Sasakian manifolds. Finally, we give several interesting corollaries.
Corollary 13.7.35: There exists a simply connected 3-Sasakian 7-manifold for
every rational homology type allowed by Corollary 13.5.3.
Our next corollary follows from the results of this section and a remarkable
theorem of Gromov [Gro81]:
Corollary 13.7.36: For any non-positive real number κ there are infinitely many
3-Sasakian 7-manifolds which do not admit metrics whose sectional curvatures are
all greater than or equal to κ.
Corollary 13.7.37: There exist Q-factorial contact Fano 3-folds X with b2 (X) = l
for any positive integer l.
The last corollary should be contrasted to the smooth case, where Mori and
Mukai have proven that b2 ≤ 10 [MM82]:
Corollary 13.7.38: If the second Betti number b2 (S(Ω)) = k > 3, the 3-Sasakian
manifolds S(Ω) are not homotopy equivalent to any homogeneous space.

13.7.7. Higher Dimensional Toric 3-Sasakian Manifolds. As in dimen-


sion 7, we have the following classification theorem of Bielawski [Bie99]:
Theorem 13.7.39: Let (M, S) be a toric 3-Sasakian manifold of dimension 4n−1.
Then (M, S) is isomorphic as a 3-Sasakian T n -manifold to a 3-Sasakian quotient
of a sphere by a torus, that is to a S(Ω) for some Ω.
This result is a weaker than the similar result of Theorem 13.7.24 as following
Bielawski we assume that (M, S) is a smooth manifold. On the other hand it is
likely that this assumption can be removed. Note that this theorem includes the
degenerate case when the quotient is a sphere or a discrete quotient of such. The
Betti numbers of a 3-Sasakian orbifold obtained by a toral quotient of a sphere were
computed by Bielawski [Bie97] using different techniques than the ones employed
in the previous section. The following result follows easily from Theorem 12.9.6:
Theorem 13.7.40: Let Ω ∈ Mk,n+1 (Z) be non-degenerate so that S(Ω) is a com-
pact 3-Sasakian orbifold of dimension 4(n − k) + 3. Then we have
 
k+i−1
(13.7.18) b2i (S(Ω)) =  , i < n + 1 − k.
k
13.7. TORIC 3-SASAKIAN MANIFOLDS 521

Furthermore, the Betti number constraints of Proposition 13.5.7(ii) can hold


for S(Ω) if and only if k = 1.
Combining Theorems 13.7.39 and 13.7.40 with Lemmas 13.7.7 and 13.7.8 which
give obstructions to smoothness gives the somewhat surprising result [BGM98b],
Theorem 13.7.41: Let (M, S) be a toric 3-Sasakian manifold.
(i) If the dimension of M is 19 or greater, then b2 (M ) ≤ 1.
(ii) If the dimension of M is 11 or 15, then b2 (M ) ≤ 4.
(iii) If b2 (M ) > 4, then the dimension of M is 7.
A corollary due to Bielawski [Bie99] is:
Corollary 13.7.42: Let (M, S) be a regular toric 3-Sasakian manifold. Then M
is one of the 3-Sasakian homogeneous spaces S 4n−1 , RP4n−1 or ¡ SU (n) ¢.
S U (n−2)×U (1)

Next we give an explicit construction of toric 3-Sasakian manifolds not elimi-


nated by Theorem 13.7.41. It is enough to show that A4,8 and A4,7 are not empty
as the rest follow by deletion of rows of the corresponding Ω ∈ A4,∗ . We shall
present two three parameter families of solutions, namely
   
1 1 1 1 1 1 2
   
2 1 1 + 2l  2 1 1 + 2l
0
−1
   
A1 =  , A2 =  ,
 1 16 1 + 2m   1 16 1 + 22n 3 
   
−1 3 2c −1 3 2c0 −1
where l, l0 , m, n ∈ Z+ , and c, c0 ∈ Z.
With the aid of MAPLE symbolic manipulation program [BGM98b], we find
Lemma 13.7.43: Let ∆ = 2(31c + 6 + 19 · 2l−1 − 7 · 2m−1 ).
¡ ¢
(i) Ω1 = 1l4 A1 is admissible if and only if c 6= 0 and is not divisible by 3,
and ∆ 6=¡ 0 and¢ is not divisible by 7, 19 nor 31.
(ii) Ω2 = 1l4 A2 is admissible if and only if c and all minor determi-
nants of A2 are non-vanishing, and c0 6≡ 0(3), l0 6≡ 0(4), c0 6≡ 5(7),
0
and 11, 19, 37, 71 do not divide det A = 19 · 22n − 63 − 148c0 − 11 · 2l , and
the following conditions hold:
0 0
gcd(3, 4c0 + 2l + 1, 2c0 − 2l − 1) = 1
2n+1 l0 l0 2n
gcd(7, 2 − 2 + 1, 3 · 2 + 2 + 4) = 1
(13.7.19)
2n l0 +4 l0 2n
gcd(19, 2 −2 − 15, 3 · 2 + 2 + 4) = 1
0 2n 0 2n
gcd(25, 32c − 3 · 2 − 3, 6c + 2 + 1) = 1.
The conditions in this proposition guarantee that the quotient spaces denoted
by S(c, l, m) and S(c0 , l0 , n) are smooth manifolds of dimension 11 and 15, respec-
tively.
It is routine to verify that the three parameter infinite family given by
c ≡ 14(21), l 6≡ 1(5), m 6≡ α(c) mod(18) ,
where 2α(c) ≡ 22(31c + 6)(18) satisfies the conditions in (i) of Lemma 13.7.43.
This gives examples in dimension 11. (Notice that as 2 is a primitive root of 19
the equation defining α(c) has a unique solution (mod 18) for each value of c.)
522 13. 3-SASAKIAN MANIFOLDS

Similarly, it is straightforward to verify that the infinite family given by


c0 = 2, l0 = 1, n ≡ 21(90) ,
satisfies the conditions (ii) of Lemma 13.7.43. We have arrived at [BGM98b]:
Theorem 13.7.44: There exist toric 3-Sasakian manifolds S of dimensions 11
and 15 with b2 (S) = 2, 3, 4. Consequently, the Betti number relations of (ii) of
Proposition 13.5.7 do not hold generally. More explicitly there are compact 11-
dimensional 3-Sasakian manifolds for which b2 6= b4 , and compact 15-dimensional
3-Sasakian manifolds for which b2 6= b6 .

13.8. Cohomogeneity One 3-Sasakian 7-Manifolds


In this section we discuss all known complete cohomogeneity one 3-Sasakian
metrics in dimension 7. Locally, the classification of such metrics follows from
Hitchin’s classification of self-dual Einstein metrics with an isometric action of
SU (2) [Hit96, Hit95a, Hit95b]. In particular, as we shall see, there are at least
two families of such manifolds: the U (2)-invariant family S(q, 1, 1) of the toric 3-
Sasakian spaces and another non-toric SO(3)-invariant family of metrics obtained
from the Hitchin’s metrics on the orbifolds Ok described in Section 12.5. Globally,
however, the question is more subtle. Nevertheless, it is quite reasonable to
Conjecture 13.8.1: Let (M, S) be a compact simply connected 3-Sasakian orbifold
of dimension 7 and of cohomogeneity one under the action of Aut(M, S). Then
(M, S) is one of the two infinite families of spaces
(i) the toric 3-Sasakian manifold S(q, 1, 1) as the Konishi orbibundle over
the Galicki-Lawson orbifold O(q, 1, 1) of Proposition 12.5.3
(ii) the Konishi bundle Sk over non-toric Hitchin orbifolds Ok of Theorem
12.5.8.
The toric 3-Sasakian 7-manifolds S(q, 1, 1) were already discussed in the pre-
vious section. The Hitchin metrics likewise have smooth cohomogeneity one 3-
Sasakian manifolds associated to them. They were recently considered in [GWZ05]
as possible candidates of 7-manifolds admitting cohomogeneity one metrics of pos-
itive sectional curvature. We describe the geometry and topology of these new
examples. We closely follow [GWZ05] in our exposition. Let us recall that Equa-
tion 1.6.9 and its associated Group Diagram 1.6.10 determine the manifold M .
We first consider several infinite families of 7-manifolds with cohomogeneity one
actions. Let G = Sp(1) × Sp(1), where we think of Sp(1) as the group of unit
quaternions. Consider the following circle subgroups of G:
i1 i2
(13.8.1) C(p,q) = {(epi1 θ , eqi1 θ ) | θ ∈ R}, C(p,q) = {(epi2 θ , eqi2 θ ) | θ ∈ R},
where (p, q) are integers, called slopes. Let Q denotes the quaternion group gener-
ated by {i1 , i2 , i3 } and H = Z2 × Z4 = {(±1, ±1), (±i1 , ±i1 )} ⊂ G.
7
Definition 13.8.2: Let M = M(p − ,q− ),(p+ ,q+ )
be defined by the group diagram
i1 i2
Q ⊂ {K− , K+ } ⊂ G, where {K− , K+ } = {C(p − ,q− )
· Q, C(p + ,q+ )
· Q}, where
(p− , q− ) as well as (p+ , q+ ) are relatively prime odd integers. Furthermore, let
Pl = M(1,1),(2l−1,2l+1) .
0
Here K± /K± = Z2 , where the second component is generated by (i2 , i2 ) on the
left and (i1 , i1 ) on the right, up to signs (of both coordinates). The embedding of
13.8. COHOMOGENEITY ONE 3-SASAKIAN 7-MANIFOLDS 523

Q ⊂ G is determined by the slopes and is always diagonal ∆Q, up to sign changes


in both coordinates.
The action of Sp(1) × Sp(1) is only almost effective, i.e., G and Q have a finite
normal, hence, central subgroup in common. For example, the manifolds Pl have
the effective group diagram
Z2 × Z2 ⊂ {O(2), O(2)} ⊂ SO(4).
It is easy to see that P1 = S 7 with the well-known cohomogeneity one action of
SO(4). Grove, Wilking, and Ziller [GWZ05] prove
Theorem 13.8.3: The manifolds M = M(p− ,q− ),(p+ ,q+ ) are 2-connected. If pq− −
6=
p+ 2 2 2 2
± q+ their cohomology ring is determined by π3 (M ) = Zk with k = (p− q+ −p+ q− )/8.
Otherwise H 3 (M, Z) = H 4 (M, Z) = Z. In particular, the manifolds Pl are all
rational homology spheres with π1 (Pl ) = π2 (Pl ) = 0 and π3 (Pl ) = Zl .
In order to describe all Konishi bundles over Hitchin orbifolds we need yet
another infinite family of cohomogeneity one 7-manifolds.
6
Definition 13.8.4: Let N = N(p − ,q− ),(p+ ,q+ )
be the cohomogeneity one 7-manifold
i1 j
defined by the group diagram H ⊂ {K− , K+ } = {C(p − ,q− )
· H, C(p+ ,q+ )
· H} ⊂ G
with (p− , q− ) as well as (p+ , q+ ) relatively prime, p+ even and p− , q− , q+ odd.
Furthermore, let Ql = M(1,1),(l,l+1) .
0
Here the component groups K± /K± are determined by the fact that (i1 , i1 ) ∈
0 +
K− and (1, −1) ∈ K0 . As in the previous example the Sp(1) × Sp(1)-action is only
almost effective. For example, the manifolds Qk have the effective group diagram
Z2 ⊂ {SO(2), O(2)} ⊂ SO(3) × SO(3) ,
where the groups K± are embedded in different blocks in each component of SO(3)×
SO(3). It is easy to see that P1 = S(1, 1, 1) with the well-known cohomogeneity
one action of SO(3) × SO(3). The integral cohomology groups of these manifolds
was given in [GWZ05].
Theorem 13.8.5: The manifolds N = N(p− ,q− ),(p+ ,q+ ) are simply connected with
2
H 2 (N, Z) = Z , H 3 (N, Z) = 0 and H 4 (N, Z) = Zk with k = p2− q+ 2
− p2+ q− .
In particular, for Ql we have H 2 (Ql , Z) = Z and H 4 (Ql , Z) = Z2l+1 . We will
be only interested in the manifolds Pl and Ql . Note that P1 and Q1 are precisely
the two simply connected homogeneous (hence, regular) compact 3-Sasakian 7-
manifolds. They both admit cohomogeneity one subactions and, from this point
of view, naturally appear as first in two series of cohomogeneity one spaces, under
actions of SO(4) and SO(3) × SO(3), respectively. It turns out that although these
two series are very different, they have quotients which can be described in a uniform
fashion. These quotients are precisely the Konishi bundles over Hitchin orbifolds
Ok described in Theorem 12.5.8. We can already see the pattern by considering the
first two examples. P1 ' S 7 is an Sp(1) bundle over S 4 . The Konishi bundle, on
the other hand, is the SO(3)-bundle over S 4 which is RP7 ' P1 /Z2 . The manifold
Q1 is and SO(3)-bundles over CP2 and it is the Konishi bundle. But Q1 admits a
smooth Z2 quotient which is a Konishi orbibundle over CP2 /Z2 = O4 . Recall from
section 12.5 that indeed, the Hitchin orbifold O4 is not simply connected and its
universal cover is CP2 with the standard Fubini-Study metric. This picture repeats
itself in higher dimensions: the Hitchin orbifolds O2l+1 are simply connected and
524 13. 3-SASAKIAN MANIFOLDS

the Konishi bundle S2l+1 −→O2l+1 admits a lift to an Sp(1)-bundle which turns out
to be Pl . The Hitchin orbifolds O2l+2 are not simply-connected, and neither are
the Konishi bundles S2l+2 −→O2l+2 . The universal cover of S2l+2 is again an SO(3)
orbibundle over the orbifold universal cover of O2l+2 .
Before we consider the Konishi orbibundles let us define the cohomogeneity one
7-manifolds Hk defined by the group diagram
0 0
(13.8.2) Z2 × Z2 ⊂ {K− · H, K+ · H} ⊂ SO(3) × SO(3) ,
0
where the identity components K± ' SO(2) depend on integers (p, q) which de-
scribe the slope of their embedding into a maximal torus of SO(3) × SO(3). They
0 0
are (1, 1) for K− embedded into the lower 2×2 block and (k, k+2) for K+ embedded
into the upper 2 × 2 block. That also determines the embedding of H = Z2 × Z2 .
In principle, it is clear that the Hitchin metrics give rise to 3-Sasakian orbifold
metrics on a seven dimensional orbifold Sk which is the total space of the SO(3)-
orbibundle Sk −→Ok . The cohomogeneity one action by SO(3) on the base admits
a lift to the total space Sk which commutes with the almost free principal orbifold
SO(3) action. The joint action by SO(3)SO(3) on Sk is, hence, an isometric co-
homogeneity one action. In general, one would expect the metric on Sk to have
orbifold singularities since the base does. However, a simple argument shows that
this is not the case [GWZ05]
Theorem 13.8.6: For each k, the total space Sk of the Konishi bundle correspond-
ing to the self-dual Hitchin orbifold Ok is a smooth 3-Sasakian manifold.

Proof. Note that the singular orbit B+ in Ok , k > 4 must be totally geodesic.
Indeed, being an orbifold singularity, one can locally lift the metric on a normal
slice D2 to RP2 to its (k − 2)-fold branched cover D̂ → D with an isometric action
by Zk−2 such that D̂/Zk−2 = D. Hence the singular orbit is a fixed point set of a
locally defined group action and thus totally geodesic. The SO(3) principle bundle
Sk is smooth over all smooth orbits in Sk . If it had orbifold singularities, they would
consist of an SO(3)SO(3) orbit which projects to B+ , and be again totally geodesic
by the same argument. This five dimensional orbit would now be 3-Sasakian with
respect to the natural SO(3) action on Sk , since it is totally geodesic and contains
all SO(3) orbits. But the quotient is 2-dimensional which contradicts the fact that
the base of such a manifold has dimension divisible by 4. ¤

The following theorem was proven by Grove, Wilking and Ziller and it estab-
lishes the relation between the Konishi orbibundles Sk over the Hitchin orbifolds
Ok and the cohomogeneity one manifolds Pl and Ql . We just quote the result and
refer the reader to the proof in [GWZ05].
Theorem 13.8.7: The manifolds Pl and Ql are equivariantly diffeomorphic to the
universal covers of the 3-Sasakian manifolds S2l+1 and S2l+2 , respectively.
In Theorem 11.7.11 we saw that there are infinitely many rational homology
spheres in every odd dimension greater than three which admit Sasaki-Einstein struc-
tures. As an immediate consequence of the Theorem 13.8.7 we get
Corollary 13.8.8: There are infinitely many rational homology 7-spheres which
admit a 3-Sasakian structure.
Open Problem 13.8.1: Note that H 2 (S(p, 1, 1), Z) = H 2 (Qp , Z) = Z, and the
torsion H 4 (S(p, 1, 1), Z) = H 4 (Qp , Z) = Z2p+1 . Metrically the two manifolds are
13.9. NON-TORIC 3-SASAKIAN MANIFOLDS IN DIMENSION 11 AND 15 525

different, one being toric and the other not. It would be interesting to calculate the
Kreck-Stolz invariants for the Qp family to see if these spaces can be diffeomorphic
to some of the S(p, 1, 1). However, this appears to be a rather difficult problem.
Question 13.8.1: The 7-manifolds Pk , Qk with their induced 3-Sasakian structure
and the Hitchin orbifolds Ok appear to be the only known inhomogeneous examples
which are not obtained via the symmetry reduction method. Is it indeed the case
that these metrics cannot be obtained as quotients?

13.9. Non-Toric 3-Sasakian Manifolds in Dimension 11 and 15


We now consider some examples of non-Abelian reductions. Example 13.6.8
suggests that one should consider reductions by the group T k × Sp(1) which is
equivalent to Abelian reductions of the diamond diagram associated to the classical
group SO(n). At the level of the quaternionic Kähler geometry such quotients are
just Abelian reductions of the real Grassmannian Gr+ n
4 (R ) of oriented 4-planes
n+1
in R . We should point out that Abelian reductions of ♦(SU (n)) can produce
only subfamilies of the toric examples discussed earlier. This is because ♦(SU (n))
is itself an S 1 reduction of ♦(Sp(n)). Hence, other than the exceptional cases,
♦(SO(n)) is clearly of special interest from this point of view. There are two cases
to consider: SO(2m), m > 3 and SO(2m + 1), m > 2. In both cases we need to
choose a proper k-dimensional subtorus T k of the maximal torus T m . We can treat
both cases simultaneously by introducing a parameter δ = {0, 1}, and choosing
T k ⊂ T m ⊂ SO(2m + δ).
Definition 13.9.1: Let Θδk,m ⊂ Mk,m (Z) × {0, 1} be a weight matrix with k =
1, . . . , m − 1 and m > 3 − δ. We denote by S(Θδk,m ) the 3-Sasakian reduction
of SO(2m + δ)/SO(2m + δ − 3) × Sp(1) by an action of T k ⊂ SO(2m + δ) de-
termined by Θδk,m . Equivalently, S(Θδk,m ) is the 3-Sasakian reduction of the unit
sphere S 4(2m+δ)−1 by the action of T k × Sp(1). Furthermore, we say that
(i) Θδk,m is non-singular if the corresponding T k -action on the level set of
the 3-Sasakian moment map is locally free,
(ii) Θδk,m is admissible if, in addition to Θk,m
δ
being non-singular, the cor-
k
responding T -action on the level set of the 3-Sasakian moment map is
free.
Generically, the quotient S(Θδk,m ) will be of dimension 4(2m − k − 4 + δ) + 3.
In the case when the weight matrix is non-singular the reduced space will automat-
ically be a compact 3-Sasakian orbifold, while admissible matrices yield smooth
3-Sasakian manifolds. It is interesting to observe, that unlike the case of toric
quotients, in each fixed dimension we only have finitely many choices for the
triple (δ; k, m). In that triple the pair (δ; m) determines the diamond diagram
♦(SO(2m + δ)) for the reduction while k is the dimension of the torus that one is
quotienting by.
Proposition 13.9.2: Let dim(S(Θδk,m )) = n = 4p + 3 with k = 1, . . . , m − 1 and
m > 3 − δ. Then for each n = 4p + 3 > 7 the number of possible triples (δ; k, m)
is p + 2, whereas, there are two for n = 7. We list the triples for the first three
dimensions.
(i) if n = 7 then (δ; k, m) = {(0; 3, 4), (1; 2, 3)},
(ii) if n = 11 then (δ; k, m) = {(0; 2, 4), (0; 4, 5), (1; 1, 3), (1; 3, 4)},
526 13. 3-SASAKIAN MANIFOLDS

(iii) if n = 15 then (δ; k, m) = {(0; 1, 4), (0; 3, 5), (0; 5, 6), (1; 2, 4), (1; 4, 5)}.
Proof. Fixed n means fixed p, and the dimension formula for the quotient
gives p = 2m − k − 4 + δ. Writing l = m − k we have m + l = p + 4 − δ. The
number of possible values of l equals the number of possible values of k. Assuming
that p > 1 we see that there are [ p+3 p+2
2 ] possible triples if δ = 0, and [ 2 ] if δ = 1.
This gives p + 2 total. When p = 1 the constraint m > 3 − δ eliminates (0, 1, 3), so
there are only two. It is easy to delineate the first several cases. ¤
In principle, it is not clear that there should exist any admissible matrices,
though it is rather obvious that there should be many non-singular ones. The
following result was established in [BGP02, Bis07].
Proposition 13.9.3: There are no admissible matrices among Θ03,4 and Θ12,3 .
On the other hand, as we saw in Theorem 12.5.12, there are non-singular ma-
trices Θ03,4 and Θ12,3 which yield two families of examples of positive self-dual and
Einstein orbifold metrics with a one-dimensional symmetry group. The combina-
torics of admissibility in these cases appears to be even more restrictive than in the
toric case. We do not present the solution to this problem here, but only show that
there are some cases when Θδk,m is admissible. In the remainder of this section we
shall consider the two special cases of Proposition 13.9.2 with k = 1: Θ11,3 and Θ01,4
that were treated in [BGP02].

13.9.1. Θ11,3 : 11-Dimensional Examples. We first consider a generaliza-


tion of the Kobak-Swann Example 13.6.8. In particular, we consider Θ11,3 =
(p1 , p2 , p3 ) ≡ p ∈ Z3 . We determine the conditions under which Θ11,3 is admis-
sible. Define the homomorphism fp : T 1 −−→SO(7)
 
1 0 0 0  
 
0 A(p t) 0 0  cos(pi t) sin(pi t)
 1 
fp (t) =   , A(pi t) =  .
0 0 A(p t) 0  − sin(p t) cos(p t)
 2  i i

0 0 0 A(p3 t)
Note that for p1 = p2 = p3 = 1 we recover the Example 13.6.8 so this case is a
very natural generalization. Furthermore, we use the same notation as in Example
13.6.8 regarding the Sp(1) reduction of S 27 . The homomorphism fp (t) yields a
circle action on the homogeneous 3-Sasakian manifold SO(7)/SO(3) × Sp(1) via
left multiplication fp (t)u with the 3-Sasakian moment map
X
(13.9.1) νp (u1 , . . . , u7 ) = pα (u2α u2α+1 − u2α+1 u2α ) .
α=1,2,3

Observe that without loss of generality we can assume all weights to be non-negative
as pi can be changed to −pi by renaming the quaternions in the associated pair
(u2i , u2i+1 ). We begin the analysis of this quotient by considering the zero level set
of the moment map Nν (p) ≡ V7,4 (R) ∩ {νp−1 (0)} ⊂ S 27 .
We want to consider a stratification of the level set Nν (p) that will allow us to
analyze the quotient space
Nν (p) ≡ V7,4 (R) ∩ {νp−1 (0)}
(13.9.2) S(Θ11,3 ) ≡ M11 (p) = .
Sp(1) × U (1)p
13.9. NON-TORIC 3-SASAKIAN MANIFOLDS IN DIMENSION 11 AND 15 527

Since Nν (p) is a submanifold of the Stiefel manifold V4 (R7 ) at most 3 quaternionic


coordinates can vanish on Nν (p). So setting various quaternionic coordinates equal
to zero determines a stratification of Nν (p) in which the strata of minimal dimension
play an important role. We call these strata vertices although, as we shall see, they
each have two connected components. We now give a series of lemmas to determine
conditions on the integers (p1 , p2 , p3 ) that guarantee that Θ11,3 be admissible. The
proofs are straightforward and can be found in [BGP02].
Lemma 13.9.4: Let 0 < p1 < p2 < p3 . At a vertex neither u1 nor any of the three
pairs of quaternions (u2i , u2i+1 ), i = 1, 2, 3 can vanish. Thus, there are precisely
eight vertices and they are all diffeomorphic to O(4).
Our analysis suggests the importance of the following strata:
S0 = {u ∈ Nν (p) | u1 = 0} ,
S1 = {u ∈ Nν (p) | u1 6= 0} ,
S2 = {u ∈ Nν (p) | some quaternionic pair (u2i , u2i+1 ) vanishes} ,
S3 = {u ∈ Nν (p) | no quaternionic pair (u2i , u2i+1 ) vanishes} .

Then Lemma 13.9.4 easily implies that


Corollary 13.9.5: Let 0 < p1 < p2 < p3 . Then
(i) S0 ∪ S1 = S2 ∪ S3 = Nν (p).
(ii) S0 ∩ S1 = S2 ∩ S3 = ∅.
(iii) S0 ∩ S2 = ∅.
(iv) S2 ⊂ S1 .
(v) S0 ⊂ S3 .
Notice that (iii) fails if pi = pj for some i 6= j. In particular it fails for the
level set Nν of Definition 13.6.12 in the previous section, and this is the reason that
the quotient M11 of 13.6.10 is not smooth. We now are ready to give necessary
conditions to guarantee a smooth quotient.
Lemma 13.9.6: Let p = (p1 , p2 , p3 ) ∈ (Z+ )3 . Then
(i) if the (p1 , p2 , p3 ) are pairwise relatively prime, then the isotropy group of
the Sp(1) × U (1)p action at every point of S1 is the identity;
(ii) if gcd(p1 ± p2 , p1 ± p3 ) = 1, then the isotropy group of the Sp(1) × U (1)p
action at every point of S0 is the identity.
It now follows immediately from Corollary 13.9.5, Lemma 13.9.6, and Definition
13.9.1 of admissibility that
Theorem 13.9.7: The weight matrix Θ11,3 ≡ p ∈ Z3+ is admissible if and only if
0 < p1 < p2 < p3 , gcd(pi , pj ) = 1 for all i < j, and gcd(p1 ± p2 , p1 ± p3 ) = 1.
Note that there are infinitely many admissible weight matrices Θ11,3 ≡ p ∈ Z3+ .
For example we can take p = (2k − 1, 2k, 2k + 1), where k ∈ Z+ . Thus, there are
infinite families of smooth quotients S(Θ11,3 ) ≡ M11 (p) and infinite families of the
associated triples M11 (p) → Z(p) → O(p) with their (orbifold) Einstein metrics.
The geometry of the smooth families M11 (p) is rather interesting. First we
observe that these spaces cannot be toric. This can be seen in several different
ways, for example by careful analysis of the associated foliations. One can also
generalize the analysis of [BGM96a] to show that the only isometries of the level
528 13. 3-SASAKIAN MANIFOLDS

set of the moment map Nν (p) ⊂ S 27 ⊂ R28 can come from the restriction of the
isometries of the Euclidean space R28 . From this we conclude
Theorem 13.9.8: Let Θ11,3 ≡ p be admissible so that M11 (p) is a smooth compact
3-Sasakian 11-manifold. Then the Lie algebra Aut(M11 (p), Sp ) of the group of 3-
Sasakian isometries of M(p) is isomorphic to R2 ⊕ sp(1). In particular, all such
quotients are non-toric.
Up to this point the topology of the manifolds M11 (p) remains largely unde-
termined.
Open Problem 13.9.1: Determine the cohomology ring of the 3-Sasakian mani-
folds M11 (p).
13.9.2. Θ01,4 : 15-Dimensional Examples. Consider now the reduction de-
fined by Θ01,4 ≡ p = (p1 , p2 , p3 , p4 ) and define the following homomorphism
 
A(p1 t) 0 0 0  
 
 0 A(p t) 0 0  cos(p t) sin(p t)
 2  i i
fp (t) =   , A(pi t) =  .
 0 0 A(p t) 0  − sin(p t) cos(p t)
 3  i i

0 0 0 A(p4 t)
As before we can choose all weights to be non-negative. Further note that at most
one of the weights can vanish. Using techniques similar to the ones described in
the previous section one can show that
Theorem 13.9.9: Let Θ01,4 ≡ (p1 , p2 , p3 , p4 ) ∈ Z4 . Then Θ01,4 is admissible if and
only if 0 ≤ p1 < p2 < p3 < p4 , gcd(pi , pj , pk ) = 1 and gcd(pi ± pj , pi ± pk ) = 1 for
any triple in p.
Proof. First, let us assume that all of the weights are non-negative. Consider
a triple, say (p1 , p2 , p3 ). Set u7 = u8 = 0. Then the analysis of the previous section
shows that the three must be distinct and that we must have gcd(p1 ± p2 , p1 ± p2 ) =
1. However, we no longer need the three weights to be pairwise relatively prime
as one cannot set two of the quaternionic pairs (u2i−1 , u2i ), i = 1, 2, 3, 4 equal to
(0, 0) at the same time. One such quaternionic pair can vanish; hence, we need
gcd(p1 , p2 , p3 ) = 1 to get a free action. The analysis in the case when p1 = 0 is
similar. Then one sees that the triple (p2 , p3 , p4 ) has to be admissible in the sense
of the Theorem 13.9.7. ¤
As before we have we denote the quotient manifold corresponding to an admis-
sible Θ01,4 = (p1 , p2 , p3 , p4 ) by M15 (p).
Open Problem 13.9.2: Determine the cohomology ring of the 3-Sasakian mani-
folds M15 (p).
Question 13.9.1: Can Θδk,m be admissible for other values of (δ; k, m) with m >
3 − δ?
CHAPTER 14

Sasakian Structures, Killing Spinors, and


Supersymmetry

The main purpose of this chapter is to provide a connection (non-rigorous) of


Sasaki-Einstein and related geometries to the theoretical physics of supersymmetric
quantum field theory, especially to string theory and M-theory. There have been
many recent attempts to describe these important notions of theoretical physics in
precise mathematical terms. This enormous task is far beyond the scope of this
book, so we refer the reader to recent monographs and references therein [DEF+ 99,
Var04, Jos01, AJPS97]. Here we content ourselves with providing the main
theorems and results (often without proofs) concerning Killing spinors, and then
give a brief heuristic discussion in Section 14.5 describing the relation with physics.
In particular, we discuss some additional properties of Sasaki-Einstein manifolds
that relate Killing spinors to supersymmetry.

14.1. The Dirac Operator and Killing Spinors


We begin with a definition of spinor bundles and the bundle of Clifford algebras
of a vector bundle. Here we only give a quick definition and refer to [LM89, Fri00]
for details. Recall that the Clifford algebra Cl(Rn ) over Rn can be defined as the
quotient algebra of the tensor algebra T (Rn ) by the two-sided ideal I generated by
elements of the form v ⊗ v + q(v) where q is a quadratic form on Rn .
Definition 14.1.1: Let E be a vector bundle with inner product h·, ·i on a smooth
manifold M , and let T (M ) denote the tensor bundle over E. The Clifford bundle
of E is the quotient bundle Cl(E) = T (M )/I(E) where I is the bundle of ideals
(two-sided) generated pointwise by elements of the form v ⊗ v + hv, vi with v ∈ Ex .
A real spinor bundle S(E) of E is a bundle of modules over the Clifford bundle
Cl(E). Similarly, a complex spinor bundle is a bundle of complex modules over
the complexification Cl(E) ⊗ C.
As vector bundles Cl(E) is isomorphic to the exterior bundle Λ(E), but their
algebraic structures are different. The importance of Cl(E) is that it contains the
spin group Spin(n), the universal (double) covering group of the orthogonal group
SO(n), so one obtains all the representations of Spin(n) by studying representations
of Cl(E). We assume that the vector bundle E admits a spin structure as discussed
in Chapter 1, so w2 (E) = 0. We are interested mainly in the case when (M, g) is a
Riemannian spin manifold and E = T M in which case we write S(M ) instead of
S(T M ). The Levi-Civita connection ∇ on T M induces a connection, also denoted
∇, on any of the spinor bundles S(M ), or more appropriately on the sections
Γ(S(M )).

529
530 14. SASAKIAN STRUCTURES, KILLING SPINORS, AND SUPERSYMMETRY

Definition 14.1.2: Let (M n , g) be a Riemannian spin manifold and let S(M ) be


any spinor bundle. The Dirac operator is the first order differential operator
D : Γ(S(M ))−−−→Γ(S(M )) defined by
Xn
Dψ = Ej · ∇Ej ψ ,
j=1

where {Ej } is a local orthonormal frame and · denotes Clifford multiplication.


The Dirac operator, of course originating with the famous Dirac equation de-
scribing fermions in theoretical physics, was brought into mathematics by Atiyah
and Singer in [AS63]. Then Lichnerowicz [Lic63a] proved his famous result that
a Riemannian spin manifold with positive scalar curvature must have vanishing Â-
genus. An interesting question on any spin manifold is: what are the eigenvectors
of the Dirac operator. In this regard the main objects of interest consists of special
sections of certain spinor bundles called Killing spinor fields or just Killing spinors
for short. Specifically, (cf. [BFGK91, Fri00])
Definition 14.1.3: Let (M, g) be a complete n-dimensional Riemannian spin man-
ifold, and let S(M ) be a spin bundle (real or complex) on M and ψ a smooth section
of S(M ). We say that ψ is a Killing spinor if for every vector field X there is
α ∈ C, called Killing number, such that
∇X ψ = αX ·ψ .
Here X ·ψ denotes the Clifford product of X and ψ. We say that ψ is imaginary
when α ∈ Im(C∗ ), ψ is parallel if α = 0 and ψ is real1 if α ∈ Re(C∗ ).
We shall see shortly that the three possibilities for the Killing number α : real,
imaginary, or 0, are the only possibilities. The name Killing spinor derives from
the fact that if ψ is a non-trivial Killing spinor and α is real, the vector field
Xn
(14.1.1) Xψ = g(ψ, Ej · ψ)Ej
j=1

is a Killing vector field for the metric g (which, of course, can be zero). If ψ is a
Killing spinor on an n-dimensional spin manifold, then
X n X n
(14.1.2) Dψ = Ej · ∇Ej ψ = αEj · Ej ·ψ = −nαψ .
j=1 j=1

So Killing spinors are eigenvectors of the Dirac operator with eigenvalue −nα. In
1980 Friedrich [Fri80] proved the following remarkable theorem:
Theorem 14.1.4: Let (M n , g) be a Riemannian spin manifold which admits a
non-trivial Killing spinor ψ. Then (M n , g) is Einstein with scalar curvature s =
4n(n − 1)α2 .
A proof of this is a straightforward curvature computation which can be found
in either of the books [BFGK91, Fri00]. It also uses the fact that a non-trivial
Killing spinor vanishes nowhere. It follows immediately from Theorem 14.1.4 that
α must be one of the three types mentioned in Definition 14.1.3. So if the Killing
number is real then (M, g) must be a positive Einstein manifold. In particular, if
1Here the standard terminology real and imaginary Killing spinors can be somewhat mis-
leading. The Killing spinor ψ is usually a section of a complex spinor bundle. So a real Killing
spinor just means that α is real.
14.1. THE DIRAC OPERATOR AND KILLING SPINORS 531

M is complete, then it is compact. On the other hand if the Killing number is pure
imaginary, Friedrich shows that M must be non-compact.
The existence of Killing spinors not only puts restrictions on the Ricci curva-
ture, but also on both the Riemannian and the Weyl curvature operators [BFGK91].
Proposition 14.1.5: Let (M n , g) be a Riemannian spin manifold. Let ψ be a
Killing spinor on M with Killing number α and let R, W : Λ2 M −→Λ2 M be the
Riemann and Weyl curvature operators, respectively. Then for any vector field X
and any 2-from β we have
(14.1.3) W(β) · ψ = 0 ;
¡ ¢
(14.1.4) (∇X W)(β) · ψ = −2α X W(β) · ψ ;
(14.1.5) (R(β) + 4α2 β) · ψ = 0 ;
¡ ¢
(14.1.6) (∇X R)(β) · ψ = −2α X R(β) + 4α2 β(X) · ψ .
These curvature equations can be used to prove (see [BFGK91] or [Fri00])
Theorem 14.1.6: Let (M n , g) be a connected Riemannian spin manifold admit-
ting a non-trivial Killing spinor with α 6= 0. Then (M, g) is locally irreducible.
Furthermore, if M is locally symmetric, or n ≤ 4, then M is a space of constant
sectional curvature equal 4α2 .
Friedrich’s main objective in [Fri80] was an improvement of Lichnerowicz’s
estimate in [Lic63a] for the eigenvalues of the Dirac operator. Indeed, Friedrich
proves that the eigenvalues λ of the Dirac operator on any compact manifold satisfy
the estimate
1 ns0
(14.1.7) λ2 ≥ ,
4n−1
where s0 is the minimum of the scalar curvature on M. Thus, Killing spinors ψ
are eigenvectors that realize equality in equation (14.1.7). Friedrich also proves the
converse that any eigenvector of D realizing the equality must be a Killing spinor
with
r
1 s0
(14.1.8) α=± .
2 n(n − 1)
Example 14.1.7: [Spheres] In the case of the round sphere (S n , g0 ) equality in
equation (14.1.7) is always attained. So normalizing such that s0 = n(n − 1), and
using the Bär’s Correspondence Theorem 14.2.1 below the number of corresponding
real Killing spinors equals the number of constant spinors on Rn+1 with the flat
metric. The latter is well known (see the appendix of [PR88]) to be 2bn/2c for each
of the values α = ± 21 , where bn/2c is the largest integer less than or equal to n/2.
Remark 14.1.1: Actually (without making the connection to Sasakian geometry)
already in [Fri80] Friedrich gives a non-spherical example of a compact 5-manifold
with a real Killing spinor: M = SO(4)/SO(2) with its homogeneous Kobayashi-
Tanno Sasaki-Einstein structure.
We now wish to relate Killing spinors to the main theme of this book, Sasakian
geometry. First notice that if a Sasakian manifold M 2n+1 admits a Killing spinor,
Theorem 14.1.4 says it must be Sasaki-Einstein, so the scalar curvature s0 =
2n(2n + 1), and equation (14.1.8) implies that α = ± 21 . We have the following
result of Friedrich and Kath [FK90]
532 14. SASAKIAN STRUCTURES, KILLING SPINORS, AND SUPERSYMMETRY

Theorem 14.1.8: Every simply connected Sasaki-Einstein manifold admits non-


trivial real Killing spinors. Furthermore,
(i) if M has dimension 4m + 1 then (M, g) admits exactly one Killing spinor
for each of the values α = ± 12 ,
(ii) if M has dimension 4m+3 then (M, g) admits at least two Killing spinors
for one of the values α = ± 12 .
Outline of Proof. (Details can be found in [FK90] or the book [BFGK91].)
Every simply connected Sasaki-Einstein manifold is spin by Theorem 11.1.2, so M
has a spin bundle S(M ). Given a fixed Sasakian structure S = (ξ, η, Φ, g) we con-
sider two subbundles E± (S) of S(M ) defined by
(14.1.9) E± (S) = {ψ ∈ S(M ) | (±2ΦX + £ξ X) · ψ = 0, ∀X ∈ Γ(T M )} .
Set ∇±
X= ∇X ± 1
2X ·.A straightforward computation shows that ∇± preserves the
subbundles E± and defines a connection there. Moreover, by standard curvature
computations it can be shown that the connection ∇± is flat in E± (S). So it has
covariantly constant sections which are precisely the Killing spinors. One then uses
some representation theory of Spin(2n + 1) to compute the dimensions of E+ (S)
and E− (S) proving the result. ¤
We have the following:
Corollary 14.1.9: Let (M, g) be a compact Sasaki-Einstein manifold of dimen-
sion 2m + 1. Then (M, g) is locally symmetric if and only if (M, g) is of constant
curvature. Moreover, Hol(g) = SO(2m + 1) and (M, g) is locally irreducible as a
Riemannian manifold.
Proof. If necessary, go to the universal cover M̃ . This is a compact simply
connected Sasaki-Einstein manifold; hence, it admits a non-trivial Killing spinor by
Theorem 14.1.8. The first statement then follows from the Theorem 14.1.6. The
second statement follows from the Berger Theorem 1.4.8. Since M has dimension
2m + 1 the only possibilities for Hol(g) are SO(2m + 1) and G2 . But the latter is
Ricci flat, so it cannot be Sasaki-Einstein. ¤
Friedrich and Kath began their investigation in dimension 5 [FK89] where
they showed that a simply-connected compact 5-manifold which admits a Killing
spinor must be Sasaki-Einstein. In dimension 7 they showed that there are exactly
three possibilities: weak G2 -manifolds, Sasaki-Einstein manifolds which are not 3-
Sasakian, and 3-Sasakian manifolds [FK90]. Later Grunewald gave a description of
6-manifolds admitting Killing spinors [Gru90]. We should add an earlier result of
Hijazi who showed that the only 8-dimensional manifold with Killing spinors must
be the round sphere [Hij86]. By 1990 a decade of research by many people slowly
identified all the ingredients of a classification of such manifolds in terms of their un-
derlying geometric structures. The pieces of the puzzle consisting of round spheres
in any dimension, Sasaki-Einstein manifolds in odd dimensions, nearly Kähler man-
ifolds in dimension 6, and weak G2 -holonomy manifolds in dimension 7 were all in
place with plenty of interesting examples to go around [BFGK91]. What remained
at that stage was to show that in even dimensions greater than 8 there is noth-
ing else but the round spheres, while in odd dimensions greater than 7 the only
such examples must be Sasaki-Einstein. The missing piece of the puzzle was fi-
nally uncovered by Bär: real Killing spinors on M correspond to parallel spinors on
14.2. REAL KILLING SPINORS, HOLONOMY AND BÄR’S CORRESPONDENCE 533

the cone C(M ) [Bär93]. A bit earlier Wang [Wan89] had shown that on a sim-
ply connected complete Riemannian spin manifold the existence of parallel spinors
correspond to reduced holonomy. This led Bär to an elegant description of the
geometry of manifolds admitting real Killing spinors (in any dimension) in terms of
special holonomies of the associated cones. We refer to the correspondence between
real Killing spinors on M and parallel spinors on the cone C(M ) (equivalently re-
duced holonomy) as Bär’s correspondence. In particular, this correspondence not
only answered the last remaining open questions, but also allowed for simple unified
proofs of most of the theorems obtained earlier.

14.2. Real Killing Spinors, Holonomy and Bär’s Correspondence


As mentioned the Bär correspondence relates real Killing spinors on a com-
pact Riemannian spin manifold (M, g) to parallel spinors on the Riemannian cone
(C(M ), ḡ). We now make this statement precise.
Theorem 14.2.1: Let (M n , g) be a compact Riemannian spin manifold and let
(C(M n ), ḡ) be its Riemannian cone. Then there is a one to one correspondence
between real Killing spinors on (M n , g) with α = ± 21 and parallel spinors on
(C(M n ), ḡ).
Proof. The existence of a parallel spinor on (C(M n ), ḡ) implies that ḡ is Ricci
flat by Theorem 14.1.4. Then by Lemma 11.1.5 (M n , g) is Einstein with scalar
curvature s = n(n − 1). So any Killing spinors must have α = ± 12 by equation
(14.1.8). As in the proof of Theorem 14.1.8, ∇± 1
X = ∇X ± 2 X· defines a connection
in the spin bundle S(M ). The connection 1-forms ω of ∇± are related to the
±

connection 1-form ω of the Levi-Civita connection by ω ± = ω ± 21 β, where β is a 1-


form called the soldering form. This can be interpreted as a connection with values
in the Lie algebra of spın(n + 1) = spın(n) ⊕ Rn , and pulls back to the Levi-Civita
connection in the spin bundles on the cone (C(M n ), ḡ). So parallel spinors on the
cone correspond to parallel spinors on (M, g) with respect to the connection ∇±
which correspond precisely to real Killing spinors with respect to the Levi-Civita
connection. ¤
Now we have the following definition:
Definition 14.2.2: We say that a Riemannian spin manifold (M, g) is of type
(p, q) if it carries exactly p linearly independent real Killing spinors with α > 0 and
exactly q linearly independent real Killing spinors with α < 0.
The following theorem has an interesting history. As mentioned above it
was Bär [Bär93] who recognized the correspondence between real Killing spinors
on (M, g) and parallel spinors on the Riemannian cone (C(M ), ḡ). The relation
between parallel spinors and reduced holonomy was anticipated in the work of
Hitchin [Hit74] and Bonan [Bon66], but was formalized in the 1989 paper of
Wang [Wan89]. It has also been generalized to the non-simply connected case in
[Wan95, MS00].
Theorem 14.2.3: Let (M n , g) be a complete simply connected Riemannian spin
manifold, and let Hol(ḡ) be the holonomy group of the Riemannian cone (C(M ), ḡ).
Then (M n , g) ¡ admits a non-trivial
n
¢ real Killing spinor with (M , g) of type (p, q) if
and only if dimM, Hol(ḡ), (p, q) is one of the 6 possible triples listed in the table
below:
534 14. SASAKIAN STRUCTURES, KILLING SPINORS, AND SUPERSYMMETRY

dim(M ) Hol(ḡ) type (p, q)


n id (2bn/2c , 2bn/2c )
4m + 1 SU (2m + 1) (1, 1)
4m + 3 SU (2m + 2) (2, 0)
4m + 3 Sp(m + 1) (m + 2, 0)
7 Spin(7) (1, 0)
6 G2 (1, 1)

Here m ≥ 1, and n > 1.

Outline of Proof. Since (M, g) is complete and has a non-trivial real Killing
spinor, it is compact by Theorem 14.1.4. It then follows from a theorem of Gallot
[Gal79] that if the Riemannian cone (C(M ), ḡ) has reducible holonomy it must be
flat. So we can apply Berger’s Theorem 1.4.8. Now Wang [Wan89] used the spinor
representations of the possible irreducible holonomy groups on Berger’s list to give
the correspondence between these holonomy groups and the existence of parallel
spinors. First he showed that the groups listed in Table 1.4.7 that are not on the
above table do not admit parallel spinors. Then upon decomposing the spin rep-
resentation of the group in question into irreducible pieces, the number of parallel
spinors corresponds to the multiplicity of the trivial representation. Wang computes
this in all but the first line of the table when (C(M ), ḡ) is flat. In this case (M, g) is
a round sphere as discussed in Example 14.1.7, so the number of linearly indepen-
dent constant spinors is (2bn/2c , 2bn/2c ). By Bär’s Correspondence Theorem 14.2.1
real Killing spinors on (M, g) correspond precisely to parallel spinors on (C(M ), ḡ).
Note that the hypothesis of completeness in Wang’s theorem [Wan89] is not neces-
sary, so that the correspondence between the holonomy groups and parallel spinors
holds equally well on Riemannian cones. However, the completeness assumption on
(M, g) guarantees the irreducibility of the cone (C(M ), ḡ) as mentioned above. ¤

Let us briefly discuss the types of geometry involved in each case of this the-
orem. As mentioned in the above proof the first line of the table corresponds to
the round spheres. The next three lines correspond to Sasaki-Einstein geometry,
so Theorem 14.2.3 generalizes the Friedrich-Kath Theorem 14.1.8 in this case. The
last of these three lines corresponds precisely to 3-Sasakian geometry by Definition
13.1.9. Finally the two cases whose cones have exceptional holonomy will be dis-
cussed in more detail in Section 14.3.1 below. Suffice it here to mention that it
was observed by Bryant and Salamon [BS89] that a cone on a weak holonomy G2
manifold has its own holonomy in Spin(7). It is interesting to note that Theorem
14.2.3 generalizes the result of Hijazi in dimension eight mentioned earlier as well
as part of the last statement in Theorem 14.1.6, namely
Corollary 14.2.4: Let (M 2n , g) be a complete simply connected Riemannian spin
manifold of dimension 2n with n 6= 3 admitting a non-trivial real Killing spinor.
Then M is isometric to the round sphere.
14.3. GEOMETRIES ASSOCIATED WITH 3-SASAKIAN 7-MANIFOLDS 535

We end this section with a brief discussion of the non-simply connected case.
Here we consider two additional cases for Hol(ḡ), namely SU (2m + 2) o Z2 and
Sp(2) × Zd . See [Wan95, MS00] for the list of possibilities.
Example 14.2.5: Hol(ḡ) = SU (2m) o Z2 . Consider the (4m − 1)-dimensional
Stiefel manifold V2 (R2m+1 ) with its homogeneous Sasaki-Einstein metric. As de-
scribed in Example 9.3.23 the quotient manifold Mσ4m−1 of V2 (R2m+1 ) by the
free involution σ induced from complex conjugation has an Einstein metric which
is “locally Sasakian”. The cone C(Mσ4m−1 ) is not Kähler and its holonomy is
Hol(ḡ) = SU (2m + 2) o Z2 . According to Wang [Wan95] C(Mσ4m−1 ) admits a spin
structure with precisely one parallel spinor if and only if m is even, and according
to Moroianu and Semmelmann [MS00] C(Mσ4m−1 ) admits exactly two spin struc-
tures each with precisely one parallel spinor if m is even. Thus, by Theorem 14.2.1
Mσ4m−1 admits exactly two spin structures each with exactly one Killing spinor if
and only if m is even.
Example 14.2.6: Consider a 3-Sasakian manifold (M 4n−1 , S) and choose a Reeb
vector field ξ(τ ). Let Cm be the cyclic subgroup of order m > 2 of the circle group
generated by ξ(τ ). Assume that m is relatively prime to the order υ(S) of S and
that the generic fibre of the fundamental 3-dimensional foliation FQ is SO(3), so
that Cm acts freely on M 4n−1 . This last condition on the generic fibre is easy to
satisfy; for example, it holds for any of the 3-Sasakian homogeneous spaces other
than the standard round sphere, as well as the biquotients described in Section
13.7.4 when all the pj ’s are odd. (To handle the case when the generic fibre is Sp(1)
we simply need to divide m by two when it is even). Since Cm is not in the center
of SO(3), the quotient M 4n−1 /Cm is not 3-Sasakian. However, Cm does preserve
the Sasakian structure determined by ξ(τ ), so M 4n−1 /Cm is Sasaki-Einstein. The
cone C(M 4n−1 /Cm ) has holonomy Sp(n) × Zm , and admits precisely n+1 m parallel
spinors if and only if m divides n + 1 [Wan95, MS00]. Thus, by Theorem 14.2.1
M 4n−1 /Cm admits precisely n+1 m Killing spinors when m divides n + 1.

14.3. Geometries Associated with 3-Sasakian 7-manifolds


It is most remarkable that to each 4n-dimensional positive QK metric (O, gO )
(even just locally) one can associate nine other Einstein metrics in dimensions 4n+k,
k = 1, 2, 3, 4. Alternatively, one could say that each 3-Sasakian metric (M, g) canon-
ically defines an additional nine Einstein metrics in various dimensions. We have
already encountered all of these metrics. First there are the four geometries of the
diamond diagram ♦(M, S) of Definition 13.3.15. Then M and Z admit additional
“squashed” Einstein metrics discussed in Section 13.3.3. Thus we get five Einstein
metrics with positive Einstein constants: (O, gO ), (M, g), (M 0 , g 0 ), (Z, h), (Z 0 , h0 ).
Of course M ' M 0 and Z ' Z 0 as smooth manifolds (orbifolds) but they are differ-
ent as Riemannian manifolds (orbifolds), hence, the notation. Let us scale all these
metrics so that the Einstein constant equals the dimension of the total space minus
1. Note that any 3-Sasakian metric already has this property. In the other four
cases this is a choice of scale which is quite natural due to Lemma 11.1.5. However,
note that this is not the scale one gets for (Z, h), and (O, gO ) via the Riemannian
submersion from (M, g). Now, in each case one can consider its Riemannian cone
which will be Ricci-flat by Lemma 11.1.5. We thus obtain five Ricci-flat metrics
on the corresponding Riemannian cones. In addition, one can also take (iterated)
sine-cone metrics (see Equation (14.4.1) below) on the same five base spaces. These
536 14. SASAKIAN STRUCTURES, KILLING SPINORS, AND SUPERSYMMETRY

metrics are all Einstein of positive scalar curvature (cf. Theorem 14.4.2). Let us
summarize all this with the following extension of ♦(M, S):
(14.3.1) C(Z 0 ) o ? _ Z0 ÂÄ / C(M 0 )
AA M0
AA xx
AA xx
AA xxx
à  |xÄ x
O bF / C(O)
}}> FF
} FF
}}} FF
FF
} }
C(Z) o ? _ o ÂÄ / C(M )
Z M
There would perhaps be nothing special about all these 10 (and many more
by iterating the sine-cone construction) geometries beyond what has already been
discussed in Chapter 13. This is indeed true when dim(M ) > 7. However, when
dim(M ) = 7, or, alternatively, when O is a positive self-dual Einstein orbifold metric
(more generally, just a local metric of this type) some of the metrics occurring in
diagram (14.3.1) have additional properties. We shall list all of them first. For the
moment, let us assume that (M, g) is a compact 3-Sasakian 7-manifold, then the
following hold:
(i) (O, gO ) is a positive self-dual Einstein manifold (orbifold). We will think
of it as the source of all the other geometries.
(ii) (C(O), dt2 + t2 gO ) is a 5-dimensional Ricci-flat cone with base O.
(iii) (Z, h) is the orbifold twistor space of O.
(iv) (Z 0 , h0 ) is a nearly-Kähler manifold (orbifold).
(v) (M, g) is the 3-Sasakian manifold.
(vi) (M 0 , g 0 ) is a 7-manifold with weak G2 structure.
(vii) (C(Z), dt2 + t2 h) is a 7-dimensional Ricci-flat cone with base Z.
(viii) (Cs (Z 0 ), dt2 + sin2 th0 ) is a 7-manifold with weak G2 structure.
(ix) (C(Z 0 ), dt2 + t2 h0 ) has holonomy contained in G2 .
(x) (C(M ), dt2 + t2 g) is hyperkähler with holonomy contained in Sp(2).
(xi) (C(M 0 ), dt2 + t2 g 0 ) has holonomy contained in Spin(7).
The cases (ii) and (vii) do not appear to have any special properties other than
Ricci-flatness. The cases (i), (iii), (v), and (x) are the four geometries of ♦(M, S).
The five remaining cases are all very interesting from the point of view of the clas-
sification of Theorem 14.2.3. Indeed Z 0 and C(Z 0 ) are examples of the structures
listed in the last row of the table while M 0 and C(M 0 ) (as well as Cs (Z 0 ) and
C(Cs (Z 0 ))) give examples of the structures listed in the fifth row. In particular,
our diagram (14.3.1) provides for a cornucopia of the orbifold examples in the first
case and smooth manifolds in the latter.
14.3.1. Weak G2 -Structures and Spin(7) Holonomy Cones. Recall, that
geometrically G2 is defined to be the Lie group acting on the imaginary octonions
R7 and preserving the 3-form
ϕ = α1 ∧ α2 ∧ α3 + α1 ∧ (α4 ∧ α5 − α6 ∧ α7 )
(14.3.2)
+ α2 ∧ (α4 ∧ α6 − α7 ∧ α5 ) + α3 ∧ (α4 ∧ α7 − α5 ∧ α6 ),
where {αi }7i=1
is a fixed orthonormal basis of the dual of R7 . A G2 structure
on a 7-manifold M is, by definition, a reduction of the structure group of the
tangent bundle to G2 . This is equivalent to the existence of a global 3-form ϕ ∈
14.3. GEOMETRIES ASSOCIATED WITH 3-SASAKIAN 7-MANIFOLDS 537

Ω3 (M ) which may be written locally as 14.3.2. Such a 3-form defines an associated


Riemannian metric, an orientation class, and a spinor field of constant length. The
following terminology is due to Gray [Gra71]:
Definition 14.3.1: Let (M, g) be a complete 7-dimensional Riemannian manifold.
We say that (M, g) has weak holonomy G2 if there exist a global 3-form ϕ ∈
Ω3 (M ) which locally can be written in terms of a local orthonormal basis as in
14.3.2, and dϕ = c ? ϕ, where ? is the Hodge star operator associated to g and c is
a constant whose sign is fixed by an orientation convention.
The equation dϕ = c ? ϕ implies that ϕ is ‘nearly parallel’ in the sense that
only a 1-dimensional component of ∇ϕ is different from zero [FG82]. Thus, a weak
holonomy G2 structure is sometimes called a nearly parallel G2 structure. The case
of c = 0 is somewhat special. In particular, it is known [Sal89] that the condition
dϕ = 0 = d ? ϕ is equivalent to the condition that ϕ be parallel, i.e., ∇ϕ = 0 which
is equivalent to the condition that the metric g has holonomy group contained in
G2 . The following theorem provides the connection with the previous discussion on
Killing spinors [Bär93]
Theorem 14.3.2: Let (M, g) be a complete 7-dimensional Riemannian manifold
with weak holonomy G2 . Then the holonomy Hol(ḡ) of the metric cone (C(M ), ḡ)
is contained in Spin(7). In particular, C(M ) is Ricci-flat and M is Einstein with
positive Einstein constant λ = 6.
Remark 14.3.1: The sphere S 7 with its constant curvature metric is isometric to
the isotropy irreducible space Spin(7)/G2 . The fact that G2 leaves invariant (up
to constants) a unique 3-form and a unique 4-form on R7 implies immediately that
this space has weak holonomy G2 .
Definition 14.3.3: Let (M, g) be a complete 7-dimensional Riemannian manifold.
We say that g is a proper G2 -metric if Hol(ḡ) = Spin(7).
We emphasize here that G2 is the structure group of M, not the Riemannian
holonomy group. Specializing Theorem 14.2.3 to dimension 7 gives the following
theorem due to Friedrich and Kath [FK90].
Theorem 14.3.4: Let (M 7 , g) be a complete simply-connected Riemannian spin
manifold of dimension 7 admitting a non-trivial real Killing spinor with α > 0 or
α < 0. Then there are four possibilities:
(i) (M 7 , g) is of type (1, 0) and it is a proper G2 -manifold,
(ii) (M 7 , g) is of type (2, 0) and it is a Sasaki-Einstein manifold, but (M 7 , g)
is not 3-Sasakian,
(iii) (M 7 , g) is of type (3, 0) and it is 3-Sasakian,
(iv) (M 7 , g) = (S 7 , gcan ) and is of type (8, 8).
Conversely, if (M 7 , g) is a compact simply-connected proper G2 -manifold then it
carries precisely one Killing spinor with α > 0. If (M 7 , g) is a compact simply-
connected Sasaki-Einstein 7-manifold which is not 3-Sasakian then M carries pre-
cisely 2 linearly independent Killing spinors with α > 0. Finally, if (M 7 , g) is a
3-Sasakian 7-manifold, which is not of constant curvature, then M carries precisely
3 linearly independent Killing spinors with α > 0.
Remark 14.3.2: The four possibilities of the Theorem 14.3.4 correspond to the
sequence of inclusions
Spin(7) ⊃ SU (4) ⊃ Sp(2) ⊃ {id} .
538 14. SASAKIAN STRUCTURES, KILLING SPINORS, AND SUPERSYMMETRY

All of the corresponding cases are examples of weak holonomy G2 metrics. If


we exclude the trivial case when the associated cone is flat, we have three types of
weak holonomy G2 geometries. Following [FKMS97] we use the number of linearly
independent Killing spinors to classify the types of weak holonomy G2 geometries.
We call these type I, II, and III corresponding to cases (i), (ii), and (iii) of Theorem
14.3.4, respectively.
We are now ready to describe the weak G2 geometry of the M 0 ,→ C(M 0 ) part
of Diagram 14.3.1 [GS96, FKMS97]:
Theorem 14.3.5: Let (M, S) be a 7-dimensional 3-Sasakian manifold. Then the
3-Sasakian metric g has weak holonomy G2 . Moreover, the second Einstein metric
g 0 given by Theorem 13.3.18 and scaled so that the Einstein constant λ = 6 has
weak holonomy G2 . In fact g 0 is a proper G2 metric.

Proof. For √ the second Einstein


√ 2 metric √g 0 we have three mutually orthonormal
1-forms α = tη , α = tη , α = tη 3 , where t is the parameter of the
1 1 2 3

canonical variation. Let {α4 , α5 , α6 , α7 } be local 1-forms spanning the annihilator


of the vertical subbundle V3 in T ∗ S such that
Φ̄1 = 2(α4 ∧ α5 − α6 ∧ α7 ) ,

Φ̄2 = 2(α4 ∧ α6 − α7 ∧ α5 ) ,
Φ̄3 = 2(α4 ∧ α7 − α5 ∧ α6 ) .
Then the set {α1 , . . . , α7 } forms a local orthonormal coframe for the metric g 0 . In
√ √3
terms of the 3-forms Υ and Θ of 13.5.1 we have ϕ = 21 tΘ + t Υ. One easily
sees that this is of the type of equation (14.3.2) and, therefore, defines a compatible
G2 -structure. Moreover, a straightforward computation gives
1√ √ 1 1
dϕ = tΩ + t(t + 1)dΥ, ?ϕ = − tdΥ − Ω .
2 2 24
√ √ √
Thus, dϕ = c ? ϕ is solved with t = 1/ 5, and c = −12/ 5. So g 0 has weak
holonomy G2 . That g 0 is a proper G2 metric is due to [FKMS97]. The idea is
to use Theorem 14.3.4. Looking at the four possibilities given in that theorem,
we see that it suffices to show that g 0 is not Sasaki-Einstein. The details are in
[FKMS97]. ¤

Example 14.3.6: 3-Sasakian 7-manifolds are plentiful as we have seen in Chapter


13. All of them give, by Theorem 14.3.5, examples of type I and type III geome-
tries. Examples of simply connected type I geometries that do not arise via Theorem
7
14.3.5 are the homogeneous Aloff-Wallach spaces Mm,n , (m, n) 6= (1, 1), discussed
in Section 13.7.4 as special case of Eschenburg biquotients [CMS96, BFGK91]
together with an isotropy irreducible homogeneous space defined as follows: Con-
sider the space H2 of homogeneous polynomials of degree 2 in three real variables
(x1 , x2 , x3 ). As dim(H2 ) = 5 it gives rise to the embedding SO(3) ⊂ SO(5). We take
M = SO(5)/SO(3). This example was used by Bryant to get the first 8-dimensional
metric with holonomy Spin(7) [Bry87]. Examples of type II geometries (Sasaki-
Einstein) are equally rich as seen in Chapter 11, Sections 11.5 and 11.7. In partic-
ular, there are hundreds of examples of type II weak holonomy G2 metrics on each
of the 28 homotopy spheres in dimension 7.
14.3. GEOMETRIES ASSOCIATED WITH 3-SASAKIAN 7-MANIFOLDS 539

7
Remark 14.3.3: According to [CMS96] the Aloff-Wallach manifold M1,1 has three
Einstein metrics. One is the homogeneous 3-Sasakian metric. The second is the
proper G2 metric of Theorem 14.3.5, while the third Einstein metric also has weak
holonomy G2 most likely of type I but we could not positively exclude type II as a
possibility.
Open Problem 14.3.1: Classify all compact 7-manifolds with weak G2 holonomy
of type I, II, or III, respectively.
Type III classification is the classification of all compact 3-Sasakian 7-manifolds,
or more generally all compact positive self-dual Einstein orbifolds. This is proba-
bly very hard. The case of 3-Sasakian 7-manifolds with vanishing aut(M, S) ap-
pears quite difficult. Type II classification (7-dimensional Sasaki-Einstein manifolds
which are not 3-Sasakian) is clearly completely out of reach at the moment. A clas-
sification of proper weak G2 structures on a compact manifold that do not arise via
Theorem 14.3.5 would be very interesting and it is not clear how hard this problem
really is.
Remark 14.3.4: The holonomy Spin(7) cone metrics are plentiful but never com-
plete. However, some of these metrics can be deformed to complete holonomy
Spin(7) ones on non compact manifolds. The first example was obtained in Bryant
and Salamon who observed that the spin bundle over S 4 with its canonical metric
carries a complete metric with holonomy Spin(7) [BS89]. Locally the metric was
later considered also in [GPP90]. More generally, spin orbibundles over positive
QK orbifolds also carry such complete orbifold metrics as observed by Bryant and
Salamon in [BS89]. Other complete examples were constructed later by physicists
[CGLP02, CGLP04, KY02a, KY02b]. Finally, the first compact examples
were obtained in 1996 by Joyce [Joy96a, Joy99]. See Joyce’s book [Joy00] for an
excellent detailed exposition of the methods and the discussion of examples.
Open Problem 14.3.2: Let (M 7 , S) be any 3-Sasakian 7-manifold introduced in
Chapter 13 and let (M 7 , g 0 ) be the associated proper weak holonomy G2 squashed
metric. Consider the two Riemannian cones (C(M ), dt2 + t2 g 0 ), (C(M ), dt2 + t2 g).
(i) When does the metric cone (C(M ), dt2 +t2 g 0 ) admit a complete holonomy
Spin(7) deformation?
(ii) When does the metric cone (C(M ), dt2 +t2 g) admit a complete holonomy
Sp(2) (hyperkähler) deformation?
The metric on the spin bundle S(S 4 ) by Bryant and Salamon is a deformation
on the Spin(7) holonomy metric on the cone over the squashed metric on S 7
[CGLP02, CGLP04], so there are examples of such deformations regarding ques-
tion (i). Regarding (ii): Recall that every compact 3-Sasakian 3-manifold is isomet-
ric to S 3 /Γ and the metric cone is the flat cone C2 /Γ. Hence, one could think of (ii)
as a 7-dimensional analogue of the similar problem whose complete solution was
described in Section 12.10. There are non-trivial examples also in the higher dimen-
sional cases. The metric cone on the homogeneous 3-Sasakian manifold S(1, 1, 1)
admits complete HK deformations, namely the Calabi metric on T ∗ CP2 of Example
12.8.5. We do not know of any other examples at the moment.

14.3.2. Nearly Kähler 6-Manifolds and G2 Holonomy Cones. In this


section we explain the geometry of the Z 0 ,→ C(Z 0 ) part of the diagram 14.3.1.
Before we specialize to dimension 6 we begin with a more general introduction.
540 14. SASAKIAN STRUCTURES, KILLING SPINORS, AND SUPERSYMMETRY

Nearly Kähler manifolds were first studied by Tachibana in [Tac59] and they appear
under the name of almost Tachibana spaces in Chapter VIII of the book [Yan65].
They were then rediscovered by Gray [Gra70] and given the name nearly Kähler
manifolds which by now is the accepted name.
Definition 14.3.7: A nearly Kähler manifold is an almost Hermitian manifold
(M, g, J, ω) such that (∇X J)X = 0 for all tangent vectors X, where ∇ is the Levi-
Civita connection and J is the almost complex structure. One says that a nearly
Kähler manifold is strict if it is not Kähler.
This definition is equivalent to the condition
(14.3.3) (∇X J)Y + (∇Y J)X = 0
for all vector fields X, Y, which is to say that J is a Killing tensor field. An
alternative characterization of nearly Kähler manifolds is given by
Proposition 14.3.8: An almost Hermitian manifold (M, g, J, ω) is nearly Kähler
if and only if
1
∇ω = dω .
3
In particular, a strict nearly Kähler structure is never integrable.
Any nearly Kähler manifold can be locally decomposed as the product of a
Kähler manifold and a strict nearly Kähler manifold. Such a decomposition is
global in the simply connected case [Nag02b]. Hence, the study of nearly Kähler
manifolds reduces to the case of strict ones. In addition every nearly Kähler mani-
fold in dimension 4 must be Kähler so that the first interesting dimension is six.
The following theorem establishes relationship between the twistor space Z−→O
of a quaternionic Kähler manifold (orbifold) and nearly Kähler geometry.
Theorem 14.3.9: Let π : (Z, h)−→(O, gO ) be the twistor space of a positive QK
manifold with its Kähler structure (J, h, ωh ). Then Z admits a strict nearly Kähler
structure (J1 , h1 , ωh1 ). If T M = V ⊕ H is the natural splitting induced by π then
(14.3.4) h|V = 2h1|V , h|H = h1|H = π ∗ (gO ) ,

(14.3.5) J|V = −J1|V , J|H = J1|H .


Theorem 14.3.9 is due to Eels and Salamon when O is 4-dimensional. The
higher dimensional analogue was established in [AGI98] (see also [Nag02b]).
Remark 14.3.5: Observe that the metric of the nearly Kähler structure of Theorem
14.3.9, in general, is not Einstein. In particular, h1 is not the squashed metric h0
introduced in the diagram 14.3.1, unless dim(Z) = 6. In six dimensions, we can
scale h1 so that it has scalar curvature s = 30 and then indeed h1 = h0 as one can
easily check.
Definition 14.3.10: Let M = G/H be a homogeneous space. We say that M is
3-symmetric if G has an automorphism σ of order 3 such that Gσ0 ⊂ H ⊂ Gσ ,
where Gσ is the fixed point set of σ and Gσ0 is the identity component in Gσ0 .
We have the following two theorems concerning nearly Kähler homogeneous
Riemannian manifolds. The first is due to Wolf and Gray in all dimensions but
six [WG68a, WG68b]. They also conjectured that the result is true for strict
nearly Kähler 6-manifolds. The Wolf-Gray conjecture was proved quite recently by
Butruille [But05, But06] which is the second theorem below.
14.3. GEOMETRIES ASSOCIATED WITH 3-SASAKIAN 7-MANIFOLDS 541

Theorem 14.3.11: Every compact homogeneous strict nearly Kähler manifold M


of dimension different than 6 is 3-symmetric.
Theorem 14.3.12: Let (M, g) be a strict nearly Kähler 6-dimensional Riemannian
homogeneous manifold. Then M is isomorphic as a homogeneous space to a finite
quotient of G/H, where G and H are one of the following:

(i) G = SU (2) × SU (2) and H = {id};


(ii) G = G2 and H = SU (3), where metrically G/H = S 6 the round sphere;
(iii) G = Sp(2) and H = SU (2)U (1), where G/H = CP3 with its nearly
Kähler metric determined by Theorem 14.3.9;
(iv) G = SU (3) and H = T 2 , where G/H is the flag manifold with its nearly
Kähler metric determined by Theorem 14.3.9.

Each of these manifolds carries a unique invariant nearly Kähler structure, up to


homothety.
In every dimension, the only known compact examples of nearly Kähler mani-
folds are 3-symmetric. On the other hand, Theorem 14.3.9 can be easily generalized
to the case of orbifolds so that there are plenty examples of compact inhomogeneous
strict nearly Kähler orbifolds in every dimension.
Theorem 14.3.13: Let M be a compact simply-connected strict nearly Kähler
manifold. Then, in all dimensions, as a Riemannian manifold M decomposes as a
product of

(i) 3-symmetric spaces,


(ii) twistor spaces of positive QK manifolds Q such that Q is not symmetric,
(iii) 6-dimensional strict nearly Kähler manifold other than the ones listed in
Theorem 14.3.12.
This theorem is due to Nagy [Nag02a], but our formulation uses the result of
Butruille together with the fact that the twistor spaces of all symmetric positive QK
manifolds are 3-symmetric. The LeBrun-Salamon conjecture can now be phrased
as follows
Conjecture 14.3.14: Any compact simply connected strict irreducible nearly Kähler
manifold (M, g) of dimension greater than 6 must be a 3-symmetric space.
In particular, the Conjecture 14.3.14 is automatically true in dimensions 4n
because of Nagy’s classification theorem and also true in dimensions 10 and 14
because all positive QK manifolds in dimension 8 and 12 are known. The third
case leads to an important
Open Problem 14.3.3: Classify all compact strict nearly Kähler manifolds in
dimension 6.
Dimension six is special not just because of the rôle they play in Theorem
14.3.13. They have several remarkable properties which we summarize in the fol-
lowing theorem.
Theorem 14.3.15: Let (M, J, g, ωg ) be a compact strict nearly Kähler 6-manifold.
Then

(i) The metric g is Einstein of positive scalar curvature.


(ii) c1 (M ) = 0 and w2 (M ) = 0.
542 14. SASAKIAN STRUCTURES, KILLING SPINORS, AND SUPERSYMMETRY

(iii) If g is scaled so it has Einstein constant λ = 5 then the metric cone


(C(M ), dt2 t + t2 g) has holonomy contained in G2 . In particular, C(M )
is Ricci-flat.
The first property is due to Matsumoto [Mat72] while the second is due to
Gray [Gra76]. The last part is due to Bär [Bär93]. In fact, nearly Kähler 6-
manifolds is the geometry of the last row of the table of Theorem 14.2.3. More
precisely we have the following theorem proved by Grunewald [Gru90]:
Theorem 14.3.16: Let (M 6 , g) be a complete simply connected Riemannian spin
manifold of dimension 6 admitting a non-trivial Killing spinor with α > 0 or α < 0.
Then there are two possibilities:
(i) (M, g) is of type (1, 1) and it is a strict nearly Kähler manifold,
(ii) (M, g) = (S 6 , gcan ) and is of type (8, 8).
Conversely, if (M, g) is a compact simply-connected strict nearly Kähler 6-manifold
of non-constant curvature then M is of type (1, 1).
Compact strict nearly Kähler manifolds with isometries were investigated in
[MNS05] where it was shown that
Theorem 14.3.17: Let (M, J, g, ωg ) be a compact strict nearly Kähler 6-manifold.
If M admits a unit Killing vector field, then up to finite cover M is isometric to
S 3 × S 3 with its standard nearly Kähler structure.
Remark 14.3.6: The first example of a non-trivial G2 holonomy metric was found
by Bryant [Bry87], who observed that a cone on the complex flag manifold U(3)/T 3
carries an incomplete metric with G2 -holonomy. The flag U(3)/T 3 is the twistor
space of the complex projective plane CP2 and as such it also has a strict nearly
Kähler structure. As explained in this section, this therefore is just one possible
example. One gets such non-trivial metrics also for the cones with bases CP3 and
S 3 × S 3 with their homogeneous strict nearly Kähler structures. Interestingly,
in some cases there exist complete metrics with G2 holonomy which are smooth
deformations of the asymptotically conical ones. This fact was noticed by Bryant
and Salamon [BS89] who constructed complete examples of G2 holonomy metrics
on bundles of self-dual 2-forms over CP2 and S 4 . Replacing the base with any
positive QK orbifold O of Section 12.5 gives complete (in the orbifold sense) metrics
on orbibundles of self-dual 2-forms over O. Locally some of these metrics were
considered in [San03]. More complete examples of explicit G2 holonomy metrics on
non-compact manifolds were obtained by Salamon [Sal04]. G2 holonomy manifolds
with isometric circle actions were investigated by Apostolov and Salamon [AS04].
The first compact examples are due to the ground breaking work of Joyce [Joy96b].

14.4. Geometries Associated with Sasaki-Einstein 5-manifolds


Like 3-Sasakian manifolds Sasaki-Einstein 5-manifolds are naturally associated
to other geometries introduced in the previous section. Of course, each such space
(M 5 , S) comes with its Calabi-Yau cone (C(M ), ḡ) and, if the Sasaki-Einstein struc-
ture S is quasi-regular, with its quotient log del Pezzo surface (Z, h). But as it
turns out, there are two more Einstein metrics associated to g. The examples of
this section also illustrate how the Theorem 14.2.3 and the Bär’s correspondence
break down when (M, g) is a manifold with Killing spinors which is, however, not
complete.
14.4. GEOMETRIES ASSOCIATED WITH SASAKI-EINSTEIN 5-MANIFOLDS 543

We begin by describing a relation between 5-dimensional Sasaki-Einstein struc-


tures and six dimensional nearly Kähler structures which was uncovered recently
in [FIMU06]. This relation involves the so-called sine cone. This is the warped
product metric on Cs (N ) = N × (0, π) of Definition 6.5.1 with φ(t) = sin2 t, so the
Riemannian metric is
(14.4.1) ḡs = dt2 + sin2 t gN .
We use the notation ḡs to distinguish this metric from the usual Riemannian cone
metric ḡ. Of course this metric is not complete, but one can compactify M obtaining
a very tractable stratified space M̄ = N × [0, π] with conical singularities at t = 0
and t = π. Observe the following simple fact which shows that the Riemannian
cone on a sine cone is always a Riemannian product.
Lemma 14.4.1: Let (M, g) be a Riemannian manifold. Then the product metric
ds2 = dx2 + dy 2 + y 2 g on R × C(M ) can be identified with the iterated cone metric
on C(Cs (M )).
Proof. Consider the map R+ × (0, π)−→R × R+ given by polar coordinate
change (r, t) 7→ (x, y) = (r cos t, r sin t), where r > 0 and t ∈ (0, π). We get
ds2 = dx2 + dy 2 + y 2 g = dr2 + r2 dt2 + r2 sin2 tg = dr2 + r2 (dt2 + sin2 tg) . ¤
So the iterated Riemannian cone (C(Cs (M )), ds2 ) has reducible holonomy
1 × Hol(C(M )). The importance of sine-cones can be seen via the following easy
consequence of Lemmas 14.4.1 and 11.1.5:
Theorem 14.4.2: Let (N m , g) be a an Einstein manifold with Einstein constant
m − 1. Then the sine-cone Cs (N ) with the metric ḡs given by Equation (14.4.1) is
Einstein with Einstein constant m.
One can iterate this construction. Upon each iteration we get an Einstein
metric in one dimension higher with an appropriately scaled Einstein constant.
However, each iteration also introduces more singularities. One exception is the
sphere metric (S m , gcan ) scaled in each iteration so that Einstein constant equals
m − 1. The sine-cone metric gives the constant curvature metric with Einstein
constant m. In each iteration one gets only a part of the sphere, but it can be
completed to the usual round sphere metric on S m+1 . The case of special interest
to us is when the Einstein manifold N m of Theorem 14.4.2 is Sasaki-Einstein. Of
course, the Einstein constant is such that the sine-cone is Einstein, but now it
inherits more structure. For example, when m = 5 the metric on Cs (N 5 ) will be
nearly Kähler as we discuss next. First compare Lemma 14.4.1 with the following
result in [Joy00], Propositions 11.1.1-2:
Proposition 14.4.3: Let (M 4 , g4 ) and (M 6 , g6 ) be Calabi-Yau manifolds. Let
(R3 , ds2 = dx2 + dy 2 + dz 2 ) and (R, ds2 = dx2 ) be the Euclidean spaces. Then
(i) (R3 × M 4 , g = ds2 + g4 ) has a natural G2 structure and g has holonomy
Hol(g) ⊂ 1l3 × SU (2) ⊂ G2 ,
(ii) (R × M 6 , g = ds2 + g6 ) has a natural G2 structure and g has holonomy
Hol(g) ⊂ 1 × SU (3) ⊂ G2 .
As long as (M 4 , g4 ) and (M 6 , g6 ) are simply connected then the products
R × M 4 and R × M 6 are simply connected G2 -holonomy manifolds with reducible
3

holonomy groups and parallel Killing spinors. Note that this does not violate The-
orem 14.2.3 as these spaces are not Riemannian cones over complete Riemannian
544 14. SASAKIAN STRUCTURES, KILLING SPINORS, AND SUPERSYMMETRY

manifolds. Using (ii) of Proposition 14.4.3 we obtain the following corollary of


Theorem 14.4.2 first obtained in [FIMU06]
Corollary 14.4.4: Let (N 5 , g) be a Sasaki-Einstein manifold. Then the sine cone
Cs (N 5 ) = N 5 × (0, π) with metric ḡs is nearly Kähler of Einstein constant λ = 5.
Furthermore ḡs approximates pure SU (3) holonomy metric near the cone points.
Using Corollary 14.4.4 we obtain a host of examples of nearly Kähler 6-manifolds
with conical singularities by choosing N 5 to be any of the Sasaki-Einstein manifolds
discussed in detail in Section 11.4. For example, in this way we obtain nearly-Kähler
metrics on N × (0, π) where N is any Smale manifold with a Sasaki-Einstein metric
such as S 5 or k(S 2 ×S 3 ), etc. Note that every simply connected strict nearly Kähler
manifold has exactly two real Killing spinors. So, as long as N 5 is simply connected
Cs (N 5 ) will have two real Killing spinors. Using Theorem 14.4.2 the examples in
Chapter 11 in other dimensions also give new Einstein metrics on Cs (N 2n+1 ). For
example, one obtains many positive Einstein metrics on Σ2n+1 ×(0, π) where Σ2n+1
is any homotopy sphere in bP2n+2 . Of course, there are no Killing spinors unless
n = 2. Returning to the case of dimension 6, a somewhat more general converse
has been obtained in [FIMU06], namely
Theorem 14.4.5: Any totally geodesic hypersurface N 5 of a nearly Kähler 6-
manifold M 6 admits a Sasaki-Einstein structure.
The method in [FIMU06] uses the recently developed notion of hypo SU (2)
structure due to Conti and Salamon [CS06b]. The study of sine cones appears to
have originated in the physics literature [BM03, ADHL03], but in one dimension
higher. Now recall the following result of Joyce (cf. [Joy00], Propositions 13.1.2-3)
Proposition 14.4.6: Let (M 6 , g6 ) and (M 7 , g7 ) be Calabi-Yau and G2 -holonomy
manifolds, respectively. Let (R2 , ds2 = dx2 + dy 2 ) and (R, ds2 = dx2 ) be Euclidean
spaces. Then
(i) (R2 × M 6 , g = ds2 + g6 ) has a natural Spin(7) structure and g has holo-
nomy Hol(g) ⊂ 1l2 × SU (3) ⊂ Spin(7),
(ii) (R×M 7 , g = ds2 +g7 ) has a natural Spin(7) structure and g has holonomy
Hol(g) ⊂ 1 × G2 ⊂ Spin(7).
Again, if (M 6 , g6 ) and (M 7 , g7 ) are simply connected so are the Spin(7)-
manifolds R2 × M 6 and R × M 7 so that they have parallel spinors. Not surprisingly,
in view of Lemma 14.4.1 and Proposition 14.4.6, the sine cone construction now re-
lates strict nearly Kähler geometry in dimension 6 to weak holonomy G2 geometry
in dimension 7. More precisely [BM03]
Theorem 14.4.7: Let (N 6 , g) be a strict nearly Kähler 6-manifold such that g
has Einstein constant λ6 = 5. Then the manifold Cs (N ) = N 6 × (0, π) with its
sine cone metric ḡs has weak holonomy G2 with Einstein constant λ7 = 6 and it
approximates pure G2 holonomy metric near the cone points.

Proof. Just as before, starting with (N 6 , g6 ) we consider its metric cone


C(N 6 ) with the metric ḡ = dy 2 + y 2 g6 and the product metric g8 on R × C(N 6 ).
With the above choice of the Einstein constant we see that g8 = dx2 + dy 2 + y 2 g6
must have holonomy Hol(g8 ) ⊂ 1 × G2 ⊂ Spin(7). By Lemma 14.4.1 g8 is a met-
ric cone on the metric g7 = dt2 + sin2 tg6 , which must, therefore, have weak G2
holonomy and the Einstein constant λ7 = 6. ¤
14.5. GEOMETRIC STRUCTURES ON MANIFOLDS AND SUPERSYMMETRY 545

Again, any simply connected weak G2 -manifold has at least one Killing spinor.
That real Killing spinor on Cs (N 6 ) will lift to a parallel spinor on C(Cs (N 6 )) =
R × C(N 6 ) which is a non-complete Spin(7)-manifold of holonomy inside 1 × G2 .
One can iterate the two cases by starting with a compact Sasaki-Einstein 5-manifold
N 5 and construct either the cone on the sine cone of N 5 or the sine cone on the
sine cone of N 5 to obtain a weak holonomy G2 manifold. We list the Riemannian
manifolds coming from this construction that are irreducible.
Proposition 14.4.8: Let (N 5 , g5 ) be a compact Sasaki-Einstein manifold which is
not of constant curvature. Then the following have irreducible holonomy groups:
(i) the manifold C(N 5 ) with the metric g6 = dt2 + t2 g5 has holonomy SU (3);
(ii) the manifold Cs (N 5 ) = N 5 × (0, π) with metric g6 = dt2 + sin2 t g5 is
strict nearly Kähler;
(iii) the manifold Cs (Cs (N 5 )) = N 5 × (0, π) × (0, π) with the metric g7 =
dα2 + sin2 α(dt2 + sin2 t g5 ) has weak holonomy G2 .
In addition we have the reducible cone metrics: C(Cs (N 5 )) = R × C(N 5 ) has
holonomy in 1×SU (3) ⊂ G2 and C(Cs (Cs (N 5 ))) = R×C(Cs (N 5 )) = R×R×C(N 5 )
has holonomy 1l2 ×SU (3) ⊂ 1×G2 ⊂ Spin(7). If N 5 is simply connected then G5 , g6
and g7 admit two Killing spinors. For a generalization involving conformal factors
see [MO07].
Remark 14.4.1: Recall Remark 14.3.2. Note that when a weak holonomy G2
metric is not complete then the type I-III classification is no longer valid. The
group Spin(7) has other subgroups than the ones listed there and we can consider
the following inclusions of (reducible) holonomies

Spin(7) ⊃ G2 × 1 ⊃ SU (3) × 1l2 × ⊃ SU (2) × 1l3 ⊂ 1l8 .

According to the Friedrich-Kath Theorem 14.3.4 the middle three cannot occur as
holonomies of Riemannian cones of complete 7-manifolds with Killing spinors. But
as the discussion of this section shows, they most certainly can occur as holonomy
groups of Riemannian cones of incomplete weak holonomy G2 metrics. These met-
rics can be still separated into three types depending on the holonomy reduction:
say the ones that come from strict nearly Kähler manifolds are generically of type
Is while the ones that come from Sasaki-Einstein 5-manifolds via the iterated sine
cone construction are of type IIs and of type IIIs when H ⊂ SU (3) is some proper
non-trivial subgroup. On the other hand, it is not clear what is the relation between
the holonomy reduction and the actual number of the Killing spinors one gets in
each case.

14.5. Geometric Structures on Manifolds and Supersymmetry


Supersymmetry has emerged in physics as an attempt to unify the way physical
theories deal with bosonic and fermionic particles. Since its birth around the early
70ties it has come to dominate theoretical high energy physics (for a historical
perspective see [KS00] with the introduction by Kane and Shifman, and for a
mathematical treatment see [Var04]). This dominance is still ongoing in spite of the
fact that after almost 40 years there is no single experimental evidence that would
directly and convincingly “prove” or “discover” the existence of supersymmetry in
nature. On the other hand, especially in the last 20 years, supersymmetry has given
546 14. SASAKIAN STRUCTURES, KILLING SPINORS, AND SUPERSYMMETRY

birth to many beautiful mathematical theories. Gromov-Witten Theory, Seiberg-


Witten Theory, Rozansky-Witten Theory as well as the Mirror Duality Conjecture
are just a few of the more famous examples of important and deep mathematics
having its origins in the physics of various supersymmetric theories.
The intricate relationship between supersymmetry and geometric structures
on manifolds was recognized along the way the physics of supersymmetry slowly
evolved from its origins: first globally supersymmetric field theories (70ties) arose,
later came supergravity theory (80ties), which evolved into superstring theory and
conformal field theory (late 80ties and 90ties), and finally into M-Theory and the
supersymmetric branes of today. At every step the “first” theory would quickly
lead to various generalizations creating many different new ones: so it is as if after
discovering plain vanilla ice cream one would quickly find oneself in an Italian ice
cream parlor confused and unable to decide which flavor was the right choice for the
hot afternoon. This is a confusion that is possibly good for one’s sense of taste, but
many physicists believe that there should be just one theory, the Grand Unified
Theory which describes our world at any level.2 An interesting way out of this
conundrum is to suggest that even if two theories appear to be completely different,
if both are consistent and admissible, they actually do describe the same physical
world and, therefore, they should be dual to one another in a certain sense. This
gave rise to various duality conjectures such as the Mirror Symmetry Conjecture
or the AdS/CFT Duality Conjecture.
The first observation of how supersymmetry can restrict the underlying geom-
etry was due to Zumino [Zum79] who discovered that globally N = 1 supersym-
metric σ-models in d = 4 dimensions require that the bosonic fields (particles) of
the theory are local coordinates on a Kähler manifold. Later Alvarez-Gaumé and
Friedman observed that N = 2 supersymmetry requires that the σ-model manifold
be not just Kähler but hyperkähler [AGF81]. This relation between globally su-
persymmetric σ-models and complex manifolds was used by Lindström and Roček
to discover the hyperkähler quotient construction in [LR83, HKLR87].
The late seventies witnessed a series of attempts to incorporate gravity into the
picture which quickly led to the discovery of various supergravity theories. Again
the N = 1 supergravity-matter couplings in d = 4 dimensions require bosonic
matter fields to be coordinates on a Kähler manifold with some special properties
[WB82] while N = 2 supergravity demands that the σ-model manifold be quater-
nionic Kähler [BW83]. The quaternionic underpinnings of the matter couplings
in supergravity theories lead to the discovery of quaternionic Kähler reduction in
[Gal87a, GL88].
At the same time manifolds with Killing spinors emerged as important play-
ers in the physics of the supergravity theory which in D = 11 dimensions was
first predicted by Nahm [Nah78] and later constructed by Cremmer, Julia and
Scherk [CJS79]. The well-know Kaluza-Klein trick applied to a D = 11 supergrav-
ity model is a way of constructing various limiting compactifications which would
2Actually, string theory of today appears to offer a rather vast range of vacua (or possible
universes). Such possible predictions have been nicknamed the string landscape [Sus03]. This
fact has been seen as a drawback by some, but not all, physicists (see more recent discussion on
landscape and swampland in [Vaf05, OV06a]). The insistence that the Universe we experience,
and this on such a limited scale at best, is the only Universe, is largely a matter of ‘philosophical
attitude’ towards science. See the recent book of Leonard Susskind on string theory and cosmic
landscape [Sus05].
14.5. GEOMETRIC STRUCTURES ON MANIFOLDS AND SUPERSYMMETRY 547

better describe the apparently four-dimensional physical world we observe. The


geometry of such a compactification is simply a Carteasian product R3,1 × M 7 ,
where R3,1 is the Minkowski space-time (or some other Lorenzian 4-manifold) and
M 7 is a compact manifold with so small a radius that its presence can be only
felt and observed at the quantum level. Many various models for M 7 were studied
in the late seventies which by the eighties had already accrued into a vast physics
literature (cf. the extensive three-volume monograph by Castellani, D’Auria and
Fré [CDF91]). Most of the models assumed a homogeneous space structure on
M 7 = G/H (see Chapter V.6 in [CDF91], for examples). Two things were of key
importance in terms of the required physical properties of the compactified theory.
First, the compact space M 7 , as a Riemannian manifold, had to be Einstein of pos-
itive scalar curvature. Second, although one could consider any compact Einstein
space for the compactification, the new theory would no longer be supersymmetric
unless (M 7 , g) admitted Killing spinor fields, and the number of them would be
exactly the number of residual supersymmetries of the compactified theory. For
that reason compactification models involving (S 7 , g0 ) were quite special as they
gave the maximally supersymmetric model. However, early on it was realized that
there are other, even homogeneous, 7-manifolds of interest. The Sp(2)-invariant
Jensen metric on S 7 , or as physicists correctly nicknamed it, the squashed 7-sphere
is one of the examples. Indeed, Jensen’s metric admits exactly one Killing spinor
field since it has weak holonomy G2 . Of course, any of the Einstein geometries in
the table of Theorem 14.2.3 can be used to obtain such supersymmetric models.
The D = 11 supergravity theory only briefly looked liked it was the Grand
Theory of Einstein’s dream. It was soon realized that there are difficulties with
getting from D = 11 supergravity to the standard model. The theory which was
to solve these and other problems was Superstring Theory and later M-Theory
(which is yet to be constructed). With the arrival of superstring theory and M-
theory, supersymmetry continues its truly remarkable influence on many different
areas of mathematics and physics: from geometry to analysis and number theory.
For instance, once again five, six, and seven-dimensional manifolds admitting real
Killing spinors have become of interest because of the so called AdS/CFT Dual-
ity. Such manifolds have emerged naturally in the context of p-brane solutions
in superstring theory. These so-called p-branes, “near the horizon” are modelled
by the pseudo-Riemannian geometry of the product adSp+2 × M , where adSp+2
is the (p + 2)-dimensional anti-de-Sitter space (a Lorentzian version of a space of
constant sectional curvature) and (M, g) is a Riemannian manifold of dimension
d = D − p − 2. Here D is the dimension of the original supersymmetric theory. In
the most interesting cases of M2-branes, M5-branes, and D3-branes D equals either
11 (Mp-branes of M-theory) or 10 (Dp-branes in type IIA or type IIB string theory).
String theorists are particularly interested in those vacua of the form adSp+2 × M
that preserve some residual supersymmetry. It turns out that this requirement im-
poses constraints on the geometry of the Einstein manifold M which is forced to
admit real Killing spinors. Depending on the dimension d, the possible geometries
of M are as follows:
where the notation (µ, µ̄), which is common in the physics literature, represents the
ratio of the number of real Killing spinors of type (p, q) to the maximal number
of real Killing spinors that can occur in the given dimension. This maximum is,
of course, realized by the round sphere of that dimension. So this table is just a
548 14. SASAKIAN STRUCTURES, KILLING SPINORS, AND SUPERSYMMETRY

d Geometry of M (µ, µ̄)


any round sphere (1, 1)
7 weak holonomy G2 ( 18 , 0)
Sasaki–Einstein ( 14 , 0)
3-Sasakian ( 38 , 0)
6 nearly Kähler ( 18 , 18 )
5 Sasaki–Einstein ( 14 , 14 )

translation of the table of Theorem 14.2.3 for the special dimensions that occur in
the models used by the physicists.
Furthermore, given a p-brane solution of the above type, the interpolation be-
tween adSp+2 × M and Rp,1 × C(M ) leads to a conjectured duality between the
supersymmetric background of the form adSp+2 × M and a (p + 1)-dimensional
superconformal field theory of n coincident p-branes located at the conical singu-
larity of the Rp,1 × C(M ) vacuum. This is a generalized version of the Maldacena
or AdS/CFT Conjecture [Mal99]. In the case of D3-branes of string theory the
relevant near horizon geometry is that of adS5 ×M , where M is a Sasaki-Einstein 5-
manifold. The D3-brane solution interpolates between adS5 × M and R3,1 × C(M ),
where the cone C(M ) is a Calabi-Yau threefold. In its original version the Malda-
cena conjecture (also known as AdS/CFT duality) states that the ’t Hooft large n
limit of N = 4 supersymmetric Yang-Mills theory with gauge group SU (n) is dual
to type IIB superstring theory on adS5 × S 5 [Mal99]. This conjecture was further
examined by Klebanov and Witten [KW99] for the type IIB theory on adS5 ×T 1,1 ,
where T 1,1 is the other homogeneous Sasaki-Einstein 5-manifold T 1,1 = S 2 × S 3
and the Calabi-Yau 3-fold C(T 1,1 ) is simply the quadric cone in C4 . Using the well-
known fact that C(T 1,1 ) is a Kähler quotient of C4 (or, equivalently, that S 2 × S 3
is a Sasaki-Einstein quotient of S 7 ), a dual super Yang-Mills theory was proposed,
representing D3-branes at the conical singularities. In the framework of D3-branes
and the AdS/CFT duality the question of what are all the possible near horizon
geometries M and C(M ) might be of importance. Much of the interest in Sasaki-
Einstein manifolds is precisely due to the fact that each such explicit metric, among
other things, provides a useful model to test the AdS/CFT duality. That is why the
Yp,q manifolds of Section 11.4.2 attracted such attention (cf. the papers mentioned
in Remarks 11.4.1).
Remark 14.5.1: [G2 holonomy manifolds unification scale and proton de-
cay] Until quite recently the interest in 7-manifolds with G2 holonomy as a source
of possible physical models was tempered by the fact the Kaluza-Klein compact-
ifications on smooth and complete manifolds of this type lead to models with no
charged particles. All this has dramatically changed in the last few years largely be-
cause of some new developments in M-theory. Perhaps the most compelling reasons
for reconsidering such 7-manifolds was offered by Atiyah and Witten who consid-
ered the dynamics on manifolds with G2 holonomy which are asymptotically conical
[AW02]. The three models of cones on the homogeneous nearly Kähler manifolds
mentioned earlier are of particular interest, but Atiyah and Witten consider other
cases which include orbifold (quotient) singularities. Among other things they point
14.5. GEOMETRIC STRUCTURES ON MANIFOLDS AND SUPERSYMMETRY 549

to a very interesting connection between Kronheimer’s quotient construction of the


ALE metrics and asymptotically conical manifolds with G2 -holonomy. To explain
the connection, consider Kronheimer’s construction for Γ = Zn+1 described in Sec-
1
tion 12.10. Suppose one chooses a circle Sk,l ' U(1) ∈ K(Zn+1 ) = U(1)n and then
one considers a 7-manifold obtained by performing Kronheimer’s HK quotient con-
struction with zero momentum level (ξ = 0) while ”forgetting” the three moment
map equations corresponding to this particular circle. An equivalent way of look-
ing at this situation is to take the Kronheimer quotient with nonzero momentum
1
ξ = a ∈ sp(1) but only for the moment map of the chosen circle Sk,l (such ξ is never
in the ”good set”) and then consider the fibration of singular Kronheimer quotients
over a 3-dimensional base parameter space. Algebraically this corresponds to a
partial resolution of the quotient singularity and this resolution depends on the
1
choice of Sk,l , hence ξ. This example was first introduced in [AW02]. It can be
shown that the 7-manifold is actually a cone on the complex weighted projective
3-space with weights (k, k, l, l), where k +l = n+1. It then follows from the physical
model considered that such a cone should admit a metric with G2 holonomy. How-
ever, unlike the homogeneous cones over the four homogeneous strict nearly Kähler
manifolds of Theorem 14.3.12, the metric in this case is not known explicitly. This
construction appears to differ from all previous geometric constructions of metrics
with G2 holonomy. One can consider similar constructions for other choices of
S 1 ⊂ K(Γ) [BB02].
In [FW03] using a specific models of M-theory compactifications on mani-
folds with G2 holonomy, Friedman and Witten address the fundamental questions
concerning the unification scale (i.e., the scale at which the Standard Model of
SU 3) × SU (2) × U (1) unifies in a single gauge group) and proton decay. The au-
thors point out that the results obtained are model dependent, but some of the
calculations and conclusions apply to a variety of different models.
Appendix A

A.1. Preliminaries on Groupoids

We briefly review some properties of groupoids [CdSW99, Mac87, Moe02,


MM03, MS88, ALR06]. Here we follow [Moe02] closely.
Definition A.1.1: A groupoid G is a small category in which every morphism
has an inverse.
So a groupoid G consists of a set G0 of objects and a set G1 of arrows or mor-
phisms together with certain structure maps. For any groupoid there are canonical
maps s : G1 −−→G0 and t : G1 −−→G0 , called the source map and target map (some-
times the latter is called the range map), respectively, and defined by sending each
arrow g : x−−→y to its source x and target y, respectively, i.e., s(g) = x, t(g) = y.
Next there is an associative multiplication, called multiplication or composition,
defined for any two arrows g, h ∈ G1 such that t(g) = s(h) with s(h ◦ g) = s(g) and
t(h ◦ g) = t(h). Alternatively, if g : x−−→y and h : y−−→z are arrows from x to y,
and from y to z, respectively, then h ◦ g : x−−→z is an arrow from x to z. Since,
composition is defined only when the target of the first arrow equals the source of
the second, we consider the fibered product G1 ×G0 G1 defined by

(A.1.1) G1 ×G0 G1 = {(h, g) ∈ G1 × G1 | s(h) = t(g)} .

Then composition defines a map m : G1 ×G0 G1 −−→G1 by m(h, g) = h ◦ g. For each


object x ∈ G0 there is an unit or identity arrow 1x ∈ G1 defined by sending x to
itself, i.e., 1x x = x. The identity is a two-sided unit for composition, i.e., g ◦ 1x = g
and 1x ◦ h = h for any g, h ∈ G1 with s(g) = x and t(h) = x. The unit arrows define
a map u : G0 −−→G1 by u(x) = 1x .
Finally for each arrow g : x−−→y in G1 there is an inverse arrow g −1 : y−−→x
which is a two-sided inverse for composition, i.e., g ◦ g −1 = 1y and g −1 ◦ g = 1x .
The inverse defines an involution ι : G1 −−→G1 by ι(g) = g −1 .
Thus, a groupoid G = (G0 , G1 , s, t, m, u, ι) consists of two sets G0 and G1
together with five structure maps s, t, m, u, ι. It is often convenient to think of G0
as a “base space” and refer to G as a groupoid over G0 , written as G1 =⇒ G0 , where
the double arrow indicates the two maps, the source and target. The composition
g ◦ h in G is often written as juxtaposition gh.
For any groupoid G we have the following subsets of interest:
(A.1.2)

Gx = {g ∈ G1 | s(g) = x}, G y = {g ∈ G1 | t(g) = y} ,


Gxy = {g ∈ G1 | s(g) = x, r(g) = y}, GZW = {g ∈ G1 | s(g) ∈ Z, t(g) ∈ W } .

551
552 APPENDIX A

Here Z and W are subsets of G0 . Notice from this point of view that the composition
map m becomes a map on the subsets Gxy × Gyz −−→Gxz and that the subset Gxx ⊂ G1
is a group which acts on Gxy from the left. Morever, the group Gyy acts on Gxy from
the right. The group Gxx of arrows from x to itself is called the isotropy group at x
and often denoted by Gx .
Definition A.1.2: A subgroupoid H of G is defined by a pair of subsets H0 ⊂
G0 and H1 ⊂ G1 such that s(H1 ) ⊂ H0 , t(H1 ) ⊂ H0 , and H1 is closed under
multiplication when it is defined and under inversion. If H0 = G0 then H is called
a wide subgroupoid. If for all x, y ∈ H0 , Hxy = Gxy then H is said to be a full
subgroupoid.
Notice that the fact that H is closed under inversion implies that u(x) = 1x is
in H1 for all x ∈ H0 , i.e., the restriction of u to H0 defines a map from H0 to H1 .
Next we give some simple examples of groupoids.
Example A.1.3: Some simple examples of groupoids are:
(i) The identity or trivial or unit groupoid: All arrows are units, so G1 = G0
as well. Any subset of this groupoid is a subgroupoid.
(ii) A group G is a groupoid with G0 = {pt}. The subgroupoids of G are just
the subgroups of G and the empty set.
(iii) The pair groupoid P : G0 = X is any set, and G1 = X × X with exactly
one arrow between any two points x, y ∈ G0 . Then a subgroupoid of P is
a relation on X which is symmetric and transitive. A wide subgroupoid
is an equivalence relation on X.
The groupoids that we are interested in have a richer structure. In fact they
are smooth manifolds. This leads to:
Definition A.1.4: A topological groupoid is a groupoid G in which both G1 and
G0 are topological spaces, and the structure maps are continuous. A Lie groupoid
is a topological groupoid for which G1 is a smooth, but not necessarily Hausdorff,
manifold, G0 is a smooth Hausdorff manifold, the above structure maps are differ-
entiable, and both the source and target maps are submersions.
For a complete treatment of Lie groupoids, see [Mac87]. In almost all cases
of interest to us, the groupoid G will be a Lie groupoid, and G1 will be Hausdorff.
By abuse of notation we sometimes refer to the manifold G1 as the groupoid G
and by the dimension of G we mean the dimension of G1 . For any Lie groupoid G
the subsets defined by A.1.2 above are smooth manifolds, and Gx = Gxx is a Lie
group. Lie groupoids exist in abundance. The simple examples in Example A.1.3
have their Lie groupoid analogues.
(i) G0 = M is any manifold and all arrows are units, so G1 = M as well. G
is called the unit Lie groupoid.
(ii) A Lie group is a Lie groupoid with G0 = {pt}.
(iii) G0 = M is any manifold, and G1 = M × M with exactly one arrow
between any two points x, y ∈ G0 . This is called the pair Lie groupoid
and sometimes it is denoted by M × M.
There are two other Lie groupoids that are of much interest to us. The first
is the holonomy groupoid G(M, F) of a foliation F that is discussed in detail in
Chapter 1. The second is the so-called action or translation groupoid.
A.1. PRELIMINARIES ON GROUPOIDS 553

Definition A.1.5: Let K be a Lie group acting smoothly on the manifold M from
the right3. We define a Lie groupoid M o K called the action or translation
groupoid as follows: the space of objects (M o K)0 is M and the space of arrows
(M o K)1 is M × K. An arrow (x, k) ∈ (M × / K)1 = M × K from x ∈ M to
y ∈ M is given by the right action y = xk. The source map is s : M × K−−→M is
projection onto the first factor, while the target map t : M × K−−→M is defined by
t(x, k) = xk. Composition is defined by (x, k) · (y, k 0 ) = (x, kk 0 ) when y = xk. The
unit is u(x) = (x, e), where e is the identity in K.
Next we give several definitions that are pertinent to understanding orbifolds
from the groupoid point of view [ALR06, Moe02].
Definition A.1.6: Let G be a Lie (topological) groupoid.
(i) G is called an étale groupoid if the source map s is a local diffeomor-
phism (homeomorphism), and it is said to be effective if the isotropy
groups Gxx act effectively on G0 ;
(ii) G is said to be proper if G1 is Hausdorff and the achor map (s, t) :
G1 −−→G0 × G0 is proper;
(iii) G is called a foliation groupoid if every isotropy group Gxx is discrete;
(iv) an orbifold groupoid is a proper étale Lie groupoid.
Note that the fact that the source map s is a local diffeomorphism (home-
omorphism) implies that the other structure maps are diffeomorphisms (homeo-
morphisms). These definitions are enough to capture the essence of an orbifold as
seen in Section 4.3. (For our orbifolds we should add that the isotropy groups act
effectively).
Following [Moe02] we now describe the important notion of ‘equivalence’ of
groupoids. Let G and H be groupoids. A homomorphism φ : G−−→H of groupoids is
a pair φ0 : G0 −−→H0 and φ1 : G1 −−→H1 of maps that commute with all the structure
maps, e.g., φ0 ◦s = s◦φ1 , etc. If G and H are Lie (topological) groupoids, we demand
that the maps are smooth (continuous). In particular, φ is an isomorphism of Lie
(topological) groupoids if the maps φ0 , φ1 are diffeomorphisms (homeomorphisms).
We call such groupoids strongly equivalent. For groupoids it is important to have a
weaker notion of equivalence. First we need
Definition A.1.7: Let φ : H−−−→G and ψ : K−−−→G be homomorphisms of Lie
(topological) groupoids. The fibered product H ×G K is the Lie (topological)
groupoid whose objects are triples (y, g, z), where y ∈ H0 , z ∈ K0 , and g : φ(y)−−→ψ(z)
is in G1 . The morphisms (y, g, z)−−−→(y 0 , g 0 , z 0 ) of H ×G K are pairs (h, k) of mor-
phisms h : y → y 0 in H1 and k : z → z 0 in K1 such that g 0 φ(h) = ψ(h)g. This can
be represented by the following diagram:
g
y φ(y) −−−−→ φ(z) z
   
   
   
h φ(h) φ(k) k
   
y y y y
g0
y0 φ(y 0 ) −−−−→ φ(z 0 ) z0
Composition in H ×G K is defined by extending he diagram in the obvious way.

3It is more common to use the left action, but we choose the action to be from the right in
order to follow the standard convention for principal bundles.
554 APPENDIX A

Definition A.1.8: A homomorphism of groupoids φ : G−−→H is called an essen-


tial or weak equivalence if:
(i) the map t ◦ π1 : H1 ×H0 G0 −−→H0 is a surjective submersion.
(ii) the square
φ
G1 −−−−→ H1
 
 
 
(s,t) (s,t)
 
y y
φ×φ
G0 × G0 −−−−→ H0 × H0
is a fibered product.
Two such groupoids are said to be Morita equivalent if there exists a third
groupoid K such that G ← K → H are weak equivalences.
Condition (i) means that every object y ∈ H0 is connected by a morphism
(h : φ(x)−−→y) ∈ H1 to an object in the image of φ. Condition (ii) implies that
φ induces a diffeomorphism G1 (x, y)−−→H1 (φ(x), φ(y)) from the subspace of all
morphisms from x to y in G1 to the subspace of all morphisms φ(x) to φ(y) in H1 .
One can check that Morita equivalence defines an equivalence relation. Although
Morita equivalent Lie groupoids can be very different, for example they can have
different dimensions and different connectivity properties, Morita equivalence does
preserve some essential geometric features as we shall see. For us the relationship
between étale Lie groupoids and orbifolds is crucial.
Next we describe the concept of an action of a topological groupoid on an
appropriate space [Mac87, CdSW99].
Definition A.1.9: Let G be a topological groupoid over X ' G0 , and let µ : Y −−→X
be a continuous map. Define the space
G1 ×X Y = {(g, y) ∈ G1 × Y | s(g) = µ(y)} .
Then the action of G on Y is defined by a map A : G1 ×X Y −−→Y, satisfying
(i) µ(A(g, y)) = t(g),
(ii) A(g ◦ h, y) = A(g, A(h, y)),
(iii) A(u(µ(y)), y) = y.
The orbit of G through y ∈ Y is the subset G(y) = {A(g, y) | g ∈ Gµ(y) } ⊂ Y.
The map µ is often called the moment map in analogy with the moment map of
symplectic geometry. The action map A(g, y), thought of as a left action, is often
written as g · y. There are two natural actions of a groupoid on itself. The first
is an action of G on its set of arrows, G1 , so we take Y = G1 , µ = t, and define
the action by A : G1 ×X G1 −−→G1 as A(g, h) = g ◦ h which is well defined since
s(g) = t(h). The second is an action on the set of objects, so we take Y = G0 and
µ = id. The action map is now defined by A(g, x) = t(g) which makes sense since
the source of g is x. Notice that the first action is always free, while the second
is generally not free. A right action is defined in a completely analogous way. An
important example in which the convention is to use the right action is [Hae84]:
Definition A.1.10: A principal G-bundle over Y is a topological space E to-
gether with a continuous surjection π : E−−→Y , and a continuous action A of G
on E with moment map µ : E−−→X such that π(A(g, e)) = π(e) for e ∈ E. The
A.2. THE CLASSIFYING SPACE OF A TOPOLOGICAL GROUPOID 555

action A is transitive on the fibres of π. Furthermore, each point y ∈ Y has an


open neighborhood U and a continuous section σ : U −−→E of π such that the “lo-
cal triviality” condition holds, that is the map h : G1 ×X U −−→π −1 (U ) defined by
h(g, y) = A(g, σ(y)) is a homeomorphism.

A.2. The Classifying Space of a Topological Groupoid


Classifying spaces are important in mathematics as spaces which capture all
objects with a particular behavior. In particular, we are interested in constructing
universal spaces for topological groupoids. Classifying spaces were first constructed
for topological groups. Good references here are the book of Husemoller [Hus66],
and the papers of Milnor [Mil56a] and Dold [Dol63]. The classifying space theory
for groupoids was developed by Haefliger [Hae62, Hae71] and independently by
Buffet and Lor [BL70]. In [Hae62] Haefliger approaches the subject through
E. Brown’s theory of representable homotopy functors, whereas, in [Hae71] and
[BL70] the authors generalize Milnor’s beautiful join construction of classifying
space for topological groups [Mil56a] to groupoids. It is this latter approach that
we present here. Another suitable approach to classifying spaces for groupoids, or
more generally to categories is Segal’s construction [Seg68, Moe02, ALR06] of
the nerve of a category. Segal shows this to be equivalent to the Milnor construction.
For any topological groupoid G we consider the superdiagonal
(A.2.1) SDn+1 (G) = {(x0 , . . . , xn ) ∈ Gn+1
1 | t(xi ) = t(xj ) for all i, j = 0, . . . n}
together with the standard Pn n-simplex ∆n with barycentric coordinates (t0 , . . . , tn ),
i.e., ti ∈ [0, 1] and i=0 ti = 1. Write a point of SDn+1 (G) × ∆n as a 2n + 2-tuple
(t0 x0 , t1 x1 , . . . , tn xn ). We denote the n + 1-fold join by
n+1 times
(A.2.2) EG(n) = G ? ··· ?G.
Then as a set EG(n) is the identification space SDn+1 (G) × ∆n / ∼, where
(t0 x0 , t1 x1 , . . . , tn xn ) ∼ (t00 x00 , t01 x01 , . . . , t0n x0n )
if ti = t0i for all i and if ti > 0 then xi = x0i . A point of EG(n) is denoted by
ht, xi = ht0 x0 , . . . , tn xn i. On EG(n) there is a natural map which by abuse of
notation is also denoted by ti : EG(n)−−→[0, 1]. It is defined to be the projection
map
(A.2.3) ht0 x0 , . . . , tn xn i 7→ ti .
On each inverse image t−1
i ((0, 1]) there are also natural projection maps xi defined
by
(A.2.4) ht0 x0 , . . . , tn xn i 7→ xi .
Notice that if any of the coordinates ti vanish then xi is not defined as a map. Now
the topology on EG(n) is defined to be the smallest topology such that the maps
ti and xi are continous. There are also inclusion maps ι : EG(n) ,→ EG(n + 1)
defined by
(A.2.5) ι(ht0 x0 , . . . , tn xn i) = ht0 x0 , . . . , tn xn , 0xn+1 i .
Notice that this is independent of the choice of xn+1 . The inclusions give rise to a
directed system and the direct limit
(A.2.6) EG = lim EG(n)
−→
556 APPENDIX A

with the direct limit topology. Notice also that we have not used all of the properties
of a topological groupoid. Indeed, we have only used the fact that G consists of
topological spaces G1 and G0 together with a continuous map t : G1 −−→G0 .
Now we define an action of the topological groupoid G on EG. The moment
map µ : EG(n)−−→G0 is defined by µ(ht, xi) = t(xi ) which is a well defined point
of G0 , since EG(n) is formed from the superdiagonal SDn+1 (G). The action A on
EG(n) is that induced by the left action of G on G1 , namely,

A(g, ht, xi) = ht, gxi .

One easily checks that this satisfies the conditions of Definition A.1.9. The inclusion
maps ι are equivariant with respect to this action giving an action on EG. It is also
clear since not all ti can vanish that this action is free on both EG(n) and EG.
The quotient spaces are denoted by BG(n) and BG, respectively, and the natural
projections πn : EG(n)−−→BG(n) and π : EG−−→BG are principal G-bundles. In
analogy with the case of principal bundles, the space BG is called the classifying
space of principal G-structures which we now describe.
Definition A.2.1: Let B be a topological space and {Ui }i∈I an open cover of B.
Then the cover {Ui } is said to be numerable if there exists a locally finite partition
of unity {ui }i∈I such that the closure of u−1
i ((0, 1]) is contained in Ui for each i ∈ I.
A principal G-bundle π : E−−→B is said to be numerable if B has a numerable
cover {Ui }i∈I such that π −1 (Ui ) is locally trivial for each i ∈ I. A numerable
principal G-bundle π : E−−→B is n-universal if for each CW -complex X with
dim ≤ n the pullback operation E 7→ f ∗ E induces a one-to-one correspondence
between homotopy classes of maps f : X−−→B and isomorphism classes of principal
G-bundles over X. In particular, E is called universal if it is n-universal for all n.
It is a standard result that a Hausdorff space X is paracompact if and only if
every open cover is numerable. Moreover, a result of Miyazaki says that any CW -
complex is paracompact. So for any CW -complex every open cover is numerable.
It is easy to see that under any continuous map numerable bundles pullback to
numerable bundles. Moreover, pulling back numerable bundles by homotopic maps
gives isomorphic bundles. We now have the generalization to groupoids of Milnor’s
well-known result. It is due to Buffet-Lor [BL70] and Haefliger [Hae71].
Theorem A.2.2: The bundle π : EG−−→BG is a numerable universal principal
G-bundle.
So BG classifies principal G-bundles as well as Haefliger cocycles, but it is only
the Morita equivalence class of G that is relevant.
Theorem A.2.3: Let G and H be Morita equivalent topological groupoids, then
there classifying spaces BG and BH are weakly homotopy equivalent.
When G is a connected topological group G, we recover the classical universal
principal G-bundle of Milnor [Mil56a] EG−−−→BG. See also [Ste51] and Appen-
dix B of [LM89]. In particular, if G is a Lie group it suffices to consider compact
Lie groups since any Lie groups is homotopy equivalent to its maximal compact
Lie group by the Iwasawa decomposition. Then isomorphism classes of principal
G bundles on a smooth manifold M are in one-to-one correspondence with ho-
motopy classes [M, BG] of maps c : M −−−→BG. Elements of the cohomology ring
H ∗ (BG, A) for a ring A are called universal characteristic classes. One obtains
A.2. THE CLASSIFYING SPACE OF A TOPOLOGICAL GROUPOID 557

characteristic classes on M by pulling back the universal characteristic classes by


a classifying map c. We have the following important examples:
(i) The (universal) Stiefel-Whitney classes wi are the generators of the coho-
mology ring H ∗ (BO(n), Z2 ), i.e., H ∗ (BO(n), Z2 ) = Z2 [w1 , w2 , . . . , wn ].
(ii) H ∗ (BSO(n), Z2 ) = Z2 [w2 , . . . , wn ].
(iii) H ∗ (BU (n), Z) = Z[c1 , . . . , cn ], where ci ∈ H 2i (BU (n), Z) is the ith Chern
class.
(iv) H ∗ (BSU (n), Z) = Z[c2 , . . . , cn ].
(v) H ∗ (BO(n), A) = A[p1 , . . . , pb n2 c ], where A is an integral domain contain-
ing 21 such as Z[ 12 ] or Q. When A = Q, pi ∈ H 4i (BO(n), A) is called the
ith rational Pontrjagin class.
(vi) H ∗ (BO(n), Z) = Z[p1 , . . . , pb n2 c ] ⊕ Image(β), where β is the Bockstein
homomorphism β : H i (BO(n), Z2 )−−→H i+1 (BO(n), Z) induced by the
2
exact sequence 0−−−→Z −−−→ Z−−−→Z2 −−−→0; here pi ∈ H 4i (BO(n), Z) is
the ith integral Pontrjagin class.
(vii) H ∗ (BSO(2n + 1), A) = A[p1 , . . . , pn ], where A is an integral domain
containing 12 .
(viii) H ∗ (BSO(2n), A) = A[p1 , . . . , pn , e]/(e2 − pn ), where A is an integral
domain containing 12 ; here e ∈ H 2n (BSO(2n), A) is called the Euler
class.
(ix) H ∗ (BSp(n), Z) = Z[s1 , . . . , sn ], where si ∈ H 4i (BSp(n), Z) is the image
of c2i under the natural map BSp(n)−−→BU (2n).
The corresponding characteristic classes of a vector bundle E on a manifold M are
obtained by pulling back these universal classes by the classifying map c. When
E = T M, the tangent bundle of M , the characteristic classes are invariants of the
manifold with certain additional structure and we write wi (M ), ci (M ), etc. For
a real vector bundle E it is common to define the integral Pontrjagin classes by
pi (E) = (−1)i c2i (E ⊗ C). Also of some interest are the integral Stiefel Whitney
classes Wi defined to be the image of the ordinary Stiefel Whitney classes under
the Bockstein homomorphism, that is Wi = β(wi−1 ). Since the kernel of β is the
set of classes in H ∗ (BO(n), Z2 ) that are the mod 2 reductions of integral classes,
Wi vanishes if and only if wi−1 is the mod 2 reduction of an integral class.
Appendix B

B.1. Reid’s List of K3 Surfaces as hypersurfaces in CP4 (w)

b2 (X) X
22 X4 ⊂ P(1, 1, 1, 1), X6 ⊂ P(1, 1, 1, 3)
21 X5 ⊂ P(1, 1, 1, 2), X12 ⊂ P(1, 1, 4, 6)
20 X8 ⊂ P(1, 1, 2, 4), X10 ⊂ P(1, 1, 3, 5)
19 X6 ⊂ P(1, 1, 2, 2), X7 ⊂ P(1, 1, 2, 3), X9 ⊂ P(1, 1, 3, 4)
17 X10 ⊂ P(1, 2, 2, 5), X18 ⊂ P(1, 2, 6, 9)
16 X8 ⊂ P(1, 2, 2, 3), X12 ⊂ P(1, 2, 3, 6), X14 ⊂ P(1, 2, 4, 7),
X16 ⊂ P(1, 2, 5, 8)
15 X9 ⊂ P(1, 2, 3, 3), X10 ⊂ P(1, 2, 3, 4), X11 ⊂ P(1, 2, 3, 5),
X12 ⊂ P(1, 2, 4, 5), X15 ⊂ P(1, 2, 5, 7), X24 ⊂ P(1, 3, 8, 12)
14 X16 ⊂ P(1, 3, 4, 8), X18 ⊂ P(1, 3, 5, 9), X22 ⊂ P(1, 3, 7, 11)
X30 ⊂ P(1, 4, 10, 15)
X12 ⊂ P(1, 3, 4, 4), X13 ⊂ P(1, 3, 4, 5), X14 ⊂ P(2, 2, 3, 7)
13 X15 ⊂ P(1, 3, 4, 7), X15 ⊂ P(1, 3, 5, 6), X20 ⊂ P(1, 4, 5, 10)
X21 ⊂ P(1, 3, 7, 10), X22 ⊂ P(1, 4, 6, 11), X28 ⊂ P(1, 4, 9, 14)
X36 ⊂ P(1, 5, 12, 18), X42 ⊂ P(1, 6, 14, 21)
12 X12 ⊂ P(2, 2, 3, 5), X16 ⊂ P(1, 4, 5, 6), X18 ⊂ P(1, 4, 6, 7)
X26 ⊂ P(1, 5, 7, 13), X30 ⊂ P(1, 6, 8, 15),
11 X12 ⊂ P(2, 3, 3, 4), X18 ⊂ P(2, 3, 4, 9), X21 ⊂ P(1, 5, 7, 8),
X24 ⊂ P(1, 6, 8, 9), X30 ⊂ P(2, 3, 10, 15),
10 X14 ⊂ P(2, 3, 4, 5), X16 ⊂ P(2, 3, 4, 7), X20 ⊂ P(2, 3, 5, 10)
X22 ⊂ P(2, 4, 5, 11), X24 ⊂ P(2, 3, 7, 12), X26 ⊂ P(2, 3, 8, 13),
X15 ⊂ P(2, 3, 5, 5), X15 ⊂ P(3, 3, 4, 5), X17 ⊂ P(2, 3, 5, 7),
9 X18 ⊂ P(2, 3, 5, 8), X20 ⊂ P(2, 4, 5, 9), X20 ⊂ P(2, 4, 5, 9)
X21 ⊂ P(2, 3, 7, 9), X24 ⊂ P(2, 3, 8, 11), X26 ⊂ P(2, 5, 6, 13)
8 X18 ⊂ P(3, 4, 5, 6), X20 ⊂ P(2, 5, 6, 7), X24 ⊂ P(3, 4, 5, 12)
X32 ⊂ P(2, 5, 9, 16), X42 ⊂ P(3, 4, 14, 21),
X19 ⊂ P(3, 4, 5, 7), X20 ⊂ P(3, 4, 5, 8), X21 ⊂ P(3, 5, 6, 7)
7 X27 ⊂ P(2, 5, 9, 11), X28 ⊂ P(3, 4, 7, 14), X30 ⊂ P(4, 5, 6, 15)
X34 ⊂ P(3, 4, 10, 17), X36 ⊂ P(3, 4, 11, 18), X48 ⊂ P(3, 5, 16, 24)
X24 ⊂ P(3, 4, 7, 10), X24 ⊂ P(4, 5, 6, 9), X30 ⊂ P(3, 4, 10, 13)
6 X32 ⊂ P(4, 5, 7, 16), X34 ⊂ P(4, 6, 7, 15, 17), X38 ⊂ P(3, 5, 11, 19)
X54 ⊂ P(4, 5, 18, 27)
X25 ⊂ P(4, 5, 7, 9), X27 ⊂ P(5, 6, 7, 9), X28 ⊂ P(4, 6, 7, 11)
5 X33 ⊂ P(3, 5, 11, 14), X38 ⊂ P(5, 6, 8, 19), X40 ⊂ P(5, 7, 8, 20)
X42 ⊂ P(2, 5, 14, 21), X44 ⊂ P(4, 5, 13, 22), X66 ⊂ P(5, 6, 22, 33)
4 X24 ⊂ P(3, 6, 7, 8), X30 ⊂ P(5, 6, 8, 11), X36 ⊂ P(7, 8, 9, 12)
X50 ⊂ P(7, 8, 10, 25),

559
560 APPENDIX B

B.2. Differential topology of 2k(S 3 × S 4 ) and 2k(S 5 × S 6 )

Table B.2.1: 2k#(S 3 × S 4 )


D2 (k)
k τk D2 (k) |bP8 |
1 1 28 1
2 3 28 1
1
3 6 14 2
1
4 10 14 2
5 15 28 1
1
6 21 4 7
1
7 28 1 28
1
8 36 7 4
9 45 28 1
10 55 28 1
1
20 210 2 14
1
48 1176 1 28
50 1275 28 1
1
100 5050 14 2
1
496 123256 1 28
1
500 125250 14 2

Table B.2.2: 2k#(S 5 × S 6 )


D3 (k)
k τk D3 (k) |bP12 |
1 1 992 1
2 3 992 1
1
3 6 496 2
1
4 10 496 2
5 15 992 1
6 21 992 1
1
7 28 248 4
1
8 36 248 4
9 45 992 1
10 55 992 1
1
31 496 2 496
1
48 1176 124 8
50 1275 992 1
1
62 1953 32 31
1
124 7750 16 62
1
248 30876 8 124
1
496 123256 4 248
1
500 125250 496 2
1
992 492528 2 496
B.3. TABLES OF KÄHLER-EINSTEIN METRICS ON HYPERSURFACES 561

B.3. Tables of Kähler-Einstein metrics on hypersurfaces CP(w)


Table B.3.1 lists the infinite series solutions appearing in Theorem 5.4.16, while
Table B.3.2 lists the sporadic ones. The last column of the tables indicate whether
the klt condition (which implies that Zw admits a Kähler-Einstein metric) holds
(Y) or is unknown (?). The computer search gives the following complete list:

Table B.3.1: Infinite Series Examples of Zw of Index 1 ≤ I ≤ 10


I w d b2 K-E
1 (2, 2k + 1, 2k + 1, 4k + 1) 8k + 4 8 Y
2 (4, 2k + 1, 2k + 1, 4k) 8k + 4 7 Y
2 (3, 3k + 1, 6k + 1, 9k + 3) 18k + 6 6 Y
2 (3, 3k + 1, 6k + 1, 9k) 18k + 3 5 ?
2 (3, 3k, 3k + 1, 3k + 1) 9k + 3 7 Y
2 (3, 3k + 1, 3k + 2, 3k + 2) 9k + 6 5 Y
2 (4, 2k + 1, 4k + 2, 6k + 1) 12k + 6 6 ?
4 (6, 6k + 3, 6k + 5, 6k + 5) 18k + 15 5 Y
4 (6, 6k + 5, 12k + 8, 18k + 9) 36k + 24 3 ?
4 (6, 6k + 5, 12k + 8, 18k + 15) 36k + 30 4 Y
6 (8, 4k + 1, 4k + 3, 4k + 5) 12k + 11 3 ?
6 (9, 3k + 2, 3k + 5, 6k + 1) 12k + 11 3 ?
562 APPENDIX B

Table B.3.2: Sporadic Examples of Zw of Index 1 ≤ I ≤ 10


I w Monomials of fw d b2 K-E
1 (1,2,3,5) z010 , z15 , z23 z1 , z32 , . . . (17)∗ 10 9 Y
1 (1,3,5,7) z015 , z15 , z23 , z32 z0 , . . . (19) 15 9 Y
1 (1,3,5,8) z0 , z1 z0 , z23 z0 , z32 , . . . (20)
16 5
16 10 Y
1 (2,3,5,9) z09 , z16 , z23 z1 , z32 , . . . (13) 18 7 Y
1 (3,3,5,5) g5 (z0 , z1 ), f3 (z2 , z3 ) 15 5 Y
1 (3,5,7,11) z06 z2 , z15 , z22 z3 , z32 z0 , . . . (8) 25 5 Y
1 (3,5,7,14) z07 z2 , z15 z0 , g2 (z22 , z3 ), . . . (9) 28 6 Y
1 (3,5,11,18) g2 (z06 , z3 ), z15 z2 , z23 z0 , . . . (10) 36 6 Y
1 (5,14,17,21) z07 z3 , z14 , z23 z0 , z32 z1 , z05 z1 z2 56 4 Y
1 (5,19,27,31) z010 z3 , z14 z0 , z23 , z32 z1 , z07 z1 z2 81 3 Y
1 (5,19,27,50) g2 (z010 , z3 ), z15 z0 , z23 z1 , z07 z12 z3 100 4 Y
1 (7,11,27,37) z010 z1 , z14 z3 , z23 , z33 z0 81 3 Y
1 (7,11,27,44) g2 (z14 , z3 ), z011 z1 , z23 z0 , z04 z13 z2 88 4 Y
1 (9,15,17,20) z05 z1 , z14 , z23 z0 , z33 60 3 Y
1 (9,15,23,23) z0 z1 , z1 z0 , z23 , z22 z3 , z2 z32 , z33
6 4
69 5 Y
1 (11,29,39,49) z08 z2 , z14 z0 , z23 , z32 z1 127 3 Y
1 (11, 49, 69, 128) z0 z2 , z15 z0 , z24 , z22 z3 , z32
17
256 2 Y
1 (13,23,35,57) z08 z1 , z14 z2 , z22 z3 , z32 z0 127 3 Y
1 (13, 35, 81, 128) z017 z1 , z15 z2 , z23 z0 , z32 256 2 Y
6 4 3 2
2 (2,3,4,5) z0 , z1 , z2 , z3 z0 , . . . (10) 12 5 ?
2 (2,3,4,7) z07 , z14 z0 , z23 z0 , z32 , . . . (11) 14 6 ?
2 (3,4,5,10) z05 z2 , z15 , z24 , z32 , . . . (9) 20 5 Y
2 (3,4,6,7) g3 (z0 , z2 ), z13 z2 , z32 z1 , . . . (8)
2
18 6 ?
2 (3,4,10,15) z010 , z15 z3 , z23 , z32 , . . . (10) 30 7 Y
2 (3,7,8,13) z0 z2 , z13 z2 , z22 z3 , z32 z0 , . . . (7)
7
29 5 ?
2 (3,10,11,19) z010 z3 , z13 z2 , z22 z3 , z32 z0 , . . . (7) 41 5 ?
2 (5,13,19,22) z07 z3 , z14 z0 , z23 , z32 z1 , z05 z1 z2 57 3 Y
2 (5,13,19,35) g2 (z07 , z3 ), z15 z0 , z23 z1 , z05 z12 z2 70 3 Y
2 (6,9,10,13) z06 , z14 , z23 z0 , z32 z2 , z03 z12 36 4 Y
2 (7,8,19,25) z0 z1 , z14 z3 , z23 , z32 z0 , z02 z13 z2
7
57 3 Y
2 (7,8,19,32) z08 z1 , g2 (z14 , z3 ), z23 z0 , . . . (7) 64 4 Y
2 (9,12,13,16) z04 z1 , z14 , z23 z0 , z33 48 3 Y
2 (9,12,19,19) z0 z1 , z1 z0 , z23 , z22 z3 , z2 z33 , z33
5 4
57 5 Y
2 (9,19,24,31) z09 , z13 z2 , z23 z0 , z32 z1 81 3 Y
2 (10,19,35,43) z07 z2 , z15 z0 , z23 , z32 z1 105 3 Y
2 (11,21,28,47) z07 z2 , z15 , z23 z1 , z32 z0 105 3 Y
2 (11,25,32,41) z0 z3 , z13 z2 , z23 z0 , z32 z1
6
107 3 Y
2 (11,25,34,43) z010 , z14 z0 , z22 z3 , z32 z1 111 3 Y
2 (11, 43, 61, 113) z015 z2 , z15 z0 , z23 z1 , z32 226 2 Y
2 (13,18,45,61) z09 z1 , z15 z2 , z23 , z32 z0 135 3 Y
B.3. TABLES OF KÄHLER-EINSTEIN METRICS ON HYPERSURFACES 563

Table B.3.2: (cont.) Sporadic Examples of Zw of Index 1 ≤ I ≤ 10


Index w Monomials of fw d b2 K-E
2 (13,20,29,47) z06 z3 , z13 z3 , z23 z1 , z32 z0 107 3 Y
2 (13,20,31,49) z07 z1 , z14 z2 , z23 z3 , z32 z0 111 3 Y
2 (13,31,71,113) z015 z1 , z15 z2 , z23 z0 , z32 226 2 Y
2 (14,17,29,41) z0 z3 , z15 z0 , z22 z3 , z32 z1
5
99 3 Y
3 (5,7,11,13) z0 z3 , z14 z0 , z23 , z32 z1 , z03 z1 z2
4
33 3 ?
3 (5,7,11,20) g2 (z04 , z3 ), z15 z0 , z23 z1 , z03 z12 40 4 Y
3 (11,21,29,37) z06 z2 , z14 z0 , z22 z3 , z32 z1 95 3 Y
3 (11,37,53,98) z013 z2 , z15 z0 , z23 z1 , z32 196 2 Y
3 (13,17,27,41) z0 z1 , z14 z2 , z22 z3 , z32 z0
6
95 3 Y
3 (13,27,61,98) z013 z1 , z15 z2 , z23 z0 , z32 196 2 Y
3 (15,19,43,74) z07 z2 , z17 z0 , z23 z1 , z32 148 2 Y
4 (5,6,8,9) z0 z3 , z14 , z23 , z32 z1 , z02 z1 z2
3
24 3 ?
4 (5,6,8,15) g2 (z03 , z3 ), z15 , z23 z1 , z02 z12 z2 30 4 ?
4 (9,11,12,17) z05 , z13 z2 , z23 z0 , z32 z1 45 3 ?
4 (10,13,25,31) z05 z2 , z15 z0 , z23 , z32 z1 75 3 Y
4 (11,17,20,27) z0 z3 , z13 z2 , z23 z0 , z32 z1
4
71 3 ?
4 (11,17,24,31) z05 z2 , z14 z0 , z22 z3 , z32 z1 79 3 Y
4 (11,31,45,83) z011 z2 , z15 z0 , z23 z1 , z32 166 2 Y
4 (13,14,19,29) z0 z2 , z13 z3 , z23 z1 , z32 z0
4
71 2 ?
4 (13,14,23,33) z05 z1 , z14 z2 , z22 z3 , z32 z0 79 3 Y
4 (13,23,51,83) z011 z1 , z15 z2 , z23 z0 , z32 166 2 Y
5 (11,13,19,25) z0 z2 , z14 z0 , z22 z3 , z32 z1
4
63 3 ?
5 (11,25,37,68) z09 z2 , z15 z0 , z23 z1 , z32 136 2 Y
5 (13,19,41,68) z09 z1 , z15 z2 , z23 z0 , z32 136 2 Y
6 (7,10,15,19) z05 z1 , z13 z2 , z23 , z32 z0 45 3 ?
6 (11,19,29,53) z07 z2 , z15 z0 , z24 , z32 106 2 Y
6 (13,15,31,53) z0 z1 , z15 z2 , z23 z0 , z32
7
106 2 Y
7 (11,13,21,38) z05 z2 , z15 z0 , z23 z1 , z32 76 2 Y
8 (7,11,13,23) z05 z1 , z13 z2 , z23 z0 , z32 46 2 ?
8 (7,18,27,37) z09 z1 , z13 z2 , z23 , z32 z0 81 3 ?
9 (7,15,19,32) z07 z1 , z13 z2 , z23 z0 , z32 64 2 ?
10 (7,19,25,41) z09 z1 , z13 z2 , z23 z0 , z32 82 2 ?
10 (7,26,39,55) z013 z1 , z13 z2 , z23 , z32 z0 117 3 ?

* (for lack of space only the total number of monomial terms in fw is indicated)

The computer program indicates that there are neither series solutions nor
sporadic solutions satisfying the hypothesis of Theorem 5.4.16 for I > 10. In fact,
an easy argument shows that there are no such solutions for sufficiently large I. 4

4The code for the C program used to generate the tables of the Theorem 5.4.16 are available
at: http://www.math.unm.edu/˜galicki/papers/publications.html.
564 APPENDIX B

B.4. Positive Breiskorn-Pham Links in Dimension 5

Table B.4.1: Positive BP Links with b2 (M ) = 0, H2 (L(a), Z) 6= 0


M L(a) NSF
M2 L(2, 3, 9, 15), L(2, 3, 3, k), k = 3(6) ∞
M3 L(2, 3, 8, 20), L(2, 3, 4, k), k = 4, 8(12) ∞
M4 L(3, 3, 3, 4), L(3, 3, 4, 9) 2
M5 L(3, 3, 5, 6), L(2, 3, 5, k), k = 6, 12, 18, 24(30), ∞
L(2, 4, 5, 8), L(2, 4, 5, 12), L(2, 4, 5, 16)
L(2, 4, 7, 8), L(2, 4, 6, 7), L(2, 3, 7, 36),
M7 L(2, 3, 7, 30), L(2, 3, 7, 24), L(2, 3, 7, 18), 10
L(2, 3, 7, 12), L(3, 3, 3, 7), L(2, 4, 4, 7),
L(2, 3, 6, 7)
M8 L(3, 3, 3, 8) 1
M9 L(2, 4, 4, 9) 1
M10 L(3, 3, 3, 10) 1
M11 L(2, 3, 11, 12), L(3, 3, 3, 11), L(2, 4, 4, 11), 5
L(2, 3, 6, 11), L(2, 4, 6, 11)
Mk , k > 12, k = 1, 5(6) L(2, 3, 6, k), L(3, 3, 3, k), L(2, 4, 4, k) 3
Mk , k > 12, k = 2, 4(6) L(3, 3, 3, k) 1
Mk , k > 12, k = 3(6) L(2, 4, 4, k) 1
2M3 L(2, 3, 5, k), k = 10, 20(30) ∞
3M3 L(2, 3, 8, 8), L(2, 3, 8, 16), L(3, 4, 4, 4), 5
L(2, 3, 7, 14), L(2, 3, 7, 28)
4M3 L(2, 3, 10, 10) 1
2M5 L(2, 5, 6, 6) 1
4M2 L(2, 3, 5, k), k = 15(30) ∞
6M2 L(2, 5, 5, 5), L(2, 3, 7, 21) 2
7M2 L(2, 3, 9, 9) 1
B.4. POSITIVE BP LINKS 565

Table B.4.2: Positive BP Links with b2 (M ) > 0, H2 (L(a), Z)tor 6= 0


M L(a) NSF
2M∞ #M2 L(2, 3, 4, k), k = 6(12) ∞
L(2, 4, 6, 9), L(2, 3, 6, 8)
2M∞ #M4 L(2, 3, 8, 18), L(2, 3, 6, 8) 2
2M∞ #M5 L(2, 3, 10, 12), L(2, 3, 6, 10) 2
2M∞ #Mk , (3, k) = 1, k > 6 L(2, 3, 6, 2k) 1
3M∞ #M3 L(2, 4, 6, 8), L(2, 4, 4, 6) 2
3M∞ #Mk , (2, k) = 1, k > 4 L(2, 4, 4, 2k) 1
4M∞ #M2 L(3, 3, 4, 6) 1
4M∞ #M3 L(2, 3, 9, 12), L(2, 3, 6, 9) 2
4M∞ #Mk , (2, k) = 1, k > 4 L(2, 3, 6, 3k) 1
4M∞ #2M2 L(2, 4, 5, 10) 1
5M∞ #2M2 L(2, 4, 6, 6) 1
6M∞ #M2 L(3, 3, 3, 6), L(2, 3, 8, 12) 2
6M∞ #Mk , k > 2 L(3, 3, 3, 3k) 1
7M∞ #Mk , k > 1 L(2, 4, 4, 4k) 1
8M∞ #Mk , k > 1 L(2, 3, 6, 6k) 1
566 APPENDIX B

Table B.4.3: Positive BP Links diffeomorphic to S 5


L(2, 2, p, q) (p, q) = 1
L(2, 3, 3, k) k = 1, 5(6)
L(2, 3, 4, k) k = 1, 2, 5, 7, 10, 11(12)
L(2, 3, 5, k) k 6= 0, 6, 10, 12, 15, 18, 20, 24(30)
L(2, 3, 7, k) 6 < k < 42, k 6= 12, 14, 18, 21, 24, 28, 30, 36
L(2, 3, 8, k) k = 10, 11, 13, 14, 17, 19, 22, 23
L(2, 3, 9, k) k = 11, 13, 17
L(2, 3, 10, k) k = 11, 13, 14
L(2, 3, 11, k) k = 11, 13
L(2, 4, 5, k) k = 6, 7, 9, 11, 13, 14, 17, 18, 19
L(2, 4, 6, k) k = 7, 11
L(2, 4, 7, k) k=9
L(2, 5, 5, k) k = 7, 9
L(2, 5, 6, k) k=7
L(3, 3, 4, k) k = 5, 7, 11
L(3, 3, 5, k) k=7
L(3, 4, 4, k) k=5

Table B.4.4: Positive BP Links with b2 (M ) > 0, H2 (L(a), Z)tor = 0


M L(a)
(n − 1)M∞ L(2, 2, p, q), (p, q) = n > 1
2M∞ L(2, 3, 3, k), k = 2, 4(6), L(2, 3, 4, k), k = 3, 9(12)
4M∞ L(2, 3, 3, k), k = 0(6)
6M∞ L(2, 3, 4, k), k = 0(12)
8M∞ L(2, 3, 5, k), k = 0(30)
M∞ L(2, 4, 6, 10)
2M∞ L(2, 3, 8, k), k = 9, 15, 21, L(2, 3, 9, k), k = 10, 14, 16
L(3, 3, 4, 10)
4M∞ L(2, 4, 5, k), k = 5, 15, L(2, 5, 5, k), k = 6, 8
6M∞ L(2, 4, 7, 7), L(3, 3, 4, k), k = 4, 8
8M∞ L(3, 3, 5, 5)
B.5. YAU-YU LINKS 567

B.5. The Yau-Yu Links in Dimensions 5

Table B.5.1: The Yau-Yu Links in Dimensions 5


Type f (z0 , z1 , z2 , z3 ) |w|/degree(f )
1 1 11
I (BP) z0a + z1b + z2c + z3d a + b + dc +
1 1 1 c−1
II z0a + z1b + z2c + z2 z3d a + b + c + cd
1 1 d−1 c−1
III z0a + z1b + z2c z3 + z2 z3d a + b + cd−1 + cd−1
1 a−1 1 c−1
IV z0a + z0 z1b + z2c + z2 z3d a + ab + c + cd
b−1 a−1 1 c−1
V z0a z1 + z0 z1b + z2c + z2 z3d ab−1 + ab−1 + c + cd
b−1 a−1 d−1 c−1
VI z0a z1 + z0 z1b + z2c z3 + z2 z3d ab−1 + ab−1 + cd−1 + dc
1 1 b−1 [b(c−1)+1]
VII z0a + z1b + z1 z2c + z2 z3d a + b + bc + bcd
VIII z0a + z1b + z1 z2c + z1 z3d + z2p z3q 1 1 b−1
a + b + bc + bd
b−1

p(b−1)
bc + q(b−1)
bc =1
(d−1) b(d−1)
IX z0 + z1 z3 + z2 z3 + z1 z3d + z1p z2q
a b c 1
a + bd−1 + c(bd−1) + b−1
bd−1
p(d−1) qb(d−1)
bd−1 + c(bd−1) = 1
[d(c−1)+1]
X z0a + z1b z2 + z2c z3 + z1 z3d 1
a + bcd+1 + [b(d−1)+1]
bcd+1
+ [c(b−1)+1]
bcd+1
1 a−1 [a(b−1)+1]
XI z0a + z0 z1b + z1 z2c + z2 z3d a + ab + abc
+ [ab(c−1)+(a−1)]
abcd
[a(b−1)+1]
XII z0a + z0 z1b + z0 z2c + z1 z3d + z1p z2q 1 a−1
a + ab + ac +
a−1
abd
p(a−1)
ab + q(a−1)
ac =1
[a(b−1)+1]
XIII z0 + z0 z1 + z1 z2 + z1 z3d + z2p z3q
a b c 1
a + a−1
ab + a−1
ac + abd
p[a(b−1)+1]
abc + q[a(b−1)+1]
abd =1
XIV z0a + z0 z1b + z0 z2c + z0 z3d + z1p z2q + z2r z3s 1
a + a−1
ab + a−1
ac + a−1
ad
p(a−1)
ab + q(a−1)
ac = 1 = r(a−1)ac + s(a−1)
ad
b(a−1)
XV z0a z1 + z0 z1b + z0 z2c + z2 z3d + z1p z2q b−1
ab−1
a−1
+
ab−1 + c(ab−1)
p(a−1) qb(a−1)
ab−1 + c(ab−1) = 1 + [c(ab−1)−b(a−1)]
cd(ab−1)
(b−1) (a−1) b(a−1) b(a−1)
XVI z0a z1 + z0 z1b + z0 z2c + z0 z3d + z1p z2q + z2r z3s ab−1 + ab−1 + c(ab−1) + d(ab−1)
p(a−1) qb(a−1) r(a−1)
ab−1 + c(ab−1) = 1 = ac + s(a−1)
ad
a(b−1) b(a−1)
XVII z0a z1 + z0 z1b + z1 z2c + z0 z3d + z1p z3q + z0r z2s b−1
ab−1 + a−1
ab−1 + c(ab−1) + d(ab−1)
p(a−1) qb(a−1) r(b−1) sa(b−1)
ab−1 + d(ab−1) = 1 = ab−1 + c(ab−1)
[b(c−1)+1]
XVIII z0a z2 + z0 z1b + z1 z2c + z1 z3d + z2p z3q abc+1 + [c(a−1)+1]
abc+1
p[a(b−1)+1]
abc+1 + qc[a(b−1)+1]
d(abc+1) =1 + [a(b−1)+1]
c(abc+1) + c[a(b−1)+1]
d(abc+1)
[b(d(c−1)+1)−1] [d(c(a−1)+1)−1]
XIX z0 z3 + z0 z1 + z2 z3 + z2 z3d
a b c
abcd−1 + abcd−1
[a(b(d−1)+1)−1] [c(a(b−1)+1)−1]
+ abcd−1 + abcd−1
568 APPENDIX B
Bibliography

[AB68] M. F. Atiyah and R. Bott, A Lefschetz fixed point formula for elliptic complexes.
II. Applications, Ann. of Math. (2) 88 (1968), 451–491. MR 0232406 (38 #731)
[AB83] , The Yang-Mills equations over Riemann surfaces, Philos. Trans. Roy. Soc.
London Ser. A 308 (1983), no. 1505, 523–615. MR 702806 (85k:14006)
[Abb84] E. Abbena, An example of an almost Kähler manifold which is not Kählerian, Boll.
Un. Mat. Ital. A (6) 3 (1984), no. 3, 383–392. MR 86a:53036
[ABC+ 96] J. Amorós, M. Burger, K. Corlette, D. Kotschick, and D. Toledo, Fundamen-
tal groups of compact Kähler manifolds, Mathematical Surveys and Monographs,
vol. 44, American Mathematical Society, Providence, RI, 1996. MR 97d:32037
[ABCC06] R. Argurio, M. Bertolini, C. Closset, and S Cremonesi, On stable non-
supersymmetric vacua at the bottom of cascading theories, J. High Energy Phys.
(2006), no. 9, 030, 21 pp. (electronic). MR 2257499
[Abe76] K. Abe, Some examples of non-regular almost contact structures on exotic spheres,
Tôhoku Math. J. (2) 28 (1976), no. 3, 429–435. MR 0423247 (54 #11227)
[Abe77] , On a generalization of the Hopf fibration. I. Contact structures on the
generalized Brieskorn manifolds, Tôhoku Math. J. (2) 29 (1977), no. 3, 335–374.
MR 0464253 (57 #4187)
[Abr01] M. Abreu, Kähler metrics on toric orbifolds, J. Differential Geom. 58 (2001), no. 1,
151–187. MR 1895351 (2003b:53046)
[ACDVP03] D. V. Alekseevsky, V. Cortés, C. Devchand, and A. Van Proeyen, Flows on
quaternionic-Kähler and very special real manifolds, Comm. Math. Phys. 238
(2003), no. 3, 525–543. MR 1993384 (2004m:53080)
[ACG06] V. Apostolov, D. M. J. Calderbank, and P. Gauduchon, Hamiltonian 2-forms in
Kähler geometry. I. General theory, J. Differential Geom. 73 (2006), no. 3, 359–412.
MR 2228318
[ADHL03] B. Acharya, F. Denef, C. Hofman, and N. Lambert, Freund-Rubin revisited,
preprint; arXiv:hep-th/0308046 (2003).
[ADM96] V. Apostolov, J. Davidov, and O. Muškarov, Compact self-dual Hermitian surfaces,
Trans. Amer. Math. Soc. 348 (1996), no. 8, 3051–3063. MR 1348147 (96j:53040)
[ADM06] V. Apostolov, T. Draghici, and A. Moroianu, The odd-dimensional Goldberg Con-
jecture, Math. Nachr. 279 (2006), no. 9-10, 948–952. MR MR2242959
[AE75] K. Abe and J. Erbacher, Nonregular contact structures on Brieskorn manifolds,
Bull. Amer. Math. Soc. 81 (1975), 407–409. MR 0417974 (54 #6019)
[AG86] V. A. Alexiev and G. T. Ganchev, On the classification of the almost contact metric
manifolds, Mathematics and mathematical education (Bulgarian) (Sunny Beach
(Sl00 nchev Bryag), 1986), Publ. House Bulgar. Acad. Sci., Sofia, 1986, pp. 155–161.
MR 872914 (88c:53035a)
[AG02] V. Apostolov and P. Gauduchon, Selfdual Einstein Hermitian four-manifolds,
Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 1 (2002), no. 1, 203–243. MR 1994808
(2004i:53048)
[AGF81] L. Alvarez-Gaumé and D. Z. Freedman, Geometrical structure and ultraviolet finite-
ness in the supersymmetric σ-model, Comm. Math. Phys. 80 (1981), no. 3, 443–451.
MR 626710 (82i:81143)
[AGI98] B. Alexandrov, G. Grantcharov, and S. Ivanov, Curvature properties of twistor
spaces of quaternionic Kähler manifolds, J. Geom. 62 (1998), no. 1-2, 1–12.
MR 1631453 (2000b:53066)

569
570 BIBLIOGRAPHY

[AGP06] C. Arezzo, A. Ghigi, and G. P. Pirola, Symmetries, quotients and Kähler-Einstein


metrics, J. Reine Angew. Math. 591 (2006), 177–200. MR 2212883
[Agr06] I. Agricola, The SRNI Lectures on non-integrable geometries with torsion, preprint;
arXiv:math.DG/0606705 (2006).
[AGV72] M. Artin, A. Grothendieck, and J. L. Verdier, Théorie des topos et cohomologie étale
des schémas. Tome 1: Théorie des topos, Springer-Verlag, Berlin, 1972, Séminaire
de Géométrie Algébrique du Bois-Marie 1963–1964 (SGA 4), Dirigé par M. Artin,
A. Grothendieck, et J. L. Verdier. Avec la collaboration de N. Bourbaki, P. Deligne
et B. Saint-Donat, Lecture Notes in Mathematics, Vol. 269. MR 50 #7130
[AH85] M. F. Atiyah and N. J. Hitchin, Low energy scattering of nonabelian monopoles,
Phys. Lett. A 107 (1985), no. 1, 21–25. MR 778313 (86e:53053)
[AH88] M. Atiyah and N. Hitchin, The geometry and dynamics of magnetic monopoles, M.
B. Porter Lectures, Princeton University Press, Princeton, NJ, 1988. MR 934202
(89k:53067)
[AHDM78] M. F. Atiyah, N. J. Hitchin, V. G. Drinfel0 d, and Yu. I. Manin, Construction of
instantons, Phys. Lett. A 65 (1978), no. 3, 185–187. MR 598562 (82g:81049)
[AHS78] M. F. Atiyah, N. J. Hitchin, and I. M. Singer, Self-duality in four-dimensional
Riemannian geometry, Proc. Roy. Soc. London Ser. A 362 (1978), no. 1711, 425–
461. MR 506229 (80d:53023)
[AJPS97] S. Albeverio, J. Jost, S. Paycha, and S. Scarlatti, A mathematical introduction to
string theory, London Mathematical Society Lecture Note Series, vol. 225, Cam-
bridge University Press, Cambridge, 1997, Variational problems, geometric and
probabilistic methods. MR 1480168 (99d:58178)
[AK94] N. A’Campo and D. Kotschick, Contact structures, foliations, and the fundamental
group, Bull. London Math. Soc. 26 (1994), no. 1, 102–106. MR 94j:57027
[AK04] D. Alekseevsky and Y. Kamishima, Quaternionic and para-quaternionic CR struc-
ture on (4n+3)-dimensional manifolds, Cent. Eur. J. Math. 2 (2004), no. 5, 732–753
(electronic). MR 2172049 (2006i:53069)
[Akh90] D. N. Akhiezer, Homogeneous complex manifolds, Several Complex Variables IV
(G.M. Khenkin, ed.), Encyclopedia of mathematical Sciences, Vol. 10, Springer-
Verlag, Berlin, 1990, Algebraic aspects of complex analysis, A translation of Sovre-
mennye problemy matematiki. Fundamentalnye napravleniya, Tom 10, Akad. Nauk
SSSR, Vsesoyuz. Inst. Nauchn. i Tekhn. Inform., Moscow, 1986 [MR 87m:32003],
Translation by J. Leiterer and J. Nunemacher, Translation edited by S. G. Gindikin
and G. M. Khenkin, pp. 223–275, 283. MR 88i:32047
[Akh95] , Lie group actions in complex analysis, Aspects of Mathematics, E27,
Friedr. Vieweg & Sohn, Braunschweig, 1995. MR 1334091 (96g:32051)
[AKL89] M. T. Anderson, P. B. Kronheimer, and C. LeBrun, Complete Ricci-flat Kähler
manifolds of infinite topological type, Comm. Math. Phys. 125 (1989), no. 4, 637–
642. MR 1024931 (91a:53073)
[Alb89] C. Albert, Le théorème de réduction de Marsden-Weinstein en géométrie cosym-
plectique et de contact, J. Geom. Phys. 6 (1989), no. 4, 627–649. MR 91k:58033
[Ale68] D. V. Alekseevskiı̆, Riemannian spaces with unusual holonomy groups, Funkcional.
Anal. i Priložen 2 (1968), no. 2, 1–10. MR 37 #6868
[Ale75] , Classification of quaternionic spaces with transitive solvable group of mo-
tions, Izv. Akad. Nauk SSSR Ser. Mat. 39 (1975), no. 2, 315–362, 472. MR 53
#6465
[Ale90] D. V. Alekseevskiı̆, Contact homogeneous spaces, Funktsional. Anal. i Prilozhen.
24 (1990), no. 4, 74–75. MR 1092805 (91j:53027)
[ALR06] A. Adem, J. Leida, and Y. Ruan, Orbifolds and stringy topology, Cambridge Uni-
versity Press (to appear), 2006.
[AM78] R. Abraham and J. E. Marsden, Foundations of mechanics, Benjamin/Cummings
Publishing Co. Inc. Advanced Book Program, Reading, Mass., 1978, Second edition,
revised and enlarged, With the assistance of Tudor Raţiu and Richard Cushman.
MR 81e:58025
[AM96a] D. V. Alekseevsky and S. Marchiafava, Hypercomplex structures on quaternionic
manifolds, New developments in differential geometry (Debrecen, 1994), Math.
BIBLIOGRAPHY 571

Appl., vol. 350, Kluwer Acad. Publ., Dordrecht, 1996, pp. 1–19. MR 1377270
(96m:53052)
[AM96b] , Quaternionic structures on a manifold and subordinated structures, Ann.
Mat. Pura Appl. (4) 171 (1996), 205–273. MR 1441871 (98j:53033)
[AMP97] L. Astey, E. Micha, and G. Pastor, Homeomorphism and diffeomorphism types of
Eschenburg spaces, Differential Geom. Appl. 7 (1997), no. 1, 41–50. MR 98h:53072
[AN54] Y. Akizuki and S. Nakano, Note on Kodaira-Spencer’s proof of Lefschetz theorems,
Proc. Japan Acad. 30 (1954), 266–272. MR 0066694 (16,619a)
[And90] M. T. Anderson, Convergence and rigidity of manifolds under Ricci curvature
bounds, Invent. Math. 102 (1990), no. 2, 429–445. MR 92c:53024
[Ano69] D. V. Anosov, Geodesic flows on closed Riemann manifolds with negative curva-
ture., Proceedings of the Steklov Institute of Mathematics, No. 90 (1967). Translated
from the Russian by S. Feder, American Mathematical Society, Providence, R.I.,
1969. MR 39 #3527
[Ara02] C. Araujo, Kähler-Einstein metrics for some quasi-smooth log del Pezzo sur-
faces, Trans. Amer. Math. Soc. 354 (2002), no. 11, 4303–4312 (electronic).
MR 2003g:32039
[Arn78] V. I. Arnold, Mathematical methods of classical mechanics, Springer-Verlag, New
York, 1978, Translated from the Russian by K. Vogtmann and A. Weinstein, Grad-
uate Texts in Mathematics, 60. MR 57 #14033b
[AS53] W. Ambrose and I. M. Singer, A theorem on holonomy, Trans. Amer. Math. Soc.
75 (1953), 428–443. MR 0063739 (16,172b)
[AS63] M. F. Atiyah and I. M. Singer, The index of elliptic operators on compact manifolds,
Bull. Amer. Math. Soc. 69 (1963), 422–433. MR 0157392 (28 #626)
[AS04] V. Apostolov and S. Salamon, Kähler reduction of metrics with holonomy G2 ,
Comm. Math. Phys. 246 (2004), no. 1, 43–61. MR 2044890 (2005e:53070)
[Ati82] M. F. Atiyah, Convexity and commuting Hamiltonians, Bull. London Math. Soc.
14 (1982), no. 1, 1–15. MR 642416 (83e:53037)
[Aub76] T. Aubin, Équations du type Monge-Ampère sur les variétés kähleriennes com-
pactes, C. R. Acad. Sci. Paris Sér. A-B 283 (1976), no. 3, Aiii, A119–A121. MR 55
#6496
[Aub78] , Équations du type Monge-Ampère sur les variétés kählériennes compactes,
Bull. Sci. Math. (2) 102 (1978), no. 1, 63–95. MR 494932 (81d:53047)
[Aub82] , Nonlinear analysis on manifolds. Monge-Ampère equations, Grundlehren
der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sci-
ences], vol. 252, Springer-Verlag, New York, 1982. MR 85j:58002
[Aub84] , Réduction du cas positif de l’équation de Monge-Ampère sur les variétés
kählériennes compactes à la démonstration d’une inégalité, J. Funct. Anal. 57
(1984), no. 2, 143–153. MR 749521 (85k:58084)
[Aub98] , Some nonlinear problems in Riemannian geometry, Springer Monographs
in Mathematics, Springer-Verlag, Berlin, 1998. MR 1636569 (99i:58001)
[AW75] S. Aloff and N. R. Wallach, An infinite family of distinct 7-manifolds admitting
positively curved Riemannian structures, Bull. Amer. Math. Soc. 81 (1975), 93–97.
MR 51 #6851
[AW02] M. Atiyah and E. Witten, M -theory dynamics on a manifold of G2 holonomy, Adv.
Theor. Math. Phys. 6 (2002), no. 1, 1–106. MR 1992874 (2004f:53046)
[Bai56] W. L. Baily, The decomposition theorem for V -manifolds, Amer. J. Math. 78 (1956),
862–888. MR 20 #6537
[Bai57] , On the imbedding of V -manifolds in projective space, Amer. J. Math. 79
(1957), 403–430. MR 20 #6538
[Bal06] W. Ballmann, Lectures on Kähler manifolds, ESI Lectures in Mathematics and
Physics, European Mathematical Society (EMS), Zürich, 2006. MR 2243012
[Ban90] A. Banyaga, A note on Weinstein’s conjecture, Proc. Amer. Math. Soc. 109 (1990),
no. 3, 855–858. MR 90m:58170
[Ban96] , Instantons and hypercontact structures, J. Geom. Phys. 19 (1996), no. 3,
267–276. MR 1397411 (97g:53025)
572 BIBLIOGRAPHY

[Ban97] , The structure of classical diffeomorphism groups, Mathematics and its


Applications, vol. 400, Kluwer Academic Publishers Group, Dordrecht, 1997.
MR 1445290 (98h:22024)
[Bar65] D. Barden, Simply connected five-manifolds, Ann. of Math. (2) 82 (1965), 365–385.
MR 32 #1714
[Bar85] R. Barlow, A simply connected surface of general type with pg = 0, Invent. Math.
79 (1985), no. 2, 293–301. MR 778128 (87a:14033)
[Bär93] Ch. Bär, Real Killing spinors and holonomy, Comm. Math. Phys. 154 (1993), no. 3,
509–521. MR 94i:53042
[Bat96a] F. Battaglia, A hypercomplex Stiefel manifold, Differential Geom. Appl. 6 (1996),
no. 2, 121–128. MR 1395025 (97g:53084)
[Bat96b] , S 1 -quotients of quaternion-Kähler manifolds, Proc. Amer. Math. Soc. 124
(1996), no. 7, 2185–2192. MR 1307492 (96i:53045)
[Bat99] , Circle actions and Morse theory on quaternion-Kähler manifolds, J. Lon-
don Math. Soc. (2) 59 (1999), no. 1, 345–358. MR 1688506 (2000h:53060)
[Bau00] H. Baum, Twistor and Killing spinors in Lorentzian geometry, Global analysis
and harmonic analysis (Marseille-Luminy, 1999), Sémin. Congr., vol. 4, Soc. Math.
France, Paris, 2000, pp. 35–52. MR 2002d:53060
[BB73] A. BiaÃlynicki-Birula, Some theorems on actions of algebraic groups, Ann. of Math.
(2) 98 (1973), 480–497. MR 0366940 (51 #3186)
[BB82] L. Bérard-Bergery, Sur de nouvelles variétés riemanniennes d’Einstein, Institut
Élie Cartan, 6, Inst. Élie Cartan, vol. 6, Univ. Nancy, Nancy, 1982, pp. 1–60.
MR 727843 (85b:53048)
[BB02] P. Berglund and A. Brandhuber, Matter from G2 manifolds, Nuclear Phys. B 641
(2002), no. 1-2, 351–375. MR 1928190 (2004f:53055)
[BBC05] M. Bertolini, F. Bigazzi, and A. L. Cotrone, Supersymmetry breaking at the end
of a cascade of Seiberg dualities, Phys. Rev. D (3) 72 (2005), no. 6, 061902, 5.
MR 2182338 (2006e:81194)
[BC01] R. L. Bishop and R. J. Crittenden, Geometry of manifolds, AMS Chelsea Publish-
ing, Providence, RI, 2001, Reprint of the 1964 original. MR 1852066 (2002d:53001)
[BCGP05] C. P. Boyer, D. M. J. Calderbank, K. Galicki, and P. Piccinni, Toric self-dual
Einstein metrics as quotients, Comm. Math. Phys. 253 (2005), no. 2, 337–370.
MR 2140252
[BD00] R. Bielawski and A. S. Dancer, The geometry and topology of toric hyperkähler
manifolds, Comm. Anal. Geom. 8 (2000), no. 4, 727–760. MR 2002c:53078
[BDM96] M. L. Barberis and I. Dotti Miatello, Hypercomplex structures on a class of solv-
able Lie groups, Quart. J. Math. Oxford Ser. (2) 47 (1996), no. 188, 389–404.
MR 1460231 (98k:53055)
[BE69] M. Berger and D. Ebin, Some decompositions of the space of symmetric tensors on
a Riemannian manifold, J. Differential Geometry 3 (1969), 379–392. MR 0266084
(42 #993)
[Bea98] A. Beauville, Fano contact manifolds and nilpotent orbits, Comment. Math. Helv.
73 (1998), no. 4, 566–583. MR 1639888 (99e:14046)
[Bea05] , Riemannian holonomy and algebraic geometry, preprint,
math.AG/9902110 (2005).
[Bel00] F. A. Belgun, On the metric structure of non-Kähler complex surfaces, Math. Ann.
317 (2000), no. 1, 1–40. MR 2002c:32027
[Bel01] , Normal CR structures on compact 3-manifolds, Math. Z. 238 (2001), no. 3,
441–460. MR 2002k:32065
[Ber55] M. Berger, Sur les groupes d’holonomie homogène des variétés à connexion affine
et des variétés riemanniennes, Bull. Soc. Math. France 83 (1955), 279–330.
MR 18,149a
[Ber66] , Remarques sur les groupes d’holonomie des variétés riemanniennes, C. R.
Acad. Sci. Paris Sér. A-B 262 (1966), A1316–A1318. MR 0200860 (34 #746)
[Ber03] , A panoramic view of Riemannian geometry, Springer-Verlag, Berlin, 2003.
MR 2002701 (2004h:53001)
BIBLIOGRAPHY 573

[Bes87] A. L. Besse, Einstein manifolds, Ergebnisse der Mathematik und ihrer Grenzgebiete
(3) [Results in Mathematics and Related Areas (3)], vol. 10, Springer-Verlag, Berlin,
1987. MR 88f:53087
[BF82] C. P. Boyer and J. D. Finley, III, Killing vectors in self-dual, Euclidean Einstein
spaces, J. Math. Phys. 23 (1982), no. 6, 1126–1130. MR 660020 (84f:53064)
[BFGK91] H. Baum, T. Friedrich, R. Grunewald, and I. Kath, Twistors and Killing spinors on
Riemannian manifolds, Teubner-Texte zur Mathematik [Teubner Texts in Math-
ematics], vol. 124, B. G. Teubner Verlagsgesellschaft mbH, Stuttgart, 1991, With
German, French and Russian summaries. MR 94a:53077
[BFH+ 05] S. Benvenuti, S. Franco, A. Hanany, D. Martelli, and J. Sparks, An infinite family
of superconformal quiver gauge theories with Sasaki-Einstein duals, J. High Energy
Phys. (2005), no. 6, 064, 26 pp. (electronic). MR 2158305 (2006h:81276)
[BFZ05] A. Butti, D. Forcella, and A. Zaffaroni, The dual superconformal theory for Lp,q,r
manifolds, J. High Energy Phys. (2005), no. 9, 018, 39 pp. (electronic). MR 2171349
(2006h:81204)
[BG67] D. E. Blair and S. I. Goldberg, Topology of almost contact manifolds, J. Differential
Geometry 1 (1967), 347–354. MR 37 #2128
[BG97a] O. Biquard and P. Gauduchon, Hyper-Kähler metrics on cotangent bundles of Her-
mitian symmetric spaces, Geometry and physics (Aarhus, 1995), Lecture Notes in
Pure and Appl. Math., vol. 184, Dekker, New York, 1997, pp. 287–298. MR 1423175
(97k:53041)
[BG97b] C. P. Boyer and K. Galicki, The twistor space of a 3-Sasakian manifold, Internat.
J. Math. 8 (1997), no. 1, 31–60. MR 98e:53072
[BG99] , 3-Sasakian manifolds, Surveys in differential geometry: essays on Einstein
manifolds, Surv. Differ. Geom., VI, Int. Press, Boston, MA, 1999, pp. 123–184.
MR 2001m:53076
[BG00a] , A note on toric contact geometry, J. Geom. Phys. 35 (2000), no. 4, 288–
298. MR 2001h:53124
[BG00b] , On Sasakian-Einstein geometry, Internat. J. Math. 11 (2000), no. 7, 873–
909. MR 2001k:53081
[BG01a] , Einstein manifolds and contact geometry, Proc. Amer. Math. Soc. 129
(2001), no. 8, 2419–2430 (electronic). MR 2001m:53071
[BG01b] , New Einstein metrics in dimension five, J. Differential Geom. 57 (2001),
no. 3, 443–463. MR 2003b:53047
[BG02] , Rational homology 5-spheres with positive Ricci curvature, Math. Res.
Lett. 9 (2002), no. 4, 521–528. MR 2003g:53044
[BG03] , New Einstein metrics on 8#(S 2 ×S 3 ), Differential Geom. Appl. 19 (2003),
no. 2, 245–251. MR 2 002 662
[BG04] D. Burns and V. Guillemin, Potential functions and actions of tori on Kähler man-
ifolds, Comm. Anal. Geom. 12 (2004), no. 1-2, 281–303. MR 2074879 (2005e:53133)
[BG05] C. P. Boyer and K. Galicki, Sasakian geometry, hypersurface singularities, and
Einstein metrics, Rend. Circ. Mat. Palermo (2) Suppl. (2005), no. 75, suppl., 57–
87. MR 2152356
[BG06a] , Einstein Metrics on Rational Homology Spheres, J. Differential Geom. 74
(2006), no. 3, 353–362. MR 2269781
[BG06b] , Erratum and addendum for “rational homology 5-spheres with positive
Ricci curvature”, Math. Res. Lett. 13 (2006), no. 3, 463–465.
[BG06c] , Highly connected manifolds with positive Ricci curvature, Geom. Topol. 10
(2006), 2219–2235(electronic).
[BG06d] , Sasakian geometry and Einstein metrics on spheres, Perspectives in Rie-
mannian Geometry, CRM Proceedings and Lecture Notes, vol. 40, Amer. Math.
Soc., Providence, RI, 2006, pp. 47–61. MR 2237105
[BGK05] C. P. Boyer, K. Galicki, and J. Kollár, Einstein metrics on spheres, Ann. of Math.
(2) 162 (2005), no. 1, 557–580. MR MR2178969 (2006j:53058)
[BGKT05] C. P. Boyer, K. Galicki, J. Kollár, and E. Thomas, Einstein metrics on exotic
spheres in dimensions 7, 11, and 15, Experiment. Math. 14 (2005), no. 1, 59–64.
MR MR2146519 (2006a:53042)
574 BIBLIOGRAPHY

[BGM93] C. P. Boyer, K. Galicki, and B. M. Mann, Quaternionic reduction and Einstein


manifolds, Comm. Anal. Geom. 1 (1993), no. 2, 229–279. MR 95c:53056
[BGM94a] , The geometry and topology of 3-Sasakian manifolds, J. Reine Angew.
Math. 455 (1994), 183–220. MR 96e:53057
[BGM94b] , Some new examples of compact inhomogeneous hypercomplex manifolds,
Math. Res. Lett. 1 (1994), no. 5, 531–538. MR 95f:53083
[BGM96a] , Hypercomplex structures on Stiefel manifolds, Ann. Global Anal. Geom.
14 (1996), no. 1, 81–105. MR 97e:53084
[BGM96b] , On strongly inhomogeneous Einstein manifolds, Bull. London Math. Soc.
28 (1996), no. 4, 401–408. MR 97a:53069
[BGM98a] , Hypercomplex structures from 3-Sasakian structures, J. Reine Angew.
Math. 501 (1998), 115–141. MR 99g:53049
[BGM98b] , A note on smooth toral reductions of spheres, Manuscripta Math. 95
(1998), no. 2, 149–158. MR 99c:53042
[BGM06] C. P. Boyer, K. Galicki, and P. Matzeu, On eta-Einstein Sasakian geometry, Comm.
Math. Phys. 262 (2006), no. 1, 177–208. MR MR2200887
[BGMR98] C. P. Boyer, K. Galicki, B. M. Mann, and E. G. Rees, Compact 3-Sasakian 7-
manifolds with arbitrary second Betti number, Invent. Math. 131 (1998), no. 2,
321–344. MR 99b:53066
[BGN02a] C. P. Boyer, K. Galicki, and M. Nakamaye, Einstein metrics on rational homology 7-
spheres, Ann. Inst. Fourier (Grenoble) 52 (2002), no. 5, 1569–1584. MR 2003j:53056
[BGN02b] , Sasakian-Einstein structures on 9#(S 2 × S 3 ), Trans. Amer. Math. Soc.
354 (2002), no. 8, 2983–2996 (electronic). MR 2003g:53061
[BGN03a] , On positive Sasakian geometry, Geom. Dedicata 101 (2003), 93–102.
MR MR2017897 (2005a:53072)
[BGN03b] , On the geometry of Sasakian-Einstein 5-manifolds, Math. Ann. 325
(2003), no. 3, 485–524. MR 2004b:53061
[BGN03c] , Sasakian geometry, homotopy spheres and positive Ricci curvature, Topol-
ogy 42 (2003), no. 5, 981–1002. MR 1 978 045
[BGO07] C. P. Boyer, K. Galicki, and L. Ornea, Constructions in Sasakian Geometry,
preprint, arXiv:math.DG/0602233, to appear in Mathematische Zeitscrift (2007).
[BGP02] C. P. Boyer, K. Galicki, and P. Piccinni, 3-Sasakian geometry, nilpotent orbits,
and exceptional quotients, Ann. Global Anal. Geom. 21 (2002), no. 1, 85–110.
MR 2003d:53076
[BGS06] C. P. Boyer, K. Galicki, and S. Simanca, Canonical Sasakian metrics,
math.DG/0604325 (2006).
[BH58] A. Borel and F. Hirzebruch, Characteristic classes and homogeneous spaces. I,
Amer. J. Math. 80 (1958), 458–538. MR 0102800 (21 #1586)
[BH99] M. R. Bridson and A. Haefliger, Metric spaces of non-positive curvature,
Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of
Mathematical Sciences], vol. 319, Springer-Verlag, Berlin, 1999. MR 1744486
(2000k:53038)
[BH02] A. Bergman and C. P. Herzog, The volume of some non-spherical horizons and
the AdS/CFT correspondence, J. High Energy Phys. (2002), no. 1, No. 30, 24.
MR 1900111 (2003e:83065)
[BHK05] S. Benvenuti, A. Hanany, and P. Kazakopoulos, The toric phases of the Y p,q quiv-
ers, J. High Energy Phys. (2005), no. 7, 021, 13 pp. (electronic). MR 2163034
(2006h:81277)
[BHPVdV04] W. P. Barth, K. Hulek, C. A. M. Peters, and A. Van de Ven, Compact complex
surfaces, second ed., Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge.
A Series of Modern Surveys in Mathematics [Results in Mathematics and Related
Areas. 3rd Series. A Series of Modern Surveys in Mathematics], vol. 4, Springer-
Verlag, Berlin, 2004. MR 2030225 (2004m:14070)
[Bie96] R. Bielawski, On the hyper-Kähler metrics associated to singularities of nilpotent
varieties, Ann. Global Anal. Geom. 14 (1996), no. 2, 177–191. MR 97e:53083
[Bie97] , Betti numbers of 3-Sasakian quotients of spheres by tori, Bull. London
Math. Soc. 29 (1997), no. 6, 731–736. MR 98j:53051
BIBLIOGRAPHY 575

[Bie99] , Complete hyper-Kähler 4n-manifolds with a local tri-Hamiltonian Rn -


action, Math. Ann. 314 (1999), no. 3, 505–528. MR 2001c:53058
[Biq96] O. Biquard, Sur les équations de Nahm et la structure de Poisson des algèbres de
Lie semi-simples complexes, Math. Ann. 304 (1996), no. 2, 253–276. MR 1371766
(97c:53066)
[Biq99] , Quaternionic contact structures, Quaternionic structures in mathematics
and physics (Rome, 1999), Univ. Studi Roma “La Sapienza”, Rome, 1999, pp. 23–30
(electronic). MR 1848655 (2002h:53079)
[Biq00] , Métriques d’Einstein asymptotiquement symétriques, Astérisque (2000),
no. 265, vi+109. MR 1760319 (2001k:53079)
[Biq02] , Métriques autoduales sur la boule, Invent. Math. 148 (2002), no. 3, 545–
607. MR 1908060 (2003i:53058)
[Bis07] L. Bisconti, Positive self-dual Einstein orbifolds with one-dimensional isometry
group.
[BK78] J.-P. Bourguignon and H. Karcher, Curvature operators: pinching estimates and
geometric examples, Ann. Sci. École Norm. Sup. (4) 11 (1978), no. 1, 71–92. MR 58
#12829
[BK06] C. Böhm and M. M. Kerr, Low-dimensional homogeneous Einstein manifolds,
Trans. Amer. Math. Soc. 358 (2006), no. 4, 1455–1468 (electronic). MR 2186982
(2006g:53056)
[BL70] J.-P. Buffet and J.-C. Lor, Une construction d’un universal pour une classe assez
large de Γ-structures, C. R. Acad. Sci. Paris Sér. A-B 270 (1970), A640–A642.
MR 0271942 (42 #6823)
[Bla70] D. E. Blair, Geometry of manifolds with structural group U (n)×O(s), J. Differential
Geometry 4 (1970), 155–167. MR 0267501 (42 #2403)
[Bla76a] , Contact manifolds in Riemannian geometry, Springer-Verlag, Berlin, 1976,
Lecture Notes in Mathematics, Vol. 509. MR 57 #7444
[Bla76b] , On the non-existence of flat contact metric structures, Tôhoku Math. J.
(2) 28 (1976), no. 3, 373–379. MR 55 #4029
[Bla77] , Two remarks on contact metric structures, Tôhoku Math. J. (2) 29 (1977),
no. 3, 319–324. MR 57 #4043
[Bla02] , Riemannian geometry of contact and symplectic manifolds, Progress
in Mathematics, vol. 203, Birkhäuser Boston Inc., Boston, MA, 2002.
MR 2002m:53120
[BLMPZ05] B. A. Burrington, J. T. Liu, M. Mahato, and L. A. Pando Zayas, Towards su-
pergravity duals of chiral symmetry breaking in Sasaki-Einstein cascading quiver
theories, J. High Energy Phys. (2005), no. 7, 019, 24 pp. (electronic). MR 2163035
[BM87] S. Bando and T. Mabuchi, Uniqueness of Einstein Kähler metrics modulo connected
group actions, Algebraic geometry, Sendai, 1985, Adv. Stud. Pure Math., vol. 10,
North-Holland, Amsterdam, 1987, pp. 11–40. MR 89c:53029
[BM93a] A. Banyaga and P. Molino, Géométrie des formes de contact complètement
intégrables de type toriques, Séminaire Gaston Darboux de Géométrie et Topolo-
gie Différentielle, 1991–1992 (Montpellier), Univ. Montpellier II, Montpellier, 1993,
pp. 1–25. MR 94e:53029
[BM93b] C. P. Boyer and B. M. Mann, The hyper-Kähler geometry of the ADHM construc-
tion and quaternionic geometric invariant theory, Differential geometry: geome-
try in mathematical physics and related topics (Los Angeles, CA, 1990), Proc.
Sympos. Pure Math., vol. 54, Amer. Math. Soc., Providence, RI, 1993, pp. 45–83.
MR 1216529 (94d:58026)
[BM96] A. Banyaga and P. Molino, Complete integrability in contact geometry, preprint
(1996), 1–25.
[BM03] A. Bilal and S. Metzger, Compact weak G2 -manifolds with conical singularities,
Nuclear Phys. B 663 (2003), no. 1-2, 343–364. MR 1986853 (2004d:53055)
[Böh98] C. Böhm, Inhomogeneous Einstein metrics on low-dimensional spheres and other
low-dimensional spaces, Invent. Math. 134 (1998), no. 1, 145–176. MR 99i:53046
[Boh03] C. Bohle, Killing spinors on Lorentzian manifolds, J. Geom. Phys. 45 (2003), no. 3-
4, 285–308. MR 1 952 661
576 BIBLIOGRAPHY

[Bon64] E. Bonan, Tenseur de structure d’une variété presque quaternionienne, C. R. Acad.


Sci. Paris 259 (1964), 45–48. MR 0166741 (29 #4014)
[Bon66] , Sur des variétés riemanniennes à groupe d’holonomie G2 ou spin (7), C.
R. Acad. Sci. Paris Sér. A-B 262 (1966), A127–A129. MR 0196668 (33 #4855)
[Bon67] , Sur les G-structures de type quaternionien, Cahiers Topologie Géom.
Différentielle 9 (1967), 389–461. MR 38 #1624
[Bon82] , Sur l’algèbre extérieure d’une variété presque hermitienne quaternionique,
C. R. Acad. Sci. Paris Sér. I Math. 295 (1982), no. 2, 115–118. MR 676377
(84e:53077)
[Boo61] W. M. Boothby, Homogeneous complex contact manifolds, Proc. Sympos. Pure
Math., Vol. III, American Mathematical Society, Providence, R. I., 1961, pp. 144–
154. MR 0124863 (23 #A2173)
[Boo80] , On compact homogeneous contact manifolds, Bull. Inst. Math. Acad. Sinica
8 (1980), no. 2-3, part 2, 341–351. MR 595542 (82a:53052)
[Bor54] A. Borel, Kählerian coset spaces of semisimple Lie groups, Proc. Nat. Acad. Sci.
U. S. A. 40 (1954), 1147–1151. MR 17,1108e
[Bor60] , Density properties for certain subgroups of semi-simple groups without
compact components, Ann. of Math. (2) 72 (1960), 179–188. MR 0123639 (23
#A964)
[Bot54] R. Bott, Nondegenerate critical manifolds, Ann. of Math. (2) 60 (1954), 248–261.
MR 0064399 (16,276f)
[Bou81] J.-P. Bourguignon, Les variétés de dimension 4 à signature non nulle dont la cour-
bure est harmonique sont d’Einstein, Invent. Math. 63 (1981), no. 2, 263–286.
MR 610539 (82g:53051)
[Bou89] N. Bourbaki, Commutative algebra. Chapters 1–7, Elements of Mathematics
(Berlin), Springer-Verlag, Berlin, 1989, Translated from the French, Reprint of the
1972 edition. MR 979760 (90a:13001)
[Bou02] F. Bourgeois, Odd dimensional tori are contact manifolds, Int. Math. Res. Not.
(2002), no. 30, 1571–1574. MR 2003f:53157
[Boy86] C. P. Boyer, Conformal duality and compact complex surfaces, Math. Ann. 274
(1986), no. 3, 517–526. MR 87i:53068
[Boy88a] , A note on hyper-Hermitian four-manifolds, Proc. Amer. Math. Soc. 102
(1988), no. 1, 157–164. MR 915736 (89c:53049)
[Boy88b] , Self-dual and anti-self-dual Hermitian metrics on compact complex sur-
faces, Mathematics and general relativity (Santa Cruz, CA, 1986), Contemp.
Math., vol. 71, Amer. Math. Soc., Providence, RI, 1988, pp. 105–114. MR 954411
(89h:53127)
[BPVdV84] W. Barth, C. Peters, and A. Van de Ven, Compact complex surfaces, Ergebnisse
der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related
Areas (3)], vol. 4, Springer-Verlag, Berlin, 1984. MR 86c:32026
[BR62] A. Borel and R. Remmert, Über kompakte homogene Kählersche Mannigfaltigkeiten,
Math. Ann. 145 (1961/1962), 429–439. MR 0145557 (26 #3088)
[BR86] M. Beltrametti and L. Robbiano, Introduction to the theory of weighted projective
spaces, Exposition. Math. 4 (1986), no. 2, 111–162. MR 88h:14052
[Bre72] G. E. Bredon, Introduction to compact transformation groups, Academic Press,
New York, 1972, Pure and Applied Mathematics, Vol. 46. MR 54 #1265
[Bre93] , Topology and geometry, Graduate Texts in Mathematics, vol. 139, Springer-
Verlag, New York, 1993. MR 94d:55001
[Bre97] , Sheaf theory, second ed., Graduate Texts in Mathematics, vol. 170,
Springer-Verlag, New York, 1997. MR MR1481706 (98g:55005)
[Bri66] E. Brieskorn, Beispiele zur Differentialtopologie von Singularitäten, Invent. Math.
2 (1966), 1–14. MR 34 #6788
[Bri68] , Rationale Singularitäten komplexer Flächen, Invent. Math. 4 (1967/1968),
336–358. MR 0222084 (36 #5136)
[Bri00] , Singularities in the work of Friedrich Hirzebruch, Surveys in differential
geometry, Surv. Differ. Geom., VII, Int. Press, Somerville, MA, 2000, pp. 17–60.
MR 1919421 (2003g:32002)
BIBLIOGRAPHY 577

[Bro73] W. Browder, Cobordism invariants, the Kervaire invariant and fixed point free
involutions, Trans. Amer. Math. Soc. 178 (1973), 193–225. MR 0324717 (48 #3067)
[Bro82] K. S. Brown, Cohomology of groups, Graduate Texts in Mathematics, vol. 87,
Springer-Verlag, New York, 1982. MR 672956 (83k:20002)
[Bru70] G. Brumfiel, The homotopy groups of BP L and P L/O. III, Michigan Math. J. 17
(1970), 217–224. MR 0271938 (42 #6819)
[Bry87] R. L. Bryant, Metrics with exceptional holonomy, Ann. of Math. (2) 126 (1987),
no. 3, 525–576. MR 89b:53084
[Bry91] , Two exotic holonomies in dimension four, path geometries, and twistor
theory, Complex geometry and Lie theory (Sundance, UT, 1989), Proc. Sym-
pos. Pure Math., vol. 53, Amer. Math. Soc., Providence, RI, 1991, pp. 33–88.
MR 1141197 (93e:53030)
[Bry96] , Classical, exceptional, and exotic holonomies: a status report, Actes de
la Table Ronde de Géométrie Différentielle (Luminy, 1992), Sémin. Congr., vol. 1,
Soc. Math. France, Paris, 1996, pp. 93–165. MR 1427757 (98c:53037)
[Bry01] , Bochner-Kähler metrics, J. Amer. Math. Soc. 14 (2001), no. 3, 623–715
(electronic). MR 1824987 (2002i:53096)
[BS89] R. L. Bryant and S. M. Salamon, On the construction of some complete metrics
with exceptional holonomy, Duke Math. J. 58 (1989), no. 3, 829–850. MR 90i:53055
[BS99] V. V. Batyrev and E. N. Selivanova, Einstein-Kähler metrics on symmetric toric
Fano manifolds, J. Reine Angew. Math. 512 (1999), 225–236. MR 2000j:32038
[BT82] R. Bott and L. W. Tu, Differential forms in algebraic topology, Graduate Texts in
Mathematics, vol. 82, Springer-Verlag, New York, 1982. MR 658304 (83i:57016)
[But05] J.-B. Butruille, Classification des variétés approximativement kähleriennes ho-
mogènes, Ann. Global Anal. Geom. 27 (2005), no. 3, 201–225. MR 2158165
(2006f:53060)
[But06] , Homogeneous nearly Kähler manifolds, preprint; arXiv:math.DG/0512655
(2006).
[BV04] N. Blažić and S. Vukmirović, Examples of self-dual, Einstein metrics of (2, 2)-
signature, Math. Scand. 94 (2004), no. 1, 63–74. MR 2032336 (2004j:53056)
[BW58] W. M. Boothby and H. C. Wang, On contact manifolds, Ann. of Math. (2) 68
(1958), 721–734. MR 22 #3015
[BW83] J. Bagger and E. Witten, Matter couplings in N = 2 supergravity, Nuclear Phys. B
222 (1983), no. 1, 1–10. MR 708440 (84k:83050)
[BZ05] A. Butti and A. Zaffaroni, R-charges from toric diagrams and the equivalence of
a-maximization and Z-minimization, J. High Energy Phys. (2005), no. 11, 019, 42
pp. (electronic). MR 2187558
[Cal56] E. Calabi, The space of kähler metrics, Proceedings of the International Congress
of Mathematicians, Vol. 1, 2 (Amsterdam 2, 1954) (Amsterdam), North-Holland,
1956, pp. 206–207.
[Cal57] , On Kähler manifolds with vanishing canonical class, Algebraic geometry
and topology. A symposium in honor of S. Lefschetz, Princeton University Press,
Princeton, N. J., 1957, pp. 78–89. MR 19,62b
[Cal79] , Métriques kählériennes et fibrés holomorphes, Ann. Sci. École Norm. Sup.
(4) 12 (1979), no. 2, 269–294. MR 83m:32033
[Cal82] , Extremal Kähler metrics, Seminar on Differential Geometry, Ann. of
Math. Stud., vol. 102, Princeton Univ. Press, Princeton, N.J., 1982, pp. 259–290.
MR 83i:53088
[Cal85] , Extremal Kähler metrics. II, Differential geometry and complex analysis,
Springer, Berlin, 1985, pp. 95–114. MR 780039 (86h:53067)
[Cam91] F. Campana, Une version géométrique généralisée du théorème du produit de
Nadel, Bull. Soc. Math. France 119 (1991), no. 4, 479–493. MR 1136848 (93h:14029)
[Cam92] , Connexité rationnelle des variétés de Fano, Ann. Sci. École Norm. Sup.
(4) 25 (1992), no. 5, 539–545. MR 1191735 (93k:14050)
[Cam04] , Orbifolds, special varieties and classification theory, Ann. Inst. Fourier
(Grenoble) 54 (2004), no. 3, 499–630. MR 2097416
[Car84a] Y. Carrière, Flots riemanniens, Astérisque (1984), no. 116, 31–52, Transversal
structure of foliations (Toulouse, 1982). MR 86m:58125a
578 BIBLIOGRAPHY

[Car84b] , Les propriétés topologiques des flots riemanniens retrouvées à l’aide


du théorème des variétés presque plates, Math. Z. 186 (1984), no. 3, 393–400.
MR 86f:53048
[CC00] A. Candel and L. Conlon, Foliations. I, Graduate Studies in Mathematics, vol. 23,
American Mathematical Society, Providence, RI, 2000. MR 2002f:57058
[CC03] , Foliations. II, Graduate Studies in Mathematics, vol. 60, American Math-
ematical Society, Providence, RI, 2003. MR 1 994 394
[CDF91] L. Castellani, R. D’Auria, and P. Fré, Supergravity and superstrings. A geomet-
ric perspective. Vol. 1-3, World Scientific Publishing Co. Inc., Teaneck, NJ, 1991,
Mathematical foundations. MR 1120022 (92k:83127a)
[CDG02] D. M. J. Calderbank, L. David, and P. Gauduchon, The Guillemin formula and
Kähler metrics on toric symplectic manifolds, J. Symplectic Geom. 1 (2002), no. 4,
767–784. MR 2039163 (2005a:53145)
[CdS01] A. Cannas da Silva, Lectures on symplectic geometry, Lecture Notes in Mathemat-
ics, vol. 1764, Springer-Verlag, Berlin, 2001. MR 1853077 (2002i:53105)
[CdS03] , Symplectic toric manifolds, Symplectic geometry of integrable Hamiltonian
systems (Barcelona, 2001), Adv. Courses Math. CRM Barcelona, Birkhäuser, Basel,
2003, pp. 85–173. MR 2000746
[CdSW99] A. Cannas da Silva and A. Weinstein, Geometric models for noncommutative alge-
bras, Berkeley Mathematics Lecture Notes, vol. 10, American Mathematical Society,
Providence, RI, 1999. MR 2001m:58013
[CEZ07] T. Chinburg, Ch. Escher, and W. Ziller, Topological properties of Eschenburg spaces
and 3-Sasakian manifolds, Mathematische Annalen (2007), to appear.
[CFdL85] L. A. Cordero, M. Fernández, and M. de León, Examples of compact non-Kähler
almost Kähler manifolds, Proc. Amer. Math. Soc. 95 (1985), no. 2, 280–286.
MR 86m:53049
[CFG86] L. A. Cordero, M. Fernández, and A. Gray, Symplectic manifolds with no Kähler
structure, Topology 25 (1986), no. 3, 375–380. MR 87j:53051
[CFG93] , The failure of complex and symplectic manifolds to be Kählerian, Differen-
tial geometry: geometry in mathematical physics and related topics (Los Angeles,
CA, 1990), Proc. Sympos. Pure Math., vol. 54, Amer. Math. Soc., Providence, RI,
1993, pp. 107–123. MR 94e:53071
[CFO07] K. Cho, A. Futaki, and H. Ono, Uniqueness and examples of compact toric Sasaki-
Einstein metrics, preprint; arXiv:math.DG/0701122 (2007).
[CG90] D. Chinea and C. Gonzalez, A classification of almost contact metric manifolds,
Ann. Mat. Pura Appl. (4) 156 (1990), 15–36. MR 1080209 (91i:53047)
[CGLP02] M. Cvetič, G. W. Gibbons, H. Lu, and C. N. Pope, Cohomogeneity one manifolds of
Spin(7) and G2 holonomy, Ann. Physics 300 (2002), no. 2, 139–184. MR 1921784
(2003m:53074)
[CGLP04] M. Cvetič, G. W. Gibbons, H. Lü, and C. N. Pope, New cohomogeneity one metrics
with Spin(7) holonomy, J. Geom. Phys. 49 (2004), no. 3-4, 350–365. MR 2076750
(2005f:53074)
[CH05] S. A. Cherkis and N. J. Hitchin, Gravitational instantons of type Dk , Comm. Math.
Phys. 260 (2005), no. 2, 299–317. MR 2177322 (2007a:53095)
[Che46] S. S. Chern, Characteristic classes of Hermitian manifolds, Ann. of Math. (2) 47
(1946), 85–121. MR 0015793 (7,470b)
[Che78] B.-Y. Chen, Some topological obstructions to Bochner-Kaehler metrics and their
applications, J. Differential Geom. 13 (1978), no. 4, 547–558 (1979). MR 570217
(81f:32037)
[Che03] J. Cheeger, Degeneration of Einstein metrics and metrics with special holonomy,
Surveys in differential geometry, Vol. VIII (Boston, MA, 2002), Surv. Differ. Geom.,
VIII, Int. Press, Somerville, MA, 2003, pp. 29–73. MR 2039985 (2005e:53060)
[Che07] I. Cheltsov, Fano varieties with many selfmaps, preprint; arXiv:math.AG/0611881
(2007).
[CJS79] E. Cremmer, B. Julia, and J. Scherk, Supergravity in eleven dimensions, Phys. Lett.
B 76 (1979), 409–412.
BIBLIOGRAPHY 579

[CK99] D. A. Cox and S. Katz, Mirror symmetry and algebraic geometry, Mathematical
Surveys and Monographs, vol. 68, American Mathematical Society, Providence, RI,
1999. MR 1677117 (2000d:14048)
[CL85] C. Camacho and N. A. Lins, Geometric theory of foliations, Birkhäuser Boston
Inc., Boston, MA, 1985, Translated from the Portuguese by Sue E. Goodman.
MR 87a:57029
[CL97] F. Catanese and C. LeBrun, On the scalar curvature of Einstein manifolds, Math.
Res. Lett. 4 (1997), no. 6, 843–854. MR 1492124 (98k:53057)
[CLOT03] O. Cornea, G. Lupton, J. Oprea, and D. Tanré, Lusternik-Schnirelmann category,
Mathematical Surveys and Monographs, vol. 103, American Mathematical Society,
Providence, RI, 2003. MR 1 990 857
[CLPP05] M. Cvetič, H. Lü, D. N. Page, and C. N. Pope, New Einstein-Sasaki spaces in five
and higher dimensions, Phys. Rev. Lett. 95 (2005), no. 7, 071101, 4. MR 2167018
[CLS90] P. Candelas, M. Lynker, and R. Schimmrigk, Calabi-Yau manifolds in weighted P4 ,
Nuclear Phys. B 341 (1990), no. 2, 383–402. MR 91m:14062
[CM74] S. S. Chern and J. K. Moser, Real hypersurfaces in complex manifolds, Acta Math.
133 (1974), 219–271. MR 0425155 (54 #13112)
[CM92] D. Chinea and J. C. Marrero, Classification of almost contact metric structures,
Rev. Roumaine Math. Pures Appl. 37 (1992), no. 3, 199–211. MR 1172273
(93g:53044)
[CM00] M. Crainic and I. Moerdijk, A homology theory for étale groupoids, J. Reine Angew.
Math. 521 (2000), 25–46. MR 1752294 (2001f:58039)
[CMS96] F. M. Cabrera, M. D. Monar, and A. F. Swann, Classification of G2 -structures, J.
London Math. Soc. (2) 53 (1996), no. 2, 407–416. MR 97e:53047
[Con74] L. Conlon, Transversally parallelizable foliations of codimension two, Trans. Amer.
Math. Soc. 194 (1974), 79–102, erratum, ibid. 207 (1975), 406. MR 51 #6844
[Cor96] V. Cortés, Alekseevskian spaces, Differential Geom. Appl. 6 (1996), no. 2, 129–168.
MR 1395026 (97m:53079)
[Cor00] , A new construction of homogeneous quaternionic manifolds and related
geometric structures, Mem. Amer. Math. Soc. 147 (2000), no. 700, viii+63.
MR 1708628 (2001e:53047)
[CP02] D. M. J. Calderbank and H. Pedersen, Selfdual Einstein metrics with torus symme-
try, J. Differential Geom. 60 (2002), no. 3, 485–521. MR 1950174 (2003m:53065)
[CR04] W. Chen and Y. Ruan, A new cohomology theory of orbifold, Comm. Math. Phys.
248 (2004), no. 1, 1–31. MR MR2104605 (2005j:57036)
[Cro01] D. Crowley, The classification of highly connected manifolds in dimension 7 and
15, Indiana University Thesis (2001).
[CS04] D. M. J. Calderbank and M. A. Singer, Einstein metrics and complex singularities,
Invent. Math. 156 (2004), no. 2, 405–443. MR 2052611 (2005h:53064)
[CS06a] , Toric self-dual Einstein metrics on compact orbifolds, Duke Math. J. 133
(2006), no. 2, 237–258. MR 2225692
[CS06b] D. Conti and S. Salamon, Generalized Killing spinors in dimension 5, preprint;
arXiv:math.DG/0508375, to appear in Trans. Amer. Math. Soc. (2006).
[CT06] J. Cheeger and G. Tian, Curvature and injectivity radius estimates for Einstein 4-
manifolds, J. Amer. Math. Soc. 19 (2006), no. 2, 487–525 (electronic). MR 2188134
(2006i:53042)
[ČV98] M. Čadek and J. Vanžura, Almost quaternionic structures on eight-manifolds, Os-
aka J. Math. 35 (1998), no. 1, 165–190. MR 1621248 (99f:57031)
[CZ06] H.-D. Cao and X.-P. Zhou, The complete proof of the Poincaré and geometrization
conjectures – application of the Hamilton-Perelman theory of the Ricci flow, Asian
J. Math. 10 (2006), no. 2, 165–492.
[Dan93] A. S. Dancer, Dihedral singularities and gravitational instantons, J. Geom. Phys.
12 (1993), no. 2, 77–91. MR 1231230 (94g:53038)
[Dan94] , A family of hyper-Kähler manifolds, Quart. J. Math. Oxford Ser. (2) 45
(1994), no. 180, 463–478. MR 1315458 (96a:53059)
[Dan96] , Scalar-flat Kähler metrics with SU(2) symmetry, J. Reine Angew. Math.
479 (1996), 99–120. MR 1414390 (97m:53080)
580 BIBLIOGRAPHY

[Dan99] , Hyper-Kähler manifolds, Surveys in differential geometry: essays on Ein-


stein manifolds, Surv. Differ. Geom., VI, Int. Press, Boston, MA, 1999, pp. 15–38.
MR 1798605 (2001k:53087)
[Dav05] L. David, The Bochner-flat cone of a CR manifold, preprint; math.DG/0512604
(2005).
[Dea04] O. Dearricott, Positive sectional curvature on 3-Sasakian manifolds, Ann. Global
Anal. Geom. 25 (2004), no. 1, 59–72. MR 2033140 (2005f:53060)
[Dea05] , Positively curved self-dual Einstein metrics on weighted projective planes,
Ann. Global Anal. Geom. 27 (2005), no. 1, 79–86. MR 2130534
[DEF+ 99] P. Deligne, P. Etingof, D. S. Freed, L. C. Jeffrey, D. Kazhdan, J. W. Morgan,
D. R. Morrison, and E. Witten (eds.), Quantum fields and strings: a course for
mathematicians. Vol. 1, 2, American Mathematical Society, Providence, RI, 1999,
Material from the Special Year on Quantum Field Theory held at the Institute for
Advanced Study, Princeton, NJ, 1996–1997. MR 1701618 (2000e:81010)
[Del75] C. Delorme, Espaces projectifs anisotropes, Bull. Soc. Math. France 103 (1975),
no. 2, 203–223. MR 0404277 (53 #8080a)
[Del88] T. Delzant, Hamiltoniens périodiques et images convexes de l’application moment,
Bull. Soc. Math. France 116 (1988), no. 3, 315–339. MR 984900 (90b:58069)
[Dem96] J.-P. Demailly, L2 vanishing theorems for positive line bundles and adjunction the-
ory, Transcendental methods in algebraic geometry (Cetraro, 1994), Lecture Notes
in Math., vol. 1646, Springer, Berlin, 1996, pp. 1–97. MR 1603616 (99k:32051)
[Dem01] , Multiplier ideal sheaves and analytic methods in algebraic geometry, School
on Vanishing Theorems and Effective Results in Algebraic Geometry (Trieste, 2000),
ICTP Lect. Notes, vol. 6, Abdus Salam Int. Cent. Theoret. Phys., Trieste, 2001,
pp. 1–148. MR 1919457 (2003f:32020)
[Der83] A. Derdziński, Self-dual Kähler manifolds and Einstein manifolds of dimension
four, Compositio Math. 49 (1983), no. 3, 405–433. MR 707181 (84h:53060)
[DF89] S. Donaldson and R. Friedman, Connected sums of self-dual manifolds and de-
formations of singular spaces, Nonlinearity 2 (1989), no. 2, 197–239. MR 994091
(90e:32027)
[DF06] S. Donaldson and J. Fine, Toric anti-self-dual 4-manifolds via complex geometry,
Math. Ann. 336 (2006), no. 2, 281–309. MR 2244374
[DGMS75] P. Deligne, P. Griffiths, J. Morgan, and D. Sullivan, Real homotopy theory of Kähler
manifolds, Invent. Math. 29 (1975), no. 3, 245–274. MR 52 #3584
[Dim92] A. Dimca, Singularities and topology of hypersurfaces, Universitext, Springer-
Verlag, New York, 1992. MR 94b:32058
[DK75] A. Durfee and L. Kauffman, Periodicity of branched cyclic covers, Math. Ann. 218
(1975), no. 2, 157–174. MR 52 #6731
[DK79] T. Duchamp and M. Kalka, Deformation theory for holomorphic foliations, J. Dif-
ferential Geom. 14 (1979), no. 3, 317–337 (1980). MR 82b:57019
[DK80] , Stability theorems for holomorphic foliations, Trans. Amer. Math. Soc. 260
(1980), no. 1, 255–266. MR 81f:57022
[DK84] , Holomorphic foliations and deformations of the Hopf foliation, Pacific J.
Math. 112 (1984), no. 1, 69–81. MR 85m:57019
[DK90] S. K. Donaldson and P. B. Kronheimer, The geometry of four-manifolds, Oxford
Mathematical Monographs, The Clarendon Press Oxford University Press, New
York, 1990, Oxford Science Publications. MR 1079726 (92a:57036)
[DK00] J. J. Duistermaat and J. A. C. Kolk, Lie groups, Universitext, Springer-Verlag,
Berlin, 2000. MR 2001j:22008
[DK01] J-P. Demailly and J. Kollár, Semi-continuity of complex singularity exponents and
Kähler-Einstein metrics on Fano orbifolds, Ann. Sci. École Norm. Sup. (4) 34
(2001), no. 4, 525–556. MR 2002e:32032
[DMR82] A. Dı́az Miranda and A. Reventós, Homogeneous contact compact manifolds and
homogeneous symplectic manifolds, Bull. Sci. Math. (2) 106 (1982), no. 4, 337–350.
MR 688194 (84d:53034)
[dMT97] S. F. B. de Moraes and C. Tomei, Moment maps on symplectic cones, Pacific J.
Math. 181 (1997), no. 2, 357–375. MR 98m:58038
BIBLIOGRAPHY 581

[DO98] S. Dragomir and L. Ornea, Locally conformal Kähler geometry, Progress in Math-
ematics, vol. 155, Birkhäuser Boston Inc., Boston, MA, 1998. MR 1481969
(99a:53081)
[DO06] O. Drăgulete and L. Ornea, Non-zero contact and Sasakian reduction, Differential
Geom. Appl. 24 (2006), no. 3, 260–270. MR MR2216940
[Dol53] P. Dolbeault, Sur la cohomologie des variétés analytiques complexes, C. R. Acad.
Sci. Paris 236 (1953), 175–177. MR 0052771 (14,673a)
[Dol63] A. Dold, Partitions of unity in the theory of fibrations, Ann. of Math. (2) 78 (1963),
223–255. MR 0155330 (27 #5264)
[Dol82] I. Dolgachev, Weighted projective varieties, Group actions and vector fields (Van-
couver, B.C., 1981), Lecture Notes in Math., vol. 956, Springer, Berlin, 1982, pp. 34–
71. MR 85g:14060
[DOR02] O. Drăgulete, L. Ornea, and T. S. Ratiu, Cosphere bundle reduction in contact ge-
ometry, J. Symplectic Geom. 1 (2002), no. 4, 695–714. MR 2039161 (2004m:53141)
[DP04] J.-P. Demailly and M. Paun, Numerical characterization of the Kähler cone of
a compact Kähler manifold, Ann. of Math. (2) 159 (2004), no. 3, 1247–1274.
MR 2113021
[DS97a] A. Dancer and A. Swann, The geometry of singular quaternionic Kähler quotients,
Internat. J. Math. 8 (1997), no. 5, 595–610. MR 1468352 (98m:53062)
[DS97b] A. S. Dancer and R. Szöke, Symmetric spaces, adapted complex structures and
hyper-Kähler structures, Quart. J. Math. Oxford Ser. (2) 48 (1997), no. 189, 27–38.
MR 1439696 (98e:53083)
[DT92] W. Y. Ding and G. Tian, Kähler-Einstein metrics and the generalized Futaki in-
variant, Invent. Math. 110 (1992), no. 2, 315–335. MR 93m:53039
[DT06] S. Dragomir and G. Tomassini, Differential geometry and analysis on CR manifolds,
Progress in Mathematics, vol. 246, Birkhäuser Boston Inc., Boston, MA, 2006.
MR 2214654
[Duc06] D. Duchemin, Quaternionic contsct structures in dimension 7, Ann. Inst. Fourier
(Grenoble) 56 (2006), no. 4, 851–885.
[Dur71] A. H. Durfee, Diffeomorphism Classification of Isolated Hypersurfaces Singularities,
Cornell University Thesis (1971).
[Dur77] , Bilinear and quadratic forms on torsion modules, Advances in Math. 25
(1977), no. 2, 133–164. MR 0480333 (58 #506)
[dWLP+ 84] B. de Wit, P. G. Lauwers, R. Philippe, S. Q. Su, and A. Van Proeyen, Gauge and
matter fields coupled to N = 2 supergravity, Phys. Lett. B 134 (1984), no. 1-2,
37–43. MR 729799 (84m:83067)
[dWLVP85] B. de Wit, Lauwers, and A. Van Proeyen, Lagrangians in N = 2 supergravity-matter
systems, Nucl. Phys. B 255 (1985), 569–601.
[dWVP92] B. de Wit and A. Van Proeyen, Special geometry, cubic polynomials and homo-
geneous quaternionic spaces, Comm. Math. Phys. 149 (1992), no. 2, 307–333.
MR 1186031 (94a:53079)
[EH79] T. Eguchi and A. J. Hanson, Self-dual solutions to Euclidean gravity, Ann. Physics
120 (1979), no. 1, 82–106. MR 540896 (80h:83011)
[Ehr53] C. Ehresmann, Introduction à la théorie des structures infinitésimales et des pseu-
dogroupes de Lie, Colloque de topologie et géométrie différentielle, Strasbourg,
1952, no. 11, La Bibliothèque Nationale et Universitaire de Strasbourg, 1953, p. 16.
MR 15,828e
[EKA90] A. El Kacimi-Alaoui, Opérateurs transversalement elliptiques sur un feuilletage rie-
mannien et applications, Compositio Math. 73 (1990), no. 1, 57–106. MR 91f:58089
[EKAH86] A. El Kacimi-Alaoui and G. Hector, Décomposition de Hodge basique pour un
feuilletage riemannien, Ann. Inst. Fourier (Grenoble) 36 (1986), no. 3, 207–227.
MR 87m:57029
[EKAN93] A. El Kacimi-Alaoui and M. Nicolau, On the topological invariance of the basic
cohomology, Math. Ann. 295 (1993), no. 4, 627–634. MR 1214951 (94c:57046)
[Eli89] Y. Eliashberg, Classification of overtwisted contact structures on 3-manifolds, In-
vent. Math. 98 (1989), no. 3, 623–637. MR 1022310 (90k:53064)
[Eli92] , Contact 3-manifolds twenty years since J. Martinet’s work, Ann. Inst.
Fourier (Grenoble) 42 (1992), no. 1-2, 165–192. MR 1162559 (93k:57029)
582 BIBLIOGRAPHY

[Eli98] , Invariants in contact topology, Proceedings of the International Congress


of Mathematicians, Vol. II (Berlin, 1998), no. Extra Vol. II, 1998, pp. 327–338
(electronic). MR 2000a:57068
[Eps72] D. B. A. Epstein, Periodic flows on three-manifolds, Ann. of Math. (2) 95 (1972),
66–82. MR 0288785 (44 #5981)
[ER44] C. Ehresmann and G. Reeb, Sur les champs d’éléments de contact de dimension
p complètement intégrables dans une variété continuement différentiable Vn , C. R.
Acad. Sci. Paris 218 (1944), 955–957. MR 0014757 (7,327g)
[Esc82] J.-H. Eschenburg, New examples of manifolds with strictly positive curvature, In-
vent. Math. 66 (1982), no. 3, 469–480. MR 83i:53061
[Esc84] , Freie isometrische Aktionen auf kompakten Lie-Gruppen mit positiv
gekrümmten Orbiträumen, Schriftenreihe des Mathematischen Instituts der Uni-
versität Münster, 2. Serie [Series of the Mathematical Institute of the University of
Münster, Series 2], vol. 32, Universität Münster Mathematisches Institut, Münster,
1984. MR 86a:53045
[Esc92] , Cohomology of biquotients, Manuscripta Math. 75 (1992), no. 2, 151–166.
MR 93e:57070
[Fan04] F. Fang, Positive quaternionic Kähler manifolds and symmetry rank, J. Reine
Angew. Math. 576 (2004), 149–165. MR 2099202 (2005i:53060)
[FG82] M. Fernández and A. Gray, Riemannian manifolds with structure group G2 , Ann.
Mat. Pura Appl. (4) 132 (1982), 19–45 (1983). MR 84e:53056
[FGG88] M. Fernández, M. J. Gotay, and A. Gray, Compact parallelizable four-dimensional
symplectic and complex manifolds, Proc. Amer. Math. Soc. 103 (1988), no. 4, 1209–
1212. MR 90a:53039
[FIMU06] M. Fernandez, S. Ivanov, V Muñoz, and L. Ugarte, Nearly hypo structures
and compact nearly Kähler 6-manifolds with conical singularities, preprint;
arXiv:math.DG/0602160 (2006).
[Fin77] R. Fintushel, Circle actions on simply connected 4-manifolds, Trans. Amer. Math.
Soc. 230 (1977), 147–171. MR 0458456 (56 #16659)
[Fin78] , Classification of circle actions on 4-manifolds, Trans. Amer. Math. Soc.
242 (1978), 377–390. MR 496815 (81e:57036)
[Fin01] J. D. Finley, III, Difficulties with the SDiff(2) Toda equation, Bäcklund and Dar-
boux transformations. The geometry of solitons (Halifax, NS, 1999), CRM Proc.
Lecture Notes, vol. 29, Amer. Math. Soc., Providence, RI, 2001, pp. 217–224.
MR 1870919 (2002k:37148)
[FIP04] M. Falcitelli, S. Ianus, and A. M. Pastore, Riemannian submersions and related
topics, World Scientific Publishing Co. Inc., River Edge, NJ, 2004. MR 2110043
[FK82] T. Friedrich and H. Kurke, Compact four-dimensional self-dual Einstein manifolds
with positive scalar curvature, Math. Nachr. 106 (1982), 271–299. MR 84b:53043
[FK89] T. Friedrich and I. Kath, Einstein manifolds of dimension five with small first
eigenvalue of the Dirac operator, J. Differential Geom. 29 (1989), no. 2, 263–279.
MR 982174 (90e:58158)
[FK90] , 7-dimensional compact Riemannian manifolds with Killing spinors,
Comm. Math. Phys. 133 (1990), no. 3, 543–561. MR 91j:58170
[FKMS97] T. Friedrich, I. Kath, A. Moroianu, and U. Semmelmann, On nearly parallel G2 -
structures, J. Geom. Phys. 23 (1997), no. 3-4, 259–286. MR 98j:53053
[FKS02] E. V. Ferapontov, D. A. Korotkin, and V. A. Shramchenko, Boyer-Finley equation
and systems of hydrodynamic type, Classical Quantum Gravity 19 (2002), no. 24,
L205–L210. MR 1956302 (2003m:83024)
[Flo91] A. Floer, Self-dual conformal structures on lCP2 , J. Differential Geom. 33 (1991),
no. 2, 551–573. MR 1094469 (92e:53049)
[FM94] R. Friedman and J. W. Morgan, Smooth four-manifolds and complex surfaces,
Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and
Related Areas (3)], vol. 27, Springer-Verlag, Berlin, 1994. MR 1288304 (95m:57046)
[FOW06] A. Futaki, H. Ono, and G. Wang, Transverse Kähler geometry of Sasaki manifolds
and toric Sasaki-Einstein manifolds, preprint: arXiv:math.DG/0607586 (2006).
[Fra87] A. Franc, Spin structures and Killing spinors on lens spaces, J. Geom. Phys. 4
(1987), no. 3, 277–287. MR 957015 (90e:57047)
BIBLIOGRAPHY 583

[Fri80] T. Friedrich, Der erste Eigenwert des Dirac-Operators einer kompakten, Rie-
mannschen Mannigfaltigkeit nichtnegativer Skalarkrümmung, Math. Nachr. 97
(1980), 117–146. MR 82g:58088
[Fri00] , Dirac operators in Riemannian geometry, Graduate Studies in Mathemat-
ics, vol. 25, American Mathematical Society, Providence, RI, 2000, Translated from
the 1997 German original by Andreas Nestke. MR 1777332 (2001c:58017)
[Fuj66] T. Fujitani, Complex-valued differential forms on normal contact Riemannian man-
ifolds, Tôhoku Math. J. (2) 18 (1966), 349–361. MR 35 #3596
[Fuj72] A. Fujimoto, Theory of G-structures, Study Group of Geometry, Department of
Applied Mathematics, College of Liberal Arts and Science, Okayama University,
Okayama, 1972, English edition, translated from the original Japanese, Publications
of the Study Group of Geometry, Vol. 1. MR 50 #1161
[Fuj87] A. Fujiki, On the de Rham cohomology group of a compact Kähler symplectic man-
ifold, Algebraic geometry, Sendai, 1985, Adv. Stud. Pure Math., vol. 10, North-
Holland, Amsterdam, 1987, pp. 105–165. MR 90d:53083
[Ful84] W. Fulton, Intersection theory, Ergebnisse der Mathematik und ihrer Grenzgebiete
(3) [Results in Mathematics and Related Areas (3)], vol. 2, Springer-Verlag, Berlin,
1984. MR 732620 (85k:14004)
[Ful93] , Introduction to toric varieties, Annals of Mathematics Studies, vol. 131,
Princeton University Press, Princeton, NJ, 1993, The William H. Roever Lectures
in Geometry. MR 94g:14028
[Fur86] M. Furushima, Singular del Pezzo surfaces and analytic compactifications of
3-dimensional complex affine space C3 , Nagoya Math. J. 104 (1986), 1–28.
MR 868434 (88m:32051)
[Fut83] A. Futaki, An obstruction to the existence of Einstein Kähler metrics, Invent. Math.
73 (1983), no. 3, 437–443. MR 84j:53072
[Fut88] , Kähler-Einstein metrics and integral invariants, Lecture Notes in Mathe-
matics, vol. 1314, Springer-Verlag, Berlin, 1988. MR 90a:53053
[FW03] T. Friedmann and E. Witten, Unification scale, proton decay, and manifolds of
G2 holonomy, Adv. Theor. Math. Phys. 7 (2003), no. 4, 577–617. MR 2039032
(2005h:81279)
[Gal79] S. Gallot, Équations différentielles caractéristiques de la sphère, Ann. Sci. École
Norm. Sup. (4) 12 (1979), no. 2, 235–267. MR 543217 (80h:58051)
[Gal87a] K. Galicki, A generalization of the momentum mapping construction for quater-
nionic Kähler manifolds, Comm. Math. Phys. 108 (1987), no. 1, 117–138.
MR 88f:53088
[Gal87b] , New matter couplings in N = 2 supergravity, Nuclear Phys. B 289 (1987),
no. 2, 573–588. MR 88j:53076
[Gal91] , Multi-centre metrics with negative cosmological constant, Classical Quan-
tum Gravity 8 (1991), no. 8, 1529–1543. MR 92i:53040
[Gei91] H. Geiges, Contact structures on 1-connected 5-manifolds, Mathematika 38 (1991),
no. 2, 303–311 (1992). MR 93e:57042
[Gei97a] , Constructions of contact manifolds, Math. Proc. Cambridge Philos. Soc.
121 (1997), no. 3, 455–464. MR 1434654 (98f:53027)
[Gei97b] , Normal contact structures on 3-manifolds, Tohoku Math. J. (2) 49 (1997),
no. 3, 415–422. MR 98h:53046
[Gei06] , Contact geometry, Handbook of differential geometry. Vol. II,
Elsevier/North-Holland, Amsterdam, 2006, pp. 315–382. MR 2194671
[GGP03] D. Grantcharov, G. Grantcharov, and Y.S. Poon, Calabi-Yau connections with tor-
sion on toric bundles, preprint; arXiv:math.DG/0306207; to appear in J. Differen-
tial Geometry (2003).
[GH78a] G. W. Gibbons and S. W. Hawking, Gravitational multi-instantons, Phys. Lett. B
78 (1978), 430–432. MR 0465052 (57 #4965)
[GH78b] P. Griffiths and J. Harris, Principles of algebraic geometry, Wiley-Interscience [John
Wiley & Sons], New York, 1978, Pure and Applied Mathematics. MR 80b:14001
[GH06] K. Guruprasad and A. Haefliger, Closed geodesics on orbifolds, Topology 45 (2006),
no. 3, 611–641. MR 2218759
584 BIBLIOGRAPHY

[GHJ03] M. Gross, D. Huybrechts, and D. Joyce, Calabi-Yau manifolds and related geome-
tries, Universitext, Springer-Verlag, Berlin, 2003, Lectures from the Summer School
held in Nordfjordeid, June 2001. MR 1963559 (2004c:14075)
[GHR84] S. J. Gates, Jr., C. M. Hull, and M. Roček, Twisted multiplets and new supersym-
metric nonlinear σ-models, Nuclear Phys. B 248 (1984), no. 1, 157–186. MR 776369
(87b:81108)
[GHS83] J. Girbau, A. Haefliger, and D. Sundararaman, On deformations of transversely
holomorphic foliations, J. Reine Angew. Math. 345 (1983), 122–147. MR 84j:32026
[GHV73] W. Greub, S. Halperin, and R. Vanstone, Connections, curvature, and cohomology.
Vol. II: Lie groups, principal bundles, and characteristic classes, Academic Press
[A subsidiary of Harcourt Brace Jovanovich, Publishers], New York-London, 1973,
Pure and Applied Mathematics, Vol. 47-II. MR 49 #1424
[GHY04] G. W. Gibbons, S. A. Hartnoll, and Y. Yasui, Properties of some five-dimensional
Einstein metrics, Classical Quantum Gravity 21 (2004), no. 19, 4697–4730.
MR 2094200 (2006b:53058)
[Gif69a] C. H. Giffen, Smooth homotopy projective spaces, Bull. Amer. Math. Soc. 75 (1969),
509–513. MR 0239607 (39 #964)
[Gif69b] , Weakly complex involutions and cobordism of projective spaces, Ann. of
Math. (2) 90 (1969), 418–432. MR 0253346 (40 #6561)
[GK05] A Ghigi and J. Kollár, Kähler-Einstein metrics on orbifolds and Einstein metrics
on Spheres, preprint; arXiv:math.DG/0507289 (2005).
[GKN] M. Godliński, W. Kopczyński, and P. Nurowski.
[GKP89] R. L. Graham, D. E. Knuth, and O. Patashnik, Concrete mathematics, Addison-
Wesley Publishing Company Advanced Book Program, Reading, MA, 1989, A foun-
dation for computer science. MR 91f:00001
[GL80] M. Gromov and H. B. Lawson, Jr., The classification of simply connected mani-
folds of positive scalar curvature, Ann. of Math. (2) 111 (1980), no. 3, 423–434.
MR 577131 (81h:53036)
[GL88] K. Galicki and H. B. Lawson, Jr., Quaternionic reduction and quaternionic orb-
ifolds, Math. Ann. 282 (1988), no. 1, 1–21. MR 89m:53075
[GL02] V. Guillemin and E. Lerman, Melrose-Uhlmann projectors, the metaplectic rep-
resentation and symplectic cuts, J. Differential Geom. 61 (2002), no. 3, 365–396.
MR 2004e:53126
[GLS96] V. Guillemin, E. Lerman, and S. Sternberg, Symplectic fibrations and multiplicity
diagrams, Cambridge University Press, Cambridge, 1996. MR 1414677 (98d:58074)
[GM74] D. Gromoll and W. Meyer, An exotic sphere with nonnegative sectional curvature,
Ann. of Math. (2) 100 (1974), 401–406. MR 51 #11347
[GM80] X. Gómez-Mont, Transversal holomorphic structures, J. Differential Geom. 15
(1980), no. 2, 161–185 (1981). MR 82j:53065
[GM88] M. Goresky and R. MacPherson, Stratified Morse theory, Ergebnisse der Mathe-
matik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)],
vol. 14, Springer-Verlag, Berlin, 1988. MR 932724 (90d:57039)
[GMSW04a] J. P. Gauntlett, D. Martelli, J. Sparks, and D. Waldram, A new infinite class of
Sasaki-Einstein manifolds, Adv. Theor. Math. Phys. 8 (2004), no. 6, 987–1000.
MR 2194373
[GMSW04b] , Sasaki-Einstein metrics on S 2 × S 3 , Adv. Theor. Math. Phys. 8 (2004),
no. 4, 711–734. MR 2141499
[GMSY06] J. P. Gauntlett, D. Martelli, J. Sparks, and S.-T. Yau, Obstructions to the existence
of Sasaki-Einstein metrics, preprint; arXiv:hep-th/0607080 (2006).
[GN92] K. Galicki and T. Nitta, Nonzero scalar curvature generalizations of the ALE hyper-
Kähler metrics, J. Math. Phys. 33 (1992), no. 5, 1765–1771. MR 93d:53056
[GO98] P. Gauduchon and L. Ornea, Locally conformally Kähler metrics on Hopf surfaces,
Ann. Inst. Fourier (Grenoble) 48 (1998), no. 4, 1107–1127. MR 2000g:53088
[GO01] G. Grantcharov and L. Ornea, Reduction of Sasakian manifolds, J. Math. Phys. 42
(2001), no. 8, 3809–3816. MR 1845220 (2002e:53060)
[God58] R. Godement, Topologie algébrique et théorie des faisceaux, Actualités Sci. Ind. No.
1252. Publ. Math. Univ. Strasbourg. No. 13, Hermann, Paris, 1958. MR 0102797
(21 #1583)
BIBLIOGRAPHY 585

[God91] C. Godbillon, Feuilletages, Progress in Mathematics, vol. 98, Birkhäuser Verlag,


Basel, 1991, Études géométriques. [Geometric studies], With a preface by G. Reeb.
MR 93i:57038
[Gol65] S. I. Goldberg, Rigidité de variétés de contact à courbure positive, C. R. Acad. Sci.
Paris 261 (1965), 1936–1939. MR 0188950 (32 #6377)
[Gol67] , Rigidity of positively curved contact manifolds, J. London Math. Soc. 42
(1967), 257–263. MR 0232325 (38 #650)
[Gol69] , Integrability of almost Kaehler manifolds, Proc. Amer. Math. Soc. 21
(1969), 96–100. MR 38 #6514
[Gol02] R. F. Goldin, An effective algorithm for the cohomology ring of symplectic reduc-
tions, Geom. Funct. Anal. 12 (2002), no. 3, 567–583. MR 1924372 (2003m:53148)
[Gom95] R. E. Gompf, A new construction of symplectic manifolds, Ann. of Math. (2) 142
(1995), no. 3, 527–595. MR 96j:57025
[Gor80] V. V. Gorbacevič, Compact homogeneous manifolds of small dimension, Geometric
methods in problems of algebra and analysis, No. 2 (Russian), Yaroslav. Gos. Univ.,
Yaroslavl0 , 1980, pp. 37–60, 161. MR 617386 (82j:53081)
[Got98] R. Goto, On hyper-Kähler manifolds of type A∞ and D∞ , Comm. Math. Phys.
198 (1998), no. 2, 469–491. MR 1672449 (2000f:53062)
[GP00] G. Grantcharov and Y-S. Poon, Geometry of hyper-Kähler connections with torsion,
Comm. Math. Phys. 213 (2000), no. 1, 19–37. MR 1782143 (2002a:53059)
[GPP90] G. W. Gibbons, D. N. Page, and C. N. Pope, Einstein metrics on S 3 , R3 , and R4
bundles, Comm. Math. Phys. 127 (1990), no. 3, 529–553. MR 1040893 (91f:53039)
[GR65] R. C. Gunning and H. Rossi, Analytic functions of several complex variables,
Prentice-Hall Inc., Englewood Cliffs, N.J., 1965. MR 0180696 (31 #4927)
[GR84] H. Grauert and R. Remmert, Coherent analytic sheaves, Grundlehren der Mathe-
matischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol.
265, Springer-Verlag, Berlin, 1984. MR 755331 (86a:32001)
[Gra59] J. W. Gray, Some global properties of contact structures, Ann. of Math. (2) 69
(1959), 421–450. MR 22 #3016
[Gra67] A. Gray, Pseudo-Riemannian almost product manifolds and submersions, J. Math.
Mech. 16 (1967), 715–737. MR 34 #5018
[Gra69a] , A note on manifolds whose holonomy group is a subgroup of Sp(n) · Sp(1),
Michigan Math. J. 16 (1969), 125–128. MR 39 #6226
[Gra69b] , Six dimensional almost complex manifolds defined by means of three-fold
vector cross products, Tôhoku Math. J. (2) 21 (1969), 614–620. MR 0261515 (41
#6128)
[Gra70] , Nearly Kähler manifolds, J. Differential Geometry 4 (1970), 283–309.
MR 0267502 (42 #2404)
[Gra71] , Weak holonomy groups, Math. Z. 123 (1971), 290–300. MR 45 #2614
[Gra76] , The structure of nearly Kähler manifolds, Math. Ann. 223 (1976), no. 3,
233–248. MR 0417965 (54 #6010)
[Gra04] G. Grantcharov, Hyper-Hermitian manifolds and connections with skew-symmetric
torsion, Clifford algebras (Cookeville, TN, 2002), Prog. Math. Phys., vol. 34,
Birkhäuser Boston, Boston, MA, 2004, pp. 167–183. MR 2025979 (2004m:53082)
[Gro61] A. Grothendieck, Éléments de géométrie algébrique. II. Étude globale élémentaire
de quelques classes de morphismes, Inst. Hautes Études Sci. Publ. Math. (1961),
no. 8, 222. MR 36 #177b
[Gro68] , Cohomologie locale des faisceaux cohérents et théorèmes de Lefschetz lo-
caux et globaux (SGA 2), North-Holland Publishing Co., Amsterdam, 1968, Aug-
menté d’un exposé par Michèle Raynaud, Séminaire de Géométrie Algébrique du
Bois-Marie, 1962, Advanced Studies in Pure Mathematics, Vol. 2. MR 57 #16294
[Gro81] M. Gromov, Curvature, diameter and Betti numbers, Comment. Math. Helv. 56
(1981), no. 2, 179–195. MR 82k:53062
[Gro99] , Metric structures for Riemannian and non-Riemannian spaces, Progress
in Mathematics, vol. 152, Birkhäuser Boston Inc., Boston, MA, 1999, Based on
the 1981 French original [ MR0682063 (85e:53051)], With appendices by M. Katz,
P. Pansu and S. Semmes, Translated from the French by Sean Michael Bates.
MR 1699320 (2000d:53065)
586 BIBLIOGRAPHY

[Gru90] R. Grunewald, Six-dimensional Riemannian manifolds with a real Killing spinor,


Ann. Global Anal. Geom. 8 (1990), no. 1, 43–59. MR 1075238 (92a:58146)
[GS66] V. Guillemin and S. Sternberg, Deformation theory of pseudogroup structures,
Mem. Amer. Math. Soc. No. 64 (1966), 80. MR 35 #2302
[GS82a] , Convexity properties of the moment mapping, Invent. Math. 67 (1982),
no. 3, 491–513. MR 664117 (83m:58037)
[GS82b] , Homogeneous quantization and multiplicities of group representations, J.
Funct. Anal. 47 (1982), no. 3, 344–380. MR 84d:58034
[GS84] , Symplectic techniques in physics, Cambridge University Press, Cambridge,
1984. MR 770935 (86f:58054)
[GS96] K. Galicki and S. Salamon, Betti numbers of 3-Sasakian manifolds, Geom. Dedicata
63 (1996), no. 1, 45–68. MR 98e:53075
[GS99] R. E. Gompf and A. I. Stipsicz, 4-manifolds and Kirby calculus, Graduate Studies
in Mathematics, vol. 20, American Mathematical Society, Providence, RI, 1999.
MR 1707327 (2000h:57038)
[GT95] H. Geiges and C. B. Thomas, Hypercontact manifolds, J. London Math. Soc. (2)
51 (1995), no. 2, 342–352. MR 96d:53048
[Gui94a] V. Guillemin, Kaehler structures on toric varieties, J. Differential Geom. 40 (1994),
no. 2, 285–309. MR 1293656 (95h:32029)
[Gui94b] , Moment maps and combinatorial invariants of Hamiltonian T n -spaces,
Progress in Mathematics, vol. 122, Birkhäuser Boston Inc., Boston, MA, 1994.
MR 1301331 (96e:58064)
[GWZ05] K. Grove, B. Wilking, and W. Ziller, Positively curved manifolds cohomogeneity
one manifolds and 3-Sasakian geometry, math.DG/0511464 (2005).
[GY70] S. I. Goldberg and K. Yano, On normal globally framed f -manifolds, Tôhoku Math.
J. (2) 22 (1970), 362–370. MR 0305295 (46 #4425)
[GY71] , Globally framed f -manifolds, Illinois J. Math. 15 (1971), 456–474.
MR 0278247 (43 #3978)
[GZ00] K. Grove and W. Ziller, Curvature and symmetry of Milnor spheres, Ann. of Math.
(2) 152 (2000), no. 1, 331–367. MR 2001i:53047
[GZ02] , Cohomogeneity one manifolds with positive Ricci curvature, Invent. Math.
149 (2002), no. 3, 619–646. MR 1923478 (2004b:53052)
[Hae58] A. Haefliger, Structures feuilletées et cohomologie à valeur dans un faisceau de
groupoı̈des, Comment. Math. Helv. 32 (1958), 248–329. MR 20 #6702
[Hae62] , Variétés feuilletées, Ann. Scuola Norm. Sup. Pisa (3) 16 (1962), 367–397.
MR 32 #6487
[Hae70] , Feuilletages sur les variétés ouvertes, Topology 9 (1970), 183–194. MR 41
#7709
[Hae71] , Homotopy and integrability, Manifolds–Amsterdam 1970 (Proc. Nuffic
Summer School), Lecture Notes in Mathematics, Vol. 197, Springer, Berlin, 1971,
pp. 133–163. MR 44 #2251
[Hae84] , Groupoı̈des d’holonomie et classifiants, Astérisque (1984), no. 116, 70–97,
Transversal structure of foliations (Toulouse, 1982). MR 86c:57026a
[Ham82] R. S. Hamilton, Three-manifolds with positive Ricci curvature, J. Differential Geom.
17 (1982), no. 2, 255–306. MR 664497 (84a:53050)
[Har77] R. Hartshorne, Algebraic geometry, Springer-Verlag, New York, 1977, Graduate
Texts in Mathematics, No. 52. MR 57 #3116
[Har92] J. Harris, Algebraic geometry, Graduate Texts in Mathematics, vol. 133, Springer-
Verlag, New York, 1992, A first course. MR 1182558 (93j:14001)
[Hat63] Y. Hatakeyama, Some notes on differentiable manifolds with almost contact struc-
tures, Osaka Math. J. (2) 15 (1963), 176–181. MR 27 #705
[Hau05] T. Hausel, Mirror symmetry and Langlands duality in the non-abelian Hodge the-
ory of a curve, Geometric methods in algebra and number theory, Progr. Math.,
vol. 235, Birkhäuser Boston, Boston, MA, 2005, pp. 193–217. MR 2166085
(2006h:14055)
[Hau06] , Betti numbers of holomorphic symplectic quotients via arithmetic Fourier
transform, Proc. Natl. Acad. Sci. USA 103 (2006), no. 16, 6120–6124 (electronic).
MR 2221039
BIBLIOGRAPHY 587

[Haw77] S. W. Hawking, Gravitational instantons, Phys. Lett. A 60 (1977), no. 2, 81–83.


MR 0465052 (57 #4965)
[HD84] A. Haefliger and Quach Ngoc Du, Appendice: une présentation du groupe fonda-
mental d’une orbifold, Astérisque (1984), no. 116, 98–107, Transversal structure of
foliations (Toulouse, 1982). MR 755164 (86c:57026b)
[HE73] S. W. Hawking and G. F. R. Ellis, The large scale structure of space-time, Cam-
bridge University Press, London, 1973, Cambridge Monographs on Mathematical
Physics, No. 1. MR 54 #12154
[HEK05] C. P. Herzog, Q. J. Ejaz, and I. R. Klebanov, Cascading RG flows from new Sasaki-
Einstein manifolds, J. High Energy Phys. (2005), no. 2, 009, 22 pp. (electronic).
MR 2140502
[Hep05a] R. Hepworth, Generalized Kreck-Stolz invariants and the topology of certain 3-
Sasakian 7-manifolds, Edinburgh University Thesis (2005).
[Hep05b] , The topology of certain 3-Sasakian 7-manifolds, math.AT/0511735 (2005).
[Her96] G. Hernandez, On hyper f -structures, Math. Ann. 306 (1996), no. 2, 205–230.
MR 1411345 (97m:53057)
[HH86] G. Hector and U. Hirsch, Introduction to the geometry of foliations. Part A, sec-
ond ed., Aspects of Mathematics, vol. 1, Friedr. Vieweg & Sohn, Braunschweig,
1986, Foliations on compact surfaces, fundamentals for arbitrary codimension, and
holonomy. MR 88a:57048
[HH02a] H. Herrera and R. Herrera, Classification of positive quaternion-Kähler 12-
manifolds, C. R. Math. Acad. Sci. Paris 334 (2002), no. 1, 43–46. MR 2003e:53056
[HH02b] , Â-genus on non-spin manifolds with S 1 actions and the classification of
positive quaternion-Kähler 12-manifolds, J. Differential Geom. 61 (2002), no. 3,
341–364. MR 1 979 364
[Hij86] O. Hijazi, Caractérisation de la sphère par les premières valeurs propres de
l’opérateur de Dirac en dimensions 3, 4, 7 et 8, C. R. Acad. Sci. Paris Sér. I
Math. 303 (1986), no. 9, 417–419. MR 862207 (88b:58142)
[Hir66] F. Hirzebruch, Topological methods in algebraic geometry, Third enlarged edi-
tion. New appendix and translation from the second German edition by R. L.
E. Schwarzenberger, with an additional section by A. Borel. Die Grundlehren der
Mathematischen Wissenschaften, Band 131, Springer-Verlag New York, Inc., New
York, 1966. MR 0202713 (34 #2573)
[Hir67] , Singularities and exotic spheres, Séminaire Bourbaki, Vol. 10, Soc. Math.
France, Paris, 1967, pp. Exp. No. 314, 13–32. MR 1 610 436
[Hir71] , Pontrjagin classes of rational homology manifolds and the signature of
some affine hypersurfaces, Proceedings of Liverpool Singularities Symposium, II
(1969/1970) (Berlin), Springer, 1971, pp. 207–212. Lecture Notes in Math., Vol.
209. MR 49 #1525
[Hit74] N. J. Hitchin, Harmonic spinors, Advances in Math. 14 (1974), 1–55. MR 50
#11332
[Hit79] , Polygons and gravitons, Math. Proc. Cambridge Philos. Soc. 85 (1979),
no. 3, 465–476. MR 520463 (80h:53047)
[Hit81] , Kählerian twistor spaces, Proc. London Math. Soc. (3) 43 (1981), no. 1,
133–150. MR 84b:32014
[Hit87a] , Monopoles, minimal surfaces and algebraic curves, Séminaire de
Mathématiques Supérieures [Seminar on Higher Mathematics], vol. 105, Presses
de l’Université de Montréal, Montreal, QC, 1987. MR 935967 (89h:58043)
[Hit87b] , The self-duality equations on a Riemann surface, Proc. London Math. Soc.
(3) 55 (1987), no. 1, 59–126. MR 887284 (89a:32021)
[Hit92] , Hyper-Kähler manifolds, Astérisque (1992), no. 206, Exp. No. 748, 3, 137–
166, Séminaire Bourbaki, Vol. 1991/92. MR 1206066 (94f:53087)
[Hit95a] , Poncelet polygons and the Painlevé equations, Geometry and analysis
(Bombay, 1992), Tata Inst. Fund. Res., Bombay, 1995, pp. 151–185. MR 97d:32042
[Hit95b] , Twistor spaces, Einstein metrics and isomonodromic deformations, J. Dif-
ferential Geom. 42 (1995), no. 1, 30–112. MR 96g:53057
588 BIBLIOGRAPHY

[Hit96] , A new family of Einstein metrics, Manifolds and geometry (Pisa, 1993),
Sympos. Math., XXXVI, Cambridge Univ. Press, Cambridge, 1996, pp. 190–222.
MR 97d:53050
[HKK+ 03] K. Hori, S. Katz, A. Klemm, R. Pandharipande, R. Thomas, C. Vafa, R. Vakil,
and E. Zaslow, Mirror symmetry, Clay Mathematics Monographs, vol. 1, American
Mathematical Society, Providence, RI, 2003, With a preface by Vafa. MR 2003030
(2004g:14042)
[HKLR87] N. J. Hitchin, A. Karlhede, U. Lindström, and M. Roček, Hyper-Kähler metrics and
supersymmetry, Comm. Math. Phys. 108 (1987), no. 4, 535–589. MR 88g:53048
[HKW05] A. Hanany, P. Kazakopoulos, and B. Wecht, A new infinite class of quiver gauge
theories, J. High Energy Phys. (2005), no. 8, 054, 30 pp. (electronic). MR 2166231
(2006h:81286)
[HL82] R. Harvey and H. B. Lawson, Jr., Calibrated geometries, Acta Math. 148 (1982),
47–157. MR 666108 (85i:53058)
[HM04] Andre Henriques and David S. Metzler, Presentations of noneffective orbifolds,
Trans. Amer. Math. Soc. 356 (2004), no. 6, 2481–2499 (electronic). MR MR2048526
[Hoe07] C. A. Hoelscher, Classification of cohomogeneity one manifolds in low dimensions,
University of Pennsylvania Ph.D. Thesis.
[Hon06] K. Honda, The topology and geometry of contact structures in dimension three,
math.GT/0601144 (2006).
[HOT63] Y. Hatakeyama, Y. Ogawa, and S. Tanno, Some properties of manifolds with contact
metric structure, Tôhoku Math. J. (2) 15 (1963), 42–48. MR 26 #4295
[HP96] P. S. Howe and G. Papadopoulos, Twistor spaces for hyper-Kähler manifolds with
torsion, Phys. Lett. B 379 (1996), no. 1-4, 80–86. MR 1396267 (97h:53073)
[HS84] P. S. Howe and G. Sierra, Two-dimensional supersymmetric nonlinear σ-models
with torsion, Phys. Lett. B 148 (1984), no. 6, 451–455. MR 769268 (86g:81121)
[HS91] A. Haefliger and É. Salem, Actions of tori on orbifolds, Ann. Global Anal. Geom.
9 (1991), no. 1, 37–59. MR 92f:57047
[HS02] T. Hausel and B. Sturmfels, Toric hyperKähler varieties, Doc. Math. 7 (2002),
495–534 (electronic). MR 2015052 (2004i:53054)
[HS82] I. Hasegawa and M. Seino, Some remarks on Sasakian geometry—applications of
Myers’ theorem and the canonical affine connection, J. Hokkaido Univ. Ed. Sect.
II A 32 (1981/82), no. 1, 1–7. MR 84j:53055
[HSY04] Y. Hashimoto, M. Sakaguchi, and Y. Yasui, Sasaki-Einstein twist of Kerr-AdS black
holes, Phys. Lett. B 600 (2004), no. 3-4, 270–274. MR 2093038 (2005h:83235)
[HSY05] , New infinite series of Einstein metrics on sphere bundles from AdS black
holes, Comm. Math. Phys. 257 (2005), no. 2, 273–285. MR 2164598
[Hus66] D. Husemoller, Fibre bundles, McGraw-Hill Book Co., New York, 1966. MR 37
#4821
[HZ74] F. Hirzebruch and D. Zagier, The Atiyah-Singer theorem and elementary number
theory, Publish or Perish Inc., Boston, Mass., 1974, Mathematics Lecture Series,
No. 3. MR 58 #31291
[IF00] A. R. Iano-Fletcher, Working with weighted complete intersections, Explicit bira-
tional geometry of 3-folds, London Math. Soc. Lecture Note Ser., vol. 281, Cam-
bridge Univ. Press, Cambridge, 2000, pp. 101–173. MR 2001k:14089
[IK72] S. Ishihara and M. Konishi, Fibred Riemannian spaces with Sasakian 3-structure,
Differential geometry (in honor of Kentaro Yano), Kinokuniya, Tokyo, 1972,
pp. 179–194. MR 48 #4964
[IP99] V. A. Iskovskikh and Yu. G. Prokhorov, Fano varieties, Algebraic geometry, V,
Encyclopaedia Math. Sci., vol. 47, Springer, Berlin, 1999, pp. 1–247. MR 1668579
(2000b:14051b)
[Ish73] S. Ishihara, Quaternion Kählerian manifolds and fibred Riemannian spaces with
Sasakian 3-structure, Kōdai Math. Sem. Rep. 25 (1973), 321–329. MR 48 #2943
[Ish74] , Quaternion Kählerian manifolds, J. Differential Geometry 9 (1974), 483–
500. MR 50 #1184
[Isk77] V. A. Iskovskih, Fano threefolds. I, Izv. Akad. Nauk SSSR Ser. Mat. 41 (1977),
no. 3, 516–562, 717. MR 80c:14023a
BIBLIOGRAPHY 589

[Isk78] , Fano threefolds. II, Izv. Akad. Nauk SSSR Ser. Mat. 42 (1978), no. 3,
506–549. MR 80c:14023b
[Isk79] , Anticanonical models of three-dimensional algebraic varieties, Current
problems in mathematics, Vol. 12 (Russian); (English translation: J. Soviet Math.
13 (1980), 745-814), VINITI, Moscow, 1979, pp. 59–157, 239 (loose errata).
MR 537685 (81i:14026b)
[Ito84] M. Itoh, Self-duality of Kähler surfaces, Compositio Math. 51 (1984), no. 2, 265–
273. MR 739738 (85m:53079)
[Ito97] , Odd-dimensional tori and contact structure, Proc. Japan Acad. Ser. A
Math. Sci. 73 (1997), no. 4, 58–59. MR 98h:53048
[IY82] H. Inoue and K. Yano, The Gromov invariant of negatively curved manifolds, Topol-
ogy 21 (1982), no. 1, 83–89. MR 82k:53091
[Jac90] H. Jacobowitz, An introduction to CR structures, Mathematical Surveys and
Monographs, vol. 32, American Mathematical Society, Providence, RI, 1990.
MR 93h:32023
[Jel01] W. Jelonek, Positive and negative 3-K-contact structures, Proc. Amer. Math. Soc.
129 (2001), no. 1, 247–256. MR 1694866 (2001c:53061)
[Jen73] G. R. Jensen, Einstein metrics on principal fibre bundles, J. Differential Geometry
8 (1973), 599–614. MR 50 #5694
[JK01a] J. M. Johnson and J. Kollár, Fano hypersurfaces in weighted projective 4-spaces,
Experiment. Math. 10 (2001), no. 1, 151–158. MR 2002a:14048
[JK01b] , Kähler-Einstein metrics on log del Pezzo surfaces in weighted projective
3-spaces, Ann. Inst. Fourier (Grenoble) 51 (2001), no. 1, 69–79. MR 2002b:32041
[Jos01] J. Jost, Bosonic strings: a mathematical treatment, AMS/IP Studies in Ad-
vanced Mathematics, vol. 21, American Mathematical Society, Providence, RI, 2001.
MR 1999211 (2004h:81002)
[Joy91] D. D. Joyce, The hypercomplex quotient and the quaternionic quotient, Math. Ann.
290 (1991), no. 2, 323–340. MR 92f:53052
[Joy92] , Compact hypercomplex and quaternionic manifolds, J. Differential Geom.
35 (1992), no. 3, 743–761. MR 1163458 (93g:53064)
[Joy95] , Explicit construction of self-dual 4-manifolds, Duke Math. J. 77 (1995),
no. 3, 519–552. MR 1324633 (96d:53049)
[Joy96a] , Compact 8-manifolds with holonomy Spin(7), Invent. Math. 123 (1996),
no. 3, 507–552. MR 1383960 (97d:53052)
[Joy96b] , Compact Riemannian 7-manifolds with holonomy G2 . I, II, J. Differential
Geom. 43 (1996), no. 2, 291–328, 329–375. MR 1424428 (97m:53084)
[Joy99] , A new construction of compact 8-manifolds with holonomy Spin(7), J.
Differential Geom. 53 (1999), no. 1, 89–130. MR 1776092 (2002a:53063)
[Joy00] , Compact manifolds with special holonomy, Oxford Mathematical Mono-
graphs, Oxford University Press, Oxford, 2000. MR 2001k:53093
[Käh33] E. Kähler, Über eine bemerkenswerte Hermitesche Metrik, Abh. Math. Sem. Ham-
burg Univ. 9 (1933), 173–186.
[Käh03] , Mathematische Werke/Mathematical works, Walter de Gruyter & Co.,
Berlin, 2003, Edited by Rolf Berndt and Oswald Riemenschneider. MR 2229234
[Kar03] Y. Karshon, Maximal tori in the symplectomorphism groups of Hirzebruch surfaces,
Math. Res. Lett. 10 (2003), no. 1, 125–132. MR 1960129 (2004f:53101)
[Kas71] T. Kashiwada, A note on a Riemannian space with Sasakian 3-structure, Natur.
Sci. Rep. Ochanomizu Univ. 22 (1971), 1–2. MR 46 #2586
[Kas01] , On a contact 3-structure, Math. Z. 238 (2001), no. 4, 829–832.
MR 2002h:53136
[Kaw78] T. Kawasaki, The signature theorem for V -manifolds, Topology 17 (1978), no. 1,
75–83. MR 57 #14072
[Kaw79] , The Riemann-Roch theorem for complex V -manifolds, Osaka J. Math. 16
(1979), no. 1, 151–159. MR 80f:58042
[Kaw81] , The index of elliptic operators over V -manifolds, Nagoya Math. J. 84
(1981), 135–157. MR 83i:58095
[Keh04] W. Kehowski, Hypernormal manifolds, University of New Mexico Thesis (2004).
590 BIBLIOGRAPHY

[Kir84] F. Kirwan, Cohomology of quotients in symplectic and algebraic geometry, Mathe-


matical Notes, vol. 31, Princeton University Press, Princeton, NJ, 1984. MR 766741
(86i:58050)
[Kir98] , Momentum maps and reduction in algebraic geometry, Differential Geom.
Appl. 9 (1998), no. 1-2, 135–171, Symplectic geometry. MR 1636303 (99e:58072)
[KKS78] D. Kazhdan, B. Kostant, and S. Sternberg, Hamiltonian group actions and dynam-
ical systems of Calogero type, Comm. Pure Appl. Math. 31 (1978), no. 4, 481–507.
MR 0478225 (57 #17711)
[KL06] B. Kleiner and J. Lott, Notes on Perelman’s papers, preprint;
arXiv:math.DG/0605667 (2006).
[KLP97] J. Kim, C. LeBrun, and M. Pontecorvo, Scalar-flat Kähler surfaces of all genera,
J. Reine Angew. Math. 486 (1997), 69–95. MR 1450751 (98m:32045)
[KM63] M. A. Kervaire and J. W. Milnor, Groups of homotopy spheres. I, Ann. of Math.
(2) 77 (1963), 504–537. MR 26 #5584
[KM98] J. Kollár and S. Mori, Birational geometry of algebraic varieties, Cambridge Tracts
in Mathematics, vol. 134, Cambridge University Press, Cambridge, 1998, With the
collaboration of C. H. Clemens and A. Corti, Translated from the 1998 Japanese
original. MR 2000b:14018
[KM99] S. Keel and J. McKernan, Rational curves on quasi-projective surfaces, Mem. Amer.
Math. Soc. 140 (1999), no. 669, viii+153. MR 1610249 (99m:14068)
[KMM92] J. Kollár, Y. Miyaoka, and S. Mori, Rational connectedness and boundedness of
Fano manifolds, J. Differential Geom. 36 (1992), no. 3, 765–779. MR 94g:14021
[KN63] S. Kobayashi and K. Nomizu, Foundations of differential geometry. Vol I, Inter-
science Publishers, a division of John Wiley & Sons, New York-Lond on, 1963.
MR 27 #2945
[KN69] , Foundations of differential geometry. Vol. II, Interscience Tracts in Pure
and Applied Mathematics, No. 15 Vol. II, Interscience Publishers John Wiley &
Sons, Inc., New York-London-Sydney, 1969. MR MR0238225 (38 #6501)
[KN77] L. H. Kauffman and W. D. Neumann, Products of knots, branched fibrations and
sums of singularities, Topology 16 (1977), no. 4, 369–393. MR 0488073 (58 #7644)
[KN90] P. B. Kronheimer and H. Nakajima, Yang-Mills instantons on ALE gravitational
instantons, Math. Ann. 288 (1990), no. 2, 263–307. MR 1075769 (92e:58038)
[KO73] S. Kobayashi and T. Ochiai, Characterizations of complex projective spaces and
hyperquadrics, J. Math. Kyoto Univ. 13 (1973), 31–47. MR 47 #5293
[Kob59a] S. Kobayashi, On the automorphism group of a certain class of algebraic manifolds.,
Tôhoku Math. J. (2) 11 (1959), 184–190. MR 0112159 (22 #3014)
[Kob59b] , Remarks on complex contact manifolds, Proc. Amer. Math. Soc. 10 (1959),
164–167. MR 0111061 (22 #1925)
[Kob61] , On compact Kähler manifolds with positive definite Ricci tensor, Ann. of
Math. (2) 74 (1961), 570–574. MR 24 #A2922
[Kob63] , Topology of positively pinched Kaehler manifolds, Tôhoku Math. J. (2) 15
(1963), 121–139. MR 27 #4185
[Kob72] , Transformation groups in differential geometry, Springer-Verlag, New
York, 1972, Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 70. MR 50
#8360
[Kob87] , Differential geometry of complex vector bundles, Publications of the Math-
ematical Society of Japan, vol. 15, Princeton University Press, Princeton, NJ, 1987,
Kanô Memorial Lectures, 5. MR 89e:53100
[Kod60a] K. Kodaira, On compact complex analytic surfaces. I, Ann. of Math. (2) 71 (1960),
111–152. MR 24 #A2396
[Kod60b] , On deformations of some complex psuedo-group structures, Ann. of Math.
(2) 71 (1960), 224–302. MR 22 #5992
[Kod86] , Complex manifolds and deformation of complex structures, Grundlehren
der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sci-
ences], vol. 283, Springer-Verlag, New York, 1986, Translated from the Japanese by
Kazuo Akao, With an appendix by Daisuke Fujiwara. MR 87d:32040
[Kod87] T. Koda, Self-dual and anti-self-dual Hermitian surfaces, Kodai Math. J. 10 (1987),
no. 3, 335–342. MR 929993 (89a:53053)
BIBLIOGRAPHY 591

[Koi83] N. Koiso, Einstein metrics and complex structures, Invent. Math. 73 (1983), no. 1,
71–106. MR 707349 (85d:58018)
[Kol92] J. Kollár, Flips and abundance for algebraic threefolds, Société Mathématique de
France, Paris, 1992, Papers from the Second Summer Seminar on Algebraic Geom-
etry held at the University of Utah, Salt Lake City, Utah, August 1991, Astérisque
No. 211 (1992). MR 94f:14013
[Kol96] , Rational curves on algebraic varieties, Ergebnisse der Mathematik und
ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results
in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Math-
ematics], vol. 32, Springer-Verlag, Berlin, 1996. MR 98c:14001
[Kol97] , Singularities of pairs, Algebraic geometry—Santa Cruz 1995, Proc. Sym-
pos. Pure Math., vol. 62, Amer. Math. Soc., Providence, RI, 1997, pp. 221–287.
MR 99m:14033
[Kol04] , Seifert Gm -bundles, preprint; arXiv:math.AG/0404386 (2004).
[Kol05] J. Kollár, Einstein metrics on five-dimensional Seifert bundles, J. Geom. Anal. 15
(2005), no. 3, 445–476. MR MR2190241
[Kol06a] , Circle actions on simply connected 5-manifolds, Topology 45 (2006), no. 3,
643–671. MR 2218760
[Kol06b] J. Kollár, Positive Sasakian structures on 5-manifolds, preprint;
arXiv:math.DG/0612524 (2006).
[Kol07] , Einstein metrics on connected sums of S 2 × S 3 , J. Differential Geom. 77
(2007), no. 2, 259–272. MR 2269781
[Kon00] H. Konno, Cohomology rings of toric hyperkähler manifolds, Internat. J. Math. 11
(2000), no. 8, 1001–1026. MR 1797675 (2001k:53089)
[Kon75] M. Konishi, On manifolds with Sasakian 3-structure over quaternion Kaehler man-
ifolds, Kōdai Math. Sem. Rep. 26 (1974/75), 194–200. MR 51 #13951
[Kos55] B. Kostant, Holonomy and the Lie algebra of infinitesimal motions of a Riemannian
manifold, Trans. Amer. Math. Soc. 80 (1955), 528–542. MR 0084825 (18,930a)
[Kot97] D. Kotschick, Orientations and geometrisations of compact complex surfaces, Bull.
London Math. Soc. 29 (1997), no. 2, 145–149. MR 1425990 (97k:32047)
[Kot98a] , Einstein metrics and smooth structures, Geom. Topol. 2 (1998), 1–10 (elec-
tronic). MR 1489351 (99h:57060)
[Kot98b] , On the Gromov-Hitchin-Thorpe inequality, C. R. Acad. Sci. Paris Sér. I
Math. 326 (1998), no. 6, 727–731. MR 1641774 (99j:53068)
[Kou76] A. G. Kouchnirenko, Polyèdres de Newton et nombres de Milnor, Invent. Math. 32
(1976), no. 1, 1–31. MR 0419433 (54 #7454)
[Kov96] A. G. Kovalev, Nahm’s equations and complex adjoint orbits, Quart. J. Math.
Oxford Ser. (2) 47 (1996), no. 185, 41–58. MR MR1380949 (97c:53070)
[KP95] J. Kim and M. Pontecorvo, A new method of constructing scalar-flat Kähler sur-
faces, J. Differential Geom. 41 (1995), no. 2, 449–477. MR 1331974 (96c:32032)
[Kra65] V. Y. Kraines, Topology of quaternionic manifolds, Bull. Amer. Math. Soc. 71
(1965), 526–527. MR 33 #6653
[Kra66] , Topology of quaternionic manifolds, Trans. Amer. Math. Soc. 122 (1966),
357–367. MR 0192513 (33 #738)
[Kro89a] P. B. Kronheimer, The construction of ALE spaces as hyper-Kähler quotients, J.
Differential Geom. 29 (1989), no. 3, 665–683. MR 992334 (90d:53055)
[Kro89b] , A Torelli-type theorem for gravitational instantons, J. Differential Geom.
29 (1989), no. 3, 685–697. MR 992335 (90d:53056)
[Kro90a] , A hyper-Kählerian structure on coadjoint orbits of a semisimple complex
group, J. London Math. Soc. (2) 42 (1990), no. 2, 193–208. MR 1083440 (92b:53031)
[Kro90b] , Instantons and the geometry of the nilpotent variety, J. Differential Geom.
32 (1990), no. 2, 473–490. MR 1072915 (91m:58021)
[Kru97] B. Kruggel, A homotopy classification of certain 7-manifolds, Trans. Amer. Math.
Soc. 349 (1997), no. 7, 2827–2843. MR 97m:55012
[Kru98] , Kreck-Stolz invariants, normal invariants and the homotopy classification
of generalised Wallach spaces, Quart. J. Math. Oxford Ser. (2) 49 (1998), no. 196,
469–485. MR 2000a:55018
592 BIBLIOGRAPHY

[Kru05] , Homeomorphism and diffeomorphism classification of Eschenburg spaces,


Q. J. Math. 56 (2005), no. 4, 553–577. MR MR2182466 (2006h:53045)
[KS58] K. Kodaira and D. C. Spencer, On deformations of complex analytic structures. I,
II, Ann. of Math. (2) 67 (1958), 328–466. MR 22 #3009
[KS60] , On deformations of complex analytic structures. III. Stability theorems for
complex structures, Ann. of Math. (2) 71 (1960), 43–76. MR 22 #5991
[KS72] A. Kumpera and D. Spencer, Lie equations. Vol. I: General theory, Princeton Uni-
versity Press, Princeton, N.J., 1972, Annals of Mathematics Studies, No. 73. MR 52
#1805
[KS88] M. Kreck and S. Stolz, A diffeomorphism classification of 7-dimensional homoge-
neous Einstein manifolds with SU(3) × SU(2) × U(1)-symmetry, Ann. of Math. (2)
127 (1988), no. 2, 373–388. MR 89c:57042
[KS93a] P. Z. Kobak and A. Swann, Quaternionic geometry of a nilpotent variety, Math.
Ann. 297 (1993), no. 4, 747–764. MR 1245417 (94j:53057)
[KS93b] T. Koda and K. Sekigawa, Self-dual Einstein Hermitian surfaces, Progress in dif-
ferential geometry, Adv. Stud. Pure Math., vol. 22, Math. Soc. Japan, Tokyo, 1993,
pp. 123–131. MR 1274944 (95b:53056)
[KS98] P. Z. Kobak and A. Swann, Exceptional hyperkähler reductions, Twistor Newsletter
44 (1998), 23–26.
[KS00] G. Kane and M. Shifman (eds.), The supersymmetric world, World Scientific Pub-
lishing Co. Inc., River Edge, NJ, 2000, The beginnings of the theory. MR 1887931
(2003g:81253)
[KSY05] H. Kihara, M. Sakaguchi, and Y. Yasui, Scalar Laplacian on Sasaki-Einstein mani-
folds Y p,q , Phys. Lett. B 621 (2005), no. 3-4, 288–294. MR 2151750 (2006h:58038)
[KT70] Y. Y. Kuo and S. Tachibana, On the distribution appeared in contact 3-structure,
Taita J. Math. 2 (1970), 17–24. MR 46 #8115
[KT87] Franz W. Kamber and Philippe Tondeur, de Rham-Hodge theory for Riemannian
foliations, Math. Ann. 277 (1987), no. 3, 415–431. MR 89d:53070
[Kuo70] Y. Kuo, On almost contact 3-structure, Tôhoku Math. J. (2) 22 (1970), 325–332.
MR 43 #3956
[Kur71] M. Kuranishi, Deformations of compact complex manifolds, Les Presses de
l’Université de Montréal, Montreal, Que., 1971, Séminaire de Mathématiques
Supérieures, No. 39 (Été 1969). MR 50 #7588
[KW99] I. R. Klebanov and E. Witten, Superconformal field theory on threebranes at a
Calabi-Yau singularity, Nuclear Phys. B 536 (1999), no. 1-2, 199–218. MR 1666725
(99k:81253)
[KY02a] H. Kanno and Y. Yasui, On Spin(7) holonomy metric based on SU(3)/U(1). I, J.
Geom. Phys. 43 (2002), no. 4, 293–309. MR 1929908 (2003m:53076)
[KY02b] , On Spin(7) holonomy metric based on SU(3)/U(1). II, J. Geom. Phys. 43
(2002), no. 4, 310–326. MR 1929909 (2003m:53077)
[Lan00] T. Lance, Differentiable structures on manifolds, Surveys on surgery theory, Vol. 1,
Ann. of Math. Stud., vol. 145, Princeton Univ. Press, Princeton, NJ, 2000, pp. 73–
104. MR 2001d:57035
[Laz04a] R. Lazarsfeld, Positivity in algebraic geometry. I, Ergebnisse der Mathematik und
ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in
Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Math-
ematics], vol. 48, Springer-Verlag, Berlin, 2004, Classical setting: line bundles and
linear series. MR 2095471 (2005k:14001a)
[Laz04b] , Positivity in algebraic geometry. II, Ergebnisse der Mathematik und ihrer
Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in
Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Math-
ematics], vol. 49, Springer-Verlag, Berlin, 2004, Positivity for vector bundles, and
multiplier ideals. MR 2095472 (2005k:14001b)
[LB70] A. Lascoux and M. Berger, Variétés Kähleriennes compactes, Lecture Notes in
Mathematics, Vol. 154, Springer-Verlag, Berlin, 1970. MR 0278248 (43 #3979)
[LB92] H. Lange and Ch. Birkenhake, Complex abelian varieties, Grundlehren der Mathe-
matischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol.
302, Springer-Verlag, Berlin, 1992. MR 94j:14001
BIBLIOGRAPHY 593

[LdM71] S. López de Medrano, Involutions on manifolds, Springer-Verlag, New York, 1971,


Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 59. MR 0298698 (45
#7747)
[LeB88] C. LeBrun, A rigidity theorem for quaternionic-Kähler manifolds, Proc. Amer.
Math. Soc. 103 (1988), no. 4, 1205–1208. MR 89h:53105
[LeB91a] , Complete Ricci-flat Kähler metrics on Cn need not be flat, Several com-
plex variables and complex geometry, Part 2 (Santa Cruz, CA, 1989), Proc. Sym-
pos. Pure Math., vol. 52, Amer. Math. Soc., Providence, RI, 1991, pp. 297–304.
MR 1128554 (93a:53038)
[LeB91b] , Explicit self-dual metrics on CP2 # · · · #CP2 , J. Differential Geom. 34
(1991), no. 1, 223–253. MR 1114461 (92g:53040)
[LeB91c] , On complete quaternionic-Kähler manifolds, Duke Math. J. 63 (1991),
no. 3, 723–743. MR 92i:53042
[LeB91d] , Scalar-flat Kähler metrics on blown-up ruled surfaces, J. Reine Angew.
Math. 420 (1991), 161–177. MR 1124569 (92i:53066)
[LeB93] , A finiteness theorem for quaternionic-Kähler manifolds with positive scalar
curvature, The Penrose transform and analytic cohomology in representation the-
ory (South Hadley, MA, 1992), Contemp. Math., vol. 154, Amer. Math. Soc., Prov-
idence, RI, 1993, pp. 89–101. MR 1 246 379
[LeB95] , Fano manifolds, contact structures, and quaternionic geometry, Internat.
J. Math. 6 (1995), no. 3, 419–437. MR 96c:53108
[LeB96] , Four-manifolds without Einstein metrics, Math. Res. Lett. 3 (1996), no. 2,
133–147. MR 1386835 (97a:53072)
[LeB99] , Four-dimensional Einstein manifolds, and beyond, Surveys in differential
geometry: essays on Einstein manifolds, Surv. Differ. Geom., VI, Int. Press, Boston,
MA, 1999, pp. 247–285. MR 1798613 (2001m:53072)
[LeB01] , Ricci curvature, minimal volumes, and Seiberg-Witten theory, Invent.
Math. 145 (2001), no. 2, 279–316. MR 1872548 (2002h:53061)
[LeB03] , Einstein metrics, four-manifolds, and differential topology, Surveys in dif-
ferential geometry, Vol. VIII (Boston, MA, 2002), Surv. Differ. Geom., VIII, Int.
Press, Somerville, MA, 2003, pp. 235–255. MR 2039991 (2005g:53078)
[Lee96] J. M. Lee, CR manifolds with noncompact connected automorphism groups, J.
Geom. Anal. 6 (1996), no. 1, 79–90. MR 1402387 (97h:32013)
[Ler02a] E. Lerman, Contact toric manifolds, J. Symplectic Geom. 1 (2002), no. 4, 785–828.
MR 2 039 164
[Ler02b] , A convexity theorem for torus actions on contact manifolds, Illinois J.
Math. 46 (2002), no. 1, 171–184. MR 2003g:53146
[Ler03a] , Geodesic flows and contact toric manifolds, Symplectic geometry of
integrable Hamiltonian systems (Barcelona, 2001), Adv. Courses Math. CRM
Barcelona, Birkhäuser, Basel, 2003, pp. 175–225. MR 2 000 747
[Ler03b] , Maximal tori in the contactomorphism groups of circle bundles over Hirze-
bruch surfaces, Math. Res. Lett. 10 (2003), no. 1, 133–144. MR 2004g:53097
[Ler04a] , Contact fiber bundles, J. Geom. Phys. 49 (2004), no. 1, 52–66. MR 2077244
[Ler04b] , Homotopy groups of K-contact toric manifolds, Trans. Amer. Math. Soc.
356 (2004), no. 10, 4075–4083 (electronic). MR 2 058 839
[Lib54] P. Libermann, Sur le problème d’équivalence de certaines structures infinitésimales,
Ann. Mat. Pura Appl. (4) 36 (1954), 27–120. MR 0066020 (16,520c)
[Lic53] A. Lichnerowicz, Sur les espaces homogènes kählériens, C. R. Acad. Sci. Paris 237
(1953), 695–697. MR 0066016 (16,519b)
[Lic57] , Sur les transformations analytiques des variétés kählériennes compactes,
C. R. Acad. Sci. Paris 244 (1957), 3011–3013. MR 0094479 (20 #996)
[Lic63a] , Spineurs harmoniques, C. R. Acad. Sci. Paris 257 (1963), 7–9. MR 0156292
(27 #6218)
[Lic63b] , Théorèmes de réductivité sur des algèbres d’automorphismes, Rend. Mat.
e Appl. (5) 22 (1963), 197–244. MR 0157329 (28 #564)
[Lie77] S. Lie, Geometrie der Berührungs transformationen, corrected ed., Chelsea Publish-
ing Co., New York, 1977, With editorial assistance by Georg Scheffers. MR 0460049
(57 #45)
594 BIBLIOGRAPHY

[LM76] R. Lutz and C. Meckert, Structures de contact sur certaines sphères exotiques, C.
R. Acad. Sci. Paris Sér. A-B 282 (1976), no. 11, Aii, A591–A593. MR 0397612 (53
#1471)
[LM87] P. Libermann and Ch.-M. Marle, Symplectic geometry and analytical mechanics,
Mathematics and its Applications, vol. 35, D. Reidel Publishing Co., Dordrecht,
1987, Translated from the French by Bertram Eugene Schwarzbach. MR 88c:58016
[LM89] H. B. Lawson, Jr. and M.-L. Michelsohn, Spin geometry, Princeton Mathematical
Series, vol. 38, Princeton University Press, Princeton, NJ, 1989. MR 91g:53001
[Loo84] E. J. N. Looijenga, Isolated singular points on complete intersections, London Math-
ematical Society Lecture Note Series, vol. 77, Cambridge University Press, Cam-
bridge, 1984. MR 747303 (86a:32021)
[Loo01] F. Loose, Reduction in contact geometry, J. Lie Theory 11 (2001), no. 1, 9–22.
MR 1828281 (2002g:53147)
[LPP04] H. Lü, D. N. Page, and C. N. Pope, New inhomogeneous Einstein metrics on sphere
bundles over Einstein-Kähler manifolds, Phys. Lett. B 593 (2004), no. 1-4, 218–226.
MR 2076718 (2005f:53063)
[LR83] U. Lindström and M. Roček, Scalar tensor duality and N = 1, 2 nonlinear σ-
models, Nuclear Phys. B 222 (1983), no. 2, 285–308. MR 710273 (85d:81113)
[LS93] C. LeBrun and M. Singer, Existence and deformation theory for scalar-flat Kähler
metrics on compact complex surfaces, Invent. Math. 112 (1993), no. 2, 273–313.
MR 1213104 (94e:53070)
[LS94] C. LeBrun and S. Salamon, Strong rigidity of positive quaternion-Kähler manifolds,
Invent. Math. 118 (1994), no. 1, 109–132. MR 95k:53059
[LT97] E. Lerman and S. Tolman, Hamiltonian torus actions on symplectic orbifolds
and toric varieties, Trans. Amer. Math. Soc. 349 (1997), no. 10, 4201–4230.
MR 98a:57043
[Lut79] R. Lutz, Sur la géométrie des structures de contact invariantes, Ann. Inst. Fourier
(Grenoble) 29 (1979), no. 1, xvii, 283–306. MR 82j:53067
[LW99] C. LeBrun and M. Wang (eds.), Surveys in differential geometry: essays on Einstein
manifolds, Surveys in Differential Geometry, VI, International Press, Boston, MA,
1999, Lectures on geometry and topology, sponsored by Lehigh University’s Journal
of Differential Geometry. MR 1798603 (2001f:53003)
[LW01] E. Lerman and C. Willett, The topological structure of contact and symplectic
quotients, Internat. Math. Res. Notices (2001), no. 1, 33–52. MR 2001j:53112
[Mab87] T. Mabuchi, Einstein-Kähler forms, Futaki invariants and convex geometry on
toric Fano varieties, Osaka J. Math. 24 (1987), no. 4, 705–737. MR 89e:53074
[Mac87] K. Mackenzie, Lie groupoids and Lie algebroids in differential geometry, London
Mathematical Society Lecture Note Series, vol. 124, Cambridge University Press,
Cambridge, 1987. MR 89g:58225
[Mal99] J. Maldacena, The large-N limit of superconformal field theories and supergravity,
Internat. J. Theoret. Phys. 38 (1999), no. 4, 1113–1133, Quantum gravity in the
southern cone (Bariloche, 1998). MR 2001h:81246
[Man93] L. Manivel, Un théorème de prolongement L2 de sections holomorphes d’un fibré
hermitien, Math. Z. 212 (1993), no. 1, 107–122. MR 94e:32050
[Mar58] A. Markov, Unsolvability of certain problems in topology, Dokl. Akad. Nauk SSSR
123 (1958), 978–980. MR 0103455 (21 #2224)
[Mar59] E. Martinelli, Varietà a struttura quaternionale generalizzata, Atti Accad. Naz.
Lincei. Rend. Cl. Sci. Fis. Mat. Nat. (8) 26 (1959), 353–362. MR 0144288 (26
#1835)
[Mar60] A. A. Markov, Insolubility of the problem of homeomorphy, Proc. Internat. Congress
Math. 1958, Cambridge Univ. Press, New York, 1960, pp. 300–306. MR 0115160
(22 #5962)
[Mar70] S. Marchiafava, Sulle varietà a struttura quaternionale generalizzata, Rend. Mat.
(6) 3 (1970), 529–545. MR 0276892 (43 #2632)
[Mar71] J. Martinet, Formes de contact sur les variétés de dimension 3, Proceedings of Liv-
erpool Singularities Symposium, II (1969/1970) (Berlin), Springer, 1971, pp. 142–
163. Lecture Notes in Math., Vol. 209. MR 0350771 (50 #3263)
BIBLIOGRAPHY 595

[Mat57a] Y. Matsushima, Sur la structure du groupe d’homéomorphismes analytiques d’une


certaine variété kählérienne, Nagoya Math. J. 11 (1957), 145–150. MR 0094478
(20 #995)
[Mat57b] , Sur les espaces homogènes kählériens d’un groupe de Lie réductif, Nagoya
Math. J. 11 (1957), 53–60. MR 19,315c
[Mat72] M. Matsumoto, On 6-dimensional almost Tachibana spaces, Tensor (N.S.) 23
(1972), 250–252. MR 0312431 (47 #989)
[Mat02] K. Matsuki, Introduction to the Mori program, Universitext, Springer-Verlag, New
York, 2002. MR MR1875410 (2002m:14011)
[McD84] D. McDuff, Examples of simply-connected symplectic non-Kählerian manifolds, J.
Differential Geom. 20 (1984), no. 1, 267–277. MR 772133 (86c:57036)
[McG77] D. McGavran, T n -actions on simply connected (n + 2)-manifolds, Pacific J. Math.
71 (1977), no. 2, 487–497. MR 0461542 (57 #1527)
[McK80] J. McKay, Graphs, singularities, and finite groups, The Santa Cruz Conference on
Finite Groups (Univ. California, Santa Cruz, Calif., 1979), Proc. Sympos. Pure
Math., vol. 37, Amer. Math. Soc., Providence, R.I., 1980, pp. 183–186. MR 604577
(82e:20014)
[McK81] , Cartan matrices, finite groups of quaternions, and Kleinian singularities,
Proc. Amer. Math. Soc. 81 (1981), no. 1, 153–154. MR 589160 (81j:20018)
[Mey73] K. R. Meyer, Symmetries and integrals in mechanics, Dynamical systems (Proc.
Sympos., Univ. Bahia, Salvador, 1971), Academic Press, New York, 1973, pp. 259–
272. MR 0331427 (48 #9760)
[MFK94] D. Mumford, J. Fogarty, and F. Kirwan, Geometric invariant theory, third ed.,
Ergebnisse der Mathematik und ihrer Grenzgebiete (2) [Results in Mathematics and
Related Areas (2)], vol. 34, Springer-Verlag, Berlin, 1994. MR 1304906 (95m:14012)
[Mil56a] J. Milnor, Construction of universal bundles. II, Ann. of Math. (2) 63 (1956),
430–436. MR 17,1120a
[Mil56b] , On manifolds homeomorphic to the 7-sphere, Ann. of Math. (2) 64 (1956),
399–405. MR 18,498d
[Mil63] , Morse theory, Based on lecture notes by M. Spivak and R. Wells. Annals
of Mathematics Studies, No. 51, Princeton University Press, Princeton, N.J., 1963.
MR 29 #634
[Mil65] , Lectures on the h-cobordism theorem, Notes by L. Siebenmann and J.
Sondow, Princeton University Press, Princeton, N.J., 1965. MR 0190942 (32 #8352)
[Mil68] , Singular points of complex hypersurfaces, Annals of Mathematics Studies,
No. 61, Princeton University Press, Princeton, N.J., 1968. MR 39 #969
[Mil75] , On the 3-dimensional Brieskorn manifolds M (p, q, r), Knots, groups, and
3-manifolds (Papers dedicated to the memory of R. H. Fox), Princeton Univ. Press,
Princeton, N. J., 1975, pp. 175–225. Ann. of Math. Studies, No. 84. MR 54 #6169
[Mil00] R. J. Milgram, The classification of Aloff-Wallach manifolds and their generaliza-
tions, Surveys on surgery theory, Vol. 1, Ann. of Math. Stud., vol. 145, Princeton
Univ. Press, Princeton, NJ, 2000, pp. 379–407. MR 2000m:57056
[MM64] H. Matsumura and P. Monsky, On the automorphisms of hypersurfaces, J. Math.
Kyoto Univ. 3 (1963/1964), 347–361. MR 0168559 (29 #5819)
[MM03] I. Moerdijk and J. Mrčun, Introduction to foliations and Lie groupoids, Cambridge
Studies in Advanced Mathematics, vol. 91, Cambridge University Press, Cambridge,
2003. MR 2012261
[MM82] S. Mori and S. Mukai, Classification of Fano 3-folds with B2 ≥ 2, Manuscripta
Math. 36 (1981/82), no. 2, 147–162. MR 83f:14032
[MNS05] A. Moroianu, P.-A. Nagy, and U. Semmelmann, Unit Killing vector fields on nearly
Kähler manifolds, Internat. J. Math. 16 (2005), no. 3, 281–301. MR 2130627
(2006e:53056)
[MO70] J. Milnor and P. Orlik, Isolated singularities defined by weighted homogeneous poly-
nomials, Topology 9 (1970), 385–393. MR 45 #2757
[MO00] P. Matzeu and L. Ornea, Local almost contact metric 3-structures, Publ. Math.
Debrecen 57 (2000), no. 3-4, 499–508. MR 1798730 (2002a:53057)
[MO07] A. Moroianu and L. Ornea, Conformally Einstein products and nearly Kähler man-
ifolds, preprint; arXiv:math.DG/0610599 (2007).
596 BIBLIOGRAPHY

[Moe02] I. Moerdijk, Orbifolds as groupoids: an introduction, Orbifolds in mathematics


and physics (Madison, WI, 2001), Contemp. Math., vol. 310, Amer. Math. Soc.,
Providence, RI, 2002, pp. 205–222. MR 2004c:22003
[Mol75] P. Molino, Sur la géométrie transverse des feuilletages, Ann. Inst. Fourier (Greno-
ble) 25 (1975), no. 2, xiv, 279–284. MR 52 #11945
[Mol77] , Théorie des G-structures: le problème d’équivalence, Springer-Verlag,
Berlin, 1977, Notes rédigées avec la collaboration de F. Toupine, Lecture Notes
in Mathematics, Vol. 588. MR 58 #24419
[Mol79] , Feuilletages riemanniens sur les variétés compactes; champs de Killing
transverses, C. R. Acad. Sci. Paris Sér. A-B 289 (1979), no. 7, A421–A423.
MR 80j:57026
[Mol88] , Riemannian foliations, Progress in Mathematics, vol. 73, Birkhäuser
Boston Inc., Boston, MA, 1988, Translated from the French by Grant Cairns,
With appendices by Cairns, Y. Carrière, É. Ghys, E. Salem and V. Sergiescu.
MR 89b:53054
[Mor75] S. Morita, A topological classification of complex structures on S 1 × S 2n−1 , Topol-
ogy 14 (1975), 13–22. MR 0405444 (53 #9237)
[Mor97] A. Moroianu, Parallel and Killing spinors on Spinc manifolds, Comm. Math. Phys.
187 (1997), no. 2, 417–427. MR 98i:58245
[Mos57] P. S. Mostert, On a compact Lie group acting on a manifold, Ann. of Math. (2) 65
(1957), 447–455. MR 0085460 (19,44b)
[Mos65] J. Moser, On the volume elements on a manifold, Trans. Amer. Math. Soc. 120
(1965), 286–294. MR 32 #409
[Mos66] E. Moskal, Contact manifolds of positive curvature, University of Illinois Thesis
(1966).
[MP97] I. Moerdijk and D. A. Pronk, Orbifolds, sheaves and groupoids, K-Theory 12
(1997), no. 1, 3–21. MR 98i:22004
[MR75] S. Marchiafava and G. Romani, Sui fibrati con struttura quaternionale generalizzata,
Ann. Mat. Pura Appl. (4) 107 (1975), 131–157 (1976). MR 53 #6558
[MS39] S. B. Myers and N. E. Steenrod, The group of isometries of a Riemannian manifold,
Ann. of Math. (2) 40 (1939), no. 2, 400–416. MR 1503467
[MS74] J. W. Milnor and J. D. Stasheff, Characteristic classes, Princeton University Press,
Princeton, N. J., 1974, Annals of Mathematics Studies, No. 76. MR 0440554 (55
#13428)
[MS85] P. Molino and V. Sergiescu, Deux remarques sur les flots riemanniens, Manuscripta
Math. 51 (1985), no. 1-3, 145–161. MR 86h:53035
[MS88] C. C. Moore and C. Schochet, Global analysis on foliated spaces, Mathemati-
cal Sciences Research Institute Publications, vol. 9, Springer-Verlag, New York,
1988, With appendices by S. Hurder, Moore, Schochet and Robert J. Zimmer.
MR 89h:58184
[MS98] D. McDuff and D. Salamon, Introduction to symplectic topology, second ed., Oxford
Mathematical Monographs, The Clarendon Press Oxford University Press, New
York, 1998. MR 2000g:53098
[MS99] S. Merkulov and L. Schwachhöfer, Classification of irreducible holonomies of
torsion-free affine connections, Ann. of Math. (2) 150 (1999), no. 1, 77–149.
MR 1715321 (2000m:53065a)
[MS00] A. Moroianu and U. Semmelmann, Parallel spinors and holonomy groups, J. Math.
Phys. 41 (2000), no. 4, 2395–2402. MR 1751897 (2001f:53099)
[MS05] D. Martelli and J. Sparks, Toric Sasaki-Einstein metrics on S 2 × S 3 , Phys. Lett.
B 621 (2005), no. 1-2, 208–212. MR 2152673
[MS06] , Toric geometry, Sasaki-Einstein manifolds and a new infinite class of
AdS/CFT duals, Comm. Math. Phys. 262 (2006), no. 1, 51–89. MR 2200882
[MSY56] D. Montgomery, H. Samelson, and C. T. Yang, Groups on E n with (n − 2)-
dimensional orbits, Proc. Amer. Math. Soc. 7 (1956), 719–728. MR 0078643
(17,1224a)
[MSY06a] D. Martelli, J. Sparks, and S.-T. Yau, The geometric dual of a-maximisation for
toric Sasaki-Einstein manifolds, Comm. Math. Phys. 268 (2006), no. 1, 39–65.
MR 2249795
BIBLIOGRAPHY 597

[MSY06b] , Sasaki-Einstein manifolds and volume minimisation, preprint; arXiv:hep-


th/0603021 (2006).
[MT06] J. W. Morgan and G. Tian, Ricci flow and the Poincare Conjecture, preprint;
arXiv:math.DG/0607607 (2006).
[MU83] S. Mukai and H. Umemura, Minimal rational threefolds, Algebraic geometry
(Tokyo/Kyoto, 1982), Lecture Notes in Math., vol. 1016, Springer, Berlin, 1983,
pp. 490–518. MR 726439 (85c:14027)
[Mum61] D. Mumford, The topology of normal singularities of an algebraic surface and a
criterion for simplicity, Inst. Hautes Études Sci. Publ. Math. (1961), no. 9, 5–22.
MR 0153682 (27 #3643)
[Mum79] , An algebraic surface with K ample, (K 2 ) = 9, pg = q = 0, Amer. J. Math.
101 (1979), no. 1, 233–244. MR 527834 (80j:14032)
[MW74] J. Marsden and A. Weinstein, Reduction of symplectic manifolds with symmetry,
Rep. Mathematical Phys. 5 (1974), no. 1, 121–130. MR 0402819 (53 #6633)
[MW88] K. Mikami and A. Weinstein, Moments and reduction for symplectic groupoids,
Publ. Res. Inst. Math. Sci. 24 (1988), no. 1, 121–140. MR 944869 (90c:58060)
[MY06] G. Marinescu and N. Yeganefar, Embeddability of some strongly pseudoconvex CR
manifolds, math.CV/0403044 (2006).
[MZ55] D. Montgomery and L. Zippin, Topological transformation groups, Interscience Pub-
lishers, New York-London, 1955. MR 0073104 (17,383b)
[MZ88] M. Miyanishi and D.-Q. Zhang, Gorenstein log del Pezzo surfaces of rank one, J.
Algebra 118 (1988), no. 1, 63–84. MR 961326 (89i:14033)
[Nad89] A. M. Nadel, Multiplier ideal sheaves and existence of Kähler-Einstein metrics of
positive scalar curvature, Proc. Nat. Acad. Sci. U.S.A. 86 (1989), no. 19, 7299–7300.
MR 1015491 (90k:32061)
[Nad90] , Multiplier ideal sheaves and Kähler-Einstein metrics of positive scalar
curvature, Ann. of Math. (2) 132 (1990), no. 3, 549–596. MR 92d:32038
[Nad91] , The boundedness of degree of Fano varieties with Picard number one, J.
Amer. Math. Soc. 4 (1991), no. 4, 681–692. MR 1115788 (93g:14048)
[Nag92] Y. Nagatomo, Rigidity of c1 -self-dual connections on quaternionic Kähler mani-
folds, J. Math. Phys. 33 (1992), no. 12, 4020–4025. MR 93m:53028
[Nag02a] P.-A. Nagy, Nearly Kähler geometry and Riemannian foliations, Asian J. Math. 6
(2002), no. 3, 481–504. MR 1946344 (2003m:53043)
[Nag02b] , On nearly-Kähler geometry, Ann. Global Anal. Geom. 22 (2002), no. 2,
167–178. MR 1923275 (2003g:53073)
[Nah78] W. Nahm, Supersymmetries and their representations, Nuclear Phys. B 135 (1978),
149–166.
[Nah82] , The construction of all self-dual multimonopoles by the ADHM method,
Monopoles in quantum field theory (Trieste, 1981), World Sci. Publishing, Singa-
pore, 1982, pp. 87–94. MR 766754 (86e:53058)
[Nak90] H. Nakajima, Moduli spaces of anti-self-dual connections on ALE gravitational
instantons, Invent. Math. 102 (1990), no. 2, 267–303. MR 1074476 (92a:58024)
[Nak98] , Quiver varieties and Kac-Moody algebras, Duke Math. J. 91 (1998), no. 3,
515–560. MR MR1604167 (99b:17033)
[Nak99] , Lectures on Hilbert schemes of points on surfaces, University Lecture Se-
ries, vol. 18, American Mathematical Society, Providence, RI, 1999. MR 1711344
(2001b:14007)
[NN57] A. Newlander and L. Nirenberg, Complex analytic coordinates in almost complex
manifolds, Ann. of Math. (2) 65 (1957), 391–404. MR 19,577a
[NT88] S. Nishikawa and P. Tondeur, Transversal infinitesimal automorphisms for har-
monic Kähler foliations, Tohoku Math. J. (2) 40 (1988), no. 4, 599–611. MR 972248
(89j:53031)
[NW04] M. Nieper-Wiβkirchen, Chern numbers and Rozansky-Witten invariants of compact
hyper-Kähler manifolds, World Scientific Publishing Co. Inc., River Edge, NJ, 2004.
MR 2110899 (2005h:53070)
[Oba56] M. Obata, Affine connections on manifolds with almost complex, quaternion or
Hermitian structure, Jap. J. Math. 26 (1956), 43–77. MR 0095290 (20 #1796a)
598 BIBLIOGRAPHY

[Oba62] , Certain conditions for a Riemannian manifold to be iosometric with a


sphere, J. Math. Soc. Japan 14 (1962), 333–340. MR 0142086 (25 #5479)
[Oba66] , Riemannian manifolds admitting a solution of a certain system of differ-
ential equations, Proc. U.S.-Japan Seminar in Differential Geometry (Kyoto, 1965),
Nippon Hyoronsha, Tokyo, 1966, pp. 101–114. MR 35 #7263
[Oda88] T. Oda, Convex bodies and algebraic geometry, Ergebnisse der Mathematik und
ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)], vol. 15,
Springer-Verlag, Berlin, 1988, An introduction to the theory of toric varieties, Trans-
lated from the Japanese. MR 922894 (88m:14038)
[Oh83] H. S. Oh, Toral actions on 5-manifolds, Trans. Amer. Math. Soc. 278 (1983), no. 1,
233–252. MR 697072 (85b:57043)
[Oku62] M. Okumura, Some remarks on space with a certain contact structure, Tôhoku
Math. J. (2) 14 (1962), 135–145. MR 26 #708
[Oku68] , On contact metric immersion, Kōdai Math. Sem. Rep. 20 (1968), 389–409.
MR 0264564 (41 #9156)
[Olm05] C. Olmos, A geometric proof of the Berger holonomy theorem, Ann. of Math. (2)
161 (2005), no. 1, 579–588. MR 2150392 (2006c:53045)
[O’N66] B. O’Neill, The fundamental equations of a submersion, Michigan Math. J. 13
(1966), 459–469. MR 0200865 (34 #751)
[O’N83] , Semi-Riemannian geometry, Pure and Applied Mathematics, vol. 103, Aca-
demic Press Inc. [Harcourt Brace Jovanovich Publishers], New York, 1983, With
applications to relativity. MR 85f:53002
[OP99] L. Ornea and P. Piccinni, Induced Hopf bundles and Einstein metrics, New devel-
opments in differential geometry, Budapest 1996, Kluwer Acad. Publ., Dordrecht,
1999, pp. 295–305. MR 99k:53099
[Opr77] V. Oproiu, Almost quaternal structures, An. Ştiinţ. Univ. “Al. I. Cuza” Iaşi Secţ. I
a Mat. (N.S.) 23 (1977), no. 2, 287–298. MR 0493860 (58 #12822)
[Opr84] , Integrability of almost quaternal structures, An. Ştiinţ. Univ. Al. I. Cuza
Iaşi Secţ. I a Mat. 30 (1984), no. 5, 75–84. MR 800155 (86k:53055)
[OR70] P. Orlik and F. Raymond, Actions of the torus on 4-manifolds. I, Trans. Amer.
Math. Soc. 152 (1970), 531–559. MR 0268911 (42 #3808)
[OR74] , Actions of the torus on 4-manifolds. II, Topology 13 (1974), 89–112.
MR 0348779 (50 #1274)
[OR77] P. Orlik and R. C. Randell, The monodromy of weighted homogeneous singularities,
Invent. Math. 39 (1977), no. 3, 199–211. MR 57 #314
[OR04] J.-P. Ortega and T. S. Ratiu, Momentum maps and Hamiltonian reduction,
Progress in Mathematics, vol. 222, Birkhäuser Boston Inc., Boston, MA, 2004.
MR 2021152 (2005a:53144)
[Orl70] P. Orlik, Weighted homogeneous polynomials and fundamental groups, Topology 9
(1970), 267–273. MR 41 #6251
[Orl72a] , On the homology of weighted homogeneous manifolds, Proceedings of
the Second Conference on Compact Transformation Groups (Univ. Massachusetts,
Amherst, Mass., 1971), Part I (Berlin), Springer, 1972, pp. 260–269. Lecture Notes
in Math., Vol. 298. MR 55 #3312
[Orl72b] , Seifert manifolds, Springer-Verlag, Berlin, 1972, Lecture Notes in Mathe-
matics, Vol. 291. MR 54 #13950
[OT87] T. Ohsawa and K. Takegoshi, On the extension of L2 holomorphic functions, Math.
Z. 195 (1987), no. 2, 197–204. MR 88g:32029
[OV05] L. Ornea and M. Verbitsky, An immersion theorem for Vaisman manifolds, Math.
Ann. 332 (2005), no. 1, 121–143. MR 2139254
[OV06a] H. Oguri and C. Vafa, On the geometry of the string landscape and the swampland,
preprint; arXiv:hep-th/0605264 (2006).
[OV06b] L. Ornea and M. Verbitsky, Embeddings of compact Sasakian manifolds,
math.DG/0609617 (2006).
[OW71a] P. Orlik and P. Wagreich, Isolated singularities of algebraic surfaces with C ∗ action,
Ann. of Math. (2) 93 (1971), 205–228. MR 44 #1662
[OW71b] , Singularities of algebraic surfaces with C ∗ action, Math. Ann. 193 (1971),
121–135. MR 45 #1914
BIBLIOGRAPHY 599

[OW75] , Seifert n-manifolds, Invent. Math. 28 (1975), 137–159. MR 50 #13596


[OW77] , Algebraic surfaces with k∗ -action, Acta Math. 138 (1977), no. 1-2, 43–81.
MR 57 #336
[OY06a] T. Oota and Y. Yasui, Explicit toric metric on resolved Calabi-Yau cone, Phys.
Lett. B 639 (2006), no. 1, 54–56. MR 2244850
[OY06b] , New example of infinite family of quiver gauge theories, preprint;
arXiv:hep-th/0610092 (2006).
[OY06c] , Toric Sasaki-Einstein manifolds and Heun equations, Nuclear Phys. B 742
(2006), no. 1-3, 275–294. MR 2219526
[Pal57] R. S. Palais, A global formulation of the Lie theory of transformation groups, Mem.
Amer. Math. Soc. No. 22 (1957), iii+123. MR 0121424 (22 #12162)
[Pal61] , On the existence of slices for actions of non-compact Lie groups, Ann. of
Math. (2) 73 (1961), 295–323. MR 0126506 (23 #A3802)
[Pal05] S. S. Pal, A new Ricci-flat geometry, Phys. Lett. B 614 (2005), no. 3-4, 201–206.
MR 2136797 (2006g:53060)
[Par86] J. Parker, 4-dimensional G-manifolds with 3-dimensional orbits, Pacific J. Math.
125 (1986), no. 1, 187–204. MR 860758 (88e:57033)
[Ped86] H. Pedersen, Einstein metrics, spinning top motions and monopoles, Math. Ann.
274 (1986), no. 1, 35–59. MR 834105 (87i:53070)
[Pen76] R. Penrose, Nonlinear gravitons and curved twistor theory, General Relativity and
Gravitation 7 (1976), no. 1, 31–52, The riddle of gravitation–on the occasion of the
60th birthday of Peter G. Bergmann (Proc. Conf., Syracuse Univ., Syracuse, N. Y.,
1975). MR 0439004 (55 #11905)
[Per02] G. Pereleman, The entropy formula for the Ricci flow and its geometric applica-
tions, preprint; arXiv:math.DG/0211159 (2002).
[Per03a] , Finite extinction time for the solutions to the Ricci flow on certain three-
manifolds, preprint; arXiv:math.DG/0307245 (2003).
[Per03b] , Ricci flow with surgery on three-manifolds, preprint;
arXiv:math.DG/0303109 (2003).
[Pet98] P. Petersen, Riemannian geometry, Graduate Texts in Mathematics, vol. 171,
Springer-Verlag, New York, 1998. MR 98m:53001
[Pha65] F. Pham, Formules de Picard-Lefschetz généralisées et ramification des intégrales,
Bull. Soc. Math. France 93 (1965), 333–367. MR 33 #4064
[Poo86] Y. S. Poon, Compact self-dual manifolds with positive scalar curvature, J. Differ-
ential Geom. 24 (1986), no. 1, 97–132. MR 857378 (88b:32022)
[PP88] H. Pedersen and Y. S. Poon, Hyper-Kähler metrics and a generalization of the
Bogomolny equations, Comm. Math. Phys. 117 (1988), no. 4, 569–580. MR 953820
(89f:53069)
[PP95] , Equivariant connected sums of compact self-dual manifolds, Math. Ann.
301 (1995), no. 4, 717–749. MR 1326765 (95m:53069)
[PP98a] H. Pedersen and Y.-S. Poon, Deformations of hypercomplex structures, J. Reine
Angew. Math. 499 (1998), 81–99. MR 99h:32009
[PP98b] J. F. Plebañski and M. Przanowski, Hermite-Einstein four-dimensional mani-
folds with symmetry, Classical Quantum Gravity 15 (1998), no. 6, 1721–1735.
MR 1628004 (99h:53063)
[PP99] H. Pedersen and Y.-S. Poon, A note on rigidity of 3-Sasakian manifolds, Proc.
Amer. Math. Soc. 127 (1999), no. 10, 3027–3034. MR 2000a:53075
[PPS98] H. Pedersen, Y. S. Poon, and A. F. Swann, Hypercomplex structures associated
to quaternionic manifolds, Differential Geom. Appl. 9 (1998), no. 3, 273–292.
MR 1661193 (99k:53062)
[PR87] R. Penrose and W. Rindler, Spinors and space-time. Vol. 1, Cambridge Monographs
on Mathematical Physics, Cambridge University Press, Cambridge, 1987, Two-
spinor calculus and relativistic fields. MR 917488 (88h:83009)
[PR88] , Spinors and space-time. Vol. 2, second ed., Cambridge Monographs on
Mathematical Physics, Cambridge University Press, Cambridge, 1988, Spinor and
twistor methods in space-time geometry. MR 944085 (89d:83010)
[Pri67] D. Prill, Local classification of quotients of complex manifolds by discontinuous
groups, Duke Math. J. 34 (1967), 375–386. MR 0210944 (35 #1829)
600 BIBLIOGRAPHY

[Pro90] Yu. G. Prokhorov, Automorphism groups of Fano 3-folds, Uspekhi Mat. Nauk 45
(1990), no. 3(273), 195–196. MR 1071944 (91j:14034)
[PS91] Y. S. Poon and S. M. Salamon, Quaternionic Kähler 8-manifolds with positive
scalar curvature, J. Differential Geom. 33 (1991), no. 2, 363–378. MR 92b:53071
[PY05] G. Prasad and S-K. Yeung, Fake projective planes, preprint;
arXiv:math.AG/0512115 (2005).
[Ran75] R. C. Randell, The homology of generalized Brieskorn manifolds, Topology 14
(1975), no. 4, 347–355. MR 54 #1270
[Rav86] D. C. Ravenel, Complex cobordism and stable homotopy groups of spheres, Pure
and Applied Mathematics, vol. 121, Academic Press Inc., Orlando, FL, 1986.
MR 860042 (87j:55003)
[Ree52] G. Reeb, Sur certaines propriétés topologiques des trajectoires des systèmes dy-
namiques, Acad. Roy. Belgique. Cl. Sci. Mém. Coll. in 8◦ 27 (1952), no. 9, 64.
MR 0058202 (15,336b)
[Rei59a] B. L. Reinhart, Foliated manifolds with bundle-like metrics, Ann. of Math. (2) 69
(1959), 119–132. MR 21 #6004
[Rei59b] , Harmonic integrals on foliated manifolds, Amer. J. Math. 81 (1959), 529–
536. MR 0107280 (21 #6005)
[Rei80] M. Reid, Canonical 3-folds, Journées de Géometrie Algébrique d’Angers, Juillet
1979/Algebraic Geometry, Angers, 1979, Sijthoff & Noordhoff, Alphen aan den
Rijn, 1980, pp. 273–310. MR 82i:14025
[Rei83] B. L. Reinhart, Differential geometry of foliations, Ergebnisse der Mathematik und
ihrer Grenzgebiete [Results in Mathematics and Related Areas], vol. 99, Springer-
Verlag, Berlin, 1983, The fundamental integrability problem. MR 85i:53038
[Rol76] D. Rolfsen, Knots and links, Publish or Perish Inc., Berkeley, Calif., 1976, Mathe-
matics Lecture Series, No. 7. MR 0515288 (58 #24236)
[RS01] J. Rosenberg and S. Stolz, Metrics of positive scalar curvature and connections with
surgery, Surveys on surgery theory, Vol. 2, Ann. of Math. Stud., vol. 149, Princeton
Univ. Press, Princeton, NJ, 2001, pp. 353–386. MR 1818778 (2002f:53054)
[RS05] Y. Rollin and M. Singer, Non-minimal scalar-flat Kähler surfaces and parabolic
stability, Invent. Math. 162 (2005), no. 2, 235–270. MR 2199006
[Ruk93] P. Rukimbira, Some remarks on R-contact flows, Ann. Global Anal. Geom. 11
(1993), no. 2, 165–171. MR 94h:53043
[Ruk94] , The dimension of leaf closures of K-contact flows, Ann. Global Anal.
Geom. 12 (1994), no. 2, 103–108. MR 95f:53072
[Ruk95a] , Chern-Hamilton’s conjecture and K-contactness, Houston J. Math. 21
(1995), no. 4, 709–718. MR 96m:53032
[Ruk95b] , Topology and closed characteristics of K-contact manifolds, Bull. Belg.
Math. Soc. Simon Stevin 2 (1995), no. 3, 349–356. MR 96g:53044
[Ruk99] , On K-contact manifolds with minimal number of closed characteristics,
Proc. Amer. Math. Soc. 127 (1999), no. 11, 3345–3351. MR 2000d:53049
[Sae87] O. Saeki, Knotted homology spheres defined by weighted homogeneous polynomials,
J. Fac. Sci. Univ. Tokyo Sect. IA Math. 34 (1987), no. 1, 43–50. MR 88c:57021
[Sal82] S. M. Salamon, Quaternionic Kähler manifolds, Invent. Math. 67 (1982), no. 1,
143–171. MR 83k:53054
[Sal84] , Quaternionic structures and twistor spaces, Global Riemannian geome-
try (Durham, 1983), Ellis Horwood Ser. Math. Appl., Horwood, Chichester, 1984,
pp. 65–74. MR 757207
[Sal86] , Differential geometry of quaternionic manifolds, Ann. Sci. École Norm.
Sup. (4) 19 (1986), no. 1, 31–55. MR 87m:53079
[Sal89] , Riemannian geometry and holonomy groups, Pitman Research Notes in
Mathematics Series, vol. 201, Longman Scientific & Technical, Harlow, 1989.
MR 90g:53058
[Sal99] , Quaternion-Kähler geometry, Surveys in differential geometry: essays on
Einstein manifolds, Surv. Differ. Geom., VI, Int. Press, Boston, MA, 1999, pp. 83–
121. MR 1798608 (2001m:53078)
BIBLIOGRAPHY 601

[Sal04] , Explicit metrics with holonomy G2 , UK-Japan Winter School 2004—


Geometry and Analysis Towards Quantum Theory, Sem. Math. Sci., vol. 30, Keio
Univ., Yokohama, 2004, pp. 50–58. MR 2131362 (2005k:53070)
[San03] O. P. Santillan, A construction of G2 holonomy spaces with torus symmetry, Nu-
clear Phys. B 660 (2003), no. 1-2, 169–193. MR 1983318 (2004d:53057)
[Sas60] S. Sasaki, On differentiable manifolds with certain structures which are closely
related to almost contact structure. I, Tôhoku Math. J. (2) 12 (1960), 459–476.
MR 23 #A591
[Sas65] , Almost contact manifolds, Part 1, Lecture Notes, Mathematical Institute,
Tôhoku University (1965).
[Sas67] , Almost contact manifolds, Part 2, Lecture Notes, Mathematical Institute,
Tôhoku University (1967).
[Sas68] , Almost contact manifolds, Part 3, Lecture Notes, Mathematical Institute,
Tôhoku University (1968).
[Sas85] , Selected papers, Kinokuniya Company Ltd., Tokyo, 1985, With a foreword
by Shiing Shen Chern, Edited by Shun-ichi Tachibana. MR 834384 (87g:01064)
[Sas72] , Spherical space forms with normal contact metric 3-structure, J. Differen-
tial Geometry 6 (1971/72), 307–315. MR 46 #831
[Sat56] I. Satake, On a generalization of the notion of manifold, Proc. Nat. Acad. Sci.
U.S.A. 42 (1956), 359–363. MR 18,144a
[Sat57] , The Gauss-Bonnet theorem for V -manifolds, J. Math. Soc. Japan 9 (1957),
464–492. MR 20 #2022
[Sat77] H. Sato, Remarks concerning contact manifolds, Tôhoku Math. J. 29 (1977), no. 4,
577–584. MR 0458334 (56 #16537)
[Sav79] I. V. Savel0 ev, Structure of the singularities of a class of complex hypersurfaces,
Mat. Zametki 25 (1979), no. 4, 497–503, 635. MR 534292 (80g:32018)
[Sav89] M. V. Saveliev, Integro-differential nonlinear equations and continual Lie algebras,
Comm. Math. Phys. 121 (1989), no. 2, 283–290. MR 985400 (90d:58061)
[Sav02] N. Saveliev, Invariants for homology 3-spheres, Encyclopaedia of Mathematical
Sciences, vol. 140, Springer-Verlag, Berlin, 2002, Low-Dimensional Topology, 1.
MR 2004c:57026
[Sch29] J. A. Schouten, Über unitäre Geometrie, Nederl. Acad. Wetensch. Indagationes
Math. 32 (1929), 457–465.
[Sch95] R. Schoen, On the conformal and CR automorphism groups, Geom. Funct. Anal.
5 (1995), no. 2, 464–481. MR 1334876 (96h:53047)
[Sco83] P. Scott, The geometries of 3-manifolds, Bull. London Math. Soc. 15 (1983), no. 5,
401–487. MR 84m:57009
[Sco05] A. Scorpan, The wild world of 4-manifolds, American Mathematical Society, Prov-
idence, RI, 2005. MR 2136212
[Seg68] G. Segal, Classifying spaces and spectral sequences, Inst. Hautes Études Sci. Publ.
Math. (1968), no. 34, 105–112. MR 0232393 (38 #718)
[Sei32] H. Seifert, Topologie dreidimensionaler gefaserter Räume, Acta Math. 60 (1932),
148–238.
[Ser55] J.-P. Serre, Un théorème de dualité, Comment. Math. Helv. 29 (1955), 9–26.
MR 0067489 (16,736d)
[SH62] S. Sasaki and Y. Hatakeyama, On differentiable manifolds with contact metric
structures, J. Math. Soc. Japan 14 (1962), 249–271. MR 25 #4458
[SH76] S. Sasaki and C. J. Hsu, On a property of Brieskorn manifolds, Tôhoku Math. J.
(2) 28 (1976), no. 1, 67–78. MR 0397614 (53 #1473)
[Sim91] S. R. Simanca, Kähler metrics of constant scalar curvature on bundles over CPn−1 ,
Math. Ann. 291 (1991), no. 2, 239–246. MR 1129363 (92m:53115)
[Sim92] , A note on extremal metrics of nonconstant scalar curvature, Israel J. Math.
78 (1992), no. 1, 85–93. MR 1194961 (93k:58054)
[Sim04] , Canonical metrics on compact almost complex manifolds, Pub-
licações Matemáticas do IMPA. [IMPA Mathematical Publications], Instituto de
Matemática Pura e Aplicada (IMPA), Rio de Janeiro, 2004. MR 2113907
602 BIBLIOGRAPHY

[Siu87] Y. T. Siu, Lectures on Hermitian-Einstein metrics for stable bundles and


Kähler-Einstein metrics, DMV Seminar, vol. 8, Birkhäuser Verlag, Basel, 1987.
MR 89d:32020
[Siu88] , The existence of Kähler-Einstein metrics on manifolds with positive anti-
canonical line bundle and a suitable finite symmetry group, Ann. of Math. (2) 127
(1988), no. 3, 585–627. MR 89e:58032
[Sja95] R. Sjamaar, Holomorphic slices, symplectic reduction and multiplicities of repre-
sentations, Ann. of Math. (2) 141 (1995), no. 1, 87–129. MR 1314032 (96a:58098)
[SL91] R. Sjamaar and E. Lerman, Stratified symplectic spaces and reduction, Ann. of
Math. (2) 134 (1991), no. 2, 375–422. MR 1127479 (92g:58036)
[Sma61] S. Smale, Generalized Poincaré’s conjecture in dimensions greater than four, Ann.
of Math. (2) 74 (1961), 391–406. MR 25 #580
[Sma62a] , On the structure of 5-manifolds, Ann. of Math. (2) 75 (1962), 38–46.
MR 25 #4544
[Sma62b] , On the structure of manifolds, Amer. J. Math. 84 (1962), 387–399.
MR 0153022 (27 #2991)
[Som75] A. J. Sommese, Quaternionic manifolds, Math. Ann. 212 (1974/75), 191–214.
MR 0425827 (54 #13778)
[Sou70] J.-M. Souriau, Structure des systèmes dynamiques, Maitrises de mathématiques,
Dunod, Paris, 1970. MR 0260238 (41 #4866)
[Sou97] , Structure of dynamical systems, Progress in Mathematics, vol. 149,
Birkhäuser Boston Inc., Boston, MA, 1997, A symplectic view of physics, Trans-
lated from the French by C. H. Cushman-de Vries, Translation edited and with a
preface by R. H. Cushman and G. M. Tuynman. MR 1461545 (98i:58103)
[Spe62] D. C. Spencer, Deformation of structures on manifolds defined by transitive, con-
tinuous pseudogroups. I. Infinitesimal deformations of structure, Ann. of Math. (2)
76 (1962), 306–398. MR 27 #6287a
[SS65] I. M. Singer and S. Sternberg, The infinite groups of Lie and Cartan. I. The tran-
sitive groups, J. Analyse Math. 15 (1965), 1–114. MR 36 #911
[SS85] B. Shiffman and A. J. Sommese, Vanishing theorems on complex manifolds,
Progress in Mathematics, vol. 56, Birkhäuser Boston Inc., Boston, MA, 1985.
MR 782484 (86h:32048)
[SSTVP88] P. Spindel, A. Sevrin, W. Troost, and A. Van Proeyen, Extended supersymmetric σ-
models on group manifolds. I. The complex structures, Nuclear Phys. B 308 (1988),
no. 2-3, 662–698. MR 967938 (89m:53138)
[ST71] M. Sebastiani and R. Thom, Un résultat sur la monodromie, Invent. Math. 13
(1971), 90–96. MR 0293122 (45 #2201)
[ST76] S. Sasaki and T. Takahashi, Almost contact structures on Brieskorn manifolds,
Tóhoku Math. J. (2) 28 (1976), no. 4, 619–624. MR 0428340 (55 #1365)
[ST80] H. Seifert and W. Threlfall, Seifert and Threlfall: a textbook of topology, Pure and
Applied Mathematics, vol. 89, Academic Press Inc. [Harcourt Brace Jovanovich
Publishers], New York, 1980, Translated from the German edition of 1934 by
Michael A. Goldman, With a preface by Joan S. Birman, With “Topology of 3-
dimensional fibered spaces” by Seifert, Translated from the German by Wolfgang
Heil. MR 575168 (82b:55001)
[Ste51] N. Steenrod, The Topology of Fibre Bundles, Princeton Mathematical Series, vol.
14, Princeton University Press, Princeton, N. J., 1951. MR 12,522b
[Ste83] S. Sternberg, Lectures on differential geometry, second ed., Chelsea Publishing
Co., New York, 1983, With an appendix by Sternberg and Victor W. Guillemin.
MR 88f:58001
[Sto74] R. E. Stong, Contact manifolds, J. Differential Geometry 9 (1974), 219–238. MR 49
#6258
[Sto92] S. Stolz, Simply connected manifolds of positive scalar curvature, Ann. of Math.
(2) 136 (1992), no. 3, 511–540. MR 1189863 (93i:57033)
[SU02] Y. Shimizu and K. Ueno, Advances in moduli theory, Translations of Mathemati-
cal Monographs, vol. 206, American Mathematical Society, Providence, RI, 2002,
Translated from the 1999 Japanese original, Iwanami Series in Modern Mathemat-
ics. MR 1865412 (2002j:14017)
BIBLIOGRAPHY 603

[Sul77] D. Sullivan, Infinitesimal computations in topology, Inst. Hautes Études Sci. Publ.
Math. (1977), no. 47, 269–331 (1978). MR 0646078 (58 #31119)
[Sus03] L. Susskind, The anthropic landscape of string theory, preprint; hep-th/0302219
(2003).
[Sus05] , The cosmic landscape: String theory and the illusion of intelligent design,
Little, Brown and Company, 2005.
[SvD30] J. A. Schouten and D. van Dantzig, Über unitäre Geometrie, Math. Ann. 103
(1930), no. 1, 319–346. MR 1512625
[SW99] B. Steer and A. Wren, Grothendieck topology and the Picard group of a complex
orbifold, Homotopy invariant algebraic structures (Baltimore, MD, 1998), Contemp.
Math., vol. 239, Amer. Math. Soc., Providence, RI, 1999, pp. 251–262. MR 1718086
(2000j:14029)
[Swa89] A. F. Swann, Quaternionic Kähler geometry and the fundamental 4-form, Pro-
ceedings of the Workshop on Curvature Geometry (Lancaster, 1989) (Lancaster),
ULDM Publ., 1989, pp. 165–173. MR 1089891 (91m:53053)
[Swa91] A. Swann, Hyper-Kähler and quaternionic Kähler geometry, Math. Ann. 289
(1991), no. 3, 421–450. MR 92c:53030
[Swa97] , Singular moment maps and quaternionic geometry, Symplectic singular-
ities and geometry of gauge fields (Warsaw, 1995), Banach Center Publ., vol. 39,
Polish Acad. Sci., Warsaw, 1997, pp. 143–153. MR 1458656 (98m:53066)
[SY91] J.-P. Sha and D. Yang, Positive Ricci curvature on the connected sums of S n × S m ,
J. Differential Geom. 33 (1991), no. 1, 127–137. MR 92f:53048
[SZ05] K. Sfetsos and D. Zoakos, Supersymmetric solutions based on Y p,q and Lp,q,r ,
Phys. Lett. B 625 (2005), no. 1-2, 135–144. MR 2168911 (2006g:53062)
[Tac59] S. Tachibana, On almost-analytic vectors in certain almost-Hermitian manifolds,
Tôhoku Math. J. (2) 11 (1959), 351–363. MR 0117753 (22 #8527)
[Tac65] , On harmonic tensors in compact Sasakian spaces, Tôhoku Math. J. (2) 17
(1965), 271–284. MR 32 #8288
[Tak78] T. Takahashi, Deformations of Sasakian structures and its application to the
Brieskorn manifolds, Tôhoku Math. J. (2) 30 (1978), no. 1, 37–43. MR 81e:53024
[Tak00] S. Takayama, Simple connectedness of weak Fano varieties, J. Algebraic Geom. 9
(2000), no. 2, 403–407. MR 1735807 (2001d:14042)
[Tam92] I. Tamura, Topology of foliations: an introduction, Translations of Mathematical
Monographs, vol. 97, American Mathematical Society, Providence, RI, 1992, Trans-
lated from the 1976 Japanese edition and with an afterword by Kiki Hudson, With
a foreword by Takashi Tsuboi. MR 93c:57021
[Tan62] S. Tanno, Note on infinitesimal transformations over contact manifolds, Tôhoku
Math. J. (2) 14 (1962), 416–430. MR 0145441 (26 #2972)
[Tan63a] , Some transformations on manifolds with almost contact and contact metric
structures, Tôhoku Math. J. (2) 15 (1963), 140–147. MR 0150716 (27 #703)
[Tan63b] , Some transformations on manifolds with almost contact and contact metric
structures. II, Tôhoku Math. J. (2) 15 (1963), 322–331. MR 0161277 (28 #4485)
[Tan64] , A remark on transformations of a K-contact manifold, Tôhoku Math. J.
(2) 16 (1964), 173–175. MR 0172210 (30 #2430)
[Tan67] , Harmonic forms and Betti numbers of certain contact Riemannian mani-
folds, J. Math. Soc. Japan 19 (1967), 308–316. MR 35 #3604
[Tan68] , The topology of contact Riemannian manifolds, Illinois J. Math. 12 (1968),
700–717. MR 38 #2803
[Tan69a] , The automorphism groups of almost contact Riemannian manifolds,
Tôhoku Math. J. (2) 21 (1969), 21–38. MR 39 #3428
[Tan69b] , Sasakian manifolds with constant φ-holomorphic sectional curvature,
Tôhoku Math. J. (2) 21 (1969), 501–507. MR 40 #4894
[Tan70] , On the isometry groups of Sasakian manifolds, J. Math. Soc. Japan 22
(1970), 579–590. MR 42 #6755
[Tan71] , Killing vectors on contact Riemannian manifolds and fiberings related to
the Hopf fibrations, Tôhoku Math. J. (2) 23 (1971), 313–333. MR 44 #4681
604 BIBLIOGRAPHY

[Tan79] , Geodesic flows on CL -manifolds and Einstein metrics on S 3 ×S 2 , Minimal


submanifolds and geodesics (Proc. Japan-United States Sem., Tokyo, 1977), North-
Holland, Amsterdam, 1979, pp. 283–292. MR 81g:58027
[Tan96] , Remarks on a triple of K-contact structures, Tohoku Math. J. (2) 48
(1996), no. 4, 519–531. MR MR1419082 (97i:53055)
[Tau92] C. H. Taubes, The existence of anti-self-dual conformal structures, J. Differential
Geom. 36 (1992), no. 1, 163–253. MR 1168984 (93j:53063)
[Tho69] E. Thomas, Vector fields on manifolds, Bull. Amer. Math. Soc. 75 (1969), 643–683.
MR 0242189 (39 #3522)
[Tho76] C. B. Thomas, Almost regular contact manifolds, J. Differential Geometry 11
(1976), no. 4, 521–533. MR 57 #4195
[Thu76] W. P. Thurston, Some simple examples of symplectic manifolds, Proc. Amer. Math.
Soc. 55 (1976), no. 2, 467–468. MR 53 #6578
[Thu79] W. Thurston, The geometry and topology of 3-manifolds, Mimeographed Notes,
Princeton University, 1979.
[Thu97] W. P. Thurston, Three-dimensional geometry and topology. Vol. 1, Princeton Math-
ematical Series, vol. 35, Princeton University Press, Princeton, NJ, 1997, Edited by
S. Levy. MR 97m:57016
[Tia87] G. Tian, On Kähler-Einstein metrics on certain Kähler manifolds with C1 (M ) > 0,
Invent. Math. 89 (1987), no. 2, 225–246. MR 88e:53069
[Tia90] , On Calabi’s conjecture for complex surfaces with positive first Chern class,
Invent. Math. 101 (1990), no. 1, 101–172. MR 91d:32042
[Tia97] , Kähler-Einstein metrics with positive scalar curvature, Invent. Math. 130
(1997), no. 1, 1–37. MR 99e:53065
[Tia99] , Kähler-Einstein manifolds of positive scalar curvature, Surveys in differ-
ential geometry: essays on Einstein manifolds, Surv. Differ. Geom., VI, Int. Press,
Boston, MA, 1999, pp. 67–82. MR 2001m:32052
[Tia00] , Canonical metrics in Kähler geometry, Lectures in Mathematics
ETH Zürich, Birkhäuser Verlag, Basel, 2000, Notes taken by Meike Akveld.
MR 2001j:32024
[TO66] S. Tachibana and Y. Ogawa, On the second Betti number of a compact Sasakian
space, Natur. Sci. Rep. Ochanomizu Univ. 17 (1966), 27–32. MR 0214096 (35
#4947)
[Tod94] K. P. Tod, Self-dual Einstein metrics from the Painlevé VI equation, Phys. Lett.
A 190 (1994), no. 3-4, 221–224. MR 1285788 (95c:83020)
[Tod95] , Scalar-flat Kähler and hyper-Kähler metrics from Painlevé-III, Classical
Quantum Gravity 12 (1995), no. 6, 1535–1547. MR 1344288 (96f:53068)
[Tod97] , The SU(∞)-Toda field equation and special four-dimensional metrics, Ge-
ometry and physics (Aarhus, 1995), Lecture Notes in Pure and Appl. Math., vol.
184, Dekker, New York, 1997, pp. 307–312. MR 1423177 (98a:53068)
[Ton97] P. Tondeur, Geometry of foliations, Monographs in Mathematics, vol. 90,
Birkhäuser Verlag, Basel, 1997. MR 98d:53037
[Tsu88] H. Tsuji, Logarithmic Fano manifolds are simply connected, Tokyo J. Math. 11
(1988), no. 2, 359–362. MR 976571 (90b:14050)
[TW79] K. P. Tod and R. S. Ward, Self-dual metrics with self-dual Killing vectors, Proc.
Roy. Soc. London Ser. A 368 (1979), no. 1734, 411–427. MR 551471 (81c:83011)
[TW03] S. Tolman and J. Weitsman, The cohomology rings of symplectic quotients, Comm.
Anal. Geom. 11 (2003), no. 4, 751–773. MR 2015175 (2004k:53140)
[TY70] S. Tachibana and W. N. Yu, On a Riemannian space admitting more than one
Sasakian structures, Tôhoku Math. J. (2) 22 (1970), 536–540. MR 43 #1086
[TY87] G. Tian and S.-T. Yau, Kähler-Einstein metrics on complex surfaces with C1 > 0,
Comm. Math. Phys. 112 (1987), no. 1, 175–203. MR 88k:32070
[Udr69] C. Udrişte, Structures presque coquaternioniennes, Bull. Math. Soc. Sci. Math. R.
S. Roumanie (N.S.) 13(61) (1969), no. 4, 487–507 (1970). MR 45 #5908
[Ust99] I. Ustilovsky, Infinitely many contact structures on S 4m+1 , Internat. Math. Res.
Notices (1999), no. 14, 781–791. MR 1704176 (2000f:57028)
[Vaf05] C. Vafa, The string landscape and the swampland, preprint; arXiv:hep-th/0509212
(2005).
BIBLIOGRAPHY 605

[Vai82] I. Vaisman, Generalized Hopf manifolds, Geom. Dedicata 13 (1982), no. 3, 231–255.
MR 690671 (84g:53096)
[Var04] V. S. Varadarajan, Supersymmetry for mathematicians: an introduction, Courant
Lecture Notes in Mathematics, vol. 11, New York University Courant Institute of
Mathematical Sciences, New York, 2004. MR 2069561 (2005g:58011)
[vC06] C. van Coevering, Toric surfaces and Sasaki-Einstein 5-manifolds, SUNY at Stony
Brook Ph.D. Thesis; see also preprint, arXiv:math.DG/0607721 (2006).
[VK99] M. Verbitsky and D. Kaledin, Hyperkähler manifolds, Mathematical Physics
(Somerville), vol. 12, International Press, Somerville, MA, 1999. MR 1815021
(2002f:53079)
[vK04] O. van Koert, Contact homology of Brieskorn manifolds, preprint;
arXiv:math.AG/0410208 (2004).
[Voi02] C. Voisin, Hodge theory and complex algebraic geometry. I, Cambridge Studies in
Advanced Mathematics, vol. 76, Cambridge University Press, Cambridge, 2002,
Translated from the French original by Leila Schneps. MR 1967689 (2004d:32020)
[Voi04] , On the homotopy types of compact Kähler and complex projective mani-
folds, Invent. Math. (2004), no. 157, 329–343.
[Wad75] A. W. Wadsley, Geodesic foliations by circles, J. Differential Geometry 10 (1975),
no. 4, 541–549. MR 53 #4092
[Wak58] H. Wakakuwa, On Riemannian manifolds with homogeneous holonomy group
Sp(n), Tôhoku Math. J. (2) 10 (1958), 274–303. MR 0106488 (21 #5220)
[Wal62] C. T. C. Wall, Classification of (n − 1)-connected 2n-manifolds, Ann. of Math. (2)
75 (1962), 163–189. MR 0145540 (26 #3071)
[Wal67a] F. Waldhausen, Eine Klasse von 3-dimensionalen Mannigfaltigkeiten. I, II, Invent.
Math. 3 (1967), 308–333; ibid. 4 (1967), 87–117. MR 0235576 (38 #3880)
[Wal67b] C. T. C. Wall, Classification problems in differential topology. VI. Classification of
(s − 1)-connected (2s + 1)-manifolds, Topology 6 (1967), 273–296. MR 35 #7343
[Wan54] H.-C. Wang, Closed manifolds with homogeneous complex structure, Amer. J. Math.
76 (1954), 1–32. MR 0066011 (16,518a)
[Wan82] M. Y. Wang, Some examples of homogeneous Einstein manifolds in dimension
seven, Duke Math. J. 49 (1982), no. 1, 23–28. MR 83k:53069
[Wan89] , Parallel spinors and parallel forms, Ann. Global Anal. Geom. 7 (1989),
no. 1, 59–68. MR 91g:53053
[Wan95] , On non-simply connected manifolds with non-trivial parallel spinors, Ann.
Global Anal. Geom. 13 (1995), no. 1, 31–42. MR 96a:53066
[War90] R. S. Ward, Einstein-Weyl spaces and SU(∞) Toda fields, Classical Quantum Grav-
ity 7 (1990), no. 4, L95–L98. MR 1045295 (91g:83019)
[WB82] E. Witten and J. Bagger, Quantization of Newton’s constant in certain supergravity
theories, Phys. Lett. B 115 (1982), no. 3, 202–206. MR 669354 (83m:81107)
[Web77] S. M. Webster, On the transformation group of a real hypersurface, Trans. Amer.
Math. Soc. 231 (1977), no. 1, 179–190. MR 0481085 (58 #1231)
[Web78] , Pseudo-Hermitian structures on a real hypersurface, J. Differential Geom.
13 (1978), no. 1, 25–41. MR 80e:32015
[Wei58] A. Weil, Introduction à l’étude des variétés kählériennes, Publications de l’Institut
de Mathématique de l’Université de Nancago, VI. Actualités Sci. Ind. no. 1267,
Hermann, Paris, 1958. MR 0111056 (22 #1921)
[Wei77a] A. Weinstein, Lectures on symplectic manifolds, American Mathematical Society,
Providence, R.I., 1977, Expository lectures from the CBMS Regional Conference
held at the University of North Carolina, March 8–12, 1976, Regional Conference
Series in Mathematics, No. 29. MR 57 #4244
[Wei77b] , Symplectic V -manifolds, periodic orbits of Hamiltonian systems, and the
volume of certain Riemannian manifolds, Comm. Pure Appl. Math. 30 (1977),
no. 2, 265–271. MR 56 #13260
[Wei79] , On the hypotheses of Rabinowitz’ periodic orbit theorems, J. Differential
Equations 33 (1979), no. 3, 353–358. MR 81a:58030b
[Wei80] , Fat bundles and symplectic manifolds, Adv. in Math. 37 (1980), no. 3,
239–250. MR 82a:53038
606 BIBLIOGRAPHY

[Wel80] R. O. Wells, Jr., Differential analysis on complex manifolds, second ed., Gradu-
ate Texts in Mathematics, vol. 65, Springer-Verlag, New York, 1980. MR 608414
(83f:58001)
[WG68a] J. A. Wolf and A. Gray, Homogeneous spaces defined by Lie group automorphisms.
I, J. Differential Geometry 2 (1968), 77–114. MR 0236328 (38 #4625a)
[WG68b] , Homogeneous spaces defined by Lie group automorphisms. II, J. Differential
Geometry 2 (1968), 115–159. MR 0236329 (38 #4625b)
[Wil72] D. L. Wilkens, Closed (s − 1)-connected (2s + 1)-manifolds, s = 3, 7, Bull. London
Math. Soc. 4 (1972), 27–31. MR 46 #6378
[Wil02] C. Willett, Contact reduction, Trans. Amer. Math. Soc. 354 (2002), no. 10, 4245–
4260 (electronic). MR 1926873 (2003m:53152)
[Win83] H. E. Winkelnkemper, The graph of a foliation, Ann. Global Anal. Geom. 1 (1983),
no. 3, 51–75. MR 85j:57043
[Wiś91] J. A. Wiśniewski, On Fano manifolds of large index, Manuscripta Math. 70 (1991),
no. 2, 145–152. MR 1085628 (92d:14033)
[Wit83] E. Witten, Global aspects of current algebra, Nuclear Phys. B 223 (1983), no. 2,
422–432. MR 717915 (86i:81103)
[Wol65] J. A. Wolf, Complex homogeneous contact manifolds and quaternionic symmetric
spaces, J. Math. Mech. 14 (1965), 1033–1047. MR 32 #3020
[Wra97] D. Wraith, Exotic spheres with positive Ricci curvature, J. Differential Geom. 45
(1997), no. 3, 638–649. MR 98i:53058
[WZ71] J. Wess and B. Zumino, Consequences of anomalous Ward identities, Phys. Lett.
37B (1971), 95–97. MR 0342064 (49 #6810)
[WZ86] M. Y. Wang and W. Ziller, Einstein metrics with positive scalar curvature, Curva-
ture and topology of Riemannian manifolds (Katata, 1985), Lecture Notes in Math.,
vol. 1201, Springer, Berlin, 1986, pp. 319–336. MR 859594 (87k:53114)
[WZ90] , Einstein metrics on principal torus bundles, J. Differential Geom. 31
(1990), no. 1, 215–248. MR 91f:53041
[Yam99] T. Yamazaki, A construction of K-contact manifolds by a fiber join, Tohoku Math.
J. (2) 51 (1999), no. 4, 433–446. MR 2001e:53094
[Yam01] , On a surgery of K-contact manifolds, Kodai Math. J. 24 (2001), no. 2,
214–225. MR 2002c:57048
[Yan63] K. Yano, On a structure defined by a tensor field f of type (1, 1) satisfying f 3 +f =
0, Tensor (N.S.) 14 (1963), 99–109. MR 0159296 (28 #2513)
[Yan65] , Differential geometry on complex and almost complex spaces, International
Series of Monographs in Pure and Applied Mathematics, Vol. 49, A Pergamon Press
Book. The Macmillan Co., New York, 1965. MR 0187181 (32 #4635)
[Yau77] S.-T. Yau, Calabi’s conjecture and some new results in algebraic geometry, Proc.
Nat. Acad. Sci. U.S.A. 74 (1977), no. 5, 1798–1799. MR 0451180 (56 #9467)
[Yau78] , On the Ricci curvature of a compact Kähler manifold and the complex
Monge-Ampère equation. I, Comm. Pure Appl. Math. 31 (1978), no. 3, 339–411.
MR 81d:53045
[YK84] K. Yano and M. Kon, Structures on manifolds, Series in Pure Mathematics, vol. 3,
World Scientific Publishing Co., Singapore, 1984. MR 86g:53001
[YY02] S. S.-T. Yau and Y. Yu, Algebraic classification of rational CR structures on
topological 5-sphere with transversal holomorphic S 1 -action in C4 , Math. Nachr.
246/247 (2002), 207–233. MR 1944558 (2003j:32042)
[YY05] , Classification of 3-dimensional isolated rational hypersurface singular-
ities with C∗ -action, Rocky Mountain J. Math. 35 (2005), no. 5, 1795–1809.
MR 2206037 (2006j:32034)
[Zhe00] F. Zheng, Complex differential geometry, AMS/IP Studies in Advanced Mathemat-
ics, vol. 18, American Mathematical Society, Providence, RI, 2000. MR 1777835
(2001i:32035)
[Zil82] W. Ziller, Homogeneous Einstein metrics on spheres and projective spaces, Math.
Ann. 259 (1982), no. 3, 351–358. MR 84h:53062
[Zum79] B. Zumino, Supersymmetry and Kähler manifolds, Phys. Lett. B 87 (1979), 203–
207.
Index

3-Sasakian diamond, 486 line, 17, 95


3-Sasakian moment map, 499 principal, 14
3-symmetric spaces, 540 spin, 18
D = 11 supergravity, 546 vector, 17
G-manifold, 38
G-structure, 23 Calabi conjecture, 102, 154
Q-Cartier, 125 Calabi metrics on T ∗ CPn , 459
Q-factorial, 125, 157 Calabi-Yau manifold, 104
Γ-atlas, 37 Calabi-Yau orbifold, 129, 165
¯
i∂ ∂-lemmas, 81 Calabi-Yau Theorem, 103, 133
transverse, 239 characteristic
étale space, 10 Euler Poincaré, 93
3-Sasakian, 473 orbifold Euler, 121
characteristic classes, 18
action, 39 characteristic foliation
free, 39 irregular, 188
local type, 44 non-regular, 188
locally free, 39 quasi-regular, 188
locally proper, 42 regular, 188
proper, 39 characteristic hyperplane, 285
transitive, 42 Chern classes, 243, 557
adjunction, 142 Chern classes, basic, 243
AdS/CFT Duality Conjecture, 547 Chern form, 92
Alekseevskian spaces, 438 class
ample, 99 characteristic, 556
automorphism Chern, 92, 557
G structure, 40 Euler, 557
integral Pontrjagin, 557
Barden invariant, 336 integral Stiefel Whitney, 557
Barden theorem, 336 orbifold canonical, 127
basic p-from, 59 orbifold characteristic, 120
basic cohomology, 59 orbifold Chern class, 121
Beauville Conjecture, 437 orbifold Euler, 121
Berger theorem, 26 orbifold first Chern class, 128
Betti numbers orbifold Pontrjagin, 121
basic, 216 rational Pontrjagin, 557
Bishop theorem, 384 real characteristic, 121
blowing-up, 97 Stiefel Whitney, 557
Boothby-Wang Theorem, 209 Todd, 92
bundle classifying space, 117, 555
associated, 17 cocycle, 15
frame, 15 cocyle
holomorphic line, 95 Haefliger, 53
holomorphic vector, 11 cohomology

607
608 INDEX

group, 118 Q, 125


commuting sheaf, 212 absolute, 126
complex structure Baily, 126
transverse, 68 big, 102
cone, 201 branch, 125
metric, 201 Cartier, 100, 124
sine, 543 class group, 124
symplectic, 202, 203 effective, 100
connection, 19 exceptional, 97
Koszul, 21 nef, 102
Levi Civita, 26 orbifold canonical, 127
linear, 24 ramification, 127
Obata, 35 Weil, 100
Oproiu, 35
torsionfree, 24 Eguchi-Hanson metric, 460
contact Eilenberg-MacLane space, 117, 118
almost, 190–192, 195, 202, 474, 475 elliptic
complex, 184 Lie algebra, 41
distribution, 182 elliptic fibration, 361
form, 180, 182 equivalence
Hamiltonian, 190 linear, 101
line bundle, 182 numerical, 101
manifold, 181 Euler characteristic
co-oriented, 183 basic, 216
basic holomorphic, 218
metric manifold, 199
holomorphic, 92, 96, 104
metric structure, 199
Euler class, 557
pseudogroup, 180
exotic
strict transformation group, 189
contact, 324
structure, 182
subbundle, 182 fan, 282
transformation, 180 fibration
infinitesimal, 181 Boothby-Wang, 188
strict, 180, 181 flow
transformation group, 189 Anosov, 70
contact type, 186 Riemannian, 69
convex polytope, 279 foliated atlas, 31
covariant derivative, 21 foliated coordinate chart, 31
exterior, 20 foliation, 31
covariantly constant, 21 characteristic, 187, 196, 198–200
cuplength, 229 Riemannian, 62
curvature, 221 simple, 51, 55
2-form, 20 singular Riemannian, 70
Φ-sectional, 225 form
Ricci, 223 Kähler, 78
Riemannian, 26, 221 function
scalar, 372 transition, 16
sectional, 221, 225, 233 fundamental basic class, 216

deformation genus, 93
characteristic foliation, 269 basic arithmetic, 218
transverse holomorphic structure, 267 basic geometric, 218
type I, 269 geometric, 96
type II, 240 Todd, 92
dimension gravitational instantons, 456
algebraic, 96 gravitational multi-instantons, 462
Iataka, 96 group
Kodaira, 96 Neron-Severi, 95
divisor, 100, 123 Picard, 95
INDEX 609

groupoid, 38, 53 leaves, 32


étale, 38, 115, 553 space of, 32
action, 552 LeBrun-Salamon Conjecture, 436
holonomy, 54 Legendre transform, 461
Lie, 552 lens space, 264
translation, 553 line bundle
canonical, 96
Hard Lefschetz Theorem link, 299
transverse, 218 local uniformizing groups, 106
Heisenberg group, 181, 229, 261, 275 local uniformizing system, 105
Hirzebruch signature log del Pezzo surfaces, 128
basic, 218 log resolution, 157
Hodge numbers, 89–91 Lusternik-Schnirelman theory, 232
basic, 218
Hodge Theorem, 89 Maldacena Conjecture, 548
holonomy manifold
covering, 56 3-Sasakian, 477
group, 22 homogeneous, 42
leaf, 54 almost Hermitian, 78
pseudogroup, 55 almost hyperkähler, 455
holonomy group (of a leaf), 55 complex, 76
holonomy groupoid, 54 complex flag, 81
homologous Fano, 98
a, 242, 248 Hermitian, 78
hypercomplex, 35, 453 Hodge, 99
hyperhamiltonian T n -space, 461 hyperkähler, 455
hyperhamiltonian action, 458 Kähler, 79
hyperkähler, 36 nearly Kähler, 539
hyperkähler manifold polarized Kähler, 96
Asymptotically Locally Euclidean, 468 quaternionic Kähler, 429
locally toric, 461 Vaisman, 250
hypersurface Wang-Ziller, 254, 255, 275, 294, 392
weighted, 139 Marchiafava-Romani class, 428
McKay correspondence, 465
index meromorphic function, 100
divisibility, 131 metric
Fano, 131 η-Einstein, 370, 417
ramification, 125 canonical Kähler, 152
injections, 106, 109 Einstein, 370
integrable, 23 Hermitian, 78, 93
irregular, 188 Kähler, 79
isotropy subgroup, 106 Lorentzian, 29, 420
warped product, 201
K-contact, 200, 203, 206, 207, 211, 221 Milnor number, 300
structure theorem, 212 minimal model program, 341
K-contact manifold, 200 mirror symmetry, 166
Kähler moment map, 277
almost, 203 contact, 284
Kähler cone, 102 hypercomplex, 454
Kähler form, 29 hyperkähler, 458
Kähler potential, 81 quaternionic Kähler, 439
Killing spinor, 530 symplectic cone, 283
Kirwan map, 278 Monge-Ampère equation, 104, 153
Kleinian singularities, 467
Nijenhuis tensor, 28, 76, 203–205
leaf normal, 203
closure, 211 number
generic, 55 Bernoulli, 313
leaf holonomy, 107 Chern, 93
610 INDEX

orbifold Picard, 131 hypercomplex, 454


Picard, 95 hyperkähler, 458

O’Neill tensor, 65, 220, 483 rational map, 97


obstruction reduction
Bishop, 384 contact, 290
Lichnerowicz, 386 hypercomplex, 454
orbibundle, 110 hyperkähler, 458
anticanionical, 127 quaternionic Kähler, 440
canonical, 127 singular symplectic, 277, 459
orbifold, 105 symplectic, 276, 277
classifying space, 117 Reeb flow, 200
developable, 107 Reeb type, 285
Fano, 128 Reeb vector field, 187, 188, 199, 208
polarized Kähler, 132 irregular, 188
quaternionic, 431 non-regular, 188
orbifold charts, 105 quasi-regular, 188
orbifold G-structure, 114 regular, 188
orbisheaf, 108 reflection, 123
multiplier ideal, 157, 158 reflection group, 123
structure, 109 regular, 188
orbit, 39 Ricci form, 86, 243
exceptional, 46 Ricci form, transverse, 243
principal, 46 Ricci tensor, 85, 223, 243, 371
regular, 46 transverse, 243
singular, 46 Ricci tensor, transverse, 224
type, 43 Riemann curvature tensor, 84
order Riemannian
K-contact, 207 curvature, 26
of an orbifold, 107 Riemannian metric, 25
ringed space, 11, 123
p-brane, 547
Painlevé VI equation, 446, 447 Sasaki group, 236
parallel translation, 22 Sasaki metrics
perturbation Einstein, 370
weighted homogeneous, 141 extremal, 378
Picard group, 125 Sasaki-Einstein manifolds, 369
of an orbifold, 130 Sasaki-Seifert structure
plurigenera, 96 positive, 245
Poincaré homology sphere, 333 Sasakian, 206
Poisson bracket, 277 Sasakian manifold, 206
polyhedral cone, 283 Sasakian structure, 206
polynomial anticanonical, 245
Brieskorn-Pham, 130, 141, 167 canonical, 245
weighted homogeneous, 139, 300 indefinite, 245
polytope, 279, 281 negative, 245
Delzant, 280 null, 245
lt, 280 positive, 245, 248, 319
presheaf, 9 Seifert bundles, 144
primitive cohomology, 91 Seifert fibred 3-manifold, 144
prolongation, 41 sheaf, 9, 108
pseudo-Riemannian metric, 29 derived functor, 118
pseudogroup, 36 locally free, 11
transitive, 37 multiplier ideal, 156
structure, 10
quasi-regular, 188 sheaf of groupoids, 38
quasi-smoothness, 140 simple normal crossings, 157
quaternionic Kähler, 35 singular locus
quotient orbifold, 106, 124, 125
INDEX 611

singularity pseudo-Riemannian, 29
Kawamata log-terminal (klt), 158 pseudogroup, 36
log-canonical, 158 quaternionic, 429
slice, 42 quaternionic Hermitian, 35, 429
Smale theorem, 336 quaternionic Kähler, 35, 429, 431
Smale-Barden classification, 335 Riemannian, 17
spin manifold, 18 Sasaki-Seifert, 242
spinor, 18 Sasakian, 206, 225, 236
stabilizer, 39 Seifert fibred, 47
startification spin, 18
smooth, 43 symplectic, 17, 30
Stiefel Whitney classes, 557 weak G2 , 536
Stiefel-Whitney class, 18 proper, 537
stratification subgroup
Whitney, 43 isotropy, 39
strictly pseudoconvex, 199 submersion, 32, 51
structure supersymmetry, 545
3-Sasakian, 477 surface
f −, 34 complex, 93
almost contact 3-structure, 475 del Pezzo, 98
almost contact metric, 195 Riemann, 93
almost hypercontact, 474
almomst hypercontact, 36 Taub-NUT metrics, 460
almost complex, 27 taut, 196
almost contact, 33, 202 tensor
normal, 203 Killing, 540
almost CR, 33, 199 Tian’s α-invariant, 155
almost Hermitian, 29 toral rank, 71, 72, 212
almost hypercomplex, 35 transversal, 54
almost hyperhermitian, 36 transverse geometry, 32
almost hyperkähler, 455 transverse holonomy groupoid, 58
almost product, 32 transverse homothety, 228
almost product Riemannian metric, 32 type
almost quaternionic, 34, 428 (p, q) differential form, 77
almost quaternionic Hermitian, 35, 428 of a Sasakian structure, 245
almost Tachibana, 540 of Killing spinors, 533
complex, 17, 76 of weak holonomy G2 , 538
complex contact, 184
UFD, 123
conformal, 30
conformal symplectic, 31 V-bundle, 110
conjugate Sasakian, 241 V-manifold, 105
contact, 180 variety
contact metric, 195 algebraic, 82
CR, 33, 199, 206 analytic, 82
Haefliger, 32, 52 irreducible, 82
Hermitian, 29 weighted, 139
homogeneous contact, 272 vector
homogeneous K-contact, 273 exponent, 141
homogeneous Sasakian, 273 vector bundle
hyper f-, 36 compelx, 92
hypercomplex, 35 vector field
hyperhermitian, 36 foliate, 60
hyperkähler, 36, 455 stransverse holomorphic, 69
K-contact, 221
Kähler, 29 weighted Sasakian sphere, 213
locally Sasakian, 265, 312 weighted sphere, 237
Lorentzian, 29 Weinstein conjecture, 231
nearly Kähler, 539 well-formed, 136, 141
612 INDEX

Wolf spaces, 434


Wolf-Gray conjecture, 540

Vous aimerez peut-être aussi