Garcia-Prada, Bourguignon, Salamon - The Many Facets of Geometry - A Tribute To Nigel Hitchin
Garcia-Prada, Bourguignon, Salamon - The Many Facets of Geometry - A Tribute To Nigel Hitchin
Edited by
Oscar Garcı́a-Prada
Jean Pierre Bourguignon
Simon Salamon
1
3
Great Clarendon Street, Oxford ox2 6dp
Oxford University Press is a department of the University of Oxford.
It furthers the University’s objective of excellence in research, scholarship,
and education by publishing worldwide in
Oxford New York
Auckland Cape Town Dar es Salaam Hong Kong Karachi
Kuala Lumpur Madrid Melbourne Mexico City Nairobi
New Delhi Shanghai Taipei Toronto
With offices in
Argentina Austria Brazil Chile Czech Republic France Greece
Guatemala Hungary Italy Japan Poland Portugal Singapore
South Korea Switzerland Thailand Turkey Ukraine Vietnam
Oxford is a registered trade mark of Oxford University Press
in the UK and in certain other countries
Published in the United States
by Oxford University Press Inc., New York
c Oxford University Press 2010
The moral rights of the authors have been asserted
Database right Oxford University Press (maker)
First published 2010
All rights reserved. No part of this publication may be reproduced,
stored in a retrieval system, or transmitted, in any form or by any means,
without the prior permission in writing of Oxford University Press,
or as expressly permitted by law, or under terms agreed with the appropriate
reprographics rights organization. Enquiries concerning reproduction
outside the scope of the above should be sent to the Rights Department,
Oxford University Press, at the address above
You must not circulate this book in any other binding or cover
and you must impose the same condition on any acquirer
British Library Cataloguing in Publication Data
Data available
Library of Congress Cataloging in Publication Data
Data available
Typeset by SPI Publisher Services, Pondicherry, India
Printed in Great Britain
on acid-free paper by
CPI Antony Rowe
ISBN 978–0–19–953492–0
1 3 5 7 9 10 8 6 4 2
To Nigel Hitchin
This page intentionally left blank
SHORT CONTENTS
FULL CONTENTS ix
PREFACE xvi
LIST OF EDITORS AND CONTRIBUTORS xvii
INDEX 421
FULL CONTENTS
PREFACE xvi
LIST OF EDITORS AND CONTRIBUTORS xvii
INDEX 421
PREFACE
Editors
Oscar Garcı́a-Prada (CSIC, Madrid, Spain)
Jean Pierre Bourguignon (IHES, France)
Simon Salamon (Politecnico di Torino, Italy)
Contributors
Jørgen Ellegaard Andersen (Aarhus Universitet, Denmark): Toeplitz opera-
tors and Hitchin’s projectively flat connection.
Sir Michael Atiyah (University of Edinburgh, UK): Mathematical work of
Nigel Hitchin.
Philip Boalch (ENS, Paris, France): Towards a non-linear Schwarz’s list.
Robert L. Bryant (MSRI, USA): Non-embedding and non-extension results in
special holonomy.
Benoit Charbonneau (Duke University, Durham, USA) and Jacques
Hurtubise (McGill University, Canada): The Nahm transform for calorons.
Simon K. Donaldson (Imperial College, London, UK): Nahm’s equations and
free-boundary problems.
Daniel S. Freed (University of Texas at Austin, USA), Michael J. Hopkins
(Harvard University), and Constantin Teleman (UC Berkeley, USA): Consis-
tent orientation of moduli spaces.
William M. Goldman (University of Maryland, USA): Higgs bundles and
geometric structures on surfaces.
Marco Gualtieri (MIT, USA): Branes on Poisson varieties.
F. Reese Harvey (Rice University, USA) and H. Blaine Lawson, Jr. (SUNY
at Stony Brook, USA): Projective linking and boundaries of positive holomorphic
chains in projective manifolds, part I.
Tamás Hausel (University of Oxford, UK): S-duality in hyperkähler Hodge
theory.
Nigel Hitchin (University of Oxford, UK): Geometry and Physics: a personal
view.
xviii List of Editors and Contributors
Nigel Hitchin
This excellent collection of chapters prompts me to recall here how their various
themes emerged in my own mathematical development. It may help the reader to
navigate a passage through them and also perhaps explain how the links between
physics and geometry which seem to underlie much of my own work came about.
I hope these background comments do not seem too self-indulgent, but they
might shed some light on the curious process of doing mathematical research –
a process which is largely hidden in the chapters we finally write.
I was never a serious student of physics. As a mathematics undergraduate
in Oxford from 1965 to 1968 I avoided the courses in relativity and quantum
mechanics that my peers were taking. Instead I would browse in the Jesus College
library trying to make sense of Weyl’s The Classical Groups or in particular
Harley Flanders’ Differential Forms with Applications to the Physical Sciences,
one which in a way defined my interests right from the start.
Although my DPhil thesis was about the Dirac operator, this was largely
in the context of the Atiyah–Singer index theorem and associated vanishing
theorems and it hardly occurred to me that physicists would be interested in
this Riemannian version. When I went to the Institute for Advanced Study in
1971–3 my neighbour in the next apartment was the physicist Andy Hanson
but we rarely spoke about our work. He once asked me what was so special
about 26 dimensions and all I could think of was that it was of the form 8k + 2
(where the mod 2 index exists). My main thoughts at that time were on the
two big unsolved problems in differential geometry: the Yamabe conjecture and
the Calabi conjecture. Yau’s presence at the Institute in the second year was
a clear influence here. My first paper, on four-dimensional Einstein manifolds
(Hitchin 1974), was heavily influenced by the Calabi conjecture and the K3
surface. There was a great lack of understanding of the Einstein equations
at the time amongst pure mathematicians, concerning both global and local
existence.
It was the conformal invariance of the Dirac operator which brought me
closer to physics. I spent the year 1973–4 in New York at the Courant Institute
and I began looking for a systematic way of finding all first-order conformally
invariant differential operators. I started to read papers of Penrose on the zero
rest-mass field equations; the conformal invariance of Maxwell’s equations on a
2 Geometry and physics
Chapter 3 by Claude LeBrun represents well this general area – he takes the
Einstein–Maxwell equation in four dimensions and links it with Kähler geometry
and topological obstructions. The work draws also on the more recent Seiberg–
Witten equations, one of the most important contributions to four-dimensional
geometry from the physicists. This of course followed on from Donaldson’s
own spectacular use of gauge theory in four-dimensional topology. I have to
admit that, when writing the 1978 paper with Atiyah and Singer, I thought
that self-dual connections on self-dual spaces was a cosy self-contained theory –
a sort of quaternionic version of Riemann surfaces. But then Taubes showed
how to construct self-dual connections in general and there was also clearly
a story to be told for algebraic surfaces. When Simon Donaldson became my
research student in 1980, I asked him to look into these questions, and the rest is
history.
Following on from the instantons, I began to look at related equations, and in
particular magnetic monopoles in Euclidean space. Here the physical intuition of
Nick Manton had acquired solid mathematical support in the existence proof of
Jaffe and Taubes, but the Bogomolny equations which describe the monopoles
were also clear cousins of the instanton equations and I wanted to use twistor
methods to describe the solutions more concretely. I began by asking what
was the natural background geometry in three dimensions to support these –
an analogue of the self-dual conformal structures in four dimensions. These
turned out to be what are known as Einstein–Weyl manifolds (after Weyl’s
early attempt to introduce the electromagnetic field into general relativity).
The important feature of these is that the space of geodesics has the structure
of a two-dimensional complex manifold, giving a version of twistor theory. I
then abandoned the general picture and concentrated on the Euclidean case and
came up with the spectral curve description of SU (2) monopoles. The solution
to the equations was determined by an algebraic curve which satisfied some
transcendental constraints. This was in some respects a geometrical version of
Corrigan and Goddard’s application of Ward’s method. I was about to move on
to something else when a preprint of Nahm fell on my desk (1982).
Nahm’s construction of SU (2) monopoles of charge k involved a system of
ordinary differential equations for three k × k matrices T1 , T2 , T3 : T1 = [T2 , T3 ],
etc. For k = 2 he showed that this could be solved with elliptic functions. Since
my spectral curve for charge 2 was elliptic there was a clear challenge to relate
directly the two approaches, which led to Hitchin (1983). In writing this, I learned
more about non-linear systems which can be linearized on the Jacobian of a
curve. Nahm’s equations – generalizations of the spinning top equations – are
examples of this type and the curve in question turns out to be the spectral
curve of the monopole as defined by the twistor approach. The Nahm approach,
however, has the specific advantage that one can see directly that the solutions
to the Bogomolny equations are nonsingular. One should note that singularity
issues dogged the early development of the theory – Ward’s charge 2 solution
had to be checked numerically and Forgács, Horváth, and Palla consigned
4 Geometry and physics
essentially the same solution to the waste bin because they thought it was
singular.
The Nahm method is the subject of Chapter 4 by Benoit Charbonneau and
Jacques Hurtubise, though not for monopoles – instead for the more sophisticated
calorons, which are instantons on S 1 × R3 . This is a more elaborate closing
of a circle of ideas than in Hitchin (1983). I am particularly pleased to see it
here since I began to make headway on the problem in 1983 and had every
intention of spending a sabbatical in Stony Brook working out the details, but
only got as far as giving a talk to Jacques and his colleagues in Montreal
when other issues intervened, of which more later. One of the most beautiful
applications of Nahm’s equations to monopoles was Simon Donaldson’s proof
(1984) that a circle bundle over the moduli space is naturally diffeomorphic
to the space of based rational maps from the projective line to itself. In his
contribution to this collection, Simon returns to Nahm’s equations but where
the matrices Ti are replaced by vector fields. He pointed out to me long ago that
Ashtekar’s Hamiltonian formulation of general relativity could be viewed this
way (I gave a minor extension to this in Hitchin (1998) in the context of hyper-
complex geometry). Donaldson’s chapter links Nahm’s equations to the physical
area of free boundary problems and their analogues on compact Riemannian
manifolds.
Arriving in Stony Brook in August 1983, I made contact with Martin Roček
in the Theoretical Physics Department. Earlier in the year Blaine Lawson had
visited Oxford and he was puzzled by a construction of hyperkähler manifolds
by Martin and his co-workers. ‘They take a group representation, turn a handle,
and you get a hyperkähler metric.’ My own interest at the time in hyperkähler
geometry was dominated by the twistor approach, so I was curious to learn what
was going on. We soon discovered that this was a straightforward generalization
of the symplectic quotient construction, which was very much in vogue in Oxford
at the time, for example, in the work of Atiyah and Bott on the Yang–Mills
equations on Riemann surfaces, and in Frances Kirwan’s work in algebraic
geometry. I worked hard contributing to the joint paper Hitchin et al. (1987a)
in trying to reconcile the physicists’ language of supersymmetric fields with the
twistor approach, but I think the reader can easily discern the borderline between
mathematics and physics. In fact Martin tried at the time to convince me to work
on another piece of geometry related to the supersymmetric sigma model, but I
wasn’t listening. It finally surfaced 25 years later in Marco Gualtieri’s 2003 thesis
as a generalized Kähler manifold.
On my return to Oxford in 1984, Atiyah and I set to work in evaluating the
hyperkähler metric on the moduli space of charge 2 monopoles – the hyperkähler
quotient construction when applied to a certain infinite-dimensional affine space
showed that every monopole moduli space has such a metric. I then began looking
at other moduli spaces with this property, where the gauge-theoretic equations
are hyperkähler moment maps. One of these could be defined on a Riemann
Geometry and physics 5
surface and became the study of Higgs bundles (the terminology is due to Carlos
Simpson). At first (as in Hitchin 1986) I thought I was just getting a hyperkähler
metric on the cotangent bundle of the moduli space of stable bundles, and I
began explaining this on the blackboard to Simon Donaldson but got stuck.
Having decided that there must be a joint stability condition it was easy to see
what should happen because one knew the equations to be satisfied and the type
of vanishing theorem that would ensue. From then on, virtually every day a new
facet of the subject seemed to appear. The paper Hitchin (1987b) by no means
wrote itself but it certainly had an inner momentum. There was one aspect
however – the integrable system – which came a little later, because it was not
really part of the hyperkähler story. I was visiting Jacques Hurtubise in 1985 and
we went up to his family cottage in the Laurentian mountains. While he was doing
some maintenance work I sat on the porch with a pen and paper and suddenly
realized that we had a completely integrable Hamiltonian system where the fibres
were recognizable abelian varieties. Until then, my experience with such objects
was confined to Nahm’s equations but here was a vast natural source of such
things.
Chapter 6 by Ramanan gives a good survey of the subject from the point
of view of algebraic geometry, including the Hecke correspondence which links
up with Chapter 7 by Witten on the very active recent work on the geometric
Langlands programme from a physics point of view. This makes essential use of
both the gauge-theoretic equations and the integrable system. Chapter 8 by Bill
Goldman is again about Higgs bundles, but focuses on the interpretation as flat
bundles and in particular on representations of the fundamental group of the
surface into a real form of the group. One of the current problems in this area is
to give an understanding of the generalized Teichmüller spaces of Hitchin (1982)
as structures on the surface modulo diffeomorphism and Goldman discusses some
recent results in this direction.
From 1979 to 1989, I was a Fellow of St Catherine’s College, and apart from
my teaching duties there I also had multiple opportunities over lunch to talk
to Graeme Segal and Michael Atiyah, who were also Fellows. I never wrote a
joint paper with Graeme, but I valued greatly the questions and remarks which
came in our discussions. Our approaches were different – I would always want
to talk about stable bundles and a finite-dimensional moduli space whereas for
him stability was not an issue and, as he says in Chapter 9, the consideration of
the category or stack of bundles is actually closer to quantum field theory which
was always his prime motivation. In Autumn 1988 we had a seminar in Oxford
run by Atiyah, Segal, and Ruth Lawrence on quantum field theory and the Jones
polynomial. During that summer, and in particular at the International Congress
of Mathematical Physicists in Swansea, Witten, aided by Atiyah and Segal, had
seen how these invariants could be viewed via Chern–Simons theory. One aspect
which I liked very much was the geometric quantization of the moduli space of
flat unitary connections on a surface. As a research student I had learned a bit
6 Geometry and physics
about quantization from another student John Rawnsley, but I never understood
pairings between different polarizations. The interpretation here as a projectively
flat connection on a bundle over Teichmüller space was more appealing, and I
set to work with the aim of giving a talk at Michael Atiyah’s 60th birthday
meeting in 1989. For some reason, I chanced upon the paper of Welters (1983)
using a heat equation in the abelian case, and before long I saw that the Poisson-
commuting functions of my integrable system were instrumental in defining this
connection for non-abelian theta functions. My construction was a mixture of
algebraic geometry and differential geometry, but Chapter 10 by Jørgen Andersen
shows how much further one can go by a deeper analysis of the non-abelian heat
equation.
In January 1990 I left Oxford for Warwick, and began to learn about the
Painlevé equations. It was really the work of Paul Tod in Oxford on self-dual four-
manifolds with SO(3) symmetry that spurred my interest. The two-monopole
metric which Atiyah and I had calculated using complete elliptic integrals was one
such example, but one knew that the general solution of the Painlevé equation
involved genuinely new transcendental functions, so there was a question about
how far one could go in hoping to get closed-form expressions for such metrics.
On the other hand twistor theory said that these could all be described by
a complex three-manifold, so I tried to approach the problem that way by
looking at threefolds with an action of the complex group SO(3, C). This led
me to some algebraic solutions which I spoke about for the 60th birthdays of
Narasimhan and Seshadri (Hitchin 1996). It also led into the classical geometry of
Poncelet’s theorem, which I had learned about earlier from Atiyah in the context
of hyperbolic monopoles. The twistor spaces in these cases were equivariant
compactifications of SO(3, C) modulo a dihedral group and the obvious ques-
tion was to find solutions of Painlevé’s sixth equation for any finite subgroup.
Chapter 11 by Philip Boalch describes this subject and its general context well.
He has given a classification of icosahedral solutions, the most challenging case.
I am only now beginning to understand the twistorial interpretation of any of
these.
It was a visit of Jean-Luc Brylinski to Warwick that got me interested in
the subject of gerbes, an interest that was later reinforced by conversations
with Dan Freed. Chapter 12 by Michael Murray recalls the enthusiasm that I
conveyed, though it was more in terms of the potential of the subject rather
than any theorems. I felt that at the right time and place I would see the
correct role of gerbes and that I should look out for that in my research. Some
years later, after I moved to Cambridge, I noticed that there was a twistor
transform analogous to the Atiyah–Ward approach to the self-dual Yang–Mills
equations and I had my student David Chatterjee write his thesis on various
aspects of gerbes, including this one. Murray’s chapter carefully leads the reader
through his bundle gerbe approach, connecting it with my own viewpoint. One of
Chatterjee’s contributions was to use the harmonic theory of the current defined
by a codimension three submanifold to define a gerbe with connection – the direct
Geometry and physics 7
geometry on the moduli spaces coming from physics. The moduli space of
complex structures on a Calabi–Yau threefold led me to the study of invariant
functionals on differential forms (Hitchin 2001b). This seemed to give a natural
setting for the central objects studied in string theory and M-theory, but also
a practical one when considering the associated flow equations which describe
the geometry in terms of an evolving hypersurface geometry. Chapter 17 by
Robert Bryant addresses the initial value problems associated with such flows,
using the methods of exterior differential systems. My approach to Calabi–Yau
complex structures was based on the fact that the action of GL(6, R) on Λ3 R6
has an open dense orbit; G2 structures appeared from a similar remark about
the group GL(7, R). In both instances I learned a lot about the background to
these facts from Bryant. I began to notice the occurrence of this type of situation
in a number of rather different places, most notably after a lecture in London by
Merkulov on the holonomy of affine connections, and I began wondering about
a geometrical use of SO(6, 6) in the spin representation.
In the Spring of 2001 I spent a sabbatical term in Madrid, visiting Oscar
Garcı́a-Prada at the Universidad Autónoma. I went with the intention of finishing
a book on hyperkähler manifolds (which has still not been written) but I
got distracted by thinking again about the common features of SL(6, R) and
SO(6, 6). The first week in May seemed like a succession of public holidays in
Madrid and it rained all the time that year but at the end of it I had the concept
of a generalized complex structure – a hybrid object that interpolates between
symplectic and complex geometry and should, I thought, have something to
contribute to mirror symmetry. Moreover the role of two-forms seemed to parallel
the physicists’ B-field. Even more, the natural setting seemed to require a gerbe in
the background. Marco Gualtieri, who wrote his thesis on the subject, addresses
this in Chapter 18, discussing the objects within this generalized area which
replace holomorphic vector bundles and submanifolds, for which he adopts the
physicists’ term brane. There are still many questions unanswered in this theory,
not least a tantalizing connection with noncommutative geometry. In this area,
the twisted de Rham complex – replacing the exterior derivative d by d + H
where H is a closed three-form – is a natural process. The K-theoretic analogue
of this uses the gerbe idea much more closely, and Chapter 19 by Dan Freed,
a continuation of a series with Mike Hopkins and Constantin Teleman, focuses
again on the equivariant twisted K-theory of a Lie group. I have much to learn
about twisted K-theory and the cycles that represent it, but it clearly fits into
the framework of the generalized geometry around which much of my current
research revolves.
If any conclusion can be drawn from this personal timeline, it is that it is
specific problems that have engaged me in research projects, and that theoretical
physics is perhaps the richest source of these. I have generally sought to learn
about an area of mathematics by seeing how it impinges on a particular problem,
which by itself might seem unimportant (the ‘Miss Marple’ approach). So
progress this way requires an ever-changing source of mathematical challenges,
References 9
of the sort that string theory seems to provide in abundance. This is surely
one of the reasons for the currently active interface between geometry and
physics.
Acknowledgements
I would like to wholeheartedly thank the contributors to this book, including
some who were unable to be at the Madrid conference. It is a wonderfully rich
cross-section of articles which appeal to my own interests. I would like to thank
Oxford University Press for agreeing to publish the book, and the editors for
their hard work in assembling it. Above all I would like to express my immense
thanks to Oscar Garcı́a-Prada for instigating the whole project – the original
2006 conference and this book based on it.
References
Donaldson, S. K. (1984). Nahm’s equations and the classification of monopoles. Com-
mun. Math. Phys. 96, 387–407.
Hitchin, N. J. (1974). Compact four-dimensional Einstein manifolds. J. Differential
Geometry 9, 435–41.
Hitchin, N. J. with Atiyah, M. F. and Singer, I. M. (1978). Self-duality in four-
dimensional Riemannian geometry. Proc. R. Soc. London Ser A 362, 425–61.
Hitchin, N. J. (1983). On the construction of monopoles. Commun. Math. Phys. 89,
145–90.
Hitchin, N. J. (1986). Metrics on moduli spaces, The Lefschetz centennial conference,
Part I (Mexico City, 1984). Contemp. Math. 58, American Mathematical Society,
Providence, RI, 157–78.
Hitchin, N. J. with Karlhede, A. Lindström, U. and Roček, M. (1987a). Hyperkähler
metrics and supersymmetry. Commun. Math. Phys. 108, 535–89.
Hitchin, N. J. (1987b). The self-duality equations on a Riemann surface. Proc. London
Math. Soc. 55(3), 59–126.
Hitchin, N. J. (1992). Lie groups and Teichmüller space. Topology 31, 449–73.
Hitchin, N. J. (1996). Poncelet polygons and the Painlevé equations, in Geometry and
Analysis. (Bombay, 1992), Oxford University Press, Bombay, India, pp. 151–85.
Hitchin, N. J. (1997). The moduli space of special Lagrangian submanifolds. Ann.
Scuola Norm. Sup. Pisa Cl. Sci. 25, 503–15.
Hitchin, N. J. (1998). Hypercomplex manifolds and the space of framings, in The
Geometric Universe: Science, Geometry and the Work of Roger Penrose. S. A.
Huggett et al. (eds.), Oxford University Press, Oxford, pp. 9–30.
Hitchin, N. J. (1999). The moduli space of complex Lagrangian submanifolds. Asian J.
Math. 3, 77–92.
Hitchin, N. J. (2001a). Lectures on special Lagrangian submanifolds, in Winter School
on Mirror Symmetry, Vector Bundles and Lagrangian Submanifolds. C. Vafa and
S. T. Yau (eds.), Studies in Advanced Mathematics 23, AMS/International Press,
Providence, RI, pp. 151–82.
Hitchin, N. J. (2001b). Stable forms and special metrics, in Global Differential Geom-
etry: The Mathematical Legacy of Alfred Gray. M. Fernández and J.A. Wolf (eds.),
10 Geometry and physics
It was a great pleasure for me to take part in the 60th birthday celebration
for Nigel Hitchin. I have known him ever since he was a graduate student in
Oxford and subsequently as my assistant at the institute in Princeton and then
as a colleague and collaborator. I have watched him mature mathematically
over the years and gradually establish a unique niche for himself on the frontier
between differential geometry and theoretical physics. This is now a very popular
and active field but Nigel occupies a singular place in it by virtue of the many
important and beautiful ideas he has introduced.
It has been said that the real measure of a mathematician’s contribution is
how long it would have taken the community to discover these results in his
absence. In Nigel’s case it is clear that he would score highly by this criterion,
since many of his results have been somewhat neglected on their first appearance
and their significance has only become apparent some years later.
Nigel has a large output with many beautiful papers which deserve to be read
and re-read. Because of the constraints of time I will limit myself to half-a-dozen
topics which I particularly like, and which show Nigel’s real knack of finding
natural and elegant ideas.
I will omit reference to the papers we have written jointly, except tangentially,
but I would like to mention our work on instantons which became known
as the ADHM construction (Atiyah et al. 1978). One morning, after a long
struggle, we finally saw the light. We adjourned to lunch in St Catherine’s College
euphoric at our success, though post-prandial analysis often punctures premature
celebration. On this occasion there was no unseen error but instead there was a
letter from Manin informing us that he and Drinfeld had just reached the same
conclusion! This was the genesis of the four-author paper and it was some years
later before Nigel or I met Drinfeld.
My first topic, which dates back to 1982, is Nigel’s introduction of the spectral
curve of a magnetic monopole (Hitchin 1982). I remember his letter to me at the
time saying he had found something which he felt I would like. He was completely
right. The result was really simple, beautiful, and important. It was at the basis
of much subsequent work on monopoles by many people, as I will recall.
Let me first review the details.
12 Mathematical work of Nigel Hitchin
DA φ = ∗FA
(DA − iφ)s = 0
FA + [φ, φ∗ ] = 0
dA φ = 0
Mathematical work of Nigel Hitchin 13
the physics of gauge theories and the formal quotienting process can be rigorously
justified.
Cases where this programme has been successfully used include the work
of Kronheimer (1989) on ALE spaces, moduli spaces of instantons on R4 and
monopoles or Higgs bundles on R3 .
A more recent application has been (Atiyah and Bielawski 2002) the construc-
tion of hyperkähler metrics on manifolds associated with nilpotent conjugacy
classes in Lie groups. The sub-regular class leads to the four-dimensional man-
ifolds studied by Kronheimer (extending the work of Brieskorn), while other
classes lead to higher dimensional generalizations.
All of this has totally transformed the subject of hyperkähler manifolds.
While compact manifolds are rare, non-compact ones are numerous and of great
interest. Before this general theory was developed Nigel wrote an elegant little
paper (Hitchin 1979) which has always been one of my favourites. It deals with
the Gibbons–Hawking gravitational instantons (Gibbons and Hawking 1978).
These are the special cases of ALE manifolds associated to the An singularities,
that is, C2 /Γ where Γ is cyclic. What Nigel did was to show how the Gibbons–
Hawking metric could be simply derived by twistor methods. These connected
naturally to the algebraic geometry used by Brieskorn and the outcome is the
hyperkähler extension of the Brieskorn theory, which includes the metric as well
as the complex structure.
Two things should be noted about this paper of Nigel’s. First it eventually
led to Kronheimer’s complete treatment of all the ADE singularities. Second
it showed in an explicit way how twistor theory could be harnessed in such
problems. In fact twistor theory lies behind much of Nigel’s work and he was
one of the first to successfully add twistor theory to other techniques in the
differential geometric study of moduli spaces. This was illustrated again in our
joint work on monopoles (Atiyah and Hitchin 1988) as well, of course, as in the
better known work on instantons (Atiyah et al. 1978). It is interesting that a
new link between twistor theory and gauge theory has been initiated by Witten
(2004).
Many mathematicians reaching the age of 60 may be slowing down, at least
in terms of producing really novel ideas. This is not the case with Nigel, so let
me end with two examples of more recent work, each of which illustrates Nigel’s
knack of identifying fruitful new geometric ideas. In both cases physicists have
been quick to latch on to Nigel’s work and exploit it.
The first is his use of stable three forms in connection with special metrics
(Hitchin 2001). A three-form ρ on a linear space V is said to be stable if it is in
an open orbit of GL(V ). There are two interesting special cases:
References
Atiyah, M. F. (1990). The Geometry and Physics of Knots. Cambridge University Press.
Atiyah M. F. and Bielawski, R. (2002). Nahm’s equations, configuration spaces and
flag manifolds. Bull. Braz. Math. Soc., New Ser. 33(2), 157–76.
Atiyah, M. F., Drinfeld, V. G., Hitchin, N.J. and Manin, Yu. I. (1978). Construction
of instantons. Phys. Lett. A 65, 185–7.
Atiyah, M. F. and Hitchin, N. J. (1988). The Geometry and Dynamics of Magnetic
Monopoles. Princeton University Press.
Gibbons, G. W. and Hawking, S. W. (1978). Grativational multi-instantons. Phys.
Lett. B 78, 430.
Hitchin, N. J. (1979). Polygons and gravitons. Math. Proc. Camb. Phil. Soc. 85, 465–76.
Hitchin, N. J. (1982). Monopoles and geodesics. Commun. Math. Phys. 83, 579–602.
Hitchin, N. J. (1987). The self-duality equations on a Riemann surface. Proc. London
Math. Soc. 55(3), 59–126.
Hitchin, N. J. (2001). Stable forms and special metrics. Contemp. Math. 288, 70–89.
Hitchin, N. J. (2003). Generalized Calabi-Yau manifolds. Q. J. Math. 54, 281–308.
Hitchin, N. J., Karlhede, A., Lindström, U. and Roček, M. (1987). Hyperkähler metrics
and supersymmetry. Commun. Math. Phys. 108, 535–89.
Kapustin, A. and Witten, E. (2007). Electro-magnetic duality and the geometric
Langlands program, hep-th/0604151.
Kronheimer, P. B. (1989). The construction of ALE spaces as hyper-Kahler quotients.
J. Diff. Geom. 29, 665–83.
Manton, N. J. and Sutcliffe, P. M. (2004). Topological Solitons. Cambridge University
Press, Cambridge.
Seiberg, N. and Witten, E. (1996). Gauge Dynamics and Compactification to Three-
dimensions in The Mathematical Beauty of Physics: A Memorial Volume for Claude
Itzykson Saclay, France, June 5–7 (Advanced Series in Mathematical Physics, Vol.
24) by Claude Itzykson, Jean Bernard Zuber, and Jean-Michel Drouffe, pp. 333–66.
Witten, E. (1991). Quantization of Chern-Simons gauge theory with complex gauge
group. Commun. Math. Phys. 137, 29–66.
Witten, E. (2004). Perturbative gauge theory as a string theory in twistor space.
Commun. Math. Phys. 252, 189–258.
III
THE EINSTEIN–MAXWELL EQUATIONS, EXTREMAL
KÄHLER METRICS, AND SEIBERG–WITTEN THEORY
Claude LeBrun
Dedicated to Nigel Hitchin on the occasion of his 60th birthday
dF = 0
dF = 0
where F is again allowed to vary over all closed two-forms in a given de Rham
class, and g is allowed to vary over all Riemannian metrics of some fixed total
volume V . The privileged status of dimension 4 becomes more pronounced in
this context, for it is only when n = 4 that the Einstein–Maxwell equations
imply that the scalar curvature s of g is constant. Indeed, this just reflects
Yamabe’s observation (Yamabe 1960) that a Riemannian metric has constant
scalar curvature
iff it is a critical point of the restriction of the Einstein–Hilbert
action sdµ to the space of volumeV metrics in a fixed conformal class. When
n = 4, the conformal invariance of |F |2 dµ thus implies that critical points of
the above functional must have constant scalar curvature; but when n = 4, by
contrast, the scalar curvature turns out to be constant only when F has constant
norm.
We have just observed that the Einstein–Maxwell equations on a four-manifold
imply that the scalar curvature is constant. But what happens in the converse
direction is far more remarkable: namely, any constant-scalar-curvature Kähler
metric on a four-manifold may be interpreted as a solution of the Einstein–
Maxwell equations. Indeed, suppose that (M 4 , g, J) is a Kähler surface, with
Kähler form ω = g(J·, ·) and Ricci form ρ = r(J·, ·). Let
s
ρ̊ = ρ − ω
4
denote the primitive part of the Ricci form, corresponding to the trace-free Ricci
tensor
s
r̊ := [r]0 = r − g.
4
Suppose that the scalar curvature s of (M, g) is constant, and set
ρ̊
F =ω+ .
2
Then (g, F ) automatically solves the Einstein–Maxwell equations. This general-
izes an observation due to Flaherty (1978) concerning the scalar-flat case.
Einstein–Maxwell equations 19
Since in our special case we have F + = ω and F − = ρ̊/2, it therefore follows that
F ◦ F = −r̊,
0
so that
r+F ◦F =0
0
as required. Moreover, since ρ is automatically closed, and its self-dual part sω/4
is closed if s is assumed to be constant, we conclude that F is indeed harmonic
on a constant-scalar-curvature Kähler surface, exactly as claimed.
But there is actually a great deal more going on here. Recall that Calabi (1982)
defined an extremal Kähler metric on a compact complex manifold (M 2m , J) to
be a Kähler metric which is a critical point of the Riemannian functional
g −→ s2 dµ
M
and
g −→ |R|2g dµg
M
2
obtained by squaring the L norms of the Ricci curvature r and the full Riemann
curvature R. Here, his observation was that the restriction of either of these
functionals to the space of Kähler metrics can be rewritten in the form
g −→ a + b s2 dµ
M
where a and b > 0 are constants depending only on the Kähler class. For example,
1 8π2
|r|g dµg =
2
s2g dµg − c2 · [ω]m−2 ,
M 2 M (m − 2)! 1
so that
8π 2 m (c1 · [ω]m−1 )2
|r|2g dµg ≥ − c21 · [ω]m−2 , (3.2)
M (m − 2)! m − 1 [ω]m
with equality iff s is constant. Thus extremal Kähler metrics turn out to minimize
these functionals too.
I would now like to point out an interesting Riemannian analog of Calabi’s
variational problem that leads to the Einstein–Maxwell equations on a smooth
compact four-manifold. To this end, first notice that the Kähler form of a Kähler
surface is self-dual and harmonic. Let us therefore introduce the following notion:
Definition 3.1 Let M be smooth compact oriented four-manifold, and let [ω] ∈
H 2 (M, R) be a deRham class with [ω]2 > 0. We will then say that a Riemannian
metric g is adapted to [ω] if the harmonic form ω representing [ω] with respect
to g is self-dual.
This allows us to introduce the Riemannian analog of a Kähler class:
Definition 3.2 Let M be smooth compact-oriented four-manifold, and let [ω] ∈
H 2 (M, R) be a deRham class with [ω]2 > 0. We then set
G = smooth metrics g on M ,
and
d 1 a
[dµ] = h dµ,
dt t=0 2 a
so that
d 2 ˙
s dµ = 2sṡdµ + s2 dµ
dt M t=0 M M
= 2s ∆haa + ∇a ∇b hab − hab r̊ab dµ
Theorem 3.1 Let (M 4 , J) be a compact complex surface, and suppose that [ω]
is a Kähler class with c1 · [ω] ≤ 0. Then any Riemannian metric g ∈ G[ω] satisfies
the inequalities
(c1 · [ω])2
s2 dµ ≥ 32π 2 (3.3)
[ω]2
(c · [ω])2
1
|r|2 dµ ≥ 8π 2 2 − c2
1 (3.4)
[ω]2
with equality iff g is constant-scalar-curvature Kähler.
In the equality case, the complex structure J˜ with respect to which g is Kähler
will typically be different from J, but must have the same first Chern class c1 ,
while its Kähler class must be a positive multiple of [ω].
We also remark that if (M, J) is not rational or ruled, the hypothesis that
c1 · [ω] ≤ 0 holds automatically, and that in this setting a Kähler metric is
extremal iff it has constant scalar curvature. In this context, the relevant constant
is of course necessarily non-positive.
By contrast, if (M, J) is rational or ruled, there will always be Kähler classes
for which c1 · [ω] > 0. When this happens, the above generalization (3.3) of (3.1)
turns out definitely not to hold for arbitrary Riemannian metrics. Instead, the
correct generalization (LeBrun 1997) is that
4π c1 · [ω]
Y[g] ≤ , (3.5)
[ω]2 /2
where the Yamabe constant Y[g] is obtained by minimizing the Einstein–Hilbert
action sdµ over all unit-volume metrics g̃ = u2 g conformal to g. Moreover, the
inequality is strict unless the Yamabe minimizer is a constant-scalar-curvature
Kähler metric, so that (3.3) is in fact violated by an appropriate conformal
rescaling of any generic Riemannian metric of positive scalar curvature.
It is also worth remarking that no sharp lower
bound in the spirit of Theorem
3.1 is currently known for the square-norm |R|2 dµ of the Riemann curvature
tensor. Deriving one would be extremely interesting and potentially very useful,
but, for reasons I will now explain, the technical obstacles to doing so seem
formidable.
24 Einstein–Maxwell equations
Here W± are the trace-free pieces of the appropriate blocks, and are called the
self-dual and anti-self-dual Weyl curvatures, respectively. The scalar curvature
s is understood to act by scalar multiplication, whereas the trace-free Ricci
curvature r̊ = r − 4s g acts on two-forms by ϕab → 2ϕ[a c r̊b]c .
When (M, g) happens to be Kähler, Λ2,0 ⊂ ker R, and the entire upper-left-
hand block is therefore entirely determined by the scalar curvature s. For Kähler
metrics, one thus obtains the identity
s2
|W+ |2 ≡ ,
24
and Gauss–Bonnet-type formulae like
2
1 s |r̊|2
(2χ + 3τ )(M ) = + 2|W+ | −
2
dµ
4π 2 M 24 2
reduce many questions about square norms of curvature to questions about the
scalar curvature alone. But for general Riemannian metrics, the norms of s and
W+ are utterly independent quantities, so if one wants to use the identity
2
s 1
|r|2 dµ = −8π 2 (2χ + 3τ )(M ) + 8 + |W+ |2 dµ (3.7)
24 2
to prove a generalization of (3.2) for Riemannian metrics, information must be
obtained concerning not only the scalar curvature, but also concerning the self-
dual Weyl curvature as well.
The curvature estimates of Theorem 3.1 are derived by means of Seiberg–
Witten theory (Witten 1994), making it clear that this really is an essentially
four-dimensional story. The complex structure J determines a spinc structure on
M with twisted spin bundles S± ⊗ L1/2 , where L−1 is the canonical line bundle
Λ2,0 of (M, J). For simplicity, suppose that c1 · [ω] < 0. For each metric g ∈ G[ω] ,
one then considers the Seiberg–Witten equations
DA Φ = 0
1
FA+ = − Φ Φ̄
2
Einstein–Maxwell equations 25
where [c1 (L)] is the orthogonal projection of c1 (L) ∈ H 2 (M, R) = Hg+ ⊕ Hg−
+
into the space Hg+ of harmonic self-dual two-forms, defined with respect to g.
Since ω is assumed to be self-dual with respect to g, we therefore have
(c1 · [ω])2
[c1 (L)+ ]2 ≥
[ω]2
by the Cauchy–Schwarz inequality. Inequality (3.3) follows. Since yet another
Cauchy–Schwarz argument shows that
2 √
s 1 1
+ |W+ | dµg ≥
2
(s − 6|W+ |)2 dµg ,
24 2 36
the second inequality and (3.7) together imply (3.4). The fact that only Kähler
metrics can saturate (3.3) or (3.4) is then deduced by examining the relevant
Weitzenböck formulae.
One might be tempted to expect the story to be similar for the norm of the
full Riemann tensor. After all, the identity
2
s
|R| dµ = −8π (χ + 3τ )(M ) + 2
2 2
+ 2|W+ | dµg
2
M M 24
certainly provides a good analog of (3.7). In the Kähler case, one has
s2
= |W+ |2 ,
24
so this simplifies to become
1
|R|2 dµ = 8π 2 (c2 − c21 ) + s2 dµg ,
M 4 M
26 Einstein–Maxwell equations
(χ + c21 )2
= 8π 2 −(χ + 3τ )(X) +
(2χ + c21 )
On the other hand, there are symplectic forms on X which are compatible with
the non-standard orientation of X; for example, the cohomology class F + εc1 is
represented by such forms if ε is sufficiently small. A celebrated theorem of
Taubes (1994) therefore tells us that the reverse-oriented version M = X of X
has a non-trivial Seiberg–Witten invariant (Kotschick 1998; Leung 1996). The
relevant spinc structure on X is of almost-complex type, and its first Chern class,
which we will denote by c̄1 , is given by
c̄1 · [ω ] (χ + 3 τ )
(c̄1 )+ = ω = − ω ,
[ω ]2 [ω ]2
(χ + 3 τ )2 (χ + 3 τ )2
|(c̄1 )+ |2 = = .
[ω ] 2 (2χ + c21 )
so that
(χ + 3 τ )2
|(c̄1 )− |2 = 2χ − 3τ + ,
(2χ + c21 )
and
(χ + 3 τ )2
|(c̄1 )− |2 − (χ − 3τ )(X) = χ(X) + ,
(2χ + c21 )
28 Einstein–Maxwell equations
Proof. Let us begin by remembering the remarkable fact (Barth, Peters, and
Van de Ven 1984; Buchdahl 1999; Siu 1983) that a compact complex surface is
of Kähler type iff it has b1 even. Consequently, for any non-Kähler-type complex
surface M , b1 is odd, and b+ = 2pg , where pg = h2,0 is the geometric genus
(Barth, Peters, and Van de Ven, 1984, Theorem IV.2.6). Since the latter is
assumed to vanish, M then has negative-definite intersection form, and Hodge
theory tells us that there are no non-trivial self-dual harmonic two-forms for any
metric g on M .
Now suppose that (g, F ) is a Riemannian solution of the Einstein–Maxwell
equations on M . Then the harmonic two-form F satisfies F + = 0, and hence
r̊ = −[F ◦ F ]0 = −2F + ◦ F − = 0.
This chapter has endeavored to convince the reader that the four-dimensional
Einstein–Maxwell equations represent a beautiful and natural generalization of
the constant-scalar-curvature Kähler condition. However, it still remains to be
seen whether solutions exist on many compact four-manifolds other than complex
surfaces of Kähler type. In this direction, my guess is that Proposition 3.3 will
actually prove to be rather misleading. For example, there are ALE Riemannian
solutions of the Einstein–Maxwell equations, constructed (Yuille 1987) via the
Israel–Wilson ansatz, on manifolds very unlike any complex surface. It thus
seems reasonable to conjecture that there are plenty of compact solutions that
in no sense arise from Kähler geometry. Perhaps some interested reader will feel
inspired to go out and find some!
Acknowledgments
The author would like to thank Maciej Dunajski, Caner Koca, and Galliano
Valent for their interesting comments and queries regarding an earlier version
References 31
References
Apostolov, V. and Tønnesen-Friedman, C. (2006). A remark on Kähler metrics of
constant scalar curvature on ruled complex surfaces. Bull. London Math. Soc. 38(3),
494–500.
Arezzo, C. and Pacard, F. (2006). Blowing up and desingularizing constant scalar
curvature Kähler manifolds. Acta Math. 196(2), 179–228.
Arezzo, C. and Pacard, F. (2009). Blowing up and desingularizing Kähler orbifolds of
constant scalar curvature II. Ann. Math. 170, 685–738. E-print math.DG/0504115.
Arezzo, C., Pacard, F. and Singer, M. (2007). Extremal metrics on blow ups. E-print
math.DG/0701028.
Atiyah, M. F. (1969). The signature of fibre-bundles, in Global Analysis (Papers in
Honor of K. Kodaira), University Tokyo Press, Tokyo, pp. 73–84.
Atiyah, M. F., Hitchin, N. J. and Singer, I. M. (1978). Self-duality in four-dimensional
Riemannian geometry. Proc. R. Soc. London Ser. A 362, 425–61.
Aubin, T. (1976). Equations du type Monge-Ampère sur les variétés kählériennes
compactes. C. R. Acad. Sci. Paris A 283, 119–21.
Barth, W., Peters, C. and Van de Ven, A. (1984). Compact Complex Surfaces. Springer-
Verlag, Berlin, Heidelberg, New York, Tokyo.
Besse, A. L. (1987). Einstein manifolds, Vol. 10 of Ergebnisse der Mathematik und ihrer
Grenzgebiete (3) [Results in Mathematics and Related Areas (3)], Springer-Verlag,
Berlin.
Buchdahl, N. (1999). On compact Kähler surfaces. Ann. Inst. Fourier (Grenoble)
49(1), 287–302.
Calabi, E. (1982). Extremal Kähler metrics, in Seminar on Differential Geometry, Vol.
102 of Ann. of Math. Stud., Princeton University Press, Princeton, NJ, pp. 259–90.
Chen, X. X. (2006). Space of Kähler metrics III on the lower bound of the Calabi energy
and geodesic distance. E-print math.DG/0606228.
Chen, X. X., LeBrun, C. and Weber, B. (2008). On conformally Kähler, Einstein
manifolds. J. Am. Math. Soc. 21(4), 1137–1168. E-print arXiv:0705.0710.
Chen, X. X. and Tian, G. (2005). Uniqueness of extremal Kähler metrics. C. R. Math.
Acad. Sci. Paris 340(4), 287–90.
Donaldson, S. K. (1986). Connections, cohomology and the intersection forms of 4-
manifolds. J. Differential Geom. 24(3), 275–341.
Donaldson, S. K. (2001). Scalar curvature and projective embeddings. I. J. Differential
Geom. 59(3), 479–522.
Fine, J. (2004). Constant scalar curvature Kähler metrics on fibred complex surfaces.
J. Differential Geom. 68(3), 397–432.
Flaherty, E. J., Jr. (1978). The nonlinear graviton in interaction with a photon. Gen.
Relativity Gravit. 9(11), 961–78.
Fujiki, A. and Schumacher, G. (1990). The moduli space of extremal compact Kähler
manifolds and generalized Weil-Petersson metrics. Publ. Res. Inst. Math. Sci.
26(1), 101–83.
Gay, D. T. and Kirby, R. (2004). Constructing symplectic forms on 4-manifolds which
vanish on circles. Geom. Topol. 8, 743–77 (electronic).
32 Einstein–Maxwell equations
Ross, J. (2006). Unstable products of smooth curves. Invent. Math. 165(1), 153–62.
Ross, J. and Thomas, R. (2006). An obstruction to the existence of constant scalar
curvature Kähler metrics. J. Differential Geom. 72(3), 429–66.
Shu, Y. (2006). Compact complex surfaces and constant scalar curvature Kähler
metrics. E-print math.DG/0612013.
Siu, Y. (1983). Every K3 surface is Kähler. Invent. Math. 73, 139–50.
Taubes, C. H. (1994). The Seiberg-Witten invariants and symplectic forms. Math. Res.
Lett. 1, 809–22.
Witten, E. (1994). Monopoles and four-manifolds. Math. Res. Lett. 1, 809–22.
Yamabe, H. (1960). On a deformation of Riemannian structures on compact manifolds.
Osaka Math. J. 12, 21–37.
Yau, S. T. (1977). Calabi’s conjecture and some new results in algebraic geometry.
Proc. Natl. Acad. USA 74, 1789–99.
Yuille, A. L. (1987). Israel-Wilson metrics in the Euclidean regime. Classical Quantum
Gravity 4(5), 1409–26.
IV
THE NAHM TRANSFORM FOR CALORONS
4.1 Introduction
One rather mysterious feature of the self-duality equations on R4 is the exis-
tence of a quite remarkable non-linear transform, the Nahm transform. It maps
solutions to the self-duality equations on R4 invariant under a closed translation
group G ⊂ R4 to solutions to the self-duality equations on (R4 )∗ invariant under
the dual group G∗ . This transform uses spaces of solutions to the Dirac equation,
it is quite sensitive to boundary conditions, which must be defined with care,
and it is not straightforward: for example, it tends to interchange rank and
degree.
The transform was introduced by Nahm as an adaptation of the original
ADHM construction of instantons (Atiyah et al. 1978) having the advantage that
it can be generalized. The series of papers (Corrigan and Goddard 1984; Nahm
1983, 1984) details Nahm’s original transform. In the monopole case, G = R,
and the transform takes the monopole to a solution to some non-linear matrix
valued ordinary differential equations, Nahm’s equations on an interval, and there
is an inverse transform giving back the monopole. Much of the mathematical
development of this transform is due to Nigel Hitchin (1983), who showed for the
SU (2) monopole how the monopole and the corresponding solution to Nahm’s
equations are both encoded, quite remarkably, in the same algebraic curve, the
spectral curve of the monopole. This work was extended to the cases of monopoles
for the other classical groups by Hurtubise and Murray (1989). This extension
illustrates just how odd the transform can be: one gets solutions to the Nahm
equations on a chain of intervals, but the size of the matrices jump from interval
to interval.
The Nahm transform for other cases of G invariance has been studied by
various authors (among others, Braam and van Baal 1989; Charbonneau 2004;
Cherkis and Kapustin 2001; Jardim 2001, 2002a, 2002b; Nye 2001; Schenk 1988;
Szabó 2005); a nice survey can be found in Jardim (2004). Here, we study the
case of ‘minimal’ invariance, under Z; the fields are referred to as calorons, and
the Nahm transform sends them to solutions of Nahm’s equations over the circle.
This case is very close to the monopole one, and indeed calorons can be considered
as Kač–Moody monopoles (Garland and Murray 1988).
The work of Nye and Singer 35
Calorons have been studied from different angles by various authors, including
Bruckmann and van Baal (2002); Bruckmann et al. (2003, 2004), Chakrabarti
(1987), Kraan (2000), Kraan and van Baal (1998a; 1998b; 1998c), Lee (1998),
Lee and Lu (1998), Lee and Yi (2003), Nógrádi (2005), Norbury (2000), and
Ward (2004). In particular, many explicit solutions have been found.
This project started while we were studying the works of Nye (2001) and
Nye and Singer (2000) on calorons; they consider the Nahm transform directly,
and do most of the work required to show that the transform is involutive. The
missing ingredients turn out to lie in complex geometry, in precisely the same
way Hitchin’s extra ingredient of a spectral curve complements Nahm’s. The
complex geometry allows us to complete the equivalence, which can then be used
to compute the moduli. It thus seems to us quite appropriate to consider this
problem in a volume dedicated to Nigel Hitchin, as it allows us to revisit some
of his beautiful mathematics, including some on calorons which has remained
unpublished.
In Section 4.2, we summarize the work of Nye and Singer towards showing
that the Nahm transform is an equivalence between calorons and appropriate
solutions to Nahm’s equations. In Section 4.3, following in large part on the work
of Garland and Murray (1988), we describe the complex geometry (‘spectral
data’) that encodes a caloron. In Section 4.4, we study the process by which
spectral data also correspond to solutions to Nahm’s equations. In Section 4.5,
we close the circle, showing the two Nahm transforms are inverses. In Section
4.6, we give a description of moduli, expounded in Charbonneau and Hurtubise
(2007).
We require that the L2 norm of the curvature of A be finite, and that in suitable
gauges, the functions Ai be O(|x|−2 ) as |x| → ∞, and that φ be conjugate to
diag(i(µ1 − j/2|x|), i(−µ1 + j/2|x|)) + O(|x|−2 ) for a positive real constant µ1
and a positive integer j (the monopole charge). We also have bounds on the
derivatives of these fields.
The boundary conditions tell us in essence that in a suitable way the connec-
tion extends to the two-sphere at infinity times S 1 ; furthermore, one can show
that the extension is to a fixed connection, which involves fixing a trivialization
at infinity; there is thus a second invariant we can define, the relative second
Chern class, which we represent by a (positive) integer k. There are then two
integer charges for our caloron, k and j, the instanton and monopole charges,
respectively.
There is a suitable group of gauge transformations acting on these fields (Nye
2001): Nye’s approach is to compactify to S 1 × B̄ 3 with a fixed trivilization over
the boundary S 1 × S 2 . The gauge transformations are those extending smoothly
to the identity on the boundary. Two solutions are considered to be equivalent
if they are in the same orbit under this group.
4.2.1.2 Solutions to Nahm’s equations on the circle
The second class of objects we consider are skew adjoint matrix-valued functions
Ti (z), i = 0, . . . , 3, of size (k + j) × (k + j) over the interval (−µ1 , µ1 ), and of
size (k) × (k) over the interval (µ1 , µ0 − µ1 ) (hence µ1 < µ0 − µ1 , so we impose
that condition) that are solutions to Nahm’s equations
dTσ(1)
+ [T0 , Tσ(1) ] = [Tσ(2) , Tσ(3) ], for σ even permutations of (123), (4.1)
dz
on the circle R/(z → z + µ0 ). These equations are reductions of the self-duality
equations to one dimension, and are invariant under a group of gauge transfor-
d
mations under which the dz + T0 transforms as a connection.
We think of the Ti as sections of the endomorphisms of a vector bundle K
whose rank jumps at the two boundary points; at these points ±µ1 we need
boundary conditions.
Case 1: j = 0. At each of the boundary points, there is a large side, with a rank
(k + j) bundle, and a small side, with a rank k bundle. We attach the two at
the boundary point using an injection ι from the small side into the large side,
and a surjection π going the other way, with π · ι the identity. One asks, from
the small side, that the Ti have well-defined limits at the boundary point. From
the large side, one has a decomposition near the boundary points of the bundle
Ck+j × [−µ1 , µ1 ] into an orthogonal sum of subbundles of rank k and j invariant
d
with respect to the connection dz + T0 and compatible with the maps ι, π. With
respect to this decomposition the Ti have the form
T̂i O ẑ (j−1)/2
Ti = , (4.2)
O(ẑ (j−1)/2 ) Ri
The work of Nye and Singer 37
Over each interval, this bundle sits inside the trivial bundle whose fibre is the
space of L2 sections of V ⊗ S− . Let P be the orthogonal projection from this
trivial bundle onto K. As elements of K decay exponentially, the operation Xi
of multiplying by the coordinate xi can be used to define operators on sections
of K by
d d
+ T0 = P · ,
dz dz
Ti = P · Xi .
Theorem 4.1 (Nye 2001, section 4.1.2, p. 108) The operators defined in
this way satisfy Nahm’s equations.
This theorem and Theorem 4.2 below fall in line with the general Nahm
transform heuristic philosophy: the curvature of the transformed object, seen
as an invariant connection on R4 , is always composed of a self-dual piece and
another piece depending on the behaviour at infinity of the harmonic spinors.
Since the latter are decaying exponentially, that other piece is zero and the
transformed object is self-dual, or equivalently once we reduce, it satisfies Nahm’s
equations (see for instance Charbonneau 2006, section 3).
3
(Tj )+ − (Tj )− ⊗ ej = u+ ⊗ u∗+ 0 .
j=1
Twistor transform for calorons/Kač–Moody monopoles 39
Here the subscript ‘0’ signifies taking the trace-free End(K) ⊗ Sl(2, C) compo-
nent inside End(K) ⊗ End(C2 ) End(K ⊗ C2 ).
One has a similar vector u− at −µ1 . Let U be the vector space spanned by
u+ , u− . Let Π : Kµ1 ⊕ K−µ1 → U be the orthogonal projection, let X be the sum
of the space of L2 sections of K with the space U , and set
D̂x,t := (Dx,t , Π) : W → X (4.6)
The kernel of this operator consists of sections s in the kernel of Dx,t , with values
at ±µ1 lying in U ⊥ ; the cokernel consists of triples (s, cµ1 uµ1 , c−µ1 u−µ1 ), where
∗
s is a section of K ⊗ C2 , lying in the kernel of Dx,t , with jump discontinuity
c+ u+ at µ1 and c− u− at −µ1 .
Theorem 4.2 (Nye 2001, p. 104) The operator D̂x,t has trivial kernel for
all (x, t), and has a cokernel of rank 2, defining a rank 2 vector bundle over
R4 , with natural time periodicity which allows one to build a vector bundle V
over S 1 × R3 , defined locally as a subbundle of the infinite-dimensional bundle
X × S 1 × R3 . Let P denote the orthogonal projection from X to V . Setting, on
sections of V ,
∂
∇i = P · , i = 1, 2, 3
∂xi
∂
∇t = P · ,
∂t
defines an SU (2) caloron.
and Murray (1988, section 2), we summarize the construction, again for SU (2)
only.
It is convenient first to recall from Hitchin (1982, section 3) the twistor space
for R3 . It can be interpreted as the space of oriented lines in R3 and it is
isomorphic to T P1 , the tangent bundle of the Riemann sphere. Let ζ be the
standard affine coordinate on P1 , and η be the fibre coordinate in T P1 associated
to the basis vector ∂/∂ζ. The incidence relation between the standard coordinates
on R3 and (ζ, η) is given by
The space T P1 comes equipped with standard line bundles O(k), lifted from P1 ,
and line bundles Lt , parameterized by t ∈ C, with transition function exp(tη/ζ)
from the open set ζ = ∞ to the open set ζ = 0. For t ∈ R, these line bundles
correspond to the standard U (1) monopoles on R3 given by a flat connection
and constant Higgs field it. Let Lt (k) := Lt ⊗ O(k). These bundles are in some
sense the building blocks for monopoles for higher gauge groups.
According to Garland and Murray, the twistor space T for S 1 × R3 parame-
trizes pairs consisting of a point in S 1 × R3 and a unit vector in R3 . Such a pair
gives a cylinder along which to integrate, and that projects to an oriented line in
R3 . There is then a C∗ fibration π : T → T P1 , which is in fact holomorphic and
the complement of the zero section in Lµ0 . It has a natural fibrewise compactifica-
tion T̃ = P(O ⊕ Lµ0 ) compactifying the cylinders into spheres. The complement
T̃ \T is a sum of two disjoint divisors T 0 and T ∞ mapping isomorphically to
T P1 .
In the monopole case, the holomorphic vector bundle was obtained by inte-
grating ∇s − iφ along real lines using a metric coordinate s along the line, and
the Higgs field φ. Here, the analogous operation is integrating ∇s − i∇t over the
cylinders to obtain a vector bundle E over T . The boundary conditions allow
us to extend the bundle to the compactification T̃ ; we denote this extension
again by E. The boundary conditions also give line subbundles over these
divisors, E10 = L−µ1 (−j) over T 0 , corresponding to solutions of ∇s − iφ which
decay as one goes to plus infinity on the cylinder, and E1∞ = Lµ1 (−j) over T ∞ ,
corresponding to solutions which decay as one goes to minus infinity.
The twistor space has a real structure τ , which acts on the bundle by τ ∗ (E) =
E , and maps the line subbundle L−µ1 (−j) over T 0 to the annihilator of the
∗
Indeed, the twistor transform for the caloron gives us a bundle E over T .
Taking a direct image (restricting to sections with poles of finite order along 0, ∞)
gives one an infinite-dimensional vector bundle F over T P1 . If w is a standard
fibre coordinate on T vanishing over T 0 and with a simple pole at T ∞ , it induces
an endomorphism W of F , and quotienting F by the O-module generated by the
image of W − w0 gives us the restriction of E to the section w = w0 , so that E
and (F, W ) are equivalent. One has more, however; the fact that the bundle E
extends to T̃ gives a subbundle F 0 of sections in the direct image which extend
over T 0 , and a subbundle F ∞ of sections in the direct image which extend
over T ∞ . One can go further and use the flags 0 = E00 ⊂ E10 ⊂ E20 = E over T 0 ,
0 = E0∞ ⊂ E1∞ ⊂ E2∞ = E over T ∞ to define for p ∈ Z and q = 0, 1 subbundles
0 ∞
Fp,q , Fp,q of F as
0
F−p,q = s ∈ F | w−p s finite at T 0 with value in Eq0 ,
∞
Fp,q = s ∈ F | w−p s finite at T ∞ with value in Eq∞ .
· · · ⊂ F−1,0
0
⊂ F−1,1
0
⊂ F0,0
0
⊂ F0,1
0
⊂ F1,0
0
⊂ F1,1
0
⊂ ··· ,
∞ ∞ ∞ ∞ ∞ ∞
· · · ⊃ F2,0 ⊃ F1,1 ⊃ F1,0 ⊃ F0,1 ⊃ F0,0 ⊃ F−1,1 ⊃ ··· . (4.7)
∞ ∞
Note that W · Fp,q
0 0
= Fp+1,q and W · Fp,q = Fp−1,q . Garland and Murray show
r Fp,0
0 ∞
and F−p+1,0 have zero intersection and sum to F away from a compact
curve S0 lying in the linear system |O(2k)|.
r F 0 and F ∞ have zero intersection and sum to F away from a compact
p,1 −p,1
curve S1 lying in the linear system |O(2k + 2j)|.
r Quotients F 0 /F 0 (p−1)µ0 +µ1
p,0 p−1,1 are line bundles isomorphic to L (j).
r Quotients F 0 /F 0 are line bundles isomorphic to Lpµ0 −µ1 (−j).
p,1 p,0
r Quotients F ∞ /F ∞ are line bundles isomorphic to L−(p−1)µ0 −µ1 (j).
p,0 p−1,1
r Quotients F ∞ /F ∞ are line bundles isomorphic to L−pµ0 +µ1 (−j).
p,1 p,0
It is worthwhile stepping back now and seeing what we have from a group
theoretic point of view. On T P1 , we have a bundle with structure group
C) of Gl(2, C)-valued loops. The flags that we have found give two
G = Gl(2,
reductions R0 , R∞ to the opposite Borel subgroups B0 , B∞ (of loops extending
to 0, ∞, respectively, in P1 , and preserving flags over these points) in G. As for
finite-dimensional groups, we have exact sequences relating the Borel subgroups,
their unipotent subgroups, and the maximal torus:
0 → U0 → B0 → T → 0,
(4.8)
0 → U∞ → B∞ → T → 0.
42 The Nahm transform for calorons
0 → U0 → M → Pr → 0
φ
Pf0 (U0 ) → Pf0 (M) −−→ Pf0 (Pr),
We are now ready to define the principal part data of the caloron. The principal
part data of the caloron is the image under φ of the class of the reduction given
as a section R∞ of Pf0 ×B0 G/B∞ = Pf0 (M):
To get back the caloron bundle from the principal part data, one takes the
coboundary δ(φ(R∞ )), in the obvious Čech sense, for the second sequence,
obtaining a class in H 1 (T P1 , A0 (U0 )). One checks that this is precisely the
class corresponding to the bundle P0 ; while this seems a bit surprising, the
construction is fairly tautological, and is done in detail in (Hurtubise and Murray
1990, section 3). In our infinite-dimensional context, the bundles are quite special,
as they, and their reductions, are invariant under the shift operator W .
Of course, this construction does not tell us what the mysterious class φ(R∞ )
actually corresponds to, but it turns out that, over a generic set of calorons, it is
quite tractable. Indeed, the principal part data, as a sheaf, is supported over the
spectral curves, and, if the curves have no common components and no multiple
components, then the principal part data amounts to the following spectral data
(Garland and Murray 1988):
r The two spectral curves, S0 and S1
r The ideal IS0 ∩S1 , decomposing as IS01 · IS10 ; the real structure interchanges
the two factors
r An isomorphism of line bundles O[−S10 ] ⊗ Lµ0 −2µ1 O[−S01 ] over S0
r An isomorphism of line bundles O[−S01 ] ⊗ L2µ1 O[−S10 ] over S1
r A real structure on the line bundle O(2k + j − 1)[−S01 ] ⊗ Lµ1 |S1 , lifting the
real involution τ on T P1
The structure of this data can be understood in terms of the Schubert structure
of G/B0 . The sheaf of principal parts, by definition, lives in the complement of
U0 , where the two flags cease to be transverse. In the case that concerns us here
(remember the invariance under W ), there are essentially two codimension one
varieties at infinity to U0 that we consider, whose pull-backs give the two spectral
curves. The decomposition of S0 ∩ S1 into two pieces is simply a pull-back of the
Schubert structure from G/B0 . For the line bundles, the basic idea is that in
codimension one, the principal parts can be understood using embeddings of P1
into G/B0 given by principal Sl(2, C)’s in G (see for instance Boyer et al. 1994,
section 4), reducing the question at least locally to understanding principal parts
for maps into P1 , a classical subject. For simple poles, the principal part of a map
into P1 is encoded by its residue. Here, the poles are over the spectral curves,
and globally, the principal part is then encoded as a section of a line bundle over
each curve; we identify these below. A way of seeing that the spectral data splits
into components localized over each curve is that we can choose appropriately
44 The Nahm transform for calorons
two parabolic subgroups P ar0 , P ar1 and project from G/B0 to G/P ar0 and
G/P ar1 .
To see what the ‘residues’ correspond to here, we exploit a description of the
bundle first given in Hurtubise and Murray (1989, proposition 1.12), the short
exact sequence
.. ..
. .
0
∞
F/ Fp−1,1 + F−p+1,0 ⊕
⊕ 0 ∞
0 F/ F p,0 + F−(p−1),0
∞
0 −→ F −→ F/ Fp,0 + F−p,1 −→ ⊕ −→ 0. (4.9)
0
0 ⊕ F/ Fp,1 + F−p,1 ∞
∞
F/ Fp,1 + F−p,0 ⊕
.. ..
. .
0
The quotients at the right are supported over the spectral curves (F/(Fp,0 +
∞ 0 ∞
F−(p−1),0 ) over S0 , and F/(Fp,1 + F−p,1 ) over S1 ) and those in the middle are
generically line bundles.
The shift operator W moves the quotients in the middle and last columns
two steps down. The quotients are in fact direct image sheaves R1 π̃∗ of sheaves
on T̃ derived from E; for example, let Ep,0,−(p−1),∞ be the sheaf of sections
of E over T with
0 poles∞ of order p at T0 and a zero1 of order p − 1 at T∞ .
The sheaf F/ Fp,0 + F−p,1 can be identified with R π̃∗ (Ep,0,−(p−1),∞ ). One
can make similar identifications for the other sheaves. In a way that parallels
Hitchin (1983), we have the following result:
Proposition 4.3
0
1. The spectral curves S0 and S1 are real (τ -invariant); the quotient F/(Fp,1 +
∞
F−p,1 ) over S1 inherits from E a quaternionic structure, lifting τ .
0 ∞ 0 ∞
2. The sheaves F/(Fp,0 + F−(p−1),0 ) and F/(Fp,1 + F−p,1 ) satisfy a vanishing
theorem:
∞
H 0 T P1 , F/ Fp,0
0
+ F−(p−1),0 ⊗ L−z (−2) = 0,
the quotients
∞
0
F/ Fp,0 + F−(p−1),0 = Lpµ0 −µ1 (2k + j)[−S10 ]|S0
The ‘residues’ are then the sheaves on the right, supported over the spectral
curves; the maps from the middle to the right-hand sheaves give the various
isomorphisms. This diagram shows that F , and hence the caloron, is encoded in
the spectral data.
More generally, we can define the spectral data composed of the curves S0 , S1 ,
the pull-backs to T P1 of the Schubert varieties in G/B0 , and the sheaves on
the right-hand side of the sequence (4.9), with isomorphisms similar to the
ones for generic spectral data given by maps from the sheaves in the middle
column.
To summarize, we have that the caloron determines principal part data, a
section of a sheaf of principal parts supported over the spectral curve; this
data determines the caloron in turn. In the generic case, the principal part
data equivalent to spectral data can be described in terms of two curves, their
intersections, and sections of line bundles over these curves. We note that these
generic calorons exist; indeed, we already know that a caloron is determined by
a solution to Nahm’s equations, and it is easy to see that generic solutions to
Nahm’s equations exist, as we shall see in Section 4.6.
46 The Nahm transform for calorons
in the product T P1 × T P1 , and is cut out by ζ − ζ , and one has the exact
sequence
ζ−ζ
0 → O(−1, −1) −−−−→ O → OF P → 0. (4.17)
Lifting Lz to T P1 × T P1 , tensoring it with this sequence and taking direct image
on the other factor gives the long exact sequence
0 → H 0 (T P1 , Lz (−1)) ⊗ O(−1) → H 0 (T P1 , Lz ) ⊗ O → V
→ H 1 (T P1 , Lz (−1)) ⊗ O(−1) · · · .
The vanishing of the cohomolology of Lz (−1) shows that V Ok .
We then have given A(ζ, z) as a map of bundles; the process of turning it into
a matrix-valued function A(ζ, z), and of showing that it evolves according to
Nahm’s equations as one tensors it by Lt , is given in Hitchin (1983, proposition
4.16). Similarly, the definition of an inner product on H 0 (T P1 , Lz ) and the proof
that the Ti are skew adjoint with respect to the inner product follow the line
given in Hitchin (1983, section 6).
this filtration can be turned onto a sum, as we have the subsheaves Li of sections
η i s, with s lifted from P1 . The construction, by limiting to a curve which is a
j-sheeted cover of P1 , essentially says that the degrees of vanishing of interest
are at most j − 1, as one is taking the remainder by division by the equation of
the curve. The bundle V over P1 × {µ1 } decomposes as a sum
This process can be applied over any point ζ, as the matrices A(ζ), A+ (ζ) have
residues (in z) conjugate to (4.18) and (4.19). Indeed, at ζ = 0, they represent
standard generators of a representation of Sl(2), and moving away from ζ = 0
simply amounts to a change of basis. Explicitly, conjugating the residues of A(ζ),
A+ (ζ) by N (ζ) takes them to the residues of ζA(1), A+ (1). Conjugating again
by a matrix T takes them to the residues of ζA(0), A+ (0), and then by N (ζ)−1
to the residues of A(0), A+ (0). Thus, for a suitable
M (ζ, z) = (Holomorphic in z, ζ) · N (ζ −1 z)T N (ζ) (4.22)
one has
⎛ ⎞
−η 0 0. . . 0 −pj (ζ)
⎜ 1 −η 0. . . 0 −pj−1 (ζ) ⎟
⎜ ⎟
⎜ −η. . . 0 −pj−2 (ζ) ⎟
M (ζ, z)(A(ζ, z) − ηI)M (ζ, z)−1 =⎜ 0 1 ⎟. (4.23)
⎜ .. .. .. .. .. .. ⎟
⎝ . . . . . . ⎠
0 0 0 . . . 1 −η − p1 (ζ)
At z = 0, it is a holomorphic section of Hom(V (−2), V ) in standard trivializa-
tions.
From Nahm’s equations to spectral data, and back 51
ds
The same procedures work when k = 0. Indeed, the singular solution to dz +
A+ s has the same behaviour. Starting from a solution to Nahm’s equations, and
integrating the connection as in Hurtubise (1989), we find that, near µ1 on the
interval (−µ1 , µ1 ), a change of basis of the form
M (ζ, z) = (Holomorphic in z, ζ) · diag(Ik×k , N (ζ −1 z)T N (ζ)) (4.24)
conjugates A1 (ζ, z) − ηI to the constant (in z) matrix
⎛ ⎞
a11 (ζ) − η a12 (ζ) . . . a1k (ζ) 0 0 0 ... 0 g1 (ζ)
⎜ a21 (ζ) a22 (ζ) − η . . . a1k (ζ) 0 0 0 ... 0 g2 (ζ) ⎟
⎜ ⎟
⎜ .. .. .. .. .. .. .. .. .. .. ⎟
⎜ . . . . . . . . . . ⎟
⎜ ⎟
⎜ ak1 (ζ) a (ζ) . . . a (ζ) − η 0 0 0 ... 0 gk (ζ) ⎟
⎜ k2 kk ⎟
⎜ f1 (ζ) f2 (ζ) . . . fk (ζ) −η 0 0 . . . 0 −pj (ζ) ⎟
⎜ ⎟. (4.25)
⎜ 0 0 ... 0 1 −η 0 . . . 0 −pj−1 (ζ) ⎟
⎜ ⎟
⎜ 0 0 . . . 0 0 1 −η . . . 0 −pj−2 (ζ) ⎟
⎜ ⎟
⎜ .. .. .. .. .. .. .. .. .. .. ⎟
⎝ . . . . . . . . . . ⎠
0 0 ... 0 0 0 0 . . . 1 −η − p1 (ζ)
Setting C(p(ζ)) to denote the companion matrix of p(ζ), we write this matrix
schematically as
⎛ 0 ⎞
A(ζ) −ηI 0 G(ζ)
A1 (ζ) − ηI = ⎝ F (ζ) ⎠. (4.26)
C(p(ζ)) − ηI
0
Let Madj denote the classical adjoint of M , so that Madj M = det(M )I. Then
det(A1 (ζ) − ηI) = det(A0 (ζ) − ηI) η j + η j−i pi (ζ)
+ (−1)j F A0 (ζ) − ηI adj G. (4.27)
The matrix A0 (ζ) is equal to the limit A0 (ζ, µ1 ). At the boundary point µ1 ,
the bundle Vµ01 for S0 is trivial, since the solution on S0 is smooth at that point;
one has
A0 (ζ)−ηI
0 → O(−2)k −−−−−−→ O k → L0µ1 → 0. (4.28)
The limit bundle Vµ11 for S1 is
Lemma 4.6 Let m, n > 0. Lifting O(mC + nP∞ ) to the fibre product and
tensoring with the sequence (4.28), then pushing down, we obtain an exact
sequence
n−1
0 −→ O (m − 2i − 2) C + (n − i − 1)P∞
i=0
↓ S(n, η)
n (1,η,η2 ,...,ηn )
O (m − 2i)C+ (n − i)P∞ −−−−−−−−−−→ O(mC + nP∞ ) → 0.
i=0
Over T P , it becomes
1
We have
(φ1 , . . . , φk ) A0 (ζ) − ηI + F (ζ)(−1)k det A0 (ζ) − ηI = 0. (4.33)
k−1
Write φi = k−1
j=0 φji (ζ)η , and (−1) det(A (ζ) − ηI) = η +
j k 0 k j
j=0 hj (ζ)η .
Decomposing (4.33) into different powers of η, we obtain
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 ··· 0 φ01 · · · φ0k h0
⎜ φ01 · · · φ0k ⎟ ⎜ ..
⎟ ⎜
.. .. ⎟ ⎜ .. ⎟
⎜
−⎜ .. .. .. ⎟ + ⎜ . . . ⎟ ⎜ ⎟
⎟ A0 (ζ) + ⎜ . ⎟ F = 0. (4.34)
⎝ . . . ⎠ ⎝φk−1,1 · · · φk−1,k ⎠ ⎝hk−1 ⎠
φk−1,1 · · · φk−1,k 0 ··· 0 1
R(ζ,η)
O(−2)⊕k ⊕ Vj−3,−j+1 / O⊕k ⊕ Vj−1,−j+1 /R
Mj−1 Mj
S(k+j,η) (1,η,...,η k+j−1 )
V2k+j−3,−j+1 / V2k+j−1,−j+1 / O(2k + j − 1).
A0 (ζ)−ηI
O(−2)⊕k / O⊕k / L0µ
O O O1
Πk+j−1,k Πk+j,k
R(ζ,η)
O(−2)⊕k ⊕ Vj−3,−j+1 / O⊕k ⊕ Vj−1,−j+1 /R
Ik+j−1,k+j I
A1 (ζ)−ηI
⊕k
O(−2) ⊕ Vj−3,−j−1 / O⊕k ⊕ Vj−1,−j+1 / L1µ
1
intersections of the spectral curves used here are more general, but the proof
goes through unchanged.
of A(0, z) = T1 + iT2 (z); there is a sheaf over the curve whose fibre over the point
η in the spectral curve given by the cokernel of A(0, z) − ηI.
A priori, it is not evident what link there is between eigenvalues on a space
of solutions to an equation defined over all of space, and the behaviour of
solutions to an equation along a single line. The link is provided by a remarkable
formulation of the Dirac equation (see for instance Donaldson and Kronheimer
1990 or Cherkis and Kapustin 2001. The idea, roughly, is to write the Dirac
equation as the equations for the harmonic elements of the complex
⎛ ⎞
∇3 + i∇0 + z ⎠
D1 = ⎝
−∇1 − i∇2 ⊕2
D2 = ∇1 + i∇2 ∇3 + i∇0 + z
L (V ) −−−−−−−−−−−−−−−−−→ L (V ) −−−−−−−−−−−−−−−−−−−−−−−−→ L2 (V ).
2 2
(4.39)
Proposition 4.7 The operators D1 , D2 commute with multiplication by w =
x1 + ix2 , and for self-dual connections, D2 D1 = 0. The kernel Kz = ker(D2 ) ∩
ker(D1∗ ) of the Dirac operator Dz∗ is naturally isomorphic to the cohomology
ker(D2 )/Im(D1 ) of the complex.
On a compact manifold, this equivalence is part of standard elliptic theory.
In the non-compact case, the analysis must be done with care. The simplest
way, for us, is to adapt Nye and Singer (2000, section 4). Using the techniques
developed by Mazzeo and Melrose (1998), they show that the Dirac operator Dz∗
is Fredholm if and only if an ‘operator on fibres along the boundary’ P = (∇0 −
z)∞ + i(η1 e1 + η2 e2 + η3 e3 ) is invertible along all the circles S 1 in S 1 × S∞
2
=
∂(S × R ). Again, the e1 are the Pauli matrices, and the constraint must hold
1 3
for all choices of constants (η1 , η2 , η3 ). Nye and Singer show that it does as long
as z does not have values ±µ1 + nµ0 . In the elliptic complex, the constraint gets
replaced by the essentially equivalent condition that the complex
⎛ ⎞
η
⎝ 3
+ i(∇0 )∞ + z ⎠
−η1 − iη2 η1 + iη2 , η3 + i(∇0 )∞ + z
⊕2
L (V |S 1 ) −−−−−−−−−−−−−−−→ L (V |S 1 ) −−−−−−−−−−−−−−−−−−−−−−→ L2 (V |S 1 )
2 2
defined over the circle be exact. The equivalence between the Euler characteristic
of the complex and the index of the Dirac operator then goes through establishing
the isomorphism.
We now turn to analysing the way a solution to Nahm’s equations is extracted
from this complex. The operator A(0, z) on the kernel of Dzx is defined by
multiplication by x1 + ix2 followed by the projection on that kernel. On the
cohomology, it is simply multiplication by w = x1 + ix2 , simplifying matters
considerably.
Suppose η is an eigenvalue of A(0, z); hence there is a representative v of
the cohomology class such that (x1 + ix2 − η)(v) = D1 (s) for some s in L2 . In
particular, along w = η in S 1 × R3 , we have (∇3 + i∇0 − z)s = 0, and so (η, ζ) =
(w, 0) belongs to the spectral curve. Conversely, if (∇3 + i∇0 − z)s = 0 along
Closing the circle 57
Notice that the derivative (∇1 + i∇2 )(σ(w)) is supported on the annulus <
|w| < 2 .
Consider a ball B centred at η in the fibre above ζ = 0 of radius 2 chosen such
that B ∩ S1 = {η}. The ball parameterizes cylinders S 1 × {w} × R, for w ∈ B,
and thus a section φ ∈ H 0 (B, F ) is in fact a section of V on S 1 × B × R satisfying
(∇3 + i∇0 + z)φ = 0. The holomorphicity condition is then (∇1 + i∇2 )φ = 0.
Let φ0 ∈ H 0 (B, Fp,1
0
|(0,η) ). Then σ(w − η)ρ(x3 )φ0 lies in L2 (S 1 × R3 , V ). Thus
the section D1 (σ(w − η)ρ(x3 )φ0 ) is a coboundary for the complex (4.39). Sim-
∞
ilarly, if φ∞ lies in F−p,1 , then σ(w − η)(ρ(x3 ) − 1)φ∞ also lies in L2 , and its
image by D1 is also a coboundary.
Suppose for simplicity S1 ∩ {ζ = 0} contains only points of multiplicity 1.
Consider now a general section φ ∈ H 0 (B, F ). Away from the spectral curve,
φ decomposes into a sum φ0 + φ∞ , and combining the two constructions above
gives a coboundary K(φ) for the complex (4.39). This decomposition has a pole
0 ∞
at η if φ(η) represents a non-trivial element in the quotient F/(Fp,1 + F−p,1 ).
However, the section
K(φ) = D1 σ (w − η)ρ(x3 )φ − ρ∞
σ (w − η) ∇3
+ i∇0 + z ρ(x3 )φ − ρ∞
=
− ∇1 + i∇2 σ (w − η)ρ(x3 )φ − ρ∞
For intersections of higher multiplicity, the case presents only notational diffi-
culties, but is conceptually identical. The choice of ζ made above just simplified
the analysis. Varying it we obtain the following result:
For this complex to be defined, we must display a bit of care in our choice of
function spaces:
r L̃22 is the space of sections s of K that are L22 over each interval and such that,
(subscripts denoting the limits on the appropriates sides of the jump points)
π(slarge ) = ssmall , π(dz slarge ) = dz ssmall when j = 1 and slarge = ι(ssmall ),
dz slarge = ι(dz ssmall ) when j > 1.
r (L̃2 )⊕2 is the space of sections (s1 , s2 ) of K ⊕ K that are in L2 over each
1 1
interval, with π(slarge ) = ssmall at the jump points when j = 1 and slarge =
ι(ssmall ) when j > 1.
r L̃2 is the space of L2 sections of K.
Closing the circle 59
3
d2z v2 = 2 dz + (T0 − iµ0 x0 ) v + 22
(Tj − ixj )v2 0. (4.42)
j=1
This map must be injective (thus bijective) because of the convexity property of
solutions given by (4.42). If Rx−1 (0, u) = (f, g+ , g− ), the desired Green’s function
v can be chosen by taking f outside of (−µ1 , µ1 ), f + g− on (−µ1 , x), and f + g+
on (x, µ1 ).
60 The Nahm transform for calorons
The last inequality follows from the fact that the solution to Nahm’s equation is
irreducible. This L21 bound on v ensures continuity, with vL∞ CvL21 , and
hence the L∞ norm of the Green’s function is bounded by a constant times the
norm of u.
r (L̃2 )⊕2 is the space of sections (s1 , s2 ) of K ⊕ K that are L2 over the
1 1
intervals, but with a jump discontinuity ∆± (s1 , s2 ) = c± v± at ±µ1 .
r L̃2 is the space of L2 sections of K.
Proposition 4.14 For solutions to Nahm’s equations and for the complex
defined using those function spaces just defined, Proposition 4.9 holds, and
furthermore, at the jump points,
Proceeding as for j > 0, one can then build a Green’s function. Using
the Green’s function in both cases j > 0 and j = 0, we have the analog of
Proposition 4.7.
From the twistor point of view, the space of global sections of E along the
real line Lx in T correspond to the fibre of the caloron bundle V̂ over the
corresponding x ∈ S 1 × R3 . On the other hand, from the Nahm point of view,
64 The Nahm transform for calorons
facts tell us that the map from caloron to spectral data is a bijection, that all
three sets of data are equivalent, and that the six transforms between them are
pairwise inverses of each other.
Let us now leave the generic set. If we now take an arbitrary solution T
to Nahm’s equations, again satisfying all the conditions, we can fit it into a
continuous family T (t) with T (0) = T with T (t) generic for t = 0; that it can
be done follows from our description of moduli given in Section 4.6. The Nahm
transform of this family is a continuous family C(t) of calorons, which in turn
produces a continuous family of solutions T̃ (t) to Nahm’s equations. We do not
know a priori whether the boundary and symmetry conditions are satisfied at
t = 0. However, for t = 0, T̃ (t) = T (t), and so T̃ (0) = T (0), and we are done; the
transform Nahm to caloron to Nahm is the identity.
If the caloron C lies in the closure of the set of generic calorons (presumably
all calorons do, but it needs to be proved, perhaps using the methods of Taubes
(1984) to show that the moduli space is connected), we again fit it as C(0) into
a family C(t) of calorons, with C(t) generic for t = 0, all of same charge. The
Nahm transform produces a family T (t) of solutions to Nahm’s equations, with
T (t) satisfying all the conditions for t = 0. We also get a family S(t) of spectral
data, which for all t corresponds to both C(t) and T (t). We need to show that
T (0) satisfy all the conditions.
Taking a limit, it is fairly clear that the symmetry condition is satisfied also
for t = 0. The boundary conditions are less obvious. From the small side, there
is no problem, as the vanishing theorem (proposition 4.3) holds at C(0), and
so the solutions to Nahm’s equations are continuous at the boundary of the
interval. From the large side, what saves us is the rigidity of representations
into SU (2); indeed, the polar parts of the solutions are given by representations,
and so are fixed, in a suitable family of unitary gauges, hence preserved in a
limit. The process, given in Section 4.4.2, of passing to a constant gauge by
solving ds/dz + A+ (z)s = 0 applies in the limit near the singular points. We
have in the limit the same type of transformations relating the basis of K in
which one has the solution to Nahm’s equations to the continuous basis obtained
from lifting up and pushing down as to obtain the sequence (4.16), with the
same process of producing endomorphisms A(z, ζ)(t), giving us again in the
limit a continuous endomorphism of a bundle V over T P1 × S 1 , with V in fact
lifted from P1 × S 1 . The restrictions V (±µ1 ) (on the large side) are of fixed
type Vµ11 = Ok ⊕ Vj−1,−j+1 for all t. The summand Vj−1,−j+1 corresponds to
∞
the subsheaf O(j) of the sheaf F/(Fp,1 0
+ F−p,1 ) ⊗ L±µ1 , which exists in the
limit. The polar part C(p(ζ, t)) of (4.26), mapping Vj−1,−j+1 to itself, has limit
C(p(ζ, 0)). By (4.27), this latter limit is determined by the spectral curves, and
so is well defined. The other summand Ok is also well defined in the limit, as it
is the piece transferred from the small side, which is still Ok in the limit because
of the vanishing theorem. The off-diagonal vanishing as one goes back to the
trivialization for Nahm’s equations is simply a consequence of the polar behaviour
of solutions to ds/dz + A+ (z)s = 0, and of the normal form for A(z, ζ)(t) as in
66 The Nahm transform for calorons
(4.25). In short, the limit has exactly the same normal form, and so the same
boundary behaviour.
Finally, there is the question of irreducibility. The reducible solutions corre-
spond to calorons for which charge has bubbled off; as our family has constant
charge, this is precluded. The limit solutions satisfies all the conditions necessary
to produce a caloron by the opposite Nahm transform; as the transform is
involutive on the generic member of the family, it is also involutive in the limit,
and so the circle closes.
In short one has the following theorem:
Theorem 4.16
4.6 Moduli
The equivalence exhibited above allows us to classify calorons by classifying
appropriate solutions to Nahm’s equations. One has to guide us through the
example of monopoles, as classified in Donaldson (1984) and Hurtubise (1989);
the (framed) monopoles for gauge group G, with symmetry breaking to a torus
at infinity, and of charge k are classified by the space Ratk (P1 , GC /B) of based
degree k rational maps from the Riemann sphere into the flag manifold GC /B.
As our calorons are Kač–Moody monopoles, we should have the same theorem,
with framed calorons equivalent to rational maps from P1 into the loop group.
Following an idea developed in Atiyah (1984), one thinks of these as bundles on
P1 × P1 , with some extra data of a flag along a line, and some framing.
The rank 2 bundle corresponding to an SU (2) caloron, again following the
example of monopoles, should be the restriction of the bundle E on T̃ corre-
sponding to the caloron, to the inverse image of a point, say ζ = 0, in P1 . This
inverse image is P1 × C; along the divisor {0} × C, the bundle has the flag E0 ,
and along {∞} × C, the bundle has the flag E∞ . We extend this bundle to
infinity, using a framing, in such a way that E∞ is a trivial subbundle, and E0
has degree −j. This bundle will be corresponding to the caloron.
Moduli 67
Acknowledgement
This research was conducted while the first author was at McGill University
and was supported by an NSERC PDF. The second author is supported by
NSERC and FQRNT grants. The diagrams in this chapter were created using
Paul Taylor’s Commutative Diagram package.
References
Adams, M. R., Harnad, J., and Hurtubise, J. (1990). Isospectral Hamiltonian flows in
finite and infinite dimensions. II. Integration of flows. Commun. Math. Phys. 134(3),
555–85.
Atiyah, M. F. (1984). Instantons in two and four dimensions. Commun. Math.
Phys. 93(4), 437–51.
Atiyah, M. F., Drinfel d, V. G., Hitchin, N. J., and Manin, Yu. I. (1978). Construction
of instantons. Phys. Lett. A 65(3), 185–7.
Boyer, C. P., Mann, B. M., Hurtubise, J. C., and Milgram, R. J. (1994). The topology
of the space of rational maps into generalized flag manifolds. Acta Math. 173(1),
61–101.
Braam, P. J. and van Baal, P. (1989). Nahm’s transformation for instantons. Commun.
Math. Phys. 122(2), 267–80.
Bruckmann, F. and van Baal, P. (2002). Multi-caloron solutions. Nuclear Phys.
B 645(1–2), 105–33.
Bruckmann, F., Nógrádi, D., and van Baal, P. (2003). Constituent monopoles through
the eyes of fermion zero-modes. Nuclear Phys. B 666(1–2), 197–229.
Bruckmann, F., Nógrádi, D., and van Baal, P. (2004). Higher charge calorons with
non-trivial holonomy. Nuclear Phys. B 698(1–2), 233–54.
Chakrabarti, A. (1987). Periodic generalizations of static, self-dual SU(2) gauge fields.
Phys. Rev. D (3) 35(2), 696–706.
Charbonneau, B. (2004). Analytic aspects of periodic instantons. PhD thesis, Sep-
tember, Massachusetts Institute of Technology, Cambridge, MA. Available at
http://hdl.handle.net/1721.1/26746.
Charbonneau, B. (2006). From spatially periodic instantons to singular monopoles.
Commun. Anal. Geom. 14(1), 183–214.
Charbonneau, B. and Hurtubise, J. (2008).Calorons, Nahm’s equations on S 1 and
bundles over P1 × P1 . Commun. Math. Phys. 280(2), 315–49.
Cherkis, S. and Kapustin, A. (2001). Nahm transform for periodic monopoles and
N = 2 super Yang–Mills theory. Commun. Math. Phys. 218(2), 333–71.
References 69
Nógrádi, D. (2005). Multi-calorons and their moduli. PhD thesis, Institute Lorentz
for Theoretical Physics, University of Leiden, Leiden, the Netherlands. arXiv:hep-
th/0511125.
Norbury, P. (2000). Periodic instantons and the loop group. Commun. Math.
Phys. 212(3), 557–69.
Nye, T. M. W. (2001). The geometry of calorons. PhD thesis, University of Edinburgh,
Edinburgh, Scotland. arXiv:hep-th/0311215.
Nye, T. M. W. and Singer, M. A. (2000). An L2 -index theorem for Dirac operators on
S 1 × R3 . J. Funct. Anal. 177(1), 203–18.
Schenk, H. (1988). On a generalised Fourier transform of instantons over flat tori.
Commun. Math. Phys. 116(2), 177–83.
Szabó, S. (2007). Nahm transform for integrable connections on the Riemann sphere.
Mém. Soc. Math. Fr. (N.S.) No. 110.
Taubes, C. H. (1984). Path-connected Yang–Mills moduli spaces. J. Differential Geom.
19(2), 337–92.
Ward, R. S. (2004). Symmetric calorons. Phys. Lett. B 582(3–4), 203–10.
V
NAHM’S EQUATIONS AND FREE-BOUNDARY PROBLEMS
Simon K. Donaldson
Dedicated to Nigel Hitchin, with gratitude and affection
5.1 Introduction
In Donaldson (1984), following up work of Hitchin (1983), the author found it
useful to express Nahm’s equations, for a matrix group, in terms of the motion
of a particle in a Riemannian symmetric space, subject to a potential field. This
point of view lead readily to an elementary existence theorem for solutions of
Nahm’s equation, corresponding to particle paths with prescribed end points.
The original motivation for this chapter is the question of formulating an
analogous theory for the Nahm equations associated to the infinite-dimensional
Lie group of area-preserving diffeomorphisms of a surface – in the spirit of
Donaldson (1993). We will see that this can be done, and that a form of the
appropriate existence theorem holds – essentially a special case of a result of
Chen. However the main focus of the chapter is not on existence proofs but on
the various formulations of the problem, and connections between them. In these
developments, one finds that the natural context is rather more general than the
original question, so we will start out of a different tack, and return to Nahm’s
equations in Section 5.5.
Consider the following set-up in Euclidean space R3 , in which we take coordi-
nates (x1 , x2 , z) – thinking of z as the vertical direction. (We will use the notation
∂ ∂
∂xi = ∂i , ∂z = ∂z .) Suppose we have a strictly positive function H(x1 , x2 ). This
defines a domain
Explicitly
ρ = ∂z θ − (∂1 θ∂1 H + ∂2 θ∂2 H),
with the right-hand side evaluated at (x1 , x2 , H(x2 , x2 )). By the maximum
principle, ρ is a positive function, since the normal derivative of θ is positive
in the positive z direction on {z = H}. We consider the following free-boundary
problem: Given a positive periodic function ρ does it arise from some periodic
H, and is H unique?
One can gain some physical intuition for this question by supposing that the
lower half-space {z ≤ 0} represents a body with an infinite specific heat capacity
fixed at temperature 0 and ΩH corresponds to a layer of ice covering this body.
We choose units so that the melting temperature of the ice is 1. Sunlight shines
vertically downwards onto the upper surface {z = H} of the ice, but with a
variable intensity so that heat is transmitted to the surface according to the
density function ρ. We suppose that the surface of the ice is sprinkled by rain,
which will instantly freeze if the surface temperature of the ice is less than 1. We
also suppose that a wind blows across the surface, instantly removing any surface
water. Then we see that the solution to our free-boundary problem represents
a static physical state, in which the upper surface of the ice is just at freezing
point, the lower surface is at the imposed sub-freezing temperature, and the
heat generated by the given sunlight flows through the ice without changing the
temperature. Physical intuition suggests that there should indeed be a unique
solution.
We can express the free-boundary problem considered above as a special case
of another question. Suppose now that we have a pair of periodic functions
H0 , H1 on R2 with H0 < H1 . Then we have a domain ΩH0 ,H1 = {H0 (x1 , x2 ) <
z < H1 (x1 , x2 )}, with two boundary components. Let θ be the harmonic function
in this domain equal to 0, 1 on {z = H0 }, {z = H1 }, respectively. Then we obtain
a pair of flux functions ρ0 , ρ1 as before. By Gauss’ theorem, these satisfy a
constraint
ρ0 dx = ρ1 dx, (5.2)
[0,1]2 [0,1]2
since [0, 1]2 is a fundamental domain for the Z2 action. Obviously, if we replace
H0 , H1 by H0 + c, H1 + c for any constant c we get the same pair ρ0 , ρ1 . We ask:
given ρ0 , ρ1 satisfying the integral constraint (5.2), is there a corresponding pair
(H0 , H1 ), and if so is the solution unique up to the addition of a constant? A
positive answer to this question implies a positive answer to the previous one, by a
simple reflection argument. (Given ρ, as in the first problem, take ρ0 = ρ1 = ρ/2.
Then uniqueness implies that the solution has reflection symmetry about θ = 1/2
and we get a solution to the first problem by changing θ to 2θ − 1.)
Of course we can also imagine a physical problem corresponding to this second
question: for example, a layer of ice in the region ΩH0 ,H1 . We can now vary
the problem by supposing that in place of ice we have a horizontally stratified
Introduction 73
material in which heat can only flow in the horizontal directions. Thus the
steady-state condition, for a temperature distribution θ(x1 , x2 , z), is
2
∂1 + ∂22 θ = 0. (5.3)
and the integral constraint (5.2) still holds. So we ask: given ρ0 , ρ1 satisfying
(5.2), is there a pair H0 , H1 and a function θ on ΩH0 ,H1 , equal to 0, 1 on the
two boundary components, which has these fluxes, and is the solution essentially
unique? (In this case, one has to relax the condition on the domain to H0 ≤ H1 .)
It is natural to extend these questions to a general compact-oriented Rie-
mannian manifold X (which would be the flat torus R2 /Z2 in the discussion
above). Write dµ for the Riemannian volume form on X. We fix a real parameter
≥ 0 and define a map ∗ from T ∗ (X × R) to Λn T ∗ (X × R) by
d ∗ dθ = (∆ θ)dzdµ,
where
∆ θ = − ∂z2 + ∆X θ,
with ∆X the standard Laplace operator on X. (We use the sign convention that
∆X is a positive operator, so when = 1 our ∆ is the standard Laplace operator
on X × R.) If θ is defined on a domain ΩH0 ,H1 , as above, we define the flux ρi
on the boundary {z = Hi } by pulling back ∗ dθ just as before. We consider a
pair of functions ρ0 , ρ1 > 0 with
ρ0 dµ = ρ1 dµ = dµ
X X X
and we ask
Question 5.1 Is there a pair H0 ≤ H1 and a function θ on the set ΩH0 ,H1 ⊂
X × R with θ = 0, 1 on the hypersurfaces {z = H0 }, {z = H1 }, with fluxes ρi and
with ∆ θ = 0? If so, is the solution essentially unique?
For any > 0 the equation ∆ θ = 0 can be transformed into the standard
Laplace equation on the product, by rescaling the z variable. When = 0 the
equation has a very different character: it is not elliptic and we obviously do not
have automatic interior regularity with respect to z.
74 Nahm’s equations and free-boundary problems
|∇X φ̇|2
φ̈ = .
1 − ∆X φ
These equations define the geodesics in H. We can read off the Levi-Civita
connection of the metric from this geodesic equation, as follows. Let φ(t) be
any path in H and ψ(t) be another function on X × [0, 1], which we regard as
a vector field along the path φ(t). Then the covariant derivative of φ along the
path is given by
dφ
Dt φ = + (Wt , ∇X ψ), (5.5)
dt
where
−1
Wt = ∇X φ̇
1 − ∆X φ
and ( , ) is the Riemannian inner product on tangent vectors to X. (We write
∇X , or sometimes just ∇, for the gradient operator on X, so Wt is a vector
field on X.) This has an important consequence for the holonomy group of the
manifold H. Observe that the tangent space to H at a point φ is the space of
functions on X endowed with the standard L2 inner product associated to the
An infinite-dimensional Riemannian manifold 75
measure
dµφ = (1 − ∆φ)dµ0 .
So, in a general way, the parallel transport along a path from φ0 to φ1 should
be an isometry from L2 (X, dµφ0 ) to L2 (X, dµφ1 ). (Here we are ignoring the
distinction between, e.g., smooth functions and L2 functions.) What we see from
(5.5) is that this isometry is induced by a diffeomorphism f : X → X with
Lνα,β (dµφ ) = 0.
To identify this vector field we introduce some notation. For vector fields v, w on
X we write v × w for the exterior product: a section of the bundle Λ2 T X. We
define a differential operator
Λ2 T X ∼
= Λn−2 T ∗ X
(using the Riemannian volume form dµ), the exterior derivative
In the proof of Theorem 5.1 we will make use of two identities. For any pair
of vector fields v, w and function f
curl (v × w) = [v, w] + (div v)w − (div w)v (5.8)
Now we write
∂ ∂
Ds = + W s , Dt = + Wt ,
∂s ∂t
where the vector fields Ws , Wt are regarded as operators on the functions on X.
So Dt Ds − Ds Dt is the operator given by the vector field
∂Ws ∂Wt
ν= − − [Ws , Wt ],
∂t ∂s
and ν is exactly the vector field να,β we need to identify. Recall that
−∇∂s φ −∇∂t φ
Ws = , Wt = .
1 − ∆φ 1 − ∆φ
So
∂Ws −1 ∂2φ 1
= ∇ + ∆∂s φ∇∂t φ.
∂t 1 − ∆φ ∂s∂t (1 − ∆φ)2
Evaluating at s = t = 0 where ∂s φ = α, ∂t φ = β we have
∂Ws ∂Wt 1
− = (∆α∇β − ∆β∇α).
∂t ∂s (1 − ∆φ)2
Write g for the function (1 − ∆φ)−1 . Combining with the Lie bracket term we
obtain
να,β = [g∇α, g∇β] + g 2 (∆α∇β − ∆β∇α).
Since
div (g∇α) = g∆α + (∇g, ∇α), div (g∇β) = g∆β + (∇g, ∇β)
we see that
as required.
In the case when X has dimension 2—as discussed in Donaldson (1999),
Mabuchi (1987), and Semmes (1992)—the space H is formally a symmetric space.
This is not true in general, since the curvature tensor is not preserved by the
action of the group G.
78 Nahm’s equations and free-boundary problems
We define a functional on H by
V (φ) = φ dµ.
X
This function is convex along geodesics in H, since the geodesic equation implies
φ̈ ≥ 0. Now introduce a real parameter ≥ 0 and consider the functional on paths
in H:
1 2
E= |φ̇| + V (φ) dt, (5.10)
2 φ
corresponding to the motion of a particle in the potential − V . The Euler–
Lagrange equations are
|∇X φ̇|2 +
φ̈ = . (5.11)
1 − ∆X φ
∆ θ = 0.
r Φ equation. Here we are given φ0 , φ1 on X, with 1 − ∆φi > 0. We seek a
function Φ on X × [0, 1], equal to φ0 , φ1 on the two boundary components,
with 1 − ∆Φ > 0 for all t and satisfying the non-linear equation
∂2Φ ∂Φ 2
(1 − ∆X Φ) − |∇ | = . (5.13)
∂t2 ∂t
As we have explained in Section 5.2, this is the same as finding a path in the
space H, with end points φ0 , φ1 , corresponding to the motion of a particle
in the potential − V .
and this is
∂i θ
∂i θ
∆X ht = − ∂i − ∂z −
i
∂z θ ∂z θ
80 Nahm’s equations and free-boundary problems
which is
∂ 2θ ∂i θ∂i ∂z θ ∂i θ∂i θ∂z ∂z θ
− i
−2 + .
i
∂z θ (∂z θ)2 (∂z θ)3
On the other hand the flux ρt is given by pulling back the differential form ∗ dθ
on the product by the map x → (x, ht (x)) and this gives
1
ρt = ∂z θ + (∂i θ)2 .
∂z θ i
So
2
∂ρ 1 i (∂i θ)
= ∂z ∂z θ + .
∂t ∂z θ ∂z θ
This is
2
∂z ∂z θ ∂i θ∂i ∂z θ i (∂i θ) ∂z ∂z θ
∂t ρt = +2 i
− .
∂z θ (∂z θ)2 (∂z θ)3
So we see that ∂t ρt = −∆X ht , since ∂z ∂z θ = − i ∂i ∂i θ.
Now the normalization (5.13) implies that there is a function φ0 on X such
that ρ0 = 1 − ∆X φ0 . For t > 0 we define φt by
t
φt = φ0 + hτ dτ.
0
We can also regard this family of functions as a single function Φ on X × [0, 1],
Then (5.15) implies that ρt = 1 − ∆X φt for each t. We have
∂2Φ 1
= ∂t h =
∂t2 ∂z θ
and
1
1 − ∆X Φ = ∂z θ + (∂i θ)2 .
∂z θ i
So
2
∂2Φ ∂i θ
2
(1 − ∆X Φ) = + .
∂t ∂z θ
i
Now since
−1
∂i ∂t Φ = ∂i ht = ∂i θ
∂z θ
we can write the above as
2
∂2Φ ∂
(1 − ∆X Φ) = + ∇X Φ ,
∂t2 ∂t
as required.
Three equivalent problems 81
the Laplace operator on X with i∂∂. We take the product with a circle, with
coordinate α, and let τ = t + iα be a complex coordinate on the Riemannn
surface S 1 × (0, 1). Then, in differential form notation, our non-linear operator
is given by
n
n
Q(A) = A00 Aii − A2i0 .
i=1 i=1
Lemma 5.1
To see the first item, observe that we can change basis in Rn ⊂ Rn+1 to reduce
to the case when the block Aij , 1 ≤ i, j ≤ n is diagonal, say, with entries bi . Then
if A ≥ 0 we have A00 bi ≥ A20i and so
Q(A) = A00 bi − A20i ≥ 0,
For the second item, we just have to observe that Q is induced from a quadratic
form of Lorentzian signature on Rn+2 by the linear map
n
π : A → (A00 , Aii , A0i ).
i=1
The hypotheses imply that π(A) and π(B) are in the same component of a hyper-
boloid defined by this Lorentzian form and the statements follow immediately
from elementary geometry of Lorentz space.
Using this lemma we can deduce the uniqueness of the solution to our Dirichlet
problem, in any dimension.
Proposition 5.1 If φ0 , φ1 ∈ H then there is at most one solution Φ of the
equation Q(Φ) = on X × [0, 1] with 1 − ∆X Φ > 0 for all t and with Φ(x, 0) =
φ0 (x), Φ(x, 1) = φ1 (x).
We show that the functional E(Φ) given by (5.10) is convex with respect to the
obvious linear structure. Thus we consider a one-parameter family Φs = Φ + sψ,
with the fixed end points. We have
1
d
E(Φs ) = 2Φ̇s ψ̇(1 − ∆X Φs ) − Φ̇2 ∆X ψ.
ds 0 X
introduced in Section 5.3 all appear in this literature. The transformation from
θ to Φ taking the harmonic function θ as a new independent variable is called in
Crank (1984, chapter 5) the “isothermal migration method.” The transformation
from the formulation in terms of θ to that in terms of U is known as the Baiocchi
transformation (Elliot and Ockendon 1982; Baiocchi and Capelo 1984; Crank
1984). The transformation of the free-boundary problem for a linear equation to
a non-linear Dirichlet problem is used in Kinderlehrer and Nirenberg (1979) to
derive fundamental regularity results.
An important feature of the U -formulation is that it admits a variational
description. Recall that we are given a function L = max(φ0 − φ1 + z, 0) on X ×
R and we seek a C 1 function U with U ≥ L satisfying the equation ∆ U = ρ0 on
the set where U > L. This can be formulated as follows. We fix a large positive
M and consider the functional
1
EM (U ) = |∇X U |2 + |∂z U |2 − ρ0 U dµdz
2
over the space of functions satisfying the constraint U ≥ L, where the integral
is taken over X × [−M, M ] in X × R (which, a posteriori, should contain the
set ΩH0 ,H1 on which U > L). Then the solution minimizes EM over all functions
U ≥ L. This can be used to give another proof of the uniqueness of the solution
to our problem. It seems likely that it could also be made the basis of an existence
proof, following standard techniques in the free-boundary literature. Now recall
that our Φ-formulation was based on a variational principle, with Lagrangian
(5.10). To relate the two, we consider any function Φ on X × [0, 1] with ∂t ∂t Φ ≥ 0
and define U by the recipe of Section 5.3. We suppose that −M < ∂t Φ(x, 0) and
∂t Φ(x, 1) < M for all x ∈ X. Then we have
1
∆X Φ(∂t Φ)2 dµ dt
0 X
are equal modulo boundary terms. We leave the full calculation as an exercise
for the reader.
86 Nahm’s equations and free-boundary problems
so T2 + iT3 moves in a single adjoint orbit in the Lie algebra of GL(n, C).
Conversely if we fix some B in this complex Lie algebra, introduce a function
g(t) taking values in GL(n, C) and define skew-Hermitian matrices Ti (t) by
dg −1
T0 + iT1 = g ,
ds
T2 + iT3 = gBg −1 ,
then two of the three Nahm equations are satisfied identically. The remaining
equation can be expressed in terms of the function h(t) = g ∗ (t)g(t), taking values
in the space H of positive definite Hermitian matrices, which we can also regard
as the quotient space GL(n, C)/U (n). This equation for h(t) is a second-order
ODE which is the Euler–Langrange equation for the Lagrangian
dh
E(h) = | |2H + VB (h)dt.
dt
Here | |H denotes the standard Riemannian metric on H. The function V on H
is
VB (h) = Tr(hBh−1 B ∗ ).
VB (h) = |gBg−1 |2 ,
Now take the compact Riemann surface Σ to be a two-torus, and identify the
space H with periodic Kähler potentials on the universal cover C. On this cover
the identity function β is holomorphic, and we see from the above that the formal
expression
Vβ = φ
C
which will generate the same equations of motion. So we see that, modulo some
blurring of the distinction between Σ and its universal cover, the functional we
have been considering is indeed analogous to that in the finite-dimensional case.
Using the transformation from the Φ equation to the θ equation, we obtain
a relation between Nahm’s equations for the Hamiltonian diffeomorphisms of a
surface and harmonic functions on R3 . This can be seen in other ways. Most
directly, we consider three one-parameter families of functions hi (t) on a surface
Σ with an area form which satisfy
dhi
= {hj , hk }, (5.22)
dt
Relation with Nahm’s equations 89
Ω2 = dτ dβdτ dβ.
where the integral is taken over the intersection of the vertical line through
(x1 , x2 , 0) with Ω. Our hypotheses imply that F ≥ 0, and F is supported on the
larger disc D1 .
The question we are lead to is the following:
Question 5.3 Suppose that D0 ⊂ D1 are topological discs in C, that ρi are
two-forms supported on Di , and that there is a non-negative function F on C,
supported on D1 , with ρ0 − ρ1 = ∆C F (where the Laplacian is defined in the
distributional sense). Do ρ0 , ρ1 arise from a unique harmonic function θ on a
domain in R3 , by the construction above?
(For simplicity we have not specified precisely what singularities should be
allowed in the forms ρi : this specification should be a part of the question.)
Hitchin showed in Hitchin (1983) that Nahm’s equations form an integrable
system. The root of this is the invariance of the conjugacy class given by
(5.21), together with the family of similar statements that arise from the SO(3)
action on the set-up. In this vein, we can write down infinitely many conserved
quantities for the solutions of (5.11) on the Riemannian manifold H. Let fλ be
an eigenfunction of the Laplacian ∆X , with eigenvalue λ > 0. Then we have
Proposition 5.3 For any > 0, if φt satisfies (5.11) then the quantity
"
λ
exp φ̇ fλ (1 − ∆φ) dµ,
X
satisfies ∆ Kλ = 0.
Acknowledgements
The author is grateful to Professors Colin Atkinson, Xiuxiong Chen, Darryl
Holm, and John Ockendon for helpful discussions.
References
Baiocchi, C. and Capelo, A. (1984). Variational and Quasivariational Inequalities.
Wiley, New York.
Chen, X.-X.(2000). The space of Kahler metrics. Journal of Differential Geometry 56,
189–234.
Crank, J. (1984). Free and Moving Boundary Problems. Oxford University Press,
Oxford.
References 91
S. Ramanan
Dedicated to Nigel Hitchin
This is an exposition of certain aspects of the theory of Higgs pairs (also called
‘Higgs bundles’ by some authors), naturally with some emphasis on topics in
which I played a part. This is an area to which fundamental contributions have
been made by Nigel Hitchin, and I take pleasure in presenting this summary on
the occasion of his sixtieth birthday.
It turns out that from the moduli point of view, it is better to think of the
moduli space as the set of S-equivalence classes of semi-stable bundles, rather
than as isomorphism classes of polystable bundles.
94 Some aspects of the theory of Higgs pairs
stable, which is generically the case, the kernel is also stable and therefore defines
a point of SU (n, d − 1). Incidentally, the kernels corresponding to two linear
forms are isomorphic if they are scalar multiples of each other. Hence the kernel
is determined by an element of P (Ea∗ ).
Two points E, E of SU (n, d) are Hecke modifications of each other if there
exist elements l ∈ P (Ea∗ ) and l ∈ P (Ea∗ ) such that ker(l) is isomorphic to ker(l ).
Thus we have a closed subvariety HC of SU (n, d) × SU (n, d) which is invariant
under the transposition of coordinates.
For a generic E ∈ SU (n, d + 1) with determinant ξ ⊗ O(a) we can thus see that
P (Ea∗ ) is contained in SU (n, d) with determinant ξ. These projective spaces are
called Hecke cycles. In particular, if we consider a projective line in P (Ea∗ ) it
gives a projective line in SU (n, d) which we call a Hecke curve. Thus SU (n, d)
contains a lot of rational curves. This is not surprising, since it is easily seen that
SU (n, d) are unirational, and indeed (not so easily) even rational if n and d are
coprime (King and Schofield 1999).
Example 6.1 Let ξ be a ‘theta characteristic’, namely, a line bundle such that
ξ 2 is isomorphic to K. In particular the degree of ξ is g − 1. Take for E the vector
bundle ξ ⊕ ξ −1 . It is of rank 2 and degree 0. Since it has ξ as a subbundle, it
is not semi-stable. Now a homomorphism
α : E → E ⊗ K can be written in the
α 1 α2
form of a (2 × 2) matrix where α1 and α4 are sections of K, α2 is a
α3 α4
section of K 2 , and α3 is a section of O. If α3 is nonzero, then ξ is not invariant
under α. Any line subbundle of E other than ξ admits a nonzero homomorphism
into ξ −1 and hence has degree at most −(g − 1). Hence for any such α, the pair
(E, α) is stable.
Remark 6.2 On the other hand, if E is stable, then any α (including 0) gives
rise to a Higgs pair (E, α).
Remark 6.4 Note that when the representation is unitary, the flat metric
gives, on local trivialization on a domain, a constant map, which is of course
harmonic.
Here we interpret mE as the m-fold direct sum of E and assume that the i’s
and j’s are disjoint. He showed these form a Tannaka category N and that its
Tannaka dual is P f (π).
Tannaka category Hg, one easily concludes that there is a natural action of C∗
on T n(π) as well.
The group of homomorphisms of π into C is isomorphic to H 1 (C, C) and the
Hodge decomposition can be described as the action of C∗ on it, acting by the
natural (identity) character on H 1,0 and by its inverse (namely, the character
λ → λ−1 on H 0,1 ).
From this point of view, the above action of C∗ on T n(π), which mimics the
Hodge action, can be thought of as the non-abelian version of the Hodge action.
It should be pointed out that this action depends on the complex structure of
the curve.
Actually, the considerations above can be generalized in two directions. Firstly,
we can take a reductive group G and consider principal G-bundles. Stability of a
principal G-bundle E will then mean that for any reduction of its structure
group G to a parabolic subgroup P (which can be interpreted to mean a
section s of the bundle E/P ), the pull-back of the tangent bundle along the
fibres is of positive degree. As for a Higgs pair, it now consists of a principal
G-bundle E and a differential form with values in the associated adjoint bundle
ad(E), namely the vector bundle associated to E via the adjoint representation
of G in its Lie algebra. We will discuss this generalization in some detail
below.
Secondly, we may take, instead of C, an arbitrary smooth, polarized projective
variety X of dimension n. This essentially means that we are given an ample line
bundle ξ on X. Then one can define the degree of a line bundle L to be the
intersection number c1 (L).ξ n−1 , and proceed as before. A Higgs field is, in the
general case, a principal G-bundle E, and a one-form α with values in ad(E)
such that the two-form [α, α] vanishes. Thus we can get an algebraic description
of the pro-reductive algebraic completion of the fundamental group, but more
importantly, an action of C∗ as well.
This would show that many (non-uniform) discrete subgroups of real semi-
simple groups cannot be the fundamental group (nor even split quotients of the
fundamental group) of smooth projective varieties. This is because one can check,
thanks to rigidity theorems on uniform discrete subgroups of Lie groups, that
these do not admit an action of C∗ on the above lines. For example, SL(n, Z),
n ≥ 3 are not.
End(E) → End(E) ⊗ K,
6.7 Quantization
If D is a (linear) differential operator of order ≤ k, then its (kth order) symbol is
a function on the cotangent bundle which is a homogeneous polynomial of degree
k on the fibres. In the above context therefore one may ask if one can associate
in a canonical fashion a basis of commuting differential operators whose symbols
are the Poisson commuting functions defined above.
Some motivation or analogy may be in order here. The Hitchin space is the
analogue of (actually the dual of) the space of invariant polynomials for the
conjugacy action of GL(n) on the vector space of (n, n) matrices. The Hitchin
map itself can be thought of as the analogue of associating to any matrix A, the
linear form on invariant polynomials which maps any f to the evaluated complex
number f (A). One can associate canonically to any homogeneous invariant
polynomial f , a bi-invariant differential operator Df on the Lie group, with
102 Some aspects of the theory of Higgs pairs
first section with the second section after translating by the divisor class a2 − a1 .
This is the compactified Jacobian of C. This can be identified with the Hitchin
fibre.
Let E be the direct image of L on C. Its fibre at a is of the form V ⊕ La1 ⊕ La2 ,
where V is a vector space of dimension n − 2. It comes with a one-dimensional
subspace, namely, the graph of the isomorphism η. The corresponding Hecke
transform is the bundle E, obtained as the direct image on C of the line bundle
on S given by (L, η).
In other words, the Hitchin fibre corresponding to the spectral curve S is the
compactified Jacobian J˜ described above. The explicit Higgs pair given rise to
by a point of J˜ is also given above. If we deal with SU (n, d) instead of U (n, d),
one has to replace the Jacobian by the Prym variety. The point is of course that
we get on this fibre, projective lines parametrized by the Jacobian, or the Prym
variety of S . These lines map to Hecke curves on SU (n, d).
Hwang and I (2004) considered the tangents to Hecke curves passing through
the generic point E of SU (n, d) and studied their lifts to the cotangent bundle.
The unions of all these are, as a simple consequence of our remark above, dense in
the Hitchin discriminant. The intersection of this discriminant with the cotangent
space at E is a cone and defines a projective subvariety of P (TE∗ ). We showed
from the above considerations that the dual of this variety is the variety of all
the tangents to Hecke curves through E.
This implies that from SU (n, d) as a variety one can geometrically extract the
Hitchin discriminant and hence the variety of Hecke tangents. The latter has as
its Albanese image, the curve C, so that we have a geometric description of the
curve – a Torelli kind of result.
Indeed, one can work a little more and show the following (Hwang and
Ramanan 2004):
Theorem 6.5 If C and C are two curves, then there exists a nonconstant
morphism from SUC (n, d) into SUC (n, d) if and only if C is isomorphic to C .
Moreover any isomorphism is actually induced by an isomorphism of C with C ,
tensoring by line bundles of order n and possibly dualizing.
ad(E) → ad(E) ⊗ K
Reductive groups and Higgs pairs 107
where the map is given by Lie bracketing with α. Using this, one can identify
the symplectic form and show that the fibres of the Hitchin morphism are
Lagrangian. This would prove the complete integrability in general (Biswas and
Ramanan 1994).
Nitsure showed in Nitsure (1991) that the Hitchin map is proper. One can seek
therefore to identify the abelian varieties which are generic fibres of the Hitchin
morphism. This has been done in a beautiful paper of Donagi (1993).
The Hitchin map has also been quantized in the general case (Frenkel 2004;
Beilinson and Drinfeld, to appear). I should point out here that the moduli space
may be gainfully replaced by the moduli ‘stack’ and it is in this set-up that the
subject is dealt with in the above references. Since the moduli stack is smooth,
many of the circumlocutions in the statements can thereby be avoided. Moreover,
in the context of the Langlands programme, the moduli stack is more natural,
but I prefer to leave things as they are, rather than launch into the theory of
algebraic stacks here.
The algebra D of differential operators on the moduli space M (G) from the
square root of the canonical bundle to itself turns out to be commutative and
their symbols give the space of functions on Ht(G). In other words, the algebra
of their symbols has as its spectrum the space Ht(G).
One may then ask for the spectrum of D itself. What is it natually isomorphic
to? This turns out to be an affine space, called the space of opers.
Here there is another subtlety. There is a canonical isomorphism (or rather
duality) between the algebra of adjoint invariant polynomials on the Lie algebra
of a semi-simple group G and those of its Langlands dual L G. The Langlands
dual is a group associated to the root system, which is dual to that of G. If G
is simply connected, L G is the adjoint group with the dual root system. Also
the centre of G and the fundamental group of L G are canonically dual to each
other. In general, the Langlands dual of G/A, where A is a central subgroup of
G, is defined to be the covering group of L G corresponding to the quotient A∗
of π1 (L G).
For simplicity, let us assume here that G is simply connected. We have
remarked that the algebra of adjoint invariant polynomials on g and L g are
isomorphic to their respective Weyl group invariant polynomials on their Cartan
algebras. The Cartan algebras are canonically dual, compatible with the Weyl
group actions.
A G-oper consists of a principal G-bundle E provided with a connection ∇, and
a reduction F of E to a Borel subgroup B of G. The connection is not supposed
to be compatible with the reduction. On the contrary, one can associate to the
connection its second fundamental form ψ which measures the extent to which
∇ does not come from an F -connection. This is defined for example, locally, by
taking a connection ∇loc on F and taking the difference ψ = ∇ − ∇loc . It is a
one-form with values in ad(E). Using the natural map g → g/b, we get a section
ϕ of K ⊗ Fρ , where Fρ is the vector bundle associated to F via the representation
of B in g/b. This form is obviously independent of ∇loc and therefore glues into
108 Some aspects of the theory of Higgs pairs
0 → ξ → P → ξ −1 → 0
given by the generator of H 1 (C, K). The bundle P can be considered an
SL(2, C) bundle with a reduction to a Borel subgroup B0 . It is clear that it
is indecomposable of degree 0. Hence by Weil’s theorem, it admits a connection.
Although P depends on the choice of the theta characteristic, the corresponding
P SL(2)-bundle is entirely canonical.
In general, that is to say, for any semi-simple Lie group G, Kostant considered
all non-trivial homomorphisms of SL(2) into it, up to conjugacy. He isolated one
such conjugacy class as particularly interesting, calling it the principal TDS. TDS
stands for three-dimensional simple Lie group. If we fix one such homomorphism
κ, we get the associated principal G-bundle κ(P ). It comes with a reduction to
a Borel subgroup B of G. Thus, we have a principal G-bundle, a B-reduction
which admits a G-connection. Its Harder–Narasimhan filtration is essentially the
reduction to B.
The set of (equivalence classes of) opers is an affine space with the torsor
vector space Ht(G). The conclusion is that the sought for spectrum of D is
the set of opers for the Langlands dual group. There does not yet seem to be
a simple understanding of this nice relationship. For further details one may
consult Beilinson and Drinfeld (to appear).
These are acted upon via the character λ → λ−2 on Cf and as characters λ →
λ2mi −2 on these one-dimensional subspaces in L.
We considered the SL(2, C)-bundle E = ξ −1 ⊕ ξ and determined all the Higgs
pairs with that as the underlying bundle. Take the Kostant homomorphism of
SL(2, C) in G and consider the associated bundle EG . We will now consider Higgs
fields on the bundle EG . Now ad(EG ) can be identified with the bundle associated
to E via the restriction of the adjoint representation of G to SL(2). Since the
bundle E comes from the line bundle ξ by the inclusion of C∗ as diagonal matrices
in SL(2), the Kostant subbundle also makes sense in ad(EG ). Clearly this bundle
is isomorphic to the direct sum ξ −2 ⊕i=l
i=1 (ξ)
2mi −2
G ) ⊗ K contains
. Thus, ad(E
the trivial bundle, and Higgs fields on EG of the form α = 1 + βi , βi ∈ Γ(K mi )
therefore make sense. Then one can show that the set of all Higgs pairs (EG , α)
with α, as above, are mapped by the Hitchin morphism isomorphically on the
Hitchin space. This is ‘the’ generalized Hitchin component. It of course depends
on the choice of ξ.
Note that there is a canonical conjugacy class of real forms Gr of G, namely,
the Chevalley or split form. Reductive representations of π in Gr give a closed
subspace of Rep(π, G). As we have seen this space is not connected. In some cases,
such as SL(2, R), there is a topological invariant (called the Toledo invariant)
which takes different values on different components. The invariants which serve
to nail down all components is, to my knowledge, not fully understood yet.
One of the components in this case is the Hitchin component we have described
above.
We have thus some understanding of the Higgs pairs corresponding to the
compact and the split forms. It is natural to consider the other real forms as
well. A real form is defined as the fixed point set of a complex anti-holomorphic
involution of G, but the fixed points in Hg(G) constitute a complex subvariety.
The relation between Higgs pairs and representations can then be understood
in the following way. Let σ be a complex conjugation defining a real form Gσ of
G. Then there is a suitable compact real form H of G which is invariant under
σ. Now extend σ|H as a complex analytic involution of the complexification
H c of H in G. Let m be the intersection of h and gσ . Then there is a Cartan
decomposition of the vector space gσ as m ⊕ p. Thus we have g = hc = mc ⊕ pc .
Then we have
It is easy to see that the diagonal automorphism with i and −i on the diagonal
takes α to −α, so that (E, α) and (E, −α) are indeed equivalent.
The involution (E, α) → (E, −α) obviously fixes the Higgs pairs given by
representations into SU (n) for all n, since α is 0 for these pairs. On the other
hand, it is easily seen that points in the Hitchin component are not fixed by this
involution if n ≥ 3. Why is this so?
A little reflection immediately gives us the clue. The two anti-holomorphic
involutions of SL(n, C) given by A → A and A → (A )−1 give rise to the two real
forms SL(n, R) and SU (n, C) as fixed points, respectively. Consider two complex
conjugations as equivalent if they differ by an inner automorphism. Then the
maps induced on the set of equivalence classes of representations coming from
either of these two real forms are clearly the same. In the case of SL(2), the two
anti-holomorphic involutions are indeed equivalent, but in the case of SL(n)’s
for n ≥ 3, they are not!
One may also consider other involutions and determine their fixed points, and
this gives rise to a nice interplay between the geometry of etale coverings of C,
the equivalence among real forms, etc. These are being studied in Garcia-Prada
and Ramanan (in preparation).
Acknowledgement
I would like finally to thank the referee for reading the typescript carefully,
pointing out many typographical errors, and catching at least one misleading
statement.
References
Beauville, A., Narasimhan, M. S. and Ramanan, S. (1989). Spectral curves and the
generalized theta divisor. J. Reine Angew. Math. 398, 801–28.
Beilinson, A. and Drinfeld, V. (2005). Opers—Quantization of Hitchin integrable
systems and Hecke eigensheaves. math. AG/0501398.
Biswas, I. and Ramanan, S. (1994). An infinitesimal study of the moduli of Hitchin
pairs. J. Lond. Math. Soc. 49, 219–31.
Bradlow, S. B., Garcia-Prada, O. and Gothen, P. B. (2003). Surface group representa-
tions and U (p, q)-Higgs bundles. J. Diff. Geom. 64, 111–70.
References 111
Corlette, K. (1988). Flat G-bundles with canonical metrics. J. Diff. Geom. 28, 361–82.
Donagi, R. (1993). Decompositions of spectral covers. Astérisque 218, 145–75.
Donaldson, S. K. (1987). Twisted harmonic maps and self-duality equations. Proc.
Lond. Math. Soc. 55, 127–31.
Donaldson, S. K. (1993). A new proof of a theorem of Narasimhan and Seshadri. J.
Diff. Geom. 18, 269–77.
Drezet M. and Narasimhan, M. S. (1989). Groupe de Picard des variétés de modules
de fibré semi-stable sur les courbes algèbriques. Inv. Math. 97, 53–94.
Eells, J. and Sampson, J. H. (1964). Harmonic mappings of Riemannian manifolds.
Am. J. Math. 86, 109–60.
Frenkel, E. (2004). Recent advances in the Langlands program. Bull. Am. Math. Soc.
41, 151–84.
Garcia-Prada, O. and Ramanan, S. Involutions on Higgs moduli space on curves, in
preparation.
Goldman, W. M. (1988). Topological components of spaces of representations. Inv.
Math. 93, 557–607.
Hitchin, N. J. (1987). The self-duality equations on a Riemann surface. Proc. Lond.
Math. Soc. 55, 59–126.
Hitchin, N. J. (1992). Lie groups and Teichmüller space. Topolgy 31, 449–73.
Hwang, J.-M. and Ramanan, S. (2004). Hecke curves and Hitchin discriminant. Ann.
Sci. Ecole Norm. Sup. 37, 801–17.
King, A. and Schofield, A. (1999). Rationality of moduli of vector bundles on curves.
Indagationes Mathematicae 10, 519–35.
Kostant, B. (1959). The principal three-dimensional subgroup and the Betti numbers
of a complex simple group. Am. J. Math. 81, 973–1032.
Kostant, B. (1963). Lie group representations on polynomial rings. Am. J. Math. 82,
327–404.
Milnor, J. W. (1957). On the existence of a connection with curvature zero. Commun.
Math. Helv. 32, 216–23.
Mumford, D. (1962). Projective invariants of projective structures and applications.
Proceedings of the International Congress on Mathematics, Stockholm, Sweden,
pp. 526–30.
Narasimhan, M. S. and Ramanan, S. (1975). Deformations of the moduli space of vector
bundles over an algebraic curve. Ann. Math. 101, 391–417.
Narasimhan, M. S. and Seshadri, (1965). C. S. Stable and unitary vector bundles on
compact Riemann surfaces. Ann. Math. 82, 540–64.
Nitsure, N. (1991). Moduli spaces of semistable pairs on a curve. Proc. Lond. Math.
Soc. 62, 275–300.
Nori, M. V. (1976). On the representations of the fundamental group. Compositio
Mathematica 33, 29–41.
Ramanan, S. (1973). The moduli space of vector bundles over an algebraic curve. Math.
Ann. 200, 69–84.
Ramanathan, A. (1975). Stable and principal bundles on a compact Riemann surface.
Math. Ann. 213, 129–52.
Ramanathan, A. (1996). Moduli for principal bundles over algebraic curves. Proc.
Indian Acad. Sci. 106, 301–28.
Seshadri, C. S. (1967). Space of unitary vector bundles on a compact Riemann surface.
Ann. Math. 85, 303–36.
112 Some aspects of the theory of Higgs pairs
Simpson, C. T. (1992). Higgs bundles and local systems. Inst. Hautes Études Sci. Publ.
Mat. 75, 5–95.
Toledo, D. (1989). Representations of surface groups in complex hyperbolic space. J.
Diff. Geom. 29, 125–33.
Uhlenbeck, K. and Yau, S. T. (1986). On the existence of Hermitian-Yang-Mills
connections in stable vector bundles. Commun. Pure Appl. Math. 39, 5257–93.
Weil, A. (1938). Généralisation de fonctions abéliennes. J. Math. Pure Appl. 17, 47–87.
Wood, J. W. (1976). Bundles with totally disconnected structure group. Commun.
Math. Helv. 51, 183–99.
VII
MIRROR SYMMETRY, HITCHIN’S EQUATIONS,
AND LANGLANDS DUALITY
Edward Witten
Dedicated to Nigel Hitchin on the occasion of his 60th birthday
1 Actually, it is best to define Y(G, C) as a geometric invariant theory quotient that para-
metrizes stable homomorphisms plus equivalence classes of semi-stable ones. This refinement
will not concern us here (see Section 7.6.1).
2 The definition of this intersection pairing depends on the choice of an invariant quadratic
form on the Lie algebra of G. It can be shown using Hitchin’s C∗ action on the moduli space
of Higgs bundles that the A-model that we define shortly is independent of this choice, up to
a natural isomorphism. The geometric Langlands duality that one ultimately defines likewise
does not depend on this choice.
3 The usual definition of Ω is such that Im Ω is cohomologically trivial, while Re Ω is not.
The fact that ω = Im Ω is cohomologically trivial is a partial explanation of the fact, mentioned
in the last footnote, that the A-model of Y is invariant under scaling of ω.
114 Mirror symmetry, Hitchin’s equations, and Langlands duality
These are the topological field theories that are relevant to the most basic
form of geometric Langlands duality. However, there is also a generalization that
is relevant to what is sometimes called quantum geometric Langlands. From the
A-model side, it is obvious that a generalization is possible, since we could use a
more general linear combination of Re Ω and Im Ω in defining the A-model. What
is less evident is that the B-model can actually be deformed, as a topological
field theory, into this family of A-models. This rather surprising fact is natural
from the point of view of generalized complex geometry (see Hitchin 2003) and
has been explained from that point of view in section 4.6 of Gualtieri (2003), as a
general statement about complex symplectic manifolds. In sections 5.2 and 11.3
of Kapustin and Witten (2007), it was shown that quantum geometric Langlands
is naturally understood in precisely this setting.
Here, however, to keep things simple, we will focus on the most basic B-model
and A-model that were just described.
Remark 7.1 As an aside, one may ask how closely related φ, known in the
present context as the Higgs field, is to the Higgs fields of particle physics. Thus,
to what extent is the terminology that was introduced in Hitchin (1987a) actually
justified? The main difference is that Higgs fields in particle physics are scalar
fields, while φ is a one-form on C (valued in each case in some representation
of the gauge group). However, although Hitchin’s equations were first written
down and studied directly, they can be obtained from N = 4 supersymmetric
gauge theory via a sort of twisting procedure (similar to the procedure that
leads from N = 2 supersymmetric gauge theory to Donaldson theory). In this
twisting procedure, some of the Higgs-like scalar fields of N = 4 super Yang–
Mills theory are indeed converted into the Higgs field that enters in Hitchin’s
equations. This gives a reasonable justification for the terminology.
116 Mirror symmetry, Hitchin’s equations, and Langlands duality
of solutions of Hitchin’s equation for the dual group L G. It turns out that the
bases of the Hitchin fibrations for G and L G can be identified in a natural way.
The resulting picture is something like this:
MH (L G, C) MH (G, C)
B
7.4 Ramification
Before getting back to stacks, however, I want to give an idea of what is called
“ramification” in the context of geometric Langlands.
A simple generalization of what we have said so far is to consider flat bundles
not on a closed oriented two-manifold C but on a punctured two-manifold C =
C\p; that is, C is C with a point p omitted.
We pick a conjugacy class C ⊂ GC , and we let Y(G, C ; C) denote the moduli
space of homomorphisms ρ : π1 (C ) → GC , up to conjugation, such that the
monodromy around p is in the conjugacy class C.
Many statements that we made before have natural analogs in this punctured
case. In particular, Y(G, C ; C) has a natural structure of a complex symplectic
manifold. It has a natural complex structure and holomorphic symplectic form
Ω. Just as in the unpunctured case, we can define a B-model of Y(G, C ; C).
Also, viewing Y(G, C ; C) as a real symplectic manifold with symplectic form
ω = Im Ω, we can define an A-model. The B-model and the A-model are both
completely independent of the complex structure of C .
Next, introduce the dual group L G and let L C denote a conjugacy class in
its complexification. Again, the space Y(L G, C ; L C) has a natural B-model and
A-model.
Based on what we have said so far, one may wonder if, for some map
between C and L C, there might be a mirror symmetry between Y(G, C ; C)
and Y(L G, C ; L C). The answer to this question is “not quite,” for a number
of reasons. One problem is that there is no natural correspondence between
conjugacy classes in GC and in L GC . A more fundamental problem is that the
B-model of Y(G, C ; C) varies holomorphically with the conjugacy class C, but
the A-model of the same space does not. To find a version of the statement
Ramification 119
that has a chance of being right, we have to add additional parameters to find a
mirror-symmetric set.
In any event, regardless of what parameters one adds, it is very difficult to
answer the question about mirror symmetry if C is viewed simply as an oriented
two-manifold with a puncture. To make progress, just as in the unramified case
(i.e. the case without punctures), it is very helpful to endow C with a complex
structure and to use Hitchin’s equations. This actually also helps us in finding the
right parameters, because an improved set of parameters appears just in trying
to give a natural formulation of Hitchin’s equations on a punctured surface. Let
z be a local parameter near the puncture and write z = reiθ . In the punctured
case, it is natural (see Simpson 1990) to introduce variables α, β, γ taking values
in the Lie algebra t of a maximal torus T ⊂ G, and consider solutions of Hitchin’s
equations on C whose behavior near z = 0 is as follows:
A = α dθ + . . . (7.2)
dr
φ=β − γ dθ + . . . . (7.3)
r
The ellipses refer to terms that are less singular near z = 0.
All the usual statements about Hitchin’s equations have close analogs in this
situation. The moduli space of solutions of Hitchin’s equations with this sort
of singularity is a hyper-Kähler manifold MH (G, C; α, β, γ). In one complex
structure, usually called J, it coincides with Y(G, C; C), where C is the conju-
gacy class that contains 4 U = exp(−2π(α − iγ)). In another complex structure,
often called I, MH (G, C; α, β, γ) parametrizes Higgs bundles (E, ϕ), where
ϕ ∈ H 0 (C , K ⊗ ad(E)) has a pole at z = 0, with ϕ ∼ 12 (β + iγ)(dz/z). More-
over, there is a Hitchin fibration, and most of the usual statements about the
unramified case—those that we have explained and those that we have omitted
here—have close analogs. For a much more detailed explanation, and references
to the original literature, see Gukov and Witten (2006).
The variables α, β, γ are a natural set of parameters for the classical geometry.
However, quantum mechanically, there is one more natural variable η (analogous
to the usual θ angles of gauge theory), as described in section 2.3 of Gukov
and Witten (2006). With the complete set of parameters (α, β, γ, η) at hand,
it is possible to formulate a natural duality statement, according to which
MH (L G, C; L α, L β, L γ, L η) is mirror to MH (G, C; α, β, γ, η), with a certain map
between the parameters, described in section 2.4 of Gukov and Witten (2006).
The main point of this map is that (α, η) = (L η, −L α). Since the monodromy U
depends on L α, this shows that the dual of the monodromy involves the quantum
parameter η that is invisible in the classical geometry. In the A-model, η becomes
the imaginary part of the complexified Kähler class.
4 For simplicity, we assume that U is regular. The more involved statement that holds in
Remark 7.2 We pause here to explain one very elementary fact about the
classical geometry that will be helpful as background for Section 7.5. In complex
structure J, a solution of Hitchin’s equations with the singularity of (7.2)
describes a flat GC bundle E → C with monodromy around the puncture p.
E can be extended over p as a holomorphic bundle, though of course not as a
flat one, and moreover from a holomorphic point of view, E can be trivialized
near p. The flat connection on E → C is then represented, in this gauge, by a
holomorphic (1, 0)-form on C (valued in the Lie algebra of GC ) with a simple
pole at p:
α − iγ
A = dz + ... , (7.4)
iz
where the omitted terms are regular at z = 0. The singularity of the connection
at z = 0 is a simple pole because the ansatz (7.2) for Hitchin’s equations only
allows a singularity of order 1/|z|. A holomorphic connection with such a simple
pole is said to have a regular singularity.
where regular terms are omitted. A meromorphic connection with a pole of degree
greater than 1 is said to have an irregular singularity.
Trying to formulate a duality statement for this situation poses, at first sight,
a severe challenge for the approach to geometric Langlands described here. Our
basic point of view is that the fundamental duality statements depend on C only
as an oriented two-manifold. A complex structure on C is introduced only as a
tool to answer certain natural questions that can be asked without introducing
the complex structure.
From this point of view, tame ramification is natural because a simple pole
in this sense has a clear topological meaning. A meromorphic connection with a
simple pole at a point p ∈ C is a natural way to encode the monodromy about
p of a flat connection on C = C\p. And this monodromy, of course, is a purely
topological notion. But what could possibly be the topological meaning of a
connection with a pole of degree greater than 1?
A closely related observation is that T1 is the residue of the pole in A at
z = 0, and so is independent of the choice of local coordinate z. However, the
coefficients T2 , . . . , Tn of the higher order poles most definitely do depend on the
choice of a local coordinate. How can we hope to include them in a theory that
is supposed to depend on C only as an oriented two-manifold?
Moreover, if the plan is to formulate a duality conjecture of a topological nature
and then prove it using Hitchin’s equations, we face the question of whether
Hitchin’s equations are compatible with an irregular singularity. Hitchin’s equa-
tions for a pair Φ = (A, φ) are schematically of the form dΦ + Φ2 = 0. If near
z = 0, we have a singularity with |Φ| ∼ 1/|z|n , then |dΦ| ∼ 1/|z|n+1 and |Φ|2 ∼
1/|z|2n . For n = 1, dΦ and |Φ|2 are comparable in magnitude, and therefore
Hitchin’s equations look reasonable. However, for n > 1, we have |Φ|2 >> |dΦ|,
and it looks like the non-linear term in Hitchin’s equations will be too strong.
Both questions, however, have natural answers. The answer to the first ques-
tion is that, despite appearances, one actually can associate to a connection
with irregular singularity something that goes beyond the ordinary monodromy
and has a purely topological meaning. The appropriate concept is an extended
monodromy that includes Stokes matrices as well as the ordinary monodromy.
Stokes matrices are part of the classical theory of ordinary differential equations
with irregular singularity (e.g. see Wasow 1965).
Assuming for brevity that the leading coefficient Tn of the singular part of
the connection is regular semi-simple, one can make a gauge transformation to
conjugate T1 , . . . , Tn to the maximal torus. Then one defines a moduli space
Y(G, C; T1 , . . . , Tn ) that parametrizes, up to a gauge transformation, pairs con-
sisting of a holomorphic GC -bundle over C and a connection with an irregular
singularity of the form described in (7.5). As shown in Boalch (2001), it turns
out that this space Y(G, C; T1 , . . . , Tn ) is in a natural way a complex symplectic
manifold, with a complex symplectic structure that depends on C only as an
oriented two-manifold. This can be proved by adapting to the present setting
the gauge theory definition of the symplectic structure, formulated in Atiyah and
122 Mirror symmetry, Hitchin’s equations, and Langlands duality
precise, we can denote this space as Y( G, C; T1 , n, ∗). The group B(G) acts via
L L
monodromies on both the B-model and the A-model of Y(G, C; T1 , n, ∗). Dually,
the corresponding braid group B(L G) acts on the B-model and the A-model
of Y(L G, C; L T1 , n, ∗). However, the two groups B(G) and B(L G) are naturally
isomorphic; indeed, modulo a choice of an invariant quadratic form, there is
a natural map from treg L reg
C to tC , so the two spaces have the same fundamental
group and a choice of basepoint in one determines a basepoint in the other, up to
homotopy. A better (but still not yet precise) question is whether there is a mirror
symmetry between Y(G, C; T1 , n, ∗) and Y(L G, C; L T1 , n, ∗) that commutes with
the braid group.
We expect as well that this mirror symmetry depends on C only as an oriented
two-manifold, and so commutes with the mapping class group. We can think of
the mapping class group of C and the braid group as playing quite parallel roles.
In fact, because of the appearance of the fiber Kp of the canonical bundle in the
last paragraph, these two groups do not simply commute with each other; the
group that acts is an extension of the mapping class group by B(G).
Four-dimensional gauge theory and stacks 123
Just as in the tamely ramified case, to get the right mirror symmetry con-
jecture, we need to extend the parameters slightly to get a mirror-symmetric
set. But we also face the fundamental question of whether Hitchin’s equations
are compatible with wild ramification. As explained above, the non-linearity of
Hitchin’s equations makes this appear doubtful at first sight. But happily, it
turns out that all is well, as shown in Biquard and Boalch (2004). The key point
is that, again with Tn assumed to be regular semi-simple, we can assume that
the singular part of the connection is abelian. Though Hitchin’s equations are
non-linear, they become linear in the abelian case, and once abelianized, they are
compatible with a singularity of any order. Using this as a starting point, it turns
out that, for any n, one can develop a theory of Hitchin’s equations with irregular
singularity that is quite parallel to the more familiar story in the unramified case.
For example, the moduli space MH of solutions of the equations is hyper-Kähler.
In one complex structure, MH parametrizes flat connections with a singularity
similar to that in (7.5); in another complex structure, it parametrizes Higgs
bundles (E, ϕ) in which ϕ has an analogous pole of order n. There is a Hitchin
fibration, and all the usual properties have close analogs.
All this gives precisely the right ingredients to use Hitchin’s equations to
establish the desired mirror symmetry between the two moduli spaces. See
Witten (2007) for a detailed explanation in which this classical geometry is
embedded in four-dimensional gauge theory. Many of the arguments are quite
similar to those given in the tame case in Gukov and Witten (2006). The
construction makes it apparent that the duality commutes with isomonodromic
deformation.
Finally, one might worry that the assumption that Tn is regular semi-simple
may have simplified things in some unrealistic way. This is actually not the
case. For one thing, the analysis in Biquard and Boalch (2004) requires only
that T2 , . . . , Tn should be simultaneously diagonalizable (in some gauge), and in
particular semi-simple, but not that Tn is regular. But even if these coefficients
are not semi-simple, there is no essential problem. In the classical theory of
ordinary differential equations, it is shown that given any such equation with
an irregular singularity at z = 0, after possibly passing to a finite cover of the
punctured z-plane and changing the extension of a holomorphic bundle over
the puncture at z = 0, one can reduce to the case that the irregular part of
the singularity has the properties assumed in Biquard and Boalch (2004). Given
this, one can adapt all the relevant arguments concerning geometric Langlands
duality to the more general case, as is explained in section 6 of Witten (2007).
5 Moreover, these singularities are worse than orbifold singularities. Orbifold singularities
would cause no difficulty. See Frenkel and Witten (2007) for a discussion of orbifold singularities
in geometric Langlands.
Four-dimensional gauge theory and stacks 125
set of fields on M but can be naturally completed to one. For example, A(x, y) is
locally a one-form tangent to the second factor in M = Σ × C; to get a four-
dimensional gauge field, we should relax the condition that A is tangent to
the second factor. Similarly, we can extend φ to an adjoint-valued one-form
on Σ × C. N = 4 super Yang–Mills theory, or rather its twisted version that is
related to geometric Langlands, is obtained by completing this set of fields to a
supersymmetric combination in a minimal fashion.
In N = 4 super Yang–Mills theory, there are no singularities analogous to
the singularities of MH (G, C). The space of gauge fields, for example, is an
affine space, and the other fields (such as φ) take values in linear spaces. The
problems with singularities that make it difficult to define a sigma model of
maps Φ : Σ → MH (G, C) have no analog in defining gauge theory on M = Σ × C
(or any other four-manifold). The relation between the two is that the two-
dimensional sigma model is an approximation to the four-dimensional gauge
theory. The approximation breaks down when one runs into the singularities of
MH (G, C). Any question that involves those singularities should be addressed
in the underlying four-dimensional gauge theory.
But away from singularities, it suffices to consider only the smaller set of fields
that describe a map Φ : Σ → MH (G, C). Many questions do not depend on the
singularities and for these questions the description via two-dimensional sigma
models and mirror symmetry is adequate.
7.6.1 Stacks
To conclude, we will make contact with the counterpart of this discussion in the
usual mathematical theory. We start with bundles rather than Higgs bundles
because this case will be easier to explain.
In the usual mathematical theory, the right-hand side of the geometric Lang-
lands correspondence is described in terms of D-modules on, roughly speaking,
the moduli space of all holomorphic GC bundles on the Riemann surface C.
However, instead of the moduli space M(G, C) of semi-stable holomorphic GC
bundles E → C, one considers D-modules on the “stack” BunG (C) of all such
bundles. The main reason for this is that to define the action of Hecke operators,
it is necessary to allow unstable bundles. Unstable bundles are related to the
nonorbifold singularities of M(G, C).
What is a stack? Roughly, it is a space that can everywhere be locally described
as a quotient. The trivial case is a stack that can actually be described globally as
a quotient. Interpreting BunG (C) as a global quotient would mean finding a pair
(Y, WC ), consisting of a smooth algebraic variety Y and a complex Lie group WC
acting on Y , with the following properties. Isomorphism classes of holomorphic
GC bundles E → C should be in one-to-one correspondence with WC orbits on
Y , and for every E → C, its automorphism group should be isomorphic to the
subgroup of WC leaving fixed the corresponding point in Y .
126 Mirror symmetry, Hitchin’s equations, and Langlands duality
A pair (Y, WC ) representing in this way the stack BunG (C) does not exist
if Y and WC are supposed to be finite-dimensional. Indeed, the GC -bundle
E → C can be arbitrarily unstable, so there is no upper bound on the dimension
of its automorphism group. So no finite-dimensional WC can contain all such
automorphism groups as subgroups.
However as shown in Atiyah and Bott (1982), taking G to be of adjoint type
for simplicity, there is a natural infinite-dimensional pair (Y, WC ). One simply
takes Y to be the space of all connections on a given G-bundle E → C which
initially is defined only topologically. One defines W to be the group of all gauge
transformations of the bundle E; thus, if E is topologically trivial, we can identify
W as the group Maps(C, G). Then we take WC to be the complexification of W ,
or in other words Maps(C, GC ). (This complexification acts on Y as follows. We
associate to a connection A the corresponding ∂¯ operator ∂¯A . Then a complex-
valued gauge transformation acts by ∂¯A → g ∂¯A g −1 .)
Suppose then that we were presented with the problem of making sense of the
supersymmetric sigma model of maps Φ : Σ → M(G, C), given the singularities
of M(G, C). (This is a practice case for our actual problem, which involves
MH (G, C) rather than M(G, C).) Our friends in algebraic geometry would tell
us to replace M(G, C) by the stack BunG (C). We interpret this stack as the pair
(Y, WC ), where Y is the space of all connections on a G-bundle E → C and WC
is the complexified group of gauge transformations. The connected components
of the stack correspond to the topological choices for E.
By a supersymmetric sigma model with target a pair (Y, WC ), with WC a
complex Lie group acting on a complex manifold Y , we mean 6 in the finite-
dimensional case a gauge-invariant supersymmetric sigma model in which the
gauge group is W (a maximal compact subgroup of WC ) and the target is Y .
Actually, to define this sigma model, we want Y to be a Kähler manifold with a
W -invariant (but of course not WC -invariant) Kähler structure. The sigma model
action contains a term which is the square of the moment map for the action
of W . This term is minimized precisely when the moment map vanishes. The
combined operation of setting the moment map to zero and dividing by W is
equivalent classically to dividing by WC .
To write down the term in the action that involves the square of the moment
map (and in fact, to write down the kinetic energy of the gauge fields) one needs
an invariant and positive definite quadratic form on the Lie algebra of W . If W
is finite-dimensional, existence of such a form is equivalent to W being compact.
However, the appropriate quadratic form also exists in the infinite-dimensional
case that W = Maps(C, G) for some space C. (An element of the Lie algebra of
W is a g-valued function on C, and the quadratic form is defined by C dµ ( , ),
where (, ) is an invariant positive-definite quadratic form on g, and µ is a suitable
measure on C.)
6 For a discussion of this construction in relation to stacks, see Pantev and Sharpe (2006).
References 127
Acknowledgement
This chapter was supported in part by NSF Grant Phy-0503584.
References
Atiyah, M. F. and Bott, R. (1982). The Yang-Mills equations over Riemann surfaces.
Phil. Trans. R. Soc. London A308, 166–86.
Atiyah, M. F. and Hitchin, N. J. (1988). The Physics and Geometry of Monopoles.
Princeton University Press, Princeton, NJ.
128 Mirror symmetry, Hitchin’s equations, and Langlands duality
Bershadsky, M., Johansen, A., Sadov, V. and Vafa C. (1995). Topological reduction of
4 − D SYM to 2 − D sigma models. Nucl. Phys. B448, 166–86.
Biquard, O. and Boalch, P. (2004). Wild non-abelian Hodge theory on curves. Compo-
sitio Mathematica 140, 179–204.
Boalch, P. (2001). Symplectic manifolds and isomonodromic deformations. Adv. Math.
163, 137–205.
Brink, L., Schwarz, J. H. and Scherk, J. (1977). Supersymmetric Yang-Mills theories.
Nucl. Phys. B121, 77.
Frenkel, E. and Gaitsgory, D. (2005). Local geometric Langlands correspondence and
affine Kac-Moody algebras. math.RT/0508382.
Frenkel, E. and Witten, E. (2007). Geometric endoscopy and mirror symmetry.
arXiv.0710.5939.
Gualtieri, M. (2003). Generalized complex geometry. DPhil thesis, Oxford University,
Oxford, math.DG/0401221.
Gukov, S. and Witten, E. (2006). Gauge theory, ramification, and the geometric
Langlands program. hep-th/0612073.
Harvey, J. A., Moore, G. W. and Strominger A. (1995). Reducing S duality to T duality.
Phys. Rev. D52, 7161–7.
Hausel, T. and Thaddeus, M. (2002). Mirror symmetry, Langlands duality, and the
Hitchin system. math.AG/0205236.
Hitchin, N. J. (1987a). The self-duality equations on a Riemann surface. Proc. London
Math. Soc. (3) 55, 67.
Hitchin, N. J. (1987b). Stable bundles and integrable systems. Duke Math. J. 54, 91–
114.
Hitchin, N. J. (2003). Generalized Calabi-Yau manifolds. Q. J. Math. 54, 281–308,
math.DG/0209099.
Hitchin, N. J., Karlhede, A., Lindstrom, U. and Rocek M. (1987). Hyperkähler metrics
and supersymmetry. Commun. Math. Phys. 108, 535–89.
Hori, K. et al., (2003). editors. Mirror Symmetry. American Mathematical Society,
Providence, RI.
Jimbo, M., Miwa, T. and Ueno, K. (1981). Monodromy preserving deformations of
linear differential equations with rational coefficients. Physica 2D, 407–48.
Kapustin, A. and Witten, E. (2007). Electric-magnetic duality and the geometric
Langlands program. Commun. Number Theory Phys. 1, 1–236, hep-th/0604151.
Montonen, C. and Olive, D. (1977). Magnetic monopoles as gauge particles?. Phys.
Lett. B72, 117.
Pantev, T. and Sharpe, E. (2006). GLSM’s for gerbes (and other toric stacks). Adv.
Theor. Math. Phys. 10, 77–121, hep-th/0502053.
Simpson, C. (1990). Harmonic bundles on noncompact curves. J. Am. Math. Soc. 3,
713–70.
Strominger, A., Yau, S.T. and Zaslow, E. (1996). Mirror symmetry is T duality. Nucl.
Phys. B479, 243–59, hep-th/9606040.
Wasow, W. (1965). Asymptotic Expansions for Ordinary Differential Equations. Wiley,
New York.
Witten, E. (2007). Gauge theory and wild ramification. arXiv:0710.0631.
VIII
HIGGS BUNDLES AND GEOMETRIC STRUCTURES
ON SURFACES
William M. Goldman
Dedicated to Nigel Hitchin on the occasion of his 60th birthday
8.1 Introduction
In the late 1980s Hitchin (1987) and Simpson (1988) discovered deep connec-
tions between representations of fundamental groups of surfaces and algebraic
geometry. The fundamental group π = π1 (Σ) of a closed orientable surface Σ of
genus g > 1 is an algebraic object governing the topology of Σ. For a Lie group
G, the space of conjugacy classes of representations π → G is a natural algebraic
object Hom(π, G)/G whose geometry, topology, and dynamics intimately relates
to the topology of Σ and the various geometries associated with G. In partic-
ular Hom(π, G)/G arises as a moduli space of locally homogeneous geometric
structures as well as flat connections on bundles over Σ.
Giving Σ a conformal structure profoundly affects π and its representations.
This additional structure induces further geometric and analytic structure on the
deformation space Hom(π, G)/G. Furthermore this analytic interpretation allows
Morse-theoretic methods to compute the algebraic topology of these non-linear
finite-dimensional spaces.
For example, when G = U(1), the space of representations is a torus of dimen-
sion 2g. Giving Σ a conformal structure denotes the resulting Riemann surface
by X. The classical Abel–Jacobi theory identifies representations π1 (X) −→ U(1)
with topologically trivial holomorphic line bundles over X. The resulting Jacobi
variety is an abelian variety, whose structure strongly depends on the Riemann
surface X. However the underlying symplectic manifold depends only on the
topology of Σ, and indeed just the fundamental group π. (Compare Goldman
1984).
Another important class of representations of π arises from introducing the
local structure of hyperbolic geometry to Σ. Giving Σ a Riemannian metric
of curvature −1 determines a representation ρ in the group G = Isom+ (H2 ) ∼ =
PSL(2, R). These representations, which we call Fuchsian, are characterized
as embeddings of π onto discrete subgroups of G. Equivalence classes of
130 Higgs bundles and geometric structures on surfaces
b2
a2 a1
b1
Figure 8.1 The pattern of identifications for a genus 2 surface. The sides of
an octagon are pairwise identified to construct a surface of genus 2. The 8
vertices identify to a single 0-cell in the quotient, and the 8 sides identify to
four 1-cells, which correspond to the 4 generators in the standard presentation
of the fundatmental group.
a2
b2
b1 a1
set. In particular Hom(π, G) inherits both the Zariski and the classical topology.
We consider the classical topology unless otherwise noted.
In terms of the standard presentation (8.1), Hom(π, G) identifies with the
subset of G2g consisting of
(α1 , β1 , . . . αg , βg )
[α1 , β1 ] · · · [αg , βg ] = 1.
8.2.3 Symmetries
The product Aut(π) × Aut(G) acts naturally by left and right compositions, on
Hom(π, G): An element
φ−1 ρ α
π −−→ π −
→G−
→ G.
One motivation for this study is that the deformation spaces provide natural
objects upon which mapping class groups act (Goldman 2006).
Abelian groups and rank 1 Higgs bundles 133
M
hyper-Kähler, relating Hom(π, G)/G to Tn,0 (X) might lead to a hyper-Kähler
geometry on Hom(π, G)/G.
Thus a neighborhood of the U(n) representations in the space of GL(n, C)
corresponds to a neighborhood of the zero-section of T ∗ Mn,0 (X). In turn,
elements in this neighborhood identify with pairs (V, Φ) where V is a semistable
holomorphic vector bundle and Φ is a tangent covector to V in the space of
holomorphic vector bundles. Such a tangent covector is with a Higgs field, by
definition, an End(V )-valued holomorphic one-form on X.
Although one can define a hyper-Kähler structure on the moduli space of
such pairs, the hyper-Kähler metric is incomplete and not all irreducible linear
representations arise. To rectify this problem, one must consider Higgs fields on
possibly unstable vector bundles.
Following Hitchin (1987) and Simpson (1988), define a Higgs pair to be a pair
(V, Φ) where V is a (not ncessarily semistable) holomorphic vector bundle and
the Higgs field Φ a End(V )-valued holomorphic one-form. Define (V, Φ) to be
stable if and only if for all Φ-invariant holomorphic subbundles W ⊂ V ,
deg(W ) deg(V )
< .
rank(W ) rank(V )
l
The Higgs bundle (V, Φ) is polystable if and only if (V, Φ) = i=1 (Vi , Φi ) where
each summand (Vi , Φi ) is stable and
deg(Vi ) deg(V )
=
rank(Vi ) rank(V )
for i = 1, . . . , l.
The following basic result follows from Hitchin (1987) and Simpson (1988),
with a key ingredient (the harmonic metric) supplied by Corlette (1988) and
Donaldson (1987):
Theorem 8.1 The following natural bijections exist between equivalences
classes:
) )
Stable Higgs pairs Irreducible representations
←→ ρ
(V, Φ) over Σ π1 (Σ) −
→ GL(n, C)
) )
Polystable Higgs pairs Reductive representations
←→ ρ
(V, Φ) over Σ π1 (Σ) −
→ GL(n, C)
When the Higgs field Φ = 0, this is just the Narasimhan–Seshadri theorem,
identifying stable holomorphic vector bundles with irreducible U(n) represen-
tations. Allowing the Higgs field Φ to be nonzero, even when V is unstable, leads
to a rich new class of examples, which can now be treated using the techniques
of geometric invariant theory.
Hyperbolic geometry 137
ψα |C = gC,α,β ◦ ψβ |C .
The resulting local hyperbolic geometry defined on the patches by the coordinate
charts is independent of the charts, and extends to a global structure on Σ. The
surface Σ with this refined structure of local hyperbolic geometry will be called
a hyperbolic surface and denoted by M . Such a structure is equivalent to a
Riemannian metric of constant curvature −1. The equivalence follows from two
basic facts:
r Any two Riemannian manifolds of curvature −1 are locally isometric.
r A local isometry from a connected subdomain of H2 extends globally to an
isometry of H2 .
φ
Suppose M1 , M2 are two hyperbolic surfaces. Define a morphism M1 − → M2 as
a map φ, which, in the preferred local coordinates of M1 and M2 , is defined
by isometries in G. Necessarily a morphism is a local isometry of Riemannian
f
manifolds. Furthermore, if M is a hyperbolic surface and Σ − → M is a local
homeomorphism, there exists a hyperbolic structure on Σ for which f is a mor-
phism. In particular every covering space of a hyperbolic surface is a hyperbolic
surface.
In more traditional terms, a morphism of hyperbolic surfaces is just a local
isometry.
138 Higgs bundles and geometric structures on surfaces
the holonomy representation of the hyperbolic surface M . The pair (devM , holM )
is unique up to the G-action defined by
g
(devM , holM ) −→ (g ◦ devM , Inn(g) ◦ holM )
for g ∈ Isom+ (H2 ).
If the hyperbolic structure is complete, that is, the Riemannian metric is
geodesically complete, then the developing map is a global isometry M̃ ≈ H2 .
In that case the π-action on H2 defined by the holonomy representation ρ is
equivalent to the action by deck transformations. Thus ρ defines a proper free
action of π on H2 by isometries. Conversely if ρ defines a proper free isometric
π-action, then the quotient
M := H2 /ρ(π)
is a complete hyperbolic manifold with a preferred isomorphism
ρ
π1 (Σ) −
→ ρ(π) ⊂ G.
This isomorphism (called a marking) determines a preferred homotopy class of
homotopy equivalences
Σ −→ M.
structure with a single branch point, that is, a point with local coordinate
given by a branched conformal mapping z −→ z k where k 1. (The nonsingular
case corresponds to k = 1.) In our example k = 2 and the singular point has a
neighborhood isometric to a hyperbolic cone of cone angle 4π.
A− − + + − − + +
1 , B1 , A1 , B1 , A2 , B2 , A2 , B2 ,
a1 , b1 , a2 , b2 ∈ PSL(2, R)
according to the pattern described in Figure 8.1. Given any two oriented geo-
desic segments in H2 of equal length, a unique orientation-preserving isometry
Figure 8.3 A regular octagon with vertex angles π/4 can be realized in the
tiling of H2 by triangles with angles π/2, π/4, and π/8. The identifications
depicted in Figure 8.1 are realized by orientation-preserving isometries. The eight
angles of π/4 fit together to form a cone of angle 2π, forming a coordinate chart
for a hyperbolic structure around that point.
140 Higgs bundles and geometric structures on surfaces
maps one to the other. Since the polygon is regular, one can realize all four
identifications in Isom+ (H2 ).
The quotient (compare Figure 8.2) contains three types of points:
r A point in the open 2-cell has a coordinate chart which is the embedding
P → H2 .
r A point on the interior of an edge has a half-disc neighborhood, which
together with the half-disc neighborhood of its part, gives a coordinate chart
for the corresponding point in the quotient.
r Around the single vertex in the quotient is a cone of angle
8(π/4) = 2π,
a1 b1 a−1 −1 −1 −1
1 b1 a2 b2 a2 b2 = 1
Figure 8.4 A regular right-angled octagon can also be realized in the tiling of
H2 by triangles with angles π/2, π/4, and π/8. The identifications depicted in
Figure 8.1 are realized by orientation-preserving isometries. The eight angles of
π/2 fit together to form a cone of angle 4π, forming a coordinate chart for a
singular hyperbolic structure, branched at one point.
Hyperbolic geometry 141
Figure 8.5 A degree 1 map from a genus 3 surface to a genus 2 surface which
collapses a handle. Such a map is not homotopic to a smooth map with branch
point singularities (such as a holomorphic map).
142 Higgs bundles and geometric structures on surfaces
M
}>
f }
}}
}}}
φ
}
Σ / M
f
moduli of hyperbolic structures on surfaces goes back at least to Fricke and Klein
(1897/1912).
The Teichmüller space T(Σ) of Σ is defined similarly, as the space of equiva-
lence classes of marked conformal structures on Σ, that is, pairs (X, f ) where X
f
is a Riemann surface and Σ − → X is a homotopy equivalence. Teichmüller used
quasiconformal mappings to parametrize T(Σ) by elements of a vector space,
define a metric on T(Σ), and prove analytically that T(Σ) is a cell. Using these
ideas, Ahlfors (1960) proved T(Σ) is naturally a complex manifold.
Since a hyperbolic structure is a Riemannian metric, every hyperbolic struc-
ture has an underlying conformal structure. The uniformization theorem asserts
that if χ(Σ) < 0, then every conformal structure on Σ underlies a unique
hyperbolic structure. The resulting identification of conformal and hyperbolic
structures identifies T(Σ) with F(Σ). As discussed below, F(Σ) identifies with
an open subset of Hom(π, PSL(2, R))/PSL(2, R) which has no apparent complex
structure. Thus the complex structure on T(Σ) is more mysterious when T(Σ)
is viewed as a space of hyperbolic structures. For a readable survey of classical
Teichmüller theory, see Bers (1972), Abikoff (1980), Nag (1988), Imayoshi and
Taniguchi (1992), Gardiner-Lakic (2000) or Hubbard (2006).
H2 / (H2 )ρ
Σ
ρ
associated to a representation π −
→ PSL(2, R) as follows. The total space is the
quotient
(H2 )ρ := (Σ̃ × H2 )/π
Moduli of hyperbolic structures and representations 145
satisfies
(Milnor 1958 and Wood 1976). Call a representation maximal if equality holds
in (8.3), that is, Euler(ρ) = ±χ(Σ).
The following converse was proved in Goldman (1980) (compare also Hitchin
1987 and Goldman 1988).
l
Euler(ρ) = χ(Σ) + ki .
i=1
π −→ PSL(2, R) → PSL(2, C)
ρ
Let π −
→ G be a representation and let
X / (X)ρ
Σ
be the corresponding flat (G, X)-bundle over Σ. Then the G-invariant Kähler
form ω on X defines a closed exterior two-form ωρ on the total space (X)ρ . Let
s
Σ− → (X)ρ be a smooth section. Then the integral
s ∗ ωρ
Σ
X
f }}>
}} φ
}}}
}
Σ / X
f
homotopy commutes.
Theorem 8.6 (Uniformization) Let X be a Riemann surface with χ(X) < 0.
Then there exists a unique hyperbolic metric whose underlying conformal struc-
ture agrees with X.
Since every hyperbolic structure possesses an underlying conformal structure,
Fricke space F(Σ) maps to Teichmüller space T(Σ). By the uniformization
theorem, F(Σ) → T(Σ) is an isomorphism. It is both common and tempting
to confuse these two deformation spaces. In the present context, however, it
seems best to distinguish between the representation/hyperbolic structure and
the conformal structure.
For example, each Fuchsian representation determines a marked hyperbolic
structure, and hence an underlying marked conformal structure. An equivalence
class of Fuchsian representations thus determines a special point in Teichmüller
space. This contrasts sharply with other representations which do not generally
pick out a preferred point in T(Σ). This preferred point can be characterized as
the unique minimum of an energy function on Teichmüller space.
The construction, due to Tromba (1992), is as follows. Given a hyperbolic
f
surface M and a homotopy equivalence X − → M , then by Eels and Sampson
F
(1964) a unique harmonic map X − → M exists homotopic to f . The harmonic
map is conformal if and only if M is the uniformization of X. In general the
nonconformality is detected by the Hopf differential Hopf(F ) ∈ H 0 (X, KX 2
),
defined as the (2, 0) part of the pull-back by F of the complexified Riemannian
metric on M . The resulting mapping
F(X) −→ H 0 (X, KX
2
)
(f, M ) −→ Hopf(F )
is a diffeomorphism.
Moduli of hyperbolic structures and representations 149
Fixing M and letting the marked complex structure (f, X) vary over T(Σ)
yields an interesting invariant discussed in Tromba (1992), and extended in
Goldman and Wentworth (2005) and Labourie (to appear). The energy of the
harmonic map F = F (f, X, M ) is a real-valued function on T(Σ). In the present
context it is the square of the L2 -norm of Hopf(F ).
L2 ∼
= KX
L −→ Eρ̃ −→ Eρ̃ /L ∼
= L−1
150 Higgs bundles and geometric structures on surfaces
H 1 (X, Hom(L−1 , L) ∼
= H 1 (X, K) ∼
=C
defining Serre duality (compare Gunning (1967)).
One resolves this difficulty by changing the question. Replace the extension
class ε by an auxiliary holomorphic object, a Higgs field
Φ ∈ H 0 (X; KX ⊗ End(E)),
for the vector bundle E := L ⊕ L−1 so that the Higgs pair (E, Φ) is stable in the
appropriate sense. In our setting the Higgs field corresponds to the everywhere
nonzero holomorphic section of the trivial holomorphic line bundle
C∼
= KX ⊗ Hom(L, L−1 ) ⊂ KX ⊗ End(E).
Now the only Φ-invariant holomorphic subbundle of E is L−1 which is negative,
and the pair (E, Φ) is stable.
into the symmetric space GL(n, C)/U(n). The metric h determines a reduction of
structure group of Eρ̃ from GL(n, C) to U(n), giving Eρ̃ a Hermitian structure.
Let A denote the unique connection on Eρ̃ which is unitary with respect to h.
The harmonic metric determines the Higgs pair (V, ∂¯V , Φ) as follows:
r The Higgs field Φ is the holomorphic (1, 0)-form ∂h ∈ Ω1 End(V ), where the
tangent space to GL(n, C)/U(n) is identified with a subspace of h∗ End(V ).
r The holomorphic structure d on V arises from conformal structure Σ and
A
the Hermitian connection A.
The Higgs pair satisfies the self-duality equations with respect to the Hermitian
metric h:
dA (Φ) = 0
F (A) + [Φ, Φ∗ ] = 0 (8.4)
Rank 2 Higgs bundles 151
Here F (A) denotes the curvature of A, and Φ∗ denotes the adjoint of Φ with
respect to h. Conversely, Hitchin and Simpson show that every stable Higgs pair
determines a Hermitian metric satisfying (8.4).
{Q ∈ H 0 (X, KX
2
) | div(Q) D} ∼
= C3(g−1)−d .
k
2 − 2g + (θi − 2π) > 0,
i=1
there exists a unique singular hyperbolic surface conformal to X with cone angle
θi at pi .
When the θi are multiples of 2π, then this structure is a branched structure
(and the above theorem follows from Hitchin (1987)). The moduli space of such
branched conformal structures forms a bundle Sd over T(Σ) where the fiber over
a marked Riemann surface Σ → X is the symmetric power Symd (X) where
1
k
d= (θi − 2π).
2π i=1
ρ0 K
π → SL(2, R) −
→G
H 0 X; KX
2
⊕H 0 X; KX
3
⊕ · · · ⊕ H 0 (X; KX
n
)
∼
= C3(g−1) ⊕ C5(g−1) ⊕ · · · ⊕ C(2n−1)(g−1) .
When n is odd, Hitchin proves there are exactly three components. The second
Stiefel–Whitney characteristic class is nonzero on exactly one component; it is
zero on two components, one of which is the Hitchin–Teichmüller component.
Theorem 8.10 The deformation space C(Σ) naturally identifies with the holo-
morphic vector bundle over T(Σ) whose fiber over a marked Riemann surface
Σ → X is H 0 X, KX 3
.
For every such representation, there exists a unique conformal structure so
that
h̃
Σ̃ −
→ SL(3, R)/SO(3)
is a conformal map, that is, the component of the Higgs field in H 0 Σ, KX 2
—the
Hopf differential Hopf(h)—vanishes. This defines the projection C(Σ) → T(Σ).
The zero section corresponds to the Fuchsian RP2 structures, that is, the RP2
structures arising from hyperbolic structures on Σ.
It is natural to attempt to generalize this as follows. For any split real form
G, and Riemann surface X with π1 (X) ∼ = π, Hitchin (1992) identifies a certain
direct sum of holomorphic line bundles VX naturally associated to X so that a
Hitchin component of Hom(π, G)/G identifies with the complex vector space
H 0 X, KX 2
⊕ H 0 (X, VX ).
However, this identification depends crucially on the Riemann surface X and
fails to be Mod(Σ)-invariant. Generalizing the Labourie–Loftin Theorem 8.10,
we conjecture that each Hitchin component of Hom(π, G)/G identifies naturally
with the total space of a holomorphic vector bundle E(Σ) over T(Σ), whose fiber
over a marked Riemann surface X equals H 0 (X, VX ).
f (x1 ) + · · · + f (xn ) = Rn .
Theorem 8.12 (Guichard 2005, Guichard 2008, Labourie 2006) ρ is
Hitchin if and only if ρ preserves hyperconvex curve.
Recently Fock and Goncharov (2006, 2007) have studied this component of rep-
resentations, using global coordinates generalizing Thurston and Penner’s shear-
ing coordinates. In these coordinates the Poisson structure admits a particularly
simple expression, leading to a quantization. Furthermore they find a positive
156 Higgs bundles and geometric structures on surfaces
found that the number of connected components of Hom(π, U(p, q)) equals
2(p + q) min(p, q) (g − 1) + gcd(p, q).
(For a survey of these techniques and other results, compare Bradlow et al. 2006
as well as their recent column Bradlow et al. 2007.)
where R2n ∼ = Sym2n−1 (R2 ) with the symplectic structure induced from R2 .
Another class of maximal representations arises from deformations of compo-
ρ0
sitions of a Fuchsian representation π −→ SL(2, R) with the diagonal embedding
n
∆
- ./ 0
SL(2, R) → SL(2, R) × · · · × SL(2, R) → Sp(2n, R).
More generally, the diagonal embedding extends to a representation
,
∆
SL(2, R) × O(n) → Sp(2n, R)
corresponding to the SL(2, R) × O(n)-equivariant decomposition of the symplec-
tic vector space
R2n = R2 ⊗ Rn
as a tensor product of the symplectic vector space R2 and the Euclidean inner
product space Rn . Deformations of compositions of Fuchsian representations
into SL(2, R) × O(2) with ∆, provide 22g more components of maximal represen-
tations.
For n > 2, these account for all the maximal components. This situation is
more complicated when n = 2. In that case, 4g − 5 components of maximal
representations into Sp(4, R) do not contain representations into smaller compact
extensions of embedded subgroups isomorphic to SL(2, R). In particular the
158 Higgs bundles and geometric structures on surfaces
Acknowledgments
Goldman gratefully acknowledges partial support from National Science Foun-
dation grants DMS-070781, DMS-0103889, and DMS-0405605. This chapter was
presented at the Geometry Conference honoring Nigel Hitchin, Consejo Superior
de Investigaciones Cientı́ficas, Madrid, Spain, 5 September 2006. We gratefully
acknowledge support from the Oswald Veblen Fund at the Institute for Advanced
Study where this work was completed.
I am grateful to Nigel Hitchin for the inspiration of these ideas, and to him and
Simon Donaldson for the opportunity to study with them at the Maths Institute
in Oxford in 1989. Over the years my knowledge of the subject has benefitted
enormously from conversations with Yves Benoist, Steve Bradlow, Marc Burger,
References 159
References
Abikoff, W. (1980). The real-analytic theory of Teichmüller spaces. Lecture Notes in
Mathematics Vol. 820, Springer-Verlag, Berlin, Germany.
Ahlfors, L. (1960). The complex analytic struture of the space of closed Rie-
mann surfaces. in Analytic Functions. Princeton University Press Princeton, NJ,
pp. 45–66.
Barbot, T., Charette, V., Drumm, T., Goldman, W. and Melnick, K. (2008). A primer
on the (2 + 1) Einstein universe in Recent developments in pseudo-Riemannian
geometry. ESI Lect. Math. Phys., Math. Soc., Zurich, 179–229.
Benoist, Y. (2000). Automorphismes des cônes convexes. Inv. Math. 141(1), 149–93.
Benoist, Y. Convexes divisibles. I, Algebraic groups and arithmetic, Tata Inst. Fund.
Res., Mumbai (2004) 339–74. II, Duke Math. J. 120(1) (2003), 97–120. III, Ann. Sci.
École Norm. Sup. (4) 38(5) (2005), 793–832. IV, Structure du bord en dimension 3,
Inv. Math. 164(2) (2006), 249–78.
Benzécri, J. P. (1960). Sur les variétés localement affines et projectives. Bull. Soc. Math.
France 88, 229–332.
Bers, L. (1960). Simultaneous uniformization. Bull. Am. Math. Soc. 66, 94–7.
Bers, L. (1972). Uniformization, moduli, and Kleinian groups. Bull. Lond. Math. Soc.
4, 257–300.
Bers, L. and Gardiner, F. (1986). Fricke spaces. Adv. Math. 62, 249–84.
Bonahon, F. (1986). Bouts des varit́és hyperboliques de dimension 3. Ann. Math. (2)
124(1), 71–158.
Bradlow, S., Garcı́a-Prada, O. and Gothen, P. (2003). Surface group representations
and U (p, q)-Higgs bundles. J. Diff. Geom. 64(1), 111–70.
Bradlow, S., Garcı́a-Prada, O. and Gothen, P. (2006). Maximal surface group repre-
sentations in isometry groups of classical Hermitian symmetric spaces. Geom. Ded.
122, 185–213.
Bradlow, S., Garcı́a-Prada, O. and Gothen, P. (2007). What is a Higgs bundle?. Notices
Am. Math. Soc. 54(8), 980–1.
Bradlow, S., Garcı́a-Prada, O. and Gothen, P. (2009). Deformations of maximal repre-
sentations in Sp(4, R). arXiv:0903.5496.
Burger, M. and Iozzi, A. (2002). Boundary maps in bounded cohomology. Geom. Funct.
Anal. 12, 281–92.
160 Higgs bundles and geometric structures on surfaces
Burger, M., Iozzi, A., Labourie, F. and Wienhard, A. (2005). Maximal representations
of surface groups: symplectic Anosov structures, Pure Appl. Math. Q., Special Issue:
In Memory of Armand Borel Part 2 of 3, 1(3), 555–601.
Burger, M. and Iozzi, A. (2007). Bounded differential forms, generalized Milnor–Wood
inequality and applications to rigidity theory. Geom. Ded. 125, 1–23.
Burger, M., Iozzi, A. and Wienhard, A. (2003). Surface group representations with
maximal Toledo invariant. C. R. Acad. Sci. Paris, Sér. I. 336, 387–90.
Burger, M., Iozzi, A. and Wienhard, A. Representations of surface groups with maximal
Toledo invariant. Preprint. math.DG/0605656.
Burger, M. and Monod, N. (2002). Continuous bounded cohomology and applications
to rigidity theory. Geom. Funct. Anal. 12, 219–80.
Buser, P. (1992). Geometry and spectra of compact Riemann surfaces. Progress in
Mathematics. Vol. 106, Birkhuser Boston, Inc., Boston, MA.
Choi, S. and Goldman, W. (1993). Convex real projective structures on closed surfaces
are closed. Proc. Am. Math. Soc. 118(2), 657–61.
Chuckrow, V. (1968). On Schottky groups with applications to Kleinian groups. Ann.
Math. (2) 88, 47–61.
Clerc, J. L. and Ørsted, B. (2003). The Gromov norm of the Kähler class and the
Maslov index. Asian J. Math. 7, 269–96.
Corlette, K. (1988). Flat G-bundles with canonical metrics. J. Diff. Geom. 28, 361–82.
Domic, A. and Toledo, D. (1987). The Gromov norm of the Kaehler class of symmetric
domains. Math. Ann. 276(3), 425–32.
Donaldson, S. (1987). Twisted harmonic maps and the self-duality equations. Proc.
Lond. Math. Soc. (3) 55(1), 127–31. MR MR 887285 (88g:58040).
Eells, J. and Sampson, J. (1964). Harmonic mappings of Riemannian manifolds. Am.
J. Math. 85, 109–60.
Farkas, H. and Kra, I. (1980). Riemann surfaces. Graduate Texts in Mathematics. Vol.
71, Springer-Verlag, Berlin/Heidelberg/New York.
Fock, V. and Goncharov, A. (2006). Moduli spaces of local systems and
higher Teichmüller theory. Publ. Math. Inst. Hautes Études Sci. 103, 1–211.
math.AG/0311149, 2003.
Fock, V. and Goncharov, A. (2007). Moduli spaces of convex projective structures on
surfaces. Adv. Math. 208(1), 249–273. math.DG/0405348, 2004.
Fricke, R. and Klein, F. Vorlesungen der Automorphen Funktionen, Teubner, Leipzig,
Germany, Vol. I (1897), Vol. II (1912).
Gardiner, F. P. and Lakic, N. (2000). Quasiconformal Teichmüller theory. Mathematical
Surveys and Monographs. Vol. 76, American Mathematical Society Providence, RI.
Garcı́a-Prada, O. and Mundet i Riera, I. (2003). Representations of the fundamental
group of a closed orientable surface in Sp(4, R). Topology 43, 831–55.
Garcı́a-Prada, O., Gothen, P. and Mundet i Riera, I. (2008). Higgs bundles and surface
group representations in the real symplectic group. Preprint arXiv:0809.0576v3.
Goldman, W. (1980). Discontinuous groups and the Euler class, Doctoral thesis.
University of California, Berkeley, CA.
Goldman, W. (1984). The symplectic nature of fundamental groups of surfaces. Adv.
Math. 54, 200–25.
Goldman, W. (1988). Topological components of spaces of representations. Inv. Math.
93(3), 557–607.
References 161
Turaev, V. G. (1985). The first symplectic Chern class and Maslov indices. (Russian)
Studies in topology, V. Zap. Nauchn. Sem. Leningrad. Otdel. Mat. Inst. Steklov.
(LOMI) 143, 110–129, 178.
Turaev, V. G. (1984). A cocycle of the symplectic first Chern class and Maslov indices.
(Russian) Funktsional. Anal. i Prilozhen. 18 (1), 43–48.
Weil, A. (1962). On discrete subgroups of Lie groups II. Ann. Math. 75, 578–602.
Wood, J. (1976). Bundles with totally disconnected structure group. Commun. Math.
Helv. 51, 183–99.
IX
LOCALITY OF HOLOMORPHIC BUNDLES,
AND LOCALITY IN QUANTUM FIELD THEORY
Graeme Segal
Dedicated to Nigel Hitchin, on his 60th birthday
Nigel Hitchin has been a close colleague of mine for most of my mathematical life,
and I have profited enormously from my contact with him. His wonderful skill
in asking the right questions and in obtaining deep results without sacrificing
concreteness or simplicity has both inspired me and filled me with envy. Having
seen, however, how many others at this meeting are better qualified than I am
to talk about Nigel’s work, I decided it would be best to keep to my own terrain,
and talk about locality in quantum field theory. All the same, many of the ideas
involved are well exemplified in the study of bundles on Riemann surfaces which
Nigel is famous for, and I shall begin there, especially as the question of locality
relates to an aspect of his work that has not been talked about so far at this
meeting, namely, its role in so-called ‘geometric Langlands theory’.
We can approach the subject by contrasting two opposite ways of looking
at holomorphic vector bundles E on a compact Riemann surface Σ. At one
extreme, if we remove any finite set σ of points from Σ then E is trivial on the
remaining surface Σ − σ, so we can think of all the ‘twisting’ of the bundle as
being concentrated into tiny neighbourhoods of the points of σ. At the other
extreme, we can try to spread the twist as evenly as possible over all of Σ.
The classical case of line bundles is very simple – perhaps misleadingly so.
On the one hand, any line bundle L can be constructed from a divisor z =
n1 z1 + · · · + nk zk – an element of the free abelian group on the set of points
of Σ – by attaching the trivial bundle L0 on Σ − {z1 , . . . , zk } to trivial bundles
Li in the neighbourhood of each zi by means of the clutching function ζini ,
where ζi is a local parameter at zi . (The resulting bundle L(z) depends up to
canonical isomorphism only on the divisor z.) On the other hand, for any choice
of Riemannian metric on Σ, any line bundle can be given a unique unitary
connection with constant curvature, so that it looks exactly the same in the
neighbourhood of any point of Σ. The isomorphism classes of holomorphic line
bundles on Σ naturally form a commutative complex Lie group Pic(Σ), and the
correspondence between the two ways of looking at a bundle amounts to the
classical theorem that Pic(Σ) is – in the holomorphic category – the free abelian
group on Σ, traditionally called its ‘Albanese variety’, that is, that the map
Noncommutative geometry 165
1 This may seem strange. The set-theoretical free abelian group F generated by the points
Σ
of Σ fits into an exact sequence
×
KΣ → FΣ → PicΣ → 1,
where KΣ is the field of rational functions on Σ. The group FΣ is the disjoint union of
(n) (n)
a sequence of finite-dimensional algebraic varieties FΣ , where FΣ consists of all Σnk xk
(n)
such that Σ|nk | = n. Furthermore, FΣ has a natural topology in which the closure of FΣ is
(m)
compact, and is the union of the FΣ for m ≤ n. But we cannot say that FΣ is any kind of
algebraic variety: in fact it is easy to see that if U is a neighbourhood of the identity element
(n)
of FΣ then any continuous f : U → C for which f |U ∩ FΣ is holomorphic for each n has to
×
be constant along the orbits of KΣ .
166 Locality of holomorphic bundles, and locality in quantum field theory
indexed by the first Chern class of E, but each connected component, though
infinite, is the closure of a single point.
To make better sense of unpromising ‘spaces’ such as BΣ a number of different
approaches are commonly used. In algebraic geometry the main candidates are
1. To regard BΣ as a stack, that is, to work with the category – actually a
groupoid – of bundles and their isomorphisms rather than just with the set
π0 (BΣ ) of isomorphism classes of objects
2. the approach of geometric invariant theory, which picks out a class of
‘stable’ bundles whose isomorphism classes do form a nice space, in fact
an algebraic variety
Here I am going to talk about the first approach, which is more obviously
related to quantum field theory, especially in the treatment of the Langlands
theory by Kapustin and Witten (2007). Nigel Hitchin’s own main tool, however,
was geometric invariant theory.
Stepping back a little from algebraic geometry, one can say that the category
of bundles on a space resembles the category of representations of a group. For
example, the Hitchin moduli space associated to a surface Σ and a compact group
G is – among other things – the space of conjugacy classes of homomorphisms
from the fundamental group π1 (Σ) to the complexified group GC . The ‘space’
of irreducible unitary representations of a group Γ is an archetypal example
in Connes’s theory of noncommutative geometry (Connes 1994). He observes
that the set of irreducible representations is the set Spec(AΓ ) of irreducible
∗-representations of a C∗ -algebra AΓ associated to Γ. If A is a commutative
algebra its irreducible representations are the algebra-homomorphisms A → C,
which form a topological space Spec(A) on which A is an algebra of continuous
complex-valued functions. Connes’s idea is to think of the geometry of the set of
irreducible representations of the group Γ as defined by the noncommutative
algebra AΓ rather than by the – often too small – commutative algebra of
continuous functions on Spec(AΓ ), which in fact is the centre of AΓ .
To relate this picture to the stacks or groupoids of algebraic geometry we
think of a groupoid as halfway between a group and a space. More precisely,
a groupoid B consists of two sets together with some maps between them: a
set B0 of objects, and a set B1 of morphisms which is the disjoint union of
the sets B(x, y) of morphisms from x to y, where x and y run through all the
objects in B0 . We are concerned here, however, with topological groupoids, for
which the sets B0 and B1 have topologies and the structural maps between them
are continuous. At one extreme, if B0 is a point then we have a topological
group; at the other, if B0 = B1 , we have no morphisms except identities, and
the groupoid is simply a space. As Connes has emphasized, a groupoid B has a
groupoid-algebra AB , which interpolates between the group-algebra of a group
and the commutative algebra of functions on a space. (The algebra AB is devised
so that an AB -module is the same thing as a functor from B to vector spaces.
For a discrete groupoid, AB has a vector space basis ef indexed by the set
Noncommutative geometry 167
where GΣi is the group of holomorphic maps from Σi to G = GLn (C), and GS
is the group of smooth maps S → G, for any bundle can be trivialized on a
non-closed surface. We then have three groupoids:
All three are weakly equivalent, and are weakly equivalent to the groupoids
obtained from any other way of decomposing Σ into pieces. According to the
‘stack’ point of view, any of them can be taken as the ‘space’ of bundles
on Σ.
The relation of equivalence of groupoids translates into Morita equivalence of
algebras. Two algebras are Morita equivalent if their categories of left modules
are equivalent. Recall that if A and B are algebras then an (A, B)-bimodule
F defines an additive functor F∗ : MB → MA , where MA and MB are the
categories of left A and B modules, by F∗ (M ) = F ⊗B M . In particular, for the
168 Locality of holomorphic bundles, and locality in quantum field theory
standard technicality.
Algebraic structures up to homotopy 169
Ck (Rd ; P ) → P
for each k > 0; but these must of course satisfy various compatibilities.
170 Locality of holomorphic bundles, and locality in quantum field theory
Notice that applying the theorem when M is an open ball tells us that P is
homotopy-equivalent to the d-fold based loop space of B d P .
The equivalences are defined by the scanning map (see McDuff 1975, 1977
and Segal 1979) which associates to an element c ∈ C(M ; P ), the section of
B d P whose value at z ∈ M is the image of c in C(M ; P )/CM −W (M ; P ) =
C(U ; P )/CV (U ; P ), where W ⊂ U are two concentric open balls around z, and
V = U − W . The theorem is very easy to prove, and not at all deep, in the form
I have stated it here: in applications the main difficulty may be to verify the
hypotheses.
In the application to holomorphic vector bundles on a surface, at a point z on
the surface, with neighbourhood U , the labelling space P is the quotient GǓ /GU ,
where GǓ is the group of holomorphic maps U − {z} → GLn (C) which extend
meromorphically over U , and GU is the group of holomorphic maps U → GLn (C).
This can be identified with GLn (C(t))/GLn (C[t]). It is (see Pressley and Segal
1986) the union of a sequence of compact finite-dimensional algebraic varieties,
and it has the homotopy type of the based loop space of GLn (C). To define the
Quantum field theory 171
topology of B̂Σ we first define a topology on the part B̂Σ,U with support in a
disjoint union U of open discs by identifying it with G∂U /GU , where G∂U is the
group of smooth maps ∂U → GLn (C) which are boundary values of meromorphic
maps U → GLn (C), and then we give B̂Σ the finest topology compatible with
these. Having checked the hypotheses of the theorem, it tells us that B̂Σ has the
homotopy type of the space of continuous maps from Σ to B 2 P BGLn (C),
that is, the homotopy type of the space of smooth bundles, as we expect from
the stack picture.
Before leaving this topic, let us mention the converse question to the one
answered by the theorem. If we are given a space Q, can we model the mapping
spaces Map(M ; Q) for varying d-manifolds M – at least up to homotopy – by
labelled configuration spaces C(M ; P )? The answer, clearly, is: if and only if the
space Q is d-connected; for the delooping B d P of any d-fold loop space P is d-
connected. If we want to model spaces of maps into less highly connected spaces
Q we would have to allow not just ‘particles’ but also configurations of higher
dimensional submanifolds – presumably, up to dimension m if Q is (d − m)-
connected. (One way to see this is to realize Q as an open manifold with a Morse
function with critical points of indices ≥ d − m, and to make maps M → Q flow
downwards along the gradient flow.)
and that the state is a unit vector ψ in H related to θ by θ(a) = ψ, a ψ. But the equivalent
description by (A, θ) seems more satisfactory to me.
5 The essential example is the Bohm–Aharonov phenomenon, when an electromagnetic field
for each finite set {x1 , . . . , xk } of distinct points of M , which are got by
multiplying in A and composing with θ. The Θk are traditionally called vacuum
expectation values.
It is not easy to say what properties the vector spaces Ox and the functions
Θk must have for them to constitute a quantum field theory. A first attempt at
an answer can be given by defining a d-dimensional theory as a rule which
1. associates a complex topological vector space HY to each compact-oriented
Riemannian manifold Y of dimension d − 1, functorially with respect to
diffeomorphisms Y → Y , and
2. associates a trace-class operator UX : HY0 → HY1 to each oriented
Riemannian cobordism X from Y0 to Y1 .
These data are constrained to satisfy two axioms:
(a) Concatenation:
UX ◦X = UX ◦ UX
when X ◦ X is the cobordism from Y0 to Y2 obtained by concatenating
X from Y0 to Y1 with X from Y1 to Y2 .
(b) Tensoring: We are given associative natural isomorphisms
∼
=
HY ⊗ HY → HY Y
UX ⊗ U X = U X X
where the inverse limit is over the ordered set of all closed balls D in M which
are neighbourhoods of x, ordered by
◦
D > D ⇐⇒ D ⊂D,
in which case we have a canonical map U ◦ : H∂D → H∂D defined by the
D−D
annular cobordism.
If x1 , . . . , xk are distinct points of M , and D1 , . . . , Dk are disjoint discs with
xi in the interior of Di , let M0 denote the manifold with boundary obtained by
Quantum field theory 173
just two points. The first is that a three-tier two-dimensional field theory seems
to have a good claim as a candidate definition of a ‘noncommutative manifold’.
For, schematically at least, a natural way to give the data of a three-tier theory
is to associate an algebra AZ to each (d − 2)-manifold, and to take CZ to be the
category of left AZ -modules; then to a cobordism Y is associated an (AZ1 , AZ0 )-
bimodule HY , which defines a functor CZ0 → CZ1 by
M → HY ⊗AZ0 M ;
and to a cobordism between cobordisms is associated a homomorphism of
bimodules. When d = 2 this simply means that we have a dual pair of linear
categories – the left and right modules for an algebra – associated to a point
with its two orientations, while the one-dimensional data expresses the cate-
gorical duality, and the two-dimensional data gives us ‘trace’ or ‘integration’
maps. The field theory even leads one naturally from categories of modules to
categories of cochain complexes of modules, if we assume that the field theory is
supersymmetric.
The idea that two-dimensional theories should replace manifolds is, of course,
the central proposal of string theory, which models space-time by a two-
dimensional conformal field theory, with the category of D-branes in space-time
as the category which the field theory associates to a point. It is also what arises
in the Kapustin–Witten treatment of geometric Langlands duality. There, one
begins from the maximally supersymmetric four-dimensional Yang–Mills theory
associated to a compact group G, and observes that, for any compact surface
Σ, a four-dimensional theory gives a two-dimensional theory by dimensional
reduction along Σ – that is, by composing with the functor M → Σ × M from i-
manifolds to (i + 2)-manifolds. The two-dimensional theory obtained from Yang–
Mills theory for G by reducing along Σ is supposed to associate to a point
the category of D-modules on the moduli space of holomorphic GC -bundles
on Σ.
My second point is even vaguer. The obvious fear, if one starts to study
three-tier theories, is that one will be impelled to believe that a d-dimensional
theory should really mean a (d + 1)-tier d-dimensional theory, which associates
a two-category to a manifold of dimension d − 3, and even worse things to
lower dimensional manifolds, until one gets to a (d − 1)-category associated to a
point.
This is not completely mad, as it works well in the one famous example afforded
by three-dimensional Chern–Simons theory for a compact group G at a given
‘level’. There, the category associated to a circle S is the category of positive
energy representations of the loop group of maps S → G at the specified level,
and the two-category associated to a point is the same thing, but remembering
its tensor structure coming from the fusion of loop group representations. (We
think of this as a two-category with just one object, whose linear category of
endomorphisms is the category of loop group representations, with fusion as its
composition law.)
References 175
References
Atiyah, M. F. and Bott, R. (1982). The Yang-Mills equations over Riemann surfaces.
Phil. Trans. Roy. Soc. London 308A, 523–615.
Connes, A. (1994). Noncommutative Geometry. Academic Press, San Diego, CA.
Kapustin, A. and Witten, E. (2007). Electro-magnetic duality and the geometric
Langlands program. Commun. Number Theory Phys. 1, 1–236. hep-th/0610210.
McDuff, D. (1975). Configuration spaces of positive and negative particles. Topology
14, 91–107.
176 Locality of holomorphic bundles, and locality in quantum field theory
10.1 Introduction
Let Z be a (2 + 1)-dimensional topological quantum field theory (TQFT). That
is, the two dimensional part of Z is a modular functor with finite label set 1 Λ:
⎧ ⎫
⎪ Category of ⎪
⎪
⎪ ⎪
⎪
⎪ (extended) closed ⎪ ⎪
⎪
⎪
⎪ ⎪ )
⎨oriented surfaces ⎪ ⎬ Category of finite
Z: with Λ-labeled → dimensional vector
⎪
⎪ ⎪
⎪
⎪
⎪ marked points with ⎪
⎪ spaces over C
⎪
⎪ ⎪
⎪
⎪
⎩ projective tangent ⎪
⎭
vectors
The three-dimensional part of the TQFT is an association of a vector
Z(M, L, λ) ∈ Z(∂M, ∂L, ∂λ)
to any compact oriented framed three-manifold M together with an oriented
framed link (L, ∂L) ⊆ (M, ∂M ) and a Λ-labeling λ : π0 (L) → Λ:
λ2
λ2
λ2
λ1
This association has to satisfy the Atiyah–Segal–Witten TQFT axioms (see, e.g.,
Witten 1989, Atiyah 1990, Segal 1992 and for a more comprehensive presentation
of the axioms see Turaev 1994).
1 The set is also equipped with an involution † and a trivial element which is preserved by
the involution.
178 Toeplitz operators and Hitchin’s projectively flat connection
λ
(d)
λ0
For the curve operators, we can derive an explicit formula using factorization:
Let Σ be the surface obtained from cutting Σ along γ and identifying the
two boundary components to two points, say {p+ , p− }. Here p+ is the point
(n)
corresponding to the “left” side of γ. For any label µ ∈ Λk we get a labeling of
the ordered points (p+ , p− ) by the ordered pair of labels (µ, µ† ).
Introduction 179
9
Zk (Σ, p, λ0 ) ∼ Z (k) (Σ , p+ , p− , p, µ, µ† , λ0 ).
(n) (d) (d)
= (10.3)
(n)
µ∈Λk
(d)
which also induces a direct sum decomposition of End(Z (k) (Σ, p, λ0 )) which is
independent of the choice of the base point.
The TQFT axioms imply that the curve operator Z (k) (γ, λ) is diagonal with
respect to this direct sum decomposition along γ. One has the formula
9
Z (k) (γ, λ) = Sλ,µ (S0,µ )−1 IdZ (k) (Σ ,p ,p ,p,µ,µ† ,λ(d) ) .
+ − 0
(n)
µ∈Λk
(n)
Here Sλ,µ is the S-matrix 2 of the theory Zk (see, e.g., Blanchet 2000 for a
derivation of this formula).
In this chapter we review some of the results we have established regarding
these representations of the mapping class group and the curve operators. For
that we need the geometric construction of these theories as discussed below. We
will here discuss the applications to TQFT of the general program of combining
the theory of Toeplitz operators with the Hitchin connection. In this introduction
we will cover the applications. The general setting will be discussed in the rest
of this chapter.
(n)
10.1.1 Geometric construction of Zk
The geometric construction of these TQFTs was proposed by Witten (1989),
where he derived via the Hamiltonian approach to quantum Chern–Simons
theory, that the geometric quantization of the moduli spaces of flat connec-
tions should give the two-dimensional part of the theory. Further, he pro-
posed an alternative construction of the two-dimensional part of the theory
2 The S-matrix is determined by the isomorphism a modular functor induces from two
different way of glueing an annulus to obtain a torus. For its definition see for example Moore
and Seiberg (1989), Walker (1991), Segal (1992), or Bakalov and Kirillov (2000) and references
therein. It is also discussed in Andersen and Ueno (2006).
180 Toeplitz operators and Hitchin’s projectively flat connection
via WZW-conformal field theory. This theory has been studied intensively.
In particular the work of Tsuchiya, Ueno, and Yamada (1989) provided the
major geometric constructions and results needed. In Bakalov and Kirillov
(2000) their results was used to show that the category of integrable highest
modules of level k for the affine Lie algebra associated to any simple Lie
algebra is a modular tensor category. Further in Bakalov and Kirillov (2000)
this result is combined with the works of Kazhdan and Lusztig (1993, 1994a,
1994b) and of Finkelberg (1996) to argue that this category is isomorphic to
the modular tensor category associated to the corresponding quantum group,
from which Reshetikhin and Turaev constructed their TQFT. Unfortunately,
these results do not allow one to conclude the validity of the geometric con-
structions of the two-dimensional part of the TQFT proposed by Witten.
However, in joint work with Ueno (Andersen and Ueno, 2006, 2007a, 2007b,
2008), we have given a proof, based mainly on the results of Tsuchiya, Ueno
and Yamada (1989), that the TUY construction of the WZW-conformal field
theory after twist by a fractional power of an abelian theory satisfies all the
axioms of a modular functor. Furthermore, we have proved that the full (2 + 1)-
dimensional TQFT that results from this is isomorphic to the one constructed
by BHMV via skein theory mentioned above. Combining this with the theo-
rem of Laszlo (Laszlo 1998), which identifies (projectively) the representations
of the mapping class groups one obtains from the geometric quantization of
the moduli space of flat connections with the ones obtained from the TUY
constructions, one gets a proof of the validity of the construction proposed
by Witten (1989).
Let us now briefly recall the geometric construction of the representations
(n,d)
Zk of the mapping class group, as proposed by Witten, using geometric
quantization of moduli spaces.
We assume from now on that the genus of the closed oriented surface Σ is
at least 2. Let M be the moduli space of flat SU (n) connections on Σ − p
with holonomy around p equal to exp(2πid/n) Id ∈ SU (n). In the case (n, d)
are coprime, the moduli space is smooth. In all cases the smooth part of the
moduli space has a natural symplectic structure ω. There is a natural smooth
symplectic action of the mapping class group Γ of Σ on M . Moreover there is a
unique prequantum line bundle (L, ∇, (·, ·)) over (M, ω). The Teichmüller space
T of complex structures on Σ parametrizes naturally and Γ-equivariantly Kähler
structures on (M, ω). For σ ∈ T , we denote (M, ω) with its corresponding Kähler
structure by Mσ .
By applying geometric quantization to the moduli space M , one gets a certain
finite rank bundle over Teichmüller space T , which we will call the Verlinde
bundle Vk at level k, where k is any positive integer. The fiber of this bundle
over a point σ ∈ T is Vk,σ = H 0 (Mσ , Lk ). We observe that there is a natural
Hermitian structure ·, · on H 0 (Mσ , Lk ) by restricting the L2 inner product on
global L2 sections of Lk to H 0 (Mσ , Lk ).
The main result pertaining to this bundle is
Introduction 181
Theorem 10.1 (Axelrod, Della Pietra, and Witten 1991; Hitchin 1990)
The projectivization of the bundle Vk supports a natural flat Γ-invariant connec-
ˆ
tion ∇.
obtained from the action of the mapping class group on the covariant constant
sections of P(Vk ) over T .
Theorem 10.2 Assume that n and d are coprime or that (n, d) = (2, 0) when
g = 2. Then we have that
∞
: {1, H} g = 2, n = 2 and d = 0
(n,d)
ker Zk =
k=1
{1} otherwise.
Large parts of the proof of this theorem, which we gave in Andersen (2006a),
apply to the general setting discussed in the following sections. In particular,
we get the general asymptotic faithfulness Theorem 10.19 discussed in the last
section of this chapter. Theorem 10.2 follows directly from Theorem 10.19 as
argued in the end of section 6 of Andersen (2006a). See also Andersen (2008a)
and Andersen and Christ (2008) for the singular case.
The main ingredient behind this is the theory of Toeplitz operators associated
to smooth functions on M . For each f ∈ C ∞ (M (d) ) and each point σ ∈ T , we
have the Toeplitz operator
(k)
Tf,σ : H 0 Mσ , Lkσ → H 0 Mσ , Lkσ
182 Toeplitz operators and Hitchin’s projectively flat connection
which is given by
(k)
Tf,σ = πσ(k) (f s)
(k)
for all s ∈ H 0 (Mσ , Lkσ ), where πσ is the orthogonal projection onto H 0 (Mσ , Lkσ )
induced from the L2 inner product on C ∞ (M, Lk ). We get a smooth section of
End(V (k) ) over T
∈ C ∞ (T , End(V (k) ))
(k)
Tf
(k) (k)
by letting Tf (σ) = Tf,σ (see Andersen 2006a). See Section 10.3 for further
discussion of the Toeplitz operators and their connection to deformation quan-
tization.
(k)
The sections Tf of End(V (k) ) over T are not covariant constant with respect
to Hitchin’s connection ∇ˆ e . However, they are asymptotic as k goes to infinity.
The precise meaning of this is stated in Theorem 10.18 of Section 10.5, where
it is proved in the general setting. Theorem 10.19 then follows in a rather
straightforward way from this asymptotic covariant constantness of the Toeplitz
operators and Theorem 10.3 due to Bordeman, Meinrenken, and Schlichenmaier
(1994).
Theorem 10.3 (Bordeman, Meinrenken, and Schlichenmaier 1994)
For any f ∈ C ∞ (M ) we have that
; ;
; (k) ;
lim ;Tf,σ ; = sup |f (x)|.
k→∞ x∈M
(k)
Since the association of the sequence of Toeplitz operators Tf,σ , k ∈ Z+ is
linear in f , we see from this theorem that this association is faithful.
for all g ∈ G.
The Hilbert space representation of the mapping class group we consider is
the following.
Let
(2,1)
Rk = End0 (Zk (Σ)),
where End0 means the traceless endomorphisms. Since the TQFTs are unitary,
there is a natural unitary structure which is mapping class group invariant on
Rk . Now define
∞
9
R̃ = Rk
k+2 prime
is the one associated to the Hermitian structure on Vk∗ ⊗ Vk induced from ·, ·
on Vk .
184 Toeplitz operators and Hitchin’s projectively flat connection
In fact the proof of this theorem is valid in the general setting discussed in the
following sections. However, the techniques used in Andersen (2007) go beyond
the techniques reviewed in this chapter and we therefore refer to Andersen (2007)
for the proof of this theorem.
To produce the almost fixed vectors, we now pick a finite subgroup Λ of SU (2)
which contains −1 ∈ SU (2) and we consider the finite subset X of M , consisting
of connections which reduce to Λ. For the dihedral group Λ, we get a non-empty
finite subset X of M this way which is invariant under the action of the mapping
(k)
class group. Let now EX be the section of End(Vk ) given by
(k)
EX = Ex(k) .
x∈X
(k) (k)
Let EX,0 be the traceless part of EX . As we have argued in section 7 of Andersen
(k)
(2007), for large enough k, EX = 0. Hence for large enough k we have a unique
(k)
(up to unit scale) unit vector in EX,0 ∈ Rk , which at σ0 agrees projectively with
(k)
EX,0 (σ0 ).
(k)
Theorem 10.7 The sequence EX,0 is an almost fixed vector for the action
of Γ on R.
Let us here give the main idea behind the proof of Theorem 10.8 and refer
to Andersen (2008b) for details. One considers the explicit expression for the
S-matrix, as given in formula (13.8.9) in Kac (1995):
µ̌+ρ̌
Sλ,µ /S0,µ = λ(e−2πi k+n ), (10.4)
where ρ is half of the sum of the positive roots and ν̌ (ν any element of Λ) is
the unique element of the Cartan subalgebra of the Lie algebra of SU (n) which
is dual to ν with respect to the Cartan–Killing form (·, ·).
From (10.4) one sees that under the isomorphism µ̌ → µ, Sλ,µ /S0,µ makes
sense for any µ̌ in the Cartan subalgebra of the Lie algebra of SU (n). Furthermore
one finds that the values of this sequence of functions (depending on k) are
asymptotic to the values of the holonomy function hγ,λ at the level k Bohr–
Sommerfeld sets of the limiting non-negative polarizations discussed above (see
Andersen 1998). From this one can deduce Theorem 10.8 (see again Andersen
2008b for details). Please see Andersen (2005) for the corresponding result in the
abelian case.
2. The mapping class φ is not finite order, but it is reducible, meaning there
exists a simple closed curve on the surface, whose non-trivial homotopy
class is preserved by some power of φ.
3. The mapping class φ is pseudo-Anosov, meaning that there exists ζ > 1,
two transverse measured foliations F s and F u on Σ and a diffeomorphism
f of Σ, which represents φ, such that
f∗ (F s ) = ζ −1 F s and f∗ (F u ) = ζF u .
In the pseudo-Anosov case, ζ is uniquely determined by φ and it is called the
streching factor for φ.
In the reducible case one continues the analysis of φ, by cutting Σ along
the preserved simple closed curve, to get a mapping class of a surface with
boundary. This mapping class is then classified in terms of the Nielsen–Thurston
classification of mapping classes of surfaces with boundary. The upshot of this is
that there is a diffeomorphism f of Σ, which represents φ and which preserves a
system of simple closed curves, and when one cuts the surface along these curves,
f induces a diffeomorphism of the resulting cut surface. For each component of
the cut surface, there is a smallest power of f , which preserves the component,
and this power of f is a diffeomorphism of the component, which is either finite
order or pseudo-Anosov (see Fathi, Laudenbach, and Poénaru 1979 for further
details regarding this).
The asymptotic faithfulness property gives us immediately the following the-
orem:
Theorem 10.10 For any mapping class φ ∈ Γ we have that there exists an
integer M such that
M
(n,d)
Zk (φ) ∈ C Id
for all k if and only if φM = 1 (or φ2M = 1, in case (n, d) = (2, 0)).
This separates the finite-order ones from the rest. In order to separate the
reducibles from the pseudo-Anosov ones, we consider the curve operators. Sup-
pose that γ is a simple closed curve on the surface Σ. When we choose an
orientation on γ, we have the holonomy function hγ, . Note that hγ, only
depends on the free homotopy class of γ. Further hγ, is constant if γ is
nul-homotopic.
Theorem 10.11 For any mapping class φ ∈ Γ and any homotopy class γ of a
simple closed curve on Σ we have that φ is reducible along γ, that is,
φ(γ) = γ
if and only if
(n,d) (n,d)
[Zk (φ), Zk (γ, )] = 0
Introduction 187
for infinitely many k and any fixed pair (n, d) (and any choice of orientation of
γ if n > 2).
Using these two conditions, we see immediately how to determine the Nielsen–
Thurston classification of a mapping class using TQFT:
(φ−1 ) = Zk
(n,d) (n,d) (n,d) (n,d)
Zk (φ)Zk (γ, )Zk (φ(γ), ). (10.5)
Conversely, suppose that
(n,d) (n,d)
Zk (φ), Zk (γ, ) = 0.
Then
(n,d) (n,d)
Zk (γ, ) = Zk (φ(γ), )
by (10.5). From Theorem 10.8 we conclude that
; ;
; (k) ;
lim ;Thγ, − Zk (φ)Thγ, Zk (φ−1 ); = 0.
(n,d) (k) (n,d)
k→∞
But then from Proposition 10.5 below and Theorem 10.3, we get that hγ, =
hφ(γ), . We now get that φ(γ) = γ by proposition 1 in Andersen (2006c). The
main idea behind proposition 1 from Andersen (2006c) is that the holonomy
functions extend to holomorphic functions on the SL(n, C)-moduli space M.
The restriction of the holonomy functions from M to the real slice M ⊂ M is
injective. Moreover if two holonomy functions for two simple closed curves agree
on M for some d, then they also agree on M for d = 0. But this space contains a
copy of the Teichmüller space of Σ where it is clear that the holonomy function
of a simple closed curve determines the curve up to homotopy on Σ. We refer to
Andersen (2006c) for the full details of this argument.
We consider it an interesting problem to provide a TQFT formula for the
streching factor ζ of any pseudo-Anosov mapping class (see Andersen, Masbaum,
and Ueno 2006 for the first steps in this direction). We expect that the asymptotic
(n,d)
expansions of the operators Zk (γ, ) should be related to work of Andersen,
Mattes and Reshetikhin (1996, 1998)
188 Toeplitz operators and Hitchin’s projectively flat connection
than the gauge theory setting in which Hitchin (1990) constructed his original
connection. But when applied to the gauge theory situation we get the corollary
that Hitchin’s connection agrees with Axelrod, Della Pietra, and Witten’s.
Hence, we start in the general setting and let (M, ω) be any compact symplectic
manifold.
Definition 10.3 A prequantum line bundle (L, (·, ·), ∇) over the symplectic
manifold (M, ω) consists of a complex line bundle L with a Hermitian structure
(·, ·) and a compatible connection ∇ whose curvature is
i
F∇ = ω.
2π
We say that the symplectic manifold (M, ω) is prequantizable if there exists a
prequantum line bundle over it.
The condition on the curvature is to be interpreted in terms of the induced
principal U (1)-connection in the circle bundle of L in the above definition. In
terms of covariant derivatives, this means
ω(X, Y ) = [∇X , ∇Y ] − ∇[X,Y ] .
Recall that the condition for the existence of a prequantum line bundle
is that [ω] ∈ Im(H 2 (M, Z) → H 2 (M, R)) and that the inequivalent choices of
prequantum line bundles (if they exist) are parametriced by H 1 (M, R) (see e.g.
Woodhouse 1992).
We shall assume that (M, ω) is prequantizable and fix a prequantum line
bundle (L, (·, ·), ∇) over (M, ω).
Assume that T is a smooth manifold which smoothly parametrizes
Kähler structures on (M, ω). This means we have a smooth 3 map I :
T → C ∞ (M, End(T M )) such that (M, ω, Iσ ) is a Kähler manifold for each σ ∈ T .
We will use the notation Mσ for the complex manifold (M, Iσ ). For each σ ∈ T
we use Iσ to split the complexified tangent bundle T MC into the holomorphic
and the anti-holomorphic parts, which we denote
Tσ = E(Iσ , i) = Im(Id −iIσ )
and
T̄σ = E(Iσ , −i) = Im(Id +iIσ ),
respectively.
3 Here a smooth map from T to C ∞ (M, W ) for any smooth vector bundle W over M means a
smooth section of πM ∗ (W ) over T × M , where π
M is the projection onto M . Likewise a smooth
p-form on T with values in C ∞ (M, W ) is by definition a smooth section of πT∗ Λp (T ) ⊗ πM∗ (W )
over T × M . We will also encounter the situation where we have a bundle W̃ over T × M and
then we will talk about a smooth p-form on T with values in C ∞ (M, W̃σ ) and mean a smooth
section of πT∗ Λp (T ) ⊗ W̃ over T × M .
190 Toeplitz operators and Hitchin’s projectively flat connection
ˆV = ∇
∇ ˆ tV − u(V )
i
V [I]∇1,0 s + ∇0,1 u(V )s = 0 (10.6)
2
for all vector fields V on T and all smooth sections s of H (k) .
This result is not surprising (see Andersen 2006b for a proof this lemma).
(k)
Observe that if this condition holds, we can conclude that the subspaces Hσ ⊂
C ∞ (M, Lk ), for all σ ∈ T , form a subbundle H (k) of H(k) .
We observe that
V [I]∇1,0 s = 0,
so u(V ) = 0 solves (10.6) along the anti-holomorphic directions on T . In other
words the (0, 1) part of the trivial connection ∇ ˆ t induces a ∂¯ operator on H (k)
and hence makes it a holomorphic vector bundle over T .
This is of course not in general the situation in the (1, 0) direction. Let us now
consider a particular u and prove that it solves (10.6) under certain conditions.
On the Kähler manifold (Mσ , ω) we have the Kähler metric and we have
the Levi-Civita connection ∇ in Tσ . We also have the Ricci potential Fσ ∈
C0∞ (M, R). Here
< =
C0∞ (M, R) = f ∈ C ∞ (M, R) | f ωm = 0
M
and the Ricci potential is the element of Fσ ∈ C0∞ (M, R) which satisfies
1 ¯
Ricσ = RicH
σ + i∂σ ∂σ Fσ ,
2
where Ricσ ∈ Ω1,1 (Mσ ) is the Ricci form and RicHσ is its harmonic part. We see
that we get this way a smooth function F : T → C0∞ (M, R).
For any G ∈ C ∞ (M, S 2 (Tσ )) we get a linear bundle map
G : Tσ∗ → Tσ
192 Toeplitz operators and Hitchin’s projectively flat connection
−ik Tr −2G∂σ F ω + ∇1,0
σ (G)ω s,
(M, ω) is n[ω] ∈ H 2 (M, Z) is then just replaced by the vanishing of the second
Stiefel–Whitney class of M (see Andersen, Gammelgaard, and Lauridsen 2008
for more details).
In Andersen (2006b) we prove the following lemma:
Lemma 10.5 There exist smooth one-forms Xr , Z and functions Yr , r =
1, . . . , R, on T with values in C ∞ (M, Tσ ) such that
1 R
∆G(V ) − ∇G(V )dF = ∇Xr (V ) ∇Yr + ∇Z(V ) (10.10)
2 r=1
as was done by Boutet de Monvel and Guillemin (1981) (in fact in a much
more general situation than the one we consider here) and others following
them. In particular the applications developed by Schlichenmaier and further by
Karabegov and Schlichenmaier to the study of Toeplitz operators in the geomet-
ric quantization setting is what will interest us here. Let us first describe the basic
setting.
For each f ∈ C ∞ (M ) we consider the prequantum operator, namely, the
differential operator Mf : C ∞ (M, Lk ) → C ∞ (M, Lk ) given by
(k)
(k)
Mf (s) = f s
(k)
Tf,σ (s) = πσ(k) (f s)
(k)
for any element s in Hσ and any point σ ∈ T . We observe that the Toeplitz
(k)
operators are smooth sections Tf of the bundle Hom(H(k) , H (k) ) and restrict
to smooth sections of End(H (k) ).
196 Toeplitz operators and Hitchin’s projectively flat connection
X(s1 , s2 ) = (∇X s1 , s2 ).
Now, calculating the Lie derivative along X of (s1 , s2 )ω m and using the above,
one obtains after integration that
Computing the Lie derivative along X̄ of (s1 , s2 )ω m and integrating, we get that
Then
−1 −1
σ g = (f ◦ φ
f ˜BT σ (g ◦ φ
) BT ) ◦φ
for all f, g ∈ Ch∞ (M ), where φ(h) = h
2+nh .
for all smooth section f of Ch and all smooth vector fields on T . Moreover
D̃ = 0 mod h.
Here ∼ means the following: For all L ∈ Z+ we have that
; ;
; L ;
; ˆ e (k) 1 ;
; = O(k −(L+1) )
(k) (k)
;∇V Tf − TV [f ] + T (l)
; D̃V f (2k + n)l ;
l=1
The proof of this theorem is given in Andersen (2006b). Let us recall formula
(19) from Andersen (2006b):
D̃(V )(f ) = V [F ]f − V [F ]˜BT f + h(E(V )(f ) − H(V )˜
BT f ) (10.19)
200 Toeplitz operators and Hitchin’s projectively flat connection
and H is the one-form on T with values in C ∞ (M ) such that H(V ) = E(V )(1).
Lemma 10.7 σ for each σ ∈ T ,
The formal operator DV is a derivation for BT
that is,
DV (f ˜BT g) = DV (f )˜BT g + f ˜
BT DV (g)
for all f, g ∈ C ∞ (M ).
Again this lemma is proved in Andersen (2006b), and in fact it follows basically
from
ˆ eV T (k) Tg(k) = ∇
∇ ˆ eV T (k) Tg(k) + T (k) ∇
ˆ eV Tg(k) .
f f f
Theorem 10.17 Assume that the formal Hitchin connection D is flat and
HΓ1 (T , D(M )) = 0,
Theorem 10.18 Let σ0 and σ1 be two points in T and Pσ0 ,σ1 be the parallel
ˆ e ) from σ0 to σ1 with respect to ∇
transport in the bundle (End(V (k) ), ∇ ˆ e . Then
where · is the operator norm on H 0 Mσ1 , Lkσ1 .
Let
d 2 @ˆ A @ A
ˆ σ s .
E(s) = |s|F − ∇σt s, s − s, ∇
dt F t
F
Recall that ∇ˆv = ∇ˆ tv − u(v) and u(v) = 1 o(v) + v [F ], hence we have for all
2k+n
sections s of Vk that
1 −F
E(s) = πe o (σt ) s, s + s, πe−F o (σt ) s .
2k + n
Hence by combining Theorem 10.3, (10.10), (10.14), and (10.15) we have proved
that
Proposition 10.4 The Hermitian structure (10.20) is asymptotically flat with
ˆ that is, for any compact subset K of T , there exists
respect to the connections ∇,
a constant C such that for all sections s of Vk over K, we have that
C
|E(s)| ≤ |s|2
k+n F
over K.
We note that this proposition implies the same proposition for sections of
End(Vk ) with respect to the induced Hermitian structure on End(Vk ) = Vk∗ ⊗ Vk ,
which we also denote ·, ·F . We denote the analogous quantity to E for the
endomorphism bundle by Ee .
Lemma 10.8 The Hermitian structure on Hk
1
s1 , s2 F = (s1 , s2 )e−F ω m
m! M
and the constant L2 -Hermitian structure on Hk
1
s1 , s2 = (s1 , s2 )ω m
m! M
are equivalent uniformly in k when restricted to Vk over any compact subset K
of T .
The proof of Lemma 10.8 can be found in section 5 of Andersen (2006a).
Proof. (Of Theorem 10.12). Let σt , t ∈ J, be a smooth one-parameter family
of complex structures such that σt is a curve in T between the two points in
question. By Lemma 10.8 the Hermitian structures ·, · and ·, ·F on Vk are
equivalent uniformly in k over compact subsets of T . It follows that there exists
Asymptotic flatness of Toeplitz operators 203
constants C1 and C2 , such that we have the following inequalities over the image
of σt in T for the operator norm · and the norm | · |F on End(Vk ):
B
· ≤ C1 | · |F ≤ C2 Pg,n (k) · , (10.21)
where Pg,n (k) is the rank of Vk given by the Verlinde formula. By the Riemann–
Roch theorem this is a polynomial in k of degree m.
Because of these inequalities, we choose an integer r bigger than m/2 and
let fr = P (r) (f ) k1 where P is a trivialization of the formal Hitchin connection
along σt , t ∈ J, which clearly exists and is unique under the condition that
Pσ = Id by Proposition 10.1. Then let nk : J →[0, ∞) be given by
nk (t) = |Θk (t)|2F
where
Θk (t) : Vk,σt → Vk,σt
is given by
(k) (k)
Θk (t) = Pσ0 ,σt T(fr )0 ,σ0 − T(fr )t ,σt .
Consequently we can apply Propositions 10.3 and 10.4 to obtain that there exists
a constant C such that
dnk C 1/2
≤ (nk + nk ).
dt k
This estimate implies that
2
Ct
nk (t) ≤ exp −1 .
2k
But by (10.21) we get that
; ;
; (k) (k) ;
;Pσ0 ,σ1 T(fr )0 ,σ0 − T(fr )1 ,σ1 ; = Θk (1) ≤ C1 nk (1)1/2 .
204 Toeplitz operators and Hitchin’s projectively flat connection
The theorem then follows from these two estimates, since (fr )0 = f and
; ;
; (k) (k) ;
;T(fr )1 ,σ1 − Tf,σ1 ; = O(k−1 ).
ρk : Γ → Aut(P(Vk )),
φ∗ Pφ(σ),σ
H 0 Mσ , Lkσ −−−−→ H 0 Mφ(σ) , Lkφ(σ) −−−−−→ H 0 Mσ , Lkσ ,
where Pφ(σ),σ : H 0 Mφ(σ) , Lkφ(σ) → H 0 Mσ , Lkσ on the horizontal arrows
refers to parallel transport in the bundle V (k) , whereas Pφ(σ),σ refers to the
parallel transport in the endomorphism bundle End(V (k) ) in the last vertical
arrow. By the definition of ρk , we see that
f ◦φ=f
if and only if
f ◦ φ = f.
; ;;;
;
(k) ;
≤ ;ρk (φ−1 ); ; ρk (φ), Tf,σ ; .
206 Toeplitz operators and Hitchin’s projectively flat connection
Lemma 10.8 and Proposition 10.4 give a uniform bound on ρk (φ−1 ). Hence
Proposition 10.5 implies that
; ;
; (k) ;
lim ;T(f −f ◦φ−1 ),σ ; = 0.
k→∞
Acknowledgement
I would like to take this opportunity to thank Professor Nigel Hitchin for
introducing me to the subject at hand and further for the many very helpful
discussions we have had over the years regarding this interesting and facinating
research field of mathematics bordering on Mathematical Physics.
References
Andersen, J. E. (1998). New polarizations on the moduli space and the Thurston
compactification of Teichmuller space. Int. J. Math. 9(1), 1–45.
Andersen, J. E. (2005). Geometric quantization and deformation quantization of abelian
moduli spaces. Commun. Math. Phys. 2005, 727–45.
Andersen, J. E. (2006a). Asymptotic faithfulness of the quantum SU (n) representations
of the mapping class groups. Ann. Math. 163, 347–68.
Andersen, J. E. (2006b). Hitchin’s connection, Toeplitz operators and symmetry invari-
ant deformation quantization. math.DG/0611126.
Andersen, J. E. (2006c). The Nielsen-Thurston classification of mapping classes is
determined by TQFT. math.QA/0605036.
Andersen, J. E. (2007). Mapping class groups do not have Kazhdan’s property (T),
math.QA/0706.2184.
Andersen, J. E. (2008a). Asymptotic faithfulness of the quantum SU (n) representations
of the mapping class groups in the singular case. In preparation.
Andersen, J. E. (2008b). Asymptotic in teichmuller space of the Hitchin connection. In
preparation.
Andersen, J. E. and Christ, M. (2008). Asymptotic expansion of the Szegö kernel on
singular algebraic varieties. In preparation.
Andersen, J. E., Gammelgaard, N. L., and Lauritsen, M. (2008). Hitchin’s connection
in half-form quantization. math.DG/0711.3995.
Andersen, J. E., Masbaum, G., and Ueno, K. (2006). Topological quantum field theory
and the Nielsen-Thurston classification of M (0, 4). Math. Proc. Camb. Philos. Soc.
141(3), 477–88.
References 207
Andersen, J. E., Mattes, J., and Reshetikhin, N. (1996). The Poisson Structure
on the moduli space of flat connections and chord diagrams. Topology 35,
1069–83.
Andersen, J. E., Mattes, J., and Reshetikhin, N. (1998). Quantization of the algebra
of chord diagrams. Math. Proc. Camb. Phil. Soc. 124, 451–67.
Andersen, J. E. and Ueno, K. (2006). Modular functors are determined by their genus
zero data. math.QA/0611087.
Andersen, J. E. and Ueno, K. (2007a). Abelian conformal field theories and determinant
bundles. Int. J. Math. 18, 919–93.
Andersen, J. E. and Ueno, K. (2007b). Constructing modular functors from conformal
field theories. J. Knot Theory Ramifications 16(2), 127–202.
Andersen, J. E. and Ueno, K. (2008). Construction of the Reshetikhin-Turaev TQFT
from conformal field theory. In preparation.
Atiyah, M. (1990). The Jones-Witten invariants of knots. Astérisque (189–190), Exp.
No. 715, 7–16. Séminaire Bourbaki, Vol. 1989/90.
Axelrod, S., Della Pietra, S., and Witten, E. (1991). Geometric quantization of Chern-
Simons gauge theory. J. Diff. Geom. 33(3), 787–902.
Bakalov, B. and Kirillov, A. (2000). Lectures on Tensor Categories and Modular
Functors, Vol. 21. AMS University Lecture Series, Providence, RI.
Bekka, B., de la Harpe, P., and Valette, A. (2007). Kazhdan’s Proporty (T). Cambridge
University Press. In press.
Blanchet, C. (2000). Hecke algebras, modular categories and 3-manifolds quantum
invariants. Topology 39(1), 193–223.
Blanchet, C., Habegger, N., Masbaum, G., and Vogel, P. (1992). Three-manifold
invariants derived from the Kauffman bracket. Topology 31, 685–99.
Blanchet, C., Habegger, N., Masbaum, G., and Vogel, P. (1995). Topological quantum
field theories derived from the Kauffman bracket. Topology 34, 883–927.
Bleiler, S. and Casson, A. (1988). Automorphisms of Sufaces after Nielsen and
Thurston. Cambridge University Press, Cambridge, UK.
Bordeman, M., Meinrenken, E., and Schlichenmaier, M. (1994). Toeplitz quantiza-
tion of Kähler manifolds and gl(N ), N → ∞ limit. Commun. Math. Phys. 165,
281–96.
Boutet de Monvel, L. and Guillemin, V. (1981). The Spectral Theory of Toeplitz
Operators, Vol. 99 of Ann. Math. Stud. Princeton University Press, Princeton, NJ.
Boutet de Monvel, L. and Sjöstrand, J. (1976). Sur la singularité des noyaux de
Bergmann et de Szegö. Asterique 34–35, 123–64.
Faltings, G. (1993). Stable G-bundles and projective connections. J. Alg. Geom. 2,
507–68.
Fathi, A., Laudenbach, F., and Poénaru, V. (1991/1979). Travaux de Thurston sur les
surfaces, Astérisque 66–67.
Finkelberg, M. (1996). An equivalence of fusion categories. Geom. Funct. Anal. 6,
249–67.
208 Toeplitz operators and Hitchin’s projectively flat connection
Freedman, M. H., Walker, K., and Wang, Z. (2002). Quantum SU(2) faithfully detects
mapping class groups modulo center. Geom. Topol. 6, 523–39.
Hitchin, N. (1990). Flat connections and geometric quantization. Commun. Math. Phys.
131, 347–80.
Kac, V. G. (1995). Infinite Dimensional Lie Algebras, 3 edn. Cambridge University
Press, Cambridge, UK.
Karabegov, A. V. and Schlichenmaier, M. (2001). Identification of Berezin–Toeplitz
deformation quantization. J. Reine Angew. Math. 540, 49–76.
Kazhdan, D. (1967). Connection of the dual space of a group with the structure of its
closed subgroups. Funct. Anal. Appl. 1, 64–65.
Kazhdan, D. and Lusztig, G. (1993). Tensor structures arising from affine Lie algebras.
I, II. J. Am. Math. Soc. 6(4), 905–47, 949–1011.
Kazhdan, D. and Lusztig, G. (1994a). Tensor structures arising from affine Lie algebras.
III. J. Am. Math. Soc. 7(2), 335–81.
Kazhdan, D. and Lusztig, G. (1994b). Tensor structures arising from affine Lie algebras.
IV. J. Am. Math. Soc. 7(2), 383–453.
Laszlo, Y. (1998). Hitchin’s and WZW connections are the same. J. Diff. Geom. 49(3),
547–76.
Moore, G. and Seiberg, N. (1989). Classical and quantum conformal field theory.
Commun. Math. Phys. 123, 177–254.
Reshetikhin, N. and Turaev, V. (1990). Ribbon graphs and their invariants derived fron
quantum groups. Commun. Math. Phys. 127, 1–26.
Reshetikhin, N. and Turaev, V. (1991). Invariants of 3-manifolds via link polynomials
and quantum groups. Invent. Math. 103, 547–97.
Roberts, J. (2001). Irreducibility of some quantum representations of mapping class
groups. J. Knot Theory Ramifications 10, 763–7.
Schlichenmaier, M. (1998). Berezin-Toeplitz quantization and conformal field theory.
PhD thesis.
Schlichenmaier, M. (2000). Deformation quantization of compact Kähler manifolds by
Berezin-Toeplitz quantization. In Conférence Moshé Flato 1999, Vol. II (Dijon), Vol.
22 of Math. Phys. Stud. pp. 289–306. Kluwer Academic Publication, Dordrecht, the
Netherlands.
Schlichenmaier, M. (2001). Berezin-Toeplitz quantization and Berezin transform. In
Long Time Behaviour of Classical and Quantum Systems (Bologna, 1999), Vol. 1 of
Ser. Concr. Appl. Math., pp. 271–87. World Scientific Publication, River Edge, NJ.
Segal, G. (1992). The Definition of Conformal Field Theory. Oxford University
Preprint.
Thurston, W. (1988). On the geometry and dynamics of diffeomorphisms of surfaces.
Bull. Am. Math. Soc. 19, 417–31.
Tsuchiya, A., Ueno, K., and Yamada, Y. (1989). Conformal field theory on uni-
versal family of stable curves with gauge symmetries. Adv. Stud. Pure Math. 19,
459–566.
References 209
Philip Boalch
Dedicated to Nigel Hitchin for his 60th birthday
11.1 Introduction
The main theme of this chapter is ‘icosahedral’ solutions of (ordinary) differential
equations, a topic that seems suitable for a 60th birthday conference. We will
however try to go beyond the icosahedron, to see what comes next, and consider
various symmetry groups each of which could be thought of as the next in a
sequence, following the icosahedral group.
To fix ideas let us give a classical example. Recall the icosahedral rotation
group of order 60:
= PSL2 (F5 ) ∼
A5 ∼ = ∆235 ∼
= a, b, c a2 = b3 = c5 = abc = 1 .
This is described via three generators a, b, and c whose product is the identity,
and so it is natural to look for ordinary differential equations (ODEs) on the
three-punctured sphere P1 (C) \ {0, 1, ∞} with monodromy group A5 . Now A5
is a three-dimensional rotation group so naturally lives in SO3 (R) which is a
subgroup of SO3 (C) which is isomorphic to PSL2 (C). Thus we are led to search
for connections
A1 A2
∇=d− + dz, Ai ∈ sl2 (C) (11.1)
z z−1
Note added in proof: Lisovyy and Tykhyy have recently announced (arXiv:0809.4873) that
the ‘Non-linear Schwarz’s list’ constructed here is in fact complete.
Introduction 211
Inhalt
No. λ µ ν Polyeder
π
1 1
aa∗ I 2 2
ν ν Regelmässige Doppelpyramide
1 1 1 1
abb II 2 3 3 6
=A Tetraeder
2 1 1 1
bbb III 3 3 3 3
= 2A
1 1 1 1
abg IV 2 3 4 12
=B Würfel und Oktaeder
2 1 1 1
bgg V 3 4 4 6
= 2B
1 1 1 1
abc VI 2 3 5 30
= C Dodekaeder und Ikosaeder
2 1 1 1
bbd VII 5 3 3 15
= 2C
2 1 1 1
bcc VIII 3 5 5 15
= 2C
1 2 1 1
acd IX 2 5 5 10
= 3C
3 1 1 2
bcd X 5 3 5 15
= 4C
2 2 2 1
ddd XI 5 5 5 5
= 6C
2 1 1 1
bbc XII 3 3 5 5
= 6C
4 1 1 1
ccc XIII 5 5 5 5
= 6C
1 2 1 7
abd XIV 2 5 3 30
= 7C
3 2 1 1
bdd XV 5 5 3 3
= 10C
A key point here is that the Gauss hypergeometric equation is rigid so the full
monodromy representation (of the fundamental group of the three-punctured
sphere into PSL2 (C)) is determined by the conjugacy classes of the monodromy
around each of the punctures. Thus in Schwarz’s list it is sufficient to list these
local monodromy conjugacy classes in order to specify the possible monodromy
representations (and from this it is easy to find a hypergeometric equation with
given monodromy). To ease recognition, to the left of the table we have listed the
triples of conjugacy classes which occur, labelling the four non-trivial conjugacy
classes of A5 by a, b, c, and d, representing rotations by 12 , 13 , 15 , and 25 of a turn,
respectively. (In the octahedral case one may also have rotations by a quarter of
a turn, which we label by g.)
These questions can now be answered and lead to two ‘non-rigid Schwarz’s
lists’, that is, to classifications of possible monodromy representations with
finite image (up to equivalence) and the construction of connections realizing
such representations. We should emphasize that the main focus has been the
construction of such connections with given monodromy representation for any
value of t (which is a tricky business in this non-rigid case), rather than just the
classification.
Example (of type (B)) . The full symmetry group of the icosahedron is the
icosahedral reflection group of order 120:
D E
H = H3 ∼ = r1 , r2 , r3 ri2 = 1, (r1 r2 )2 = (r2 r3 )3 = (r3 r1 )5 = 1
⊂ O3 (R) ⊂ GL3 (C).
This is generated by three reflections (whose product is not the identity) and
so it is natural to look for connections on rank 3 bundles over a four-punctured
sphere with monodromy H (generated by three reflections about three of the
punctures – that is, connections of the form (B) with each of the three residues
Bi having trace 12 so the corresponding reflections are of order 2). There turn out
to be three inequivalent triples of generating reflections of H, two of which are
related by an outer automorphism. The problem is to write down connections
1 That is, arbitrary automorphisms of the form ‘one plus rank 1’, not necessarily of order 2
or orthogonal.
Introduction 213
of the desired form for any value t of the final pole position. One triple of
generating reflections is intimately related to K. Saito’s flat structure (1993)
for H (or icosahedral Frobenius manifold) and appears in Dubrovin’s article
(1995, appendix E). The other two triples were dealt with around 1997: see
Dubrovin and Mazzocco (2000); one is similar to the first case (since related to
it by an outer automorphism) but the final triple turned out to be much trickier,
and writing out the family of connections in this case involved a specific elliptic
curve which took about 10 pages of 40 digits integers to write down (see the
preprint version of op. cit. on the mathematics arxiv). We will eventually see
below that this elliptic solution is in fact equivalent to a solution with a simple
parametrization, agreeing with Hitchin’s philosophy that ‘nice problems should
have nice solutions’.
Remark. Before moving on to the third generalization let us add some other
historical comments. The ‘non-naive’ generalizations of the Gauss hypergeomet-
ric equation are the equations satisfied by the n Fn−1 hypergeometric functions
(the Gauss case being that of n = 2). The corresponding Schwarz’s list appears
in Beukers and Heckman (1989). In terms of connections this amounts to consid-
ering connections (11.1) on rank n vector bundles, still with three singularities
on P1 , but with A1 of rank n − 1 and A2 of rank 1; these connections are still
rigid.
Some work in the non-rigid case has been done (besides that we will recall
below) by considering generalizations of the hypergeometric equation as an
equation (rather than as a connection); for example, the algebraic solutions of
the Lamé equation were studied in Beukers and van der Waall (2004) (Lamé
equations are basically the second order Fuchsian equations with four singular
points on P1 such that three of the local monodromies are of order 2). In general
connections of type (A) with such monodromy representations will not come
from a Lamé equation (since upon choosing a cyclic vector the corresponding
equations will in general have additional apparent singularities; this can also be
seen by counting dimensions). Indeed it turns out (op. cit.) that Lamé equations
only have finite monodromy for special configurations of the four poles.
The Painlevé property means that any local solution y(t) defined in a disc
in the three-punctured sphere P1 \ {0, 1, ∞} extends to a meromorphic function
on the universal cover of P1 \ {0, 1, ∞}. It is this property that enables us to
speak of the monodromy of PVI solutions. Concerning solutions there is a basic
trichotomy (see Watanabe (1998)):
⎧
⎪
⎨a ‘new’ transcendental function, or
A solution of PVI is either a solution of a first-order Riccati equation, or
⎪
⎩
an algebraic function.
α = (θ4 − 1)2 /2, β = −θ12 /2, γ = θ32 /2, and δ = (1 − θ22 )/2.
Before going into more detail let us also mention one further property of PVI :
it admits a group of symmetries isomorphic to the affine Weyl group of type
F4 (see Okamoto (1987) or the exposition in Boalch (2006)). Indeed treating
θ = (θ1 , . . . , θ4 ) ∈ C4 as the set of parameters for PVI is useful since the affine
F4 Weyl group of symmetries acts in the standard way on this C4 . (We will see
below that these four parameters may also be interpreted as coordinates on the
moduli space of cubic surfaces.)
216 Towards a non-linear Schwarz’s list
Here G = SL2 (C) does not act on B and acts by diagonal conjugation on the
residues Ai . In general this quotient will not be well behaved, but it has a natural
Poisson structure and the generic symplectic leaves will be smooth complex
symplectic surfaces. Clearly M∗ is trivial as a bundle over B (projecting onto the
configuration of poles), but the nonabelian Gauss–Manin connection is different
to this trivial connection and was computed about 100 years ago by Schlesinger
(essentially in the way stated above it seems). The non-linear connection is given
by Schlesinger’s equations, which in the case at hand are
dA1 [A2 , A1 ] dA3 [A2 , A3 ]
= , =
dt t dt t−1
together with a third equation for dA2 /dt easily deduced from the fact that A4
remains constant. If the residues of the connection satisfy these equations then
the corresponding monodromy representation remains constant as t varies. (They
are
easily derived from the vanishing
of the curvature of the ‘full’ connection
d − A1 z + A2 z−t + A3 z−1 .)
dz dz−dt dz
full generality by R. Fuchs (1905) (whose father L. Fuchs was also the father of
‘Fuchsian equations’).
P1
such that
r t is a Belyi map (i.e. its branch locus is a subset of {0, 1, ∞}).
r y, when viewed as a function of t away from the ramification points of t,
solves PVI for some value of the four parameters.
In principle it is straightforward to go between the two definitions, but in practice
it is useful to look for a good model of Π (and the model given by the closure of
the zero locus of the polynomial F is usually a bad choice).
(s − 1)(s + 2) (s − 1)2 (s + 2)
y= , t= , and θ = (2, 1, 1, 2)/3,
s(s + 1) (s + 1)2 (s − 2)
θ1 = λ1 − µ1 , θ2 = λ2 − µ1 , θ3 = λ3 − µ1 , and θ4 = µ3 − µ2 ,
The implication of this for algebraic solutions should now be clear: the mon-
odromy of a PVI solution is also described by an action of the free group F2 on
(conjugacy classes of) triples of three-dimensional complex reflections (r1 , r2 , r3 )
(with the same formula as before, just replace Mi by ri ). Thus in this context the
‘obvious’ topological solutions (i.e. finite F2 orbits) come from taking a triple of
generating reflections of a finite complex reflection group in GL3 (C). Such finite
complex reflection groups were classified by Shephard and Todd (1954) and apart
from the familiar real orthogonal reflection groups there is an infinite family plus
four exceptional complex groups, the Klein reflection group (of order 336, a two-
fold cover of Klein’s simple group isomorphic to PSL2 (F7 ) → PGL3 (C)), two
Hessian groups, and the Valentiner group (of order 2,160, a six-fold cover of
A6 → PGL3 (C)).
The infinite family of groups and the two Hessian groups do not seem to
lead to interesting new solutions, but by computing the F2 orbits (determining
the topology of Π) it is easy to see that the Klein group yields a genus 0
degree 7 solution and the Valentiner group has three inequivalent triples of
generating reflections, each leading to genus 1 solutions with degrees 15,15,
and 24, respectively. These are new solutions, previously undetected. (The 24
appearing here led to a certain amount of trepidation, given that the 10 page
elliptic solution of Dubrovin and Mazzocco 2000 had degree 18.)
11.4.1 Construction
Of course finding the topological solution is not the same as finding an explicit
isomonodromic family of connections; one needs to solve a family of Riemann–
Hilbert problems inverting the transcendental Riemann–Hilbert map for each
value of t. (Indeed my original plan was to just prove the existence of new
interesting solutions, in Boalch (2003), but a certain stubbornness, and some
inspiration from reading about Klein’s work finding explicit 3 × 3 matrices
generating his simple group, convinced us to go further.)
The two main steps in the method we finally got to work are as follows. (This
is a generalization of the method used by Dubrovin and Mazzocco 2000.)
1. Jimbo’s asymptotic formulae. Jimbo (1982) found an exact for-
mula for the leading asymptotics at t = 0 of the branch of the PVI solution
y(t) corresponding to any sufficiently generic linear monodromy representation
(M1 , M2 , M3 ). (This formula was obtained by considering the degeneration of
the isomonodromic family of connections (A) as t → 0; in the limit the four-
punctured sphere degenerates into a stable curve with two components, each
Beyond Platonic Painlevé VI solutions 223
B
0 1 ∞
with three marked points. The connections (A) degenerate into hypergeometric
systems on each component, with known monodromy (Figure 11.1). Since these
are rigid it is easy to solve their Riemann–Hilbert problems explicitly and this
gives the leading asymptotics of the isomonodromic family and thus of the PVI
solution.)
This is useful for us because, as Jimbo mentions, one may substitute the leading
asymptotics back into the PVI equation to get arbitrarily many terms of the
precise asymptotic expansion of the solution at 0. If the solution is algebraic,
then this is its Puiseux expansion, a sufficient number of terms of which will
determine the entire solution.
It turns out there was a typo in Jimbo (1982), which meant the entire method
did not work (indeed the fact it did not work led to the questioning of Jimbo’s
formula and hence the correction in Boalch 2005). (Note the special parameters
of Dubrovin and Mazzocco 2000 are not covered by Jimbo’s result; rather they
adapted the argument of Jimbo 1982 to their case.)
2. Relating (A) and (B). Since Jimbo’s formula requires a monodromy
representation of a connection of type (A), and we are starting with a triple of
3 × 3 complex reflections (the monodromy representation of a connection of type
(B)), the second step is that we need to see how to go between these two pictures
(on both the De Rham and Betti sides of the Riemann–Hilbert correspondence).
This will be described in the following subsection.
The idea is to describe a transcendental map from gln (C) to GLn (C) in two
different ways (the two paths down the left and the right from the top to the
bottom of the figure).
Choose n distinct complex numbers a1 , . . . , an and define A0 = diag
(a1 , . . . , an ). Roughly speaking (on a dense open patch) the left-hand column
arises by defining Ai = Ei A (setting to zero all but the ith row of A) and
Beyond Platonic Painlevé VI solutions 225
Ai
constructing the logarithmic connection d − z−ai
dz having rank 1 residues
at each ai . Then taking the monodromy of this yields n complex reflections ri
(and if bases of solutions are chosen carefully one can naturally define vectors ei
and one-forms αi such that ri = 1 + ei ⊗ αi and that the ei form a basis). Then
the map to GLn (C) is given by taking the product of rn · · · r1 of these reflections,
written in the ei basis.
Now the key algebraic fact, which dates back at least to Killing (1889) (see
Coleman 1989), is that any such product of complex reflection lies in the big cell
of GLn (C) and so may be factored as the product of a lower triangular and an
upper triangular matrix. We write this product as u−1 − hu+ with u± ∈ U± the
unipotent triangular subgroups, and h ∈ H diagonal:
rn · · · r2 r1 = u−1
− hu+ . (11.2)
Further, although this relation between the reflections and u± looks to be highly
non-linear, one can relate them in an almost linear fashion: the matrix hu+ − u−
is the matrix with entries αi (ej ).
On the other hand it turns out that the same map can
be defined by taking
A0
the Stokes data of the irregular connection d − z2 + z dz. Indeed the map on
A
the right-hand side generalizes (Boalch 2002) to any complex reductive group
G in place of GLn (C), but only for GLn (C) is the alternative ‘logarithmic’
viewpoint available. Thus u± are also the two Stokes matrices of this irregular
connection (the natural analogue of monodromy data for such connections); the
exact definition is not important here. (The element h is the so-called formal
monodromy, explicitly it is simply exp(2πiΛ) where Λ is the diagonal part of
A.) The two connections are related (see Balser, Jurkat, and Lutz 1981) by
the Fourier–Laplace transform: this is more than just formal, and by relating
bases of solutions on both sides the stated relation between the Stokes and
monodromy data is obtained. (In both cases the resulting element of GLn (C)
is the monodromy around z = ∞ in a suitable basis.) In summary we see that
the ‘Betti’ incarnation of the Fourier–Laplace transform is the relation of Killing–
Coxeter.
Now to apply this in the current context we consider the effect of adding a
scalar λ to A ∈ gln (C). On the right-hand side this corresponds to tensoring
the irregular connection by the logarithmic connection d − λdz/z on the trivial
line bundle, and Balser, Jurkat, and Lutz (1981) showed that the Stokes data is
changed only by scaling h by s := exp(2πiλ), fixing u± . On the logarithmic side
this corresponds to a non-trivial convolution operation, changing the monodromy
representation in a non-trivial way. Of course using the Killing–Coxeter identity
we now see precisely how the complex reflections vary. (It is perhaps worth
noting that this scalar shift is essentially the inverse of the spectral parameter
introduced by Killing 1889, p. 20, appearing in the characteristic polynomial of
the Killing–Coxeter matrix (11.2): det(u−1 − shu+ − 1) = det(shu+ − u− ).)
If we set n = 3 then the logarithmic connections appearing are of the form
(B), upon taking a1 , a2 , a3 = 0, t, 1. Then we may choose the scalar shift such
226 Towards a non-linear Schwarz’s list
that the resulting element of GL3 (C) has 1 as an eigenvalue. This implies that
the connections are reducible and we can take the irreducible rank 2 sub- or
quotient connection. Projecting to sl2 gives the desired connection of type (A)
(see Boalch 2005). (Note that there is a choice involved here, of which eigenvalue
to shift to 1.)
11 2 0 2 b2 c2 32 10 0 3 d4
12 2 0 2 b2 d2 33 12 0 0 abcd
13 2 0 2 c2 d2 34 12 1 1 a b c2
14 3 0 1 b c2 d 35 12 1 1 a b d2
15 3 0 1 b c d2 36 12 1 1 b2 c d
16 4 0 2 a c3 37 15 1 2 b3 c
17 4 0 2 a d3 38 15 1 2 b3 d
18 4 0 2 c3 d 39 15 1 2 b2 c2
19 4 0 2 c d3 40 15 1 2 b2 d2
20 5 0 1 b2 c d 41 18 1 3 b4
21 5 0 2 c2 d2 42 20 1 1 a b2 c
22 6 0 1 b c2 d 43 20 1 1 a b2 d
23 6 0 1 b c d2 44 20 1 3 a2 c2
24 8 0 1 a c2 d 45 20 1 3 a2 d 2
25 8 0 1 a c d2 46 24 1 2 a b3
26 9 1 2 b c3 47 30 2 2 a2 b c
27 9 1 2 b d3 48 30 2 2 a2 b d
28 10 0 2 a2 c d 49 36 3 3 a2 b2
29 10 0 2 b3 c 50 40 3 3 a3 c
30 10 0 2 b3 d 51 40 3 3 a3 d
31 10 0 3 c4 52 72 7 3 a3 b
that most of the rows of this table have some representative (in their affine F4
orbit) to which Jimbo’s formula maybe applied (on every branch). Thus we could
start working down the list constructing new solutions. An initial goal was to get
to solution 33: this solution purports to be on none of the reflection hyperplanes
and the folklore was that all explicit solutions to Painlevé equations must lie on
some reflection hyperplane. The folklore was wrong:
‘Generic’ solution/Icosahedral solution 33
9s(s2 + 1)(3s − 4)(15s4 − 5s3 + 3s2 − 3s + 2)
y=− ,
(2s − 1)2 (9s2 + 4)(9s2 + 3s + 10)
11.5 Pullbacks
In his 1884 book on the icosahedron (see Klein 1956), Klein showed that all
second-order Fuchsian differential equations with finite monodromy are (essen-
tially) pullbacks of a hypergeometric equation along a rational map f :
P1
2 2 3 7
1 2 3 1
f ·· 2
· 2 3 1
1 2 1 1
2 3 7
P1
(where a, b, and c are lifts to SL2 (C) of standard generators of ∆237 with cba = 1),
which we know a priori lives in a finite F2 orbit. One finds immediately that the
Final steps 231
orbit through the conjugacy class of this triple has size 18 and constitutes a
genus 1, degree 18 topological PVI solution.
Now it turns out that Jimbo’s formula may be applied to every branch of
this solution, and proceeding as before we obtain (Boalch 2006c) the solution
explicitly:
Elliptic 237 solution
1 3 s8 −2 s7 −4 s6 − 204 s5 −536 s4 −1738 s3 −5064 s2 −4808 s−3199 u
y= − ,
2 4 (s6 + 196 s3 + 189 s2 + 756 s + 154) (s2 + s + 7) (s + 1)
9
1 s − 84 s6 − 378 s5 − 1512 s4 − 5208 s3 − 7236 s2 − 8127 s − 784 u
t= − 2 2 ,
2 432 s (s + 1) (s2 + s + 7)
where u2 = s (s2 + s + 7) and θ = (2/7, 2/7, 2/7, 1/3). (This solution, or rather
an inequivalent ‘Galois conjugate’ of it, has also been obtained independently
by Kitaev 2006, p. 219 by directly computing such a family of rational maps –
apparently also influenced by Doran’s list.)
Substituting this into the formula of Theorem 11.2 with λ = (1, 1, 1)/2 and µ =
(1, 3, 5)/6 now gives explicitly the third (and trickiest) family of connections of
type (B) with monodromy the icosahedral reflection group.
This can be pushed further with more tweaking to get up to degree 24 (row
46 in Table 11.2), that is, to obtain the largest Valentiner solution (Boalch
2006a) (the main further tricks used are described in (Boalch 2006c, appen-
dix C). In particular this finishes the construction of all elliptic icosahedral
solutions. Intriguingly, one finds that the resulting elliptic icosahedral Painlevé
curves Π become singular only on reduction modulo the primes 2, 3, and 5
(except for rows 44 and 45 – we will see another reason in the following
subsection that these are abnormal). Similarly the elliptic Painlevé curve related
to the 237 triangle group becomes singular only on reduction modulo 2, 3,
and 7.
11.7 Conclusion
Thus in conclusion we have filled in a number of rows of what could be called
the non-linear Schwarz’s list. Whether or not there will be other rows remains
to be seen. So far this list of known algebraic solutions to PVI takes the following
shape (we will use the letters d and g to denote the degree and genus of solutions,
and consider solutions up to equivalence under Okamoto’s affine F4 symmetry
group. Some non-trivial work has been done to establish which of the published
solutions are equivalent to each other and which were genuinely new). See also
Boalch (2006d).
First there are the rational solutions (d = 1), studied by Mazzocco (2001)
and Yuan and Li (2002), which fit into the set of Riccati solutions classified by
Watanabe (1998). (Beware that ‘rational’ here means the solution is a rational
function of t, which implies, but is by no means equivalent to, having a rational
parameterization.)
Then there are three continuous
√ families of solutions g = 0, d = 2, 3, 4.
The degree 2 family is y = t which, as one may readily verify, solves PVI
for a family of possible parameter values. Similarly the degree 3 tetrahedral
solution, and the degree 4 octahedral and dihedral solutions (of Dubrovin 1995
and Hitchin 1995a, 2003) fit into such families, as discussed in Ben Hamed and
Gavrilov (2005); Boalch (2006a) and Cantat and Loray (2007). In general in
such a family y(t) may depend on the parameters of the family. Ben Hamed
and Gavrilov (2005) showed that any family with y(t) not depending on the
parameters is equivalent to one of the above cases and recently Cantat and Loray
(2007) showed that any solution with two, three, or four branches is in such
family.
Next there is one discrete family (d, g unbounded, θ = (0, 0, 0, 1) ∼
(1, 1, 1, 1)/2). Indeed this PVI equation was solved completely by Picard (1889,
p. 299), Fuchs (1905), and in a different way by Hitchin (1995b). Algebraic
(determinantal) formulae for the algebraic solutions amongst these appear in
Hitchin (1995a), using links with the Poncelet problem – in this framework they
are dihedral solutions (controlling connections of type (A) with binary dihedral
monodromy).
Finally there are 45 exceptional solutions, which collapse down to 30 if we
identify solutions related by quadratic transformations. The possible genera are
0, 1, 2, 3, 7, and the highest degree is 72. Of these 30 solutions 7 have previously
234 Towards a non-linear Schwarz’s list
References
Andreev, F. V. and Kitaev, A. V. (2002). Transformations RS24 (3) of the ranks ≤ 4 and
algebraic solutions of the sixth Painlevé equation. Commun. Math. Phys. 228(1),
151–76. nlin.SI/0107074.
Balser, W., Jurkat, W. B. and Lutz, D. A. (1981). On the reduction of connection
problems for differential equations with an irregular singularity to ones with only
regular singularities, I., SIAM J. Math. Anal. 12(5), 691–721.
Ben Hamed, B. and Gavrilov, L. (2005). Families of Painlevé VI equations having a
common solution. Int. Math. Res. Not. No. 60, 3727–52. math/0507002.
Beukers, F. and Heckman, G. (1989). Monodromy for the hypergeometric function
n Fn−1 . Invent. Math. 95(2), 325–54.
Beukers, F. and van der Waall, A. (2004). Lamé equations with algebraic solutions. J.
Differential Equations 197(1), 1–25.
Boalch, P. P. (2002). G-bundles, isomonodromy and quantum Weyl groups. Int. Math.
Res. Not. No. 22, 1129–66, math.DG/0108152.
Boalch, P. P. (2003). Painlevé equations and complex reflections. Ann. Inst. Fourier
53(4), 1009–22.
Boalch, P. P. (2005). From Klein to Painlevé via Fourier, Laplace and Jimbo. Proc.
London Math. Soc. 90(3), no. 3, 167–208, math.AG/0308221.
Boalch, P. P. (2006a). The fifty-two icosahedral solutions to Painlevé VI. J. Reine
Angew. Math. 596, 183–214, math.AG/0406281.
Boalch, P. P. (2006b). Six results on Painlevé VI. Sémin. Congr., Vol. 14, Society of
Mathematics France, Paris, math.AG/0503043, pp. 1–20.
Boalch, P. P. (2006c). Some explicit solutions to the Riemann–Hilbert problem. IRMA
Lect. Math. Theor. Phys. 9, 85–112, math.DG/0501464.
Boalch, P. P. (2006d). Survey of the known algebraic solutions of Painlevé VI.
Two talks given at the Newton Institute 71pp of transparencies on their webpage
www.tinyurl.com/boalchINI1, 2.
Boalch, P. P. (2007). Higher genus icosahedral Painlevé curves. Funk. Ekvac. (Kobe)
50(1), 19–32, math.AG/0506407.
Cantat, S. and Loray, F. (2007). Holomorphic dynamics, Painlevé VI equation and
character varieties, arXiv:0711.1579. Preprint Nov.
Cayley, A. (1849). On the triple tangent planes of surfaces of the third order. Dublin
Math. J. No. 4, 118–38.
Coleman, A. J. (1989). Killing and Coxeter transformations of Kac-Moody algebras.
Invent. Math. 95, 447–77.
Dettweiler M. and Reiter, S. (2000). An algorithm of Katz and its application to the
inverse Galois problem. J. Symbolic Comput. 30(6), 761–98.
Dettweiler M. and Reiter, S. (2007). Painlevé equations and the middle convolution.
Adv. Geom. 7(3), 317–30. (See also preprint dated October 1, 2004).
Doran, C. F. (2001). Algebraic and geometric isomonodromic deformations. J. Differ-
ential Geom. 59(1), 33–85.
Dubrovin, B. (1995). Geometry of 2D topological field theories. Integrable Systems and
Quantum Groups, M. Francaviglia and S. Greco (eds.), Vol. 1620, Springer Lecture
Notes in Mathematics, pp. 120–348. hep-th/9407018.
References 235
Michael K. Murray
Dedicated to Nigel Hitchin on the occasion of his 60th birthday
Over the years I have had many interesting mathematical conversations with
Nigel and regularly came away with a solution to a problem or a new idea.
While preparing this chapter I was trying to recall when he first told me
about gerbes. I thought for awhile that age was going to get the better of
my memory as many conversations seemed to have blurred together. But then
I discovered that the annual departmental research reports really do have
their uses. In July 1992 I attended the ‘Symposium on gauge theories and
topology’ at Warwick and reported in the 1992 Departmental Research Report
that:
I . . . had discussions with Nigel Hitchin about ‘gerbes’. These are a generalisation
of line bundles . . .
This year I began some work on a geometric construction called a bundle gerbe.
These provide a geometric realisation of the three dimensional cohomology of a
manifold.
My sincere thanks to Nigel for introducing me to gerbes and for the many
other fascinating insights into mathematics that he has given me over the
years.
238 An introduction to bundle gerbes
12.1 Introduction
The theory of gerbes began with Giraud (1971) and was popularized in the book
by Brylinski (1993). A short introduction by Nigel Hitchin (2003) in the ‘What
is a gerbe?’ series can be found in the Notices of the AMS. Gerbes provide
a geometric realization of the three-dimensional cohomology of a manifold in
a manner analogous to the way a line bundle is a geometric realization of
two-dimensional cohomology. Part of the reason for their recent popularity is
applications to string theory in particular the notion of the B-field. Strings on
a manifold are elements in the loop space of the manifold and we would expect
their quantization to involve a Hermitian line bundle on the loop space arising
from a two class on the loop space. That two class can arise as the transgression
of some three class on the underlying manifold. Gerbes provide a geometrization
of this process. String theory however is not the only application of gerbes and
we refer the interested reader to the related work of Hitchin (2001, 2006) which
applies gerbes to generalized geometry and to reviews such as (Carey et al.
2000) and (Mickelsson 2006) which give applications of gerbes to other problems
in quantum field theory.
As with everything else in the theory of gerbes, the relationship of bundle
gerbes to gerbes is best understood by comparison with the case of Hermitian
line bundles or equivalently U (1) (principal) bundles. There are basically three
ways of thinking about U (1) bundles over a manifold M :
Note that we are slightly abusing the definition of gerbe here as what we are
considering are gerbes with band the sheaf of smooth functions from M into U (1).
There are more general kinds of gerbes on M just as there are more general kinds
of sheaves on M beyond those arising as the sheave of sections of a Hermitian
line bundle.
Recall some of the basic facts about U (1) bundles on a manifold M :
A gerbe is an attempt to generalize all the above facts about U (1) bundles to
some new kind of mathematical object in such a way that the characteristic class
is in three-dimensional cohomology. Obviously for consistency other dimensions
then have to change. In particular the curvature should be a three-form and
holonomy should be over two-dimensional submanifolds. It turns out to be useful
to consider the general case of any dimension of cohomology which we call a
p-gerbe. For historical reasons a p-gerbe has a characteristic class in H p+2 (M, Z)
so the interesting values of p are −2, −1, 0, 1, . . . with U (1) bundles corresponding
to p = 0.
A p-gerbe then is some mathematical object which represents (p + 2)-
dimensional cohomology. To make completely precise what representing (p + 2)-
dimensional cohomology means would take us to far afield from the present
topic, but we give a sketch here to motivate the behaviour we are looking for
in p-gerbes. To this end we will assume our p-gerbes P live in some category G
and there is a (forgetful) functor Π : G → Man the category of manifolds. The
functor Π and the category G have to satisfy
Clearly by construction the category of U (1) bundles, with the forgetful functor
which assigns to a U (1) bundle its base manifold, is an example of a 0-gerbe.
Before we consider other examples we need some facts about bundles with
structure group an abelian Lie group H. If P → M is an H bundle on M
then by choosing local sections of P for an open cover U = {Uα | α ∈ I} we
can construct transition functions hαβ : Uα ∩ Uβ → H and in the usual way this
defines a class c(P ) ∈ H 1 (M, H) where here we abuse notation and write H for
what is really the sheaf of smooth functions with values in H. It is a standard
fact that isomorphism classes of H bundles are in bijective correspondence with
H 1 (M, H) in this manner. If P → M is an H bundle we can define its dual as
follows. Let P ∗ be isomorphic to P as a manifold with projection to M and for
convenience let p∗ ∈ P ∗ denote p ∈ P thought of as an element of P ∗ . Then define
a new H action on P ∗ by p∗ h = (ph−1 )∗ . It is ovious that if hαβ are transition
functions for P then h∗αβ = h−1 ∗
αβ are transition functions for P . In particular we
have that c(P ∗ ) = −c(P ) if we write the group structure on H 1 (M, H) additively.
If Q is another H bundle we can form the fibre product P ×M Q and let H act
on it by (p, q)h = (ph, qh−1 ). Denote the orbit of (p, q) under this action by [p, q]
and define an H action by [p, q]h = [p, qh] = [ph, q]. The resulting H bundle
is denoted by P ⊗ Q → M . If hαβ are transition functions for P and kαβ are
transition functions for Q then hαβ kαβ are transition functions for P ⊗ Q and
thus c(P ⊗ Q) = c(P ) + c(Q). Notice that these constructions will not generally
work for non-abelian groups because in such a case the action of H on P ∗ is not
a right action and the action on P ⊗ Q is not even well-defined.
Introduction 241
Example 12.3 It is clear from the above example that maps φ : M → U (1)
are also −1-gerbes with a connective structure. The dual and product are
just pointwise inverse and pointwise product. The class is the degree and the
connective structure is included automatically as part of φ.
We can also forget that there is a natural connective structure and just regard
maps φ : M → U (1) as −1-gerbes. In that case the natural notion of isomorphism
between two maps φ, χ : M → U (1) would be equality. However two such maps
have the same degree if and only if they are homotopic. So the notion of
242 An introduction to bundle gerbes
12.2 Background
We will be interested in surjective submersions π : Y → M which we regard as
generalizations of open covers. In particular if U = {Uα | α ∈ I} is an open cover
we have the disjoint union
YU = {(x, α) | x ∈ Uα } ⊂ M × I
with projection map π(x, α) = x. The surjective morphism π : YU → M is called
the nerve of the open cover U.
A morphism of surjective submersions π : Y → M and p : X → M is a map
ρ : Y → X covering the identity, that is, p ◦ ρ = π. Any surjective submersion
π : Y → M admits local sections so there is an open cover U of M and local
sections sα : Uα → Y of π. These local sections define a morphism s : YU → Y
by s(x, α) = sα (x). Indeed any morphism YU → Y will be of this form. If V =
{Vα | α ∈ J} is a refinement of U, that is, there is a map ρ : J → I such that for
every α ∈ J we have Vα ⊂ Uρ (α), we have a morphism of surjective submersions
YV → YU defined by (α, x) → (ρ(α), x).
Given a surjective morphism π : Y → M we can form the p-fold fibre product
Y [p] = {(y1 , . . . , yp ) | π(y1 ) = · · · = π(yp )} ⊂ Y p .
The submersion property of π implies that Y [p] is a submanifold of Y p . There
are smooth maps πi : Y [p] → Y [p−1] , for i = 1, . . . , p, defined by omitting the ith
element. We will be interested in two particular examples.
[p]
Example 12.5 If U is an open cover of M then the pth fibre product YU is the
disjoint union of all the ordered p-fold intersections. For example, if U = {U1 , U2 }
[2]
is an open cover of M then YU is the disjoint union of U1 ∩ U2 , U2 ∩ U1 , U1 ∩ U1 ,
and U2 ∩ U2 .
Example 12.6 If P → M is a principal G bundle then P → M is a surjective
submersion. It is easy to show that P [p] = P × Gp−1 . In particular P [2] = P × G
and we shall need later the related fact that there is a map g : P [2] → G defined
by p1 g(p1 , p2 ) = p2 .
Bundle gerbes 243
We can rephrase the existence and associativity of the bundle gerbe multi-
plication to an equivalent pair of conditions in the following way. The bundle
gerbe multiplication gives rise to a section s of δ(P ) → Y [3] . Moreover δ(s)
is a section of δ(δ(P )) → Y [4] . But δ(δ(P )) is canonically trivial so it makes
sense to ask that δ(s) = 1. This is the condition of associativity. The family of
spaces {Y [p] | p = 1, 2, . . . } is an example of a simplicial space (Dupont 1978)
and by comparing to Brylinski and McLaughlin (1994) we see that a bundle
gerbe is the same thing as a simplicial line bundle over this particular simplicial
space.
Example 12.7 If we replace Y in the definition by YU for some open cover U of
M we obtain the definition of gerbe given by Hitchin (2001) and by his student
Chatterjee (1998). This consists of choosing an open cover U of M and a family
of U (1) bundles P : Uα ∩ Uβ such that over triple overlaps we have sections
∗
sαβγ ∈ Γ(Uα ∩ Uβ ∩ Uγ | Pβγ ⊗ Pαγ ⊗ Pαβ )
and we require that δ(s) = 1 in the appropriate way.
Example 12.8 The simplest example of a line bundle is given by the clutching
construction on the two sphere S 2 . If U0 and U1 are the open neighbourhoods of
the north and south hemispheres we take the transition function g : U0 ∩ U1 →
U (1) to have winding number 1. As there are only two open sets there is no
condition on triple overlaps and we obtain the U (1) bundle over S 2 of chern
class 1. In a similar fashion we can consider U0 and U1 to be open neighbourhoods
of the north and south hemispheres of the three-sphere S 3 . Their intersection is
retractable to the two-sphere so we can choose over this the U (1) bundle P of
chern class 1. Again there are no additional conditions and we obtain the gerbe
of degree 1 over S 3 .
Example 12.9 Hitchin and Chatterjee also consider a gerbe as in Example
12.7 but with the added requirement that each Pαβ is trivial in the form Pαβ =
[2]
Uα ∩ Uβ × U (1). Writing elements of the disjoint union YU as (α, β, x) where
x ∈ Uα ∩ Uβ we see that the bundle gerbe multiplication must take the form
((α, β, x), z) ⊗ ((β, γ, x), w) → ((α, γ, x), zwgαβγ (x))
for some gαβγ : Uα ∩ Uβ ∩ Uγ → U (1) and will be associative precisely when gαβγ
is a co-cycle.
We will refer to gerbes of the forms in Examples 12.7 and 12.9 as Hitchin–
Chatterjee gerbes. The connection with bundle gerbes is simple. For clarity we
define
Definition 12.2 A bundle gerbe (P, Y ) over M is called local if Y = YU for
some open cover U of M .
We then obviously have
Bundle gerbes 245
12.3.1 Pullback
If f : N → M then we can pullback Y → M to f ∗ (Y ) → N with a map
fˆ: f ∗ (Y ) → Y covering f . There is an induced map fˆ[2] : f ∗ (Y ) → Y [2] . Let
[2]
such that σαβ (x) ∈ P(sα (x),sβ (x)) . Over triple overlaps we have
m(σαβ (x), σβγ (x)) = gαβγ (x)σαγ (x) ∈ P(sα (x),sγ (x))
for gαβγ : Uα ∩ Uβ ∩ Uγ → U (1). This defines a co-cycle which is the Dixmier–
Douady class:
DD((P, Y )) = [gαβγ ] ∈ H 2 (M, U (1)) = H 3 (M, Z).
Example 12.10 If U is an open cover and gαβγ a U (1) co-cycle then we can
build a Hitchin–Chatterjee gerbe or local bundle gerbe of the type considered in
Example 12.9. It is easy to see that this has Dixmier–Douady class given by the
Čech class [gαβγ ].
Notice that this example shows that every class in H 3 (M, Z) arises as the
Dixmier–Douady class of some Hitchin–Chatterjee gerbe or of some (local)
bundle gerbe.
It is straightforward to check that if f : N → M and (P, Y ) is a bundle gerbe
over M then f ∗ (DD(P, Y )) = DD(f ∗ (P, Y )). Moreover we have
1. DD((P, Y )∗ ) = −DD(P, Y )
2. DD((P, Y ) ⊗ (Q, X)) = DD(P, Y ) + DD(Q, X)
We will defer the question of triviality of a bundle gerbe until the next section
and consider next the notion of a connective structure on a bundle gerbe.
12.4 Triviality
Recall that a U (1) bundle P → M is trivial if it is isomorphic to the bundle
M × U (1) or, equivalently, has a global section. This occurs if and only if P → M
has zero Chern class. If sa : Uα → P are local sections then P is determined by
a transition function g : Uα ∩ Uβ → U (1) given by sα = sβ gαβ and P → M is
trivial if and only if there exist hα : Uα → U (1) such that
gαβ = hβ h−1
α .
248 An introduction to bundle gerbes
In an analogous way Hitchin (2001) and Chatterjee (1998) define a gerbe Pαβ →
Uα ∩ Uβ to be trivial if there are U (1) bundles Rα → Uα and isomorphisms
φαβ : Rα ⊗ Rβ∗ → Pαβ on double-overlaps in such a way that the multiplication
becomes the obvious contraction
Rα ⊗ Rβ∗ ⊗ Rβ ⊗ Rγ∗ → Rα ⊗ Rγ∗ .
In the bundle gerbe formalism this idea takes the following form (Murray and
Stevenson 2000). Let R → Y be a U (1) bundle and let δ(R) → Y [2] be defined
as above. Note that δ(R) has a natural associative bundle gerbe multiplication
given by
δ(R)(y1 ,y2 ) ⊗ δ(R)(y2 ,y3 ) = Ry1 ⊗ Ry∗2 ⊗ Ry2 ⊗ Ry∗3 Ry1 ⊗ Ry∗3 = δ(R)(y1 ,y3 ) .
Definition 12.3 A bundle gerbe (P, Y ) over M is called trivial if there is a
U (1) bundle R → Y such that (P, Y ) is isomorphic to (δ(R), Y ). In such a case
we call a choice of R and the isomorphism δ(R) P a trivialization of (P, Y ).
Example 12.12 Let (P, Y ) be a bundle gerbe and assume that Y → M admits
a global section s : M → Y . Define R → Y by Ry = P(s(π(y)),y) . Then we have
an isomorphism
∗
δ(R)(y1 ,y2 ) = P(s(π(y2 )),y2 ) ⊗ P(s(π(y1 )),y1 )
By replacing σαβ by σαβ /hαβ we can assume that gαβγ = 1. Let Yα = π −1 (Uα )
and define Rα → Yα by letting the fibre of Rα over y ∈ Yα be P(y,sα (π(y))) .
Construct an isomorphism χαβ (y) from the fibre of Rα over y to the fibre of
Triviality 249
and using σαβ (π(y)) ∈ P(sα (π(y))sβ (π(y))) . Because σβγ σαγ σαβ = 1 we can show
that χαβ (y) ◦ χβγ (y) = χαγ (y) and hence the Rα clutch together to form a global
U (1) bundle R → Y . It is straightforward to check that δ(R) = P .
Consider now a U (1) bundle R → Y and assume that δ(R) → Y [2] has a section
s with δ(s) = 1 with respect to the canonical trivialization of δ(δ(R)) → Y [3] . The
section s is called descent data for R and is equivalent to R being the pullback
of a U (1) bundle on M . Indeed s constitutes a family of isomorphisms
This should be compared to the case of U (1) bundles (0-gerbes) where two
trivializations or sections of the bundle differ by a map into U (1) which is a −1-
gerbe. The general pattern is that we expect two trivializations of a p-gerbe to
differ by a (p − 1)-gerbe. Notice also that whereas any two trivial U (1) bundles
are isomorphic there are many trivial bundle gerbes which are not isomorphic.
This leads us to the notion of stable isomorphism.
Definition 12.4 If (P, Y ) and (Q, X) are bundle gerbes over M we say they are
stably isomorphic (Murray and Stevenson 2000) if (P, Y )∗ ⊗ (Q, X) is trivial. A
choice of a trivialization is called a stable isomorphism from (P, Y ) to (Q, X).
We have
Proposition 12.4 Bundle gerbes (P, Y ) and (Q, X) over M are stably isomor-
phic if and only if DD(P, Y ) = DD(Q, X). The Dixmier–Douady class defines
a bijection between stable isomorphism classes of bundle gerbes on M and
H 3 (M, Z).
250 An introduction to bundle gerbes
Proof. Bundle gerbes (P, Y ) and (Q, X) over M are stably isomorphic if and
only if (P, Y )∗ ⊗ (Q, X) is trivial which occurs if and only if −DD(P, Y ) +
D(Q, X) = 0. We have already seen that every three class arises as the Dixmier–
Douady class of some bundle gerbe on M .
It follows that the correct notion of equivalence for bundle gerbes is stable
isomorphism. It is actually possible to compose stable isomorphisms and the
details are given in work of Stevenson (2000) where the structure of the two
category of all bundles gerbes on M is discussed (see also Waldorf 2007).
Note that we also have
Proposition 12.5 Every bundle gerbe is stably isomorphic to a Hitchin–
Chatterjee gerbe.
12.4.1 Holonomy
Consider now a bundle gerbe (P, Y ) with connective structure (A, f ) over a
surface Σ. Because H 3 (Σ, Z) = 0 we know that (P, Y ) is trivial. So there is a
U (1) bundle R → Y with δ(R) = P . Choose a connection a for R and note that
δ(a) is a connection for P using the isomorphism δ(R) = P . But δ(δ(a)) is flat
so δ(a) is a bundle gerbe connection. Hence A = δ(a) + α for a one-form α on
Y [2] with δ(α) = 0. Using the exactness of the fundamental complex we can solve
α = δ(α ) and hence show that δ(a + α ) = A. So without loss of generality we
can choose a connection a on R with δ(a) = A. Consider the two-form f − Fa .
This satisfies δ(f − Fa ) = FA − Fδ(a) = 0 so f − FA = π ∗ (µa ) for some two-form
µa on Σ. Define the holonomy of (A, f ) over Σ by
hol((A, f ), Σ) = exp µa
Σ
and note that this is independent of the choice of trivialization R and connection
a. Indeed any two trivializations with connection will differ by a U (1) bundle
on M with connection and the corresponding µa will differ by the curvature
of the connection on that U (1) bundle. But the integral of the curvature of a
U (1) bundle over a closed surface is in 2πiZ so the two definitions of holonomy
agree.
If (P, Y ) is a bundle gerbe with connective structure (A, f ) on a general
manifold M and Σ ⊂ M is a submanifold we can pull (P, Y ) and (A, f ) back
to Σ and define hol((A, f ), Σ) as above. In this more general setting if X ⊂ M is
a three-dimensional submanifold with boundary ∂X also a submanifold of M we
can trivialize (P, Y ) over all of X and repeat the construction above. We then
have dµa = ω, the three-curvature of (A, f ), and thus
hol((A, f ), ∂X) = exp ω
X
Assume that we have a local description for (P, Y ) and (A, f ) as in Section
12.3.4 in terms of gαβγ , Aαβ , and fα . Then there is a remarkable formula
(Gawedzki and Reis 2002) for the holonomy, first proposed in Alvarez (1985)
‘
and also Gawedzki (1988) and subsequently derived by a number of authors,
‘
which can be described as follows. Choose a triangulation $ of Σ and a map
χ : $ → I such that for any simplex σ ∈ $ we have σ ⊂ Uχ(σ) . Write σ2 , σ 1 ,
and σ 0 for two-, one-, and zero-dimensional simplices, that is, faces, edges, and
vertices, respectively. Then
hol((A, f ), Σ) =
F
exp fχ(σ 2 ) exp Aχ(σ2 )χ(σ 1 ) gχ(σ2 )χ(σ 1 )χ(σ 0 ) (σ 0 ).
σ2 σ2 σ 1 ⊂σ 2 σ1 σ 0 ⊂σ1 ⊂σ 2
gαβ : Uα ∩ Uβ → G
in the usual way. Because these double overlaps are all contractible we can choose
lifts of each gαβ to ĝαβ : Uα ∩ Uβ → Ĝ. Note that
−1
αβγ = ĝβγ ĝαγ ĝαβ
takes values in U (1). It is not difficult to show that αβγ is a U (1) valued co-cycle
and that the class
The bundle gerbe (Q, Y ) is called the lifting bundle gerbe of Y (Murray 1996).
It is easy to check that the lifting bundle gerbe has Dixmier–Douady class
precisely the obstruction to lifting the bundle Y → M . Indeed if we follow
through the construction in Section 12.3.3 we find that σ̂αβ = ĝαβ . It follows
from the discussion above that the lifting bundle gerbe is trivial if and only if
the bundle Y → M lifts to Ĝ. In fact this follows directly because a lift Ŷ → Y
will be a U (1) bundle over Y and actually a trivialization of the lifting bundle
gerbe defined above.
Examples of bundle gerbes 253
gαβ : Uα ∩ Uβ → U (H)
gαβ : Uα ∩ Uβ → U (H)
satisfying
−1
gβγ gαγ gαβ = αβγ 1U (H) .
0 → Zn → SU (n) → P U (n) → 0.
In this case the lifting bundle gerbe is really a Zn ⊂ U (1) bundle gerbe and the
Dixmier–Douady class is a torsion class in the image of the Bockstein map
For further details on Zn bundle gerbes see Carey et al. (2000). Note finally
that if Y → M is a P U (n) bundle then the lifting bundle gerbe has finite-
dimensional fibres so the fact that its Dixmier–Douady class is torsion is implied
by Proposition 12.6.
254 An introduction to bundle gerbes
Similarly
W(g,w,z) ⊕ W(g,z,w) = Cn
so that
det W(g,w,z) ⊗ det W(g,z,w) = C
and
∗
det W(g,w,z) = det W(g,z,w) .
There are a number of other cases that can be dealt with in a similar fashion and
putting all these facts together gives a bundle gerbe multiplication on P → Y [2] .
A construction of the curving on (P, Y ) appears in Murray and Stevenson (2008).
It follows that
Γ(g) = exp ĝ ∗ (ω) ∈ U (1)
X
depends only on g.
If we choose a suitable connection and curving (A, f ) on the basic gerbe on
G, so that it has curvature ω then we see that
Γ(g) = hol(Σ, g∗ (A, f )).
The gerbe approach to the Wess–Zumino–Witten term has two advantages.
Firstly, it removes the topological restriction on M of two-connectedness
256 An introduction to bundle gerbes
necessary in Witten’s definition so that the map g can be extended to the three-
manifold X. Secondly, we can use the local formula for the holonomy given in
Section 12.4.1. For details see Gawedzki and Reis (2002).
‘
12.6.2 Faddeev–Mickelsson anomaly
We follow Carey and Murray (1996) and Segal (1985). Let X be a compact,
Riemannian, spin, three-manifold and denote by A the space of connections on
a complex vector bundle over M and by G the space of gauge transformations.
For any A ∈ A the chiral Dirac operator DA coupled to A has discrete spectrum.
Let
Y = {(A, t) | t ∈
/ spec(DA )}
be considered as a submersion over A. Note that G acts on Y and Y /G → A/G
is another submersion. Following Segal (1985) for (A, s) ∈ Y we can decompose
the Hilbert space H of coupled spinors into a direct sum of eigenspaces of DA for
eigenvalues greater than s and a direct sum of eigenspaces of DA for eigenvalues
− +
less than s. Denote these by H(A,s) and H(A,s) , respectively. We can then form
the Fock space
H H ∗
−
F(A,s) = +
H(A,s) ⊗ H(A,s)
where V(A,t,s) is the sum of all the eigenspaces of DA for eigenvalues between t
and s. If we use the canonical isomorphism
H H
∗
V(A,t,s) ⊗ det V(A,t,s) = V(A,t,s)
for a suitable Hilbert space H̃. However there is a more direct approach as follows.
Let P(A,s,t) be the unitary frame bundle of det V(A,t,s) . Notice that if r < s < t
then
V(A,r,t) ⊕ V(A,t,s) = V(A,r,s)
so that
det V(A,r,t) ⊗ det V(A,t,s) = det V(A,r,s)
gives rise to a bundle gerbe multiplication similar to the SU (n) case in Example
12.14. Again G acts so this descends to a bundle gerbe on A/G. If this bundle
gerbe is trivial then there is a line bundle L → Y such that
det V(A,t,s) = L(A,s) ⊗ L∗(A,t)
and moreover as it is the bundle gerbe on A/G which is trivial these are G
equivariant isomorphisms. It follows that
F(A,s) = F(A,t) ⊗ L(A,s) ⊗ L∗(A,t)
or
F(A,s) ⊗ L∗(A,s) = F(A,t) ⊗ L∗(A,t) ,
and these are G-equivariant isomorphisms. Hence F(A,s) ⊗ L∗(A,s) descends to a
G-equivariant Hilbert bundle on A whose projectivization is P. It follows that
there is a Hilbert bundle on A/G whose projectivization is P/G.
relationship with bundle gerbes is discussed in Murray (1996) and in more detail
in Stevenson (2000). In this same context I would like to thank Larry Breen
for pointing out that in the work of Ulbrich (1990, 1991) the notion of cocycle
bitorsors can be interpreted as a form of bundle gerbe.
In the definitions and theory above we could replace U (1) by any abelian
group H and there would only be obvious minor modifications such as the
Dixmier–Douady class being in H 2 (M, H). If we want to replace H by a non-
abelian group things become more difficult as we mentioned in the introduction
because we cannot form the product of two H bundles if H is non-abelian.
To get around this difficulty we need to replace principal bundles by principal
bibundles which have a left and right group actions. The resulting theory becomes
more complicated although closer to Giraud’s original aim of understanding non-
abelian cohomology. For details see Aschieri et al. (2005) and Breen and Messing
(2005).
We have motivated gerbes by the idea of replacing the transition function gαβ
by a U (1) bundle Pαβ on double-overlaps. It is natural to consider what happens
if we take the next step and replace Pαβ by a bundle gerbe on each double-
overlap. This gives rise to the notion of bundle 2-gerbes whose characteristic
class is a four class. In particular there is associated to any principal G bundle
P → M a bundle 2-gerbe whose characteristic class is the pontrjagin class of P .
For details of the theory see Stevenson (2000, 2004) and for an application to
Chern–Simons theory see Carey et al. (2005).
It is clear that we can continue on in this fashion and consider bundle 2-gerbes
on double-overlaps and more generally inductively use p-gerbes on double-
overlaps to define (p + 1)-gerbes. However the theory becomes increasingly com-
plex for a reason that we have not paid much attention to in the discussion above.
In the case of U (1) bundles the transition function has to satisfy one condition,
the co-cycle identity. In the case of gerbes the U (1) bundles over double-overlaps
satisfy two conditions, the existence of a bundle gerbe multiplication and its
associativity. In the case of bundle 2-gerbes there are three conditions and the
complexity continues to grow in this fashion.
Acknowledgements
The author acknowledges the support of the Australian Research Council and
would like to thank the organizers of the conference for the invitation to
participate. The referee is also thanked for useful remarks on the manuscript.
References
Alvarez, O. (1985). Topological quantization and cohomology. Commun. Math. Phys.
100, 279–309.
Aschieri, P., Cantini, L., and Jurčo, B. (2005). Nonabelian bundle gerbes, their differ-
ential geometry and gauge theory. Commun. Math. Phys. 254(2), 367–400.
References 259
Murray, M. K. (1996). Bundle gerbes. J. London Math. Soc. (2) 54(2), 403–16.
Murray, M. K. and Stevenson, D. (2000). Bundle gerbes: stable isomorphism and local
theory. J. London Math. Soc. (2) 62(3), 925–37.
Murray, M. K. and Stevenson, D. (2003). Higgs fields, bundle gerbes and string
structures. Commun. Math. Phys. 243(3), 541–55.
Murray, M. K. and Stevenson, D. (2008). The basic bundle gerbe on unitary groups.
J. Geometry Phys. 58(11), 1571–90.
Pressley, A. and Segal, G. (1986). Loop Groups. Oxford Mathematical Monographs,
The Clarendon Press, Oxford.
Segal, G. (1985). Faddeev’s anomaly in Gauss’s Law. Preprint, Oxford University,
Oxford.
Stevenson, D. (2000). The geometry of bundle gerbes, PhD thesis, University of
Adelaide, Adelaide, Australia 2000. math.DG/0004117.
Stevenson, D. (2004). Bundle 2-gerbes. Proc. London Math. Soc. (3) 88(2), 405–35.
Ulbrich, K.-H. (1990). On the correspondence between gerbes and bouquets. Math.
Proc. Camb. Philos. Soc. 108(1), 1–5.
Ulbrich, K.-H. (1991). On cocycle bitorsors and gerbes over a Grothendieck topos.
Math. Proc. Camb. Philos. Soc. 110(1), 49–55.
Waldorf, K. (2007). More morphisms between bundle gerbes. Theory and Applications
of Categories 18(9), 240–273.
Witten, E. (1984). Non-abelian bosonization in two dimensions. Commun. Math. Phys.
92(4), 455–72.
XIII
PROJECTIVE LINKING AND BOUNDARIES OF
POSITIVE HOLOMORPHIC CHAINS IN
PROJECTIVE MANIFOLDS, PART I
13.1 Introduction
In 2000 H. Alexander and J. Wermer published the following result:
Link(Γ, Z) ≥ 0
It is interesting to note that while Wermer’s theorem holds for curves with only
weak differentiability properties (see Alexander and Wermer 1998 or Dinh and
Lawrence 2003 for an account), its projective analogue fails even for C ∞ -curves.
On the other hand there is much evidence for the following:
Conjecture 13.1 Every real analytic curve in Pn is stable.
This brings us to the notion of projective linking numbers. Suppose that Γ ⊂
Pn is a compact-oriented smooth curve, and let Z ⊂ Pn − Γ be an algebraic
subvariety of codimension 1. The projective linking number of Γ with Z is defined
to be
LinkP (Γ, V ) ≡ N • Z − deg(Z) ω
N
where ω is the standard Kähler form on Pn and N is any integral 2-chain with
∂N = Γ in Pn . Here Z is given the canonical orientation, and • : H2 (Pn , Γ) ×
H2n−2 (Pn − Γ) → Z is the topologically defined intersection pairing. This defin-
ition is independent of the choice of N (see Section 13.3). The associated reduced
linking number is defined to be
J P (Γ, Z) ≡ 1
Link LinkP (Γ, Z).
deg(Z)
The basic result proved here is the following:
Theorem 13.4 Let Γ be a oriented, stable, real analytic curve in Pn with a pos-
itive integer multiplicity on each component. Then the following are equivalent:
1. Γ is the boundary of a positive holomorphic 1-chain of mass ≤ Λ in Pn .
J P (Γ, Z) ≥ −Λ for all algebraic hypersurfaces Z in Pn − M .
2. Link
Introduction 263
Condition (2) in Theorem 13.4 has several equivalent formulations. The first
is in terms of projective winding numbers. Given a holomorphic section σ ∈
H 0 (Pn , O(d)), the projective winding number of σ on Γ is defined as the integral
WindP (Γ, σ) ≡ dC logσ,
Γ
and we set
P (Γ, σ) ≡ 1 WindP (Γ, σ).
Wind
d
Another formulation involves the cone P SH ω (Pn ) of quasi-
plurisubharmonic functions. These are the upper semi-continuous
functions f : Pn → [−∞, ∞) for which ddC f + ω is a positive (1,1)-current
on Pn .
In Harvey and Lawson 2006c we prove that if Conjecture 13.2 holds for curves
in P2 , then all the results above continue to hold for real analytic Γ of any odd
dimension 2p − 1. (No stability hypothesis is needed.)
Remark 13.1 There are several quite different characterizations of the bound-
aries of general (i.e. not necessarily positive) holomorphic chains in projective
and certain quasi-projective manifolds (see, e.g., Harvey and Lawson 1977, 2005,
Dolbeault 1983, and Dolbeault and Henkin 1994, 1997).
Note. To keep formulas simple throughout the chapter we adopt the convention
that
i
dC = (∂ − ∂).
2π
for all global sections σ ∈ H 0 (Pn , O(d)) and all d > 0. This set of points is
G
denoted K.
The set KG is independent of the choice of metric on O(1).
The projective hull possesses interesting properties. It and its generalizations
function in projective and Kähler manifolds much as the polynomial hull and
its generalizations function in affine and Stein manifolds. The following were
established in Harvey and Lawson (2006a):
1. If Y ⊂ Pn is an algebraic subvariety and K ⊂ Y , then K G ⊂ Y . That is, KG is
contained in the Zariski closure of K. Furthermore, if Y ⊂ Pn is a projective
manifold and λ = O(1) Y , the λ-projective hull of K ⊂ Y , defined as in
(13.1) with O(1) replaced by λ, agrees with K. G The same is true of λk for
any k.
2. If K is contained in an affine open subset Ω ⊂ Pn , then (K) G poly,Ω ⊆ K.
G
3. If K = ∂C, where C is a holomorphic curve with boundary in P , then n
C ⊆ K.G
G −
4. {K} − K satisfies the maximum modulus principle for holomorphic func-
tions on open subsets of Pn − K. ({K} G − denotes the closure of K.)
G
G
5. {K} −
− K is 1-pseudoconcave in the sense of Dinh and Lawrence (2003).
In particular, for any open subset U ⊂ Pn − K, if the Hausdorff 2-measure
G − ∩ U ) < ∞, then K
H 2 (K G − ∩ U is a complex analytic subvariety of dimen-
sion 1 in U .
6. If K is a real analytic curve, then the Hausdorff dimension of K G is 2.
7. KG is pluripolar if and only if K is pluripolar. 1 Thus there exist smooth
closed curves Γ ⊂ P2 with Γ G = P2 . However, real analytic curves are always
pluripolar.
The projective hull has simple characterizations in both affine and homo-
geneous coordinates. For example, if π : Cn+1 − {0} −→ Pn is the standard
projection, and we set S(K) ≡ π−1 (K) ∩ S 2n+1 , then
< =
G
K = π S(K)poly − {0}
poly is the polynomial hull of S(K) in C
n+1
where S(K) .
There is also a best constant function C : KG → R+ defined at x to be the
G the set
least C = C(x) for which the defining property (13.1) holds. For x ∈ K,
−1
π (x) ∩ S(K)poly is a disk of radius ρ(x) = 1/C(x). One deduces that K G is
compact if C is bounded.
1 A set K is pluripolar if it is locally contained in the −∞ set of a plurisubharmonic function.
266 Projective linking and boundaries of positive holomorphic chains
Definition 13.2 G → R+ is
A compact subset K ⊂ Pn is called stable if C : K
bounded.
Combining a classical argument of E. Bishop with (5) and (6) above gives the
following:
Remark 13.2 The close parallel between polynomial and projective hulls is
signaled by the existence of a projective Gelfand transformation, whose relation
to the classical Gelfand transform is analogous to the relation between Proj of
a graded ring and Spec of a ring in modern algebraic geometry. To any Banach
graded algebra A∗ = d≥0 Ad (a normed graded algebra which is a direct sum
of Banach spaces) one can associate a topological space XA∗ and a Hermitian
line bundle λA∗ → XA∗ with the property that A∗ embeds as a closed subalgebra
9
A∗ ⊆ Γcont XA∗ , λdA∗
d≥0
This parallels the affine case where, for a compact subset K ⊂ Cn , the Gelfand
spectrum of the closure of the polynomials on K in the sup-norm corresponds
to the polynomial hull of K.
For now we shall be concerned with the case p = 1. Here there is a naturally
related notion of projective winding number defined as follows:
Definition 13.4 Suppose Γ ⊂ Pn is a smooth, closed, oriented curve, and let
σ be a holomorphic section of O() over Pn which does not vanish on Γ. Then
the projective winding number of σ on Γ is defined to be
WindP (Γ, σ) ≡ dC logσ (13.4)
Γ
Proposition 13.2 Let Γ and σ be as in Definition 13.4, and let Z be the divisor
of σ. Then
by (13.5).
It will be useful in what follows to normalize these linking numbers:
Definition 13.5 For Γ and Z as above, we define the reduced projective linking
number to be
J P (Γ, Z) ≡ 1
Link LinkP (Γ, Z)
deg(Z)
P (Γ, Z) ≡ WindP (Γ, Z)/
and the reduced projective winding number to be Wind
deg(Z).
Note that if Z = Div(σ) for σ ∈ H 0 (Pn , O()), then by Proposition 13.2 we
have
J P (Γ, Z) =
Link dC logσ .
1
(13.6)
Γ
and for any divisor Z of degree which does not meet Γ we have
LinkP (Γ, Z) − LinkP (Γ, Pn−1 ) = N • Z = LinkCn (Γ, Z),
the classical linking number of Γ and Z.
Remark 13.4 (Relation to sparks and differential characters) The
de Rham–Federer approach to differential characters (cf. Gillet and Soulé 1989,
Harris 1989, and Harvey, Lawson, and Zweck 2003) is built on objects called
sparks. These are generalized differential forms (or currents) α which satisfy the
equation
dα = R − φ
Quasi-plurisubharmonic functions 269
ddC u + ω ≥ 0 on X. (13.9)
Theorem 13.7 Let ω denote the standard Kähler form on Pn . Then the
G of a compact subset K ⊂ Pn , is exactly the subset of points
projective hull K
270 Projective linking and boundaries of positive holomorphic chains
13.7. Sufficiency follows from the fact that such functions are the extreme points
of the cone P SH ω (Pn ).
One importance of quasi-plurisubharmonic functions is that they enable us to
establish Poisson–Jensen measures for points in K G (Harvey and Lawson 2006a,
Theorem 11.1).
= −M(T ) ≥ −Λ
as asserted.
SK ≡ {u ∈ C ∞ (Pn ) : ddC u + ω ≥ 0 on K}
where P1,1 ⊂ E2 (Pn ) denotes the convex cone of positive currents of bidimen-
sion (1,1) on Pn . Note that CK is a weakly closed convex subset which con-
tains 0 (since the set of T ∈ P1,1 with M(T ) ≤ 1 and supp T ⊂ K is weakly
compact).
Suppose that K is a weakly closed convex subset containing 0 in a topological
vector space V . Then the polar of K is defined to be the subset of the dual space
given by
(K0 )0 = K.
Proposition 13.5
SK = (CK )0
and so u ∈ (CK )0 .
(13.13)
Projective Alexander–Wermer theorem for curves 273
Recall from Section 13.4 that for each Λ ∈ R we have the compact set
G
Γ(Λ) = {x ∈ Pn : u(x) ≤ sup u + Λ ∀u ∈ P SH ω (Pn )}
Γ
G = K Γ(Λ).
and that Γ G The following lemma is established in Harvey and Lawson
Λ
(2006a, 18.7):
Lemma 13.2 Let u be a C ∞ function which is defined and quasi-
G − . Fix Λ > 0. Then there is a C ∞
plurisubharmonic on a neighborhood of {Γ}
, which is defined and quasi-plurisubharmonic on all of Pn and agrees
function u
G
with u on a neighborhood of Γ(Λ).
G − denotes the closure of Γ.
Here {Γ} G By our stability assumption, Γ G = Γ(Λ)
G
G G −
for some Λ, and therefore Γ = {Γ} . However, we shall keep the notation {Γ}G −,
when appropriate, in order to prove the more general result mentioned in Remark
13.6 below.
The lemma above leads to the following:
Proposition 13.6 If Γ satisfies condition (3) in Proposition 13.3, that is, if
(−dC Γ)(u) = dC u ≥ −Λ
Γ
for all u ∈ P SH ω (P ), then there exists T ∈ P1,1 with M(T ) ≤ Λ and supp T ⊂
n
G − , such that
{Γ}
ddC T = −dC Γ (13.14)
Proof. If this is not true, then it must fail on some compact neighborhood K of
G − . (Otherwise, there exists a sequence of positive currents {Tj }j , satisfying
{Γ}
(13.14), with M (Tj ) ≤ Λ and supp Tj ⊂ Kj where Kj are compact neighborhoods
squeezing down to {Γ} G − . By the standard compactness theorem for positive
currents, there would then be a convergent subsequence whose limit T would
satisfy the conclusion of Proposition 13.6.)
By (13.13) and the Bipolar Theorem, we then conclude that there is a smooth
function u which is quasi-plurisubharmonic on K with −dC Γ(u) < −Λ. Applying
Lemma 13.2 contradicts the hypothesis.
Now let T be the current given by Proposition 13.6. Let V denote the projective
hull of Γ and recall from Harvey, Lawson, and Wermer (Theorem 4.1) that V has
the following strong regularity. There exists a Riemann surface S with finitely
many components, a compact region W ⊂ S with real analytic boundary, and a
holomorphic map ρ : S → Pn which is generically injective and satisfies
1. ρ(W ) = V .
2. ρ is an embedding on a tubular neighborhood of ∂W in S.
3. ρ(∂W ) is a union of components of the support of Γ.
274 Projective linking and boundaries of positive holomorphic chains
Γ
Proof. That (1) ⇔ (2) ⇒ (3) is clear. To see that (3) ⇒ (2) we use results
of Demailly. Fix u ∈ P SH ω (X) and consider the positive closed (1,1)-current
T ≡ ddC u + ω. Note that [T ] = [ω] = c1 (λ) ∈ H 2 (X; Z). It follows from Demailly
Relative holomorphic cycles 277
The remainder of the argument replicates the one given for Proposition 13.4.
For a compact subset K ⊂ X the authors introduced the notion of the λ-hull
KG λ of K and showed that K Gλ = KG for any embedding X → PN by sections of
some power of λ. We shall say that K is λ-stable if K G λ is compact.
M
Theorem 13.11 Let Γ = a=1 mα Γα be an embedded, oriented, real analytic
closed curve with integer multiplicities in X, and assume Γ is λ-stable. Then
there exists a positive holomorphic 1-chain T in X with dT = Γ and M(T ) ≤ Λ
if and only if any of the equivalent conditions of Proposition 13.6 is satisfied.
Proof. If dT = Γ and M(T ) ≤ Λ, then (1) follows as in the proof of Theorem
13.8 above.
For the converse we recall from Harvey and Lawson (2006a, (4.4)) that the
G λ of a compact subset K ⊂ X consists exactly of the points x where the
hull K
extremal function
ΛK (x) ≡ sup{u(x) : u ∈ P SH ω (X) and u ≤ 0 on X}
is finite. This enables one to directly carry through the arguments for Theorem
13.9 in this case.
Pn − γ.
278 Projective linking and boundaries of positive holomorphic chains
where →−
γ 1 , ..., −
→
γ are the connected components of γ with a chosen orientation
and the mk ’s are integers which we can assume to be positive. We are assuming
that (13.19) holds for any positive algebraic class u. If Γ = 0, the desired
conclusion is immediate, so we assume that Γ = 0. Now let Z ⊂ Pn − γ be any
positive algebraic hypersurface, and note that
τ •Z τ •Z
0 ≤ = − τ (ω) + τ (ω) = LinkP (Γ, Z) + τ (ω).
degZ degZ
Therefore, by Theorem 13.9, there exists a positive holomorphic 1-chain T with
dT = Γ and M (T ) ≤ τ (ω). Note that δ([T ] − τ ) = 0 and ([T ] − τ )(ω) ≥ 0. Hence,
[T ] − τ is a positive class in H2 (Pn ; Z) and is represented by a positive algebraic
1-cycle, say W . Therefore, τ = [T + W ] is represented by a positive holomorphic
chain as claimed.
This result leads to a nice duality between the cone in H2 (Pn , γ; Z) of those
classes which are represented by positive holomorphic chains, and the cone in
H2n−2 (Pn − γ; Z) of classes represented by positive algebraic hypersurfaces. This
duality, in fact, extends to more general projective manifolds. All this is discussed
in detail in Harvey and Lawson (2006d).
Acknowledgment
This chapter was partially supported by the NSF.
References
Alexander, H. and Wermer, J. (1998). Several Complex Variables and Banach Algebras,
Springer-Verlag, New York.
Alexander, H. and Wermer, J. (2000). Linking numbers and boundaries of varieties.
Ann. Math. 151, 125–150.
Demailly, J.-P. (1982). Estimations L2 pour l’opérateur ∂ d’un fibré vectoriel holomor-
phe semi-positif au-dessus d’une variété kählérienne complète. Ann. Sci. E. N. S.
15(4), 457–511.
Demailly, J.-P. (1982). Courants positifs extrêmaux et conjecture de Hodge. Invent.
Math. 69(4), 347–74.
Dinh ,T.-C. and Lawrence, M. (2003). Polynomial hulls and positive currents. Ann.
Fac. Sci. de Toulouse 12, 317–34.
References 279
Harvey, F. R. and Polking, J. (1975). Extending analytic objects. Commun. Pure Appl.
Math. 28, 701–27.
Schaefer, H. H. (1999). Topological Vector Spaces, Springer Verlag, New York.
Wermer, J. (1958). The hull of a curve in Cn . Ann. Math. 68, 550–61.
Wermer, J. (2001). The argument principle and boundaries of analytic varieties.
Operator Theory: Adv. Appl. 127, 639–59.
XIV
SKYRMIONS AND NUCLEI
Nicholas S. Manton
Dedicated to Nigel Hitchin on the occasion of his 60th birthday
14.1 Skyrmions
T. H. R. Skyrme (1961, 1962) proposed that the interior of a nucleus is dominated
by a non-linear semiclassical medium formed from the three pion fields, and
he introduced the Skyrme model, a version of the Lorentz-invariant, non-linear
sigma model, in which the pion fields π = (π1 , π2 , π3 ) are combined into an
SU (2)-valued scalar field
U = (1 − π · π)1/2 1 + iπ · τ , (14.1)
where τ are the Pauli matrices. There is an associated current, with spatial
components Ri = (∂i U )U † . For static fields, the energy in the Skyrme model is
given by
< =
1 1 1
E= − Tr(R R
i i ) − Tr([R i , R j ][R i , R j ]) + m2
Tr(1 − U ) d3 x ,
12π 2 2 16
(14.2)
and the vacuum is U = 1. E is invariant under translations and rotations in R
3
1 : O(3) 2 : Dµh 3 : Td 4 : Oh
B K E/B
1 O(3) 1.2322
2 D∞h 1.1791
3 Td 1.1462
4 Oh 1.1201
5 D2d 1.1172
6 D4d 1.1079
7 Yh 1.0947
8 D6d 1.0960
p(z)
R(z) = (14.6)
q(z)
where p and q are polynomials with no common root, and a radial profile function
f (r) satisfying f (0) = π and f (∞) = 0. One should think of R as a smooth map
from a two-sphere in space (at a given radius) to a two-sphere in the target SU (2)
(at a given distance from the identity). By standard stereographic projection, the
point z corresponds to the Cartesian unit vector
1
Gz =
n (z + z̄, i(z̄ − z), 1 − |z|2 ) . (14.7)
1 + |z|2
nR(z) · τ ) ,
U (r, z) = exp(if (r)G (14.9)
generalizing the hedgehog formula (14.5). The ansatz is a suspension of the map
R : S 2 → S 2 to produce a map U : R → S 3 , the suspension points being the
3
The baryon number, B, of this Skyrme field equals the topological degree of
the rational map R : S 2 → S 2 , which is the higher of the algebraic degrees of
p and q.
An SU (2) Möbius transformation on the domain S 2 of the rational map
corresponds to a spatial rotation, whereas an SU (2) Möbius transformation
on the target S 2 corresponds to a rotation of n G R , and hence to an isospin
rotation of the Skyrme field. Thus if a rational map R has some symmetry
(i.e. a rotation of the domain can be compensated by a rotation of the target),
then the resulting Skyrme field has that symmetry (i.e. a spatial rotation can be
compensated by an isospin rotation).
An important feature of the rational map ansatz is that when one substitutes
it into the Skyrme energy function (14.2), the angular and radial parts decouple.
Rational map ansatz 285
B K I E/B
look like polyhedra with holes in the directions given by the zeros of W , and
why there are 2B − 2 such holes, precisely the structures seen in Figure 14.1.
As an example, the icosahedrally symmetric degree 7 map in (14.12) has
Wronskian
W (z) = 28z(z 10 + 11z 5 − 1), (14.14)
which is proportional to one of the icosahedral Klein polynomials, and vanishes
at the twelve face centres of a dodecahedron (including z = ∞).
The solutions we have described so far are for m = 0, but it is found that
qualitatively similar solutions with 10–20% higher energy exist for m up to 1
and a little beyond, provided B ≤ 7. There is, however, a qualitative change for
B > 7, as we discuss next.
14.3.1 B = 8
When m = 0, the B = 8 Skyrmion is a hollow polyhedron with D6h symmetry,
with no obvious relation to a pair of cubic B = 4 Skyrmions, and this solution
persists at m = 1. However, motivated by the α-particle model, one might expect
Skyrmions and α-particles 287
Figure 14.2 Baryon density contours for (a) two B = 4 cubes with one of the
cubes rotated by 90◦ around the line joining them, (b) the B = 8 truncated
octahedron, and (c) the relaxed B = 8 Skyrmion with m = 1.
14.3.2 B = 12
In the α-particle model, three α-particles form an equilateral triangle. This moti-
vates the search for a triangular B = 12 solution in the Skyrme model, composed
of three B = 4 cubes. A configuration with approximate D3h symmetry can be
obtained with each cube related to its neighbour by a spatial rotation through
120◦ followed by an isorotation by 120◦ . The isorotation cyclically permutes the
values of the pion fields on the faces of the cube, so that these values match
on touching faces, and the cubes attract. It is fairly easy to see that around the
centre of the triangle the field has a winding equivalent to a single Skyrmion, and
is unstable to this Skyrmion moving down and merging with the bottom face
of the triangle; in fact, it fills a hole in the baryon density. This C3 -symmetric
Skyrmion with E/B = 1.288 is presented in Figure 14.3.
288 Skyrmions and nuclei
Figure 14.3 Top and bottom views of the B = 12 Skyrmion with triangular
symmetry.
The top view looks similar to the minimal energy B = 11 Skyrmion with
m = 0, and suggests that the initial arrangement of three cubes can also be
viewed as the m = 1 version of this B = 11 Skyrmion with a single Skyrmion
placed inside at the origin. Such a field configuration can be constructed with
exact D3h symmetry using the double rational map ansatz (Manton and Piette
2001) (see Appendix). This involves a D3h -symmetric outer map of degree 11,
Rout , and a spherically symmetric degree 1 inner map, Rin , together with an
overall radial profile function. The maps are
z 2 (1 + az 3 + bz 6 + cz 9 )
Rout (z) = (14.15)
c + bz 3 + az 6 + z 9
1
Rin (z) = − , (14.16)
z
where a = −2.47, b = −0.84, and c = −0.13. Note that the orientation of Rin has
to be chosen compatibly with the D3h symmetry of Rout . Numerically relaxing
the field yields a solution with exact D3h symmetry, but as we discussed above,
this is only a saddle point and not a local energy minimum.
In Battye and Sutcliffe (2006) another B = 12 solution with C3 symmetry
was found, with E/B = 1.289. It is a general observation that rearrangements
of clusters have only a tiny effect on the energy of a Skyrmion, so as B increases
one expects an increasingly large number of local minima with extremely close
energies.
Rearranged solutions are analogous to the rearrangements of the α-particles
which model excited states of nuclei. An example is the Skyrme model analogue
of the three α-particles in a chain configuration modelling the 7.65 MeV excited
state of 12 C (Morinaga 1956; Friedrich et al. 1971). This is obtained from three
B = 4 cubes placed next to each other in a line, with the middle cube twisted
relative to the other two by 90◦ around the axis of the chain. The relaxed solution
is displayed in Figure 14.4 and has E/B = 1.285. This is apparently the lowest
energy of the B = 12 solutions, but note that the energy difference we are trying
Skyrmions and α-particles 289
Figure 14.4 B = 12 Skyrmion formed from three cubes in a line, with the
middle cube being rotated by 90◦ around the line of the cubes.
12
to understand, 7.65 MeV, is less than 0.1% of the total energy of a C nucleus,
smaller than the numerical errors in the Skyrmion energies.
14.3.3 B = 16
There is a tetrahedrally symmetric B = 16 solution which is an arrangement
of four B = 4 cubes. It is created using the double rational map ansatz as a
starting point, with an outer map of degree 12, Rout , and an inner map of degree
4, Rin , with compatible symmetries, and an overall radial profile function. There
is a Td -symmetric map Rout , and this can be combined with the Oh -symmetric
map familiar from the B = 4 Skyrmion, giving Td symmetry overall. The maps
are
ap3+ + bp3−
Rout = (14.17)
p2+ p−
p+
Rin = , (14.18)
p−
√
where p± (z) = z 4 ± 2 3 iz 2 + 1, a = −0.53, and b = 0.78. Letting the field U
relax, preserving the Td symmetry, results in the solution displayed in Figure
14.5a, in which U = −1 at 16 points clustered into groups of four close to the
centre of each cube. The solution has E/B = 1.288.
This tetrahedral solution is only a saddle point. It is energetically more
favourable for the two cubes on a pair of opposite edges of the tetrahedron
to open out, leading to the D2d symmetric solution in Figure 14.5b, which has
the slightly lower energy, E/B = 1.284. An α-particle molecule of similar shape
has also been found, where it is termed a bent rhomb (Bauhoff et al. 1984).
A stable tetrahedral solution would be preferable, since the closed shell
structure of 16 O is known to be compatible with clustering into a tetrahedral
arrangement of four α-particles. Moreover, the 16 O ground state and the excited
states at 6.1 MeV and 10.4 MeV, with spin/parity 0+ , 3− , and 4+ , and some
higher states, look convincingly like a rotational band for a tetrahedral intrinsic
structure (Dennison 1954; Robson 1979).
290 Skyrmions and nuclei
14.3.4 B = 24
Here one may begin with six B = 4 cubes on the vertices of an octahedron, all
with the same spatial orientation (see Figure 14.6a). The cubes above and below
the four in a square are given an isospin rotation by 180◦ , as this aids their
attraction. The relaxed solution, shown in Figure 14.6b, resembles the B = 16
bent square Skyrmion of Figure 14.5b, joined with the B = 8 Skyrmion of Figure
14.2c. The fact that these B = 16 and B = 8 Skyrmions appear as substructures
is further evidence that they are the minimal energy arrangements of B = 4
cubes. The energy per baryon of this B = 24 Skyrmion is 1.282.
Figure 14.6 (a) Initial configuration of six cubes in a square bipyramid arrange-
ment and (b) the final relaxed B = 24 Skyrmion.
Skyrmions and α-particles 291
Figure 14.7 (a) Initial configuration of four cubes in a square with three extra
cubes placed above the square but not aligned with the cubes below and (b) the
relaxed B = 28 Skyrmion.
14.3.5 B = 28
A starting configuration is shown in Figure 14.7a, with four cubes in a square
and three more placed above, all with the same space and isospace orientations.
Figure 14.7b displays the solution after relaxation. This B = 28 Skyrmion clearly
resembles the cubic B = 32 Skyrmion (see below) with one B = 4 cube removed.
The energy per baryon is 1.279, which is slightly lower than might have been
expected given the energies for B = 16 and B = 24.
14.3.6 B = 32
Even for relatively small values of the pion mass parameter m, the B = 32
Skyrmion is cubic, and has lower energy than the minimal energy, hollow
polyhedral structure (Battye and Sutcliffe 2006). The solution may be thought
of as eight B = 4 cubic Skyrmions placed on the vertices of a cube, each with the
same spatial and isospin orientations, and it may also be created by cutting out a
cubic B = 32 chunk from the infinite, triply periodic Skyrme crystal (Baskerville
1996).
Alternatively, it may be obtained beginning with the double rational map
ansatz. One places a B = 4 cube inside a B = 28 configuration with cubic
symmetry using the maps
p+ (ap6+ + bp3+ p3− − p6− )
Rout = (14.19)
p− (p6+ − bp3+ p3− − ap6− )
p+
Rin = , (14.20)
p−
where a = 0.33 and b = 1.64, and p± (z) are as before. This is displayed in Figure
14.8a. Numerical relaxation yields the solution in Figure 14.8b, which is the
B = 32 Skyrmion, with E/B = 1.274.
292 Skyrmions and nuclei
Figure 14.8 (a) Initial condition of the cubic B = 4 Skyrmion inside a cubic
B = 28 configuration and (b) the final relaxed B = 32 Skyrmion, which is a
chunk of the Skyrme crystal.
found a simple formula for determining whether the FR factor associated with
this symmetry operation is +1 or −1, in terms of the angles of rotation and the
baryon number.
Recently, a recalibration of the Skyrme model around the properties of the
6
Li nucleus has been performed (Manton and Wood 2006). This is better suited
to nuclear physics applications. The motivation for the choice of 6 Li is that it
is a small nucleus of isospin zero, with a pair of levels that can reasonably be
interpreted as a rotational band. Because the isospin is zero, the electric charge
density is half the baryon density. The ground state of spin 1 and the first
excited state of spin 3 are separated by just a few MeV, whereas the mass of
the 6 Li nucleus is approximately 5600 MeV. The rotational motion is therefore
quite non-relativistic, not leading to strong pion radiation, nor to significant
Skyrmion deformation, which could affect the calibration. The B = 6 Skyrmion
is also well-known, and can be approximated by the rational map ansatz, which is
useful when estimating the energy and size. In the new calibration, the conversion
from Skyrme units to physical energy and length units is fixed by fitting the 6 Li
mass and charge radius. m is determined as usual from the physical pion mass,
138 MeV, but because of the change of units, m is roughly doubled and is now
m = 1.125, which is in the range where the solutions described in Section 14.3,
constructed from B = 4 cubes, are favoured.
In the new calibration, the masses of Skyrmions and nuclei are rather similar
to those in the traditional calibration of Adkins and Nappi (1984), but the length
scale is roughly doubled. This means that Skyrmion moments of inertia are four
times larger, and hence the energy splitting of rotational levels is four times
smaller. These moments of inertia (both rotational and isorotational) have been
estimated for a range of baryon numbers, using the rational map ansatz where
appropriate, and used to determine the energies of spin and isospin excited
states for several nuclei (Manko et al. 2007). The results appear much closer
to experimental nuclear energy levels than those found previously. For example,
the spectrum of excited states of the double cube B = 8 Skyrmion matches well
the experimental spectrum of 8 Be and its isobars 8 Li, 8 B, etc. There are also
some detailed suggestions for a Skyrme model understanding of the physically
observed spin states for B = 5 and B = 7, which, as mentioned above, have so
far been problematic.
14.5 Conclusions
For small baryon numbers, the rational map ansatz gives good approximations to
Skyrmions, some of which have the symmetries of Platonic solids. The ansatz has
recently been shown to be helpful both in classifying the allowed spin and isospin
states of quantized Skyrmions, and also for estimating the energies, radii, and
moments of inertia of Skyrmions. The Skyrme model with pion mass parameter
of order 1 has qualitatively new solutions for baryon numbers B ≥ 8. These
are not hollow polyhedra, the solutions found for m = 0 or m = 0.528 (the
Conclusions 295
traditional value), but more dense structures often with clear clustering into
B = 4 cubes, the Skyrme model analogue of α-particle molecules. This gives an
improved fit of the model to nuclei like 8 Be and 12 C, provided the energy scale
and especially the length scale of the model are recalibrated. The double rational
map ansatz gives some insight into the shapes of Skyrmions for baryon numbers
in the range 8 ≤ B ≤ 32, and an improved version of this ansatz is proposed
(see Appendix).
The interpretation of the Skyrmions discussed in this chapter has been that of
nucleons and nuclei. However, there has been considerable interest in recent years
in studying variants of non-linear scalar field theories in the context of condensed
matter physics, the classic example being the continuum limit of the Heisenberg
model of ferromagnets. The Skyrmions that occur in such theories are often
localized in two dimensions, which means they can become extended vortex-like
structures in a three-dimensional material. For a recent analysis of one model and
further references, see Rössler et al. (2006). If the topology of the model is as in
the original Skyrme model, then Skyrmions with three-dimensional localization
should occur in the interior of the material, and it would be interesting if evidence
could be found for the polyhedral structures of higher charge Skyrmions shown
in Figure 14.1.
A problem with the double rational map ansatz is that U takes the value −1
on the entire sphere at radius r0 , and true Skyrmions do not have this behaviour.
But the formulation above suggests how this problem can be avoided. One simply
needs to relax the non-overlapping character of the product of U1 and U2 by
allowing both f1 and f2 to decrease freely from π at r = 0 to 0 at r = ∞, without
further constraint at an intermediate radius r0 . One would expect f1 to decrease
more rapidly than f2 if the Skyrmion has a genuine inner and outer structure.
This version of the product ansatz preserves the joint rotational symmetries of
U1 and U2 , but not any inversion or reflection symmetries. As usual with the
product ansatz, the field U is sensitive to the order in which U1 and U2 are
multiplied, and the loss of reflection symmetry is related to this.
It is easy to verify that a small perturbation of the profiles f1 and f2 away
from their initial, non-overlapping form will tend to reduce the energy. This is
because the trajectory of the field U along a generic radial line will no longer
pass through U = −1, but will rather take a short cut, which reduces the radial
derivative of U without a significant increase in the angular derivatives, and
also reduces the potential energy. The non-generic lines are those for which
Rin (z) = Rout (z), and there are B of these, counted with multiplicity. Therefore,
U = −1 at B points, the number required topologically if they all have positive
multiplicity. The points lie on these special lines and are all at the same distance
from the origin. Further investigations are needed. With S. Krusch, the author
is attempting to understand this improved double rational map ansatz more
systematically, in order to find its optimal form. Alternatively, one could try ad
hoc ansätze for f1 and f2 , perhaps with one scale parameter each, and seek to
minimize the energy of U numerically.
Acknowledgements
This chapter is largely based on Battye et al. (2007) and part of Manton and
Sutcliffe (2004). I would like to acknowledge the contributions of Richard Battye
and Paul Sutcliffe to these joint works, and especially thank them for all the
figures that are reproduced here.
This is also an occasion to express my debt to Nigel Hitchin’s mathemati-
cal insight. His work on monopoles, in particular, has inspired several of the
developments discussed here.
References
Adkins, G. S. and Nappi, C. R. (1984). The Skyrme model with pion masses. Nucl.
Phys. B 233, 109.
Adkins, G. S., Nappi, C. R., and Witten, E. (1983). Static properties of nucleons in
the Skyrme model. Nucl. Phys. B 228, 552.
Baskerville, W. K. (1996). Making nuclei out of the Skyrme crystal. Nucl. Phys. A 596,
611.
References 297
Battye, R. A. and Sutcliffe, P. M. (1997). Symmetric Skyrmions. Phys. Rev. Lett. 79,
363.
Battye, R. A. and Sutcliffe, P. M. (1998). A Skyrme lattice with hexagonal symmetry.
Phys. Lett. B 416, 385.
Battye, R. A. and Sutcliffe, P. M. (2001). Solitonic fullerene structures in light atomic
nuclei. Phys. Rev. Lett. 86, 3989.
Battye, R. A. and Sutcliffe, P. M. (2002). Skyrmions, fullerenes and rational maps. Rev.
Math. Phys. 14, 29.
Battye, R. A. and Sutcliffe, P. M. (2005). Skyrmions and the pion mass. Nucl. Phys. B
705, 384.
Battye, R. A. and Sutcliffe, P. M. (2006). Skyrmions with massive pions. Phys. Rev. C
73, 055205.
Battye, R. A., Houghton, C. J., and Sutcliffe, P. M. (2003). Icosahedral Skyrmions.
J. Math. Phys. 44, 3543.
Battye, R. A., Krusch, S., and Sutcliffe, P. M. (2005). Spinning Skyrmions and the
Skyrme parameters. Phys. Lett. B 626, 120.
Battye, R. A., Manton, N. S., and Sutcliffe, P. M. (2007). Skyrmions and the α-particle
model of nuclei. Proc. R. Soc. A 463, 261.
Bauhoff, W., Schultheis, H., and Schultheis, R. (1984). Alpha cluster model and the
spectrum of 16 O. Phys. Rev. C 29, 1046.
Blatt, J. M. and Weisskopf, V. F. (1952). Theoretical Nuclear Physics. Wiley, New York,
p. 292.
Braaten, E. and Carson, L. (1998). Deuteron as a toroidal Skyrmion. Phys. Rev. D 38,
3525.
Brink, D. M., Friedrich, H., Weiguny, A., and Wong, C. W. (1970). Investigation of the
alpha-particle model for light nuclei. Phys. Lett. B 33, 143.
Carson, L. (1991). B = 3 nuclei as quantized multi-Skyrmions. Phys. Rev. Lett. 66,
1406.
Dennison, D. M. (1954). Energy levels of the 16 O nucleus. Phys. Rev. 96, 378.
Finkelstein, D. and Rubinstein, J. (1968). Connection between spin, statistics and kinks.
J. Math. Phys. 9, 1762.
Forest, J. L., Pandharipande, V. R., Pieper, S. C., Wiringa, R. B., Schiavilla, R.,
and Arriaga, A. (1996). Femtometer toroidal structures in nuclei. Phys. Rev. C 54,
646.
Friedrich, H., Satpathy, L., and Weiguny, A. (1971). Why is there no rotational band
based on the 7.65 MeV 0+ state in 12 C?. Phys. Lett. B 36, 189.
Giulini, D. (1993). On the possibility of spinorial quantization in the Skyrme model.
Mod. Phys. Lett. A 8, 1917.
Hitchin, N. J., Manton, N. S., and Murray, M. K. (1995). Symmetric monopoles.
Nonlinearity 8, 661.
Houghton, C. J. and Magee, S. (2006). A zero-mode quantization of the Skyrmion.
Phys. Lett. B 632, 593.
Houghton, C. J., Manton, N. S., and Sutcliffe, P. M. (1998). Rational maps, monopoles
and Skyrmions. Nucl. Phys. B 510, 507.
Irwin, P. (2000). Zero mode quantization of multi-Skyrmions. Phys. Rev. D 61,
114024.
Jarvis, S. (2000). A rational map for Euclidean monopoles via radial scattering. J. Reine
Angew. Math. 524, 17.
298 Skyrmions and nuclei
Krusch, S. (2003). Homotopy of rational maps and the quantization of Skyrmions. Ann.
Phys. 304, 103.
Krusch, S. (2006). Finkelstein–Rubinstein constraints for the Skyrme model with pion
masses. Proc. R. Soc. A 462, 2001.
Manko, O. V. and Manton, N. S. (2007). On the spin of the B = 7 Skyrmion. J. Phys.
A 40, 3683.
Manko, O. V., Manton N. S., and Wood, S. W. (2007). Light nuclei as quantized
Skyrmions. Phys. Rev. C 76, 055203.
Manton, N. S. (1994). Skyrmions and their pion multipole moments. Acta Phys. Pol.
B 25, 1757.
Manton, N. S. and Piette, B. M. A. G. (2001). Understanding Skyrmions using rational
maps. Prog. Math. 201, 469.
Manton, N. and Sutcliffe, P. (2004). Topological Solitons. Cambridge University Press,
Cambridge, UK, Chapter 9.
Manton, N. S. and Wood, S. W. (2006). Reparametrizing the Skyrme model using the
lithium-6 nucleus. Phys. Rev. D 74, 125017.
Manton, N. S., Schroers, B. J., and Singer, M. A. (2004). The interaction energy of
well-separated Skyrme solitons. Commun. Math. Phys. 245, 123.
Morinaga, H. (1956). Interpretation of some of the excited states of 4n self-conjugate
nuclei. Phys. Rev. 101, 254.
Robson, D. (1979). Evidence for the tetrahedral nature of 16 O. Phys. Rev. Lett. 42,
876.
Rössler, U. K., Bogdanov, A. N., and Pfleiderer, C. (2006). Spontaneous Skyrmion
ground states in magnetic metals. Nature 442, 797.
Segal, G. (1979). The topology of the space of rational maps. Acta Math. 143, 39.
Skyrme, T. H. R. (1961). A nonlinear field theory. Proc. R. Soc. A 260, 127.
Skyrme, T. H. R. (1962). A unified field theory of mesons and baryons. Nucl. Phys. 31,
556.
Von Oertzen, W., Freer, M., and Kanada-En’yo, Y. (2006). Nuclear clusters and nuclear
molecules. Phys. Rep. 432, 43.
Walhout, T. S. (1992). Quantizing the four-baryon Skyrmion. Nucl. Phys. A 547, 423.
Williams, J. G. (1970). Topological analysis of a nonlinear field theory. J. Math. Phys.
11, 2611.
Wuosmaa, A. H., Betts, R. R., Freer M., and Fulton, B. R. (1995). Recent advances in
the study of nuclear clusters. Ann. Rev. Nucl. Part. Sci. 45, 89.
XV
MIRROR SYMMETRY OF FOURIER–MUKAI
TRANSFORMATION FOR ELLIPTIC CALABI–YAU
MANIFOLDS
15.1 Introduction
Mirror symmetry conjecture says that for any Calabi–Yau (CY) manifold M
near the large complex/symplectic structure limit, there is another CY manifold
X, called the mirror manifold, such that the B-model superstring theory on
M is equivalent to the A-model superstring theory on X, and vice versa.
Mathematically speaking, it roughly says that the complex geometry of M is
equivalent to the symplectic geometry of X, and vice versa. It is conjectured
(Strominger et al. 1996) that this duality can be realized as a Fourier-type
transformation along fibers of special Lagrangian fibrations on M and X, called
the SYZ mirror transformation F SY Z .
Suppose M has an elliptic fibration structure,
p : M → S,
then there is another manifold W with a dual elliptic fibration over S. Since
any elliptic curve is isomorphic to its dual, we have actually M ∼ = W provided
that p has irreducible fibers. There is also a Fourier–Mukai transformation (FM
transform) FcxFM
between the complex geometries of M and W . On the level of
derived category of coherent sheaves, Fcx FM
is an equivalence of categories. On
the level of cycles, this can be described as a spectral cover construction and it
is a very powerful tool in the studies of holomorphic vector bundles over M .
In this chapter we address the following two questions: (i) What is the SYZ
transform of the elliptic fibration structure on M ? (ii) What is the SYZ transform
of the FM transform Fcx FM
?
The answer to the first question is a twin Lagrangian fibration structure on
the mirror manifold X, coupled with a superpotential. To simplify the matter, we
will ignore the superpotential in our present discussions. Similarly there is a twin
Lagrangian fibration structure on the mirror manifold Y to W . We will explain
several important properties of twin Lagrangian fibrations. In particular, we show
that the twin Lagrangian fibration on Y is dual to the twin Lagrangian fibration
300 Mirror symmetry of Fourier–Mukai transformation
Similarly the mirror of the complex cycle M itself is a special Lagrangian section
for X → B. This heuristic reasoning is only expected to hold true asymptotically
near the large complex structure limit (LCSL) (Strominger et al. 1996). We can
apply this SYZ transform to other coherent sheaves on M . For example, suppose
302 Mirror symmetry of Fourier–Mukai transformation
namely, by F ∗ . On the mirror side, it says that given any Lagrangian fiber p−1 (c)
to p : X → C, it should intersect an one real parameter family of π −1 (b)’s where
b ∈ B.
Furthermore if m ∈ M and S1 , S2 ∈ W satisfying Ext∗OM (S1 , Om ) = 0 =
Ext∗OM (S2 , Om ), then m ∈ F1 = F2 with Fi = SuppSi . On the mirror side, this
means that if π −1 (b) ∩ p−1 (ci ) = 0 for i = 1, 2 then π(p−1 (c1 )) = π(p−1 (c2 )).
Similarly if π−1 (bi ) ∩ p−1 (c) = 0 for i = 1, 2 then p(π −1 (b1 )) = p(π −1 (b2 )).
Namely, π(p−1 (c))’s in B form a fibration over some space D, which is also
the base space of a fibration on C given by p(π −1 (b))’s. That is,
X→C
↓ ↓
B →D
We might also expect that these fibrations on B and C over D are both affine
fibrations. Such a structure on X will be called a twin Lagrangian fibration on
X. On the first sight, it seems that these two Lagrangian fibrations on X are
on equal footing. But these arguments are only valid outside singular fibers
of M → S. Recall that these two Lagrangian fibrations on X are mirror to
holomorphic fibrations id : M → M and p : M → S. The two base manifolds are
quite different in nature: M is CY but S is not.
The Lagrangian fibers to X → C are mirror to the smooth elliptic curve fibers
of M → S. The situation near a singular elliptic curve fiber could be quite
different. Their locus in S, called the discriminant locus D, causes S fails to be
CY because of the formula KS−1 = 12 1
D. In particular KS−1 is an effective divisor
on S, which is indeed ample in many cases, namely S, is a Fano manifold.
There is a version of the mirror symmetry conjecture for Fano manifolds, and
their mirror involve Lagrangian fibrations together with a holomorphic function,
called the superpotential. It is reasonable to expect that the Lagrangian fibration
X → C should also interact with this superpotential corresponding to S. We
hope to come back to further discuss this issue in the future.
15.2.4.1 Large complex structure limits
Mirror symmetry for M is expected to work well near the LCSL. In terms of
Hodge theory, it means that M is a member of an one-parameter family of CY
n-folds Mt with 0 < |t| < 1 such that the monodromy operator T : H n (M ) →
H n (M ) is of maximally unipotent, that is, N = log(I − T ) satisfies N n = 0
but N n+1 = 0. On the mirror side (Deligne 1997), this corresponds to the hard
Lefschetz action L = ∧ωX : ⊕H p,p (X) → ⊕H p,p (X) which satisfies Ln = 0 but
Ln+1 = 0.
We now assume that M has an elliptic fibration structure. In our above
discussions, we need to require each member Mt in the family also have an
elliptic fibration. When n ≥ 3 the existence of an elliptic fibration is invariant
under deformations of complex structures. Wilson (1997) showed that if M is
304 Mirror symmetry of Fourier–Mukai transformation
P = O (∆ − σM × W − M × σW ),
Besides working on the level of derived categories, we can also study the FM
transform of a stable bundle, the so-called spectral cover construction (Donagi
1997, 1998; Friedman et al. 1997, 1999). The basic idea is any stable bundle over
an elliptic curve is essentially a direct sum of line bundles. In the family situation,
a stable bundle on M gives a multi-section for p : W → S, together with a line
bundle over it. Such construction is important in describing the moduli space of
stable bundles and it can also be generalized to construct principal G-bundles
on M , which play an important role in the duality between F-theory and String
theory (Friedman et al. 1997).
Lagrangian cycles are the objects that form the sophisticated Fukaya–Floer
category F uk (X), where morphisms are Floer homology groups which counts
holomorphic disks bounding cycles of Lagrangian submanifolds. We could also
generalize the notion of Lagrangian cycles to allow L to be a stratified Lagrangian
submanifold and to allow L to be a higher rank flat bundle.
Suppose (X, ωX ) and (Y, ωY ) are symplectic manifolds of dimensions 2m and
2n, respectively. Then (X × Y, ωX − ωY ) is again a symplectic manifold. Given
any Lagrangian cycle (P, P) ∈ C (X × Y ) we can construct a Fourier-Mukai type
transformation, or simply symplectic FM transform, defined as follow (Weinstein
1979):
Fsym
FM
: C (X) → C (Y )
Fsym
FM
(L, L) = L̂, L̂ .
P is locally determined by
P : xi = xi (u, y) and vk = v k (u, y) for all i, k.
Then (L × Y ) ∩ P ⊂ X × Y is locally given by
⎧ i
⎪
⎪ x = xi (u, 0) , y i = 0,
⎨
(L × Y ) ∩ P :
⎪
⎪ uk = uk , v k = vk (u, 0) .
⎩
(i.e. no restriction)
Therefore its projection to Y becomes
L̂ : uj = uj v k = vk (u, 0) .
The Lagrangian condition for P ⊂ X × Y is
∂xi ∂xj ∂v k ∂v l
+ =0 and + = 0,
∂y j ∂y i ∂ul ∂uk
for any (u, y) and for all i, j = 1, ..., m and k, l = 1, ..., n. By setting y = 0 for the
second equation, this implies that L̂ is a Lagrangian submanifold in Y .
To obtain the flat bundle L̂ over L̂, in the generic case, we can identify L̂ ⊂ Y
with (L × Y ) ∩ P ⊂ X × Y . The flat bundle L on L induced one on L̂ by pullback
to L × Y and then restrict to L̂. By restriction, P also determine a flat connection
on L̂. The tensor product of these two flat bundles is defined as L̂ over L̂. More
work will be needed to handle the pushforward in the non-generic case though.
Examples of Lagrangian cycles in products of symplectic manifolds include
(i) the graph of any symplectic map f : X → Y , that is, f ∗ ωy = ωX ; (ii) if M
(respectively N ) is a Lagrangian submanifold in X (respectively Y ), then P =
M × N is obviously a Lagrangian submanifold in X × Y ; and (iii) let C be any
coisotropic submanifold in (X, ωX ). It induces a canonical isotropic foliation on
C and such that its leaf space C/∼ has a natural symplectic structure, provided
that it is smooth and Hausdorff. This is called the symplectic reduction. Then the
natural inclusion C ⊂ X × (C/∼) is a Lagrangian submanifold in the product
symplectic manifold. Thus we can use this to obtain a symplectic FM transform
Fsym
FM
: C (X) → C (C/ ∼).
where Tvert X is the vertical tangent bundle. Using this and the Lagrangian
condition, we have a canonical identification between Tvert X and the pullback
cotangent bundle:
Tvert X ∼
= π ∗ T ∗ B.
C∼
= B∗
C → V → V ∗ B ∗ .
ω
V →C
↓ ↓
B→D
The fibers of the columns (respectively, rows) are Tb and Tb /Tbc (respectively,
Tc and Tc /Tbc ). The Lagrangian conditions implies the existence of a natural
isomorphism Tc /Tbc ∼
∗
= (Tb /Tbc ) , namely, the two affine bundles B → D and
C → D are fiberwise dual to each other.
The usual SYZ transform which switches the fibers of a Lagrangian fibration
V → B to their duals will interchange complex geometry and symplectic geome-
try. In order to stay within the symplectic geometry, we should take the fiberwise
dual to both Lagrangian fibrations V → B and V → C.
Taking dual to both Tb and Tc has the same effect as taking dual
to (Tb + Tc ) /Tbc while keeping Tbc fixed. This gives us the following new
Symplectic FM transform and twin Lagrangian fibrations 309
commutative diagram:
U → C
↓ ↓
B → D
Here B (respectively C ) is the total space of taking fiberwise dual to the
fibration B → D (respectively C → D). The fiber Tb of U → B is obtained by
taking dual along the base of the fibration Tb → Tb /Tbc . This is the same as
taking fiberwise dual, up to conjugation with the duality of total spaces. The
∗
fiber of B → D is (Tc /Tbc ) and likewise for C → D.
A more intrinsic way to describe this double dual process is as follow: The
fiber of V → D is a coisotropic subspace in V , say Vd . The symplectic reduction
Vd /∼ (see e.g. Weinstein 1979) is another symplectic vector space. Then U is
obtained by replacing Vd / ∼ by its dual symplectic space from
V.
In terms of an explicit coordinate system on V given by xi , xα , yi , yα with
1 ≤ i ≤ n − q and n − q + 1 ≤ α ≤ n, then we have
y i , yα y
V →C → αi
xα , xi x
↓ ↓ with coordinates
↓ ↓
B→D
xα , xi → xi
and
yi , yα∗ yα∗
U → C α∗ i →
x ,x xi
↓ ↓ with coordinates
↓ ↓
B → D
xα∗ , xi → xi
The Lagrangian conditions actually imply that B ∼
= C, C ∼
= B, and U ∼
=V.
We now return back to the general symplectic manifolds situation.
15.4.2.3 Twin Lagrangian fibrations
First we recall that every fiber of a Lagrangian fibration, say π : X → B, has a
natural affine structure.
Definition 15.2 Let (X, ω) be a symplectic manifold of dimension 2n and
π : X → B and p : X → C are two Lagrangian fibrations on X with Lagrangian
sections. We call this a twin Lagrangian fibration of index q if for general b ∈ B
and c ∈ C with p−1 (c) ∩ π −1 (b) nonempty, then it is an affine subspace of π −1 (b)
π p
of codimension q. We denote such a structure as B ← X → C.
Since π −1 (b) is the union of affine subspaces p−1 (c) ∩ π −1 (b), they form an
affine foliation of π −1 (b) of codimension q. In the above definition we assume that
when p−1 (c) ∩ π −1 (b) is nonempty, then it is a codimension q affine submanifold
of π −1 (b). We did not assume that this affine structure is compatible with the
one on p−1 (c) which comes from the other Lagrangian fibration p. Nevertheless,
310 Mirror symmetry of Fourier–Mukai transformation
under suitable assumptions, we will show that this is indeed the case and the
above definition is symmetric with respect to B and C. Furthermore we have a
commutative diagram of affine morphisms:
X→C
↓ ↓.
B →D
π p
Claim 15.2 Suppose that B ← X → C is a twin Lagrangian fibration. Then
B admits a natural (singular) foliation whose leaves are q-dimensional subspaces
π(p−1 (c))’s with c ∈ C.
Proof. Since p−1 (c) ∩ π −1 (b) is always of codimension k in p−1 (c) if nonempty,
π(p−1 (c)) is a q dimensional subspace in B. Suppose b is a smooth point in
π(p−1 (c)), we claim that its tangent space Tb (π(p−1 (c))) ⊂ Tb B is independent of
the choice of c. Assuming this, we obtain a q-dimensional (singular) distribution
on B. Furthermore π(p−1 (c))’s with c ∈ C are leaves of this distribution, thus
we have the required foliation on B.
To prove the claim, we recall that π −1 (b) has a natural affine structure and
π (b) admits an affine foliation whose leaves are p−1 (c) ∩ π −1 (b) with c ∈ C, by
−1
the assumption of a twin Lagrangian fibration. The key observation is these imply
that for any x ∈ p−1 (c) ∩ π −1 (b), under the natural identification Tx∗ (π −1 (b))
Tb B, the conormal bundle of p−1 (c) ∩ π −1 (b) in π −1 (b) at x is the same linear
subspace in Tb B of dimension q. Furthermore this coincides with π(Tx (p−1 (c))).
Hence the result.
We will assume that this foliation on B is indeed a fibration which we denote
as
p : B → D.
As a corollary of the above claim, we have the following immediate result:
π p
Corollary 15.1 Suppose that B ← X → C is a Lagrangian twin fibration.
Then we have a commutative diagram
X→C
↓ ↓
B →D
where p : B → D is the above fibration fibered by π p−1 (c) ’s on B.
π p
Claim 15.3 Suppose that B ← X → C is a twin Lagrangian fibration with
sections and π is a proper map. Then D has a natural affine structure and
p : B → D is an affine morphism.
Proof. Since π : X → B is a Lagrangian fibration, its general fiber π −1 (b) has
a natural affine structure, thus as an affine manifold π −1 (b) Rn /Λ for some
Symplectic FM transform and twin Lagrangian fibrations 311
This implies that, outside the singular locus of p : B → D, there exists a rank q
ε
vector bundle Rq → E → D with affine gluing functions, together with a multi-
section EZ such that p is affine isomorphic to the projection E/EZ → D. In
particular B = E/EZ .
Given a general x ∈ X with b = π (x), c = p (x), and d = p (b) = π (b), we write
Xd = p−1 (c) ∩ π −1 (b). From π and p both being Lagrangian fibrations on X,
we have
∗ ∗
0 → NX d /π
−1 (b),x → Tb B → Tx Xd → 0
and
∗ ∗
0 → NX d /p
−1 (c),x → Tc C → Tx Xd → 0.
As we have shown earlier NXd /p−1 (c),x can be identified with the fiber of E →
∗
D over d ∈ D, denoted as Ed . Moreover the exact sequence 0 → NX d /π
−1 (b),x →
∗ ∗ ∗
Tb B → Tx Xd → 0 is equivalent to 0 → ε E → T E → ε T D → 0 for the vector
bundle ε : E → D. This implies that Tc C is naturally isomorphic to the tangent
space of E ∗ at d.
Notice that the affine
structure
of the fibers of the Lagrangian fibration p :
X → C is given by T p−1 (c) p∗ (Tc∗ C). When p is proper, these data also
determine an affine structure on C. Thus having such a natural identification
between Tc C and Td (E ∗ ), all the affine structures involved are compatible with
each other. Such a claim can be checked by a direct diagram chasing method.
Thus we have obtained the following result:
π p
Claim 15.4 Suppose B ← X → C is a twin Lagrangian fibration with π and
p π
p proper. Then C ← X → B is also a twin Lagrangian fibration and
X→C
↓ ↓
B →D
is a commutative diagram of affine morphisms.
312 Mirror symmetry of Fourier–Mukai transformation
fibers. Such a Y is called the dual twin Lagrangian fibration to X. It is clear that
the dual twin Lagrangian fibration to Y is X again.
As we have explained in the linear situation, the fibration B → D (respectively
C → D) is the dual fibration to B → D (respectively C → D). Furthermore the
Lagrangian conditions imply that there are natural identifications B ∼ = C and
C ∼= B and also
X∼
= Y.
It is interesting to know whether this identification will continue to hold true if
superpotentials are also included in our discussions.
FM FM
Since X ∼ = Y, we choose the Lagrangian cycle Psym , Psym on X × Y given
by the diagonal Lagrangian submanifold together with the trivial flat bundle
over it. We call this the Lagrangian Poincaré cycle on X × Y . The symplectic
FM transform Fsym FM
for the twin Lagrangian fibration on X and its dual twin
Lagrangian fibration on Y is defined using this Lagrangian cycle:
Fsym
FM
: C (X) → C (Y )
Fsym
FM
(L, L) = L̂, L̂ .
µ = (µ1 , µ2 , µ3 ) : X → Im H = R3
µ : X → Rn ⊗ R3 .
p : B → D.
SYZ transformation of FM transform 315
Since the slits for all the singular points in the middle triangle in ∂∇ are
vertical with respect to p, we obtain an affine fibration around there. But p will
cease to be an affine fibration around other singular points of B. However if
we choose the defining function f for M appropriately, then we can arrange all
other singular points to be far away from this middle triangle. Thus we obtain an
affine fibration on a large portion of B. As a matter of fact, if we allow M to be
singular, then we can make p to be an affine fibration on the whole B\Sing (B).
This picture can be generalized to certain higher dimensional manifolds, for
instance CY hypersurfaces in toric varieties.
W ←
Mirror
−−−−−−manifolds
−−−−−−→ Y
Indeed our arguments only work in the large structure limit but we expect that
a modified version of such will work in more general situations. To be precise,
we are only dealing with the flat situation: We assume that M is CY n-fold
with a q-dimensional Abelian varieties fibration p : M → S which is compatible
with a Lagrangian fibration π : M → B in the sense that the restriction of π to
any smooth Abelian variety fiber of p determines a Lagrangian fibration on the
Abelian variety. We are going to impose a very restrictive assumption, which we
expect to be the first-order approximation of what happens at the large structure
limit. Namely, M is the product of an Abelian variety with S.
316 Mirror symmetry of Fourier–Mukai transformation
1 n
Let√ z , ..., z be a local complex coordinate system on M such that z =
x + −1y with y 1 , ..., yn (respectively x1 , ..., xn ) be an affine coordinate system
on fibers (respectively base) of the special Lagrangian fibration on M → B.
According to the SYZ proposal, the mirror manifold X to M is the total space
of the dual Lagrangian torus fibration. For the affine coordinate system y 1 , ..., yn
on the torus fiber, we denote the dual coordinate system on its dual torus as
y1 , ..., yn . Thus X has a Darboux coordinate system given by y1 , ..., yn , x1 , ..., xn .
We assume the q-dimensional Abelian variety fibration M → S is compatible
with the Lagrangian fibration structure in the sense that z n−q+1 , ..., z n gives
a complex affine coordinate system on fibers to M → S. In particular, the
dual coordinate system on the dual Abelian variety is given by zn−q+1 , ..., zn .
Therefore such a coordinate system on M induces a similar coordinate system on
the dual Abelian variety fibration W → S, labeled (z 1 , ..., z n−q , zn−q+1 , ..., zn ).
Note that on a fiber of the Abelian variety fibration M → S, yn−q+1 , ..., yn
(respectively, xn−q+1 , ..., xn ) are those coordinates belonging to Lagrangian
fibers (respectively, base) for the special Lagrangian fibration on M . Since fibers
are much smaller in scale when comparing to the base near the LCSL (Strominger
et al. 1996), when we take the dual Abelian variety, yn−q+1 , ..., yn will become
much larger in scale when comparing to xn−q+1 , ..., xn . As a consequences,
the special Lagrangian fiber (respectively, base) coordinates for the CY man-
ifold Y are y 1 , ..., y n−q ,xn−q+1 , ..., xn (respectively, x1 , ..., xn−q , yn−q+1 , ..., yn ).
Therefore the special Lagrangian fibration on the mirror manifold Y to W has
fiber (respectively, base) coordinates as y1 , ..., yn−q , xn−q+1 , ..., xn (respectively,
x1 , ..., xn−q , yn−q+1 , ..., yn ).
Next we need to check that Y is indeed the total space of the dual twin
Lagrangian fibration to X. For the Abelian variety fiber in M , with coordi-
nates xn−q+1 , ...xn , y n−q+1 , ...y n , its mirror Lagrangian cycle in X has coor-
dinates xn−q+1 , ...xn , y1 , ...yn−q and it is the fiber of the other Lagrangian
fibration on X. Similarly, for the Abelian variety fibers in W , with coordinates
xn−q+1 , ...xn , yn−q+1 , ...yn , its mirror Lagrangian cycle in Y has coordinates
y1 , ...yn−q , yn−q+1 , ...yn and it is the fiber of the other Lagrangian fibration on
Y . It is now clear that X and Y are dual twin Lagrangian fibrations. All these
fiber/base coordinates systems are summarized in the following diagram:
yi yα y yi
M: → X : αi
←
xi xα x xα
) )
yi xα xα yi
W : i ←
→ Y : i
x yα x yα
Our conventions are i = 1, ..., n − q and α = n − q + 1, ..., n. For each space in the
above diagram, the coordinates in the top (respectively, bottom) row are for the
SYZ special Lagrangian fibers (respectively, base). Also the coordinates in the left
SYZ transformation of FM transform 317
(respectively, right) column are for the fibers (respectively, base) of the Abelian
variety fibrations for M or W and the other Lagrangian fibrations for X or Y .
Proof. We will continue to use the same coordinate systems as before. The
Poincaré sheaf Pcx for the dual Abelian variety fibrations
M → S and W → S
has support M ×S W ⊂ M × W which has coordinates xi , y i , xα , y α , xα , yα i,α
(where xi and y i are diagonal coordinates in the product space). Over M ×S
W , Pcx is a complex line bundle, indeed only a divisorial sheaf, with an U (1)
connection:
√
FM
Dcx = d + −1 α (xα dxα + xα dxα − yα dy α − y α dyα ) .
The reason that the signs for terms involving x and y are different is the following:
For Abelian varieties, the dual complex coordinates are z α and zα , which induces
the dual coordinates for xα , y α as xα , −yα because Re (z α zα ) = xα xα − y α yα .
To apply the SYZ mirror transform to Pcx from the special Lagrangian fibra-
tion M × W → B × B ∗ to the one X × Y → B × B ∗ , we first need to describe
318 Mirror symmetry of Fourier–Mukai transformation
(M × W ) × (X × Y ) ⊂ (M × W ) × (X × Y ) .
B×B ∗
To describe the coordinates on this space, we need to rename the (x, y) coordi-
nates on W and Y to (u, v) coordinates. That is,
W Y
v i uα uα vi
SYZ
←−−−mirror
−−−→
ui vα ui v α
C →P SY Z → (M × W ) × (X × Y )
B×B ∗
is given by
√ i
DSY Z = d + −1 y dyi + yi dy i + y α dyα + yα dy α
√ i
− −1 v dvi + vi dv i + uα duα + uα duα .
Note that we have used different signs for the SYZ transform between M and X
and SYZ transform between W and Y .
In this newly named coordinates on W , the universal connection on the
Poincaré bundle
C → Pcx
FM
→ M ×S W
is given by
√
Dcx
FM
=d− −1 α (uα dxα + xα duα − vα dy α − y α dvα ) .
To apply the SYZ transform, we need to first pullback the bundle Pcx FM
(with
its connection Dcx ) from M ×S W ⊂ M × W to (M × W ) ×B×B (X × Y ) ⊂
FM ∗
Part (i) with i = 1, ..., n − q: This part is easy because the Poincaré bundle for
the
i FM itransform
does
not involve here. In fact it is the trivial line bundle over
u = x ∩ v i = y i ⊆ M × W . In these coordinates, the SYZ Poincaré bundle
P SY Z has support (M × W ) ×B×B ∗ (X × Y ) ⊂ (M × W ) × (X × Y ), which in
our coordinate systems means the xi ’s coordinates for M and X are the same
SYZ transformation of FM transform 319
Part (ii) with α = n − q + 1, ..., n: In these coordinates, the support for the SYZ
Poincaré bundle P SY Z imposes a constraint which says that the xα ’s coordinates
for M and X are the same and the vα ’s coordinates for W and Y are the same.
Now we need to integrate the directions y α ’s and uα ’s. First we rearrange the
terms
√
Dcx
FM
⊗ DSY Z = d − −1 α (uα dxα + xα duα − vα dy α − y α dvα )
√
+ −1 α (y α dyα + yα dy α − uα duα − uα duα )
√
= d + −1 α ((−vα + yα ) dy α + (xα − uα ) duα )
√
+ −1 α (uα dxα − y α dvα + y α dyα − uα duα ) .
When we integrate along the y α ’s and uα ’s directions, the terms uα dxα , y α dvα ,
y dyα , and uα duα has non-trivial Fourier modes in these variables and therefore
α
Fsym
FM
◦ F SY Z = F SY Z ◦ Fcx
FM
Complex Symplectic
(M ) F SY Z
geometry ←−−−→ geometry (X)
L L
⏐ ⏐
⏐
FM +
⏐ FM
Fcx + Fsym
Complex Z Symplectic
F−SY
(W ) ← −−→ geometry (Y )
geometry
Since Fsym
FM
is essentially an identity transformation, we can regard the FM
transform for elliptic CY as a square of the SYZ transforms! Note that even
though one can identify M with W and X with Y , nevertheless, F(M,X) SY Z
and
F(W,Y ) do not correspond to each other under these identifications of spaces. For
SY Z
Equivalently,
(πX )∗ P(M,X)
SY Z
⊗ Psym
FM
= (πW )∗ Pcx
FM
⊗ P(W,Y
SY Z
),
Conclusions and discussions 321
over M × Y . However,
Psym
FM
= (πM )∗ (πW )∗ Pcx
FM
⊗ P(M
SY Z
×W,X×Y )
Hence
(πX )∗ P(M,X)
SY Z
⊗ Psym
FM
= (πX )∗ P(M,X)
SY Z
⊗ (πM )∗ (πW )∗ Pcx
FM
⊗ P(M,X)
SY Z
⊗ P(W,Y
SY Z
)
= (πW )∗ Pcx
FM
⊗ P(W,Y
SY Z
) ⊗ (πM )∗ (πX )∗ P(M,X) ⊗ P(M,X)
SY Z SY Z
= (πW )∗ Pcx
FM
⊗ P(W,Y
SY Z
).
The last equality holds because SYZ transforms are involutive. Hence the
result.
Acknowledgments
The second author has known Nigel for almost forty years, having met him at
the Institute of Advanced Study. In those days, we had many dinners together,
and I (S.-T. Yau) always enjoyed talking to him. His many original ideas in
mathematics influenced me tremendously. This is especially the case for K3
surfaces, in which his deep insight affected my thinking of the Calabi–Yau
conjecture.
The research of N.C. Leung is partially supported by RGC Earmarked grants
of Hong Kong. The work of S.-T. Yau is supported in part by NSF grants DMS-
0354737, DMS-0306600, and DMS-0628341.
References
Andreas, B., Curio, G., Ruiperez, D. H. and Yau, S.-T. (2001). Fibrewise T-duality for
D-branes on elliptic Calabi-Yau. J. High Energy Phys. No. 3, Paper 20, 13 pp.
Bridgeland, T. and Maciocia, A. (2002). Fourier-Mukai transforms for K3 and elliptic
fibrations. J. Algebraic Geom. 11(4), 629–57.
Candelas, P., de la Ossa, X. C., Green, P. S. and Parkes, L. (1991). A pair of Calabi-Yau
manifolds as an exactly soluble superconformal theory. Nucl. Phys. B 359, 21–74.
Deligne, P. (1997). Local behavior of Hodge structures at infinity. In Mirror Symmetry,
II, pp. 683–699, AMS/IP Studies in Advanced Mathematics, Vol. 1, American
Mathematical Society, Providence, RI.
Donaldson, S. K. (1985). Anti self-dual Yang-Mills connections over complex algebraic
surfaces and stable vector bundles. Proc. London Math. Soc. (3) 50(1), 1–26.
Donagi, R. (1998). Taniguchi lectures on principal bundles on elliptic fibrations. In
Integrable Systems and Algebraic Geometry (Kobe/Kyoto, 1997), pp. 33–46, World
Scientific Publishing, River Edge, NJ.
Donagi, R. (1997). Principal bundles on elliptic fibrations. Asian J. Math. 1(2), 214–23.
Donagi R. and Pantev, T. (2008). Torus fibrations, gerbes, and duality. With an
appendix by Dmitry Arinkin. Mem. Amer. Math. Soc. 193(901), vi–90.
Friedman, R., Morgan, J. and Witten, E. (1997). Vector bundles and F theory.
Commun. Math. Phys. 187(3), 679–743.
Friedman, R., Morgan, J. and Witten, E. (1999). Vector bundles over elliptic fibrations.
J. Algebraic Geom. 8(2), 279–401.
Givental, A. (1996). Equivariant Gromov-Witten invariants. Int. Math. Res. Notices,
No. 13, 613–63.
Gross M. and Siebert, B. (2003). Affine manifolds, log structures, and mirror symmetry.
Turkish J. Math. 27(1), 33–60.
Hu, Y., Liu, C. H. and Yau, S.-T. (2002). Toric morphisms and fibrations of toric
Calabi-Yau hypersurfaces. Adv. Theor. Math. Phys. 6(3), 457–506.
Kontsevich, M. (1995). Homological algebra of mirror symmetry. In Proceedings of the
1994 International Congress of Mathematicians, Vol. I, Birkäuser, Zürich, p. 120.
Kontsevich, M. and Soibelman, Y. (2006). Affine structures and non-Archimedean
analytic spaces. In The Unity of Mathematics, pp. 321–385, Progress of Mathematics,
Vol. 244, Birkhauser Boston, Boston, MA.
References 323
Tamás Hausel
To Nigel Hitchin for his 60th birthday
16.1 Introduction
In this chapter we survey the motivations, related results, and progress made
towards the following problem, raised by Hitchin in 1995:
Problem 16.1 What is the space of L2 harmonic forms on the moduli space
of Higgs bundles on a Riemann surface?
The moduli space MdDol (SLn ) of stable rank n Higgs bundles with fixed
determinant of degree d on a Riemann surface was introduced and studied
in Hitchin (1987), Nitsure (1991), and Simpson (1991). The Betti numbers of
this space for n = 2 were determined in Hitchin (1987b) while for n = 3 in
Gothen (1994). The above problem raised two new directions to study. First
is the Riemannian geometry of MdDol (SLn ), or more precisely the asymptotics
of the natural hyperkähler metric, and its connection with Hodge theory. The
second one, which can be considered the topological side of Problem 16.1,
is to determine the intersection form on the middle-dimensional compactly
supported cohomology of MdDol (SLn ). While the first question seems still
out of reach, although we will report on some modest progress below, the
second is more approachable and we offer a conjecture at the end of this
survey.
Problem 16.1 was motivated by S-duality conjectures emerging from the string
theory literature about Hodge theory on certain hyperkähler moduli spaces,
which are close relatives of MdDol (SLn ).
In the physics literature S-duality stands for a strong–weak duality between two
quantum field theories. The interest from the physics point of view is that it gives
a tool to study physical theories with a large coupling constant via a conjectured
equivalence with a theory with a small coupling constant where perturbative
methods give a good understanding. The S-duality conjecture relevant for us
is based on the Montonen–Olive electromagnetic duality proposal from 1977 in
four-dimensional Yang–Mills theory (Montonen and Olive 1977). It was noted in
(Witten and Olive 1978) that this duality proposal is more likely to hold in a
supersymmetric version of the theory, and in Osborn (1979) it was argued that
N = 4 supersymmetry is a good candidate. Hyperkähler Hodge theory is relevant
Introduction 325
Conjecture 16.1 0 is
The dimension of the space of L2 harmonic forms on Mk
<
0 0 d = mid
dim Hd M k =
φ(k) d = mid,
where φ(k) = ki=1 δ1(i,k) is the Euler φ function, and mid = 2k − 2 is half of
0 .
the dimension of Mk
Similar S-duality arguments led Vafa and Witten (1994) to get a conjecture
on the space of L2 harmonic forms on a certain smooth completion Mφk,c1 ,
constructed in Kronheimer (1990) and Nakajima (1998), of the moduli space
of U (n) Yang–Mills instantons of first Chern class c1 , energy k, and framing
φ on one of Kronheimer’s ALE spaces, which are four-dimensional complete
hyperkähler manifolds, with an asymptotically locally Euclidean metric.
Conjecture 16.2 The dimension of the space of L2 harmonic forms on Mφk,c1
is
0 d = mid
dim Hd Mφk,c1 = k,c1 k,c1
mid
dim im Hcpt Mφ →H mid
Mφ d = mid,
is a modular form, which, as was speculated in Vafa and Witten (1994), might
be a consequence of S-duality.
This chapter will introduce the reader to various mathematical aspects of these
three problems and offer mathematical techniques and results relating to them.
326 S-duality in hyperkähler Hodge theory
The main result of Hitchin et al. (1987) is that the natural Riemannian metric
on the smooth points of this quotient is hyperkähler.
Now we list three important examples of this construction, where the original
hyperkähler manifold M and Lie group G are both infinite dimensional.
is finite. Write
A = A1 dx1 + A2 dx2 + A3 dx3 + A4 dx4
in a fixed gauge, where Ai ∈ Ω0 (R4 , ad(P )). Let G = Ω(R4 , Ad(P )) be the gauge
group of P . An element g ∈ G acts on A ∈ M by the formula g(A) = g −1 Ag +
g−1 dg, preserving the hyperkähler structure. One finds that the hyperkähler
1 Some colleagues even suggest, due to the success of this construction, that HyperKähLeR
µH (A) = 0 ⇔ FA = ∗FA
µH (A, φ) = 0 ⇔ FA = ∗dA φ
F (A) = −[Φ, Φ∗ ],
µH (A, Φ) = 0 ⇔
dA Φ = 0
are then equivalent with Hitchin’s self-duality equations. There are no solutions
of finite energy on R2 , but as the equations are conformally invariant, we can
replace R2 with a genus g compact Riemann surface C in the above definitions,
and define M(C, P ) = µ−1H (0)/G, the Hitchin moduli space, which has a natural
hyperkähler metric by construction. There are different ways to think about
this space with the different complex structures, which will be explained in
Section 16.5.2.
is just the forgetful map. (In the compact case these maps are isomorphisms,
which gives the Hodge theorem mentioned above.) Thus
∗
im(Hcpt (M ) → H ∗ (M )) (16.3)
0 ) = dim(H mid (M
χL2 (M 0 )) = φ(k).
k k
for g > 1. This thus gave the surprising result that there are no L2 harmonic
forms on M1Dol (SL2 ) plainly by topological reasons. The technique used in the
proof of (16.5) was imitating Kirwan’s proof (1992) of Mumford’s conjecture on
the cohomology ring of the moduli space of stable rank 2 bundles of degree 1 on
the Riemann surface C. Therefore the extension of (16.5) to higher rank Higgs
bundle moduli spaces MdDol (SLn ) was not straightforward.
The next advance towards Sen’s Conjecture 16.1 came in 2000. Hitchin (2000)
showed that Hd (M ) = 0 unless d = dim(M )/2 for a complete hyperkähler mani-
fold M of linear growth. Examples include all our hyperkähler quotients discussed
in this chapter. The proofs in Hitchin (2000) use techniques inspired by Jost and
Zuo’s extension (2000) of ideas of Gromov (1991). It is interesting to note that
some of the proofs in Hitchin (2000) also exploit the operators in hyperkähler
Hodge theory, which are relevant in N = 4 supersymmetry. Using the symmetries
of the Atiyah–Hitchin metric (Hitchin 2000) proves Sen’s conjecture for k = 2,
that up to a scalar the only L2 harmonic form on M 0 is Sen’s two-form.
2
A more topological approach was introduced in Hausel et al. (2004). Hausel
et al. (2004) proves for fibered boundary manifolds M
Hmid (M ) ∼ mid
= im IHm mid
(M )→IHm̄ (M ) , (16.6)
H j (X, C) = F 0 ⊇ F 1 ⊇ · · · ⊇ F m ⊇ F m+1 = 0.
We can define mixed Hodge numbers obtained from these two filtrations by the
following formula:
hp,q;j (X) := dimC GrpF Grp+q
W
H j (X)C . (16.7)
the mixed Hodge polynomial. By virtue of its definition it has the property that
the specialization
P (M ; t) = H(M ; 1, 1, t)
E(M ; x, y) = E(xy).
Applications of mixed Hodge theory 333
P H ∗ (M ) := Wn H n (M ) ⊂ H ∗ (M ),
Let Γ be a quiver (oriented graph) with vertex set I = {1, . . . , n} and edges
E ⊂ I × I. Let
v = (v1 , . . . , vn ), w = (w1 , . . . , wn ) ∈ NI
be two-dimensional vectors and Vi and Wi corresponding complex vector spaces,
that is, dim(Vi ) = vi and dim(Wi ) = wi . We define the vector spaces
9 9
Vv,w = Hom(Vt(a) , Vh(a) ) ⊕ Hom(Vi , Wi )
a∈E i∈I
with derivative
F
: gl(v) := gl(Vi ) → gl(Vv ).
i∈I
as an affine GIT quotient. Alternatively one can construct the manifold underly-
ing M(v, w) as a hyperkähler quotient of Vv,w × V∗v,w by the maximal compact
subgroup U (v) ⊂ GL(v). This shows that M(v, w) possesses a hyperkähler
metric. The holomorphic symplectic quotient we presented above is the one where
the arithmetic technique of Section 16.4 is applicable. Before we explain that,
let us recall the following fundamental theorem of Nakajima (1998) about the
cohomology of these Nakajima quiver varieties:
Theorem 16.2 Assume that the quiver Γ has no edge-loops. Then there is an
irreducible representation of the Kac–Moody algebra g(Γ) of highest weight w
on ⊕v H mid (M(v, w)). In particular the Weyl–Kac character formula gives the
middle Betti numbers of Nakajima quiver varieties. Furthermore the intersection
form on Hcmid (M(v, w)) is definite, thus χL2 (M(v, w)) equals the middle Betti
number of M(v, w).
Remark 16.1 When Γ is an affine Dynkin diagram M(v, w) could be identi-
fied with one of the spaces Mφk,c1 of certain Yang–Mills instantons on a ALE
Applications of mixed Hodge theory 335
space XΓ . In Kac (1990) he explains that the Weyl–Kac character formula for
an affine Dynkin diagram has certain modular properties. This was the line of
argument in Vafa and Witten (1994) that (16.1) is a modular form provided
Conjecture 16.2 holds.
Theorem 16.3 For any quiver Γ, the Betti numbers of the Nakajima quiver
varieties are given by the following generating function, with the notation as in
Hausel (2006, theorem 3):
M M
(i,j)∈E t−2λ
i ,λj
i∈I t−2λ
i ,(1wi )
Tv M
M M mk (λi )
v∈NI λ∈P(v) i∈I t−2λ
i ,λi
k j=1 (1−t2j )
−d(v,w) v
Pt (M(v, w))t T = M j ,
t−2λi ,λ
v∈NI Tv (i,j)∈E
M i i M M mk (λi )
i∈I t−2λ ,λ k j=1 (1−t2j )
v∈NI λ∈P(v)
(16.10)
AΓ (v, 0) = mv (16.11)
equals with the multiplicity of the weight v in the Kac–Moody algebra g(Γ).
This can be proved, as sketched above and announced in Hausel (2006), to be a
consequence of (16.10) and the above-mentioned results of Nakajima and Hua.
Remark 16.3 When the quiver is affine ADE and the RHS becomes an infinite
product (indications that this can happen are the infinite product in Hausel
2006, section 3 and the infinite products in the recent Sasaki) we could get
an alternative proof of the modularity of (16.1) in the Vafa–Witten S-duality
conjecture.
336 S-duality in hyperkähler Hodge theory
In the remaining part of this survey we will motivate and study another
application of the technique in Section 16.4, which will be less powerful as the
mixed Hodge structure will fail to be pure, but will also open new interesting
directions by the study of this more complicated mixed Hodge structure.
A−1 −1 −1 −1 2πid
1 B1 A1 B1 · · · Ag Bg Ag Bg = e Id //GLn
n
We will also consider the varieties MdDol (SLn ), MdDR (SLn ), and M0B (SLn ),
which can be defined by replacing GLn with SLn in the above definitions.
Moreover M0Dol (GL1 ), M0DR (GL1 ), and M0B (GL1 ) turn out to be Abelian
groups. Then M0Dol (GL1 ), M0DR (GL1 ), and M0B (GL1 ) will act on MdDol (GLn ),
MdDR (GLn ), and MdB (GLn ), respectively, by an appropriate form of ten-
sorization. Finally we denote the corresponding (affine GIT) quotients by
MdDol (P GLn ), MdDR (P GLn ) and MdB (P GLn ). In our case, when (d, n) = 1,
they will turn out to be orbifolds. For more details on the construction of these
varieties see Hausel (2005).
In the next section we explain the original motivation to consider the
E-polynomials of these three complex algebraic varieties. The motivation is
mirror symmetry, and most probably the same S-duality we discussed in the
Introduction in connection with the Hodge cohomology of the moduli spaces of
Yang–Mills instantons in four dimension and magnetic monopoles in three. S-
duality ideas relating to mirror symmetry for Hitchin spaces have appeared in
the physics literature (Bershadsky et al. 1995; Kapustin and Witten 2007).
MdDol (P
⏐ GLn ) MDol⏐(SLn )
d
⏐χP GL ⏐χSL
+ n + n
HP GLn ∼
= HSLn
the generic fibers of the Hitchin maps χP GLn and χSLn are dual Abelian
varieties.
where B e and B̂ d are certain gerbes on the corresponding Hitchin spaces and
the E-polynomials above are stringy E-polynomials for orbifolds twisted by the
relevant gerbe as defined in Hausel and Thaddeus (2003).
Morally, this conjecture should be related to the S-duality considerations of
Kapustin and Witten (2007) and in turn to the geometric Langlands programme
of Beilinson and Drinfeld (1995). However the lack of global analytical inter-
pretation of the mixed Hodge numbers (16.7) appearing in Conjecture 16.3
prevents a straightforward physical interpretation. Nevertheless the agreement
of certain Hodge numbers for Hitchin spaces for Langlands dual groups is an
interesting direction from a purely mathematical point of view. In particular,
if we change our focus from MDR and MDol to MB we will uncover some
surprising connections to the representation theory of finite groups of Lie type.
way to formulate it is to say that certain differences between the character tables
of P GLn (Fq ) and its Langlands dual SLn (Fq ) are governed by mirror symmetry.
This kind of consideration could be interesting because the character tables of
P GLn (Fq ) or more generally those of GLn (Fq ) have been known for a long time
starting with the work of Green (1955), while the character tables of SLn (Fq )
have just recently been completed (Bonnafé 2006; Shoji 2006). It is especially
enjoyable to follow the effect of the mirror symmetry proposal of Conjecture 16.4
by comparing the character tables of GL2 (Fq ) and SL2 (Fq ) first calculated a
hundred years ago by Jordan 1907 and Schur 1907.
H(MB (P GL2 ); x, y, t)
(q 2 t3 + 1)2g q 2g−2 t4g−4 (q2 t + 1)2g 1 q 2g−2 t4g−4 (qt + 1)2g
= + −
(q 2 t2 − 1)(q 2 t4 − 1) (q 2 − 1)(q 2 t2 − 1) 2 (qt2 − 1)(q − 1)
1 q 2g−2 t4g−4 (qt − 1)2g
− ,
2 (q + 1)(qt2 + 1)
where q = xy and the four terms correspond to the four types of irreducible
characters of GL(2, Fq ). When g = 3 this equals
be proved to follow from the master conjecture in Hausel et al. (in preparation),
which expresses the mixed Hodge polynomials of all the character varieties M̄µB
as a generating function generalizing the Cauchy formula for Macdonald poly-
2
nomials.
µIt also has the following consequence on the topological L cohomology
χL2 M̄B of (16.4).
Conjecture 16.6 The topological L2 cohomology of the manifold M̄µB is
given by
χL2 (M̄µB ) = 0, when g > 1 (16.13)
χL2 (M̄µB ) = 1, when g = 1 (16.14)
χL2 (M̄µB ) = mv , when g = 0, (16.15)
where mv is the multiplicity of the weight v in the Kac–Moody algebra g(Γ),
which are encoded by the Kac denominator formula for the star-shaped quiver Γ.
When g > 1 and the parabolic type is µ = ((n)), that is, we have only one
puncture with central conjugacy class, then one can identify M̄µB = MdB (P GLn ),
with some d such that (d, n) = 1. In this case (16.13) says that
χL2 MdB (P GLn ) = 0,
which appeared as (Hausel and Rodriguez-Villegas Conjecture 4.5.1). It follows
from the mirror symmetry Conjecture 16.3 that
d d
mid
Hcpt MB (SLn ) ∼ mid
= Hcpt MB (P GLn )
and then the intersection forms also agree. This and (16.15) then imply that
(16.5) holds for any n, that is, that the intersection form on the compactly
supported cohomology of MdB (SLn ) is trivial. This gives a conjectural answer
to the topological side of Problem 16.1.
When g = 1 the conjectured (16.14) follows from Conjecture 16.5 and the
observation that the coefficient of q in the A-polynomial AΓ (q) for a g = 1 comet-
shaped quiver Γ is always 1.
When g = 0 the varieties MµB = M̄µB coincide. Conjecture 16.5 then implies
that
χL2 (MµB ) = AΓ (v, 0).
Conjecture (16.15) is a combination of this and the equality AΓ (v, 0) = mv , that
is, Kac’s conjecture 1, in Kac (1983), which, as discussed in Remark 16.2, follows
from Theorem 16.3.
Finally one can define M̄µDol the moduli space of stable parabolic P GLn -
Higgs bundles with quasi-parabolic type µj ∈ P(n) and generic weights at the
jth puncture on the Riemann surface (Boden and Yokogawa 1996, Garcı́a-Prada
et al. 2007). Then one can prove that M̄µB is diffeomorphic to M̄µDol . Thus
Conjecture 16.6 also calculates the intersection form on the compactly supported
342 S-duality in hyperkähler Hodge theory
cohomology of the moduli space M̄µDol of stable parabolic P GLn -Higgs bundles
of any rank.
Example 16.1 Consider the genus 0 Riemann surface P1 with four punctures.
Consider the moduli space Mtoy of stable rank 2 parabolic Higgs bundles on
P1 , with generic parabolic weights on the full parabolic flag at the punctures
(see Boden and Yokogawa 1996). This is a complex surface and the intersection
form on Hc2 (Mtoy ) was discussed in Hausel (1998, example 2 for theorem 7.13).
Hc2 (Mtoy ) is five-dimensional but χL2 (Mtoy ) is only four. (The cohomology class
of the generic fiber of the Hitchin map is the one in the kernel.)
Mtoy is diffeomorphic to the character variety M̄µB where g = 0 and µ =
((1, 1),
(1,1), (1, 1), (1, 1)). Thus by Conjecture 16.6 we should be able to calculate
χL2 M̄µB in terms of the representation theory of the corresponding quiver Γ.
The corresponding quiver Γ in this case will be the affine D̃4 Dynkin diagram,
with v = (2, 1, 1, 1, 1) the minimal positive imaginary root. Its multiplicity mv
in the affine Kac–Moody algebra associated to Γ is known to be 4. Alternatively
it is known (Kac 1983 example b to conjecture 2) that AΓ (v, q) = q + 4, which
by (16.11) gives mv = 4. Thus indeed χL2 (MµB ) = mv = 4 checking (16.15) in
this case via Hausel (1998, Example 2 for Theorem 7.13).
Acknowledgements
This chapter is a write-up of the author’s talk at the Geometry Conference in
Honour of Nigel Hitchin in Madrid in September 2006. Problem 16.1 was raised
by Nigel Hitchin in 1995, then the author’s PhD supervisor, as a project for the
author’s PhD thesis. This chapter would like to show the impact of this modest-
looking question on the author’s subsequent research. The author’s research has
been supported by a Royal Society University Research Fellowship, NSF grant
DMS-0604775 and an Alfred Sloan Fellowship. The visit to Madrid was supported
by a Royal Society International Joint Project between the United Kingdom and
Spain.
References
Atiyah, M. F. (1978). Geometry of Yang-Mills fields. In Mathematical Problems in
Theoretical Physics (Proceeding of the International Conference, University of Rome,
Rome, 1977), Vol. 80 of Lecture Notes in Physics. pp. 216–21. Springer, Berlin,
Germany.
Atiyah, M. F. and Hitchin, N. J. (1985). Low energy scattering of nonabelian monopoles.
Phys. Lett. A 107(1), 21–5.
Atiyah, M. F. and Hitchin, N. J. (1988). The Geometry and Dynamics of Magnetic
Monopoles. M. B. Porter Lectures. Princeton University Press, Princeton, NJ.
Beilinson, A. and Drinfeld, V. Quantization of Hitchin’s integrable system and Hecke
eigensheaves. (384 pages, unpublished preprint, ca. 1995).
Bershadsky, M., Johansen, A., Shadov, V. and Vafa, C. (1995). Topological reduction
of 4D SYM to 2D σ-models. Nuclear Phys. B 448(1–2), 166–86.
References 343
Hausel, T. and Thaddeus, M. (2003). Relations in the cohomology ring of the moduli
space of rank 2 Higgs bundles. J. Am. Math. Soc. 16(2), 303–27.
Hitchin, N. J. (1987a). Monopoles, Minimal Surfaces and Algebraic Curves, Vol. 105
of Séminaire de Mathématiques Supérieures. Presses de l’Université de Montréal,
Montreal, QC.
Hitchin, N. J. (1987b). The self-duality equations on a Riemann surface. Proc. London
Math. Soc. (3) 55(1), 59–126.
Hitchin, N. J. (1987c). Stable bundles and integrable systems. Duke Math. J. 54(1),
91–114.
Hitchin, N. J. (2000). L2 -cohomology of hyperkähler quotients. Commun. Math. Phys.
211(1), 153–65.
Hitchin, N. J., Karlhede, A., Lindström, U. and Roček, M. (1987). Hyper-Kähler metrics
and supersymmetry. Commun. Math. Phys. 108(4), 535–89.
Hua, J. (2000). Counting representations of quivers over finite fields. J. Algebra 226(2),
1011–33.
Jordan, H. E. (1907). Group-characters of various types of linear groups. Am. J. Math.
29(4), 387–405.
Jost, J. and Zuo, K. (2000). Vanishing theorems for L2 -cohomology on infinite coverings
of compact Kähler manifolds and applications in algebraic geometry. Commun. Anal.
Geom. 8(1), 1–30.
Joyce, D. D. (2000). Compact Manifolds with Special Holonomy. Oxford Mathematical
Monographs. Oxford University Press, Oxford.
Kac, V. G. (1983). Root systems, representations of quivers and invariant theory. In
Invariant Theory (Montecatini, 1982), Vol. 996 of Lecture Notes in Mathematics.
pp. 74–108. Springer, Berlin, Germany.
Kac, V. G. (1990). Infinite-Dimensional Lie Algebras. 3rd edition. Cambridge Univer-
sity Press, Cambridge.
Kapustin, A. and Witten, E. (2007). Electric-magnetic duality and the geometric Lang-
lands program. Commun. Number Theory Phys. 1(1), 1–236. arXiv:hep-th/0604151.
Kirwan, F. (1992). The cohomology rings of moduli spaces of bundles over Riemann
surfaces. J. Am. Math. Soc. 5(4), 853–906.
Kronheimer, P. B. (1989). The construction of ALE spaces as hyper-Kähler quotients.
J. Differential Geom. 29(3), 665–83.
Kronheimer, P. B. and Nakajima, H. (1990). Yang-Mills instantons on ALE gravita-
tional instantons. Math. Ann. 288(2), 263–307.
Montonen, C. and Olive, C. D. I. (1977). Magnetic monopoles as gauge particles? Phys.
Lett. B B72, 117–20.
Mumford, D., Fogarty, J. and Kirwan, F. (1994). Geometric Invariant Theory. Vol. 34 of
Ergebnisse der Mathematik und ihrer Grenzgebiete (2), 3rd edition. Springer-Verlag,
Berlin, Germany.
Nakajima, H. (1998). Quiver varieties and Kac-Moody algebras. Duke Math. J. 91(3),
515–60.
Nitsure, N. (1991). Moduli space of semistable pairs on a curve. Proc. London Math.
Soc. (3) 62(2), 275–300.
Osborn, H. (1979). Topological charges for N=4 supersymmetric gauge theories and
monopoles of spin 1. Phys. Lett. B 83, 321.
Sasaki, T. O(−2) blow-up formula via instanton calculus on C 2 /Z and Weil conjecture.
2
arXiv:hep-th/0603162.
References 345
Schur, I. (1907). Untersuchungen über die darstellung der endlichen gruppen durch
gebrochene lineare substitutionen. J. Reine Angew. Math. 132, 85.
Segal, G. and Selby, A. (1996). The cohomology of the space of magnetic monopoles.
Commun. Math. Phys. 177(3), 775–87.
Sen, A. (1994). Dyon-monopole bound states, self-dual harmonic forms on the multi-
monopole moduli space, and SL(2, Z) invariance in string theory. Phys. Lett. B
329(2–3), 217–21.
Shoji, T. (2006). Lusztig’s conjecture for finite special linear groups. Represent. Theory
10, 164–222 (electronic).
Simpson, C. (1997). The Hodge filtration on nonabelian cohomology. In Algebraic
Geometry—Santa Cruz 1995, Vol. 62 of Proceedings of the Symposium on Pure
Mathematics. pp. 217–81. American Mathematical Society, Providence, RI.
Simpson, C. T. (1991). Nonabelian Hodge theory. In Proceedings of the International
Congress of Mathematicians, Vols. I, II (Kyoto, 1990), pp. 747–56. Mathematical
Society Japan, Tokyo, Japan.
Strominger, A., Yau, S.-T. and Zaslow, E. (1996). Mirror symmetry is T -duality.
Nuclear Phys. B 479(1–2), 243–59.
Vafa, C. and Witten, E. (1994). A strong coupling test of S-duality. Nuclear Phys. B
431(1–2), 3–77.
Witten, E. and Olive, D. (1978). Supersymmetry algebras that include topological
charges. Phys. Lett. B 78, 97.
XVII
NON-EMBEDDING AND NON-EXTENSION RESULTS
IN SPECIAL HOLONOMY
Robert L. Bryant
Dedicated to Nigel Hitchin with great admiration on the occasion of his 60th birthday
17.1 Introduction
In the early analyses of metrics with special holonomy in dimensions 7
and 8, particularly in regards to their existence and generality, heavy use was
made of the Cartan–Kähler theorem, essentially because the analyses were
reduced to the study of overdetermined PDE systems whose natures were
complicated by their diffeomorphism invariance. The Cartan–Kähler theory is
well suited for the study of such systems and the local properties of their
solutions. However, the Cartan–Kähler theory is not particularly well suited
for studies of global problems for two reasons: first, it is an approach to
PDE that relies entirely on the local solvability of initial value problems
and, second, the Cartan–Kähler theory is only applicable in the real-analytic
category.
Nevertheless, when there are no other adequate methods for analyzing the
local generality of such systems, the Cartan–Kähler theory is a useful tool and it
has the effect of focusing attention on the initial value problem as an interesting
problem in its own right. The point of this chapter is to clarify some of the
existence issues involved in applying the initial value problem to the problem
of constructing metrics with special holonomy. In particular, the role of the
assumption of real-analyticity will be discussed and examples will be constructed
to show that one cannot generally avoid such assumptions in the initial value
formulations of these problems.
The general approach can be outlined as follows: As is well known (cf. Bryant
1987), the problem of understanding the local Riemannian metrics in dimen-
sion n whose holonomy is contained in a specified connected group H ⊂ SO(n)
is essentially equivalent to the problem of understanding the H-structures in
dimension n whose intrinsic torsion vanishes, or, equivalently, that are parallel
with respect to the Levi-Civita connection of the Riemannian metric associated
to the underlying SO(n)-structure. In this chapter, an n-manifold M endowed
with an H-structure B → M with vanishing intrinsic torsion will be said to be
Introduction 347
expressed in terms of the second fundamental form can be found, for example, in Conti and
Salamon (2007).
348 Non-embedding and non-extension results in special holonomy
3 These three cases do not exhaust the possibilities; for the example of SU (3) ⊂ SO(6),
because the algebra is simpler. Also, because other approaches, based on the
existence of local holomorphic coordinates, have been employed in this case,
there is an instructive comparison to be made between those methods and the
Cartan–Kähler approach. For this reason, I go into the SU (2)-case in some detail.
I hope that the reader will find this as interesting as I have.
17.2 Beginnings
17.2.1 Holonomy
Let (M n , g) be a connected Riemannian n-manifold.
17.2.1.1 Parallel transport
To g, one associates its Levi-Civita connection ∇, which defines, for a piecewise-
C 1 curve γ : [0, 1] → M , a parallel transport
Pγ∇ : Tγ(0) M → Tγ(1) M, (17.1)
which is a linear g-isometry between the two tangent spaces.
17.2.1.2 Group structure
In Schouten (1918), he considered the set
Hx = Pγ∇ γ(0) = γ(1) = x ⊆ O(Tx M ) (17.2)
and called its dimension the number of degrees of freedom of g.
It is easy to establish the identities
−1
Pγ̄∇ = Pγ∇ and Pγ∇2 ∗γ1 = Pγ∇2 ◦ Pγ∇1 (17.3)
where γ̄ is the reverse of γ and γ2 ∗γ1 is the concatenation of paths γ1 and γ2
satisfying γ1 (1) = γ2 (0).
Consequently, Hx ⊂ O(Tx M ) is a subgroup and
−1
Hγ(1) = Pγ∇ Hγ(0) Pγ∇ . (17.4)
In particular, fixing a linear isometry u : Tx M → Rn , the conjugacy class
of Hu = uHx u−1 in O(n) is well-defined, independent of the choice of x ∈ M
or the isometry u : Tx M → Rn . By abuse of terminology, we say that H is the
holonomy of the metric g if H ⊂ O(n) is a group conjugate to some (and hence
any) of the groups Hu .
For later reference, if u : Tx M → Rn is fixed, we let
Bu = u ◦ Pγ∇ γ(1) = x . (17.5)
This Bu is an Hu -subbundle of the orthonormal coframe bundle of g, that is, it is
an Hu -structure on M . By its very construction, it is invariant under ∇-parallel
translation and, since ∇ is torsion-free, it follows that this Hu -structure admits
a torsion-free compatible connection.
350 Non-embedding and non-extension results in special holonomy
where
1. The tautological one-form ω takes values in R4
2. The connection one-form θ takes values in su(2) ⊂ so(4)
3. The curvature function R takes values in W4 , the five-dimensional
(real)
irreducible representation of SU (2) that lies in Hom Λ2 (R4 ), su(2)
4. The derived curvature function R takes values in V5 , the six-dimensional
complex irreducible representation of SU (2) that lies in Hom(R4 , W4 )
Calculation shows that the subspace V5 is an involutive tableau in Hom(R4 , W4 ),
with character sequence (s1 , s2 , s3 , s4 ) = (5, 5, 2, 0). Since the last non-zero char-
acter of this tableau is s3 = 2, Cartan’s generalization of the third fundamental
theorem of Lie applies to the structure equation (17.6) to yield his third obser-
vation above.
Υ2 + i Υ3 = dz 1 ∧ dz 2 and ¯
Υ1 = 12 i ∂ ∂φ, (17.9)
∂2φ
gφ = dz i ◦ dz̄ j . (17.12)
∂z i ∂ z̄ j
arguments given in Bryant (1987) for the cases G2 ⊂ SO(7) and Spin(7) ⊂ SO(8). I omit the
calculations in this case, since they are straightforward.
354 Non-embedding and non-extension results in special holonomy
the second factor. Also, to save writing, I will write η for π1∗ η where π1 : N × GL(3, R) → N
is the projection onto the first factor.
Hyper-Kähler four-manifolds 355
Let I be the ideal on X generated by {dΥ1 , dΥ2 , dΥ3 }. One calculates that
dΥ = t γ − (tr γ)I3 ∧ ∗ω ω + γ ∧ ω ∧ dt. (17.23)
Consequently, I is involutive, with characters (s0 , s1 , s2 , s3 , s4 ) = (0, 0, 3, 6, 0).
Since d(∗η η) = 0, the locus L = {0} × N × {I3 } ⊂ X is a regular, real-analytic
integral manifold of the real-analytic ideal I. (Note that L is just a copy of N .)
By the Cartan–Kähler theorem, L lies in a unique four-dimensional I-integral
manifold M ⊂ X. The Υi thus pull back to M to be closed and to define the
desired SU (2)-structure forms Υi on M inducing η on L = N .
coordinates x : U → R3 satisfying
d ∗η dx = 0. (17.26)
Now, fix a constant C and consider a coframing η = h(x)−1 dx on U ⊂ R3
where h : U → GL(3, R) is a mapping satisfying the first-order, quasi-linear
system:
d(∗η η) = 0, ∗η (t η ∧ dη) = 2C, and d ∗η dx = 0. (17.27)
The system (17.27) consists of seven equations for the nine unknown entries of h.
Calculation shows this first-order system to be underdetermined elliptic (i.e.
its symbol is surjective at every real covector ξ). By standard theory, it has
smooth local solutions that are not real-analytic.
Taking a non-real-analytic solution h, the resulting η will not be real-analytic
in the x-coordinates, which, by construction, are η-harmonic. Thus, such an η is
not real-analytic in any local coordinate system.
Remark 17.2 (“Flow” interpretation) The condition d(dt∧ω + ∗ω) = 0
has sometimes been described as an “SU (2)-flow” on coframings of N . In fact,
this closure condition can be written in the “evolutionary” form
d
ω = ∗ω (dω) − 1
2 ∗ω (t ω ∧ dω) ω. (17.28)
dt
By Theorem 17.1, if η on N 3 is real-analytic and satisfies d(∗η η) = 0, then (17.28)
has a solution in a neighborhood of t = 0 in R × N that satisfies the initial
condition
ω t=0 = η. (17.29)
One does not normally think of evolution equations as having to have real-
analytic initial data. However, Theorem 17.2 shows that some such regularity
assumption must be made. 8
Remark 17.3 (Co-closed coframings with specified metric) Given a
Riemannian three-manifold (N, g), it is an interesting question as to when there
exist (either locally or globally) a g-orthonormal coframing η = (η1 , η2 , η3 ) that
is co-closed. 9
One can formulate this as an EDS for the section η : N → B of the g-
orthonormal coframe bundle B → N . The natural EDS for this is not involutive,
but a slight extension is and is worth describing here.
8 On the other hand, the reader should be aware that real-analyticity, while sufficient, is
certainly not necessary for the initial-value problem to have a (local) solution. For example, it
suffices to let N be a smooth hypersurface in an SU (2)-manifold M 4 that is not real-analytic
with respect to the real-analytic structure provided by the holomorphic coordinates of the
underlying integrable complex structure. The induced co-closed coframing on N will then be
non-real-analytic initial data for which the initial-value problem will have a (local) solution.
9 Of course, closed g-orthonormal coframing exist locally if and only if g is flat.
Hyper-Kähler four-manifolds 357
in the real-analytic case, the general integral depends on two functions of two
variables.
Note, by the way, that when QS is positive (or negative) definite, the
linearization around such a solution is elliptic and hence such co-closed
coframings are as regular as the metric g. In particular, this always happens
when R > 0, that is, when the scalar curvature is positive. 10 Thus, for a real-
analytic metric with positive scalar curvature, all of its co-closed coframings are
real-analytic.
17.4 G2 -Manifolds
For background on the group G2 ⊂ SO(7) and G2 -manifolds, the reader can
consult Bryant (1987), Salamon (1989), Joyce (2000), and Bryant (2006). I will
generally follow the notation in Bryant (2006).
The crucial point is that the group G2 can be defined as the stabilizer
in GL(7, R) of the three-form
where eijk = ei ∧ej ∧ek and the ei are a basis of linear forms on R7 . The Lie
group G2 is connected, has dimension 14, preserves the metric and orientation
for which the ei are an oriented orthonormal basis, and acts transitively on the
unit sphere S 6 ⊂ R7 . The G2 -stabilizer of e1 is the subgroup SU (3) ⊂ SO(6)
that preserves the two-form e23 + e45 + e67 and the three-form
e246 − e257 − e347 − e356 = Re (e2 + i e3 ) ∧ (e4 + i e5 ) ∧ (e6 + i e7 ) . (17.38)
The GL(7, R)-orbit of φ in Λ3 (R7 ) is an open (but not convex) cone Λ3+ (R7 ) ⊂
Λ (R7 ) that consists precisely of the three-forms on R7 whose stabilizers are
3
10 It is interesting to note that, when R > 0, the ideal I on the six-manifold B is algebraically
where π : F(M )/G2 → M is the natural basepoint projection. Let I be the dif-
ferential ideal on F (M )/G2 = Λ3+ (T ∗ M ) generated by dσ and dτ. The following
result is proved in Bryant (1987):
17.4.1 Hypersurfaces
The group G2 acts transitively on S 6 ⊂ R7 , with stabilizer SU (3). Hence, an
oriented N 6 ⊂ M inherits a canonical SU (3)-structure, which is determined by
the (1, 1)-form ω and (3, 0)-form Ω = φ + i ψ defined by
then
Thus, when the right-hand side of this equation is constant, it follows by elliptic
regularity that N 6 is a real-analytic submanifold of the real-analytic (M 7 , σ).
Thus, if the given SU (3)-structure on N satisfying (17.45) and (17.49) is
not real-analytic, it cannot be induced by an embedding into a G2 -holonomy
manifold M .
It remains to construct a non-analytic example satisfying (17.45)
and (17.49).
Here is why it is somewhat delicate: Since dim GL(6, R)/SU (3) = 28, a choice
of an SU (3)-structure (ω, Ω) on N 6 depends on 28 functions of 6 variables.
Modulo diffeomorphisms, this leaves 22 functions of 6 variables. On the other
hand, the equations
d Re(Ω) = 0, d( 12 ω 2 ) = 0, and ∗ ω ∧ d Im(Ω) = C (17.51)
constitute 15 + 6 + 1 = 22 equations for the SU (3)-structure.
Thus, the equations to be solved are “more determined” than in the analogous
SU (2) case. Nevertheless, their diffeomorphism invariance still allows one to
construct the desired example, as will now be shown.
Say that a three-form
1 φ ∈ Ω3 (N 6 ) is elliptic if, at each point, it is linearly
equivalent to Re dz ∧dz ∧dz 3 . This is a open pointwise condition on φ (i.e.
2
17.5 Spin(7)-manifolds
For background on the group Spin(7) ⊂ SO(8) and Spin(7)-manifolds, the reader
can consult Bryant (1987), Salamon (1989), and Joyce (2000). I will generally
follow the notation in Bryant (1987).
Spin(7)-manifolds 363
The main point is that the group Spin(7) ⊂ SO(8) is the GL(8, R)-stabilizer
of the four-form Φ0 ∈ Λ4 (R8 ), defined by
Φ0 = e0 ∧ φ + ∗φ φ (17.61)
dΦ = 0. (17.62)
17.5.1 Hypersurfaces
Spin(7) acts transitively on S 7 and the stabilizer of a point is G2 . An oriented
hypersurface N 7 ⊂ M 8 inherits a G2 -structure σ ∈ Ω3+ (M ) that is defined by the
rule
σ=n Φ (17.64)
where n is the oriented normal vector field along N . It also satisfies
∗σ σ = N ∗ Φ. (17.65)
The structure equations show that
∗σ σ ∧ dσ = 28H (17.66)
where H is the mean curvature of N in (M, gΦ ).
Theorem 17.7 If σ ∈ Ω3+ (N 7 ) is real-analytic and satisfies d(∗σ σ) = 0, then σ
is induced by an immersion of N into a Spin(7)-manifold (M, Φ).
Proof. The argument in this case is entirely analogous to the SU (2) and G2
cases already treated.
Recall the definitions of σ and τ on F(N 7 )/G2 and, on X = R × F(N )/G2 ,
define
Φ = dt ∧ σ + τ. (17.67)
Let I be the ideal on R × F(N )/G2 generated by dΦ. The same calculation
used to prove Theorem 17.6 (see Bryant 1987) then yields that I is involutive,
with character sequence
(s0 , s1 , . . . , s8 ) = (0, 0, 0, 0, 1, 4, 10, 20, 0). (17.68)
Since d(∗σ σ) = 0, the G2 -structure σ defines a regular I-integral L ⊂ X within
the locus t = 0.
Since σ is assumed to be real-analytic, it follows that X, I, and L are
real-analytic with respect to the obvious induced analytic structures on the
appropriate underlying manifolds. Thus, the Cartan–Kähler theorem applies to
show that L lies in a unique I-integral M 8 ⊂ X. The form Φ then pulls back
to M to be a closed Φ ∈ Ω4s (M ) which induces the given σ on N L ⊂ M defined
by t = 0.
Theorem 17.8 There exist non-real-analytic G2 -structures σ ∈ Ω3+ (N 7 ) that
satisfy
d(∗σ σ) = 0 (17.69)
but that are not induced from a Spin(7)-immersion.
In fact, if a non-analytic G2 -structure σ satisfies (17.69) and
∗σ σ ∧ dσ = C (17.70)
Spin(7)-manifolds 365
Φ = dt ∧ S−1 (τ ) + τ (17.74)
Acknowledgments
Thanks to Duke University for its support via a research grant and to the NSF
for its support via DMS-0604195. Finally, I would like to thank the referee
for a careful reading of the manuscript and for helpful suggestions regarding
references.
References
Besse, A. (1987). Einstein manifolds. Ergebnisse der Mathematik und ihrer Grenzgebiete
(3) [Results in Mathematics and Related Areas (3)]. 10. Springer-Verlag, Berlin,
Germany. MR 0867684 (88f:53087).
Borel, A. and Lichnerowicz, A. (1952). Groupes d’holonomie des variétés riemanniennes.
C. R. Acad. Sci. Paris 234, 1835–37. MR 48133 (13,986b).
Bryant, R. (1987). Metrics with exceptional holonomy. Ann. Math. (2) 126, 525–76.
MR 916718 (1989b:53084).
Bryant, R. et al ., (1991). Exterior Differential Systems. MSRI Series 18, Springer-
Verlag, New York. MR 1083148 (1992h:58007).
Bryant, R. (2006). Some remarks on G2 -structures. Proceedings of Gökova Geometry-
Topology Conference 2005. pp. 75–109, Gökova Geometry/Topology Conference
(GGT), Gökova, MR 2282011 (2007k:53019) arXiv:math.DG/0305124.
Cartan, É. (1925). La géometrie des espaces de Riemann. Mémorial des Sciences
Mathematiques Fasc. IX.
Conti, D. and Salamon, S. (2007). Generalized Killing spinors in dimension 5. Trans.
Am. Math. Soc. 359, 5319–43.
DeTurck, D. and Kazdan, J. (1981). Some regularity theorems in Riemannian geometry.
Ann. Sci. Éc. Norm. Sup. 14, 249–60. MR 644518 (83f:53018).
Hitchin, N. (2001). Stable forms and special metrics. Global Differential Geometry: The
Mathematical Legacy of Alfred Gray (Bilbao, 2000). pp. 70–89, Contemporary Math-
ematics, Vol. 288, American Mathematical Society, Providence, RI. MR 1871001
(2003f:53065).
References 367
Marco Gualtieri
Dedicated to Nigel Hitchin on the occasion of his 60th birthday
18.1 Introduction
In this chapter we shall take a second look at a classical structure in differential
and algebraic geometry, that of a holomorphic Poisson structure, which is a
complex manifold with a holomorphic Poisson bracket on its sheaf of regular
functions. The structure is determined, on a real smooth manifold M , by the
choice of a pair (I, σI ), where I is an integrable complex structure tensor and
σI is a holomorphic Poisson tensor. We shall view (I, σI ) not as we normally do
but instead as a generalized complex structure, in the sense of Hitchin (2003). In
so doing, we shall obtain a new notion of equivalence between the pairs (I, σI )
which does not imply the holomorphic equivalence of the underlying complex
structures.
In studying this equivalence relation we are naturally led to an unexpected
connection to generalized Kähler geometry, as defined in Gualtieri (2004),
and to a method for constructing certain examples of these structures which
extends the recent work of Hitchin constructing bi-Hermitian metrics on Del
Pezzo surfaces (Hitchin 2007); in particular we obtain similar families of bi-
Hermitian metrics on all smooth Poisson Fano varieties, and in fact on any
smooth Poisson variety admitting a positive Poisson line bundle. We therefore
give an explicit construction of a subclass of the generalized Kähler struc-
tures proven to exist by the generalized Kähler stability theorem of Goto
(2007).
In both these efforts we shall find it useful to introduce an extension of the
notion of connection on a vector bundle, to allow differentiation not only in the
tangent but also in the cotangent directions; we call such a structure a generalized
connection. We also show that in the presence of a generalized metric, there is
a canonical connection D which plays the role of the Levi-Civita connection in
Kähler geometry: namely, we show that (J , G) is generalized Kähler if and only
if DJ = 0.
In the final section we make some speculative comments concerning the rela-
tionship between generalized Kähler geometry and non-commutative geometry,
a topic we hope to clarify in the future.
Gerbe trivializations 369
π∗ π
0 / T∗ / E / T / 0, (18.1)
in this splitting is
[X + ξ, Y + η]H = [X, Y ] + LX η − iY dξ + iY iX H,
where H ∈ Ω3cl (M )
is defined by (18.2). The splitting also defines an anti-
orthogonal automorphism C : E −→ E defined by C(X + ξ) = X − ξ, which
satisfies C(C± ) = C∓ . It also has the property, for Z, W ∈ C ∞ (E):
[CZ, CW ]H = C([Z, W ]−H ).
Theorem 18.1 Let C+ ⊂ E be a maximal positive-definite subbundle, i.e. a
generalized metric, as above. Let C : E −→ E be the map defined above. Write
Z = Z+ + Z− for the orthogonal projections of Z ∈ C ∞ (E) to C± . Then the
operator
DZ (W ) = [Z− , W+ ]+ + [Z+ , W− ]− + [CZ−, W− ]− + [CZ+, W+ ]+ (18.5)
defines a generalized connection on E, preserving both ·, · and the positive-
definite metric G. We call this the generalized Bismut connection
Proof. Using the properties of the Courant bracket and the orthogonality C+ =
⊥
C− , we have immediately the property Df Z W = f DZ W , for f ∈ C ∞ (M ). We
also have
DZ (f W ) = f DZ (W ) + (Z− f )W+ + (Z+ f )W− + (Z− f )W− + (Z+ f )W+
= f DZ (W ) + (Zf )W,
proving that D is a generalized connection.
It is clear from (18.5) that C± are preserved by the connection, since DZ W has
nonzero component in C± if and only if W does. Hence we obtain a decomposition
D = D+ ⊕ D− ,
where D± are generalized connection on C± respectively.
To prove that D preserves the canonical metric ·, · as well as the metric G,
we show that D ± preserve the induced metrics on C± . Let V, W ∈ C ∞ (C+ ), and
Z ∈ C ∞ (E). Then
Z+ V, W = (CZ+ )V, W
= −[CZ+ , V ], W ] − V, [CZ+ , W ]
= DZ+ V, W + V, DZ+ W
Also, we have
Z− V, W = [Z− , V ], W + V, [Z− , W ]
= DZ− V, W + V, DZ− W .
Summing these two results, we obtain that D + preserves the metric on C+ ; the
same argument holds for C− , completing the proof.
Generalized connections 373
χ±
Z W = DZ W = [Z∓ , W± ]± + [CZ± , W± ]±
= [Z∓ + (CZ)∓ , W± ]±
= 0,
∇± ±
X Y = 2π± DX Y±
±
= 4π± DX∓
Y±
= 4π± [X∓ , Y± ]± .
We may easily compute the torsion T ± of the connections ∇± , for vector fields
X, Y, Z:
X Y − ∇Y X − [X, Y ], Z)
= 2g(∇+ +
= 2H(X, Y, Z).
A similar calculation for x, y, z ∈ C ∞ (C− ) gives T (x, y, z) = 2H(x, y, z) as well,
yielding the result.
As we have explained, the generalized Bismut connection D is completely
determined by a usual connection on T ⊕ T ∗ . Using the fact that the Bismut
connections satisfy ∇± = ∇ ± 12 g −1 H for ∇ the Levi-Civita connection, we may
write D explicitly with respect to the splitting E = T ⊕ T ∗ , and for X ∈ C ∞ (T ),
as follows:
1 2 −1
∇ ∧ g (iX H)
DX = 1 X 2
2 iX H ∇∗X
The significance of this connection in the context of generalized geometry was
first understood and investigated by Ellwood. Here we simply view it as a
generalized connection 1 mainly for the purpose of highlighting its properties
and defining its torsion tensor.
version.
Generalized holomorphic bundles and branes 375
describe how the structures in the previous two sections may be made compatible
with J .
Corollary 18.1 The real vector field X = −iχ associated to any Hermitian
generalized holomorphic line bundle preserves the Poisson structure Q, that is, it
is a Poisson vector field. Furthermore its Poisson cohomology class [X] ∈ HQ
1
(M )
is independent of the Hermitian metric.
Generalized holomorphic bundles and branes 377
0 / N ∗S / τS / TS / 0.
∗
N1,0 C, so that we may form φ = Ω−1 θ ∈ H 0 (C, N1,0 ⊗ End(V )), making (V, φ)
into a brane for the complex structure.
On the other hand, if (V, θ) is a stable Higgs bundle, then by the existence
theorem (Hitchin 1987) for solutions to Hitchin’s equations we obtain a complex
flat connection ∇ on V , rendering (C, V, ∇) into a symplectic brane with respect
to either the real or imaginary parts of Ω.
(18.9)
π0 π1
M0 o S / M1
where we use the fact that sections pulled back from opposite factors M0 , M1
Courant commute in the product. Applying the projections to the final formula,
we obtain
p0 ([Z, Z ]) = [p0 (Z), p0 (Z )] and p1 ([Z, Z ]) = [p1 (Z), p1 (Z )],
as required.
We now describe the general form of a generalized complex brane when it
is supported on the whole manifold M ; these are usually called “space-filling
branes.” We first observe that the requirement that M be a generalized complex
submanifold of itself places a very strong constraint on J .
Proposition 18.5 (M, J ) is a generalized complex submanifold of itself if and
only if there exists an integrable isotropic splitting E = T ⊕ T ∗ of the Courant
algebroid with respect to which J has the form
I Q
J = , (18.11)
−I ∗
where I is a usual complex structure on the manifold and σ = P + iQ, for P =
IQ, is a holomorphic Poisson structure, that is, it satisfies [σ, σ] = 0.
Proof. Compatibility of the splitting with J forces J T = T , which holds if
and only if J is upper triangular, and the orthogonality of J together with the
fact J 2 = −1 guarantees that I is an almost complex structure and that Q is
a bivector of type (2, 0) + (0, 2). The −i eigenbundle of J is then the direct
∗
sum of T0,1 with the graph of σ : T1,0 −→ T1,0 . This is closed (involutive) for the
Courant bracket if and only if T0,1 is integrable and [σ, σ] = 0, as required.
In the splitting E = T ⊕ T ∗ for which J has the form (18.11), we see that
τS = T M , and further that = T1,0 , so that -modules are precisely holomorphic
bundles with respect to the complex structure I.
F I + I ∗ F + F QF = 0, (18.14)
Definition 18.7 Fix a real manifold M with real Poisson structure Q. Let G
be the groupoid whose objects are holomorphic Poisson structures (Ii , σi ) on M
with fixed imaginary part given by Im(σi ) = Q, and whose morphisms Hom(i, j)
consist of real closed two-forms Fij ∈ Ω2cl (M, R) such that
Ij − Ii = QFij ,
(18.15)
Fij Ij + Ii∗ Fij = 0.
∗
where Γσ = {ξ + σ(ξ) : ξ ∈ T1,0 }. Let ∂ V : C ∞ (V ) −→ C ∞ (L∗ ⊗ V ) be a gen-
eralized holomorphic structure. Decomposing using (18.19) and identifying
∗
Γσ = T1,0 , we write ∂ V = ∂ + ∂ , where ∂ : C ∞ (V ) −→ C ∞ (T0,1 ∗
⊗ V ) is a
usual holomorphic structure and ∂ : C ∞ (V ) −→ C ∞ (T1,0 ⊗ V ) satisfies, for
f ∈ C ∞ (M, C) and s ∈ C ∞ (V ),
∂ (f s) = f ∂ s + Zf ⊗ s,
where Zf = σ(df ) is the Hamiltonian vector field of f . This is equivalent to
2
condition (18.17). Furthermore ∂ V = 0 implies that ∂ is holomorphic and
defines a Poisson module structure via
{f, s} = ∂ ∂f s,
since (∂ )2 = 0 implies ∂ ∂{f,g} = [∂ ∂f , ∂ ∂g ], which is equivalent to (18.18), as
required.
Letting Lk = ker(Jk + i1), we see from (18.12) that exp(Fij ) takes Li to Lj .
Hence the map on generalized holomorphic bundles induced by the isomor-
phism (18.16) may be described as composition with exp(Fij )
∂ i → eFij ◦ ∂ i .
This map may be made more explicit in terms of the associated generalized
connections. Choose a Hermitian structure on the Ji -holomorphic bundle (i.e. σi -
Poisson module), and let D = ∇ + χ be the extension of ∂ i as in Proposition 18.2.
Then Fij acts on D via
D → D = ∇ + χ,
(18.20)
∇ = ∇ + Fij (χ).
which then defines a σj -Poisson module. It is important to note that the σi -
Poisson module, which is Ii -holomorphic, inherits via (18.20) a Ij -holomorphic
structure, without the presence of any holomorphic map between (M, Ij ) and
(M, Ii ).
Given this result, it is natural to ask how restrictive the condition of admitting
a Poisson module structure actually is. The following is a simple result describing
the complete obstruction to the existence of a Poisson module structure on a
holomorphic line bundle.
Proposition 18.9 Let M be a holomorphic Poisson manifold, and let V be a
∗
holomorphic line bundle on M . Then the Atiyah class of V , α ∈ H 1 (T1,0 ), com-
bines with the Poisson structure σ ∈ H 0 (∧2 T1,0 ) to give the class σα ∈ H 1 (T1,0 ).
If σα = 0, then there is a well-defined secondary characteristic class fα ∈ Hσ2 (M )
in Poisson cohomology. V admits a Poisson module structure if and only if
both classes {σα, fα } vanish. The space of Poisson module structures is affine,
modeled on Hσ1 (M ).
384 Branes on Poisson varieties
where Xi are holomorphic Poisson vector fields (since ∂d{f,g} = [∂df , ∂dg ]) such
that
Xi − Xj = Zlog gij . (18.22)
The Hamiltonian vector fields Zlog gij = σ(d log gij ) are a Čech representative for
the image of the Atiyah class under σ. Therefore, (18.22) holds if and only if
σα = 0 ∈ H 1 (T1,0 ). If σα = 0, then we may solve (18.22) for some holomorphic
vector fields X̃i . We can modify these by a global holomorphic vector field so
that they are each Poisson if and only if the global bivector field fσ defined
by fσ |Ui = [X̃i , σ] vanishes in Poisson cohomology, that is, fσ = [Y, σ] for Y ∈
H 0 (T1,0 ), in which case Xi = X̃i − Y |Ui defines a Poisson module structure as
required.
Given any holomorphic Poisson vector field Z ∈ Hσ1 (M ) and Poisson module
structure ∂, the sum ∂ + Z defines a new Poisson module structure. Con-
versely, two Poisson module structures ∂ , ∂ must satisfy ∂ − ∂ ∈ Hσ1 (M ), as
claimed.
{f, ρ} = LZf ρ.
= eσc Ωc ,
showing that (E0 , σ0 ) and (Ec , σc ) are isomorphic as generalized complex
manifolds.
The first eight terms cancel since Dx (J y) = J Dx y, and the last four terms
cancel since TD is of type (2, 1) + (1, 2). Therefore J is integrable, as
claimed.
Definition 18.8 Let (I, J, Q, F ) be a solution to the system (18.13), that is, it
defines two global Courant trivializations compatible with a generalized complex
structure, separated by the two-form F . Then
g = − 12 F (I + J)
If F is positive, then (g, I), (g, J) are both Hermitian structures. Let ωI = gI,
ωJ = gJ be their associated two-forms. Then we have the following:
Relation to generalized Kähler geometry 387
Theorem 18.3 Let (I, J, Q, F ) be as above, and let F be positive. Then the
pair
1 1 J + I −(ωJ−1 − ωI−1 )
1
JB = ,
2 b1 ωJ − ωI −(J ∗ + I ∗ )
−b 1
(18.23)
1 1 J − I −(ωJ−1 + ωI−1 ) 1
JA =
2 b1 ωJ + ωI −(J ∗ − I ∗ ) −b 1
defines a generalized Kähler structure on the standard Courant algebroid (T ⊕
T ∗ , [·, ·]0 ), for b ∈ Ω2 (M, R) given by
b = − 12 F (J − I).
Proof. It is easily verified that JA2 = JB2 = −1 and that [JA , JB ] = 0. To show
integrability, we first observe that JA has the form of a pure symplectic structure;
indeed, with the definitions above,
−F −1
JA = .
F
We see therefore that JA is integrable since dF = 0.
The structure JB is also integrable, as follows. Let LB = ker(JB − i) and Let
LB = L+ ⊕ L− be its decomposition into ±i eigenspaces for JA . Then
L+ = {X + (b + g)X : X ∈ TJ1,0 }
L− = {X + (b − g)X : X ∈ TI1,0 }
It follows from the definitions of b, g that b + g = −F J whereas b − g = F I. As
a result we have
L+ = {X − iF X : X ∈ TJ1,0 }
L− = {X + iF X : X ∈ TI1,0 }
D = ∇0 + iX,
where X is a real Q-Poisson vector field such that ∂X 1,0 = σI F0 , giving rise
to the real equations
LX Q = 0,
(18.25)
LX I0 = QF0 .
3. Let ϕt be the time t flow of the vector field X. Then we may transport F0 by
the flow, yielding the cohomologous family of two-forms Ft = ϕ∗−t F0 , which
satisfies
Ḟt = LX Ft = diX Ft .
I˙t = LX It = QFt ,
by (18.25). Note that Ft is type (1, 1) with respect to It . Also note that Ft
is the curvature of the family of connections
t
∇t = ∇0 + iX Fs ds,
0
= tQF t ,
390 Branes on Poisson varieties
X− X+
`AA >
AA }}
AA }}}
AA }
}}
C
line bundle Lk with connection ∇k , and all coinciding on the vanishing locus C
of Q. As a result, we obtain an infinite family of embeddings
k
F Fϕ Fϕ
··· / (M, I0 ) / (M, I1 ) / ··· / (M, Ik+1 ) / ···
cFF O : gggggg3
FF t g
FF tt ggggg
FF tttt
gggg ggggg
FF tt gggggg
gtgggg
C
Where the arrows on the top row indicate morphisms in the sense of the groupoid
of Definition 18.7. This may provide an alternative interpretation of Van den
Bergh’s construction of the twisted homogeneous coordinate ring (see Stafford
and van den Bergh 2001): let L = L0 |C , and let Lϕ = (ϕ−1 )∗ L. Then define the
vector spaces
i i+1 j−1
Hom(i, j) = H 0 (C, Lϕ ⊗ Lϕ ⊗ · · · ⊗ Lϕ )
and define a Z>0 -graded algebra structure on
9
A• = Hom(0, k), (18.28)
k≥0
Acknowledgements
I would like to thank Nigel Hitchin for many illuminating discussions, and also
the organizers of his birthday conference, especially Luis Álvarez-Cónsul and
Oscar Garcı́a-Prada, for a stimulating event. I would also like to thank Mike
Artin, Gil Cavalcanti, Izzet Coskun, and Pavel Etingof for helpful discussions.
Finally, I thank Yicao Wang for pointing out an error in a previous draft.
References
Apostolov, V., Gauduchon, P., and Grantcharov, G. (1999). Bihermitian structures on
complex surfaces, Proc. London Math. Soc. 79(3), 414–28.
Artin, M., Tate, J., and Van den Bergh, M. (1991). Modules over regular algebras of
dimension 3, Invent. Math. 106(2), 335–88.
Bondal, A. I. and Polishchuk, A. E. (1993). Homological properties of associative
algebras: the method of helices, Izv. Ross. Akad. Nauk Ser. Mat. 57(2), 3–50.
Brylinski, J.-L. (1993). Loop Spaces, Characteristic Classes and Geometric Quantiza-
tion, Vol. 107 of Progress in Mathematics, Birkhäuser Boston Inc., Boston, MA.
Bursztyn, H. and Radko, O. (2003). Gauge equivalence of Dirac structures and sym-
plectic groupoids, Ann. Inst. Fourier (Grenoble) 53(1), 309–37.
Courant, T. (1990). Dirac manifolds, Trans. Am. Math. Soc. 319, 631–61.
Ellwood, I. T. (2007). NS-NS fluxes in Hitchin’s generalized geometry, J. High Energy
Phys. 12, 084, 1–24.
Evens, S., Lu, J.-H., and Weinstein, A. (1999). Transverse measures, the modular
class and a cohomology pairing for Lie algebroids, Q. J. Math. Oxford Ser. (2)
50(200), 417–36.
Goto, R. (2007). Deformations of generalized complex and generalized Kähler struc-
tures, arXiv:0705.2495v2.
Gualtieri, M. (2004). Generalized complex geometry, DPhil thesis, Oxford University,
Oxford, math.DG/0401221.
Gualtieri, M. (2007). Generalized complex geometry, arXiv:math/0703298v2.
Hitchin, N. (1987). The self-duality equations on a Riemann surface, Proc. London
Math. Soc. (3) 55(1), 59–126.
Hitchin, N. (2003). Generalized Calabi-Yau manifolds, Q. J. Math. 54, 281–308.
Hitchin, N. (2006a). Brackets, forms and invariant functionals, Asian J. Math.
10(3), 541–60.
Hitchin, N. (2006b). Instantons, Poisson structures and generalized Kähler geometry,
Commun. Math. Phys. 265(1), 131–64.
Hitchin, N. (2007). Bihermitian metrics on Del Pezzo surfaces, J. Symplectic Geom.
5, 1–7.
Kapustin, A. (2005). A-branes and noncommutative geometry, hep-th/0502212.
Kapustin, A. and Orlov, D. (2003). Remarks on A-branes, mirror symmetry, and the
Fukaya category, J. Geom. Phys. 48, 84.
Lichnerowicz, A. (1977). Les variétés de Poisson et leurs algèbres de Lie associées,
J. Diff. Geom. 12(2), 253–300.
Lindström, U., Roček, M., von Unge, R., and Zabzine, M. (2007). Generalized Kähler
manifolds and off-shell supersymmetry, Commun. Math. Phys. 269(3), 833–49.
394 Branes on Poisson varieties
Liu, Z.-J., Weinstein, A., and Xu, P. (1997). Manin triples for Lie bialgebroids, J. Diff.
Geom. 45, 547–74.
Polishchuk, A. (1997). Algebraic geometry of Poisson brackets, J. Math. Sci.
(New York) 84(5), 1413–44. Algebraic Geometry, 7.
Roytenberg, D. (1999). Courant algebroids, derived brackets and even symplectic
supermanifolds, PhD thesis, University of California, Berkeley, CA.
Simpson, C. T. (1992). Higgs bundles and local systems, Inst. Hautes Études Sci. Publ.
Math. (75), 5–95.
Stafford, J. T. and van den Bergh, M. (2001). Noncommutative curves and noncom-
mutative surfaces, Bull. Am. Math. Soc. (N.S.) 38(2), 171–216 (electronic).
Ševera, P. (1998–2000). Letters to A. Weinstein concerning Courant algebroids. Unpub-
lished.
Weinstein, A. (1997). The modular automorphism group of a Poisson manifold,
J. Geom. Phys. 23, 379–94.
XIX
CONSISTENT ORIENTATION OF MODULI SPACES
and ours the moduli spaces in question sit inside infinite-dimensional function
spaces, and the virtual tangent bundle to the moduli space extends to a virtual
bundle on the function space. Thus it suffices to orient over the function space,
and this becomes a universal problem. Presumably our methods apply to his
situation as well, but we have not worked out the details.
It is a pleasure and an honor to dedicate this chapter to Nigel Hitchin. We
greatly admire his mathematical taste, style, and influence. ¡Feliz cumpleaños y
que cumplas muchos más!
FX (19.1)
DD
s zzz DD t
z DD
zz DD
z} z !
FY0 FY1
398 Consistent orientation of moduli spaces
FX ◦X (19.2)
u II
u II r
uu
r
u II
uu II
zuu I$
FX FX
u II II
uu II t s uu II t
s
uuu II
II uuu II
uu II uu II
uz u $ uz u I$
FY0 FY1 FY2
the physical point of view the differential equations are the BPS (Bogomolńyi-
Prasad-Sommerfield) equations of supersymmetry; from a mathematical point
of view they define the minima of a calculus of variations functional. In this
chapter we consider pure gauge theories. Fix a compact Lie group G and for any
manifold M let FM denote the stack of G-connections on M . Define MM as the
stack of flat G-connections on M . If we choose a set {mi } ⊂ M of “basepoints,”
one for each
M Ncomponent
of M , then
OMM is represented by the product of
groupoids i Hom π1 (M, mi ), G //G . A basic property of flat connections is
the gluing law (see (19.2)).
Lemma 19.1 Suppose X : Y0 → Y1 and X : Y1 → Y2 are bordisms of smooth
manifolds. Then MX ◦X is the fiber product of
MX MX (19.3)
GG
GG t
s www
GG w
GG ww
# w{ w
MY
Roughly speaking, this says that given flat connections on X, X and an isomor-
phism of their restrictions to Y , one can construct a flat connection on X ◦ X
and every flat connection on X ◦ X comes this way.
Replace the infinite-dimensional correspondence diagram (19.1) with the finite-
dimensional correspondence diagram of flat connections:
MX (19.4)
FF
xx FF t
xx
s
x FF
x FF
x| x "
MY0 MY1
Whereas the path integral linearizes (19.1) using measure theory, we propose
instead to linearize (19.4) using algebraic topology. Let E be a generalized
cohomology theory. To every closed (n − 1)-manifold we assign the abelian group
AY = E • (MY ).
To a morphism X : Y0 → Y1 we would like to attach a homomorphism ZX :
AY0 → AY1 defined as the push–pull
ZX := t∗ ◦ s∗ : E • (MY0 ) −→ E • (MY1 ) (19.5)
in E-cohomology. Whereas the path integral requires measures consistent under
gluing to define integration t∗ , in our topological setting we require orientations
of t consistent with gluing to define pushforward t∗ . The consistency of orienta-
tions under gluing ensures that (19.5) defines a TQFT which satisfies the gluing
law (functoriality).
400 Consistent orientation of moduli spaces
MY ∼
= G//G
19.1.3 Remarks
r Let X be the “pair of pants” with the two legs incoming and the single waist
outgoing. Then restriction to the outgoing boundary is the map t : (G ×
G)//G → G//G induced by multiplication µ : G × G → G. So ZX = t∗ ◦ s∗ ,
which defines the ring structure in a two-dimensional TQFT, is pushforward
by multiplication on G. Therefore, we do construct the Pontrjagin product
τ +dim G
on KG (G)—here τ is the twisting and there is a degree shift as well—
and have implicitly used an isomorphism of twistings µ∗ τ → τ ⊗ 1 + 1 ⊗ τ
which, since the TQFT guarantees an associative product, satisfies a com-
patibility condition for triple products. This isomorphism and compatibility
are embedded in our consistent orientation construction.
r We do not use the theorem (Abrams 1996, Moore and Segal 2006) which
constructs a two-dimensional TQFT from a Frobenius ring. Rather, our a
priori construction manifestly produces a TQFT which satisfies the gluing
law, and we deduce the Frobenius ring as a derived quantity.
r Three-dimensional Chern–Simons theory is defined on a bordism category of
manifolds which carry an extra topological structure. For oriented manifolds
this extra structure is described as a trivialization of p1 or signature, or a
certain sort of framing. (For spin manifolds it is described as a string struc-
ture or, since we are in sufficiently low dimensions, an ordinary framing.)
The two-dimensional reduction constructed here factors through the bordism
categories of oriented manifolds.
Orientation and twisting 401
where Ωo(V )+q (M ) is the space of smooth sections of the ungraded vector bundle
P r+q ∗
T M ⊗ o(V ). The cohomology of (19.6) is the twisted de Rham cohomology
o(V )+•
HdR (M ). Let o(M ) = o(T M ). Then integration is a map
: Ωo(M
c
)
(M ) −→ R. (19.7)
M
H o(M ) (M ) −→ Z.
Orientation and twisting 403
19.2.2 K-theory
This discussion applies to any multiplicative cohomology theory. 1 The only issue
is to determine the twisting of a real vector bundle in that theory. For complex
K-theory there are many possible models for the twisting τV of a vector bundle
V → M . In the Donovan–Karoubi picture (Donovan and Karoubi 1970) τV is
represented by the bundle of complex Z/2Z-graded Clifford algebras defined
by V . A bundle of algebras A → M of this type is considered trivial if A =
End(W ) for a Z/2Z-graded complex vector bundle W → M , that is, if A is
Morita equivalent to the trivial bundle of algebras M × C. The equivalence class
of τV is
[τV ] = rank V, w1 (V ), W3 (V ) ∈ H 0 (M ; Z/2Z) × H 1 (M ; Z/2Z) × H 3 (M ; Z).
(19.10)
Only torsion classes in H 3 (M ; Z) are realized by bundles of finite-dimensional
algebras, but we have in mind a larger model which includes nontorsion
classes. (Such models are developed in Atiyah and Segal 2006, Freed et al.
2007a, and Murray 1996 among other works.) There is a Whitney sum
isomorphism
∼
=
τV1 ⊕V2 −−−→ τV1 + τV2 ; (19.11)
the sum of twistings is realized by the tensor product of algebras. A spinc
structure on V induces an orientation, that is, a Morita equivalence
∼
=
τrank V −−−→ τV . (19.12)
An A-twisted vector bundle is a vector bundle with an A-module structure; it
represents an element of twisted K-theory.
The Whitney formula (19.11) allows us to attach a twisting to any virtual real
vector bundle: set
τ−V = −τV . (19.13)
d d
0 −→ Ω0X (gP ) −−−
A
→ Ω1X (gP ) −−−
A
→ Ω2X (gP ) −→ 0, (19.17)
the de Rham complex with coefficients in the adjoint bundle associated to P . This
is an elliptic complex. Its symbol σ satisfies the reality condition σ(−ξ) = σ(ξ)
for ξ ∈ T X, since (19.17) is a complex of real differential operators (Atiyah and
Singer 1971). Recall that the symbol of any complex differential operator lies
in Kcv0
(T X) ∼= K 0 (X T X ). The reality condition gives a lift σ ∈ KR0 (X iT X ),
where the imaginary tangent bundle iT X carries the involution of complex
conjugation (Atiyah 1966). Bott periodicity asserts that V ⊕ iV is canonically
KR-oriented for any real vector bundle V with no degree shift. In the language
of twistings of KR this means
(KR) (KR)
τV + τiV = 0. (19.18)
406 Consistent orientation of moduli spaces
Therefore
(KR) (19.18) (KR) (19.13)
−→ KR−τiT X (X) −−−−
Thom
KR0 (X iT X ) −−−
∼ ∼
−−→ KRτT X (X) −−−−
∼
−−→
= = =
(KR)
−τ−T X
KO0 (X −T X )
Thom
KR (X) −−−
∼
−→ (19.19)
=
from which we locate the symbol σ ∈ KO0 (X −T X ). Note that by Atiyah duality
KO0 (X −T X ) ∼
= KO0 (X) and the KO-homology group is well-known to be the
home of real elliptic operators.
Now (19.17) is a universal operator: its symbol is constructed from the exterior
algebra of T X and the adjoint representation of G; it does not depend on details
of the manifold X. Thus it is pulled back from a universal symbol. Let Vn →
BSOn denote the universal oriented n-plane bundle. The universal symbol lives
on the bundle V2 → BSO2 × BG, and by (19.19) we identify it as an element
σuniv ∈ KO0 (BSO2−V2 ∧ BG+ ). (19.20)
Here BG+ is the space BG with a disjoint basepoint adjoined and “∧” is the
smash product. Introduce the notation
M T SOn = BSOn−Vn
for this Thom spectrum and so write
σuniv ∈ KO 0 (M T SO2 ∧ BG+ ).
If f : X → BSO2 × BG is a classifying map for T X and P , and f˜ : X −T X →
MTSO2 ∧ BG+ the induced map on Thom spectra, then σ = f˜∗ σuniv . It is in
this sense that (19.20) is a universal symbol.
Remark 19.2 We digress to explain M T SOn in more detail. Let Grn+ (RN ) be
the Grassmannian of oriented n-planes in RN and
0 → Vn → N → QN −n → 0
the exact sequence of real vector bundles over Grn+ (RN ) in which Vn is the
universal subbundle and QN −n the universal quotient bundle. Denote the Thom
space of the latter as
ZN := Grn+ (RN )QN −n .
Then the suspension ΣZN is the Thom space of QN −n ⊕ 1 → Grn+ (RN ). But
QN −n ⊕ 1 is the pullback of QN +1−n → Grn+ (RN +1 ) under the natural inclusion
Grn+ (RN ) → Grn+ (RN +1 ), and in this manner we produce a map ΣZN → ZN +1 .
Whence the spectrum M T SOn = {ZN }N ≥0 . The notation identifies M T SOn
as an unstable version of the Thom spectrum M SO and also alludes to its
appearance in the work of Madsen and Tillmann (2001). There are analogous
spectra M T On , M T Spinn , M T Stringn , etc. If F : S N → ZN is transverse to
the 0-section, then X := F −1 (0-section) is an n-manifold and the pullback
Universal orientations and consistent orientations 407
b
/ M T SOn / / (BSOn )+ .
Σ−1 M T SOn−1 (19.24)
Then identify BSOn−1 as the unit sphere bundle S(Vn ) and write (19.26) in
terms of Thom spaces:
Here 0 is the vector bundle of rank zero. Now add −Vn to each of the vector
bundles in (19.27) and note that the restriction of Vn to BSOn−1 is Vn−1 ⊕ 1.
Consider the diagram
b q
Σ−1 M T SO1 ∧ BG / M T SO2 ∧ BG / (M T SO2 , Σ−1 M T SO1 ) ∧ BG
TTTT
TTTT σ̄univ
TTTT
TTTT σ̄univ
TTTT
*
ko
(19.28)
The top row is a cofibration. From (19.24) we can replace (M T SO2 , Σ−1 M T SO1 )
with (BSO2 )+ .
Lemma 19.3 Define σ̄univ in (19.28) to be the map (BSO2 )+ ∧ BG → ko
induced by the standard representation of SO2 smash with the reduced adjoint
Universal orientations and consistent orientations 409
Proof. Recalling the isomorphisms in (19.19), and using the fact that the
universal symbol σ̄univ is canonically associated to a representation of SO2 × G,
we locate σ̄univ ∈ KRSO 0
2 ×G
(−R2 )c ∼ = KRSO 0
2 ×G
(iR2 )c . (Recall that the involu-
2 ∼ 2
tion on C = R ⊕ iR is complex conjugation and the subscript “c” denotes
2
QN −n / QN −n
/ Gr+ (RN )
+
Grn−1 (RN −1 ) n
f˙ : M −→ M T SO1 ∧ BG
410 Consistent orientation of moduli spaces
The symbol σ of an elliptic operator lives in the third group, and Atiyah and Bott
construct a lift to the second group from a local boundary condition. (The image
of σ in the last group is an obstruction to the existence of local boundary condi-
tions; the image of the first group in the second measures differences of boundary
conditions.) The relative boundary conditions on the twisted de Rham complex
are universal, so the corresponding lift of the symbol occurs in the universal
setting. Recall from the proof of Lemma 19.3 the exact sequence (19.28), now
extended one step to the left:
0
KRG (iR)c −→ KRSO
0
2 ×G
(C2 )c −→ KRSO
0
2 ×G
(iR2 )c −→ KRG
0
(iR2 )c
The group G acts trivially in all cases. The Atiyah–Bott procedure applied
to the relative boundary conditions gives a lift of σ̄univ ∈ KRSO 0
2 ×G
(iR2 ))c
0 2
to KRSO2 ×G (C )c . Recall that σ̄univ , constructed in the proof of Lemma 19.3,
is also a lift of σ̄univ . But by periodicity we find KRG 0
(iR)c ∼
= KRG 0
(−R)c ∼=
0 ∼ 1
KOG (−R)c = KOG (pt) which vanishes by Anderson (1964). Thus the lift
of σ̄univ is unique and σ̄univ computes the relative twisted de Rham cohomology.
This completes the proof.
Universal orientations and consistent orientations 411
q r
M T SO2 ∧ BG / (M T SO2 , Σ−1 M T SO1 ) ∧ BG / M T SO1 ∧ BG
TTTT
TTTT σ̄univ
TTTT
−λ
TTTT σ̄univ
TTTT
) τ
/ picg K
ko
(19.31)
The top row is a cofibration, the continuation of the top row of (19.28) in the
Puppe sequence. Recall that a universal orientation is a null homotopy of τ ◦
σ̄univ = τ ◦ σ̄univ ◦ q.
λ : BG −→ B Picg K (19.32)
on spaces.
σ̄univ −λ
τ
k / ko / picg K
(19.33)
and specifying a homotopy which makes the left square commute. There is
a natural choice: smash the K-theory Thom class U : M T U1 M T SO2 → k
with the complexified reduced adjoint representation ḡC . This is the universal
rewriting of de Rham on a Riemann surface in terms of Dolbeault, at least on the
symbolic level. In terms of the proof of LemmaP 19.3, the map0 σ̄univ , 2restricted
to M T SO2 , is the exterior algebra complex ( • C2 , ) in KRSO 2
(iR )c . Write
R2 = LR for the complex line L = C and C2 ∼= R2 ⊗ C ∼ = L ⊕ L. Then the symbol
complex at θ ∈ iR2 ,
(θ) (θ)
C / L⊕L / L⊗L ,
which defines the K-theory Thom class. Tensor with the complexified reduced
adjoint representation ḡC to complete the argument.
η
To compute the level of µ we factorize τ as ko → Σ−1 ko → picg K, where the
first map is multiplication by η : S 0 → S −1 and the second is projection to a
Postnikov section. The map η fits into the Bott sequence k → ko → Σ−1 ko, and
so we extend (19.33) to the left:
(19.34)
The homotopy which expresses the commutativity of the right-hand square
induces the map α in this diagram, and the map −λ induced in (19.33) is the
suspension of α. We claim that there is a unique α, up to homotopy, which makes
the left square in (19.34) commute. For the difference of any two choices for α is a
map Σ−1 M T SO1 ∧ BG → Σ−1 ko, and the homotopy classes of such maps form
Universal orientations and consistent orientations 413
the group KO1 (BG) which vanishes (Anderson 1964). It is easy to find a map α
as follows. Since Σ−1 M T SO1 S −2 (Lemma 19.2(1)), the upper left map is
the inclusion of the bottom cell of M T SO2 ∧ BG. The composite Σ−1 M T SO1 ∧
BG Σ−2 BG → k factors as Σ−2 BG → Σ−2 k → k, where the first map is the
double desuspension of ḡC and the second Bott periodicity. Choose α to be the
double desuspension of ḡ, the real reduced adjoint representation. This completes
the proof.
Since equivalence classes of universal orientations form a torsor for the
group O(G) in (19.23), the canonical universal orientation identifies the torsor
of universal orientations with O(G). Notice the natural map
: O(G) = [ΣM T SO2 ∧ BG, picg K] −→ [M T SO1 ∧ BG, picg K]
∼
= [BG, B Picg K] (19.35)
from universal orientations to levels. If g ∈ O(G), then the level of µ + g is
(g) − h. If G is connected, simply connected, and simple, then [BG, B Picg K] ∼
=
H 4 (BG; Z) ∼= Z and h is the dual Coxeter number of G times a generator. Then
g → (g) − h is a version of the ubiquitous “adjoint shift.”
Remark 19.3 For any G the top homotopy group of Map(ΣM T SO2 , picg K)
and of B Picg K is π4 , which is infinite cyclic. So there is a homomorphism
of H 4 (BG; Z) into the domain and codomain of (19.35), and on these subspaces
is an isomorphism. This means that we can change a consistent orientation by
an element of H 4 (BG; Z), and the level changes by the same amount.
da
0 / Ω0 (g ) / Ω1 (g ) / 0 (19.39)
Y0 Q Y0 Q
This is the map (19.5) with the twistings induced from the universal
orientation.
A universal orientation induces consistent orientations on the outgoing restric-
tion maps of bordisms. In other words, if X : Y0 → Y1 and X : Y1 → Y2 are
Families of surfaces, twistings, and anomalies 415
MX ◦X
s KK
r ss KK r
ss KK
ss KK
sy s K%
MX MX
JJ JJ
s ttt JJ t tt
t JJ t
tt JJ s
tt JJ
tt JJ tt JJ
ty t J% ty t J%
MY0 M Y1 MY2
satisfy
∗
(t r )∗ ◦ (sr)∗ = [t∗ ◦ s ] ◦ [t∗ ◦ s∗ ].
This follows from Lemma 19.1 and the “Fubini property”
(t r )∗ = t∗ r∗ (19.41)
of pushforward. The orientation of t induces orientations of r and t r , since
the diamond is a fiber product. At stake in (19.41) is the consistency of the
orientations, which is ensured by the use of a universal orientation. The details
of this argument 4 will be given on another occasion.
One caveat: since MX , M∂X are stacks we can only pushforward along repre-
sentable maps, and this forces every component of X to have a nonempty out-
going boundary. As mentioned at the end of section 19.1, the partial topological
quantum field theory obtained from the push–pull construction extends to a full
theory using the invertibility of the (co)pairing attached to the cylinder (Freed
et al. 2003, section 17).
field theory, and these field theories factor through the group completion of bordism, that is,
the Madsen–Tillmann space.
416 Consistent orientation of moduli spaces
a section F (X/S) of Hom F (Y0 /S), F (Y1 /S) → S; the topological invariance
and gluing law of the TQFT imply thatthis section is flat. In other words,
F (X/S) ∈ H 0 S; Hom F (Y0 /S), F (Y1 /S) . It is natural, then, to postulate that
a TQFT in families gives more, namely, classes of all degrees:
F (X/S) ∈ H • S; Hom F (Y0 /S), F (Y1 /S) . (19.42)
These classes are required to satisfy gluing laws and topological invariance as
well as naturality under base change.
The idea of a TQFT in families—at least in two dimensions—was introduced
in the mid-1990s. In two dimensions it is often formulated in a holomorphic
language (e.g. Kontsevich and Manin 1994), and classes are required to extend
to the Deligne–Mumford compactification of the moduli space of Riemann
surfaces.
Our push–pull construction works for families of surfaces—with a twist. The
purpose of this section is to alert the reader to the twist. 5
Suppose X → S is a family of bordisms from Y0 → S to Y1 → S, where Yi →
S are fiber bundles of oriented one-manifolds. Then the moduli stacks of flat
connections form a correspondence diagram over S:
MX/S
II
s uuu II t
uu II
uu II
uz u I$
MY0 /S MY1 /S
JJ t
JJ tt
JJ tt
JJ tt π1
JJ t
tz t
π0
$
S
exposition.
Families of surfaces, twistings, and anomalies 417
The degree shift is now incorporated into the twist τX/S , and there may be non-
trivial contributions to the twist from w1 and W3 of H• (X/S) → S as well. (The
degree shift and twistings vanish canonically if dim G is even.)
Example 19.1 Consider the disjoint union X of two 2-disks. The boundary
circles are outgoing, as above. Suppose that dim G is odd. For a single disk
τ +1
the pushforward t∗ (1) in (19.38) lands in KG (G) and is the unit 1 in the
Verlinde ring. Thus for the disjoint union of two disks, t∗ (1) is the image of 1 ⊗ 1
(τ,τ )
under the external product KG τ +1
(G) ⊗ KGτ +1
(G) → KG×G (G × G). Now con-
sider the family X → S with fiber X and base S = S in which the monodromy
1
exchanges the two disks. The flat bundle H• (X/S) → S has rank 2 and non-
trivial w1 . According to (19.43), then, t∗ (1) for the family lives in the twisted
∗
group K (τ,τ )+π τX/S (M∂X/S ). On each fiber of π : M∂X/S → S we recover the
class 1 ⊗ 1 above. But upon circling the loop S = S 1 this class changes sign in the
π ∗ τX/S -twisted K-group. Said differently, the diffeomorphism which exchanges
the disks acts by a sign on 1 ⊗ 1. Of course, one might predict this from the
τ +1
sign rule in graded algebra: the Verlinde ring KG (G) is in odd degree, so upon
exchanging the factors of 1 ⊗ 1 one picks up a sign. It shows up here as an extra
twisting.
Acknowledgments
The work of Daniel S. Freed, Michael J. Hopkins, and Constantin Tele-
man was supported by NSF grant DMS-0603964, DMS-0306519, and EPSRC
GR/SO6165/01, respectively.
We thank Veronique Godin, Jacob Lurie, Ib Madsen, and Ulrike Tillmann for
enlightening conversations.
References
Abrams, L. (1996). Two-dimensional topological quantum field theories and Frobenius
algebras. J. Knot Theory Ramifications 5(5), 569–87.
Anderson, D. W. (1964). The real K-theory of classifying spaces. Proc. Nat. Acad. Sci.
51(4), 634–6.
Atiyah, M. F. (1966). K-theory and reality. Q. J. Math. Oxford Ser. (2) 17, 367–86.
Atiyah, M. F. and Bott, R. (1964). The index problem for manifolds with boundary.
Differential Analysis, Bombay Colloquium, 1964. Oxford University Press, London,
pp. 175–86.
Atiyah, M. and Segal, G. (2006). Twisted K-Theory and Cohomology, Inspired by
S. S. Chern, Nankai Tracts Mathematics, Vol. 11, World Scientific Publication,
Hackensack, NJ, pp. 5–43, arXiv:math/0510674.
Atiyah, M. F. and Singer, I. M. (1971). The index of elliptic operators. V, Ann. Math.
(2) 93, 139–49.
Costello, K. (2007). Topological conformal field theories and Calabi-Yau categories.
Adv. Math. 210(1), 165–214, arXiv:math/0412149.
Donaldson, S. K. (1987). The orientation of Yang-Mills moduli spaces and 4-manifold
topology. J. Differential Geom. 26(3), 397–428.
Donaldson, S. K. and Kronheimer, P. B. (1990). The geometry of four-manifolds. Oxford
Mathematical Monographs. Oxford Science Publications, New York.
Donovan, P. and Karoubi, M. (1970). Graded Brauer groups and K-theory with local
coefficients. Inst. Hautes Études Sci. Publ. Math. No. 38, 5–25.
Freed, D. S. (2009). Remarks on Chern–Simons theory. Bull. Amer. Math. Soc. 46,
221–254, arXiv:0808.2507.
Freed, D. S., Hopkins, M. J. and Teleman, C. Loop groups and twisted K-theory I,
arXiv:0711.1906.
Freed, D. S., Hopkins, M. J. and Teleman, C. Loop groups and twisted K-theory II,
arXiv:math/0511232.
Freed, D. S., Hopkins, M. J. and Teleman, C. Loop groups and twisted K-theory III,
arXiv:math/0312155.
Freed, D. S., Hopkins, M. J. and Teleman, C. (2007b). Twisted equivariant K-theory
with complex coefficients. J. Topol. 1, arXiv:math/0206257.
Galatius, S., Madsen, I., Tillmann, U. and Weiss, M. (2009). The homotopy type of
the cobordism category. Acta Math. 202, 195–239, arXiv:math/0605249.
Gilkey, P. B. (1995). Invariance Theory, the Heat Equation, and the Atiyah-Singer Index
Theorem, 2nd edn., Studies in Advanced Mathematics. CRC Press, Boca Raton, FL.
Hitchin, N. (2001). Lectures on special Lagrangian submanifolds. Winter School on
Mirror Symmetry, Vector Bundles and Lagrangian Submanifolds (Cambridge, MA,
References 419
Data Dolbeault
principal part, 43 cohomology, 336
spectral, 35, 45 complex, 396
de Rham Donaldson theory, 124
class, 18, 241 Double coset space, 167
cohomology, 200, 328, 401–402 Dual
complex, 8, 243, 375, 396, 405–410 bundle, 238–240
functor, 216 group, 114, 165
Deck transformation, 402 Duality
Deformation Alexander, 267, 277
complex, 2, 405, 414 Atiyah, 406
infinitesimal, 92 Langlands, 113–127, 174
isomonodromic, 122–123, 215–222 Poincaré, 329
of complex structures, 303 Serre, 150
of geometric structures, 130–132 strong-weak, 324
quantization, 197–198 T–, 116
space, 142, 155 Dubrovin–Mazzocco solution, 231
Degree, 402
of a map, 241, 281, 292 E polynomial, 332–333, 338
Del Pezzo surface, 388–390 Eigenfunction, 90
Deligne–Mumford compactification, 416 Eilenberg–MacLane space, 404
Derived category, 300, 304–306 Einstein
Descent data, 249 –Hilbert action, 23
Desuspension, 408, 413 –Maxwell equations, 3, 17–30
Determinant line bundle, 93 –Weyl manifolds, 3
Developing map, 138 equations, 1
Diffeomorphism, 75, 148, 186, 353, 359, 370, manifold, 1
379, 388, 397, 417 metric, 2, 28
area-preserving, 71 universe, 158
Differential Elliptic
form, 14, 76, 80, 134, 241, 267, 328, complex, 56
351–356, 363, 369–372, 386, curve, 213, 231, 302–305, 385
388–389 fibration, 299–321
Hopf, 148, 151 function, 3, 7
ideal, 357 integral, 6, 215
operator, 75, 101, 117, 181–206 operator, 406–407
quadratic, 116 regularity, 355
Dihedral Embedding, 348–366, 392
group, 6 Energy, 74, 281–282, 293, 326, 395
solution, 220, 233 of a harmonic map, 148–149
symmetry, 285–287 Equation(s)
Dimensional reduction, 13, 215, 327, Bogomolny, 3, 12, 118, 283
397 BPS, 398
Dirac Cauchy–Riemann, 348
equation, 34, 55 Dirac, 34, 55
operator, 1–2, 25, 37, 256 Einstein, 1
structure, 370–371, 377 Einstein–Maxwell, 3, 17–30
Dirichlet problem, 71, 83 Euler–Lagrange, 74, 78, 87
Discriminant Fuchsian, 213, 218, 228–229
Hitchin, 103 heat, 6
locus, 303 Hitchin, 113–127
Distribution, 271 hypergeometric, 210–213, 228–230
Divisor, 48, 66, 164, 267, 276, 303 KdV, 12
at infinity, 52 Lamé, 213
class, 92 Lax, 46
normal crossing, 332 Maxwell, 1, 17
Dixmier–Douady class, 245–254, 258 monad, 67
Dodecahedron, 286, 293 Monge–Ampère, 82, 352
Index 425
Nahm, 3–5, 7, 36–39, 46, 67, 71–90 Form, see Differential form
Painlevé, 6, 214–234 Formal
Riccati, 215 connection, 199–200
Schlesinger, 217 trivialization, 188
Seiberg–Witten, 3, 17–30 Fourier
self-duality, 2, 34, 327 –Laplace transform, 224–225
Skyrme, 281 –Mukai
Yang–Mills, 2–6, 13, 215, 327 transform, 299–321
Equivariant K-theory, 395 Framing, 400
Euler Fredholm operator, 37, 56
φ function, 325 Fricke space, 142–148
–Lagrange equations, 74, 78, 87 Frobenius
characteristic, 329, 413 manifold, 213
class, 104, 144–145 ring, 395, 400, 414
Exact sequence, 54 Fubini property, 415
Exponential mapping, 365 Fuchsian
Exterior equation, 213, 218, 228–229
algebra, 406 representation, 129, 137–148, 154
differential system, 8, 348–366 Fukaya–Floer category, 300
Extremal Kähler metric, 17–30 Fullerene, 283
Function
F4 , 227–228 Green, 59, 271
Faddeev–Mickelsson anomaly, 256 harmonic, 274
Family Morse, 171
of elliptic complexes, 407 plurisubharmonic, 269, 273
of surfaces, 415–417 Poisson commuting, 6, 101
Fano spaces, 58, 61
manifold, 303–305, 368 theta, 6–7
Poisson manifold, 390 Fundamental
Federer theorem, 274 form, second, 347
Fibration group, 5, 92, 98, 113, 122, 129–130, 138,
elliptic, 299–321 158, 215–218, 230, 327, 395
Lagrangian, 299–321 Fusion ring, 395
of groupoids, 237
special Lagrangian, 337 G2 , 348
Fibre product, 240–242, 399 manifold, 15, 358–365
Filtration, 49 structure, 8
weight, 332 Galois group, 94
Finite field, 331–332, 340 Gauge
Flag, 41 group, 86, 114, 327
Flat theory, 123, 188, 398–399
bundle, 13, 113, 118, 147, 307 Gauss
connection, 5, 96, 115, 179–181, 213, 217, –Bonnet formula, 24
319, 333, 375–379, 399–400, 416 –Manin connection, 213–217
line bundle, 133 theorem, 72
structure, 213, 347 Gelfand transformation, 266
Floer homology group, 306 Generalized
Flop, 304 cohomology theory, 399
Flow complex structure, 8, 15, 114, 368–392
geodesic, 158 connection, 368–374
of a vector field, 389 holomorphic bundle, 374–377
of hypersurfaces, 8 Kähler manifold, 4, 368
of line bundles, 49 submanifold, 392
of maps, 171 Genus, 13, 180
of sheaves, 46 geometric, 30
of structures, 348 of Painlevé curve, 219
Fock space, 256 Geodesic, 3, 78
Foliated structure, 158 flow, 158
426 Index
star operator, 12, 17, 73, 114, 134, 328, Calabi–Yau, 315
354 complex, 218
structure, 130
theory, 98, 134, 328–330 Icosahedral
theory, hyperkähler, 324–342 group, 210–220, 232
Holomorphic solution, 6, 232–234
p chain, 261 symmetry, 282, 286
bundle, 115, 117, 120, 164–217 Immersion, 347
curve, 300 Index theorem, 1, 404–407
differential form, 134 Infinite-dimensional
line bundle, 129, 264, 302, 305, 383 Lie group, 13
map, 124, 165–167, 190 manifold, 74
Poisson structure, 368–392 Instanton, 2–4, 14, 34, 325, 337
Poisson variety, 380–388 bubbling, 37
representation, 165 Integer cohomology, 402
section, 263, 267, 276, 376, 391–392 Integrable system, 5–6, 100, 116, 130, 337
vector bundle, 151, 175 Integral manifold, 353, 358
Holonomy, 239–241, 247, 250–256, 300, 337, Intersection
346–366, 400 cohomology, 330
function, 187 form, 340–342
group, 74 pairing, 113, 267
representation, 138, 145, 154 Intrinsic torsion, 346–366
Homogeneous space, 353 Invariant
Homology Seiberg–Witten, 27
class, 264 Toledo, 146–147
Floer, 306 Involution, 54, 405
Homotopy, 241 anti-holomorphic, 110
class, 186–187 hyperelliptic, 181
equivalence, 404 Involutive
null, 407–411 ideal, 354–357
theory, 168–175 tableau, 351
Hopf differential, 148, 151 Isomonodromic
Hull deformation, 122–123, 215–222
λ, 277 family, 229
projective, 262–265, 269 Isothermal migration, 85
Hurewicz theorem, 133 Israel–Wilson ansatz, 30
Hyperbolic
affine sphere, 154 Jacobian, 101, 305
cone, 139–145 variety, 92, 129, 134
geometry, 129 Jimbo’s formula, 222–223
space, 137 Jones polynomial, 5, 178, 188
structure, 137–144, 151 Jordan curve, 146
surface, 137 Jump
triangle group, 229 of holomorphic structure, 49
Hypercohomology, 100 of isomorphism class, 165
Hyperconvex curves, 155–156
Hyperelliptic involution, 181 K-theory, 156, 168, 395–417
Hypergeometric twisted, 8
connection, 212 K3 surface, 1–2, 314
equation, 210–213, 228–230 Kähler
Hyperkähler class, 20, 119
Hodge theory, 324–342 cone, 29
manifold, 4–8, 13, 119, 314, 351–358 form, 147, 267, 326
metric, 324–342 manifold, 126, 130, 189, 197, 265,
quotient, 13, 127, 326–330 269–270
structure, 89, 115, 135–136 metric, 89, 188–194, 275–276
Hypersurface, 8, 73, 347–366 potential, 74, 87
algebraic, 263, 275, 278 surface, 18
428 Index
Kac–Moody isomorphism, 43
algebra, 40, 342 simplicial, 244
extension, 257 Linear
group, 254 category, 168, 173
monopole, 66 system, 41
KdV equation, 12 Link, 177
Killing–Coxeter matrix, 224–225 Linking number, 261–263, 267–269,
Klein 275–276
connection, 226 affine, 268
polynomial, 286 projective, 262
theorem, 229 reduced, 262
Knot, 188 Local
Kodaira dimension, 29 bundle gerbe, 244–246
Kontsevich conjecture, 301 Locality, 164–175
Locally product metric, 350
L2 Logarithmic connection, 215–225
cohomology, 341 Loop, 293
harmonic form, 7, 324–342 bundle, 257
Hermitian structure, 202 group, 66, 395
inner product, 74 space, 171–175, 238, 254, 404
norm, 88 Lorentzian space, 84
norm of curvature, 20, 36
section, 38, 58, 180–181 Möbius transformation, 284
solution, 12, 44, 64 Macdonald polynomial, 341
Lagrangian, 85 Magnetic monopole, see Monopole
cycle, 307–308 Manifold, see also Variety
fibration, 299–321 affine, 308–316
fibre, 100, 116 ALE, 14, 29, 325, 330
Poincaré cycle, 313–319 Calabi–Yau, 7–8, 300–321, 337
submanifold, 378–379, 388 Einstein, 1
subspace, 158 Einstein–Weyl, 3
torus, 116 Fano, 303–305, 368
Lamé equation, 213 Frobenius, 213
Langlands G2 , 15, 358–365
correspondence, 168 generalized Kähler, 4, 368
dual group, 165 hyperkähler, 4–8, 13, 119, 314, 351–358
duality, 113–127, 174 infinite-dimensional, 74
programme, 5, 13, 164, 338–339 integral, 353–358
Laplacian, 71, 90, 192 Kähler, 126, 130, 189, 197, 265, 269–270
Large complex structure limit, 301–304, 315 non-commutative, 174
Lattice, 308 one-, 400–401, 416
Lax equations, 46 projective, 261–278
Legendre transform, 81 quasi-projective, 264
Level, 411 Riemannian, 172, 397
Levi-Civita connection, 191, 346, 349, 356, seven-, 358
368–374 Stein, 265
Lie SU (2), 348–357
algebra, 86, 180, 185 symplectic, 113, 118–122, 129, 134,
bracket, 76 188–189
derivative, 75, 196 three-, 177, 188, 255–256
group, 13, 113, 125, 129, 130, 178, 240, two-, 113–127, 400–401, 405–409, 413
254–255, 258, 326, 346, 395–417 Map
Line bundle, 7, 92, 129, 165, 188–189, 238, conformal, 155
305, 388, 417 covering, 167
canonical, 13, 122, 300 developing, 138
flat, 133 harmonic, 148
Hermitian, 266, 376 holomorphic, 124, 165–167, 190
holomorphic, 302, 305 meromorphic, 171
Index 429