BMS Particlesin Three Dimensions
BMS Particlesin Three Dimensions
BMS Particlesin Three Dimensions
in Three Dimensions
Blagoje Oblak∗
arXiv:1610.08526v3 [hep-th] 26 Jan 2018
Abstract
∗
Current e-mail: [email protected]
Université Libre de Bruxelles
BMS Particles
in Three Dimensions
Blagoje Oblak
— electronic version —
This thesis was defended behind closed doors on June 22nd, and publicly on June
24th, in front of the following jury at Université Libre de Bruxelles:
President: Prof. Marc Henneaux
Secretary: Prof. Petr Tinyakov
Advisor: Prof. Glenn Barnich
The work presented in this thesis was carried out in the group of Mathematical Physics of Fun-
damental Interactions at Université Libre de Bruxelles (Belgium), and at the Department
of Applied Mathematics and Theoretical Physics of the University of Cambridge (United
Kingdom). It was supported by the Fonds de la Recherche Scientifique – F.N.R.S., of which
the author is a Research Fellow (aspirant), by the Wiener-Anspach Foundation, and by the
International Solvay Institutes.
Printed in June 2016 by Presses Universitaires de Bruxelles ASBL — Avenue Franklin Roo-
sevelt 50, C.P. 149, B-1050, Belgium.
Foreword xiii
How to read this thesis ...................................................................................... xiii
Acknowledgements............................................................................................. xiv
1 Introduction 1
1.1 Asymptotic BMS symmetry ...................................................................... 1
1.2 Global BMS and extended BMS ............................................................... 2
1.3 Holography ................................................................................................ 6
1.4 BMS particles and soft gravitons .............................................................. 8
1.5 Plan of the thesis....................................................................................... 10
I Quantum symmetries 13
2 Quantum mechanics and central extensions 15
2.1 Symmetries and projective representations ............................................... 15
2.1.1 Quantum mechanics .......................................................................... 15
2.1.2 Symmetry representation theorem...................................................... 16
2.1.3 Projective representations ................................................................. 18
2.1.4 Central extensions ............................................................................ 19
2.1.5 Topological central extensions............................................................ 20
2.1.6 Classifying projective representations ................................................. 21
2.2 Lie algebra cohomology ............................................................................. 22
2.2.1 Cohomology ..................................................................................... 22
2.2.2 Central extensions ............................................................................ 25
2.3 Group cohomology .................................................................................... 27
2.3.1 Cohomology ..................................................................................... 27
2.3.2 Central extensions ............................................................................ 29
3 Induced representations 31
3.1 Wavefunctions and measures ..................................................................... 32
3.1.1 Measures .......................................................................................... 32
vii
Contents viii
4 Semi-direct products 61
4.1 Representations and particles.................................................................... 61
4.1.1 Semi-direct products ......................................................................... 61
4.1.2 Momenta .......................................................................................... 63
4.1.3 Orbits and little groups ..................................................................... 64
4.1.4 Particles ........................................................................................... 66
4.1.5 Exhaustivity theorem ........................................................................ 70
4.2 Poincaré particles ...................................................................................... 71
4.2.1 Poincaré groups ................................................................................ 71
4.2.2 Orbits and little groups ..................................................................... 73
4.2.3 Particles ........................................................................................... 77
4.2.4 Massive characters ............................................................................ 78
4.2.5 Massless characters ........................................................................... 81
*4.2.6 Wigner rotations and entanglement.................................................... 83
4.3 Poincaré particles in three dimensions ...................................................... 85
4.3.1 Poincaré group in three dimensions .................................................... 85
4.3.2 Particles in three dimensions ............................................................. 87
4.3.3 Characters ........................................................................................ 89
*4.4 Galilean particles....................................................................................... 90
4.4.1 Bargmann groups ............................................................................. 90
4.4.2 Orbits and little groups ..................................................................... 92
4.4.3 Particles ........................................................................................... 94
4.4.4 Characters ........................................................................................ 95
12 Conclusion 381
Bibliography 385
Index 407
Foreword
This thesis collects thoughts and results that originate from a four-year long research
project in theoretical physics. The main topic is representation theory and its appli-
cation to quantum gravity, in particular in the context of BMS symmetry.
The original contributions of this thesis are based on the following publications:
• G. Barnich and B. Oblak, “Holographic positive energy theorems in three-di-
mensional gravity,” Class. Quant. Grav. 31 (2014) 152001, 1403.3835.
• G. Barnich and B. Oblak, “Notes on the BMS group in three dimensions: I.
Induced representations,” JHEP 06 (2014) 129, 1403.5803.
• G. Barnich and B. Oblak, “Notes on the BMS group in three dimensions: II.
Coadjoint representation,” JHEP 03 (2015) 033, 1502.00010.
• B. Oblak, “Characters of the BMS Group in Three Dimensions,” Commun. Math.
Phys. 340 (2015), no. 1, 413–432, 1502.03108.
• G. Barnich, H. A. González, A. Maloney, and B. Oblak, “One-loop partition
function of three-dimensional flat gravity,” JHEP 04 (2015) 178, 1502.06185.
• B. Oblak, “From the Lorentz Group to the Celestial Sphere,” Notes de la Sep-
tième BSSM, U.L.B. (2015). 1508.00920.
• A. Campoleoni, H. A. González, B. Oblak, and M. Riegler, “Rotating Higher
Spin Partition Functions and Extended BMS Symmetries,” JHEP 04 (2016)
034, 1512.03353.
xiii
Foreword xiv
Acknowledgements
Professional community
This thesis could not have been completed without the help and support of a number
of people. First and foremost I am indebted to my supervisor, Prof. Glenn Barnich, for
suggesting the topic in the first place and guiding me through its completion. Working
with Glenn has been both a delight and a challenge; the former for his contagious pas-
sion for physics, and the latter for his sharp critical mind and unwavering skepticism,
leading to numerous lively discussions about the nature of our work and of science
altogether. Our complementary approaches have made our collaboration all the more
fruitful; I am grateful to him for preserving my complete freedom while guiding me
along the thesis. He has been a teacher, a friend, and a model of independence and
scientific integrity that I hope to emulate myself.
I also wish to thank my other collaborators. The earliest ones in the history of
my Ph.D. are Sophie De Buyl, Stéphane Detournay and Antonin Rovai, with whom I
spent a few weeks in Harvard in the Spring of 2013. I am grateful to them for their
friendship and our many relaxed discussions about physics and other matters, includ-
ing world economy, Belgian movies and American food. Later on my path crossed that
of Hernán González, starting with several ultra local journal clubs on quantum gravity,
and ending with common projects. I am grateful to him for sharing his enthusiasm
and almost convincing me that three-dimensional gravity is, in fact, gravity. My grat-
itude also goes to Hamid Afshar, Daniel Grumiller, Alexander Maloney, Max Riegler,
and especially Andrea Campoleoni, for fruitful and enjoyable scientific collaborations;
I hope there will be many more in the future.
physics would be impossible without the precious help and efficiency of the adminis-
trative staff of the service and the Solvay Institutes — Dominique Bogaerts, Fabienne
De Neyn, Marie-France Rogge, Isabelle Van Geet, and Chantal Verrier. To them,
thank you for your unshakeable goodwill in the face of the fiercest of administrative
challenges.
Since October 2015 I have had the chance to meet a number of new colleagues at
the Deparment of Applied Mathematics and Theoretical Physics of the University of
Cambridge. This would not have been possible without the support of Harvey Re-
all, whom I wish to thank warmly for this incredible opportunity. I am also grateful
to Siavash Golkar, Shahar Hadar, Sasha Hajnal-Corob, Kai Roehrig, Arnab Rudra,
Joshua Schiffrin and Piotr Tourkine for making my time there a daily enjoyment.
Still on the international side, I would like to thank Matthias Gaberdiel and Bianca
Dittrich for giving me the opportunity to give seminars at their respective institutions.
Both of these visits were a delight, and I had great pleasure in sharing some thoughts
about my research with them and their colleagues.
My friends have also contributed greatly to this thesis, often unknowingly. Above
all I wish to thank David Alaluf, Mitia Duerinckx, Jihane Elyahyioui, Geoffrey Mullier,
Olmo Nieto-Silleras, and Roxane Verdikt for our many adventures together, includ-
ing holidays and unforgettable parties (in the Balkans and elsewhere). As a matter of
fact, Mitia Duerinckx has been a great help in understanding some of the mathematics
used in this thesis, always patiently replying to my many e-mails with questions about
Hilbert spaces, measures and the like. More generally, I am grateful to my friends from
the mathematics department of ULB for their friendship and sometimes professional
interactions, including Charel Antony, Cédric De Groote, Julien Meyer and Patrick
Weber. Finally, I am indebted to Thierry Maerschalk for helping me out long ago with
some cute LATEX tricks, many of which were used in this thesis.
Foreword xvi
On the other side of the Channel I wish to thank the flurry of people whom I
have had the chance to meet during the last year of my Ph.D. and who have allowed
me to take helpful breaks away from physics. This includes Abhimanyu Chandra,
Robert Cochrane, Marius Leonhardt, Dušan Perović, Frank Schindler, Bianca Schor,
and Aaron Wienkers. I hope to meet them again in the future, despite our living in
different corners of the world.
Finally, I wish to thank Vanessa Drianne, who has been in an entangled state
between Brussels and Cambridge for most of the past year and whose support has
been critical for my well-being in the last stages of writing the thesis.
Introduction
1
Chapter 1. Introduction 2
Asymptotic symmetries of gravitational systems have been studied for about fifty
years by now. Their first appearance in the literature is also the one that motivates
the present work. Indeed, it was observed in the sixties by Bondi, van der Burg,
Metzner [1, 2] and Sachs [3, 4] that the presence of gravitation in an asymptotically
flat space-time leads to a symmetry group that is much, much larger than standard
Poincaré. The group that they found turned out to be an infinite-dimensional exten-
sion of the Poincaré group, and is known today as the Bondi-Metzner-Sachs group, or
BMS group for short.
The BMS group considered by the authors of [2–4] consists of two pieces: the
first is the standard Lorentz group of special relativity, and the second is an infinite-
dimensional Abelian group of so-called supertranslations.1 In abstract mathematical
notation, its structure can be written symbolically as
The notation n used here means that elements of the BMS group are pairs consisting
of a Lorentz transformation and a supertranslation, and that Lorentz transformations
act non-trivially on supertranslations. In the same way, the Poincaré group is
Groups of the form (1.1) or (1.2) are known as semi-direct products. They are
ubiquitous in physics, and many of the conclusions of this thesis rely on this structure.
Bondi coordinates
Consider Minkowski space-time, endowed with inertial coordinates xµ in terms of which
the metric reads
ds2 = ηµν dxµ dxν , with (ηµν ) = diag(−1, +1, +1, +1). (1.3)
Now suppose we wish to study, say, outgoing massless particles sent by an observer
located at the spatial origin. For this purpose we introduce retarded Bondi coordinates
1/2 x1 + ix2
r ≡ xi xi u ≡ x0 − r.
, z≡ , (1.4)
r + x3
1
The terminology of “super-things” here has nothing to do with supersymmetry: “super-object”
simply means that a certain object, which one is familiar with in the finite-dimensional context of
special relativity, gets extended in an infinite-dimensional way in the BMS group.
Chapter 1. Introduction 3
Figure 1.1: The coordinates u and r in space-time. The time coordinate x0 points
upwards. The wavy red line represents an outgoing radial massless particle emitted at
r = 0 and moving to some non-zero distance r away from the observer at r = 0; the
particle moves along one of the generators of the light cone given by u = const. The
drawing is three-dimensional, so the circle of radius r in this picture would actually
be a sphere (spanned by the coordinate z) in a four-dimensional space-time.
The notion of asymptotic flatness allows one to define the associated asymptotic
symmetry group. The latter consists, roughly speaking, of diffeomorphisms of space-
time that preserve the asymptotic behaviour of the metric.2 Bondi et al. [2–4] found
2
More precisely, the asymptotic symmetry group is the quotient of the group of diffeomorphisms
that preserve the asymptotic behaviour of the metric by its normal subgroup consisting of so-called
Chapter 1. Introduction 5
This property relies on the isomorphism3 SO(3, 1)↑ ∼ = SL(2, C)/Z2 , which ex-
presses Lorentz transformations in terms of SL(2, C) matrices. Lorentz trans-
formations also act on the coordinates r and u at infinity by angle-dependent
rescalings, but this subtlety is unimportant at this stage.
• The second family consists of angle-dependent translations of retarded time,
where α(z, z̄) is any (smooth) real function on the sphere. In this language,
Poincaré space-time translations are reproduced by functions α which are linear
combinations of the functions
1 − z z̄ z + z̄ i(z − z̄)
1, , , .
1 + z z̄ 1 + z z̄ 1 + z z̄
(In terms of polar coordinates θ and ϕ, this corresponds to the spherical har-
monics Y00 (θ, ϕ) and Y1,m (θ, ϕ) with m = −1, 0, 1.) This is the main surprise
discovered by Bondi et al. It states that asymptotic symmetries (as opposed to
isometries) enhance the Poincaré group to an infinite-dimensional group with
an infinite-dimensional Abelian normal subgroup consisting of transformations
(1.7). These transformations are the supertranslations alluded to in eq. (1.1).
Extended BMS
The group of asymptotic symmetry transformations (1.6) and (1.7) is the original
BMS group discovered in [2–4]. It consists of globally well-defined, invertible trans-
formations of null infinity, so from now on we call it the global BMS group. This
slight terminological alteration is rooted in one of the most intriguing aspects of BMS
symmetry. Indeed, in their work, Bondi et al. observed that asymptotic symmetries
include conformal transformations (1.6), but in principle one may even include trans-
formations generated by arbitrary (generally singular) conformal Killing vector fields
on the celestial spheres. Only six of those vector fields generate the invertible Möbius
transformations (1.6); the remaining ones are singular. Upon including these extra
generators, the global conformal transformations (1.6) are enhanced to arbitrary local
conformal transformations
z 7→ f (z) + O(1/r), (1.8)
where f (z) is any meromorphic function. Despite their singularities, these transfor-
mations do preserve the asymptotic behaviour of the metric and may therefore qualify
trivial diffeomorphisms. We will return to this in section 8.1.
3
O(3, 1) is the Lorentz group in four dimensions and SO(3, 1)↑ is its largest connected subgroup.
Chapter 1. Introduction 6
It was recently suggested by Barnich and Troessaert [5, 6] that extended (as op-
posed to global) BMS symmetry is the true, physically relevant symmetry of asymp-
totically flat gravitational systems in four dimensions (see also footnote 17 of [7]).
This proposal is motivated by a similar symmetry enhancement occurring in two-
dimensional conformally-invariant systems: while their global symmetry algebra is
finite-dimensional, they turn out to enjoy a much richer infinite-dimensional symme-
try. This observation first appeared in a seminal paper by Belavin, Polyakov and
Zamolodchikov [8] and triggered the development of two-dimensional conformal field
theory (CFT).
Thus, the truly thrilling aspect of extended BMS symmetry is the prospect of
applying conformal field-theoretic techniques to gravitational phenomena in four di-
mensions. This reduction from four to two dimensions is reminiscent of holograms, and
indeed the notion of “holography” in quantum gravity is one of the main motivations
that led to these considerations.
1.3 Holography
The elementary concept of holography in quantum gravity is simple: it is the state-
ment that gravitational phenomena occurring in a certain space-time manifold can be
described equivalently in terms of some lower-dimensional, “dual” theory. This idea is
originally due to ’t Hooft [9] and Susskind [10], who were led to it by model-independent
considerations. In particular, holography is compatible with the Bekenstein-Hawking
entropy formula [11,12], according to which the entropy of a black hole is proportional
to the area of its horizon. (The keyword here is “area”, as opposed to the “volume”
expected on the basis of standard thermodynamics.)
AdS/CFT
In practice, the first genuine illustration of holography in a concrete model of quantum
gravity — namely string theory — was exhibited by Maldacena [13], initiating what
4
Strictly speaking, local conformal transformations do not span a group but a semi-group, and
the same applies to extended BMS. This abuse of terminology is pretty common, and it will be
inconsequential for the discussion of this introduction.
Chapter 1. Introduction 7
The proposal that BMS symmetry might account for gravitational physics in
asymptotically flat space-times is similar in spirit to AdS/CFT. The problem is that
most known holographic constructions rely on the key assumption that the bulk space-
time is endowed with a negative cosmological constant, i.e. that it is of the AdS type.
This leads to a natural question: how should one deal with holography in flat space?
In this thesis we argue that the BMS group in three dimensions [20], or BMS3 ,
provides such a toy model. We shall see that it displays the extended structure (1.9) in
a simplified and controlled setting, and successfully accounts for many aspects of three-
dimensional asymptotically flat gravity, both classically and quantum-mechanically.
The BMS3 group is the main actor of this work and we will use it to develop our
intuition on flat space holography in general, including the four-dimensional case.
In hindsight this observation is not too surprising. Indeed, Klein’s Erlangen pro-
gramme [21] posits that geometric statements can be recast in the language of group
theory. This point of view has led to numerous developments in mathematics through-
out the twentieth century (including e.g. the work of Poincaré on special relativity).
Since general relativity is essentially the dynamics of pseudo-Riemannian geometry,
it is natural that the programme should apply to it as well provided one identifies
the correct symmetry group. In particular, holography may sometimes be seen as an
Erlangen programme in disguise.5
Since the Poincaré group is a subgroup of BMS, it provides a first rough picture
of what one should expect from BMS representations. Indeed, irreducible unitary
representations of Poincaré are, by definition, particles: they are classified by their
5
See e.g. the Wikipedia page https://en.wikipedia.org/wiki/Erlangen_program.
Chapter 1. Introduction 9
mass and spin [23] and their Hilbert space accounts for the available one-particle states.
The BMS group is expected to generalize this notion in a way that incorporates certain
gravitational effects. It should describe the quantum states of a particle, plus some
extra degrees of freedom accounting for the fact that BMS is only an asymptotic, rather
than an exact, symmetry group. Guided by this picture, we introduce the following
terminology:
This thesis is devoted to the description and classification of such particles in three
space-time dimensions.
Accordingly, the classification of BMS particles that we expose in this thesis may
be thought of as a classification of all possible ways to dress a Poincaré particle with
soft gravitons. A word of caution is in order: since we will be working in three space-
time dimensions, the gravitational field will have no local degrees of freedom so there
will be no genuine gravitons. In particular, the name “soft graviton” is ambiguous, as
there is no actual graviton whose zero-energy limit would be a soft particle. However,
asymptotic symmetries precisely account for soft graviton degrees of freedom, so we
shall adopt the viewpoint that any system with non-trivial asymptotic symmetries
does indeed have non-trivial soft degrees of freedom. This amounts to using the words
“soft graviton” as a synonym for the more standard “topological” or “boundary degree
of freedom”. In particular, three-dimensional gravitational systems generally do have
highly non-trivial asymptotic symmetries [14, 37] and therefore possess soft degrees of
freedom in this sense. In this language, the statement that three-dimensional gravity
has no bulk degrees of freedom turns into the fact that the only non-trivial degrees of
freedom of three-dimensional gravity are soft.
Remark. Unitary representations of the globally well-defined BMS group (1.1) have
already been classified by McCarthy and others in [38–40], and it was indeed suggested
in [41] that BMS symmetry is relevant to particle physics in that it provides a better
definition of the notion of “particle”. However, these representations appear to miss the
fact that supertranslations create soft gravitons when acting on the vacuum, which
is crucial for the application of BMS symmetry to soft theorems. In this sense the
Chapter 1. Introduction 10
Quantum symmetries
The first part of the thesis deals with the implementation of symmetries in quantum
mechanics through projective unitary representations, which are worked out in detail
for the Poincaré groups and the Bargmann groups. It consists of four chapters.
Quantum symmetries generally act in a projective way, which is to say that the
group operation of the underlying symmetry group is represented up to certain con-
stant phases. The presence of such phases is captured by central extensions of the
symmetry group. Accordingly, chapter 2 is devoted to central extensions and to the
more general notion of group and Lie algebra cohomology. Chapter 3 then explains
how one can build Hilbert spaces of wavefunctions on a homogeneous space endowed
with a unitary action of a symmetry group. This involves the important notion of
induced representations, which we discuss in detail.
Finally, chapter 5 describes the relation between classical and quantum symmetries
through geometric quantization. In a nutshell this relation is obtained by defining a
space of wavefunctions on what is known as a coadjoint orbit of a symmetry group. For
semi-direct products this approach reproduces the classification of representations by
momentum orbits and leads to a group-theoretic version of the world line formalism.
In chapter 8 we show how Virasoro symmetry emerges in AdS3 gravity with Brown-
Henneaux boundary conditions, after explaining some basic notions on asymptotic
symmetries in general. We also show that the phase space of AdS3 gravity is embedded
as a hyperplane at constant central charge in the space of the coadjoint representa-
tion of two copies of the Virasoro group. As an application we relate highest-weight
representations of the Virasoro algebra to the quantization of gravitational boundary
degrees of freedom.
partition function coincides with the character of a unitary representation of the BMS3
group [48, 49]. For higher spins in three dimensions we similarly obtain characters of
flat non-linear WN algebras [50]. Along the way we describe unitary representations
of these algebras [51] and show that they differ qualitatively from earlier proposals in
the literature. We end by describing certain supersymmetric extensions of the BMS3
group, their representations, and their characters.
Note that our notation is not the standard Dirac notation of bras and kets: a
vector in H is denoted as Ψ (not |Ψi), and its dual is the linear form hΨ|·i on H .
1
Recall that a metric space is complete if any Cauchy sequence converges.
15
Chapter 2. Central extensions 16
where Ψ is some non-zero state vector. The vanishing vector does not represent a
quantum state, so the set of mutually inequivalent pure states is the projective space
PH = (H \0)/C. It is the set of one-dimensional subspaces of H . Stated differently,
the set of distinct states in H is the quotient of the unit sphere in H by the equivalence
relation
Ψ ∼ eiθ Ψ for all θ ∈ R. (2.4)
We shall denote by [Ψ] the resulting equivalence class of Ψ. For example, in a two-level
system where H = C2 , the set of inequivalent states is CP 1 ∼ = S 2.
Now let the system be in a state [Ψ]. If  is an observable and if λ is one of its
eigenvalues with eigenvector Φ say, the probability of finding the value λ is
|hΦ|Ψi|2
Prob(λ, Â, [Ψ]) = . (2.5)
hΦ|ΦihΨ|Ψi
(We are assuming for simplicity that the eigenvalue λ is not degenerate.) Note that
this expression is independent, as it should, of the choice of both the representative Ψ
of the state [Ψ], and the eigenvector Φ.
Remark. In quantum mechanics, one generally assumes that the Hilbert space is
separable, i.e. that it admits a countable basis. Any such space is isometric to the space
`2 (N) of square-integrable sequences of complex numbers — so there really exists only
one infinite-dimensional separable Hilbert space. This is not to say that all separable
Hilbert spaces describe the same quantum system, because the definition of a system
also involves the set of observables that act on it — and identical Hilbert spaces may
well come with very different operator algebras.
Accordingly, the framework suited for the study of symmetries is group theory.
In this thesis we will be concerned with Lie groups, consisting of symmetry transfor-
mations that depend smoothly on a certain number of real parameters. This number is
the dimension of the group. In part I of the thesis, all Lie groups are finite-dimensional.
Remark. The notion of symmetry can be relaxed in such a way that not all pairs of
symmetry transformations are allowed to be composed together. The resulting set of
symmetry transformations then spans a groupoid rather than a group (see e.g. [58,59]).
This relaxed notion of symmetry is relevant to gauge theories [60], and in particular to
BMS symmetry in four dimensions [18]. However, standard group theory suffices for
all symmetry considerations in three-dimensional gravity (and in particular for BMS3 ),
so we will not deal with groupoids in this thesis.
for all normalized vectors Φ, Ψ, Φ0 , Ψ0 such that Φ0 ∈ S([Φ]) and Ψ0 ∈ S([Ψ]). The
key result on symmetries in quantum mechanics is the following [22]:
for all λ, µ ∈ C and all Φ, Ψ ∈ H . A proof of this theorem can be found in chapter 2
(appendix A) of [53].
Note that symmetries represented by antiunitary operators only arise when the
symmetry group is disconnected. For example, in Lorentz-invariant theories, time-
reversal is always represented in an antiunitary way (see e.g. [61]). In this work
we will restrict attention to connected symmetry groups, in which case all symmetry
operators are linear and unitary. In particular they satisfy Û † = Û −1 , where Hermitian
conjugation is defined by (2.2).
Chapter 2. Central extensions 18
T : G → GL(H ) : g 7→ T [g]
[T ] : G → GL(H )/C∗ : f 7→ T [f ]
(2.8)
is a homomorphism. Here GL(H )/C∗ is the projective group of H , i.e. the quotient
of the linear group of H by its normal subgroup consisting of multiples of the iden-
tity. For any operator O in GL(H ), the symbol [O] denotes its class in the projective
group. Throughout this thesis, any map T satisfying this property will be called a pro-
jective representation. In quantum mechanics, symmetries are represented by unitary
projective representations, i.e. projective representations whose operators are unitary.
From now on, if we wish to stress that a representation is not projective, we will
call it exact. Quantum mechanics tells us that exact representations are overrated: the
truly important ones are generally projective. This seemingly anecdotal observation is
at the core of the richest aspects of the representation theory of the Virasoro algebra,
and it will also play a key role for BMS3 particles. For instance, all interesting two-
dimensional conformal field theories are such that the conformal group is represented
projectively in their Hilbert space, and how exactly this phenomenon takes place is
measured by the central charge. For this reason, this whole chapter is devoted to the
various ways in which projective effects occur; they are accounted for by group and
Lie algebra cohomology.
3
Throughout this thesis representations of groups are denoted by the letters R, S, T , etc. The
letter G denotes a group whose elements are written f , g, h, etc. The identity in G is denoted e.
Chapter 2. Central extensions 19
Remark. Since we are focussing on Lie groups, the representations of interest are
continuous in the sense that the map G × H → H : (f, Ψ) 7→ T [f ] · Ψ is continuous.
From now on it is understood that all representations are continuous.
whose elements are pairs (f, λ), endowed with a group operation
(f, λ) · (g, µ) = f · g, λ + µ + C(f, g) . (2.11)
The group (2.10) is called a central extension of the group G. We will study this
notion in much greater detail in section 2.3. For now let us only work out the basic
consequences of this structure and its relation to representation theory.
The question then is whether G admits central extensions to begin with. For any
group, an obvious type of central extension always exists. Namely, suppose K is a real
function on G and define C : G × G → R by
This question leads to group (and Lie algebra) cohomology, studied in detail in
sections 2.2 and 2.3. For now we simply point out that central extensions may arise
via two distinct mechanisms. The first is algebraic in that it follows from the local
group structure of G, or equivalently from the commutation relations of its Lie algebra.
In short, in some cases, the Lie algebra g of G can be enlarged into a bigger algebra
g which contains extra generators commuting with those of g (see eq. (2.27) below).
b
The group corresponding to this enlarged algebra then is a central extension of G.
The second mechanism is topological in the sense that it is due to the global structure
of G. We now describe this topological mechanism in some more detail.
Owing to the fact that the map T is continuous, the phase φ(γ) only depends on
the homotopy class of γ. In addition, if γ1 and γ2 are two closed paths starting at f , we
can concatenate them into a single path γ1 · γ2 (which is γ1 at double speed followed by
γ2 at double speed); the phase φ must be compatible with this operation in the sense
that eiφ(γ1 ) · eiφ(γ2 ) = eiφ(γ1 ·γ2 ) . Thus, any one-dimensional unitary representation of the
fundamental group of G, multiplying an exact unitary representation of G, produces
a projective unitary representation of G.
Remark. One might be worried by the fact that only one-dimensional unitary repre-
sentations of the fundamental group are allowed to appear in this construction. Indeed,
if the fundamental group was non-Abelian, it would generally admit no non-trivial
one-dimensional unitary representation. Fortunately, it turns out that the fundamen-
tal group of any finite-dimensional Lie group is a discrete commutative group, whose
irreducible unitary representations are necessarily one-dimensional.
4
Beware: a manifold being multiply connected means that it has a non-trivial fundamental group,
and not that it has several connected components.
Chapter 2. Central extensions 21
Figure 2.1: The group U(1) is diffeomorphic to a circle S 1 , whose universal cover is the
real line R. The projection R → S 1 ∼= R/Z is obtained by identifying points of R that
differ by some periodicity, typically θ ∼ θ + 2π. In particular, paths in R which are
not closed may be projected on closed paths in S 1 . As an application we can picture
topological projective representations: if T is projective and if γ is a closed path in
the circle, the sequence T [γ(t)] may not be a closed path in the space of operators.
The example just described occurs in Nature. Indeed, fermions provide a well-
known example of projective representations, as already suggested above by the case
s = 1/2. By the spin-statistics theorem, all fermions have half-integer spins, and
therefore transform according to a projective representation of the Lorentz group.
The latter is multiply connected (its fundamental group is Z2 ), which is why it admits
projective representations in the first place. We will return to the representation theory
of the Lorentz group (as a subgroup of Poincaré) in much greater detail in section 4.2.
In the cases where arbitrary real values of spin are allowed by quantum mechanics,
as for example in three space-time dimensions, the particles whose spin is neither an
integer nor a half-integer are known as anyons. We will encounter this phenomenon
in section 10.1 when dealing with BMS3 particles.
that problem:
2.2.1 Cohomology
Let g be a Lie algebra with Lie bracket [·, ·]. We recall that a representation of g in
V is a linear map T : g → End(V) such
a vector space that T [X] ◦ T [Y ] − T [Y ] ◦
T [X] = T [X, Y ] for all Lie algebra elements X, Y . 5
for all X1 , ..., Xk+1 in g; the hat denotes omission. Note that the representation T of
g in V appears explicitly in this definition. In particular, when T is trivial, formula
(2.15) simplifies since its last line disappears.
Isomorphic Lie algebras have the same cohomology for any choice of the represen-
tation T . Thus, cohomology is a way to associate invariants with Lie algebras: if two
algebras have different cohomology spaces, then they cannot be isomorphic. This is
analogous to, say, de Rham cohomology in differential geometry, as manifolds with
different de Rham cohomologies cannot be diffeomorphic.
H1 (g) ∼
= g/[g, g] , (2.19)
Definition. A Lie algebra g is perfect if g = [g, g], i.e. if any Lie algebra element can
be written as the bracket of two other elements.
It follows from (2.19) that g is perfect if and only if H1 (g) vanishes. We will use
this property in section 2.2.2 when defining central extensions.
Remark. Here we have been using the word “deformation” in a vague way, but there
is an exact definition of the notion of deformations. Namely, a (true) deformation of
a Lie algebra g is a Lie algebra g̃ that coincides with g as a vector space, but whose
brackets are
˜[X, Y ˜] = [X, Y ] + c(X, Y ) (2.22)
where [·, ·] is the bracket in g while c is a g-valued two-cocycle on g,7 such that the
image of c belongs to its kernel. The latter condition means that c X, c(Y, Z) = 0 for
all Lie algebra elements X, Y, Z; together with the fact that c is a cocycle, this ensures
that (2.22) is a Lie bracket.
Examples
For finite-dimensional semi-simple Lie algebras, cohomology is trivial:
7
It is understood that the relevant representation of g in this case is the adjoint, T [X]·Y ≡ [X, Y ].
Chapter 2. Central extensions 25
where cab = c(ta , tb ), while all brackets with Z vanish. The cocycle condition on c then
becomes the requirement
for the coefficients cab . Note that this construction can be readily generalized to
g = g ⊕ RN , in which case there are N central generators
multiple central extensions b
Z1 , ..., ZN .
Chapter 2. Central extensions 26
For example, as on page 25, consider the Abelian Lie algebra R2 and let c be a
non-zero antisymmetric bilinear form on R2 . We then define the three-dimensional
Heisenberg algebra as the algebra R3 = R2 ⊕ R whose elements are pairs (X, λ),
endowed with the Lie bracket (2.26). Since c is non-trivial, so is the central extension.
If we choose a basis {Q, P } of R2 such that c(Q, P ) = 1 and if we call Z the central
element (0, 1), the commutation relations of the Heisenberg algebra take the form
[Q, P ] = Z. (2.29)
Property (2.24) says that there is only one linearly independent central extension of
R2 , i.e. that Heisenberg algebras built using different (non-zero) two-cocycles c are
mutually isomorphic. This can be generalized to higher dimensions: by seeing R2n as
an Abelian Lie algebra and taking c an arbitrary non-zero 2n-form on R2n , the Lie
algebra defined by the bracket (2.26) is the (2n + 1)-dimensional Heisenberg algebra.
Definition. A central extension b g of g is universal if, for any other central extension
g0 of g, there exists a unique isomorphism of Lie algebras b
b g0 ∼
=bg.
As it turns out, any perfect Lie algebra admits a universal central extension. By
virtue of (2.19), this is to say that any algebra such that H1 (g) = 0 admits a universal
central extension. For example we will see in chapter 6 that the Virasoro algebra is
the universal central extension of the Lie algebra of vector fields on the circle.
Chapter 2. Central extensions 27
2.3.1 Cohomology
Let G be a Lie group, T : G → GL(V) a representation of G in a vector space V.
(dk C)(g1 , ..., gk+1 ) ≡ T [g1 ] · C(g2 , ..., gk+1 ) + (−1)k+1 C(g1 , ..., gk )
k
X (2.31)
+ (−1)i C(g1 , ..., gi gi+1 , ..., gk+1 )
i=1
The differential (2.31) satisfies the key property (2.16), so the usual machinery of
homological algebra applies: one defines a k-cocycle as a closed k-cochain, that is,
a cochain C such that dk C = 0. One also defines a k-coboundary to be an exact k-
cochain, i.e. one that can be written as the differential of a (k − 1)-cochain. As before
any coboundary is trivially a cocycle, so one defines the k th cohomology space of G with
values in V as the quotient of the space of k-cocycles by the space of k-coboundaries:
Interpretation
As in the case of Lie algebras, cohomology spaces are invariants that measure the flex-
ibility of a group structure; isomorphic Lie groups have the same cohomology. This
interpretation is simplest to illustrate with the cohomology spaces of lowest degree.
Given a one-cocycle S, one defines the associated affine module as the space V ⊕ R
acted upon by the following representation Tb of G:
Tb [f ] · (v, λ) ≡ T [f ] · v + λ S(f ), λ . (2.33)
Proposition. Let G be a Lie group, g its Lie algebra. Let V be a vector space, T a
smooth representation of G in V, and T the representation of g corresponding to T
by differentiation. Then, for any non-negative integer k, there is a homomorphism
given by
" #
∂k X
i1 ...ik C eti1 Xi1 , ..., etik Xik
δC(X1 , ..., Xk ) ≡
∂t1 ...∂tk 1≤i1 <...<ik ≤k t1 =0,...,tk =0
for all X1 , ..., Xk in g, with eX the exponential of X ∈ g and i1 ...ik the Levi-Civita
symbol with k indices (and 12...k ≡ +1). For k = 2 this can be rewritten as
∂2 h tX sY
sY tX
i
δC(X, Y ) = C e ,e − C e ,e . (2.35)
∂t ∂s t=0, s=0
We will use formula (2.35) in section 6.2 to relate the Virasoro algebra to the
Virasoro group. The key point here is that any differentiable group cocycle C admits
an algebraic analogue δC. The converse problem is to start from a Lie algebra cocycle,
say c, and ask whether there exists a group cocycle whose differential is c. This is
the problem of integrating Lie algebra cocycles to group cocycles, and it is generally
much more complicated than differentiation. However, for “sufficiently connected” Lie
groups, the Van Est theorem states that integration is trivial because group and Lie
algebra cohomologies coincide (see e.g. [56]). In particular, when the universal cover
of a group is homotopic to a point, the cohomology of the universal cover coincides
with that of the Lie algebra.
Example. Let us find the group corresponding to the (2n + 1)-dimensional Heisen-
berg algebra. Consider the Abelian additive group G = Rn × Rn (whose elements are
pairs of column vectors (α, β)) and define the Heisenberg group as
1 αt λ
n
G≡
b 0 In β α, β ∈ R , λ ∈ R
(2.36)
0 0 1
where In denotes the n × n identity matrix and αt is the transpose of α. The group
operation in G
b is given by matrix multiplication and can be written as
(α, β, λ) · (α0 , β 0 , λ0 ) = α + α0 , β + β 0 , λ + λ0 + αt · β 0
(2.37)
By differentiation, one can associate with C a Lie algebra cocycle given by (2.35). For
example, when n = 1 (and writing elements of the Lie algebra R2 as pairs X = (x, y)),
0 0
(2.35) ∂ 2
δC (x, y), (x , y ) = (tx · sy 0 − sx0 · ty)|t=0, s=0 = xy 0 − yx0 .
∂t ∂s
This is a non-zero antisymmetric bilinear form on R2 , hence defining the Heisenberg
algebra of (2.29). Note that this is an example of “cocycle integration”: we have found
the explicit group two-cocycle whose differential defines the Heisenberg Lie algebra.
As in the algebraic case, there is a simple criterion for knowing when a group
admits a universal central extension. A group is said to be perfect if it coincides with
the group of its commutators, i.e. if any f ∈ G can be written as f = ghg −1 h−1
for some g, h ∈ G. It turns out that any perfect group admits a universal central
extension. In chapters 6 and 9 we will see that both Diff(S 1 ) and BMS3 are perfect
groups, so that their central extensions are universal.
Chapter 3
Induced representations
In the previous chapter we learned how to deal with projective representations: given
a symmetry group, we are to find its universal cover and its most general central ex-
tension. Exact representations of this central extension then account for all projective
representations of the original group. The remaining problem then is to write down
explicit representations, so our goal in this chapter is to build Hilbert spaces of wave-
functions acted upon by a group of unitary transformations. Guided by group actions
on homogeneous spaces, we will be led to the method of induced representations. Their
basic principle is very simple: starting from a representation of some subgroup H of a
group G, one induces a representation of G that acts on wavefunctions which live on
the quotient space G/H.
The plan of this chapter is as follows. In section 3.1 we review some basics of
measure theory and the ensuing construction of Hilbert spaces of square-integrable
wavefunctions. Section 3.2 is concerned with measures on homogeneous spaces and
introduces quasi-regular representations — the simplest examples of induced represen-
tations. In section 3.3 we display the basic formulas of induced representations and
list some of their elementary properties. Along the way we define a basis of plane
waves, later to be interpreted as particles with definite momentum. This basis is then
used in section 3.4 to compute characters. Finally, section 3.5 is devoted to systems
of imprimitivity. All these notions are crucial prerequisites for chapter 4.
31
Chapter 3. Induced representations 32
It would be illusory to present a complete account of the rich theory of induced rep-
resentations, so we refer to Barut and Raczka [62] or Mackey [63] for a more thorough
exposition. For some background on measure theory, see e.g. [64, 65].
3.1.1 Measures
When defining a quantum-mechanical system, one of the key ingredients is a prescrip-
tion for computing scalar products. For the spaces of wavefunctions that we wish
to consider, this requires being able to evaluate integrals of functions on a manifold.
Integration, in turn, relies on the existence of a measure.
Measures
Let M be a set. Roughly speaking, a measure is a function µ that associates a non-
negative number with essentially any subset U of M. That number, denoted µ(U ),
“measures” the size of U . Strictly speaking, not all subsets of M can be measured:
there exists a family of subsets of M, called “measurable sets”, and only those subsets
can actually be measured. The measure µ then is a map
where R̄+ denotes the set of non-negative real numbers supplemented with +∞. In
order to qualify as a measure, this map needs to satisfy certain conditions; in particular,
it must be σ-additive: if U1 , U2 , etc. are disjoint measurable sets, then
+∞
! +∞
[ X
µ Ui = µ(Ui ) when Ui ∩ Uj = ∅ ∀ i, j. (3.2)
i=1 i=1
In other words, the total measure of a set consisting of several disconnected compo-
nents must be the sum of the measures of the individual components. A measure µ
on M is said to be finite if µ(M) is finite; it is σ-finite if M is a countable union of
measurable sets with finite measure (any finite measure is trivially σ-finite).
For instance, the standard translation-invariant Lebesgue measure on the real line
R is defined so that µ([a, b]) = b − a for any closed interval [a, b] ⊂ R; the measure
takes the same value for open or half-open intervals. In particular, R is a countable
union of intervals of finite length, so the Lebesgue measure is σ-finite. This definition
is readily generalized to Rn .
Chapter 3. Induced representations 33
Borel measures
Throughout this chapter and the next ones, we systematically endow M with a topol-
ogy. One can take advantage of this structure when defining a measure:
Thus, Borel measures are compatible with the topology of M. In particular, any
continuous function F : M → N is Borel-measurable. From now on, all measures are
understood to be Borel. When M is a smooth manifold, the data of a Borel measure
is equivalent to that of a volume form on M. For simplicity, we always assume that
M is a manifold.
Integrals
Measures can be used to integrate functions.1 Let µ be a Borel measure on M and
U ⊆ M a Borel set. When V is a topological vector space and F : M → V : q 7→ F(q)
is a measurable function, the (Lebesgue) integral of F over U relative to the measure
µ is written as2 Z Z
F(q) dµ(q) or F dµ .
U U
In these terms, the measure µ(U ) of a Borel set U is the integral of the function
F(q) = 1 over U : Z
µ(U ) = dµ(q). (3.3)
U
The word “measure” often also refers to the quantity dµ appearing in this expression.
where M 2 is a positive parameter (the mass squared) while q = (q1 , ..., qD−1 ) is the
spatial momentum in D space-time dimensions.
(·|·) : E × E → C : v, w 7→ (v|w) ,
which we take to be linear in its second argument and antilinear in the first one. When
E = C we simply set (v|w) = v ∗ w.
Wavefunctions
Definition. Let M be a topological space, µ a Borel measure on M, E a complex
Hilbert space with scalar product (·|·). Then an E-valued square-integrable wavefunc-
tion is a measurable map Ψ : M → E such that
Z
dµ(q) Ψ(q)|Ψ(q) < +∞.
M
It is tempting to turn L2 (M, µ, E) into a Hilbert space by declaring that the scalar
product of two wavefunctions is the integral of their product over M, but there is a
problem: wavefunctions need not be continuous. In particular, functions that vanish
everywhere on M except at some countable number of points, are strictly speaking
non-zero vectors in L2 even though all their would-be scalar products vanish. In
the language of conformal field theory, those are “null states”. In order to cure this
pathology, one introduces the following notion:
This space is also simply called the (E-valued) L2 space on M relative to the measure µ.
Strictly speaking we should write the left-hand side of this definition as k[Ψ]k2 , where
[Ψ] ∈ L2 is the class3 of Ψ ∈ L2 . However, we will systematically abuse notation
by choosing arbitrarily a representative Ψ of a class [Ψ], and we use the word “wave-
function” to refer both to actual functions Ψ : M → E and to the corresponding
equivalence classes in L2 .
It can be shown that the space L2 is a complete normed vector space, i.e. a Banach
space, with respect to the norm (3.6). In addition the space of (equivalence classes of)
smooth functions with compact support is dense in L2 (M, µ, E), so any wavefunction
can be approximated with arbitrary precision by a smooth function.
With this definition we can start interpreting L2 (M, µ, E) as the space of states of
some quantum system. In Dirac notation we would write wavefunctions as Ψ ≡ |Ψi,
which is indeed suggested by the notation (3.7). The quantum state defined by such
a wavefunction is a ray (2.3)consisting of all functions M → E that are equal almost
everywhere to some constant multiple of Ψ. (Again, the notation [·] in (2.3) does not
3
This class has nothing to do with the ray (2.3) despite the identical notation.
Chapter 3. Induced representations 36
Remark. When dealing with unitary representations of the BMS3 group in part
III, we will need to describe square-integrable wavefunctions on infinite-dimensional
manifolds (see section 10.1). Until then we will not discuss this issue.
Equivalent measures can be vastly different, yet they are still pretty much the same
with regard to measure theory:
In addition, any other function ρ̃ satisfying this property coincides with ρ almost ev-
erywhere on M. The function ρ is called the Radon-Nikodym derivative of ν with
respect to µ.4 A proof of this theorem can be found in [66].
For example, we mentioned below (3.3) that when dn x is the Lebesgue measure
on Rn , any non-negative function ρ gives rise to a new measure ρ(x)dn x. The Radon-
Nikodym derivative of that measure with respect to the Lebesgue measure then coin-
cides with the function ρ. In particular, when ρ(x) only vanishes on a set of Lebesgue
measure zero, the measures dn x and ρ(x)dn x are equivalent.
Remark. When µ and ν are equivalent measures, one has dµ(q)/dν(q) ∼ [dν(q)/dµ(q)]−1 ,
i.e. the Radon-Nikodym of µ with respect to ν is (almost everywhere) the inverse of
the Radon-Nikodym of ν with respect to µ.
Isomorphic L2 spaces
The notion of equivalent measures allows us to address the question raised above,
namely whether the Hilbert spaces L2 (M, µ, E) and L2 (M, ν, E) differ if the measures
µ and ν differ.
There exist many important examples of group actions in physics: the space Rn
can be seen as an Abelian group acting on itself by the addition of vectors; the sphere
S 2 is acted upon by rotations. More generally, any group representation is a linear
action of a group on a vector space; in particular the energy-momentum of a particle
in Minkowski space is acted upon linearly by the Lorentz group.
Op ≡ {f · p|f ∈ G} . (3.13)
Gp ≡ {f ∈ G|f · p = p} . (3.14)
Homogeneous spaces
Definition. An action of a group G on M is said to be transitive when for any two
points p, q ∈ M there exists a group element f such that f · p = q. The space M is
5
As before elements of G are written as f , g, etc. and the identity is denoted e.
Chapter 3. Induced representations 39
In particular a homogeneous space coincides with the orbit of any of its points
under the group action: M = Op for any p ∈ M. It follows that any homogeneous
space can be written as a coset space (3.15).
The simplest example of a G-homogeneous space is the group G itself, with the
action given by left multiplication:
g 7−→ Lf (g) = f g. (3.16)
The stabilizer in that case is trivial. Note that right multiplication
g 7−→ Rf (g) = gf (3.17)
is not quite a group action since Rf ◦Rg = Rgf does not coincide with Rf g . This can be
cured by considering right multiplication by inverse elements, i.e. g 7→ Rf −1 g = gf −1 .
In the aforementioned example of Rn , seen as an Abelian group acting on itself by
the addition of vectors, left and right multiplications coincide. (This is true for any
Abelian group.) The sphere S 2 is a more interesting example of homogeneous space,
since it is acted upon transitively by the group of rotations SO(3) but has a non-trivial
stabilizer SO(2), and is therefore diffeomorphic to the quotient SO(3)/SO(2). More
generally, one has a family of diffeomorphisms S n ∼ = SO(n + 1)/SO(n). Homogeneous
spaces will play a central role in representation theory, so we will encouter many more
examples of transitive actions later in this thesis.
Quasi-invariant measures
Given a homogeneous space M, we wish to integrate functions over it and we ask
whether there exists an invariant measure. It turns out that this is not always the
case (see e.g. [62]), so one introduces the following weaker notion of invariance:
is equivalent to µ.
∂f n
dµf (x) = dn [f (x)] = d x, (3.21)
∂x
where |∂f /∂x| is the Jacobian of f . Thus the Radon-Nikodym derivative of a quasi-
invariant measure can also be seen as a generalization of the Jacobian.
Each operator T [f ] is manifestly linear, and the fact that this is indeed a representation
follows from the fact that the map q 7→ f · q is a group action. The interpretation of
formula (3.22) is simple: if the wavefunction Ψ is sharply centred around some point k
of M, then the operator T [f ] maps Ψ on a new wavefunction, now centred around the
point f · k. In chapter 4 the space M will consist of the allowed momenta of a particle,
Ψ will be the particle’s wavefunction (in momentum space), and the map q 7→ f · q
will be an action by boosts or rotations.
The far right-hand side generally does not coincide with the original scalar product
(3.7) because it involves the Radon-Nikodym derivative (3.19). Thus, in order to
ensure unitarity, we need to correct formula (3.22) by a factor that compensates the
non-trivial transformation law of µ:
where we relied on the fact that q 7→ f · q is a group action. Now using (3.20) and the
definition (3.24), we can rewrite this as
1/2
(T [f g]Ψ)(q) = ρg−1 (f −1 · q)ρf −1 (q) Ψ g −1 · (f −1 · q)
(3.25)
1/2 −1
= [ρf −1 (q)] (T [g]Ψ) (f · q) = T [f ] ◦ T [g] · Ψ (q) , (3.26)
which proves that (3.24) is indeed a representation. To complete the proof we also have
to show that T is unitary for the scalar product (3.7) with (Φ(q)|Ψ(q)) = Φ∗ (q)Ψ(q).
Repeating the computation (3.23) we now find that the Radon-Nikodym derivative in
(3.24) yields an extra term in the integrand. Using (3.20) and the fact that ρe = 1, this
term cancels the Radon-Nikodym derivative in (3.23) so (3.24) is indeed unitary.
Remark. When the homogeneous space M coincides with the group G and is en-
dowed with the invariant Haar measure, formula (3.22) defines a unitary representation
of G known as the regular representation. Quasi-regular representations extend this
concept by trading the base manifold G for an arbitrary homogeneous space M.
Accordingly, the representations Tµ and Tν are unitarily equivalent. This is to say that
the quasi-regular representation (3.24) is essentially independent of the measure µ.
(d log ρ)f g (q) = log ρg−1 (f −1 · q) + log ρf −1 (q) − log ρg−1 f −1 (q) ,
dµ(f · q)
= eΨ(f ·q)−Ψ(q) , i.e. e−Ψ(q) dµ(q) = e−Ψ(f ·q) dµ(f · q) , (3.29)
dµ(q)
which says that the quasi-invariant measure µ is actually an invariant measure in
disguise! Indeed, the measure ν defined by dν(q) = e−Ψ(q) dµ(q) is invariant by virtue
of (3.29). In other words, the first cohomology of G with values in the space of func-
tions on M classifies the inequivalent quasi-invariant measures on M, two measures
being equivalent if they are related to one another by a function that multiplies them.
In particular, the first cohomology vanishes if all quasi-invariant measures on M are
equivalent to an invariant measure.
We can also rephrase this in the language of representation theory: the quasi-
regular representation (3.24) is a (multiplicative) affine module (2.33) on top of the
original representation (3.22). Two such modules are equivalent if the corresponding
measures are equivalent.
M∼
= G/H . (3.30)
Now consider the point p ∈ M, identified with the identity coset eH in G/H. Since
M is a homogeneous space one can map p on any other point q ∈ M, with a group
element g ∈ G such that g · p = q. Given q, this group element is only defined up to
multiplication from the right by an element of H, since h · p = p for any h ∈ H.
As given here, formula (3.32) comes a bit out of the blue, so it is worth analysing
its elementary features. First note that, if we denote the trivial representation of a
group by the symbol 1, then IndG H (1) is the quasi-regular representation (3.24) of G
G
on G/H while Ind{e} (1) is the regular representation. Formula (3.32) extends these
constructions by including a non-trivial action of H in an internal space E. Before
interpreting (3.32) any further, we now verify that it is a consistent definition.
Consistency
Up to the term involving S, the right-hand side of (3.32) coincides with the quasi-
regular representation (3.24), so the only potential problem could arise from the in-
sertion of S. But since the map (3.31) is a family of standard boosts, we have
gq−1 f gf −1 ·q · p = gq−1 · f · gf −1 ·q · p = gq−1 · f · f −1 · q = gq−1 · q = p ,
and using the fact that S is a representation of H, one can mimick the sequence of equa-
tions (3.25)-(3.26) for T given by (3.32), which implies that it is indeed a representa-
tion. As forunitarity, it follows from the fact that S is unitary: S[h]Φ(q) S[h]Ψ(q) =
Φ(q) Ψ(q) for all wavefunctions Φ, Ψ, any point q ∈ M and any h ∈ H.
Chapter 3. Induced representations 46
Interpretation
The basic interpretation of the induced representation (3.32) is the same as for (3.24):
Ψ(f −1 · q) represents the fact that the wavefunction Ψ is “boosted” by f , while the
√
factor ρf −1 ensures unitarity. The new ingredient is the combination
S gq−1 f gf −1 ·q ≡ Wq [f ].
(3.35)
Its appearance represents the fact that, in contrast to the quasi-regular representation,
wavefunctions take their values not in C, but in some more general “internal” Hilbert
space E carrying a representation S of H.
Robustness
Formula (3.32) depends not only on the inducing data (the group G, its subgroup H
and a spin representation S), but also on the measure µ on M ∼ = G/H and on the
choice of a family of standard boosts gq . Naively, one expects all these parameters to
affect the representation. We now show that this is not the case.
As far as the measure is concerned, one readily verifies that two induced repre-
sentations defined with the same inducing data and the same standard boosts but
Chapter 3. Induced representations 47
T [f ] ◦ V = V ◦ T 0 [f ] ∀ f ∈ G. (3.37)
Proof. The fact that V is a unitary operator follows from unitarity of S. Property
(3.37) then follows from the definitions (3.32) and (3.36).
A corollary of these observations on robustness is that one may unambiguously say
“the” representation of G induced by the representation S of H, without any reference
to the measure or to the choice of standard boosts.
Let E be a Hilbert space with scalar product (·|·). We call conjugation the map
∗ ∗
C : E → Ects : v 7→ (v|·), where Ects denotes the space of continuous linear functionals7
on E. Then, if S is a unitary representation acting on E, its conjugate representation
is S ≡ C ◦ S ◦ C −1 . In the context of induced representations, one can then show that8
IndG G
H (S) ∼ IndH (S), i.e. the representation induced by the conjugate of S is unitarily
equivalent to the conjugate of the representation induced by S.
7
Recall that any continuous linear functional on a Hilbert space E is a scalar product (v|·) for
some fixed vector v ∈ E.
8
The symbol ∼ denotes unitary equivalence of representations.
Chapter 3. Induced representations 48
One can similarly show that induced representations behave well under direct sums
and tensor products thanks to the unitary equivalences
IndG G G
H (S1 ⊕ S2 ) ∼ IndH (S1 ) ⊕ IndH (S2 ) ,
(3.38)
IndG G G
H (S1 ⊗ S2 ) ∼ IndH (S1 ) ⊗ IndH (S2 ) .
One should also check that induction itself is a “good” operation on representations.
This is guaranteed by the theorem of induction in stages: let H1 be a closed subgroup
of H2 , which itself is a closed subgroup of G. Let S be a unitary representation of H1 .
Then one has the following unitary equivalence of representations:
H2
IndG G
H1 (S) ∼ IndH2 IndH1 (S) .
Delta functions
Let M ∼ = G/H be a homogeneous space, µ a quasi-invariant measure on M. Pick
a point k ∈ M. We define the Dirac distribution δk at k associated with µ as the
distribution such that hδk , ϕi ≡ ϕ(k) for any test function ϕ on M. Equivalently, we
introduce a “Dirac delta function” δ(k, ·) such that
Z
hδk , ϕi = dµ(q)δ(k, q)ϕ(q) ≡ ϕ(k). (3.39)
M
Thus the distribution δk acts on a test function ϕ(·) by integrating it against the delta
function δ(k, ·).
Note that the definition of the delta function δ relies on the measure µ since the
combination dµ(q)δ(k, q) is G-invariant by design. To make this explicit, let us denote
by δµ the delta function associated with µ. If ρ is some positive function on M and
dν(q) = ρ(q)dµ(q) is a new measure, then the delta function δν associated with ν
differs from δµ by a factor ρ:
δµ (k, q)
δν (k, q) = . (3.40)
ρ(q)
In particular, since µ is quasi-invariant under G, for any f ∈ G we have
δµ (k, q)
δµ (f · k, f · q) = (3.41)
ρf (q)
Chapter 3. Induced representations 49
The best known Dirac distribution is the one associated with the translation-
invariant Lebesgue measure on Rn . We will encounter this delta function repeatedly
so we denote it by δ (n) to distinguish it from other Dirac distributions. With that
notation, the delta function associated with the Lorentz-invariant measure (3.4) is
p
δ(k, q) = M 2 + q2 δ (D−1) (k − q). (3.42)
Scalar products of plane waves can be evaluated thanks to the definition (3.7).
Using the fact that the basis of ek ’s is orthonormal and the definition (3.39) of the
delta function, one finds
hΨk,` |Ψk0 ,`0 i = δ(k, k 0 )δ``0 . (3.44)
This property allows us to see in which sense plane waves form a “basis” of the Hilbert
space H = L2 (M, µ, E). Indeed, any wavefunction Φ : M → E can be written as
N
X
Φ(q) = Φ` (q) e` where Φ` (k) = hΨk,` |Φi .
`=1
Removing the argument q, this says that any wavepacket Φ is a superposition of plane
waves:
Z XN Z N
X
Φ= dµ(k) Φ` (k)Ψk,` = dµ(k) hΨk,` |Φi Ψk,` . (3.45)
M `=1 M `=1
9
We are assuming that E is a separable Hilbert space; N may be infinite.
Chapter 3. Induced representations 50
where I is the identity operator. In the more common (but less precise) Dirac notation
this would be a sum of projectors |Ψk,` ihΨk,` |. For example, for a Rparticle on the real
line, the Dirac formR of this completeness relation would read I = R dx|xihx| in posi-
tion space, or I = R dk|kihk| in momentum space. We will show in section 3.5 that
the existence of a family of projectors associated with M is one of the key properties
of induced representations.
From the construction of plane waves we see that induced representations are just
an upgraded version of one-particle quantum mechanics, with extra freedom in the
choice of the space M, the group G, and the spin states contained in E. This obser-
vation will guide us in developing our intuition of induced representations, especially
in part III of the thesis.
where we used (3.41) and the property (3.20). Note what we have achieved: in the
original definition (3.32) the argument of Ψ changes between the left and the right-
hand sides; here, by contrast, the argument will be the same, but what changes is the
label specifying the momentum of the plane wave. Indeed, using (3.47) in formula
(3.32), we find
q
T [f ] · Ψk,` (q) = ρf −1 (q)ρf (k) S gq−1 f gf −1 ·q · Ψf ·k,` (q).
Since the plane wave on the right-hand side contains a delta function δ(f · k, q), we
can replace all q’s in this expression
p by f · k and
p remove the argument from both sides.
Using once more (3.20) in ρf −1 (q)ρf (k) = ρf (k), we end up with
q
T [f ] · Ψk,` = ρf (k) S gf−1
·k f gk · Ψf ·k,` (3.48)
Chapter 3. Induced representations 51
This formula is the simplest rewriting of the induced representation (3.32). The only
extra improvement we could still add is to write as (S[· · · ])``0 the matrix element of
the operator S[· · · ] between the states e` and e`0 , whereupon (3.48) becomes
q
T [f ] · Ψk,` = ρf (k) S gf−1
·k · f · gk `0 ,` · Ψf ·k,`0 (3.49)
3.4 Characters
In this section we describe the characters associated with induced representations.
We start by motivating and defining characters in general terms, before proving the
Frobenius formula. We end by discussing the relation between characters and fixed
point theorems.
When a system is invariant under time translations, for instance, the correspond-
ing symmetry generator is the Hamiltonian operator H. The information about its
spectrum is captured by the canonical partition function 10
Z(β) = Tr e−βH ,
(3.50)
where β is the inverse of the temperature. If the system admits extra symmetries such
as, say, rotations, one can look for the maximal set of mutually commuting symmetry
10
The notation “Z” stands for the German word Zustandssumme, meaning “sum over states.”
Chapter 3. Induced representations 52
generators Qa and switch on their chemical potentials µa .11 The spectrum of these
new operators, together with H, is then contained in the grand canonical partition
function !
Xr
Z(β, µ1 , ..., µr ) = Tr exp − β H − µa Q a . (3.51)
a=1
Now suppose we take β to be purely imaginary (while keeping the µa ’s real) in this
expression. Then the operator inside the trace is unitary, since it is an exponential
of anti-Hermitian operators. In fact, it is a symmetry transformation acting in the
Hilbert space according to some unitary representation T , so we can write
Z(β, µ1 , ..., µr ) = Tr (T [f ])
for some element f belonging to the symmetry group G. This motivates the following
definition:
χ : G → C : f 7→ χ[f ] ≡ Tr (T [f ]) . (3.52)
This definition ensures that χ[f ] is independent of the basis of H used to evaluate
it. As an application, recall that two group elements f and f 0 are conjugate if there
exists an element g ∈ G such that f 0 = gf g −1 , and that the conjugacy class of f is
[f ] ≡ g f g −1 g ∈ G .
Thus, formula (3.52) ensures that characters are class functions in the sense that χ[f ]
only depends on the conjugacy class of f , and not on f itself:
Remark. The definition (3.52) suggests that characters are functions on G. While
this is true for finite-dimensional representations, it is not true in infinite-dimensional
ones. In fact, characters should not be seen as functions, but rather as distributions
[79]. Similarly to our dealing with Dirac distributions as if they were “delta functions”,
we will not take such mathematical subtleties into account.
where we “sum over momenta” thanks to the measure µ on M. Now using (3.48) and
the scalar products (3.44), we find
Z q N
X
χ[f ] = dµ(k) ρf (k) hψk,` |S[gf−1
·k · f · gk ]ψf ·k,` i (3.55)
M `=1
Z q N
(3.44) X
S[gf−1
= dµ(k) ρf (k) δ(k, f · k) ·k · f · gk ] ``
.
M `=1
Here the delta function δ(k, f ·k) allows us to trade f ·k for k. In particular the Radon-
Nikodym derivative ρf (k) = dµ(f · k)/dµ(k) reduces to unity. One then recognizes the
sum N
P
`=1 (S[· · · ])`` ≡ χS [· · · ] as the character of S, and eq. (3.54) follows.
The Frobenius formula (3.54) embodies the geometrization of representation the-
ory mentioned at the end of section 3.3.4: the trace of an operator has now become
an integral over a (subset of a) homogeneous space. That integral can be interpreted
as a sum of characters of S. Before studying this formula further, we need to check
that it satisfies the basic properties of a character.
First, since it is the character of an induced representation and since the latter
is independent (up to unitary equivalence) of the choice of the measure µ, the same
should be true of expression (3.54). To see that this is indeed the case, recall that
the combination dµ(k)δ(k, ·) is invariant under changes of measures (as follows from
the definition (3.39) of the Dirac distribution), which then implies invariance of the
character. Note in particular that the Radon-Nikodym derivative of the measure µ
does not appear in (3.54). Secondly, induced representations are independent of the
choice of standard boosts gq ; using the fact that the character of S is a class function,
one readily verifies that (3.54) is also independent of that choice. Finally, recall from
(3.53) that characters are class functions; using (3.20) one verifies that this is indeed
the case with formula (3.54). Note one crucial implication of this fact: because of
the term χS [gk−1 f gk ] in (3.54), the character χ[f ] vanishes if f is not conjugate to an
element of H. In other words the character of the induced representation IndG H (S) is
supported on the points of G whose conjugacy class intersects H.
Chapter 3. Induced representations 54
Remark. Since the character χS of the spin representation is a class function, one is
naively tempted to pull the term χS [gk−1 f gk ] out of the integral (3.54), as the notation
suggests that f is conjugate to gk−1 f gk . This is not true, because for generic g, g 0 ∈ G
one has χS [g −1 f g] 6= χS [g 0−1 f g 0 ]. As a consequence, the integral (3.54) is generally
non-trivial.
Figure 3.2: A manifold M acted upon by a rotation around some axis. The points
that belong to the axis are the only ones left fixed by the rotation, and are therefore
the only ones that contribute to the integral of formula (3.54).
Remark. The relation between characters of group representations and fixed point
theorems is much deeper and more general than the superficial description given here.
Indeed one can show [82] (see also [79]) that (3.54) coincides with the Lefschetz number
of T [f ] when the latter is seen as an endomorphism acting on a space of E-valued
sections on M. In turn, the fact that T [f ] is derived by (3.32) from a diffeomorphism
action of f on M turns out to imply that its Lefschetz number is given by the Atiyah-
Bott fixed point theorem [83].
consequential to the remainder of the thesis and may be skipped in a first reading.
We saw in eq. (3.46) that the identity operator I can be written as an integral of
projectors Ψk hΨk |·i = |Ψk ihΨk |. This leads to a seemingly random idea: why not
combine these projectors into more general operators? For example, if U is any Borel
subset of M, we can associate with it a projection operator
Z N
X
PU ≡ dµ(k) hΨk,` |·i Ψk,` . (3.56)
U `=1
In that language the identity operator is I = PM . As it turns out this idea is one of
the key properties of induced representations, and sparked the whole development of
the theory by Mackey in the fifties [84–86]. In particular it leads to the imprimitivity
theorem, which roughly states that any representation that admits a suitable family
of projectors (3.56) is necessarily induced. An important corollary of that result is the
fact that all irreducible unitary representations of semi-direct products are induced.
The plan of this section is the following. We first define the notion of systems of
imprimitivity as suitable families of projection operators, and show that any induced
representation admits such a family. We then state (without proof) the imprimitivity
theorem, which we eventually use to define a restricted notion of equivalence for in-
duced representations. The presentation is based on [62, 63], but our approach will be
heuristic at times; we refer to [87, 88] for a mathematically rigorous presentation.
Projection-valued measures
Let us put (3.56) in a more general context. Observe that, given the Borel set U , the
projector PU acts on wavefunctions by setting them to zero everywhere outside of U :
Ψ(q) if q ∈ U,
PU · Ψ (q) = (3.57)
0 otherwise.
• The map P is σ-additive in the sense that, if U1 , U2 , etc. are disjoint Borel sets,
PU1 ∪U2 ∪··· = PU1 + PU2 + · · · . (3.59)
The terminology here is inspired by measure theory: the map (3.58) is an operator
analogue of (3.1) and property (3.59) corresponds to (3.2). Thus a projection-valued
measure measures the “size” of a subset U not by a real number µ(U ), but by the rank
of a projector PU .
at any k ∈ M. Here we have used both our Pnotation and the standard Dirac one;
N
we have also introduced an operator Ik = `=1 |Ψ k,` ihΨk,` | such that the identity
operator in H is an integral I = M dµ(k)Ik . Analogously to (3.3), the operator PU is
R
the integral of dP over U . For a particle on the real line, for example, the projection-
valued measure in momentum space would read dP = dk |kihk| with k ∈ R.
Systems of imprimitivity
There is one property that makes the projection-valued measure (3.56) very special.
Namely, the transitive action of G on M gives rise to an action (3.48) on wavefunctions;
the latter, in turn, yields an action on the projectors |Ψk ihΨk |. So the fact that (3.56)
acts in the space of a representation provides a relation between the geometry of M
and the action of G on the projection-valued measure, which motivates the following
definition:
In this language the projectors (3.56) imply that any induced representation has a
transitive system of imprimitivity:
for all k ∈ M and any f ∈ G. We shall refer to (3.56) as the canonical system of
imprimitivity of the induced representation T .
Proof. Transitivity is obvious, so the only subtlety is proving (3.61). Let us pick a
group element f ∈ G and a Borel set U ⊆ M. We start from the definition (3.56) and
relate T [f ] ◦ PU ◦ T [f ]−1 to Pf ·U using formula (3.48) and the fact that T is unitary:
Z N
X
T [f ] ◦ PU ◦ T [f ] −1
= dµ(k)ρf (k) S[gf−1 −1
·k f gk ]Ψf ·k,` hS[gf ·k f gk ]Ψf ·k,` |·i.
U `=1
Here the sum over ` allows us to cancel the two S[· · · ]’s by unitarity. Using also
dµ(k)ρf (k) = dµ(f · k) and renaming the integration variable, the right-hand side
boils down to Pf ·U .
Remark. The word “imprimitive” means “which is not primitive” and was introduced
by Galois [89] in the context of permutation groups. The action of a group on a set
shuffles the elements of this set, and the action is imprimitive if these permutations
preserve some (non-trivial) partition of the set. In the present case the group G acts
on the Hilbert space H by the induced representations (3.48), and property (3.62)
says that this action preserves the partition of H into isomorphic subspaces Ek ∼ =E
with definite momentum k.
U ◦ PUS ◦ U −1 = PU (3.63)
The complete proof of this theorem can be found in [90] and is reproduced in [62].
Given the representation T and the system of imprimitivity P , the key subtlety is to
construct a Hilbert space E and a representation S of H in E. We will not dwell on
this proof any further, but turn now to some of its applications.
Chapter 3. Induced representations 58
In particular, if S1 and S2 are equivalent (in the usual sense), then so are the induced
representations T1 and T2 (regardless of their systems of imprimitivity). The converse
is not true, since the “if and only if” of (3.65) also involves the systems of imprimitivity
associated with T1 and T2 . In other words, saying that two induced representations
are equivalent without saying anything about their systems of imprimitivity is not
sufficient to conclude that they are induced from the same spin representation S.
As mentioned below eq. (3.38), irreducibility of S does not generally imply irre-
ducibility of the corresponding induced representation. In the next chapter we shall
state a stronger result for semi-direct products, but for now we display a theorem that
provides a slightly weaker criterion for the irreducibility of induced representations.
The latter theorem shows that a suitable notion of irreducibility is preserved along
the induction process, since an irreducible S will lead to an induced representation T
and a system of imprimitivity P which, together, will be considered irreducible in the
above sense. But the theorem does not say that an induced representation T on its
own is irreducible if it is induced from an irreducible S.
Chapter 4
Semi-direct products
In this chapter we introduce semi-direct products such as the Poincaré group, the
Galilei group and the Bargmann group. We describe their irreducible unitary rep-
resentations, which are induced from representations of their translation subgroup
combined with a so-called little group. We interpret these representations as particles
propagating in space-time and having definite transformation properties under the
corresponding symmetry group. This picture will be instrumental in our study of the
BMS3 group.
The plan is as follows. In section 4.1 we define semi-direct products and intro-
duce the key notions of momentum orbits, little groups and particles. We also explain
why irreducible unitary representations are always induced, and describe these rep-
resentations in general terms. The remaining sections are devoted to applications of
these considerations. In section 4.2 we describe relativistic particles, i.e. unitary rep-
resentations of the Poincaré group, with a particular emphasis in section 4.3 on the
three-dimensional setting (which will be useful when dealing with BMS3 ). Section
4.4 is devoted to non-relativistic particles, i.e. unitary representations of Bargmann
groups. Useful references include [62, 63] for the general theory, and [53, 91, 92] for its
application to Poincaré.
61
Chapter 4. Semi-direct products 62
Elements of G are then called rotations or boosts while elements of A are translations.
Note that, since A is a vector group, the inverse (4.3) of (f, α) is (f, α)−1 =
f , −σf −1 (α) . Relation (4.4) also simplifies to (g, β) · (e, α) · (g, β)−1 = (e, σg α).
−1
From now on the words “semi-direct product” and the notation G n A will always refer
to a group (4.1) with A a vector group. (This is why the second factor in (4.1) was
denoted “A” in the first place.)
• The Euclidean group in n space dimensions takes the form (4.1) where rotations
span the group G = O(n) while translations belong to A = Rn , with the action
σ of rotations on translations given by the vector representation of O(n).
• The Poincaré group in D space-time dimensions takes the form (4.1) where rota-
tions and boosts span the Lorentz group O(D−1, 1) while space-time translations
span A = RD (which is sometimes written RD−1,1 ); the action σ is the vector
representation of the Lorentz group.
• The BMS groups (1.1)-(1.9) all take the form (4.1) with G a specific non-Abelian
group and A an Abelian vector group of so-called “supertranslations”. A similar
structure will hold in three space-time dimensions.
Note that the definition of G n A singles out the normal subgroup A, so G and A
live on unequal footings. In particular the Lie algebra of G n A contains a non-trivial
Abelian ideal and is not semi-simple. This implies that, in contrast to simple Lie
groups, the representations of G n A must somehow distinguish the roles of G and A
by making them act on the carrier space in radically different ways. We will illustrate
this in the pages that follow (see e.g. formula (4.23)).
4.1.2 Momenta
Suppose we wish to build unitary representations of a semi-direct product G nσ A.
Where should we start? A simple approach is to note that the restriction to A of any
unitary representation of GnA is a (reducible) unitary representation of A. So instead
of directly looking for representations of G n A, let us consider the simpler problem of
building unitary representations of the group of translations, A.
where the right-hand side is a composition of linear operators. Assuming that A has
a countable basis so that α and β have components αi , β i , the derivative of (4.9) with
respect to β i yields
∂j R[α] = i − i∂j R[0] R[α] (4.10)
where each (−i∂j R[0]) is Hermitian by unitarity. Hence R[α] = exp[i(−i∂j R[0])αj ],
which can be diagonalized into a direct sum of multiplicative operators (4.8).
Remark. In proving that all irreducible unitary representations of A takes the form
(4.8), we relied crucially on eq. (4.9). The latter assumes that R is an exact repre-
sentation of A, which is not a restrictive assumption as long as there exists no central
extension of G n A that turns A into a non-Abelian group. The Poincaré groups, the
Bargmann groups and the BMS3 group all satisfy this property, so one may safely
assume that A is Abelian even upon switching on central extensions. By contrast, the
symmetry group of warped conformal field theories [93] is a semi-direct product whose
central extension makes translations non-Abelian [52].
Call H the Hilbert space of the representation T . Suppose there exists a subspace
E of H where translations are represented by multiplicative operators (4.8) with a
certain momentum p:
where IE is the identity operator in E. We shall refer to this property by saying that the
representation T “contains the momentum p”. Now pick some group element f ∈ G.
By virtue (4.6) and since T is a representation, one has
T [(e, α)] · T [(f, 0)] = T [(f, 0)] · T [(e, σf −1 α)]. (4.14)
1
More precisely the momentum vector is obtained by raising the indices of the covector p thanks
to some metric on A left invariant by G, but we will keep referring to p as the “momentum vector”.
2
Our notation here is not mathematically precise; we refer to [62] for a more rigorous treatment.
Chapter 4. Semi-direct products 65
One can then act with both sides of this equation on the space E; the last term on the
right-hand side produces a multiplicative operator (4.13) with α replaced by σf −1 α.
This operator is a c-number and therefore commutes with T [(f, 0)]. We conclude that,
on the space T [(f, 0)] · E ≡ E 0 , all translations are again represented by multiplicative
operators, but now with an additional insertion of σf −1 in the phase hp, αi:
This motivates the following definition for the action of boosts on momenta:
i.e. hσf∗ (p), αi ≡ hp, σf −1 αi for all translations α. This defines a representation σ ∗ of
G in the space of momenta, known as the dual representation corresponding to σ. To
reduce clutter, we will often denote it by
In terms of the dual representation (4.17) we can rewrite (4.15) as T [(e, α)] E 0 =
ihf ·p,αi
e IE 0 , where E 0 = T [(f, 0)] · E. Thus, whenever the representation T contains a
momentum p, compatibility with the structure of G n A implies that it also contains
the boosted momentum f · p, where f is any element of G. This is the answer to
the question (4.12): if there exists a momentum p such that (4.13) holds, then the
representation also contains all momenta that belong to the orbit (3.13) of p under G,
Op ≡ {f · p | f ∈ G} . (4.18)
Gp ≡ {f ∈ G | f · p = p} . (4.19)
The notion of orbits is perhaps the one most important concept needed to under-
stand representations of semi-direct products. We will encounter it repeatedly later
Chapter 4. Semi-direct products 66
on. Orbits hint at a geometrization of representation theory analogous to the one men-
tioned at the end of section 3.3, and therefore suggest that representations of G n A
are closely related to induced representations. In the next pages we will confirm this
intuition by showing how to associate representations of G n A with a given orbit.
Note that the action (4.16) of G on the space of momenta leaves the pairing (4.7) in-
variant in the sense that hf ·p, σf αi = hp, αi. This has an important implication: when
A is finite-dimensional it is isomorphic to its dual, so (4.7) defines a non-degenerate
bilinear form on A and the action σ ∗ of G on momenta is equivalent to σ. We shall see il-
lustrations of this in the Poincaré groups. By contrast, when A is infinite-dimensional,
σ ∗ may not be equivalent to σ despite the property hf · p, σf αi = hp, αi. This obser-
vation will be relevant to the BMS3 group in part III.
4.1.4 Particles
We now explain how to build irreducible unitary representations of GnA starting from
a momentum orbit Op . Inspired by the Poincaré group, we refer to such representations
as particles. We start by describing scalar particles and identify them with induced
representations of G n A. This identification will then allow us to introduce spin.
Scalar particles
Let p ∈ A∗ be a momentum with orbit (4.18). The latter is a homogeneous space
and therefore admits a quasi-invariant measure µ. Let then H = L2 (Op , µ, C) be the
Hilbert space of square-integrable wavefunctions in momentum space,
Ψ : Op → C : q 7→ Ψ(q). (4.20)
where (f, α) belongs to G n A and where T [(f, α)] is some unitary operator. Linearity
implies that the result should be proportional to Ψ, so
T [(f, α)] · Ψ (q) = (some number) × Ψ(some point on Op )
where the unknown quantities may depend on q, f and α. Note that the quantity
multiplying Ψ(· · · ) on the right-hand side must be a number, as opposed to an operator,
because Ψ takes its values in C (this will change upon adding spin). Now recall that
the reason for introducing orbits in the first place was to represent translations by
multiplicative operators (4.8). Accordingly the translation α in (4.21) should produce
a momentum-dependent phase factor:
We call this representation a scalar particle with momentum orbit Op . Note how
translations and rotations have radically different roles: translations multiply wave-
functions by momentum-dependent phase factors, while boosts move them around on
the orbit by changing their argument. In particular, pure translations act as
where Iq is the identity operator in the fibre at q. This is precisely the anticipated
expression (4.11).
Chapter 4. Semi-direct products 68
T = IndGnA
Gp nA (S) (4.27)
Spinning particles
To generalize (4.23), let R be an irreducible, unitary representation of Gp in some
space E and consider the spin representation
This reduces to (4.26) when R is trivial, and the corresponding induced representation
of G n A is
GnA GnA ihp,·i
T = IndG p nA
(S) = Ind Gp nA e R . (4.29)
Its action on wavefunctions is analogous to (3.32) and generalizes (4.23):
q
(T [(f, α)] · Ψ) (q) = ρf −1 (q) eihq,αi R[gq−1 f gf −1 ·q ] · Ψ(f −1 · q) , (4.30)
We call the representation (4.30) a spinning particle with spin R and momenta
belonging to Op . It is an irreducible unitary representation of G n A acting on the
Hilbert space H = L2 (Op , µ, E). The operator
R[gq−1 f gf −1 ·q ] ≡ Wq [f ] (4.31)
is the Wigner rotation (3.35) associated with f and q. It is the transformation that
corresponds to f in the space of internal degrees of freedom E at q and it entangles
momentum and spin degrees of freedom. The decomposition (4.25) still holds in the
spinning case, with Iq the identity operator in the fibre Eq ∼
= E at q.
From this point on, the whole machinery of induced representations applies to
unitary representations (4.30) of G n A. In particular they are independent of the
choice of the quasi-invariant measure µ and of the family of standard boosts gq . The
plane waves (3.43) provide a basis of the Hilbert space and represent one-particle states
with definite momentum and definite spin. They transform under G n A according to
q
T [(f, α)] · Ψk,` = ρf (k) eihf ·k,αi R gf−1
·k · f · gk · Ψf ·k,` . (4.32)
This is just formula (3.48) applied to (4.29); it is the plane wave analogue of (4.30).
Using this, one can go on and evaluate characters along the lines that led to the
Frobenius formula (3.54). One finds
Z
dµ(k) δ(k, f · k) eihk,αi χR [gk−1 f gk ] ,
χ[(f, α)] = Tr T [(f, α)] = (4.33)
Op
This theorem has enormous practical value: it provides the classification of all
irreducible unitary representations of a semi-direct product G nσ A when A is a vector
group. This classification can be performed thanks to the following algorithm:
2. We call set of orbit representatives a set of momenta that exhaust all orbits
in a non-redundant way, in the sense that (i) each orbit contains one of the
representatives, and (ii) different representatives belong to different orbits. Find
a set of orbit representatives, compute the little group of each representative, and
find standard boosts connecting each representative to the points of its orbit.
3. For each representative p with little group Gp , classify all irreducible unitary
representations of Gp . Given such a representation R, the associated induced
representation of G n A is (4.30).
We will illustrate this classification for the Poincaré groups in section 4.2 and for the
Bargmann groups in section 4.4, and of course for BMS3 in chapter 10.
which is precisely the statement (3.62) that the projection-valued measure dµ(q)Iq is
a system of imprimitivity for T on A∗ . The imprimitivity theorem then implies that
the representation T is induced. The last step of the proof consists in showing that,
if T is irreducible, then the measure µ in (4.34) localizes to a single momentum orbit.
We refer to [62] for details.
Chapter 4. Semi-direct products 71
The plan is the following. First we define the Poincaré group as a semi-direct
product of the Lorentz group with the group of space-time translations. Then we turn
to the classification of its momentum orbits and describe the corresponding particles.
We also compute their characters and end with the observation that Lorentz transfor-
mations generally entangle momentum and spin degrees of freedom.
The classification of relativistic particles was first performed by Wigner [23], and
their relation to wave equations was worked out in [94]. These results are among the
foundations of quantum mechanics and field theory; see e.g. [53, 62, 91, 92].
We write (α, α) ≡ α2 for any α. The sign of α2 determines whether α is time-like, null
or space-like, corresponding respectively to α2 < 0, α2 = 0 or α2 > 0.
Chapter 4. Semi-direct products 72
In this section we focus on the connected Lorentz group, i.e. the proper orthochro-
nous Lorentz group SO(D −1, 1)↑ . The latter satisfies an important property known as
standard decomposition: any proper, orthochronous Lorentz transformation is a prod-
uct f = R1 · Λ · R2 , where R1 and R2 are spatial rotations and Λ is a pure boost [95].
In what follows we often refer to SO(D − 1, 1)↑ simply as “the Lorentz group”.
Poincaré groups
Definition. The Poincaré group or inhomogeneous Lorentz group in D space-time
dimensions is the semi-direct product
IO(D − 1, 1) ≡ O(D − 1, 1) n RD (4.39)
Chapter 4. Semi-direct products 73
Figure 4.2: The four connected components of the Lorentz group. The upper left
component is the proper orthochronous Lorentz group SO(D−1, 1)↑ . It can be mapped
on the other components using parity and time reversal. In particular the proper
Lorentz group is generated by SO(D − 1, 1)↑ together with time reversal, while the
orthochronous Lorentz group is generated by SO(D − 1, 1)↑ together with parity.
The Poincaré group turns out to have no algebraic central extensions, so its only
non-trivial projective transformations are of topological origin. The Poincaré Lie al-
gebra is generated by D(D − 1)/2 Lorentz generators and D translation generators;
we will not display their brackets here.
where the components of (α, ·) are those of α lowered with the Minkowski metric.
Using I, one verifies that the action σ ∗ of Lorentz transformations on momenta is
equivalent to their action σ on translations:
σf∗ = I ◦ σf ◦ I −1 . (4.43)
The connected Poincaré group has six distinct families of momentum orbits, which
we now describe. Further details can be found e.g. in [91, 92].
• Let p = 0 be the vanishing momentum. Its orbit O0 = {0} contains a single
point. Its little group is the whole Lorentz group.
• Let p be a timelike momentum with positive energy, p0 > 0. Its orbit Op is
massive with positive energy and consists of momenta q satisfying
where we have introduced the mass squared M 2 ≡ −p2 . We can choose as orbit
representative the rest frame momentum
Gp = SO(D − 1) (4.46)
consisting of proper Lorentz transformations that leave the time coordinate fixed.
In particular, the orbit is diffeomorphic to the quotient
Op ∼
= SO(D − 1, 1)↑ /SO(D − 1) ∼
= RD−1 (4.47)
and its points can be labelled by the spatial components of momentum (since
the zeroth component is then determined by eq. (4.44)). Massive orbits with
different masses are disjoint.
Chapter 4. Semi-direct products 75
• Let p be a time-like momentum with negative energy, p0 < 0. Its orbit is massive
with negative energy and consists of momenta q satisfying (4.44) with q0 < 0.
A typical orbit representative is (4.45) with M < 0, and orbits with different
masses are disjoint. The little group is a again SO(D − 1).
• Let p be a null momentum with positive energy, p0 > 0. Its orbit Op is massless
with positive energy. It consists of momenta q satisfying (4.44) with M 2 = 0. A
typical orbit representative is
where the energy E is positive; different values of E yield the same orbit. Note
that there is no rest frame for massless particles. The little group of (4.48) is
isomorphic to the Euclidean group
Gp ∼
= SO(D − 2) n RD−2 = ISO(D − 2). (4.49)
Op ∼
= SO(D − 1, 1)↑ /ISO(D − 2) ∼
= R × S D−2 (4.50)
and its points can be labelled by the spatial components of momentum (since
the zeroth component is then determined by q 2 = 0).
The little group is the lower-dimensional Lorentz group SO(D − 2, 1)↑ consisting
of transformations that leave the spatial coordinate xD−1 fixed. In particular the
orbit is diffeomorphic to the quotient
O∼
= SO(D − 1, 1)↑ /SO(D − 2, 1)↑ ∼
= R × S D−2 . (4.52)
Tachyonic orbits with different negative values of M 2 are disjoint. Note that
rotations always allow us to map p on −p, which is why any tachyonic orbit
representative can be written as (4.51).
This enumeration exhausts all Poincaré momentum orbits. Among the six families of
orbits, three contain only one orbit: the trivial orbit and the two massless orbits. The
remaining three families all contain infinitely many orbits labelled by a non-vanishing
mass squared, corresponding to massive particles and tachyons. These orbits and their
representatives are schematically depicted in fig. 4.3.
Chapter 4. Semi-direct products 76
(a) (b)
Figure 4.3: On the left, fig. (a) represents a few momentum orbits of the Poincaré
group in three dimensions, embedded in R3 with the vertical axis corresponding to p0
and the two horizontal axes (not represented in the figure) corresponding to spatial
components of momentum. Orbits can be massive, massless or tachyonic depending
on whether M 2 is positive, vanishing or negative, respectively. The cross in the middle
is the trivial orbit of p = 0, consisting of a single point. On the right, fig. (b) is a
schematic representation of momentum orbits: each point of the diagram corresponds
to an orbit representative, where massive orbits are represented by a vertical line,
tachyonic ones by a horizontal line, and discrete orbits (the two massless ones and the
trivial one) by dots. This schematic representation will be useful in parts II and III
for the interpretation of BMS3 supermomentum orbits.
To complete the description of orbits we now display standard boosts for massive
particles (the other cases are less important for our purposes so we skip them). We
take as orbit representative the momentum
p (4.45) of a particle at rest, and look for a
family of boosts gq such that gq · p = ( M 2 + q2 , q) that depend continuously on q.
One readily verifies that the matrices [96]
p !
1 + q2 /M 2 qj /M
gq = qq
p (4.53)
qi /M δij + qi 2j 1 + q2 /M 2 − 1
Remark. The little groups displayed in (4.46) and (4.49) hold for the connected
Poincaré group (4.40). If we replace the latter by its universal cover (4.41), then the
little groups are replaced by their double covers (assuming that D ≥ 4). In particular
Chapter 4. Semi-direct products 77
the little group of massive particles becomes Spin(D−1) while that of massless particles
becomes Spin(D − 2) n RD−2 , with the convention that Spin(2) is the double cover of
SO(2). Note that Spin(3) = SU(2).
4.2.3 Particles
According to the exhaustivity theorem of section 4.1.5, the momentum orbits in fig.
4.3 roughly classify relativistic particles. The states of each particle are wavefunc-
tions on its momentum orbit, valued in a spin representation of the little group and
transforming under Poincaré transformations according to formula (4.30). Provided
we know all irreducible unitary representations of all little groups, we have effectively
classifed all irreducible unitary representations of the Poincaré group.
Vacuum
Vacuum representations of Poincaré are those whose orbit O0 = {0} is trivial and is
left invariant by the whole Lorentz group. In that case a spin representation is a (pro-
jective) irreducible unitary representation R of SO(D − 1, 1)↑ . The latter is simple but
non-compact, so its only finite-dimensional irreducible unitary transformation is the
trivial one; the corresponding induced representation of Poincaré is trivial as well. All
other irreducible unitary representations of the Lorentz group are infinite-dimensional;
the corresponding induced representations of Poincaré are such that translations act
trivially, while Lorentz transformations act non-trivially on an infinite-dimensional
Hilbert space E of spin-like degrees of freedom. These representations can be inter-
preted as “vacua with spin” but are generally discarded as unphysical.
Massive particles
The momenta of a massive particle with mass M span an orbit (4.44) with little group
(4.46). The spin representation R then is a finite-dimensional, irreducible, generally
projective unitary representation of SO(D − 1) specified by some highest weight λ.
For example, when D = 4, R is a highest-weight representation of SO(3) = SU(2)/Z2
with spin s ≥ 0; the latter is either an integer or a half-integer. The carrier space
of R has dimension 2s + 1 and is generated by states | − si, | − s + 1i, ..., |s − 1i, |si
with definite spin projection along a prescribed axis. In that case the highest weight
λ coincides with s. The higher-dimensional case is analogous except that the number
of coefficients specifying λ is the rank b(D − 1)/2c of SO(D − 1). We will illustrate
this point in section 11.1 when dealing with partition functions of higher-spin fields in
Minkowski space.
Massless particles
The description of massless particles is analogous to that of massive ones, up to the
key difference that the massless little group is the Euclidean group (4.49). It is a
semi-direct product (4.1) with an Abelian normal subgroup, so the exhaustivity theo-
rem ensures that its irreducible unitary representations are induced and classified by
momentum-like orbits of their own. From the Poincaré viewpoint each induced repre-
sentation of (4.49) is a spin representation for a massless particle.
In that context one makes the distinction between two types of massless particles:
particles with discrete spin are those given by spin representations of (4.49) with van-
ishing Euclidean momentum. These are Euclidean analogues of the “spinning vacua”
described earlier, except that they are finite-dimensional. They amount to making
the action of RD−2 in (4.49) trivial, and coincide with (projective) irreducible uni-
tary representations of SO(D − 2). Thus massless particles with discrete spin have
a finite-dimensional space of spin degrees of freedom. By contrast, massless particles
with continuous or infinite spin are those whose spin representations of (4.49) have
non-trivial Euclidean momentum. The space of spin degrees of freedom is infinite-
dimensional in that case, since it consists of wavefunctions on a Euclidean momentum
orbit SO(D − 2)/SO(D − 3) ∼ = S D−3 . Particles with continuous spin are generally
discarded on the grounds that they are unphysical, although they have recently been
described in a field-theoretic framework [97–99].
Tachyons
Tachyons are particles moving faster than light. Their little group is SO(D − 2, 1)↑ . It
is simple and non-compact, so tachyons either have no spin at all, or have continuous
spin. They are generally considered as unphysical.
Odd dimensions
For odd D we erase the last row and column of (4.54). Then the Frobenius formula
(4.33) localizes to the unique rotation-invariant point of the momentum orbit, namely
the momentum at rest p = (M, 0, 0, ..., 0). This allows us to simplify (4.33) by setting
k = p in the exponential and the little group character, and pulling them out of the
integral. Denoting by λ the spin of the particle (it is a highest weight for SO(D − 1)),
we find Z
iM α0 (D−1)
χ[(f, α)] = e χλ [f ] dµ(k) δ(k, f · k) (4.55)
Op
where the replacement of k by p has projected the translation α on its time component
(D−1)
α0 . The little group character χλ [f ] is some function of the angles θ1 , ..., θr that we
do not need to write down at this stage (in practice it follows from the Weyl character
formula and is displayed in eq. (11.151) below). To obtain (4.55) it only remains to
evaluate the integral of the delta function. As coordinates on the orbit we choose the
spatial components of momentum, in terms of which the Lorentz-invariant measure on
Op is (3.4) and the corresponding delta function is (3.42). We thus get
dD−1 k p 2
dµ(k) δ(k, q) = p M + k2 δ (D−1) (k − q) = dD−1 k δ (D−1) (k − q) , (4.56)
2 2
M +k
where the multiplicative factors of the measure and its delta function cancel out. Note
that the same cancellation would have taken place for any measure µ proportional to
dD−1 k, in accordance with the fact that induced representations are insensitive to the
choice of measure. Applied to (4.55), the cancellation (4.56) allows us to write
Z
iM α0 (D−1)
χ[(f, α)] = e χλ [f ] dD−1 k δ (D−1) (k, f · k) (4.57)
RD−1
where f · k denotes the action of the spatial submatrix of (4.54) on k. The integral
can be written as
Z
1
dD−1 k δ (D−1) (I − f ) · k =
(4.58)
RD det(I − f )
Chapter 4. Semi-direct products 80
Note that for a Euclidean time translation α0 = iβ, this quantity may be seen as
the partition function of a relativistic particle in a rotating frame (albeit with purely
imaginary angular velocity).
is the Lefschetz number of the operator T [(f, α)]. If f was a number e−βω , (4.61)
would coincide with the partition function of a harmonic oscillator with frequency ω
at temperature 1/β.
Even dimensions
For even D the rotation f is exactly given by (4.54). Then the situation is more
complicated because the integral (4.33) localizes to a line rather than a point, as in
fig. 3.2. To make things simple we take α = (α0 , 0, ..., 0) to be a pure time translation.
Formula (4.55) is then replaced by
Z √
(D−1) D−1 iα0 M 2 +k2 (D−1)
χ[(f, α)] = χλ [f ] d ke δ (k − f · k) (4.62)
RD−1
where we have already implemented the simplification (4.56). The SO(D−1) character
(D−1)
χλ has been pulled out of the integral because, for D even, boosts along the
direction kD−1 commute with rotations (4.54). It remains once more to integrate the
delta function in (4.62). As far as the first D − 2 components of k are concerned, the
computation is the same as in the odd-dimensional case and results in a factor (4.58)
given by (4.59). But the last component of k is untouched by (4.54), so (4.62) becomes
r Z +∞ √
(D−1)
Y 1 0 M 2 +k2
χ[(f, α)] = χλ [f ] dk eiα δ (1) (k − k) , (4.63)
j=1
|1 − eiθj |2 −∞
where the length scale L is an infrared regulator. The integral in (4.63) then gives
Z +∞ √
0 2 2
dk eiα M +k = 2M K1 (−iM α0 )
−∞
where K1 is the first modified Bessel function of the second kind. In conclusion we get
r
ML 0 (D−1)
Y 1
χ[(f, α)] = K1 (−iM α ) χλ [f ] , (4.65)
π j=1
|1 − eiθj |2
whose Wick-rotated version can now be seen as the rotating partition function of a
particle trapped in a box of height L.
Time translations
All characters written above diverge when one of the angles θj goes to zero. These
divergences are infrared since they are due to delta functions evaluated at zero in
momentum space, and can be regularized as in (4.64). A case of particular interest is
the character of a pure time translation, whose Wick rotation is a canonical partition
function (3.50). Using once more the Frobenius formula (4.33) and the cancellation
(4.56), and letting α = (α0 , 0, ..., 0) be a pure time translation, we find
Z √
D−1 iα0 M 2 +k2 (D−1)
χ[(e, α)] = N d ke δ (0) (4.66)
RD−1
where N ≡ dim(E) is the number of spin degrees of freedom of the particle. The
infrared-divergent delta function can be seen as the spatial volume of the system,
Z
(D−1) 1 V
δ (0) = D−1
dD−1 x = .
(2π) RD−1 (2π)D−1
Using spherical coordinates we can then rewrite (4.66) as
Z +∞
NV 2π (D−1)/2 D−2
√
iα0 M 2 +k2
χ[(e, α)] = k dk e
(2π)D−1 Γ((D − 1)/2) 0
(D−2)/2
2N V 2πM
= M KD/2 (−iM α0 ) , (4.67)
(2π)D−1 −iα0
where KD/2 denotes once more a modified Bessel function of the second kind. This
is the character of a pure time translation in a massive Poincaré representation. For
α0 = iβ purely imaginary, it becomes the canonical partition function of a massive
relativistic particle,
(D−2)/2
−βH
2N V 2πM
Tr e massive particle
= M KD/2 (βM ). (4.68)
(2π)D−1 β
Even dimensions
For even D the Lorentz transformation (4.54) belongs to the little group of a massless
particle since it leaves invariant the momentum vector (E, 0, ..., 0, E). The character
computation then is the same as in the even-dimensional massive case; formula (4.63)
still holds with M = 0 and χ(D−1) replaced by the character χ(D−2) of a representation
of SO(D − 2) instead of SO(D − 1). Note that for even D these two groups have
the same rank r = b(D − 1)/2c, so there is no restriction on the values of the angles
θ1 , ..., θr (this will change for odd D). Using the regulator (4.64) and the fact that
Z +∞
0 2
dk ei|k|(α +iε) = − 0 ,
−∞ iα
one finds the character
r
iL (D−2) Y 1
χ[(f, α)] = 0
χ λ [f ] iθj |2
. (4.69)
πα j=1
|1 − e
Odd dimensions
For odd D the transformation (4.54) (with the last row and column suppressed) is
no longer an element of the little group of (E, 0, ..., 0, E) so its character vanishes if
θr 6= 0. This is consistent with the fact that SO(D − 2) has lower rank than SO(D − 1)
when D is odd. Accordingly we now take θr = 0 in (4.54), being understood that
the last row and column are suppressed. From there on the character computation is
identical to the cases treated above, except that the infrared divergence of the integral
becomes worse and requires two regulators L, L0 :
Z
0
√ LL0
2 2
dkdq eiα k +q δ (1) (k − k)δ (1) (q − q) = − . (4.70)
R2 2π(α0 )2
Massless characters in odd dimension D thus read
r−1
LL0
Y
(D−2) 1
χ[(f, α)] = χλ [f ] − (4.71)
2π(α0 )2 j=1 |1 − eiθj |2
(D−2)
where it is understood that θr = 0 in (4.54) and χλ is a character of SO(D − 2).
Note that this expression is not the massless limit of (4.60) because in general θr 6= 0 in
the latter formula. However, upon setting θr = 0 in (4.57) and regulating the resulting
double infrared divergence as in (4.70), the limit M → 0 does produce an expression of
the form (4.71), albeit with a reducible representation of SO(D − 2). We shall return
to this in section 11.1. Characters of time translations can be treated as in the massive
case and coincide, up to spin multiplicity, with the massless limit of (4.67).
Chapter 4. Semi-direct products 83
Wigner rotations
Consider a particle with mass M and spin representation R. We wish to understand
the action of the Wigner rotation (4.31) for an arbitrary momentum q belonging to its
orbit, and for a boost
cosh γ − sinh γ 0 · · · 0
− sinh γ cosh γ 0 · · · 0
f =
0 0 1 · · · 0
(4.72)
.. .. .. . . ..
. . . . .
0 0 0 ··· 1
with rapidity γ in the direction x1 . Since the gq ’s are standard boosts (4.53), the
combination gq−1 f gf −1 ·q is a sequence of three pure boosts in a plane, so we may
safely take D = 3 without affecting the outcome of the computation. The momentum
q then reads p
M 2 + Q2
q = Q cos ϕ (4.73)
Q sin ϕ
for some angle ϕ and some positive number Q. After a mildly cumbersome but straight-
forward computation, one finds a Wigner rotation matrix
1 0 0
gq−1 f gf −1 ·q = 0 cos θ − sin θ (4.74)
0 sin θ cos θ
whose entries are given by
M sin(ϕ) sinh(γ)
sin θ = − q p ,
Q2 sin2 (ϕ) + (Q cos(ϕ) cosh(γ) + M 2 + Q2 sinh(γ))2
p (4.75)
Q cosh(γ) + M 2 + Q2 cos(ϕ) sinh(γ)
cos θ = q p .
Q2 sin2 (ϕ) + (Q cos(ϕ) cosh(γ) + M 2 + Q2 sinh(γ))2
This is a pure rotation, as it should. It represents the fact that a boost acting on a
particle with non-zero momentum is seen, from the rest frame of the particle, as a
boost combined with a rotation (4.74) rather than a pure boost. The rotation only
affects spin degrees of freedom; scalar particles are insensitive to it.
The Wigner rotation (4.75) is responsible for the phenomenon of Thomas precession
[106] (see also [96], section 11.8). The latter is visible in atomic physics, where the spin
of an electron orbiting around a nucleus undergoes a slow precession due to the fact
that the electron’s acceleration is a sequence of boosts directed towards the nucleus.
Chapter 4. Semi-direct products 84
Momentum/spin entanglement
In (3.8) we saw that a space of E-valued wavefunctions on Op is a tensor product of
E with the scalar space L2 (Op , µ, C). For a relativistic particle, the former consists
of spin degrees of freedom while the latter accounts for momenta (or positions after
Fourier transformation). For example, any state of a massive particle with spin 1/2
takes the form (3.9) and describes the separate propagation of the two spin states |+i
and |−i. (For simplicity we use the Dirac notation until the end of this section.)
Hilbert space factorizations such as (3.8) are seldom preserved by unitary maps.
Indeed, if H = A ⊗ B and |Ψi ∈ H is a state with unit norm, the reduced density
matrix associated with |Ψi and acting in B is ρ ≡ TrA |ΨihΨ|. When U is a unitary
operator in H , it is generally not true that the reduced density matrix of U · |Ψi is
unitarily equivalent to ρ. In particular, U does not preserve the degree of entanglement
between A and B. Accordingly one may ask [102] whether Poincaré representations
spoil the splitting (3.8). To answer this, consider for definiteness a massive spin 1/2
particle in four dimensions. We start from a normalized E-valued wavefunction
Ψ(q) = ψ(q)|+i, i.e. |Ψi = |ψi ⊗ |+i (4.76)
where ψ is some complex-valued wavefunction while |+i is one of the two members of
an orthonormal basis |+i, |−i of E. This state represents a particle with spin up (say
along the x3 axis) propagating with a momentum probability distribution dµ(q)|ψ(q)|2 .
(For definiteness we take the measure µ to be the Lorentz-invariant expression (3.4).)
The corresponding reduced density matrix obtained by tracing over spin degrees of
freedom acts on the scalar Hilbert space L2 (Op , µ, C) and reads
ρ = h+| |ψi|+ihψ|h+| |+i + h−| |ψi|+ihψ|h+| |−i = |ψihψ| ,
which is a pure state. Now let us act on (4.76) with a Lorentz transformation f .
According to (4.30), and writing U ≡ T [(f, 0)], the resulting wavefunction is
(4.76)
(U · Ψ) (q) = Wq [f ] · Ψ(f −1 · q) = ψ(f −1 · q) Wq [f ]|+i (4.77)
where Wq [f ] is the Wigner rotation (4.31). Denoting φ(q) ≡ ψ(f −1 · q) we now find
that the entries of the reduced density matrix of (4.77) are
∗ ∗
ρ̃(q, q 0 ) = φ(q)χ+ (q) φ(q 0 )χ+ (q 0 ) + φ(q)χ− (q) φ(q 0 )χ− (q 0 )
(4.78)
where we have defined χ± (q) ≡ h±|Wq [f ]|+i. In general expression (4.78) is not equal
to a product ψ̃(q)ψ̃ ∗ (q 0 ) (for some complex wavefunction ψ̃), so the state (4.78) is not
pure! In particular the boosted state (4.77) is generally entangled with respect to the
splitting (3.8), even though the original state (4.76) was not. The reason for this is
that the Wigner rotation (4.74) generally has non-vanishing +− entries. Note that
the functions χ± satisfy |χ+ |2 + |χ− |2 = 1 by virtue of the fact that Wigner rotations
are unitary, so formula (4.78) indeed defines a density matrix.
such that Wq [f ] does not depend on q. In both situations the splitting (3.8) is robust
against symmetry transformations. An example of momentum-independent Wigner
rotations will be provided by the Bargmann group below. Thus the entanglement of
spin and momentum due to Wigner rotations is a purely relativistic effect.
Lemma. The group SL(2, R) is connected, but not simply connected. It is homotopic
to a circle and its fundamental group is isomorphic to the group of integers Z:
π1 (SL(2, R)) ∼
= Z. (4.80)
Proof. Since the determinant of (4.79) is non-zero, the vectors (a, b) and (c, d) in R2
are linearly independent. We can thus find linear combinations of these vectors that
span an orthonormal basis of R2 . In other words there exists a real matrix
ᾱ 0
K̄ =
β̄ γ̄
such that, for any SL(2, R) matrix S of the form (4.79), the product
ᾱa ᾱb
O ≡ K̄S =
β̄a + γ̄c β̄b + β̄d
belongs
√ to the orthogonal group O(2). We can make ᾱ positive by setting ᾱ−1 =
a2 + b2 and we can set γ̄ = 1/ᾱ so that O ∈ SO(2). Any matrix S ∈ SL(2, R) can
therefore be decomposed uniquely as
−1 x 0
S = K̄ O ≡ KO, with O ∈ SO(2) and K = (4.81)
y 1/x
Chapter 4. Semi-direct products 86
for some y ∈ R and x ∈ R strictly positive.4 This shows that SL(2, R) is connected
and homotopic to its maximal compact subgroup consisting of rotations
cos θ − sin θ
, θ ∈ R. (4.82)
sin θ cos θ
SO(2, 1)↑ ∼
= SL(2, R)/Z2 ≡ PSL(2, R) (4.83)
where the Z2 subgroup of SL(2, R) consists of the identity matrix and its opposite. In
particular, the fundamental group of the connected Lorentz group in three dimensions
is isomorphic to Z.
Proof. Our goal is to build a homomorphism
Im(φ) ∼
= SL(2, R)/Ker(φ). (4.85)
Let A be the Lie algebra sl(2, R). Each matrix α ∈ sl(2, R) can be written as a linear
combination
α = αµ tµ (4.86)
where the αµ ’s are real coefficients and the matrices
1 0 1 1 0 1 1 1 0
t0 ≡ , t1 ≡ , t2 ≡ (4.87)
2 −1 0 2 1 0 2 0 −1
A → A : α 7→ f αf −1 . (4.89)
4
This is a rewriting of the Iwasawa decomposition.
Chapter 4. Semi-direct products 87
This action preserves the determinant since det(f ) = 1, so according to (4.88) it may
be seen (for each f ) as a Lorentz transformation. This motivates the definition of a
homomorphism (4.84) given by
f tµ f −1 = tν φ[f ]ν µ ∀ µ = 0, 1, 2. (4.90)
Remark. For future reference note that the homomorphism (4.90) explicitly reads
1 2
(a + b2 + c2 + d2 ) 21 (a2 − b2 + c2 − d2 ) −ab − cd
a b 2
φ = 21 (a2 + b2 − c2 − d2 ) 12 (a2 − b2 − c2 + d2 ) −ab + cd , (4.91)
c d
−ac − bd bd − ac ad + bc
where the argument of φ is an SL(2, R) matrix. This will be useful in section 9.1.
SL(2, R) n R3 (4.92)
where the action of SL(2, R) on R3 is given by (4.89). The latter is in fact the adjoint
representation so we can also rewrite (4.92) as
where sl(2, R)Ab is the Lie algebra of SL(2, R), seen as an Abelian vector group. This
observation will turn out to be crucial in part III of this thesis. We stress that (4.93)
is not the universal cover of the Poincaré group in three dimensions, since SL(2, R) is
homotopic to a circle. This implies that (4.93) admits topological projective represen-
tations (which are equivalent to exact representations of its universal cover). There
are no algebraic central extensions.
Let us prove that these are the correct little groups. We shall use the fact that the
action of Lorentz transformations on momenta is equivalent to its action on transla-
tions, which in turn is equivalent to the adjoint representation of SL(2, R) according to
the definition (4.93). From that point of view a momentum (p0 , p1 , p2 ) is represented
by a matrix η µν pµ tν where ηµν is the Minkowski metric in D = 3 dimensions and the
tµ ’s are given by (4.87). Explicitly the matrix is
1 p2 −p0 + p1
p= . (4.94)
2 p0 + p1 −p2
In that language the little group of p is the set of SL(2, R) matrices that commute with
(4.94). It immediately follows that the little group of p = 0 is SL(2, R). For massive
orbits we move to a rest frame where p0 = M and p1 = p2 = 0; the only matrices
leaving p fixed then are rotations (4.82). For tachyons we take p0 = p1 = 0, p2 6= 0
and find that the little group consists of matrices of the form
x
e 0
± , x ∈ R, (4.95)
0 e−x
Note that the little groups listed in table 4.1 are sensitive to the cover chosen in
(4.93). Had we chosen the standard connected Poincaré group SO(2, 1)↑ n R3 , the
little groups for massless particles and tachyons would be quotiented by Z2 and would
reduce to R. For the universal cover of the Poincaré group, the little groups would
instead get decompactified, so e.g. U(1) would be replaced by R. This has important
implications for the spin of relativistic particles in three dimensions.
Chapter 4. Semi-direct products 89
Massive particles
The properties of massive particles in three dimensions are the same as in section 4.2.3.
In particular their momentum orbits take the form
Op ∼
= SO(2, 1)↑ /U(1) ∼
= SL(2, R)/S 1 . (4.97)
The only subtlety is that the group of spatial rotations now is U(1) ∼ = SO(2), so the
spin of a massive particle is a one-dimensional irreducible unitary representation of the
form (2.13) labelled by some number s. If the double cover (4.93) was the universal
cover of the Poincaré group, that number would be restricted to integer or half-integer
values. However the fact that (4.93) is homotopic to a circle implies that s may take
any real value. Thus massive particles in three dimensions can be anyons [110, 111].
The same phenomenon will occur with massive BMS3 particles.
Remark. Wigner rotations do occur in three dimensions, but they do not lead
to momentum/spin entanglement when the space of spin degrees of freedom is one-
dimensional.
Massless particles
The spin properties of massless particles in three dimensions are also somewhat pecu-
liar compared to those of their higher-dimensional cousins. Their little group R × Z2
can be seen as a Euclidean group in one dimension, where Z2 plays the role of rotations
while R is spanned by Euclidean translations. If the latter is represented non-trivially,
one obtains an analogue of “continuous spin” particles in three dimensions, although
in the present case the space of spin degrees of freedom is actually finite-dimensional.
By contrast, when R is represented trivially, the spin representation boils down to an
irreducible unitary representation of Z2 . The latter has exactly two irreducible unitary
representations (the trivial one and the fundamental one), so we conclude that “dis-
crete spin” massless particles in three dimensions can only be distinguished by their
statistics (bosonic or fermionic); they have no genuine spin. This is consistent with
the fact that massless field theories in three dimensions either have no local degrees
of freedom at all (such as in gravity or Chern-Simons theory), or have only scalar or
Weyl fermion degrees of freedom.
4.3.3 Characters
For future reference, we now list characters of irreducible unitary representations of
the Poincaré group in three dimensions. The results of sections 4.2.4 and 4.2.5 apply,
so the character of a rotation by θ combined with an arbitrary translation α in a
Poincaré representation with mass M and spin s is given by formula (4.60),
0 +isθ 1 0 1
χ[(rotθ , α)] = eiM α iθ 2
= eiM α +isθ 2 , (4.98)
|1 − e | 4 sin (θ/2)
where we have replaced the little group character by χλ [f ] = eisθ . In part III we shall
encounter the BMS3 generalization of this expression. Similarly the character (4.68)
Chapter 4. Semi-direct products 90
Characters of massless particles with discrete spin are given by formula (4.71) with
D = 3, r = 1 and χλ = ±1.
where the dots on the right-hand side denote either matrix multiplication, or the ac-
tion of a matrix on a column vector. The largest connected subgroup of (4.100) is
obtained upon replacing O(D − 1) by SO(D − 1); its universal cover is obtained by
replacing SO(D − 1) by its universal cover, Spin(D − 1).
The intricate structure (4.100) translates the fact that space and time live on
different footings in Galilean relativity. Thus the analogue of a Lorentz transformation
now is a pair (f, v), while space-time translations are pairs (α, t). Boosts and rotations
span a group SO(D − 1) n RD−1 while space-time translations span an Abelian group
RD . In particular each boost is a velocity vector v acted upon by rotations according
to the matrix representation of O(D − 1). Since time is absolute in Galilean relativity,
the last entry on the right-hand side of (4.101) is a sum s + t without influence of
Chapter 4. Semi-direct products 91
(e, v, α, 0) (4.102)
The Lie algebra of the Galilei group is generated by (D − 1)(D − 2)/2 rotation
generators, (D − 1) boost generators, (D − 1) spatial translation generators, and one
generator of time translations. We will not display their Lie brackets here.
Bargmann groups
The Galilei group turns out to admit a non-trivial algebraic central extension:
whose elements are 5-tuples (f, v, α, t, λ) where (f, v, α, t) belongs to the Galilei group
(4.100) while λ is a real number. The group operation is
1 2
(f, v, α, s, λ)·(g, w, β, t, µ) = (f, v, α, s)·(g, w, β, t), λ+µ+v·f ·β + v t (4.104)
2
where the first entry on the right-hand side is given by (4.101) while v · β ≡ v i β i is
the Euclidean scalar product of v and β; in particular, v2 ≡ v i v i .
This central extension says that the Abelian subgroup of boosts and translations
(4.102) gets extended into a Heisenberg group (2.37):
(e, v, α, 0, λ) · (e, w, β, 0, µ) = e, v + w, α + β, 0, λ + µ + v · β .
In other words, in quantum mechanics, spatial translations and boosts do not com-
mute. Note that, even in the Bargmann group, the normal subgroup of (centrally
extended) space-time translations
(e, 0, α, t, λ) (4.105)
remains Abelian. Hence the exhaustivity theorem of section 4.1.5 applies to the
Bargmann group: all Galilean particles are induced representations.
For D ≥ 4 space-time dimensions, eq. (4.104) is the only algebraic central exten-
sion of the Galilei group. But for D = 3, the Galilei group admits three non-trivial
differentiable central extensions [113], one of which is the one displayed in (4.104).
We will not take these extra central extensions into account. As regards topological
central extensions, the Galilean situation is identical to that of the Poincaré group.
Thus Bargmann(3) has a fundamental group Z and admits infinitely many topological
projective representations, while for D ≥ 4 the fundamental group of Bargmann(D)
is Z2 , leading either to exact representations or to representations up to a sign.
Chapter 4. Semi-direct products 92
Remark. The Bargmann group is a limit of the Poincaré group as the speed of light
goes to infinity (see e.g. [53]), known more accurately as an Inönü-Wigner contraction
[115]. We will not describe this procedure here, although we will encounter a very
similar one in part III when showing that the BMS3 group is an ultrarelativistic limit
of two Virasoro groups.
Generalized momenta
The Abelian normal subgroup of (4.103) consists of centrally extended translations
(4.105). Its dual space consists of generalized momenta
(p, E, M ) (4.106)
paired with translations according to5
h(p, E, M ), (α, t, λ)i = hp, αi − Et − M λ (4.107)
where hp, αi ≡ pi αi , i = 1, ..., D − 1. Accordingly, p is dual to spatial translations
and represents the actual momentum of a particle; E is dual to time translations and
represents the particle’s energy; finally M is a central charge dual to the central entries
λ in (4.105). Working in units such that ~ = 1, eq. (4.104) says that λ has dimensions
[distance] × [velocity] so the pairing (4.107) implies that M is a mass scale:
[energy] × [time]
[M ] = = [mass]. (4.108)
[distance] × [velocity]
In fact we will see below that M is the mass of a non-relativistic particle. Note that
M lives on a different footing than p and E, which is why the relation E = M c2 is
invisible in Galilean relativity.
According to the structure (4.103), the group G acting on translations is the Eu-
clidean group spanned by rotations and boosts. Its action is given by
1 2
σ(f,v) (α, t, λ) = f · α + vt, t, λ + v · f · α + v t (4.109)
2
by virtue of (4.104). The pairing (4.107) then yields the action σ ∗ of boosts and
rotations on generalized momenta:
∗
σ(f,v) (p, E, M ), (α, t, λ) =
(4.16)
= (p, E, M ), σ(f −1 ,−f −1 ·v) (α, t, λ)
D 1 E
= (p, E, M ), f −1 · α − f −1 · vt, t, λ − v · α + v2 t
2
(4.107)
1
= p, f −1 · α − f −1 · vt − Et − M λ − v · α + v2 t , (4.110)
2
5
The minus signs are conventional, and included for later convenience.
Chapter 4. Semi-direct products 93
where we have used the fact that rotations preserve Euclidean scalar products. We can
then use the Euclidean analogue of the isomorphism (4.42) to identify (RD−1 )∗ with
RD−1 and rewrite the pairing hp, αi = pi αi as a scalar product p · α = pi αi = pi αi ,
where indices are raised and lowered thanks to the Euclidean metric. This allows us
to rewrite v · α as hv, αi in (4.110), and leads to
∗
1 2
σ(f,v) (p, E, M ) = f · p + M v, E + v · f · p + M v , M . (4.111)
2
One may recognize here the non-relativistic transformations laws of momentum and
energy under rotations and boosts. The mass M is left unchanged, as was to be
expected for a central charge.
Orbits
Let us classify orbits of generalized momenta under the transformations (4.111). Since
the mass M is invariant, it is a constant quantity specifying each orbit; orbits with dif-
ferent masses are disjoint. In particular, the orbits differ greatly depending on whether
M vanishes or not.
where the sphere S D−2 is spanned by all momenta f · p while R is spanned by the
values of energy. The little group is
and consists of rotations leaving p invariant together with boosts that are orthogonal
to p. Note that this is the same little group (4.49) as for relativistic massless particles.
∗
p2
σ(e,−p/M ) (p, E, M ) = 0, E + , M (4.114)
2M
so any massive particle admits a rest frame. If we call E0 ≡ E + p2 /2M , the orbit of
(4.114) under rotations and boosts is a parabola
M v2
D−1
O(0,E,M ) = M v, E0 + ,M v ∈ R ⊂ RD−1 × R. (4.115)
2
Chapter 4. Semi-direct products 94
As orbit representative we can take the generalized momentum in the rest frame,
(0, E0 , M ) (4.116)
in accordance with the fact that the orbit (4.115) is diffeomorphic to the quotient
space SO(D − 1) n R D−1
/SO(D − 1) = R∼ D−1
. Note again that this is exactly the
same little group (4.46) as for relativistic massive particles. Finally, pure boosts
provide a continuous family of standard boosts on the orbit (4.115) of (4.116). Note
that energy is bounded from below on the orbit if and only if M > 0.
4.4.3 Particles
According to the exhaustivity theorem of section 4.1.5, all irreducible unitary repre-
sentations of Bargmann groups are induced, and they are classified by momentum
orbits. Each such representation consists of wavefunctions on an orbit, representing
the quantum states of a non-relativistic particle.
For example, the spin of a massive Galilean particle is an irreducible unitary repre-
sentation of SO(D − 1). The space of states of the particle then consists of wavefunc-
tions on the orbit (4.115) taking values in the space of the spin representation. Scalar
products of wavefunctions are defined as usual by (3.7), where µ is some measure on
the orbit. For convenience one can pick the standard Lebesgue measure dD−1 q, which
is left invariant by both rotations and boosts since (4.111) says that they act on (4.115)
as Euclidean transformations q 7→ f · q + M v.
Using the group operation (4.101) we read off the Wigner rotation
where we have also used the fact that the measure dD−1 q is invariant to cancel its
Radon-Nikodym derivative. This result differs from the Poincaré transformations of
Chapter 4. Semi-direct products 95
relativistic particles in two key respects. First, Galilean Wigner rotations (4.119) are
momentum-independent, so in contrast to (4.75) they do not entangle momentum and
spin. In fact, there is no Thomas precession for non-relativistic particles. The second
difference is the presence of the mass M : formula (4.120) is an exact representation of
the Bargmann group (4.104), but because M 6= 0 it is a projective representation of
the centreless Galilei group (4.100). This can be seen by noting that for a pure boost
v and a spatial translation α, eq. (4.120) gives
T [v] · T [α] = e−iM v·α T [α] · T [v], (4.121)
which says that boosts and spatial translations do not commute. In part III we will
encounter a similar phenomenon with the BMS3 group, whose dimensionful central
charge will coincide with the Planck mass.
4.4.4 Characters
We now evaluate characters of massive non-relativistic particles. Let M > 0 and
choose a spin λ, specifying an irreducible unitary representation of the little group
SO(D − 1). For definiteness we take the rest frame energy E0 = 0 in (4.115). In order
for the character (4.33) to be non-zero we must set v = 0. Eq. (4.119) then allows us
(D−1)
to pull the little group character χR = χλ out of the momentum integral:
Z
(D−1) 2
χ[(f, 0, α, t, λ)] = e−iM λ χλ [f ] dD−1 k δ (D−1) (k − f · k) eik·α−ik t/2M . (4.122)
RD−1
For simplicity we set λ = 0 from now on and neglect writing this entry. We take f to
be a rotation (4.54) with the first row and column suppressed and all angles θ1 , ..., θr
non-zero, r = b(D − 1)/2c. If D is odd we also erase the last row and column. We
treat separately even and odd dimensions.
If D is odd, then the only fixed point of f on (4.115) is the tip k = 0. The integral
of (4.122) localizes and (4.58) yields
r
Y 1
χ[(f, 0, α, t)] = χλ [f ] . (4.123)
j=1
|1 − eiθj |2
If D is even, then f leaves fixed the whole axis kD−1 as in fig. 3.2. Integrating first
over the rotated coordinates k1 , ..., kD−2 in (4.122) and writing kD−1 ≡ k, we find
r Z +∞
(D−1)
Y 1 D−1 2
χ[(f, 0, α, t)] = χλ [f ] iθ 2
dk δ(0) eikα −ik t/2M . (4.124)
j=1
|1 − e | −∞
j
Here the term δ(0) = δ(k − k) is an infrared divergence that we regularize as in (4.64)
with a length scale L. Denoting αD−1 ≡ x, we are left with the integral
Z +∞ 1/2
ikx−ik2 t/2M 2πM 2
dk e = eiM x /2t (4.125)
−∞ it
Chapter 4. Semi-direct products 96
The only dependence of this expression on α appears through the component αD−1 ≡
x, because we picked a rotation f leaving fixed the direction kD−1 . For a general ro-
tation, the component of α appearing in the character would be its projection on the
axis left fixed by f .
which we can interpret as the trace of the operator eiP x−iHt in the Hilbert space of a
free massive particle on the real line:
Z +∞
iP x−iHt
dy y + x|e−iHt |y .
Tr T [(e, x, t)] = Tr e = (4.127)
−∞
The second comment concerns the relation between Bargmann characters and
Poincaré characters. For even D, (4.126) is the non-relativistic analogue of (4.65)
but the functions appearing in the two results are different. By contrast, for odd
D, the Bargmann character (4.123) coincides with its Poincaré analogue (4.60). This
may be seen as a consequence of the phenomenon (4.58), whose effect is to localize
the computation of the character to the region of momentum space surrounding the
momentum at rest, that is, the non-relativistic region. By contrast, when the localiza-
tion is not complete as is the case for even D, the momenta in the integral (4.125) are
arbitrarily large and relativistic effects become important. This produces a difference
between Bargmann and Poincaré characters. It is particularly apparent for characters
of Euclidean time translations, which in the non-relativistic case are given by
Z (D−1)/2
NV D−1 −βk2 /2M M
χ[(e, 0, 0, −iβ)] = d ke = NV
(2π)D−1 RD−1 2πβ
where N is the dimension of the spin representation. This is the non-relativistic version
of (4.68). For D = 3 (and N = 1) it reduces to V M/(2πβ), which is the non-relativistic
limit of (4.99).
Chapter 5
In the previous chapters we have seen how representation theory leads to geometric
objects such as orbits. The purpose of this chapter is to describe the opposite phe-
nomenon: starting from a coadjoint orbit of a group G, we will obtain a representation
by quantizing the orbit. This construction will further explain why orbits of momenta
classify representations of semi-direct products. In addition it will turn out to be a
tool for understanding gravity in parts II and III.
The plan is as follows. We start in section 5.1 with basic reminders on symplectic
manifolds with symmetries, including their momentum maps. Along the way we intro-
duce the notion of coadjoint orbits, which will turn out to be crucial for the remainder
of this thesis. Section 5.2 is then devoted to the quantization of symplectic manifolds,
and describes in particular the relation between representation theory and symplectic
geometry. In section 5.3 we reformulate geometric quantization in terms of action
principles that describe the propagation of a point particle on a group manifold. The
two last sections of the chapter are concerned with applications of these considerations
to semi-direct products: in section 5.4 we describe the coadjoint orbits and world line
actions of such groups in general, while in section 5.5 we illustrate these results with
the Poincaré group and the Bargmann group.
Our language in this chapter will be slightly different than in the previous ones,
as we rely on differential-geometric tools that were unnecessary for our earlier con-
siderations. Useful references include [43, 55] for symplectic geometry, [116, 117] for
quantization, as well as the (sadly unpublished) Modave lecture notes [118].
Remark. The presentation adopted here is self-contained, but fairly dense. We urge
the reader who is not acquainted with differential geometry to only read sections 5.1.1
and 5.1.2, then go directly to part II of the thesis. In doing so one will miss the
symplectic aspects of our later considerations, but the other points of our presentation
should remain accessible.
97
Chapter 5. Coadjoint orbits and quantization 98
One can verify that any left-invariant vector field is given by ξg = (Lg )∗e X for some
tangent vector X ∈ Te G. Thus the space of left-invariant vector fields is isomorphic to
the tangent space of G at the identity. We shall denote by ζX the left-invariant vector
field on G given by (ζX )g = (Lg )∗e X.
Definition. The Lie algebra of G is the vector space g = Te G endowed with the Lie
bracket
[X, Y ] ≡ [ζX , ζY ]e (5.1)
where the bracket on the right-hand side is the usual Lie bracket of vector fields eval-
uated at the identity.
One can show that the bracket (5.1) is such that ζ[X,Y ] = [ζX , ζY ]. As a corollary,
any smooth homomorphism of Lie groups F : G → H is such that its differential
F∗e at the identity is a homomorphism of Lie algebras. When interpreting G as a
symmetry group, the elements of its Lie algebra are seen as “infinitesimal” symmetries,
i.e. transformations near the identity. In practice the Lie algebra structure of g is often
displayed in terms of a basis {ta |a = 1, ..., dim g} of g with Lie brackets
In that context the coefficients fab c ∈ R are known as the structure constants of g in
the basis {ta }.
1
Recall that the differential of a smooth map F : M → N at p ∈ M is the map F∗p : Tp M →
d
TF (p) N : γ̇(0) 7→ dt F(γ(t)) t=0 , where γ(t) is a path in M such that γ(0) = p.
Chapter 5. Coadjoint orbits and quantization 99
Exponential map
Definition. Let X ∈ g, and let γX be the integral curve2 of the corresponding left-
invariant vector field ζX such that γX (0) = e. Then the exponential map of G is
exp : g → G : X 7→ exp[X] ≡ γX (1). (5.3)
One can
P verifynthat, for matrix groups, this definition reduces to the standard Taylor
X
series n∈N X /n!. We often denote exp[X] ≡ e .
One can verify that this is indeed a representation of G. Using (5.4), one also
shows that it satisfies the identity
eAdf X = f eX f −1 (5.7)
where eX is the exponential map of G. Note that the adjoint representation of any
Abelian Lie group is trivial. Finally, the adjoint representation of the Lie algebra g is
defined as the differential of (5.5) at the identity:
d
adX (Y ) ≡ AdetX (Y ) t=0 = [X, Y ] . (5.8)
dt
In (4.16) we saw how to define dual representations. Let us apply this to the
adjoint representation (5.5): we write the dual space of g as g∗ , which consists of
linear forms p : g → R : X 7→ hp, Xi. When interpreting G as a symmetry group,
the elements of the dual of g can be seen as “momenta”, or more generally conserved
vectors, associated with the symmetries. In particular the number hp, Xi then is
the Noether charge associated with the symmetry generator X when the system has
“momentum” p.
2
An integral curve of a vector field ξ on a manifold M is a path γ(t) on M such that γ̇(t) = ξγ(t) .
Chapter 5. Coadjoint orbits and quantization 100
Definition. Let G be a Lie group with Lie algebra g. Then the coadjoint represen-
tation of G is the homomorphism
i.e. Ad∗f (p), X ≡ hp, Adf −1 (X)i for all p ∈ g∗ and any X ∈ g. From now on we refer
to elements of g and g∗ as adjoint and coadjoint vectors, respectively.
Wp ≡ Ad∗g (p) g ∈ G
the coadjoint orbit of p. It is a homogeneous space for the coadjoint action of G. Note
that the coadjoint representation of any Abelian group is trivial, so its coadjoint orbits
are single points. By contrast, coadjoint orbits of non-Abelian groups are generally
non-trivial (except if p = 0). We will see in section 5.4.3 that the coadjoint orbits of
semi-direct products contain their momentum orbits.
The dual of the infinitesimal adjoint representation (5.8) is the differential of (5.9)
at the identity, i.e. the coadjoint representation of the Lie algebra g:
d (5.10)
ad∗X (p) ≡ (Ad∗etX (p))|t=0 = −p ◦ adX = −p ◦ [X, ·] . (5.11)
dt
Remark. The adjoint and coadjoint representations of a group G are generally in-
equivalent. In fact they are equivalent if and only if g admits a non-degenerate bilinear
form (which is the case e.g. for semi-simple Lie groups).
This map is called the Poisson bracket on M, and the pair M, {·, ·} is a Poisson
manifold or a phase space.
The Poisson bracket endows the space of functions C ∞ (M) with a structure of Lie
algebra; the Leibniz identity implies in addition that the map
tonian vector field associated with H the (unique) vector field ξH on M such that
Now consider the set of all Hamiltonian vector fields on M; at a point p ∈ M, they
span a subspace of the tangent space Tp M. By taking this span for all p ∈ M, one
obtains a subbundle of the tangent bundle T M (i.e. a distribution on M). Because
brackets of Hamiltonian vector fields are Hamiltonian, Frobenius’ theorem implies
that Hamiltonian vector fields yield a foliation of M into so-called symplectic leaves.
Two points belong to the same leaf if they can be joined by the integral curve of a
Hamiltonian vector field. In the example of R3 mentioned above, symplectic leaves are
planes z = const. This leads to the definition of symplectic manifolds.
4
A derivation of an algebra A is a linear map D : A → A : a 7→ D(a) that satisfies the Leibniz
rule D(a · b) = D(a) · b + a · D(b).
Chapter 5. Coadjoint orbits and quantization 102
µ≡ ω
| ∧ {z
... ∧ ω} . (5.14)
dim(M)/2 times
The symplectic leaves described above are prime examples of symplectic mani-
folds: they are endowed with a symplectic form ω such that ω(ξF , ξG ) ≡ {F, G}; this
condition determines ω unambiguously because symplectic leaves are, by definition,
spanned by the integral curves of Hamiltonian vector fields. Another common exam-
ple is the phase space M = R2n of a non-relativistic particle in Rn , with coordinates
(q 1 , ..., q n , p1 , ..., pn ) and symplectic form
for any vector V ∈ T(q,α) T ∗ Q. Then ω ≡ −dθ is the canonical symplectic form on
T ∗ Q. In the example (5.15), Q = Rn .
Thus the differential of the projection (5.16) projects V on its part tangent to Q. The
definition (5.17) then implies that θ = pi dq i , so
ω = −dθ = dq i ∧ dpi (5.18)
is definitely a closed, non-degenerate two-form. It coincides locally with (5.15).
Remark. The Darboux theorem states that any point of a symplectic manifold has
a neighbourhood with local coordinates (q i , pj ) such that the symplectic form reads
(5.18). Thus any symplectic manifold is locally equivalent to a cotangent bundle.
Hamiltonian vector fields can be used to define Poisson brackets in the same way
as on symplectic leaves of Poisson manifolds: for any two functions F, G we write
{F, G} ≡ ω(ξF , ξG ). (5.20)
In terms of this bracket the definition (5.19) is equivalent to our earlier definition of
Hamiltonian vector fields in (5.12). In local coordinates the definition (5.19) reads
ωij ξFi = ∂j F ⇔ ξFi = ∂j Fω ji (5.21)
where ωij are the components of ω and ω ij is the matrix inverse of ωij . Accordingly,
the bracket (5.20) can be written as {F, G} = −ω ij ∂i F∂j G.
Note that symplectic manifolds only contain kinematical data: they tell us the
available combinations of “positions” and “momenta” — those combinations are clas-
sical states. Classical observables then are real-valued functions on phase space. Once
we declare that a certain observable H is the Hamiltonian, time evolution is given
locally by the equations of motion ẋ = {x, H}.
Symplectomorphisms
Definition. Let (M, ω) and (N , Ω) be symplectic manifolds. A symplectomorphism
(or canonical transformation) from M to N is a diffeomorphism φ : M → N that
preserves the symplectic structure in the sense that6
φ∗ Ω = ω. (5.22)
6
Recall that the pullback of a tensor field T of rank k on a manifold N by a map φ : M → N is
defined by (φ∗ T )p (v1 , ..., vk ) ≡ Tφ(p) (φ∗p v1 , ..., φ∗p vk ) for any p ∈ M and all v1 , ..., vk ∈ Tp M.
Chapter 5. Coadjoint orbits and quantization 104
One can associate a Hamiltonian vector field (5.12) with any function F on a
Poisson manifold. In the present case one has the following result:
Corollary. The symplectic leaves of the Kirillov-Kostant bracket are the coadjoint
orbits of G. In particular, all (finite-dimensional) coadjoint orbits have even dimension.
Proof. By the above proposition, the Hamiltonian vector field ξF associated with a
function F and evaluated at a point p ∈ g∗ is
Now, given an adjoint vector X ∈ g, we can always find a real function F on g∗ such
that F∗p = X. Accordingly eq. (5.26) implies that the integral curves of all possible
Hamiltonian vector fields going through p span the coadjoint orbit of p.
For future reference it is useful to rewrite the Kirillov-Kostant bracket in terms
of a basis {ta } of g with Lie brackets (5.2). Any adjoint vector can then be written
as X = X a ta . If {(ta )∗ |a = 1, ..., n} denotes the corresponding dual basis of g∗ , so
that h(ta )∗ , tb i = δba , any coadjoint vector can be written as p = pa (ta )∗ with real
components pa . This defines global coordinates {pa |a = 1, ...n} on g∗ , where each pa
is a real function on g∗ that associates with a coadjoint vector p its component pa .
To apply (5.23) we note that the differential (pa )∗ of pa acts on the basis vectors ∂p∂ c
according to
∂ ∂p
a
(pa )∗ = = δac . (5.27)
∂pc ∂pc
But since g∗ is a vector space we can identify Tp g∗ with g∗ by declaring that ∂/∂pc
coincides with (tc )∗ , so in fact the differential satisfies (pa )∗ ((tc )∗ ) = δac . With this
identification the differential (pa )∗ belongs to the dual of the dual, (g∗ )∗ = g, and may
be seen as an adjoint vector. Property (5.27) says that this adjoint vector is precisely
ta . The Poisson bracket follows:
In parts II and III we will see that the Poisson brackets of three-dimensional gravity
coincide with Kirillov-Kostant brackets for suitable asymptotic symmetry groups.
where X, Y ∈ g.
Here ad∗X q and ad∗Y q are “infinitesimal displacements” of q and represent generic
tangent vectors of Wp at q. One can verify that (5.29) is closed and non-degenerate on
Wp , so it is indeed a symplectic form. In addition it is invariant under the coadjoint
action of G in the sense that (Ad∗f )∗ (ω) = ω for all f ∈ G. Thus each coadjoint orbit
of G is a homogeneous space equipped with a G-invariant symplectic structure. In
this sense it is a symmetric phase space. We will see below that, for instance, each
coadjoint orbit of the Poincaré group coincides with the space of classical states of a
relativistic particle with definite mass and (classical) spin.
where the Lie bracket on the left-hand side is that of vector fields, while the bracket
on the right is that of g, given by (5.1).
For example, the representations (5.8) and (5.11) are infinitesimal generators of the
adjoint and coadjoint representations of G, respectively.9 In this language the tangent
space at q of an orbit (3.13) consists of all vectors of the form (ξX )q , where X spans
the Lie algebra g. The flow of ξX is R × M → M : (t, q) 7→ etX · q. In what follows
we study group actions where the manifold M is symplectic.
9
Property (5.31) does not contradict the fact that the adjoint and coadjoint representations of
g are actual representations, i.e. for example that adX adY − adY adX = ad[X,Y ] . Indeed, the vector
fields in (5.31) are derivations acting on functions on M, while adX and ad∗X are generally understood
as linear operators acting on g and g∗ , respectively.
Chapter 5. Coadjoint orbits and quantization 107
Momentum maps
Let (M, ω) be a symplectic manifold. An action of G on M is symplectic if each map
q 7→ f · q is a symplectomorphism, in which case G is a symmetry group of M. If ξX
denotes the infinitesimal generator (5.30) of a symplectic action, then LξX ω = 0.
J : M → g∗ : p 7→ J (p) (5.32)
The definition (5.33) can be compared to that of Hamiltonian vector fields, eq.
(5.19), and is equivalent to the statement
Here ξX is the infinitesimal generator (5.30), while ξJX is the Hamiltonian vector field
associated with the function JX and {·, ·} is the Poisson bracket (5.20). Thus the
function JX generates the transformation corresponding to X ∈ g in phase space, in
the sense that for any function F ∈ C ∞ (M) we have {JX , F} = −ξX (F) = −δX F.
From this observation we can derive an important corollary: if X, Y ∈ g and if we
consider the corresponding functions JX and JY , their Poisson bracket acts on classical
observables according to
{JX , JY }, F = JX , {JY , F} − JY , {JX , F}
(5.34) (5.31) (5.34)
= [ξX , ξY ](F) = −ξ[X,Y ] (F) = J[X,Y ] , F (5.35)
where in the first equality we have used the Jacobi identity. Since this property is
true for any function F, it is tempting to remove it from both ends of the equation
and conclude that the momentum map provides a representation of the Lie algebra
g. However, this hasty argument overlooks one crucial possibility, namely the fact
that brackets of momentum maps may include a central term that commutes with all
functions on phase space (see [43] or appendix 5 of [119]). Thus we conclude that:
for some real two-cocycle c on g. If the phase space has several connected components,
there may be several different cocycles (one for each connected component).
Chapter 5. Coadjoint orbits and quantization 108
Remark. Not all symplectic group actions have a momentum map, as there may be
no function JX such that (5.34) holds. If such a function exists for each X ∈ g, then
the action does admit a momentum map and is said to be Hamiltonian. Note that,
if J and J 0 are momentum maps for the same group action, then (5.33) implies that
they differ by a constant coadjoint vector (provided M is connected).
Noether’s theorem
The momentum map gives the conserved quantity J (p) ∈ g∗ associated with each
classical state p ∈ M. As anticipated earlier, coadjoint vectors may thus be seen as
“conserved vectors” for symmetric phase spaces. This interpretation stems from the
following fundamental result:
for any time t belonging to the domain of the curve. In other words the dim g com-
ponents of the coadjoint vector J (γ(t)) are conserved when the equations of motion
γ̇ = (ξH )γ are satisfied.
Proof. The Hamiltonian H is invariant under G, so dtd H(etX · p) = 0 for any p ∈ M.
Since any integral curve of ξX takes the form etX · p for some initial condition p, this
is to say that ξX (H) = ξJX (H) = 0, so H is constant along integral curves of JX ;
equivalently, JX is constant along integral curves of ξH .
In a translation-invariant system the momentum map associates a momentum vec-
tor with any point in phase space. Similarly, in a rotation-invariant system it coincides
with angular momentum. Finally, in a two-dimensional conformal field theory, it co-
incides with the stress tensor of a given field configuration. We will illustrate these
statements below. In the remainder of this section we build momentum maps for
specific families of symplectic manifolds.
path γ(t) in Wp can be written as γ(t) = Ad∗f (t) p for some path f (t) in G. If γ(0) = q
and γ̇(0) = ad∗Y q for some Y ∈ g, then we find
(5.29)
ωq ad∗X q, γ̇(0) = ωq ad∗X q, ad∗Y q = hq, [X, Y ]i = had∗Y q, Xi
for any X ∈ g. Since ad∗X q is the infinitesimal generator ξX of the coadjoint action
of
G on Wp , the far left-hand side of this equation coincides with (iξX ω)q γ̇(0) . As a
consequence the momentum map (5.33) should be such that
d
had∗Y q, Xi = = hJ∗q ad∗Y q, Xi
hJ (γ(t)), Xi t=0
dt
for all X ∈ g. This implies that the differential J∗q : Tq Wp → g∗ is just the inclusion,
and determines J up to a constant coadjoint vector. In particular:
→ g∗ : q 7→ q,
J : Wp ,− (5.38)
is a momentum map for the coadjoint action of G on (Wp , ω) when ω is the Kirillov-
Kostant symplectic form (5.29).
This result implies that the action of a Lie group on its coadjoint orbits is always
Hamiltonian. In fact one can show that any symplectic manifold endowed with a tran-
sitive Hamiltonian action of some group G is a covering space of a coadjoint orbit of
G. In this sense coadjoint orbits are “universal” homogeneous phase spaces.
Note that the momentum map (5.38) automatically realizes g symmetry without
central extensions. Indeed, if X, Y belong to g and if JX , JY are the corresponding
momentum maps, then at a point p ∈ g∗ we find
(5.23) (5.38)
{JX , JY }(p) = hp, [(JX )∗p , (JY )∗p ]i = hp, [X, Y ]i = J[X,Y ] (p).
This is exactly the statement (5.36) with a vanishing cocycle c. However, one should
keep in mind that the group G itself may be centrally extended.
JX ≡ hθ, ξX i (5.39)
defines a momentum map (5.32) that satisfies (5.36) with a vanishing cocycle c = 0.
Proof. Since the action leaves θ invariant, one has LξX θ = 0 for any X ∈ g. Writing
the Lie derivative as Lξ = d ◦ iξ + iξ ◦ d and using ω = −dθ, this is equivalent
to dhθ, ξX i = −iξX dθ = iξX ω. One may recognize this as the definition (5.33) of a
Chapter 5. Coadjoint orbits and quantization 110
momentum map given by (5.39). In order to prove that (5.36) is satisfied with a
vanishing cocycle c, we evaluate the Poisson bracket
1 1 1
{JX , JY } = [ξY hθ, ξX i − ξX hθ, ξY i] = ω(ξX , ξY ) − hθ, [ξX , ξY ]i.
2 2 2
Here ω(ξX , ξY ) = {JX , JY } by virtue of (5.20) and (5.34), while eqs. (5.31) and (5.39)
imply that hθ, [ξX , ξY ]i = −J[X,Y ] . Eq. (5.36) follows with c = 0.
Let us now apply this result to the cotangent bundle T ∗ Q of a manifold Q. Let
φ : Q → Q be a diffeomorphism. We define the associated point transformation as
φ ◦ π ◦ φ̄ = π (5.41)
(φ̄)∗ θ = θ . (5.42)
Proposition (5.42) ensures that this action is symplectic and even preserves the Liou-
ville one-form. Accordingly we can apply (5.39) to build its momentum map:
Thus a complex line bundle consists of infinitely many copies of the complex plane
C, one at each point of M, glued together in a smooth way (see fig. 4.1 with Op
replaced by M). The bundle locally looks like the direct product of M with C. When
this is true globally, i.e. when L is diffeomorphic to M × C, the line bundle is said to
be trivial.
Definition. Let L be a complex line bundle over M, VectC (M) the space of complex
vector fields on M. A connection for L is a map
which is linear on Γ(M, L) and VectC (M), and satisfies the property ∇F ξ Ψ = F ∇ξ Ψ
as well as the Leibniz rule ∇ξ (F Ψ) = ξ(F) Ψ + F ∇ξ Ψ for any F ∈ C ∞ (M, C). The
linear operator ∇X is known as the covariant derivative along ξ.
which is a two-form defined for all ξ, ζ ∈ VectC (M) and any section Ψ by
R(ξ, ζ)Ψ ≡ ∇ξ ∇ζ − ∇ζ ∇ξ − ∇[ξ,ζ] Ψ. (5.45)
When the curvature vanishes the connection is said to be flat. Any trivial vector
bundle admits a flat connection, but the converse is not true: there exist non-trivial
bundles with flat connections.
Hermitian structures
Since we eventually wish to interpret sections as wavefunctions, we need to define their
scalar products.
Now let L be a complex line bundle over M endowed with a connection ∇ and a
Hermitian structure (5.46). We say that ∇ is Hermitian if it is compatible with the
Hermitian structure in the sense that
where (Φ|Ψ) is the function M → C whose value at q is Φ(q) Ψ(q) . Condition (5.47)
is the Hermitian analogue of the condition of metric-compatibility for connections on
the tangent bundle. In the realm of quantum mechanics, property (5.47) will allow us
to define self-adjoint operators.
Furthermore the constant function F(p) = 1 must be mapped on the identity opera-
tor, i.e. 1̂ = I. Thus, the problem is to find a quantum/classical correspondence that
fulfills these criteria.
The solution turns out to be given by so-called geometric quantization and consists
of two steps: prequantization and polarization. Here we describe these steps for the
simple case of cotangent bundles endowed with the symplectic form (5.18), so that
ω = −dθ. More general symplectic manifolds are treated in section 5.2.3.
Prequantization
As a first attempt at quantization, let us consider the space of complex wavefunctions
on M. Their scalar products are then given by (3.7) where one may choose µ to be
the Liouville volume form (5.14). To define a linear correspondence between classical
and quantum observables, one can try to use the Hamiltonian vector fields (5.12):
? ?
F 7−→ F̂ = −i~ ξF . (5.49)
We seem to be stuck: how are we to define F̂ in such a way that both condition
(5.48) and the requirement 1̂ = I be satisfied? The way out turns out to be the further
Chapter 5. Coadjoint orbits and quantization 114
where θ is such that ω = −dθ. Indeed, using (5.13) one can verify that the com-
mutators of operators (5.51) close according to (5.48), and furthermore the constant
observable F = 1 is represented, as it should, by the identity operator F̂ = I. Thus,
provided θ exists, the prescription (5.51) is a consistent quantization of the algebra of
classical observables on M.
F̂ = −i~∇ξF + F (5.52)
Polarization
Since the symplectic form is exact, the map (5.51) provides a globally well-defined
quantization prescription and our job here is almost done. But there is still a problem:
the would-be wavefunctions Ψ : M → C depend at this stage on all coordinates of
M = T ∗ Q. For example, on R2n we would have Ψ = Ψ(q i , pj ). In particular, in
the current situation one could easily devise a wavefunction with arbitrarily accurate
values of position and momentum, violating Heisenberg uncertainty. The purpose of
polarization is to cure this pathology by cutting in half the number of coordinates on
which wavefunctions are allowed to depend.
Polarization also affects quantum observables since they must preserve the polar-
ization while still satisfying the commutation relations (5.48). As a result, the space
of quantizable classical observables is a subset of the full space C ∞ (M). For instance,
in R2n with Darboux coordinates q i , pj (i, j = 1, ..., n), the classical observables whose
quantization preserves the polarization ∂p Ψ = 0 all take the form
for some functions F j , G. Observables which are not of this form do not preserve
the polarization and are therefore not quantizable in this sense. One should keep
in mind, however, that this does not mean that all quantum operators acting in the
polarized Hilbert space are forced to take the form (5.53). Rather, quantizable classical
observables give rise to a vector space of Hermitian quantum operators, and the full
algebra of quantum observables is generated by sums and products of these operators.
For example, the non-relativistic Hamiltonian p̂2 is obtained by squaring the operator
that quantizes the classical observable p, although there exists no quantizable classical
observable whose quantization would yield the operator p̂2 . In this way one essentially
recovers standard quantum mechanics from the quantization of the phase space T ∗ Q.
that provide a linear correspondence between classical and quantum observables. The
problem then is to glue together operators defined on different open sets. On any
non-empty intersection Uj ∩ Uk one has dθj = dθk so there exists a function Gjk on
Uj ∩ Uk such that
θj − θk = dGjk . (5.56)
Using (5.55) one can then show that the multiplicative operator
F̂ k
= `kj ◦ F̂ j
◦ `−1
kj on Uj ∩ Uk (5.58)
for any classical observable F ∈ C ∞ (M). This result indicates that the action of F̂
on functions depends on whether one defines it on Uj or on Uk . It is an ambiguity in
the definition of the operator corresponding to F, which threatens the consistency of
the construction based on (5.55). The way out is think of F̂ as a differential operator
acting not on functions, but on sections of a complex line bundle over M. Indeed,
if the line bundle is chosen properly, one may hope that its transition functions for
some local trivialization associated with the open covering {Ui } coincide with the mul-
tiplication maps (5.57), so that the local formula (5.55) provides globally well-defined
differential operators acting on sections.
One is thus led to the problem of determining whether there exists a line bundle
whose transition functions take the form (5.57) for the covering {Ui |i ∈ I}, in such
a way that the operator (5.55) can be written globally as (5.52) for a connection ∇
whose local connection one-forms are the θi ’s. This can be addressed in the framework
of Čech cohomology, which we will not describe here. The bottom line is that such a
line bundle with such a connection exists if and only if the cohomology class of ω/2π~
2
is integral in the cohomology space Hde Rham (M, R), i.e. if
h ω i
2
∈ Hde Rham (M, Z). (5.59)
2π~
This quantization condition is equivalent to demanding that the integral of ω/2π~ over
any closed two-surface be an integer.11 The only quantizable symplectic manifolds are
those that satisfy this requirement.
The reason why we did not see this condition in the case of cotangent bundles is
that their symplectic form is globally exact, so that its cohomology class vanishes and
the requirement (5.59) is trivially satisfied. In fact one can show that the curvature
two-form (5.45) of the connection determined by (5.55) is R = iω/~, consistently with
the fact that the curvature of any line bundle is integral. In particular the connection
used to define quantum operators (5.52) for cotangent bundles is flat.
Provided the quantization condition (5.59) is satisfied, one can endow the space of
sections of the line bundle with a Hermitian structure and use it to define the scalar
product (3.7) thanks to the Liouville volume form (5.14). One can show that the Her-
mitian structure can always be chosen in a way (5.47) compatible with the connection
determined by ω, so that all operators (5.55) are Hermitian.This completes the first
step of geometric quantization, i.e. prequantization.
not be directly visible in our later considerations, we skip its presentation and refer
instead to [116, 118] for a much more thorough discussion.
We will assume that the action of G is Hamiltonian, with a momentum map (5.32).
We also assume that we have chosen a certain value for Planck’s constant ~ and that
ω/2π~ is integral in the sense (5.59). Then (M, ω) is quantizable and prequantization
can be carried out independently of the group action. In particular, for each adjoint
vector X ∈ g there is a classical observable JX given by (5.33), and the corresponding
operator (5.52) is
JˆX = −i~∇ξX + JX (5.60)
where we have used property (5.34) to replace ξJX by the infinitesimal generator (5.30).
By virtue of (5.36) the map X 7→ JˆX is a homomorphism, possibly up to a central
extension. Thus the assignment (5.60) provides a (generally projective) representation
of the Lie algebra g, acting on a space of sections on M.
The subtlety arises with polarization, since then the wavefunctions of the system
satisfy extra conditions which may not be preserved by (5.60). To avoid such patholo-
gies one has to choose a G-invariant polarization. In that case each operator (5.60)
is a well-defined Hermitian operator acting on polarized wavefunctions, and one ob-
tains a projective, unitary representation of g. It was shown by Kostant [120] that,
when the action of G on M is transitive, the homomorphism X 7→ JˆX exponentiates
to a unitary representation of the group G. This is true in particular when M is a
coadjoint orbit [117]. In addition, when G is semi-simple, compact or solvable, the
representations obtained in this way are irreducible. Thus geometric quantization does
produce unitary representations of groups, which is the conclusion we were hoping to
obtain.
manifolds endowed with different symplectic structures. If ω and λω (with λ > 0) are
two symplectic forms on M, then large λ assigns a larger measure to a given portion
of (M, λω) than to the same portion in (M, ω). In this sense large λ is a semi-classical
regime with respect to (M, ω), with 1/λ playing the role of the coupling constant. In
the case of coadjoint orbits, by linearity, Wp is diffeomorphic to Wλp for any λ 6= 0,
but the definition (5.29) ensures that the symplectic form on Wλp is “larger” (for λ > 1
say) than that on Wp . Thus, for λ large enough the quantization of Wλp can be treated
semi-classically. Note that this intuition breaks down if the orbit is invariant under
scalings, i.e. Wλp = Wp .
where the notation −d−1 ω means “whatever one-form θ such that ω = −dθ”. This is
a purely kinematical Hamiltonian action associated with the symplectic form ω. For
example, when M = R2n with ω = dq i ∧ dpi = −d(pi dq i ), expression (5.61) is globally
well-defined and reads Z 1
S[q i (t), pj (t)] = dt pj (t)q̇ j (t) (5.62)
0
which is the standard reparameterization-invariant kinetic term of any Hamiltonian
action. There is no term involving p2 or any other combination of q’s and p’s because
there is no Hamiltonian at this stage.
For a generic symplectic form ω the definition (5.61) is not enough: one needs an
action principle that makes sense for any path in M, regardless of exactness. So let
{Ui |i ∈ I} be a contractible open covering of M such that ω|Ui = −dθi for each i ∈ I.
Chapter 5. Coadjoint orbits and quantization 119
We can then write an action (5.61) on each Ui , but we can also attempt to define S[γ]
for any path γ by Z
S[γ] ≡ − d−1 ω . (5.63)
γ
We refer to this functional as the geometric action for (M, ω) evaluated on the path
γ; its definition follows from the geometry of M. In particular, when a group G acts
on M by symplectomorphisms, the action automatically has global G symmetry. In
section 5.5 we will interpret (5.63) as the action of a point particle in space-time.
The action (5.63) can be evaluated as follows. Given a path γ, we can cover its
image by open sets Uj , with j ∈ J ⊂ I. If only one Uj suffices we can simply use the
original definition (5.61) to evaluate the action. If there are two open sets, say U1 and
U2 , then we call γj the portion of the path γ contained in Uj (for j = 1, 2) and γ12 the
portion contained in U1 ∩ U2 . We can then define
Z Z Z
S[γ] ≡ θ1 + θ2 − θ1 (5.64)
γ1 γ2 γ12
where the last term removes the overcounting due to a double integration on U1 ∩ U2 .
There is a subtlety in this expression: we chose to write ω|U1 ∩U2 = −dθ1 in the last
term, but we could equally well have chosen ω = −dθ2 ; this would have given a different
compensating term in (5.64), hence a different value for the action! This is a problem
at first sight, but one may recall that the action as such need not be a single-valued
functional on the space of paths in phase space. The truly important quantity is the
complex number
eiS[γ]/~ (5.65)
which determines the path integral measure and leads to transition amplitudes in the
quantum theory. Thus we are free to have a multivalued action as long as all ambi-
guities are integer multiples of 2π~. This is in effect a quantization condition on the
parameters of the action.
Given an action (5.63) that satisfies the quantization condition, one can choose
a Hamiltonian H ∈ C ∞ (M) and compute transition amplitudes using path integrals
with the action functional
Z Z T
−1
S[γ] = − d ω − dt H(γ(t)). (5.67)
γ 0
Note that this expression is no longer invariant under time reparameterizations for
generic choices of the Hamiltonian function.
Remark. The geometric actions (5.63) associated with coadjoint orbits of centrally
extended loop groups describe certain families of Wess-Zumino-Witten models [124].
In that context the single-valuedness of (5.65) leads to the quantization of the Kac-
Moody level [127–129].
for any group element g. When G is a matrix group, any f can be written as a
matrix and the entries of f define local coordinates on G. One can then think of
df as the matrix whose entries are the differentials of these coordinates, and the left
Maurer-Cartan form can be written as
Θf = f −1 · df . (5.71)
One can similarly define a right Maurer-Cartan form (Rf −1 )∗f , where R denotes right
multiplication (3.17). Its expression for a matrix group is df · f −1 .
Chapter 5. Coadjoint orbits and quantization 121
for all vector fields ξ, ζ on G, where [·, ·] denotes the Lie bracket (5.1) in g.
Proof. Recall that the exterior derivative of Θ is such that, for all vector fields ξ, ζ,
where [·, ·] is the Lie bracket of vector fields. If ξ and ζ are left-invariant, they can be
written as ξg = (Lg )∗e X and ζg = (Lg )∗e Y for some adjoint vectors X, Y . Then (5.69)
implies that Θ(ξ) = X is constant on g, and (5.73) reduces to
By left-invariance we may write Θ([ξ, ζ]) = [Θ(ξ), Θ(ζ)] where the bracket on the
right-hand side now is the Lie bracket (5.1) of g. Eq. (5.74) then takes the form (5.72)
save for the fact that ξ and ζ are left-invariant. This condition can be relaxed upon
recalling that the span of left-invariant vector fields at a point g ∈ G is the whole
tangent space Tg G.
ω ≡ π∗ω (5.76)
where ω is the Kirillov-Kostant symplectic form (5.29). One may think of ω as the
analogue of (5.29) on the group G.
d d
Ad∗γ(t) (p)
π∗g (v) = π(γ(t)) t=0
= . (5.79)
dt dt t=0
Chapter 5. Coadjoint orbits and quantization 122
In turn we can write γ = g · γ0 (t) where γ0 (0) = e is the identity. Then γ̇0 (0) belongs
to the Lie algebra Te G = g of G and is given by
d d (5.69)
g −1 · γ(t) t=0 =
γ̇0 (0) = Lg−1 γ(t) = (Lg−1 )∗g (v) = Θg (v) (5.80)
dt dt t=0
Eq. (5.77) follows upon plugging this result (and its analogue for w) in (5.78).
Formula (5.77) is sometimes rewritten as
1
ω= hp, [Θ ∧, Θ]i (5.81)
2
where [Θ ∧, Θ] is the g-valued two-form such that [Θ ∧, Θ]g (v, w) ≡ 2[Θg (v), Θg (w)]
for all tangent vectors v, w ∈ Tg G. In what follows we call ω the symplectic form on
G since it is related by (5.76) to the Kirillov-Kostant symplectic form (in particular
dω = 0), but one should keep in mind that this terminology is actually incorrect:
The kernel of ω consists of vectors v = (Lg )∗e X such that (5.82) vanishes for any
Y ∈ g, which is to say that ad∗X p = 0. The latter property holds if and only if X
belongs to the Lie algebra of the stabilizer of p.
This lemma confirms that ω is not a symplectic form because its components do
not form an invertible matrix. The rank of ω is dim(G) − dim(Gp ), where Gp is the
stabilizer of p. This number coincides (as it should) with the dimension of the coadjoint
orbit of p, which proves by the way that the original form (5.29) on G/Gp ∼ = Wp is
invertible.
where the exterior derivative goes through the coadjoint vector p by linearity. Thus
the action (5.83) becomes
Z Z T
dt p, Θf (t) f˙(t) .
S[f (t)] = hp, Θi = (5.84)
f (t) 0
It describes the dynamics of paths f (t) ∈ G and may be seen as the (kinetic piece of
the) action of a non-linear Sigma model. When G is a simple matrix group, adjoint
and coadjoint vectors can be identified so that hp, ·i = Tr[X·] for some X ∈ g, and
(5.71) allows us to recast the integrand of (5.84) in the form Tr Xf −1 f˙ .
A key subtlety with (5.84) is that the group variable f (t) is the group element
that appears in a coadjoint action Ad∗f (t) p, as in (5.68). The latter coadjoint vector is
invariant under multiplication of f (t) from the right by any (generally time-dependent)
group element h(t) belonging to the stabilizer of p. This means that (5.84) should be
invariant under gauge transformations f (t) 7→ f (t) · h(t), and therefore describes a
gauged non-linear Sigma model. Let us check that (5.84) does indeed admit such a
symmetry. Using the Leibniz rule we find
Z TD E Z T D E
S[f ·h] = ˙
p, Θf (t)h(t) (Rh )∗f (t) f (t) + p, Θf (t)h(t) (Lf (t) )∗h(t) ḣ(t) (5.85)
0 0
where the adjoint vector paired with p in the first term can be rewritten as
Θf (t)h(t) (Rh )∗f (t) f˙(t) = Adh−1 Θf (t) f˙(t)
thanks to the definitions (5.69) and (5.6). This implies that the first term of (5.85)
coincides with the original action (5.84). As for the second term in (5.85), we use
left-invariance of Θ to rewrite it as a Sigma model action evaluated on a path wholly
contained in the stabilizer Gp :
Z T D E
S[h(t)] = dt p, Θh(t) ḣ(t) . (5.86)
0
The counterpart of h(t) in the coadjoint orbit of p is the constant path Ad∗h(t) p = p,
but in the Sigma model it carries a generally non-vanishing action (5.86). Thus gauge-
invariance of (5.84) may be true, but is not obvious at this stage since the gauge-
transformed action (5.85) differs from (5.84) by the extra term (5.86). To reconcile
this observation with the much desired gauge-invariance of (5.84), we note that the
exterior derivative of the integrand of (5.86) vanishes. Indeed, for all v, w ∈ Th Gp the
Maurer-Cartan equation (5.72) yields
(5.11)
d hp, Θih (v, w) = − hp, [Θh (v), Θh (w)]i = ad∗Θh (v) p, Θh (w) = 0
Chapter 5. Coadjoint orbits and quantization 124
where the last equality follows from the fact that Θh (v) belongs to the Lie algebra
of the stabilizer of p. Thus the integrand of (5.86) is closed, and is therefore locally
exact. In particular, for a path h(t) located in a sufficiently small neighbourhood of
the identity in H, there exists a function F(t) such that
D E
p, Θh(t) ḣ(t) = Ḟ(t) (5.87)
for any t ∈ [0, T ]. The integral (5.86) of this quantity is a boundary term, so the action
functional (5.84) is indeed gauge-invariant, albeit up to boundary terms that can be
cancelled by requiring for instance that initial and final configurations be fixed.
Here µνρ is the completely antisymmetric tensor such that 012 = +1, and indices
are raised and lowered using the Minkowski metric ηµν = diag(− + +). For future
reference we also note that, in the complex basis
for m, n = −1, 0, 1. On account of the isomorphism (4.83) this can also be seen as the
Lorentz algebra in three dimensions.
Chapter 5. Coadjoint orbits and quantization 125
Note that the fact that coadjoint orbits of SL(2, R) coincide with Poincaré mo-
mentum orbits in three dimensions follows from the structure G nAd gAb of the double
cover (4.93) of the Poincaré group. We will encounter a similar structure in the BMS3
group, albeit with an infinite-dimensional group G.
with orbit representative p = h(t0 )∗ and stabilizer U(1). We denote the “mass” of
the orbit by h rather than M because its quantization will eventually correspond to
a representation of sl(2, R) with highest weight h (or more precisely h + 1/2). The
restriction of (5.92) to the orbit gives rise to the Kirillov-Kostant symplectic form
(5.29), which we now evaluate.
We can label the points of (5.93) by their “spatial components” (q1 , q2 ). In order to
write down (5.29) in these coordinates, we need a dictionary between the components
xµ of X and those of the corresponding vector field ad∗X q in terms of the coordinates q1 ,
q2 . We first evaluate ad∗X q for X = xµ tµ ; using (5.88), for any adjoint vector Y = y µ tµ
we find had∗X q, Y i = −hq, xµ y ν µν ρ tρ i. For q belonging to (5.93) one obtains
q
∗
adX q = − q1 X + q2 X (t ) + − h2 + q12 + q22 X 2 − q2 X 0 (t1 )∗
2 1 0 ∗
q (5.94)
2 2 2 1 0 2 ∗
+ h + q1 + q2 X + q1 X (t ) .
ω = dQ ∧ dP. (5.99)
Hence the Kirillov-Kostant symplectic form on the orbit (5.93) is globally exact and
the quantization condition (5.59) is trivially satisfied for any value of h.
Thus the quantization of the orbit Wp with the Hamiltonian (5.100) is a quantum
harmonic oscillator on the line!
where the symbol “Ad” on the right-hand side denotes the adjoint representation of
G. (More generally, in case of ambiguous notations, the argument of a group action
determines which group it refers to.) In particular, rotation generators transform ac-
cording to the adjoint representation of G, while translations are subject to mixed
transformations involving both the finite action σ and its differential Σ.
From (5.105) one can read off the Lie bracket in g A A upon using (5.8):
(X, α), (Y, β) = [X, Y ], ΣX β − ΣY α . (5.106)
The presence of Σ on the right-hand side justifies calling g A A a semi -direct sum.
Note that, if A was non-Abelian, the second entry on the right-hand side would include
a bracket of generators of A.
The structure of the algebra (5.106) can be made more transparent by choosing a
basis. Let ta be a basis of g satisfying the brackets (5.2), and let αi be a basis of A
(here a = 1, ..., dim g and i = 1, ..., dim A). Introducing the basis elements
that generate the semi-direct sum g A A, the Lie bracket (5.106) yields
where gai k pk ≡ Σta pi so that the coefficients (ga )i k are the entries of the matrix
representing the linear operator Σta : A → A in the basis αi . The brackets (5.107)
make the semi-direct structure manifest since the bracket [j, p] gives p’s while the
bracket [p, p] vanishes on account of the fact that A is Abelian. This structure will
appear repeatedly in this thesis.
g∗ ⊕ A∗ . (5.108)
Its elements are pairs (j, p) where j ∈ g∗ and p ∈ A∗ , paired with adjoint vectors
according to
(j, p), (X, α) = hj, Xi + hp, αi (5.109)
where the first pairing h·, ·i on the right-hand side is that of g∗ with g while the
second one pairs A∗ with A. Note that A∗ is precisely the space of momenta (see
Chapter 5. Coadjoint orbits and quantization 129
section 4.1), while g∗ is dual to infinitesimal rotations and may be seen as a space of
angular momentum vectors. This is consistent with the general interpretation of coad-
joint vectors as conserved quantities (see section 5.1.2) and justifies the notation (j, p).
Definition. The cross product of translations and momenta is the bilinear map A ×
A∗ → g∗ : (α, p) 7→ α × p given for any X ∈ g by
hα × p, Xi ≡ hp, ΣX αi . (5.110)
The notation is justified by the fact that × coincides with the vector product when
G n A is the Euclidean group in three dimensions.
(5.109)
Ad∗(f,α) (j, p), (X, β) = hj, Adf −1 Xi + hp, σf −1 β + σf −1 ΣX αi . (5.112)
In the first term of the right-hand side we recognize the coadjoint representation of G;
the part of the second term involving β is the transformation law (4.16) of momenta;
the last term involves the cross product (5.110) of σf∗ p with α. Collecting all these terms
and removing the argument (X, β), we conclude that the coadjoint representation of
G n A is
Ad∗(f,α) (j, p) = Ad∗f j + α × σf∗ p , σf∗ p
(5.113)
where we keep the notation σf∗ p instead of the simpler f · p to avoid confusion. Note
that the coadjoint action of the translation group A affects only angular momenta,
since the transformation of p only involves f ∈ G. The translation α contributes a
term α×σf∗ p, which for trivial f boils down to the cross product α×p; this contribution
can be identified with a combination of orbital angular momentum and the centre of
mass vector, while the spin angular momentum is contained in Ad∗f j. We will return
to this interpretation below.
where Σ∗X p ≡ −p ◦ ΣX . We will use this formula below when dealing with the Kirillov-
Kostant symplectic form.
Chapter 5. Coadjoint orbits and quantization 130
It remains to understand the geometry of these fibres and the relation between fibres
at different points. Note that in the degenerate case p = 0 the orbit W(j,0) is simply the
coadjoint orbit of j under G; in particular W(0,0) contains only one point. Accordingly
we take p 6= 0 until the end of this section.
So let us describe an orbit W(0,p) . With j = 0 the first entry of the right-hand side
of (5.113) reduces to
α × σf∗ p. (5.116)
Keeping q = σf∗ p fixed, the set spanned by angular momenta of this form is
A × q ≡ {α × q|α ∈ A} ⊂ g∗ (5.117)
and coincides with the set of orbital angular momenta that can be reached by acting
with translations on a particle with momentum q. The geometric interpretation of
(5.117) is as follows. Recall first that the tangent space of Op at q can be identified
with the space of “small displacements” of q generated by infinitesimal boosts:
Tq Op = Σ∗X q X ∈ g ⊂ A∗ .
(5.118)
Chapter 5. Coadjoint orbits and quantization 131
Here Σ∗X q = 0 if and only if X belongs to the Lie algebra gq of the little group Gq ,
so (5.118) is isomorphic to the coset space g/gq . It follows that the cotangent space
Tq∗ Op at q is the annihilator of gq in g∗ ,
Lemma. The cotangent space (5.119) coincides with the set (5.117):
τq : A → g0q : α 7→ α × q (5.121)
mapping a translation on the associated orbital angular momentum. The rank of this
map is dim(A) − dim[Ker(τq )], where
The elements of this kernel are translations constrained by dim(g) − dim(gq ) inde-
pendent conditions (the subtraction of dim(gq ) comes from the quotient by gq ). This
implies that dim[Ker(τq )] = dim(A) − dim(g) + dim(gq ), from which we conclude that
the rank of τq is
dim[Im(τq )] = dim(g) − dim(gq ) = dim(g0q ).
It follows that τq is surjective, which was to be proven.
We have just shown that the span (5.117) at each q ∈ Op is the cotangent space of
Op at q. Since j = 0, this analysis exhausts all points of W(0,p) and we conclude that
Spinning orbits
We now turn to spinning orbits, which generally contain no point with vanishing
total angular momentum. To begin, we pick a coadjoint vector (j, p) and restrict our
attention to rotations f that belong to the little group Gp . The resulting span is
n o
Ad∗f j + α × p, p f ∈ Gp , α ∈ A
(5.125)
and is a subset of the full orbit (5.115). In general Ad∗f (j) 6= j because the little group
Gp need not be included in the stabilizer of j for the coadjoint action of G. Noting
that the cross product (5.110) satisfies the property Ad∗f (α × p) = σf α × σf∗ p, and
using the fact that f fixes p, we rewrite (5.125) as
Ad∗f (j + β × p) , p f ∈ Gp , β ∈ A
(5.126)
where Gj is the stabilizer of j for the coadjoint action of G and all the remaining
6 0.
notation is as before. This extends (5.124) to the case j =
The rewriting (5.126) allows us to see that translations along β can modify at will
all components of j that point along directions in the annihilator g0p . The only piece
of j that is left unchanged by the action of translations is its restriction to gp ,
j gp
≡ jp . (5.128)
where we recognize the cotangent space (5.120) and where Wjp ⊂ g∗p denotes the coad-
joint orbit of jp ∈ g∗p under the little group Gp . This is in fact our main conclusion:
when W(j,p) is seen as a fibre bundle over Op , the fibre at p is a product (5.129) of the
cotangent space of Op at p with the coadjoint orbit of the projection jp of j under the
action of the little group of p.
Inspecting (5.129), note in particular how the little group orbit Wjp factorizes from
the cotangent space A×p due to translations. This splitting is reminiscent of the repre-
sentation (4.28) of Gp n A, where the operators representing f ∈ Gp and α ∈ A live on
very different footings (and actually commute). Recall that we used this representation
to induce an irreducible representation (4.29) of the full group G n A. What we see in
(5.129) is the classical analogue of this little group representation; upon quantization,
the sub-orbit (5.129) will precisely produce a representation of the form (4.28), and its
extension to the full orbit W(j,p) will correspond to the induction (4.29). In particu-
lar the projection (5.128) is a classical definition of spin. We shall return to this below.
Chapter 5. Coadjoint orbits and quantization 133
The arguments that led from (5.125) to the result (5.129) can be run at any other
point q on Op , except that the little group is Gq instead of Gp . Thus the fibre above
any point q = σf∗ p ∈ Op is a product of the cotangent space of Op at q with the
Gq -coadjoint orbit W(Ad∗f j)q , where (Ad∗f j)q denotes the restriction of Ad∗f j to gq . But
little groups at different points of Op are isomorphic: if one chooses standard boosts
gq ∈ G such that σg∗q (p) = q, then Gq = gq · Gp · gq−1 and gq = Adgq gp . Therefore
W(Ad∗f j)q is diffeomorphic to Wjp for any q = σf∗ p ∈ Op ; the relation between the fibres
above q and p is given by the coadjoint action of G n A.
According to our earlier observations, the bundle of little group orbits associated
with (j, p) is really the same as the coadjoint orbit W(j,p) , except that the cotangent
spaces at each point of Op are “neglected” since translations do not appear in (5.130).
Thus B(j,p) is a fibre bundle over Op , the fibre Fq at q ∈ Op being a coadjoint orbit of
the little group Gq . The relation between fibres at different points of Op is given by
the coadjoint action of G n A, or explicitly
(k, q) ∈ Fq iff ∃ f ∈ G such that k = Ad∗f j q and q = σf∗ p.
Conversely, suppose that two elements p ∈ A∗ and j0 ∈ g∗p are given. The group G
can be seen as a principal Gp -bundle over Op , equipped with a natural Gp -action by
multiplication from the left in each fibre. In addition Gp acts on the coadjoint orbit
Wj0 , so one can define an action of Gp on G × Wj0 by
g∈Gp
(f, k) ∈ G × Wj0 7−→ g · f, Ad∗g (k) .
The corresponding bundle of little group orbits is defined as the associated bundle
B(j0 ,p) ≡ (G × Wj0 ) /Gp . (5.131)
Thus one can associate a bundle of little group orbits (5.130) with each coadjoint or-
bit of G n A; conversely, starting from any bundle of little group orbits as defined in
(5.131), one can build a coadjoint orbit of G n A by choosing any j ∈ g∗ such that
jp = j0 and taking the orbit W(j,p) . In other words the classification of coadjoint or-
bits of GnA is equivalent to the classification of bundles of little group orbits [132,133].
Equivalently, W(j,p) is a fibre bundle over the cotangent bundle T ∗ Op , the fibre above
(q, α × q) ∈ T ∗ Op being a coadjoint orbit of Gq . To exhaust all coadjoint orbits of
G n A, one can proceed as follows:
1. Pick an element p ∈ A∗ and compute its momentum orbit Op under the action
σ ∗ of G; let Gp be the corresponding little group.
2. Pick jp ∈ g∗p and compute its coadjoint orbit under the action of Gp .
The set of all orbits Op and of all coadjoint orbits of the corresponding little groups
classifies the coadjoint orbits of G n A. Put differently, suppose one has classified the
following objects:
1. The orbits of G for the action σ ∗ , with an exhaustive set of orbit representatives
pλ ∈ A∗ and corresponding little groups Gλ , with λ ∈ I some index such that
Opλ and Opλ0 are disjoint whenever λ 6= λ0 ;
2. The coadjoint orbits of each Gλ , with an exhaustive set of orbit representatives
jλ,µ ∈ g∗λ , µ ∈ Jλ being some index such that Wjλ,µ and Wjλ,µ0 are disjoint
whenever µ 6= µ0 .
Then the set
(jλ,µ , pλ ) λ ∈ I, µ ∈ Jλ ⊂ g∗ ⊕ A∗
(5.133)
is an exhaustive set of orbit representatives for the coadjoint orbits of G n A. The
(generally continuous) indices λ, µ label the orbits uniquely. This algorithm is a clas-
sical analogue of the classification of representations described in section 4.1, since it
classifies the phase spaces of all “particles” associated with G n A.
Quantization
Suppose we wish to quantize a coadjoint orbit W(j,p) of G n A; let ω be its Kirillov-
in g A A is (5.106), the sym-
Kostant symplectic form (5.29). Since the Lie bracket
∗
plectic form evaluated at the point Adf j + α × q, q ≡ (κ, q) in W(j,p) reads
In the two last terms of this expression we recognize the Liouville symplectic form
(5.136) on the cotangent bundle T ∗ Op when α × q is seen as an element of Tq∗ Op
thanks to (5.120). On the other hand the first term of (5.137) looks like the natural
symplectic form (5.29) on the G-coadjoint orbit of j. In particular, when X and Y
belong to the Lie algebra gq of the little group at q, the second term in (5.137) vanishes
and the first one reduces to
where we use the notation (5.128). This is the natural symplectic form on the coadjoint
orbit W(Ad∗f j)q , so if we see W(j,p) as a fibre bundle over T ∗ Op with typical fibre Wjp ,
restricting the symplectic form (5.137) to a fibre gives back the symplectic form on the
little group’s coadjoint orbit. This observation actually follows from a more general
result, which states that the coadjoint orbits of a semi-direct product are obtained
by symplectic induction from the coadjoint orbits of its little groups. Symplectic in-
duction is the classical analogue of the method of induced representations that yields
irreducible unitary representations of semi-direct products. We will not dwell on the
details of this construction and refer e.g. to [133,137] for a much more thorough treat-
ment.
For quantization to be possible, the symplectic form (5.137) must be integral in the
sense (5.59). But the Liouville two-form (5.136) is exact, so its de Rham cohomology
class vanishes and demanding that (5.137) be integral reduces to demanding integrality
of the symplectic form on the coadjoint orbit of the little group. We conclude (see
e.g. [134] for the proof):
Provided the little group orbit Wjp is quantizable, one obtains a unitary repre-
sentation R of the little group Gp acting on polarized sections of a line bundle over
Wjp . These sections are spin states; as in section 4.1, we denote their Hilbert space
by E. Upon declaring that the polarized sections on T ∗ Op depend only on the coor-
dinates of the momentum orbit Op , polarized sections on the whole orbit W(j,p) can
be seen as E-valued wavefunctions in momentum space. Assuming that there exists a
quasi-invariant measure µ on Op , the Hilbert space H obtained by quantizing W(j,p)
becomes a tensor product (3.8) of E with the space of square-integrable functions
Op → C. This exactly reproduces the construction of section 4.1.
Chapter 5. Coadjoint orbits and quantization 136
Let us describe this in more detail in the scalar case j = 0, so that W(0,p) = T ∗ Op .
Then the Kirillov-Kostant symplectic form coincides with the canonical symplectic
form on T ∗ Op and the operator (5.60) representing a Lie algebra element (X, α) is
d
(Σ∗X q) · Ψ ≡ − Ψ(σe∗−tX q) .
dt t=0
Thus all observables Jˆ(X,α) are polarized and can be quantized so as to satisfy (5.48).
Remark. This theorem says nothing about the exhaustivity of the procedure: it does
not guarantee that all induced representations can be obtained by quantization. In fact
it is easy to work out explicit examples where certain induced representations cannot
follow from geometric quantization, for instance if the little group is not connected. In
this sense geometric quantization is somewhat weaker than the full theory of induced
representations exposed in section 4.1.
We start by evaluating the left Maurer-Cartan form (5.69) for a semi-direct product
with multiplication (4.6). In order to describe a vector tangent to G n A at the point
(f, α), consider a path in G n A given by
γ(t) = g(t), β(t) (5.140)
with g(0) = f , β(0) = α and γ̇(0) ≡ v. Using the group operation (4.6) in G n A, we
then find
(5.69) d
h i
f −1 g(t), σf −1 β(t) − σf −1 α
Θ(f,α) (v) = = Θf ⊕ σf −1 (v) , (5.141)
dt t=0
where on the far right-hand side Θ denotes the Maurer-Cartan form on G. The direct
sum refers to the fact that the tangent space T(f,α) (G n A) is isomorphic to Tf G ⊕ A.
Using (5.141) we can now write the Sigma model action (5.84) associated with the
orbit of a coadjoint vector (j, p) ∈ g∗ ⊕ A∗ :
Z T Z T
(5.109)
S[f (t), α(t)] = dt j, Θf (t) (f˙(t)) + dt σf∗(t) p, α̇(t) . (5.142)
0 0
where hσ ∗ p, dαi is the one-form on G n A that gives hσf∗ p, βi when evaluated at (f, α)
and acting on a vector (v, β). Note that this is just the sum of the Sigma model action
(5.84) on G with a purely kinetic scalar action functional
Z
Sscalar [f (t), α(t)] = hσ ∗ p, dαi (5.144)
(f (t),α(t))
Chapter 5. Coadjoint orbits and quantization 138
describing a point particle propagating in A along a path α(t) with momentum q(t) =
σf∗(t) p. In particular the group A is now interpreted as “space-time”. Expression (5.144)
also has a gauge symmetry with gauge group (5.124), and it is invariant under redef-
initions of the time parameter. As in (5.67), adding a Hamiltonian generally spoils
reparameterization symmetry. In the example of the Poincaré group below the con-
dition p(t) ∈ Op will be a constraint generating time reparameterizations. Note that
this condition only applies to momenta q(t) ∈ A∗ , while the position of the particle,
α(t) ∈ A, is completely unconstrained.
As an example consider the (double cover of the) Poincaré group in three di-
mensions, (4.93). Its momentum orbits coincide with SL(2, R) coadjoint orbits, and
the little groups are stabilizers of SL(2, R) coadjoint vectors. All stabilizers are one-
dimensional and Abelian, except for the trivial orbit whose little group is SL(2, R).
This implies that all Poincaré coadjoint orbits are cotangent bundles of momentum
orbits, except in the case p = 0 for which W(j,0) coincides with the coadjoint orbit
of j under SL(2, R). The set of coadjoint orbit representatives for Poincaré can be
obtained by following the algorithm outlined above (5.133).
The action principle describing a scalar world line is (5.144). We choose a basis
eµ of RD such that each translation can be written as α = αµ eµ . The dual basis
Chapter 5. Coadjoint orbits and quantization 139
consists of momenta (eµ )∗ such that hp, αi = pµ αµ for p = pµ (eµ )∗ . The argument of
the action functional (5.144) is a path in G n A, which we denote (f (τ ), x(τ )) in order
to distinguish the time parameter τ along the world line from the time coordinate
t = x0 . With the coordinates pµ just described we have σf∗(τ ) p µ ≡ pµ (τ ) for some
orbit representative p, and the action becomes
Z T
S[p(τ ), x(τ )] = dτ pµ (τ )ẋµ (τ ) with a constraint pµ (τ )pµ (τ ) = −M 2 ∀ τ,
0
where indices are raised and lowered using the Minkowski metric. The constraint
accounts for the fact that momenta must belong to a massive orbit. It can be incor-
porated in the action thanks to a Lagrange multiplier N (τ ):
Z T h i
S[p(τ ), x(τ ), N (τ )] = dτ pµ (τ )ẋµ (τ ) − N (τ ) pµ (τ )pµ (τ ) + M 2 . (5.145)
0
φ ≡ pµ pµ + M 2 = 0 (5.146)
Note how the non-trivial dynamics emerges from the fact that momenta span an orbit,
even though we haven’t included any Hamiltonian.
ẋµ ẋµ
2
−M = . (5.149)
4N 2
Since our goal is to describe a massive particle, its trajectory must be time-like so we
require that ẋµ remains inside the light-cone at any time τ , which gives ẋµ ẋµ < 0.
This implies that (5.149) has two real solutions N ; we choose the positive one,
p
−ẋµ ẋµ
N= . (5.150)
2M
Together with (5.148) this defines an invertible Legendre transformation from the
space of positions and velocities {(xµ , ẋµ )} to the space of positions and constrained
momenta supplemented with a Lagrange multiplier,
n o
(xµ , pµ , N ) x ∈ RD , p ∈ RD such that p2 = −M 2 , N > 0 .
Chapter 5. Coadjoint orbits and quantization 140
One can also run the argument in reverse and recover the Hamiltonian action from
the Lagrangian one. In doing so one discovers that the mass shell condition is a
primary constraint generating time reparameterizations while N (τ ) is a lapse function
along the world line. The canonical Hamiltonian then reads H = 2N (p2 + M 2 ) and
vanishes on the constraint surface, as usual for generally covariant systems.
Remark. Starting from the Hamiltonian action (5.145), one can evaluate the as-
sociated transition amplitude as a path integral. This computation was performed
in [144–147] and the result turns out to coincide with the Feynman propagator of a
free scalar field with mass M . This observation is one of the starting points of the
world line formalism of quantum field theory [148–150], where scattering amplitudes
are reformulated in terms of point particles propagating in space-time.
In order to write down the action (5.144) we use the same trick as in (5.145) to
express the integrand in components and absorb the constraint p(τ ) ∈ Op with a
Lagrange multiplier N (τ ). (The time parameter along the world line is once again
denoted as τ , in order to distinguish it from the time coordinate t.) Using the pairing
(4.107) the world line action reads
Z T 2
i p
S x(τ ), t(τ ), p(τ ), E(τ ), N (τ ) = dτ pi ẋ − E ṫ − N −E (5.152)
0 2M
where i = 1, ..., D − 1. In principle we should also include a time-dependent central
term λ(τ ) (recall the last entry of (4.105)), but one readily verifies that its contribution
Chapter 5. Coadjoint orbits and quantization 141
to the action is a boundary term so we neglect it from now on. This being said, note
that the presence of the central extension is crucial in giving rise to the constraint
E ≈ p2 /2M obtained by varying N . The equations of motion obtained by varying E
give N = ṫ, so we can once more interpret N (τ ) as a lapse function along the world
line. Plugging the solution of the equations of motion of N and E into (5.152), we get
Z T
p2
i
S x(τ ), t(τ ), p(τ ) = dτ pi ẋ − ṫ ,
0 2M
which we recognize as the action of a free non-relativistic particle moving in RD−1 ,
written in a reparameterization-invariant way (see e.g. chapter 4 of [151]). By express-
ing the action as an integral over the “real time” t = t(τ ), we find
Z T
p2
i
S[x(t), p(t)] = dt pi ẋ − (5.153)
0 2M
where the dot now denotes differentiation with respect to t.
where Ĥ = p̂2 /2M is the Hamiltonian and Jˆ is the angular momentum operator
Jˆ = x̂1 p̂2 − x̂2 p̂1 . (5.155)
The trace (5.154) can be interpreted as the partition function of a free non-relativistic
particle in a frame that rotates at imaginary angular velocity iθ/β. There are at least
two equivalent ways to evaluate it. The first is to compute a time-sliced path integral
Z β
p2
Z
j 1 2
Z(β, θ) = DxDp exp − dτ −ipj ẋ + − iθ(x p2 − x p1 ) (5.156)
0 2M
x(β)=x(0)
where DxDp is the standard path integral measure of quantum mechanics. In the ar-
gument of the exponential we recognize the Euclidean section of (5.153) supplemented
by a term proportional to θJ. Expression (5.156) may thus be seen as the canonical
partition function (3.50) of a system with effective Hamiltonian Ĥeff = Ĥ − iθβ J. ˆ The
second way is to realize that the operator Jˆ generates rotations in the plane. Thus
if we introduce a basis of states |x1 , x2 i localized at (x1 , x2 ), the trace (5.154) is a
(finite-dimensional) integral
Z
Z(β, θ) = dx1 dx2 Rθ · (x1 , x2 ) e−βH x1 , x2 (5.157)
R2
Chapter 5. Coadjoint orbits and quantization 142
where Rθ ·(x1 , x2 ) denotes the action of a rotation by θ on the vector (x1 , x2 ). From this
second viewpoint, the partition function is a trace over transition amplitudes between
initial and final states that are rotated with respect to each other. Since transition
amplitudes can be written as path integrals, expression (5.157) is a path integral in
disguise and takes the same form as (5.156) up to two key differences: (i) the term
iθJ no longer appears in the exponential, and (ii) the periodicity condition on paths
is x(β) = Rθ · x(0) instead of x(β) = x(0).
The two methods just described give identical results, but we pick the second one
for simplicity. Recall that the propagator of a free massive particle on a plane is (in
Dirac notation)
iM |x0 − x|2
0 −iHt M
x ,t e x, 0 = exp (5.158)
2πit 2t
where | · | is the Euclidean norm. From this we find the Euclidean propagator
−βH M M 2
Rθ · x, t e x, 0 = exp − (1 − cos θ)x (5.159)
2πβ 2β
For θ 6= 0 (modulo 2π) this is just a Gaussian integral and the result is precisely a
character (4.123) with r = 1. We conclude that the space obtained by quantizing a
massive coadjoint orbit of the Bargmann group coincides with the Hilbert space of a
free, massive, non-relativistic particle.
Remark. Having seen the computation of the trace of e−βH+iθJ in Bargmann rep-
resentations, one may wonder if the result can be analytically continued to the grand
canonical partition function
Z(β, Ω) = Tr e−β(H−ΩJ)
(5.161)
In the first part of this thesis we have introduced some general tools for dealing with
symmetries in quantum mechanics. Our goal is to eventually apply these tools to the
BMS3 group in three dimensions. Accordingly, in this chapter and the two next ones
we address a necessary prerequisite for these considerations by studying the central
extension of the group of diffeomorphisms of the circle, i.e. the Virasoro group. The
latter is part of the asymptotic symmetry group of many gravitational systems, where
it essentially consists of conformal transformations of celestial circles. It also accounts
for the symmetries of two-dimensional conformal field theories and thus illuminates
certain aspects of holography in general, and AdS3 /CFT2 in particular.
A word of caution is in order at the outset regarding the interpretation of the Vira-
soro group from a gravitational viewpoint. While diffeomorphisms in general relativity
are generally thought of as gauge redundancies, the group Diff(S 1 ) that we shall study
here should by no means be understood in that sense. On the contrary, it should
be interpreted as a global space-time symmetry group on a par with SL(2, R) or the
Poincaré group. In fact, in the BMS3 case, Diff(S 1 ) will be an infinite-dimensional
extension of the Lorentz group in three dimensions. Accordingly this chapter and the
next one may be seen as a detailed investigation of a group that extends Lorentz sym-
metry in an infinite-dimensional way.
Our plan for this chapter is the following. In section 6.1 we define the group Diff(S 1 )
of diffeomorphisms of the circle as an infinite-dimensional Lie group, and we describe
its adjoint representation, its Lie algebra Vect(S 1 ), and its coadjoint representation.
Section 6.2 is devoted to its cohomology; in particular we introduce the Gelfand-Fuks
cocycle and its integral, the Bott-Thurston cocycle, which respectively define the Vira-
soro algebra and the Virasoro group. In section 6.3 we study the Schwarzian derivative,
which will lead to a unified picture of Virasoro cohomology. Finally, in section 6.4 we
define the Virasoro group and work out its adjoint and coadjoint representations; the
latter coincides with the transformation law of two-dimensional CFT stress tensors
under conformal transformations.
Regarding references, the holy book on the Virasoro group is [56] while [55] is a
pedagogical introduction to infinite-dimensional group theory. Some familiarity with
two-dimensional CFT may come in handy at this stage; see e.g. [152–154].
145
Chapter 6. The Virasoro group 146
In the same way that any finite-dimensional manifold looks locally like Rn , one
would like to find the prototypical infinite-dimensional topological vector space V
such that infinite-dimensional manifolds be locally homeomorphic to V. As it turns
out, taking V to be a Fréchet space leads to a well-defined theory of differentiation
and smoothness, which can then be used to define Fréchet manifolds. Roughly speak-
ing, Fréchet spaces are vector spaces that generalize Banach spaces. For example the
space C ∞ (M) of smooth functions on a finite-dimensional manifold M is a Fréchet
space (but not a Banach space). A Lie-Fréchet group then is a group endowed with
a structure of Fréchet manifold such that multiplication and inversion are smooth.
For instance the group Diff(M) of diffeomorphisms of a compact finite-dimensional
manifold M is a Lie-Fréchet group. From now on we refer to infinite-dimensional
Lie-Fréchet groups simply as “infinite-dimensional groups”.
In the remainder of this section we deal with the group of diffeomorphisms of the
circle as an infinite-dimensional Lie(-Fréchet) group. In particular we will think of its
Lie algebra as its tangent space at the identity, identified with the space of left-invariant
vector fields, from which the remaining definitions will follow.
depicted in fig. 2.1. This allows us to think of S 1 as the quotient R/2πZ of the real line
by the equivalence relation ϕ ∼ ϕ + 2π, since the kernel of p consists of translations
of R by integer multiples of 2π.
for any ϕ ∈ R, where prime denotes differentiation with respect to ϕ. The group
operation (6.2) then becomes
f ·g =f ◦g (6.8)
where f, g correspond to F, G according to (6.6). From now on we always describe
Diff(S 1 ) in terms of diffeomorphisms of R satisfying the properties (6.7). By the way,
this is why we have kept writing group elements as “f ” throughout this thesis.
Note that the diffeomorphism F does not determine f uniquely: one can add to
f (ϕ) an arbitrary constant multiple of 2π without affecting F = eif . This ambiguity
can be removed by requiring e.g. that f (0) belongs to the interval [0, 2π[. One says that
f is a lift of F , and there are infinitely many lifts for a given F . For our purposes it is
only important that giving f determines F = eif uniquely, so that we can consistently
write all orientation-preserving diffeomorphisms of the circle in the form (6.7).
For each t, ft satisfies (6.7) and therefore defines a diffeomorphism of the circle. At
t = 0 it coincides with f while at t = 1 it is the identity. See fig. 6.2.
For the purposes of representation theory it is important to know the fundamental
group of Diff+ (S 1 ), as it determines whether Diff+ (S 1 ) has topological central exten-
sions. In that context the key result is the following:
π1 Diff+ (S 1 ) ∼
= Z. (6.10)
Figure 6.1: The two connected components of Diff(S 1 ) are related by parity. Compare
with the connected components of the Lorentz group in fig. 4.2.
Proof. We follow [156]. The key to the proof is to realize that Diff+ (S 1 ) is homotopic
to its subgroup Isom+ (S 1 ) of orientation-preserving isometries of the circle (for the
standard flat metric). Since Isom+ (S 1 ) is a group U(1) of rigid rotations, it will follow
that Diff+ (S 1 ) has the homotopy type of a circle and therefore has a fundamental
group Z. So let us prove the homotopy equivalence Diff+ (S 1 ) ∼ Isom+ (S 1 ). Call
Diff+ 1
0 (S ) the group of orientation-preserving diffeomorphisms of the circle leaving the
point ϕ = 0 fixed. Since isometries of S 1 are rotations, there exists a decomposition
Diff+ (S 1 ) = Diff+ 1 + 1
0 (S ) · Isom (S ). (6.11)
Indeed, any diffeomorphism of the circle is the composition of a rigid rotation with a
diffeomorphism leaving ϕ = 0 fixed; both Diff+ 1 + 1
0 (S ) and Isom (S ) are groups and their
intersection only contains the identity. Now note that any diffeomorphism preserving
ϕ = 0 admits a unique lift f such that f (0) = 0 and f (2π) = 2π, so we can think
of Diff+ 1
0 (S ) as the set of 2πZ-equivariant diffeomorphisms of R that fix the point
ϕ = 0; this identification is one-to-one. It only remains to observe that the maps
(6.9) define a homotopy whose effect at t = 1 is to retract the whole Diff+ 1
0 (S ) on the
identity. As a result Diff0 (S ) is homotopic to a point, and so by (6.11) Diff+ (S 1 ) is
+ 1
homotopic to a circle. Unwinding this circle gives rise to the group of 2πZ-equivariant
diffeomorphisms of R, which therefore span the universal cover of Diff+ (S 1 ).
This lemma confirms the interpretation of Diff+ (S 1 ) as an infinite-dimensional ana-
logue of PSL(2, R), since the latter is also homotopic to a circle (see section 4.3). In
particular formula (6.11) is the Diff(S 1 ) analogue of the Iwasawa decomposition (4.81)
of SL(2, R). Since Diff+ 1
0 (S ) has the homotopy type of a point, the group Diff (S )
+ 1
In what follows we focus on the universal cover Diff g + (S 1 ) rather than Diff(S 1 )
or Diff+ (S 1 ), except if explicitly stated otherwise. To reduce clutter we will abuse
notation by writing Diff(S 1 ) for the universal cover, instead of the more accurate
g + (S 1 ). Accordingly, from now on elements of Diff(S 1 ) are diffeomorphisms
notation Diff
f , g, etc. of the real line satisfying the properties (6.7). The inverse of f will be denoted
f −1 and is such that f (f −1 (ϕ)) = f −1 (f (ϕ)) = ϕ.
Chapter 6. The Virasoro group 150
Figure 6.2: The homotopy (6.9) turns a diffeomorphism f (ϕ) (here leaving the point
ϕ = 0 fixed) into the identity. It implies both that the group Diff+ (S 1 ) is connected
and that it is homotopic to a circle.
dh i (6.13) dh i
f γt f −1 (ϕ) f f −1 (ϕ) + tX f −1 (ϕ)
Adf (X) (ϕ) = = .
dt t=0 dt t=0
(6.14)
Since t is “small” in this expression, we can Taylor expand
where we have used f (f −1 (ϕ)) = ϕ. The derivative of the latter equation implies
1
f 0 (f −1 (ϕ)) = (6.16)
(f −1 )0 (ϕ)
which can be plugged into (6.15) and thus provides the adjoint representation
X (f −1 (ϕ))
Adf (X) (ϕ) = . (6.17)
(f −1 )0 (ϕ)
Chapter 6. The Virasoro group 151
We recognize here the transformation law of the component X(ϕ) of a vector field
∂
X(ϕ) (6.19)
∂ϕ
Take once more a path γt in Diff(S 1 ) that satifies (6.13). Picking a vector field
Y ∈ Vect(S 1 ) and a point ϕ ∈ [0, 2π], let us evaluate
Y (γt−1 (ϕ))
(5.8) d (6.17) d
(adX (Y )) (ϕ) = Adγt (Y ) t=0 (ϕ) = . (6.20)
dt dt (γt−1 )0 (ϕ) t=0
Here (6.13) implies (γt−1 )0 (ϕ) = 1 − tX 0 (ϕ) as well as Y (γt−1 (ϕ)) = Y (ϕ) − tX(ϕ)Y 0 (ϕ)
to first order in t. Plugging these expressions in (6.20) we obtain adX Y = −XY 0 +Y X 0 ,
where it is understood that both sides are evaluated at the same point ϕ. We conclude
that the Lie bracket of Vect(S 1 ), seen as the Lie algebra of the group Diff(S 1 ), is the
opposite of the standard Lie bracket of vector fields:
Proposition. The Lie algebra of the group Diff(S 1 ) is the space Vect(S 1 ) of vector
fields on the circle, with the Lie bracket (6.21) given by the opposite of the standard
Lie bracket of vector fields.
In what follows we will bluntly neglect the sign subltety and endow Vect(S 1 ) with
the usual bracket
∂
[X, Y ] = X(ϕ)Y 0 (ϕ) − Y (ϕ)X 0 (ϕ) . (6.22)
∂ϕ
This is a harmless abuse of conventions and may be seen as an alternative definition
of the Lie bracket for groups of diffeomorphisms. With that abuse the Lie algebra of
Diff(S 1 ) becomes Vect(S 1 ) with the usual Lie bracket of vector fields.
Chapter 6. The Virasoro group 152
Witt algebra
Since all functions on the circle can be expanded in Fourier series, any vector field is
a (generally infinite) complex linear combination of generators
`m ≡ eimϕ ∂ϕ , m ∈ Z. (6.23)
The brackets (6.22) of these generators can be written as
i[`m , `n ] = (m − n)`m+n , (6.24)
where one may recognize the Witt algebra of conformal field theory. It is an infinite-
dimensional extension of the sl(2, R) algebra (5.90) spanned by `−1 , `0 , `1 . The latter
consists of vector fields X(ϕ)∂ϕ with X(ϕ) = X0 + X1 cos ϕ + X2 sin ϕ for Xµ ∈ R, and
generates diffeomorphisms of the form (6.3). In particular the constant vector field `0
generates rigid rotations of the circle.
or equivalently as
α(ϕ)
(f · α)(f (ϕ)) ≡ . (6.28)
(f 0 (ϕ))h
This reduces to (6.18) for h = −1 and to (6.25) for h = 1. If we think of f (ϕ) as
a “conformal transformation” of the circle, then eq. (6.28) coincides with the trans-
formation law of a (chiral) primary field of weight h. It provides a representation of
Diff(S 1 ) in the space Fh (S 1 ). This representation is infinite-dimensional and generally
non-unitary because the would-be “scalar product”
Z 2π
dϕ α(ϕ)β(ϕ) , α, β ∈ Fh (S 1 ) (6.29)
0
is not left invariant by (6.28) for generic values of h. The only exception is the case
of spinor fields, h = 1/2. One can think of (6.28) as a Diff(S 1 ) generalization of the
various finite-dimensional (but non-unitary) irreducible representations of the Lorentz
group. The number h can then be thought of as a spin label, in the same way that
finite-dimensional Lorentz representations correspond to transformation laws of rela-
tivistic fields with definite spin. (Beware: the word “spin” here does not refer to the
notion of “spin” encountered in representations of semi-direct products. These two no-
tions are related in that the Lorentz spin of a quantum field determines the Poincaré
spin of the corresponding particles, but they are nevertheless different concepts.)
From (6.28) one can read off the transformation law of densities under infinitesimal
diffeomorphisms, that is, under vector fields on the circle. Taking f (ϕ) = ϕ + X(ϕ)
in (6.28) one finds, to first order in ,
We then define
(f · α)(ϕ) − α(ϕ)
X · α(ϕ) ≡ − (6.31)
and obtain
X · α = Xα0 + h αX 0 . (6.32)
As before, one may recognize here the infinitesimal transformation law of a primary
field α of weight h under an conformal transformation generated by X.
p : Fh (S 1 ) → R : α 7→ hp, αi . (6.33)
where p(ϕ) is a smooth function on the circle. We will call the space of such distribu-
tions the smooth or regular dual of Fh (S 1 ). As a vector space, it is isomorphic to the
space C ∞ (S 1 ) of smooth functions on the circle. Note that any distribution can be
obtained as the limit of a sequence of regular distributions, so in this sense we are not
missing anything even when restricting attention to regular distributions. The regular
dual of Fh (S 1 ) will be denoted as Fh (S 1 )∗ . The notation in (6.33) is justified by the
fact that, in BMS3 , p will be an infinite-dimensional supermomentum vector dual to
supertranslations.
Since Fh (S 1 ) carries a representation (6.28), it is natural to ask how the dual repre-
sentation (4.16) acts on the regular dual. By definition, one has hf · p, αi = hp, f −1 · αi
for any diffeomorphism f , any density α ∈ Fh (S 1 ) and any smooth distribution
p ∈ Fh (S 1 )∗ . Using (6.27) and the pairing (6.34), we get
Z 2π
1
hf · p, αi = dϕ p(ϕ)(f 0 (ϕ))h α(f (ϕ)) . (6.35)
2π 0
Thus the duals of densities with weight h are densities with weight 1 − h, and
vice-versa. One can apply this to the examples encountered above:
• the duals of spinor fields (h = 1/2) are spinor fields (i.e. F1/2 (S 1 ) is self-dual).
Note that, for all values of h except h = 1/2, the conformally invariant pairing (6.34)
is not a scalar product since its arguments are densities whose transformation laws
under Diff(S 1 ) differ. This should be contrasted with the finite-dimensional examples
Chapter 6. The Virasoro group 155
encountered in chapter 4, where the existence of an invariant bilinear form led to the
equivalence of σ ∗ and σ. We shall see below that this difference is crucial for coadjoint
orbits of the Virasoro group (section 7.1) and hence for the supermomentum orbits of
the BMS3 group (see part III).
Coadjoint representation
Since the adjoint representation of Diff(S 1 ) is the transformation law (6.18) of vector
fields, we now know that the coadjoint representation of Diff(S 1 ) is the transformation
law of quadratic densities:
p(ϕ)
(Ad∗f p)(f (ϕ)) = (6.36)
(f 0 (ϕ))2
In CFT terminology, vector fields are infinitesimal conformal transformations and their
duals are (quasi-)primary fields with weight h = 2, that is, CFT stress tensors. Indeed
formula (6.36) is the transformation law of a stress tensor p(ϕ) if we think of the map
ϕ 7→ f (ϕ) as a conformal transformation. Similarly, the duals of functions are primary
fields with weight h = 1, i.e. currents. From now on we sometimes refer to Diff(S 1 )-
invariance as “conformal invariance”. Note that at this point we haven’t included any
central charge yet; this will change once we turn to the Virasoro group.
In particular, when t = is “small” one finds that ϕ() takes the form (6.12) with
initial condition ϕ(0) = ϕ. Equation (6.38) is an ordinary differential equation in one
dimension and X(ϕ) is smooth, so given an initial condition ϕ(0), the solution exists
and is unique. We define the flow of X as the one-parameter family of diffeomorphisms
that maps a “time” t and an initial condition ϕ on the point ϕ(t) obtained by solving
(6.38) with this initial condition. If we call this solution ϕ̃(t, ϕ), then the flow of X is
For example the flow of the constant vector field X(ϕ) = 1 is given by ϕ̃(t) = ϕ + t
and consists of rigid rotations by t, as already anticipated above. Using the notion of
flow, one can define an exponential map for Diff(S 1 ):
Chapter 6. The Virasoro group 156
In any finite-dimensional Lie group, the exponential map (5.3) is a local diffeomor-
phism, so any group element belonging to a suitable neighbourhood of the identity can
be written as the exponential of an element of the Lie algebra. However, this is not
so for groups of diffeomorphisms: one can show that the exponential map (6.40) does
not define a local chart on Diff(S 1 ) in that it is neither locally injective, nor locally
surjective. The idea of the proof is to build and explicit family of diffeomorphisms that
are arbitrarily close to the identity but cannot be written as exponentials of vector
fields. This being said, the exponential map is always well-defined on a Lie-Fréchet
group, even when it is not locally surjective. See [55, 56] for details.
With this motivation, the present section is devoted to the cohomology of Diff(S 1 )
and its Lie algebra. These considerations will eventually lead to the definition of
the Virasoro algebra, so we refer to them as “Virasoro cohomology”. We will start
by describing the real cohomology groups of Vect(S 1 ) and of Diff(S 1 ), then turn to
cohomologies whose cochains are primary fields on the circle. The results summarized
here are discussed at greater length in [56].
Remark. We use the notation and conventions of chapter 2, and all cochains are
required to be smooth. Lie algebra cochains are denoted by lowercase sans serif letters
such as b, c, s, etc. while group cochains are denoted by uppercase letters B, C, S, etc.
H1 Vect(S 1 ) = 0.
(6.42)
In other words there is no non-trivial real one-cocycle on Vect(S 1 ). The second coho-
mology of Vect(S 1 ) is far more interesting:
m3
c(`m , `n ) = −i δm+n,0 . (6.44)
12
Proof. Let c be a real two-cocycle on Vect(S 1 ). Then dc = 0 where d is the Chevalley-
Eilenberg differential (2.15) for the trivial representation T . In terms of the basis
(6.23), the statement dc = 0 is tantamount to
Here we can interpret the left-hand side as the differential of the one-cochain k =
c(`0 , ·), so the left-hand side is exact while the right-hand side is a Lie derivative1
where we used dc = 0. Since the left-hand side of (6.46) is exact, we conclude that Lie
derivation with respect to `0 leaves the cohomology class of c invariant. (In geometric
terms `0 generates rotations, so this says that the cohomology class of c is invariant
under rotations.) This allows us to turn c into a rotation-invariant cocycle. Indeed,
let b be a one-cochain and define c̃ ≡ c + db, which has the same cohomology class as
c. The Lie derivative of c̃ with respect to `0 now is
(6.46)
L`0 c̃ = L`0 c + L`0 db = dk + d(i`0 (db)) = d k + db(`0 , ·) . (6.48)
In order to make c̃ invariant under rotations, we need to choose b such that (6.48)
vanishes. One verifies that the definition
i
b(`m ) = c(`0 , `m ) for m 6= 0 (6.49)
m
1
We denote by i the interior product of cochains.
Chapter 6. The Virasoro group 158
satisfies this requirement for any b(`0 ). Thus, from now on we work only with the
rotation-invariant cocycle c̃ and we rename it into c for simplicity. Then we have
c(`0 , `m ) = 0, and eq. (6.46) becomes
c([`0 , `m ], `n ) + c(`m , [`0 , `n ]) = 0 (6.50)
for all integers m, n. The Lie brackets (6.24) then yield
(m + n) c(`m , `n ) = 0 (6.51)
and thus imply that c(`m , `n ) = 0 whenever m + n is non-zero. Writing c(`m , `n ) =
cm δm+n,0 for some coefficients cm = −c−m , we are left with the task of determining the
cm ’s with m > 0. Returning to the cocycle identity (6.45) with p = −m − 1 and using
once more the brackets (6.24), we find
(2 + m)cm − (2m + 1)c1
cm+1 = (6.52)
m−1
for m ≥ 2. This shows that all cm ’s are determined recursively by c1 and c2 . In
particular, we now know that the cohomology space H2 Vect(S 1 ) is at most two-
dimensional; the choices cm = m3 and cm = m are indeed two linearly independent
solutions of the recursion relations (6.52). Now note that, if c is a coboundary c = dk
for some one-cochain k, then
(2.15) (6.24)
c(`m , `n ) = dk(`m , `n ) = k([`m , `n ]) = −i(m − n)k(`m+n )
so that c(`m , `−m ) = −2imk(`0 ) always depends linearly on m. Accordingly, the
solution cm = m of the recursion relations (6.52) yields a trivial cocycle, while cm = m3
is non-trivial. We conclude that, up to a coboundary, any non-trivial two-cocycle on
Vect(S 1 ) reads
c(`m , `n ) = N m3 δm+n,0 (6.53)
2 1
for some normalization N 6= 0. In particular, H Vect(S ) is one-dimensional.
The first step is to fix the kind of cochains one wants to study. For Vect(S 1 ) it is
natural to consider local real cochains
c : Vect(S 1 )k → R : (X1 , ..., Xk ) 7→ c(X1 , ..., Xk ) (6.54)
that take the form of an integral over S 1 of some “cochain density” C :
Z 2π
(n ) (n )
c(X1 , ..., Xk ) = dϕ C X1 (ϕ), X10 (ϕ), ..., X1 1 (ϕ), ..., Xk (ϕ), ..., Xk k (ϕ) .
0
Here the word local is used in the same sense as in field theory. The Gelfand-Fuks
cocycle (6.43) is of that form, with a density C (X, Y ) ∝ XY 000 . With this restriction on
k
the allowed cochains, one can study the resulting cohomology groups Hloc Vect(S 1 ) .
The result is as follows:
Chapter 6. The Virasoro group 159
k
Proposition. The real local cohomology groups Hloc Vect(S 1 ) are all trivial except
if k is equal to 0, 2 or 3, in which case the cohomology group is one-dimensional:
(
k R if k ∈ {0, 2, 3}
Vect(S 1 ) =
Hloc (6.55)
0 otherwise.
0
The generator of Hloc is the class of any non-zero constant function on Vect(S 1 ); that
2 3
of Hloc is the Gelfand-Fuks cocycle (6.43). Finally Hloc is generated by the class of
the Godbillon-Vey cocycle
Z 2π X Y Z
dϕ det X 0 Y 0 Z 0 (6.56)
0 X 00 Y 00 Z 00
Note that the full, generally non-local, cohomology groups of Vect(S 1 ) do not
coincide with (6.55) because they contain classes of wedge products of the Gelfand-Fuks
and Godbillon-Vey cocycles. For example c ∧ c is a non-trivial, non-local four-cocycle
on Vect(S 1 ) when c is the Gelfand-Fuks cocycle.
Remark. In this work the Godbillon-Vey cocycle (6.56) will be unimportant. How-
ever, it does play a key role in a specific context, as it was shown in [157] that it
is responsible for the unique non-trivial gauge-invariant deformation of a higher-spin
Chern-Simons action in three dimensions with gauge algebra Vect(S 1 ) Aad∗ Vect(S 1 )∗ .
Cocyclic recipes
We start by describing a general algorithm for building two-cocycles on a group [54,56].
Let M be an orientable manifold endowed with a volume form µ. For any orientation-
preserving diffeomorphism f : M → M, we define a function T[f −1 ] on M by
−1 ]
f ∗ µ ≡ eT[f µ. (6.57)
is a W-valued two-cocycle on G.
Proof. We need to show that dC = 0 for the Chevalley-Eilenberg differential (2.31),
given that W is acted upon by G according to the representation T . Using the fact
that Ω is bilinear and antisymmetric together with property (6.59), one readily verifies
by brute force that this is indeed the case.
We consider the circle S 1 endowed with the flat volume form µ = dϕ. Under a
0
diffeomorphism ϕ 7→ f (ϕ) we have (f ∗ µ)ϕ = d(f (ϕ)) = f 0 (ϕ)dϕ = elog(f (ϕ)) µ, so
(6.57) provides a C ∞ (S 1 )-valued twisted derivative
2
The fact that the same letter denotes the cocycle T and the representation T is merely a nota-
tional coincidence.
Chapter 6. The Virasoro group 161
To apply the construction (6.60), we also need to find a bilinear antisymmetric map
Ω : C ∞ (S 1 ) × C ∞ (S 1 ) → R which is invariant under Diff(S 1 ) in the sense that (6.59)
holds when T is the trivial representation while S is the action (6.58) of Diff(S 1 ) on
functions. A natural guess is
Z 2π Z
∞ 1 ∞ 1 0
Ω : C (S ) × C (S ) → R : (F, G) 7→ dϕ F(ϕ)G (ϕ) = FdG , (6.62)
0 S1
Lemma. If T is the twisted derivative (6.61), then for all f, g ∈ Diff(S 1 ) one has
Z Z
T[f ]dT[f ◦ g] = T[(f ◦ g)−1 ]dT[g −1 ] . (6.66)
S1 S1
Proof. We use two key properties: the first is the fact that T is a one-cocycle with
respect to the action of Diff(S 1 ) on C ∞ (S 1 ), so
and the second is a property that follows from the definition (6.61) and eq. (6.16):
Thanks to this lemma we can write the Bott-Thurston cocycle (6.64) in a more
convenient way, without f −1 ’s all around the place:
Z Z
1 0 0 1
C(f, g) = − log(f ◦ g) d log(g ) = − T[(f ◦ g)−1 ]dT[g −1 ]. (6.69)
48π S 1 48π S 1
This is the definition that we will be using from now on.
At this stage the Bott-Thurston cocycle seems to be coming out of the blue. How-
ever it turns out that (6.69) is, in fact, a very natural quantity. We will explain this
in greater detail in section 6.3, but for now we simply note the following relation:
d2 h i
C etX , esY − C esY , etX
c(X, Y ) = − . (6.70)
dt ds t=0, s=0
to first order in t, s. Relation (6.70) follows. It also follows that the Bott-Thurston
cocycle is non-trivial, since the Gelfand-Fuks cocycle is non-trivial.
Remark. The bilinear map (6.62) is a non-trivial two-cocycle on the Abelian Lie
algebra C ∞ (S 1 ) of smooth functions on the circle. It defines a central extension of
C ∞ (S 1 ) that can be interpreted in several ways: either as an infinite-dimensional
Heisenberg algebra, or as a u(1) Kac-Moody algebra. This kind of central extension
occurs for instance in the realm of warped conformal field theories [52, 93].
for some weight h. The functional C depends on the Xi ’s and finitely many of their
derivatives, all evaluated at the same point ϕ. We denote the corresponding cohomol-
ogy spaces by Hk Vect(S 1 ), Fλ (S 1 ) . In order to avoid technical considerations we
state the results without proof and refer to [56] for details.
Chapter 6. The Virasoro group 163
The result (6.71) implies that the non-trivial primary cohomology of Vect(S 1 ) is
localized only on non-negative integers with weights h ∈ N. Here we briefly describe
the non-trivial first cohomology groups (k = 1) for the cases h = 0, 1, 2 that will be
useful below.
The latter may be recognized as the infinitesimal cocycle corresponding to the twisted
derivative (6.61) by differentiation.
The notation w is because this cocycle is relevant to certain aspects [52] of warped
conformal symmetry [93].
One can go on and similarly classify all cohomology groups with higher weight h. Since
we will not need these results here, we refrain from displaying them (see e.g. [159,160]).
Instead, we now relate the cocycles (6.72), (6.73) and (6.74) to one-cocycles on Diff(S 1 ).
where d denotes the exterior derivative on the circle. As mentioned above this
“warped derivative” has been used recently [52] to describe certain aspects of
warped conformal field theories.
R Note that in these terms the Bott-Thurston
cocycle (6.63) is C(f, g) ∝ T[f ] ⊗ W[f ◦ g].
The one-cocycle (6.74) can similarly be related to the F2 -valued Schwarzian derivative
on Diff(S 1 ), although the integration is somewhat less trivial than in the two cases
just described. The Schwarzian derivative is crucial for our upcoming considerations,
so the whole next section is devoted to it.
Many references use the notation {f ; ϕ}, but we will stick to S[f ](ϕ) instead.
where Ad∗ denotes the coadjoint representation (6.36) of Diff(S 1 ). Upon defining
S̃[f ] ≡ S[f −1 ]dϕ2 , one obtains a map that associates a quadratic density with any
diffeomorphism f , and which satisfies
d h i 1
C(f, etX ) + C(f etX , f −1 ) ≡− hS[f ], Xi (6.79)
dt t=0 12
for any f ∈ G and any adjoint vector X ∈ g. (The normalization is chosen so that S
eventually coincides with the Schwarzian derivative.)
−1 d h
(6.79) −1 −1 tX −1 −1 tX
i
S[(f g) ], X = C(g f , e ) + C(g f e , f g) . (6.80)
dt t=0
Using the cocycle identity (6.65) together with property (5.7), one can then show by
brute force that (6.81) coincides with the right-hand side of (6.80).
Chapter 6. The Virasoro group 166
In the case of the group Diff(S 1 ), the Souriau construction yields the Schwarzian
derivative from the Bott-Thurston cocycle. Let us check this explicitly: taking g(ϕ) =
ϕ + tX(ϕ) in (6.69), one finds
Z 2π " 00 2 #
000
t f f
C(f, etX ) = − dϕ − X(ϕ),
48π 0 f0 f0
Z 2π " 00 2 #
000
t f f
C(f ◦ etX , f −1 ) = (t-independent) − dϕ − 2 X(ϕ)
48π 0 f0 f0
to first order in t. It then follows that relation (6.79) holds when S is the Schwarzian
derivative (6.76) and h·, ·i is the pairing (6.34) of Vect(S 1 ) with its dual. Note that by
taking f (ϕ) = ϕ + sY (ϕ) with small s, the Schwarzian derivative reduces to S[f ] =
sY 000 . Upon differentiating with respect to s in the right-hand side of (6.79), we recover
precisely the Gelfand-Fuks cocycle (6.43).
Virasoro universality
At this point the cohomological constructions of the previous pages are starting to
fit in a global picture of Virasoro cohomology: eq. (6.70) relates the Bott-Thurston
cocycle to the Gelfand-Fuks cocycle (6.43), while (6.79) relates it to the Schwarzian
derivative, which in turn is the integral of the infinitesimal cocycle (6.74). In addition
the integral of the latter with a vector field on the circle reproduces the Gelfand-Fuks
cocycle. The common feature of all these expressions is the occurrence of third deriva-
tives such as f 000 or X 000 , which will indeed play a key role in the sequel (and give rise
to the term m3 in (6.44)).
In this sense, all these cocycles are really one and the same quantity, albeit ex-
pressed in very different ways. Depending on one’s viewpoint, one may decide that
the most fundamental quantity is the Gelfand-Fuks cocycle, or the Schwarzian deriva-
tive, or Bott-Thurston. Our point of view will be that the Bott-Thurston cocycle is
the most fundamental of them all, since it yields the other ones by differentiation:
Bott-Thurston
C
SouriauSouriau
y
Schwarzianderivative S
differentiatedif f erentiate
y
infinitesimalSchwarzian s
pair with Vect(S 1 )pair with Vect(S 1 )
y
Gelfand-Fuks c
SL(2, R)/Z2 = PSL(2, R). Our goal here is to explore this relation. Accordingly
we start with a short detour through one-dimensional projective geometry, before
recovering the Schwarzian derivative as a quantity that measures the extent to which
diffeomorphisms deform the projective structure. Along the way we will encounter the
expression of Lorentz transformations in terms of diffeomorphisms of the circle.
Denoting by [(x, y)] the equivalence class of (x, y) ∈ R2 , the projective line is thus
In topological terms the projective line is a circle centred at the origin in R2 with
antipodal points identified. This is to say that RP 1 ∼ = S 1 /Z2 , where Z2 acts on S 1 by
rotations. Since any group Zn acting on the circle by rotations of 2π/n is such that
S 1 /Zn ∼
= S 1 , the projective line is actually diffeomorphic to a circle:
RP 1 ∼
= S 1. (6.83)
The diffeomorphism (6.83) can be made explicit in terms of well chosen coordinates.
Indeed, in terms of (6.82), the projective line is a union RP 1 = {[(ζ, 1)]|ζ ∈ R} ∪
{[(1, 0)]}, so the projective coordinate
ζ ≡ x/y (6.84)
is a local coordinate on RP 1 that misses only one point, namely the class of (1, 0). In
this sense the projective line is a real line R with an extra “point at infinity”.
This is exactly the same situation as with the stereographic coordinate on a circle.
For later convenience we define this coordinate in terms of an angular coordinate ϕ on
the circle by
eiϕ + 1
ζ = −cot(ϕ/2) = iϕ (6.85)
ie − i
(see fig. 6.3). The diffeomorphism (6.83) is then obtained by identifying this stereo-
graphic coordinate with the projective coordinate (6.84).
Chapter 6. The Virasoro group 168
Figure 6.3: The stereographic coordinate (6.85) is obtained by projecting the points of
a unit circle in the (X, Y ) plane on the Y axis, along a straight line that goes through
the “east pole” (1, 0). Upon writing X = cos ϕ and Y = sin ϕ and declaring that
the projective coordinate ζ is minus the Y coordinate of the projection, one obtains
(6.85). The minus sign is included so as to preserve the orientation of the circle in the
sense that dζ/dϕ > 0.
Projective transformations
The projective line inherits a symmetry from the linear action of GL(2, R) on R2 .
Explicitly, an invertible matrix
a b
(6.86)
c d
acts on the coordinate (6.84) as a projective transformation
aζ + b
ζ 7→ . (6.87)
cζ + d
Upon identifying the projective coordinate (6.84) with the stereographic coordinate
(6.85), the transformation (6.87) can be reformulated in terms of the angular coordi-
nate ϕ. Using (6.87) and the inverse of (6.85), we find that projective transformations
act on eiϕ according to
Aeiϕ + B
eiϕ 7→ eif (ϕ) = (6.88)
B̄eiϕ + Ā
where the complex coefficients
1 1
A = (a + ib − ic + d), B = (a − ib − ic − d) (6.89)
2 2
Chapter 6. The Virasoro group 169
are such that |A|2 − |B|2 = 1. The family of transformations (6.88) spans a subgroup
PSL(2, R) of Diff(S 1 ) that we already anticipated in (6.3). In section 9.1 we shall
interpret that subgroup as the Lorentz group acting on null infinity. For infinitesimal
parameters A = 1 + i and B = ε (with ∈ R and ε ∈ C), formula (6.88) becomes an
infinitesimal diffeomorphism
This is an sl(2, R) transformation generated by the vector fields `0 , `1 and `−1 men-
tioned below (6.24). Conversely, any transformation (6.88) belongs to the flow of an
sl(2, R) vector field.
Remark. The relation between S 1 and RP 1 discussed here has a complex analogue
CP 1 ∼= S 2 , where CP 1 is the complex projective line. In this generalization the
projective coordinate (6.84) becomes a complex coordinate z and coincides with the
stereographic coordinate (1.4) of S 2 . The projective transformations of CP 1 then are
Möbius transformations (1.6), i.e. Lorentz transformations in four dimensions.
(ζ1 − ζ3 )(ζ2 − ζ4 )
[ζ1 , ζ2 , ζ3 , ζ4 ] ≡ .
(ζ1 − ζ2 )(ζ3 − ζ4 )
This number is a projective invariant, as one can verify by direct computation. Now
take a diffeomorphism f : S 1 → S 1 , where we think of S 1 as a projective line (6.83).
This diffeomorphism can be written in terms of the projective coordinate (6.84), giving
rise to a map ζ 7→ f(ζ). The explicit correspondence between f and f follows from
(6.85) and reads
f −cot(ϕ/2) = −cot f (ϕ)/2 . (6.90)
In general, f is not a projective transformation (6.87) and therefore spoils the projective
structure of S 1 . This can be measured by taking a point with coordinate ζ on RP 1
together with three other nearby points, then evaluating the change in their cross-ratio
under the action of f. Let therefore ζ1 = ζ + , ζ2 = ζ + 2 and ζ3 = ζ + 3 be three
points close to ζ. These points move under the action of f, and one can show that
where S is the Schwarzian derivative (6.76). Thus we have recovered the Schwarzian
derivative as a measuring device that tells us how much the diffeomorphism f spoils
the projective structure of RP 1 . From this one concludes:
In order to go from (6.92) to S[f ](ϕ) we use the cocycle identity (6.77) repeatedly.
First we write
(6.77) 2
S[f ](ϕ) = S[log(eif (ϕ) )](ϕ) = S[eif (ϕ) ](ϕ) + if 0 (ϕ)eif (ϕ) S[log](eif (ϕ) ) , (6.93)
(6.76) 1
where S[log](x) = 2x2
. The first term on the far right-hand side of (6.93) involves
(6.77) 2 (6.92) (6.76) 1
S[eif (ϕ) ](ϕ) = S[eiϕ ](ϕ) + (eiϕ )0 S[eif (ϕ) ](eiϕ ) = S[eiϕ ](ϕ) = ,
2
which finally gives
1 2
1 − f 0 (ϕ)
S[f ](ϕ) = (6.94)
2
when f (ϕ) is given by (6.88). We will put this formula to use in the next chapter.
Note that these observations can be generalized to infinitely many other families of
diffeomorphisms of the circle. Indeed, pick a positive integer n ∈ N∗ and take formula
(6.88) with ϕ replaced by nϕ and f (ϕ) replaced by nf (ϕ):
Aeinϕ + B
einf (ϕ) = , |A|2 − |B|2 = 1. (6.95)
B̄einϕ + Ā
This defines a diffeomorphism of the circle, and the family of such diffeomorphisms
also spans a subgroup of Diff(S 1 ) which is locally isomorphic to SL(2, R). The dif-
ference with respect to the case n = 1 discussed above is that the corresponding Lie
algebra sl(2, R) is generated by `0 , `n and `−n , and that the actual group spanned by
such transformations is an n-fold cover of PSL(2, R); we shall denote this cover by
SL(n) (2, R)/Z2 ≡ PSL(n) (2, R). One can also verify that the Schwarzian derivative of
the diffeomorphism f defined by (6.95) satisfies
n2
1 − (f 0 (ϕ))2 ,
S[f ](ϕ) = (6.96)
2
generalizing the case n = 1 of (6.94).
Chapter 6. The Virasoro group 171
Remark. One can show that the restriction of the Bott-Thurston cocycle (6.69)
to the PSL(2, R) subgroup (6.88) coincides with the unique non-trivial two-cocycle on
PSL(2, R). The latter acts on the hyperbolic plane H2 by isometries of the form (6.87),
where ζ ∈ C has positive imaginary part, and the two-cocycle associates with two
such transformations f, g the area of the triangle with corners i, f −1 (i) and g−1 ◦ f −1 (i).
See [56] for details.
Adjoint representation
Since the group G b consists of pairs (f, λ) where f ∈ G and λ ∈ R, its Lie algebra b
g
consists of pairs (X, λ) where X ∈ g. The adjoint representation of G b then follows
from (5.6): for X ∈ g, f ∈ G and λ, µ ∈ R we find
It then remains to compute the two entries on the right-hand side (6.97). The
first entry yields the adjoint representation of G, while the second is precisely the
expression (6.79) defining the Souriau cocycle S associated with C. We conclude that
the adjoint representation of Gb reads
The first entry on the right-hand side is the same as expression (5.8) in G; accordingly
it boils down to the standard Lie bracket of g, which we denote as [X, Y ]. The second
entry involves the differential of the Souriau cocycle,
d
s[X] ≡ S[etX ] t=0
, (6.100)
dt
paired with Y ∈ g. We therefore define a two-cocycle c on g by
1
c(X, Y ) ≡ − hs[X], Y i (6.101)
12
and the bracket of bg takes the centrally extended form (2.26):
h i 1
(X, λ), (Y, ρ) = [X, Y ], c(X, Y ) = [X, Y ], − hs[X], Y i . (6.102)
12
The fact that (6.101) is indeed a two-cocycle is inherited from the Souriau cocycle.
Note that the central terms λ, µ commute with everything, as they should. In terms
of Lie algebra generators the bracket (6.102) takes the general form (2.27).
Coadjoint representation
The Lie algebra bg is spanned by pairs (X, λ), so its dual consists of pairs (p, c) where
p belongs to g∗ while c ∈ R is a real number, paired with adjoint vectors according to
where the pairing h·, ·i on the right-hand side is that of g∗ with g. The number c is
known as a central charge. The coadjoint transformation law of (p, c) follows from
the definition (5.10). In particular, since central elements act trivially in the adjoint
c∗ c∗
representation (6.97), we can safely write Ad (f,λ) ≡ Adf for any (f, λ) ∈ G, where the
b
∗
hat on top of Ad indicates that we refer to a representation of the centrally extended
group. If then (X, λ) ∈ b g and (p, c) ∈ bg∗ , one obtains
D E
c ∗ (p, c), (X, λ) (6.98)
D 1 −1
E
Ad f = (p, c), Ad f −1 , λ − S[f ], X
12
(6.103) c
= hp, Adf −1 Xi + cλ − S[f −1 ], X
12
where the pairing h·, ·i on the right-hand side of the last equation is the centreless
pairing of g∗ with g. In particular the first term is simply the coadjoint representation
of G. Removing the dependence on X, we conclude that
∗
∗ c −1
Adf (p, c) = Adf p − S[f ], c .
c (6.104)
12
Chapter 6. The Virasoro group 173
Note that the central charge c is left invariant by the coadjoint representation, as it
should. Crucially, it also appears in the first entry and thus affects the transformation
law of p. In abstract terms, formula (6.104) is the affine G-module (2.33) associated
with the Souriau cocycle.
d 1 ).
We shall denote the Virasoro group by Diff(S
The adjoint representation of a group yields the Lie brackets (5.8) of its algebra.
In the present case this definition leads to an awkward sign (6.21), which we absorb
by declaring that the Lie bracket of the Virasoro algebra is defined by
d hc i
(X, λ), (Y, µ) ≡ − AdetX (Y, µ) . (6.107)
dt t=0
Chapter 6. The Virasoro group 174
With this definition formula (6.99) holds up to an overall minus sign on the right-hand
side. Using then the infinitesimal Schwarzian derivative (6.74), the pairing (6.34)
allows us to recognize the Gelfand-Fuks cocycle (6.43) in hs[X], Y i. Thus the Lie
bracket of the algebra of the Virasoro group takes the form (6.102), or explicitly
(X, λ), (Y, µ) = [X, Y ], c(X, Y ) (6.108)
d 1 ) = Vect(S 1 ) ⊕ R en-
Definition. The Virasoro algebra is the Lie algebra Vect(S
dowed with the Lie bracket (6.108). It is the universal central extension of Vect(S 1 ).4
where the `m ’s are given by (6.23). The bracket (6.108) then yields [Z, Z] = [Z, Lm ] =
0, as well as
(6.108)
i[Lm , Ln ] = i[(`m , 0), (`n , 0)] = (i[`m , `n ], ic(`m , `n )) .
Using the Witt algebra (6.24) and eq. (6.44), we can rewrite this as
Z 3
i[Lm , Ln ] = (m − n)Lm+n + m δm+n,0 , (6.110)
12
which is indeed the standard expression of the Virasoro algebra [152–154]. In this form
it can be seen as a central extension of the Witt algebra (6.24), with a central term
involving the celebrated m3 δm+n,0 . As mentioned below (6.44), the latter is a remnant
of the third derivative of Y in the Gelfand-Fuks cocycle (6.43), while the δm+n,0 follows
from the integration over the circle and reflects the fact that the cocycle is invariant
under rotations.
Remark. The generator Z of eq. (6.109) should rightfully be called the “central
charge” of the Virasoro algebra, since it is a Lie algebra element that commutes with
everything. However, in keeping with the standard physics terminology, we will also
use the word “central charge” to refer to the dual of Z, which is just a real number c
(see the coadjoint representation below). This ambiguous terminology should not lead
to any confusion.
d 1 )∗ .
We refer to pairs (p, c) as Virasoro coadjoint vectors; they span the space Vect(S
It is worth mentioning that coadjoint vectors are crucial physical quantities in all
theories enjoying Diff(S 1 ) symmetry, and in particular all conformal field theories in
two dimensions. Indeed the function p(ϕ) is nothing but (the chiral component of) a
CFT stress tensor, while c is a CFT central charge. Expression (6.111) then coincides
(up to central terms) with the Noether charge associated with a symmetry generator
X(ϕ)∂ϕ , seen as an infinitesimal conformal transformation. More precisely, in a CFT
on a Lorentzian cylinder, the coordinate ϕ would be replaced by one of the two light-
cone coordinates x± and p(ϕ) would become p(x+ ) or p̄(x− ). This is consistent with
the interpretation of coadjoint vectors as Noether currents, thanks to the momentum
maps of section 5.1.
Coadjoint representation
The transformation law of Virasoro coadjoint vectors follows from the definition (5.10).
In particular, since central elements act trivially in the adjoint representation (6.97),
c∗ c∗
we may write Ad (f,λ) ≡ Adf for any (f, λ) belonging to the Virasoro group. If then we
d 1 ) be an adjoint vector and (p, c) ∈ Vect(S
let (X, λ) ∈ Vect(S d 1 )∗ be a coadjoint one,
formula (6.104) still holds upon letting S be the Schwarzian derivative. The central
charge c is left invariant by the coadjoint representation, as it should, but it also affects
the transformation law of p(ϕ). Accordingly, from now on we often write the coadjoint
representation of the Virasoro group without including a second slot for the central
charge, since the latter is invariant. With this simplified notation formula (6.104) boils
down to
c ∗ p = Ad∗ p − c S[f −1 ].
Ad (6.112)
f f
12
For future reference it will be useful to rewrite this in detail, in terms of functions on
the circle. Evaluating both sides of the equation at a point ϕ on the circle, we obtain
by virtue of the centreless coadjoint action (6.36). The formulas are much simpler if
we evaluate eq. (6.112) at f (ϕ); using the cocycle identity (6.77) we find
c ∗ p (f (ϕ)) =
1 h c i
Adf p(ϕ) + S[f ](ϕ) . (6.114)
(f 0 (ϕ))2 12
c ∗ p)(ϕ) − p(ϕ)
(Ad
c∗ p(ϕ) ≡ − f
adX
as in (6.31), we end up with the coadjoint representation of the Virasoro algebra:
c∗ p = Xp0 + 2X 0 p − c X 000 ,
ad (6.115)
X
12
where both sides are evaluated at the same point.5 This is the Virasoro version of
eq. (6.105). In the homogeneous term we recognize the primary transformation law
(6.32), while the central term involves the infinitesimal Schwarzian (6.74).
d 1 )∗ .
functions on Vect(S
Now recall the basis (6.109) of the Virasoro algebra and let {(Lm )∗ , Z ∗ } denote
the corresponding dual basis, such that hL∗m , Ln i = δmn and hZ ∗ , Zi = 1. Using the
pairing (6.111) we find that, as coadjoint vectors,
Thus, when writing a quadratic density as a Fourier series (6.116), the parameters
d 1 )∗ defined with respect to the basis (6.117):
pm , c are actually coordinates on Vect(S
X
p(ϕ)dϕ2 , c = pm L∗m + c Z ∗ .
m∈Z
Accordingly, eq. (5.28) implies that the Kirillov-Kostant Poisson bracket of these co-
ordinates reproduces the Lie brackets (6.110):
c 3
i{pm , pn } = (m − n)pm+n + m δm+n,0 , (6.118)
12
while all Poisson brackets involving the central charge c vanish. The key difference
between (6.118) and (6.110) is that the latter is an abstract Lie bracket, while the
former is its phase space realization.
The bracket (6.118) is well-known to physicists. Indeed, the standard way to intro-
duce the Virasoro algebra in CFT textbooks is to expand the stress tensor in modes as
in (6.116), and then compute their Poisson brackets. Upon quantization, the operator
[
i{·, ·} coincides with the commutator [·, ·] and the resulting quantum commutators
span a Virasoro algebra (6.110)-(6.118), generally with a non-zero central charge c.
d 1 )∗ that maps
Note that each coordinate pm can be seen as the function on Vect(S
(p, c) on h(p, c), Lm i. As mentioned below (6.111), the object h(p, c), Lm i may be
thought of as the Noether charge associated with the symmetry generator Lm , so the
Poisson bracket (6.118) can be interpreted as a Poisson bracket of Noether charges.
We shall see in chapters 8 and 9 that the Poisson brackets of surface charges in three-
dimensional gravity coincide with the Kirillov-Kostant brackets on the dual of suitable
asymptotic symmetry algebras (albeit with definite values of the central charges).
Chapter 7
In this chapter we classify the coadjoint orbits of the Virasoro group. Aside from
their usefulness in the study of conformal symmetry, they are crucial for our purposes
because they will turn out to coincide with the supermomentum orbits that classify
BMS3 particles. As we shall see, despite being infinite-dimensional, these orbits be-
have very much like the finite-dimensional coadjoint orbits of SL(2, R).
The plan is as follows. In section 7.1 we describe the problem and explain how
it can be addressed in terms of two invariant quantities, namely the conjugacy class
of a certain monodromy matrix and the winding number of a related curve taking its
values in a circle. We then use this approach in section 7.2 to display explicit orbit
representatives. Finally, section 7.3 is devoted to a discussion of energy positivity in
the Virasoro context.
Coadjoint orbits of the Virasoro group were first classified in [161, 162] and are
described in many later papers [121,163–165] and textbooks [55,56]. The presentation
of this chapter follows [166].
179
Chapter 7. Virasoro orbits 180
Diff(S 1 ) = Diff
g + (S 1 ), whose coadjoint representation is given by eq. (6.36).
Let us pick a coadjoint vector p(ϕ)dϕ2 , c = 0 and denote its coadjoint orbit by
W(p,0) . For now, suppose for simplicity that p(ϕ) is strictly positive for all ϕ ∈ [0, 2π].
One can then verify that the integral
√ 1
Z 2π p
M≡ dϕ p(ϕ) (7.1)
2π 0
is invariant under the action (6.36) of Diff(S 1 ) on p. This actually follows from the
fact that p is a quadratic density, so its square root is a one-form and can be integrated
on the circle in a Diff(S 1 )-invariant way. With this notation the diffeomorphism
Z ϕ r
p(φ)
f (ϕ) ≡ dφ (7.2)
0 M
maps p(ϕ) on the constant coadjoint vector f · p = M . Thus any strictly positive
coadjoint vector p belongs to the orbit of a constant M given by (7.1), which is the
“mass” associated with p(ϕ). The stabilizer of p(ϕ) = M is the set of diffeomorphisms
f such that M = M/(f 0 (ϕ))2 . Since we set f 0 > 0 to preserve orientation, the only
solution is f 0 = 1 and the stabilizer of p = M is the group U(1) of rigid rotations
f (ϕ) = ϕ + θ. Thus the orbit of any strictly positive coadjoint vector is diffeomorphic
to the quotient space Diff(S 1 )/S 1 . The same analysis applies, up to signs, to strictly
negative coadjoint vectors. Note that Diff(S 1 )/S 1 has codimension one in Diff(S 1 ).
Of course, coadjoint vectors may well vanish at certain points of the circle, and
in particular they can change sign; the previous analysis must then be modified. For
example, suppose p(ϕ) is everywhere non-negative, but vanishes at the point ϕ = 0.
We will say that p(ϕ) has a “double zero” at ϕ = 0, since both p(ϕ) and p0 (ϕ) vanish
there. Then the integral (7.1) is still invariant on the orbit of p, but it is no longer
true that p(ϕ) can be mapped on a constant because the corresponding would-be
diffeomorphism (7.2) is degenerate: its derivative vanishes at ϕ = 0. We conclude
that the orbit of p is now specified by two invariant statements: first, the fact that
√
the integral of p takes the value (7.1), and second, the fact that p(ϕ) has one double
zero. More generally, if p(ϕ) is everywhere non-negative but has N double zeros at
the points ϕ = ϕ1 , ..., ϕN , then the N integrals
Z ϕi+1 p
dϕ p(ϕ)
ϕi
A similar treatment can be applied to coadjoint vectors that change sign on the
circle, i.e. functions p(ϕ) having simple zeros (where p0 does not vanish). The number
of such points is always even since p(ϕ) p is 2π-periodic, so let us suppose p(ϕ) has
2N 0 simple zeros. Then the integral of |p(ϕ)| between any two consecutive zeros is
Chapter 7. Virasoro orbits 181
(where ϕi and ϕi+1 are any two consecutive zeros), together with the ordering of these
invariants, the specification of whether the points ϕi and ϕi+1 are simple or double
zeros, and the sign of p on a given interval, say [ϕ1 , ϕ2 ]. The orbit of p then has
codimension N + 2N 0 in Diff(S 1 ); in particular there exist orbits with arbitrarily high
codimension.
As we can see here, centreless coadjoint orbits are somewhat messy: they can be
specified by an arbitrarily large number of parameters. Besides, we haven’t even dis-
cussed the case of coadjoint vectors p(ϕ) that vanish on a whole open set in S 1 — these
have an infinite-dimensional little group and their orbits have infinite codimension in
Diff(S 1 ). In particular, coadjoint orbits can have arbitrary (even or odd) codimension
in Diff(S 1 ). This is in sharp contrast with finite-dimensional Lie groups, where all
coadjoint orbits are even-dimensional since they are symplectic manifolds. In the case
of Diff(S 1 ), coadjoint orbits are still symplectic, but they need not satisfy “codimension
parity”: the fact that a given orbit has codimension N does not imply that there are
no orbits with codimension N ± 1. We shall see that this pathology does not occur
when the Virasoro central charge is non-zero, where all orbits have codimension one
or three.
While this task was relatively easy in the centreless case thanks to the integral in-
variants (7.3), it turns out to be much more complicated in the centrally extended case;
the remainder of this chapter is devoted to orbits at non-zero central charge, where
the classification will rely on elaborate techniques involving monodromy matrices. For
now we simply describe the most elementary aspects of some of these orbits.
Chapter 7. Virasoro orbits 182
Stabilizers
Suppose we are given a coadjoint vector p(ϕ)dϕ2 , c . Since the central charge is
invariant, the orbit of (p, c) under the Virasoro group can be represented as
n o
W(p,c) = Adc ∗ p f ∈ Diff(S 1 ) , (7.4)
f
∗
c p is given by (6.114). It is an infinite-dimensional manifold, so obtaining
where Ad f
information on its geometry sounds at first like an impossible task. Accordingly,
instead of actually trying to picture the orbit as such, let us look for the stabilizer Gp
of p, which is a subgroup of Diff(S 1 ) such that
W(p,c) ∼ d 1 )/(Gp × R) ∼
= Diff(S = Diff(S 1 )/Gp . (7.5)
Given p(ϕ), this is a highly non-linear differential equation for f (ϕ); if we could actu-
ally solve it, we would know the stabilizer.
To make things simpler let us look only for the Lie algebra of the stabilizer, rather
than the stabilizer itself. This algebra is spanned by vector fields X that leave p(ϕ)
invariant, which according to (6.115) amounts to the requirement
c 000
Xp0 + 2X 0 p − X = 0. (7.7)
12
This is already a lot easier than eq. (7.6): it is a linear third order equation for the
function X(ϕ), assuming that the function p(ϕ) is known. A number of important
consequences follow from this equation. The first is that, for non-zero c, it admits at
most three linearly independent solutions:
This is already a sharp difference with respect to the centreless case, where sta-
bilizers had arbitrarily high dimension. If we were on a line rather than a circle, we
would actually conclude from (7.7) that the stabilizer is always three-dimensional; but
the requirement of periodicity restricts the space of allowed solutions X for a given p,
as we shall see momentarily.
wherepA is real while B and C are generally complex coefficients, being understood
that 24p0 /c is purely imaginary when p0 < 0. For generic values of p0 , the only 2π-
periodic solution of this type is a constant X(ϕ) = const. In that case the stabilizer
is one-dimensional, and consists of rigid rotations of the circle. But there also exist
exceptional values of p0 whose stabilizer is larger, namely
n2 c
p0 = − (7.9)
24
where n ∈ N∗ is a positive integer. At such values the exponentials in (7.8) are e±inϕ
and the corresponding vector field X is automatically 2π-periodic (and real upon set-
ting C = B ∗ ). Thus, for exceptional constants (7.9), the stabilizer is three-dimensional.
Its Lie algebra is isomorphic to sl(2, R); we will see below that the stabilizer itself is
an n-fold cover of PSL(2, R). In particular, orbits of generic constants p0 are radically
different from orbits of exceptional constants (7.9). The situation is depicted in fig. 7.1.
Until the next section it will be convenient to think of the coordinate ϕ as span-
ning the real line R, without identification ϕ ∼ ϕ + 2π (this will be justified below).
Accordingly we now reinstate the notation Diff g + (S 1 ) for the universal cover of the
group of orientation-preserving diffeomorphisms of the circle and we think of it as a
subgroup of Diff+ (R). Functions on the circle then are 2π-periodic functions on R.
We also sometimes use the words “conformally invariant” or “conformally equivalent”
g + (S 1 )-invariant or Diff
to refer to objects that are Diff g + (S 1 )-equivalent, respectively.
Chapter 7. Virasoro orbits 184
Figure 7.1: The map of Virasoro coadjoint orbits with constant representatives. The
open dots labelled by n = 1, 2, 3, ... indicate the location of the exceptional points
−c/24, −4c/24, −9c/24, etc.
c ∂2
∆(p,c) ≡ − + p(ϕ) (7.11)
6 ∂ϕ2
where ϕ is a coordinate on the real line, p(ϕ) is 2π-periodic, and the operator ∆(p,c)
acts on suitable densities on the real line. The normalization in front of ∂ϕ2 is chosen
g + (S 1 ),
so as to ensure that the operator has good transformation properties under Diff
as we shall see below. Note that the crucial term ∂ϕ2 disappears if c = 0, which is why
the considerations that follow apply exclusively to orbits with non-zero central charge.
for the real-valued function ψ(ϕ) on the real line. With the notation (7.11) this is just
the statement ∆(p,c) · ψ = 0.
Lemma. If ψ(ϕ) is a density with weight −1/2 on the real line, then Hill’s equation
g + (S 1 ).
(7.12) is invariant under Diff
Proof. To simplify formulas, let us act on Hill’s operator with a diffeomorphism f −1
rather than f so that f −1 · ψ (ϕ) = ψ(f (ϕ))(f 0 (ϕ))−1/2 . Then f −1 maps the left-hand
side of Hill’s equation (7.12) on
c
− ∂ϕ2 ψ(f (ϕ))(f 0 (ϕ))−1/2 + p(f (ϕ))ψ(f (ϕ))(f 0 (ϕ))3/2
6 (7.14)
c
− S[f ](ϕ) (f 0 (ϕ))−1/2 ψ(f (ϕ))
12
where the term with a second derivative can be written as
1
∂ϕ2 ψ(f (ϕ))(f 0 (ϕ))−1/2 = ψ 00 (f (ϕ))(f 0 (ϕ))3/2 − ψ(f (ϕ))(f 0 (ϕ))−1/2 S[f ](ϕ) .
2
Here the term involving the Schwarzian derivative cancels that of (7.14); the latter
expression can therefore be rewritten as
3/2 h c 00 i
f 0 (ϕ) − ψ (ϕ) + p(ϕ)ψ(ϕ) .
6
Provided ψ solves (7.12), this vanishes.
As a consequence of this lemma, the map that associates Hill’s equations with
g + (S 1 )-invariant. Thus, Hill’s equation is an
Virasoro coadjoint vectors (p, c) is Diff
invariant associated with each coadjoint orbit of the Virasoro group, and classifying
g + (S 1 )-inequivalent Hill’s equations.
Virasoro orbits is equivalent to classifying all Diff
1
Note that Hill’s operator (7.11) coincides with a Sturm-Liouville operator with periodic potential.
Chapter 7. Virasoro orbits 186
Monodromy
Hill’s equation (7.12) is a second-order linear differential equation, so its solutions span
a two-dimensional vector space. Let ψ1 and ψ2 be two linearly independent solutions.
We define their Wronskian as
ψ1 ψ2
W ≡ det = ψ1 ψ20 − ψ2 ψ10 . (7.15)
ψ10 ψ20
The Wronskian is constant on the real line (W 0 = 0) by virtue of Hill’s equation.
Furthermore W does not vanish since ψ1 and ψ2 are linearly independent. (Conversely,
if the Wronskian does not vanish, then the solutions ψ1 , ψ2 are linearly independent.)
Thus we can always choose
W [ψ1 , ψ2 ] = −1 . (7.16)
We will refer to this equality as the “Wronskian condition” and to the solutions that
satisfy it as being “normalized”. Note that the Wronskian (7.15) is invariant under
g + (S 1 ) when the ψi ’s transform as densities of weight −1/2, regardless of them
Diff
solving Hill’s equation:
Proof. Let ψ1 , ψ2 be two linearly independent solutions of (7.12), and define ψ̃i (ϕ) ≡
ψi (ϕ + 2π) for i = 1, 2. Then the Wronskian associated with ψ̃1,2 takes the same value
as that of ψ1,2 ; furthermore, the functions ψ̃i solve the same Hill’s equation as the
functions ψi since p(ϕ) is 2π-periodic. This implies that there exists a real matrix
M such that (7.18) holds for any ϕ ∈ R. Since the ψi ’s and the ψ̃i ’s have the same
Wronskian, M must have unit determinant.
Thus we can associate a monodromy matrix with any Virasoro coadjoint vector
and any pair of (normalized) solutions of the corresponding Hill’s equation. From now
on we use the notation
ψ1
Ψ≡ (7.19)
ψ2
for the “solution vector” associated with the basis of solutions ψ1 , ψ2 . Relation (7.18)
then becomes Ψ(ϕ + 2π) = M · Ψ(ϕ). If we were to choose another normalized basis
Chapter 7. Virasoro orbits 187
is invariant under changes of bases. It depends only on the function p(ϕ), and not on
the choice of solutions Ψ.
We have shown above that Hill’s equation is invariant under Diffg + (S 1 ) in the sense
that, if ψ solves the equation with a potential p(ϕ), then f · ψ solves the same equation
with a potential f · p. In addition we have seen in (7.17) that this transformation
preserves the Wronskian condition, so that normalized solutions remain normalized
g + (S 1 ). Accordingly, if Ψ is a normalized solution vector for the potential p,
under Diff
then f · Ψ is a normalized solution vector for f · p. And now comes the key argument:
g + (S 1 ) on Ψ are linear, the monodromy
since both Hill’s equation and the action of Diff
matrix of f · Ψ coincides with that of Ψ. We therefore conclude:
Theorem. Let c 6= 0 and denote by [M](p,c) the conjugacy class of any monodromy
matrix M associated with the Hill’s equation (7.12) specified by p(ϕ) and c. Then
there is a well-defined map
that associates with a coadjoint orbit W(p,c) the equivalence class [M](p,c) of the corre-
sponding monodromy matrix. In particular, Virasoro coadjoint vectors with the same
central charge but non-conjugate monodromy matrices do not belong to the same orbit.
This result illustrates the power of Hill’s operators. It provides a rough classi-
fication of Virasoro orbits by allowing us to distinguish orbits with non-conjugate
monodromies and may be seen as an infinite-dimensional analogue of the classification
of coadjoint orbits of SL(2, R) according to the value of the “mass squared”. In partic-
ular the trace Tr(M) is a conformally invariant quantity. However, the classification is
not precise in that two orbits whose monodromy matrices are conjugate may well be
different: the map (7.20) need not be injective (and we shall see below that it is not).
To make further progress we need to investigate Hill’s equation in more detail.
For future reference, note the following: thanks to the fact that the conjugacy class
of the monodromy matrix is independent of the choice of a solution vector Ψ for Hill’s
equation associated with (p, c), one can write its trace as a Wilson loop
Z 2π
0 1
Tr(M) = Tr P exp dϕ (7.21)
0 6p(ϕ)/c 0
where P denotes path ordering. This quantity is conformally invariant, so one can
replace (p, c) by any coadjoint vector (q, c) belonging to its orbit without affecting the
Chapter 7. Virasoro orbits 188
The same formula holds for negative p0 , with cosh(ix) = cos(x). We will put it to use
in section 10.1 when defining the mass of BMS3 particles.
all solve the stabilizer equation (7.7). These products are generally not 2π-periodic
and therefore do not represent vector fields on the circle, but one can show that there
always exist either one or three 2π-periodic linear combinations of these products.
This confirms our earlier observation that the stabilizer of all orbits is either one- or
three-dimensional. Note that, being −1/2-densities on the circle, the products (7.23)
were bound to be densities of weight −1, i.e. vector fields.
Let us now see how the stabilizer Gp of (p, c) is described in the Hill language. If
f ∈ Gp and if Ψ is a normalized solution vector of Hill’s equation associated with (p, c),
the action of Gp on Ψ is such that f · Ψ provides another normalized solution vector
for the same equation. Accordingly there exists some (constant) SL(2, R) matrix Af
such that
f · Ψ = A−1
f Ψ. (7.24)
In addition we have seen that the action of Diff(S 1 ) leaves the monodromy matrix
invariant, so the monodromy of f ·Ψ coincides with the monodromy M of Ψ. Combining
this statement with (7.24) we conclude that A−1 f MAf = M, which is to say that Af
belongs to the stabilizer GM of M with respect to conjugation. In addition the inversion
A−1
f in (7.24) ensures that Af g = Af Ag , so we conclude:
A : Gp → GM : f 7→ A(f ) ≡ Af (7.25)
This map relates the stabilizer of p to that of the corresponding monodromy matrix.
In particular it allows us to classify the conformally inequivalent solutions of Hill’s
equation at fixed (p, c). Indeed, the set of normalized solution vectors of Hill’s equation
Chapter 7. Virasoro orbits 189
GM /Im(A) (7.26)
where Im(A) is the image of (7.25). Two solution vectors are conformally equivalent if
and only if they belong to the same orbit under Gp , i.e. if they define the same point
in (7.26).
Remark. The fact that the products of half-densities (7.23) solving Hill’s equation
produce integer densities solving the stabilizer equation (7.7) is reminiscent of the fact
that the “square” of two Killing spinors is a Killing vector. This correspondence is
exactly realized in three-dimensional gravity: eq. (7.7) turns out to coincide with the
Killing equation expressed in terms of a suitable component X of a vector field on
space-time, while Hill’s equation (7.12) corresponds to the Killing spinor equation for
a suitable spinor component (see e.g. eq. (16) in [167]).
Let ψ1 and ψ2 be normalized solutions of Hill’s equation (7.12). They have non-zero
g + (S 1 ), but their ratio
weight under Diff
ψ1 (ϕ)
η(ϕ) ≡ (7.27)
ψ2 (ϕ)
g + (S 1 ) as a function (i.e. a zero-weight density). It blows up at
transforms under Diff
the zeros of ψ2 , so it is more convenient to think of it as a curve
η : R → RP 1 : ϕ 7→ η(ϕ) (7.28)
whose expression is (7.27) in terms of the projective coordinate (6.84). The points
where η diverges are then mapped by η on the “point at infinity” in RP 1 . Since RP 1
is diffeomorphic to the circle (6.83), we can also think of η as a path in S 1 whose
expression in stereographic coordinates is (7.27).
Winding numbers
One can think of η(ϕ) as a path in the circle with a “time parameter” ϕ. Eq. (7.31)
then says that η 0 (ϕ) > 0, so η(ϕ) always spins around the circle in the same direction.
We therefore introduce the following terminology:
Definition. The winding number n ∈ N of η(ϕ) is the number of laps around the
circle performed by η in a “time interval” of length 2π.
We will illustrate the computation of the winding number in the next section, when
describing explicit Virasoro orbit representatives. For now note that η(ϕ) transforms
g + (S 1 ) as a function, so its winding number is conformally invariant:
under Diff
Proposition. Let c 6= 0 and let n(p,c) ∈ N be the winding number of the curve η
associated with the Hill’s equation (7.12) specified by p(ϕ) and c. Then there is a
well-defined map
In particular, Virasoro coadjoint vectors with the same central charge but different
winding numbers do not belong to the same orbit.
Chapter 7. Virasoro orbits 191
This supplements our previous observation (7.20) that the conjugacy classes of
monodromy matrices yield a rough classification of Virasoro coadjoint orbits. In fact,
these two invariants together provide the complete classification of Virasoro orbits.
Indeed one can show that the map that associates a pair ([M], n) with each Virasoro
coadjoint orbit is injective, provided [M] is the conjugacy class of the monodromy
matrix and n is the winding number. Note however that the map is not surjective, as
some pairs ([M], n) do not belong to its image. We now verify this by brute force by
describing orbit representatives.
Each conjugacy class of SL(2, R) is contained in one of these three families, but each
family contains several conjugacy classes. The elliptic and hyperbolic families contain
infinitely many conjugacy classes since they depend on a continuous parameter (the
trace of M). Note that the trace of M determine the properties of its eigenvalues:
M is elliptic ↔ distinct complex eigenvalues;
M is parabolic ↔ degenerate real eigenvalue ±1;
M is hyperbolic ↔ distinct real eigenvalues.
We now determine the conjugacy classes contained in each family. The computations
are very similar to those of section 4.3 where we determined the orbits of momenta for
the Poincaré group in three dimensions.
Proof. In the elliptic family, the eigenvalues of M are complex conjugates of one another
with non-zero imaginary part. Since det(M) = 1, they can be written as e±2πiω where
ω belongs to the open interval ]0, 1[ without loss of generality, but differs from 1/2.
Let v ∈ C2 be an eigenvector of M such that M · v = e2πiω v. This vector is complex
and linearly independent of its complex conjugate v̄; the latter is an eigenvector of M
with eigenvalue e−2πiω . Then v + v̄ and i(v − v̄) are linearly independent real vectors;
we can choose the norm of v in such a way that the (real) matrix S expressing M in
the basis {v + v̄, i(v − v̄)} has unit determinant. Then SMS −1 takes the form (7.33).
The stabilizer consists of all matrices that commute with (7.33) and is readily seen to
consist of rotations.
The stabilizer of the first two matrices is the whole group SL(2, R), while the stabilizer
R × Z2 of the four remaining ones consists of triangular matrices (4.96).
Proof. When M is parabolic, its eigenvalues are either both 1 or both −1. Let λ be
the eigenvalue of M and let v ∈ R2 be a (real) eigenvector of M. Let v 0 be another
vector such that {v, v 0 } is a basis of R2 , and choose the normalization of v and v 0 in
such a way that the matrix S expressing M in this basis has unit determinant. Then
−1 λ x
SMS = (7.35)
0 λ
where λ = ±1 and x is an arbitrary real number. For x = 0 we find the first two
matrices in the list (7.34), each of which is alone in its conjugacy class. For non-zero
x, note that
y 0 1 ±1 1/y 0 1 ±y 2
= , (7.36)
0 1/y 0 1 0 y 0 1
so for λ = +1, M is conjugate to the second matrix in (7.34) if x > 0 and to the
third one if x < 0, in both cases with an overall plus sign. The situation is similar
when λ = −1, but with the minus sign. The proof ends with the observation that all
matrices in (7.34) belong to disjoint conjugacy classes. The stabilizer is obtained by
direct computation.
them so that the matrix S expressing M in the basis {v, v 0 } has unit determinant.
Then SMS −1 takes the form (7.37) with e2πω = λ or e2πω = 1/λ. The ordering of
0 1
eigenvalues can be changed thanks to the SL(2, R) matrix , so we are free to
−1 0
pick ω > 0, and this specifies uniquely the conjugacy class of the matrix M. Finding
the stabilizer is straightforward.
From now on we say that a Virasoro orbit is elliptic, parabolic or hyperbolic if the
associated monodromy matrix is of one of those three types, respectively. In addition
we will distinguish parabolic orbits associated with ±I from parabolic orbits associated
with the four other matrices in (7.34) by referring to the former as “degenerate” and
to the latter as “non-degenerate”.
is strictly positive and 2π-periodic; it is a vector field on the circle, since it is a quadratic
combination of −1/2-densities such as (7.23). In fact, it belongs to the Lie algebra of
the stabilizer of q since it solves equation (7.7). In addition it is invariant under the
action of the stabilizer of M and is therefore a well-defined functional of q(ϕ), which
justifies the notation Xq . Conversely, Xq determines q(ϕ) since Hill’s equation implies
" 0 2 #
00 00 00
c ψ1 ψ1 + ψ2 ψ2 (7.38) c 1 qX 1 X q 1
q= = − − 2 . (7.39)
6 Xq 6 2 Xq 4 Xq Xq
We would have obtained the same formula upon using eq. (7.30) with η = ψ1 /ψ2 . Our
g + (S 1 ) such that q is obtained by acting
goal now is to build a diffeomorphism gq ∈ Diff
with gq on a suitable orbit representative p. In the language of induced representa-
tions, the maps gq will be “standard boosts” on the orbit of p.
where the notation “p0 ” will be justified below. This number is well-defined since
g + (S 1 ) since Xq is a vector field. We
Xq (ϕ) never vanishes, and it is invariant under Diff
can then define a diffeomorphism f ∈ Diffg + (S 1 ) by
Z ϕ
2π dφ
f (ϕ) ≡ p . (7.41)
6|p0 |/c 0 Xq (φ)
This quantity is the inverse of the sought-for standard boost since eq. (7.39) can be
written as
c
q(ϕ) = p0 (f 0 (ϕ))2 − S[f ](ϕ) , (7.42)
12
which we recognize as the coadjoint action (6.113) of
gq ≡ f −1 (7.43)
Proposition. Let (q, c) with c > 0 be a Virasoro coadjoint vector with elliptic
monodromy. Then it belongs to the orbit of a constant coadjoint vector (p, c) with
p(ϕ) = p0 , where the value of p0 is determined by q(ϕ) according to (7.40) with Xq
given by (7.38) in terms of normalized solutions of the Hill’s equation of (q, c). In
addition, the diffeomorphism gq defined as the inverse of (7.41) is a standard boost for
the orbit of p in the sense that
gq · p = q (7.44)
where the dot denotes the coadjoint action (6.114).
(7.9) will turn out to have degenerate parabolic monodromy (see below).
Thus different values of p0 generally define disjoint orbits since their monodromy
matrices (7.33) are not conjugate. However, at this stage we cannot tell whether
p q 2
c
p0 and − |p0 | + 6
N (7.49)
belong to different orbits when N ∈ N since their monodromy matrices coincide (their
angles differ by 2πN ). This issue is settled by the winding number (7.32): the curve
(7.27) associated with the solutions (7.47) is
In conclusion, the orbits of two generic constants p0 and p̃0 are disjoint if and only
if these constants differ. We have thus recovered the lower part (p0 < 0) of fig. 7.1. As
a bonus we can now assign a monodromy matrix determined by (7.46), and a winding
number (7.51), with each point on that part. In particular the integers n written on
the left of the p0 axis can be interpreted as winding numbers for constants p0 located
between −(n + 1)2 c/24 and −n2 c/24.
Stabilizers
To conclude the description of orbits of generic constants p0 < 0, it remains to find
their stabilizer. As anticipated in (7.8), one shows that the stabilizer of p0 is the group
U(1) of rigid rotations f (ϕ) = ϕ + θ (or more precisely its universal cover R when
g + (S 1 )). The coadjoint orbit of (p0 , c) can thus be written as
dealing with Diff
Remark. The stabilizer U(1) coincides with the stabilizer of the monodromy matrix
(7.33), so the quotient (7.26) consists of a single point. This implies that all conformally
inequivalent normalized solutions of the Hill equation associated with (p0 , c) can be
obtained by acting with rotations on the solution (7.47).
As in the elliptic case we can choose constant coadjoint vectors as orbit representa-
tives. The corresponding Hill’s equation then reads (7.45) and admits the normalized
solutions (7.47), but the monodromy matrix is ±I. Such a monodromy matrix M only
occurs when p0 takes the exceptional form (7.9) for some strictly positive integer n, in
which case
n2 c
p0 = − and M = (−1)n I. (7.53)
24
By contrast, elliptic orbits never contain an exceptional constant; this is a sharp dif-
ference between elliptic and degenerate parabolic orbits.
The monodromy matrix (7.53) implies that two exceptional constants specified by
integers n, n0 can belong to the same orbit only if n and n0 have the same parity; but
at this stage we cannot tell if two orbits with the same parity are disjoint. As in the
elliptic case we can address this question by studying the winding number of the curve
(7.27) associated with the solutions (7.47). One can verify that the winding number
coincides with the number n specified by p0 = −n2 c/24, which implies that any two
orbits of exceptional constants specified by different values of n > 0 are disjoint. In
conclusion, we have now recovered the dots in the lower part of fig. 7.1, and the values
of n displayed there coincide with winding numbers. In particular the orbit at n = 1
will be called the vacuum orbit from now on; in the context of Riemann surfaces, it is
known as universal Teichmüller space [56, 168]. Note that the orbit of p0 = 0 does not
have degenerate parabolic monodromy, and so has not yet been accounted for by our
classification of Hill’s equations.
Stabilizers
We now study the stabilizers of orbits of exceptional constants p0 = −n2 c/24. We saw
below (7.8) that the stabilizer is three-dimensional for such values, and is generated
Chapter 7. Virasoro orbits 197
The Lie algebra of the stabilizer is therefore isomorphic to sl(2, R), but different values
of n define non-conjugate embeddings of sl(2, R) in Vect(S 1 ). In fact one can verify
using (6.96) that the finite diffeomorphisms that span the stabilizer of p0 = −n2 c/24
(and that reduce to (7.54) close to the identity) are projective transformations (6.95)
spanning a group PSL(n) (2, R) (the n-fold cover of PSL(2, R)). In conclusion:
Lemma. The stabilizer of p0 = −n2 c/24 for the coadjoint action of Diff g + (S 1 ) (resp.
Diff+ (S 1 )) is the group PSL
g (n) (2, R) (resp. PSL(n) (2, R)) spanned by diffeomorphisms
g (n) (2, R) is the universal cover of the n-fold cover of
f (ϕ) given by (6.95), where PSL
PSL(2, R) = SL(2, R)/Z2 . The coadjoint orbit of (p0 , c) can be written as
W 2
∼
= Diff g (n) (2, R) ∼
g + (S 1 )/PSL = Diff+ (S 1 )/PSL(n) (2, R) (7.55)
− n24c ,c
The Lie algebra of the stabilizer is generated by the vector fields (7.54).
In section 9.3 we will interpret the orbit of p0 = −c/24 as the set of gravitational
perturbations around Minkowski space. In that context the little group PSL(2, R) will
be seen as the Lorentz group in three dimensions. The remaining exceptional values
p0 = −n2 c/24 (with n ≥ 2) will be interpreted as conical excesses where one turn
around the origin of space spans an angle 2πn.
such that ψ2 has no zeros on the real line. Then η(ϕ) is smooth and, by virtue of
(7.37), we have
η(ϕ + 2π) = e4πω η(ϕ) (7.56)
so η(ϕ) is monotonically increasing on R (since ω > 0). As in the case of elliptic orbits,
our goal is to find a “standard boost” gq whose inverse gq−1 ≡ f will map q(ϕ) on a
suitably chosen orbit representative. To do so we define
1 (7.27) 1 ψ1 (ϕ)
f (ϕ) ≡ log(η(ϕ)) = log (7.57)
2ω 2ω ψ2 (ϕ)
g + (S 1 ) by virtue of (7.56). Now if we set
which belongs to Diff
c ω2
p0 ≡ , (7.58)
6
we can use (7.30) and the cocycle identity (6.77) to write
c
q(ϕ) = p0 (f 0 (ϕ))2 − S[f ] . (7.59)
12
As in (7.42) we recognize the coadjoint action of gq = f −1 on the constant coadjoint
vector p(ϕ) = p0 > 0, and thus conclude:
Proposition. Let (q, c) with c > 0 be a Virasoro coadjoint vector with hyperbolic
monodromy and zero winding number. Then it belongs to the orbit of a constant
coadjoint vector (p0 , c), where p0 > 0 is determined by the monodromy matrix accord-
ing to (7.58). In addition the diffeomorphism gq defined as the inverse of (7.57) is a
standard boost for the orbit of p in the sense (7.44).
Note that the definition (7.58) coincides with eq. (7.46) for p0 > 0. Roughly
speaking, “hyperbolic orbits are an analytic continuation of elliptic orbits to imaginary
values of the monodromy parameter ω”. This is analogous to the fact that tachyonic
momentum orbits may be seen as massive orbits with imaginary mass.
Stabilizers
At p = p0 , Hill’s equation (7.12) reads − 6c ψ 00 + |p0 |ψ = 0 where we stress that the sign
of the potential term is opposite to the one in (7.45). A basis of solutions satisfying
the Wronskian condition (7.16) is provided by
1 1
ψ1± (ϕ) = ± √ eωϕ , ψ2± (ϕ) = ± √ e−ωϕ (7.60)
2ω 2ω
where ω > 0 is given by (7.46). The corresponding monodromy matrix is (7.37).
We have seen in (7.8) that the stabilizer is one-dimensional for p0 > 0, and that
it consists of rotations of the circle. Thus the stabilizer of p0 is a group U(1) of rigid
rotations (or more precisely its universal cover R when dealing with Diff g + (S 1 )). In
particular the orbit of (p0 , c) for p0 > 0 and c > 0 is diffeomorphic to
In terms of fig. 7.1, we have now completed our understanding of almost the whole
real line p0 ∈ R, since we now know that the orbits that pass through p0 > 0 are of
hyperbolic type without winding. The only remaining mystery is the orbit of p0 = 0,
and of course all the orbits that do not contain constant representatives.
c ψ100 c ψ200
p≡ = (7.62)
6 ψ1 6 ψ2
is smooth for any constant c > 0. If in addition there exists a monodromy matrix M
such that (7.18) holds, then p(ϕ) is 2π-periodic and ψ1 , ψ2 are solutions of the corre-
sponding Hill’s equation. This procedure provides a way to build Virasoro coadjoint
vectors out of functions ψi ; in particular, in order to prove that there exist Virasoro
orbits with hyperbolic monodromy and non-zero winding number, it suffices to find
two normalized functions ψi satisfying these criteria, and the identification of the cor-
responding Virasoro coadjoint vectors will follow.
Thus, let ω > 0 be a strictly positive real number and let n > 0 be a positive
integer. Let us define the positive function
2
2 2ω
Fn,ω (ϕ) ≡ cos (nϕ/2) + sin(nϕ/2) + cos(nϕ/2) (7.63)
n
as well as
r
eωϕ 2 ω
ψ1 (ϕ) ≡ p sin(nϕ/2) + cos(nϕ/2) , (7.64)
Fn,ω (ϕ) n n
r
e−ωϕ 2
ψ2 (ϕ) ≡ p cos(nϕ/2). (7.65)
Fn,ω (ϕ) n
Chapter 7. Virasoro orbits 200
Since Fn,ω is strictly positive, the ψi ’s are smooth functions. They satisfy the Wron-
skian condition (7.16) and their monodromy matrix is (7.37). Their ratio is
ω
η(ϕ) = e2ωϕ tan(nϕ/2) +
n
and describes a path on the circle with varying velocity and winding number n. It
follows that the function p(ϕ) defined by (7.62) is a Virasoro coadjoint vector with
hyperbolic monodromy (7.37) and winding number n > 0. It is explicitly given by
c ω2 c n2 + 4ω 2 c n2
p(ϕ) = + − 2 (ϕ)
(7.66)
6 12 Fn,ω (ϕ) 8 Fn,ω
in terms of the function (7.63). We have thus built explicit orbit representatives with
hyperbolic monodromy and non-zero winding number.
It is worth spending some time to interpret formula (7.66). Let us take ω small
and expand p around ω = 0. To first order in ω, we get
n2 c nc
p(ϕ) = − + ω sin(nϕ) + O(ω 2 ) . (7.67)
24 3
The leading term in p is an exceptional constant −n2 c/24, so we can think of (7.66) as
a deformation of that constant. The term of order one in ω in (7.67) is proportional to
sin(nϕ), which is one of the elements of the Lie algebra of the stabilizer of −n2 c/24.
This ensures that the deformation does not belong to the orbit of −n2 c/24. Indeed,
all deformations that do belong to that orbit take the form
2
∗ n c (6.115) c
n2 X 0 + X 000
adX −
c = −
24 12
for some vector field X, where the term n2 X 0 + X 000 annihilates the contribution of the
modes sin(nϕ) or cos(nϕ).
Figure 7.2: A partial map of Virasoro orbits, including orbits of constant coadjoint
vectors together with tachyonic orbits. Compare to fig. 7.1.
this stage that any coadjoint vector satisfying these properties can be mapped
on (7.66). However, this turns out to be the case; in this sense, the orbit rep-
resentatives (7.66) exhaust all orbits with hyperbolic monodromy and non-zero
winding. See [166] for the proof.
• The Lie algebra of the stabilizer of (7.66) is spanned by the periodic linear com-
binations of the functions (7.23). As it turns out, the only periodic combination
in this case is the product ψ1 ψ2 . The latter has 2n simple zeros inside [0, 2π[
and generates a non-compact group R. In addition the function (7.66) is peri-
odic with period 2π/n, so the stabilizer must contain a group Zn consisting of
rotations by integer multiples of 2π/n. In fact, one can show (see [166]) that the
stabilizer of p in Diff+ (S 1 ) is isomorphic to a product R × Zn , while its stabilizer
g + (S 1 ) is R × T2π/n where T2π/n is the group of trans-
in the universal cover Diff
lations of the real line by integer multiples of 2π/n. We conclude that the orbit
of (7.66) is diffeomorphic to
Zero winding
At zero winding we proceed as in the elliptic and n = 0 hyperbolic cases, i.e. we
look for standard boosts. Let therefore (q, c) be a Virasoro coadjoint vector such that
a normalized solution vector Ψ = (ψ1 ψ2 )t associated with the corresponding Hill’s
equation has non-degenerate parabolic monodromy and zero winding number. The
monodromy matrices in (7.34) imply that
ψ1 (ϕ + 2π) = ± ψ1 (ϕ) + εψ2 (ϕ) , ψ2 (ϕ + 2π) = ±ψ2 (ϕ) (7.69)
and (7.31) implies that ε must actually be equal to +1. The opposite sign corresponds
to changing the orientation in the space of solutions of Hill’s equation, so with our
choice of orientation for ψ1 , ψ2 , only the value ε = +1 gives rise to an admissible
monodromy matrix. Then the function
is a 2πZ-equivariant diffeomorphism of the real line, and property (7.30) implies that
c
q(ϕ) = − S[f ](ϕ) .
12
As in eqs. (7.42) and (7.59), we recognize the coadjoint action of gq ≡ f −1 :
Proposition. Let (q, c) with c > 0 be a Virasoro coadjoint vector with non-dege-
nerate parabolic monodromy and vanishing winding number. Then it belongs to the
orbit of (0, c) and the inverse of the diffeomorphism (7.71) is a standard boost in the
sense of eq. (7.44).
Thus we have finally found the orbit of p0 = 0 ! It was the only point of fig. 7.1 that
was still eluding us. We now know that its orbit has parabolic type. The corresponding
stabilizer is the group U(1) of rigid rotations (as for all positive or generic constants
p0 ), and there are two conformally inequivalent solutions of Hill’s equation at p0 = 0,
namely ψ1± = ±ϕ, ψ2± (ϕ) = ±1. The orbit can be represented as a quotient space
W(0,c) ∼ g + (S 1 )/R ∼
= Diff = Diff+ (S 1 )/S 1
Non-zero winding
At non-zero winding our strategy will be similar to that used in the hyperbolic case
with winding: we rely on the fact that formula (7.62) always defines a 2π-periodic func-
tion p(ϕ) when ψ1 and ψ2 satisfy the Wronskian condition and admit a well-defined
monodromy, which allows us to build orbit representatives.
Thus, pick a number ε ∈ {±1} and let n ∈ N∗ be a non-zero winding number. Let
us define the positive function
ε
Hn,ε (ϕ) ≡ 1 + sin2 (nϕ/2) (7.72)
2π
as well as
1 εϕ 2
ψ1 (ϕ) ≡ p sin(nϕ/2) − cos(nϕ/2) , (7.73)
Hn,ε (ϕ) 2π n
1
ψ2 (ϕ) ≡ p sin(nϕ/2) . (7.74)
Hn,ε (ϕ)
Since the function Hn,ε is strictly positive, the ψi ’s are smooth functions. They satisfy
the Wronskian condition (7.16) and their monodromy matrix is one of the four matrices
on the right in the list (7.34), with the off-diagonal entry coinciding with ε and the
overall ±1 = (−1)n . The curve (7.27) corresponding to this basis of solutions is
εϕ 2
η(ϕ) = − cot(nϕ/2)
2π n
and has winding number n. This is all as in the hyperbolic case below eq. (7.65). It
follows that the function p(ϕ) defined by (7.62) is a Virasoro coadjoint vector with
non-degenerate parabolic monodromy and winding number n > 0, explicitly given by
c n2 c n2 (1 + ε/2π)
p(ϕ) = − 2 (ϕ)
. (7.75)
12 Hn,ε (ϕ) 8 Hn,ε
are conjugate in SL(2, R) for any positive real number λ. Accordingly we could just
as well have chosen the representatives of non-degenerate parabolic conjugacy classes
to involve an arbitrary positive parameter ε; the limit ε → 0 then may be taken since
it does not affect the conjugacy class of the monodromy matrix. The corresponding
coadjoint vector is (7.75) and its expansion to first order in ε reads
n2 c ε
p(ϕ) = − 1+ (1 + 2 cos ϕ) + O(ε2 ). (7.77)
24 2π
Chapter 7. Virasoro orbits 204
As in (7.67), the leading term is an exceptional constant (7.9) and we can think of
(7.77) as a deformation thereof. The deformation is designed so that it does not belong
to the orbit of −n2 c/24. When dealing with BMS3 supermomentum orbits in section
10.1, we will interpret (7.75) as the supermomentum of a massless BMS3 particle.
Accordingly, from now on we refer to non-degenerate parabolic orbits with non-zero
winding as massless orbits. Note that the statement that the matrices (7.76) are con-
jugate is tantamount to saying that massless orbits are scale-invariant.
To conclude our analysis we state (without proof) a few features of massless orbits:
• One can show that the orbit representatives (7.75) are exhaustive in that any
coadjoint vector belonging to a massless orbit can be brought in that form by a
suitable diffeomorphism. See appendix C of [166].
• The Lie algebra of the stabilizer of (7.75) is generated by the vector field X = ψ22 ,
which has n double zeros. In fact, as in the hyperbolic case, the stabilizer is
isomorphic to R × Zn , but the generator of the R part of that group is not
the same as in the hyperbolic case. The orbit is diffeomorphic to a quotient of
Diff+ (S 1 ) by this stabilizer, or equivalently a quotient of Diff
g + (S 1 ) by R × T2π/n
where T2π/n is the same discrete translation group as in (7.68).
g + (S 1 ) transformations, the solution (7.73)-(7.74) is the unique solution
• Up to Diff
of Hill’s equation with non-degenerate parabolic monodromy
n 1 ε
(−1) (7.78)
0 1
The schematic drawings of figs. 7.1 and 7.2 represent Virasoro orbits. The only
orbits which are not accounted for by these pictures are massless ones; in order to in-
clude them we use the same trick as in fig. 4.3b, where massless orbits are represented
by two dots near the origin (one with positive energy, the other with negative energy).
We will use the same notation here, except that such a pair of massless orbits occurs
for all positive integers n ∈ N∗ . With this convention, fig. 7.2 turns into the complete
map of Virasoro coadjoint orbits displayed in fig. 7.3.
Each point in that map represents an orbit representative; different points corre-
spond to different representatives and define disjoint orbits. All orbits are now ac-
counted for since the orbit representatives are exhaustive. The vertical line represents
orbits that contain a constant orbit representative:
Chapter 7. Virasoro orbits 205
Figure 7.3: The map of Virasoro coadjoint orbits at positive central charge. Note the
similarity with fig. 4.3b. Roughly speaking, the map consists of an infinity of copies
of Poincaré momentum orbits glued together and labelled by the winding number n.
Locally (near a node n), the two pictures look identical. This is not surprising given
that Poincaré momentum orbits in three dimensions coincide with SL(2, R) coadjoint
orbits, which in turn are classified similarly to the conjugacy classes of SL(2, R) that
were instrumental for Virasoro coadjoint orbits. This hints that there exists a relation
between Virasoro and Poincaré symmetry; we shall see in part III that this relation is
embodied by the BMS3 group.
• For generic p0 < 0 the orbit has elliptic monodromy determined by eq. (7.46).
Its winding number is given by (7.51), so the points of fig. 7.3 located between n
and n + 1 have winding number n (while points such that −c/24 < p0 < 0 have
zero winding number).
• For exceptional values p0 = −n2 c/24 with n ∈ N∗ , the orbit has degenerate
parabolic monodromy determined by (7.53). Its winding number is n. In partic-
ular, the orbit at n = 1 is the vacuum orbit.
• For p0 > 0, the orbit has hyperbolic monodromy with zero winding, and the
conjugacy class of the monodromy matrix is determined by (7.46).
On the other hand, the points of fig. 7.3 that do not belong to the vertical axis represent
orbits that do not contain any constant representative:
• Each pair of dots surrounding a tachyonic line at n represents the two massless
orbits with winding number n. The orbit representatives are given by (7.75)
and involve a discrete parameter ε = ±1 that determines the corresponding
monodromy matrix (7.78).
Focussing for definiteness on the multiply connected group Diff+ (S 1 ), the stabilizers
of Virasoro orbits are as follows:
Orbit Stabilizer
Vacuum-like p0 = −n2 c/24, n ≥ 1 PSL(n) (2, R)
Elliptic U(1)
Hyperbolic, zero winding U(1)
Non-degenerate parabolic, zero winding U(1)
Massless, winding n ≥ 1 R × Zn
Tachyonic, winding n ≥ 1 R × Zn
Table 7.1: Virasoro coadjoint orbits and their stabilizers.
In the universal cover of the Virasoro group the first four entries of the right column
would be replaced by their universal covers, while the two last ones would be replaced
by R×T2π/n . This should be compared with (and is very similar to) the list of Poincaré
little groups in table 4.1. Note that, at n = 1, the Virasoro stabilizers are quotients
by Z2 of their Poincaré counterparts. This is because table 4.1 lists the little groups
given by the double cover (4.93) of the Poincaré group.
Remark. Fig. 7.3 may be misleading since it suggests that all Virasoro orbits of
constant coadjoint vectors are of a similar type, which is clearly not the case since
orbits of constants p0 > 0 are hyperbolic while those of (generic) constants p0 < 0
are elliptic. In this sense, the map of orbits would have been more accurate if we had
represented the orbits of p0 > 0 by a horizontal line to suggest that they have the same
type of monodromy as the tachyonic orbits; see e.g. fig. 1 of [166]. Our convention
in fig. 7.3 is motivated instead by the fact that the value of p0 essentially measures
energy (see below), so that higher points in fig. 7.3 have higher energy.
the Virasoro energy functional, before showing that the Schwarzian derivative satisfies
an “average lemma” which will play a key role for this functional’s boundedness. We
then show that the only orbits with energy bounded from below are either orbits of
constants p0 ≥ −c/24, or the massless orbit at winding n = 1 and monodromy ε = −1.
To reduce clutter we return to our earlier abusive notation by writing as Diff(S 1 ) the
universal cover of the group of orientation-preserving diffeomorphisms of the circle.
Relevant references include [56, 166] as usual.
Now consider a CFT with central charge c > 0 and let Ω be the space of its
stress tensors p(ϕ); in general Ω is a certain subset of the space F2 (S 1 ) of quadratic
densities. Since any quantum system with a well-defined vacuum is expected to have
energy bounded from below, the map
Ω → R : p 7→ E[p] (7.81)
should be bounded from below. In addition, consistency with conformal symmetry
implies that Ω is a union of Virasoro coadjoint orbits. One is thus led to the following
question:
Which of the Virasoro coadjoint orbits of fig. 7.3 have
(7.82)
energy bounded from below under conformal transformations?
In the sequel we will refer to orbits with energy bounded from below as orbits “with
positive energy”, although their energy (7.79) may actually be negative for some field
configurations p(ϕ).
Note that all orbits have energy unbounded from above. Indeed the term involving
the Schwarzian derivative in (7.80) can be written as
Z 2π Z 2π
c dϕ c
S[f ](ϕ) = − dϕ S[f −1 ](ϕ) (7.83)
24π 0 f 0 (ϕ) 24π 0
Chapter 7. Virasoro orbits 208
where we have renamed the integration variable from ϕ to f −1 (ϕ), then used (6.16)
and the cocycle identity (6.77). Since the Schwarzian derivative can be written as
00 0 2
f 1 f 00
S[f ](ϕ) = − , (7.84)
f0 2 f0
we can also recast (7.83) in the form
Z 2π −1 00 2
c (f )
dϕ .
48π 0 (f −1 )0
This can be made arbitrarily large for suitable choices of f , which proves that the
energy functional E is unbounded from above on any Virasoro orbit.
Average lemma. Let f ∈ Diff g + (S 1 ) and let S[f ](ϕ) be its Schwarzian derivative
(6.76) at ϕ. Then the average of the Schwarzian derivative satisfies the inequality
Z 2π Z 2π
1
dϕ 1 − (f 0 (ϕ))2 ,
dϕ S[f ](ϕ) ≤ (7.85)
0 0 2
with equality if and only if f (ϕ) is a projective transformation of the form (6.88).
Proof. We consider the functional
Z 2π
1 0 2
I[f ] ≡ − dϕ (f (ϕ)) + S[f ](ϕ) . (7.86)
0 2
Our goal is to show that this quantity is bounded from below and that its minimum
value is −π. By (7.84), it only depends on f 0 and f 00 . A convenient way to express
this dependence is to define
(6.16) 1
Y (ϕ) ≡ f 0 (f −1 (ϕ)) = . (7.87)
(f −1 )0 (ϕ)
Since f is a 2πZ-equivariant, orientation-preserving diffeomorphism, Y (ϕ) is strictly
positive and 2π-periodic. In terms of Y we can rewrite (7.86) as
1 2π
0
(Y (ϕ))2
Z
I[Y ] = dϕ − Y (ϕ) (7.88)
2 0 Y (ϕ)
where the integrand is well-defined since Y > 0. Let us denote the minimum and
maximum of Y (ϕ) by
is non-negative and vanishes only at the points where Y reaches its minimum or its
maximum. Now consider the obvious inequality
r !2
|Y 0 | mM
√ − m+M −Y − ≥ 0. (7.90)
Y Y
where we have introduced the function G(z) to reduce clutter below. To see the
use of this, consider a function Y (ϕ) of the following shape (the general case follows
straightforwardly):
Figure 7.4: The function Y (ϕ) is 2π-periodic and strictly positive. Here we choose
it with four local extrema, the global minimum being Y (ϕ1 ) = m and the global
maximum Y (ϕ4 ) = M .
This function has two local minima at ϕ1 and ϕ3 and two local maxima at ϕ2
and ϕ4 (the numbers of local minima and maxima coincide since Y (ϕ) is smooth and
Chapter 7. Virasoro orbits 210
We conclude that I[Y ] is bounded from below by the value −π, which is exactly the
inequality (7.85). It only remains to find the conditions under which (7.85) becomes
an equality. For this to be the case, the inequalities (7.90), (7.93) and (7.94) must all
be saturated; this occurs when Y (ϕ) satisfies the following three conditions:
• In order to saturate (7.90), it satisfies the differential equation
Y 02 = (m + M )Y − Y 2 − mM. (7.95)
• In order to saturate (7.93), Y (ϕ) has only one minimum and one maximum,
where it takes the values m and M , respectively.
• In order to saturate the second inequality of (7.94), M = 1/m.
To solve (7.95) we use (7.87) and rewrite the equation in terms of f −1 . Using M = 1/m
the derivative of (7.95) becomes
Y 0 1 − ((f −1 )0 )2 − 2 S[f −1 ] = 0,
(7.96)
which is equivalent to (6.94). We have shown below (7.54) that the only f ’s satisfying
this property are those that belong to the group of projective transformations (6.88),
which concludes the proof.
Proposition. The vacuum orbit, containing the point pvac = −c/24, has energy
bounded from below:
c
E[f · pvac ] ≥ E[pvac ] = − . (7.97)
24
The minimum of energy is located at pvac .
Proof. We consider formula (7.80) with p(ϕ) = pvac = −c/24. Renaming the integra-
tion variable from ϕ to f −1 (ϕ) and using eqs. (6.16) and (6.77), we find
Z 2π
c 1 −1 0
2 −1
E[f · pvac ] = dϕ − (f ) (ϕ) − S[f ](ϕ)
24π 0 2
which we recognize as the functional (7.86) evaluated at f −1 . The average lemma (7.85)
then implies that E[f · pvac ] ≥ −c/24, with equality if and only if f is a projective
transformation (6.88). Our earlier result (7.55) ensures that such transformations
precisely span the stabilizer of pvac , so the minimum of energy is reached at pvac .
Let us turn to other orbits containing a constant representative p(ϕ) = p0 . The
key will be to rewrite their energy functional as the vacuum energy functional, plus
another term. Starting from formula (7.80) we obtain
p0 + c/24 2π dϕ
Z
E[f · p0 ] = + E[f · pvac ] (7.98)
2π 0 f 0 (ϕ)
Proposition. If p0 ≥ −c/24, then the orbit of (p0 , c) has energy bounded from
below, with the energy minimum located at p0 :
c
p0 ≥ − ⇒ E[f · p0 ] ≥ E[p0 ] = p0 .
24
Chapter 7. Virasoro orbits 212
Now what happens when p0 is lower than −c/24? In that case energy is un-
bounded, as can be shown by finding a family of diffeomorphisms that lower the energy
indefinitely. Indeed, consider the matrix
cosh(γ/2) sinh(γ/2)
∈ SL(2, R) (7.99)
sinh(γ/2) cosh(γ/2)
where we have used the fact that the integral of f 0 over S 1 is normalized to 2π. The
integral of (7.101) then yields
c
E[f · p0 ] = (p0 + c/24) cosh γ − , (7.102)
24
Chapter 7. Virasoro orbits 213
and this can become arbitrarily negative when p0 < −c/24. In conclusion:
Thus, when p0 < −c/24, fig. 7.5 is no longer valid because the energy functional can
reach arbitrarily low values in certain directions. The orbit then looks like an infinite-
dimensional saddle instead of the hyperboloid represented in fig. 7.5.
Note that the matrix (7.99) can be interpreted as the SL(2, R) group element
that represents a Lorentz boost with rapidity γ in three dimensions3 thanks to the
isomorphism (4.83), which also explains our choice of normalization. In that context,
formula (7.102) is the transformation law of the energy of a particle with mass p0 +c/24
under Lorentz boosts. We will return to this interpretation in part III.
We will not prove this proposition here and refer instead to [166]. Roughly speak-
ing, the proof follows from a construction very similar to the one used in the proof
of the average lemma (7.85), except that it crucially relies on the parameters n = 1,
ε = −1. The proof of the fact that the infimum of energy is never reached on the
orbit follows from the construction of a one-parameter family of points belonging to
the orbit in such a way that they converge to the constant −c/24 without ever quite
reaching it.
3
Rapidity is related to velocity v by γ = arctanh(v).
Chapter 7. Virasoro orbits 214
One might think that the other orbits without constant representatives behave
in a similar way, i.e. that they also have energy bounded from below. However, for
any such orbit, it is possible to build a one-parameter family of orbit elements whose
energy can be arbitrarily low, similarly to constant representatives p0 < −c/24. We
refer again to [166] for explicit constructions. Thus one concludes that
Figure 7.7: The map of Virasoro coadjoint orbits at positive central charge. Orbits
with energy bounded from below are coloured in red. Those are orbits of constants
p0 ≥ −c/24, plus the unique massless orbit with monodromy (7.78) such that ε = −1
and winding number n = 1. All other orbits have energy unbounded from below.
Chapter 8
In this chapter we explore a physical model where the Virasoro group plays a key role,
namely three-dimensional gravity on Anti-de Sitter (AdS) backgrounds and its puta-
tive dual two-dimensional conformal field theory (CFT). These considerations will be
a basis and a guide for our study of asymptotically flat space-times in part III.
The plan is the following. Section 8.1 is a prelude where we recall a few basic
facts about (three-dimensional) gravity, in particular regarding the notion of asymp-
totic symmetries. Section 8.2 is then devoted to three-dimensional space-times whose
metric approaches that of Anti-de Sitter space at spatial infinity; this includes Brown-
Henneaux boundary conditions and their asymptotic symmetries, which will turn out
to consist of two copies of the Virasoro group. In section 8.3 we describe the phase
space of AdS3 gravity as a hyperplane at fixed central charges in the space of the
coadjoint representation of two Virasoro groups. Finally, in section 8.4 we describe
unitary highest-weight representations of the Virasoro algebra and relate them to the
quantization of the AdS3 phase space.
217
Chapter 8. Symmetries of gravity in AdS3 218
freedom, then turn to a discussion of boundary conditions and the ensuing boundary
terms that one adds to the action in order to make the variational principle well-
defined. This finally leads to the concept of asymptotic symmetries and the important
observation that the Poisson brackets of surface charges that generate these symmetries
generally contain central extensions.
Upon varying the action (8.1) and neglecting all boundary terms (which we shall
talk about later), one obtains the vacuum Einstein’s equations with a cosmological
constant:
1
Rµν − Rgµν + Λgµν = 0 , i.e. Rµν = 2Λgµν . (8.2)
2
What is special about three-dimensional manifolds is that their Ricci curvature wholly
determines their Riemann tensor independently of the equations of motion:
1
Rλµνρ = gλν Rµρ − gλρ Rµν − gµν Rλρ + gµρ Rλν − R(gλν gµρ − gλρ gµν ) . (8.3)
2
Then the Einstein equations (8.2) imply that, at each point of space-time, the on-shell
Riemann tensor is that of a maximally symmetric manifold with curvature determined
by the cosmological constant:
space. Importantly, this is not to say that the only solution of three-dimensional
gravity is empty space. For example, any quotient of Minkowski space by some dis-
crete group solves Einstein’s equations, but is not globally isometric to Minkowski.
Thus global aspects are essential: even though all solutions of Einstein’s equations
are locally isometric, they are generally not globally isometric and therefore represent
physically distinct field configurations. In this sense the absence of local degrees of
freedom in three-dimensional gravity does not prevent the overall absence of degrees
of freedom: it only means that the actual, physical degrees of freedom of the theory
cannot be captured by a local analysis, but require instead a global one, taking into
account topological properties of the space-time manifold. Field theories of this type,
having no local degrees of freedom but still globally non-trivial, are called topological
field theories.
Note that the absence of local degrees of freedom is confirmed by the Hamiltonian
formalism [185]: picking a time direction in M, one can split the metric field into a
lapse N , a shift N i and a spatial metric gij with conjugate momenta π ij , the indices
i, j ∈ {1, 2} labelling spatial directions. The lapse and shift play the role of Lagrange
multipliers enforcing the constraints that generate reparameterizations of time and
spatial diffeomorphisms, respectively. One thus obtains three dynamical Lagrange
variables gij with three conjugate momenta π ij , subject to three first-class constraints.
These constraints can be solved by choosing three gauge-fixing conditions (this is
the statement that “first-class constraints count twice”), which reduces the number of
physical degrees of freedom of three-dimensional Einstein gravity to 21 (3×2−3−3) = 0,
as expected.
Φ(r, x) = O(r# ) as r → +∞
Chapter 8. Symmetries of gravity in AdS3 220
where Φ is some field and the coefficient # depends on the choice of fall-off conditions.
In writing this it is understood that ∂r Φ is of order O(r#−1 ) at infinity.2
The influence of fall-offs is visible at the level of the action principle. Indeed, it
is understood that the action of the theory should be plugged in an exponential eiS ,
which is then to be integrated over field configurations in a path integral so as to
produce quantum-mechanical transition amplitudes. In the classical limit, the leading
contribution to the path integral should be due to on-shell field configurations; but
for this to be true the integrand must be differentiable, which is to say that the
functional derivative δS/δΦ(x) is a local quantity. This, in turn, is only true provided
the variation of the action contains no boundary terms. For instance, the variation of
the Einstein-Hilbert action (8.1) is given by
√
Z
1 1
δSEH = d3 x −g Rµν − Rgµν + Λgµν δg µν
16πG M 2
(8.5)
√ √
Z
1 3 µν α µα λ
+ d x ∂α −gg δΓµν − −gg δΓλµ .
16πG M
The first term of this expression is the integral of the variation of the metric multiplying
the vacuum Einstein equations, as expected. The second term is the integral of a total
divergence and is therefore equal, by Stokes’ theorem, to the flux of a vector field
through the boundary ∂M of M. Depending on one’s choice of fall-off conditions for
the metric, this boundary term may or may not vanish. If it does vanish, then the
pure bulk action (8.1) can be legally plugged into a path integral. If it does not, then
(8.1) is not differentiable and cannot be inserted as such in a path integral, which is
to say that the semi-classical limit of a path integral involving only the action (8.1)
is not given by on-shell field configurations. Accordingly, in order for the theory to
have a well-defined semi-classical limit given by the equations of motion (8.2), one is
generally forced to modify the pure bulk action (8.1) as
Z
S[gµν ] = SEH [gµν ] + d2 x L(gµν , ∂gµν , ...) . (8.6)
∂M
This explains, in terms of the action, how boundary conditions affect the definition
of the theory. A few remarks are in order:
A naive guess is to simply apply the Noether procedure. For a field theory which
is left invariant by certain symmetry transformations generated by some parameters
a , a = 1, ..., N , with field and space-time transformations of the general form
x 7→ x + δ x, Φ 7→ Φ + δ Φ ,
the N Noether currents jaµ can be obtained by “gauging” the symmetry, that is, replac-
ing the rigid parameters a by arbitrary functions a (x) on space-time. The variation
of the action then takes the form
Z
δS = − dD x jaµ ∂µ a (8.7)
M
from which one can read off the definition of the currents jaµ . Their conservation follows
from the fact that δS ≈ 0 on-shell, and the corresponding conserved Noether charges
are the fluxes of these currents through a space-like slice Σ of space-time:
Z
Qa = (dD−1 x)µ jaµ , (8.8)
Σ
where (dD−1 x)µ is proportional to µα1 ...αD−1 dxα1 ...dxαD−1 . Equivalently, (dD−1 x)µ ∝
dD−1 x · nµ where nµ is the future-pointing time-like unit vector field orthogonal to Σ,
and indices are moved thanks to the space-time metric.
Figure 8.1: A space-time manifold M with an embedded space-like slice Σ and future-
pointing time-like normal vector n.
The problem with gauge symmetries now becomes apparent. Indeed, in that case
the symmetry parameters a are already gauged, which is to say that the right-hand
side of (8.7) vanishes. This in turn implies that the Noether currents associated with
gauge transformations all vanish! In particular, there seems to be no way of defining
conserved charges of the form (8.8) for a gauge symmetry; this problem is the key
difference between gauge symmetries and rigid symmetries.
The solution is provided by the following observation: the Noether current defined
by (8.7) is not unique, as one can add to it the divergence of a two-form without
affecting the left-hand side. In other words eq. (8.7) does not specify the Noether
Chapter 8. Symmetries of gravity in AdS3 223
current jaµ uniquely, since the modified current j̃aµ = jaµ + ∂ν kaµν , where kaµν = −kaνµ ,
satisfies the same property provided the antisymmetric tensor k falls off fast enough at
infinity. The corresponding Noether charge (8.8) is left unaffected by this modification
provided the integral of k on the boundary of Σ vanishes; if that integral does not
vanish, however, the charge receives an additional surface contribution of the form
Z
Qsurface = (dD−2 x)µν k µν (8.9)
∂Σ
where (dD−2 x)µν is proportional to µνα1 ...αD−2 dxα1 ...dxαD−2 . As we have just argued,
the would-be Noether charges of a gauge theory can only receive surface contributions
such as (8.9) since the corresponding Noether current vanishes up to the divergence
of a two-form.
At first sight this means that the situation is even worse than expected, since the
Noether charges of gauge theories are apparently ill-defined: there is no a priori way
to associate a k µν with a given symmetry generator, so the surface integral (8.9) can
take any value. But in fact, this also suggests a solution to the problem: instead of
trying to build a conserved current j µ , one can associate, with a gauge symmetry, a
(D − 2)-form k µν and define the corresponding charge by (8.9). If k µν is conserved on-
shell in the sense that ∇µ k µν ≈ 0, then the corresponding charge (8.9) is conserved by
time evolution. In that context, the field k µν is called a superpotential and its integral
(8.9) over the boundary of Σ is known as the associated surface charge.3 For example,
in electrodynamics, the superpotential coincides with the strength tensor F µν and the
corresponding surface charge is the flux of the electric field at infinity, that is, the total
electric charge. Its conservation follows from the fact that ∂µ F µν vanishes on-shell by
virtue of Maxwell’s equations.
Thus the computation of conserved charges for gauge symmetries boils down to the
problem of associating a conserved superpotential with a given gauge transformation,
and understanding to what extent that superpotential is unique.
Asymptotic symmetries
While the definition (8.7) of the Noether current associated with a global symmetry
transformation is straightforward, that of the superpotential associated with a gauge
transformation is much more involved; see e.g. [176,178,199]. Here we simply summa-
rize the main ideas so as to apply them later to the specific case of three-dimensional
gravity. The contruction consists of several steps:
1. Define the theory by choosing a bulk action, imposing certain fall-off conditions
on the field content, and possibly adding a boundary term to the bulk action
such that the full action is differentiable.
2. Find, among all possible gauge transformations, those that preserve the fall-off
conditions. Such gauge transformations are said to be allowed, as opposed
to the gauge transformations that spoil the fall-off conditions and are therefore
3
The term “superpotential” here has nothing to do with supersymmetry.
Chapter 8. Symmetries of gravity in AdS3 224
• Non-trivial gauge transformations are global symmetries that map a field config-
uration on a physically different one. They fall off at infinity much slower than
trivial gauge transformations and change the state of the system when acting on
it. For example, in electrodynamics, non-trivial gauge transformations at spatial
infinity take the form δAµ (x) = ∂µ (x) with (x) = const.4 This corresponds to
a global U(1) symmetry and the associated charge is the electric charge.
Note that infinitesimal gauge transformations are always endowed with a Lie bra-
cket. Accordingly, they span a Lie algebra. The notions introduced above then lead
to the following terminology:
Definition. The asymptotic symmetry algebra of a theory is the quotient of the alge-
bra of allowed gauge transformations by its ideal consisting of trivial transformations.
Figure 8.2: Gauge transformations fall in three classes: forbidden transformations are
those that do not preserve the fall-off conditions of the theory; allowed transformations
are those that do, although they generally change the state of the system; trivial
transformations are those that preserve the fall-off conditions and leave the state of
the system unchanged. In this sense trivial gauge transformations are actual gauge
redundancies, and the global symmetry group of the system is the quotient of the
group of allowed transformations by its subgroup of trivial gauge transformations.
Since Poisson brackets satisfy the Jacobi identity, eq. (5.35) still holds: for any two
infinitesimal gauge transformations ξ, ζ and any field configuration Φ, we have
{Q[ξ], Q[ζ]} , Φ = Q [ξ, ζ] , Φ . (8.11)
It is tempting to remove the Poisson brackets from both sides of this equality and
conclude that surface charges provide an exact representation of the asymptotic sym-
metry algebra. However, this naive removal would overlook the crucial point (5.36)
that surface charges generally close according to a (classical) central extension of the
algebra of asymptotic symmetry generators:
{Q[ξ], Q[ζ]} = Q [ξ, ζ] + c(ξ, ζ) . (8.12)
Here c(ξ, ζ) is a real-valued two-cocycle that acts trivially on any field and is therefore
invisible in eq. (8.11). The point of the seminal paper [14] was to show that such non-
trivial central extensions do arise in asymptotic symmetries of gravitational systems.
Chapter 8. Symmetries of gravity in AdS3 226
Then three-dimensional Anti-de Sitter space (or simply AdS3 ) is the submanifold of
R2,2 given by
for some parameter `2 > 0, equipped with the induced metric of R2,2 . The parameter
` is called the AdS radius. The manifold (8.14) is diffeomorphic to a product S 1 × R2
where the circle is time-like; in particular it contains closed time-like curves. Its
isometry group is O(2, 2) and acts transitively according to xµ 7→ Λµ ν xν , where xµ
denotes the coordinates (x, y, u, v) and Λ is a 4×4 matrix that preserves the “Minkowski
metric” (8.13). The stabilizer for this action is isomorphic to O(2, 1), so there is a
diffeomorphism
AdS3 ∼= O(2, 2)/O(2, 1) ∼ = SO(2, 2)/SO(2, 1).
Killing vectors
The Killing vectors that generate isometries of AdS3 can be found thanks to the em-
bedding (8.14), where “Lorentz” transformations are generated by the six independent
vector fields
ξ1 = u∂v − v∂u , ξ2 = x∂y − y∂x , ξ3 = u∂y + y∂u ,
ξ4 = v∂x + x∂v , ξ5 = u∂x + x∂u , ξ6 = v∂y + y∂v .
The combinations of signs appearing here are due to the metric (+ + − −) in (8.13).
Upon defining
`0 ≡ 12 (ξ1 + ξ2 ), `¯0 ≡ 12 (ξ1 − ξ2 ),
`1 ≡ 21 (ξ3 + ξ4 − iξ5 + iξ6 ), `¯1 ≡ 12 (−ξ3 + ξ4 − iξ5 − iξ6 ), (8.17)
`−1 ≡ 12 (ξ3 + ξ4 + iξ5 − iξ6 ), `¯−1 ≡ 12 (−ξ3 + ξ4 + iξ5 + iξ6 ),
one finds the following Lie brackets for m, n = −1, 0, 1:
i[`m , `n ] = (m − n)`m+n , i[`¯m , `¯n ] = (m − n)`¯m+n , i[`m , `¯n ] = 0 . (8.18)
This exhibits the isomorphism so(2, 2) ∼ = sl(2, R) ⊕ sl(2, R),5 upon identifying the Lie
brackets (5.90). Note that the generator of time translations is ∂t = 1` (`0 + `¯0 ) while
the generator of rotations is ∂ϕ = `0 − `¯0 .
5
Strictly speaking we have displayed this isomorphism here for the complexification of so(2, 2),
but it also holds for real Lie algebras.
Chapter 8. Symmetries of gravity in AdS3 228
Figure 8.4: The universal cover of three-dimensional Anti-de Sitter space-time, diffeo-
morphic to R3 . It is equivalent to the interior of a solid cylinder, which may be seen
as the Penrose diagram of AdSg 3 . The time coordinate t is directed along the axis of
the cylinder while r is a radial coordinate, and ϕ is a 2π-periodic coordinate on the
circle. The spatial boundary r → +∞ is a two-dimensional time-like cylinder spanned
by the coordinates (ϕ, t), or equivalently by the light cone coordinates x± .
Spatial infinity
The region r → +∞ is a cylinder spanned by coordinates (ϕ, t) at space-like infinity.
It is the spatial boundary ∂M of AdS3 . In that region the metric (8.16) is
`2 2
2
`2 2
dt
2
ds ∼ 2 dr − r 2
− dϕ 2
= dr − r2 dx+ dx− (8.19)
r `2 r2
where we have introduced the light-cone coordinates
t
x± ≡ ± ϕ. (8.20)
`
For large r the Killing vector fields (8.17) are asymptotic to
+ 1 + − 1 −
`m ∼ eimx ∂+ − imeimx r∂r , `¯m ∼ eimx ∂− − imeimx r∂r
2 2
where m = −1, 0, 1. They generate global conformal transformations of the cylinder
at infinity, including time translations `0 + `¯0 = `∂t and rotations `0 − `¯0 = ∂ϕ . These
expressions have the general form
1 1
ξ ∼ X(x+ )∂+ − ∂+ X(x+ )r∂r , ξ¯ ∼ X̄(x− )∂− − ∂− X̄(x− )r∂r (8.21)
2 2
where the functions X and X̄ are 2π-periodic. Brown-Henneaux boundary conditions
will be such that vector fields of the form (8.21) are asymptotic symmetry generators
for arbitrary functions X, X̄.
As a starting point we ask what is the minimum amount of metrics that we wish to
include. A natural choice is to take pure AdS3 together with conical deficits, which
are obtained by cutting out a wedge out of the middle of AdS3 and identifying its two
sides. Concretely, consider the manifold described by coordinates r ∈ [0, +∞[ , ϕ ∈ R,
t ∈ R subject to the identifications
(r, ϕ, t) ∼ r, ϕ + 4πω, t − 2πA (8.22)
for some A ∈ R and ω > 0. (The normalization of ω is chosen for later convenience.)
For ω = 1/2 and A = 0 this reduces to the identifications that define pure AdS3 .
For 0 < ω < 1/2 it is a conical deficit; for ω > 1/2 it is a conical excess. Since
this is a global (topological) identification, the resulting pseudo-Riemannian manifold
still solves Einstein’s vacuum equations everywhere, except at the origin. In fact, the
metric (8.16) with identifications (8.22) is the solution of Einstein’s equations coupled
to the stress tensor of a point mass at the origin. Using the change of coordinates
A ϕ
t0 ≡ t + ϕ, r0 ≡ r , ϕ0 ≡ , (8.23)
2ω 2ω
the space-time metric can be rewritten as
r02 dr02
ds = − 1 + 2 (dt0 − Adϕ0 )2 +
2
+ 4ω 2 r02 dϕ02 (8.24)
` 1 + r02 /`2
where now there are no identifications on t0 , while ϕ0 is 2π-periodic. The term A dt0 dϕ0
suggests that A is proportional to angular momentum, as will indeed be the case below.
Note that the integral curves of ∂ϕ0 contain closed time-like curves unless
02 A2 `2
|A| ≤ 2ω` and r ≥ . (8.25)
4ω 2 `2 − A2
Thus, the space-time manifold has no pathologies only in the region where r0 is large
enough (and in particular in the asymptotic region r0 → +∞), and provided the pa-
rameter A is not too large compared to ω`. Accordingly, from now on we refer to the
solutions (8.24) with 0 < ω < 1/2 and |A| = 2ω` as extreme conical deficits.
In order to find boundary conditions that genuinely describe AdS3 space-times, one
would like the asymptotic symmetry algebra to at least include so(2, 2). If in addition
the phase space is to contain conical deficits (8.24), one is led to act with so(2, 2)
transformations on such conical deficit metrics so that, if ξ is an AdS3 Killing vector
and gµν is the metric of a conical deficit, the fall-off conditions are satisfied by the
infinitesimally transformed metric
gµν + Lξ gµν . (8.26)
Here Lξ gµν generally does not vanish because ξ may not be a Killing vector for the
conical deficit. In terms of cylindrical coordinates (r, ϕ, t), one thus obtains metrics
that satisfy the fall-off conditions [14]
`2
+ O(r−4 ) O(r−3 ) O(r−3 )
grr grϕ grt r2
(gµν ) = gϕr gϕϕ gϕt ∼ O(r−3 ) r2 + O(1) O(1) . (8.27)
2
gtr gtϕ gtt −3 r
O(r ) O(1) − `2 + O(1)
Chapter 8. Symmetries of gravity in AdS3 230
In practice, we will impose an extra gauge-fixing condition that simplifies the compu-
tation of asymptotic symmetries. Namely, it turns out that the mixed components grϕ
and grt can always be set to zero (identically) by applying a trivial differomorphism —
2
one whose surface charges all vanish. The subleading corrections to grr = r`2 + O(r−4 )
can similarly be set to zero. We refer to this gauge choice as the Fefferman-Graham
gauge. It leads to the following definition [6]:
r→+∞ `2 2
ds2 dr + r2 ηab + O(1) dxa dxb
∼ 2
(8.28)
r
with ηab dxa dxb the two-dimensional Minkowski metric on the cylinder. Then we say
that (M, ds2 ) is asymptotically Anti-de Sitter in the sense of Brown-Henneaux (in the
Fefferman-Graham gauge), with a cosmological constant Λ = −1/`2 .
From now on, when dealing with AdS3 gravity, we always restrict our attention to
metrics satisfying the Brown-Henneaux boundary conditions (8.28). For practical pur-
poses we will mostly describe the time-like cylinder in terms of light-cone coordinates
x± , in which case the label a in (8.28) takes the values ± and the Minkowski metric
on the cylinder is ηab dxa dxb = −dx+ dx− . Note that asymptotically AdS3 space-times
need not be (and generally are not) globally diffeomorphic to AdS3 ; in particular there
may be singularities in the bulk, as the definition (8.28) only requires r to be larger
than some lower limiting value. In the following pages we establish the main properties
of this family of metrics, including their asymptotic symmetry algebra.
Remark. The fact that one is allowed to choose the Fefferman-Graham gauge with-
out losing any information is a general property of locally asymptotically Anti-de Sitter
space-times [200]. It is related to the Fefferman-Graham expansion of AdS metrics and
the ambient construction of conformal structures [201], where conformal manifolds are
built as boundaries, or celestial spheres, of higher-dimensional bulk manifolds.
in terms of light-cone coordinates (8.20). Here the first condition follows from the fact
that the components grr = `2 /r2 and gr± = 0 are fixed, while the components gab are
allowed to fluctuate by terms of order r0 at infinity.
Chapter 8. Symmetries of gravity in AdS3 231
Lemma. Let gµν be a metric that is asymptotically AdS3 in the sense (8.28) and let
ξ be a vector field that satisfies the properties (8.29). Then
1
ξ = X(x+ )∂+ + X̄(x− )∂− − ∂+ X(x+ ) + ∂− X̄(x− ) r∂r + (subleading)
(8.30)
2
where X(x+ ) and X̄(x− ) are two arbitrary (smooth) 2π-periodic functions while the
subleading terms take the form
Z +∞ 0
`2 dr ab 0 ±
− ∂a (∂+ X + ∂− X̄) g (r , x )∂b =
2 r r0
(8.31)
`2 −4
= 2 ∂− (∂+ X + ∂− X̄)∂+ + ∂+ (∂+ X + ∂− X̄)∂− + O(r ) .
2r
These formulas associate an asymptotic Killing vector ξ with an asymptotically AdS3
metric gµν and a vector field X(x+ )∂+ + X̄(x− )∂− on the cylinder; the dependence of
ξ on the latter is linear.
Proof. Let gµν be an asymptotically AdS3 metric (8.28). We first note that the re-
quirement Lξ grr = 0 imposes ∂r ξ r = ξ r /r, whose solution is
for some function F on the cylinder. On the other hand the condition Lξ gr± = 0 yields
2
∂r ξ c = −g ca `r ∂a F, which is solved by
+∞
dr0 ab 0 ±
Z
a a ± 2 ±
ξ = X (x ) + ` ∂b F(x ) g (r , x ) (8.33)
r r0
where X a ∂a is an arbitrary vector field on the cylinder. Note that the integral over
r0 converges since gab (r, x± ) = r2 ηab + O(1) by virtue of (8.28), so that the inverse is
ab
g ab = ηr2 + O(r−4 ). Plugging this in the integral of (8.33) we find explicitly
`2 ab
ξ a = X a (x± ) + η ∂b F(x± ) + O(r−4 ) . (8.34)
2r2
In light-cone coordinates (8.20), the two-dimensional Minkowski metric reads
η++ η+− 0 −1/2 ab 0 −2
(ηab ) = = and (η ) =
η−+ η−− −1/2 0 −2 0
Note that the asymptotic Killing vector (8.30) takes the anticipated form (8.21)
and thus provides the generalization we were hoping to find. We will denote by ξ(X,X̄)
the asymptotic Killing vector field determined by the functions X(x+ ) and X̄(x− ).
One can decompose these functions in Fourier modes and define the vector fields
whose Lie brackets take the form (8.18) up to subleading corrections, with indices
m, n ranging over all integer values. Thus, asymptotically, the finite-dimensional
isometry algebra so(2, 2) of AdS3 is enhanced to two commuting copies of the infinite-
dimensional Witt algebra (6.24). In fact we can already anticipate the result:
At this stage, we have not yet proven this claim since we do not know whether
all asymptotic Killing vector fields (8.30) have non-vanishing surface charges on the
phase space; this will be done in the following pages. Also note that we are being
slightly sloppy in (8.37), since the diffeomorphisms generated by (8.30) affect the radial
coordinate. Hence formula (8.37) only holds up to 1/r corrections; it is accompanied
by transformations of the radial coordinate that we do not bother writing down, but
that do preserve the limit r → +∞ in that they map r on a positive O(1) multiple of
itself.
With this improved action it makes sense to solve Einstein’s equations in the space
of metrics (8.28). We will not review this computation here and refer to [6, 179, 180]
for details. The bottom line is that the general solution of the equations of motion
with Brown-Henneaux boundary conditions in the Fefferman-Graham gauge reads
`2 2 4G` − −
− 4G`
ds2 = dr − rdx +
− p̄(x )dx rdx − p(x +
)dx +
(8.38)
r2 r r
where p(x+ ) and p̄(x− ) are arbitrary, 2π-periodic functions of their arguments. The
factors of 4G` are introduced for later convenience. We will study this space of so-
lutions in greater detail below. For now, we only note that it is endowed with a
well-defined action of asymptotic symmetry transformations. Indeed, we define the
variation of p and p̄ under the action of an asymptotic Killing vector (8.30) by
Lξ(X,X̄) ds2 ≡ 4G` δX p(x+ ) (dx+ )2 + 4G` δX̄ p̄(x− ) (dx− )2 + (subleading),
and this variation preserves the structure of the solution (8.38). In particular, observe
that ξ(X,X̄) is an exact Killing vector if the variations δX p and δX̄ p̄ vanish. Using (8.30)
one finds
c 3 c̄ 3
δX p = X∂+ p + 2p∂+ X − ∂ X, δX̄ p̄ = X̄∂− p̄ + 2p̄∂− X̄ − ∂ X̄ (8.39)
12 + 12 −
where c = c̄ is the Brown-Henneaux central charge
3`
c = c̄ = . (8.40)
2G
The transformations (8.39) are exactly those of the components of a CFT stress ten-
sor under conformal transformations; they coincide with the coadjoint representation
(6.115) of the Virasoro algebra when seeing p(x+ ) and p̄(x− ) as Virasoro coadjoint
vectors. In that context the condition for ξ(X,X̄) to be an exact Killing vector is equiv-
alent to the statement that (X, X̄) belongs to the stabilizer of (p, p̄). We refrain from
interpreting these results any further for now; we will return to them in section 8.3.
Note that at this stage there is actually no reason to call (8.40) a central charge: even
though it does appear in (8.39) exactly as the inhomogeneous term of the coadjoint
representation (6.115), the specific value (8.40) is irrelevant since changing the normal-
ization of p or p̄ would change the value of c and c̄. The importance of the parameter
(8.40) will become apparent only from the algebra of surface charges.
solution for which all surface charges vanish, which we take to be the degenerate conical
deficit at p = p̄ = 0,
`2
ḡ = 2 dr2 − r2 dx+ dx− . (8.41)
r
With this normalization one can show that the conserved superpotentials correspond-
ing to Brown-Henneaux asymptotic symmetries are such that the surface charge (8.9)
associated with the vector field ξ(X,X̄) on the solution (p, p̄) is
Z 2π
1
dϕ p(x+ )X(x+ ) + p̄(x− )X̄(x− )
Q(X,X̄) [p, p̄] = (8.42)
2π 0
where ϕ = 12 (x+ − x− ). (See e.g. [6] for an explicit computation.)
This charge can be interpreted in two ways: first, as the Noether charge associated
with a conformal transformation (X, X̄) in a two-dimensional CFT on the cylinder
with stress tensor (p, p̄); second, as the pairing (6.111) between the direct sum of two
Virasoro algebras and its dual. This is consistent with the fact that the transformation
law (8.39) coincides with the coadjoint representation of Virasoro. In particular, the
charge associated with time translations corresponds to the asymptotic Killing vector
∂t = (∂+ + ∂− )/`; it is the ADM mass of the system, or equivalently the Hamiltonian
Z 2π
1
dϕ p(x+ ) + p̄(x− )
M [p, p̄] = (8.43)
2π` 0
and it coincides (up to a factor 1/`) with the sum of two Virasoro energy functionals
(7.79). Similarly the charge associated with rotations, generated by the asymptotic
Killing vector ∂ϕ = ∂+ − ∂− , is the angular momentum
Z 2π
1
dϕ p(x+ ) − p̄(x− )
J= (8.44)
2π 0
and coincides with the difference of two Virasoro energy functionals. With this nor-
1
malization, pure AdS3 (8.47) has mass M = − 8G ; all its other surface charges vanish.
In the last line we have introduced the bracket [X, Y ] defined as the usual Lie bracket
of vector fields on the line, while c(X, Y ) is the Gelfand-Fuks cocycle (6.43) expressed
in the coordinate x+ . This is a Virasoro algebra (6.108), with a classical central
extension! The same computation would hold in the barred (antichiral) sector, while
chiral and antichiral charges commute. Thus we conclude:
Chapter 8. Symmetries of gravity in AdS3 235
The Poisson bracket algebra (8.45) can also be rewritten in terms of more conven-
tional Virasoro generators. If we define the charges
their Poisson brackets close according to two copies of the Virasoro algebra (6.118),
up to the renaming p → L. The central charges take the definite value c = c̄ = 3`/2G.
In particular, the normalization of the homogeneous term of the bracket fixes the
normalization of the Brown-Henneaux central charge, confirming the fact that it is an
unambiguous parameter specifying the phase space. In this language the mass (8.43)
and the angular momentum (8.44) are
1
M = (L0 + L̄0 ) , J = L0 − L̄0 ,
`
as in a two-dimensional conformal field theory. In particular, pure AdS3 has L0 =
L̄0 = −c/24, as does a CFT vacuum on the cylinder. Note that the Brown-Henneaux
central charge is essentially the Planck mass measured in units of the inverse of the
AdS3 radius. Equivalently, it is the inverse of the coupling constant of the system, so
the semi-classical limit corresponds to c → +∞.
`2 2 `2 − − `2 +
ds2AdS +
= 2 dr − rdx + dx rdx + dx . (8.47)
r 4r 4r
To verify that this is indeed pure AdS3 , note that the change of coordinates
`
r = earcsinh(r̄/`) (8.48)
2
Chapter 8. Symmetries of gravity in AdS3 236
brings this metric into the manifest AdS3 form (8.16) (up to the bar on the coordinate
r̄) by virtue of the identity
dr dr̄
=√ .
r `2 + r̄2
The angular momentum vanishes while the ADM mass of the solution is
1 c 1
Mvac = (L0 + L̄0 ) = − =− . (8.49)
` 12` 8G
This is the energy of the vacuum state of a two-dimensional CFT on the cylinder.
A2
p0 + p̄0 1 2
M= =− 4ω + 2 .
` 8G `
Extreme conical deficits are solutions of this type for which either p0 or p̄0 vanishes,
or equivalently for which |A| = 2ω`. Conical excesses are solutions for which p0 , p̄0 are
of the form (8.50) with |A| ≤ 2ω but ω > 1/2, and the line separating deficits from
excesses is a section of parabola
` 2G 2
`M = − − J , |J| ≤ `/4G (8.51)
8G `
whose endpoints are tangent to the lines `M = |J|. The solution at p0 = p̄0 = 0 is the
degenerate conical deficit (8.41) that we used to normalize charges. Note that conical
excesses with an angle of 2πn around the origin correspond to ω = n/2; for fixed n,
the set of such excesses is again a section of parabola in the (J, `M ) plane specified by
` 2 2G J 2
`M = − n − ,
8G ` n2
which generalizes (8.51). Note that, for vanishing angular momentum (A = 0), eq.
(8.50) yields p0 = p̄0 = −c ω 2 /6 in terms of the Brown-Henneaux central charge. This
is precisely the relation (7.46) between constant elliptic Virasoro coadjoint vectors and
their monodromy matrix.
When p0 and p̄0 are positive constants, the metric (8.38) turns out to be that of
a BTZ black hole with mass M = (p0 + p̄0 )/` and angular momentum J = p0 − p̄0
written in Fefferman-Graham coordinates [179]:
`2 2 2G` −
− 2G`
ds2BTZ +
= 2 dr − rdx − (`M − J)dx rdx − +
(`M + J)dx . (8.52)
r r r
Chapter 8. Symmetries of gravity in AdS3 237
Figure 8.5: The zero-mode solutions of AdS3 gravity with Brown-Henneaux boundary
conditions. The origin of the coordinate system (J, `M ) is the degenerate conical
deficit (8.41); the AdS3 metric is located below, on the `M axis, at the lower tip of
the shaded square. BTZ black holes are located in the wedge |J| ≤ `M . Conical
deficits and excesses are located in the lower wedge |J| ≤ −`M , respectively above
and below the parabola (8.51). All metrics such that |J| > `|M | contain closed time-
like curves at arbitrarily large radius. Anticipating the results of section 8.3.3, we
have shaded the solutions whose orbit has energy bounded from below under Brown-
Henneaux transformations; those are all BTZ black holes, the AdS3 metric, and all
conical deficits such that p0 , p̄0 ≥ −c/24. Certain solutions with energy bounded from
below are pathological in that they contain closed time-like curves at infinity — those
are the two diagonal strips surrounding the BTZ wedge.
The lightest BTZ black hole at M = J = 0 is the degenerate conical deficit (8.41).
This is strikingly different from four-dimensional black holes: in the latter case, the
lightest black hole is typically empty space, whereas in three dimensions the lightest
black hole is separated from AdS3 by a classical mass gap. The metrics that fill this
gap are conical deficits, i.e. metrics of point particles, so one can loosely say that a
particle turns into a black hole when its mass is higher than the threshold c/24`, which
Chapter 8. Symmetries of gravity in AdS3 238
is essentially the Planck mass. We will encounter a similar phenomenon in flat space,
though in that case black holes will be replaced by cosmological space-times.
Remark. Since three-dimensional gravity has no local degrees of freedom, all solu-
tions of Einstein’s equations in three dimensions are locally isometric to AdS3 and can
therefore be realized as quotients of AdS3 . In particular, the BTZ metric (8.52) has
no curvature singularity at r = 0, where, as everywhere else, it is locally isometric
to AdS3 . So how can it be a black hole? The answer to this question was clarified
in [17], where it was noted that the point r = 0 is a singularity in the causal sense
even though it is a regular point in the metric sense. Note that black holes obtained
as regular identifications of AdS also exist in higher dimensions [213, 214].
According to this viewpoint, the space of solutions (8.38) is really the phase space
of AdS3 gravity with Brown-Henneaux boundary conditions. The purpose of this sec-
tion is to analyse some of its properties and to relate them with holography. Thus,
we interpret points of phase space as CFT stress tensors, describe and interpret their
transformation law under Brown-Henneaux transformations, and derive a positive en-
ergy theorem for AdS3 gravity. Quantization is relegated to section 8.4.
gravity coincides with the space of CFT stress tensors on the cylinder at fixed central
charges. The finite transformation laws of these stress tensors under conformal trans-
formations are given by the coadjoint representation of the Virasoro group, eq. (6.114).
The Poisson structure on the phase space of AdS3 gravity is determined by the
requirement (8.10) ensuring that surface charges generate the correct transformation
laws when acting on the fields of the theory. For Brown-Henneaux boundary conditions
this leads to the Poisson brackets of charges (8.45), which coincides with the Kirillov-
Kostant Poisson bracket (6.118). Hence we conclude:
Theorem. The covariant phase space of AdS3 gravity with Brown-Henneaux bound-
ary conditions is a hyperplane at fixed central charges (8.40) embedded in the space
of the coadjoint representation of the direct product of two Virasoro groups, and en-
dowed with the corresponding Kirillov-Kostant Poisson structure.
A loose way to interpret this theorem is to say that AdS3 gravity is group theory:
the whole phase space of the system is determined by the structure of the Virasoro
group, save for the fact that the value of the central charge is fixed by the coupling
constant. We will encounter a similar phenomenon in the next chapter when dealing
with asymptotically flat gravity. This being said, the occurrence of a Virasoro coad-
joint representation should not come as too big a surprise. Indeed, it is always true
that the charges associated with certain symmetries transform under the coadjoint
representation of the symmetry group (since these charges are nothing but momen-
tum maps). Accordingly, the surface charges of AdS3 gravity were bound to involve
the coadjoint representation of the Virasoro group. The only surprise is that Virasoro
coadjoint vectors exactly coincide with the functions specifying the metric, instead of
being some complicated non-linear combinations of the entries of the metric and their
derivatives. In particular, note that the set of on-shell Brown-Henneaux metrics (8.38)
is a vector space.
Remark. It is not strictly true that the whole phase space of AdS3 gravity coincides
with the dual of two Virasoro algebras. Indeed, this conclusion is entirely based
on the asymptotic solutions (8.38), but completely overlooks the fact that some of
these solutions cannot be extended arbitrarily far into the bulk. This subtlety leads
to (finitely many) additional directions in the complete phase space of the theory,
as discussed in [217]. We will ignore this detail since it plays a minor role for our
purposes.
zero-mode solutions were described in section 8.2.6, but a generic par (p, p̄) is definitely
not a zero-mode since p and p̄ may have some non-trivial profile on the circle. If we pick
one such solution at random, we can generate infinitely many other ones by acting on
it with asymptotic symmetry transformations. The resulting manifold is the product
Chapter 8. Symmetries of gravity in AdS3 240
This is a good point to introduce a terminology which has come to be more or less
standard [181,218,219]: a metric (p, p̄) obtained by acting on pure AdS3 with a certain
asymptotic symmetry transformation is known as a (classical) boundary graviton. This
nomenclature is then extended to any metric obtained from a zero-mode solution by
an asymptotic symmetry transformation. The name is justified by the fact that three-
dimensional gravity has no local (bulk) degrees of freedom, but does have non-trivial
topological (boundary) degrees of freedom visible in the arbitrariness of the pair (p, p̄)
that specifies a solution of the equations of motion.
If our goal is to classify all solutions (8.38), then orbits provides a natural organiz-
ing criterion: rather than classifying the solutions, we can classify their orbits under
asymptotic symmetries. Since we know the classification of Virasoro coadjoint orbits,
we may claim to control the full covariant phase space of AdS3 gravity. In particu-
lar, the classification of zero-mode solutions in fig. 8.5 is a first step towards the full
classification: each point in the plane (J, `M ) defines the orbit of the corresponding
zero-mode solution, and different points define distinct orbits. However, as we have
seen in section 7.2, not all orbits have constant representatives: there exist infinitely
many conformally inequivalent on-shell metrics that cannot be brought into zero-mode
solutions by asymptotic symmetry transformations. Thus the complete classification
of AdS3 metrics is essentially a product of two copies of fig. 7.3, where zero-mode
solutions are those where both p and p̄ belong to the vertical axis of the figure. This
classification foliates the covariant phase space of AdS3 gravity into disjoint orbits of
the asymptotic symmetry group.
Chapter 8. Symmetries of gravity in AdS3 241
Figure 8.6: A schematic representation of the AdS3 phase space foliated into orbits
of the asymptotic symmetry group. Solutions belonging to the same symplectic leaf
are related to one another by asymptotic symmetry transformations, i.e. “boosts”, but
there are no boosts that connect different leaves.
Remark. The relation between AdS3 gravity and orbits of the Virasoro group has
recently been the object of renewed interest, as it was realized that a similar structure
arises in many other contexts. To the author’s knowledge, the first explicit mention of
that relation appears in [182, 183, 220]; it is also hidden between the lines in [184, 221].
The relation was later studied in [45, 181] due to its implications for positive energy
theorems, while [222] (see also [223]) is devoted to the geometric properties of metrics
corresponding to non-constant pairs (p, p̄).
These arguments can be interpreted as a positive energy theorem for AdS3 gravity
[45, 181]. They imply in particular that, in the diagram of zero-mode solutions of fig.
8.5, all BTZ black holes belong to orbits with energy bounded from below, while all
conical excesses belong to orbits with unbounded energy. The pure AdS3 metric also
belongs to an orbit with energy bounded from below, while the only conical deficits
whose energy is bounded from below under asymptotic symmetries are those located in
Chapter 8. Symmetries of gravity in AdS3 242
the square −c/24 ≤ p0 , p̄0 ≤ 0. The absolute minimum of energy among all solutions
with energy bounded from below is realized by AdS3 space-time.
Remark. Positive energy theorems in general relativity have a long history; in short,
the problem is to show that energy is bounded from below in a suitably defined phase
space of metrics. This problem is classically addressed in four-dimensional asymp-
totically flat space-times, where positivity of energy was first proved in [224]. A
supersymmetry-based proof can also be found in [225], while the case of the Bondi
mass was settled shortly thereafter by various authors — see e.g. [226] for a list of
references. Note that a relation between positive energy theorems [227] and Virasoro
orbits was already suggested in footnote 8 of [121], albeit in a very different con-
text. In section 9.3 we will encounter a positive-energy theorem in three-dimensional
asymptotically flat space-times.
Remark. This section is our first encounter with representations of Lie algebras in
this thesis, so our language and notations will be somewhat different from those of
part I. The link between the language of part I and Lie algebra representations will be
established through induced modules, in section 10.2.
Equivalently, Lm = i`m in terms of the basis (5.89) and the brackets (5.90) become
with m, n = −1, 0, 1. In any unitary representation T of sl(2, R), the real generators
tµ are represented by anti-Hermitian operators acting in a suitable Hilbert space. In
terms of the generators (8.54), this is to say that the Hermiticity conditions
hold in a unitary representation. From now on we will abuse notation and neglect
writing the representation T , so that T [Lm ] ≡ Lm . This abuse is common in physics,
so it should not lead to any misunderstanding. Until the end of this chapter we also
use the Dirac notation instead of the less standard notation of part I.
Since the group SL(2, R) is simple but non-compact, all its non-trivial unitary
representations are infinite-dimensional. Fortunately, the complexification of sl(2, R)
coincides with that of su(2), and we definitely know how to build unitary highest-
weight representations of the latter. Let us therefore use the same approach for sl(2, R):
we start from a (normalized) highest-weight state |hi belonging to the Hilbert space
of the representation, such that
where h is the highest weight. The conditions (8.56) imply that the operator repre-
senting L0 is Hermitian, so h must be real. The interpretation of (8.57) is that |hi
has energy h if we think of L0 as the Hamiltonian, while L1 is an annihilation operator.6
In order to produce a representation we must also act with operators L−1 on the
highest-weight state. This leads to descendant states of the form
where the non-negative integer N is the level of the descendant. Each descendant
state is an eigenstate of L0 with eigenvalue h + N , so L−1 is analogous to a creation
operator. We then declare that the carrier space H of the representation is spanned by
all linear combinations of descendant states. Since we want H to be a Hilbert space,
descendant states with different levels must be orthogonal because their eigenvalues
under L0 differ. Furthermore, all descendant states must have non-negative norm
squared. Using the Hermiticity conditions (8.56), this amounts to the requirement
2 (8.56)
0 ≤ (L−1 )N |hi = h [(L−1 )N ]† (L−1 )N h = h (L1 )N (L−1 )N h .
To ensure that this condition holds, we evaluate scalar products of descendant states.
Thanks to the commutation relations (8.55) one finds
N
Y −1
N N
h (L1 ) (L−1 ) h = N! (2h + k) (8.59)
k=0
6
The terminology is somewhat backwards, since the highest weight h is actually the lowest eigen-
value of L0 in the space of the representation; this terminological clash is standard.
Chapter 8. Symmetries of gravity in AdS3 244
where we have used hh|hi = 1. Thus, all descendant states have strictly positive norm
squared if and only if h > 0. If h = 0, then the representation is trivial.
Representations by quantization
The representations just described can be identified with representations obtained by
quantizing suitable coadjoint orbits of SL(2, R). Indeed, note that the Lie algebra
(8.55) admits a quadratic Casimir operator
1 (8.54)
C = L20 − (L1 L−1 + L−1 L1 ) = −t20 + t21 + t22 . (8.60)
2
This operator is proportional to the identity in any irreducible representation of
sl(2, R). In the highest-weight representation (8.58) it takes the value
which proves by the way that the representation is irreducible. The fact that (8.60)
takes a constant value is reminiscent of the “mass shell” condition defining the coad-
joint orbit (5.93), and indeed the highest-weight representation just displayed is the
quantization of such an orbit for h > 0. The only subtlety is that the value of the
Casimir (8.61) is not quite h2 , but h(h − 1); the two numbers coincide in the semi-
classical limit h → +∞, and the h − 1 of (8.61) may be seen as a quantum correction
of the classical result.
is complicated by the fact that all coadjoint orbits (at non-zero central charge) are
infinite-dimensional; in addition, the only Casimir operators of the Virasoro algebra
are functions of its central charge [228]. Accordingly, we start with a few comments
regarding Virasoro geometric quantization, before turning to the construction of its
highest-weight representations and the evaluation of the associated characters.
Semi-classical regime
If one believes in the orbit method, geometric quantization applied to the coadjoint
orbits of the Virasoro algebra should produce unitary Virasoro representations. This
viewpoint was adopted in [121, 130, 131], with the conclusion that the quantization of
orbits with positive energy and constant representatives indeed provides highest-weight
representations in the large c limit. By contrast, the limit of small c is much more
elusive, and at present it is not known if the discrete series of Virasoro representations
at c ≤ 1 can be obtained by geometric quantization (see e.g. [229]). From a gravita-
tional point of view, the Virasoro central charge (8.40) is the AdS radius in units of the
Planck length, so large c corresponds to the semi-classical regime. This is confirmed
by symplectic geometry: the Kirillov-Kostant symplectic form (5.29) evaluated at a
constant Virasoro coadjoint vector (p0 , c) is
(6.111) c
ω(p0 ,c) (ξm )p0 , (ξn )p0 = −im 2p0 + m2 δm+n,0 (8.62)
12
where ξm = ad∗Lm is the vector field on W(p0 ,c) that generates the coadjoint action
of the Virasoro generator Lm given by (6.109). The occurrence of c confirms that
the regime of large c is semi-classical in the sense that a large volume is assigned to
any portion of phase space. Conversely, small c corresponds to the non-perturbative
regime, where quantum corrections may alter classical results in a radical way.
This heuristic argument is consistent with the fact that geometric quantization is
relatively well established at large c, but poorly understood at small c. Since appli-
cations to three-dimensional gravity rely on the semi-classical limit anyway, from now
on we restrict ourselves to the regime of large c. This assumption turns out to greatly
simplify representation theory, and allows us to think of highest-weight Virasoro rep-
resentations as quantizations of Virasoro orbits with constant representatives.
This being said, to our knowledge there is as yet no strict mathematical proof of
the fact that geometric quantization of Virasoro orbits produces highest-weight repre-
sentations, despite numerous attempts in the literature (see e.g. [230,231]). Our view-
point here will be pragmatic, and we shall assume that the representations obtained by
quantizing such orbits are indeed highest-weight representations. This assumption will
be supported, among other observations, by the fact that Virasoro characters match
suitable gravitational partition functions (see section 8.4.4).
Highest-weight representations
The basis of the Virasoro algebra given by (6.109) is such that, in any unitary repre-
sentation, the operators representing the generators Lm + L−m , i(Lm − L−m ) and Z
Chapter 8. Symmetries of gravity in AdS3 246
where we abuse notation by denoting the basis element Lm and the operator that
represents it with the same symbol. In any irreducible representation the Hermitian
central operator Z is proportional to the identity with a coefficient c ∈ R, so we may
write the commutation relations of the operators representing the generators (8.63) as
c
[Lm , Ln ] = (m − n)Lm+n + m(m2 − 1)δm+n,0 . (8.65)
12
Here the existence of the sl(2, R) subalgebra (8.55) is manifest, since the contribution
of the central extension vanishes for m = −1, 0, 1.
To begin, in analogy with (8.57), one defines the highest-weight state of the repre-
sentation to be a normalized state |hi such that7
The state |hi is also called a primary state. Its definition ensures that it has energy
h under the Hamiltonian L0 , while the operators Lm with m > 0 are annihilation
operators. In analogy with (8.58) one also defines descendant states
Thus one can interpret the operators L−m with m > 0 as creation operators. We will
discuss the gravitational interpretation of this representation in section 8.4.4.
7
The actual value of the quantum weight h may differ from the classical parameter defined in
(8.66) by corrections of order O(1/c), so from now on it is understood that h refers to the quantum
value. This subtlety will have very little effect on our discussion.
Chapter 8. Symmetries of gravity in AdS3 247
Using the commutation relations (8.65) of the Virasoro algebra, one verifies that
each descendant (8.68) is an eigenstate of L0 with eigenvalue
n
X
h+ ki ≡ h + N
i=1
where the non-negative integer N is the level of the descendant. One then declares
that the space H of the Virasoro representation is the Verma module spanned by all
linear combinations of descendant states. As in the case of sl(2, R), descendants with
different levels have different eigenvalues under L0 . According to (8.64) the latter must
be Hermitian if the representation is to be unitary, so scalar products of descendants
with different levels vanish. However, in contrast to sl(2, R), there are in general many
different descendant states with the same level. More precisely, at large central charge
c, the number of different descendants at level N is the number p(N ) of partitions of
N in distinct positive integers (e.g. p(0) = p(1) = 1, p(2) = 2, p(3) = 3, p(4) = 5, etc.).
Note that the representation whose carrier space is spanned by the descendant
states (8.68) is an induced representation of the Virasoro algebra, i.e. an induced
module. Indeed, the conditions (8.67) define a one-dimensional representation of the
subalgebra generated by L0 and the Lm ’s with m > 0, and the prescription (8.68) is
the algebraic analogue of the statement that wavefunctions live on a quotient space
G/H (recall section 3.3). By the way, a similar interpretation holds for the highest-
weight representations of sl(2, R) displayed in section 8.4.1. We will return to this
observation in section 10.2.
Remark. It was recently shown [237] that all irreducible unitary representations
of the Virasoro group with a spectrum of L0 bounded from below are highest-weight
representations of the type described here. In this sense, highest-weight representations
exhaust all unitary representations of the Virasoro algebra.
Vacuum representation
In sl(2, R), the representation with highest weight h = 0 is trivial since all descendant
states are null by virtue of eq. (8.59). We now describe the Virasoro representation
obtained by setting h = 0, to which we refer as the vacuum representation. It is ob-
tained by quantizing the vacuum Virasoro orbit, containing the point pvac = −c/24.
The highest weight state |0i of that representation satisfies the properties
which ensures that the vacuum state |0i is invariant under the sl(2, R) subalgebra
generated by L−1 , L0 and L1 ; it is the quantum counterpart of the statement that the
stabilizer of the vacuum orbit is the group PSL(2, R). Crucially, the vacuum is not
invariant under the higher-mode generators L−2 , L−3 etc. so that the representation
whose carrier space is spanned by all descendant states
is non-trivial. This representation is unitary for all c > 0, and it is free of null states
(i.e. irreducible) whenever c > 1. It is in many ways analogous to the standard
highest-weight representation generated by the descendant states (8.68), but the con-
dition L1 |0i = 0 makes it slightly smaller than generic highest-weight representations.
It may seem puzzling that the vacuum state is not left invariant by all Virasoro
generators but only by a subset thereof as in (8.69). The reason is that, at non-zero
central charge, it is impossible to define a non-zero state that is annihilated by all
Virasoro generators. Indeed, if there was such a state |0̃i, then we would have
(8.65)
c c
0 = h0̃|Ln L−n |0̃i = h0̃| 2nL0 + n(n2 − 1) |0̃i = n(n2 − 1) 0̃|0̃ ,
12 12
Chapter 8. Symmetries of gravity in AdS3 249
where the notation is almost the same as in eq. (5.102). The parameter τ is a complex
number with positive imaginary part so that, if L0 is interpreted as the Hamiltonian,
then Im(τ ) is the inverse temperature and the character itself is a canonical partition
function. The normalization factor −c/24 is conventional.
X +∞
X
h+N −c/24 h−c/24
χ(τ ) = p(N )q =q p(N )q N . (8.72)
N =0 N =0
We now rewrite this in a more convenient way thanks to the following result:
Proof. We follow [238], to which we refer for a careful treatment of the convergence
issues that will not be addressed here. To prove (8.73) we consider its right-hand side
Chapter 8. Symmetries of gravity in AdS3 250
and expand each individual term of the infinite product as a geometric series:
+∞ +∞ X+∞
Y 1 Y
nk 2 3
2 4
n
= q = 1 + q + q + q + · · · 1 + q + q + · · · ···
n=1
1 − q n=1 k=0
+∞ X
X N X +∞
X
k1 +k2 +···+kn
=1+ q = p(N )q N .
| {z }
N =1 n=1 k1 ,k2 ,...,kn N =0
k1 +k2 +···+kn =N qN
1≤k1 ≤k2 ≤···≤kn
| {z }
p(N )
where in the second equality we have introduced the Dedekind eta function
+∞
Y
1/24
η(τ ) ≡ q (1 − q n ). (8.75)
n=1
The result (8.74) can be seen as an infinite product of sl(2, R) characters (5.102) with
parameters nτ . Equivalently, since sl(2, R) representations coincide with harmonic
oscillators whose partition functions were written in (5.102), one can think of a Virasoro
representation as an infinite collection of harmonic oscillators. This suggests that
Virasoro characters can be interpreted as quantum field theory partition functions,
which will be confirmed in section 8.4.4 below. Note also the presence of the ubiquitous
factor 1 − q n in the denominator.
Remark. Our derivation of (8.74) relied on the fact that the eigenvalue h + N of L0
has degeneracy p(N ). This is only true provided there are no null states, i.e. provided
c > 1. By contrast, for c ≤ 1, null states generally do exist and are modded out of the
Hilbert space of the representation. This leads to a smaller degeneracy of eigenvalues
of L0 , hence to a character that is strikingly different from (8.74). We will not display
characters at c ≤ 1 here; see e.g. [239, 240] for explicit formulas.
Vacuum representation
The character of the vacuum Virasoro representation can be evaluated in the same way
as for generic highest-weight representations. The only subtlety is that the vacuum is
sl(2, R)-invariant, leading to a reduced number of descendant states (8.70). Explicitly,
let ∆(N ) denote the degeneracy of the eigenvalue N in the space spanned by the
vacuum descendants. For N ≥ 2, this degeneracy is the number of partitions of N
in positive integers which are strictly greater than one (thus ∆(0) = 1 by convention
but ∆(1) = 0, ∆(2) = ∆(3) = ∆(4) = ∆(5) = 1, ∆(6) = 2, etc.). Then the vacuum
character is
+∞
X +∞
X
N −c/24 −c/24
χ(τ ) = ∆(N )q =q ∆(N )q N . (8.76)
N =0 N =0
Chapter 8. Symmetries of gravity in AdS3 251
Note how the product in the denominator of (8.77) is truncated (no term n = 1) owing
to the sl(2, R) symmetry of the vacuum state. In particular the vacuum character
(8.77) is not just the limit h → 0 of the generic character (8.74).
First let us make the proposal more precise: the orbit of a metric (p, p̄) is a coad-
joint orbit (8.53) of the product of two Virasoro groups with central charges (8.40). For
simplicity let us assume that the metric is a zero-mode and that its energy is bounded
from below under asymptotic symmetry transformations, so p(x+ ) = p0 ≥ −c/24 and
p̄(x− ) = p̄0 ≥ −c̄/24. The non-zero modes belonging to the orbit (8.53) may be seen
as classical analogues of the descendant states (8.68). Upon defining h ≡ p0 + c/4 and
h̄ ≡ p̄0 + c̄/24, one expects that the geometric quantization of the orbit (8.53) pro-
duces the tensor product of two highest-weight representations of the Virasoro algebra
labelled by (h, c) and (h̄, c̄).8 The same would be true by quantizing the orbit of AdS3 ,
except that the result would be the tensor product of two vacuum representations.
where q ≡ e2πiτ and q̄ is its complex conjugate. This is precisely the character of the
tensor product of two Virasoro vacuum representations, which confirms that quantized
boundary gravitons around AdS3 span an irreducible highest-weight representation of
vir ⊕ vir. The same computation can be performed for orbifolds of AdS3 obtained by
imposing identifications of the form ϕ ∼ ϕ + 2π/N in terms of cylindrical coordinates,
where N ∈ N∗ . The corresponding metric is that of a conical deficit labelled by the
Virasoro coadjoint vectors p0 = p̄0 = −c/(24N 2 ). For N ≥ 2 the resulting one-loop
partition function is found to be
+∞
2h
Y 1
Zgrav,N (β, θ) = |q| (8.80)
n=1
|1 − q n |2
c
where h = 24 (1 − 1/N 2 ). This is again the character of the tensor product of two
highest-weight representations of the Virasoro algebra with weights h = h̄, which
confirms the interpretation of Virasoro modules as particles dressed with boundary
gravitons.
In chapter 10 we will develop a similar interpretation for BMS3 particles, which will
then be confirmed in chapter 11 by the matching of BMS3 characters with one-loop
partition functions for asymptotically flat gravity and higher-spin theories.
Remark. In [184] it was conjectured that the one-loop partition function (8.79) is
exact because it is the only expression compatible with Virasoro symmetry. Higher
loop corrections would then renormalize the Brown-Henneaux central charge but would
leave the q-dependent one-loop determinant unaffected. This conjecture was used to
evaluate a Farey tail sum representing the putative full, non-perturbative, partition
function of AdS3 gravity (see also [242]). To our knowledge the one-loop exactness of
(8.79) is still an unproven statement.
Chapter 8. Symmetries of gravity in AdS3 253
This being said, one should keep in mind that the comparison to Goldstone bosons
should be handled with care. Indeed, spontaneously broken internal symmetries are
always such that states obtained by acting with broken symmetry generators on the
vacuum do not belong to the Hilbert space. This statement is the Fabri-Picasso the-
orem (see e.g. [243]), and it follows from the fact that the norm of the state Q|0i has
an infrared (volume) divergence whenever Q generates a broken global internal sym-
metry. If the Fabri-Picasso theorem was to hold for asymptotic symmetries, then the
descendant states (8.81) would make no sense. Fortunately the situation of asymptotic
symmetries is different because the charges that generate them are surface charges (8.9)
rather than Noether charges. As a result, when Q is a broken asymptotic symmetry
generator, the norm squared of Q|0i is an integral over a compact manifold, and is
therefore finite. In this sense spontaneously broken asymptotic symmetries behave in
a way radically different from spontaneously broken internal symmetries.
Part III
BMS3 symmetry
and gravity in flat space
This part contains the original contributions of the thesis and is de-
voted to Bondi-Metzner-Sachs (BMS) symmetry in three dimensions. It
starts with an introductory chapter where the definition of the BMS3 group
is motivated by asymptotic symmetry considerations. We then move on
to the quantization of BMS3 symmetry and show that irreducible unitary
representations of the BMS3 group, i.e. BMS3 particles, are classified by
supermomentum orbits that coincide with coadjoint orbits of the Virasoro
group. We also evaluate the associated characters and show that they
coincide with one-loop partition functions of the gravitational field at finite
temperature and angular potential. Finally we extend this matching to
higher-spin theories and supergravity in three dimensions.
Chapter 9
The structure is as follows. In section 9.1 we show how BMS3 symmetry emerges
from an asymptotic symmetry analysis. Section 9.2 is devoted to the abstract mathe-
matical definition of the BMS3 group and its central extension, including their adjoint
and coadjoint representations. In section 9.3 we describe the phase space of three-
dimensional asymptotically flat gravity embedded in the space of the coadjoint repre-
sentation of BMS3 . Finally, in section 9.4 we show how BMS3 symmetry can be seen
as a flat limit of Virasoro symmetry.
This chapter is mostly based on [45–47], although the first section follows the earlier
references [6, 37]. As usual, more specialized references will be cited in due time.
257
Chapter 9. Classical BMS3 symmetry 258
The isometry group of Minkowski space is the Poincaré group (4.39) with D = 3:
IO(2, 1) = O(2, 1) n R3 . Its elements are pairs (f, α) acting transitively on R3 ,
xµ 7→ f µ ν xν + αµ , (9.2)
While inertial coordinates are the most common in Minkowski space, a different set
of coordinates will be more convenient for the description of BMS3 symmetry. Namely,
as in (1.4), we define retarded Bondi coordinates (r, ϕ, u) by
p x + iy
r≡ x2 + y 2 , eiϕ ≡ , u ≡ t−r, (9.3)
r
whose range is r ∈ [0, +∞[ , u ∈ R, and ϕ ∈ R with the identification ϕ ∼ ϕ + 2π.
In that context the coordinate u is known as retarded time. We will also refer to the
coordinates (r, ϕ, t) as cylindrical coordinates; they are analogous to (8.15) in AdS3 .
In terms of cylindrical and Bondi coordinates, the Minkowski metric (9.1) reads
Killing vectors
The Killing vector fields that generate Poincaré transformations (9.2) are simplest to
write down in inertial coordinates, where they have the general form
ξ(x) = αρ + X µ xν µν ρ ∂ρ .
(9.5)
Here αµ and X µ are two arbitrary, constant vectors generating translations and Lorentz
transformations, respectively, while µνρ is the completely antisymmetric tensor such
that 012 = 1 (indices are raised and lowered with the Minkowski metric). In partic-
ular, the component α0 is responsible for time translations, while α1 and α2 generate
translations in the directions x = x1 and y = x2 , respectively. The component X 0 is
responsible for spatial rotations while X 1 and X 2 give rise to boosts in the directions
Chapter 9. Classical BMS3 symmetry 259
x1 and x2 , respectively.
α0 (ϕ)
ξTranslation = α(ϕ)∂u − ∂ϕ + α00 (ϕ)∂r (9.6)
r
where the function α(ϕ) is related to the translation vector αµ by
Note that both (9.6) and (9.8) depend on functions on the circle; already at this stage
it is tempting to speculate that there exist boundary conditions such that asymptotic
symmetry generators take that form with arbitrary functions (X, α) on the circle. The
BMS boundary conditions below will do just that.1
The structure of the algebra spanned by the vector fields (9.6) and (9.8) can be
made more transparent by a suitable choice of basis. Thus we define the complexified
Poincaré generators
jm ≡ ξLorentz , pm ≡ ξTranslation
X(ϕ)=eimϕ α(ϕ)=eimϕ
with m, n = −1, 0, 1. The Lie algebra of the BMS3 group will extend these brackets
by allowing arbitrary integer values of m, n, in the same way that the Witt algebra
(6.24) extends sl(2, R). Note that the structure G n g of the Poincaré group (4.93) is
manifest in these relations, as the bracket of j’s with p’s takes exactly the same form
as the bracket of j’s with themselves.
null infinity. It is the upper cone on the boundary of the Penrose diagram of fig. 1.1.
This is due to our choice of coordinates: instead of (9.3) we could have defined ad-
vanced Bondi coordinates, with advanced time given by v = t + r instead of u = t − r.
As a result we would have found that the region r → +∞ is past null infinity, but up
to this difference the whole construction would have been the same. In this thesis we
use retarded Bondi coordinates throughout, but it is always understood that a parallel
construction exists in terms of advanced Bondi coordinates. In particular the region
r → +∞ will always be future null infinity, denoted I + .
Future null infinity is the region of space-time where all light rays eventually escape;
similarly past null infinity is the origin of all incoming light rays. In optical terms, if we
were living in a three-dimensional space-time, the region that we would see around us
would be a circle on our past light-cone. As the distance from us to the circle increases,
the latter approaches past null infinity. A similar (time-reversed) interpretation holds
for future null infinity, and justifies the following terminology:
Definition. The future celestial circle at retarded time u associated with the Bondi
coordinates (r, ϕ, u) is the circle spanned by the coordinate ϕ on future null infinity,
and at fixed time u. Similarly the past celestial circle at advanced time v is the circle
at fixed time v on past null infinity.
This definition is illustrated in fig. 1.2. From now on the words “celestial circle”
always refer to a future celestial circle. As we shall see, BMS3 symmetry will refor-
mulate and generalize the action of Poincaré transformations on celestial circles, and
more generally on null infinity.
Poincaré transformations on I +
Since we have rewritten Minkowski Killing vectors in Bondi coordinates, it is worth
asking whether one can write finite Poincaré diffeomorphisms (9.2) (as opposed to
infinitesimal vector fields) in Bondi coordinates. The answer is obviously yes, but the
result is not particularly illuminating because the linear nature of the transformations
is hidden when writing them in terms of (r, ϕ, u). Fortunately, Bondi coordinates are
designed so that things simplify at null infinity; in particular it turns out that Poincaré
transformations preserve the limit r → +∞ in the sense that (i) they map r on a pos-
itive multiple of itself and (ii) they affect ϕ and u but leave them finite. Accordingly
Poincaré transformations are well-defined at null infinity and one may ask how they
act on the coordinates (ϕ, u) spanning I + . The procedure for finding this action is
explained in greater detail in [109].
For definiteness we focus on the connected Poincaré group (4.40). We can use the
isomorphism (4.83) to describe Lorentz transformations in terms of SL(2, R) matrices,
the correspondence being given explicitly by (4.91). One then finds that a pure space-
time translation xµ 7→ xµ + αµ acts on null infinity according to
(u, ϕ) 7→ u + α(ϕ), ϕ (translation) (9.11)
where the function α(ϕ) is related to the components αµ by (9.7). In particular,
pure time translations act on Bondi coordinates as u 7→ u + α0 , without affecting the
Chapter 9. Classical BMS3 symmetry 261
transformation (6.88) of the celestial circle, with parameters A, B given by (6.89). For
instance, spatial rotations act as ϕ 7→ ϕ + θ, leaving all other coordinates untouched.
This is analogous to the four-dimensional situation described in eqs. (1.6)-(1.7). Upon
performing simultaneously a translation α and a Lorentz transformation f , the trans-
formation of (u, ϕ) reads
(u, ϕ) 7→ f 0 (ϕ)u + α(f (ϕ)) , f (ϕ) (9.12)
where f takes the form (6.88) while α is given by (9.7). As we shall see below, the
BMS3 group acts on I + in the same way, except that f (ϕ) will be an arbitrary dif-
feomorphism of the circle and that α(ϕ) will be an arbitrary function on the circle.
Analogous results hold at past null infinity.
Note that in (9.12) we are abusing notation slightly. Indeed, a Poincaré transfor-
mation is a diffeomorphism of the whole space-time (not just null infinity) and acts
on all three coordinates r, ϕ, u. In particular the transformation law (9.12) only holds
up to corrections of order 1/r. These corrections vanish in the limit r → +∞ and
leave out only the leading piece displayed in (9.12), but they matter for the extension
of Poincaré (or BMS) transformations from the boundary into the bulk.
which is the flat limit (` → +∞) of eq. (8.24). In these terms there are no identifica-
tions on t0 , and ϕ0 is 2π-periodic. As in the AdS3 case the cross-term Adt0 dϕ0 suggests
that A is proportional to angular momentum, as will indeed be the case below. In
contrast to AdS3 , however, the region of large r0 is always free of closed time-like
curves since the condition for the integral curves of ∂ϕ0 to be space-like now simply
yields r02 > A2 /(4ω 2 ), without condition on the ratio of A and ω. This is a flat limit
of the more stringent conditions (8.25) encountered in AdS3 . We refer for instance
to [244, 245] for a more thorough study of conical deficits.
Now suppose we wish to find boundary conditions that include such conical deficits.
If we want the asymptotic symmetry group to contain the Poincaré group, we are
forced to include in the phase space all metrics obtained by performing rotations,
Chapter 9. Classical BMS3 symmetry 262
translations and boosts of conical deficits. This is the same argument as in section
8.2, where we derived Brown-Henneaux boundary conditions. It leads to a class of
metrics with prescribed asymptotic behaviour at null infinity, analogous to eq. (8.27)
in the AdS3 case. Some of the subleading components of the metric can then be set
to zero identically as a gauge choice, which leads to the following definition:
Then we say that (M, ds2 ) is asymptotically flat at future null infinity in the BMS
gauge. A parallel construction exists at past null infinity.
The BMS gauge condition is the flat space analogue of the Fefferman-Graham gauge
used in (8.28). We stress that it is truly a gauge condition in the sense of asymptotic
symmetries: the diffeomorphism used to bring a metric from a general asymptotically
flat form into the BMS gauge is trivial, as it does not affect the surface charges of
the metric. By contrast, the diffeomorphisms that change the physical state of the
system are generated by non-zero surface charges and span the asymptotic symmetry
group of the system, which will turn out to be the BMS3 group. From now on, when
dealing with asymptotically flat gravity, we always restrict our attention to metrics
satisfying the BMS boundary conditions (9.14). Note that asymptotically flat space-
times need not be (and generally are not) globally diffeomorphic to Minkowski space;
the definition (9.14) only requires r to be larger than some lower limiting value. Note
also that there is no restriction on the sign of the fluctuating components in the metric
(9.14); in particular the term of order r0 multiplying du2 may be positive.
together with
in terms of retarded Bondi coordinates. Here (9.15) follows from the fact that the
components grr = grϕ = 0 and gϕϕ = r2 are fixed in the BMS gauge (9.14); by
contrast the components guu , guϕ and gur are allowed to fluctuate by terms of order
r0 , r0 and r−1 respectively.
Chapter 9. Classical BMS3 symmetry 263
Lemma. Let gµν be an asymptotically flat metric in the sense (9.14) and let ξ be a
vector field that satisfies (9.15) and (9.16). Then
where X(ϕ) and α(ϕ) are two arbitrary (smooth) 2π-periodic functions, while the
subleading terms take the form
Z +∞ 0
0 00 dr
(α + uX ) gur ∂ϕ
r r02
Z +∞ 0
0 00 dr 1 0 00 (9.18)
+ ∂ϕ (α + uX ) gur + 2 (α + uX )guϕ ∂r =
r r02 r
1 1
= (α0 + uX 00 )∂ϕ + (α00 + uX 000 )∂r + O(r−2 ) .
r r
These formulas uniquely associate an asymptotic Killing vector
field ξ with an asymp-
totically flat metric gµν and two functions X(ϕ), α(ϕ) on the celestial circle; the
dependence of ξ on these functions is linear.
Proof. Let gµν be an asymptotically flat metric (9.14). First note that the condition
Lξ grr = 0 yields ∂r ξ u = 0, so ξ u is r-independent. On the other hand the condition
Lξ grϕ = 0 gives a differential equation ∂r ξ ϕ = − r12 gru ∂ϕ ξ u , which is solved by
+∞
dr0
Z
ϕ u
ξ = X(u, ϕ) + ∂ϕ ξ gr0 u (9.19)
r r02
where X(u, ϕ) is an arbitrary function on the cylinder at null infinity. The integral
over r0 converges since gru = −1 + O(1/r) by virtue of (9.14), so that ξ ϕ = X(u, ϕ) +
1
∂ ξ u + O(1/r2 ). At this point we introduce a function α(u, ϕ) defined by
r ϕ
(prime denotes partial differentiation with respect to ϕ), which is allowed by virtue of
the fact that ξ u is r-independent. In these terms the condition Lξ gϕϕ = 0 gives
1
ξ r = −r∂ϕ ξ ϕ − guϕ (α0 + uX 00 ) (9.21)
r
where ξ ϕ is given by (9.19). Since we now know that the most general solution ξ
of (9.15) is determined by two functions X(u, ϕ) and α(u, ϕ) on null infinity, we can
use the remaining conditions (9.16) to constrain these functions. Using first Lξ gur =
O(1/r), we find
∂u ξ u = X 0 , (9.22)
which upon rewriting ξ u as (9.20) says that the combination ∂u α+u∂u X 0 vanishes. The
requirement Lξ guϕ = O(1) then yields ∂u X = 0, which is to say that X(u, ϕ) = X(ϕ)
only depends on the coordinate ϕ on the celestial circle. Plugging this back into (9.22)
then yields ∂u α = 0 as well. Formula (9.17) follows, while the subleading terms (9.18)
are produced by (9.19) and (9.21).
Chapter 9. Classical BMS3 symmetry 264
Note that the asymptotic Killing vectors (9.17) precisely take the anticipated form
(9.6)-(9.8) and generalize Poincaré transformations in an infinite-dimensional way. In
particular the asymptotic symmetry group contains all space-time translations (cor-
responding to α(ϕ) of the form (9.7)) and all Lorentz transformations (corresponding
to X(ϕ) of the form (9.9)). We shall denote by ξ(X,α) the asymptotic Killing vector
determined by the functions X(ϕ) and α(ϕ). One verifies that the Lie brackets of such
vector fields read
ξ(X,α) , ξ(Y,β) = ξ([X,Y ],[X,β]−[Y,α]) + (subleading) (9.23)
where the brackets in the subscript on the right-hand side are understood to be stan-
dard Lie brackets on the circle, e.g. [X, α] ≡ Xα0 − αX 0 . The subleading terms can be
neglected because they will turn out not to contribute to the surface charges; alterna-
tively, as in the AdS3 case, they can be absorbed by a redefinition of the Lie bracket
such that the algebra is realized everywhere in the bulk [60, 176].
As one can verify, formula (9.23) implies that their Lie brackets take the form (9.10)
with arbitrary integer labels m, n, up to subleading corrections.
Remark. One should keep in mind that Bondi coordinates are global, since the
definition (9.3) covers all points of Minkoswki space-time. Thus the fact that Bondi
coordinates allow one to describe either only future or only past null infinity (and not
both) does not mean that they cover only “half” of the space-time. A similar comment
applies to BMS symmetry, whose definition in terms of space-time relies on a choice
of coordinates that favours future over past null infinity (or vice-versa). Despite this
2
The prefix “super” has nothing to do with supersymmetry, but stresses the fact that special-
relativistic quantities are extended in an infinite-dimensional way.
Chapter 9. Classical BMS3 symmetry 265
asymmetry, it was recently realized that (for well-behaved asymptotically flat space-
times [246]) the two definitions of BMS can be related by an “antipodal identification”,
which leads to the application of BMS symmetry to scattering phenomena [19, 24, 26,
27, 247–258]. A related question (as yet unsolved) is whether BMS symmetry can be
defined at spatial infinity [259].
Accordingly, it makes sense to ask about the general solution of Einstein’s vacuum
equations in the BMS gauge. It was shown in [6] that this solution reads
where p(ϕ) and j(ϕ) are arbitrary, 2π-periodic functions of ϕ. Upon evaluating surface
charges we will see that p(ϕ) and j(ϕ) are densities of energy and angular momentum
at null infinity, respectively. As in the earlier AdS3 case (8.38), the normalization
factors involving Newton’s constant G are included for later convenience.
The transformation law of the solution (9.25) under the action of asymptotic Killing
vectors follows from the definition
Lξ(X,α) ds2 ≡ 8G δ(X,α) p(ϕ) du2 + 8G δ(X,α) j(ϕ) + u δ(X,α) p0 (ϕ) dudϕ
(9.26)
where the functions X(ϕ) and α(ϕ) determine the vector field (9.17). Evaluating the
Lie derivative (9.26) one finds
c2
δ(X,α) j = Xj 0 + 2X 0 j + αp0 + 2α0 p − α000 , (9.27)
12
0 0 c2 000
δ(X,α) p = Xp + 2X p − X (9.28)
12
where c2 is a dimensionful central charge proportional to the Planck mass [37]:
3
c2 = . (9.29)
G
In this language the asymptotic vector field ξ(X,α) is an exact Killing vector field for
the metric (j, p) if both variations (9.27)-(9.28) vanish. The subscript “2” in (9.29) will
be justified below.
The transformation law of p in (9.28) coincides with that of a CFT stress tensor
under a conformal transformation generated by X; it is the coadjoint representation
Chapter 9. Classical BMS3 symmetry 266
With this normalization one can show that the surface charge (8.9) associated with
the vector field ξ(X,α) on the solution (j, p) is [6]
Z 2π
1
Q(X,α) [j, p] = dϕ j(ϕ)X(ϕ) + p(ϕ)α(ϕ) . (9.31)
2π 0
It can be interpreted as the pairing of the bms3 algebra, consisting of pairs (X, α),
with its dual consisting of pairs (j, p). In particular, even though we haven’t defined
the BMS3 group at this stage, we already know that the space of solutions (9.25)
belongs to its coadjoint representation. The charge associated with time translations
correponds to the asymptotic Killing vector ∂u ; it is the Hamiltonian of the system,
Z 2π
1
M = P0 = dϕ p(ϕ) , (9.32)
2π 0
and it allows us to interpret p(ϕ) as the energy density carried by the gravitational
field at (future) null infinity. Thus p(ϕ) is the Bondi mass aspect associated with the
metric (9.25) and its zero-mode (9.32) is the Bondi mass. More generally the charges
associated with supertranslations (X = 0) take the form
Z 2π
1
Q(0,α) [j, p] = dϕ p(ϕ)α(ϕ) . (9.33)
2π 0
In the same way, the charge associated with rotations corresponds to the asymptotic
Killing vector ∂ϕ ; it is the angular momentum
Z 2π
1
J = J0 = dϕ j(ϕ) . (9.34)
2π 0
Chapter 9. Classical BMS3 symmetry 267
We can interpret j(ϕ) as the density of angular momentum carried by the gravitational
field at null infinity; it is the angular momentum aspect associated with the metric
(9.25). More generally all superrotation charges take the form
Z 2π
1
Q(X,0) [j, p] = dϕ j(ϕ)X(ϕ)
2π 0
and generalize centre of mass charges. With this normalization Minkowski space (9.4)
has energy M = −1/8G and all its other surface charges vanish.
Using the infinitesimal transformation laws (9.27)-(9.28) and integrating by parts one
can then show that
Q(X,α) [j, p], Q(Y,β) [j, p] = Q([X,Y ],[X,β]−[Y,α]) [j, p] + c2 [c(X, β) − c(Y, α)] , (9.36)
for all m ∈ Z, generalizing the Hamiltonian (9.32) and angular momentum (9.34).
Then the bracket (9.36) yields the algebra
i{Jm , Jn } = (m − n)Jm+n ,
c2 3
i{Jm , Pn } = (m − n)Pm+n + m δm+n,0 , (9.37)
12
i{Pm , Pn } = 0 .
Remark. The BMS boundary conditions given here are the flat analogue of Brown-
Henneaux boundary conditions. In this sense they are the “standard” fall-offs for
three-dimensional asymptotically flat gravity. However, it is likely that other consistent
boundary conditions exist in Einstein gravity — for instance adapting to flat space the
free AdS3 boundary conditions of [207]. In addition one can devise BMS-like boundary
conditions for other theories of gravity, such as topologically massive gravity [261,262],
bigravity [263], conformal gravity [264] or new massive gravity [265,266]. In particular,
in parity-breaking theories such as TMG, one typically finds that the Virasoro algebra
spanned by superrotations Jm develops a non-zero central charge c1 . Aside from this
comment we will have very little to say about these alternative possibilities.
3
Indeed, changing the normalization of p would also change the value of the central charge that
ensures that the bracket {J , P} takes the canonical form in eq. (9.37).
Chapter 9. Classical BMS3 symmetry 269
Figure 9.1: The zero-mode solutions of asymptotically flat gravity with BMS3 fall-offs.
The origin of the coordinate system (J, M ) is the null orbifold (9.30); the Minkowski
metric is located below, on the M axis, right between conical deficits and conical
excesses. Flat space cosmologies are located in the region M > 0. Conical deficits are
such that −c2 /24 < M < 0 while excesses have M < −c2 /24. Anticipating section
9.3.3, we have shaded the solutions whose orbit has energy bounded from below under
BMS3 transformations; those are all flat space cosmologies and all conical excesses,
plus Minkowski space. Note that this figure is a flat limit of fig. 8.5, as the slope of
the curve `M = J in the plane (J, M ) goes to zero when ` → +∞.
Solutions having 0 > p0 > −c2 /24 are conical deficits for all values of j0 , with
a deficit angle 2π(1 − 2ω) given by (7.46). In particular, solutions with p0 = 0 are
degenerate conical deficits, and the solution p0 = j0 = 0 is the null orbifold (9.30) that
we used to normalize charges. Solutions having p0 < −c2 /24 are conical excesses with
an excess angle 2π(2ω − 1) given again by (7.46). For p0 = −c2 n2 /24 the excess angle
is 2π(n − 1).
Zero-mode solutions with positive p turn out to describe flat space cosmologies,
sometimes also called shifted boost orbifolds [267, 268]. They represent a (2 + 1)-
dimensional universe that undergoes a big crunch followed by a big bang, where the
transition between the contracting and expanding phases is smooth only if j 6= 0.
When j = 0 these solutions can be thought of as a compactification of the three-
dimensional Milne universe. They can also be seen as limits of the interior region of
BTZ black holes as the AdS3 radius goes to infinity. The lightest flat space cosmology
has p0 = 0 and is separated from Minkowski space-time pvac = −c2 /24 by a classical
mass gap; the latter is filled by conical deficits. This is very similar to the mass gap
separating BTZ black holes from AdS3 .
Chapter 9. Classical BMS3 symmetry 270
Note that, in contrast to AdS3 , no cosmic censorship is needed to ensure the absence
of closed time-like curves at infinity (although closed time-like curves generally do exist
in the bulk). In fact, the whole classification of flat zero-mode metrics may be seen as
a limit ` → +∞ of that of zero-mode metrics in AdS3 . The family of flat zero-mode
solutions is plotted in fig. 9.1.
The plan of this section is as follows. Motivated by the structure of the Poincaré
group (4.93), we start by defining a notion of “exceptional semi-direct products” (gen-
erally centrally extended) and work out their adjoint and coadjoint representations.
We then use asymptotic symmetries to motivate the definition of the BMS3 group and
its central extension, which turn out to be exceptional semi-direct products based on
the Virasoro group. Finally, we write down the adjoint representation, the Lie algebra
and the coadjoint representation of the (centrally extended) BMS3 group. Through-
out the section, these structures are compared to their Poincaré counterparts and to
three-dimensional asymptotically flat gravity. Note that the material presented here
relies heavily on chapters 2, 4 and 6.
where gAb denotes the Lie algebra of G seen as an Abelian vector group acted upon by
G according to the adjoint representation. Its group operation is given by (4.6) with
the action σ replaced by the adjoint.
G Ad gAb
b nc b (9.39)
(f, λ; α, µ) (9.40)
where λ, µ are real numbers, being understood that the pair (f, λ) belongs to G b while
(α, µ) belongs to bgAb . The notation emphasizes the fact that “centrally extended rota-
tions” (f, λ) play a role radically different
from “centrally extended translations” (α, µ).
In fact the notation (f, λ), (α, µ) would be more accurate, but to reduce clutter we
stick to (9.40).
(f, λ; α, µ) · (g, ρ; β, ν) =
(6.98)
1 (9.41)
= f · g, λ + ρ + C(f, g) ; α + Adf β, µ + ν − hS[f ], βi
12
g Aad
b c bgAb (9.42)
Chapter 9. Classical BMS3 symmetry 272
The adjoint representation of the group (9.39) follows from formula (5.105). Start-
ing for simplicity with the centreless group (9.38), it is given by
Ad(f,α) (X, β) = Adf X, Adf β − adAdf X α = Adf X, Adf β − [Adf X, α] (9.43)
where the “Ad” on the right denotes the adjoint representation of G alone.4 In the
second equality we abuse notation by writing a bracket between Adf X ∈ g and α ∈
gAb , being understood that we use the Lie bracket of g and interpret the result as an
element of gAb . The Lie bracket of the centreless Lie algebra g Aad gAb follows:
(X, α), (Y, β) = [X, Y ], adX β − adY α = [X, Y ], [X, β] − [Y, α] , (9.44)
in accordance with the general formula (5.106). Note that this is precisely the form of
the Lie bracket (9.23) of BMS3 asymptotic Killing vectors.
c (f,α) (X, λ; β, µ) =
Ad
1 1 1
= Adf X, λ − hS[f ], Xi ; Adf β − [Adf X, α], µ − hS[f ], βi + hs[Adf X], αi
12 12 12
(9.45)
where Ad on the right-hand side denotes the adjoint representation of G and [·, ·] is
the Lie bracket of g. On the left-hand side we have neglected central terms in the
subscript of the adjoint representation, since they act trivially.
From the adjoint representation one can read off, by differentiation, the Lie bracket
of the centrally extended algebra (9.42). One expects the contribution of central terms
to include a cocycle c given by (6.101), and indeed one finds
(X, λ; α, µ), (Y, ρ; β, ν) = [X, Y ], c(X, Y ); [X, β] − [Y, α], c(X, β) − c(Y, α) (9.46)
where we abuse notation as in (9.44). Already note that the last entry precisely takes
the form of the central extension in the Poisson bracket (9.36) of flat surface charges.
The appearance of the same cocycle c in both central entries of (9.46) is due to the
exceptional semi-direct product structure of (9.39). It implies that, when written in
terms of generators, the brackets of rotations with translations take the same form as
the brackets of rotations with themselves, including central terms. Explicitly, suppose
gAb
we are given a basis of b gAb consisting of non-central generators
Ja ≡ (ja , 0; 0, 0) , Pa ≡ (0, 0; pa , 0)
4
In case of identical notations, the subscript indicates which group we are referring to.
Chapter 9. Classical BMS3 symmetry 273
where the ja ’s and pa ’s respectively generate g and gAb , together with two central
elements
Z1 ≡ (0, 1; 0, 0) , Z2 ≡ (0, 0; 0, 1) . (9.47)
Suppose also that the Lie brackets of Ja ’s take the form (2.27) with some structure
constants fab c and some central coefficients cab , and let us choose the basis elements
Pa such that their bracket with Ja ’s takes the same form as the bracket of Ja ’s with
themselves. This is allowed by the exceptional semi-direct product structure. Then
the bracket (9.46) implies that the commutation relations of b gAb gAb are
[Ja , Jb ] = fab c Jc + cab Z1 ,
[Ja , Pb ] = fab c Pc + cab Z2 , (9.48)
[Pa , Pb ] = 0 .
The fact that J ’s act on P’s according to the adjoint representation is now manifest
since the structure constants of the two first lines are identical. Note in particular that
the central generator Z1 pairs rotations with themselves, while Z2 pairs rotations with
translations. The centrally extended BMS3 algebra (9.37) illustrates this phenomenon,
as does the Poincaré algebra (9.10), albeit without central extension.
Note that the definition of (9.39) rules out all central extensions in the bracket
[P, P] of (9.48), and indeed we will show in section 9.2.5 that such central extensions
never take place in the centrally extended BMS3 algebra. However, for other semi-
direct product groups, such extensions may occur; an example is the symmetry group
of warped conformal field theories [93], Diff(S 1 ) n C ∞ (S 1 ).
Coadjoint representation
g∗ ⊕ b
The space of coadjoint vectors dual to the algebra (9.42) is a direct sum b g∗ , or more
g∗ ⊕ b
accurately b g∗Ab . Following the notation of section 5.4, its elements are quadruples
(j, c1 ; p, c2 ) (9.49)
where (j, c1 ) is a centrally extended angular momentum dual to b g, while (p, c2 ) is a
centrally extended momentum dual to b gAb . The real numbers c1 , c2 are central charges;
the first pairs rotation generators with themselves, while the second pairs rotations
with translations. The pairing of (9.49) with b gAb gAb is
(j, c1 ; p, c2 ), (X, λ; α, µ) = hj, Xi + hp, αi + c1 λ + c2 µ , (9.50)
where the two pairings h·, ·i on the right-hand side are those of g∗ with g and g∗Ab with
gAb , respectively. This is a centrally extended generalization of (5.109).
Formula (9.54) looks a bit scary but it is crucial for our purposes, so let us briefly
point out two of its important features. First, the central charges c1 , c2 are left invariant
by the action of the group, as expected. Second, note that the transformation law of
momentum is
c2
f · p = Ad∗f p − S[f −1 ] , (9.55)
12
where the Ad∗ on the right-hand side is that of G (not G). b This formula says that p is
invariant under translations (since it is unaffected by α) and that its transformation
law is blind to the central charge c1 , but not to c2 . In fact, eq. (9.55) is the coadjoint
representation (6.104) of the centrally extended group G b at central charge c2 . As
a corollary we can already conclude that the orbits of momenta labelling unitary
representations of (9.39) are coadjoint orbits of the group G b at fixed central charge c2 ;
there is no need to master the much more complicated transformation law of angular
momentum in (9.54) in order to classify such representations. This will have key
consequences for the BMS3 group below.
Remark. Property (9.51) explains why we refer to the map (5.110) as a cross product.
Indeed, the (double cover of the) Euclidean group in three dimensions is an exceptional
semi-direct product SU(2) nAd su(2)Ab . Since the coadjoint representation of SU(2)
is equivalent to the adjoint, one may identify vectors with covectors and the cross
product (9.51) for the Euclidean group can be rewritten as α × p = adα p = [α, p].
Here the Lie bracket is that of su(2), so in components one has (α × p)i = ijk αj pk ,
which is the standard definition of the cross product in mechanics.
Chapter 9. Classical BMS3 symmetry 275
(f,α)
7−→ f 0 (g(ϕ)) g 0 (ϕ)u + β(g(ϕ)) + α f (g(ϕ)) , f (g(ϕ)) .
where Adf β denotes the adjoint representation (6.17) of Diff(S 1 ) acting on β, that is,
the transformation law of a vector field β(ϕ)∂ϕ under f (ϕ):
(Adf β) f (ϕ)
= f 0 (ϕ)β(ϕ). (9.57)
Definition. The centreless BMS group in three dimensions is the exceptional semi-
direct product
BMS3 ≡ Diff(S 1 ) nAd Vect(S 1 )Ab (9.58)
where Diff(S 1 ) is the group of diffeomorphisms of the circle while Vect(S 1 )Ab is its
Lie algebra, seen as an Abelian vector group acted upon by Diff(S 1 ) according to the
adjoint representation. Its elements are pairs (f, α); its group operation follows from
the general definition (4.6) and is given by
(f, α) · (g, β) = f ◦ g, α + Adf β (9.59)
where Ad is the action (9.57) of Diff(S 1 ) on vector fields. With this definition the
action (9.12) of BMS3 on null infinity reproduces the group operation (9.59).
The BMS3 group is infinite-dimensional and has the announced form (9.38), with
G = Diff(S 1 ). Since we saw in section 6.1 that PSL(2, R) is a subgroup of Diff(S 1 ),
the Poincaré group is obviously a subgroup of BMS3 . We therefore introduce officially
the following terminology:
Definition. In the BMS3 group (9.58), elements of Diff(S 1 ) are known as superro-
tations while elements of Vect(S 1 )Ab are called supertranslations.
Remark. The name “superrotation” has come to be standard, but the geometric
interpretation of Diff(S 1 ) makes the terminology “superboosts” somewhat more ap-
propriate. Indeed, recall from section 6.1 that the group Diff(S 1 ) is homotopic to a
circle, so that the only superrotations spanning a compact group are those conjugate
to rigid rotations f (ϕ) = ϕ + θ. The other one-parameter subgroups of Diff(S 1 ) are
all non-compact and should be seen as boost groups.
Definition. The universal cover of the BMS group in three dimensions is the excep-
tional semi-direct product
]+
BMS g+ 1 1
3 ≡ Diff (S ) nAd Vect(S )Ab (9.61)
]+
In particular, exact representations of BMS 3 generally correspond to projective rep-
resentations of BMS3 . The groups BMS3 , BMS+
+ ]+
3 and BMS3 are well-defined infinite-
dimensional Lie-Fréchet groups. In what follows, motivated by quantum-mechanical
applications, we always focus (implicitly) on the universal cover. Accordingly we abuse
notation and denote the universal cover simply by BMS3 , neglecting the superscript
“+” and the tilde.
Definition. The centrally extended BMS group in three dimensions is the exceptional
semi-direct product
d 1 ) n c Vect(S
[ 3 ≡ Diff(S
BMS d 1) (9.62)
Ad
The centreless BMS3 group (9.58) is perfect, in the same way as Diff(S 1 ); this
implies that it admits a universal central extension. As it turns out, this is precisely
[ 3 (see section 9.2.5 for the proof):
achieved by BMS
Theorem. The centrally extended BMS3 group (9.62) is the universal central exten-
sion of the centreless BMS3 group (9.58).
Chapter 9. Classical BMS3 symmetry 278
d 3 ≡ Vect(S
bms d 1 )Ab .
d 1 ) Ac Vect(S (9.64)
ad
Adjoint representation
The adjoint representation of the centreless BMS3 group is given by formula (9.43),
where the adjoint action of Diff(S 1 ) is the transformation law of vector fields (6.18).
An important subtlety is that the Lie bracket appearing on the right-hand side is
that of the Lie algebra of Diff(S 1 ) and is therefore the opposite (6.21) of the standard
bracket of vector fields. Accordingly, in terms of the usual Lie bracket of vector fields
on the circle one would write the adjoint representation of BMS3 as
Ad(f,α) (X, β) = Adf X, Adf β + [Adf X, α] , (9.65)
with a plus sign instead of a minus sign in the second entry of (9.43). The centrally
extended generalization of that expression is provided by eq. (9.45), where S is the
Schwarzian derivative (6.76), s is its infinitesimal version (6.74), and h·, ·i is the stan-
dard pairing (6.34). Again, when writing the adjoint representation in terms of the
standard Lie bracket of vector fields, the sign in front of the bracket of the third entry
of (9.45) is a plus instead of a minus. Since we will not explicitly need the adjoint
[ 3 group, we do not display it here.
representation of the BMS
Lie brackets
From the adjoint representation one can read off the Lie bracket of the bms
d 3 algebra.
In order to absorb the minus sign of (6.21) we define the bracket to be
d c
(X, λ; α, µ), (Y, ρ; β, ν) ≡ − Ad (etX ,tα) (Y, ρ; β, ν) .
dt t=0
Chapter 9. Classical BMS3 symmetry 279
The Lie algebra structure can be made more apparent by writing the bracket
(9.46) in a suitable basis. As in (9.24) we define the complex superrotation and
supertranslation generators of the centreless bms3 algebra,
where the index m runs over all integers. Their brackets take the form (9.10). The
corresponding basis of the centrally extended bms
d 3 algebra is
(9.66)
Jm ≡ jm , 0; 0, 0 = eimϕ ∂ϕ , 0; 0, 0 ,
(9.66) (9.67)
Pm ≡ 0, 0; pm , 0 = 0, 0; eimϕ (dϕ)−1 , 0 ,
together with two central elements (9.47), i.e. Z1 = (0, 1; 0, 0) and Z2 = (0, 0; 0, 1). In
these terms the centrally extended bracket (9.46) yields
Z1 3
i[Jm , Jn ] = (m − n)Jm+n + m δm+n,0 ,
12
Z2 3
i[Jm , Pn ] = (m − n)Pm+n + m δm+n,0 , (9.68)
12
i[Pm , Pn ] = 0 .
Up to central terms this is of the same form as the asymptotic symmetry algebra (9.10),
and it is consistent with the general form (9.48) for centrally extended exceptional
semi-direct products. In the first line we see that superrotations close according to
a Virasoro algebra (6.110) with central generator Z1 , while the second line shows
that brackets of superrotations with supertranslations take the Virasoro form with a
different central element Z2 . The algebra of surface charges (9.37) takes that form,
with definite values c1 = 0, c2 = 3/G for the central generators Z1 , Z2 .
Remark. The canonical Poincaré subgroup of BMS3 is the one spanned by super-
rotations (6.88) and supertranslations (9.7), or equivalently the one generated by ba-
sis elements jm , pm with m = −1, 0, 1. But in fact, BMS3 admits infinitely many
other Poincaré subgroups: each of them has a Lie algebra spanned by jn , j0 , j−n and
pn , p0 , p−n , consisting of superrotations of the form (6.95) and supertranslations
whose only non-vanishing Fourier modes are the zero-mode and the nth modes.
Definition. Let (j, p) be a coadjoint vector for the BMS3 group. Then p = p(ϕ)dϕ2
is called a supermomentum while j = j(ϕ)dϕ2 is an angular supermomentum.
The embedding of the Poincaré algebra in bms3 suggests an interpretation for the
lowest Fourier modes of
X X
j(ϕ) = jm e−imϕ and p(ϕ) = pm e−imϕ . (9.69)
m∈Z m∈Z
Indeed, p0 is dual to time translations and should be interpreted as the energy as-
sociated with p(ϕ); similarly, the components p1 and p−1 = p∗1 are complex linear
combinations of the spatial components of momentum. As for j0 , it is the angular
momentum associated with j(ϕ), while j1 and j−1 are centre of mass charges. More
generally, the function p(ϕ) should be seen as an energy density on the circle — es-
sentially a stress tensor — while j(ϕ) is an angular momentum density on the circle.
In particular, it is natural to give dimensions of energy to the function p(ϕ) and the
central charge c2 , while the function j(ϕ) and the central charge c1 are dimensionless.
This interpretation is confirmed by the surface charges (9.31), since p(ϕ) is a Bondi
mass aspect while j(ϕ) is an angular momentum aspect; furthermore the central charge
(9.29) is indeed a mass scale.
Coadjoint representation
As in the Virasoro case, one should think of the pair (j, p) as the stress tensor of a
BMS3 -invariant theory; its transformations under BMS3 then coincide with the coad-
joint representation, given for centrally extended exceptional semi-direct products by
formula (9.54). In that expression, the central charges are invariant (as they should)
while the transformation law of supermomentum coincides with the coadjoint repre-
sentation (6.114) of the Virasoro group at central charge c2 :
1 h c2 i
f · p f (ϕ) = 2 p(ϕ) + S[f ](ϕ) , (9.70)
f 0 (ϕ) 12
Chapter 9. Classical BMS3 symmetry 281
where S denotes the Schwarzian derivative (6.76). We stress once more that supermo-
mentum is left invariant by supertranslations, as it should.
Kirillov-Kostant bracket
A prerequisite for showing that the asymptotically flat phase space is a coadjoint
representation is to understand the Kirillov-Kostant Poisson bracket of the asymptotic
symmetry
∗ ∗ group.
[ 3 ; we proceed as in section 6.4. Thus let
Let us do this here for BMS
∗ ∗
Jm , Pm , Z1 , Z2 be the dual basis corresponding to (9.67) and (9.47). Writing any
coadjoint vector as
X
j(ϕ)dϕ2 , c1 ; p(ϕ)dϕ2 , c2 = jm Jm∗ + pm Pm
∗
+ c1 Z1∗ + c2 Z2∗ ,
m∈Z
d 3 ∗ . Their
the components {jm , pm , c1 , c2 } are global coordinates on the dual space bms
Poisson brackets (5.28) take the form
c1 3
i{jm , jn } = (m − n)jm+n + m δm+n,0 ,
12
c2
i{jm , pn } = (m − n)pm+n + m3 δm+n,0 , (9.71)
12
i{pm , pn } = 0 .
The bms3 algebra is perfect: it is equal to its Lie bracket with itself. This can
be seen, for instance, by noting that the right-hand sides of the brackets (9.10) span
all possible bms3 generators. Accordingly it follows from (2.19) that the first real
Chapter 9. Classical BMS3 symmetry 282
cohomology of bms3 vanishes: H1 (bms3 ) = 0. Since the same is true of the centreless
BMS3 group, its central extension is universal, and it only remains to establish the
second cohomology of bms3 .
where c is the Gelfand-Fuks cocycle (6.43). Their expression in the basis (9.66) is
m3
c1 (jm , jn ) = c2 (jm , pn ) = −i δm+n,0 , (9.73)
12
while their other components vanish. As a consequence, the Lie algebra (9.64) is
the universal central extension of (9.63), and the group (9.62) is the universal central
extension of (9.58).
Proof. The fact that the cocycle c1 is the only non-trivial cocycle pairing superrotation
generators with themselves follows from the fact that infinitesimal superrotations span
a Witt subalgebra of bms3 . The considerations of section 6.2 then carry over directly
to bms3 . Now let us ask whether there exists a two-cocycle c such that c(pm , pn ) 6= 0.
The cocycle identity (2.21) with trivial T implies
!
c(j0 , [pm , pn ]) + c̃(pm , [pn , j0 ]) + c̃(pn , [j0 , pm ]) = (m + n)c̃(pm , pn ) = 0 ,
where we used the Lie brackets (9.10). This yields c̃(pm , pn ) = c̃m δm+n,0 where the
coefficients c̃m = −c̃−m are to be determined. We now attempt to find a recursion
relation for these coefficients; using the cocycle identity
the bms3 algebra (9.10) implies (2m − 1)c̃1 + (m − 2)c̃m = 0. Since this must be true
for all integer values of m we conclude that c̃1 = 0, which in turn implies c̃m = 0
for all m ∈ Z. Thus, any two-cocycle on the bms3 algebra has vanishing components
c̃(pm , pn ) = 0. Finally, suppose that c is a two-cocycle on the bms3 algebra and let
us ask whether one can have c(jm , pn ) 6= 0. As in (6.46) we start by ensuring that
the cocycle c is rotation-invariant by adding to it a suitable coboundary. Consider
therefore the cocycle relation
The left-hand side can be interpreted as the differential of the one-cochain k = c(j0 , ·)
evaluated at (jm , pn ), while the right-hand side is the Lie derivative of c with respect
to j0 . Since the left-hand side is exact we know that the cohomology class of c is
left invariant by rotations; in particular we can add to c the differential db of the
one-cochain
i
b(jm ) ≡ 0 , b(pm ) ≡ c(j0 , pm ) ,
m
which is such that
Lj0 (c + db)(jm , pn ) = 0 . (9.74)
Chapter 9. Classical BMS3 symmetry 283
From now on we simply write c to denote c + db. Then, analogously to (6.50), eq.
(9.74) implies (m + n)c(jm , pn ) = 0 by virtue of the brackets (9.10). In particular we
can now write c(jm , pn ) = cm δm+n,0 and it only remains to find the coefficients cm . For
this we derive a recursion relation using the cocycle identity
which implies
(2m + 1)c−1 + (m − 1)c−m−1 + (m + 2)cm = 0 (9.75)
by virtue of the bms3 algebra (9.10). In particular we have c0 = 0 and c1 = −c−1 ,
which then gives cm = −c−m and the recursion relation (9.75) can be rewritten as
Using the fact that the Poisson algebra of surface charges (9.37) coincides with
the Kirillov-Kostant bracket (9.71), we conclude that the (covariant) phase space of
three-dimensional asymptotically flat gravity with BMS boundary conditions is a hy-
perplane c1 = 0, c2 = 3/G embedded in the space of the coadjoint representation of the
[ 3 group and endowed with its Kirillov-Kostant Poisson bracket. This observation
BMS
is the flat space analogue of the statement that the subleading components of an AdS
space-time metric contain one-point functions of the dual CFT stress tensor. As in
AdS, this observation should not come as a surprise. Indeed, the coadjoint representa-
tion of BMS3 was bound to appear in the transformation law of the momentum map
of the system, and it just so happens that this map is determined by the entries of the
metric (9.25). The truly surprising aspect of this observation is the fact that it is the
entries of the metric, and not some non-linear combinations thereof, that determine the
momentum map. In particular, as in AdS3 , the set of solutions (9.25) is a vector space.
In view of these results, one may ask whether the subleading components of asymp-
totically flat metrics can be interpreted as the components of the stress tensor of some
dual theory, similarly to AdS/CFT. The notion of “dual theory” appears to be elusive
in the asymptotically flat case, essentially because the metric becomes degenerate at
null infinity, but the question can be answered regardless of this complication. In-
deed, whatever the dual theory is, it must be such that its stress tensor transforms
under the coadjoint representation of the BMS3 group (generally with some non-zero
central charges), by virtue of the very nature of momentum maps. Accordingly, the
stress tensor T of any BMS3 -invariant theory is necessarily such that Tuu = p(ϕ) is a
supermomentum generating supertranslations, while Tuϕ = j(ϕ) is an angular super-
momentum generating superrotations.
port our claim that the functions (j, p) coincide with the components of a “dual” stress
tensor.
The fact that the phase space of flat gravity coincides with (a hyperplane in) the
coadjoint representation of BMS [ 3 allows us to use the orbits (9.77) as an organizing
principle. As in fig. 8.6, the space of solutions is foliated into disjoint BMS [ 3 orbits,
each of which is a symplectic manifold. Since the classification of coadjoint orbits of
[ 3 follows from the results of section 5.4, we may claim to control the full covariant
BMS
phase space of asymptotically flat gravity. In particular the classification of zero-mode
solutions in fig. 9.1 is a first step towards the full classification: each point in the plane
(J, M ) determines an orbit (9.77), and different points define disjoint orbits. Since
not all orbits have constant representatives, fig. 9.1 is an incomplete representation
of the full phase space of the system. The complete picture would involve the BMS3
analogue of fig. 7.3. Note that the relation between metrics and BMS [ 3 orbits hints
that the quantization of asymptotically flat gravity produces unitary representations
of BMS3 . We will investigate this proposal in chapters 10 and 11.
The answer follows from the fact that asymptotically flat metrics transform un-
der BMS3 according to the coadjoint representation (9.54). For our purposes the key
property of that formula is the fact that the transformation law of p is blind to super-
translations. In this sense the positive energy theorem in three-dimensional flat space
is even simpler than in AdS3 :
Chapter 9. Classical BMS3 symmetry 286
Positive energy theorem. The asymptotically flat metric (j, p) has energy boun-
ded from below under BMS3 transformations if and only if p belongs to a Virasoro
coadjoint orbit (at central charge c2 = 3/G) with energy bounded from below. This is
to say that either p is superrotation-equivalent to a constant p0 ≥ −c2 /24, or p belongs
to the unique massless Virasoro orbit with bounded energy.
As a corollary, we now know that all conical deficits and all flat space cosmologies
have energy bounded from below under BMS3 transformations. By contrast, all conical
excesses have energy unbounded from below.
In this section we explore this flat limit from the point of view of group theory,
starting from the definition of the AdS3 asymptotic symmetry group as a set of con-
formal transformations of a time-like cylinder. We describe the limit at the level of
groups, then at the level of Lie algebras, and finally at the level of the coadjoint repre-
sentation. We end by pointing out a different contraction that produces the Galilean
conformal algebra in two dimensions. Considerations related to flat limits of unitary
representations are relegated to section 10.2.
For instance, if F(ϕ) = ϕ + θ and F̄(ϕ) = ϕ + θ̄ are rotations, this condition says that
θ − θ̄ goes to zero at least as fast as 1/` in the large ` limit. With this choice the trans-
formation law of u reduces to u 7→ f 0 (ϕ)u + α(f (ϕ)). Including (9.79), we have thus
reproduced the BMS3 transformations (9.12) from a flat limit of Diff(S 1 )×Diff(S 1 ). In
this sense the centreless BMS3 group (9.58) is a flat limit of the asymptotic symmetry
group of AdS3 with Brown-Henneaux boundary conditions. In particular superrota-
tions arise in the form (9.80) while supertranslations (9.81) measure how fast F̄(ϕ)
Chapter 9. Classical BMS3 symmetry 288
Note that the condition (9.82) does not imply that there are less elements in the
BMS3 group than in the group Diff(S 1 ) × Diff(S 1 ). Indeed, both groups are infinite-
dimensional Lie groups consisting of two spaces of functions on the circle and have the
same cardinality in this sense.
and thus coincides with the leading non-radial components of the asymptotic Killing
vector field (9.17). The Lie brackets of such vector fields satisfy the centreless bms3
algebra; the latter is thus a flat limit of the direct sum of two Witt algebras. Note
that from this perspective the fact that supertranslations have dimensions of length
follows from the fact (9.86) that α(ϕ) is proportional to `.
The limit from Witt to bms3 can also be formulated in terms of commutation
−
relations. Indeed, let `m = eimx ∂+ and `¯m = eimx ∂− denote the generators of two
+
The terminology here is consistent with the fact that, on the cylinder, `0 − `¯0 generates
rotations while `0 + `¯0 generates time translations. In the basis (9.87), the commutation
relations of the direct sum of two Witt algebras take the form
1
i[jm , jn ] = (m − n)jm+n , i[jm , pn ] = (m − n)pm+n , (m − n)jm+n .
i[pm , pn ] =
`2
(9.88)
In the limit ` → +∞ the last bracket vanishes and the algebra reduces to (9.10),
reproducing bms3 as expected. The same argument can be applied to the direct sum
of two Virasoro algebras and gives rise to the centrally extended bms
d 3 algebra (see eq.
(9.93) below).
This observation can be used to define “flat limits” of Lie algebras in general terms.
Consider indeed the Lie algebra g ⊕ g, whose generators we denote ta and t̄a , with
identical commutation relations (5.2) in both sectors:
The limiting procedure that turns the sum of two Witt algebras into bms3 is an ex-
ample of Inönü-Wigner contraction [115], similar to the relation between the Poincaré
group and the Galilei group. Conversely, the direct sum of two Witt algebras is a
deformation of bms3 . The same construction can be used to show that the Poincaré
algebra is a flat limit of the AdS3 isometry algebra, so(2, 2) ∼
= sl(2, R) ⊕ sl(2, R). We
will return to this in section 10.2.
Chapter 9. Classical BMS3 symmetry 290
Using (9.85) and (9.86) one then verifies that, in the limit ` → +∞,
T (x+ )X (x+ ) + T̄ (x− )X̄ (x− ) = j(ϕ)X(ϕ) + p(ϕ)α(ϕ) + u(pX)0 (ϕ),
which coincides up to a total derivative with the integrand of the flat surface charge
(9.31). In other words the surface charges of flat space gravity are large ` limits of
those of AdS3 . In mathematical terms this is to say that the flat limit of the coadjoint
representation of the direct product of two Virasoro groups is the coadjoint represen-
tation of the (centrally extended) BMS3 group.
This phenomenon also allows us to relate the central charges of the Virasoro algebra
to those of bms
d 3 . (We could have done this in terms of abstract Lie algebra generators,
but for comparison with three-dimensional gravity we do it here in terms of coadjoint
vectors.) Let us consider two Virasoro algebras with central charges c and c̄ that
depend on ` as
c = A` + B + O(1/`), c̄ = A` + B̄ + O(1/`)
where A, B and B̄ are `-independent. Then the definitions
c + c̄
c1 ≡ lim (c − c̄), c2 ≡ lim (9.93)
`→+∞ `→+∞ `
allows us to write the flat limit of the algebra in the bms
d 3 form (9.71) in terms of
generators (jm , pm ) related to Virasoro generators (`m , `¯m ) by (9.87). This is the
centrally extended analogue of the flat limit described in (9.91). Note that for the
Brown-Henneaux central charges (8.40) the prescription (9.93) yields c1 = 0 and c2 =
3/G, which are indeed the standard values for asymptotically flat space-times.
Remark. The fact that flat space holography can be studied as a flat limit of the
AdS/CFT correspondence is an old idea [282, 283]; see also [284–286]. Here we have
described its group-theoretic formulation. It should be noted, however, that there is
no known limiting construction that yields BMS symmetry in four dimensions from
some corresponding asymptotic symmetry in AdS4 .
Chapter 9. Classical BMS3 symmetry 291
The redefinitions (9.90) suggest a contraction of Witt algebras that differs from
the flat limit (9.87). Namely, instead of performing the involution (9.92) before taking
the limit ` → +∞, one can define
1
j̃m ≡ `¯m + `m , p̃m ≡ (`¯m − `m ) . (9.94)
`
In contrast to (9.87), this redefinition has nothing to do with the flat limit of AdS3 ,
but the limit ` → +∞ still gives rise to an algebra with commutation relations (9.10)
upon renaming jm → j̃m and pm → p̃m . The key difference is that now the generator
of time translations is j̃0 (since it coincides with `0 + `¯0 ) while p̃0 generates rotations
(since it is proportional to `0 − `¯0 ). More generally, with the redefinition (9.94), the
generators of would-be supertranslations do not commute while those of would-be su-
perrotations do commute. This is the opposite of the behaviour of superrotations and
supertranslations in three-dimensional Einstein gravity.
At some point the isomorphism gca2 ∼ = bms3 led to the proposal that flat space
holography (in three space-time dimensions) is described by a Galilean conformal field
theory [288, 292]. In view of the geometric interpretation of superrotations and super-
translations described above, this sounds suspicious: the Galilean conformal algebra is
a version of the bms3 algebra “rotated by 90 degrees” where the roles of the Hamiltonian
and angular momentum are exchanged. In particular the flat limit of AdS3 /CFT2 , if it
exists, should not give rise to a Galilean conformal field theory since the gravitational
flat limit (9.87) of two Witt algebras gives rise to standard bms3 , in which p0 generates
time translations. Nevertheless, at the level of classical symmetries, there is essentially
no distinction between bms3 and gca2 ; the two are interchangeable. This coincidence
led to many publications concerned with flat space holography and attempting to
describe its dual theory as a Galilean conformal field theory; see e.g. [293–295] and
references therein. One of the goals of this thesis is to explain why the dual theory
Chapter 9. Classical BMS3 symmetry 292
This being said, we stress that discarding Galilean conformal field theories as pu-
tative duals for asymptotically flat gravity does not rule out all the conclusions of
the substantial literature on flat space holography approached from the Galilean side.
Rather, the point we wish to make is that those computations that did work in flat
space while relying on gca2 symmetry would have worked equally well in the language
of BMS3 symmetry. More precisely, any computation that holds for gca2 but does
not rely on its realization as a quantum symmetry algebra also holds for bms3 , and
therefore for asymptotically flat gravity.
Chapter 10
This chapter is devoted to irreducible unitary representations of the BMS3 group, i.e.
BMS3 particles, which we classify and interpret. As we shall see, the classification is
provided by supermomentum orbits that coincide with coadjoint orbits of the Virasoro
group. Upon identifying supermomentum with the Bondi mass aspect of asymptot-
ically flat metrics, we will be led to interpret BMS3 particles as relativistic particles
dressed with gravitational degrees of freedom.
The plan is as follows. In section 10.1 we classify BMS3 particles according to orbits
of supermomenta under superrotations. We also describe and interpret the resulting
Hilbert spaces of wavefunctions, which we relate to the quantization of (coadjoint)
orbits of asymptotically flat metrics under BMS3 . Section 10.2 is devoted to the de-
scription of BMS3 particles as representations of the (centrally extended) bms3 algebra
and their relation to highest-weight representations of the Virasoro algebra; we also
briefly touch upon Galilean representations. Finally, in section 10.3 we evaluate char-
acters of BMS3 particles. To lighten the notation, from now on the words “BMS3
group” or “bms3 algebra” implicitly refer to their centrally extended versions (except
if stated otherwise). We also abuse notation by writing Diff(S 1 ) to refer either to
Diff+ (S 1 ) or to Diff
g + (S 1 ), depending on the context.
Most of the results exposed in this chapter have been reported in the papers [46–
48, 51]. The relation between BMS3 particles and gravitational one-loop partition
functions [49, 50] will be described in the next chapter. Note that the considerations
that follow rely heavily on the material of chapter 4.
293
Chapter 10. Quantum BMS3 symmetry 294
The plan of this relatively long section is the following. We first describe the su-
permomentum orbits and little groups that classify BMS3 particles. We shall see that
these orbits are in fact coadjoint orbits of the Virasoro group, which will allow us to
define massive, massless and tachyonic BMS3 particles. We also discuss the existence
of integration measures on supermomentum orbits, since such measures are required
to define scalar products of wavefunctions. We then describe the states represented
by such wavefunctions and interpret them as particles dressed with quantized gravita-
tional degrees of freedom, in accordance with the relation between asymptotically flat
metrics and the coadjoint representation of BMS3 . We also apply this interpretation
to the vacuum representation and to spinning BMS3 particles, and we conclude by
discussing the extension of our considerations to four space-time dimensions.
Supermomentum orbits
The key ingredient in the description of BMS3 particles is the dual of the space of
d 1 )∗ . Following the terminology of section 9.2, its elements
supertranslations, Vect(S Ab
are centrally extended supermomenta
p(ϕ)dϕ2 , c2
(10.1)
paired with centrally extended supertranslations (α, λ) according to (6.111) with the
replacements X → α and c → c2 . As mentioned below (9.69), p(ϕ) has dimensions
of energy; its three lowest Fourier modes form a Poincaré energy-momentum vector
(in particular the zero-mode is the energy of p). More generally p(ϕ) is an energy
density on the circle while the central charge c2 is an energy scale. Supermomentum
transforms as a Virasoro coadjoint vector (9.70) under superrotations, so the allowed
supermomenta of a BMS3 particle span a coadjoint orbit of the Virasoro group at
central charge c2 . This is the first key conclusion of this section:
Chapter 10. Quantum BMS3 symmetry 295
When interpreting p(ϕ) as the Bondi mass aspect of an asymptotically flat metric
(9.25), the orbit Op is a subset of the orbit (9.77) of the metric under BMS3 transfor-
mations. In that context the central charge c2 coincides with the Planck mass (9.29).
Accordingly, from now on we restrict our attention to centrally extended supermo-
menta whose central charge c2 is strictly positive.
In this definition the word “generic” refers to the fact that p0 should not take
one of the discrete exceptional values −n2 c2 /24. Indeed the orbits containing such
exceptional constants are better thought of as BMS3 generalizations of the trivial rep-
resentation of Poincaré; we will return to this interpretation below.
In these definitions the terms “monodromy” and “winding number” refer to the
Virasoro invariants defined in section 7.1. They are the BMS3 generalization of the
mass squared in the Poincaré group.
Little groups
The little groups of BMS3 particles coincide with the stabilizers of the corresponding
Virasoro coadjoint orbits. Here, for comparison with the Poincaré little groups of
section 4.3, we list the little groups obtained by using the central extension of the
multiply connected BMS3 group (9.60). The list of orbits is that of section 7.2 and
their little groups are summarized in table 7.1:
Chapter 10. Quantum BMS3 symmetry 296
• For a massive BMS3 particle, the stabilizer is the group U(1) of spatial rotations.
• For a vacuum-like BMS3 particle whose supermomentum at rest takes the value
−n2 c2 /24, the little group is an n-fold cover of the Lorentz group in three di-
mensions, PSL(n) (2, R) (with n ≥ 1).
This list should be compared with table 4.1. The representations of these little groups
will lead to a notion of BMS3 spin. Note that when dealing with the universal cover
(9.61) of BMS3 , all compact directions of the above little groups get decompactified
so that U(1) is replaced by R, PSL(n) (2, R) is replaced by its universal cover, and Zn
is replaced by the group T2π/n ∼
= Z of translations of R by integer multiples of 2π/n.
Theorem. A BMS3 particle has energy bounded from below if and only if its super-
momenta span one of the Virasoro orbits coloured in red in fig. 7.7.
Recall that Poincaré particles with positive energy fall in exactly three classes,
two of which contain only one momentum orbit: massive particles, massless particles,
and the trivial orbit. The theorem tells us that essentially the same conclusion holds
for BMS3 particles, since all supermomentum orbits with energy bounded from below
belong to one of the three following classes:
• one of the massive orbits located above the vacuum and containing a constant
supermomentum p0 > −c2 /24,
From now on, when referring to BMS3 particles we always implicitly refer only to
particles with energy bounded from below (except if explicitly stated otherwise). Note
that, in contrast with Virasoro representations, BMS3 particles with unbounded energy
may provide unitary representations of BMS3 . Furthermore the energy spectrum of
any BMS3 particle is continuous.
Defining mass
The starting point is the observation that the vacuum supermomentum is pvac =
−c2 /24, while the supermomentum at rest of any massive BMS3 particle is located
above that vacuum value.
Intepreting supermomentum
In order to develop our intuition about the supermomentum vector p(ϕ), it is use-
ful to rewrite standard Poincaré momenta in terms of functions on the circle. The
supermomentum of a Poincaré particle with mass M typically takes the form
q c2
p(ϕ) = M 2 + p2x + p2y + px cos ϕ + py sin ϕ − (10.4)
24
Chapter 10. Quantum BMS3 symmetry 298
and represents a particle moving in space with a spatial momentum (px , py ). Note the
extra factor −c2 /24 at the end, which ensures that M actually coincides with the mass
of the particle.
One can then use (9.70) to act with superrotations on (10.4) and obtain various
boosted momenta. In particular,
p 2 Lorentz transformations take the form (6.88) and
act on the components ( M + px + p2y , px , py ) according to the vector representation.
2
cosh(γ/2)eiϕ + sinh(γ/2)
eif (ϕ) = (10.5)
sinh(γ/2)eiϕ + cosh(γ/2)
where γ is the rapidity (in terms of standard velocity, γ = arctanh(v)). Since this
superrotation is of the projective form (6.88), its Schwarzian derivative satisfies (6.94)
and the corresponding transformation (9.70) of the supermomentum p can be rewritten
as
c2 1 h c2 i
f · p (f (ϕ)) + = 0 p(ϕ) + . (10.6)
24 (f (ϕ))2 24
This says that the combination p + c2 /24 transforms under Lorentz transformations
as a centreless coadjoint vector of Diff(S 1 ). In particular, the supermomentum of a
massive particle at rest transforms according to (10.6) with p(ϕ) = M − c2 /24. As a
result, since f 0 (ϕ) is given by (7.101), the energy E[γ] of the boosted particle is
M 2π dϕ
Z
E[γ] = = M cosh γ
2π 0 f 0 (ϕ)
M 2π dϕ
Z
px [γ] = cos(f (ϕ)) = M sinh γ.
2π 0 f 0 (ϕ)
The boosted spatial momentum along the y direction vanishes, as it should. This
confirms that Lorentz transformations act on a supermomentum at rest exactly as in
standard special relativity (albeit in three space-time dimensions).
From these considerations we can now draw a general conclusion on the supermo-
mentum of a (massive) BMS3 particle. Typically, the function p(ϕ) will have some
non-trivial profile on the circle; for instance the momentum of a particle moving fast
in the x direction is represented by a function p(ϕ) which is larger than −c2 /24, has a
bump around the point ϕ = 0, and almost vanishes in the neighbourhood of the oppo-
site point ϕ = π. If the particle is obtained by a pure Poincaré boost from a particle
at rest, the only non-vanishing components of its supermomentum are its three lowest
Fourier modes, p0 , p1 , p−1 (as in eq. (10.4)). Upon switching on superrotations, the
supermomentum of the particle acquires extra Fourier modes (p2 , p3 , etc.) that dress
the original Poincaré momentum with additional fluctuations. In the upcoming pages
we will interpret these extra degrees of freedom as being of gravitational origin.
Chapter 10. Quantum BMS3 symmetry 299
(a) (b)
Figure 10.1: Two possible supermomenta of a massive BMS3 particle. In (a) the
supermomentum is that of a boosted Poincaré particle, given by eq. (10.4); it is a
function on the circle located above the line −c2 /24 and its only non-zero Fourier
modes are the three lowest ones. In (b) the function is dressed with extra non-vanishing
Fourier modes, which results in more wiggles. These extra Fourier modes account for
gravitational degrees of freedom that do not appear in the pure Poincaré case.
It is worth stressing the crucial importance of c2 for the conclusions of the pre-
vious pages. For one thing, the whole classification of supermomentum orbits and
the ensuing definition of massive/massless BMS3 particles only makes sense because
c2 is non-zero. If c2 happened to vanish, none of these results would hold since the
corresponding supermomentum orbits would be Virasoro coadjoint orbits at vanishing
central charge, and we saw in section 7.1 that these orbits are radically different from
(and arguably much uglier than) their centrally extended peers. This is not to say that
c2 must be non-zero in order for the supermomentum (10.1) to yield a representation
of the BMS3 group; in principle, representations associated with orbits having c2 = 0
are just as acceptable as representations in which c2 6= 0. However the application to
gravity, and the ensuing interpretation of representations as particles, relies crucially
on the fact that c2 = 3/G does not vanish. Note that the change of monodromy
occurring at M = c2 /24 suggests that something radical happens with BMS3 particles
whose mass is higher than that bound. This bifurcation reflects the fact that the met-
ric of the gravitational field surrounding the particle changes from that of a conical
deficit (when M < 1/8G) to that of a flat cosmology (when M > 1/8G).
Chapter 10. Quantum BMS3 symmetry 300
If the answer is affirmative, then the measure is a functional one since Op consists of
functions on the circle.
A conjecture
Our viewpoint regarding the problem (10.7) will be pragmatical: path integral mea-
sures are used on a daily basis in quantum mechanics, and their efficiency in correctly
predicting the values of physical observables is firmly established. Thus, if one is
willing to define Hilbert spaces of square-integrable functions thanks to functional
measures, their application to BMS3 particles is as acceptable as in quantum physics.
In particular one may hope that Virasoro coadjoint orbits do admit quasi-invariant
measures:
Conjecture. Let Op be a Virasoro coadjoint orbit with non-zero central charge (and
energy bounded from below). Then there exists a Borel measure dµ(q) on Op which
is quasi-invariant under the action of the Virasoro group (where q ∈ Op ).
In the remainder of this thesis we will rely on this conjecture in order to define the
Hilbert space of a BMS3 particle (at least one with bounded energy). The conjecture
does not say how the measure dµ(q) is actually defined, but this is not a problem
since the results of section 3.2 imply that induced representations based on different
quasi-invariant measures are unitarily equivalent. Thus, assuming that the conjecture
is true, we do not really need to know anything specific about the measure.2 In par-
ticular the character computation of section 10.3, although relying on an unknown
measure, will produce an unambiguous result.
Aside from these basic observations, we will have very few concrete things to say
about the measure. Nevertheless the lines that follow are devoted to a brief review of
the literature on Virasoro measure theory, with the intent of further motivating the
validity of the conjecture. The reader who is not interested in mathematical subtleties
is free to go directly to section 10.1.4.
1
We recall that the definition of quasi-invariant measures was given in section 3.2.
2
Note that the existence of a quasi-invariant measure dµ(q) on Op implies the existence of infinitely
many other ones, since one can always multiply the measure by a strictly positive smooth function
ρ(q) and obtain a new measure ρ(q)dµ(q).
Chapter 10. Quantum BMS3 symmetry 301
The main motivation for defining Virasoro measures comes from representations
of the Virasoro algebra and conformal field theory; the hope is that such measures
could provide a rigorous prescription for the geometric quantization of Virasoro or-
bits. This approach to the problem is adopted for instance in [230, 231, 298, 299].
In [300] the authors tackle the issue with a similar motivation, though with different
methods; in particular it is suggested there that a measure might be provided by an
infinite-dimensional version of the Liouville measure (5.14) obtained by taking the
Kirillov-Kostant symplectic form (8.62) to an infinite power.
conjecture.
The space of states of a BMS3 particle carries an action of BMS3 by unitary trans-
formations. Since we are assuming that the particle is scalar, the action of BMS3 on
H is given by formula (4.23):
q
T [(f, α)] · Ψ (q) = ρf −1 (q) eihq,αi Ψ(f −1 · q) .
(10.10)
Here (f, α) is an element of the BMS3 group (9.58), with f (ϕ) a superrotation and
α(ϕ) a supertranslation. The point q ∈ Op is a supermomentum vector and its pairing
hq, αi with α is given by (6.34). The function ρf (q) is the (unknown) Radon-Nikodym
derivative (3.19) of the measure µ; if by chance the measure happens to be invariant,
then one can set ρf (q) = 1. Finally, the action f · q appearing in the argument of the
wavefunction on the right-hand side is the BMS3 generalization of the action of boosts
on momenta; it is given by formula (9.70), which is the coadjoint representation of the
Virasoro group at central charge c2 .
Chapter 10. Quantum BMS3 symmetry 303
Remark. Wavefunctionals are common in quantum field theory. Indeed, the quan-
tum state of a typical field theory is a wavefunctional Ψ[φ(x)], where φ(x) is a spatial
field configuration. The truly striking aspect of (10.8) is not quite the fact that it
belongs to a space of wavefunctionals, but rather that it provides an irreducible rep-
resentation of the symmetry group.
where the two lines differ by a term proportional to hS[f ], αi, with S the Schwarzian
derivative (6.76). This phenomenon is analogous to the statement that boosts and
translations do not commute in the Bargmann group (4.103). In practice it means
that the wavefunction obtained by acting first with (e, α), then by (f, 0), differs from
the one obtained by acting first with (f, 0), then with (e, Adf α), by a constant complex
phase that can be evaluated using (9.55):
h c i
2
T [(f, 0)] · T [(e, α)] = exp −i hS[f ], αi T [(e, Adf α)] · T [(f, 0)] . (10.12)
12
This is indeed the statement (2.7) that the representation T is projective, when seen
as a representation of the centreless BMS3 group (9.58). It is the BMS3 analogue of
the Galilean result (4.121).
Plane waves
As in section 3.3 we can describe the representation (10.10) in terms of a basis of
one-particle states with definite (super)momentum on the orbit Op . Let therefore
δ denote the Dirac distribution associated with the measure µ and defined by the
requirement (3.39). In the present case µ is a functional measure, so δ is a functional
delta distribution. For k ∈ Op , we define the plane wave state with supermomentum
k as
Ψk (q) ≡ δ(k, q) (10.13)
which is now a functional analogue of eq. (3.43). It is a typical asymptotic state in a
scattering experiment. The scalar products of plane waves are given by (3.44):
Strictly speaking, plane waves are not square-integrable, hence do not belong to the
space of states of a BMS3 particle. They should therefore be understood in the weaker
sense that any wavepacket (3.45) can be written as an infinite sum of plane waves.
With this word of caution, one may say that plane waves form a “basis” of the space
of states of a BMS3 particle. Their transformation law under BMS3 transformations
is given by eq. (4.32),
q
T [(f, α)] · Ψk = ρf (k) eihf ·k,αi Ψf ·k , (10.15)
except that we have removed the spin representation R since the particle considered
here has vanishing spin. This formula reflects the fact that a wavefunction with mo-
mentum k boosted by a superrotation f becomes a wavefunction with momentum f ·k.
In short, all results of chapters 3 and 4 remain valid, up to the fact that manifolds
become spaces of functions while functions become functionals.
Leaking wavefunctions
The transformation law (10.10) is an infinite-dimensional generalization of a scalar
representation (with mass M ) of the Poincaré group in three dimensions. Indeed, by
restricting one’s attention to the Poincaré subgroup of BMS3 , one obtains a (highly
reducible) unitary representation of Poincaré. The latter contains the standard scalar
irreducible representation with mass M , but it also contains an uncountable infinity
of other representations with higher mass. These extra representations arise because
the action of Lorentz transformations on the supermomentum orbit Op ∼ = Diff(S 1 )/S 1
is not transitive; in fact, the set of Lorentz-inequivalent supermomenta is an infinite-
dimensional manifold
PSL(2, R)\Diff(S 1 )/S 1 , (10.16)
which is a double quotient of Diff(S 1 ). The quotient on the left is taken with respect
to the Lorentz group PSL(2, R).
This observation can be rephrased in a more intuitive way: it says that the super-
momentum orbit Op = Diff(S 1 )/S 1 contains infinitely many finite-dimensional sub-
manifolds SL(2, R)/S 1 obtained by acting on the orbit with Lorentz transformations.
Each submanifold is a standard momentum orbit (4.97) for massive Poincaré particles
in three dimensions. Any two of those submanifolds are mutually Lorentz-inequivalent;
the set of such inequivalent sub-orbits is the space (10.16).
Figure 10.2: A supermomentum orbit Op ∼ = Diff(S 1 )/S 1 with embedded Lorentz sub-
orbits represented as curvy lines. Lorentz transformations move points along these sub-
orbits. Transitions from one sub-orbit to another are only possible with superrotations
that do not belong to the Lorentz subgroup of Diff(S 1 ). One can define an equivalence
relation on the supermomentum orbit by declaring that two points are equivalent if
they belong to the same Lorentz sub-orbit. The quotient of Op by that equivalence
relation is the double quotient (10.16).
Figure 10.3: The supermomentum orbit of fig. 10.2, now with a wavefunction on top.
The wavefunction is roughly supported on a Lorentz sub-orbit, but not quite: it leaks
into directions that cannot be achieved with Lorentz transformations.
Gravitational dressing
We have just explained that any supermomentum orbit contains infinitely many Poin-
caré-inequivalent sub-orbits. It is natural to wonder how the extra directions of the
orbit — those that do not lie along Lorentz generators — are to be interpreted. In
other words: how should one think of the fact that the wavefunction of a BMS3 par-
ticle leaks into directions that are forbidden by Poincaré transformations?
A natural guess is suggested by the very origin of the BMS3 group: as we showed
in section 9.1, it is the asymptotic symmetry group of Minkowskian space-times (in
three dimensions). Asymptotic symmetries may be thought of as generalizations of
isometries that incorportate gravitational fluctuations. Now, the isometry group of
Minkowski space-time is the Poincaré group, and its unitary representations are par-
ticles in the usual sense. Accordingly,
A BMS particle is a Poincaré particle
(10.17)
dressed with gravitational degrees of freedom.
relation ∼ such that Ψ ∼ Ψ0 if there exists a Poincaré transformation (f, α) for which
Ψ0 = T [(f, α)]Ψ. Then the space of Poincaré-inequivalent states is the quotient
H/∼ . (10.18)
The latter can also be seen as the set of quantum states obtained by acting on a
state at rest with supertranslations and superrotations that do not belong to the
Poincaré subgroup. In this sense it is the three-dimensional analogue of the space of soft
gravitons in four dimensions, as follows from the recently discovered relation [19, 247]
between asymptotic symmetries and soft theorems in gauge theories. Thus,
A BMS particle is a particle dressed with soft gravitons.
Note that the terminology of “soft gravitons” is a bit dangerous here, since three-
dimensional Einstein gravity has no local degrees of freedom, hence no genuine (bulk)
gravitons. We already pointed out this subtlety in the introduction of the thesis, and
our point of view remains the same: owing to the relation between soft gravitons and
asymptotic symmetries, any theory with non-trivial asymptotic symmetries can be
interpreted as a theory containing soft degrees of freedom, regardless of the presence
of bulk degrees of freedom. In this sense three-dimensional gravity is a toy model for
soft gravitons.
L2 (M × N , µM × µN ) ∼
= L2 (M, µM ) ⊗ L2 (N , µN ). (10.21)
Let us use this result to compare massive particles, with supermomentum orbits
Diff(S 1 )/S 1 , to the vacuum whose orbit is Diff(S 1 )/PSL(2, R). Since both of these
infinite-dimensional manifolds are homotopic to a point, we can relate them as
Diff(S 1 )/S 1 ∼
= PSL(2, R)/S 1 × Diff(S 1 )/PSL(2, R) .
(10.22)
This is to say that the massive supermomentum orbit is a direct product M × N . The
first factor of the product is the Poincaré momentum orbit (4.97) of a massive particle
in three dimensions, which suggests that a BMS3 particle is equivalent to a relativistic
particle “times” the vacuum BMS3 representation. This can be made precise using
(10.21): if the measure µ used to define the scalar product of wavefunctions for a mas-
sive BMS3 particle factorizes into a product on Diff(S 1 )/PSL(2, R) and PSL(2, R)/S 1 ,
then the Hilbert space of a massive BMS3 particle factorizes into
HBMS ∼
= HPoinc ⊗ Hvac (10.23)
where HPoinc is the space of states of a massive Poincaré particle and Hvac is that of
the BMS3 vacuum representation.
One can reformulate the statement (10.23) in the basis of plane wave states (10.13).
Indeed, on a massive supermomentum orbit Op = Diff(S 1 )/S 1 , the diffeomorphism
(10.22) allows us to write any supermomentum q as a pair q = (qPoinc , qvac ) where
qPoinc is a momentum vector with three components belonging to the Poincaré sub-
orbit OPoinc = PSL(2, R)/S 1 , while qvac is a supermomentum that belongs to the
vacuum BMS3 orbit Ovac = Diff(S 1 )/PSL(2, R). Then, under the assumption that the
measure µ on Op disintegrates into a product of measures on PSL(2, R)/S 1 and Ovac ,
any plane wave (10.13) can be written as a tensor product
Ψk = ΨkPoinc ⊗ Ψkvac
Chapter 10. Quantum BMS3 symmetry 309
where the left and right factors of the product are plane waves on the orbits OPoinc and
Ovac , respectively. A generic state of a BMS3 particle is an infinite linear combination
of such factorized plane waves. Note that none of our upcoming conclusions rely on
this phenomenon, so in the sequel we will not necessarily assume that the measure µ
disintegrates into a product.
We recall from (4.28) that spin is the label that specifies the representation of
the little group chosen for the description of a particle. In the case of BMS3 parti-
cles at non-zero c2 , the little groups were described at the end of subsection 10.1.1.
They are all either one-dimensional Abelian groups such as U(1) or R (possibly up
to discrete factors), or n-fold covers of the Lorentz group PSL(2, R). All these are
finite-dimensional Lie groups and their unitary representations are known, so writing
down generic spinning representations of BMS3 is mostly a technical problem.
For definiteness, let us consider a massive BMS3 particle. Its little group U(1) con-
sists of spatial rotations, exactly as for massive Poincaré particles in three dimensions.
All irreducible unitary representations of U(1) are of the form (2.13), with s an integer.
However, the BMS3 group (9.58) has the same homotopy type as Diff(S 1 ), which is
homotopic to a circle. This implies that it admits topological projective representa-
tions of the type described in section 2.1, which can be classified by considering exact
(non-projective) representations of the universal cover (9.61) of BMS3 . In the latter
case, the little group of massive particles gets unwrapped from U(1) to R, whose uni-
tary representations are now labelled by an arbitrary real spin s. Since those are the
physically relevant representations, we conclude that the spin of a massive BMS3 par-
ticle is generally an arbitrary real number. In particular, most massive BMS3 particles
are anyons. This is the same conclusion as in the Poincaré group in three dimensions.
Now suppose we fix a certain value of mass M and spin s ∈ R and ask what are
the states of the corresponding BMS3 particle. We denote the spin s representation of
the little group by R; it is a one-dimensional representation of the form (2.13). Thus
the Hilbert space of the BMS3 particle consists again of complex-valued wavefunctions
(10.8), but their transformation law under BMS3 contains an extra term with respect
Chapter 10. Quantum BMS3 symmetry 310
to the scalar representation (10.10). That extra term involves the Wigner rotation
(4.31) associated with the superrotation f and the supermomentum q,
Wq [f ] = R[gq−1 f gf −1 ·q ] , (10.24)
where the superrotations gq (q ∈ Op ) are standard boosts such that gq · p = q. Thus we
run into the problem of finding standard boosts for a massive supermomentum orbit.
Incidentally we have already defined such standard boosts, though not in the same
language. Indeed, eq. (7.44) is precisely the definition of standard boosts for elliptic
Virasoro coadjoint orbits. These boosts are built as follows:
1. Take the supermomentum q(ϕ) with central charge c2 ; write down the associated
Hill’s equation (7.12) with the replacement (p, c) → (q, c2 ).
2. Find two linearly independent solutions ψ1 , ψ2 of Hill’s equation satisfying the
Wronskian condition (7.16).
3. Define a vector field Xq (ϕ) by (7.38).
4. Define a diffeomorphism f of S 1 by (7.41), with p0 = M − c2 /24
5. The standard boost associated with q is gq = f −1 .
This procedure is somewhat convoluted, but it does provide a family of standard boosts
on the orbit of a massive supermomentum, as desired. We refrain here from actually
computing these boosts.
Equipped with standard boosts one can write down the transformation law of
massive BMS3 particles with non-zero spin, given by eq. (4.30):
q
T [(f, α)] · Ψ (q) = ρf −1 (q) eihq,αi Wq [f ] · Ψ(f −1 · q) ,
where the notation is the same as in (10.10) up to the insertion of the Wigner rotation
(10.24). This can also be rewritten in terms of plane waves (10.13) as
q
T [(f, α)] · Ψk = ρf (k) eihf ·k,αi R gf−1
·k f gk · Ψf ·k
(recall eq. (4.32)). The interpretation of all these formulas is the same as before, up
to the extra Wigner rotation. In particular, a spinning BMS3 particle is a spinning
Poincaré particle dressed with soft gravitons.
The first suggestion that BMS representations might be relevant to particle physics
appeared in [3]. A little later McCarthy and collaborators set out to study the rep-
resentation theory of the global BMS group in full detail, with scattering amplitudes
as a motivation [41]. Owing to the semi-direct product structure of (1.1) the strategy
was to build induced representations à la Wigner in terms of orbits and little groups,
as described in this thesis in chapter 4. In particular it was shown in [38, 39] that all
little groups are compact, leading to the conclusion that BMS particles in four dimen-
sions cannot have continuous spin, in contrast to their Poincaré counterparts. This
was followed by the observation that the restriction of a BMS representation to its
Poincaré subgroup is reducible and consists of a tower of Poincaré particles with dif-
ferent spins [40]; in [309] this spin mixing was interpreted as being due to the presence
of the gravitational field. It was also shown in [310] that certain BMS representations
studied earlier in [311, 312] were in fact reducible induced representations.
Along the way it was realized that the absence of continuous-spin particles exhib-
ited in [38, 39] was due to a delicate choice of topology, and that different topologies
lead to radically different conclusions, including particles with continuous spin [313].
These continuous-spin particles were then interpreted as scattering states in [269,314].
Finally, the whole construction was put on firm mathematical ground in [315], where
the theory of induced representations was extended to groups of the type G n A with
an infinite-dimensional Abelian group A. As a corollary, it was shown in [316] that
induced representations à la Wigner exhaust all irreducible unitary representations of
the global BMS group (1.1). This analysis was later completed by the proof that the
global BMS group has no non-trivial central extensions, and therefore admits no pro-
jective representations other than those originating from its non-trivial topology [317].
Along the way, it was realized by Strominger and collaborators that BMS symmetry
does have highly non-trivial implications for the S-matrix, in the form of soft graviton
theorems [19, 24, 26, 27, 253]. This discovery sparked a flurry of papers discussing the
applications of the BMS group (and its gauge-theoretic generalizations [247–249, 251,
255–258]) to scattering amplitudes [348, 349], memory effects [250, 252, 254] and more
recently to black holes [350].
BMS4 particles?
Despite recent progress, we still seem to be quite far from having truly understood
BMS symmetry in four dimensions. Representation theory provides an easy way to
illustrate the problem. Indeed, one of the cornerstones of the relation between BMS
symmetry and soft theorems is the fact that supertranslations generate soft graviton
states when acting on the vacuum. Accordingly, if the global BMS group (1.1) is cor-
rect, then the representations considered by McCarthy in [38, 39] should account for
this effect: they should represent Poincaré particles dressed with soft gravitons.
However, it is easy to see that this is not the case. To illustrate this point, consider
the analogue of the global BMS group (1.1) in three dimensions,
In particular the Hilbert space of any irreducible unitary representation of the group
(10.25) coincides with the space of states of a Poincaré particle; it consists of wave-
functions on a finite-dimensional momentum orbit. In fact, the only difference between
the representations of this group and those of Poincaré would be that translations are
paired with supermomenta according to the functional formula (6.34) rather than a
finite-dimensional product pµ αµ . In particular the vacuum representation would be
trivial and there would be no way for quantum supertranslations to create soft gravi-
ton states upon acting on the vacuum.
These observations exhibit an important point: the global BMS group (1.1) cannot
be the end of the story. It is too small to account for soft graviton degrees of freedom,
and it must be extended in some way. Unfortunately the argument does not tell us
how BMS symmetry should be extended. To our knowledge, two proposals have been
formulated so far, both suggesting a superrotational extension of the Lorentz group.
The first is the aforementioned idea of turning superrotations into a semi-group of
local conformal transformations of celestial spheres [5, 6, 18, 336, 351]; the second sug-
gests that superrotations should instead span a group Diff(S 2 ) of diffeomorphisms of
the sphere [352, 353]. It appears that there are currently no definitive arguments for
selecting one proposal over the other; see however [354], where it is argued that finite
singular conformal transformations of celestial spheres in four dimensions are patho-
logical.
The fact that BMS symmetry in four dimensions is ill-defined is a call for further
developments. This thesis is one of them: it aims at understanding a three-dimensional
toy model and using it as a guide for the realistic problem. Indeed, many properties
that we have encountered in our investigation of BMS3 should remain true in BMS4 .
In particular the semi-direct product structure G n A appears to be a robust feature
and implies that BMS particles are classified by supermomentum orbits that coincide
with orbits of the Bondi mass aspect under asymptotic symmetry transformations.
Furthermore the occurrence of a central extension pairing superrotations with super-
translations has also been observed in BMS4 [18]. However, a sharp difference between
BMS3 and BMS4 is that, in the former, supertranslations do not create new states when
acting on the vacuum. A possibly related difference is that the language suited to the
study of BMS4 appears to be that of groupoids rather than groups. We will not have
much more to say about this here, and return now to our study of BMS3 .
same construction to the bms3 algebra. The presentation is adapted from [51].
Poincaré algebra
In three dimensions, the Lie algebra of the Poincaré group is spanned by three Lorentz
generators jm and three translation generators pm (m = −1, 0, 1) with Lie brackets
(9.10). As in the case (8.54) of sl(2, R), it is more convenient to use a different
complexified basis Jm = ijm , Pm = ipm , in terms of which the brackets become
These conventions are such that, in any unitary representation, the operators repre-
senting Poincaré generators satisfy Hermiticity conditions of the type (8.56):
As in section 8.4, we abuse notation by denoting with the same letter both the abstract
generators Jm , Pn and the operators that represent them.
The Poincaré algebra has two quadratic Casimir operators: the mass squared
Induced modules
Irreducible unitary representations of the Poincaré group are obtained by considering
the Lorentz orbit of a momentum p and building a Hilbert space of wavefunctions on
that orbit. A basis of this space is provided by plane waves (10.13), where the Dirac
distribution is determined by the choice of measure on the orbit. (For instance one
can take the Lorentz-invariant measure (3.4).) Their transformation laws are given
by (10.15). Here, in order to make the link with the standard Dirac notation, we
denote such plane waves by Ψk ≡ |k, si for any k ∈ Op , where s is the spin of the
representation.
Chapter 10. Quantum BMS3 symmetry 315
For future comparison with bms3 , we focus on a relativistic particle with mass
M > 0. Its little group U(1) consists of spatial rotations generated by J0 . If we call p
the momentum of the representation in the rest frame, then the corresponding plane
wave Ψp ≡ |M, si satisfies
P0 |M, si = M |M, si , P−1 |M, si = P1 |M, si = 0 , J0 |M, si = s|M, si . (10.30)
From now on we call |M, si the rest frame state of the representation. Any other plane
wave Ψk = |k, si with boosted momentum k ∈ Op can be obtained by acting on |M, si
with a Lorentz transformation gk , where gk is a standard boost. In this sense the rest
frame state determines all the properties of the representation, in the same way that
highest-weight representations are determined by their highest-weight state. Note,
however, that the conditions (10.30) that define |M, si are not of the same form as the
highest-weight conditions (8.57) in that they involve both positive and negative modes.
Let us now understand how the conditions (10.30) induce a representation of the
Poincaré algebra. They define a one-dimensional representation of the subalgebra gen-
erated by {Pm , J0 }. This subalgebra consists of infinitesimal translations and spatial
rotations, i.e. it is a semi-direct sum u(1) A R3 where u(1) is generated by J0 while
R3 is generated by the Pm ’s. Thus the prescription (10.30) is a Lie-algebraic version
of the spin representation (4.28) for the case of a little group U(1) with R[θ] = eisθ .
Guided by our experience of induced representations, we can attempt to induce a
representation of the full Poincaré algebra out of the one-dimensional representation
(10.30); the result is known as an induced module (see e.g. section 10.7 of [355]). We
thus declare that the carrier space H of the representation is spanned by all states
obtained by acting on the rest frame state with operators that do not appear in the
conditions (10.30):
|k, l i = (J−1 )k (J1 )l |M, si , (10.31)
where k, l are non-negative integers. Such states are infinitesimally boosted states
analogous to the descendant states (8.58) that span Verma modules for the Virasoro
algebra. By definition, they form a basis of the space H . The latter provides a
Poincaré representation as it should, since acting from the left on the states (10.31)
yields linear operators on H whose commutators coincide with (10.26). Moreover, the
Casimir operators (10.28) and (10.29) have the same eigenvalue on each state (10.31),
since they commute by construction with all elements of the algebra. This readily
implies that the representation is irreducible.
Note that unitarity is not obvious in this picture: if one did not know that the in-
duced module follows from a manifestly unitary representation of the Poincaré group
in terms of wavefunctions, there would be no straightforward way to define a scalar
product on the space H spanned by the states (10.31), even after enforcing the Her-
miticity conditions (10.27). In fact, the norm squared of any plane wave state is strictly
infinite because of the delta function in (10.14). This is strikingly different from the
highest-weight representations of section 8.4, where the highest-weight conditions were
enough to evaluate the norm squared (8.59) of all descendant states.
Remark. The definition of the infinitesimally boosted states (10.31) follows from
the general construction of induced modules, as follows. Let g be a Lie algebra with
Chapter 10. Quantum BMS3 symmetry 316
The Poincaré algebra (10.26) can be recovered from sl(2, R) ⊕ sl(2, R) as a flat
limit of the type described in section 9.4. Thus we introduce a length scale ` (to be
identified with the AdS radius) and define new generators as in (9.87):
1
Jm ≡ Lm − L̄−m , Pm ≡ (Lm + L̄−m ) . (10.32)
`
The resulting algebra is (9.88) without i’s on the left-hand side, and its limit ` → +∞
reproduces the Poincaré algebra (10.26). In addition the quadratic Casimir (8.60) can
be combined with its barred counterpart C, ¯ producing
2 ¯ = M2 + O(`−2 ), 1
C − C¯ = S,
C + C (10.33)
`2 `
where M2 and S are the Poincaré Casimirs (10.28) and (10.29).
The matching of Casimir operators suggests that the contraction also relates Poin-
caré modules to sl(2, R) ⊕ sl(2, R) representations. Concretely, consider the tensor
product of two highest-weight representations (8.57) of sl(2, R) with weights h, h̄:
L1 |h, h̄i = L̄1 |h, h̄i = 0 , L0 |h, h̄i = h|h, h̄i , L̄0 |h, h̄i = h̄|h, h̄i . (10.34)
h + h̄
M≡ , s ≡ h − h̄ , (10.35)
`
Chapter 10. Quantum BMS3 symmetry 317
h + h̄
P0 |h, h̄i = |h, h̄i , J0 |h, h̄i = (h − h̄)|h, h̄i (10.36)
`
in terms of operators (10.32). Similarly, in terms of J’s and P ’s, the condition that
L1 and L̄1 annihilate the highest-weight state becomes
1
P±1 ± J±1 |h, h̄i = 0 . (10.37)
`
This allows us to reformulate the whole representation of sl(2, R) ⊕ sl(2, R) in terms
of operators Jm , Pn ; it results in expressions of the form
X (n) X (n)
Pn |k, li = Pk0 ,l0 ; k,l (M, s, `)|k 0 , l0 i , Jn |k, li = Jk0 ,l0 ; k,l (M, s)|k 0 , l0 i (10.38)
k0 ,l0 k0 ,l0
where the states |k, li take the form (10.31) with the identification |M, si ≡ |h, h̄i,
while P(n) and J(n) are infinite matrices. Owing to the definition (10.32) and property
(10.37), only negative powers of ` appear in (10.38). It follows that the matrix elements
(n) (n)
Pk0 ,l0 ; k,l and Jk0 ,l0 ; k,l have a well-defined limit ` → ∞. This limit coincides with the
result that one would find in a Poincaré module spanned by states (10.31), provided
that the conformal weights scale as
M` + s M` − s
h= + λ + O(1/`), h̄ = + λ + O(1/`), (10.39)
2 2
where λ is an arbitrary parameter independent of `. Thus, in the flat limit, the sl(2, R)
highest-weight conditions (10.37) are turned into a rest frame condition (10.30) and
the Poincaré Casimirs M2 and S take the values M 2 and M s, respectively. In short,
Poincaré modules are flat limits of sl(2, R) ⊕ sl(2, R) modules.
Note again that unitarity is subtle: starting from scalar products of sl(2, R) ⊕
sl(2, R) descendants, the flat limit gives rise to scalar products of states (10.31) that
diverge like positive powers of `. Indeed the norms (8.59) diverge when ` → +∞, ow-
ing to the fact that h is proportional to ` in (10.39). Equivalently, the wavefunctions
corresponding to states (10.31) become (derivatives of) delta functions in the flat limit;
from this point of view ` is an infrared regulator. Nevertheless, upon recognizing these
divergent scalar products as delta functions (10.14), one concludes that the Poincaré
module is a unitary representation in disguise.
can be identified with a quotient of U(sl(2, R)) ⊗ C as discussed in the remark of page
315. The main difference with respect to Poincaré is the splitting of the algebra as
n− ⊕ h ⊕ n+ , where n± are nilpotent subalgebras, which allows one to evaluate scalar
products by enforcing the Hermiticity conditions (8.56).
4
Recall that we use the same notation for both the bms3 algebra and its central extension.
5
I am indebted to Axel Kleinschmidt for this observation.
Chapter 10. Quantum BMS3 symmetry 319
Aside from the mass operator (10.3), any function of the BMS3 central charges
c1 , c2 is clearly a Casimir. To our knowledge, whether this list exhausts all possible
bms3 Casimirs is an open question, though it seems plausible that it does since the
only Casimirs of the Virasoro algebra are functions of its central charges [228]. In
particular, it is not clear whether there exists a bms3 Casimir whose value specifies
the spin of a BMS3 particle, analogously to the Poincaré combination (10.29).
We now describe induced modules for the bms3 algebra (10.41), built analogously
to the Poincaré modules above and classified by their mass and spin (and central
charges). We discuss separately generic massive modules and the vacuum module,
and end by showing how they can all be obtained as ultrarelativistic limits of highest-
weight representations of Virasoro algebras.
Massive modules
Consider a BMS3 particle with mass M > 0 and spin s. Its supermomentum orbit
contains a constant p = M − c2 /24; the corresponding plane wave state Ψp ≡ |M, si
is such that
P0 |M, si = M |M, si, Pm |M, si = 0 for m 6= 0, J0 |M, si = s|M, si. (10.42)
Thus |M, si is a supermomentum eigenstate with vanishing eigenvalues under Pm ,
m 6= 0. In analogy with (10.30), we call |M, si the rest frame state of the module.
As in the Poincaré case above, unitarity is hidden in this picture because there is
no straightforward way to compute scalar products of states (10.43). In fact, since
|M, si is a delta function, all such states strictly have infinite norm. This is because
realistic states of BMS3 particles are smeared wavefunctions that consist of infinite
linear combinations of plane waves. Unitarity can then be recognized in the fact that
acting with (finite) superrotations on |M, si produces a “basis” of plane waves that
generate a space of square-integrable wavefunctionals on the supermomentum orbit.
In particular the representation is automatically irreducible in the sense that all basis
states are obtained by acting with symmetry transformations on the single state |M, si.
Vacuum module
Recall that the BMS3 vacuum is the scalar representation whose supermomentum
orbit Diff(S 1 )/PSL(2, R) contains the vacuum configuration p = pvac = −c2 /24. The
Chapter 10. Quantum BMS3 symmetry 320
Here the condition P0 |0i = 0 says that the vacuum has zero mass for the normalization
(10.40), while the extra conditions J±1 |0i = 0 enforce Lorentz-invariance. They reflect
the fact that the little group of the vacuum is the whole Lorentz group, rather than
the group of spatial rotations that occurs for massive particles.
If we were dealing with the Poincaré algebra, the requirements (10.44) would pro-
duce a trivial representation. Here, by contrast, there exist non-trivial “boosted vacua”
of the form (10.43), where now the ni ’s are integers different from −1, 0, 1. These vacua
are Lie-algebraic analogues of the boundary gravitons described earlier. The fact that
the vacuum is not invariant under the full BMS3 symmetry, but only under its Poincaré
subgroup, suggests that the boosted states (10.43) (with all ni ’s 6= −1, 0, 1) can be in-
terpreted as Goldstone-like states associated with broken symmetry generators; see the
discussion surrounding (8.81). Note that, in contrast to the realistic four-dimensional
case, BMS3 supertranslations do not create new states when acting on the vacuum.
with 1 ≤ n1 ≤ n2 ≤ ... ≤ nk and 1 ≤ n̄1 ≤ ... ≤ n̄l . Since we eventually wish to take
the ultrarelativistic limit of this representation, we will be interested in large values of
h and h̄, where the representation is irreducible and unitary thanks to the standard
Hermiticity conditions (8.64).
As in the Poincaré case, one can define new generators (10.32), now including
also the central charges c1 , c2 defined by (9.93). In particular, the space of Virasoro
descendants can be rewritten in the basis (10.43) with the identification |M, si ≡ |h, h̄i,
where M and s are the eigenvalues of P0 and J0 related to h and h̄ by (10.35). The
change of basis from descendant states (10.46) to infinitesimally boosted states (10.43)
is invertible because none of the Jn ’s annihilate the highest-weight state. The resulting
Virasoro representation takes a form analogous to (10.38), where now each state is
labelled by the quantum numbers ni of (10.43) and the matrices P(n) and J(n) also
Chapter 10. Quantum BMS3 symmetry 321
depend on the central charges (9.93). As before, only negative powers of ` enter P(n)
via the highest-weight conditions (10.45) written in the new basis:
1
P±n ± J±n |h, h̄i = 0. (10.47)
`
A limit ` → ∞ performed at fixed M , s and c1 , c2 (rather than fixed h, h̄ say) then
yields a massive bms3 module of the type described above. In particular, the limit
maps the highest-weight state (10.45) on the rest frame state (10.42). In this sense
bms3 modules are high-energy limits of tensor products of Virasoro modules, since h
and h̄ go to infinity in the flat limit. By the way, this provides an intuitive picture of
why the energy spectrum of BMS3 particles is continuous: the typical distance between
two consecutive eigenvalues of P0 = (L0 + L̄0 )/` is 1/`, which shrinks to zero when `
goes to infinity.
J˜0 |M̃ , s̃i = s̃|M̃ , s̃i, P̃0 |M̃ , s̃i = M̃ |M̃ , s̃i (10.48)
and
J˜m |M̃ , s̃i = P̃m |M̃ , s̃i = 0 for m > 0 . (10.49)
Chapter 10. Quantum BMS3 symmetry 322
Again, one concludes in that case that the representation is unitary if and only if
c̃2 = 0 and c̃1 ≥ 0.
The highest-weight representations of the type just described which are unitary (i.e.
have M̃ = 0 and c̃2 = 0) are special cases of induced representations of BMS3 as de-
scribed in section 10.1. Indeed, consider the vanishing supermomentum (p̃, c̃2 ) = (0, 0).
Its orbit under superrotations is trivial and its little group is the whole Virasoro group,
so the corresponding induced representation is entirely determined by its spin s̃. The
latter labels a unitary highest-weight representation of Virasoro, with central charge
c̃1 say. At the Lie-algebraic level this spin representation takes the form of a highest-
weight representation (10.48)-(10.49) with M̃ = 0 and c̃2 = 0. There is an analogue of
this construction in the Poincaré group: the vanishing momentum vector p = 0 has a
trivial orbit and its little group is the whole Lorentz group, so the corresponding in-
duced representation of Poincaré is just a unitary representation of the Lorentz group;
it is a “vacuum with spin” of the type mentioned in section 4.2.
The difference betwen the BMS3 vacuum (10.44) and the Galilean vacuum (10.51)
implies sharp differences for all quantum systems enjoying such symmetries, since it
affects the definition of normal ordering. For example, the normal-ordered product
: J2 P−3 : equals J2 P−3 in a BMS3 -invariant theory, while in a Galilean conformal field
theory one has : J˜2 P̃−3 : = P̃−3 J˜2 . We shall see explicit illustrations of this phenomenon
in section 11.3 below, when dealing with non-linear higher-spin symmetry algebras. It
suggests in particular that theories enjoying bms3 symmetry or gca2 symmetry differ
greatly at the quantum level, despite the isomorphism bms3 ∼ = gca2 .
we let the generators Lm , L̄n satisfy the algebra (8.65) with central charges c, c̄ respec-
tively, and we consider a highest-weight representation of the type (10.34)-(10.45). In
order to take the non-relativistic limit (9.94), we introduce a length scale ` and define
1
J˜n ≡ L̄n + Ln ,
P̃n ≡ L̄n − Ln . (10.52)
`
We stress that the combinations of Lm ’s appearing here differ from those of the ultra-
relativistic limit (10.32). In particular, J˜0 now generates time translations while P̃0
generates spatial translations; the parameter ` should no longer be interpreted as the
AdS3 radius, and there is no mixing between positive and negative modes. In these
terms, the limit ` → +∞ of the direct sum of two Virasoro algebras reduces to a gca2
algebra (10.41) with tildes on top of all generators, including the central charges
c̄ − c
c̃1 = c̄ + c , c̃2 = . (10.53)
`
Note that, up to central charges, the same redefinitions applied to sl(2, R) ⊕ sl(2, R)
reproduce the Poincaré algebra iso(2, 1); this is a non-relativistic limit to be contrasted
with the ultrarelativistic limit described in section 10.2.1.
In analogy with section 10.2.2, let us rewrite the tensor product of two highest-
weight representations of Virasoro in terms of the operators (10.52). Given the weights
h, h̄ we define
h̄ − h
s̃ ≡ h̄ + h , M̃ ≡
`
which we stress is radically different from the ultrarelativistic redefinition (10.35).
Upon identifying |M̃ , s̃i ≡ |h, h̄i, the highest-weight state satisfies (10.48) and (10.49).
These conditions hold for any value of `, including the limit ` → +∞. The descendant
states (10.50) then provide a representation of the sum of two Virasoro algebras, which
in the non-relativistic limit ` → +∞ becomes a generically non-unitary representation
of gca2 . Unitarity is recovered if M̃ = c̃2 = 0 and s̃, c̃1 ≥ 0. Again, this is strikingly
different from the ultrarelativistic contraction described above.
Remark. The difference between gca2 modules and bms3 modules has been known,
albeit in disguise, ever since the nineties. Namely, the tensionless limit of string theory
gives rise to so-called null strings [357], whose worldsheet is a null surface and thus
provides a stringy generalization of null geodesics. It was observed in [358] that the
algebra of constraints arising from worldsheet reparameterization invariance of null
strings is the bms3 algebra, although the name “BMS” was not used at the time.6
In the same paper the authors observed that a suitable normal-ordering prescription
gives rise to a consistent quantization of the null string in any space-time dimension,
and systematically results in a continuous mass spectrum. This result is the stringy
analogue of the bms3 modules described in section 10.2.2. Only later was it realized
that a different normal ordering prescription [359, 360] gives rise to the same criti-
cal dimension as in standard string theory (26 for the bosonic string and 10 for the
superstring), but that the resulting spectrum is massless and discrete; in fact, the spec-
trum then coincides with the massless part of the spectrum of standard string theory.
6
This occurrence of the bms3 algebra predates its gravitational description [20] by a decade!
Chapter 10. Quantum BMS3 symmetry 324
The latter result is the stringy analogue of the gca2 modules described here and the
critical dimension is analogous to the requirement c̃2 = 0 that ensures unitarity for
non-relativistic modules; see [361–363] for a recent account of these results.
7
Here δ denotes the delta function associated by (3.39) with the measure µ.
Chapter 10. Quantum BMS3 symmetry 325
To evaluate the character it only remains to integrate the delta function. This
requires local coordinates on the orbit in a neighbourhood of p, which can be obtained
by Fourier-expanding each supermomentum k(ϕ) as in eq. (6.116). The Fourier modes
∗
kn = k−n then transform under rotations f (ϕ) = ϕ + θ according to
kn 7→ [f · k]n = kn einθ .
As we shall see, the character that follows from this transformation is divergent due
to the fact that the group is infinite-dimensional. To cure this divergence we consider
complex rotations rather than real ones and introduce a complex parameter
1
τ≡ (θ + i) (10.56)
2π
where > 0. We then define the transformation of supermomentum Fourier modes
under complex rotations to be
kn e2πinτ if n > 0,
kn 7→ [f · k]n = k0 if n = 0, (10.57)
kn e2πinτ̄ if n < 0.
We will see below that this modification can be justified by thinking of BMS3 repre-
sentations as high-energy limits of Verma modules. Note that this prescription leaves
room for “Euclidean” rotations (i.e. rotations by an imaginary angle) while preserving
the reality condition (kn )∗ = k−n .
The problem now is to express the measure µ and the corresponding delta function
δ in terms of Fourier modes. On the massive supermomentum orbit Op ∼ = Diff(S 1 )/S 1 ,
the non-zero Fourier modes of k(ϕ) determine its energy k0 . This is analogous to the
statement that thepenergy of a relativistic particle is determined by its momentum
according to E = M 2 + k2 . Let us prove this in a neighbourhood of the supermo-
mentum at rest, p = M − c2 /24, by acting on it with an infinitesimal superrotation X
that we Fourier-expand as X
X(ϕ) = i Xn e−inϕ .
n∈Z
Here the variation of the zero-mode, δp0 , vanishes for any choice of X. By contrast, all
other Fourier modes are acted upon in a non-trivial way and can therefore take arbi-
trary values by a suitable choice of X. This implies that (at least in a neighbourhood
of p) the non-zero modes of supermomenta provide local coordinates on Op . In terms
of the Fourier decomposition (6.116), this is to say that when k(ϕ) = p + ε(δX p)(ϕ),
the non-zero modes kn coincide with εδpn (while k0 = p to first order in ε).
Chapter 10. Quantum BMS3 symmetry 326
where theQprefactor is unknown. In quantum mechanics one would write the infinite
product n∈Z∗ dkn as a path integral measure Dk, with the extra rule that the zero-
mode of k is not to be integrated over. The definition of the delta function (3.39)
associated with µ ensures that
Y
δ(q, k) = (Some k-dependent prefactor)−1 × δ(qn − kn ),
n∈Z∗
where the δ on the right-hand side is the usual Dirac distribution in one dimension.
Crucially, the prefactor appearing in front of the delta function is the inverse of the
prefactor of the measure (10.59). As in eq. (4.56) this implies that the combination
dµ(k)δ(k, ·) is invariant under changes of measures, and it allows us to rewrite the
character (10.55) as
Z
isθ iα0 (M −c2 /24)
Y Y
χ[(f, α)] = e e dkn δ (kn − [f · k]n )
R2∞ n∈Z∗ n∈Z∗
Z Y +∞ +∞ 2
(10.57) 0 (M −c
Y
eisθ eiα 2 /24)
δ kn (1 − e2πinτ )
= dkn , (10.60)
R∞ n=1 n=1
where we have replaced the real angle θ by its complex counterpart 2πτ given by
(10.56). Denoting q ≡ exp[2πiτ ] and evaluating the integral, the character of a massive
BMS3 particle finally reduces to
0 (M −c /24) 1
χ[(f, α)] = eisθ eiα 2
Q+∞ n 2
. (10.61)
n=1 |1 − q |
Remark. At this stage, and in contrast to conformal field theory, the coefficient
τ should not be seen as a modular parameter. The small parameter in (10.56)
was merely introduced to ensure convergence of the determinant arising from the
integration of the delta function in (10.60). This being said, the occurrence of the
Dedekind eta function is compatible with the modular transformations used in [337,
341] to derive a Cardy-like formula reproducing the entropy of flat space cosmologies.
Chapter 10. Quantum BMS3 symmetry 327
Here the term n = 1 coincides (4.98), while the contribution of higher Fourier modes
is due, loosely speaking, to the infinitely many Poincaré subgroups of BMS3 . This is
analogous to the fact that the Virasoro character (8.74) may be seen as a product of
infinitely many SL(2, R) characters (5.102) labelled by an integer n.
Writing the modular parameter in the form (8.78) with an `-independent β and intro-
ducing a mass M and a spin s defined by (10.35), the large ` limit of the quantities
appearing in the right-hand side of (10.62) is
1
τ∼ (θ + i), q h−c/24 q̄ h̄−c̄/24 ∼ eiθ(s−c1 /24) e−β(M −c2 /24) . (10.63)
2π
(Here the imaginary part of τ goes to zero, but we keep writing it as > 0 to reproduce
the regularization used in (10.61).) Thus the flat limit of (10.62) is
L0 −c/24 L̄0 −c/24
1
lim Tr q q̄ = eiθ(s−c1 /24) e−β(M −c2 /24) Q+∞ n 2
,
n=1 |1 − q |
`→+∞
and coincides (up to a redefinition of spin) with the BMS3 character (10.61) for a
supertranslation whose zero-mode is a Euclidean time translation, α = iβ. The left-
hand side of this expression can be interpreted as a trace
(10.63)
lim Tr q L0 −c/24 q̄ L̄0 −c/24 = Tr eiθ(J0 −c1 /24) e−β(P0 −c2 /24) = χ[(f, α)], (10.64)
`→+∞
We could also have guessed that such a simplification would occur thanks to di-
mensional arguments. Indeed, both M and c2 are dimensionful parameters labelling
BMS3 representations, so their values can be tuned at will by a suitable choice of
units. Accordingly, in contrast to Virasoro highest weight representations, one should
not expect to find sharp bifurcations in the structure of BMS3 particles as M and c2
vary. In this sense formula (10.61) is a universal character.
As before, the quantity we wish to compute is χ[(f, α)], where α is any super-
translation. Since the little group is now larger than U(1), one can obtain non-trivial
characters even when f is not conjugate to a rotation. We will not consider such
cases here and stick instead to our earlier convention that f (ϕ) = ϕ + θ is a rotation
by θ 6= 0. (Equivalently we may take f to be merely conjugate to a rotation since
the character is a class function.) Then the integral of the Frobenius formula (10.54)
localizes to the unique rotation-invariant point pvac on the orbit and the character can
be written as Z
−iα0 c2 /24
χvac [(f, α)] = e dµ(k) δ(k, f · k). (10.65)
Ovac
Here µ is some quasi-invariant measure on the vacuum orbit. Using Fourier expan-
sions (6.116), we can think of Fourier modes as redundant coordinates on the orbit;
the subtlety is to understand which of these modes should be modded out so as to
provide genuine, non-redundant local coordinates on Ovac .
little group is PSL(2, R). Thus, in a neighbourhood of pvac , we can use the higher
Fourier modes pn with |n| ≥ 2 as local coordinates. In particular the measure µ now
takes the form
+∞
Y
dµ(k) = (Some k-dependent prefactor) × dkn dk−n ,
n=2
where the prefactor is again unknown, but eventually irrelevant since it is cancelled
by the prefactor of the corresponding delta function. The vacuum character (10.65)
thus boils down to
Z +∞ +∞
−iα0 c2 /24
Y Y
χvac [(f, α)] = e dkn dk−n δ(kn − [f · k]n )δ(k−n − [f · k]−n )
R2∞−2 n=2 n=2
Z +∞ +∞ 2
(10.57) −iα0 c2 /24
Y Y
n
= e dkn δ (kn (1 − q )) ,
R∞−1 n=2 n=2
0c 1
χvac [(f, α)] = e−iα 2 /24
Q+∞ n 2
. (10.66)
n=2 |1 − q |
The results described in this chapter first appeared in [49–51]. They are Minkow-
skian analogues of earlier observations on partition functions in AdS3 [69,70, 241] that
we already referred to in section 8.4. Note that our language in this chapter will be
somewhat different than in the previous ones, as we will rely much more heavily on
331
Chapter 11. Partition functions and characters 332
quantum field theory. On the other hand the group-theoretic tools that we will be
using are essentially the same as in chapter 10.
where the exponent # depends on the nature of the fields that were integrated out. The
quantity δ 2 S/δφ(x)δφ(y) appearing in this expression is a differential operator acting
on sections of a suitable vector bundle over RD /Z. The evaluation of the one-loop con-
tribution to the partition function thus boils down to that of a functional determinant.
Heat kernels are well suited for the computation of functional determinants on
quotient spaces. Indeed, suppose Γ is a discrete subgroup of the isometry group of
RD , acting freely on RD . Introducing the equivalence relation x ∼ y if there exists a
γ ∈ Γ such that γ(x) = y, we define the quotient manifold RD /Γ as the corresponding
set of equivalence classes. Given a differential operator ∆ on RD , it naturally induces
a differential operator on RD /Γ, acting on fields that satisfy suitable (anti)periodicity
conditions. Because the heat equation (11.4) is linear, the heat kernel on the quotient
space can be obtained from the heat kernel on RD by the method of images:
D
X
K R /Γ (t, x, x0 ) = K t, x, γ(x0 ) .
(11.6)
γ ∈Γ
Here, abusing notation slightly, x and x0 denote points both in RD and in its quotient.
In writing (11.6) we are assuming, for simplicity, that the tensor structure of K is
trivial, but as soon as K carries tensor or spinor indices (i.e. whenever the fields under
consideration have non-zero spin), the right-hand side involves Jacobians that account
D
for the non-trivial transformation law of K. Once K R /Z is known, the determinant of
D
the operator −∆ + M 2 is given by (11.3) with K replaced by K R /Z and RD replaced
by RD /Z.
for D odd or D even, respectively. Being isometries of flat space, these transformations
are linear maps in Cartesian coordinates, so their nth power coincides with the Jacobian
matrix ∂γ n (x)µ /∂xν that will be needed later for the method of images. Throughout
this chapter we take all angles θ1 , ..., θr to be non-vanishing and combine them in a
vector θ~ = (θ1 , ..., θr ). We now display the computation of one-loop partition functions
on RD /Z for bosonic higher-spin fields.
Fµ1 ···µs ≡ ∆ φµ1 ...µs − ∂(µ1 | ∂ λ φ|µ2 ...µs )λ + ∂(µ1 ∂µ2 φµ3 ...µs )λ λ . (11.10)
Parentheses denote the symmetrization of the indices they enclose, with the minimum
number of terms needed and without any overall factor. The massless action (11.9) has
a gauge symmetry φµ1 ...µs 7→ φµ1 ...µs + ∂(µ1 ξµ2 ...µs ) , where ξµ2 ...µs is a symmetric tensor
field. When s = 2 the action reduces to that of a metric perturbation hµν around a
flat background. We refer for instance to [368] for many more details on this topic.
Chapter 11. Partition functions and characters 335
Massive case
Applying e.g. the techniques of [70] to the presentation of the Euclidean action of a
massive field of spin s of [366], one finds that the partition function is given by
1 1
log Z = − log det(−∆(s) + M 2 ) + log det(−∆(s−1) + M 2 ) , (11.11)
2 2
where ∆(s) is the Laplacian ∂µ ∂ µ acting on periodic,1 symmetric, traceless tensor fields
with s indices on RD /Z. We denote the heat kernel associated with (−∆(s) + M 2 ) on
RD by Kµs ,νs (t, x, x0 ), where µs and νs are shorthands that denote sets of s symmetrized
indices. The differential equation (11.4) with initial condition (11.5) for Kµs ,νs (t, x, x0 )
then reads
where Iµs ,νs is an identity matrix with the same tensor structure as Kµs ,νs . Sets of
repeated covariant or contravariant indices denote sets of indices that are symmetrized
with the minimum number of terms required and without multiplicative factors, while
contractions involve as usual a covariant and a contravariant index. For instance the
tracelessness condition on the heat kernel amounts to
where |x − x0 | is the Euclidean distance between x and x0 , while the spin-s identity
matrix is
b 2s c
X (−1)n 2n n! [D + 2(s − n − 2)]!! n s−2n n
Iµs , νs = δµµ δµν δνν . (11.15)
n=0
s! [D + 2(s − 2)]!!
Note that the dependence of this heat kernel on the space-time points x, x0 and on
Schwinger proper time t is that of a scalar heat kernel, and completely factorizes from
its spin/index structure which is entirely accounted for by the matrix I.2 This simpli-
fication is the reason why heat kernel computations are simpler in flat space than in
AdS or dS.
To determine the heat kernel associated with the operator (−∆(s) + M 2 ) on RD /Z,
we use the method of images (11.6), taking care of the non-trivial index structure.
Denoting the matrix (11.8) by Jα β (it is the Jacobian of the transformation x 7→ γ(x)),
the spin-s heat kernel on the quotient space RD /Z is
D
X
KµRs , ν/Zs (t, x, x0 ) = (J n )α β ...(J n )α β Kµs , βs t, x, γ n (x0 ) ,
(11.16)
n∈Z
1
More precisely, the field at time τ + β is rotated by θ~ with respect to the field at time τ .
2
Note also that the scalar heat kernel coincides with the propagator of a free particle in RD ,
whose expression for D = 2 was written in eq. (5.158).
Chapter 11. Partition functions and characters 336
where we recall again that repeated covariant or contravariant indices are meant to
be symmetrized with the minimum number of terms required and without multiplica-
tive factors, while repeating a covariant index in a contravariant position denotes a
contraction. Accordingly, eq. (11.3) gives the determinant of (−∆(s) + M 2 ) on RD /Z:
Z +∞ Z
(s) 2 dt D
− log det(−∆ + M ) = dD x (δ µα )s KµRs , α/Zs (t, x, x)
0 t RD /Z
Z +∞ Z
X
n µβ n µβ dt 1 1
−M 2 t− 4t |x−γ n (x)|2
= (J ) · · · (J ) Iµs ,βs dD x D/2
e .
n∈Z 0 t RD /Z (4πt)
(11.17)
In this series the term n = 0 contains both an ultraviolet divergence (due to the singular
behaviour of the integrand as t → 0) and an infrared one (due to the integral of a
constant over RD /Z), proportional to the product βV where V is the spatial volume
of the system. This divergence is a quantum contribution to the vacuum energy, which
we ignore from now on. The only non-trivial one-loop contribution then comes from
the terms n 6= 0 in (11.17). Using
r
X
|x − γ n (x)|2 = n2 β 2 + 4 sin2 (nθi /2)(x2i + yi2 )
i=1
in terms of the coordinates introduced around (11.7), the integrals over t and x give
rise to a divergent series
(
X 1 χs [nθ~ ] e−|n|βM if D odd,
− log det(−∆(s) + M 2 ) = r ×
|n| Q ML
K1 (|n|βM ) if D even,
n ∈ Z∗ |1 − einθj |2 π
j=1
R +∞ (11.18)
where K1 is the first modified Bessel function of the second kind, L ≡ −∞ dz is an
infrared divergence (4.64) that arises in even dimensions because the z axis is left fixed
by the rotation (11.8), and
s
χs [nθ~ ] ≡ (J n )µβ ...(J n )µβ Iµs , βs ≡ (J n )µβ Iµs , βs
(11.19)
is the full mixed trace of Iµs ,νs . As such, expression (11.18) makes no sense because
the sum over n diverges. To cure this problem, one needs to choose a regularization
procedure. Motivated by the similar situation already encountered in eq. (10.56), for
now we choose to regulate the series by a naive replacement: we let j , j = 1, ..., r be
small positive parameters and replace θj by θj ± ij in all positive powers of e±iθj . As
a result, expression (11.18) is replaced by the convergent series
(
(s) 2
X 1 ~
χs [nθ,~ ] e−|n|βM if D odd,
− log det(−∆ + M ) = r ×
|n| Q ML
K1 (|n|βM ) if D even,
n ∈ Z∗ |1 − ein(θj +ij ) |2 π
j=1
(11.20)
~
where χs [nθ,~ ] is still given by (11.19), except that now all factors e ±iθj
appearing in
±i(θj ±ij )
the Jacobians are replaced by e .
Chapter 11. Partition functions and characters 337
The regularization described here is motivated by the fact that, for odd D, the
resulting expressions look very much like flat limits of AdS one-loop determinants, in
which case the parameters j ∝ β/` are remnants of the inverse temperature (with `
the AdS radius). The subtlety, however, is that the exact matching of the flat limit
of AdS with combinations such as (11.20) requires some of the j ’s to be multiplied
by certain positive coefficients; thus eq. (11.20) is not quite the same as the flat limit
of its AdS counterpart — we will illustrate this point for D = 3 in section 11.1.3. As
for even values of D, the situation is even worse since the flat limit of the AdS result
contains an infrared divergence; it is not obvious how this divergence can be regular-
ized so as to reproduce the combination L · K1 of (11.20), though apart from this the
other terms of the expression indeed coincide with the flat limit of their AdS counter-
parts. From now on we will use the i prescription systematically, often omitting to
indicate it explicitly. We will keep it only in the final results, and in section 11.1.3 we
will introduce a refined regularization such that partition functions in D = 3 exactly
reproduce characters of suitable asymptotic symmetry algebras, while also matching
the flat limit of their AdS peers.
We can now display the one-loop partition function (11.11). Using expression
(11.20) for the one-loop determinant together with property (11.21), we find
(D) (D)
X+∞
n−1 χλs [nθ~ ] − χλs−1 [nθ~ ] e−nβM
~
Z(β, θ ) = exp
r ×
Q (D) ~ 0] − χ (D) ~ 0] M L K1 (nβM )
|1 − einθj |2 χλs [nθ, [n θ,
n=1
λs−1 π
j=1
(11.22)
where the upper (resp. lower) line corresponds to the case where D is odd (resp. even).
Remarkably, the differences of SO(D) characters appearing here can be simplified:
according to eqs. (11.154a) and (11.155), the difference of two SO(D) characters with
weights (s, 0, ..., 0) and (s − 1, 0, ..., 0) is a (sum of) character(s) of SO(D − 1):
(D) (D)
χλs [θ~ ] − χλs−1 [θ~ ] (D odd)
(D−1)
= χλs [θ~ ] . (11.23)
(D) ~ (D) ~
χλs [θ, 0] − χλs−1 [θ, 0] (D even)
Chapter 11. Partition functions and characters 338
Since the rank of SO(D − 1) is r = b(D − 1)/2c, the right-hand side of this equality
makes sense regardless of the parity of D, and the partition function (11.22) boils
down to
e−nβM
+∞ (D−1) ~
(
X 1 χλs [nθ,~ ] (D odd)
~
Z(β, θ ) = exp
× . (11.24)
r
n Q in(θ +i )
ML
K1 (nβM ) (D even)
n=1 |1 − e j j |2 π
j=1
Note that the function of nθ~ and nβ appearing here in the sum over n is essentially
the character (4.60)-(4.65) of a Poincaré particle with mass M and spin λs ; we will
return to this observation in section 11.1.4. An analogous result holds in Anti-de Sitter
space [369–371].
Massless case
We now turn to the one-loop partition function associated with the Euclidean Fronsdal
action (11.9), describing a massless field with spin s. The extra ingredient with respect
to the massive case is the gauge symmetry φµ1 ...µs 7→ φµ1 ...µs + ∂(µ1 ξµ2 ...µs ) . This forces
one to fix a gauge and introduce ghost fields that absorb the gauge redundancy [70],
which adds two more functional determinants to the massive result (11.11) and leads
to the following expression for the one-loop term of the partition function:
1 1
log Z = − log det(−∆(s) ) + log det(−∆(s−1) ) − log det(−∆(s−2) ) . (11.25)
2 2
As before, ∆(s) is the Laplacian on RD /Z acting on periodic, traceless, symmetric fields
with s indices. The functional determinants can be evaluated exactly as in the massive
case, upon setting M = 0. In particular, using limx→0 xK1 (x) = 1, the massless version
of the functional determinant (11.20) is
(
X 1 ~ ]
χs [nθ,~ 1 if D odd,
(s)
− log det(−∆ ) = r × L (11.26)
|n| Q if D even,
n ∈ Z∗ |1 − ein(θj +ij ) |2 π|n|β
j=1
which has been regularized as in the massive case. The matching (11.21) between χs
and a character of SO(D) remains valid, but a sharp difference arises upon including
all three functional determinants in (11.25). Indeed, the combination of χs ’s now is
(11.21)-(11.23) (D−1) (D−1)
χs [nθ~ ] − 2χs−1 [nθ~ ] + χs−2 [nθ~ ] = χλs [nθ~ ] − χλs−1 [nθ~ ] . (11.27)
It is tempting to use (11.23) once more to rewrite this as a character of SO(D − 2),
and indeed this is exactly what happens for even D because in that case the rank of
SO(D − 1) equals that of SO(D − 2):
+∞ (D−2) ~
~
X 1 χλs [nθ,~ ] L
Z(β, θ ) = exp
r
(even D). (11.28)
n Q in(θ +i ) πnβ
n=1 |1 − e j j |2
j=1
Chapter 11. Partition functions and characters 339
If D is odd, however, the rank decreases by one unit in going from SO(D − 1) to
SO(D − 2), so expression (11.27) contains one angle too much to be a character of
SO(D − 2). In fact, when D = 3, the right-hand side of (11.27) is the best we can
hope to get; for s ≥ 2 it takes the form
(1) (2) (2)
χλs [nθ] ≡ χλs [nθ] − χλs−1 [nθ] = eisnθ − ei(s−1)nθ + c.c. (11.29)
where we have used the character χs [θ] = eisθ + e−isθ for parity-invariant unitary
representations of SO(2) and “c.c.” means “complex conjugate”. (For lower spins one
(1) (1)
has χλ0 [θ] ≡ 1 and χλ1 [θ] ≡ 2 cos θ − 1.) Hence the partition function given by (11.25)
becomes
" +∞ #
(0)
X1 1
Z(β, θ) = e−S exp eisn(θ+i) − ei(s−1)n(θ+i) + c.c.
in(θ+i) |2
(D = 3)
n=1
n |1 − e
(11.30)
upon using the crude regularization described below eq. (11.20). For the sake of
generality we have included a spin-dependent classical action S (0) , whose value is
a matter of normalization and is generally taken to vanish, except for spin two (see
below). In the more general case where D is odd and larger than three, a simplification
does occur on the right-hand side of (11.27): as we show in appendix 11.A.3, the
difference (11.27) can be written as a sum of SO(D−2) characters with angle-dependent
coefficients (see eq. (11.154b)). Indeed, let us define
where |Aij | denotes the determinant of an r × r matrix. Then the rotating one-loop
partition function for a massless field with spin s in odd space-time dimension D ≥ 5
reads
r
P r ~ (D−2)
+∞ A (nθ,~ ) χλs [nθ1 , ..., nθk , ..., nθr ,~ ]
d
~
X 1 k=1 k
Z(β, θ ) = exp
r
(odd D ≥ 5),
n Q in(θ +i ) 2
n=1 |1 − e j j |
j=1
(11.32)
where the hat on top of an argument denotes omission.
Note that the massless partition functions (11.28) and (11.32) are related to the
massless limit of (11.24). Indeed, as we show in appendix 11.A.4, it turns out that
P (D−2)
r r ~
k=1 Ak (θ ) χλj [θ1 , ..., θbk , ..., θr ] for odd D,
Xs
(D−1) ~
χ λs [ θ ] = (11.33)
χ(D−2) [θ~ ] for even D.
j=0 λj
Accordingly, the massless limit of a massive partition function with spin s is a product
of massless partition functions with spins ranging from 0 to s,
s
Y
lim ZM,s = Zmassless,j , (11.34)
M →0
j=0
Chapter 11. Partition functions and characters 340
consistently with the structure of the massive action [366]. This result stresses again
the role of the functions Ark (θ~ ) defined in (11.31): when the space-time dimension
is odd, one needs angle-dependent coefficients because the rank of the little group of
massless particles is smaller than the maximum number of angular velocities, so that
a single SO(D − 2) character cannot account for all of them. By the way, the results
(11.33) and (11.34) also hold in dimension D = 3, provided one takes the “characters”
(1) (1) (1)
χλs [θ] to be of the form (11.29) with χλ0 [θ] = 1 and χλ1 [θ] = 2 cos θ − 1.
where the new regularization (11.35) has ensured that the summand decomposes as
the sum of a chiral and an anti-chiral piece in q. (This was not the case with the
rough regularization of eq. (11.20)!) In order to write down the full partition function,
it only remains to assign a value to the classical action S (0) ; a convention that has
come to be standard in the realm of three-dimensional gravity is to set S (0) = 0 for any
spin s 6= 2 (vacuum expectation values are assumed to vanish), while S (0) = −β/8G
for spin two (with G the Newton constant in three dimensions). This choice ensures
covariance of the on-shell action under modular transformations [49, 372], in analogy
with the similar choice in AdS3 [241]. All in all, combining the value of S (0) with the
series (11.36) and renaming j into n, one finds that the three-dimensional partition
function (11.30) can be written as
+∞
δs,2
βc2 Y 1
Z(β, θ) = e 24 , c2 = 3/G. (11.37)
n=s
|1 − ein(θ+i) |2
This expression is the flat limit of the analogous higher-spin partition function in
AdS3 [70]. But most importantly for our purposes, taking s = 2 in this formula, we
recognize the vacuum BMS3 character (10.66). This is our first key conclusion in this
chapter: it confirms that boundary gravitons in flat space form an irreducible unitary
representation of the BMS3 group of the type described in chapter 10. The case s > 2
will be studied in section 11.2, with similar conclusions.
The result (11.37) can be generalized to orbifolds in flat space: upon declaring
that the angular coordinate ϕ of (9.4) is identified as ϕ ∼ ϕ + 2π/N with some integer
N > 1, one obtains a flat three-dimensional conical deficit. One can then evaluate heat
Chapter 11. Partition functions and characters 341
kernels on that background by computing a sum over images (11.6), where Γ is now a
group Z × ZN whose two generators enforce (i) the thermal identifications (11.7), and
(ii) the orbifolding ϕ ∼ ϕ + 2π/N . An important technical subtlety is that, in order
to evaluate a partition function with temperature 1/β and angular potential θ on that
background, the angle appearing in (11.7) must be θ/N rather than θ. The rest of the
computation is straightforward, and one finds that the one-loop partition function of
gravity can be written as
+∞
−βp0
Y 1
ZN (β, θ) = e ,
n=1
|1 − ein(θ+i) |2
where p0 = −c2 /(24N 2 ). Comparing with (10.61), we recognize the (Euclidean) char-
c2
acter of a BMS3 particle with mass M = 24 (1 − 1/N 2 ), which is indeed the mass one
would obtain by writing the conical deficit metric in BMS form (9.25). Note in par-
ticular that the sum over images of the
Q orbifolding group ZN converts the truncated
product of (11.37) into a full product +∞
n=1 (· · · ).
The result (11.37) first appeared in [49] and parallels earlier observations in AdS3
[241]. It is tempting to conjecture that formula (11.37) is one-loop exact, since it is the
only expression compatible with BMS3 symmetry. This being said we will not need
to assume one-loop exactness in this thesis and we will not attempt to prove it. The
remainder of this chapter is devoted to various extensions of this matching.
Remark. In [339] it was shown that the one-loop partition function (11.37) with s =
2, and hence the vacuum BMS3 character (10.66), can be reproduced using quantum
Regge calculus. In that context the truncation of the product over n = 2, 3, ... is a
consequence of triangulation-invariance in the bulk.
For massless fields, the situation is a bit more complicated. For even D the massless
Poincaré character (4.69) is the limit M → 0 of its massive counterpart (4.65), and
the one-loop partition function (11.28) can again be written as an exponential (11.38).
But in odd space-time dimensions, SO(D − 2) has lower rank than SO(D − 1), so the
rotation (11.8) is not, in general, conjugate to an element of the massless little group:
it has one angle too much, and whenever all angles θ1 , ..., θr are non-zero, the character
(4.33) vanishes. The only non-trivial irreducible character arises when at least one of
the angles θ1 , ..., θr vanishes, say θr = 0, in which case the massless character takes the
form (4.71). However, comparison with (11.32) reveals a mismatch: the partition func-
tion does not take the form (11.38) in terms of the massless characters (4.71); in field
theory, all r angles θi may be switched on simultaneously! To accommodate for this
one can resort to the angle-dependent coefficients Ark (θ) ~ introduced in (11.31), whose
origin can again be understood through the massless limit of the character (4.60).
Using relation (11.33), the product of massless partition functions with spins ranging
from zero to s can be written as (11.38), where the characters on the right-hand side
are massless limits of massive Poincaré characters. However, it is unclear whether the
quantities appearing in the exponent of (11.32) can be related directly to Poincaré
characters without invoking a massless limit.
Asymptotic symmetries
Asymptotic symmetries of higher-spin theories in three dimensions were first studied
in AdS3 [377–380], and are similar to the Brown-Henneaux asymptotic symmetries of
gravity described in chapter 8. Here we focus on models including fields with spin
ranging from 2 to N ; this setup can be described as an sl(N, R) ⊕ sl(N, R) Chern-
Simons action with a principally embedded sl(2, R)⊕sl(2, R) gravitational subalgebra.
When N = 3, the asymptotic symmetries are generated by gauge transformations
specified by four arbitrary, 2π-periodic functions X(x+ ), ξ(x+ ) and X̄(x− ), ξ(x ¯ −)
One can think of the pair (X, ξ) as an element of the asymptotic symmetry algebra,
so the charge (11.39) is the pairing between this algebra and its dual space. This
generalizes the pairing (6.34) of the Virasoro algebra with CFT stress tensors, and
(p, ρ) may be seen as a coadjoint vector of the symmetry algebra. Its infinitesimal
transformation law extends (8.39) and turns out to be [378]
c 000
δ(X,ξ) p = Xp0 + 2X 0 p − X + 2 ξρ0 + 3 ξ 0 ρ , (11.40a)
12
δ(X,ξ) ρ = Xρ0 + 3X 0 ρ + 2ξp000 + 9ξ 0 p00 + 15ξ 00 p0 + 10ξ 000 p
c 192
− ξ (5) − ξpp0 + ξ 0 p2 ,
(11.40b)
12 c
where prime denotes differentiation with respect to x+ , and c = 3`/2G is the Brown-
Henneaux central charge. Analogous formulas hold in the anti-chiral sector. Since X
generates conformal transformations, this implies that p is a (chiral) quasi-primary
field with weight 2 while ρ is a primary with weight 3. Together with the surface
charges (11.39), these transformation laws yield the Poisson bracket (8.10):
Q(X,ξ) [ p, ρ], Q(Y,ζ) [ p, ρ] = − δ(X,ξ) Q(Y,ζ) [p, π] . (11.41)
Formula (11.40) turns out to coincide with the coadjoint representation of a Poisson
algebra known as the W3 algebra, and indeed one finds that the bracket (11.41) re-
produces the non-linear bracket of a W3 algebra with central charge c (see eq. (11.43)
below). Similar considerations apply to models including fields with spin ranging from
2 to N [378, 380]; the resulting asymptotic symmetry algebra is the direct sum of two
copies of WN .
Quantum W3 algebra
As in the purely gravitational case (8.46), the classical asymptotic symmetry alge-
bra given by the surface charges (11.39) can be written in terms of modes
and their barred counterparts in the right-moving sector. The normalization is such
that pure AdS3 has all charges vanishing except L0 = L̄0 = −c/24. Using (11.40),
one finds that the Poisson brackets (11.41) of the charges (11.42) take the form of a
Chapter 11. Partition functions and characters 345
classical W3 algebra:
c 3
i{Lm , Ln } = (m − n)Lm+n + m δm+n,0 ,
12
i{Lm , Wn } = (2m − n)Wm+n , (11.43)
96 c
i{Wm , Wn } = (m − n)(2m2 + 2n2 − mn)Lm+n + Λm+n + m5 δm+n,0 ,
c 12
P
where Λm ≡ p∈Z Lm−p Lp is a non-linear term and the first line is the usual Vira-
soro algebra. The same brackets hold in the right-moving sector, so as announced the
asymptotic symmetry algebra is a direct sum of two classical W3 algebras. Under quan-
tization the Poisson brackets are turned into commutators according to i{·,[ ·} = [ˆ·,ˆ·]
and the charges Lm , Wn become operators Lm , Wn which, in any unitary representa-
tion, satisfy the Hermiticity conditions
It is also customary to normalize the Virasoro generators Lm so that the vacuum state
c
has vanishing eigenvalue under L0 , i.e. to rename Lm + 24 δm,0 into Lm . As a result the
commutators of the operators Lm , Wn yield the quantum W3 algebra
c
[Lm , Ln ] = (m − n)Lm+n + (m3 − m)δm+n, 0 , (11.44a)
12
[Lm , Wn ] = (2m − n)Wm+n , (11.44b)
96
[Wm , Wn ] = (m − n)(2m2 + 2n2 − mn − 8)Lm+n + (m − n) : Λm+n :
c + 22/5
c
+ (m2 − 4)(m3 − m)δm+n,0 , (11.44c)
12
whose non-linear terms are normal-ordered according to the prescription
X X 3
: Λm : ≡ Lm−p Lp + Lp Lm−p − (m + 3)(m + 2)Lm . (11.45)
p≥−1 p<−1
10
Here the term linear in Lm ensures that the operator : Λm : is quasi-primary with
respect to the action of Lm ’s. Note how the denominator of the structure constant of
the non-linear term in (11.44c) involves a shifted central charge c + 22/5 instead of
the classical c in the last line of (11.43). Analogous commutation relations hold in the
barred sector.
This reduces to the product of Virasoro characters (10.62) when N = 2. The vacuum
character of WN ⊕ WN similarly reads
N +∞
!
Y Y 1
Trvac q L0 −c/24 q̄ L̄0 −c̄/24 = n |2
, (11.47)
s=2 n=s
|1 − q
where the truncated product arises because the vacuum state is left invariant by the
wedge algebra sl(N, R). This reduces to the vacuum character (8.79) when N = 2.
As mentioned earlier, it was shown in [241] that the one-loop partition function of
gravitons in AdS3 at temperature 1/β and angular potential θ is a vacuum character
(8.79) with modular parameter (8.78). This result was later extended to higher-spin
theories [69, 70], whose one-loop partition functions on thermal AdS3 coincide with
vacuum WN ⊕ WN characters (11.47) upon including the contribution of fields with
spins s = 2, 3, ..., N . These results confirm the interpretation of irreducible unitary
representations of asymptotic symmetry groups as particles dressed with boundary
degrees of freedom. The purpose of the remainder of this chapter is to describe the
similar matching that occurs in asymptotically flat theories.
W3 algebra, i.e. according to (11.40), albeit with a central charge c2 = 3/G instead of c.
The Poisson brackets satisfied by the surface charges (11.48) are given as usual by
(11.41) and are most easily expressed in terms of generators
Note that with this normalization pure Minkowski space has all its charges vanishing,
except P0 = −1/8G. One then finds that the Jm ’s and Pm ’s close according to a
bms3 algebra (9.37) with central charge c2 = 3/G, while brackets involving higher-spin
charges take the form
These formulas show that, up to central terms, the brackets of (J , K)’s with (P, Q)’s
take the same form as the brackets of (J , K)’s with themselves. This is the situation
described in (9.48) and it implies that, similarly to (9.64), the asymptotic symmetry
algebra is an exceptional semi-direct sum
where W3 is the classical W3 algebra (11.43) and (W3 )Ab denotes an Abelian Lie
algebra isomorphic, as a vector space, to W3 . This algebra is centrally extended, as
the bracket between generators of W3 and those of (W3 )Ab includes a central charge
c2 ; there is also a central charge c1 specific to the left (non-Abelian) W3 subalgebra of
(11.51), but it is not switched on in parity-preserving theories. We shall return to this
structure in section 11.3, upon describing its quantum version.
symplectic leaves in the terminology of section 5.1. In the case of the Virasoro algebra
(which corresponds to WN with N = 2) this concept coincides with that of a coadjoint
orbit of the Virasoro group. For higher N the symplectic leaves of WN algebras are
still well defined [376] despite the lack of a straightforward definition of the group
that corresponds to the WN algebra. These leaves may be seen as intersections of the
coadjoint orbits of sl(N )-Kac Moody algebras with the constraints that implement
the Hamiltonian reduction to WN algebras. Accordingly, it should be possible to build
unitary representations of flat WN algebras as Hilbert spaces of wavefunctions defined
on their symplectic leaves, which we assume as usual to admit quasi-invariant mea-
sures. In most of the remainder of this chapter, we test that proposal by showing how
it can be used to evaluate characters that coincide with higher-spin one-loop parti-
tion functions. Note that the non-linearities appearing in the brackets of the algebra
(11.51) imply an extra complication for representation theory in that one has to devise
a suitable normal-ordering prescription. In the standard W3 case we displayed this
normal ordering in (11.45). In the flat case we will address this issue in section 11.3.
The complete classification of the symplectic leaves of the W3 algebra has been
worked out in [375, 376]; according to our proposal this settles the classification of
irreducible unitary representations of the flat W3 algebra, in the same way that Vira-
soro coadjoint orbits classify BMS3 particles. Instead of describing the details of this
classification, we focus from now on on orbits of constant supermomenta, which can be
classified thanks to the infinitesimal transformation laws (11.40) given by the algebra.
To describe such an orbit, let us pick a pair (p, ρ) where p(ϕ) = p0 and ρ(ϕ) = ρ0 are
constants, and act on it with an infinitesimal higher-spin superrotation (X, ξ). Then,
all terms involving derivatives of p or ρ in eq. (11.40) vanish, and we find
c2 000
δ(X,ξ) p0 = 2 X 0 p0 − X + 3 ξ 0 ρ0 , (11.52a)
12
c2 (5) 192 0 2
δ(X,ξ) ρ0 = 3 X 0 ρ0 + 10 ξ 000 p0 − ξ − ξ p0 . (11.52b)
12 c2
The little group for (p0 , ρ0 ) consists of higher-spin superrotations leaving it invariant.
Its Lie algebra is therefore spanned by pairs (X, ξ) such that the right-hand sides of
eqs. (11.52) vanish:
c2
2 X 0 p0 − X 000 + 3 ξ 0 ρ0 = 0 , (11.53a)
12
c2 (5) 192 0 2
3 X 0 ρ0 + 10 ξ 000 p0 − ξ − ξ p0 = 0 . (11.53b)
12 c2
The solutions of these equations depend on the values of p0 and ρ0 . Here we take
ρ0 = 0 for simplicity, i.e. we only consider cases where all higher-spin charges are
switched off. Then, given p0 , eqs. (11.53) become two decoupled differential equations
for the functions X(ϕ) and ξ(ϕ), leading to three different cases:
• For generic values of p0 , the only pairs (X, ξ) leaving (p0 , 0) invariant are con-
stants, and generate a little group U(1) × R.
• For p0 = −n2 c2 /96 where n is a positive odd integer, the pairs (X, ξ) leaving
(p0 , 0) invariant take the form
X(ϕ) = A, ξ(ϕ) = B + C cos(nϕ) + D sin(nϕ), (11.54)
Chapter 11. Partition functions and characters 349
where A, B, C and D are real numbers. The corresponding little group is the
n-fold cover of GL(2, R).
• For p0 = −n2 c2 /24 = −(2n)2 c2 /96 where n is a positive integer, the Lie algebra
of the little group is spanned by
X(ϕ) = A + B cos(nϕ) + C sin(nϕ),
(11.55)
ξ(ϕ) = D + E cos(nϕ) + F sin(nϕ) + G cos(2nϕ) + H sin(2nϕ),
where A, B, ..., H are real coefficients. The little group is thus an n-fold cover
of SL(3, R). In particular, p0 = −c2 /24 realizes the absolute minimum of energy
among all supermomenta belonging to orbits with energy bounded from below.
It is thus the supermomentum of the vacuum state, and indeed, upon using
c2 = 3/G, the field configuration that corresponds to it is the metric of Minkowski
space (with the spin-3 field set to zero on account of ρ0 = 0).
These results extend our earlier observations on Virasoro orbits in section 7.1.
This is a natural spin-3 extension of the spin-2 (BMS3 ) massive character (10.61), in
the same way that (11.46) generalizes Virasoro characters. It is also a flat limit of
(11.46) for N = 3.
The choice n = 1 specifies the vacuum representation of the flat W3 algebra; taking α
to be a Euclidean time translation by iβ, we get
+∞
! +∞
!
Y 1 Y 1
χvac [(rotθ , α = iβ, a = 0)] = eβc2 /24 in(θ+i) |2
· in(θ+i) |2
.
n=2
|1 − e n=3
|1 − e
(11.60)
This is one of our key results in this chapter. Indeed, comparing with eq. (11.37), we
recognize the product of the (suitably regularized) rotating one-loop partition func-
tions of massless fields with spins two and three in three-dimensional flat space. It
provides a first non-trivial check of our proposal for the construction of unitary repre-
sentations of flat WN algebras.
All the induced representations described above are unitary by construction, pro-
vided one can define (quasi-invariant) measures on the corresponding orbits. In anal-
ogy with representations of the bms3 algebra, they can also be described as induced
modules that generalize those of section 10.2; we will turn to them in section 11.3.
in analogy with (11.51). The surface charges generating these symmetries should co-
incide with the pairing of the Lie algebra of (11.61) with its dual space, and they
Chapter 11. Partition functions and characters 351
are likely to satisfy a centrally extended algebra. Since the presence of higher-spin
fields does not affect the value of the central charge in three-dimensional AdS grav-
ity [377, 378], one expects the central charge in this case to be the usual c2 = 3/G
appearing in mixed brackets. This structure was indeed observed for N = 4 in [356].
We now argue that this proposal must hold for any N by showing that the vacuum
character of (11.61), computed along the lines followed above for flat W3 , reproduces
the product of one-loop partition functions of fields of spin 2, 3, ..., N .
where the coefficients Ai , Bij , Cij are real. In principle one can obtain such symmetry
generators by looking for the little group of the vacuum as in (11.53), using for instance
the explicit formulas of [380]. Yet, a simpler way to derive the same result is to look
for the higher-spin isometries of the vacuum in the first-order formulation, in which
the theory is described by a Chern-Simons action with gauge algebra sl(N, R) Aad
(sl(N, R))Ab (see e.g. [356,382,383]). In this language, and in terms of retarded Bondi
coordinates (r, ϕ, u), the vacuum field configuration takes the form
hr i
−1 −1
Aµ (x) = b(r) g(u, ϕ) ∂µ [g(u, ϕ)b(r)] , b(r) = exp P−1 , (11.63)
2
where g(u, ϕ) is a field valued in SL(N, R) n sl(N, R), given by
1 1
g(u, ϕ) = exp P1 + P−1 u + J1 + J−1 ϕ (11.64)
4 4
in terms of Poincaré generators that satisfy the commutation relations (10.26). The
isometries of this field configuration are generated by gauge parameters of the form
(g · b)−1 Ta (g · b), where Ta is any of the basis elements of the gauge algebra. Upon
expanding g −1 Ta g as a position-dependent linear combination of gauge algebra gener-
ators, the function multiplying the lowest weight generator coincides with the corre-
sponding asymptotic symmetry parameter (see e.g. [378] for details). The latter can
be obtained as follows.
For convenience, we diagonalize the Lorentz piece of the group element (11.64) as
1
exp J1 + J−1 ϕ = SeiJ0 ϕ S −1 (11.65)
4
Chapter 11. Partition functions and characters 352
where S is some SL(2, R) matrix. Then the gauge parameters that generate the little
group of the vacuum configuration can be written as
`
X
1 m (`) 1
exp − J1 + J−1 ϕ α Wm exp J1 + J−1 ϕ (11.66a)
4 m=−`
4
`
X
−iJ0 ϕ
=Se αm S −1 Wm(`) S eiJ0 ϕ S −1 , (11.66b)
m=−`
(`)
where the αm ’s are certain real coefficients, while the Wm ’s (with 2 ≤ ` ≤ N and
(2)
−` ≤ m ≤ `) generate the sl(N, R) algebra (including Jm ≡ Wm ). Note that the
matrix S preserves the conformal weight since it is an exponential of sl(2, R) generators,
so that
X ` X `
m (`) −1
α S Wm S = α̃m Wm(`) (11.67)
m=−` m=−`
for some coefficients α̃j obtained by acting on the αm ’s with an invertible linear map.
(`)
Since each generator Wm has weight m under J0 , eq. (11.66b) can be rewritten as
`
X `
X `
X
imϕ m −1 (`)
e α̃ S Wm(`) S = β mn
Wn(`) eijϕ = eimϕ β m` W` + · · · (11.68)
m=−` m,n=−` m=−`
(`)
for some coefficients β mn . In the last step we omitted all terms proportional to Wm ’s
with m < `; the important piece is the term that multiplies the highest-weight gen-
(`)
erator W` : it is the function on the circle that generates the asymptotic symmetry
(`)
corresponding to the generator `m=−` αm Wm that we started with in (11.66a). Since
P
the β m` ’s are related to the αm ’s by an invertible linear map, and since there are 2` + 1
linearly independent generators of this type, the isometries of the vacuum exactly span
the set of functions of the form (11.62). This is what we wanted to prove; there are
N 2 − 1 linearly independent asymptotic symmetry generators of this form, and they
span the Lie algebra sl(N, R).
The character associated with the vacuum representation of (11.61) can then be
worked out exactly as in the cases N = 2 and N = 3 discussed above: using the
Fourier modes of the N − 1 components of supermomentum as coordinates on the
orbit, we need to mod out the redundant modes. For the vacuum orbit, these are the
modes ranging from −(s − 1) to (s − 1) for the sth component. The integral over the
localizing delta function in the Frobenius formula (4.33) then produces a character
N +∞
!
Y Y 1
χ[(rotθ , a1 = iβ)] = eβc2 /24 (11.69)
s=2 n=s
|1 − ein(θ+i)2 |
where we implicitly set to zero all higher-spin supertranslations except the gravita-
tional one, a1 = iβ. Comparing with (11.37), we recognize the product of one-loop
partition functions of massless higher-spin fields with spins ranging from 2 to N , in-
cluding a classical contribution. This result confirms, on the one hand, our conjecture
Chapter 11. Partition functions and characters 353
(11.61) for the asymptotic symmetry algebras of generic higher-spin theories in three-
dimensional flat space, and on the other hand it provides another consistency check of
our proposal for the characterization of unitary representations of flat WN algebras.
It is also a flat limit of the vacuum WN ⊕ WN character displayed in (11.47).
Ultrarelativistic contraction
The quantum W3 algebra is spanned by two sets of generators Lm and Wm (m ∈ Z)
whose commutation relations were displayed in (11.44). Consider now a direct sum
W3 ⊕ W3 , where the generators and the central charge of the other copy of W3 will
be denoted with a bar on top (L̄m , W̄m and c̄). Introducing a length scale ` to be
interpreted as the AdS3 radius, we define new generators Pm and Jm as in (10.32), as
well as
1
Km ≡ Wm − W̄−m , Qm ≡ Wm + W̄−m . (11.70)
`
We also define central charges c1 and c2 as in (9.93). In the limit ` → ∞, and provided
the central charges scale in such a way that both c1 and c2 are finite, one finds that
Chapter 11. Partition functions and characters 354
where the non-linear terms Ωm and Θm are quadratic operators given by (11.50), with
the exact same ordering (and calligraphic letters replaced by usual capital letters):
X X
Ωm ≡ (Pm−p Jp + Jm−p Pp ) , Θm ≡ Pm−p Pp . (11.72)
p∈Z p∈Z
The commutation relations (11.71) are quantum analogues of the Poisson brackets
(11.49), including a central charge c1 and with operators normalized so that the vac-
uum has zero eigenvalue under P0 . One can check that with the definition (11.50),
the brackets given by (10.41) and (11.71) satisfy Jacobi identities, so the generators
Jm , Km , Pn , Qn span a well-defined non-linear Lie algebra. We call it the quantum
flat W3 algebra. In any unitary representation, its generators satisfy the Hermiticity
conditions
(Qm )† = Q−m , (Km )† = K−m . (11.73)
supplemented with (10.27) for m ∈ Z.
The expressions (11.72) for the quadratic terms follow from the identities
`2 `
: Λm : + : Λ̄m : = Θm + O(`) , : Λm : − : Λ̄m : = Ωm + O(1) (11.74)
2 2
where : Λm : is the normal-ordered quadratic term (11.45) of the quantum W3 algebra,
while : Λ̄m : is its right-moving counterpart. Note, in particular, that both the linear
term in (11.45) and the mixing between positive and negative modes in (10.32)-(11.70)
are necessary to reorganize the sum of quadratic terms with the precise order of (11.72).
We shall see in section 11.3.2 that (11.50) is a normal-ordered polynomial with respect
to the natural vacuum in induced modules of the quantum flat W3 algebra.
Galilean contraction
In order to compare the quantum flat W3 algebra (11.71) with other results in the
literature [356], we now consider the non-relativistic limit of the quantum direct sum
Chapter 11. Partition functions and characters 355
Massive modules
Let us take p0 = M − c2 /24 with M > 0; this corresponds to a massive representation
of the flat W3 algebra. Assuming also that ρ0 is generic, the little group is U(1) × R
3
An interesting problem is to understand if these algebras are merely different because of an
unfortunate choice of basis, or if they are genuinely distinct in the sense that they are not isomorphic.
We will not address this issue here.
Chapter 11. Partition functions and characters 356
and the spin of the representation is therefore a pair (s, σ) ∈ R2 . Now, the plane wave
at rest Ψ(p0 ,ρ0 ) ≡ |M, ρ0 i is a state that satisfies
Here M and s are the mass and spin labels encountered in (10.42), while ρ0 and σ
are their spin-3 counterparts. As before we call |M, ρ0 i the rest frame state of the
representation and we normalize the operator P0 so that the vacuum has vanishing P0
eigenvalue.
As usual, unitarity is somewhat hidden in the induced module picture but can be
recognized in the fact that the state |M, ρ0 i is a plane wave, and that acting on it
with finite higher-spin superrotations generates an orthonormal basis of plane wave
states for the carrier space of the representation. Irreducibility can be inferred from
the same argument that we used for bms3 : by construction, a supermomentum orbit
is a homogeneous space for the action of superrotations, and this carries over to the
higher-spin setting. This implies that W3 superrotations can map any plane wave
state on any other one, which in turn implies that the space of the representation has
no non-trivial invariant subspace.
Vacuum module
The vacuum module of the flat W3 algebra can be built in direct analogy to its bms3
counterpart discussed around (10.44). The only subtlety is the enhancement of the
little group, which leads to additional conditions on superrotations. Indeed the vacuum
state |0i is now an eigenstate of all modes Pm and Qm with zero eigenvalue, and satisfies
in addition
These conditions ensure that the vacuum is invariant under the sl(3, R) wedge algebra
of the W3 subalgebra (which includes in particular the Lorentz algebra). The corre-
sponding module can then be built as usual by acting with higher-spin superrotation
generators on the vacuum state and producing states of the form (11.78), where now
Chapter 11. Partition functions and characters 357
all li ’s must be different from −1, 0, 1 and all ki ’s must be different from −2, −1, 0, 1, 2.
We stress that the li ’s and ki ’s can be positive or negative, in sharp contrast to the
non-relativistic modules investigated in [356].
The definition of the flat W3 vacuum allows us to interpret the quadratic terms
(11.72) as being normal-ordered. Indeed, their expectation values vanish in the vacuum
|0i:
h0|Θn |0i = h0|Ωn |0i = 0. (11.80)
These considerations appear to be a robust feature of “flat W algebras”: ultrarelativis-
tic contractions of WN ⊕ WN algebras always take the form
and therefore contain an Abelian ideal, where the semi-direct sum ensures that the
structure constants of the non-linear terms are always proportional to inverse powers
of the central charge. Indeed, for a non-linear operator of nth order the structure
1
constants are of order cn−1 at large c. When expanding them in powers of the con-
traction parameter `, this implies that the leading term is proportional to `1−n thanks
to (9.93). In order to obtain a finite expression, it is thus necessary that the resulting
non-linear operator consists of at least n − 1 Abelian generators. Terms of this kind
always have a vanishing expectation value in the rest frame vacuum state, although
the precise ordering in the polynomial should be fixed by other means, e.g. by defining
the algebra via a contraction of the quantum algebra or by imposing Jacobi identities.
Thus the conditions (11.77) with M = ρ0 = s = σ = 0, together with (11.79), provide
a valid definition of the vacuum for all quantum flat WN algebras.
By contrast, for a highest-weight vacuum of the type (10.49), the quadratic oper-
ators Ωm given by (11.72) generally have non-vanishing vacuum expectation values.
Thus the extra non-linear structure introduced by higher spins exhibits the fact that
the natural representations in the ultrarelativistic limit are the induced ones discussed
above, rather than the highest-weight ones of [287,356]. This difference emphasizes the
physical distinction between ultrarelativistic and Galilean limits: the former is adapted
to gravity, and more generally to models of fundamental interactions, where unitarity
is a key requirement. In particular, flat space holography (at least in the framework
of Einstein gravity) is expected to rely on the unitary construction described in this
thesis. By contrast, the Galilean viewpoint is suited to condensed matter applications,
and more generally to situations where unitarity need not hold — as was indeed ar-
gued in [287]. We stress that this difference is a genuine quantum higher-spin effect:
it is not apparent at the classical level, and it does not occur in pure gravity either.
supersymmetric BMS3 groups and their unitary representations, and note that super
BMS3 multiplets contain towers of infinitely many particles with increasing spins.
Finally, we show that the resulting characters match suitable combinations of bosonic
and fermionic one-loop partition functions.
To compute the partition function for ψ, ψ̄ one has to evaluate a path integral
(11.1) with the integration measure DψDψ̄ and S the action (11.82) or its massive
analogue. The fermionic fields live on RD /Z as defined by the group action (11.7), but
in contrast to bosons, they satisfy anti periodic boundary conditions along the thermal
cycle. For a massive field, one thus finds that the partition function is given by
1 1
log Z = log det(−∆(s+1/2) + M 2 ) − log det(−∆(s−1/2) + M 2 ) , (11.84)
2 2
where ∆(s+1/2) is the Laplacian acting on antiperiodic, symmetric, γ-traceless spinor
fields with s indices on RD /Z. For massless fields, the gauge symmetry enhancement
requires gauge-fixing and ghosts, leading to [387]
1 1
log Z = log det(−∆(s+1/2) ) − log det(−∆(s−1/2) ) + log det(−∆(s−3/2) ) . (11.85)
2 2
Eqs. (11.84) and (11.85) are fermionic analogues of the bosonic formulas (11.11) and
(11.25). To evaluate the functional determinants, we rely once more on heat kernels
and the method of images described in section 11.1.1.
Chapter 11. Partition functions and characters 359
The heat kernel KAB µs ,νs associated with the operator (−∆(s+1/2) + M 2 ) on RD is
the unique solution of
(11.86)
AB
Here K µs ,νs is a bispinor in the indices A and B, and a symmetric bitensor in the
indices µs and νs . (We use again the shorthand µs to denote a set of s symmetrized
indices.) It is also γ-traceless in the sense that
where the factor (−1)n comes from antiperiodic boundary conditions, J is the matrix
(11.8), and U is a 2bD/2c × 2bD/2c matrix acting on spinor indices in such a way that
J α β γ β = U γ α U −1 . (11.91)
In particular, a rotation by 2π around any given axis maps the field ψ on −ψ, in
accordance with the fact that spinors represent SO(D) up to a sign. Note that, using
an explicit D-dimensional representation of the γ matrices, one gets
r
Y
n bD/2c
Tr(U ) = 2 cos(nθi /2) . (11.92)
i=1
Chapter 11. Partition functions and characters 360
Now, plugging (11.90) into formula (11.3) for the determinant of −∆(s+1/2) , one obtains
a sum of integrals which can be evaluated exactly as in the bosonic case. The only
difference with respect to bosons comes from the spin structure, and the end result is
(
X (−1)n χ
(F ) ~
s [n θ,~
] e−|n|βM D odd
−log det(−∆(s+1/2) +M 2 ) = r ×
|n| Q M L
K1 (|n|βM ) D even
n ∈ Z∗ |1 − ein(θj +ij ) |2 π
j=1
(11.93)
where we have discarded a volume divergence independent of all chemical potentials
(as in eq. (11.20)), and where
) ~
χ(F ] = (J µα )s Tr I(F
)
s [nθ,~ µs ,αs (11.94)
is the fermionic analogue of (11.19), with the same rough regularization as in eq.
(11.20) (a more careful regularization will be described below for D = 3). This result
(F )
takes the same form as (11.20), up to the replacement of χs by χs and the occurrence
of (−1)n due to antiperiodicity. In appendices 11.B.1 and 11.B.2, we show that
χ(D) ~
(F ) [nθ ] for odd D,
(F ) ~ 11.B.1&11.B.2 λs
χs [nθ] = (11.95)
χ(D) ~
(F ) [nθ, 0] for even D,
λs
where the term on the right-hand side is the character of an irreducible representation
(F )
of SO(D) with highest weight λs = (s + 1/2, 1/2, ..., 1/2), written here in the dual
basis of the Cartan subalgebra of so(D) described above (11.21).
Having computed the required functional determinants on RD /Z, we can now write
down the partition functions given by (11.84) and (11.85). In the massive case, the
difference of Laplacians acting on fields with spins (s + 1/2) and (s − 1/2) produces
the difference of two factors (11.95), with labels s and s − 1. It turns out that formula
(F )
(11.23) still holds if we replace λs and λs−1 by their fermionic counterparts, λs and
(F )
λs−1 . (The proof of this statement follows the exact same steps as in the bosonic case
described in appendix 11.A.3, up to obvious replacements that account for the change
in the highest weight vector.) Accordingly, the rotating one-loop partition function of
a massive field with spin s + 1/2 is
∞ (D−1) ~ ] (
X (−1)n+1 χλ(F ) [nθ,~ e−nβM (D odd)
Z(β, θ~ ) = exp r
s
× ML
.
n Q in(θ +i ) 2 K 1 (nβM ) (D even)
n=1 |1 − e j j | π
j=1
(11.96)
In the massless case we must take into account one more difference of characters,
(F )
namely (11.27) with λs replaced by λs . For D ≥ 4, this difference can be written as
a combination of SO(D−2) characters (the proof is essentially the same as in appendix
11.A.3), and the partition function of a massless field with spin s + 1/2 exactly takes
the form (11.28) or (11.32) (for D even or odd, respectively) with an additional factor
(F )
of (−1)n+1 in the sum over n, and the replacement of λs by λs . One can also verify
that relation (11.34) remains true for fermionic partition functions.
Chapter 11. Partition functions and characters 361
Upon using this expression in the summand of (11.97) instead of the naive combination
of exponentials written there, the series in the exponential becomes
+∞ +∞
(−1)n+1 q n(s+1/2) − q n(s+1/2) q̄ n + c.c. (−1)n+1 q n(s+1/2)
X X
= + c.c.
n=1
n |1 − q n |2 n=1
n 1 − qn
+∞
X
= log(1 + q j+1/2 ) + c.c.
j=s
in terms of q = ei(θ+i) . As in the bosonic case (11.36), the regularization (11.98) has
ensured that log Z splits as the sum of a chiral and an anti-chiral function of q. After
renaming j into n, the end result is the following expression for the partition function
of a field with spin s + 1/2 in three dimensions:
+∞
2
Y
Z= 1 + ei(n+1/2)(θ+i) , (11.99)
n=s
which can also be recovered as the flat limit of the corresponding AdS result [387]. The
remainder of this chapter is devoted to relating this partition function to the vacuum
characters of various supersymmetric extensions of the BMS3 group.
and whose odd part is a Γ0 -module such that the differential of the Γ0 action be the
bracket between even and odd elements of γ [391]. Then a super semi-direct product
is a super Lie group of the form [392, 393]
G nσ A, g A (A + A) , (11.100)
where GnA is a standard (bosonic) semi-direct product group with Lie algebra g A A,
while g A (A + A) is a super Lie algebra whose odd subalgebra A is a G-module such
that the bracket between elements of g and elements of A be the differential of the
action of G on A, and such that [A, A] = 0 and {A, A} ⊆ A. By virtue of this
definition, the action of G on A is compatible with the super Lie bracket:
{g · S, g · T } = σg {S, T } ∀ S, T ∈ A , (11.101)
It was shown in [392, 393] that all irreducible, unitary representations of a super
semi-direct product are induced in essentially the same sense as for standard, bosonic
groups. In particular, they are classified by the orbits and little groups of G nσ A, as
explained in section 4.1. However, there are two important differences with respect to
the purely bosonic case:
1. Unitarity rules out all orbits on which energy can be negative, so that the mo-
mentum orbits giving rise to unitary representations of the supergroup form a
subset of the full menu of orbits available in the purely bosonic case. More
precisely, given a momentum p ∈ A∗ , it must be such that
where T (A) is the tensor algebra of A. Quotienting this algebra by its ideal
generated by the radical of A, one obtains a non-degenerate Clifford algebra C¯p .
Since A is a G-module, there exists an action of the little group Gp on C¯p ; let us
denote this action by a 7→ g · a for a ∈ C¯p and g ∈ Gp . To obtain a representation
of the full supergroup (11.100), one must find an irreducible representation τ of
C¯p and a representation R0 of Gp acting in the same space, and compatible with
τ in the sense that
τ [g · a] = R0 [g] · τ [a] · (R0 [g])−1 . (11.104)
Chapter 11. Partition functions and characters 363
For finite-dimensional groups, the pair (τ, R0 ) turns out to be unique up to mul-
tiplication of R0 by a character of Gp (and possibly up to parity-reversal). Given
such a pair, we call it the fundamental representation of the supersymmetric little
group.
The Clifford algebra (11.103) leads to a replacement of the irreducible, “spin” rep-
resentations of the little group, by generally reducible representations R0 ⊗ R. This
is the multiplet structure of supersymmetry: the restriction of an irreducible unitary
representation of a supergroup to its bosonic subgroup is generally reducible, and the
various irreducible components account for the combination of spins that gives rise
to a susy multiplet. In the Poincaré group, an irreducible supermultiplet contains
finitely many spins; by contrast, we will see below that super-BMS3 multiplets contain
infinitely many spins. Apart from this difference, the structure of induced representa-
tions of super semi-direct products is essentially the same as in the bosonic case: they
consist of wavefunctions on an orbit, taking their values in the space of the represen-
tation R0 ⊗ R. In particular, the Frobenius formula (10.54) for characters remains
valid, up to the replacement of R by R0 ⊗ R.
Here [X, Y ] is the standard Lie bracket of vector fields and the dot denotes the natural
action of vector fields on F−1/2 (S 1 ), so that X · T is the −1/2-density with component
1
X · T ≡ XT 0 − X 0 T. (11.106)
2
(This is formula (6.32) with h = −1/2.) Upon expanding the functions X(ϕ) and
S(ϕ) in Fourier modes, one recovers the standard N = 1 supersymmetric extension of
the Witt algebra. Choosing S(ϕ) to be periodic or antiperiodic leads to the Ramond
or the Neveu-Schwarz sector of the superalgebra, respectively.
The central extension of sVect(S 1 ) is the super Virasoro algebra, svir. Its elements
are triples (X, S, λ) where (X, S) ∈ sVect(S 1 ) and λ ∈ R, with a super Lie bracket
(X, S, λ), (Y, T, µ) ≡ [X, Y ] + S ⊗ T, X · T − Y · S, c(X, Y ) + h(S, T ) , (11.107)
By expanding the functions X and S in Fourier modes, one obtains the usual commu-
tation relations of N = 1 super Virasoro. Explicitly, defining the generators
The super Virasoro algebra is (half of) the asymptotic symmetry algebra of three-
dimensional supergravity with Brown-Henneaux boundary conditions [394, 395] (see
also [396]). In that context the vector field X is one of the components of an asymp-
totic Killing vector field (8.30), while S is one of the components of an asymptotic
Killing spinor. The fact that the quantization of three-dimensional supergravity pro-
duces super Virasoro representations was verified in [69] by showing that the one-loop
partition function of supergravity on thermal AdS3 coincides with the vacuum char-
acter of two super Virasoro algebras. In the remainder of this section our goal is to
describe the flat analogue of these results.
where Vect(S 1 )Ab ⊕ F−1/2 may be seen as an Abelian version of sVect(S 1 ). Again,
choosing periodic/antiperiodic boundary conditions for F−1/2 yields the Ramond/Ne-
veu-Schwarz sector of the theory (respectively). Central extensions can be included as
in (9.62) and lead to a supersymmetric version of the centrally extended algebra (9.64).
The elements of the resulting super Lie algebra sbmsd 3 are 5-tuples (X, λ; α, S, µ), where
(X, α, S) belongs to sbms3 and λ, µ are real numbers, with a super Lie bracket that
extends (9.46):
h o
(X, λ; α, S, µ), (Y, κ; β, T, ν) =
= [X, Y ], c(X, Y ); [X, β] − [Y, α], X · T − Y · S; c(X, β) − c(Y, α) + h(S, T ) .
(11.111)
Here c is again the Gelfand-Fuks cocycle (6.43) while h is given by (11.108). Upon
introducing generators analogous to (9.67), the central charges (9.47) and Qr ≡
Chapter 11. Partition functions and characters 365
(0, 0; 0, eirϕ (dϕ)−1/2 , 0), one finds the brackets (9.68) supplemented with
m
i[Jm , Qr } = − r Qm+r , (11.112a)
2
i[Pm , Qr } = 0 , (11.112b)
Z2 2
[Qr , Qs } = Pr+s + r δr+s,0 . (11.112c)
6
The indices r, s are integers/half-integers in the Ramond/Neveu-Schwarz sector, re-
spectively. Note that the centrally extended bracket of supercharges only involves the
central charge Z2 that pairs superrotations with supertranslations.
where we use the notation (11.100). This structure appears to be ubiquitous in three-
dimensional, asymptotically flat supersymmetric higher-spin theories.
Here h·, ·i is the pairing (6.111) of centrally extended supermomenta with centrally
extended supertranslations. Using the super Lie bracket (11.111), we find
Z 2π
1 c2
h(p0 , c2 ), {S, S}i = dϕ p0 (S(ϕ))2 + (S 0 (ϕ))2 . (11.116)
2π 0 6
Since the term involving (S 0 )2 can be made arbitrarily large while keeping S 2 arbitrar-
ily small, a necessary condition for (p0 , c2 ) to be admissible is that c2 be non-negative.
Already note that this condition did not arise in the bosonic BMS3 group. The admis-
sibility condition on p0 , on the other hand, depends on the sector under consideration:
• In the Ramond sector, S(ϕ) is a periodic function on the circle. In particular,
S(ϕ) = const. is part of the supersymmetry algebra, so for expression (11.116)
to be non-negative for any S, we must impose p0 ≥ 0.
• In the Neveu-Schwarz sector, S(ϕ) is antiperiodic (i.e. S(ϕ + 2π) = −S(ϕ)) and
can be expanded in Fourier modes as
X
S(ϕ) = sn+1/2 ei(n+1/2)ϕ . (11.117)
n∈Z
and the admissibility condition amounts to requiring all coefficients in this series
to be non-negative, which gives
c2
p0 ≥ − . (11.119)
24
Characters
The Fock space representations just described can be used to evaluate characters. For
example, in the massive case with spin s one finds the supersymmetric little group
character
+∞
Y
iθJ0 isθ iθ/2 3iθ/2 2iθ isθ
1 + ei(n−1/2)(θ+i) ,
tr e =e 1+e +e +e + ··· = e
n=1
(11.122)
where we have added a small imaginary part to θ to ensure convergence of the prod-
uct; the trace is taken in the fermionic Fock space associated with the “highest-weight
state” |0i. The vacuum case is similar, except that the product would start at n = 2
rather than n = 1 (and s = 0). Note that (11.122) explicitly breaks parity invari-
ance; this can be fixed by replacing the parity-breaking Fock space representations τ
described above by parity-invariant tensor products τ ⊗ τ̄ , where τ̄ is the same as τ
with the replacement of Qr by Q−r . The trace of a rotation operator in the space of
Chapter 11. Partition functions and characters 368
Comparing with (11.37) and (11.99), we recognize the product of the (suitably reg-
ularized) partition functions of two massless fields with spins 2 and 3/2, that is, the
one-loop partition function of N = 1 supergravity in three-dimensional flat space.
where r and ` are integers or half-integers, depending on the sector under consideration
(Ramond or Neveu-Schwarz, respectively). In order for the orbit to be admissible
in the sense of (11.102), the value of M must be chosen so as to ensure that all
coefficients on the right-hand side of (11.124) are non-negative. In particular, the
vacuum value M = 0 is admissible in the Neveu-Schwarz sector, in which case the
anticommutators {τ [Qr ], τ [Q−r ]} vanish for |r| = 1/2, ..., s − 1/2. Thus, in the Neveu-
Schwarz vacuum, the Clifford algebra (11.124) degenerates and τ must really be seen as
Chapter 11. Partition functions and characters 369
which generalizes (11.122). The character for τ ⊗ τ̄ is the squared norm of this expres-
sion, and the resulting vacuum character of the hypersymmetric BMS3 group is
+∞
|1 + ei(n+1/2)(θ+i) |2
Q
n=s
χhyper
vac
BMS
[(rotθ , iβ)] = eβc2 /24 +∞
. (11.126)
Q
|1 − eim(θ+i) |2
m=2
As announced earlier, this coincides with the (suitably regularized) one-loop partition
function of asymptotically flat gravity coupled to a massless field with spin s + 1/2.
We have thus completed our overview of the relation between BMS3 characters and
one-loop partition functions in three dimensions.
where " m #
s
X Y Tr[(J n )k ] k
hs (J n ) = m
. (11.128)
m1 ,...,ms ∈ N k=1
m k !k k
m1 +2m2 +...+sms = s
one can show by recursion (see e.g. [397, p. 24f]) that the polynomial (11.129) can
equivalently be written as in (11.128):
s
X Y (λk + ... + λk )mk
1 D
hs (λ1 , ..., λD ) = . (11.131)
m1 ,...,ms ∈ N k=1
mk ! k mk
m1 +2m2 +...+sms = s
We shall use this relation later. To prove (11.127), we start with the following:
Proof. The left-hand side of (11.132) can be seen as a trace over symmetric tensor
powers of J. Indeed, δ µα Jµα = Tr(J) is clear; as for 21 (δ µα )2 (Jµα )2 , one gets
2
1 µα 2 1 1X
(δ ) (Jµα )2 = Tr(J)2 + Tr J 2 = Tr S 2 (J) = Tr J i Tr S 2−i (J) ,
2 2 2 i=1
(11.133)
where S k (J) is the k th symmetric tensor power of J. One then defines recursively
s
1 µα s s s 1X
Tr J i Tr S s−i (J) ,
(δ ) (Jµα ) = Tr (S (J)) = (11.134)
s! s i=1
so that s!1 (δ µα )s (Jµα )s is just a trace in the sth symmetric tensor power of the D-
dimensional vector space V on which Jµα acts as a linear operator. Now consider an
eigenbasis {e1 , ..., eD } for Jµα , with J · ek = λk ek . Since s!1 (Jµα )s is the sth symmet-
ric tensor power of Jµα one can construct an eigenbasis for s!1 (Jµα )s by symmetrizing
ek1 ⊗ ek2 ⊗ ... ⊗ ekD , with k1 ≤ k2 ≤ ... ≤ kD . These eigenvectors have eigenvalues
λl1 λl2 ...λlD , and since (δ µα )s s!1 (Jµα )s is the trace of s!1 (Jµα )s , relation (11.132) follows
upon using the second expression of hs (λ1 , ..., λD ) in (11.129).
We can now turn to the proof of (11.127). To this end we fix conventionally the
number of terms entering the contraction of two symmetrized expressions as follows.
Objects with lower indices are symmetrized with the minimum number of terms re-
quired and without overall normalization factor, while objects with upper indices are
not symmetrized at all, since the symmetrization is induced by the contraction. This
specification is needed because terms with lower and upper indices in a contraction
Chapter 11. Partition functions and characters 371
may have a different index structure and therefore the number of terms needed for
their symmetrization may be different. For instance
δ µµ Tµs , αs = 2 δαα Tµs−2 , αs−2 , δ αα Tµs ,αs = 2 δµµ Tµs−2 ,αs−2 . (11.137)
In terms of the tensors Tµs ,αs , the mixed trace (11.19) can be written as
b 2s c
(−1)m s! [D + 2 (s − m − 2)]!!
1 s
~ = Tµs ,βs δ µβ +
X
χs [nθ] × (11.138)
s! m=1
2m m! (s − 2m)! [D + 2 (s − 2)]!!
µµ m µβ s−2m ββ m
× (δ ) (δ ) (δ )
[ 2s ]
(11.137) 1 [s] X (−1)m [D + 2 (s − m − 2)]!!
= T + × (11.139)
s! m=1
2m−1 m! (s − 2m)! [D + 2 (s − 2)]!!
× (δ µµ )m (δ µβ )s−2m (δ ββ )m−1 δµµ Tµs−2 ,βs−2 . (11.140)
To compute the trace of the (δ µµ )m (δ µβ )s−2m (δ ββ )m−1 terms, we first change our sym-
s!
metrization from δµµ Tµs−2 ,βs−2 (which contains 2(s−2)! terms) to the aforementioned
product of δ’s. In doing so one has to introduce a factor accounting for the number of
terms in each structure as
s!
δµµ Tµs−2 , βs−2 terms, (11.141a)
2(s − 2)!
s−2m m−1 s! (s − 2)!
(δ µµ )mu δ µβ δ ββ m
× m−1 terms, (11.141b)
2 m! 2 (m − 1)!(s − 2m)!
which implies
s
b2c
~ 1 [s] X (−1)m 2m−1 (m − 1)! [D + 2 (s − m − 2)]!!
χs [nθ] = T + ×
s! m=1
[(s − 2)!]2 [D + 2 (s − 2)]!! (11.142)
m s−2m m−1 µµ µs−2 ,βs−2
×δµµ δµβ δββ δ T .
Shifting m → m + 1 in the upper sum one can see that both sums are identical apart
from the overall sign and the lower extremum. Thus (11.144) boils down to
~ = 1 T [s] −
χs [nθ]
1
T [s−2] = hs (λ1 , λ2 , ..., λD ) − hs−2 (λ1 , λ2 , ..., λD ) , (11.146)
s! (s − 2)!
where λ1 , ..., λD are the eigenvalues of J n . (These eigenvalues are e±inθj for j = 1, ..., r,
and one or two unit eigenvalues depending on whether D is odd or even, respectively.)
This leads to the desired result: since traces of powers of J n can be written as
Odd D
We consider the Lie algebra so(D) = so(2r + 1), with rank r. Choosing a basis of
C2r+1 such that the Lie algebra so(2r + 1)C can be written in terms of complex ma-
trices, we may choose the Cartan subalgebra to be the subalgebra h of so(2r + 1)C
consisting of diagonal matrices. As a basis of h we choose the matrices Hi whose
entries all vanish, except the (i, i) and (r + i, r + i) entries which are 1 and −1,
respectively (with i = 1, ..., r). In our convention (11.7), the operator Hi generates ro-
tations in the plane (xi , yi ). Then, calling Li the elements of the dual basis (such that
hLi , Hj i = δij ), a dominant weight is one of the form λ = λ1 L1 + ...λr Lr ≡ (λ1 , ..., λr )
Chapter 11. Partition functions and characters 373
with λ1 ≥ ... ≥ λr ≥ 0.
Let λ be a dominant weight for so(2r + 1). According to formula (24.28) in [398],
the character of the irreducible representation of so(2r + 1) with highest weight λ is
λi +r−i+ 21 −(λi +r−i+ 12 )
qj − qj
(2r+1)
[q1 , ..., qr ] = Trλ q1H1 · · · qrHr =
χλ , (11.148)
r−i+ 12 −(r−i+ 12 )
qj − qj
where q1 , · · · qr are arbitrary complex numbers4 , Trλ denotes a trace taken in the space
of the representation, and |Aij | denotes the determinant of the matrix A with rows i
and columns j. This expression is a corollary of the Weyl character formula. Using
proposition A.60 and Corollary A.46 of [398], it can be rewritten as
(2r+1)
χλ [q1 , ..., qr ] = |hλi −i+j − hλi −i−j | , (11.149)
where hj = hj q1 , ..., qn , q1−1 , ..., qn−1 , 1 is a complete homogeneous symmetric polyno-
where λi = s δi1 . Thus for odd D the difference of symmetric polynomials in (11.127)
is just a character of SO(D).
Even D
We now turn to the Lie algebra so(2r + 2), with rank r + 1. As in the odd case
we choose a basis of C2r+2 such that we can write the Lie algebra so(2r + 2) in
terms of complex matrices and the Cartan subalgebra is generated by r + 1 diagonal
matrices Hi whose entries all vanish, except (Hi )ii = 1 and (Hi )r+1+i,r+1+i = −1.
We call Li the elements of the dual basis, and with these conventions a weight
λ = λ1 L1 + ... + λr+1 Lr+1 ≡ (λ1 , ..., λr+1 ) is dominant if λ1 ≥ λ2 ≥ ... ≥ λr ≥ |λr+1 |.
Let λ be a dominant weight for so(2r + 2). Then formula (24.40) in [398] gives the
character of the associated highest-weight representation as
h i
(2r+2) Hr+1
χλ [q1 , ..., qr+1 ] = Trλ q1H1 · · · qr+1
−(λi +r+1−i) −(λi +r+1−i)
qjλi +r+1−i + qj + qjλi +r+1−i − qj (11.151)
= ,
−(r+1−i)
qjr+1−i + qj
4
Eventually these numbers will be exponentials of angular potentials, so they are fugacities asso-
ciated with the rotation generators Hi .
Chapter 11. Partition functions and characters 374
where we use the same notations as in (11.148), except that now i, j = 1, ..., r +1. Note
that the second term in the numerator of this expression vanishes whenever λr+1 = 0
−(λ +r+1−i)
(because the (r + 1)th row of the matrix qjλi +r+1−i − qj i vanishes). Since this
is the case that we will be interested in, we may safely forget about that second term
from now on. Alternatively, for the mixed traces (11.19) that we need, we may take
qj = einθj for j = 1, ..., r and qr+1 = 1 without loss of generality, so that this second
term vanishes again. Using proposition A.64 of [398], one can then rewrite (11.151) as
(2r+2)
[q1 , ..., qr , 1] = |hλi −i+j − hλi −i−j | ,
χλ (11.152)
where hj = hj q1 , ..., qr , 1, q1−1 , ..., qr−1 , 1 . Finally, using the same arguments as for
odd D, one easily verifies that the determinant on the right-hand side of (11.152)
reduces once more to hs − hs−2 for a highest weight λs = (s, 0, ..., 0). Writing again
qj = einθj , one concludes that, for even D,
Here θ~ = (θ1 , .., θr ), λs is the weight with components (s, 0, ..., 0) in the basis defined
above equations (11.148) and (11.151), and the hat denotes omission of an argument,
while the coefficients Ark are the quotients of determinants defined in (11.31). Note
that, when one of the angles θ1 , ..., θr vanishes, say θ` = 0, then Ark = δk` and relation
(11.154b) reduces to
(2r) (2r) (2r−1)
χλs [θ~ ] − χλs−1 [θ~ ] = χλs [θ1 , ..., θb` , ..., θr ] . (11.155)
θ` =0 θ` =0
so that in particular
r
~ = |B |θk =0 .
Ark (θ) (11.157)
|B r |
Chapter 11. Partition functions and characters 375
to denote the determinant of the r × r matrix missing the angle θk of any of the
matrices defined in (11.156). As a preliminary step towards the proof, we list the four
following identities:
|Ar |
Qr = 2r−1 |B r | , (11.159a)
j=1 sin (θj /2)
(r−1)(r−2) Y
| cos [(r − i)θj ] | = 2 2 (cos(θi ) − cos(θj )) , (11.159b)
1≤i<j≤r
r
|B r |θk =0 Y
= 2r−1 (−1) k+1
sin (θj /2) , (11.159c)
Ar−1 [θk ] j=1
j6=k
r
X
r
|B | = |B r |θk =0 . (11.159d)
k=1
with some irrelevant real coefficients ck , and that the contribution of the second term of
this expression to the determinant | cos[(r−i)θj ]| vanishes by linear dependence. Equa-
tion (11.159c) then follows from (11.159a) and (11.159b), while property (11.159d) can
again be proved by induction on r.
Thanks to eqs. (11.159), we can tackle the proof of (11.154). Equation (11.154a)
(2r+1)
is easy: using expression (11.150) for the character χλs , we can write the difference
of characters on the left-hand side of (11.154a) as
r
(−1)k+1 2 cos[(s + r − 1)θk ] sin (θk /2) Ar−1 [θk ]
P
(2r+1) (2r+1) k=1
χλs − χλs−1 = . (11.161)
|Ar |
Property (11.159a) then allows us to reduce this expression to the quotient of denom-
inators appearing in the middle of eq. (11.153) (with the replacement of r + 1 by r
(2r)
and all angles non-zero), which is indeed the sought-for character χλs [θ~ ].
Equation (11.154b) requires more work. Using once more the expression in the
middle of (11.153), we first rewrite the left-hand side of (11.154b) as
r
(−1)k+1 (−2 sin[(s + r − 23 )θk ] sin (θk /2) B r−1 [θk ]
P
(2r) (2r) k=1
χλs − χλs−1 = . (11.162)
|B r |
Chapter 11. Partition functions and characters 376
Let us now recover this expression as a combination of characters of SO(2r − 1): using
formula (11.150) and the identities (11.159), one finds
r
X (2r−1)
χ λs [θ1 , ..., θbk , ..., θr ]|B r |θk =0
k=1
r r
(11.159c) X
k+1 r−1
Y
= (−1) 2 sin (θj /2) ×
k=1 j=1
j6=k
k−1
X
× (−1)j+1 sin[(s + r − 32 )θj ]Ar−2 [θj , θk ]
j=1
r
X
j
+ (−1) sin[(s + r − 3
)θ ]Ar−2 [θj , θk ]
2 j
j=k+1
r
(11.159a) X
= (−1)k+1 22r−4 sin[(s + r − 23 )θk ] sin (θk /2) ×
k=1
k−1
X r
Y
j r−2
× (−1) B [θj , θk ] sin2 (θi /2)
j=1 i=1
i∈{j,k}
/
r
X Yr
j+1 r−2 2
+ (−1) B [θj , θk ] sin (θi /2)
j=k+1 i=1
i∈{j,k}
/
r
(11.159b) X
= (−1)k+1 (−2) sin[(s + r − 32 )θk ] sin (θk /2)
k=1
" k−1 r
#
X X
(−1)j B r−1 [θk ] θj =0
+ B r−1 [θk ] θj =0
j=1 j=k+1
r
(11.159d) X
= (−1)k+1 (−2) sin[(s + r − 23 )θk ] sin (θk /2) B r−1 [θk ]. (11.163)
k=1
This coincides with the numerator of the right-hand side of (11.162), so identity
(11.154b) follows with Ark given by (11.157).
Here λj is the weight (j, 0, ..., 0) as explained above (11.21) or below (11.149), and
~ is the quotient (11.31) or (11.157). Since the proofs of these two identities are
Ark (θ)
very similar, we will only display the proof of (11.164a).
b s−1
2
c
X (−1)m+1 [D + 2(s − m − 2)]!!
+ ×
m=0
2m m!(s − 2m − 1)![D + 2(s − 1)]!!
× Tr[Tµs−1 ,βs−1 γµ γ µ (δ µµ )m (δ µβ )s−2m−1 (δ ββ )m U n ], (11.166)
where T [s] is the notation (11.136). In the first term of this expression, we shift the
symmetrization on the δ’s i.e. we exchange upper and lower indices while taking
into account the change in multiplicities of the terms involved; in all other terms, we
compute one contraction with δ ββ . Eq. (11.166) then simplifies to
) ~ 1 [s] 1 s−1 n
χ(F
s [nθ] = T Tr[U n ] − Tr[T µs−1 ,βs−1 γµ γ µ δµβ U ] (11.167)
s! [(s − 1)!]2 [D + 2(s − 1)]
b 2s c
X (−1)m 2m−1 (m − 1)![D + 2(s − m − 1)]!! µs−2 ,βs−2 µµ m s−2m m−1
+ 2 [D + 2(s − 1)]!!
T δ δµµ δµβ δββ Tr[U n ]
m=1
[(s − 2)!]
(−1)m+1 2m − 1(m − 1)![D + 2(s − m − 2)]!!
µs−2 ,βs−2 µµ m s−2m−1 m n
+ Tr[T δ δµµ δµβ δββ γµ γβ U ] .
[(s − 2)!]2 [D + 2(s − 1)]!!
Chapter 11. Partition functions and characters 378
Odd D
The character of a half-spin representation of so(2r + 1) with a dominant highest
weight λ = (λ1 + 21 , λ2 + 12 , ..., λr + 21 ) is [399, p.258f]
r
!
|sin [(λ + r − i + 1) θ ]| |sin [(λi + r − i + 1) θj ]|
i j =
(2r+1)
Y
2 cos θ2i
χλ [θ1 , ..., θr ] = 1
.
sin r − i + 2 θj i=1
|sin [(r − i + 1) θj ]|
(11.170)
Owing to expression (11.92) for the trace of U n , the second equality in (11.95) is
equivalent to
|sin [(λi + r − i + 1) θj ]|
hs (J) − hs−1 (J) = (11.171)
|sin [(r − i + 1) θj ]|
for λi = sδi1 . To prove this, consider the difference of the bosonic character (11.150)
and the right-hand side of (11.171):
sin[(λi + r − i + 12 )θj ] |sin[(λi + r − i + 1)θj ]|
1
− . (11.172)
sin[(r − i + 2 )θj ] |sin[(r − i + 1)θj ]|
Chapter 11. Partition functions and characters 379
and plugging this property in (11.174) one sees that (11.172) is just hs−1 (J) − hs−2 (J).
Since the first term of (11.172) equals hs (J)−hs−2 (J) by virtue of (11.150), this proves
(11.171).
Even D
The character of an irreducible representation of so(2r + 2) with (dominant) highest-
weight λ = (λ1 + 1/2, ..., λr+1 + 1/2) can be written as [399, p.258-259]
3
r+1 3
(2r+2) cos λ i + r − i + θ j
Y cos λi + r − i + θj
2
2 cos θ2i 2 ,
χλ [θ1 , ..., θr ] = = 3
|cos [(r − i + 1) θj ]| i=1
cos r − i + 2 θj
(11.176)
where we are including the possibility of a non-zero angle θr+1 (while in (11.95) we
take θr+1 = 0). Taking into account (11.92), proving the second equality in (11.95)
amounts to showing that
cos λi + r − i + 32 θj
hs (J) − hs−1 (J) = (11.177)
cos r − i + 23 θj
θr+1 =0
for λi = sδi1 . To prove this we proceed as in the odd-dimensional case: the difference
of the bosonic character (11.153) and the right-hand side of (11.177),
can be written as
r+1 r+1
k+1 r k+1 1 r
P P
(−1) cos[(s + r)θk ]B [θk ] (−1) 2 cos[(s + r + 2 )θk ]B [θk ]
k=1 k=1
−
|B r+1 | |B r+1 |
θr+1 =0
(11.179)
upon expanding the determinants along the first row and using the notation (11.156)-
(11.173). One can then verify that this reduces to hs−1 (J) − hs−2 (J) by the same
argument as in the odd-dimensional case. By virtue of the second equality in (11.153),
this proves (11.177).
5
See e.g. [399, p.259].
Chapter 12
Conclusion
Quantum symmetries
The overarching theme of this thesis has been group theory and its application to
quantum systems with symmetries. Accordingly, parts I and II of this thesis were
devoted to a broad overview of group representations and geometry. In particular we
have motivated and introduced central extensions, defined induced representations,
applied them to semi-direct products and relativistic symmetry groups, and explained
how classical mechanical systems with symmetries become symmetric quantum sys-
tems upon “replacing Poisson brackets by commutators”. We have also applied some of
these tools to the Virasoro group — the symmetry group of two-dimensional conformal
field theories — and used it to analyse certain properties of three-dimensional gravity
on Anti-de Sitter backgrounds.
BMS3 particles
Part III of the thesis was devoted to the application of group theoretic methods to
the study of asymptotically flat quantum gravity in three dimensions. In that context
our weapon of choice has been the BMS group in three dimensions, which we have
introduced as an asymptotic symmetry group in chapter 9, before working out its ab-
stract definition independently of gravity. We have seen in particular that it enjoys
an exceptional structure of the type G n g, where G is the Virasoro group spanned
by superrotations while g is its Lie algebra, spanned by supertranslations. A crucial
implication of this structure was that irreducible unitary representations of the BMS3
group, i.e. BMS3 particles, have supermomenta that span coadjoint orbits of the Vira-
soro group. This observation has allowed us to classify all such representations thanks
to the classification of Virasoro orbits exposed earlier, in chapter 7.
381
Chapter 12. Conclusion 382
This being said, one should not be overly enthusiastic about what we have achieved:
in essence we have worked out the flat space analogue of results that were mostly al-
ready known in the framework of AdS3 /CFT2 . In fact, in many cases our results were
flat limits of their AdS peers, although the BMS3 approach often led us to consider
Chapter 12. Conclusion 383
slightly different questions and use somewhat different methods than those suggested
by conformal symmetries. Thus we have indeed made progress in our understanding
of flat space holography, but whether this opens new doors towards quantum gravity
is a whole other matter.
There are two simple arguments that show why our work is not quite quantum grav-
ity. First, what we have studied are irreducible unitary representations of asymptotic
symmetry groups, while realistic gravitational systems are expected to form highly
reducible representations. In essence, saying that we have studied quantum gravity
by studying irreducible representations of BMS3 would be tantamount to saying that
relativistic one-particle quantum mechanics is the same as quantum field theory, which
is of course untrue. Secondly, one should realize that our study of the symmetries of
gravity hasn’t taught us anything about the microscopic details of gravity itself. For
instance, the fact that the phase space of gravity forms the coadjoint representation of
the asymptotic symmetry group is merely a restatement of the fact that momentum
maps belong to the coadjoint representation, which is a robust feature of all symmetric
phase spaces; it does not tell us anything about the details of gravity.
A look forward
Our observations on BMS particles in three dimensions have allowed us to describe
dressed particles in a group-theoretic framework. While we haven’t described inter-
acting particles in this thesis, it is likely that our methods do apply to such cases as
well. In particular, describing scattering phenomena in terms of BMS particles instead
of standard (naked) particles should incorporate soft graviton contributions. In three
dimensions this could presumably be used to describe, say, the merger of two parti-
cles into a flat space cosmology; optimistically, BMS3 representations might then even
account for gravitational quantum corrections to such amplitudes! In four dimensions
the situation is much less well understood, for reasons that we alluded to earlier. In
that case the problem is much more basic, since the very definition of BMS symmetry
is elusive — let alone its quantum representations.
A related project is the description of BMS world lines. The reader may recall that
we described in chapter 5 a general procedure for building world line actions associ-
ated with arbitrary Lie groups, and that the application of this method to the Poincaré
group resulted in relativistic world lines. It is tempting to ask what happens when
that approach is applied to the BMS3 group; the answer is very natural: the resulting
action principle describes world lines propagating in the space of supertranslations,
which can equivalently be seen as relativistic world lines dressed with gravitational
degrees of freedom. Yet another way to think of these world-lines is to interpret them
as two-dimensional field theories dual to three-dimensional asymptotically flat grav-
ity, and indeed one finds that their partition functions coincide with gravitational
(one-loop) partition functions. These considerations should appear soon in a separate
publication [400]. Note that, as before, the application of these ideas to BMS in four
dimensions is much more problematic due to the lack of a proper definition of BMS4
symmetry.
Chapter 12. Conclusion 384
There is undoubtedly much more to be done in the future, both in toy models such
as three-dimensional gravity and in real-world, four-dimensional systems. In the wake
of the experimental observation of gravitational waves [186], it is likely that new tools
and methods will soon be required to understand and study gravity, both classically
and quantum-mechanically. On a more philosophical note, it is remarkable that a
question seemingly as simple as “what is a particle?” has an answer as intricate and
rich as what we have been attempting to describe in this thesis. We hope to have
contributed to a partial solution to the problem, and look forward to investigating
some of its future applications.
[1] H. Bondi, “Gravitational Waves in General Relativity,” Nature 186 (1960), no. 4724,
535–535. (Cited on pages 2, 257, and 311.)
[2] H. Bondi, M. G. J. van der Burg, and A. W. K. Metzner, “Gravitational waves in
general relativity. 7. Waves from axisymmetric isolated systems,” Proc. Roy. Soc. Lond.
A269 (1962) 21–52. (Cited on pages 2, 4, 5, 257, and 311.)
[3] R. Sachs, “Asymptotic symmetries in gravitational theory,” Phys. Rev. 128 (1962)
2851–2864. (Cited on pages 2, 4, 5, 257, and 311.)
[4] R. K. Sachs, “Gravitational waves in general relativity. 8. Waves in asymptotically flat
space-times,” Proc. Roy. Soc. Lond. A270 (1962) 103–126. (Cited on pages 2, 4, 5,
257, and 311.)
[5] G. Barnich and C. Troessaert, “Symmetries of asymptotically flat 4 dimensional space-
times at null infinity revisited,” Phys. Rev. Lett. 105 (2010) 111103, 0909.2617. (Cited
on pages 6, 312, and 313.)
[6] G. Barnich and C. Troessaert, “Aspects of the BMS/CFT correspondence,” JHEP 05
(2010) 062, 1001.1541. (Cited on pages 6, 217, 230, 233, 234, 257, 265, 266, 312,
and 313.)
[7] T. Banks, “A Critique of pure string theory: Heterodox opinions of diverse dimensions,”
hep-th/0306074. (Cited on pages 6 and 312.)
[8] A. A. Belavin, A. M. Polyakov, and A. B. Zamolodchikov, “Infinite Conformal Symme-
try in Two-Dimensional Quantum Field Theory,” Nucl. Phys. B241 (1984) 333–380.
(Cited on page 6.)
[9] G. ’t Hooft, “Dimensional reduction in quantum gravity,” in Salamfest 1993:0284-296,
pp. 0284–296. 1993. gr-qc/9310026. (Cited on pages 6 and 312.)
[10] L. Susskind, “The World as a hologram,” J. Math. Phys. 36 (1995) 6377–6396,
hep-th/9409089. (Cited on pages 6 and 312.)
[11] J. D. Bekenstein, “Black holes and entropy,” Phys. Rev. D7 (1973) 2333–2346. (Cited
on pages 6 and 312.)
[12] S. W. Hawking, “Particle Creation by Black Holes,” Commun. Math. Phys. 43 (1975)
199–220. [,167(1975)]. (Cited on pages 6 and 312.)
[13] J. M. Maldacena, “The Large N limit of superconformal field theories and supergravity,”
Int. J. Theor. Phys. 38 (1999) 1113–1133, hep-th/9711200. [Adv. Theor. Math. Phys.
2, 231 (1998)]. (Cited on pages 6, 238, and 312.)
[14] J. D. Brown and M. Henneaux, “Central Charges in the Canonical Realization of
Asymptotic Symmetries: An Example from Three-Dimensional Gravity,” Commun.
Math. Phys. 104 (1986) 207–226. (Cited on pages 7, 9, 217, 225, and 229.)
385
Bibliography 386
[15] A. Strominger, “Black hole entropy from near horizon microstates,” JHEP 02 (1998)
009, hep-th/9712251. (Cited on page 7.)
[16] M. Bañados, C. Teitelboim, and J. Zanelli, “The Black hole in three-dimensional space-
time,” Phys. Rev. Lett. 69 (1992) 1849–1851, hep-th/9204099. (Cited on pages 7
and 217.)
[17] M. Bañados, M. Henneaux, C. Teitelboim, and J. Zanelli, “Geometry of the (2+1) black
hole,” Phys. Rev. D48 (1993) 1506–1525, gr-qc/9302012. [Erratum: Phys. Rev.D88,
069902 (2013)]. (Cited on pages 7, 217, and 238.)
[18] G. Barnich and C. Troessaert, “BMS charge algebra,” JHEP 12 (2011) 105, 1106.0213.
(Cited on pages 8, 17, and 313.)
[19] A. Strominger, “On BMS Invariance of Gravitational Scattering,” JHEP 07 (2014)
152, 1312.2229. (Cited on pages 8, 9, 253, 265, 307, and 312.)
[20] A. Ashtekar, J. Bicak, and B. G. Schmidt, “Asymptotic structure of symmetry reduced
general relativity,” Phys. Rev. D55 (1997) 669–686, gr-qc/9608042. (Cited on pages 8,
257, and 323.)
[21] F. C. Klein, “A comparative review of recent researches in geometry,” ArXiv e-prints
(July, 2008) 0807.3161. English translation by M. W. Haskell. (Cited on page 8.)
[22] E. P. Wigner, Gruppentheorie und ihre Anwendung auf die Quantenmechanik der Atom-
spektren. Springer, 1931. (Cited on pages 8 and 17.)
[23] E. P. Wigner, “On Unitary Representations of the Inhomogeneous Lorentz Group,”
Annals Math. 40 (1939) 149–204. [Reprint: Nucl. Phys. Proc. Suppl. 6, 9 (1989)].
(Cited on pages 9 and 71.)
[24] T. He, V. Lysov, P. Mitra, and A. Strominger, “BMS supertranslations and Weinberg’s
soft graviton theorem,” JHEP 05 (2015) 151, 1401.7026. (Cited on pages 9, 265,
and 312.)
[25] S. Weinberg, “Infrared photons and gravitons,” Phys. Rev. 140 (1965) B516–B524.
(Cited on page 9.)
[26] F. Cachazo and A. Strominger, “Evidence for a New Soft Graviton Theorem,”
1404.4091. (Cited on pages 9, 265, and 312.)
[27] D. Kapec, V. Lysov, S. Pasterski, and A. Strominger, “Semiclassical Virasoro symmetry
of the quantum gravity S-matrix,” JHEP 08 (2014) 058, 1406.3312. (Cited on pages 9,
265, and 312.)
[28] F. Bloch and A. Nordsieck, “Note on the Radiation Field of the electron,” Phys. Rev.
52 (1937) 54–59. (Cited on page 9.)
[29] P. A. M. Dirac, “Gauge invariant formulation of quantum electrodynamics,” Can. J.
Phys. 33 (1955) 650. (Cited on page 9.)
[30] V. Chung, “Infrared Divergence in Quantum Electrodynamics,” Phys. Rev. 140 (1965)
B1110–B1122. (Cited on page 9.)
[31] T. W. B. Kibble, “Coherent states and infrared divergences,” Lect. Theor. Phys. 11D
(1969) 387–478. (Cited on page 9.)
[32] T. W. B. Kibble, “Coherent soft-photon states and infrared divergences. ii. mass-shell
singularities of green’s functions,” Phys. Rev. 173 (1968) 1527–1535. (Cited on page 9.)
[33] T. W. B. Kibble, “Coherent soft-photon states and infrared divergences. iii. asymptotic
states and reduction formulas,” Phys. Rev. 174 (1968) 1882–1901. (Cited on page 9.)
Bibliography 387
[34] T. W. B. Kibble, “Coherent soft-photon states and infrared divergences. iv. the scat-
tering operator,” Phys. Rev. 175 (1968) 1624–1640. (Cited on page 9.)
[35] P. P. Kulish and L. D. Faddeev, “Asymptotic conditions and infrared divergences in
quantum electrodynamics,” Theor. Math. Phys. 4 (1970) 745. [Teor. Mat. Fiz. 4, 153
(1970)]. (Cited on page 9.)
[36] D. Zwanziger, “Reduction formulas for charged particles and coherent states in quan-
tum electrodynamics,” Phys. Rev. D7 (1973) 1082–1099. (Cited on page 9.)
[37] G. Barnich and G. Compère, “Classical central extension for asymptotic symmetries at
null infinity in three spacetime dimensions,” Class. Quant. Grav. 24 (2007) F15–F23,
gr-qc/0610130. (Cited on pages 9, 257, and 265.)
[38] P. J. McCarthy, “Representations of the Bondi-Metzner-Sachs Group. I. Deter-
mination of the Representations,” Proceedings of the Royal Society of London A:
Mathematical, Physical and Engineering Sciences 330 (1972), no. 1583, 517–535,
http://rspa.royalsocietypublishing.org/content/330/1583/517.full.pdf.
(Cited on pages 9, 311, and 312.)
[39] P. J. McCarthy, “Structure of the Bondi-Metzner-Sachs Group,” Journal of Mathemat-
ical Physics 13 (1972), no. 11, 1837–1842. (Cited on pages 9, 311, and 312.)
[40] P. J. McCarthy, “Representations of the Bondi-Metzner-Sachs Group. II. Properties
and Classification of the Representations,” Proceedings of the Royal Society of London
A: Mathematical, Physical and Engineering Sciences 333 (1973), no. 1594, 317–336,
http://rspa.royalsocietypublishing.org/content/333/1594/317.full.pdf.
(Cited on pages 9 and 311.)
[41] P. J. McCarthy, “Asymptotically flat space-times and elementary particles,” Phys. Rev.
Lett. 29 (1972) 817–819. (Cited on pages 9 and 311.)
[42] J. Rotman, An Introduction to the Theory of Groups. Graduate Texts in Mathematics.
Springer New York, 1999. (Cited on page 10.)
[43] R. Abraham and J. Marsden, Foundations of Mechanics. AMS Chelsea publishing.
AMS Chelsea Pub./American Mathematical Society, 1978. (Cited on pages 10, 39,
97, 106, 107, 110, and 151.)
[44] J. Lee, Introduction to Smooth Manifolds. Graduate Texts in Mathematics. Springer,
2003. (Cited on page 10.)
[45] G. Barnich and B. Oblak, “Holographic positive energy theorems in three-dimensional
gravity,” Class. Quant. Grav. 31 (2014) 152001, 1403.3835. (Cited on pages 11, 217,
241, and 257.)
[46] G. Barnich and B. Oblak, “Notes on the BMS group in three dimensions: II. Coadjoint
representation,” JHEP 03 (2015) 033, 1502.00010. (Cited on pages 11, 127, 217, 257,
280, 293, and 312.)
[47] G. Barnich and B. Oblak, “Notes on the BMS group in three dimensions: I. Induced
representations,” JHEP 06 (2014) 129, 1403.5803. (Cited on pages 11, 257, and 293.)
[48] B. Oblak, “Characters of the BMS Group in Three Dimensions,” Commun. Math. Phys.
340 (2015), no. 1, 413–432, 1502.03108. (Cited on pages 12, 78, 293, 312, and 324.)
[49] G. Barnich, H. A. González, A. Maloney, and B. Oblak, “One-loop partition function of
three-dimensional flat gravity,” JHEP 04 (2015) 178, 1502.06185. (Cited on pages 12,
293, 324, 331, 340, and 341.)
Bibliography 388
[50] A. Campoleoni, H. A. González, B. Oblak, and M. Riegler, “Rotating Higher Spin Par-
tition Functions and Extended BMS Symmetries,” JHEP 04 (2016) 034, 1512.03353.
(Cited on pages 12, 78, 293, 324, and 331.)
[51] A. Campoleoni, H. A. González, B. Oblak, and M. Riegler, “BMS Modules in Three
Dimensions,” Int. J. Mod. Phys. A31 (2016), no. 12, 1650068, 1603.03812. (Cited on
pages 12, 293, 314, and 331.)
[52] H. Afshar, S. Detournay, D. Grumiller, and B. Oblak, “Near-Horizon Geometry and
Warped Conformal Symmetry,” JHEP 03 (2016) 187, 1512.08233. (Cited on pages 12,
64, 162, 163, and 164.)
[53] S. Weinberg, The Quantum Theory of Fields. Cambridge University Press, 1995. (Cited
on pages 15, 17, 61, 71, and 92.)
[54] V. Ovsienko and S. Tabachnikov, Projective Differential Geometry Old and New: From
the Schwarzian Derivative to the Cohomology of Diffeomorphism Groups. Cambridge
Tracts in Mathematics. Cambridge University Press, 2004. (Cited on pages 15, 159,
and 166.)
[55] B. Khesin and R. Wendt, The Geometry of Infinite-Dimensional Groups. A series of
modern surveys in mathematics. Springer Berlin Heidelberg, 2008. (Cited on pages 15,
97, 105, 145, 151, 156, and 179.)
[56] L. Guieu and C. Roger, L’algèbre et le groupe de Virasoro. Publications du CRM,
Université de Montréal, 2007. (Cited on pages 15, 29, 145, 151, 156, 158, 159, 162,
163, 171, 179, 196, 207, and 363.)
[57] G. Tuynman and W. Wiegerinck, “Central extensions and physics,” Journal of Geom-
etry and Physics 4 (1987), no. 2, 207 – 258. (Cited on page 15.)
[58] A. Weinstein, “Groupoids: Unifying Internal and External Symmetry - A Tour Through
Some Examples,” Notices of the AMS 43 (1996) 744–752. (Cited on page 17.)
[59] M. Crainic and R. L. Fernandes, “Lectures on integrability of Lie brackets,” Lectures
on Poisson Geometry, Geom. Topol. Monogr. 17 (2011) 1–107. (Cited on page 17.)
[60] G. Barnich, “A Note on gauge systems from the point of view of Lie algebroids,” AIP
Conf. Proc. 1307 (2010) 7–18, 1010.0899. (Cited on pages 17, 232, and 264.)
[61] M. Peskin and D. Schroeder, An Introduction to Quantum Field Theory. Advanced
book classics. Addison-Wesley Publishing Company, 1995. (Cited on page 17.)
[62] A. Barut and R. Rączka, Theory of Group Representations and Applications. World
Scientific, 1986. (Cited on pages 32, 40, 46, 47, 55, 57, 58, 61, 64, 70, 71, and 314.)
[63] G. W. Mackey, Induced representations of groups and quantum mechanics. Publicazioni
della Classe di Scienze della Scuola Normale Superiore di Pisa. W. A. Benjamin, 1968.
(Cited on pages 32, 55, and 61.)
[64] W. Rudin, Principles of Mathematical Analysis. International series in pure and applied
mathematics. McGraw-Hill, 1976. (Cited on page 32.)
[65] W. Rudin, Real and complex analysis. Mathematics series. McGraw-Hill, 1987. (Cited
on page 32.)
[66] H. Royden and P. Fitzpatrick, Real Analysis. Featured Titles for Real Analysis Series.
Prentice Hall, 2010. (Cited on page 37.)
[67] G. Barnich, F. Brandt, and M. Henneaux, “Local BRST cohomology in gauge theories,”
Phys. Rept. 338 (2000) 439–569, hep-th/0002245. (Cited on page 43.)
Bibliography 389
[68] M. Nakahara, Geometry, Topology and Physics. Graduate student series in physics.
Taylor & Francis, 2003. (Cited on page 44.)
[69] J. R. David, M. R. Gaberdiel, and R. Gopakumar, “The Heat Kernel on AdS(3) and
its Applications,” JHEP 04 (2010) 125, 0911.5085. (Cited on pages 46, 331, 346,
and 364.)
[70] M. R. Gaberdiel, R. Gopakumar, and A. Saha, “Quantum W -symmetry in AdS3 ,”
JHEP 02 (2011) 004, 1009.6087. (Cited on pages 46, 331, 335, 338, 340, and 346.)
[71] R. Camporesi, “Harmonic analysis and propagators on homogeneous spaces,” Phys.
Rept. 196 (1990) 1–134. (Cited on page 46.)
[72] R. Camporesi and A. Higuchi, “On the Eigenfunctions of the Dirac operator on spheres
and real hyperbolic spaces,” J. Geom. Phys. 20 (1996) 1–18, gr-qc/9505009. (Cited
on page 46.)
[73] P. K. Aravind, “The Wigner angle as an anholonomy in rapidity space,” American
Journal of Physics 65 (1997), no. 7, 634–636. (Cited on page 47.)
[74] M. S. Williamson, M. Ericsson, M. Johansson, E. Sjoqvist, A. Sudbery, and V. Vedral,
“Global asymmetry of many-qubit correlations: A lattice gauge theory approach,”
Phys. Rev. A84 (2011) 032302, 1102.5609. (Cited on page 47.)
[75] J.-P. Antoine, Irreversibility and Causality Semigroups and Rigged Hilbert Spaces: A
Selection of Articles Presented at the 21st International Colloquium on Group Theoret-
ical Methods in Physics (ICGTMP) at Goslar, Germany, July 16–21, 1996, ch. Quan-
tum mechanics beyond Hilbert space, pp. 1–33. Springer Berlin Heidelberg, Berlin,
Heidelberg, 1998. (Cited on page 50.)
[76] R. de la Madrid, “The role of the rigged Hilbert space in quantum mechanics,” European
Journal of Physics 26 (Apr., 2005) 287–312, quant-ph/0502053. (Cited on page 50.)
[77] J. Dieudonné and P. Dugac, Abrégé d’histoire des mathématiques: 1700-1900. Her-
mann, 1986. (Cited on page 52.)
[78] T. Gowers, J. Barrow-Green, and I. Leader, The Princeton Companion to Mathematics.
Princeton University Press, 2010. (Cited on page 52.)
[79] S. Sternberg, Group Theory and Physics. Cambridge University Press, 1995. (Cited
on pages 52 and 54.)
[80] H. Joos and R. Schrader, “On the primitive characters of the Poincaré group,” Comm.
Math. Phys. 7 (1968), no. 1, 21–50. (Cited on pages 52 and 78.)
[81] A. A. Kirillov, Elements of the Theory of Representations. Grundlehren der mathema-
tischen Wissenschaften. Springer-Verlag, 1976. (Cited on page 52.)
[82] M. F. Atiyah and R. Bott, “A Lefschetz Fixed Point Formula for Elliptic Complexes: II.
Applications,” Annals of Mathematics 88 (1968), no. 3, 451–491. (Cited on page 54.)
[83] M. F. Atiyah and R. Bott, “A Lefschetz Fixed Point Formula for Elliptic Complexes:
I,” Annals of Mathematics 86 (1967), no. 2, 374–407. (Cited on page 54.)
[84] G. W. Mackey, “On Induced Representations of Groups,” American Journal of Math-
ematics 73 (1951), no. 3, 576–592. (Cited on page 55.)
[85] G. W. Mackey, “Induced Representations of Locally Compact Groups I,” Annals of
Mathematics 55 (1952), no. 1, 101–139. (Cited on page 55.)
[86] G. W. Mackey, “Induced Representations of Locally Compact Groups II. The Frobenius
Reciprocity Theorem,” Annals of Mathematics 58 (1953), no. 2, 193–221. (Cited on
page 55.)
Bibliography 390
[106] L. Thomas, “The motion of the spinning electron,” Nature 117 (April, 1926) 514–514.
(Cited on page 83.)
[107] B. Binegar, “Relativistic Field Theories in Three-dimensions,” J. Math. Phys. 23 (1982)
1511–1517. (Cited on page 85.)
[108] D. R. Grigore, “The Projective unitary irreducible representations of the Poincaré
group in (1+2)-dimensions,” J. Math. Phys. 34 (1993) 4172–4189, hep-th/9304142.
(Cited on page 85.)
[109] B. Oblak, “From the Lorentz Group to the Celestial Sphere,” Notes de la Septième
BSSM, U.L.B. (2015). 1508.00920. (Cited on pages 87 and 260.)
[110] J. M. Leinaas and J. Myrheim, “On the theory of identical particles,” Nuovo Cim. B37
(1977) 1–23. (Cited on page 89.)
[111] F. Wilczek, “Quantum Mechanics of Fractional Spin Particles,” Phys. Rev. Lett. 49
(1982) 957–959. (Cited on page 89.)
[112] J. de Azcárraga and J. Izquierdo, Lie Groups, Lie Algebras, Cohomology and Some
Applications in Physics. Cambridge Monographs on Mathematical Physics. Cambridge
University Press, 1998. (Cited on page 90.)
[113] S. K. Bose, “The galilean group in 2+1 space-times and its central extension,” Com-
munications in Mathematical Physics 169 no. 2, 385–395. (Cited on pages 90 and 91.)
[114] E. Inönü and E. P. Wigner, “Representations of the Galilei group,” Il Nuovo Cimento
(1943-1954) 9 (2007), no. 8, 705–718. (Cited on page 90.)
[115] E. Inönü and E. P. Wigner, “On the contraction of groups and their representations,”
Proc. Nat. Acad. Sci. 39 (1953) 510–524. (Cited on pages 92 and 289.)
[116] N. Woodhouse, Geometric Quantization. Oxford mathematical monographs. Clarendon
Press, 1997. (Cited on pages 97, 111, and 117.)
[117] A. A. Kirillov, Lectures on the Orbit Method. Graduate studies in mathematics. Amer-
ican Mathematical Society, 2004. (Cited on pages 97, 111, and 117.)
[118] S. De Buyl, S. Detournay, and Y. Voglaire, “Symplectic Geometry and Ge-
ometric Quantization,”. Available at http://www.ulb.ac.be/sciences/ptm/pmif/
Rencontres/geosymplgeoquant4.pdf. (Cited on pages 97, 111, and 117.)
[119] V. I. Arnold, Mathematical Methods of Classical Mechanics. Graduate Texts in Math-
ematics. Springer New York, 1997. (Cited on page 107.)
[120] B. Kostant, Lectures in Modern Analysis and Applications III, ch. Quantization and
unitary representations, pp. 87–208. Springer Berlin Heidelberg, Berlin, Heidelberg,
1970. (Cited on page 117.)
[121] E. Witten, “Coadjoint Orbits of the Virasoro Group,” Commun. Math. Phys. 114
(1988) 1. (Cited on pages 179, 242, and 245.)
[122] A. Alekseev, L. Faddeev, and S. Shatashvili, “Quantization of symplectic orbits of
compact Lie groups by means of the functional integral,” Journal of Geometry and
Physics 5 (1988), no. 3, 391 – 406. (Cited on pages 118 and 124.)
[123] A. Alekseev and S. Shatashvili, “Path integral quantization of the coadjoint orbits of
the Virasoro group and 2-d gravity,” Nuclear Physics B 323 (1989), no. 3, 719–733.
(Cited on page 118.)
[124] H. Aratyn, E. Nissimov, S. Pacheva, and A. H. Zimerman, “Symplectic actions on
coadjoint orbits,” Phys. Lett. B240 (1990) 127. (Cited on pages 118, 120, and 124.)
Bibliography 392
[144] C. Teitelboim, “Quantum Mechanics of the Gravitational Field,” Phys. Rev. D25
(1982) 3159. (Cited on pages 138 and 140.)
[145] M. Henneaux and C. Teitelboim, “Relativistic Quantum Mechanics of Supersymmetric
Particles,” Annals Phys. 143 (1982) 127. (Cited on pages 138 and 140.)
[146] M. Henneaux and C. Teitelboim, “First and second quantized point particles of any
spin,” in 2nd Meeting on Quantum Mechanics of Fundamental Systems (CECS) San-
tiago, Chile, December 17-20, 1987. 1987. (Cited on pages 138 and 140.)
[147] A. Yu. Alekseev and S. L. Shatashvili, “Propagator for the Relativistic Spinning Parti-
cle via Functional Integral Over Trajectories,” Mod. Phys. Lett. A3 (1988) 1551–1559.
(Cited on pages 138 and 140.)
[148] R. P. Feynman, “An Operator Calculus Having Applications in Quantum Electrody-
namics,” Phys. Rev. 84 (Oct, 1951) 108–128. (Cited on page 140.)
[149] M. J. Strassler, “Field theory without Feynman diagrams: One loop effective actions,”
Nucl. Phys. B385 (1992) 145–184, hep-ph/9205205. (Cited on page 140.)
[150] M. J. Strassler, The Bern-Kosower rules and their relation to quantum field theory.
PhD thesis, Stanford U., Phys. Dept., 1993. (Cited on page 140.)
[151] M. Henneaux and C. Teitelboim, Quantization of gauge systems. Princeton University
Press, 1992. (Cited on page 141.)
[152] P. H. Ginsparg, “Applied conformal field theory,” in Les Houches Summer School in
Theoretical Physics: Fields, Strings, Critical Phenomena Les Houches, France, June
28-August 5, 1988. 1988. hep-th/9108028. (Cited on pages 145 and 174.)
[153] P. Di Francesco, P. Mathieu, and D. Senechal, Conformal Field Theory. Graduate Texts
in Contemporary Physics. Springer-Verlag, New York, 1997. (Cited on pages 145, 174,
and 247.)
[154] R. Blumenhagen and E. Plauschinn, Introduction to Conformal Field Theory: With
Applications to String Theory. Lecture Notes in Physics. Springer Berlin Heidelberg,
2009. (Cited on pages 145 and 174.)
[155] A. Kriegl and P. W. Michor, The convenient setting of global analysis. No. 53. American
Mathematical Soc., 1997. (Cited on page 146.)
[156] J. Lurie, “Topics in Geometric topology.” MIT, 2009. Available at http://www.math.
harvard.edu/~lurie/937.html. (Cited on page 149.)
[157] G. Barnich, G. Giribet, and M. Leston, “Chern-Simons action for inhomogeneous Vi-
rasoro group as extension of three dimensional flat gravity,” J. Math. Phys. 56 (2015),
no. 7, 071701, 1505.02031. (Cited on page 159.)
[158] R. Bott, “On the characteristic classes of groups of diffeomorphisms,” Enseign. Math
23 (1977), no. 3-4, 209–220. (Cited on page 161.)
[159] L. Goncharova, “Cohomology of Lie algebras of formal vector fields on the line,” Uspekhi
Matematicheskikh Nauk 27 (1972), no. 5, 231–232. (Cited on page 163.)
[160] D. Fuks, “Cohomology of infinite-dimensional Lie algebras,” Sov. Math., Consultants
Bureau, New York (1986). (Cited on page 163.)
[161] V. F. Lazutkin and T. F. Pankratova, “Normal forms and versal deformations for Hill’s
equation,” Functional Analysis and Its Applications 9 (October, 1975) 306–311. (Cited
on page 179.)
Bibliography 394
[162] A. A. Kirillov, “Orbits of the group of diffeomorphisms of a circle and local Lie super-
algebras,” Functional Analysis and Its Applications 15 (April, 1981) 135–137. (Cited
on page 179.)
[163] G. Segal, “Unitary representations of some infinite-dimensional groups,” Comm. Math.
Phys. 80 (1981), no. 3, 301–342. (Cited on pages 179 and 246.)
[164] I. Bakas, “Conformal invariance, the KdV equation and coadjoint orbits of the Virasoro
algebra,” Nuclear Physics B 302 (1988), no. 2, 189 – 203. (Cited on page 179.)
[165] F. Gay-Balmaz, “On the classification of the coadjoint orbits of the Sobolev Bott-
Virasoro group,” Journal of Functional Analysis 256 (2009), no. 9, 2815 – 2841. (Cited
on page 179.)
[166] J. Balog, L. Fehér, and L. Palla, “Coadjoint orbits of the Virasoro algebra and the
global Liouville equation,” Int. J. Mod. Phys. A13 (1998) 315–362, hep-th/9703045.
(Cited on pages 179, 201, 204, 206, 207, 208, 213, and 214.)
[167] G. Barnich, L. Donnay, J. Matulich, and R. Troncoso, “Asymptotic symmetries and
dynamics of three-dimensional flat supergravity,” JHEP 08 (2014) 071, 1407.4275.
(Cited on pages 189, 284, 361, 364, 365, and 366.)
[168] S. Nag and A. Verjovsky, “Diff(S 1 ) and the Teichmüller spaces,” Comm. Math. Phys.
130 (1990), no. 1, 123–138. (Cited on page 196.)
[169] R. Schwartz, “A projectively natural flow for circle diffeomorphisms,” Inventiones math-
ematicae 110 (December, 1992) 627–647. (Cited on page 208.)
[170] R. Schwartz, “On the Integral Curve of a Linear Third Order O.D.E.,” Journal of
Differential Equations 135 (1997), no. 2, 183 – 191. (Cited on page 208.)
[171] S. Tabachnikov, “Variations on R. Schwartz’s inequality for the Schwarzian derivative,”
ArXiv e-prints (June, 2010) 1006.1339. (Cited on page 208.)
[172] R. P. Geroch, “Structure of the gravitational field at spatial infinity,” J. Math. Phys.
13 (1972) 956–968. (Cited on page 217.)
[173] R. Geroch, Asymptotic Structure of Space-Time, ch. Asymptotic Structure of Space-
Time, pp. 1–105. Springer US, Boston, MA, 1977. (Cited on pages 217 and 311.)
[174] A. Ashtekar, “Asymptotic Structure of the Gravitational Field at Spatial Infinity,” in
General Relativity and Gravitation II, A. Held, ed., vol. 2, p. 37. 1980. (Cited on
page 217.)
[175] L. Abbott and S. Deser, “Charge definition in non-abelian gauge theories,” Physics
Letters B 116 (1982), no. 4, 259 – 263. (Cited on page 217.)
[176] G. Barnich and F. Brandt, “Covariant theory of asymptotic symmetries, conservation
laws and central charges,” Nucl. Phys. B633 (2002) 3–82, hep-th/0111246. (Cited
on pages 217, 221, 223, 224, 232, and 264.)
[177] G. Barnich, “Boundary charges in gauge theories: Using Stokes theorem in the bulk,”
Class. Quant. Grav. 20 (2003) 3685–3698, hep-th/0301039. (Cited on page 217.)
[178] G. Barnich and G. Compère, “Surface charge algebra in gauge theories and thermody-
namic integrability,” J. Math. Phys. 49 (2008) 042901, 0708.2378. (Cited on pages 217
and 223.)
[179] M. Bañados, “Three-dimensional quantum geometry and black holes,”
hep-th/9901148. [AIP Conf. Proc. 484, 147 (1999)]. (Cited on pages 217,
233, and 236.)
Bibliography 395
[180] K. Skenderis and S. N. Solodukhin, “Quantum effective action from the AdS / CFT cor-
respondence,” Phys. Lett. B472 (2000) 316–322, hep-th/9910023. (Cited on pages 217
and 233.)
[181] A. Garbarz and M. Leston, “Classification of Boundary Gravitons in AdS3 Gravity,”
JHEP 05 (2014) 141, 1403.3367. (Cited on pages 217, 240, and 241.)
[182] T. Nakatsu, H. Umetsu, and N. Yokoi, “Three-dimensional black holes and Liouville
field theory,” Prog. Theor. Phys. 102 (1999) 867–896, hep-th/9903259. (Cited on
pages 217 and 241.)
[183] J. Navarro-Salas and P. Navarro, “Virasoro orbits, AdS(3) quantum gravity and en-
tropy,” JHEP 05 (1999) 009, hep-th/9903248. (Cited on pages 217 and 241.)
[184] A. Maloney and E. Witten, “Quantum Gravity Partition Functions in Three Dimen-
sions,” JHEP 02 (2010) 029, 0712.0155. (Cited on pages 217, 241, and 252.)
[185] C. Teitelboim, “How commutators of constraints reflect the space-time structure,” An-
nals Phys. 79 (1973) 542–557. (Cited on page 219.)
[186] Virgo, LIGO Scientific Collaboration, B. P. Abbott et al., “Observation of Gravi-
tational Waves from a Binary Black Hole Merger,” Phys. Rev. Lett. 116 (2016), no. 6,
061102, 1602.03837. (Cited on pages 219 and 384.)
[187] S. Deser, R. Jackiw, and S. Templeton, “Topologically Massive Gauge Theories,” An-
nals Phys. 140 (1982) 372–411. [Annals Phys.281,409(2000)]. (Cited on page 219.)
[188] E. A. Bergshoeff, O. Hohm, and P. K. Townsend, “Massive Gravity in Three Dimen-
sions,” Phys. Rev. Lett. 102 (2009) 201301, 0901.1766. (Cited on page 219.)
[189] R. Penrose, Group Theory in Non-Linear Problems: Lectures Presented at the NATO
Advanced Study Institute on Mathematical Physics, held in Istanbul, Turkey, August
7–18, 1972, ch. Relativistic Symmetry Groups, pp. 1–58. Springer Netherlands, Dor-
drecht, 1974. (Cited on page 221.)
[190] J. W. York, “Role of Conformal Three-Geometry in the Dynamics of Gravitation,”
Phys. Rev. Lett. 28 (Apr, 1972) 1082–1085. (Cited on page 221.)
[191] G. W. Gibbons and S. W. Hawking, “Action integrals and partition functions in quan-
tum gravity,” Phys. Rev. D 15 (May, 1977) 2752–2756. (Cited on pages 221 and 232.)
[192] A. Achucarro and P. K. Townsend, “A Chern-Simons Action for Three-Dimensional
anti-De Sitter Supergravity Theories,” Phys. Lett. B180 (1986) 89. (Cited on
page 221.)
[193] E. Witten, “(2+1)-Dimensional Gravity as an Exactly Soluble System,” Nucl. Phys.
B311 (1988) 46. (Cited on page 221.)
[194] M. Blagojevic, Gravitation and Gauge Symmetries. Series in High Energy Physics,
Cosmology and Gravitation. CRC Press, 2001. (Cited on page 221.)
[195] G. Lucena Gómez, “Higher-Spin Theories - Part II : enter dimension three,” PoS Mo-
daveVIII (2012) 003, 1307.3200. (Cited on page 221.)
[196] L. Donnay, “Asymptotic dynamics of three-dimensional gravity,” PoS Modave2015
(2016) 001, 1602.09021. (Cited on page 221.)
[197] O. Coussaert, M. Henneaux, and P. van Driel, “The Asymptotic dynamics of three-
dimensional Einstein gravity with a negative cosmological constant,” Class. Quant.
Grav. 12 (1995) 2961–2966, gr-qc/9506019. (Cited on page 221.)
Bibliography 396
[217] J. Kim and M. Porrati, “On a Canonical Quantization of 3D Anti de Sitter Pure
Gravity,” JHEP 10 (2015) 096, 1508.03638. (Cited on page 239.)
[218] C. Troessaert, “Poisson Structure of the Boundary Gravitons in 3D Gravity with Neg-
ative Λ,” Class. Quant. Grav. 32 (2015), no. 23, 235019, 1507.01580. (Cited on
page 240.)
[219] G. Compère, P.-J. Mao, A. Seraj, and M. M. Sheikh-Jabbari, “Symplectic and Killing
symmetries of AdS3 gravity: holographic vs boundary gravitons,” JHEP 01 (2016)
080, 1511.06079. (Cited on page 240.)
[220] E. J. Martinec, “Conformal field theory, geometry, and entropy,” hep-th/9809021.
(Cited on page 241.)
[221] E. Witten, “Three-Dimensional Gravity Revisited,” 0706.3359. (Cited on page 241.)
[222] M. M. Sheikh-Jabbari and H. Yavartanoo, “On 3d Bulk Geometry of Virasoro Coad-
joint Orbits: Orbit invariant charges and Virasoro hair on locally AdS3 geometries,”
1603.05272. (Cited on page 241.)
[223] A. Seraj, Conserved charges, surface degrees of freedom, and black hole entropy. PhD
thesis, 2016. 1603.02442. (Cited on page 241.)
[224] R. Schon and S.-T. Yau, “Proof of the positive mass theorem. 2.,” Commun. Math.
Phys. 79 (1981) 231–260. (Cited on page 242.)
[225] E. Witten, “A Simple Proof of the Positive Energy Theorem,” Commun. Math. Phys.
80 (1981) 381. (Cited on page 242.)
[226] R. M. Wald, General Relativity. 1984. (Cited on page 242.)
[227] D. R. Brill and S. Deser, “Variational methods and positive energy in general relativity,”
Annals Phys. 50 (1968) 548–570. (Cited on page 242.)
[228] B. Feigin and D. Fuks, “Casimir operators in modules over virasoro algebra,” Dokl.
Akad. Nauk SSSR 269 (1983), no. 5, 1057–1060. (Cited on pages 245 and 319.)
[229] H.-S. La, P. C. Nelson, and A. S. Schwarz, “Remarks on Virasoro model space,” Conf.
Proc. C9003122 (1990) 259–265. (Cited on page 245.)
[230] H. Airault and P. Malliavin, “Unitarizing probability measures for representations of
Virasoro algebra,” Journal de Mathématiques Pures et Appliquées 80 (2001), no. 6,
627 – 667. (Cited on pages 245 and 301.)
[231] H. Airault, P. Malliavin, and A. Thalmaier, “Support of Virasoro unitarizing measures,”
Comptes Rendus Mathematique 335 (2002), no. 7, 621–626. (Cited on pages 245
and 301.)
[232] B. Feigin and D. Fuks, “Verma modules over the Virasoro algebra,” Funkts. Anal.
Prilozh. 17 (1983), no. 3, 91–92. (Cited on pages 246 and 328.)
[233] B. Feigin and D. Fuchs, “Verma modules over the Virasoro algebra,” in Topology, L. D.
Faddeev and A. A. Malcev, eds., vol. 1060 of Lecture Notes in Mathematics, pp. 230–
245. Springer Berlin Heidelberg, 1984. (Cited on page 246.)
[234] B. Feigin and D. Fuchs, Representations of the Virasoro algebra. Reports: Matematiska
Institutionen. Department, Univ., 1986. (Cited on page 246.)
[235] V. G. Kac, Lie Algebras and Related Topics: Proceedings of a Conference Held at New
Brunswick, New Jersey, May 29–31, 1981, ch. Some problems on infinite dimensional
Lie algebras and their representations, pp. 117–126. Springer Berlin Heidelberg, Berlin,
Heidelberg, 1982. (Cited on page 246.)
Bibliography 398
[236] P. Goddard, A. Kent, and D. Olive, “Unitary representations of the Virasoro and
super-Virasoro algebras,” Comm. Math. Phys. 103 (1986), no. 1, 105–119. (Cited on
page 246.)
[237] H. Salmasian and K.-H. Neeb, “Classification of positive energy representations of the
Virasoro group,” 1402.6572. (Cited on page 248.)
[238] T. Apostol, Introduction to Analytic Number Theory. Undergraduate Texts in Mathe-
matics. Springer New York, 1998. (Cited on page 249.)
[239] A. Wassermann, “Direct proofs of the Feigin-Fuchs character formula for unitary rep-
resentations of the Virasoro algebra,” ArXiv e-prints (Dec., 2010) 1012.6003. (Cited
on pages 250 and 328.)
[240] A. Wassermann, “Kac-Moody and Virasoro algebras,” ArXiv e-prints (Apr., 2010)
1004.1287. (Cited on pages 250 and 328.)
[241] S. Giombi, A. Maloney, and X. Yin, “One-loop Partition Functions of 3D Gravity,”
JHEP 08 (2008) 007, 0804.1773. (Cited on pages 252, 331, 332, 340, 341, and 346.)
[242] C. A. Keller and A. Maloney, “Poincare Series, 3D Gravity and CFT Spectroscopy,”
JHEP 02 (2015) 080, 1407.6008. (Cited on page 252.)
[243] I. J. R. Aitchison and A. J. G. Hey, Gauge Theories in Particle Physics. Graduate
Student Series in Physics. Taylor & Francis, 2004. (Cited on page 253.)
[244] S. Deser, R. Jackiw, and G. ’t Hooft, “Three-Dimensional Einstein Gravity: Dynamics
of Flat Space,” Annals Phys. 152 (1984) 220. (Cited on page 261.)
[245] S. Deser and R. Jackiw, “String Sources in (2+1)-dimensional Gravity,” Annals Phys.
192 (1989) 352. (Cited on page 261.)
[246] D. Christodoulou and S. Klainerman, “The global nonlinear stability of the Minkowski
space,” Séminaire Équations aux dérivées partielles (Polytechnique) (1993) 1–29.
(Cited on page 265.)
[247] A. Strominger, “Asymptotic Symmetries of Yang-Mills Theory,” JHEP 07 (2014) 151,
1308.0589. (Cited on pages 265, 307, and 312.)
[248] T. He, P. Mitra, A. P. Porfyriadis, and A. Strominger, “New Symmetries of Massless
QED,” JHEP 10 (2014) 112, 1407.3789. (Cited on pages 265 and 312.)
[249] V. Lysov, S. Pasterski, and A. Strominger, “Low’s Subleading Soft Theorem as a
Symmetry of QED,” Phys. Rev. Lett. 113 (2014), no. 11, 111601, 1407.3814. (Cited
on pages 265 and 312.)
[250] A. Strominger and A. Zhiboedov, “Gravitational Memory, BMS Supertranslations and
Soft Theorems,” JHEP 01 (2016) 086, 1411.5745. (Cited on pages 265 and 312.)
[251] D. Kapec, V. Lysov, and A. Strominger, “Asymptotic Symmetries of Massless QED in
Even Dimensions,” 1412.2763. (Cited on pages 265 and 312.)
[252] S. Pasterski, A. Strominger, and A. Zhiboedov, “New Gravitational Memories,”
1502.06120. (Cited on pages 265 and 312.)
[253] D. Kapec, V. Lysov, S. Pasterski, and A. Strominger, “Higher-Dimensional Supertrans-
lations and Weinberg’s Soft Graviton Theorem,” 1502.07644. (Cited on pages 265
and 312.)
[254] S. Pasterski, “Asymptotic Symmetries and Electromagnetic Memory,” 1505.00716.
(Cited on pages 265 and 312.)
[255] T. He, P. Mitra, and A. Strominger, “2D Kac-Moody Symmetry of 4D Yang-Mills
Theory,” 1503.02663. (Cited on pages 265 and 312.)
Bibliography 399
[273] A. D. Helfer, “A phase space for gravitational radiation,” Comm. Math. Phys. 170
(1995), no. 3, 483–502. (Cited on page 280.)
[274] E. E. Flanagan and D. A. Nichols, “Conserved charges of the extended Bondi-Metzner-
Sachs algebra,” 1510.03386. (Cited on page 280.)
[275] J. Garecki, “Canonical angular supermomentum tensors in general relativity,” Journal
of Mathematical Physics 40 (1999), no. 8, 4035–4055. (Cited on page 280.)
[276] J. Garecki, “Superenergy and angular supermomentum tensors in general relativity,”
Reports on Mathematical Physics 44 (1999), no. 1, 95–100. (Cited on page 280.)
[277] G. Barnich, A. Gomberoff, and H. A. González, “A 2D field theory equivalent to 3D
gravity with no cosmological constant,” Springer Proc. Math. Stat. 60 (2014) 135–138,
1303.3568. (Cited on page 284.)
[278] G. Barnich, A. Gomberoff, and H. A. González, “Three-dimensional Bondi-Metzner-
Sachs invariant two-dimensional field theories as the flat limit of Liouville theory,”
Phys. Rev. D87 (2013), no. 12, 124032, 1210.0731. (Cited on pages 284 and 312.)
[279] H. A. González and M. Pino, “Boundary dynamics of asymptotically flat 3D gravity
coupled to higher spin fields,” JHEP 05 (2014) 127, 1403.4898. (Cited on page 284.)
[280] G. Barnich, L. Donnay, J. Matulich, and R. Troncoso, “Super-BMS3 invariant boundary
theory from three-dimensional flat supergravity,” 1510.08824. (Cited on pages 284,
361, and 364.)
[281] G. Barnich, A. Gomberoff, and H. A. González, “The Flat limit of three dimensional
asymptotically anti-de Sitter spacetimes,” Phys. Rev. D86 (2012) 024020, 1204.3288.
(Cited on pages 286 and 312.)
[282] L. Susskind, “Holography in the flat space limit,” hep-th/9901079. [AIP Conf. Proc.
493, 98 (1999)]. (Cited on page 290.)
[283] J. Polchinski, “S matrices from AdS space-time,” hep-th/9901076. (Cited on
page 290.)
[284] M. Gary, S. B. Giddings, and J. Penedones, “Local bulk S-matrix elements and CFT
singularities,” Phys. Rev. D80 (2009) 085005, 0903.4437. (Cited on page 290.)
[285] R. N. Caldeira Costa, “Aspects of the zero Λ limit in the AdS/CFT correspondence,”
Phys. Rev. D90 (2014), no. 10, 104018, 1311.7339. (Cited on page 290.)
[286] C. Krishnan, A. Raju, and S. Roy, “A Grassmann path from AdS3 to flat space,” JHEP
03 (2014) 036, 1312.2941. (Cited on pages 290 and 312.)
[287] A. Bagchi, R. Gopakumar, I. Mandal, and A. Miwa, “GCA in 2d,” JHEP 08 (2010)
004, 0912.1090. (Cited on pages 291, 317, 321, 322, and 357.)
[288] A. Bagchi, “The BMS/GCA correspondence,” 1006.3354. (Cited on pages 291
and 312.)
[289] A. Bagchi and I. Mandal, “Supersymmetric Extension of Galilean Conformal Algebras,”
Phys. Rev. D80 (2009) 086011, 0905.0580. (Cited on page 291.)
[290] A. Bagchi and I. Mandal, “On Representations and Correlation Functions of Galilean
Conformal Algebras,” Phys. Lett. B675 (2009) 393–397, 0903.4524. (Cited on
page 291.)
[291] A. Bagchi and R. Gopakumar, “Galilean Conformal Algebras and AdS/CFT,” JHEP
07 (2009) 037, 0902.1385. (Cited on page 291.)
Bibliography 401
http://rspa.royalsocietypublishing.org/content/335/1602/301.full.pdf.
(Cited on page 311.)
[310] M. Crampin and P. J. McCarthy, “Representations of the Bondi-Metzner-Sachs Group.
IV. Cantoni Representations are Induced,” Proceedings of the Royal Society of Lon-
don A: Mathematical, Physical and Engineering Sciences 351 (1976), no. 1664, 55–
70, http://rspa.royalsocietypublishing.org/content/351/1664/55.full.pdf.
(Cited on page 311.)
[311] V. Cantoni, “A Class of Representations of the Generalized Bondi-Metzner Group,”
Journal of Mathematical Physics 7 (1966), no. 8, 1361–1364. (Cited on page 311.)
[312] V. Cantoni, “Reduction of Some Representations of the Generalized Bondi-Metzner
Group,” Journal of Mathematical Physics 8 (1967), no. 8, 1700–1706. (Cited on
page 311.)
[313] L. Girardello and G. Parravicini, “Continuous spins in the Bondi-Metzner-Sachs group
of asymptotic symmetry in general relativity,” Phys. Rev. Lett. 32 (1974) 565–568.
(Cited on page 311.)
[314] M. Crampin, “Physical significance of the topology of the Bondi-Metzner-Sachs group,”
Phys. Rev. Lett. 33 (1974) 547–550. (Cited on page 311.)
[315] A. Piard, “Unitary representations of semi-direct product groups with infinite dimen-
sional Abelian normal subgroup,” Reports on Mathematical Physics 11 (1977), no. 2,
259 – 278. (Cited on page 311.)
[316] A. Piard, “Representations of the Bondi-Metzner-Sachs Group with the Hilbert Topol-
ogy,” Rept. Math. Phys. 11 (1977) 279–283. (Cited on page 311.)
[317] U. Cattaneo, “Borel multipliers for the Bondi-Metzner-Sachs group,” J. Math. Phys.
20 (1979) 2257–2263. (Cited on page 311.)
[318] M. A. Awada, G. W. Gibbons, and W. T. Shaw, “Conformal supergravity, Twistors
and the Super BMS group,” Annals Phys. 171 (1986) 52. (Cited on page 311.)
[319] R. P. Geroch, “Null infinity is not a good initial data surface,” J. Math. Phys. 19 (1978)
1300–1303. (Cited on page 311.)
[320] A. Ashtekar and R. O. Hansen, “A unified treatment of null and spatial infinity in gen-
eral relativity. I - Universal structure, asymptotic symmetries, and conserved quantities
at spatial infinity,” J. Math. Phys. 19 (1978) 1542–1566. (Cited on page 311.)
[321] A. Ashtekar, “Asymptotic Quantization of the Gravitational Field,” Phys. Rev. Lett.
46 (1981) 573–576. (Cited on page 311.)
[322] A. Ashtekar, Asymptotic quantization: Based on 1984 Naples lectures. 1987. (Cited
on page 311.)
[323] S. S. Gubser, I. R. Klebanov, and A. M. Polyakov, “Gauge theory correlators from
noncritical string theory,” Phys. Lett. B428 (1998) 105–114, hep-th/9802109. (Cited
on page 312.)
[324] R. Bousso, “Holography in general space-times,” JHEP 06 (1999) 028,
hep-th/9906022. (Cited on page 312.)
[325] R. Bousso, “The Holographic principle for general backgrounds,” Class. Quant. Grav.
17 (2000) 997–1005, hep-th/9911002. (Cited on page 312.)
[326] G. Arcioni and C. Dappiaggi, “Exploring the holographic principle in asymptot-
ically flat space-times via the BMS group,” Nucl. Phys. B674 (2003) 553–592,
hep-th/0306142. (Cited on page 312.)
Bibliography 403
[367] C. Fronsdal, “Massless Fields with Integer Spin,” Phys. Rev. D18 (1978) 3624. (Cited
on page 334.)
[368] R. Rahman, “Higher Spin Theory - Part I,” PoS ModaveVIII (2012) 004, 1307.3199.
(Cited on page 334.)
[369] G. W. Gibbons, M. J. Perry, and C. N. Pope, “Partition functions, the Bekenstein
bound and temperature inversion in anti-de Sitter space and its conformal boundary,”
Phys. Rev. D74 (2006) 084009, hep-th/0606186. (Cited on pages 338 and 342.)
[370] R. Gopakumar, R. K. Gupta, and S. Lal, “The Heat Kernel on AdS,” JHEP 11 (2011)
010, 1103.3627. (Cited on pages 338 and 342.)
[371] M. Beccaria and A. A. Tseytlin, “On higher spin partition functions,” J. Phys. A48
(2015), no. 27, 275401, 1503.08143. (Cited on page 338.)
[372] A. Bagchi, S. Detournay, D. Grumiller, and J. Simon, “Cosmic Evolution from Phase
Transition of Three-Dimensional Flat Space,” Phys. Rev. Lett. 111 (2013), no. 18,
181301, 1305.2919. (Cited on page 340.)
[373] F. A. Dolan, “Character formulae and partition functions in higher dimensional con-
formal field theory,” J. Math. Phys. 47 (2006) 062303, hep-th/0508031. (Cited on
page 342.)
[374] J. Balog, L. Fehér, L. O’Raifeartaigh, P. Forgacs, and A. Wipf, “Toda Theory and
W Algebra From a Gauged WZNW Point of View,” Annals Phys. 203 (1990) 76–136.
(Cited on page 343.)
[375] B. A. Khesin and B. Z. Shapiro, “Nondegenerate curves on S**2 and orbit classification
of the Zamolodchikov algebra,” Commun. Math. Phys. 145 (1992) 357–362. (Cited
on pages 343 and 348.)
[376] Z. Bajnok and D. Nogradi, “Geometry of W algebras from the affine Lie algebra point
of view,” J. Phys. A34 (2001) 4811–4830, hep-th/0012190. (Cited on pages 343, 344,
and 348.)
[377] M. Henneaux and S.-J. Rey, “Nonlinear Winf inity as Asymptotic Symmetry of Three-
Dimensional Higher Spin Anti-de Sitter Gravity,” JHEP 12 (2010) 007, 1008.4579.
(Cited on pages 343 and 351.)
[378] A. Campoleoni, S. Fredenhagen, S. Pfenninger, and S. Theisen, “Asymptotic symme-
tries of three-dimensional gravity coupled to higher-spin fields,” JHEP 11 (2010) 007,
1008.4744. (Cited on pages 343, 344, and 351.)
[379] M. R. Gaberdiel and T. Hartman, “Symmetries of Holographic Minimal Models,” JHEP
05 (2011) 031, 1101.2910. (Cited on page 343.)
[380] A. Campoleoni, S. Fredenhagen, and S. Pfenninger, “Asymptotic W-symmetries in
three-dimensional higher-spin gauge theories,” JHEP 09 (2011) 113, 1107.0290.
(Cited on pages 343, 344, and 351.)
[381] H. Afshar, A. Bagchi, R. Fareghbal, D. Grumiller, and J. Rosseel, “Spin-3 Grav-
ity in Three-Dimensional Flat Space,” Phys. Rev. Lett. 111 (2013), no. 12, 121603,
1307.4768. (Cited on pages 346 and 353.)
[382] H. A. González, J. Matulich, M. Pino, and R. Troncoso, “Asymptotically flat space-
times in three-dimensional higher spin gravity,” JHEP 09 (2013) 016, 1307.5651.
(Cited on pages 346, 351, and 353.)
[383] A. Campoleoni, “Higher Spins in D = 2 + 1,” Subnucl. Ser. 49 (2013) 385–396,
1110.5841. (Cited on page 351.)
Bibliography 406
[384] L. P. S. Singh and C. R. Hagen, “Lagrangian formulation for arbitrary spin. 2. The
fermion case,” Phys. Rev. D9 (1974) 910–920. (Cited on page 358.)
[385] R. R. Metsaev, “Gauge invariant formulation of massive totally symmetric fermionic
fields in (A)dS space,” Phys. Lett. B643 (2006) 205–212, hep-th/0609029. (Cited on
page 358.)
[386] J. Fang and C. Fronsdal, “Massless Fields with Half Integral Spin,” Phys. Rev. D18
(1978) 3630. (Cited on page 358.)
[387] T. Creutzig, Y. Hikida, and P. B. Ronne, “Higher spin AdS3 supergravity and its dual
CFT,” JHEP 02 (2012) 109, 1111.2139. (Cited on pages 358 and 361.)
[388] O. Fuentealba, J. Matulich, and R. Troncoso, “Extension of the Poincaré group
with half-integer spin generators: hypergravity and beyond,” JHEP 09 (2015) 003,
1505.06173. (Cited on pages 361 and 368.)
[389] O. Fuentealba, J. Matulich, and R. Troncoso, “Asymptotically flat structure of hyper-
gravity in three spacetime dimensions,” JHEP 10 (2015) 009, 1508.04663. (Cited on
pages 361 and 368.)
[390] I. Mandal, “Supersymmetric Extension of GCA in 2d,” JHEP 11 (2010) 018,
1003.0209. (Cited on page 361.)
[391] P. Deligne and I. Study, Quantum Fields and Strings: A Course for Mathematicians,
vol. 2 of Quantum Fields and Strings: A Course for Mathematicians. American Math-
ematical Soc., 1999. (Cited on page 362.)
[392] C. Carmeli, G. Cassinelli, A. Toigo, and V. S. Varadarajan, “Unitary representations of
super Lie groups and applications to the classification and multiplet structure of super
particles,” Commun. Math. Phys. 263 (2006) 217–258, hep-th/0501061. [Erratum:
Commun. Math. Phys. 307, 565 (2011)]. (Cited on page 362.)
[393] C. Carmeli, Super Lie Groups: Structure and Representations. PhD thesis, 2006. Avail-
able on http://www.ge.infn.it/~carmeli/publication/publication.html. (Cited
on page 362.)
[394] M. Bañados, K. Bautier, O. Coussaert, M. Henneaux, and M. Ortiz, “Anti-de Sitter
/ CFT correspondence in three-dimensional supergravity,” Phys. Rev. D58 (1998)
085020, hep-th/9805165. (Cited on page 364.)
[395] M. Henneaux, L. Maoz, and A. Schwimmer, “Asymptotic dynamics and asymptotic
symmetries of three-dimensional extended AdS supergravity,” Annals Phys. 282 (2000)
31–66, hep-th/9910013. (Cited on page 364.)
[396] O. Coussaert and M. Henneaux, “Supersymmetry of the (2+1) black holes,” Phys. Rev.
Lett. 72 (1994) 183–186, hep-th/9310194. (Cited on pages 364 and 366.)
[397] I. MacDonald, Symmetric Functions and Hall Polynomials. Second edition in Oxford
Mathematical Monographs. Oxford Science Publications, 1995. (Cited on page 370.)
[398] W. Fulton and J. Harris, Representation Theory: A First Course. Graduate Texts in
Mathematics. Springer, New York, 1991. (Cited on pages 372, 373, 374, and 378.)
[399] D. E. Littlewood, The Theory of Group Characters and Matrix Representations of
Groups. American Mathematical Soc., 1950. (Cited on pages 378 and 379.)
[400] G. Barnich, A. Gomberoff, H. González, B. Oblak, and P. Salgado-Rebolledo, in prepa-
ration. (Cited on page 383.)
Index
407
Index 408