Sei sulla pagina 1di 229

Vacua and p-branes

in
Maximal Supergravities
Marcus Sebastian Bremer
Department of Physics
Imperial College of Science, Technology and Medicine
A thesis submitted in partial fullment of the requirements for the degree of Doctor of
Philosophy of the University of London
September 1999
Dedicated to the memory of Lonny E. Bayer
Abstract
In the rst part of this thesis, we present a careful treatment of the Mp-brane
sources that couple to various D = 11 supergravity backgrounds, including both the
extreme and the non-extreme Mp-brane background, and the AdS
p+2
S
9p
vacuum
backgrounds. As a preliminary, we derive the action that describes the consistent
coupling of the M5-brane to D = 11 supergravity, using an approach that dispenses
with the need to introduce a dual 6-form potential.
In the the second part of this thesis, we present an exhaustive classication of
the Freund-Rubin vacuum solutions of the maximal, massless supergravity theo-
ries in 2 D 11. These Freund-Rubin vacua are characterised by a spacetime
product manifold of several Einstein submanifolds, covariantly constant p-form eld
strengths and constant scalar elds. Vacuum solutions of a given spacetime decom-
position and a xed number of participating eld strengths are found to assemble
into multiplets of the Weyl subgroup of the global Cremmer-Julia symmetry group.
The residual supersymmetries of the Freund-Rubin vacua are determined by eval-
uating the integrability conditions of the Killing spinor equation. We argue that
each supersymmetric vacuum emerges in the near horizon limit of an aristocratic
p-brane solution.
iii
Declaration
The work leading to this thesis has been carried out in the Theoretical Physics
Group at Imperial College of Science, Technology and Medicine, London, between
October 1995 and September 1999.
The work presented in this thesis is original, unless indicated otherwise in the
text. Some of the work in this thesis has beneted from ideas developed in collab-
oration with H. L u, C.N. Pope and K.S. Stelle and reported in Dirac quantisation
conditions and Kaluza-Klein reduction, Nucl.Phys. B529 (1998) 259-294, and in
collaboration with M.J. Du, H. L u, C.N. Pope and K.S. Stelle and reported in In-
stanton cosmology and domain walls from M-theory and string theory, Nucl.Phys.
B543 (1999) 321-364. A short summary of the former article has also been pub-
lished under the same title in Fortschr.Phys. 47 (1999) 1-3, 117-123. However, no
part of this thesis is based directly on these publications. The initial concept for the
classication of vacua in the second part of this thesis has been developed in col-
laboration with R.E. Clark, who is also responsible for implementing the computer
algorithm used there and in some supersymmetry calculations.
iv
Acknowledgements
I am grateful to my supervisor K.S. Stelle for initially guiding me into my chosen area
of research and then continually providing me with encouragement, opportunities
and challenges throughout my PhD studies. I have enjoyed many interesting and
helpful conversations with him, and I owe a great deal of my understanding of the
issues presented in this thesis to him.
I am further indebted to my collaborators M.J. Du, H. L u, C.N. Pope and R.E.
Clark for many helpful discussions, and to J. Russo for useful comments on an earlier
version of what has become the rst part of this thesis. I also thank the members
of the Theoretical Physics group in Imperial College for providing a congenial and
relaxed atmosphere and the Theory Division in CERN and the High Energy Group
at SISSA for hospitality during various stages of my studies.
I would also like to thank my parents, Gabriele and Bernd Bremer, and my
brother Tobias Bremer for their love, support and encouragement over the years
and Glykeria Karra for her companionship.
I am grateful for the nancial support of the EEC in the form of a Marie-Curie
Fellowship (TMR contract ERBFMBI-CT97-2344), which I held at Imperial College
from September 1997 to August 1999.
v
Contents
I Mp-brane sources in D = 11 supergravity 4
1 Introduction 5
2 M5-brane coupling to D = 11 supergravity 9
2.1 M5-brane coupling from the Dirac 6-brane . . . . . . . . . . . . . . 9
2.2 Normalisation issues in M5-brane coupling . . . . . . . . . . . . . . 20
2.3 M5-brane solution of the coupled eld equations . . . . . . . . . . . 23
3 Spherically symmetric Mp-brane sources 29
3.1 Coupling Mp-brane sources to extreme Mp-branes . . . . . . . . . . 29
3.2 Probing the Mp-brane with photons and test branes . . . . . . . . . 38
3.3 Coupling Mp-brane sources to non-extreme Mp-branes . . . . . . . . 44
3.4 Patching D = 11 supergravity solutions with Mp-brane sources . . . 52
4 Conclusions 61
II Freund-Rubin vacua in maximal supergravities 63
5 Introduction 64
1
CONTENTS 2
5.1 Aristocratic p-branes and vacuum interpolation . . . . . . . . . . . . 64
5.2 An approach to the classication of vacuum solutions in D 11 . . 68
6 Vacua in D = 11 supergravity 71
6.1 Review of the Freund-Rubin vacua . . . . . . . . . . . . . . . . . . . 71
6.2 A generalised Freund-Rubin ansatz . . . . . . . . . . . . . . . . . . . 75
6.3 M
d
1
K
d
2
K
d
N
vacua . . . . . . . . . . . . . . . . . . . . . . 80
6.4 The generalised Freund-Rubin ansatz in D < 11 . . . . . . . . . . . . 86
7 Vacua in D = 10 IIA and IIB supergravity 90
7.1 IIA supergravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
7.2 IIB supergravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
8 Vacua in D 9 supergravities 107
8.1 CJ global symmetries of maximal, massless supergravities in D 9 . 107
8.2 D = 9 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
8.3 D = 8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
8.4 D = 7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
8.5 D = 6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
8.6 D = 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
8.7 D = 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
8.8 D = 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
9 Supersymmetric vacua in D 11 147
9.1 Integrability conditions for Killing spinors . . . . . . . . . . . . . . . 147
9.2 Supersymmetric vacua in D = 11 . . . . . . . . . . . . . . . . . . . . 155
CONTENTS 3
9.3 Supersymmetric vacua in D = 10 . . . . . . . . . . . . . . . . . . . . 162
9.4 Supersymmetric vacua in D 9 . . . . . . . . . . . . . . . . . . . . . 167
9.5 Identifying interpolating p-branes . . . . . . . . . . . . . . . . . . . . 173
10 Conclusions 177
A Notation and conventions 179
B Maximal massless supergravities in 3 D 11 181
B.1 Toroidal Kaluza-Klein reduction of D = 11 supergravity . . . . . . . 182
B.2 IIB supergravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198
C U(1) brations and Hopf reduction 205
C.1 Hopf reduction of U(1) brations . . . . . . . . . . . . . . . . . . . . 205
C.2 Supersymmetry of U(1) brations . . . . . . . . . . . . . . . . . . . 209
Bibliography 214
Part I
Mp-brane sources in D = 11
supergravity
4
Chapter 1
Introduction
In the last couple of years, evidence has been accumulated that points to the ex-
istence of an 11-dimensional quantum theory, called M-theory, which underlies the
known 10-dimensional string theories, see e.g. Ref. [1] for a review. The funda-
mental objects in M-theory are an electrically charged 2-brane and a magnetically
charged 5-brane. They will be collectively referred to as Mp-branes. The properties
of the Mp-branes can be studied from two complementary perspectives. From the
spacetime perspective, they emerge as supersymmetric p-brane solutions of D = 11
supergravity [4], which is the eective eld theory that approximates M-theory at
low energies. The supergravity solutions describing a single static Mp-brane have
been found some time ago [2, 3]. They are the basic building blocks from which
more complicated Mp-brane solutions, describing e.g. static intersections of several
Mp-branes, can be constructed, see e.g. Ref. [5] for a review.
Alternatively, the Mp-branes can be treated as dynamical objects in their own
rights. The degrees of freedom of the M2-brane are described by the 3-dimensional
N = 8 scalar supermultiplet [7], which contains eight scalars and eight spinors.
5
CHAPTER 1. INTRODUCTION 6
The corresponding reparameterisation invariant, supersymmetric action with local
-symmetry has been constructed some time ago [8]. The degrees of freedom of the
M5-brane are described by the 6-dimensional (2, 0) tensor supermultiplet [9], which
contains ve scalars, a chiral 2-form potential with self-dual 3-form potential and
two Weyl spinors of the same chirality. The self-duality of the 3-form eld strength
presents a serious problem for the construction of a manifestly reparameterisation
invariant action for the M5-brane, which has only been overcome recently, when
equivalent supersymmetric actions with local -symmetry have been proposed in
Refs. [10, 11].
In a non-trivial 11-dimensional background, the local -symmetry of the Mp-
brane action forces the background elds to satisfy the eld equations of D = 11
supergravity. This suggests that the Mp-branes can also be viewed as the matter
sources to which the D = 11 supergravity elds couple. The Lagrangian formulation
of the coupled M2-brane/supergravity theory presents no particular problem, since
the M2-brane is electrically charged and hence couples naturally to the usual 3-
form potential of D = 11 supergravity. In contrast, the Lagrangian formulation of
the coupled M5-brane/supergravity theory is less trivial. Since the M5-brane is
magnetically charged, it couples naturally to a dual 6-form potential, whose eld
strength is the dual of the eld strength of the 3-form potential. However, the
presence of the Chern-Simons term in D = 11 supergravity Lagrangian, which is
required for supersymmetry, makes it impossible to dualise the 3-form potential in
the standard way, i.e. D = 11 supergravity cannot be reformulated in terms of a
6-form potential only [6].
The observation at the end of the last paragraph leaves room for two distinct
CHAPTER 1. INTRODUCTION 7
approaches to M5-brane coupling to D = 11 supergravity. Accepting that the M5-
brane couples naturally to the dual 6-form potential, one may try to reformulate
standard D = 11 supergravity in a way that treats both the 3-form potential and
the 6-form potential as independent variables. This approach has been followed in
Refs. [13, 12]. Given the duality-symmetric formulation of D = 11 supergravity
with both a 3-form potential and a 6-form potential, coupling the M5-brane to
this theory does not present any particular problems. However, what does cause
problems is truncating the extra degrees of freedom represented by the dual 6-form
potential. As shown in Refs. [13, 12], although commented upon explicitly only in
Ref. [12], this truncation is possible only if in addition one introduces the higher-
dimensional analogue of the Dirac string, i.e. the Dirac 6-brane, which ends on the
M5-brane.
An alternative approach to M5-brane coupling is suggested by the observation
that whereas one can consistently truncate the 6-form potential in the duality-
symmetric approach to M5-brane coupling described above, one cannot eliminate
the Dirac 6-brane altogether. This approach, which we shall develop here, dispenses
with the 6-form potential and instead determines the correct interaction terms by
requiring that the location of the Dirac 6-brane in spacetime can be chosen arbitrar-
ily.
The remainder of the rst part of this thesis is organised as follows. Chapter
2 deals mainly with M5-brane coupling to D = 11 supergravity. We begin by
deriving an action describing the consistent coupling of a bosonic M5-brane source
to the standard elds of D = 11 supergravity. This is achieved by requiring the
gauge invariance of the action and, in addition, that the equations of motion of
CHAPTER 1. INTRODUCTION 8
the Dirac 6-brane are satised irrespective of its location in spacetime. After a
brief discussion of normalisation issues in M5-brane coupling, we show that the
coupled M5-brane/supergravity eld equations admit the M5-brane supergravity
background as a solution, as is required for consistency.
Chapter 3 consists of a more careful treatment of the Mp-brane source that cou-
ples to the Mp-brane supergravity background. We construct the spherically sym-
metric distribution of Mp-branes that couples to the extreme Mp-brane supergravity
background and show that the spacetime in the interior of the source may be taken to
be strictly at. A truly point-like source is obtained as the (invariant) zero-radius
limit of the spherically symmetric source distribution. We then continue to probe
the geometry of the Mp-brane spacetime that explicitly includes the source with
test branes and photons. The Mp-brane source that couples to the non-extreme, or
black, Mp-brane supergravity background is constructed in the subsequent section,
where we also discuss the generalisation of the ADM mass-renormalisation phe-
nomena to black Mp-branes. Finally, we briey discuss Mp-brane sources in other
supergravity backgrounds, in particular in the AdS
p+2
S
9p
vacua of D = 11
supergravity. We also show how Mp-brane sources can be used to patch together
distinct (source-free) solutions of D = 11 supergravity.
Chapter 2
M5-brane coupling to D = 11
supergravity
2.1 M5-brane coupling from the Dirac 6-brane
The bosonic sector of D = 11 supergravity [4] consists of the metric g
MN
and a
3-form gauge potential A
(3)
with 4-form eld strength F
(4)
= dA
(3)
. The action can
be written as
S
11
=
1
2
2
_
M
11
R
1
2
F
(4)
F
(4)

1
6
F
(4)
F
(4)
A
(3)
, (2.1.1)
where M
11
denotes the 11-dimensional spacetime and is the gravitational constant
in D = 11.
The equations of motion which follow from (2.1.1) admit electrically charged
(p = 2) and magnetically charged (p = 5) p-brane solutions that preserve one
half of the maximum number of rigid spacetime supersymmetries [2, 3]. These
supersymmetric p-branes of D = 11 supergravity are henceforth referred to as Mp-
9
CHAPTER 2. M5-BRANE COUPLING TO D = 11 SUPERGRAVITY 10
branes. For a general review of the properties of p-brane solutions in supergravity
theories, see e.g. Refs. [14, 15, 16].
The degrees of freedom of the M2-brane are described by the 3-dimensional
N = 8 scalar supermultiplet [7], which contains eight scalars and eight spinors. The
eight scalars are interpreted as the Goldstone bosons associated with the broken
translational symmetries in the directions normal to the M2-brane, while the eight
spinors are the Goldstone fermions associated with the broken supersymmetries. The
corresponding reparameterisation invariant, supersymmetric action with local -
symmetry has been constructed some time ago [8]. Here, we shall focus exclusively on
the bosonic sector of the M2-brane worldvolume theory. In the Minkowski vacuum
of D = 11 supergravity, the bosonic M2-brane is described by a Nambo-Goto action
for the embedding coordinates X
M
= X
M
(),
S
M2
= T
2
_
W
3
d
3

_
det g

, (2.1.2)
where

are coordinates on the M2-brane worldvolume W


3
and T
2
is the M2-brane
tension. The induced worldvolume metric is dened by g

X
M

X
N

MN
.
The degrees of freedom of the M5-brane are described by the 6-dimensional (2, 0)
tensor supermultiplet [9], which contains ve scalars, a chiral 2-form potential with
self-dual 3-form potential and two Weyl spinors of the same chirality. The scalars and
the spinors are again interpreted as the Goldstone modes associated with the broken
translational symmetries and supersymmetries respectively, while the chiral 2-form
potential is needed to complete the supermultiplet. The self-duality of the 3-form
eld strength presents a serious problem for the construction of a manifestly repa-
rameterisation invariant action for the M5-brane. A supersymmetric action with
CHAPTER 2. M5-BRANE COUPLING TO D = 11 SUPERGRAVITY 11
local -symmetry was constructed in Ref. [10] by forgoing manifest reparameterisa-
tion invariance. A manifestly reparameterisation invariant supersymmetric action
with local -symmetry was proposed in Ref. [11], based on methods developed ear-
lier in Ref. [17]. The supercovariant equations of motion for the worldvolume elds
have been obtained independently in Ref. [18]. These dierent formulations of the
M5-brane worldvolume theory have subsequently been shown to be equivalent [19].
Here, we restrict our attention to the bosonic sector of the M5-brane world-
volume theory and use the PST formulation [17], which employs a single auxiliary
scalar eld a() to achieve manifest reparameterisation invariance. The bosonic
M5-brane worldvolume elds are the embedding coordinates Y
M
= Y
M
() and the
chiral 2-form potential b
(2)
whose 3-form eld strength h
(3)
= db
(2)
satises a gener-
alised self-duality condition. The action for the bosonic M5-brane in the Minkowski
vacuum of D = 11 supergravity can be written as
S
M5
= T
5
_
W
6
d
6

_
det(g

+ (
v
(1)
h
(3)
)

)+
1
2
v
(1)
h
(3)
(v
(1)
h
(3)
), (2.1.3)
where

are coordinates on the M5-brane worldvolume W


6
and T
5
is the M5-
brane tension. The 1-form v
(1)
is related to the auxiliary scalar eld a() by v
(1)
=
da (g

a)

1
2
and
v
(1)
denotes the interior product
1
with v
(1)
. The Hodge dual
is dened with respect to the induced worldvolume metric g

Y
M

Y
N

MN
.
In the PST formulation, the self-duality condition for h
(3)
arises as a consequence
of the equation of motion for b
(2)
and a local bosonic symmetry of the action (2.1.3).
The auxiliary scalar eld does not represent a dynamical degree of freedom of the
worldvolume theory; its equation of motion is implied by that for b
(2)
. We refer to
1
For the denition of the interior product, and its relation to the wedge product and the Hodge
dual, see e.g. [20].
CHAPTER 2. M5-BRANE COUPLING TO D = 11 SUPERGRAVITY 12
the original work [17] for more details.
In a non-trivial D = 11 supergravity background, the Mp-branes do not only
couple to the spacetime metric g
MN
, via the induced metric g

as in (2.1.2) and
(2.1.3), but also to the 3-form potential A
(3)
. Because the M2-brane carries electric
charge, it couples minimally to A
(3)
and the corresponding action is [8]
S
11/M2
= T
2
_
W
3
A
(3)
. (2.1.4)
It is more dicult to see immediately how the M5-brane couples to A
(3)
as it is
magnetically charged and hence would couple naturally to a 6-form potential A
(6)
whose 7-form eld strength is the Hodge dual of F
(4)
. However, it is not possible to
reformulate D = 11 supergravity in terms of a 6-form potential only [6]. Further-
more, it has been realised [21, 22] that even upon introducing A
(6)
, the M5-brane
also couples to A
(3)
. These problems have been addressed in Refs. [12, 13] by refor-
mulating standard D = 11 supergravity to include both the 3-form potential A
(3)
and its dual 6-form potential A
(6)
as independent elds to which the M5-brane cou-
ples. As it happens, the dual potential can be consistently truncated, leaving the
coupling of the M5-brane to the standard D = 11 supergravity elds.
We want to show next how to arrive at the consistent coupling of the M5-brane to
D = 11 supergravity without having to introduce an independent 6-form potential
rst. We shall follow closely a method originally developed by Dirac [23] in the
context of magnetic monopole coupling to D = 4 electromagnetism. Subsequently,
this method has been extended in Ref. [24] to magnetically charged p-branes in
D dimensions. The existence of the Chern-Simons term in D = 11 supergravity
complicates the issue somewhat, but the basic idea remains the same. The M5-
CHAPTER 2. M5-BRANE COUPLING TO D = 11 SUPERGRAVITY 13
brane couples directly to the 4-form eld strength F
(4)
via the higher-dimensional
analogue of the Dirac-string, i.e. the Dirac 6-brane. Intuitively, the worldvolume
of the Dirac 6-brane represents the set of spacetime points on which the potential
A
(3)
is singular. Notice that there is an implicit gauge dependence in this description
of the Dirac 6-brane, because a gauge transformation A
(3)
=
(3)
with singular 3-
form parameter
(3)
can change the location of the Dirac 6-brane [25, 26]. In order to
avoid this gauge dependence, it must be possible to place the Dirac 6-brane anywhere
in spacetime. Dierent spacetime locations of the Dirac 6-brane then correspond to
dierent gauge choices. It is in this sense that the location of the Dirac 6-brane
represents a gauge freedom of the theory.
We continue to show that the coupling of the M5-brane to D = 11 supergravity
is uniquely determined by requiring (i) that the equations of motion of the Dirac
6-brane do not determine its spacetime location, and (ii) the invariance of the action
under regular gauge transformations.
Let


= (

, ) be coordinates on the worldvolume W


7
of the Dirac 6-brane.
The Dirac 6-brane ends on the M5-brane, which means that the boundary of its
worldvolume is the worldvolume of the M5-brane, i.e. W
7
= W
6
. We assume that
W
7
is dened by the equation = 0, so that we may use

as coordinates on W
6
.
The embedding coordinates Z
M
= Z
M
(

) of the Dirac 6-brane are constrained to


agree with the embedding coordinates Y
M
= Y
M
() of the M5-brane on W
6
, i.e.
Z
M
(, 0) = Y
M
(). We also require that the Dirac 6-brane worldvolume intersects
every closed 4-dimensional surface which entirely surrounds the M5-brane in at
least one point. One can understand this requirement by viewing the Dirac 6-
brane worldvolume alternatively as the set of spacetime points on which the 4-form
CHAPTER 2. M5-BRANE COUPLING TO D = 11 SUPERGRAVITY 14
eld strength F
(4)
fails to be exact. As every closed 4-dimensional surface which
entirely surrounds the M5-brane captures the total magnetic ux of the M5-brane,
F
(4)
= dA
(3)
cannot hold everywhere on this surface and hence every such surface
must have a non-zero intersection with the Dirac 6-brane worldvolume.
We dene the M5-brane and Dirac 6-brane p-form currents J
(6)
and G
(7)
by
J
M
1
M
6
(x) =
_
W
6

11
(x Y )
_
(det g
MN
)
dY
M
1
dY
M
6
, (2.1.5)
G
M
1
M
7
(x) =
_
W
7

11
(x Z)
_
(det g
MN
)
dZ
M
1
dZ
M
7
. (2.1.6)
Note that since W
7
= W
6
and Z
M
= Y
M
on W
6
, these currents are related by
d G
(7)
= J
(6)
. This shows that only the M5-brane current J
(6)
is conserved. It
follows from the denitions (2.1.6) and (2.1.13) that
_
W
6
w
(6)
=
_
M
11
J
(6)
w
(6)
,
_
W
7
w
(7)
=
_
M
11
G
(7)
w
(7)
, (2.1.7)
where w
(6)
and w
(7)
are arbitrary spacetime p-forms. We also observe that J
(6)
and
G
(7)
are normal to the M5-brane worldvolume and the Dirac 6-brane worldvolume
respectively, in the sense that the contraction of J
(6)
(G
(7)
) with any vectoreld
tangential to W
6
(W
7
) vanishes.
The rationale for introducing the Dirac 6-brane is that it permits us to couple
the M5-brane directly to the eld strength via

F
(4)
= dA
(3)
+ 2
2
T
5
G
(7)
. (2.1.8)
Note that A
(3)
is dened (up to regular gauge transformations A
(3)
= d
(2)
) for
a given Dirac 6-brane conguration. The potentials corresponding to two dierent
Dirac 6-brane congurations G
(7)
and

G
(7)
are related by the gauge transformation
dA
(3)
(G
(7)


G
(7)
), which requires a singular 3-form parameter
(3)
= A
(3)
CHAPTER 2. M5-BRANE COUPLING TO D = 11 SUPERGRAVITY 15
[25, 26]. It is in this sense that one can move the Dirac 6-brane by a singular
gauge transformation
2
. The variation of A
(3)
under a singular gauge transformation
is dened such that the modied eld strengths

F
(4)
remains gauge invariant.
The M5-brane current J
(6)
acts as a source term in the Bianchi identity for the
modied eld strength (2.1.8,
d

F
(4)
= 2
2
T
5
J
(6)
. (2.1.9)
To obtain the equations of motion for the potential we replace the kinetic term in
the action (2.1.1) according to

1
2
F
(4)
F
(4)

1
2


F
(4)


F
(4)
. (2.1.10)
However, we do not modify the Chern-Simons term in (2.1.1). The action resulting
from this replacement will be denoted by

S
11
in the following.
The simplest action which describes M5-brane/supergravity coupling is just

S
11
+ S
M5
, where S
M5
, the action for the M5-brane worldvolume elds, is given
by (2.1.3). The consistency check for this theory is whether we can arbitrarily
choose the spacetime location of the Dirac 6-brane. Varying the action

S
11
+ S
M5
with respect to the Dirac 6-brane coordinates Z
M
we nd that

[
1
Z
M
1


7
]
Z
M
7
(d

F
(4)
)
M
1
M
7
P
= 0 on W
7
. (2.1.11)
We use the equation of motion for A
(3)
to show that (2.1.11) is equivalent to

[
1
Z
M
1


7
]
Z
M
7
(F
(4)
F
(4)
)
M
1
M
7
P
= 0 on W
7
. (2.1.12)
2
Another way of seeing why moving the Dirac 6-brane requires a singular gauge transformation
is to decompose the action of this transformation in two parts, the rst which removes the Dirac
6-brane altogether and the second which re-introduces the Dirac 6-brane at a dierent location.
Each of these operations is clearly singular.
CHAPTER 2. M5-BRANE COUPLING TO D = 11 SUPERGRAVITY 16
It follows that in general we cannot chose the Dirac 6-brane location arbitrarily and
hence we conclude that the simplest action does not describe the consistent coupling
of the M5-brane to D = 11 supergravity.
We can improve on this situation by adding the term
T
5
2
_
W
7
A
(3)
F
(4)
(2.1.13)
to the action

S
11
+S
M5
. The equation of motion of the Dirac 6-brane equation then
requires that

[
1
Z
M
1


7
]
Z
M
7
_
d

F
(4)
+
1
2
F
(4)
F
(4)
_
M
1
M
7
P
= 0 on W
7
. (2.1.14)
Using the equation of motion for A
(3)
, one can show that (2.1.14) is equivalent to

[
1
Z
M
1


7
]
Z
M
7
(F
(4)
G
(7)
)
M
1
M
7
P
= 0 on W
7
, (2.1.15)
which is identically satised because the vector elds

Z
M

M
are tangential to the
Dirac 6-brane worldvolume W
7
, while G
(7)
is normal to W
7
. The Dirac 6-brane can
therefore be placed anywhere in spacetime.
Unfortunately, (2.1.13) is not gauge invariant because W
7
,= . Under the gauge
transformation A
(3)
= d
(2)
, (2.1.13) transforms into

T
5
2
_
W
6
d
(2)
A
(3)
. (2.1.16)
The only way to restore gauge invariance is to demand [27, 21] that the worldvolume
2-form b
(2)
transform under the above gauge transformation as b
(2)
=
(2)
and to
add
T
5
2
_
W
6
db
(2)
A
(3)
(2.1.17)
CHAPTER 2. M5-BRANE COUPLING TO D = 11 SUPERGRAVITY 17
to the action. However, because the worldvolume 3-form eld strength h
(3)
= db
(2)
now transforms non-trivially under A
(3)
= d
(2)
, the M5-brane action (2.1.3) is no
longer gauge invariant. In order to restore the gauge invariance of the M5-brane
action we need to redene h
(3)
in (2.1.3) to be the gauge invariant combination
h
(3)
= db
(2)
A
(3)
[27, 21, 13].
It remains to be checked that the Dirac 6-brane can be placed arbitrarily in
spacetime. Its equation of motion is again (2.1.14), which is now equivalent to

[
1
Z
M
1


7
]
Z
M
7
(h
(3)
J
(6)
+F
(4)
G
(7)
)
M
1
M
7
P
= 0 on W
7
(2.1.18)
since the equation of motion for A
(3)
has been altered by the additions to the action.
The second term in (2.1.18) vanishes identically, as explained above. To see that
the rst term in (2.1.18) is also zero, note that J
(6)
is non-zero only on W
6
= W
7
.
But on the boundary of the Dirac 6-brane worldvolume,

Z
M
vanishes. It follows
that the Dirac 6-brane equation of motion is satised identically and independently
of the choice of the Dirac 6-brane location in spacetime.
We have therefore demonstrated that the action that describes the consistent
coupling of the M5-brane to D = 11 supergravity is given by the sum of the actions
(2.1.1) (with F
(4)


F
(4)
in the kinetic term, but not in the FFA term), (2.1.3)
(with h
(3)
= db
(2)
A
(3)
) and
S
11/M5
=
T
5
2
_
W
6
db
(2)
A
(3)
+
T
5
2
_
W
7
A
(3)
F
(4)
. (2.1.19)
This action has been derived by requiring that the Dirac 6-brane is non-dynamical
and can be placed arbitrarily in spacetime, and that the action be invariant under
(regular) gauge transformations. It is in fact equivalent to the action found in Ref.
[12], after the truncation of the dual 6-form potential, except for the fact that we
CHAPTER 2. M5-BRANE COUPLING TO D = 11 SUPERGRAVITY 18
use the PST formulation to describe the dynamics of the chiral worldvolume 2-
form, whereas the self-duality of the corresponding worldvolume eld strength was
imposed by hand in Ref. [12]. In contrast, it appears as if the authors of Ref. [13]
in their work on M5-brane coupling to D = 11 supergravity adopt dierent, and
not necessarily compatible, normalisation conventions for the 3-form potential in the
pure supergravity action (2.1.1) and the interaction terms (2.1.19). We attempt to
clarify these normalisation issues in Section 2.2.
It is also possible to couple both the M2-brane and the M5-brane consistently to
D = 11 supergravity. In principle, the M2-brane could have a non-zero boundary on
the M5-brane worldvolume, which would act as a 1-brane source for the M5-brane
worldvolume 2-form b
(2)
[27, 28]. Here, we shall ignore this possibility and focus
on Mp-branes without boundaries. The modications to the action required for a
non-zero M2-brane boundary on the M5-brane have been discussed in Ref. [12]. In
the present case, the equations of motion for the Mp-brane/supergravity theory are
derived from
S =

S
11
+S
M2
+S
M5
+S
11/M2
+S
11/M5
, (2.1.20)
where S
M2
, S
11/M2
and S
11/M5
are given by equations (2.1.2), (2.1.4) and (2.1.19)
and

S
11
and S
M5
are the appropriately modied versions of (2.1.1) and (2.1.3).
The eld equations for the D = 11 metric and the 3-form potential in the presence
of an M2-brane source and an M5-brane source are
R
MN

1
2
g
MN
R =
1
2
_
(

F
(4)
)
2
MN

1
2
g
MN
(

F
(4)
)
2
_
, (2.1.21)

2
T
2
T
(M2)
MN

2
T
5
T
(M5)
MN
d

F
(4)
+
1
2
F
(4)
F
(4)
= 2
2
T
2
J
(3)
(2.1.22)
CHAPTER 2. M5-BRANE COUPLING TO D = 11 SUPERGRAVITY 19
2
2
T
5
(h
(3)
J
(6)
+F
(4)
G
(7)
) ,
d

F
(4)
= 2
2
T
5
J
(6)
, (2.1.23)
where expressions such as (

F
(4)
)
2
MN
and (

F
(4)
)
2
are understood to include the appro-
priate permutation factors.
The spacetime energy-momentum tensors T
(M2)
MN
and T
(M5)
MN
of the M2-brane and
the M5-brane are given by
T
MN
(M2)
(x) =
_
W
3
d
3
_
det g

X
M

X
N

11
(x X)

det g
MN
, (2.1.24)
T
MN
(M5)
(x) =
_
W
6
d
6
_
det g

Y
M

Y
N

11
(x Y )

det g
MN
, (2.1.25)
where the M5-brane worldvolume energy-momentum tensor T

is dened by
T

=
2
T
5
_
det g

S
M5
g

. (2.1.26)
The M2-brane 3-form current J
(3)
is given by
J
M
1
M
2
M
3
(x) =
_
W
3

11
(x X)
_
(det g
MN
)
dX
M
1
dX
M
2
dX
M
3
. (2.1.27)
It satises
_
W
3
w
(3)
=
_
M
11
J
(3)
w
(3)
(2.1.28)
for any spacetime 3-form w
(3)
. Since the M2-brane worldvolume W
3
has no bound-
ary, J
(3)
is conserved, i.e. d J
(3)
= 0.
The brane-wave equations for the M2-brane and M5-brane embedding coor-
dinates are

__
det g

X
N
g
PN
(X)
_

1
2
_
det g

X
M

X
N

P
g
MN
(X)
=
1
3!

1
X
M
1

2
X
M
2

3
X
M
3
(F
(4)
(X))
M
1
M
2
M
3
P
, (2.1.29)
CHAPTER 2. M5-BRANE COUPLING TO D = 11 SUPERGRAVITY 20

__
det g

Y
N
g
PN
(Y )
_

1
2
_
det g

Y
M

Y
N

P
g
MN
(Y )
=
1
3!3!

6
(h
(3)
)
[
1

4
Y
M
4

5
Y
M
5

6
]
Y
M
6
(F
(4)
(Y ))
M
4
M
5
M
6
P

1
6!

1
Y
M
1

6
Y
M
6
(

F
(4)
(Y ))
M
1
M
6
P
. (2.1.30)
The equation of motion of the Dirac 6-brane is now equivalent to

[
1
Z
M
1


7
]
Z
M
7
(J
(3)
) = 0 onW
7
. (2.1.31)
In order to satisfy this equation we need to impose that that the Dirac 6-brane must
not cross the M2-brane, since J
(3)
is non-zero only on the M2-brane worldvolume.
This is the analogue of the so-called Dirac Veto, which states that in the presence of
an electrically charged particle, the Dirac string attached to a magnetic monopole
must not cross the particle.
For completeness we also mention the equations of motion of the other M5-brane
worldvolume elds. The dynamics of the chiral 2-form b
(2)
are determined by the
Bianchi identity for the eld strength h
(3)
= db
(2)
A
(3)
,
dh
(3)
= F
(4)
, (2.1.32)
where a pullback of F
(4)
to the M5-brane worldvolume is understood, together with
a generalised self-duality condition for h
(3)
. The latter follows in the PST formalism
from the equation of motion for b
(2)
. As mentioned before, the auxiliary scalar eld
is non-dynamical. For more details, see Refs. [17, 13].
2.2 Normalisation issues in M5-brane coupling
In the previous section, we have shown that the action that describes the consistent
coupling of the M2-brane and the M5-brane to D = 11 supergravity can be written
CHAPTER 2. M5-BRANE COUPLING TO D = 11 SUPERGRAVITY 21
as
S =
1
2
2
_
M
11
R
1
2


F
(4)


F
(4)

1
6
F
(4)
F
(4)
A
(3)
(2.2.1)
T
2
_
W
3
d
3

_
det g

+T
2
_
W
3
A
(3)
(2.2.2)
T
5
_
W
6
d
6

_
det(g

+ (
v
(1)
h
(3)
)

) +
1
2
v
(1)
h
(3)
(v
(1)
h
(3)
)
+
T
5
2
_
W
6
db
(2)
A
(3)
+
T
5
2
_
W
7
A
(3)
F
(4)
(2.2.3)
where F
(4)
= dA
(3)
and

F
(4)
= F
(4)
+ 2
2
T
5
G
(7)
,
h
(3)
= db
(2)
A
(3)
,
v
(1)
=
da

a
(2.2.4)
It is also understood the the worldvolume indices , range over 0, 1, 2 in the M2-
brane source sector, and over 0, 1, , 5 in the M5-brane source sector.
The normalisation of the 3-form potential A
(3)
is xed by convention within the
pure supergravity sector (2.2.1). We have adopted a normalisation such that the
kinetic term for A
(3)
is given relative to the Einstein-Hilbert term by
1
2
2
_
R
1
2
F
(4)
F
(4)
+
_
(2.2.5)
The numerical coecients of the interaction terms in the Mp-brane source sectors
(2.2.2) and (2.2.3) are then xed relative to this initial normalisation choice by the
requirements of gauge invariance and the arbitrariness of the spacetime location of
the Dirac 6-brane
3
.
3
To be precise, these requirements x the numerical coecients of the gauge eld dependent
interaction terms. The relative normalisations of the various terms within each Mp-brane source
sector are determined in the supersymmetric context by the existence of a local -symmetry. As will
be shown in Section 3.3, these relative normalisations can be altered in a purely bosonic context.
CHAPTER 2. M5-BRANE COUPLING TO D = 11 SUPERGRAVITY 22
Another common normalisation convention for the gauge eld is dened by
1
4
2
R
1
2
F
(4)
F
(4)
+ , (2.2.6)
It appears that this normalisation has been adopted by the authors of Ref. [13],
c.f. their equation (2.1), in their work on the coupling of Mp-branes to D = 11
supergravity.
In order to facilitate a comparison of our results with the results of Ref. [13], we
shall now reformulate the action S, as given above, so as to conform with the normal-
isation convention (2.2.6). Accordingly, we rst perform a homogeneous rescaling of
the entire action by a factor of
1
2
and then redene the 3-form potential according
to
A
(3)
2A
(3)
(2.2.7)
To compensate for the overall rescaling of the action in the Mp-brane source sectors,
we also redene the source tensions according to
T
2
2 T
2
, T
5
2 T
5
. (2.2.8)
It is not dicult to see that the action S, as given above, can be written in terms
of the rescaled gauge eld and the rescaled source tensions as
S =
_
M
11
1
4
2
R
1
2


F
(4)


F
(4)


3
F
(4)
F
(4)
A
(3)
(2.2.9)
T
2
_
W
3
d
3

_
det g

+ 2T
2
_
W
3
A
(3)
(2.2.10)
T
5
_
W
6
d
6

_
det(g

+ (
v
(1)
h
(3)
)

) +
1
2
v
(1)
h
(3)
(v
(1)
h
(3)
)
+ T
5
_
W
6
db
(2)
A
(3)
+ 2
2
T
5
_
W
7
A
(3)
F
(4)
(2.2.11)
CHAPTER 2. M5-BRANE COUPLING TO D = 11 SUPERGRAVITY 23
where F
(4)
and v
(1)
are dened as before, while

F
(4)
= F
(4)
+ 2T
5
G
(7)
,
h
(3)
= db
(2)
2A
(3)
.
(2.2.12)
We would like to point out that the numerical coecients of the gauge eld
dependent interaction terms in (2.2.10) and (2.2.11) dier from those presented in
Ref. [13], c.f. their equations (5.1) and (5.28). The denition of the worldvolume
eld strength h
(3)
in (2.2.12) is also dierent from the one used in Ref. [13], c.f. their
equation (5.13). We maintain that the action presented here describes the consistent
coupling of the M2-brane and the M5-brane to D = 11 supergravity. This claim will
be supported in the next section by the demonstration that the eld equations that
follow from the coupled Mp-brane/supergravity action admit both the M2-brane
and the M5-brane supergravity backgrounds as solutions.
2.3 M5-brane solution of the coupled eld equations
The one-half supersymmetric, extreme Mp-brane solution of D = 11 supergravity is
given in the isotropic coordinate system by [2, 3]
ds
2
11
= H

8p
9
p

dx

dx

+H
p+1
9
p

mn
dy
m
dy
n
, (2.3.1)
F
(4)
= dH
1
2
dx
0
dx
1
dx
2
, (M2brane) (2.3.2)
F
(4)
= dH
1
5
dx
0
dx
5
, (M5brane) (2.3.3)
where x

are Minkowski coordinates on the Mp-brane worldvolume and y


m
are
Euclidean coordinates on the (10p)-dimensional transverse space. The function H
p
,
CHAPTER 2. M5-BRANE COUPLING TO D = 11 SUPERGRAVITY 24
which characterises the solution, is a harmonic function of the transverse coordinates,

mn

n
H
p
= 0, lim
|y|
H
p
= 1. (2.3.4)
where [y[ =

mn
y
m
y
n
. The Mp-brane solution exhibits a manifest Poincare sym-
metry along the worldvolume directions x

. The single-charge Mp-brane that we


shall be focusing on also has an additional SO(10 p) symmetry in transverse
space.
It is well known that Laplaces equations (2.3.4) admits non-trivial solutions only
on the punctured Euclidean plane. Therefore, the Mp-brane (2.3.1) and (2.3.2), or
(2.3.3), appears to solve the source free (T
2
= T
5
= 0) supergravity eld equations
(2.1.21)-(2.1.23) everywhere except at the location of the pole of the harmonic func-
tion H
p
. In the case of the M2-brane, this has been taken as an indication that in
order to solve the eld equations everywhere, it is necessary to introduce an M2-
brane source [2]. The introduction of the M2-brane source was also natural from the
point of view of interpreting the M2-brane as an elementary, as opposed to solitonic,
object in D = 11 supergravity, analogous to the fundamental string [29] in D = 10
type IIA supergravity.
In Ref. [2] it was shown that a static M2-brane source, orientated along the x

directions and located at y


m
) = const in transverse space, not only solves the brane-
wave equation (2.1.29) in the M2-brane supergravity background (2.3.1) and(2.3.2),
but also introduces the required -function source terms in the supergravity eld
equations. In particular, (2.3.4) is modied to

mn

n
H
2
= 2
2
T
2

(8)
(y y)). (2.3.5)
in the presence of the M2-brane source with tension T
2
. The solution of (2.3.5)
CHAPTER 2. M5-BRANE COUPLING TO D = 11 SUPERGRAVITY 25
subject to the boundary condition lim
|y|
H
2
= 1 is [2]
H
2
= 1 +
k
2
[y y)[
6
, (2.3.6)
where k
2
=
2
2
T
2
6
7
and
d
is the volume of the unit d-sphere.
In contrast, the M5-brane is usually interpreted as a (non-singular) solitonic
object in D = 11 supergravity, and hence the issue of whether or not to introduce
an M5-brane source does not arise naturally. Nevertheless, we shall show next
that it is possible to couple a static M5-brane source to the M5-brane supergravity
background. We take the metric to be given by (2.3.1), with H
5
a function of the
transverse coordinates only, and replace (2.3.3) by

F
(4)
= dH
1
5
dx
0
dx
5
(2.3.7)
where

F
(4)
= dA
(3)
+ 2
2
T
5
G
(7)
is the modied eld strength which describes the
coupling of the M5-brane source with tension T
5
to the gauge eld via the Dirac
6-brane with 7-form current G
(7)
. The expression (2.3.7) can be dualised to give

F
(4)
=
1
4!

p
H
5

pq

qm
1
m
4
dy
m
1
dy
m
4
. (2.3.8)
The M5-brane source is assumed to be orientated along the x

directions and located


at y
m
) = const in transverse space,
Y

() =

, Y
m
() = y
m
), (2.3.9)
where

are the coordinates on the M5-brane worldvolume W


6
. Recall that we can
place the Dirac 6-brane that ends on the M5-brane arbitrarily in spacetime. We
have found it convenient to choose the Dirac 6-brane coordinates as
Z

(, ) =

, Z
m
(, ) = +y), (2.3.10)
CHAPTER 2. M5-BRANE COUPLING TO D = 11 SUPERGRAVITY 26
where (

, ) are the coordinates on the Dirac 6-brane worldvolume W


7
such that
= 0 denes W
7
= W
6
. It is apparent that the embedding coordinates of the
Dirac 6-brane agree with those of the M5-brane on W
7
.
The M5-brane current J
(6)
and the Dirac 6-brane current G
(7)
, dened in (2.1.5)
and (2.1.6), are here explicitly given by
J
(6)
=
(5)
(y y)) dy
1
dy
5
, (2.3.11)
G
(7)
=
1
5
5

m=1
()
m
(y
m
y
m
))

n=m
(y
n
y
n
)) (2.3.12)
dy
1
dy
m1
dy
m+1
dy
5
.
Consider rst the eld equations for the M5-brane worldvolume elds. We note
that both

F
(4)
and G
(7)
are normal to the M5-brane worldvolume. This implies
that the pullback of F
(4)
=

F
(4)
2
2
G
(7)
to W
6
vanishes. The Bianchi identity
(2.1.32) for h
(3)
thus reduces to dh
(3)
= 0, which we solve by setting h
(3)
= 0.
Clearly, this is also a (trivial) solution of the self-duality condition. For h
(3)
= 0, the
M5-brane worldvolume energy-momentum tensor T

given by (2.1.22) simplies to


T

= g

. It is then easy to verify that the brane-wave equation (2.1.30) in the


M5-brane supergravity background (2.3.1) and (2.3.7) is satised for any function
H
5
of the transverse coordinates only.
Next, we turn to the supergravity eld equations (2.1.21)-(2.1.23) (with T
2
= 0).
It is possible to show that they reduce to the single equation,

mn

n
H
5
= 2
2
T
5

(5)
(y y)), (2.3.13)
which is analogous to (2.3.5). The solution of equation (2.3.13) subject to the
CHAPTER 2. M5-BRANE COUPLING TO D = 11 SUPERGRAVITY 27
boundary condition lim
|y|
H
5
= 1 is
H
5
= 1 +
k
5
[y y)[
3
, (2.3.14)
where k
5
=
2
2
T
5
3
4
.
This calculation establishes that also the M5-brane supergravity background
(2.3.1) and (2.3.7) with a harmonic function given by (2.3.14) can be interpreted as
the eld conguration to the exterior of an innite static M5-brane source, orien-
tated along the x

directions and located at y


m
) = const in transverse space.
We have just demonstrated that the Mp-brane supergravity background is (for
an appropriate choice of the harmonic function H
p
) also a solution of D = 11
supergravity coupled to a single, static Mp-brane source. By construction, the source
is located at y
m
= y
m
) in transverse space. In general, the location of the Mp-brane
in transverse space coincides with the location of the pole of the harmonic function.
The use of Euclidean coordinates y
m
in transverse space suggests that the pole of the
harmonic function is point-like. In other words, the Mp-brane source appears to be
localised at a point in transverse space. However, it is well known that the Euclidean
coordinates are ill-suited for the description of the Mp-brane spacetime in the vicinity
of the pole of the harmonic function [9, 30]. While the pole of the harmonic function
(or, equivalently, the transverse location of the Mp-brane source) indeed appears
to be point-like from the coordinate perspective, the equation [y y)[ = 0 in fact
denes an entire surface in transverse space. A simple way to verify this is to
calculate the invariant volume of the surface dened by x

= const, [y y)[ =
R = const. One nds that even in the limit R 0, the volume of this surface
does not vanish. Therefore, what appears to be a localised Mp-brane source from
CHAPTER 2. M5-BRANE COUPLING TO D = 11 SUPERGRAVITY 28
the coordinate perspective is in fact a distribution of Mp-branes, spread out over
the surface [y y)[ = 0 in transverse space. This will be demonstrated explicitly
in the next chapter, where we shall also show how to construct a truly point-like
Mp-brane source as the zero-radius limit of a spherically symmetric Mp-brane source
distribution.
A separate issue concerns the motivation for introducing the Mp-brane source in
the rst place. In the case of the M2-brane, the source appeared to be necessary
to account for the singularity of the harmonic function at the location of the pole
[2]. Subsequently, however, it has been realised [9, 30] that the singularity of the the
(source-free) Mp-brane spacetime at the location of the pole is merely a coordinate
singularity and that the surface [y y)[ = 0 is in fact a non-singular degenerate
horizon. It is possible to analytically continue the Mp-brane spacetime through the
horizon, whence it becomes apparent that in the M2-brane spacetime the horizon
clothes a time-like curvature singularity at the M2-brane core. In contrast, the
region in the interior of the M5-brane horizon is isometric to the exterior region.
The M5-brane spacetime is therefore entirely non-singular [31].
Chapter 3
Spherically symmetric Mp-brane
sources
3.1 Coupling Mp-brane sources to extreme Mp-branes
We have seen in Section 2.3 that the use of Euclidean coordinates in transverse space
obscures the fact that the source that couples naturally to the Mp-brane supergravity
background is a distribution of Mp-branes, rather than a truly point-like source
that is localised in transverse space. The aim of the the present chapter is to make
this more explicit, by using a variety of coordinate systems which permit a well-
dened description of the spacetime geometry in the vicinity of the location of the
source.
The extreme Mp-brane supergravity background can be rewritten in a coordinate
system that makes the SO(10 p) symmetry of the solution explicit by introducing
hyperspherical coordinates (r,
i
) in transverse space, where r is an isotropic radial
29
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 30
coordinate and
i
are angular coordinates on the unit (9 p)-sphere:
ds
2
11
= H

8p
9
p

dx

dx

+H
p+1
9
p
_
dr
2
+r
2
d
2
9p
_
, (3.1.1)
F
(4)
= dH
1
2
dx
0
dx
1
dx
2
, (M2brane) (3.1.2)
F
(4)
= dH
1
5
dx
0
dx
5
. (M5brane) (3.1.3)
In (3.1.1), d
2
9p
is the standard metric on the unit (9 p)-sphere, while H
p
is a
harmonic function of the radial coordinate only. In the absence of Mp-brane sources,
H
p
is determined by

r
_
r
9p

r
H
p
_
= 0, lim
r
H
p
= 1. (3.1.4)
The Mp-brane is characterised by its ADM mass /
p
per unit p-volume and an
electric or magnetic Page charge Q
e/m
. For the asymptotically at metric of the
form (3.1.1), the ADM mass can simply be read o the asymptotic behaviour of the
harmonic function, c.f. Ref. [32],
/
p
=

9p
2
2
lim
r
_
r
9p

r
H
p
_
. (3.1.5)
The conserved electric and magnetic Page charges are quite generally [33] given by
the integrals
Q
e
=
1
2
2
_
F
(4)
+
1
2
A
(3)
F
(4)
, (3.1.6)
Q
m
=
1
2
2
_
F
(4)
, (3.1.7)
where, in the present context, the integration surface is the (9 p)-sphere at trans-
verse innity.
It is not dicult to verify that for the extreme Mp-brane given above we have
/
p
= Q
e/m
. The equality of the ADM mass and Page charge can also be viewed as
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 31
a consequence of the partial supersymmetry preservation of the bosonic Mp-brane
supergravity background, see e.g. Ref. [31].
In the hyperspherical coordinate system, the horizon is at r = 0, which is also
commonly interpreted as the location of the Mp-brane. We shall show next how
to couple a spherically symmetric Mp-brane source distribution to this supergravity
background and nd that it is possible to place the Mp-brane source at arbitrary
radial distance from the horizon.
Accordingly, we assume that the supergravity elds are given by (3.1.1) and
(3.1.2), or (3.1.3), with F
(4)
replaced by

F
(4)
in the case of the M5-brane. The
function H
p
= H
p
(r) remains to be determined. In this supergravity background, a
single, static M2-brane source, orientated along the x

directions, satises the brane-


wave equation (2.1.29) irrespective of its location in transverse space. The same is
true of a single, static M5-brane source orientated along the x

directions, after
solving the equations of motion for the chiral worldvolume 2-form b
(2)
by h
(3)
= 0.
The calculations required to establish this are analogous to those presented in Section
2.3 and will therefore not be repeated here.
However, it is clear that a single Mp-brane source, located at a constant radial
distance r) from the origin and at some point
i
) on the sphere in transverse space,
does not couple to the Mp-brane supergravity background, because it necessarily
breaks the explicit SO(10 p) symmetry of the background
1
. So, rather than
considering a single Mp-brane source, we introduce a continuous distribution of static
Mp-branes (orientated along the x

directions), located at a xed radial distance r)


1
Recall that SO(10p) acts transitively on the (9p)-sphere. Therefore, no point
i
on S
9p
is left invariant by the action of SO(10 p).
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 32
from the origin, but uniformly spread out over the sphere in transverse space.
Since each of the Mp-branes in the distribution satises its own brane-wave
equation, the introduction of an Mp-brane source distribution, as opposed to a
single Mp-brane source, mainly consists in replacing the single Mp-brane source
currents J
(r,)
(p+1)
and energy-momentum tensors T
(r,)
MN
in (2.1.21)-(2.1.23) by
those appropriate for a distribution of Mp-brane sources,
J
r
(p+1)
=
_
d
9p
()) ()) J
(r,)
(p+1)
, (3.1.8)
T
r
MN
=
_
d
9p
()) ()) T
(r,)
MN
, (3.1.9)
where ()) is the density function of the distribution. The density function for a
uniform distribution of Mp-branes is constant
()) =
1

9p
, (3.1.10)
where we have chosen a normalisation such that the total charge of the distribution
is equal to the charge of a single Mp-brane.
Due to the modied source terms (3.1.8) and (3.1.9), the equation which deter-
mines the harmonic function H
p
is now not just the coordinate transformed version
of (2.3.5) or (2.3.13), but rather

r
_
r
9p

r
H
p
_
=
2
2
T
p

9p
(r r)). (3.1.11)
The continuous, everywhere nite solution of (3.1.11) subject to the boundary con-
dition lim
r
= 1 is
H
p
= 1 +k
p
_
(r r))
r
8p
+
(r) r)
r)
8p
_
, (3.1.12)
where k
p
=
2
2
T
p
(8p)
9p
.
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 33
The ADM mass (3.1.5) of the Mp-brane spacetime with the harmonic function
(3.1.12) is given in terms of the tension of the Mp-brane source by /
p
= T
p
. In
particular, the ADM mass is independent of the location r) of the source, which
shows that even for given ADM mass, we can place the Mp-brane source at an
arbitrary radial distance r) from the horizon.
For non-zero r), the Mp-brane with the harmonic function (3.1.12) coincides
with the source-free Mp-brane solution of D = 11 supergravity [2, 3] only in the
exterior (r > r)) of the source. In the interior of the source (r < r)), the harmonic
function (3.1.12) is constant, implying that the interior solution can be identied
with the Minkowski vacuum of D = 11 supergravity. The change in geometry across
the location of the source is accompanied by a -function curvature singularity at
r = r). This, however, is the only singularity of the solution. In particular, the
interior region of the spacetime is entirely non-singular.
As has been mentioned before, the isotropic coordinate system does not cover
the entire Mp-brane spacetime [9, 30]. A particularly simple way of seeing this is to
consider the invariant radial measures for the class of Mp-branes with SO(10 p)
symmetry in transverse space that are related to one another by a transformation
of the radial coordinate. In principle, there are two such invariant radial measures
[34]. The rst is given by an integral over the (square root) of the radial component
of the metric
R
inv
(r) =
_
r
0
dr

_
g
rr
(r

), (3.1.13)
while the second is derived from the volume of the sphere (at radius r) in transverse
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 34
space

R
inv
(r) =
_
1

9p
_
d
9p

_
det g
ij
(r)
_ 1
9p
. (3.1.14)
In general, i.e. in the exterior of the Mp-brane source, these radial measures are
not the same, R
inv
(r) ,=

R
inv
(r) for r > r). They do agree, however, in the interior
of the source, R
inv
(r) =

R
inv
(r) for r r), because the spacetime geometry is at
there. It follows that the invariant radial distance (from the origin) to the location
r) of the Mp-brane source, as measured by (3.1.13), coincides with the radius of
the sphere in transverse space at the location of the source, as measured by (3.1.14).
Hence, one can describe the invariant location of the Mp-brane source equivalently as
the invariant radial distance to the location of the source, or as the invariant radius
of the sphere over which the source is distributed. This implies that truly point-like
Mp-brane sources, i.e. spherically symmetric sources distributed over a sphere with
zero radius, can only be located at the core of the Mp-brane spacetime, i.e. at zero
invariant radius. Conversely, an Mp-brane source located at nite invariant radius
appears necessarily spread out over a sphere of nite radius from the invariant point
of view.
Using (3.1.13) or (3.1.14), it is easy to calculate that the invariant location of
the Mp-brane source is related to its coordinate location in the isotropic coordinate
system of (3.1.1) by
R
inv
)
8p
= r)
8p
+k
p
. (3.1.15)
The limit r) 0
+
corresponds to coupling the Mp-brane source at the horizon.
In this limit, the harmonic function (3.1.12) becomes
2
H
p
= 1 + k
p
r
(8p)
for
2
Consistent limits of -functions are taken according to the rules lim
x0
+ (x) = 1 and
lim
x0
(x) = 0.
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 35
r > 0. This harmonic function is naturally associated with the solution of the
source-free supergravity eld equations, c.f. equations (3.1.4) and (2.3.6). We also
see from (3.1.15) that as the source approaches the horizon, its invariant radius
tends to R
inv
) = k
1
8p
p
. This establishes the claim made at the end of Section
2.3 that the source-free Mp-brane supergravity solution, expressed in either of the
isotropic transverse coordinates systems (r,
i
) or y
m
, can be interpreted as the
eld conguration to the exterior of a spherically symmetric distribution (with non-
zero invariant radius) of static Mp-branes located at the horizon of the Mp-brane
spacetime.
It is of interest to express the Mp-brane spacetime in a coordinate system for
which the coordinate location of the source coincides with its invariant location.
Accordingly, we introduce a so-called Schwarzschild radial coordinate by

8p
= r
8p
+k
p
(r r)) +k
p
r
8p
r)
8p
(r) r). (3.1.16)
In the exterior of the source, (3.1.16) reduces to the usual transformation [9, 30]
between the Schwarzschild and the isotropic radial coordinates. In the interior of the
source, however, the two radial coordinates are related by a simple rescaling, because
the spacetime geometry is at there. This coordinate transformation is continuous
across the location of the source. In particular, the Schwarzschild coordinate location
of the source ) is given in terms of the isotropic coordinate location r) by
)
8p
lim
rr

8p
= r)
8p
+k
p
, (3.1.17)
which, together with (3.1.15), implies that ) = R
inv
). In other worlds, the
Schwarzschild coordinate location of the source coincides with its invariant loca-
tion.
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 36
The Mp-brane with harmonic function given by (3.1.12) can be rewritten in terms
of the Schwarzschild radial coordinate as
ds
2
11
=
8p
9
p

dx

dx

+
_
1 k
p
( ))

8p
_
2
d
2
+
2
d
2
9p
, (3.1.18)
F
(4)
= d
2
dx
0
dx
1
dx
2
, (M2brane) (3.1.19)

F
(4)
= d
5
dx
0
dx
5
, (M5brane) (3.1.20)
where

p
= 1 k
p
_
( ))

8p
+
() )
)
8p
_
. (3.1.21)
and ) is dened in (3.1.17).
The exterior solution ( > )) now exhibits a degenerate horizon at

= k
1
8p
p
,
provided that the source is located at ) <

. For convenience, we shall continue


to refer to the surface =

as the horizon, even if it is covered by the source


() >

) and, strictly speaking, is only a would-be horizon. The interior solution


( < )) can again be identied with the Minkowski vacuum of D = 11 supergravity.
The only singularity of this solution is a -function curvature singularity at the
location of the source. This is also where the radial component of the metric is
discontinuous
3
.
The Schwarzschild radial coordinate can be analytically continued through the
horizon. This makes it possible to consider Mp-brane sources coupled inside the
horizon. The limit ) 0
+
corresponds to coupling the Mp-brane source at the core
of the spacetime, i.e. at zero invariant distance from the origin. In this limit, the Mp-
brane (3.1.18) and (3.1.19), or (3.1.20) is characterised by a single function
p
= 1
3
This discontinuity is a consequence of the fact that the Jacobian of the transformation (3.1.16)
is discontinuous across the location of the source.
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 37
k
p

(8p)
for > 0. This function is naturally associated with the Mp-brane solution
(in Schwarzschild coordinates) of the source-free supergravity eld equations. This
solution therefore acquires the interpretation of the eld conguration to the exterior
of a static Mp-brane source located at the core of the spacetime. The Mp-brane
source is truly point-like, since the invariant radius of the spherically symmetric
source distribution vanishes as the source approaches the core of the spacetime.
In the Schwarzschild coordinate system, the horizon is located at a non-zero
distance

= k
1
8p
p
from the coordinate origin. In principle, it should be possible
to place the Mp-brane source exactly at the horizon, while still being able to de-
scribe the interior spacetime region. However, in the limit that the Mp-brane source
approaches the horizon, i.e. as p)

, we see from (3.1.21) that

p
( ))
_
1
k
p

8p
_
(3.1.22)
so that the metric components along the worldvolume directions vanish for < ).
A calculation of the tangent space components of the Riemann curvature tensor,
however, reveals that the spacetime in the interior region remains strictly at
4
. It is
therefore only the Schwarzschild coordinate system that is ill-suited for a description
of the interior spacetime region when the Mp-brane source is located at the horizon.
In the next section, we probe the Mp-brane spacetime with test branes and photons,
in order to gain a better understanding of the eects of the source location on the
spacetime geometry.
4
This observation corrects statements made in Refs. [15, 35] to the eect that the singularity at
the core of the M2-brane spacetime persists even after the introduction of an M2-brane source at
the horizon.
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 38
3.2 Probing the Mp-brane with photons and test branes
The coordinate systems we have used so far have failed to aord us an adequate
description of the spacetime geometry when the Mp-brane source is located exactly
at the horizon. As an alternative to nding a better set of coordinates, we explore
the spacetime geometry in the presence of the Mp-brane source by considering the
motion of test photons and test branes. Similar methods have been used in [36]
to ascertain the singularity structure of certain p-brane spacetimes in D = 10. We
follow the probe along a radial trajectory from some point in the exterior of the
source into the core of the spacetime and back out again, keeping the source location
initially arbitrary. Strikingly, we nd that when the source is located exactly at the
horizon, the probe traverses the entire at interior spacetime region in zero proper
time, so that the interior region is completely inaccessible from the probes point of
view.
Consider rst the motion of a light-like particle, i.e. a photon, in the background
spacetime of an Mp-brane source, given by (3.1.18). Let X
M
= X
M
() be the coor-
dinates of the photon, where is an ane parameter along the photons worldline.
The photon follows a null geodesic in the Mp-brane spacetime and its coordinates
are thus determined by
d
d
_
g
MN
(X)

X
N
_
=
1
2

M
g
PQ
(X)

X
P

X
P
, (3.2.1)
where the dot denotes a derivative with respect to . We restrict our attention to
purely radial motion and set
X
0
() = t(), X

() = (), (3.2.2)
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 39
while all the other coordinates of the photon remain constant throughout its motion.
Evaluating the null geodesic equation (3.2.1) for M = 0, we easily nd that
E =
8p
9
p

t (3.2.3)
is a constant of motion. In addition, we may use that ds
2
11
= 0 on the photons
worldline, whence
0 =
8p
9
p

t
2
+
2
()

2
, (3.2.4)
where
p
is given by (3.1.21) and we have dened
()
= 1 k
p
( ))
(8p)
.
Equations (3.2.3) and (3.2.4) jointly imply that
= E
()

8p
18
p
. (3.2.5)
The total ane parameter distance along the worldline of the photon from some
radial distance
0
(in the exterior of the Mp-brane source) to the core of the spacetime
at = 0 and back to
0
is therefore given by the integral
=
1
E
_
0

0
d

8p
18
p

()
+
1
E
_

0
0
d

8p
18
p

()
, (3.2.6)
where we have chosen the minus (plus) sign for in the rst (second) term, as
appropriate for inbound (outbound) motion.
Because of the discontinuity of the function
()
at = ), the evaluation of the
integrals in (3.2.6) requires a careful limiting procedure. In a fairly obvious notation,
we set
_

0
0
lim
0
+
_

0
+lim
0
+
_

0
+
,
_
0

0
lim
0
+
_
+

0
+lim
0
+
_
0

.
(3.2.7)
This limiting procedure has been chosen to guarantee that the ane parameter
distance of the inbound trip is equal to the ane parameter distance of the outbound
trip, as required by symmetry.
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 40
Using (3.2.7), the integrals in (3.2.6) can be evaluated as
=
2)
E
_
1
_

_
8p
_
+
2
E
_

0

d
_
1
_

_
8p
_

10+p
18
, (3.2.8)
where the rst (second) term represents the contribution of the motion through the
interior (exterior) spacetime region and we have used that
8p

= k
p
.
Note that in the limit )

, i.e. in the limit that the Mp-brane source


approaches the horizon, the rst term in (3.2.8) vanishes, while the second term
remains nite. In other words, if the Mp-brane source is located exactly at the
horizon, it takes the photon zero ane parameter time to cross the entire at interior
region of the spacetime. Once the inbound photon has crossed the horizon, it re-
emerges instantaneously as an outbound photon at opposite angles. The interior
region of the spacetime is therefore completely inaccessible from the point of view
of the photon.
Consider now the motion of a test M2-brane (a test M5-brane) in the background
spacetime of an M2-brane source (an M5-brane source), given by (3.1.18) with
p = 2 (p = 5). Let X
M
= X
M
() be the coordinates of the test brane, where

are coordinates on the test branes worldvolume. The motion of the test brane
is determined by the appropriate brane-wave equation (2.1.29), or (2.1.30). For
simplicity, assume that the test brane remains either parallel or anti-parallel to
the source brane throughout its motion. For the M5-brane probe, this implies in
particular that we may set the eld strength h
(3)
of the chiral worldvolume 2-form
b
(2)
to zero.
Recall that the net force between two static parallel Mp-branes vanishes, due to
an exact cancellation of the attractive gravitational force and the repulsive electro-
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 41
magnetic force, see e.g. Ref. [14]. Two parallel Mp-branes that are initially at rest
therefore remain at rest. Reversing the orientation of one of the Mp-branes eectively
amounts to changing the sign of its charge. This implies that the electro-magnetic
force between two static anti-parallel Mp-branes is attractive, and hence that the
net force between two anti-parallel Mp-branes is also attractive. Even if the anti-
parallel Mp-branes were at rest initially, they invariably start accelerating towards
one another.
We have found it convenient to assume that the test brane is initially at rest.
Hence, we choose the anti-parallel orientation for the test brane to get a non-trivial
motion in the spacetime background of the static Mp-brane source. Note that for the
anti-parallel orientation, we have
01p
= 1 in the brane-wave equations (2.1.29)
and (2.1.30). We restrict our attention to purely radial motion and set
X
0
() = t(), X
1
() =
1
, X
p
() =
p
, X

() = (), (3.2.9)
where =
0
is proper time and the angular coordinates of the test brane remain
constant throughout its motion. The derivative with respect to will be denoted
by a dot in the following.
Evaluating the brane-wave equations (2.1.29) and (2.1.30) (with h
(3)
= 0) for
P = 0, we nd that
E =
p
_
_
_
_

t
_

t
2

8p
9
p

2
()

2
+ 1
_
_
_
_
(3.2.10)
is a constant of motion. In addition, we may use that d
2
= ds
2
11
on the test
branes worldvolume, whence
1 =
8p
9
p

t
2

2
()

2
. (3.2.11)
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 42
Equations (3.2.10) and (3.2.11) jointly imply that
= E

()

p
_
1
2
E

p
. (3.2.12)
For a test brane initially at rest ( (0) = 0) in the exterior region of the source brane,
the constant of motion E is given in term of the initial radial distance (0) =
0
of
the test brane from the origin by
E = 2
_
1
_

0
_
8p
_
. (3.2.13)
As explained above, the test brane accelerates from rest towards the source
brane until it reaches the location of the source brane. Provided that it survives the
encounter with the source brane, the test brane continues to traverse the entire at
interior spacetime region at constant velocity. Eventually, the it re-emerges from the
interior region at opposite angles to its entry point. From there on, the test brane
decelerates until it comes to rest again at its original radial distance from the origin.
The proper time that elapses during the round trip of the test brane is given by
=
1
E
_
0

0
d

p

()
_
1
2
E

p
_

1
2
+
1
E
_

0
0
d

p

()
_
1
2
E

p
_

1
2
, (3.2.14)
where we have chosen the minus (plus) sign for in the rst (second) term, as
appropriate for inbound (outbound) motion.
The integrals in (3.2.14) have to be evaluated using the limiting procedures
(3.2.7). We then nd that
=
2)
E
1
_

_
8p

1
2
E
_
1
_

_
8p
_
+
2
E
_

0

1
2
E
_
1
_

_
8p
_
, (3.2.15)
where the rst (second) term represents the contribution of the motion through the
interior (exterior) spacetime region and we have used that
8p

= k
p
.
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 43
Note that in the limit )

, i.e. in the limit that the Mp-brane source


approaches the horizon, the rst term in (3.2.8) vanishes, while the second term
remains nite. In other words, if the Mp-brane source is located exactly at the
horizon, it takes the test brane zero proper time to cross the entire at interior region
of the spacetime. In this sense, the test branes behave analogously to the photon
probe discussed earlier. Once the inbound test brane has crossed the horizon, it re-
emerges instantaneously as an outbound test brane at opposite angles. The interior
region of the spacetime is therefore completely inaccessible from the point of view
of the test brane.
The fact the both the photon probe and the test branes cross the interior space-
time region instantaneously according to their own proper time when the source
brane is located at the horizon opens up an interesting interpretation of what hap-
pens from the probes point of view. As the probe crosses the horizon it experiences
the -function curvature singularity due to the presence of the source brane. So it
can distinguish the spacetime region it traverses according to before the singularity
and after the singularity. However, because of the rotational symmetry of the Mp-
brane spacetime, the spacetime region the probe traverses before the singularity
is isometric to the region it travels through after the singularity. Therefore, from
the probes point of view, the source at the horizon partitions the spacetime into two
distinct but isometric regions. This interpretation is consistent with the observation
that a distant observer who follows the probes inbound motion using coordinate
time never sees the probe reaching the source, since it is located at the horizon. A
fortiori, the distant observer never sees the probe re-emerging at opposite angles
in the isometric spacetime region beyond the location of the source.
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 44
It is interesting to note that the motion of the probe in the M5-brane spacetime
with a source at the horizon seems very much alike the motion of the probe in the
M5-brane spacetime without a source, since in the latter the spacetime region in
the interior of the horizon is known to be isometric to the spacetime region in the
exterior of the horizon [31]. Introducing an M5-brane source at the horizon is rather
innocuous from the probes point of view. Apart from the appearance of a -function
curvature singularity, not much else appears to have changed.
3.3 Coupling Mp-brane sources to non-extreme Mp-branes
The degeneracy of the horizon of the extreme Mp-brane can be lifted by breaking
the Poincare symmetry on the Mp-brane worldvolume to IRE
p
, where E
d
denotes
the Euclidean group in d dimensions. The non-supersymmetric, non-extreme, or
black, Mp-brane is given in the Schwarzschild coordinate system by [3, 43, 44]
ds
2
11
=
8p
9
p
_

(+)

()
dt
2
+dx
2
_
+
1

(+)

()
d
2
+
2
d
2
9p
, (3.3.1)
where

(+/)
= 1
_

+/

_
8p
,
p
=
()
. (3.3.2)
The black Mp-brane exhibits a non-singular outer event horizon at =
+
.
For
+
>

, the outer horizon clothes an inner event horizon at =


+
, which,
however, coincides with a curvature singularity. Note that for
+
<

, this singu-
larity is naked. There is an additional curvature singularity at the core of the black
Mp-brane spacetime at = 0. In the limit that
+

, i.e. in the limit that the


two horizon coalesce, the singularity at the inner horizon disappears and the black
Mp-brane become the extreme Mp-brane.
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 45
The ADM mass /
p
per unit p-volume of the black Mp-brane is given in terms
of the parameters
+/
by
/
p
=

9p
2
2
_
(8 p)
8p

+ (9 p)(
8p
+

8p

)
_
. (3.3.3)
In addition, the black M2-brane carries an electric Page charge Q
e
dened by (3.1.6),
while the black M5-brane carries a magnetic Page charge Q
m
dened by (3.1.7).
Instead of presenting the expression for the eld strength F
(4)
that supports the
black Mp-brane, we content ourselves here with the value of the Page charge,
Q
e/m
=
(8 p)
9p
2
2
(
+

)
8p
2
. (3.3.4)
By considering the ratio //Q
e/m
as a function of
+
/

it is not dicult to establish


that

/ Q
e/m

/ Q
e/m
(3.3.5)
It is of interest to determine whether the black Mp-brane can be reinterpreted
as a solution of D = 11 supergravity coupled to an appropriate Mp-brane source,
and to determine the source action. It can be veried that the equations of mo-
tion for the coupled Mp-brane/supergravity theory (2.1.21)-(2.1.23) (with the source
terms replaced by (3.1.8) and (3.1.9)) do not admit the black Mp-brane supergrav-
ity background as a solution. However, this should not be too surprising, as we
are trying to couple (the bosonic part of) a supersymmetric Mp-brane source to a
non-supersymmetric supergravity background.
To see how to do better, recall that as consequence of (partial) spacetime super-
symmetry preservation the ADM mass (per unit p-volume) of the extreme Mp-brane
is equal to its electric or magnetic Page charge. This corresponds directly to the
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 46
equality of the mass parameter (tension) T
p
and charge parameter Q
p
= T
p
of the
Mp-brane source which couples to the extreme Mp-brane supergravity background.
From a worldvolume point of view, the equality of the mass and charge parameters
of the Mp-brane source can also be viewed as a consequence of the local -symmetry
of the supersymmetric Mp-brane action. In contrast, we see from (3.3.5) that the
ADM mass of the black Mp-brane is generically not equal to its Page charge. It
is therefore natural to expect that the Mp-brane source which couples to the black
Mp-brane supergravity background should have independent mass and charge pa-
rameters. Note that the action that we shall present for the black Mp-brane source
is not to be interpreted as the bosonic sector of a supersymmetric action with local
-symmetry.
It is easy to modify the M2-brane action to allow for independent mass and
charge parameters by changing the coecient in the WZ term (2.1.4). The action
for the black M2-brane is given by
S
M2(black)
= T
2
_
W
3
d
3
_
det g

+Q
2
_
W
3
A
(3)
. (3.3.6)
Modifying the M5-brane action requires a bit more work. In the PST formu-
lation, one has to be careful to preserve the bosonic symmetries of the M5-brane
action, as they are instrumental in deriving the self-duality condition of the world-
volume eld strength h
(3)
. We nd that the action describing a black M5-brane
with independent mass and charge parameters is given by
S
M5(black)
= T
5
_
W
6
d
6

det(g

+
Q
5
T
5
(
v
(1)
h
(3)
)

Q
5
2
_
W
6
v
(1)
h
(3)
(v
(1)
h
(3)
)
+
Q
5
2
_
W
6
db
(2)
A
(3)
+
Q
5
2
_
W
7
A
(3)
F
(4)
, (3.3.7)
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 47
Using the methods of [17], it can be veried that the generalised self-duality condition
of h
(3)
= db
(2)
A
(3)
follows from the equations of motion for b
(2)
, just as in the Q
5
=
T
5
case. The worldvolume 1-form v
(1)
is given in terms of an auxiliary scalar eld
a() by v
(1)
= da (g

a)

1
2
. The auxiliary eld is non-dynamical in the sense
that its equation of motion is implied by the self-duality of h
(3)
. We should also point
out that the modied eld strength that describes the coupling of the M5-brane to
the gauge eld via the Dirac 6-brane is now given by

F
(4)
= dA
(3)
+ 2
2
Q
5
G
(7)
.
A rather lengthy calculation shows that the equations of motion that follow
from D = 11 supergravity coupled to a spherically symmetric distribution of static
Q
p
,= T
p
Mp-brane sources admit the black Mp-brane spacetime as a solution. The
functions that appear in the metric (3.3.1) are given by

p
= 1
8p

_
( ))

8p
+
() )
)
8p
_
, (3.3.8)

()
= 1
8p

( ))

8p
, (3.3.9)

(+)
= 1
8p
+
( ))

8p

8p
() )

8p
. (3.3.10)
The parameters
+/
and are given in terms of the mass and charge parameters
of the Mp-brane source and its coordinate location ) by

8p

=

k
p
e
X()
, (3.3.11)

8p
+
=

k
p
e
X()
, (3.3.12)

8p
= )
8p

8p
+

8p

)
8p

8p

, (3.3.13)
where

k
p
=
2
2
Q
p
(8p)
9p
. The quantity e
X()
is dened as the positive root of the
quadratic equation
e
2X()

k
p
)
8p
_
1
Q
2
p
T
2
p
_
e
X()

Q
2
p
T
2
p
= 0. (3.3.14)
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 48
and explicitly given by
e
X()
=

k
p
2 T
2
p
)
8p
_
_
_
_
T
2
p
Q
2
p
_
+

_
_
T
2
p
Q
2
p
_
2
+
_
2 T
p
Q
p
)
8p

k
p
_
2
_
_
_.
(3.3.15)
For completeness, we also present the gauge eld conguration for the black Mp-
brane solution in the presence of the Mp-brane source:
F
(4)
=
2
2
Q
2

7
( ))
(7)
, (M2brane) (3.3.16)

F
(4)
=
2
2
Q
5

4
( ))
(4)
, (M5brane) (3.3.17)
where
(p)
denotes the volume p-form of the unit p-sphere.
In the region exterior to the Mp-brane source ( > )), this solution coincides
with the source-free black Mp-brane spacetime (3.3.1). It exhibits a non-singular
horizons at =
+
and a singular horizon at =

, provided that the source


does not cover the horizons () <
+/
). In contrast to the extreme Mp-brane
spacetime, the geometry in the interior region ( < )) is not at. In particular,
the curvature singularity at the core of the black Mp-brane at = 0 persists. There
is an additional -function curvature singularity at the location of the source.
The ADM mass (3.3.3) and the Page charge (3.3.4) of the black Mp-brane are
functions of the source tension T
p
, the source charge Q
p
and the source location ),
/
p
= Q
p
_
e
X()
+
9p
8p
_
e
X()
e
X()
__
,
Q
e/m
= Q
p
.
(3.3.18)
The value of the Page charge of the black Mp-brane therefore xes the charge pa-
rameter Q
p
of the source. However, the source tension T
p
that simultaneously solves
(3.3.18) and (3.3.14) for given ADM mass and Page charge is a non-trivial function
of the source location ), or vice versa. It follows that the values of source tension
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 49
and the source location cannot be chosen independently. Rather, they have to be
adjusted jointly for a given value of the ADM mass. This is in contrast to the situ-
ation encountered in the coupling of an Mp-brane source to the extreme Mp-brane,
where the ADM mass determined the source tension independently of the source
location.
Alternatively, we may view the ADM mass and the Page charge as being a pos-
teriori determined by the source parameters T
p
, Q
p
and ), via equations (3.3.18).
This point of view has several interesting consequences which we shall explore now.
Firstly, we note that in the limit Q
p
T
p
in which the black Mp-brane source
becomes (the bosonic sector of) a supersymmetric Mp-brane source, the function
X()) in (3.3.15) vanishes. It then follows from the denitions (3.3.11) and (3.3.12)
of
+/
that

+
=

=

k
1
8p
p
= k
1
8p
p
=
2
2
T
p
(8 p)
9p
, (3.3.19)
while the parameter dened in (3.3.13) vanishes. Therefore, the entire black
Mp-brane solution becomes precisely
5
the extreme Mp-brane solution discussed in
Section 3.1. This observation also implies that one cannot couple a supersymmetric
Mp-brane source to the black Mp-brane supergravity background, as claimed above.
More interestingly, it is possible to project the black Mp-brane solution onto the
extreme Mp-brane solution in yet another way. The deviation from extremality of
the black Mp-brane is eectively measured by the function X()). The roots of
X()) correspond to situations in which the non-extreme black Mp-brane coincides
with an extreme Mp-brane, since whenever X()) = 0 we see from (3.3.18) that
5
It is not dicult to show that for Q
p
= T
p
the expressions (3.3.16) and (3.3.17) for F
(4)
are the
Hodge duals of (3.1.19) and (3.1.20) respectively.
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 50
/
p
= Q
e/m
, or, equivalently, from (3.3.11) and (3.3.12) that
+
=

. We have just
seen that one of the roots of X()) is at Q
p
= T
p
. To determine the other roots,
we use its dening equation (3.3.14), which for X()) = 0 implies that
_
1

k
p
)
8p
__
1
Q
2
p
T
2
p
_
= 0 (3.3.20)
It follows that X()) vanishes also if the coordinate location of the source is chosen
as ) =

k
1
8p
p
. Since
+
=

=

k
1
8p
p
= ) in that case, the Mp-brane source is lo-
cated at the horizon of the extreme Mp-brane spacetime. We stress that in contrast
to the Q
p
T
p
extremal limit discussed previously, the )

k
1
8p
p
extremal limit
does not imply any functional relation between the mass and charge parameters of
the source. Indeed, the source tension T
p
is entirely unobservable, since the ADM
mass and the Page charge in (3.3.18) are both equal to the source charge Q
p
. This is
the Mp-brane analogue of the mass renormalisation phenomenon rst reported by
ADM some thirty years ago [45]. They showed that the observable mass of the non-
extreme Reissner-Nordstrm black hole solution of D = 4 Einstein-Maxwell theory,
coupled to a spherically symmetric distribution of electrically charged particles, be-
comes independent of the source mass, and given only in terms of the source charge,
in the limit that the source distribution approaches the horizon
6
.
Another issue of interest that arises from the source parameter dependence of
the ADM mass and the Page charge is whether the singularity of the black Mp-brane
6
In their original paper [45], ADM reported the renormalisation of the black hole mass in the
limit that the spherically symmetric source distribution becomes point-like, meaning that it had
zero coordinate radius. However, as pointed out in Ref. [34], ADM were using an isotropic radial
coordinate, for which the zero coordinate radius limit corresponds to placing a source with non-zero
invariant radius at the horizon.
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 51
spacetime at the inner horizon can ever be naked. This point is most easily discussed
by considering the source parameter dependence of the coordinate locations
+/
of
the horizons directly. Recall that the inner horizon at =

is singular provided
that the Mp-brane source does not cover it, i.e. provided that ) <

. Furthermore,
this singularity is naked if it is not clothed by the non-singular outer horizon at
=
+
, i.e. if
+
<

. In the following, we exploit the inequalities,


T
p
> Q
p
T
p
< Q
p
)

k
1
8p
p
)

k
1
8p
p

+
)
+

k
1
8p
p

k
1
8p
p

+

k
1
8p
p

k
1
8p
p

+
)
(3.3.21)
which follow from the denitions (3.3.11)-(3.3.15).
It is apparent from (3.3.21) that the black Mp-brane spacetime develops a naked
singularity only if T
p
< Q
p
and the source is located inside the horizons. In con-
trast, if T
p
> Q
p
, the source either covers both would-be horizons, or the outer
horizon clothes the singularity at the inner horizon. Therefore, in order to avoid the
occurrence of a naked singularity in the black Mp-brane spacetime, the mass and
charge parameters of the Mp-brane source should satisfy T
p
Q
p
.
The general picture that emerges from (3.3.21) is that for ) >

k
1
8p
p
, the Mp-
brane source covers both would-be horizons, which are at this stage still non-
singular. As the Mp-brane source approaches ) =

k
1
8p
p
, the source location coin-
cides with the location of both horizons. In this limit, the ADM mass of the black
Mp-brane is renormalised to be independent of the source tension. The entire
black Mp-brane solution becomes indistinguishable from an extreme Mp-brane so-
lution, albeit with ADM mass equal to Q
p
rather than T
p
. Pulling the Mp-brane
source further inside the horizon towards ) <

k
1
8p
p
immediately results in a devi-
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 52
ation from extremality in which the degeneracy of the horizon is lifted. The inner
horizon at =

is singular, while the outer horizon at =


+
is regular. The sin-
gularity is naked only if T
p
< Q
p
. In the limit that the Mp-brane source approaches
the core of the spacetime, i.e. as ) 0, we nd that
/+
0, while
+/
,
for T
p
> Q
p
and T
p
< Q
p
respectively, in such a way that
+

Q
2
8p
p
remains
constant. In either case, one of the horizons approaches the core of the spacetime,
while the other horizon is pushed out to innity.
3.4 Patching D = 11 supergravity solutions with Mp-
brane sources
Consider again the supergravity backgrounds with explicit P
p+1
SO(10 p) sym-
metry given by (3.1.1) and (3.1.2), or (3.1.3), where P
d
denotes the Poincare group
in d dimensions. Even though this supergravity background is commonly associ-
ated with an Mp-brane, it in fact serves as a blueprint for a more general class of
solutions. The dependence of the metric and the eld strength on the harmonic func-
tion H
p
= H
p
(r) is essentially dictated by the explicit P
p+1
SO(10 p) symmetry
and the requirement that the background preserve a non-zero amount of residual
supersymmetry. It is the choice of harmonic function that determines the overall
spacetime geometry, rather than the general functional dependence of the solution
on the harmonic function.
In Table 3.1, we have summarised how the values of the integration constants a
and k in the harmonic function H = a +k r
(8p)
inuence the spacetime geometry
of the supergravity background. The manifest P
p+1
SO(10 9) symmetry of the
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 53
H = a +k r
(8p)
Spacetime Symmetry SuSy
a > 0, k = 0 Minkowski P
11
1
a > 0, k > 0 Mp-brane P
p+1
SO(10 p)
1
2
a = 0, k > 0 AdS
p+2
S
9p
SO(p + 1, 2) SO(10 p) 1
Table 3.1: Harmonic functions and spacetime geometries.
blueprint is enlarged for k 0 or a = 0 to the full symmetry group of the Minkowski
or the AdS
p+2
S
9p
vacuum of D = 11 supergravity.
In the asymptotic limit r , respectively the near horizon limit r 0, the
harmonic function of the Mp-brane approaches that of the Minkowski vacuum, re-
spectively the AdS
p+2
S
9p
vacuum. The Mp-brane therefore interpolates between
the Minkowski vacuum at innity and the AdS
p+2
S
9p
vacuum at the horizon
[9, 30]. Furthermore, it has been realised that for k ,= 0, the value of the integration
constant a can be altered by applying a combination of dimensional reduction and
oxidation techniques with the duality symmetries of M-theory and string theory
[37, 38, 39, 40, 41]. In particular, the Mp-brane solution can be mapped onto the
AdS
p+2
S
9p
vacuum, and vice versa, by switching the constant a o or on.
In Section 3.1, we have shown that a static Mp-brane source, orientated along
the x

directions and localised at an arbitrary point (r),


i
)) in transverse space,
solves the brane-wave equation in the supergravity background (3.1.1) and (3.1.2),
or (3.1.3) for any function H
p
= H
p
(r) with a non-trivial dependence on the radial
coordinate. This shows that it is possible to place a static Mp-brane source at
arbitrary radial distance from the horizon in the Mp-brane and the AdS
p+2
S
9p
supergravity backgrounds, as has rst been pointed out in Ref. [42]. However, we
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 54
have also argued that it is not possible to solve the supergravity eld equations with
an SO(10 p) symmetric ansatz in the presence of a localised Mp-brane source,
since the location of the source in transverse space necessarily breaks the SO(10p)
symmetry of the supergravity background.
In order to solve the supergravity eld equations exactly, we had to introduce a
distribution of Mp-branes, uniformly spread out over the (9p)-sphere at coordinate
radius r). The harmonic function H
p
that characterises the solution in the presence
of the Mp-brane source satises

r
_
r
9p

r
H
p
_
=
2
2
T
p

9p
(r r)), (3.4.1)
where T
p
is the total charge, or tension of the Mp-brane source. The continuous,
everywhere nite solution of (3.4.1) is given by
H
p
= a +(r r))
_
k
p
r
8p
_
+(r) r)
_
k
p
r)
8p
_
, (3.4.2)
where
k
p
=
2
2
T
p
(8 p)
9p
(3.4.3)
and a is an integration constant. Since lim
r
H
p
= a, the value of a determines
the asymptotic geometry of the spacetime. As in the source-free case, its value can
be shifted arbitrarily by using the duality symmetries of the theory.
The harmonic function (3.4.2) can also be written as
H
p
= (r r))
_
a +
k
p
r
8p
_
+(r) r)
_
a +
k
p
r)
8p
_
, (3.4.4)
which implies that the exterior solution (r > r)) is an Mp-brane for a = 1, or an
AdS
p+2
S
9p
vacuum for a = 0, while the interior solution (r < r)) is strictly
at and can be identied with the Minkowski vacuum. This establishes that it
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 55
is possible to couple a spherically symmetric distribution of Mp-branes, located at
arbitrary radial r) distance from the origin, not only to the Mp-brane spacetime,
but also to the AdS
p+2
S
9p
vacuum
7
.
So far, we have only considered a single Mp-brane source. The presence of the
Mp-brane source (at non-zero radius) naturally partitions the overall spacetime in
an interior and an exterior region. In this section, we shall show more generally
that the introduction of N Mp-brane sources at dierent radii partitions the overall
spacetime in N+1 distinct regions, each characterised by its own harmonic function.
Therefore, one may view the role of an Mp-brane source in D = 11 supergravity as
permitting a consistent patching of distinct (source-free) spacetime regions.
Accordingly, we place N Mp-brane sources with tensions T
i
p
, orientated along
the x

directions and uniformly distributed over the (9 p)-sphere at radius r


i
), in
the supergravity background given by (3.1.1) and (3.1.2), or (3.1.3). Without loss of
generality, we assume that the Mp-brane sources are ordered such that r
i
) < r
i+1
).
The harmonic function H
p
that characterises the solution satises the multi-source
generalisation of (3.4.1),

r
_
r
9p

r
H
p
_
=
N

i=1
2
2
T
i
p

9p
(r r
i
)), (3.4.5)
7
It is interesting to note that the invariant radial distance of the Mp-brane source from the
origin of the AdS
p+2
S
9p
spacetime, as dened by the measure (3.1.13), is independent of its
coordinate distance r and just given by k
1
8p
p
. This can be understood by recalling that the
invariant distance of the Mp-brane source from the origin coincides with the radius of the sphere
over which the Mp-brane source is spread out, since the geometry in the interior of the source is
at. The radius of the (9 p)-sphere in the AdS
p+2
S
9p
vacuum is clearly constant, and indeed
equal to k
1
8p
p
.
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 56
The continuous, everywhere nite solution of (3.4.5) is given by
H
N
p
= a +
N

i=1
(r r
i
))
_
k
i
p
r
8p
_
+(r
i
) r)
_
k
i
p
r
i
)
8p
_
, (3.4.6)
where
k
i
p
=
2
2
T
i
p
(8 p)
9p
(3.4.7)
and a is an integration constant that determines the asymptotic geometry of the
spacetime.
The presence of the Mp-brane sources naturally partitions the overall spacetime
into N+1 distinct regions, as is evident by the dependence of the harmonic function
(3.4.6) on the source locations r
i
) via the -functions. Let 1
i
denote the spacetime
region dened by r
i
) r < r
i+1
) and let 1
0
be the innermost region dened
by 0 r < r
1
). We dene the Page charge Q
e/m
i
for the region 1
i
by (3.1.6) and
(3.1.7), where the integration is to be performed over a (9 p)-sphere lying in the
region 1
i
. Note that the total amount of charge enclosed in the volume bound by
two (9p)-spheres in adjacent regions 1
i1
and 1
i
is simply equal to the charge T
i
p
of the Mp-brane source, located at that separates the two regions. The page charges
Q
e/m
i
must therefore satisfy the charge conservation equations
Q
e/m
i
Q
e/m
i1
= T
i
p
, i = 1, . . . , N. (3.4.8)
By assumption, the innermost spacetime region 1
0
is at, so that its Page charge
vanishes, Q
e/m
0
= 0. We conclude from (3.4.8) and Q
e/m
0
= 0 that
Q
e/m
i
=
i

j=1
T
j
p
. (3.4.9)
Therefore, the Page charge Q
e/m
i
measures the total amount of charge deposited by
the Mp-brane sources in the spacetime region 0 r < r
i+1
).
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 57
Next, we determine the harmonic function that characterises the solution in the
region 1
i
. Let H
i
p
= H
N
p
(r 1
i
) denote the harmonic function (3.4.6) restricted to
the region 1
i
. It is not dicult to infer form (3.4.6) that
H
i
p
= H
i1
p
+
k
i
p
r
8p

k
i
p
r)
8p
, H
0
p
= a +
N

i=1
k
i
p
r)
8p
. (3.4.10)
It then follows from (3.4.10), (3.4.7) and (3.4.9) that H
i
p
can be written as
H
i
p
= a
i
+
k
i
r
8p
, (3.4.11)
where the constants a
i
and k
i
are given by
a
i
= a +
N

j=i+1
k
i
p
r
i
)
, (3.4.12)
k
i
=
2
2
Q
e/m
i
(8 p)
9p
. (3.4.13)
For future reference, we observe that the constants a
i
satisfy
a
i
a
i1
=
k
i
p
r)
8p
, (3.4.14)
while the charge conservation equation (3.4.8) can be rewritten in terms of the
constants k
i
as
k
i
k
i1
= k
i
p
, (3.4.15)
where k
i
p
is given in terms of the tension T
i
p
of the Mp-brane source at r
i
) by (3.4.7).
Let us briey summarise what we have shown so far. We have demonstrated that
the introduction of N Mp-brane sources located at distinct radii r
i
) partitions the
overall spacetime into N+1 regions. The solutions within each region is characterised
by a harmonic function H
i
p
= a
i
+k
i
r
(8p)
. The constants a
i
satisfy the matching
condition (3.4.14), which ensures that the harmonic function is continuous across the
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 58
locations of the Mp-brane sources. The constants k
i
satisfy the matching condition
(3.4.15), which guarantees charge conservation across the locations of the Mp-brane
sources. The matching conditions (3.4.14) and (3.4.15) uniquely dene the entire
solution, once we specify the value of the constants a
N
and k
0
. These are determined
by the behaviour of the harmonic function is the asymptotic (r ) limit and the
near horizon (r 0) limit, since
H
p
a
N
as r ,
H
p

k
0
r
8
p
as r 0.
(3.4.16)
Note also that a non-trivial behaviour of the harmonic function near the origin
eectively indicates the presence of an Mp-brane source in the vicinity of the origin.
This Mp-brane source near the origin can be viewed as the r) 0 limit of an
Mp-brane source which is initially located at a non-zero distance r) from the origin.
Therefore, if the innermost Mp-brane source is located at a non-zero distance from
the origin, we may without loss of generality set k
0
= 0 (as we have done above).
We conclude this section with an example, which is adapted from Ref. [42], where
it was used to argue that it is possible to place an Mp-brane source at arbitrary radial
distance from the origin in the AdS
p+2
S
9p
vacuum. It aords us with a situation
in which the techniques presented here can be applied to give exact results.
Consider N+1 static, spherically symmetric Mp-brane sources with equal tension
T
p
, orientated along the x

directions and assume that they are initially stacked on


top of one another at some non-zero radial distance r). Now move N of the Mp-
brane sources away form the remaining one and place them in the vicinity of the
origin. The corresponding harmonic function is a special case of (3.4.6) and given
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 59
by
H

p
= a+(r)
_
Nk
p
r
8p
_
+(r)
_
Nk
p

8p
_
+(rr))
_
k
p
r
8p
_
+(r)r)
_
k
p
r)
8p
_
,
(3.4.17)
where is the innitesimal distance of the N Mp-branes from the origin and k
p
is
dened in (3.4.3). In the limit 0
+
, i.e. as the stack of Mp-branes is moved to
the origin, the harmonic function (3.4.17) can be written as
H
p
= (r r))
_
a +
(N + 1) k
p
r
8p
_
+(r) r)
_
a +
k
p
r)
8p
+
Nk
p
r
8p
_
(3.4.18)
This harmonic function is characteristic of the supergravity solution describing a
static test Mp-brane placed at radial distance r) in the background provided
by N Mp-brane sources at the origin. The background itself is dened by the
harmonic function

H
p
= a +
Nk
p
r
8p
(3.4.19)
and is therefore AdS
p+2
S
9p
if a = 0, or an Mp-brane spacetime if a = 1.
We can see explicitly from (3.4.18) how the introduction of the test brane per-
turbs the background. Not surprisingly, the total charge in the exterior region
(r > r)) is increased by one unit. In the interior region (r < r)), the harmonic
function of the background is modied by the addition of the constant
k
p
r
8p
. This
implies that even if the background is AdS
p+2
S
9p
so that a = 0, the harmonic
function in the interior has a constant part and is more akin to that of an Mp-brane
spacetime.
As the number of source branes is increased, the perturbation of the background
due to the test brane becomes more and more negligible and in the limit N
CHAPTER 3. SPHERICALLY SYMMETRIC MP-BRANE SOURCES 60
, T
p
0 such that NT
p
const, the harmonic function (3.4.18) tends to the
harmonic function (3.4.18) of the background.
There is, however, yet another way of minimising the perturbation of the back-
ground due to the test brane. In the limit that the test brane is pushed out to
(transverse) innity, i.e. as r) , the harmonic function (3.4.18) also tends to
(3.4.19). To see exactly what happens in this limit, assume rst that the back-
ground is asymptotically at (a = 1). At rst, it seems as if we have encountered
a paradox, since it is not possible to couple a static test brane source (with non-
zero charge) consistently to the Minkowski vacuum. To resolve this paradox, recall
that the charge of the test brane is uniformly distributed over the (9 p)-sphere in
transverse space. As the test brane is pushed out to innity, the physical volume, as
measured by the Minkowski metric, of the (9 p)-sphere over which the test brane
is spread out becomes innite. Therefore, the physical charge density of the test
brane vanishes. We have eectively removed the test brane from the spacetime by
pushing it out to innity.
In contrast, if a = 0 and the background is asymptotically AdS
p+2
S
9p
, the
physical volume of the (9p)-sphere over which the test brane is spread out remains
constant as the test brane is pushed out to innity. The only thing that happens in
this case it that the test brane is located at the boundary of the AdS space. As is
explained in more detail in Ref. [42], the test brane, for p = 2, can then be identied
with the membrane at the end of the universe that has been constructed in Ref.
[7].
Chapter 4
Conclusions
In the preceding chapters, we have presented a careful treatment of the Mp-brane
sources that couple to various D = 11 supergravity backgrounds, including both the
extreme and the non-extreme Mp-brane background, and the AdS
p+2
S
9p
vacuum
backgrounds. As a preliminary, we have also derived the action that describes the
consistent coupling of the M5-brane to D = 11 supergravity, using an approach that
dispenses with the dual 6-form potential, but instead highlights the importance of
introducing the Dirac 6-brane consistently.
We have shown that the Mp-brane source that couples to the extreme Mp-brane
supergravity background is a spherically symmetric distribution of Mp-branes, rather
than a single Mp-brane source. This source could be placed at arbitrary radial
distance from the horizon, and even inside the horizon. We have also constructed a
point-like source as the (invariant) zero-radius limit of the spherically symmetric
source distribution.
The introduction of the Mp-brane source modies the global structure of the
spacetime. Whereas the source free Mp-brane spacetime interpolates along an in-
61
CHAPTER 4. CONCLUSIONS 62
nite throat between the Minkowski vacuum at transverse innity and the AdS
p+2

S
9p
vacuum at the horizon, the introduction of the source at nite (isotropic) co-
ordinate radius eectively places a cap on the throat. The spacetime region in the
interior of the source can be identied with Minkowski space. We have gained fur-
ther information about the Mp-brane spacetime that explicitly includes the source
by studying the motion of with test branes and photons in this background. We
found that if the source is located exactly at the horizon, the spacetime region in
the interior of the source becomes inaccessible from the point of view of the probe.
By allowing the Mp-brane source to have independent mass and charge param-
eters, we were able to construct the source distribution that couples to the non-
extreme, or black, Mp-brane supergravity backgrounds. In contrast to the extreme
case, the location of the source that couples to the black Mp-brane background can-
not be chosen arbitrarily. Rather, it is dependent on the conserved ADM mass and
the Page charge of the black Mp-brane. Alternatively, one may view the ADM mass
as a function of the source location. This point of view lead to the generalisation of
the ADM mass-renormalisation phenomena [45] to black Mp-branes.
The coupling of Mp-brane sources to other supergravity backgrounds, in partic-
ular to the AdS
p+2
S
9p
Freund-Rubin vacua, has also been discussed briey. In
this context, we have commented on how Mp-brane sources can be used to patch
together distinct (source-free) solutions of D = 11 supergravity.
Part II
Freund-Rubin vacua in maximal
supergravities
63
Chapter 5
Introduction
5.1 Aristocratic p-branes and vacuum interpolation
The M2-brane and the M5-brane in D = 11 are part of a distinguished family of
p-brane solutions of supergravity theories, whose members are sometimes referred
to as aristocratic p-branes [46]. The dening feature of an aristocratic p-brane is the
existence of a regular horizon, near which all the elds that participate in the p-brane
solution are non-singular. Now, it is well known [30] that there exists a frame, i.e.
a conformal rescaling of the metric, such that the metric and the gauge eld(s) are
non-singular near the p-brane horizon. However, in general, the dilaton(s) diverge
near the horizon, so that in practice the aristocratic p-branes may be identied with
regular dilaton p-branes, i.e. with p-brane solutions whose dilatonic scalars tend to
constants near the horizon.
There is no dilaton in D = 11 supergravity, so that the M2-brane and the
M5-brane are aristocratic by birth. Similarly, the self-dual D3-brane solution of
IIB supergravity [47] does not involve the IIB dilaton and is also aristocratic by
64
CHAPTER 5. INTRODUCTION 65
birth. These p-branes are indeed the only aristocratic single charge p-brane solutions
of maximal supergravity theories. The lower-dimensional maximal supergravities
in D 9, however, admit a number of p-brane solutions which may be termed
aristocratic by marriage, as they are supported by more than one eld strength
and can be interpreted as bound states (with zero binding energy) of several single
charge p-branes [48, 49, 50]. They are the dyonic string [51] in D = 6, the 3-charge
black hole and the 3-charge string [52] in D = 5 and the 4-charge black hole [53]
in D = 4. To be more precise, there are in fact 5 aristocratic strings in D = 6, 45
aristocratic black holes and strings each in D = 5 and 630 aristocratic black holes
in D = 4, which dier from one another in being supported by distinct sets of eld
strengths, but assemble into an irreducible multiplet under the Weyl subgroup of
the Cremmer-Julia global symmetry group [83].
These aristocratic p-branes share a number of interesting features which are in-
timately linked with the regularity of the dilatonic scalars near the horizon. In each
case, they interpolate between two vacuum solutions of the D-dimensional super-
gravity theory [9, 38, 54], namely D-dimensional Minkowski space at (transverse)
innity and a constant scalar solution of the form AdS
p+2
S
D2p
with covari-
antly constant eld strengths at the horizon. The latter solution may be regarded
as a vacuum, as it is the ground state of the compactication of the D-dimensional
supergravity theory on S
Dp2
to a (p+2)-dimensional spacetime. The aristocratic
p-branes also exhibit supersymmetry enhancement near the horizon [38, 54, 58, 59],
where the number of unbroken supersymmetries of the bulk p-brane solution are
doubled.
The lower-dimensional aristocratic p-branes are interpreted from the 11-dimensional
CHAPTER 5. INTRODUCTION 66
perspective as harmonic intersections [55, 56, 57] of M2-branes, M5-branes, waves
and monopoles, depending on eld strengths that support the lower-dimensional p-
brane, see e.g. [83]. More recently, it was found [60, 38] that certain non-standard
harmonic intersections [61, 62] of p-branes are also aristocratic in the above sense,
i.e. the dilatonic scalar(s) are everywhere regular. These non-standard intersec-
tions of p-branes can also be interpreted as solitons that interpolate between the
Minkowski vacuum at innity and a certain compactifying vacuum solution of the
form AdS
2/3
S
2/3
S
2/3
T
d
in the near horizon region close to all p-branes
in the intersection. Just as the other aristocratic p-branes, they also exhibit super-
symmetry doubling in the near horizon region [38].
An interesting question is whether the list of known aristocratic p-branes is
complete. Our approach to answering this question is based on the observation that
the duality symmetries of string theory and M-theory, combined with Kaluza-Klein
reduction/oxidation techniques, make it possible to map the aristocratic p-brane
solution onto the corresponding near horizon vacuum solution, by switching o
the constant(s) in the harmonic function(s) that characterise the p-brane [37, 38, 39,
40, 41]. This mapping is invertible, so that one may recover the aristocratic p-brane
solution from the corresponding vacuum solution, by switching on the constant(s)
in the harmonic function(s). This observation implies that an alternative to looking
for aristocratic p-branes is to look for the vacuum solutions which feature in their
near horizon limits. In contrast to the p-brane solutions themselves, these vacua
have a rather simple structure. In all known cases, the D-dimensional spacetime is
a product manifold, with up to three maximally symmetric, non-at, submanifolds,
and possibly additional toroidal factors. The eld strengths are covariantly constant
CHAPTER 5. INTRODUCTION 67
and the dilatonic scalars are constant. In fact, vacuum solutions of this type can be
viewed as generalisations of the Freund-Rubin AdS
4/7
S
7/4
vacuum solutions [65]
of D = 11 supergravity, which emerge in the near horizon limit of the M2-brane or
the M5-brane [9].
In the remainder of the second part of this thesis, we shall present a classication
of Freund-Rubin vacua in the maximal, massless, supergravity theories in D 11.
In the next section, we detail our approach to the classication of vacuum solutions.
Chapter 6 then starts o with a review of the original Freund-Rubin vacua in D =
11. This leads us naturally to consider a generalised Freund-Rubin ansatz, which
forms the basis for the subsequent classication of vacuum solutions of D = 11
supergravity with more than two Einstein submanifolds. We shall also show how
to adapt this generalised Freund-Rubin ansatz to the eld content of the maximal,
massless supergravity theories in D < 11.
The Freund-Rubin vacuum solutions of IIA and IIB supergravity are discussed
in Chapter 7. It is shown that all the IIA vacua have a higher-dimensional inter-
pretation as Hopf reductions of D = 11 vacua, which extends results obtained some
time ago in Ref. [77]. We also argue that the IIB vacua assemble naturally into
multiplets of the Weyl subgroup [78] of the global SL(2, IR) symmetry group of IIB
supergravity, even though it turns out that the vacuum solutions of IIB supergravity
that a priori cannot be toroidally reduced to D < 10 are singlets of the Weyl group.
Chapter 8 deals with the vacuum solutions of the maximal, massless supergrav-
ity theories in D 9. The actual classication of the lower-dimensional vacua is
preceded by a discussion of the equations of motions for the bosonic elds of these
supergravities and their global Cremmmer-Julia symmetries. Particular emphasis
CHAPTER 5. INTRODUCTION 68
is laid on the Weyl subgroup of the Cremmer-Julia symmetry group, which eects
permutations of the eld strengths of a given degree among themselves, as it serves
as the classifying symmetry for the Freund-Rubin vacua that we are seeking. The
Weyl multiplets of vacua in D 9 are then presented in turn.
The residual spacetime supersymmetries of every vacuum solution that we have
classied are determined in Chapter 9, by considering the integrability conditions of
an appropriate Killing spinor equation in the vacuum background. The number of
independent integrability conditions that have to be considered is reduced dramati-
cally by the observation that most of the vacuum solutions that we have found are
inter-related via Kaluza-Klein dimensional reduction/oxidation and/or Weyl trans-
formations. Finally, we comment on the identication of the aristocratic p-branes
that interpolate between the Freund-Rubin vacua and Minkowski space.
5.2 An approach to the classication of vacuum solu-
tions in D 11
In this section, we detail our approach to the classication of vacuum solutions in
maximal supergravity theories. To begin with, let us concentrate on vacuum solu-
tions of D = 11 supergravity. We have found it helpful to keep in mind the interpre-
tation of a vacuum solution as the ground state of a spontaneous compactication of
D = 11 supergravity on some (11 D)-manifold K
11D
to a D-dimensional space-
time M
D
. These vacuum solutions that can therefore be alternatively referred to as
compactifying solutions.
At rst sight, this line of thought may not appear to be particularly useful,
CHAPTER 5. INTRODUCTION 69
since the concept of spontaneous compactication presupposes the existence of a
stable solution of the 11-dimensional eld equations [63]. However, suppose that
it is known that the compactication of D = 11 supergravity on K
11D
admits a
consistent truncation to the massless sector and, moreover, that the embedding of
the D-dimensional massless eld into the 11-dimensional elds is known explicitly.
In other words, suppose that the Kaluza-Klein ansatz for the consistent dimensional
reduction of D = 11 supergravity on K
11D
to a D-dimensional supergravity theory
is known. In that case, the ground state of the compactication on K
11D
, i.e. the
vacuum in D = 11, may be obtained by oxidising a suitable D-dimensional vacuum,
obtained by solving the eld equations of the D-dimensional supergravity theory.
For an arbitrary compactifying manifold K
11D
, the truncation of the D-dim-
ensional elds to the massless sector is most likely to be inconsistent [64]. However,
if K
11D
is the (11 D)-torus, the truncation of the D-dimensional elds to the
massless sector is known to be consistent and the toroidal dimensional reduction of
D = 11 supergravity yields IIA supergravity in D = 10, or the maximal, massless
supergravity theory in D 9. The Kaluza-Klein ansatz for the toroidal reduction of
D = 11 supergravity is reviewed in Appendix B.1. The vacua of D = 11 supergravity
that are the ground states of toroidal compactications can therefore be obtained
alternatively by oxidising vacuum solutions of IIA supergravity, or the maximal
massless supergravity theories in D 9.
The preceding paragraphs suggest the following approach to the classication
of vacua in maximal supergravities in D 11. We start o in D = 11, classifying
only those vacuum solutions of the 11-dimensional eld equations that cannot be
toroidally reduced to D < 11. Next, we move on to D = 10, classifying only
CHAPTER 5. INTRODUCTION 70
those vacuum solutions of the IIA eld equations that cannot be toroidally reduced
to D < 10. Separately, we also want to consider vacua in type IIB supergravity.
However, since the S
1
reduction of type IIB supergravity yields a 9-dimensional
supergravity theory that is equivalent, up to eld redenitions, to the supergravity
theory obtained from the S
1
reduction of type IIA supergravity, it is again sucient
to classify only those vacuum solutions of the IIB eld equations that cannot be
toroidally reduced to D < 10. The T-duality mapping between the type IIA and
type IIB supergravities in D = 9 is reviewed in Appendix B.2. Finally, we repeat
a similar classication of vacua in every maximal, massless supergravity theory in
D 9.
The classication of vacua in maximal supergravities in D 11 has now been
reduced to the classication of a certain class of vacuum solutions, namely those that
cannot be further toroidally reduced, in each supergravity theory in D 11. To
construct these vacuum solutions, we shall use an ansatz that is a straightforward
generalisation of the Freund-Rubin ansatz [65] that leads to the AdS
4/7
S
7/4
vacua
in D = 11. This will be made more precise in the following chapter.
Chapter 6
Vacua in D = 11 supergravity
6.1 Review of the Freund-Rubin vacua
It has been known for a long time [65] that D = 11 supergravity admits solutions
of the form M
4
K
7
and M
7
K
4
, where M
4/7
is a Lorentzian signature Einstein
manifold and K
7/4
is a Euclidean signature Einstein manifold. The M
4
K
7
solution
is the ground state of the spontaneous compactication of D = 11 supergravity on
K
7
to D = 4, see e.g. Ref. [66] for a review. Similarly, the M
7
K
4
solution is
the ground state of the spontaneous compactication of D = 11 supergravity on K
4
to D = 7 [67]. The M
4/7
K
7/4
solutions can therefore be regarded as vacua of
D = 11 supergravity. The vacuum solutions with the maximal spacetime symmetry,
and indeed supersymmetry, are the AdS
4
S
7
and AdS
7
S
4
vacua that emerge in
the near horizon limit of the M2-brane and the M5-brane. The 4 + 7 split of the
11-dimensional spacetime is not arbitrary, but rather determined by the requirement
71
CHAPTER 6. VACUA IN D = 11 SUPERGRAVITY 72
of a non-zero vacuum expectation value (VEV) of the 4-form eld strength
1
.
It is instructive to review the construction of the M
4/7
K
7/4
vacua in a bit
more detail. The equations of motion for the 11-dimensional metric g
MN
and the
3-form gauge potential A
(3)
can be written as
1
MN
= S
MN
(F
(4)
), (6.1.1)
d F
(4)
=
1
2
F
(4)
F
(4)
, (6.1.2)
dF
(4)
= 0, (6.1.3)
where F
(4)
= dA
(3)
and
S
MN
=
1
2
_
(F
(4)
)
2
MN

1
3
g
MN
(F
(4)
)
2
_
. (6.1.4)
We start by discussing the M
4
K
7
solution. The Freund-Rubin ansatz for the
metric and the gauge eld is [65]
ds
2
11
= g

(x) dx

dx

+g
mn
(y) dy
m
dy
n
, (6.1.5)
F
(4)
= u
(4)
F
(4)
= u
(7)
, (6.1.6)
where x

and y
m
are coordinates on M
4
and K
7
, u is a constant and

(4)
=
1
4!

4
dx

1
dx

4
=
_
det g

dx
0
dx
3
, (6.1.7)

(7)
=
1
7!

m
1
m
7
dy
m
1
dy
m
7
=
_
det g
mn
dy
1
dy
7
(6.1.8)
are the volume forms on M
4
and K
7
.
1
If the VEV of the 4-form eld strength is zero, there cannot be a natural split of the 11-
dimensional spacetime into a product of two or more submanifolds. The corresponding vacua are
Ricci at 11-manifolds, Minkowski spacetime being the maximally symmetric, and indeed maximally
supersymmetric, one.
CHAPTER 6. VACUA IN D = 11 SUPERGRAVITY 73
Notice that since u is a constant, the components F
M
1
M
4
of the 4-form eld
strength are covariantly constant. Moreover, the non-zero components of F
(4)
point
only along the directions of the 4-manifold. Alternatively, in the language of dier-
ential forms, F
(4)
is a harmonic 4-form that lies entirely on M
4
. It is in this sense
that F
(4)
has a constant VEV (with magnitude u) on the 4-manifold M
4
.
The gauge eld equations (6.1.2) and (6.1.3) are identically satised for the
ansatz (6.1.6), since F
(4)
is a harmonic 4-form and F
(4)
F
(4)
= 0. Furthermore, the
energy-momentum tensor (6.1.4) becomes block-wise proportional to the metric.
The eld equation for the metric (6.1.1) therefore simplies to
1

=
1
3
u
2
g

, 1
mn
=
1
6
u
2
g
mn
. (6.1.9)
These equations imply that M
4
is a Lorentzian signature Einstein manifold with
scalar curvature
2
(M
4
) =
1
9
u
2
, while K
7
is a Euclidean signature Einstein man-
ifold with scalar curvature (K
7
) =
1
36
u
2
. The maximally symmetric solution of
(6.1.9) is AdS
4
S
7
.
The M
7
K
4
solution is obtained analogously. The Freund-Rubin ansatz for
the metric is of the form (6.1.5), with the understanding that x

and y
m
are now
coordinates on M
7
and K
4
. The ansatz for the gauge eld is replaced by
F
(4)
= u
(4)
F
(4)
= u
(7)
, (6.1.10)
where u is a constant and the volume forms
(7)
and
(4)
are dened similarly to
(6.1.7) and (6.1.8). It is evident that F
(4)
is again a harmonic 4-form with a constant
VEV (of magnitude u) on the 4-manifold K
4
. Indeed, the only dierence between
2
The scalar curvature of a d-dimensional Einstein manifold with metric g

is dened by
R

= (d 1) g

and hence related to the Ricci scalar R of the d-manifold by R = d(d 1) .


CHAPTER 6. VACUA IN D = 11 SUPERGRAVITY 74
the M
7
K
4
ansatz and the M
4
K
7
ansatz is an interchange of the signatures of
the 4-manifold and the 7-manifold.
The gauge eld equations are satised identically for the ansatz (6.1.10), while
the Einstein equation reduces to
1

=
1
6
u
2
g

, 1
mn
=
1
3
u
2
g
mn
. (6.1.11)
It follows that M
7
is a Lorentzian signature Einstein manifold with scalar cur-
vature (M
7
) =
1
36
u
2
, while K
4
is a Euclidean signature Einstein manifold with
scalar curvature (K
4
) =
1
9
u
2
. The maximally symmetric solution of (6.1.11) is
AdS
7
S
4
.
As has been mentioned already, the M
4/7
K
7/4
solution is the ground state of the
compactication of D = 11 supergravity on K
4/7
. Let us focus on compactications
on the round 4-sphere, or the round 7-sphere. In that case, the truncation of the
lower-dimensional elds to the massless sector, i.e. the Kaluza-Klein dimensional
reduction of D = 11 supergravity on S
4
and S
7
, is believed to be consistent [68, 67]
and yields gauged maximal SO(5) supergravity in D = 7 [69] and gauged maximal
SO(8) supergravity in D = 4 [70] respectively. The AdS S vacuum solutions of
D = 11 supergravity then imply the existence of an AdS vacuum solution of the
gauged supergravities in D = 7 and D = 4.
Conversely, given the consistency of the S
4
and S
7
reductions, any solution of
D = 7 and D = 4 gauged supergravity can be reinterpreted as a solution of D = 11
supergravity. In principle, it should therefore be possible to obtain the AdSS vacua
in D = 11 by oxidation of the AdS vacua in D = 7 and D = 4. In practice, however,
the oxidation of the 7-dimensional or 4-dimensional solution to D = 11 presupposes
CHAPTER 6. VACUA IN D = 11 SUPERGRAVITY 75
that the embedding of the lower-dimensional elds into the 11-dimensional elds is
known explicitly. This embedding is highly non-trivial for a general reduction of
D = 11 supergravity on S
4
or S
7
, in which all the lower-dimensional massless elds
are retained, in contrast to the embedding of the lower-dimensional massless elds
for a toroidal reduction of D = 11 supergravity. The full non-linear Kaluza-Klein
ansatz for the general reduction of D = 11 supergravity on S
4
has only recently
been constructed [71], while that for the S
7
reduction is not yet known explicitly.
The situation simplies dramatically if one retains only (and all) those elds in
the spherical reduction that are singlets under the isometry group of the compactify-
ing sphere. The consistency of this truncation is guaranteed, since the uncharged
elds that have been retained cannot act as sources for the charged elds that
have been discarded [64]. These types of spherical reductions of D = 11 super-
gravity have been considered in Ref. [72], where they have been called breathing
mode reductions, because the 11-dimensional metric reduces to just the metric and
a breathing mode scalar in the lower dimension. In Ref. [72], it has also been
shown that the resulting lower dimensional theories in D = 7 and D = 4 admit AdS
vacua, with constant breathing mode scalar, that oxidise to the AdS S vacua in
D = 11.
6.2 A generalised Freund-Rubin ansatz
The Freund-Rubin ansatz used in the construction of the M
4/7
K
7/4
vacua is
characterised by a 4 + 7 split of the 11-dimensional spacetime, the properties of
the Einstein submanifolds M
4/7
and K
7/4
and a harmonic 4-form eld strength F
(4)
CHAPTER 6. VACUA IN D = 11 SUPERGRAVITY 76
with a non-zero, constant VEV on the 4-manifold. In this section, we generalise the
Freund-Rubin ansatz by assuming that the 11-dimensional spacetime is the product
of more than two Einstein submanifolds.
To begin with, consider an 11-dimensional spacetime that is topologically the
product manifold M
d
1
K
d
2
K
d
N
, where

N
i=1
d
i
= 11. Here, M
d
(K
d
)
denotes a d-manifold that admits a Lorentzian (Euclidean) signature Einstein metric,
satisfying
1

= (d 1)g

, (6.2.1)
where is the scalar curvature of the d-manifold. Suppose for a moment that a stable
solution of the 11-dimensional eld equations of the form M
d
1
K
d
2
K
d
N
exists. We may then interpret this product spacetime as the ground state of the
compactication of D = 11 supergravity on K
d
2
K
d
N
to a d
1
-dimensional
spacetime M
d
1
. As explained in more detail in Section 5.2, we may initially ignore
vacuum solutions that correspond to ground states of toroidal compactications of
D = 11 supergravity. In the following, we shall therefore assume that the Euclidean
signature Einstein manifolds are topologically distinct from torii. In fact, we shall
also assume that the Lorentzian signature Einstein manifold is topologically not a
torus, even after a periodic identication of the time coordinate
3
. The minimum
dimensionality of each submanifold is thus equal to two.
3
Vacuum solutions of D = 11 supergravity with M
d

= T
(1,d1)
(after a periodic identication
of the time coordinate) can be reinterpreted via Kaluza-Klein dimensional reduction as vacuum
solutions of the maximal massless Euclidean signature supergravity theory in D = 11 d [41, 73].
We hope to be able to report on the classication of vacua in Euclidean signature supergravities in
the future.
CHAPTER 6. VACUA IN D = 11 SUPERGRAVITY 77
Let us now return to the discussion of the Freund-Rubin ansatz in the case when
the 11-dimensional spacetime is the product of more than two submanifolds. For-
mally, the Freund-Rubin ansatz can be formulated in a way that is independent
of the decomposition of the 11-dimensional spacetime. In eect, it merely states
that the metric should be block-diagonal, and the 4-form eld strength should be
harmonic. We shall show next that the existence of a harmonic 4-form restricts the
possible decompositions of the 11-dimensional spacetime severely. At this stage, it is
necessary to make further assumptions about the properties of the Einstein subman-
ifolds. In particular, we restrict our attention to manifolds that are harmonically
trivial, in the sense that the set of harmonic forms on the d-manifold is generated
by the volume d-form and constants
4
.
The block-diagonal Freund-Rubin ansatz for the 11-dimensional metric is
ds
2
11
=
N

i=1
ds
2
d
i
(6.2.2)
where ds
2
d
i
is the Einstein metric on the i
th
submanifold. The volume forms on the
submanifolds are dened similarly to (6.1.7) and (6.1.8). Since the 11-dimensional
metric is block-diagonal, the harmonic forms on the 11-dimensional spacetime are
generated by the volume forms on the Einstein submanifolds. In other words, any
harmonic form can be written as a linear combination (with constant coecients)
of volume forms and wedge products of volume forms.
The Freund-Rubin ansatz for the eld strength requires that there exist a har-
monic 4-form. Now recall that since we may restrict our attention initially to
vacua that a priori cannot be toroidally reduced, each submanifolds is at least
4
This excludes, for example, Einstein-Kahler manifolds (other than the CP
1

= S
2
) and torii
from our considerations. The latter, however, have been excluded anyway.
CHAPTER 6. VACUA IN D = 11 SUPERGRAVITY 78
2-dimensional. This implies that the 11-dimensional product spacetime admits a
harmonic 4-form if, and only if, at least one submanifold is 4-dimensional and/or at
least two submanifolds are 2-dimensional. The decompositions of the 11-dimensional
spacetime that permit us to set up a Freund-Rubin ansatz are therefore given by
the solutions of the combinatorial problem
N

i=1
d
i
= 11, d
i
2, (d
i
= 2 and d
j
= 2, i ,= j) or d
i
= 4. (6.2.3)
The solutions of (6.2.3) are most easily enumerated by computer and are given in
Table 6.1.
N = 2 N = 3 N = 4 N = 5
4 + 7 2 + 2 + 7 2 + 2 + 2 + 5 2 + 2 + 2 + 2 + 3
2 + 4 + 5 2 + 2 + 3 + 4
3 + 4 + 4
Table 6.1: Spacetime decompositions for D = 11 vacuum ansatze.
For each spacetime decomposition in Table 6.1, the Freund-Rubin ansatz for the
eld strength is of the form
F
(4)
=

(4)
, F

(4)
= u

(4)
, (6.2.4)
where labels the harmonic volume 4-forms

(4)
and the u

are constants. The


corresponding energy-momentum tensor (6.1.4) is given by
S
MN
(F
(4)
) =
1
2

_
(F

(4)
)
2
MN

1
3
g
MN
(F

(4)
)
2
_
. (6.2.5)
Let us now substitute the ansatz for the metric (6.2.2) and the eld strength
(6.2.4) into the 11-dimensional eld equations. Since F
(4)
is harmonic, we nd that
CHAPTER 6. VACUA IN D = 11 SUPERGRAVITY 79
(6.1.2) and (6.1.3) reduces to a purely algebraic constraint equation for the expansion
coecients u

F
(4)
F
(4)
= 0

<
u

(4)
e

(4)
= 0. (6.2.6)
It is also not dicult to convince oneself that energy-momentum tensor (6.2.5)
is block-wise proportional to the metric. Therefore, the assumption that the sub-
manifolds are Einstein manifolds is consistent with the eld equation for the metric
(6.1.1). Indeed, the latter only serves to determine the scalar curvature of each sub-
manifold in terms of the coecients u

in the harmonic expansion of F


(4)
, according
to
(d
i
1)
i
=
1
2

u
2

(i)


1
3
_
(e

(4)
)
2
, (6.2.7)
where
i
is the scalar curvature of the i
th
submanifold,
(i)

= 1 if F

(4)
has a non-zero
VEV on that submanifold and
(i)

= 0 otherwise. Note that the scalar curvature of


the Minkowski signature submanifold is negative denite, since in that case
(i)

= 1
implies (e

(4)
)
2
= 1, while
(i)

= 0 implies (e

(4)
)
2
= +1. In contrast, the scalar
curvature of a Euclidean signature submanifold can be positive, negative or even
zero, depending on the values of the u

.
In summary, we have shown that the spacetime decompositions for product vac-
uum solutions of D = 11 supergravity that a priori cannot be toroidally reduced to
D < 11 are determined by requiring the existence of a Freund-Rubin ansatz. This
generalises the result that the Freund-Rubin ansatz singles out the 4 + 7 decom-
position if the 11-dimensional spacetime is the product of two submanifold. The
Freund-Rubin ansatz for each spacetime decomposition reduces the 11-dimensional
eld equation to a set of algebraic equations that constrain the coecients in the
CHAPTER 6. VACUA IN D = 11 SUPERGRAVITY 80
harmonic expansion of the eld strength, and determine the scalar curvatures of the
submanifolds. The solution of these algebraic equations denes the corresponding
product vacuum solution. In the next section, we shall discuss the Freund-Rubin
vacua in D = 11 associated with each of the spacetime decompositions in Table 6.1.
6.3 M
d
1
K
d
2
K
d
N
vacua
We have found it convenient to represent the Freund-Rubin ansatz associated with
one of the spacetime decomposition in Table 6.1 in a pictorial way. Each term in
the harmonic expansion (6.2.4) of the eld strength is represented by one of the
following basic tablatures
5
in Figures 6.1 and 6.2.
M
4

F

(4)
: u


K
4

F

(4)
: u


Figure 6.1: The F

(4)
= u

(4)
and F

(4)
= u

(4)
tablatures.
M
2
K
2

F

(4)
: u


K
2
K
2

F

(4)
: u


Figure 6.2: The F

(4)
= u

(2)

(2)
and F

(4)
= u

1
(2)

2
(2)
tablatures.
The tablature representing the entire Freund-Rubin ansatz is constructed from
these basic building blocks by including a separate row, labelled by u

, for each
term in the harmonic expansion F
(4)
=

(4)
. This tablature therefore looks
like a matrix, with columns labelled by the submanifolds and rows labelled by the
expansion coecients. The entries in this matrix are dots, which represent the
5
Here,
(d)
and
(d)
denote the volume forms of a d-dimensional submanifold with Lorentzian,
respectively Euclidean, signature.
CHAPTER 6. VACUA IN D = 11 SUPERGRAVITY 81
volume forms on the submanifolds. Each row contains either a single lled dot in
the column of a 4-manifold, or two lled dots in the columns of two 2-manifolds.
In either case, the lled dots indicate the directions along which the eld strength
has a non-zero, constant VEV, i.e. non-zero, constant tangent space components.
Notice also that it is easy to calculate the scalar curvature, given by (6.2.7), of
each submanifold from the information given the tablature. The sum over (at
xed i) in (6.2.7) corresponds to a sum over rows in the tablature (within the
i
th
column). A lled dot in the
th
row (representing the contribution of F

(4)
) is
equivalent to
(i)

= 1, while an empty dot is equivalent to


(i)

= 0. A corollary of
this observation is that if two columns in the tablature are identical, then by (6.2.7)
we have (d
i
1)
i
= (d
j
1)
j
, i.e. the constant of proportionality between the
Ricci tensor and the metric is the same for both submanifolds.
The original Freund-Rubin vacua presuppose a 4 + 7 decomposition of the 11-
dimensional spacetime. They are depicted in Figure 6.3. The product structure of
M
4
K
7
F
(4)
: u
M
7
K
4
F
(4)
: u
Figure 6.3: The M
4
K
7
and M
7
K
4
vacua.
the 11-dimensional spacetime is indicated in the top row of the tablature, where
it is understood that the 11-dimensional metric is block-diagonal. The dots in the
tablature represent the volume forms on the submanifolds. A lled dot indicates
that the eld strength has a VEV on the corresponding 4-manifold, with constant
CHAPTER 6. VACUA IN D = 11 SUPERGRAVITY 82
magnitude u. The scalar curvatures of the submanifolds in Figure 6.3 are given by
M
4
K
7
: M
7
K
4
:
(M
4
) =
1
9
u
2
, (M
7
) =
1
36
u
2
,
(K
7
) =
1
36
u
2
, (K
4
) =
1
9
u
2
,
(6.3.1)
so that the maximally symmetric vacua of D = 11 supergravity with two submani-
folds are AdS
4
S
7
and AdS
7
S
4
.
We continue to discuss the vacuum solutions of D = 11 supergravity involving
three submanifold. Consider rst the two Freund-Rubin ansatze associated with the
2 + 2 + 7 decompositions of the 11-dimensional spacetime. They are represented
by the tablatures in Figures 6.4. Notice that in Figure 6.4, the columns of the 2-
M
2
K
2
K
7
F
(4)
: u
M
7
K
2
K
2
F
(4)
: u
Figure 6.4: The 2 + 2 + 7 Freund-Rubin ansatze.
manifolds are identical. This implies that the constant of proportionality between
the Ricci tensor and the metric is the same for both 2-manifolds. Therefore, we see
that the 2+2+7 vacua can be understood as particular instances of the M
4/7
K
7/4
vacua for which the Einstein 4-manifold is itself the product of two Einstein 2-
manifolds.
Similar comments apply to the three Freund-Rubin ansatze associated with the
2 + 4 + 5 decomposition of the 11-dimensional spacetime. They are depicted in
Figure 6.5, whence it is apparent that the constants of proportionality between the
Ricci tensor and the metric is the same for the 2-manifold and the 5-manifold. All
the 2 + 4 + 5 vacua are therefore particular instances of the M
4/7
K
7/4
vacua for
CHAPTER 6. VACUA IN D = 11 SUPERGRAVITY 83
M
2
K
5
K
4
F
(4)
: u
M
4
K
2
K
5
F
(4)
: u
M
5
K
2
K
4
F
(4)
: u
Figure 6.5: The 2 + 4 + 5 Freund-Rubin ansatze.
which the Einstein 7-manifold is itself the product of an Einstein 2-manifold and an
Einstein 5-manifold.
In addition, there are two more Freund-Rubin ansatze that also involve three
submanifolds. They require a 3+4+4 decomposition of the 11-dimensional spacetime
and are represented by the tablatures in Figure 6.6. Here, we encounter for the
M
3
K
4
K
4
F
(4)
: u
1

u
2

M
4
K
4
K
3
F
(4)
: u
1

u
2

Figure 6.6: The 3 + 3 + 4 Freund-Rubin ansatze.
rst time a eld strength conguration for which the Chern-Simons contribution
to the gauge eld equation does not vanish trivially. Indeed, in the present case,
F
(4)
F
(4)
= 0 translates from (6.2.6) into u
1
u
2
= 0. Once we set one of the expansion
coecients to zero, which corresponds to replacing one of the lled dots in Figure
6.6 by an empty one, the 3 + 3 + 4 vacua are easily seen to be particular examples
of the M
4/7
K
7/4
vacua, for which the Einstein 7-manifold is itself the product of
an Einstein 3-manifold and an Einstein 4-manifold.
Let us now turn to the vacuum solutions of D = 11 supergravity involving four
submanifolds. The three Freund-Rubin ansatze associated with the 2 + 2 + 3 + 4
CHAPTER 6. VACUA IN D = 11 SUPERGRAVITY 84
decomposition of the 11-dimensional spacetime are depicted in Figure 6.7. These
M
2
K
2
K
3
K
4
F
(4)
: u
1

u
2

M
3
K
2
K
2
K
4
F
(4)
: u
1

u
2

M
4
K
2
K
2
K
3
F
(4)
: u
1

u
2

Figure 6.7: The 2 + 2 + 3 + 4 Freund-Rubin ansatze.
ansatze solve the 11-dimensional eld equations only if u
1
u
2
= 0. Once we set one
of the expansion coecients to zero, the 2 + 2 + 3 + 4 vacua can all be interpreted
as particular instances of the M
4/7
K
7/4
vacua for an appropriate choice of the
Einstein 4-manifold or the Einstein 7-manifold. In the interest of classifying distinct
families of vacuum solutions, we shall no longer distinguish between vacua that dier
merely by the choice of Einstein submanifolds.
The remaining Freund-Rubin ansatze involving four submanifolds both require
a 2 + 2 + 2 + 5 decomposition of the 11-dimensional spacetime. They yield new
vacuum solutions of D = 11 supergravity and are represented by the tablatures in
Figure 6.8.
M
2
K
2
K
2
K
5
F
(4)
: u
1

u
2

u
3

M
5
K
2
K
2
K
2
F
(4)
: u
1

u
2

u
3

Figure 6.8: The M
2
K
2
K
2
K
5
and M
5
K
2
K
2
K
2
vacua.
CHAPTER 6. VACUA IN D = 11 SUPERGRAVITY 85
The scalar curvatures of the submanifolds in Figure 6.8 are given by
M
2
K
2
K
2
K
5
: M
5
K
2
K
2
K
2
:
(M
2
) =
1
6
u
2
1

1
3
(u
2
2
+u
2
3
), (M
5
) =
1
24
(u
2
1
+u
2
2
+u
2
3
),
(K
2
) =
1
3
u
2
1
+
1
6
u
2
2

1
3
u
2
3
, (K
2
) =
1
3
(u
2
2
+u
2
3
)
1
6
u
2
1
,
(K
2
) =
1
3
u
2
1
+
1
6
u
2
3

1
3
u
2
2
, (K
2
) =
1
3
(u
2
1
+u
2
3
)
1
6
u
2
2
,
(K
5
) =
1
24
(u
2
2
+u
2
3
)
1
24
u
2
1
, (K
2
) =
1
3
(u
2
1
+u
2
2
)
1
6
u
2
3
.
(6.3.2)
The scalar curvatures of the Lorentzian signature submanifolds are negative def-
inite, implying that the maximally symmetric solution for M
d
is AdS
d
. In contrast,
the sign of the scalar curvature of each of the Euclidean signature submanifolds de-
pends on the relative values of the constants u

. The maximally symmetric solution


for K
d
is the d-sphere S
d
if (K
d
) > 0, the hyperbolic d-manifold H
d
if (K
d
) < 0,
or the d-torus T
d
if (K
d
) = 0.
The maximum number of submanifolds that can be involved in a Freund-Rubin
ansatz that a priori cannot be toroidally reduced to D < 11 is ve. The two
Freund-Rubin ansatze associated with the 2 + 2 + 2 + 2 + 3 decomposition of the
11-dimensional spacetime lead to new vacuum solutions of D = 11 supergravity.
They are represented in Figure 6.9, where the constants u

are constrained by
u
12
u
34
+u
13
u
24
+u
14
u
23
= 0 (6.3.3)
to ensure that the Chern-Simons contribution to the gauge eld equation vanishes.
The scalar curvatures of the submanifolds in Figure 6.9 are most conveniently
presented in term of the 3-vectors
x = (u
12
, u
13
, u
14
),
y = (u
34
, u
24
, u
23
),
(6.3.4)
CHAPTER 6. VACUA IN D = 11 SUPERGRAVITY 86
M
2
K
2
K
2
K
2
K
3
F
(4)
: u
12

u
13

u
14

u
23

u
24

u
34

M
3
K
2
K
2
K
2
K
2
F
(4)
: u
12

u
13

u
14

u
23

u
24

u
34

Figure 6.9: The M
2
K
2
K
2
K
2
K
3
and M
3
K
2
K
2
K
2
K
2
vacua.
which have to be orthogonal according to (6.3.3). We then nd that
M
2
K
2
K
2
K
2
K
3
: M
3
K
2
K
2
K
2
K
2
:
(M
2
) =
1
6
(2x
2
+y
2
), (M
3
) =
1
12
(x
2
+y
2
),
(K
2
) =
1
6
(2y
2
+x
2
3(x
2
1
+y
2
1
)), (K
2
) =
1
6
(2x
2
y
2
),
(K
2
) =
1
6
(2y
2
+x
2
3(x
2
2
+y
2
2
)), (K
2
) =
1
6
(2y
2
x
2
+ 3(x
2
1
y
2
1
)),
(K
2
) =
1
6
(2y
2
+x
2
3(x
2
3
+y
2
3
)), (K
2
) =
1
6
(2y
2
x
2
+ 3(x
2
2
y
2
2
)),
(K
3
) =
1
12
(x
2
y
2
), (K
2
) =
1
6
(2x
2
y
2
+ 3(x
2
3
y
2
3
)),
(6.3.5)
where x
2
= x x and y
2
= y y. As expected, the scalar curvatures of the Lorentzian
signature manifolds are negative denite, while the sign of the scalar curvature of
each of the Euclidean signature submanifolds is in general indenite.
6.4 The generalised Freund-Rubin ansatz in D < 11
The present approach to the classication of vacua in maximal, massless supergrav-
ities is based on the idea that it is sucient to classify, in every dimension D 11,
CHAPTER 6. VACUA IN D = 11 SUPERGRAVITY 87
only those vacuum solutions that cannot be further toroidally reduced to a lower
dimension. We have just presented a classication of those vacua in D = 11. They
have been constructed by considering all decompositions of the 11-dimensional space-
time that admit a Freund-Rubin ansatz. In this section, we adapt the Freund-Rubin
ansatz to the eld content of the maximal, massless supergravities in D < 11.
Recall that the bosonic sector of the maximal, massless supergravities in D <
11 consists of the metric, the scalar sector and a number of (p 1)-form gauge
potential of various degrees, with (p 2)-form eld strengths. The scalar sector
may be divided further into dilatonic scalars and axionic scalars, where the latter
are distinguished from the former by the fact that they can be regarded as 0-form
potentials for 1-form eld strengths. Table 6.2 lists the degrees of the eld strengths
in each maximal, massless supergravity theory in D < 11. Note that we have
D = 10 IIA p = 2, 3, 4
D = 10 IIB p = 3, 5
D = 8, 9 p = 2, 3, 4
D = 6, 7 p = 2, 3
D = 4, 5 p = 2
D = 2, 3
Table 6.2: p-form eld strengths in D < 11 supergravities.
assumed that higher degree eld strengths in D < 7 have been dualised to give lower
degree eld strengths whenever possible. After all such dualisations, the bosonic
sector of the maximal, massless supergravities in D = 2, 3 consists only of the
metric and the scalar elds.
CHAPTER 6. VACUA IN D = 11 SUPERGRAVITY 88
The vacuum solutions of the lower dimensional supergravities are constructed
by starting o with a D-dimensional spacetime that is topologically the product
manifold M
d
1
K
d
2
K
d
N
, where

N
i=1
d
i
= D. The submanifolds admit an
Einstein metric and are assumed to be harmonically trivial, in the sense described
above. As we are only seeking vacuum solutions that a priori cannot be toroidally
reduced to a lower dimension, we may without loss of generality impose that d
i
2.
We take the Freund-Rubin ansatz to be that the D-dimensional metric is block-
diagonal, the scalar elds are constants and the p-form eld strengths are harmonic
6
.
The existence of at least one harmonic volume p-form, where the possible values for
p are given in Table 6.2, restricts the spacetime decompositions that admit a Freund-
Rubin ansatz to the solutions of a combinatorial problem analogous to (6.1).
The Freund-Rubin ansatz for each allowed spacetime decomposition is then con-
6
This is not the only possible adaptation of the Freund-Rubin ansatz to the eld content of the
lower dimensional supergravities. The ambiguity arises in the ansatz for the axionic scalars. We
have chosen to treat them on an equal footing with the dilatonic scalars, by requiring that the entire
scalar sector be constant. However, since the axionic scalars are 0-form potentials for 1-form eld
strengths, one could impose the weaker condition that these 1-form eld strengths be harmonic. In
other words, there exists a choice between setting the 1-form eld strengths for the axionic scalars
to zero, or merely requiring them to be covariantly constant, but possibly non-zero. As will be
shown elsewhere in more detail, the vacuum solution of (D+ 1)-dimensional massless supergravity
with a harmonic 1-form eld strength are naturally interpreted as vacuum solutions of the massive
supergravity theory in D dimensions that is obtained by the Scherk-Schwarz dimensional reduction
[74, 75] of the (D + 1)-dimensional theory on a circle. On the level of the Freund-Rubin ansatz
this is easily seen by noting that the existence of a harmonic 1-form forces the (D+1)-dimensional
spacetime to be of the form M
D
S
1
. The VEV of the 1-form eld strength on the circle plays the
role of the mass parameter in the D-dimensional theory.
CHAPTER 6. VACUA IN D = 11 SUPERGRAVITY 89
sidered in turn. The p-form eld strengths are expanded in the basis of the harmonic
volume p-forms. As in D = 11, the energy-momentum tensor for the eld strengths
is block-wise proportional to the metric. The eld equation for the metric thus only
serves to determine the scalar curvatures of the submanifolds in terms of the coef-
cients in the harmonic expansion of the eld strengths. The equations of motion
for the gauge elds and the Bianchi identities reduce to the requirement that the
various Chern-Simons and Kaluza-Klein contributions to the eld equations vanish.
This requirement translates into a set of algebraic constraint equations that the
expansion coecients have to satisfy. Additional constraint equations come from
requiring that the truncation of the dilatonic and axionic scalars be consistent with
their equations of motion. This set of constraint equations is then solved to give
the constant scalar, vacuum solution associated with the spacetime decomposition
under consideration.
Chapter 7
Vacua in D = 10 IIA and IIB
supergravity
7.1 IIA supergravity
The eld content of the bosonic sector of IIA supergravity consists of the metric,
the dilaton , and the p-form potentials A
(3)
, A
(2)
and /
(1)
with associated eld
strengths F
(4)
, F
(3)
and T
(2)
. The Cremmer-Julia (CJ) global symmetry group of
IIA supergravity is O(1, 1)

= IR. It leaves the metric invariant, shifts the dilaton by


a constant and rescales the p-form potentials accordingly.
The eld equations of IIA supergravity are given in Appendix B.1. After trun-
cating the dilaton, the equations of motion for the gauge elds are
d F
(4)
= F
(4)
F
(3)
, dF
(4)
= F
(3)
T
(2)
,
d F
(3)
= F
(4)
T
(2)
+
1
2
F
(4)
F
(4)
, dF
(3)
= 0,
d T
(2)
= F
(4)
F
(3)
, dT
(2)
= 0.
(7.1.1)
90
CHAPTER 7. VACUA IN D = 10 IIA AND IIB SUPERGRAVITY 91
The eld equation for the metric can be written as
1
MN
=
1
2
_
(F
(4)
)
2
MN

3
8
g
MN
(F
(4)
)
2
_
(7.1.2)
+
1
2
_
(F
(3)
)
2
MN

1
4
g
MN
(F
(4)
)
2
_
+
1
2
_
(T
(2)
)
2
MN

1
8
g
MN
(T
(2)
)
2
_
.
The consistency of the truncation of the dilaton requires in addition that
0 =
1
2
(F
(4)
)
2
(F
(3)
)
2
+
3
2
(T
(2)
)
2
. (7.1.3)
For notational convenience, we have chosen to set the dilaton to zero. Since
the value of the dilaton act as the modulus of the constant scalar vacua of IIA
supergravity, we have chosen to select the preferred point in the moduli space where
the dilaton vanishes. However, this choice does not impair the generality of our
discussion, as we can use the CJ symmetry group to map the zero dilaton solution
to any other point in the moduli space.
Consider next the combinatorial problem whose solution determines the decom-
positions of the 10-dimensional spacetime that admit Freund-Rubin ansatze. In the
context of IIA supergravity, it can be stated as follows:
N

i=1
d
i
= 10, d
i
2, d
i
= 2, or d
i
= 3, or d
i
= 4, (7.1.4)
where, as explained in the previous chapter, the restrictions on the dimensionali-
ties of the submanifolds are imposed to guarantee the existence of harmonic volume
p-forms, with p = 2, 3, 4, and to avoid having circular directions that can be com-
pactied.
We have considered Freund-Rubin anatze based on the decompositions of the
10-dimensional spacetime given in Table 7.1. For each of these decompositions, we
have expanded eld strengths F
(4)
, F
(3)
and T
(2)
in the basis of harmonic volume
CHAPTER 7. VACUA IN D = 10 IIA AND IIB SUPERGRAVITY 92
N = 2 N = 3 N = 4 N = 5
2 + 8 2 + 2 + 6 2 + 2 + 2 + 4 2 + 2 + 2 + 2 + 2
3 + 7 2 + 3 + 5 2 + 2 + 3 + 3
4 + 6 2 + 4 + 4
3 + 3 + 4
Table 7.1: Spacetime decompositions for IIA vacuum ansatze.
p-forms and then tried to solve the truncated eld equations (7.1.1) and (7.1.3) for
the expansion coecients. Finally, the scalar curvatures of the Einstein submani-
folds have been determined using the metric eld equation (7.1.2). In the following
subsections we present and discuss those constant scalar vacuum solutions of IIA
supergravity that a priori cannot be toroidally reduced to D < 10.
M
2
K
4
K
4
and M
4
K
4
K
2
vacua
We have found two vacuum solutions of IIA supergravity with three submanifolds.
They require a 2 + 4 + 4 decomposition of the 10-dimensional spacetime and are
represented by the tablatures in Figure 7.1. The constants u and v are related by
M
2
K
4
K
4
F
(4)
: u
T
(2)
: v
M
4
K
4
K
2
F
(4)
: u
T
(2)
: v
Figure 7.1: The M
2
K
4
K
4
and M
4
K
4
K
2
IIA vacua.
u
2
= 3v
2
to ensure the consistency of the truncation of the dilaton. The scalar
CHAPTER 7. VACUA IN D = 10 IIA AND IIB SUPERGRAVITY 93
curvatures of the submanifolds in Figure 7.1 are given by
M
2
K
4
K
4
: M
4
K
4
K
2
:
(M
2
) =
1
2
v
2

1
6
u
2
, (M
4
) =
1
9
u
2
,
(K
4
) =
1
18
u
2
, (K
4
) =
1
18
u
2
,
(K
4
) =
1
9
u
2
, (K
2
) =
1
2
v
2
+
1
6
u
2
,
(7.1.5)
The maximally symmetric vacua of IIA supergravity with three submanifold are
therefore AdS
2
H
4
S
4
and AdS
4
S
4
S
2
. The latter solution has been considered
originally in the context of the spontaneous compactication of IIA supergravity to
D = 4 [76].
By oxidising these IIA vacua to D = 11, it is not dicult to establish that they
can be interpreted as so-called Hopf reductions of certain 11-dimensional vacuum
solutions that fall within the family of 4 + 7 vacua depicted in Figure 6.3. Hopf
reductions can be understood as Kaluza-Klein reductions of U(1) brations, where
the reduction coordinate is the coordinate of the U(1) bre. They are described in
more detail in Appendix C.1.
In particular, the AdS
2
K
4
K
4
vacuum oxidises to a M
7
K
4
vacuum for which
the 7-manifold is AdS
3
K
4
and the AdS
3
is parameterised as a U(1) bration over
an AdS
2
base. A simple way to conrm this identication is to use equation (C.1.6),
which relates the scalar curvature of the base manifold to the scalar curvature of the
U(1) bration. We see from (7.1.5) that the scalar curvature of the AdS
2
is given
precisely by (C.1.6), where
1
12
u
2
is the scalar curvature of the AdS
3
.
The tablature notation allows for an intuitive visualisation of the Hopf reduction
or oxidation procedure. The M
3
K
4
K
4
vacuum solution in D = 11 is represented
in Figure 6.6, with e.g. u
1
= 0. To set ourselves up for the Hopf reduction, we
CHAPTER 7. VACUA IN D = 10 IIA AND IIB SUPERGRAVITY 94
consider M
3
= AdS
3
, parameterised as a U(1) bration over AdS
2
, and then Hopf
reduce according to the scheme depicted in Figure 7.2.
D = 11
Hopf
D = 10
AdS
3
K
4
K
4
F
(4)
: u
AdS
2
K
4
K
4
F
(4)
: u
T
(2)
: v
Figure 7.2: Hopf reduction of AdS
3
K
4
K
4
to AdS
2
K
4
K
4
.
Similarly, the M
4
K
4
S
2
vacuum oxidises to a M
4
K
7
vacuum for which the
7-manifold is K
4
S
3
and the S
3
is represented as a U(1) bration over S
2
. This
Hopf oxidation is discussed in [77], and can be conrmed by noting that the scalar
curvature of the S
2
in (7.1.5) ts equation (C.1.6), where =
1
12
u
2
is the scalar
curvature of the S
3
. The details of this Hopf reduction are given schematically in
Figure 7.3.
D = 11
Hopf
D = 10
M
4
K
4
S
3
F
(4)
: u
M
4
K
4
S
2
F
(4)
: u
T
(2)
: v
Figure 7.3: Hopf reduction of M
4
K
4
S
3
to M
4
K
4
S
2
.
CHAPTER 7. VACUA IN D = 10 IIA AND IIB SUPERGRAVITY 95
M
2
K
2
K
2
K
4
and M
4
K
2
K
2
K
2
vacua
We have found two vacuum solution of IIA supergravity with four submanifolds.
They require a 2 +2 +2 +4 decomposition of the 10-dimensional spacetime and are
represented by the tablatures in Figure 7.4. The consistency of the truncation of
M
2
K
2
K
2
K
4
F
(4)
: u
T
(2)
: v
1

v
2

v
3

M
4
K
2
K
2
K
2
F
(4)
: u
T
(2)
: v
1

v
2

v
3

Figure 7.4: The M
2
K
2
K
2
K
4
and M
4
K
2
K
2
K
2
IIA vacua.
the dilaton requires that the constants u and v

satisfy
M
2
K
2
K
2
K
4
: M
4
K
2
K
2
K
2
:
u
2
= 3v
2
1
3(v
2
2
+v
2
3
), u
2
= 3(v
2
1
+v
2
2
+v
2
3
).
(7.1.6)
The scalar curvatures of the submanifolds in Figure 7.4 are given by
M
2
K
2
K
2
K
4
: M
4
K
2
K
2
K
2
:
(M
2
) =
1
2
v
2
1

1
6
u
2
, (M
4
) =
1
9
u
2
,
(K
2
) =
1
2
v
2
2

1
6
u
2
, (K
2
) =
1
2
v
2
1
+
1
6
u
2
,
(K
2
) =
1
2
v
2
3

1
6
u
2
(K
2
) =
1
2
v
2
2
+
1
6
u
2
(K
4
) =
1
9
u
2
, (K
2
) =
1
2
v
2
3
+
1
6
u
2
.
(7.1.7)
The maximally symmetric IIA vacua with four submanifold are therefore AdS
2
K
2

K
2
S
4
, where K
2
is the determined according to Table C.1, and AdS
4
S
2
S
2
S
2
.
The latter solution is known from the study of spontaneous compactications of IIA
supergravity to D = 4 [76].
From the 11-dimensional perspective, these IIA vacua are interpreted as Hopf
reductions of certain vacuum solutions that fall within the family of 4 + 7 vacua
CHAPTER 7. VACUA IN D = 10 IIA AND IIB SUPERGRAVITY 96
depicted in Figure 6.3. In particular, the oxidation of the AdS
2
K
2
K
2
K
4
vacuum yields a M
7
K
4
vacuum for which the 7-manifold is the U(1) bration

Q
7
(v
1
, v
2
, v
3
) over the base AdS
2
K
2
K
2
, in the notation of Appendix C.1. The
oxidation of the M
4
S
2
S
2
S
2
vacuum was considered in [77] and shown to
yield a M
4
K
7
vacuum, where K
7
is the U(1) bration Q
7
(v
1
, v
2
, v
3
) over the base
S
2
S
2
S
2
. These identications are conrmed by noting that in both cases the
scalar curvatures
i
of the 2-manifolds in (7.1.7) are related to the scalar curvature
of the U(1) bration by equation (C.1.6), with =
1
36
u
2
or
1
36
u
2
the scalar
curvature of the

Q
7
(v
1
, v
2
, v
3
), respectively the Q
7
(v
1
, v
2
, v
3
), manifold. The details
of these Hopf reductions are given schematically in Figures 7.5 and 7.6.
D = 11
Hopf
D = 10

Q
7
K
4
F
(4)
: u
AdS
2
K
2
K
2
K
4
F
(4)
: u
T
(2)
: v
1

v
2

v
3

Figure 7.5: Hopf reduction of

Q
7
(v
1
, v
2
, v
3
) K
4
to AdS
2
K
2
K
2
K
4
.
D = 11
Hopf
D = 10
M
4
Q
7
F
(4)
: u
M
4
S
2
S
2
S
2
F
(4)
: u
T
(2)
: v
1

v
2

v
3

Figure 7.6: Hopf reduction of M
4
Q
7
(v
1
, v
2
, v
3
) to M
4
S
2
S
2
S
2
.
CHAPTER 7. VACUA IN D = 10 IIA AND IIB SUPERGRAVITY 97
M
2
K
2
K
2
K
2
K
2
vacua
The vacuum solutions of IIA supergravity with ve submanifolds require a 2 +
2 + 2 + 2 + 2 decomposition of the 10-dimensional spacetime. To determine the
M
2
K
2
K
2
K
2
K
2
vacua, we have used the Freund-Rubin ansatz depicted
in Figure 7.7. The constraint equations for the coecients u

and v

follow from
imposing
0 = F
(4)
T
(2)
+
1
2
F
(4)
F
(4)
,
0 =
1
2
(F
(4)
)
2
+
3
2
(T
(2)
)
2
,
(7.1.8)
in order to satisfy the gauge eld equations, and to ensure the consistency of the
truncation of the dilaton.
M
2
K
2
K
2
K
2
K
2
F
(4)
: u
12

u
13

u
14

u
15

u
23

u
24

u
25

u
34

u
35

u
45

T
(2)
: v
1

v
2

v
3

v
4

v
5

Figure 7.7: The M
2
K
2
K
2
K
2
K
2
vacuum ansatz.
In the present case, the complexity of the constraint equations that are implied by
(7.1.8) makes it dicult to obtain the general solution for the expansion coecients
u

and v

, and hence the general IIA vacuum solution associated with the 2 +
2 + 2 + 2 + 2 decomposition of the 10-dimensional spacetime. Here, we content
ourselves with discussing those 2+2+2+2+2 IIA vacua that can be obtained from
various D = 11 vacuum solutions via Hopf reduction. We are not aware of any other
CHAPTER 7. VACUA IN D = 10 IIA AND IIB SUPERGRAVITY 98
IIA vacua that fall within the Freund-Rubin ansatz in Figure 7.7, but we cannot
exclude their existence. In the following, we shall be using the notation established
in Appendix C.1.
The vacuum solutions of D = 11 supergravity that lead to IIA vacua with ve
submanifolds fall within the family of the 2+2+2+5 vacua, or the 2+2+2+2+3
vacua. For the purpose of Hopf reducing these vacua to D = 10, we assume that
the odd-dimensional Einstein manifold is given by a U(1) bration over a product
of Einstein 2-manifolds. The Hopf reduction along the U(1) bre direction results
in replacing the U(1) bration by its base manifold and a Kaluza-Klein gauge eld
with non-zero winding numbers over each 2-manifold in the base.
Consider rst the 2+2+2+2+3 vacuum solutions of D = 11 supergravity, which
are depicted in Figure 6.9. We assume that the 3-manifold is given by a U(1) bra-
tion over an Einstein 2-manifold. For the Lorentzian signature 3-manifold, which has
negative denite curvature, this implies that M
3
=

Q
3
(v), where

Q
3
(v) denotes the
parameterisation of AdS
3
as a U(1) bration over AdS
2
. For the Euclidean signa-
ture 3-manifold, which here is assumed to have positive curvature, this implies that
K
3
= Q
3
(v), where Q
3
(v) denotes the parameterisation of S
3
as a U(1) bration
over S
2
. The Hopf reduction of these D = 11 vacua along the U(1) bre direction
generates the 2 +2 +2 +2 +2 vacuum solutions of IIA supergravity represented by
the tablatures in Figure 7.8, where the constants u

and v are constrained by


M
2
K
2
K
2
(top) : M
2
K
2
K
2
(bottom) :
x y = 0, x y = 0,
x
2
= y
2
+ 3v
2
, 3v
2
= x
2
+y
2
.
(7.1.9)
The 3-vectors x and y have been dened previously in (6.3.4) and x
2
= x x and
CHAPTER 7. VACUA IN D = 10 IIA AND IIB SUPERGRAVITY 99
y
2
= y y.
M
2
K
2
K
2
K
2
K
2
F
(4)
: u
12

u
13

u
14

u
23

u
24

u
34

T
(2)
v
M
2
K
2
K
2
K
2
K
2
F
(4)
: u
12

u
13

u
14

u
23

u
24

u
34

T
(2)
: v
Figure 7.8: M
2
K
2
K
2
K
2
K
2
IIA vacua.
The scalar curvatures of the submanifolds in Figure 7.8 are given by
M
2
K
2
K
2
(top) : M
2
K
2
K
2
(bottom) :
(M
2
) =
1
6
(2x
2
+y
2
), (M
2
) =
1
6
(x
2
+y
2
)
1
2
v
2
,
(K
2
) =
1
6
(2y
2
+x
2
3(x
2
1
+y
2
1
)), (K
2
) =
1
6
(2x
2
y
2
),
(K
2
) =
1
6
(2y
2
+x
2
3(x
2
2
+y
2
2
)), (K
2
) =
1
6
(2y
2
x
2
+ 3(x
2
1
y
2
1
)),
(K
2
) =
1
6
(2y
2
+x
2
3(x
2
3
+y
2
3
)), (K
2
) =
1
6
(2y
2
x
2
+ 3(x
2
2
y
2
2
)),
(K
2
) =
1
6
(x
2
y
2
) +
1
2
v
2
, (K
2
) =
1
6
(2y
2
x
2
+ 3(x
2
3
y
2
3
)).
(7.1.10)
Consider now the 2+2+2+5 vacuum solutions of D = 11 supergravity, which are
depicted in Figure 6.8. We assume that the 5-manifold is given by a U(1) bration
over the product of two Einstein 2-manifold. For the Lorentzian signature 5-manifold
(which has negative curvature), this implies that M
5
=

Q
5
(v
1
, v
2
), where

Q
5
(v
1
, v
2
)
is a U(1) bration over AdS
2
K
2
. For the Euclidean signature 5-manifold, which
CHAPTER 7. VACUA IN D = 10 IIA AND IIB SUPERGRAVITY 100
here is assumed to have positive curvature, this implies that K
5
= Q
5
(v
1
, v
2
), where
Q
5
(v
1
, v
2
) is a U(1) bration over S
2
S
2
. The Hopf reduction of these D = 11
vacua along the U(1) bre direction generates the 2+2+2+2+2 vacuum solutions
of IIA supergravity represented by the tablatures in Figure 7.9, where the constants
u

and v
i
are constrained by
M
2
K
2
K
2
(top) : M
2
K
2
K
2
(bottom) :
u
2
2
+u
2
3
= u
2
1
+ 3v
2
1
+ 3v
2
2
, 3v
2
1
= 3v
2
2
+u
2
1
+u
2
2
+u
2
3
.
(7.1.11)
M
2
K
2
K
2
K
2
K
2
F
(4)
: u
1

u
2

u
3

T
(2)
v
1

v
2

M
2
K
2
K
2
K
2
K
2
F
(4)
: u
1

u
2

u
3

T
(2)
: v
1

v
2

Figure 7.9: M
2
K
2
K
2
K
2
K
2
IIA vacua.
The scalar curvatures of the submanifolds in Figure 7.9 are given by
M
2
K
2
K
2
(top) : M
2
K
2
K
2
(bottom) :
(M
2
) =
1
6
u
2
1

1
3
(u
2
2
+u
2
3
), (M
2
) =
1
6
(u
2
1
+u
2
2
+u
2
3
)
1
2
v
2
1
,
(K
2
) =
1
3
u
2
1
+
1
6
u
2
2

1
3
u
2
3
, (K
2
) =
1
6
(u
2
1
+u
2
2
+u
2
3
) +
1
2
v
2
2
,
(K
2
) =
1
3
u
2
1
+
1
6
u
2
3

1
3
u
2
2
, (K
2
) =
1
3
(u
2
2
+u
2
3
)
1
6
u
2
1
,
(K
2
) =
1
6
(u
2
2
+u
2
3
)
1
6
u
2
1
+
1
2
v
2
1
, (K
2
) =
1
3
(u
2
1
+u
2
3
)
1
6
u
2
2
,
(K
2
) =
1
6
(u
2
2
+u
2
3
)
1
6
u
2
1
+
1
2
v
2
2
, (K
2
) =
1
3
(u
2
1
+u
2
2
)
1
6
u
2
3
.
(7.1.12)
CHAPTER 7. VACUA IN D = 10 IIA AND IIB SUPERGRAVITY 101
We conclude this section by discussing the 2 + 2 + 2 + 2 + 2 vacuum solution of
IIA supergravity depicted in Figure 7.10, where the constants v
i
are satisfy
v
2
1
= v
2
2
+v
2
3
+v
2
4
+v
2
5
. (7.1.13)
M
2
K
2
K
2
K
2
K
2
T
(2)
: v
1

v
2

v
3

v
4

v
5

Figure 7.10: AdS
2
S
2
S
2
S
2
S
2
IIA vacuum.
The scalar curvatures of the submanifolds in Figure 7.10 are given by
(M
2
) =
1
2
v
2
1
,
(K
2
) =
1
2
v
2
i
, i = 2, . . . , 5.
(7.1.14)
Therefore, we can identify this particular IIA vacuum with AdS
2
S
2
S
2
S
2
S
2
.
Since this IIA vacuum involves only the metric and the Kaluza-Klein eld strength
T
(2)
, it oxidises to a purely gravitational solution of the 11-dimensional eld equa-
tions with Ricci at metric. More precisely, the 11-dimensional spacetime can be
interpreted as a U(1) bration over AdS
2
S
2
S
2
S
2
S
2
. As shown in Appendix
C.1, such Ricci at U(1) brations over AdS
2
S
2
S
2
exist quite generally
in (2n + 1 5)-dimensional spacetimes.
7.2 IIB supergravity
The eld content of the bosonic sector of IIB supergravity comprises the metric, a
dilatonic scalar , an axionic scalar , two 2-form gauge potentials A
NS
(2)
and A
R
(2)
CHAPTER 7. VACUA IN D = 10 IIA AND IIB SUPERGRAVITY 102
and with associated eld strengths F
NS
(3)
and F
R
(3)
and a 4-form gauge potential B
(4)
whose eld strength H
(5)
is self-dual. Type IIB supergravity has a global SL(2, IR)
symmetry under which the 2-form potentials transform as a doublet, the 4-form
potential is a singlet and the scalars parameterise the coset space O(2, IR)/SL(2, IR).
The equations of motion of the IIB supergravity elds are given in Appendix
B.2. After truncating the scalar sector, the gauge eld equations can be written as
H
(5)
= H
(5)
, dH
(5)
= F
NS
(3)
F
R
(3)
,
d F
NS
(3)
= H
(5)
F
R
(3)
, dF
NS
(3)
= 0,
d F
R
(3)
= H
(5)
F
NS
(3)
, dF
R
(3)
= 0.
(7.2.1)
The eld equation for the metric is given by
1
MN
=
1
4
(H
(5)
)
2
MN
+
1
2
_
(F
NS
(3)
)
2
MN

1
4
g
MN
(F
NS
(3)
)
2
_
+
1
2
_
(F
R
(3)
)
2
MN

1
4
g
MN
(F
R
(3)
)
2
_
.
(7.2.2)
The consistency of the truncation of the scalar sector requires in addition that
0 = F
R
(3)
F
NS
(3)
,
0 = (F
NS
(3)
)
2
(F
R
(3)
)
2
.
(7.2.3)
For notational convenience, we have again chosen to set the scalar elds to zero.
The values of the scalar elds are the moduli of the constant scalar vacua that we
are seeking. From this point of view, we have therefore chosen to place ourselves at
the distinguished point in the moduli space of constant scalar vacua where all the
scalars vanish. Other points in this moduli space can be reached by acting on the
zero-scalar solution with an element of the SL(2, IR) global symmetry group.
The classical symmetry group of a constant scalar solution of IIB supergravity
is the SO(2, IR) subgroup of SL(2, IR). It is interpreted as the stability group of
CHAPTER 7. VACUA IN D = 10 IIA AND IIB SUPERGRAVITY 103
the corresponding point in the moduli space, as it does not change the values of the
scalar elds. In general, the elements of this stability group are moduli-dependent.
However, for the particular choice = = 0, they are given by the usual SO(2, IR)
matrices. A particularly important subgroup of the SO(2, IR) stability group is the
discrete group S
2
consisting of the identity and the element represented by the matrix
(
0 1
1 0
). For vanishing scalars, the non-trivial action of this subgroup is simply to
interchange the 3-form eld strengths according to
F
NS
(3)
F
R
(3)
, F
R
(3)
F
NS
(3)
. (7.2.4)
The truncated eld equations (7.2.1) and (7.2.3) are easily seen to be left invariant
by this interchange.
This discrete S
2
symmetry group has been identied in Ref. [78] as the Weyl
group of the SL(2, IR) symmetry group of IIB supergravity. There, it was also
shown that the p-brane solutions of IIB supergravity with a xed number of charges
form irreducible multiplets under the action of this Weyl group. We have found
that the constant scalar vacuum solutions of IIB supergravity also naturally assemble
themselves into irreducible Weyl multiplets. It turns out, however, that all the vacua
that transform non-trivially under the S
2
Weyl group can be toroidally reduced to
D < 10, and hence we shall not discuss them here. Nevertheless, a similar situation
obtains in D 9. We shall show below that the constant scalar vacua of the lower
dimensional supergravities naturally form irreducible multiplets under the action of
the Weyl subgroup of the global CJ symmetry group.
Consider next the combinatorial problem whose solution determines the decom-
positions of the 10-dimensional spacetime that admit Freund-Rubin ansatze. In the
CHAPTER 7. VACUA IN D = 10 IIA AND IIB SUPERGRAVITY 104
context of IIB supergravity, it can be stated as follows:
N

i=1
d
i
= 10, d
i
2, d
i
= 3, or d
i
= 5, (7.2.5)
where the restrictions on the dimensionalities of the submanifolds are imposed to
guarantee the existence of harmonic volume p-forms, with p = 3, 5, and to avoid
having circular directions that can be compactied.
We have considered Freund-Rubin anatze based on the decompositions of the
10-dimensional spacetime given in Table 7.2. For each of these decompositions, we
N = 2 N = 3 N = 4
3 + 7 2 + 3 + 5 2 + 2 + 3 + 3
5 + 5 3 + 3 + 4
Table 7.2: Spacetime decompositions for IIB vacuum ansatze.
have expanded eld strengths F
NS
(3)
, F
R
(3)
and H
(5)
in the basis of harmonic volume p-
forms and then tried to solve the truncated eld equations (7.2.1) and (7.2.3) for the
expansion coecients. Finally, the scalar curvatures of the submanifolds have been
determined using the Einstein equation (7.2.2). It turns out that constant scalar
vacua of IIB supergravity that a priori cannot be toroidally reduced to D < 10
involve only the 5-form eld strength, and hence are singlets of the Weyl group S
2
.
M
5
K
5
vacua
We start o by discussing the IIB vacuum associated with the 5 + 5 decomposition
of the 10-dimensional spacetime. It is represented by the tablature in Figure 7.11.
The self-duality constraint reduces the number of independent coecients in
CHAPTER 7. VACUA IN D = 10 IIA AND IIB SUPERGRAVITY 105
M
5
K
5
H
(5)
: u
u
Figure 7.11: The M
5
K
5
IIB vacuum.
the harmonic expansion of H
(5)
from two to one. The scalar curvatures of the
submanifolds in Figure 7.11 are given by
(M
5
) =
1
16
u
2
, (K
5
) =
1
16
u
2
. (7.2.6)
The corresponding maximally symmetric vacuum solution of IIB supergravity is
therefore AdS
5
S
5
, which has been found originally in Ref. [79].
M
2
K
2
K
3
K
3
and M
3
K
3
K
2
K
2
vacua
The other vacuum solutions of IIB supergravity that involve only the 5-form eld
strength require a 2+2+3+3 decomposition of the 10-dimensional spacetime. They
are represented by the tablatures in Figure 7.12.
M
2
K
2
K
3
K
3
H
(5)
: u
1

u
1

u
2

u
2

M
3
K
3
K
2
K
2
H
(5)
: u
1

u
1

u
2

u
2

Figure 7.12: The M
2
K
2
K
3
K
3
and M
3
K
3
K
2
K
2
IIB vacua.
The number of independent coecients u

in the harmonic expansion of H


(5)
is
reduced from four to two by the self-duality constraint. The scalar curvatures of the
CHAPTER 7. VACUA IN D = 10 IIA AND IIB SUPERGRAVITY 106
submanifolds in Figure 7.12 are given by
M
2
K
2
K
3
K
3
: M
3
K
3
K
2
K
2
:
(M
2
) =
1
2
(u
2
1
+u
2
2
), (M
3
) =
1
2
(u
2
1
+u
2
2
),
(K
2
) =
1
2
(u
2
1
+u
2
2
), (K
3
) =
1
2
(u
2
1
+u
2
2
),
(K
3
) =
1
2
u
2
1
+
1
2
u
2
2
(K
2
) =
1
2
u
2
1
+
1
2
u
2
2
(K
3
) =
1
2
u
2
1

1
2
u
2
2
, (K
2
) =
1
2
u
2
1

1
2
u
2
2
.
(7.2.7)
The maximally symmetric vacuum solutions of IIB supergravity with four subman-
ifolds are therefore AdS
2
S
2
S
3
H
3
and AdS
3
S
3
S
2
H
2
, where we have
assumed that u
2
> u
1
.
Chapter 8
Vacua in D 9 supergravities
8.1 CJ global symmetries of maximal, massless super-
gravities in D 9
The eld content of the bosonic sector of the maximal, massless supergravity theory
in D 9 consists of the metric, a (11 D)-component vector

of dilatonic scalars,
the axionic scalars A
a
(0)
n
and A
(0)mnp
and a set of p-form gauge potential A
(3)
, A
(2)m
,
A
(1)mn
and /
a
(1)
with eld strengths F
(4)
, F
(3)a
, F
(2)ab
and T
a
(2)
. These D-dimensional
elds arise naturally out of the dimensional reduction of the 11-dimensional metric
and 3-form potential. Indeed, the indices that label the D-dimensional elds range
over 1, , (11 D) and refer to the internal toroidal directions that have been
compactied in the reduction from 11 dimensions to D dimensions, where 1 refers
to the reduction step from D = 11 to D = 10 etc. See Appendix B.1 for more
details.
In D 8, it is possible to dualise the p-form gauge potentials with p D/2 1
107
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 108
and replace them with p-form potentials, where p = D p 2. The eld strength
associated with the dual potential is simply the Hodge dual of the eld strength
associated with the original potential, up to exponentials of dilatons. The main
benet of performing these dualisations is that they enlarge the global symmetry
group of the D-dimensional supergravity theory. In particular, it has been shown
[80] that the full CJ global symmetry group in D dimension emerges only after all
possible dualisations have been performed. In D 8, we shall therefore dene the
eld strengths G
(D4)
, G
a
(D3)
, G
ab
(D2)
and (
(D2)a
for the dual potentials by
D = 8 : G
(4)
= e
a

F
(4)
,
D = 7 : G
(3)
= e
a

F
(4)
,
D = 6 : G
(2)
= e
a

F
(4)
,
D = 5 : G
(1)
= e
a

F
(4)
, G
a
(2)
= e
a
a

F
a
(3)
,
D = 4 : G
a
(1)
= e
a
a

F
a
(3)
,
D = 3 : G
ab
(1)
= e
a
ab

F
ab
(2)
, (
(1)a
= e

b
a

T
(2)a
.
(8.1.1)
We emphasise that in the process of dualisation the sign of the dilaton vector
is reversed, in the sense that if c is the dilaton vector associated with the original
eld strength F
(n)
, then

c = c is the dilaton vector associated with the dual eld
strength G
(Dn)
= F
(n)
. One way of seeing this is to recall that the dualisation
process interchanges the role of the equations of motion and the Bianchi identity.
The eld strength F
(n)
is given in terms of its potential A
(n1)
by F
(n)
= dA
(n1)
+ ,
where the ellipsis refers to Kaluza-Klein modications. The equation of motion for
A
(n1)
and the Bianchi identity for F
(n)
are of the form
d
_
e
c

F
(n)
_
= 0 + , dF
(n)
= 0 + , (8.1.2)
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 109
where the ellipses refer to Kaluza-Klein and/or Chern-Simons contributions. After
the dualisation, they become
dG
(Dn)
= 0 + , d
_
e
c

G
(D
n
)
_
= 0 + (8.1.3)
and are respectively interpreted as the Bianchi identity for the dual eld strength
and the equation of motion of the dual potential. By comparing the equations of
motion for the original potential and its dual potential, it is evident that the dilaton
vector associated with G
(Dn)
is equal to minus the dilaton vector associated with
F
(n)
.
Given the dualisation of all higher degrees eld strengths to lower degree eld
strengths, the CJ global symmetry group of the maximal, massless supergravity
theory in D dimensions is
1
E
11D
[81, 82, 80]. In Table 8.1, we have listed the E
11D
groups in 3 D 9, together with their maximal compact subgroups K(E
11D
).
The scalars, including the axions that come from the dualisation of (D 1)-form
eld strengths in D 5, parameterise the coset space K(E
11D
)/E
11D
, while the
higher degree eld strengths form certain linear representations of E
11D
.
The structure of the CJ global symmetry groups is already encoded in the alge-
braic relations that exist among the dilaton vectors (B.1.13). It has been shown that
the dilaton vectors associated with the axionic scalars (again, including those that
come from the dualisation of (D1)-form eld strengths in D 5) are in one-to-one
correspondence with the positive root vectors of the E
11D
algebra [80]. In particular,
the simple roots of the E
11D
algebra have been identied as

b
12
,

b
23
, . . . ,

b
(10D)(11D)
and a
123
, where a
123
is included only in D 8. The dilaton vectors associated with
1
The exceptional groups E
n
are here understood to be in their maximally non-compact form
E
n(+n)
.
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 110
E
11D
K(E
11D
)
D = 9 GL(2, IR) O(2, IR)
D = 8 SL(3, IR) SL(2, IR) O(3, IR) O(2, IR)
D = 7 SL(5, IR) SO(5, IR)
D = 6 SO(5, 5, IR) SO(5, IR) SO(5, IR)
D = 5 E
6(+6)
USp(8)
D = 4 E
7(+7)
SU(8)
D = 3 E
8(+8)
SO(16)
Table 8.1: CJ global symmetry groups in 3 D 9.
the p-form eld strengths, including possibly the dilaton vectors associated with the
duals of (Dp)-form eld strengths, are the weight vectors of the E
11D
representa-
tion under which these p-form eld strengths transform irreducibly [78, 80]. In Table
8.2, we group these dilaton vectors in each dimension into sets of weight vectors of
E
11D
representations. The bold face subscript for each set of weight vectors refers
to the corresponding E
11D
representation
2
.
The eld equations of D-dimensional supergravity are derived in Appendix B.1.
2
For the sets of dilaton vectors in D = 9, the subscripts refer to representations of the SL(2, IR)
subgroup of GL(2, IR). The two trivial representations, and the two fundamental representations,
are distinguished by their weights under the IR subgroup of GL(2, IR)

= SL(2, IR) IR.


CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 111
p = 2 p = 3 p = 4
D = 9 a
12

1
,

b
a

2
a
a

2
a
1
D = 8 a
ab
,

b
a

(3,2)
a
a

(3,1)
a, a
(1,2)
D = 7 a
ab
,

b
a

10
a
a
, a
5
D = 6 a
ab
,

b
a
, a
16
a
a
, a
a

10
D = 5 a
ab
,

b
a
, a
a

27
D = 4 a
ab
,

b
a
, a
ab
,

b
a

56
Table 8.2: Dilaton weight vectors of E
11D
representations.
After truncating the scalar sector, the gauge eld equations are
()
D4
d F
(4)
= X
(D3)
, dF
(4)
= F
(3)a
T
a
(2)
,
()
D3
d F
a
(3)
= F
(4)
T
a
(2)
+X
a
(D2)
, dF
(3)a
= F
(2)ab
T
b
(2)
,
()
D2
d F
ab
(2)
= 2 F
[a
(3)
T
b]
(2)
+X
ab
(D1)
, dF
(2)ab
= 0,
()
D2
d T
(2)a
= F
(4)
F
(3)a
+F
d
(3)
F
(2)da
, dT
a
(2)
= 0,
(8.1.4)
where X
(D3)
, X
a
(D2)
and X
ab
(D1)
are the dimension-dependent Chern-Simons con-
tributions to the equations of motion. They are given explicitly in Appendix B.1.
The eld equation for the metric can be written as
1
MN
= S
MN
(F
(4)
) +

d
S
MN
(F
(3)d
) +

c<d
S
MN
(F
(2)cd
) +

d
S
MN
(T
d
(2)
), (8.1.5)
where
S
MN
(F
(n)
) =
1
2
_
(F
(n)
)
2
MN

n1
D2
g
MN
(F
(n)
)
2
_
. (8.1.6)
The consistency of the truncation of the axions A
(0)mnp
and A
a
(0)
m
and the dilatons
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 112

requires in addition that


0 = 3 F
[ab
(2)
T
c]
(2)
+X
abc
(D)
,
0 = F
b
(3)
F
(3)a
F
bc
(2)
F
(2)ac
+T
(2)a
T
b
(2)
, a ,= b,
0 =a (F
(4)
)
2
+

a
a
a
(F
(3)a
)
2
+

a<b
a
ab
(F
(2)ab
)
2
+

b
a
(T
a
(2)
)
2
,
(8.1.7)
where the D-forms X
abc
(D)
are given explicitly for each 3 D 9 in Appendix B.1.
Note also that we consistently truncate the 4-form eld strength in D = 4 and the
3-form eld strengths in D = 3, since we want to restrict our attention to massless
supergravities.
In D = 5, 4, 3, the dualisation of the (D 1)-from eld strengths introduces
1-form eld strengths, whose potentials are axionic scalars. We also want to set
these axions to zero. Equivalently, we truncate the (D 1)-form eld strengths in
these dimensions. It is easy to see from the gauge eld equations (8.1.4) that the
consistency of this truncation imposes the additional constraints
D = 5 : 0 = F
(3)a
T
a
(2)
,
D = 4 : 0 = F
(2)ab
T
b
(2)
.
(8.1.8)
For notational convenience, we have again chosen to set the scalar elds to zero,
thereby selecting a preferred point in the moduli space of constant scalar vacuum
solutions. Other points in this moduli space are reached by acting on the zero-scalar
solution with an element of the E
11D
CJ global symmetry group. The classical
symmetry group of the zero-scalar vacua is the K(E
11D
) subgroup of E
11D
, since
the elements of K(E
11D
) do not change the values of the scalar elds. In other
words, K(E
11D
) is the stability group of each point in the moduli space of constant
scalar vacua.
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 113
A particularly important subgroup of the K(E
11D
) symmetry group of the con-
stant scalar vacuum solutions is the Weyl group of E
11D
, in the following denoted
by W(E
11D
). The Weyl group of E
11D
can be viewed as the symmetry group
of the weight vectors of E
11D
representations. To discuss this in more detail, let
e

: = 1, , 11 D denote the set of simple roots of the E


11D
algebra. The
Weyl reection of a weight vector c about the simple root e

is dened by

(c) =c
1
2
(e

c) e

. (8.1.9)
In the present context, the simple roots are

b
12
,

b
23
, . . . ,

b
(10D)(11D)
, a
123
, where
a
123
is included only in D 8, and the weight vectors that we are interested in are
given in Table 8.2. Using (B.1.13) and (B.1.6), it is possible to verify that the Weyl
reections

map each set of dilaton weight vectors in Table 8.2 onto itself. Note
also that

(c)) = c, since e

= 4 for all simple roots. It follows that the


permutations of the weight vectors of the E
11D
representations generated by the
Weyl reections about the simple roots form a group. This group is precisely the
Weyl group of E
11D
, as identied by its action on the weight vectors of the E
11D
representations [78].
For zero-scalar vacuum solutions, the action of the Weyl group on the eld
strengths can simply be read o its corresponding action on the dilaton vectors,
since there is a unique association of a eld strength (or a dualised eld strength)
with every dilaton weight vector. In other words, the Weyl group permutes the eld
strengths, up to changes of sign, in the same way it permutes the associated dilaton
vectors [78]. It is possible to verify explicitly in each dimension that the truncated
eld equations (8.1.4) and (8.1.7) (and (8.1.8) in D = 5, 4) are invariant under the
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 114
simultaneous Weyl group action on the eld strengths and the dilaton vectors.
The importance of the Weyl group for the purpose of classifying constant scalar
vacuum solutions derives from the observation that since its action merely permutes
the eld strengths of a given degree among themselves, it a fortiori do not change the
number of non-zero eld strengths participating in the vacuum solution. Therefore,
acting with the Weyl group on a vacuum solution involving n
p
p-form eld strengths
lls out an entire multiplet of constant scalar vacua with the same number of non-
zero p-form eld strengths. It follows that the Weyl group symmetry is the classifying
symmetry for constant scalar vacua with xed numbers n
p
of non-zero p-form eld
strengths.
It is usually straightforward to identify the smallest set of eld strengths that
is required to support a given vacuum solution. Acting with the Weyl group on
this solution lls out the corresponding Weyl multiplet. Acting with an element of
the symmetry group K(E
11D
) of zero-scalar vacua on this Weyl multiplet generates
another Weyl multiplet of zero-scalar vacua, but each solution in the new Weyl
multiplet involves, in general, more than the minimum number of eld strengths. In
the following, we shall list only those Weyl multiplets of constant scalar vacua with
the smallest set of non-zero eld strengths, with the understanding that zero-scalar
vacua with more participating eld strengths are generated by the action of the
vacuum symmetry group K(E
11D
), while vacua with non-zero, but constant, scalar
elds are generated by the action of the E
11D
CJ symmetry group.
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 115
8.2 D = 9
The p-form eld strengths in D = 9 are F
(4)
, F
(3)a
, F
(2)12
and T
a
(2)
, where a = 1, 2.
The CJ global symmetry group in D = 9 is E
2

= GL(2, IR). The 4-form eld
strength F
(4)
and the 2-form eld strength F
(2)12
are singlets of E
2
, while the two
3-form eld strengths F
(3)a
and the two 2-form eld strengths T
a
(2)
transform as
doublets.
We have considered vacuum ansatze based on the decompositions of the 9-
dimensional spacetime given in Table 8.3. For each of these decompositions, we
N = 2 N = 3 N = 4
2 + 7 2 + 2 + 5 2 + 2 + 2 + 3
3 + 6 2 + 3 + 4
4 + 5 3 + 3 + 3
Table 8.3: Spacetime decompositions for D = 9 vacuum ansatze.
have expanded the eld strengths F
(4)
, F
(3)a
, F
(2)12
and F
a
(2)
in the basis of harmonic
volume p-forms and then tried to solve the truncated eld equations (8.1.4) and
(8.1.7) for the expansion coecients. Finally, the scalar curvatures of the submani-
folds have been determined using the Einstein equation (8.1.5). In the following, we
present those constant scalar vacuum solution of D = 9 supergravity that a priori
cannot be toroidally reduced to D < 9. The indices i, j are understood to refer
to a permutation of 1, 2, so that i ,= j.
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 116
M
2
K
2
K
5
and M
5
K
2
K
2
vacua
We have found two vacuum solutions of D = 9 supergravity associated with the
2 + 2 + 5 decomposition of the 9-dimensional spacetime. They are singlets of the
Weyl group W(E
2
) and are represented by the tablatures in Figure 8.1.
M
2
K
2
K
5
F
(4)
: u
F
(2)12
: w
1

w
2

M
5
K
2
K
2
F
(4)
: u
F
(2)12
: w
1

w
2

Figure 8.1: The M
2
K
2
K
5
and M
5
K
2
K
2
vacua.
The truncation of the dilatonic scalars is consistent provided that
M
2
K
2
K
5
: M
5
K
2
K
2
:
u
2
= 2 (w
2
1
w
2
2
), u
2
= 2 (w
2
1
+w
2
2
).
(8.2.1)
The scalar curvatures of the submanifolds in Figure 8.1 are
M
2
K
2
K
5
: M
5
K
2
K
2
:
(M
2
) =
1
2
w
2
1

1
4
u
2
, (M
5
) =
1
16
u
2
,
(K
2
) =
1
2
w
2
2

1
4
u
2
, (K
2
) =
1
2
w
2
1
+
1
4
u
2
,
(K
5
) =
1
16
u
2
, (K
2
) =
1
2
w
2
2
+
1
4
u
2
.
(8.2.2)
The maximally symmetric 2 + 2 + 5 vacuum solutions of D = 9 supergravity are
therefore AdS
2
K
2
S
5
, where K
2
is determined according to Table C.1, and
AdS
5
S
2
S
2
. The latter solution has been found previously in Ref. [91] by
dierent means.
By oxidising these 9-dimensional vacuum solutions to D = 11, it is not dicult
to establish that can be interpreted as toroidal reductions of the 2 + 2 + 2 + 5
vacua depicted in Figure 6.8. In particular, the M
2
K
2
K
5
vacuum oxidises to a
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 117
M
2
K
2
K
2
K
5
vacuum for which one of the (Euclidean signature) 2-manifolds has
zero curvature, and hence is a 2-torus. Indeed, the constraint in u
2
= 2 (w
2
1
w
2
2
),
which permits a consistent truncation of the 9-dimensional dilatons, follows from
the 11-dimensional perspective from setting the scalar curvature of one of the 2-
manifolds to zero, c.f. equation (6.3.2). The details of the toroidal reduction are
summarised in Figure 8.2.
D = 11
KK
D = 9
M
2
K
2
T
2
K
5
F
(4)
: u
w
1

w
2

M
2
K
2
K
5
F
(4)
: u
F
(2)12
: w
1

w
2

Figure 8.2: Toroidal reduction of M
2
K
2
T
2
K
5
to M
2
K
2
K
5
.
Similarly, the M
5
K
2
K
2
vacuum oxidises to a M
5
K
2
K
2
T
2
vacuum,
as depicted in Figure 8.3. Again, it is not dicult to appreciate that the constraint
D = 11
KK
D = 9
M
5
K
2
K
2
T
2
F
(4)
: u
w
1

w
2

M
5
K
2
K
2
F
(4)
: u
F
(2)12
: w
1

w
2

Figure 8.3: Kaluza-Klein reduction of M
5
K
2
K
2
T
2
to M
5
K
2
K
2
.
u
2
= 2 (w
2
1
+ w
2
2
), which permits the consistent truncation of the 9-dimensional
dilatons, also insures that the choice of the 2-torus as one of the submanifolds is
consistent with the 11-dimensional Einstein equations.
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 118
Alternatively, we can rst T-dualise these 9-dimensional vacua and then oxidise
them to vacuum solution of IIB supergravity in D = 10. The 10-dimensional type
IIB elds are given in terms of the 9-dimensional (type IIA) elds that we are using
here by equation (B.2.7) in Appendix B.2. As pointed out there, the eld strength
F
(2)12
, which comes from the reduction of the 4-form eld strength in D = 11, is
mapped under T-duality to the Kaluza-Klein eld strength whose potential arises
from the reduction of the IIB metric in D = 10. This suggests that the 9-dimensional
vacuum solutions are Hopf reductions, rather than ordinary toroidal reductions, of
IIB vacuum solutions. Indeed, a more detailed analysis along the lines of Ref. [91]
shows that that upon oxidation to IIB supergravity in D = 10, the M
2
K
2
K
5
and M
5
K
2
K
2
vacua are interpreted as M
5
K
5
vacua for which one of the
5-manifolds is a U(1) bration over the product of two Einstein 2-manifolds. The
pictorial representations of the Hopf reduction that relates the 9-dimensional and
IIB vacua are given in Figures 8.4 and 8.5.
D = 10 (IIB)
Hopf
D = 9 (IIA)

Q
5
K
5
H
(5)
: u
u
AdS
2
K
2
K
5
F
(4)
: u
F
(2)12
: w
1

w
2

Figure 8.4: Hopf reduction of

Q
5
(w
1
, w
2
) K
5
to AdS
2
K
2
K
5
.
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 119
D = 10 (IIB)
Hopf
D = 9 (IIA)
M
5
Q
5
H
(5)
: u
u
M
5
S
2
S
2
F
(4)
: u
F
(2)12
: w
1

w
2

Figure 8.5: Hopf reduction of M
5
Q
5
(w
1
, w
2
) to M
5
S
2
S
2
.
M
3
K
3
K
3
vacua
We have found a Weyl doublet of vacuum solutions of D = 9 supergravity with three
submanifolds. It requires a 3 + 3 + 3 decomposition of the 9-dimensional spacetime
and is represented by the tablature in Figure 8.6, where
u
2
1
= u
2
2
+u
2
3
(8.2.3)
to ensure the consistency of the truncation of the dilatons.
M
3
K
3
K
3
F
(3)i
: u
1

u
2

u
3

Figure 8.6: The M = 2 Weyl multiplet of M
3
K
3
K
3
vacua.
The scalar curvatures of the submanifolds in Figure 8.6 are
(M
3
) =
1
4
u
2
1
, (K
3
) =
1
4
u
2
2
, (K
3
) =
1
4
u
2
3
. (8.2.4)
Since the Weyl tensor vanishes identically in three dimensions, an Einstein 3-manifold
is also maximally symmetric. The 3 +3 +3 vacuum solution of D = 9 supergravity
can therefore be identied with AdS
3
S
3
S
3
.
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 120
M
2
K
2
K
2
K
3
and M
3
K
2
K
2
K
2
vacua
We have found two vacuum solutions of D = 9 supergravity associated with the
2 + 2 + 2 + 3 decomposition of the 9-dimensional spacetime. They are singlets of
the Weyl group W(E
2
) and represented by the tablatures in Figure 8.7, where the
constants u

and w

are constrained by
M
2
K
2
K
2
K
3
: M
3
K
2
K
2
K
2
:
u
1
w
1
+u
2
w
2
+u
3
w
3
= 0, u
1
w
1
+u
2
w
2
+u
3
w
3
= 0,
u
2
1
+ 2w
2
1
= u
2
2
+u
2
3
+ 2(w
2
2
+w
2
3
), u
2
1
+u
2
2
+u
2
3
= 2(w
2
1
+w
2
2
+w
2
3
).
(8.2.5)
M
2
K
2
K
2
K
3
F
(4)
: u
1

u
2

u
3

F
(2)12
w
1

w
2

w
3

M
3
K
2
K
2
K
2
F
(4)
: u
1

u
2

u
3

F
(2)12
w
1

w
2

w
3

Figure 8.7: The M
2
K
2
K
2
K
3
and M
3
K
2
K
2
K
2
vacua.
A comparison of Figures 8.7 and 6.9 suggests strongly that the 2+2+2+3 vacua
in D = 9 oxidise to 2+2+2+2+3 vacua in D = 11 for which one of the 2-manifolds
is a 2-torus. To support this identication, it is helpful to rewrite (8.2.5) in terms
of the 3-vectors
M
2
K
2
K
2
K
3
: M
3
K
2
K
2
K
2
:
x = (u
3
, u
2
, w
1
), x = (w
1
, w
2
, w
3
),
y = (w
3
, w
2
, u
1
), y = (u
1
, u
2
, u
3
)
(8.2.6)
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 121
as
M
2
K
2
K
2
K
3
: M
3
K
2
K
2
K
2
:
x y = 0, x y = 0,
2y
2
+x
2
= 3(x
2
3
+y
2
3
), 2x
2
= y
2
,
(8.2.7)
where x
2
= x x and y
2
= y y. These 3-vectors become precisely the 3-vectors
dened in (6.3.4) when we oxidise the D = 9 vacua to D = 11. It then evident
from (8.2.7) and (6.3.5) that the truncation of the dilatons in D = 9, which imposes
the second equation in (8.2.7), is equivalent to setting the scalar curvature of the
appropriate 2-manifold in D = 11 to zero.
The identication of the 2 + 2 + 2 + 3 vacua in D = 9 as toroidal reductions
of the 2 + 2 + 2 + 2 + 3 vacua in D = 11 implies that the scalar curvatures of the
Einstein submanifolds in Figure 8.7 can be written as
M
2
K
2
K
2
K
3
: M
3
K
2
K
2
K
2
:
(M
2
) =
1
2
(2x
2
+y
2
), (M
3
) =
1
4
x
2
,
(K
2
) =
1
2
(x
2
3
+y
2
3
(x
2
1
+y
2
1
)), (K
2
) =
1
2
(x
2
+ (x
2
1
y
2
1
)),
(K
2
) =
1
2
(x
2
3
+y
2
3
(x
2
2
+y
2
2
)), (K
2
) =
1
2
(x
2
+ (x
2
2
y
2
2
)),
(K
3
) =
1
12
(x
2
y
2
), (K
2
) =
1
2
(x
2
+ (x
2
3
y
2
3
)).
(8.2.8)
8.3 D = 8
The p-form eld strengths in D = 8 are F
(4)
, F
(3)a
, F
(2)ab
and T
a
(2)
, where a = 1, 2, 3.
The CJ global symmetry group in D = 8 is E
3

= SL(3, IR) SL(2, IR). The 4-form
eld strength F
(4)
and its dual F
(4)
transform as a singlet under SL(3, IR) and as
a doublet under SL(2, IR). The 3-form eld strengths F
(3)a
transform as a triplet
under SL(3, IR) and as a singlet under SL(2, IR). The 2-form eld strengths F
(2)ab
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 122
and T
a
(2)
are in the (3, 2) representation of E
3
.
We have considered vacuum ansatze based on the decompositions of the 8-
dimensional spacetime given in Table 8.4. For each of these decompositions, we
N = 2 N = 3 N = 4
2 + 6 2 + 2 + 4 2 + 2 + 2 + 2
3 + 5 2 + 3 + 3
4 + 4
Table 8.4: Spacetime decompositions for D = 8 vacuum ansatze.
have expanded the eld strengths F
(4)
, F
(3)a
, F
(2)ab
and F
a
(2)
in the basis of harmonic
volume p-forms and then tried to solve the truncated eld equations (8.1.4) and
(8.1.7) for the expansion coecients. Finally, the scalar curvatures of the submani-
folds have been determined using the Einstein equation (8.1.5). It turns out that all
the constant scalar vacuum solutions of D = 8 supergravity that a priori cannot be
toroidally reduced to D < 8 are readily interpreted as toroidal or Hopf reductions of
vacuum solutions in D = 9, D = 10 or D = 11, or as Weyl transformations thereof.
We shall therefore discuss them directly from a point of view that highlights their
higher-dimensional origin. In the following, the indices i, j, k are understood to
refer to a permutation of 1, 2, 3, so that they are all distinct.
M
2
K
3
K
3
and M
3
K
3
K
2
vacua
The vacuum solutions of D = 8 supergravity with three submanifolds require a 2+3+
3 decomposition of the 8-dimensional spacetime. They organise themselves naturally
into multiplets of the Weyl group W(E
3
) with multiplicity M = 6. These Weyl
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 123
multiplets can be constructed either by solving the 8-dimensional eld equations
directly, or by Hopf reducing one of the 3 + 3 + 3 vacua in D = 9 and then acting
with the Weyl group on the resulting 8-dimensional vacuum to ll out the entire
multiplet. Here, we shall present the 2 + 3 + 3 vacua in D = 8 from the latter
perspective.
To begin with, consider the 3 + 3 + 3 vacuum solution of D = 9 supergravity,
which is depicted in Figure 8.6, supported by, say, F
(3)1
. Let the Lorentzian signature
3-manifold be given by AdS
3
, parameterised as a U(1) bration over AdS
2
. We can
then Hopf reduce this 3+3+3 vacuum along the U(1) bre direction to get a 2+3+3
vacuum in D = 8. The details of the Hopf reduction are summarised in Figure 8.8.
D = 9
Hopf
D = 8
AdS
3
K
3
K
3
F
(3)1
: u
1

u
2

u
3

AdS
2
K
3
K
3
F
(3)1
: u
2

u
3

F
(2)31
: u
1

T
3
(2)
: u
1

Figure 8.8: Hopf reduction of AdS
2
K
3
K
3
to AdS
2
K
3
K
3
.
Given the one particular 2 + 3 + 3 vacuum in Figure 8.8, we may act on it with
the Weyl group W(E
3
) to ll out an entire multiplet of 2 + 3 + 3 vacua in D = 8
that are supported by the same number of eld strengths. This Weyl multiplet is
depicted in Figure 8.9, where
w
2
= u
2
1
+u
2
2
. (8.3.1)
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 124
M
2
K
3
K
3
F
(3)i
: u
1

u
2

F
(2)ij
: w
T
j
(2)
: w
Figure 8.9: The M = 6 Weyl multiplet of M
2
K
3
K
3
vacua.
The scalar curvatures of the submanifolds in Figure 8.9 are
(M
2
) = w
2
, (K
3
) =
1
4
u
2
1
, (K
3
) =
1
4
u
2
2
. (8.3.2)
These vacuum solutions of D = 8 supergravity are therefore of the form AdS
2

S
3
S
3
.
Alternatively, we can consider the 3+3+3 vacuum solution of D = 9 with one of
the Euclidean signature 3-manifolds given by S
3
, parameterised as a U(1) bration
over S
2
. The Hopf reduction of this 3 +3 +3 vacuum along the U(1) bre direction
yields another 2 + 3 + 3 vacuum in D = 8. The details of the Hopf reduction are
summarised in Figure 8.10.
D = 9
Hopf
D = 8
M
3
K
3
S
3
F
(3)1
: u
1

u
2

u
3

M
3
K
3
S
2
F
(3)1
: u
1

u
2

F
(2)31
: u
3

T
3
(2)
: u
3

Figure 8.10: Hopf reduction of M
3
K
3
S
3
to M
3
K
3
S
2
.
The particular 2 + 3 + 3 vacuum in Figure 8.10 again serves as a seed for an
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 125
entire Weyl multiplet of 2 + 3 + 3 vacua in D = 8 that are supported by the same
number of eld strengths. This Weyl multiplet is depicted in Figure 8.11, where
u
2
1
= u
2
2
+w
2
. (8.3.3)
M
3
K
3
K
2
F
(3)i
: u
1

u
2

F
(2)ij
: w
T
j
(2)
: w
Figure 8.11: The M = 6 Weyl multiplet of M
3
K
3
K
2
vacua.
The scalar curvatures of the submanifolds in Figure 8.11 are
(M
3
) =
1
4
u
2
1
, (K
3
) =
1
4
u
2
2
, (K
2
) = w
2
. (8.3.4)
These vacuum solutions of D = 8 supergravity are therefore of the form AdS
3

S
3
S
2
.
M
2
K
2
K
2
K
2
vacua
The vacuum solutions of D = 8 supergravity with four submanifolds require a 2 +
2 +2 +2 decompositions of the 8-dimensional spacetime. They organise themselves
naturally into multiplets of the Weyl group W(E
3
) with multiplicities M = 1 or M =
6. These Weyl multiplets can be constructed either by solving the 8-dimensional eld
equations directly, or by dimensionally reducing certain vacua in D = 9, D = 10
and D = 11 and then acting with the Weyl group on the resulting 8-dimensional
vacuum to ll out the entire multiplet. Here, we shall present the 2+2+2+2 vacua
in D = 8 from the latter perspective.
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 126
To begin with, consider the M
2
K
2
K
2
K
2
K
3
vacuum in D = 11, which
is depicted in Figure 6.9. We see from (6.3.5) that it is possible to set the scalar
curvature of the 3-manifold to zero and hence identify the 3-manifold with a 3-torus.
The toroidal reduction of this 2+2+2+2+3 vacuum in D = 11 yields the 2+2+2+2
vacuum in D = 8 represented in Figure 8.12.
M
2
K
2
K
2
K
2
F
(4)
: u
12

u
13

u
14

u
23

u
24

u
34

Figure 8.12: An M
2
K
2
K
2
K
2
vacuum.
The constants u

are constrained by
x y = 0, x
2
1
+x
2
2
+x
2
3
= y
2
1
+y
2
2
+y
3
3
. (8.3.5)
where the 3-vectors x and y have been dened previously in (6.3.4). The rst
constraint in (8.3.5) ensures that the Chern-Simons contributions to the gauge eld
equations vanish, while the second constraint can be viewed equivalently as the zero-
curvature condition for the 3-manifold in D = 11, or as guaranteeing the consistency
of the truncation of the dilatons in D = 8.
The scalar curvatures of the submanifolds in Figure 8.12 can be deduced from
(6.3.5) and (8.3.5). Explicitly, they are given by
M
2
K
2
K
2
K
2
:
(M
2
) =
1
2
(x
2
1
+x
3
2
+x
2
3
),
(K
2
) =
1
2
(x
2
2
+x
2
3
y
2
1
),
(K
2
) =
1
2
(x
2
1
+x
2
3
y
2
2
),
(K
2
) =
1
2
(x
2
1
+x
3
2
y
2
3
).
(8.3.6)
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 127
This 2 + 2 + 2 + 2 vacuum solution of D = 8 supergravity is therefore of the form
AdS
2
K
2
K
2
K
2
, where K
2
is determined according to Table C.1.
Consider next the 2 +2 +2 +2 +2 vacuum in D = 10 depicted in Figure 7.10. If
we set one the constants to zero, say v
5
= 0, it is of the form AdS
2
S
2
S
2
S
2
T
2
.
A toroidal reduction therefore yields another 2 + 2 + 2 + 2 vacuum in D = 8. That
8-dimensional vacuum is part of the Weyl multiplet represented in Figure 8.13, where
v
2
1
= v
2
2
+v
2
3
+v
2
4
. (8.3.7)
M
2
K
2
K
2
K
2
T
i
(2)
: v
1

v
2

v
3

v
4

M
2
K
2
K
2
K
2
F
(2)jk
: v
1

v
2

v
3

v
4

Figure 8.13: An M = 6 Weyl multiplet of AdS
2
S
2
S
2
S
2
vacua.
The scalar curvatures of the submanifolds in Figure 8.13 are given by
(M
2
) =
1
2
v
2
1
,
(K
2
) =
1
2
v
2

, = 2, 3, 4.
(8.3.8)
Therefore, we can identify these vacuum solutions of D = 8 supergravity with AdS
2

S
2
S
2
S
2
.
The remaining vacuum solutions of D = 8 supergravity with four submanifolds
are most easily obtained via Hopf reduction from D = 9 vacua. Consider rst the
2 + 2 + 5 vacua in D = 9, which are depicted in Figure 8.1. Let the 5-manifold be
given by a U(1) bration over the product of two Einstein 2-manifolds. The Hopf
reduction of these 9-dimensional vacua along the U(1) bre direction yields vacuum
solutions of D = 8 supergravity that are part of the Weyl multiplet depicted in
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 128
Figure 8.14, where
u
2
= 2 (w
2
1
+w
2
2
), v
2
1
= v
2
2
+w
2
1
+w
2
2
. (8.3.9)
M
2
K
2
K
2
K
2
F
(4)
: u
F
(2)ij
: w
1

w
2

T
k
(2)
: v
1

v
2

M
2
K
2
K
2
K
2
F
(4)
: u
F
(2)ij
: v
1

v
2

T
k
(2)
: w
1

w
2

Figure 8.14: An M = 6 Weyl multiplet of M
2
K
2
K
2
K
2
vacua.
The scalar curvatures of the submanifolds in Figure 8.14 are
M
2
K
2
K
2
K
2
:
(M
2
) = (w
2
1
+w
2
2
)
1
2
v
2
2
,
(K
2
) =
1
2
(v
2
1
u
2
)
(K
2
) = w
2
1
+
1
2
w
2
2
,
(K
2
) = w
2
2
+
1
2
w
2
1
.
(8.3.10)
These vacuum solutions of D = 8 supergravity are therefore of the form AdS
2

K
2
S
2
S
2
, where K
2
is determined according to Table C.1.
Consider next the M
3
K
2
K
2
K
2
vacua in D = 9, which are depicted in
Figure 8.7. Let the Lorentzian signature 3-manifold be given by AdS
3
, parameterised
as a U(1) bration over AdS
2
. The Hopf reduction of this 9-dimensional vacuum
along the U(1) bre direction yields a vacuum solution of D = 8 supergravity that
is part of the Weyl multiplet depicted in Figure 8.15.
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 129
M
2
K
2
K
2
K
2
F
(4)
: u
1

u
2

u
3

F
(2)ij
: w
1

w
2

w
3

T
k
(2)
: v
M
2
K
2
K
2
K
2
F
(4)
: u
1

u
2

u
3

F
(2)ij
: v
T
k
(2)
: w
1

w
2

w
3

Figure 8.15: An M = 6 Weyl multiplet of M
2
K
2
K
2
K
2
vacua.
The constants u

, w

and v in Figure 8.15 are constrained by


u
1
w
1
+u
2
w
2
+u
3
w
3
= 0,
u
2
1
+u
2
2
+u
2
3
= 2(w
2
1
+w
2
2
+w
2
3
),
v
2
= w
2
1
+w
2
2
+w
2
3
.
(8.3.11)
The scalar curvatures of the Einstein submanifolds in Figure 8.15 can be written
as
(M
2
) = v
2
,
(K
2
) =
1
2
(v
2
+w
2
1
u
2
1
),
(K
2
) =
1
2
(v
2
+w
2
2
u
2
2
),
(K
2
) =
1
2
(v
2
+w
2
3
u
2
3
).
(8.3.12)
These vacuum solutions of D = 8 supergravity are therefore of the form AdS
2

K
2
S
2
S
2
, where K
2
is determined according to Table C.1.
Consider nally the M
2
K
2
K
2
K
3
vacua in D = 9, which are depicted in
Figure 8.7. Let the Euclidean signature 3-manifold be given by S
3
, parameterised
as a U(1) bration over S
2
. The Hopf reduction of this 9-dimensional vacuum along
the U(1) bre direction yields a vacuum solution of D = 8 supergravity that is part
of the Weyl multiplet depicted in Figure 8.16.
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 130
M
2
K
2
K
2
K
2
F
(4)
: u
1

u
2

u
3

F
(2)ij
: w
1

w
2

w
3

T
k
(2)
: v
M
2
K
2
K
2
K
2
F
(4)
: u
1

u
2

u
3

F
(2)ij
: v
T
k
(2)
: w
1

w
2

w
3

Figure 8.16: An M = 6 Weyl multiplet of M
2
K
2
K
2
K
2
vacua.
The constants u

, w

and v in Figure 8.16 are constrained by


u
1
w
1
+u
2
w
2
+u
3
w
3
= 0,
u
2
1
+ 2w
2
1
= u
2
2
+u
2
3
+ 2(w
2
2
+w
2
3
),
w
2
1
= v
2
+w
2
2
+w
2
3
.
(8.3.13)
The scalar curvatures of the Einstein submanifolds in Figure 8.16 can be written
as
(M
2
) =
1
2
(u
2
2
+u
2
3
+w
2
2
+w
2
3
),
(K
2
) =
1
4
(u
2
1
+u
2
2
u
2
3
+w
2
2
),
(K
2
) =
1
4
(u
2
2
u
2
2
+u
2
3
+w
2
3
),
(K
2
) = v
2
.
(8.3.14)
These vacuum solutions of D = 8 supergravity are therefore of the form AdS
2

K
2
K
2
S
2
, where K
2
is determined according to Table C.1.
8.4 D = 7
The p-form eld strengths in D = 7 are F
(4)
, F
(3)a
, F
(2)ab
and T
a
(2)
, where a = 1, . . . , 4.
The CJ global symmetry group in D = 7 is E
4

= SL(5, IR). The 3-form eld
strengths F
(3)a
and F
(4)
transform as a quintuplet under SL(5, IR). The 2-form
eld strengths F
(2)ab
and T
a
(2)
are in the

10 representation of SL(5, IR).
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 131
We have considered vacuum ansatze based on the decompositions of the 7-
dimensional spacetime given in Table 8.5. For each of these decompositions, we
N = 2 N = 3
2 + 5 2 + 2 + 3
3 + 4
Table 8.5: Spacetime decompositions for D = 7 vacuum ansatze.
have expanded the eld strengths F
(4)
, F
(3)a
, F
(2)ab
and F
a
(2)
in the basis of harmonic
volume p-forms and then tried to solve the truncated eld equations (8.1.4) and
(8.1.7) for the expansion coecients. Finally, the scalar curvatures of the subman-
ifolds have been determined using the Einstein equation (8.1.5). It turns out that
all the constant scalar vacuum solutions of D = 7 supergravity that a priori cannot
be toroidally reduced to D < 7 are readily interpreted as toroidal or Hopf reduc-
tions of vacuum solutions in D = 8 or D = 9, or as Weyl transformations thereof.
We shall therefore discuss them directly from a point of view that highlights their
higher-dimensional origin. In the following, the indices i, j, k, l are understood
to refer to a permutation of 1, 2, 3, 4, so that they are all distinct.
M
2
K
2
K
3
and M
3
K
2
K
2
vacua
The vacuum solutions of D = 7 supergravity with three submanifolds require a
2 + 2 + 3 decomposition of the 7-dimensional spacetime. They organise themselves
naturally into multiplets of the Weyl group W(E
4
) with multiplicity M = 15. Dis-
tinct Weyl multiplets are distinguished by the signatures of the submanifolds and/or
the number of eld strengths involved in the vacuum solution. These Weyl multi-
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 132
plets can be constructed either by solving the 7-dimensional eld equations directly,
or by dimensionally reducing one of the 2 +2 +2 +3 vacua in D = 9, or one of the
2 + 3 + 3 vacua in D = 8, and then acting with the Weyl group on the resulting
7-dimensional vacuum to ll out the entire multiplet. Here, we shall present the
2 + 2 + 3 vacua in D = 7 from the latter perspective.
To begin with, consider the M
3
K
3
K
2
vacuum in D = 8, which is de-
picted in Figure 8.11. Let the Lorentzian signature 3-manifold be given by AdS
3
,
parameterised as a U(1) bration over AdS
2
. Alternatively, we can consider the
M
2
K
3
K
3
vacuum solution of D = 8, which is depicted in Figure 8.9, with
one of the Euclidean signature 3-manifolds given by S
3
, parameterised as a U(1)
bration over S
2
. The Hopf reduction of either of these 8-dimensional vacua along
the U(1) bre direction yields a vacuum solution of D = 7 supergravity that is part
of the Weyl multiplet depicted in Figure 8.17, where
w
2
= u
2
+v
2
. (8.4.1)
M
2
K
2
K
3
F
(3)i
: u
F
(2)ij
: w
F
(2)ik
: v
T
j
(2)
: w
T
k
(2)
: v
M
2
K
2
K
3
F
(4)
: u
F
(2)ij
: w
F
(2)ik
: v
F
(2)kl
: w
F
(2)jl
: v
Figure 8.17: The M = 15 Weyl multiplet of M
2
K
2
K
3
vacua.
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 133
The scalar curvatures of the submanifolds in Figure 8.17 are
(M
2
) = w
2
, (K
2
) = v
2
, (K
3
) =
1
4
u
2
. (8.4.2)
These vacuum solutions of D = 7 supergravity are therefore of the form AdS
2

S
2
S
3
.
Next, consider again the M
3
K
3
K
2
vacuum in D = 8, which is depicted in
Figure 8.11. This time, however, let the Euclidean signature 3-manifolds given by S
3
,
parameterised as a U(1) bration over S
2
. The Hopf reduction this 8-dimensional
vacua along the U(1) bre direction yields a vacuum solution of D = 7 supergravity
that is part of the Weyl multiplet depicted in Figure 8.18, where
u
2
= v
2
+w
2
. (8.4.3)
M
3
K
2
K
2
F
(3)i
: u
F
(2)ij
: w
F
(2)ik
: v
T
j
(2)
: w
T
k
(2)
: v
M
3
K
2
K
2
F
(4)
: u
F
(2)ij
: w
F
(2)ik
: v
F
(2)kl
: w
F
(2)jl
: v
Figure 8.18: The M = 15 Weyl multiplet of M
3
K
2
K
2
vacua.
The scalar curvatures of the submanifolds in Figure 8.18 are
(M
3
) =
1
4
u
2
, (K
2
) = w
2
, (K
2
) = v
2
. (8.4.4)
These vacuum solutions of D = 7 supergravity are therefore of the form AdS
3

S
2
S
2
.
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 134
Consider next the M
3
K
2
K
2
K
2
vacuum in D = 9, which is depicted
in Figure 8.7, with e.g. w
3
= 0. It is easily veried that we may consistently set
the scalar curvature of the third 2-manifold to zero, in which case the 2 +2 +2 + 3
vacuum is of the form M
3
K
2
K
2
T
2
. The toroidal reduction this 9-dimensional
vacuum yields a vacuum solution of D = 7 supergravity that is part of the Weyl
multiplet depicted in Figure 8.19, where
u
2
= w
2
1
+w
2
2
. (8.4.5)
M
3
K
2
K
2
F
(3)i
: u
F
(2)ij
: w
1

w
2

T
j
(2)
: w
1

w
2

M
3
K
2
K
2
F
(4)
: u
F
(2)ij
: w
1

w
2

F
(2)kl
: w
1

w
2

Figure 8.19: The M = 15 Weyl multiplet of M
3
K
2
K
2
vacua.
The scalar curvatures of the submanifolds in Figure 8.19 are
(M
3
) =
1
4
u
2
, (K
2
) = w
2
1
, (K
2
) = w
2
2
. (8.4.6)
These vacuum solutions of D = 7 supergravity are therefore of the form AdS
2

S
2
S
3
.
Alternatively, consider the M
2
K
2
K
2
K
3
vacuum in D = 9, which is
depicted in Figure 8.7, with e.g. w
3
= 0. We may again consistently set the scalar
curvature of the third 2-manifold to zero, in which case the 2 +2 +2 +3 vacuum is
of the form M
2
K
2
T
2
K
3
. The toroidal reduction this 9-dimensional vacuum
yields a vacuum solution of D = 7 supergravity that is part of the Weyl multiplet
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 135
depicted in Figure 8.20, where
w
2
1
= u
2
+w
2
2
. (8.4.7)
M
2
K
2
K
3
F
(3)i
: u
F
(2)ij
: w
1

w
2

T
j
(2)
: w
1

w
2

M
2
K
2
K
3
F
(4)
: u
F
(2)ij
: w
1

w
2

F
(2)kl
: w
1

w
2

Figure 8.20: The M = 15 Weyl multiplet of M
2
K
2
K
3
vacua.
The scalar curvatures of the submanifolds in Figure 8.20 are
(M
2
) = w
2
1
, (K
2
) = w
2
2
, (K
3
) =
1
4
u
2
. (8.4.8)
These vacuum solution of D = 7 supergravity are therefore of the form AdS
2
S
2

S
3
.
8.5 D = 6
The p-form eld strengths in D = 6 are F
(4)
, F
(3)a
, F
(2)ab
and T
a
(2)
, where a = 1, . . . , 5.
The CJ global symmetry group in D = 6 is E
5

= SO(5, 5). The 3-form eld strengths
F
(3)a
and F
a
(3)
are in the fundamental representation of SO(5, 5), whereas the 2-form
eld strengths F
(4)
, F
(2)ab
and T
a
(2)
are in the 16 representation of SO(5, 5).
We have considered vacuum ansatze based on the decompositions of the 6-
dimensional spacetime given in Table 8.6. For each of these decompositions, we
have expanded the eld strengths F
(4)
, F
(3)a
, F
(2)ab
and F
a
(2)
in the basis of harmonic
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 136
N = 2 N = 3
2 + 4 2 + 2 + 2
3 + 3
Table 8.6: Spacetime decompositions for D = 6 vacuum ansatze.
volume p-forms and then tried to solve the truncated eld equations (8.1.4) and
(8.1.7) for the expansion coecients. Finally, the scalar curvatures of the submani-
folds have been determined using the Einstein equation (8.1.5). It turns out that all
the constant scalar vacuum solutions of D = 6 supergravity that a priori cannot be
toroidally reduced to D < 6 are readily interpreted as toroidal or Hopf reductions
of vacuum solutions in D = 7, D = 8 or D = 9, or as Weyl transformations thereof.
We shall therefore discuss them directly from a point of view that highlights their
higher-dimensional origin. In the following, the indices i, j, k, l, m are understood
to refer to an even permutation of 1, 2, 3, 4, 5, so that they are all distinct.
M
3
K
3
vacua
The vacuum solutions of D = 6 supergravity with two submanifolds require a 3 +3
decomposition of the 6-dimensional spacetime. They organise themselves naturally
into a multiplet of the Weyl group W(E
5
) with multiplicity M = 5. This Weyl
multiplets can be constructed either by solving the 6-dimensional eld equations
directly, or by dimensionally reducing one of the 3 +3 +3 vacua in D = 9, and then
acting with the Weyl group on the resulting 6-dimensional vacuum to ll out the
entire multiplet. Here, we shall present the 3 + 3 vacua in D = 6 from the latter
perspective.
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 137
Consider the 3 +3 vacuum in D = 9 that is depicted in Figure 8.6. It is possible
to set the scalar curvature of one of the Euclidean signature 3-manifolds to zero,
e.g. by setting u
3
= 0, whence we have a vacuum solution of D = 9 supergravity of
the form M
3
K
3
T
3
. The toroidal reduction this 9-dimensional vacuum yields a
vacuum solution of D = 6 supergravity that is part of the Weyl multiplet depicted
in Figure 8.21.
M
3
K
3
F
(3)i
: u
u
Figure 8.21: The M = 5 Weyl multiplet of M
3
K
3
vacua.
The scalar curvatures of the submanifolds in Figure 8.21 are
(M
3
) =
1
4
u
2
, (K
3
) =
1
4
u
2
. (8.5.1)
These vacuum solutions of D = 6 supergravity can therefore be identied with
AdS
3
S
3
.
M
2
K
2
K
2
vacua
The vacuum solutions of D = 6 supergravity with three submanifolds require a
2 + 2 + 2 decomposition of the 6-dimensional spacetime. They organise themselves
naturally into multiplets of the Weyl group W(E
5
) with multiplicity M = 16, M =
40 or M = 75. Distinct Weyl multiplets are distinguished by the signatures of the
submanifolds and/or the number of eld strengths involved in the vacuum solution.
These Weyl multiplets can be constructed either by solving the 6-dimensional eld
equations directly, or by dimensionally reducing one of the 2 + 2 + 2 + 2 vacua in
D = 8, or one of the 2 +2 +3 vacua in D = 7, and then acting with the Weyl group
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 138
on the resulting 6-dimensional vacuum to ll out the entire multiplet. Here, we shall
present the 2 + 2 + 2 vacua in D = 6 from the latter perspective.
To begin with, consider the M
3
K
2
K
2
vacuum in D = 7 that is depicted in Fig-
ure 8.18. Let the Lorentzian signature 3-manifold be given by AdS
3
, parameterised
as a U(1) bration over AdS
2
. Alternatively, we can consider the M
2
K
2
K
3
vacuum solution of D = 7, which is depicted in Figure 8.17, with the Euclidean
signature 3-manifolds given by S
3
, parameterised as a U(1) bration over S
2
. The
Hopf reduction of either of these 7-dimensional vacua along the U(1) bre direction
yields a vacuum solution of D = 6 supergravity that is part of the Weyl multiplet
depicted in Figure 8.22, where
u
2
= v
2
+w
2
. (8.5.2)
The scalar curvatures of the submanifolds in Figure 8.22 are
(M
2
) = u
2
, (K
2
) = v
2
, (K
3
) = w
2
. (8.5.3)
These vacuum solutions of D = 6 supergravity can therefore be identied with
AdS
2
S
2
S
2
.
Consider next the M
2
K
2
K
3
vacuum in D = 7 that is depicted in Figure
8.20. Let the Euclidean signature 3-manifold be given by S
3
, parameterised as a
U(1) bration over S
2
. The Hopf reduction this 7-dimensional vacuum along the
U(1) bre direction yields a vacuum solution of D = 6 supergravity that is part of
the Weyl multiplet depicted in Figure 8.23, where
w
2
1
= w
2
2
+u
2
. (8.5.4)
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 139
M
2
K
2
K
2
F
(2)ij
: u
F
(2)ik
: v
F
(2)il
: w
T
j
(2)
: u
T
k
(2)
: v
T
l
(2)
: w
M
2
K
2
K
2
F
(4)
: u
F
(2)ij
: v
F
(2)ik
: w
T
m
(2)
: u
F
(2)kl
: v
F
(2)jl
: w
M
2
K
2
K
2
F
(2)ij
: u
F
(2)ik
: v
F
(2)il
: w
F
(2)kl
: u
F
(2)lj
: v
F
(2)jk
: w
Figure 8.22: The M = 40 Weyl multiplet of M
2
K
2
K
2
vacua.
The scalar curvatures of the submanifolds in Figure 8.23 are
(M
2
) = w
2
1
, (K
2
) = w
2
2
, (K
2
) = u
2
. (8.5.5)
These vacuum solutions of D = 6 supergravity can therefore be identied with
AdS
2
S
2
S
2
.
Consider next the M
3
K
2
K
2
vacuum in D = 7 that is depicted in Figure
8.19. Let the Lorentzian signature 3-manifold be given by AdS
3
, parameterised as
a U(1) bration over AdS
2
. The Hopf reduction this 7-dimensional vacuum along
the U(1) bre direction yields a vacuum solution of D = 6 supergravity that is part
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 140
M
2
K
2
K
2
F
(4)
: u
T
m
(2)
: u
F
(2)ij
: w
1

w
2

F
(2)kl
: w
1

w
2

M
2
K
2
K
2
F
(2)ij
: u
T
j
(2)
: u
F
(2)ik
: w
1

w
2

T
k
(2)
: w
1

w
2

M
2
K
2
K
2
F
(2)ij
: u
F
(2)kl
: u
F
(4)
: w
1

w
2

T
m
(2)
: w
1

w
2

M
2
K
2
K
2
F
(2)ij
: u
F
(2)kl
: u
F
(2)ik
: w
1

w
2

F
(2)jk
: w
1

w
2

Figure 8.23: The M = 75 Weyl multiplet of M
2
K
2
K
2
vacua.
of the Weyl multiplet depicted in Figure 8.24, where
u
2
= w
2
1
+w
2
2
. (8.5.6)
The scalar curvatures of the submanifolds in Figure 8.24 are
(M
2
) = u
2
, (K
2
) = w
2
1
, (K
2
) = w
2
2
. (8.5.7)
These vacuum solutions of D = 6 supergravity can therefore be identied with
AdS
2
S
2
S
2
.
Consider next the 2 + 2 + 2 + 2 vacuum in D = 8 that is depicted in Figure
8.13. It is possible to set the scalar curvature of one of the Euclidean signature
2-manifolds to zero, e.g. by setting v
4
= 0, whence we have a vacuum solution of
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 141
M
2
K
2
K
2
F
(4)
: u
T
m
(2)
: u
F
(2)ij
: w
1

w
2

F
(2)kl
: w
1

w
2

M
2
K
2
K
2
F
(2)ij
: u
T
j
(2)
: u
F
(2)ik
: w
1

w
2

T
k
(2)
: w
1

w
2

M
2
K
2
K
2
F
(2)ij
: u
F
(2)kl
: u
F
(4)
: w
1

w
2

T
m
(2)
: w
1

w
2

M
2
K
2
K
2
F
(2)ij
: u
F
(2)kl
: u
F
(2)ik
: w
1

w
2

F
(2)jk
: w
1

w
2

Figure 8.24: The M = 75 Weyl multiplet of M
2
K
2
K
2
vacua.
D = 8 supergravity of the form AdS
2
S
2
S
2
T
2
. The toroidal reduction this
8-dimensional vacuum yields a vacuum solution of D = 6 supergravity that is part
of the Weyl multiplet depicted in Figure 8.25, where
v
2
1
= v
2
2
+v
2
3
(8.5.8)
The scalar curvatures of the submanifolds in Figure 8.25 are
(M
2
) =
1
2
v
2
1
, (K
2
) =
1
2
v
2
2
, (K
2
) =
1
2
v
2
3
. (8.5.9)
These vacuum solutions of D = 6 supergravity can therefore be identied with
AdS
2
S
2
S
2
.
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 142
M
2
K
2
K
2
F
(4)
: v
1

v
2

v
3

M
2
K
2
K
2
F
(2)ij
: v
1

v
2

v
3

M
2
K
2
K
2
T
i
(2)
: v
1

v
2

v
3

Figure 8.25: The M = 16 Weyl multiplet of M
2
K
2
K
2
vacua.
8.6 D = 5
The p-form eld strengths in D = 5 are F
(3)a
, F
(2)ab
and T
a
(2)
, where a = 1, . . . , 6.
The CJ global symmetry group in D = 6 is E
6
. The 2-form eld strengths F
a
(3)
,
F
(2)ab
and T
a
(2)
are in the 27 representation of E
6
.
M
2
K
3
and M
3
K
2
vacua
As we are seeking vacuum solutions of D = 5 supergravity that a priori cannot be
toroidally reduced to D < 5, we only need to consider the Freund-Rubin ansatze
based on the 2 +3 decomposition of the 5-dimensional spacetime. It turns out that
all the 2 + 3 vacua are readily interpreted as Hopf reductions of the 3 + 3 vacuum
solutions in D = 6, or as Weyl transformations thereof. We shall therefore discuss
them directly from a point of view that highlights their higher-dimensional origin.
In the following, the indices i, j, k, l, m, n are understood to refer to an even
permutation of 1, 2, 3, 4, 5, 6, so that they are all distinct.
To begin with, consider the 3 +3 vacuum solution of D = 6 supergravity, which
is depicted in Figure 8.6. Let the Lorentzian signature 3-manifold be given by AdS
3
,
parameterised as a U(1) bration over AdS
2
. The Hopf reduction this 6-dimensional
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 143
vacuum along the U(1) bre direction yields a vacuum solution of D = 5 supergravity
that is part of the Weyl multiplet depicted in Figure 8.26.
M
2
K
3
F
(3)i
: u
F
(2)ij
: u
T
j
(2)
: u
M
2
K
3
F
(2)ij
: u
F
(2)kl
: u
F
(2)lm
: u
Figure 8.26: The M = 45 Weyl multiplet of M
2
K
3
vacua.
The scalar curvatures of the submanifolds in Figure 8.26 are
(M
2
) = u
2
, (K
3
) =
1
4
u
2
. (8.6.1)
These vacuum solution of D = 5 supergravity are therefore of the form AdS
2
S
3
.
Alternatively, we can consider the 3 + 3 vacuum solution in D = 6 with the
Euclidean signature 3-manifolds given by S
3
, parameterised as a U(1) bration over
S
2
. The Hopf reduction this 6-dimensional vacuum along the U(1) bre direction
yields a vacuum solution of D = 5 supergravity that is part of the Weyl multiplet
depicted in Figure 8.27.
M
3
K
2
F
(3)i
: u
F
(2)ij
: u
T
j
(2)
: u
M
3
K
2
F
(2)ij
: u
F
(2)kl
: u
F
(2)lm
: u
Figure 8.27: The M = 45 Weyl multiplet of M
3
K
2
vacua.
The scalar curvatures of the submanifolds in Figure 8.27 are
(M
3
) =
1
4
u
2
, (K
2
) = u
2
. (8.6.2)
These vacuum solutions of D = 5 supergravity are therefore of the form AdS
2
S
3
.
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 144
8.7 D = 4
The p-form eld strengths in D = 4 are F
(2)ab
and T
a
(2)
, where a = 1, . . . , 7. The CJ
global symmetry group in D = 4 is E
7
. The 2-form eld strengths F
(2)ab
and T
a
(2)
,
together with their duals F
ab
(2)
and T
(2)a
are in the 56 representation of E
7
.
M
2
K
2
vacua
As we are seeking vacuum solutions of D = 4 supergravity that a priori cannot be
toroidally reduced to D < 4, we only need to consider the Freund-Rubin ansatze
based on the 2 +2 decomposition of the 5-dimensional spacetime. It turns out that
all the 2 + 2 vacua are readily interpreted as dimensional reductions of the 2 + 3
vacuum solutions in D = 5, the 2+2+2 vacua in D = 6, or as Weyl transformations
thereof. We shall therefore discuss them directly from a point of view that highlights
their higher-dimensional origin. In the following, the indices i, j, k, l, m, n, p are
understood to refer to an even permutation of 1, 2, 3, 4, 5, 6, 7, so that they are
all distinct.
To begin with, consider the M
3
K
2
vacuum in D = 5 that is depicted in Figure
8.27. Let the Lorentzian signature 3-manifold be given by AdS
3
, parameterised as
a U(1) bration over AdS
2
. Alternatively, we can consider the M
2
K
3
vacuum
solution of D = 5, which is depicted in Figure 8.26, with the Euclidean signature
3-manifolds given by S
3
, parameterised as a U(1) bration over S
2
. The Hopf
reduction of either of these 5-dimensional vacua along the U(1) bre direction yields
a vacuum solution of D = 6 supergravity that is part of the Weyl multiplet depicted
in Figure 8.28.
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 145
M
2
K
2
F
(2)ij
: u
F
(2)ik
: u
T
j
(2)
: u
T
k
(2)
: u
M
2
K
2
F
(2)ij
: u
F
(2)kl
: u
F
(2)lm
: u
T
p
(2)
: u
M
2
K
2
F
(2)ij
: u
F
(2)kl
: u
F
(2)ik
: u
F
(2)jl
: u
M
2
K
2
F
(2)ij
: u
F
(2)kl
: u
F
(2)lm
: u
T
p
(2)
: u
Figure 8.28: The M = 630 Weyl multiplet of M
2
K
2
vacua.
The scalar curvatures of the submanifolds in Figure 8.28 are
(M
3
) = u
2
, (K
3
) = u
2
. (8.7.1)
These vacuum solution of D = 4 supergravity can therefore be identied with AdS
2

S
2
.
Consider next the 2 +2 +2 vacua in D = 6 that is depicted in Figure 8.25. It is
possible to set the scalar curvature of one of the Euclidean signature 2-manifolds to
zero, e.g. by setting v
3
= 0, whence we have a vacuum solution of D = 6 supergravity
of the form AdS
2
S
2
T
2
. The toroidal reduction this 6-dimensional vacuum yields
a vacuum solution of D = 4 supergravity that is part of the Weyl multiplet depicted
in Figure 8.29.
The scalar curvatures of the submanifolds in Figure 8.29 are
(M
2
) =
1
2
v
2
1
, (K
2
) =
1
2
v
2
2
. (8.7.2)
CHAPTER 8. VACUA IN D 9 SUPERGRAVITIES 146
M
2
K
2
F
(2)ij
: v
v
M
2
K
2
T
i
(2)
: v
v
Figure 8.29: The M = 28 Weyl multiplet of M
2
K
2
vacua.
These vacuum solutions of D = 4 supergravity can therefore be identied with
AdS
2
S
2
.
8.8 D = 3
The bosonic sector of the maximal, massless supergravities in D = 3 consists only of
the metric and the scalar sector, given that all possible dualisations of higher degree
potentials have been performed. The constant scalar vacuum solutions of D = 3
supergravity are therefore rather trivial. They simply have a Ricci at metric and
the maximally symmetry solution is 3-dimensional Minkowski space.
Similar comments apply to the constant scalar vacuum solutions of the maximal,
massless supergravity theory in D = 2. The bosonic sector of this theory is obtained
by toroidal reduction of D = 3 supergravity, see e.g. Ref. [83]. It again consists only
of the metric and the scalars. The constant scalar vacua in D = 2 have a Ricci at
metric and the maximally symmetry solution is 2-dimensional Minkowski space.
Chapter 9
Supersymmetric vacua in D 11
9.1 Integrability conditions for Killing spinors
The vacuum solutions that we have classied in the previous chapters are purely
bosonic solutions of maximal supergravity theories with 32 supersymmetries. As
such, they may or may not preserve some fraction of these 32 supersymmetries. The
number of unbroken supersymmetries is given by the number of linearly indepen-
dent, non-trivial solutions of the appropriate Killing spinor equation, obtained by
setting the supersymmetry variations of the fermions in the bosonic supergravity
background to zero. The Killing spinor equation for D = 11 supergravity is re-
viewed in Appendix B.1, where it is also shown how to obtain the Killing spinor
equations for IIA supergravity and the maximal supergravities in 3 D 9 from
the 11-dimensional one by Kaluza-Klein dimensional reduction. The Killing spinor
equation for IIB supergravity is presented in Appendix B.2. These appendices also
contain our conventions regarding the -matrix representations of the D = (1, 10)
and D = (1, 9) Cliord algebras.
147
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 148
The Killing spinor equation is a rst order dierential equation, which typically
states that the appropriate supercovariant derivative of the spinorial parameter of
the supersymmetry transformation vanishes. The commutator of two supercovari-
ant derivatives, which itself is not a dierential operator, is in general non-zero.
However, when acting on a Killing spinor, i.e. a spinor which corresponds to an un-
broken supersymmetry of the bosonic supergravity background, it must vanish. In
D dimensions, this results in a total of
1
2
D(D1) non-trivial algebraic integrability
conditions that Killing spinors have to satisfy
1
. An alternative to solving the Killing
spinor equation explicitly is therefore to evaluate the constraints that are imposed
on Killing spinors by these integrability conditions. If the integrability conditions do
not admit non-trivial solutions, the supergravity background does not admit Killing
spinors and hence breaks all the supersymmetries. Here, we take it that the exis-
tence of a non-trivial solution of the integrability conditions is also sucient proof
of the existence of a non-trivial Killing spinor.
For a generic bosonic supergravity background, it is often rather impractical to
ascertain the existence of Killing spinors by evaluating these integrability conditions.
For the vacuum solutions that we are concerned with here, however, the task of
evaluating the integrability conditions is simplied considerably by the fact that the
metric is block-diagonal and that all other elds are covariantly constant. This is
most easily demonstrated with a number of simple examples.
To begin with, consider the original Freund-Rubin vacuum solutions of D = 11
1
In the lower-dimensional supergravities obtained from D = 11 supergravity by toroidal Kaluza-
Klein reduction, additional algebraic conditions have to be imposed on the Killing spinors in order
to ensure that they are independent of the coordinates that have been compactied.
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 149
supergravity, which are depicted in Figure 6.3. The Killing spinor equation in D = 11
can be written as
D
A
= /
A
,
/
A
=
1
12
1
24
F
B
1
B
4
(
B
1
B
4
A
+ 8
[B
1
B
2
B
3

B
4
]
A
) ,
(9.1.1)
where is the Majorana spinor parameter of the supersymmetry variation and D
A

its (ordinary) covariant derivative.


The integrability conditions for the Killing spinor equation follow from the iden-
tity
[D
A
, D
B
] =
1
4
1
AB
CB

CD
, (9.1.2)
where 1
AB
CB
are the tangent space components of the Riemann curvature tensor.
We now observe that for the Freund-Rubin vacua, the components F
A
1
A
4
of the
4-form eld strength are covariantly constant. Therefore, the matrix /
A
dened
in (9.1.1) is also covariantly constant. This observation allows us to evaluate the
commutator of two covariant derivatives as
[D
A
, D
B
] = [/
A
, /
B
] (9.1.3)
whence (9.1.2) implies that
[/
A
, /
B
] =
1
4
1
AB
CB

CD
. (9.1.4)
Since the 11-dimensional metric is block-diagonal, the integrability conditions
(9.1.4) simplify further to
[/

, /

] =
1
4
1

,
[/
a
, /
b
] =
1
4
1
ab
cd

cd
,
[/

, /
a
] = 0,
(9.1.5)
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 150
where , , . . . and a, b, . . . are tangent space indices for M
4/7
and K
7/4
respectively.
Let us focus rst on the M
4
K
7
vacuum, for which the matrices /
A
evaluate
as
/

=
u
6

0123

, = 0, , 3
/
a
=
u
12

0123

a
, a = 4, , 10
(9.1.6)
where the numerical indices on the -matrices refer to tangent space directions and
u is the scale setting constant of the M
4
K
7
vacuum solution. It is not dicult to
show that (9.1.5) becomes
2u
2
9

= 1

,
u
2
18

ab
= 1
ab
cd

cd
. (9.1.7)
At this stage, we have to be more specic about the Einstein submanifolds. For
maximally symmetric submanifolds, i.e. M
4
= AdS
4
and K
7
= S
7
, with scalar cur-
vatures given in (6.3.1), the non-zero components of the Riemann curvature tensor
are
1

=
2u
2
9

]
, 1
ab
cd
=
u
2
18

a
[c

b
d]
. (9.1.8)
It is apparent that the integrability conditions (9.1.7) are satised identically for
this choice of submanifolds. In eect, we have just re-derived the result that the
AdS
4
S
7
vacuum solution of D = 11 supergravity preserves the maximal number
of supersymmetries.
The M
7
K
4
vacuum is treated analogously. The matrices /
A
now evaluate as
/

=
u
12

789

, = 0, , 6
/
a
=
u
6

789

a
, a = 7, , 10
(9.1.9)
where the numerical indices on the -matrices refer to tangent space directions and
denotes the number 10 and u is the scale setting constant of the M
7
K
4
vacuum
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 151
solution. We nd that the integrability conditions (9.1.5) become
u
2
18

= 1

,
2u
2
9

ab
= 1
ab
cd

cd
. (9.1.10)
If we choose the submanifolds M
7
and K
4
to be maximally symmetric, with
scalar curvatures given in (6.3.1), the non-zero components of the Riemann curvature
tensor are
1

=
u
2
18

]
, 1
ab
cd
=
2u
2
9

a
[c

b
d]
. (9.1.11)
The integrability conditions (9.1.10) are again satised identically for this choice of
submanifolds. Hence, we have just veried that the AdS
4
S
7
vacuum solution of
D = 11 supergravity also preserves the maximal number of supersymmetries.
Another simple example is provided by the M
5
K
5
vacuum solution of IIB
supergravity, which is depicted in Figure 7.11. The relevant Killing spinor equation
can be written as
D
A
= /
A
, /
A
=
i
8
1
120
H
B
1
B
5

B
1
B
5
A
, (9.1.12)
where is the Weyl spinor parameter of the supersymmetry variation, satisfying

= ,

=
019
, (9.1.13)
and D
A
its (ordinary) covariant derivative.
For the M
5
K
5
vacuum, the components H
B
1
B
5
of the 5-form eld strength
are covariantly constant and the 10-dimensional metric is block-diagonal. It follows
that the integrability conditions for the Killing spinor equation are of the form
(9.1.5), where , , . . . and a, b, . . . now denote tangent space indices for M
5
and K
5
respectively.
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 152
In the M
5
K
5
IIB supergravity background, we nd that the matrices /
A
are
equal to
/

= i
u
8

56789

, = 0, , 4
/
a
= i
u
8

01234

a
, a = 5, , 9
(9.1.14)
where the numerical indices on the -matrices refer again to tangent space directions
and u is the scale setting constant of the M
5
K
5
vacuum solution. It is not dicult
to show that (9.1.5) becomes
u
2
8

= 1

,
u
2
8

ab
= 1
ab
cd

cd
. (9.1.15)
For maximally symmetric submanifolds, i.e. M
5
= AdS
5
and K
5
= S
5
, with
scalar curvatures given in (7.2.6), the non-zero components of the Riemann curvature
tensor are
1

=
u
2
8

]
, 1
ab
cd
=
u
2
8

a
[c

b
d]
. (9.1.16)
It is evident that the integrability conditions (9.1.15) are satised identically for
this choice of submanifolds. This conrms the well known result that the AdS
5
S
5
vacuum solution of IIB supergravity preserves the maximal number of supersymme-
tries.
The integrability conditions for the existence of Killing spinors in the super-
gravity background of a generic D-dimensional vacuum solution can be analysed
similarly. As shown in Appendix B.1, the Killing spinor equation for IIA supergrav-
ity in D = 10, or the maximal supergravity theory in 3 D 9, can be written
as
D
A
= /
A
, A = 0, 1, , (D 1),
/
a
= 0, a = 1, , (11 D),
(9.1.17)
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 153
The integrability conditions for the Killing spinor equation follow from imposing
[D
A
, D
B
] =
1
4
1
AB
CD

CD
. (9.1.18)
In the supergravity background of each of the vacuum solutions that we have
classied in the previous chapters, the scalar elds vanish. In that case, /
A
and
/
a
are given by
/
A
=
1
4(D2)
1
24
F
C
1
C
4
_
3
C
1
C
4
A
+ 4(D 5)
[C
1
C
2
C
3

C
4
]
A
_
(9.1.19)
+
1
4(D2)
1
6

a
F
a
C
1
C
2
C
3
_
2
C
1
C
2
C
3
A
+ 3(D 4)
[C
1
C
2

C
3
]
A
_

1
4(D2)
1
2

a<b
F
ab
C
1
C
2
_

C
1
C
2
A
+ 2(D 3)
[C
1

C
2
]
A
_

ab

+
1
4(D2)

a
T
a
C
1
C
2
_

C
1
C
2
A
+ 2(D 3)
[C
1

C
2
]
A
_

a
,
/
a
=
1
2
T
C
1
C
2
a

C
1
C
2
+
1
3
1
24
F
C
1
C
4

C
1
C
4

a
(9.1.20)
+
1
3
1
6

d
F
d
C
1
C
2
C
3

C
1
C
2
C
2
(
da
+ 2
da
)
+
1
3
1
4

c,d
F
cd
C
1
C
2

C
1
C
2
(
cda
+ 4
c

da
) .
In addition, the components of the p-form eld strengths are covariantly constant
and the metric is block-diagonal for each constant scalar vacuum. Assume that the
D-dimensional spacetime is the product of N submanifolds, and let A
i
, B
i
, . . . be
indices for the i
th
submanifold. The identity (9.1.18) then implies the integrability
conditions
[/
A
i
, /
B
i
] =
1
4
1
A
i
B
i
C
i
D
i

C
i
D
i
, i 1, , N,
[/
A
i
, /
A
j
] = 0, i ,= j 1, , N.
(9.1.21)
In the following sections, we determine the amount of unbroken supersymmetry
for each of the constant scalar vacuum solutions that we have classied, by evaluating
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 154
the integrability conditions (9.1.21) of the relevant Killing spinor equation. This task
is greatly simplied by two observations:
1. Kaluza-Klein dimensional reduction does not break any supersymmetries, im-
plying that a lower-dimensional vacuum preserves as many supersymmetries
as the higher-dimensional vacuum to which it oxidises
2
.
2. Weyl transformations commute with supersymmetry transformations [78], im-
plying that dierent vacua that are part of the same Weyl multiplet preserve
the same the number of supersymmetries.
As has been highlighted in the previous chapters, most of the lower-dimensional
vacua can indeed be obtained by toroidal or Hopf reduction of a higher-dimensional
vacuum, or are in the same Weyl multiplet as a vacuum solution that oxidises to a
higher-dimensional vacuum. For these vacuum solutions, we do not need to deter-
mine the Killing spinor integrability conditions explicitly. If, on the one hand, the
higher-dimensional vacuum breaks all the supersymmetries, then so does the lower-
dimensional one. If, on the other hand, the higher-dimensional vacuum admits
non-trivial Killing spinors, the unbroken supersymmetries of the lower-dimensional
vacuum correspond to those (higher-dimensional) Killing spinors that are indepen-
dent of the compactied coordinates.
2
Although Kaluza-Klein dimensional reduction does not break any supersymmetries per se, it
might well be that some of the higher-dimensional Killing spinors depends explicitly on the coor-
dinates of the compactifying torus. The corresponding supersymmetries of the higher-dimensional
solution are broken by requiring that the lower-dimensional Killing spinor be independent of the
toroidal coordinates that have been compactied. See e.g. Refs. [84, 91].
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 155
9.2 Supersymmetric vacua in D = 11
We begin our search for supersymmetric vacuum solutions of D = 11 supergravity
by considering again the 4 +7 vacua depicted in Figure 6.3. In the previous section,
we have shown that the integrability conditions for the Killing spinor equation in
these backgrounds can be written as
M
4
K
7
: M
7
K
4
:
2u
2
9

= 1

,
u
2
18

= 1

,
u
2
18

ab
= 1
ab
cd

cd
,
2u
2
9

ab
= 1
ab
cd

cd
,
(9.2.1)
where , , . . . and a, b, . . . are tangent space indices for M
4/7
and K
7/4
respectively.
We have also shown that if the Einstein submanifolds are maximally symmetric, the
integrability conditions (9.2.1) are satised identically, so that the AdS
4/7
S
7/4
vacua are also maximally supersymmetric.
In Chapter 6, we have also found vacuum solutions of D = 11 supergravity that
dier from the 4 + 7 vacua in the choice of Einstein submanifolds. In particular,
the 4-manifold may itself be the product of two Einstein 2-manifolds and/or the
7-manifold may itself be the product of an Einstein 3-manifold and an Einstein 4-
manifold. Here, we want to argue briey that all these vacua necessarily break all
the supersymmetries.
The integrability conditions for the Killing spinor equation in these backgrounds
are again of the form (9.2.1), since the 4-form eld conguration is the same for
all 4 + 7 vacua, irrespective of the choice of Einstein manifolds. We now observe
that if and refer to distinct submanifolds of M
4/7
, the 1

components of
the Riemann tensor vanish. Thus, the Killing spinor has to satisfy

= 0, which
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 156
does not admit non-trivial solutions since

is invertible. Similarly, if a and b refer


to distinct submanifolds of K
7/4
, the Killing spinor has to satisfy
ab
= 0, which
does not admit non-trivial solutions either since
ab
is invertible. It follows that the
4 + 7 vacuum solutions of D = 11 supergravity for which either the 4-manifold or
the 7-manifold is itself the product of two submanifolds are not supersymmetric.
Let us now turn to the 2 + 2 + 2 + 5 vacuum solutions of D = 11 supergravity,
which are depicted in Figure 6.8. We shall show presently that they break all the
supersymmetries, irrespective of the choice of Einstein submanifolds.
Consider rst the M
2
K
2
K
2
K
5
vacuum. We nd that the matrices /
A
,
dened in (9.1.1), are given by
/

=
1
12
(u
1

2345
2u
2

0145
2u
3

0123
)

, = 0, 1,
/
a
=
1
12
(2u
1

2345
u
2

0145
2u
3

0123
)
a
, a = 2, 3,
/
a
=
1
12
(2u
1

2345
2u
2

0145
u
3

0123
)
a
, a

= 4, 5,
/
a
=
1
12
(u
1

2345
u
2

0145
u
3

0123
)
a
, a

= 6, 7, , 10,
(9.2.2)
where , a, a

and a

are tangent space indices for M


2
, K
2
, K
2
and K
5
respectively.
Since the 11-dimensional metric is block-diagonal, the Killing spinors in the M
2

K
2
K
2
K
5
vacuum background have to satisfy, among other things,
[/

, /
a
] = [/

, /
a
] = [/

, /
a
] = 0. (9.2.3)
It is not dicult to show that (9.2.3) leads to three independent conditions on the
Killing spinor, which are equivalent to requiring that the matrices
A = u
2
1
2u
1
u
2

0123
2u
1
u
3

0145
,
B = u
2
2
+ 2u
1
u
2

0123
+ 2u
2
u
3

2345
,
C = u
2
3
2u
1
u
3

0145
+ 2u
2
u
3

2345
(9.2.4)
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 157
admit zero eigenvectors.
To determine the eigenvalues of A, B and C, we use the Hamilton-Cayley the-
orem, which states the a matrix satises its own characteristic equation. A short
calculation establishes that
_
(Au
2
1
)
2
+ 4u
2
1
(u
2
2
+u
2
3
)
2
_
2
= 64u
4
1
u
2
2
u
2
3
,
_
(B u
2
2
)
2
4u
2
2
(u
2
3
u
2
1
)
2
_
2
= 64u
4
2
u
2
1
u
2
3
,
_
(C u
2
3
)
2
4u
2
3
(u
2
2
u
2
1
)
2
_
2
= 64u
4
3
u
2
2
u
2
2
,
(9.2.5)
The Hamilton-Cayley theorem allows us to replace A, B and C in (9.2.5) by their
eigenvalues. We then nd that A, B and C admit zero eigenvectors for real values
of u
1
, u
2
and u
3
only if
u
1
= 0, u
2
2
= 4u
2
3
, u
2
3
= 4u
2
2
, (9.2.6)
which does not admit non-trivial solutions. Hence, the integrability conditions for
the Killing spinor equation in the M
2
K
2
K
2
K
5
background cannot be satised.
We continue to present a similar analysis of the Killing spinor integrability condi-
tions in the M
5
K
2
K
2
K
2
vacuum background. We now nd that the matrices
/
A
are given by
/

=
1
12
(u
1

789
+u
2

569
+u
3

5678
)

, = 0, 4,
/
a
=
1
12
(u
1

789
+ 2u
2

569
+ 2u
3

5678
)
a
, a = 5, 6,
/
a
=
1
12
(2u
1

789
+u
2

569
+ 2u
3

5678
)
a
, a

= 7, 8,
/
a
=
1
12
(2u
1

789
+ 2u
2

569
+u
3

5678
)
a
, a

= 9, 10,
(9.2.7)
where , a, a

and a

are tangent space indices for M


5
, K
2
, K
2
and K
2
respectively.
The Killing spinors in the M
5
K
2
K
2
K
2
vacuum background again have
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 158
to satisfy at least the conditions (9.2.3), which in the present case lead to the re-
quirement that the matrices
A = u
2
1
+ 2u
1
u
2

5678
+ 2u
1
u
3

569
,
B = u
2
2
+ 2u
1
u
2

5678
+ 2u
2
u
3

789
,
C = u
2
3
2u
1
u
3

569
+ 2u
2
u
3

789
(9.2.8)
admit zero eigenvectors. It is not dicult to show that these matrices satisfy
_
(Au
2
1
)
2
4u
2
1
(u
2
2
+u
2
3
)
2
_
2
= 64u
4
1
u
2
2
u
2
3
,
_
(B u
2
2
)
2
4u
2
2
(u
2
3
+u
2
1
)
2
_
2
= 64u
4
2
u
2
1
u
2
3
,
_
(C u
2
3
)
2
4u
2
3
(u
2
2
+u
2
1
)
2
_
2
= 64u
4
3
u
2
2
u
2
2
,
(9.2.9)
The Hamilton-Cayley theorem allows us to replace A, B and C in (9.2.9) by their
eigenvalues. We then nd that A, B and C admit zero eigenvectors for real values
of u
1
, u
2
and u
3
only if
u
1
= 2(u
2
u
3
), u
2
= 2(u
3
u
1
), u
3
= 2(u
1
u
2
), (9.2.10)
which does not admit non-trivial solutions. Hence, the integrability conditions for
the Killing spinor equation in the M
5
K
2
K
2
K
2
background cannot be satised.
We have just established that both the M
2
K
2
K
2
K
5
and the M
5
K
2
K
2

K
2
vacuum solutions of D = 11 supergravity necessarily break all supersymmetries,
irrespective of the choice of Einstein submanifolds. An important corollary of this
result is that all lower-dimensional vacuum solutions that can be obtained from the
M
2/5
K
2
K
2
K
5/2
vacua by a combination of Kaluza-Klein dimensional reduction
and lower-dimensional Weyl transformations are not supersymmetric either. These
lower-dimensional vacua, and their oxidation pathways to the M
2/5
K
2
K
2
K
5/2
vacuum, are depicted schematically in Figure 9.1.
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 159
M
2/5
K
2
K
2
K
5/2
(Fig. 6.8)
H T
M
2
K
2
K
2
K
2
K
2
(Fig. 7.9)
T
T
M
2/5
K
2
K
5/2
(Fig. 8.1)
T H
M
2
K
2
K
2
K
2
(Fig. 8.14)
Figure 9.1: The M
2/5
K
2
K
2
K
5/2
tree of non-supersymmetric vacua.
In Figure 9.1, H refers to a Hopf reduction step from D + 1 to D dimensions,
while T refers to a toroidal reduction step from D+1 to D dimensions. Both types
of Kaluza-Klein reduction, of course, presume a suitable choice of submanifold in
the higher-dimensional vacuum that is being dimensionally reduced. However, as
argued above, the choice of submanifold does not eect the supersymmetry of the
higher-dimensional vacuum solution, so that all the vacua in Figure 9.1 are non-
supersymmetric.
The remaining vacuum solutions of D = 11 supergravity are of the form M
2/3

K
2
K
2
K
2
K
3/2
. They are depicted in Figure 6.9. Let i = 1, , 5 label the
submanifold and let A
i
, B
i
, . . . be tangent space indices for the i
th
submanifold. The
Killing spinor integrability conditions in the M
2/3
K
2
K
2
K
2
K
3/2
vacuum
background can be written as
[/
A
i
, /
B
i
] =
1
4
1
A
i
B
i
C
i
D
i

C
i
D
i
, i 1, , 5
[/
A
i
, /
A
j
] = 0, i ,= j 1, , 5,
(9.2.11)
where the matrices /
A
are dened in (9.1.1). They depend on the parameters u

CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 160


that dene the 2 + 2 + 2 + 2 + 3 vacua, but will not be given explicitly here.
Recall that any Einstein 2-manifold is trivially maximally symmetric, while any
Einstein 3-manifold is maximally symmetric because the Weyl tensor vanishes in
three dimensions. Thus, the non-zero components of the Riemann curvature tensor
for the 2 + 2 + 2 + 2 + 3 vacua are
1
A
i
B
i
C
i
D
i
= 2
i

A
i
[C
i

B
i
D
i
]
, (9.2.12)
where
i
is the scalar curvature of the i
th
submanifold. The values of the
i
are
given in (6.3.5). It is possible to show that the commutators in (9.2.11) are of the
form
[/
A
i
, /
B
i
] =
A
i
B
i
/
ii
, [/
A
i
, /
A
j
] =
A
i
A
j
/
ij
, (9.2.13)
Since the matrices
AB
are all invertible, the 55 integrability conditions (9.2.11)
reduce to the 15 independent integrability conditions

/
ij

_
/
ij
+
1
2

i

ij
_
= 0, i, j 1, , 5. (9.2.14)
It is rather complicated to evaluate these integrability conditions any further
using analytical techniques. We have used an explicit representation of the (3232)
-matrices in D = (1, 10) to calculate the matrices

/
ij
in (9.2.14) with the help
of a computer. By evaluating the determinants of these matrices, we have veried
that it is not possible to choose the (real) parameters u

of the 2 + 2 + 2 + 2 + 3
vacua such that every matrix

/
ij
admits zero eigenvectors.
We have thus established that the M
2/3
K
2
K
2
K
2
K
3/2
vacuum solution of
D = 11 supergravity break all supersymmetries. In addition, we may conclude that
all lower-dimensional vacuum solutions that can be obtained from these vacua by
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 161
a combination of Kaluza-Klein dimensional reduction and lower-dimensional Weyl
transformations are not supersymmetric either. These lower-dimensional vacua, and
their oxidation pathways to the M
2/3
K
2
K
2
K
2
K
3/2
vacuum, are depicted
schematically in Figure 9.2.
M
2/3
K
2
K
2
K
2
K
3/2
(Fig. 6.9)
H T
M
2
K
2
K
2
K
2
K
2
(Fig. 7.8)
T
T
M
2/3
K
2
K
2
K
3/2
(Fig. 8.7)
T H T
M
2
K
2
K
2
K
2
(Fig. 8.16/Fig. 8.15)
T
T
M
2/3
K
2
K
3/2
(Fig. 8.20/Fig. 8.19)
T H
M
2
K
2
K
2
(Fig. 8.23/Fig. 8.24)
Figure 9.2: The M
2/3
K
2
K
2
K
2
K
3/2
tree of non-supersymmetric vacua.
The M
2
K
2
K
2
K
2
K
3
vacuum solution can also be toroidally reduced to
the M
2
K
2
K
2
K
2
vacuum depicted in Figure 8.12, if the 3-manifold is chosen
to be the 3-torus.
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 162
9.3 Supersymmetric vacua in D = 10
IIA supergravity
We have demonstrated in Section 7.1 that almost
3
every Freund-Rubin vacuum so-
lution of IIA supergravity oxidises to one of the Freund-Rubin vacua in D = 11.
It is therefore economical to oxidise the IIA vacua to D = 11 and evaluate the
integrability conditions for the Killing spinors in the appropriate D = 11 vacuum
background. For each 11-dimensional Killing spinor that is independent of the co-
ordinate that is being compactied in the reduction from D = 11 to D = 10, there
exists an unbroken supersymmetry of the corresponding IIA vacuum.
Consider rst the M
2/4
K
4
K
4/2
vacuum solutions of IIA supergravity de-
picted in Figure 7.1. They oxidise to M
4/7
K
7/4
vacua in D = 11 for which the
Einstein 7-manifold is itself the product of an Einstein 3-manifold and an Einstein
4-manifold. We have shown in Section 9.2 that these 11-dimensional vacua break
all supersymmetries. It follows that their 10-dimensional descendants are also non-
supersymmetric.
Consider next the M
2/4
K
2
K
2
K
4/2
vacua depicted in Figure 7.4. They
oxidise to M
4/7
K
7/4
vacua in D = 11 for which the Einstein 7-manifold is a U(1)
bration over the 3-fold product of Einstein 2-manifolds. In the notation of Ap-
pendix C.1, M
7
=

Q
7
(v
1
, v
2
, v
3
) or K
7
= Q
7
(v
1
, v
2
, v
3
). The part of the integrability
conditions of the 11-dimensional Killing spinor equation that pertain to the choice
3
The exception is the AdS
2
S
2
S
2
S
2
S
2
vacuum in Figure 7.10, which oxidises to a Ricci
at U(1) bration in D = 11.
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 163
of the 7-manifold can be written as (c.f. equation (9.2.1))
2
A

B
=

1
A

C

D

C

D
, (9.3.1)
where

A,

B, . . . are tangent space indices for 7-manifold and is its scalar curvature.
Killing spinor integrability conditions of the form (9.3.1) for the 7-manifolds

Q
7
(v
1
, v
2
, v
3
) and Q
7
(v
1
, v
2
, v
3
) have been analysed in detail in Appendix C.2. There,
it is shown that the Lorentzian signature U(1) bration

Q
7
(v
1
, v
2
, v
3
) (with negative
scalar curvature) necessarily breaks all the supersymmetries. It follows that the
M
2
K
2
K
2
K
4
vacuum solution of IIA supergravity, which is obtained by Hopf
reducing the

Q
7
(v
1
, v
2
, v
3
) K
4
vacuum solution of D = 11 supergravity along the
U(1) bre direction, is non-supersymmetric.
In contrast, the Euclidean signature U(1) bration Q
7
(v
1
, v
2
, v
3
) (with positive
scalar curvature) preserves
1
4
of the maximal number of supersymmetries (in D = 7),
provided that v
1
= v
2
= v
3
= 2

v. Therefore, an 11-dimensional vacuum


solution of the form M
4
Q
7
(v, v, v) may preserve
1
4
of the maximal number of
supersymmetries (in D = 11), depending on the choice of M
4
. However, the results
of Appendix C.2 imply that the 11-dimensional Killing spinors in a supersymmetric
M
4
Q
7
(v, v, v) vacuum background necessarily depend on the U(1) bre coordi-
nate. It follows that all the supersymmetries are lost in the Hopf reduction of this
vacuum to D = 10, and hence that the M
4
K
2
K
2
K
2
vacuum solutions of IIA
supergravity are non-supersymmetric.
Of the remaining IIA vacua that have been classied in Section 7.1, there is only
the M
2
K
2
K
2
K
2
K
2
vacuum depicted in Figure 7.10 that does not oxidise
to a D = 11 vacuum that has already been shown to be non-supersymmetric. The
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 164
oxidation of this IIA vacuum is interpreted in D = 11 as a Ricci at U(1) bra-
tion over the base AdS
2
S
2
S
2
S
2
S
2
. As shown in Appendix C.2, such
Ricci at U(1) brations are quite generally not supersymmetric. Therefore, the
corresponding M
2
K
2
K
2
K
2
K
2
vacuum solution of IIA supergravity is
not supersymmetric either. By implication, all lower-dimensional vacua that can be
obtained from this IIA vacuum via a combination of Kaluza-Klein dimensional re-
duction an lower-dimensional Weyl-transformations are non-supersymmetric. These
lower-dimensional vacua, and their oxidation pathways to the M
2
K
2
K
2
K
2
K
2
vacuum, are depicted schematically in Figure 9.3.
M
2
K
2
K
2
K
2
K
2
(Fig. 7.10)
T
T
M
2
K
2
K
2
K
2
(Fig. 8.13)
T
T
M
2
K
2
K
2
(Fig. 8.25)
T
T
M
2
K
2
(Fig. 8.29)
Figure 9.3: The M
2
K
2
K
2
K
2
K
2
tree of non-supersymmetric vacua.
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 165
IIB supergravity
We have already discussed the integrability conditions for the Killing spinor equation
in the M
5
K
5
vacuum background and conrmed that the AdS
5
S
5
vacuum
solution of IIB supergravity is maximally supersymmetric. Let us therefore turn
directly to the 2+2+3+3 vacuum solutions of IIB supergravity, which are depicted
in Figure 7.12. We shall show presently that they break all the supersymmetries.
Consider rst the M
2
K
2
K
3
K
3
vacuum. We nd that the matrices /
A
,
dened in (9.1.12), are given by
/

=
i
8
(u
1

23789
u
2

23456
)

, = 0, 1,
/
a
=
i
8
(u
1

01456
+u
2

01789
)
a
, a = 2, 3,
/
a
=
i
8
(u
1

23789
u
2

01789
)
a
, a

= 4, 5, 6,
/
a
=
i
8
(u
1

01456
+u
2

23456
)
a
, a

= 7, 8, 9,
(9.3.2)
where , a, a

and a

are tangent space indices for M


2
, K
2
, K
3
and K
3
respectively.
Since the 10-dimensional metric is block-diagonal, the Killing spinors in the
M
2
K
2
K
3
K
3
vacuum background have to satisfy, among other things,
[/

, /
a
] = [/

, /
a
] = 0. (9.3.3)
Keeping in mind the chirality condition (9.1.13), one can show that (9.3.3) leads to
two independent constraints on the Killing spinor, which are equivalent to requiring
that the matrices
A = u
2
1
u
1
u
2

0123
, (Au
2
1
)
2
= u
2
1
u
2
2
,
B = u
2
2
+u
1
u
2

0123
, (B u
2
2
)
2
= u
2
1
u
2
2
,
(9.3.4)
admit zero eigenvectors. The Hamilton-Cayley theorem allows us to replace A and
B in the second column of (9.3.4) by their eigenvalues. It is evident that A and B
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 166
do not admit zero eigenvectors for non-zero real values of u
1
and u
2
. Hence, the
integrability conditions for the Killing spinor equation in the M
2
K
2
K
3
K
3
background cannot be satised.
We continue to present a similar analysis of the Killing spinor integrability condi-
tions in the M
3
K
3
K
2
K
2
vacuum background. We now nd that the matrices
/
A
are given by
/

=
i
8
(u
1

34589
+u
2

34567
)

, = 0, 1, 2
/
a
=
i
8
(u
1

01267
+u
2

01289
)
a
, a = 3, 4, 5
/
a
=
i
8
(u
1

34589
u
2

01289
)
a
, a

= 6, 7
/
a
=
i
8
(u
1

01267
u
2

34567
)
a
, a

= 8, 9,
(9.3.5)
where , a, a

and a

are tangent space indices for M


3
, K
3
, K
2
and K
2
respectively.
The Killing spinors in the M
3
K
3
K
2
K
2
vacuum background again have
to satisfy at least the conditions (9.3.3), which in the present case lead to the re-
quirement that the matrices
A = u
2
1
+u
1
u
2

6789
, (Au
2
1
)
2
= u
2
1
u
2
2
,
B = u
2
2
+u
1
u
2

6789
, (B u
2
2
)
2
= u
2
1
u
2
2
,
(9.3.6)
admit zero eigenvectors. The Hamilton-Cayley theorem allows us to replace A and B
in the second column of (9.3.6) by their eigenvalues. It is evident that A and B admit
zero eigenvectors only if u
1
= u
2
. In that case, however, the IIB vacuum solution is
of the form M
3
K
3
T
4
, c.f. equation (7.2.7). One can show that the 6-dimensional
vacuum solution, obtained by compactifying the toroidal directions, is part of the
Weyl multiplet of M
3
K
3
vacua depicted in Figure 8.21. The supersymmetry of
the M
3
K
3
vacua will be discussed below. Here, we simply note that if we require
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 167
u
1
,= u
2
, so that the IIB vacuum cannot be toroidally reduced to D < 10, then the
integrability conditions for the Killing spinor equation in the M
3
K
3
K
2
K
2
background cannot be satised.
9.4 Supersymmetric vacua in D 9
Every vacuum solution of the lower-dimensional supergravities in D 9 that does
not oxidise to a non-supersymmetric vacuum in D = 11 can be obtained from one
of the M
3
K
3
K
3
vacua in D = 9 by a combination of Kaluza-Klein dimensional
reduction and lower-dimensional Weyl transformations. This is indicated schemati-
cally in Figures 9.4 and 9.5, where we have identied the Einstein 3-manifolds and
the Einstein 2-manifolds with the appropriate maximally symmetric spaces.
(AdS
3
S
3
S
3
)1
2
(Fig. 8.6)
H H
(AdS
2
S
3
S
3
)1
4
(Fig. 8.9)
(AdS
3
S
3
S
2
)1
4
(Fig. 8.11)
H H H
(AdS
2
S
2
S
3
)1
4
(Fig. 8.17)
(AdS
3
S
2
S
2
)1
4
(Fig. 8.18)
H H
(AdS
2
S
2
S
2
)1
4
(Fig. 8.22)
Figure 9.4: The M
3
K
3
K
3
tree of supersymmetric vacua, Part I.
As has been mentioned already, the mapping that takes any one of the lower-
dimensional vacua in Figure 9.4 or in Figure 9.5 to the appropriate M
3
K
3
K
3
vacuum in D = 9 is supersymmetry preserving, since it consists entirely of a combi-
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 168
(AdS
3
S
3
T
3
)1
2
(Fig. 8.6)
T
T
T
(AdS
3
S
3
)1
2
(Fig. 8.21)
H H
(AdS
2
S
3
)1
4
(Fig. 8.26)
(AdS
3
S
2
)1
4
(Fig. 8.27)
H H
(AdS
2
S
2
)1
4
(Fig. 8.28)
Figure 9.5: The M
3
K
3
K
3
tree of supersymmetric vacua, Part II.
nation of lower-dimensional Weyl transformations and Kaluza-Klein oxidation. To
be more precise, the lower-dimensional vacua preserve at most as many supersym-
metries as the M
3
K
3
K
3
vacuum to which the oxidise (or to which they oxidise
after the appropriate Weyl transformation). The lower-dimensional Killing spinors
correspond to the subset of 9-dimensional Killing spinors that are independent of the
coordinates that are being compactied in the Kaluza-Klein reduction from D = 9
to D < 9 dimensions.
We shall show presently that the integrability conditions for the Killing spinor
equation in the M
3
K
3
K
3
vacuum background are satised, provided that a
projection condition is imposed on the Killing spinors. This projection condition re-
duces the number of linearly independent Killing spinors by a factor of one half. The
M
3
K
3
K
3
vacua therefore preserve half of maximal number of rigid supersymme-
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 169
tries. We shall then identify the dependence of the Killing spinors on the coordinates
that are being compactied in the Kaluza-Klein reduction from D = 9 to D < 9.
In particular, we nd that the Killing spinors may be taken to be independent of
the toroidal reduction directions, while they are independent of the Hopf reduction
directions only if an additional projection condition is satised. This establishes
that the lower-dimensional vacua that are obtained from the M
3
K
3
K
3
vacua
by toroidal reduction steps only (and possibly lower-dimensional Weyl transforma-
tions) also preserve one half of the supersymmetries, while the lower-dimensional
vacua that are obtained form the M
3
K
3
K
3
vacua by Kaluza-Klein reduction
that includes one or more Hopf reduction steps (and possibly lower-dimensional Weyl
transformations) preserve one quarter of the supersymmetries. The fraction of un-
broken supersymmetries in these vacuum backgrounds is indicated by the bold-face
subscripts in Figures 9.4 and 9.5.
M
3
K
3
K
3
F
(3)1
: u
1

u
2

u
3

Figure 9.6: Representative of the M
3
K
3
K
3
multiplet of vacua in D = 9.
To determine the residual supersymmetries of the Weyl multiplet of M
3
K
3
K
3
vacua, consider the particular M
3
K
3
K
3
vacuum depicted in Figure 9.6, where
u
2
1
= u
2
2
+u
2
3
. (9.4.1)
The integrability conditions for the Killing spinor equation in this 9-dimensional
vacuum background are of the general form (9.1.21). In the present context, the
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 170
non-zero commutators are
[/

, /

] =
u
2
1
8

, , = 0, 1, 2,
[/
a
, /
b
] =
u
2
2
8

ab
, a, b = 3, 4, 5,
[/
a
, /
b
] =
u
2
3
8

a

b
, a

, b

= 6, 7, 8,
(9.4.2)
where , are tangent space indices for M
3
and a, b, and a

, b

are tangent space in-


dices for the two Euclidean signature 3-manifolds. We have also used the expressions
for the scalar curvatures of the Einstein 3-manifolds, as given in (8.2.4).
The Killing spinors have to satisfy two additional conditions which ensure that
they are independent of the coordinates that have been compactied in the Kaluza-
Klein reduction from D = 11 to D = 9. These conditions can be written as
/
1
= /
2
= 0 (9.4.3)
where the matrices /
1
and /
2
, dened in (9.1.20), are given by
/
1
=
2
3
(u
1

012
+u
2

345
+u
3

678
) , (9.4.4)
/
2
=
1
3
(u
1

012
+u
2

345
+u
3

678
)
9
. (9.4.5)
It is not dicult to see that (9.4.3) reduce to the one independent projection
condition
T = 0, T =
1
2
_
1
u
1
u
1

012345

u
3
u
1

012678
_
. (9.4.6)
Recalling that u
2
1
= u
2
2
+ u
2
3
, we see that the operator T is indeed a projector, i.e.
T
2
= T.
Let us now return to the integrability conditions (9.4.2). It can be shown that
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 171
the matrices /
A
, dened in (9.1.19), are equivalent to
/

=
u
1
4

012

,
/
a
=
u
2
4

345

a
,
/
a
=
u
3
4

678

a
,
(9.4.7)
when restricted to act on the space of spinors satisfying the projection condition
(9.4.6). A short calculation is then sucient to verify that the integrability condi-
tions (9.4.2) are satised without further constraints on the Killing spinor.
In summary, we have shown that the integrability conditions for the Killing
spinor equation in the M
3
K
3
K
3
vacuum background, supported by the eld
strength F
(3)1
, are satised, provided that the Killing spinor satises the projection
condition (9.4.6). Since the projector T has rank 16, it admits 32 16 = 16 zero
eigenvectors. The particular M
3
K
3
K
3
vacuum solution that we have considered,
and hence the entire Weyl multiplet of M
3
K
3
K
3
vacua, therefore preserves one
half of the maximum number of supersymmetries.
We continue to identify the dependence of the 9-dimensional Killing spinors on
the coordinates that are being compactied in Kaluza-Klein reduction from D = 9
to D < 9. To this end, we need to consider the actual Killing spinor equation, which
can be written as
D

=
u
1
4

012

, = 0, 1, 2
D
a
=
u
2
4

345

a
, a = 3, 4, 5
D
a
=
u
3
4

678

a
, a

= 6, 7, 8
(9.4.8)
for spinors satisfying the projection condition (9.4.6).
Assume initially that u
1
, u
2
and u
3
are all non-zero, in which case the 9-dimensional
vacuum is of the form AdS
3
S
3
S
3
. We can consider each of these 3-manifolds
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 172
independently parameterised as a U(1) bration over the appropriate 2-manifold.
Let 2 refer to the U(1) bre direction in the AdS
3
, and 5 and 8 refer to the U(1)
bre direction in the two 3-spheres respectively. The results of Appendix C.2 imply
that the covariant derivatives along these directions are
D
2
=
2
+
u
1
4

01
,
D
5
, =
5

u
2
4

34
,
D
8
=
8

u
3
4

67
,
(9.4.9)
where denotes an ordinary partial derivative. It follows from (9.4.8) and (9.4.9)
that

2
=
u
1
4

01
(1 +

) ,

5
=
u
2
4

34
(1 +

) ,

8
=
u
3
4

67
(1 +

) ,
(9.4.10)
and hence that we may take to be simultaneously independent of all three U(1)
bre direction by imposing
T

= 0, T

=
1
2
(1 +

). (9.4.11)
Notice that the operator T

is a projector. Furthermore, T

commutes with
the projector T dened in (9.4.6). Both T and T

admit 16 zero eigenvectors, but


since (TT

) has rank 8, only 8 of them are joint zero eigenvectors. Restricting the
9-dimensional Killing spinor to be independent of one or more U(1) bre directions
therefore breaks
3
4
of the maximal number of supersymmetries. This establishes
that all the vacuum solutions in Figure 9.4 that are obtained from the AdS
3
S
3

S
3
vacua by a combination of Hopf reduction steps and lower-dimensional Weyl
transformations preserve one quarter of the maximum number of supersymmetries.
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 173
Assume now that, say, u
3
= 0, in which case the 9-dimensional vacuum is of the
form AdS
3
S
3
T
3
. The part of the Killing spinor equation that pertains to the
T
3
factor is now given by
D
a
= 0. (9.4.12)
whence it is evident that the Killing spinor may be taken to be independent of
the toroidal directions. By implication, the Weyl multiplet of AdS
3
S
3
vacua
in Figure 9.5 preserves one half of the maximum number of supersymmetries. An
argument analogous to the one presented above shows that further Hopf reductions
of the AdS
3
S
3
vacuum breaks another half of the supersymmetries. The vacuum
solutions of D = 5 and D = 4 supergravity in Figure 9.5 therefore preserve one
quarter of the maximal number of supersymmetries.
9.5 Identifying interpolating p-branes
The present study of Freund-Rubin vacua was motivated in part by the observation
that these vacuum solutions emerge in the near horizon limit of aristocratic p-
brane solutions, i.e. p-brane solutions with everywhere regular dilatonic scalars.
Since the p-brane solution can be mapped onto its near horizon limit, and vice
versa, by a series of duality transformations and dimensional reduction/oxidation
steps, the classication of Freund-Rubin vacua implies a classication of aristocratic
p-branes. In this section, we comment briey on how to identify the aristocratic p-
brane solution associated with a given vacuum solution. We restrict our attention to
the supersymmetric Freund-Rubin vacua (with maximally symmetric submanifolds),
which we have listed again in Table 9.1 for convenience, and hope to be able to discuss
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 174
the non-supersymmetric vacua elsewhere.
D Spacetime SuSy Multiplicity p-brane
D = 11 AdS
4
S
7
1 1 M2-brane
D = 11 AdS
7
S
4
1 1 M5-brane
D = 10 AdS
5
S
5
1 1 D3-brane
D = 9 AdS
3
S
3
S
3 1
2
2 see text
D = 8 AdS
3
S
3
S
2 1
4
6 see text
D = 8 AdS
2
S
3
S
3 1
4
6 see text
D = 7 AdS
3
S
2
S
2 1
4
15 see text
D = 7 AdS
2
S
2
S
3 1
4
15 see text
D = 6 AdS
2
S
2
S
2 1
4
40 see text
D = 6 AdS
3
S
3 1
2
5 dyonic string
D = 5 AdS
3
S
2 1
4
45 3-charge string
D = 5 AdS
2
S
3 1
4
45 3-charge BH
D = 4 AdS
2
S
2 1
4
630 4-charge BH
Table 9.1: Supersymmetric Freund-Rubin vacua in D 11.
The supersymmetric vacua in Table 9.1 fall into two groups, according to whether
the spacetime geometry contains a single sphere factor, or two such sphere factors. It
has been known for some time now [9] that the AdS S vacua in D = 11, 10 emerge
in the near horizon limit of the aristocratic single-charge p-branes in D = 11, 10.
Similarly, the AdSS vacua in D = 6, 5, 4 are readily interpreted as the near horizon
limits of the multi-charge aristocratic p-branes in these dimensions, see e.g. Ref. [91].
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 175
A simple consistency check for the latter identication is that the multiplicities of
the AdS S vacua in D = 6, 5, 4 agree precisely with the multiplicities of the
corresponding aristocratic p-branes.
In contrast, the AdS S S vacua in D = 9, 8, 7, 6 cannot be interpreted as
near horizon limits of aristocratic p-branes in these dimensions. Rather, each of
these vacua has to be oxidised to (at least) one dimension higher, where it may
(depending on the eld strengths that support the vacuum in question) acquire the
interpretation of the near horizon limit of a non-standard harmonic intersection of p-
branes [60, 38]. This is best illustrated with an example. Consider the AdS
3
S
3
S
3
vacuum depicted in Figure 9.6. The 9-dimensional eld conguration can be written
as
ds
2
9
= e
u
1

_
dt
2
+dx
2
_
+d
2
+
_
2
u
2
_
2
d
2
3
+
_
2
u
3
_
2
d

2
3
, (9.5.1)
F
(3)1
= u
1

(3)
(
u
1
2
) +u
2
_
2
u
2
_
3

(3)
+u
3
_
2
u
3
_
3

(3)
, (9.5.2)
where d
2
3
and d

2
3
are line elements on (distinct) unit 3-spheres,
(3)
and
(3)
are
the associated volume 3-forms and w
(3)
(
u
1
2
) is the volume 3-form of the AdS
3
with
inverse radius
u
1
2
. The constants are related by u
2
1
= u
2
2
+u
2
3
.
We now oxidise this solution to D = 10. Let z denote the coordinate that has
been compactied in the reduction from D = 10 to D = 9. Following [60, 38], we
also perform the coordinate transformation
=
2
u
1
ln
r r

Q
1
Q
2
, z =
2
u
1
_
_
Q
1
Q
2
ln r
_
Q
2
Q
1
ln r
_
, (9.5.3)
where Q
1

4
u
2
2
and Q
2

4
u
2
3
, implying that
u
2
1
4
=
1
Q
1
+
1
Q
2
. The resulting 10-
dimensional metric can be written as
ds
2
10
=
r
2
r
2
Q
1
Q
2
_
dt
2
+dx
2
_
+
Q
1
r
2
_
dr
2
+r
2
d
2
3
_
+
Q
2
r
2
_
d r
2
+ r
2
d

2
3
_
. (9.5.4)
CHAPTER 9. SUPERSYMMETRIC VACUA IN D 11 176
It has been identied in Ref. [60] as the metric of a D = 10, N = 1 (or N = 2A)
supergravity p-brane conguration representing the intersection of two 5-branes over
a line, with an additional string along the intersection direction, in the region close
to the core of both 5-branes. Restoring the constants in the harmonic functions
leads to a non-dilatonic (and hence aristocratic) p-brane solution in D = 10 of the
form
ds
2
10
= (H
1
H
2
)
1
_
dt
2
+dx
2
_
+H
1
_
dr
2
+r
2
d
2
3
_
+H
2
_
d r
2
+ r
2
d

2
3
_
,
F
(3)1
= d(H
1
H
2
)
1
dt dx + 2Q
1

(3)
+ 2Q
2

(3)
, (9.5.5)
where H
1
= 1 +
Q
1
r
2
and H
2
= 1 +
Q
2
r
2
. Note that it does not appear to be possible
to Kaluza-Klein reduce this p-brane solution in D = 10, as opposed to its near
horizon limit (9.5.4), to a p-brane solution in D = 9. The aristocratic p-brane
associated with the AdS
3
S
3
S
3
vacuum lives in D = 10, and only its near
horizon limit can be Kaluza-Klein reduced to D = 9.
In principle, one can determine the aristocratic p-brane solutions associated with
every AdS S S vacuum in D = 9, 8, 7, 6 by following a method analogous to the
one reviewed above. Partial steps in this direction have already been undertaken
in Ref. [38], where a number of aristocratic p-branes in D = 11 and D = 10 we
presented, whose near horizon limits toroidally reduce to particular members of
the Weyl multiplets of AdS S S vacua in D = 9, 8, 7, 6. We postpone a detailed
discussion of this issue to a future publication.
Chapter 10
Conclusions
In the preceding chapters, we have presented an exhaustive classication of the
Freund-Rubin vacuum solutions of each maximal, massless supergravity theory in
2 D 11. These Freund-Rubin vacua are characterised by a spacetime product
manifold of several Einstein submanifolds, covariantly constant p-form eld strengths
and constant scalar elds. The possible spacetime decompositions that can serve as
the basis for a generalised Freund-Rubin ansatz are determined by the degrees of
the eld strengths that feature in the supergravity theory. The particular nature
of the Freund-Rubin ansatz reduces solving the supergravity eld equations to an
algebraic problem.
The actual classication of vacua has been simplied by the observation that
since the maximal supergravity theories in D 11 (with the exception of IIB su-
pergravity in D = 10) are related by toroidal Kaluza-Klein reduction, it is sucient
to consider, in each dimension, only those vacuum solutions that do not contain
toroidal factors in the spacetime geometry. In addition, the global Cremmer-Julia
symmetries of the lower-dimensional supergravities allowed us to reduce the number
177
CHAPTER 10. CONCLUSIONS 178
of participating eld strengths in a vacuum solution to the minimal set consistent
with the truncation of the scalars. Vacuum solutions of a given spacetime decom-
position and a xed number of participating eld strengths were found to assemble
naturally into multiplets of the Weyl subgroup of the Cremmer-Julia global symme-
try group.
The fact that the lower-dimensional vacua form irreducible Weyl multiplets, to-
gether with the observation that certain vacua in dierent dimensions are related
by Kaluza-Klein dimensional reduction, greatly facilitated the determination of the
unbroken supersymmetries of the Freund-Rubin vacua that we have classied. We
have shown that all vacuum solutions of IIA supergravity can be obtained via Hopf
reduction from vacua in D = 11, while vacuum solutions of the supergravities in
D 9 can be mapped, by a combination of lower-dimensional Weyl transformation
and dimensional oxidation, either to one of the vacua in D = 11, or to a M
3
K
3
K
3
vacuum solution in D = 9. Since Kaluza-Klein reduction/oxidation does not break
any supersymmetries per se and Weyl transformations commute with supersymme-
try transformations in every dimension, the number of unbroken supersymmetries
of the lower-dimensional vacua can be inferred from those of the higher-dimensional
vacua to which they are related, and the dependence of the higher-dimensional
Killing spinors on the reduction coordinate(s). The supersymmetric Freund-Rubin
vacua (with maximally symmetric submanifolds) are given in Table 9.1. We also
have argued that each of these supersymmetric vacua emerges in the near horizon
limit of an aristocratic p-brane solution.
Appendix A
Notation and conventions
We explain the index notation we employ and summarise the various sign conventions
we have adopted.
Indices: We use indices as indicated in Table A.1. In the context of toroidal com-
Letters Index Type Range
M, N, P, . . . spacetime 0, 1, . . . , (D 1)
I, J, K, . . . space 1, . . . , (D 1)
m, n, p, . . . internal space 1, 2, . . . , (11 D)
A, B, C, . . . tangent space 0, 1, . . . , (D 1)
a, b, c, . . . internal tangent space 1, 2, . . . , (11 D)
, , , . . . worldvolume 0, 1, . . . , p
Table A.1: Index conventions.
pactication of D = 11 supergravity to D < 11 dimensions, we distinguish the
11-dimensional spacetime indices by a hat, and split them into D-dimensional
179
APPENDIX A. NOTATION AND CONVENTIONS 180
spacetimes indices and internal indices for the compactifying torus accord-
ing to

M = (M, m), where M = i corresponds to

M = i, whereas m = i
corresponds to

M = 11 i. Equivalently, the internal index m labels the co-
ordinate that is being reduced upon in the reduction step from D = 12 m to
D = 11 m. The 11-dimensional tangent space indices are split analogously.
Signature: We use the mostly plus signature convention for the Minkowski metric

MN
and its inverse
MN
in D spacetime dimensions, i.e.

MN
=
MN
= diag (1, +1, +1, , +1
. .
(D 1) entries
). (A.0.1)
(Anti-)Symmetrisation: (Anti-)Symmetrisation of a set of indices is denoted by
round brackets ( ) and square brackets [ ] respectively and dened with
unit strength, e.g.
X
(M
1
M
p
)
=
1
p!
(X
M
1
M
p
+ symmetric permutations) (A.0.2)
X
[M
1
M
p
]
=
1
p!
(X
M
1
M
p
+ antisymmetric permutations). (A.0.3)
-symbols: The -symbols
M
1
M
D
and
M
1
M
D
in D dimensions are totally anti-
symmetric tensor densities, dened by

M
1
M
D
=
[M
1
M
D
]
(A.0.4)

M
1
M
D
=
[M
1
M
D
]
, (A.0.5)
together with
12D
=
12D
= 1. They satisfy the following useful contrac-
tion identity

P
1
P
i
M
i+1
M
D

P
1
P
i
N
i+1
N
D
= i! (D i)!
[M
i+1
N
i+1

M
D
]
N
D
. (A.0.6)
Appendix B
Maximal massless supergravities
in 3 D 11
In the rst section of this appendix, we review the construction of the bosonic
sectors of IIA supergravity [85] in D = 10 and the maximal, massless supergravity
theories in 3 D 9 by toroidal Kaluza-Klein reduction of the bosonic sector
of D = 11 supergravity [4]. We shall be using the formalism developed in Ref.
[86], see also Refs. [80, 87]. This review not only establishes the notation we shall
be using for the bosonic elds of the lower dimensional supergravities, but also
allows us to present their equations of motion, which we have been unable to nd
elsewhere, at least in D 9. The lower dimensional eld equations are derived by
substituting the Kaluza-Klein ansatz for the 11-dimensional elds directly in the 11-
dimensional eld equations, rather than by varying the D-dimensional Lagrangian.
This method has the advantage of preserving the manifest gauge invariance of the
11-dimensional eld equations, and thus avoids the occurrence of bare potentials
181
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 182
in the lower dimensional eld equations. We also discuss the Kaluza-Klein reduction
of the 11-dimensional Killing spinor equation, i.e. the equation which determines the
supersymmetries of a purely bosonic solution of the 11-dimensional eld equations,
and hence obtain the Killing spinor equations that determine the supersymmetries
of a purely bosonic solution of the lower-dimensional eld equations.
In the second section of this appendix, we present a summary of the bosonic
sector of IIB supergravity [79, 88] in D = 10. We also review briey the T-duality
mapping between the 9-dimensional supergravities obtained by dimensionally re-
ducing IIA and IIB supergravity on a circle. The Killing spinor equations that
determine the supersymmetries of a purely bosonic IIB supergravity background
are also discussed.
B.1 Toroidal Kaluza-Klein reduction of D = 11 super-
gravity
The Kaluza-Klein ansatz
The bosonic sector of D = 11 supergravity consists of the metric g
M

N
and a 3-form
potential

A
(3)
with associated 4-form eld strength

F
(4)
= d

A
(3)
. The action for these
elds can be written as the integral over an 11-form Lagrangian

L
(11)
, where

L
(11)
=

1
1
2

F
(4)


F
(4)

1
6

F
(4)


F
(4)


A
(3)
(B.1.1)
and denotes the Hodge dual with respect to the D = 11 metric and all 11-
dimensional quantities are distinguished by a hat.
As a preliminary to discussing the toroidal reduction of D = 11 supergravity to
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 183

M,

A = 0, 1, , 9, 10
0 1 (D 1) D (D + 1) 9 10
0 1 (D 1) (11 D) (10 D) 2 1
M, A = 0, 1, , (D 1) m, a = 1, 2, , (11 D)
Table B.1: KK split of 11-dimensional indices.
D < 11, we split the 11-dimensional spacetime and tangent space indices according
to

M = (M, m) and

A = (A, a). The assignment of numerical values to the indices
is explained in Table B.1. Notice that the internal indices label the reduction steps
from D = 11 to D dimensions.
The Kaluza-Klein ansatz for the 11-dimensional metric and the 3-form potential
is [86, 87]
d s
2
11
= e
s

(ds
2
D
+

a
e

f
a

(h
a
)
2
), (B.1.2)

A
(3)
= A
(3)
+A
(2)m
dz
m
+
1
2
A
(1)mn
dz
m
dz
n
(B.1.3)
+
1
6
A
(0)mnp
dz
m
dz
n
dz
p
,
where the internal vielbein h
a
is given by
h
a
=
a
m
dz
m
+/
a
(1)
+/
a
(0)
m
dz
m
. (B.1.4)
Here, it is understood that the D-dimensional elds are independent of the coordi-
nates z
m
that parameterise the internal (11 D)-torus.
The internal indices labelling the potentials A
(1)mn
and A
(0)mnp
are naturally
antisymmetric, whereas the Kaluza-Klein 0-form potentials /
a
(0)
m
are dened only
for a < m, since they come from the reduction of a Kaluza-Klein 1-form potential
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 184
/
a
(1)
that had to be already present in D + 1 dimensions. In the following, we take
/
a
(0)
m
to vanish for a m.
The quantity

is an (11 D)-component vector of dilatonic scalars, whereas s
and

f
a
are constant (11 D)-component vectors. Explicitly, they are given by
s = (s
1
, . . . , s
11D
),

f
a
= ( 0, . . . , 0
. .
a1
, (10 a)s
a
, s
a+1
, . . . , s
11D
), (B.1.5)
where s
a
=
_
2/((10 a)(9 a)). They satisfy satisfy the dot product relations
s s =
2(11D)
9(D2)
, s

f
a
=
2
D2
,

f
a


f
b
= 2
ab
+
2
D2
. (B.1.6)
In addition, it is not dicult to see that

f
a
= 9s. (B.1.7)
One can also show from (B.1.6) that for any vector v,

a
(

f
a
v)
2
= 2v v + 9(s v)
2
. (B.1.8)
Upon substitution of the Kaluza-Klein ansatz (B.1.2) and (B.1.3) into the La-
grangian (B.1.1), one nds that

L
(11)
= L
(D)
dz
1
dz
(11D)
, (B.1.9)
The D-form Lagrangian L
(D)
for the toroidal compactication of D = 11 supergrav-
ity to D dimensions is given by [86]
L
(D)
= R
1
2
d


1
2
e
a

F
(4)
F
(4)

1
2

a
e
a
a

F
a
(3)
F
(3)a

1
2

a<b
e
a
ab

F
ab
(2)
F
(2)ab

1
2

a
e

b
a

T
(2)a
T
a
(2)
(B.1.10)

1
2

a<b<c
e
a
abc

F
abc
(1)
F
(1)abc

1
2

a<b
e

b
ab

T
(1)a
b
T
a
(1)b
+L
FFA
(D)
,
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 185
where denotes the Hodge dual with respect to the D-dimensional metric.
The L
FFA
(D)
term is the contribution to the D-dimensional Lagrangian that comes
from the reduction of the
1
6

F
(4)


F
(4)


A
(3)
term in the 11-dimensional Lagrangian
(B.1.1). The explicit form of L
FFA
(D)
is dimension dependent and may be found in
Ref. [86].
The various eld strengths that appear in the Lagrangian (B.1.10) are dened
by
d

A
(3)
= F
(4)
+F
(3)a
h
a
+
1
2
F
(2)ab
h
ab
+
1
6
F
(1)abc
h
abc
, (B.1.11)
dh
a
= T
a
(2)
+T
a
(1)b
h
b
, (B.1.12)
where h
a
1
a
n
h
a
1
h
a
n
. The internal indices labelling the eld strengths
F
(2)ab
and F
(1)abc
are antisymmetric, while T
a
(1)b
vanishes for a b. We have also
adopted the convention that the dualisation of a eld strength carrying internal
indices reverses the position of the indices, in the sense that lower indices are turned
into upper indices, and vice versa.
The quantities a, a
a
, a
ab
, a
abc
,

b
a
and

b
ab
are constant (11 D)-component
dilaton vectors. They are given in terms of s and

f
a
by
F
(4)
: a = 3s, T
a
(2)
:

b
a
=

f
a
,
F
(3)a
: a
a
= 3s +

f
a
, T
a
(1)b
:

b
ab
=

f
a
+

f
b
,
F
(2)ab
: a
ab
= 3s +

f
a
+

f
b
,
F
(1)abc
: a
abc
= 3s +

f
a
+

f
b
+

f
c
,
(B.1.13)
where we have also included explicitly the eld strength with which the dilaton
vectors are associated.
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 186
Bosonic eld equations of 3 D 11 supergravity
It is a generic feature of the Kaluza-Klein reduction method that the various eld
strengths that appear in the D-dimensional Lagrangian (B.1.10) are not simply the
exterior derivatives of the corresponding potentials. Rather, they acquire non-linear
Kaluza-Klein modications as well. The explicit expressions for the eld strengths in
terms of the potentials are obtained by substituting the Kaluza-Klein ansatz (B.1.3)
and (B.1.4) into the denitions (B.1.11) and (B.1.12) respectively, and may be found
in Ref. [80]. In contrast, the L
FFA
(D)
contribution to the D-dimensional Lagrangian
is most easily given, not in terms of the Kaluza-Klein modied eld strengths, but
rather in terms of the potentials and their exterior derivatives.
The equations of motion of the various potentials that are obtained by varying
the D-dimensional Lagrangian contain, in general, bare, or undierentiated, po-
tentials. Of course, the gauge invariance of the D-dimensional Lagrangian, which is
inherited from the gauge invariance of the 11-dimensional Lagrangian, guarantees
the gauge invariance of the D-dimensional equations of motion. We have veried
that it is possible to write the equations of motion in a manifestly gauge invariant
manner, i.e. entirely in terms of the eld strengths
1
. This has been achieved by
replacing the exterior derivative of a potential with the corresponding eld strength,
minus the Kaluza-Klein modications, whenever it occurs in the equations of mo-
tion. In this process, it turns out that all the terms in the equations of motion that
contain bare potentials cancel exactly.
1
A gauge invariant rst order formulation of the equations motion has been developed in Ref.
[87]. However, this rst order formulation is not well suited for the purpose of actually solving the
equations of motion.
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 187
However, this process in not only laborious, but also relies on the cancellation
of all terms containing bare potentials. Fortunately, there exists a quicker route to
obtaining manifestly gauge invariant D-dimensional equations of motion. It is well
known [64] that the Kaluza-Klein ansatz (B.1.2) and (B.1.3) is a consistent trun-
cation of D = 11 supergravity, in the sense that the equations of motion that are
obtained by varying the D-dimensional Lagrangian are equivalent to the equations
of motion that are obtained by substituting the ansatz directly in the 11-dimensional
elds equations. The consistency of the truncation therefore guarantees that the so-
lutions of the D-dimensional eld equations are also solutions of the 11-dimensional
eld equations. The gauge invariance of the 11-dimensional eld equations is pre-
served automatically throughout the truncation, since the equations of motion, un-
like the Lagrangian, can be expressed without reference to the potentials.
Accordingly, consider the 11-dimensional eld equations that follow from (B.1.1):

1
A

1
2

A

B
1 =

T
A

B
(

F
(4)
), (B.1.14)
d

F
(4)
=
1
2

F
(4)


F
(4)
, (B.1.15)
d

F
(4)
= 0, (B.1.16)
where

T
A

B
(

F
(4)
) =
1
2
_
(

F
(4)
)
2

1
2

A

B
(

F
(4)
)
2
_
, (B.1.17)
are the tangent space components of the 11-dimensional energy-momentum tensor
2
of the gauge eld.
The Kaluza-Klein ansatz for the 11-dimensional metric is given in (B.1.2). The
2
Expressions such as (

F
(4)
)
2

B
and (

F
(4)
)
2
are understood to include the appropriate permutation
factors.
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 188
vielbein 1-forms e

A
= ( e
A
, e
a
) are thus
e
A
= e
1
2
s

e
A
, e
a
= e

h
a
, (B.1.18)
where e
A
are the D-dimensional vielbein 1-forms and
a
=
1
2
(s

f
a
). We nd that
the 11-dimensional connection 1-forms

B
, satisfying
d e

A
+

B
e

B
= 0,
A

B
=
B

A
, (B.1.19)
are given by

AB
=
AB

1
2
s
_

e
B

e
A
_

1
2

d
e

b
d

T
d
AB
h
d
, (B.1.20)

a
A
= e
1
2

b
a

_
1
2
T
a
AC
e
C
+
a

h
a
_
(B.1.21)
+
1
2

d
e
1
2

b
d

_
e
1
2

b
ad

T
ad
A
+e
1
2

b
da

T
da
A
_
h
d
,

ab
=
1
2
_
e
1
2

b
ab

T
ab
C
e
1
2

b
ba

T
ba
C
_
e
C
, (B.1.22)
where
AB
are the D-dimensional connection 1-forms and the Kaluza-Klein eld
strengths T
a
(2)
=
1
2
T
a
CD
e
C
e
D
and T
a
(1)b
= T
a
Cb
e
C
are dened in (B.1.12). The
tangent space indices (A, a) are raised/lowered with the tangent space metrics
AB
and
ab
.
The connection 1-forms w

B
determine the 11-dimensional curvature 2-forms

B
via

B
=
1
2

B

C

D
e

C
e

D
= d

B
+

B
, (B.1.23)
where

1

B

C

D
are the tangent space components of the 11-dimensional Riemann
curvature tensor. The tangent space components

1
A

B
of the 11-dimensional Ricci
tensor are then dened by

1
A

B
=

1

D

B
. We nd that

1
A

B
is given by
e
s


1
AB
= 1
AB

1
2
s
2

AB

1
2


1
2

d
e

b
d

(T
d
(2)
)
2
AB
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 189

1
2

c<d
e

b
cd

(T
c
(1)d
)
2
AB
, (B.1.24)
e
s


1
Aa
=
1
2
e

1
2

b
a

C
(e

b
a

T
CAa
) +
1
2
e

1
2

b
a

d
e

b
d

T
d
AC
T
C
da
, (B.1.25)
e
s


1
ab
=
1
2
e

1
2

b
ab

C
(e

b
ab

T
C ab
)
1
2
e

1
2

b
ba

C
(e

b
ba

T
C ba
)
+
1
2

d
e
1
2
(

b
ad
+

b
bd
)

T
Dad
T
Dd
b

1
2

d
e
1
2
(

b
da
+

b
db
)

T
D
da
T
d
Db
+
1
4
e
1
2
(

b
a
+

b
b
)

T
CDa
T
CD
b

a

ab
, (B.1.26)
where 1
AB
are the tangent space components of the D-dimensional Ricci tensor.
The 11-dimensional Ricci scalar

1 =

B

1
A

B
is therefore equal to
e
s


1 = 1s
2


1
2
(

)
2

1
2

d
e

b
d

(T
d
(2)
)
2

1
2

c<d
e

b
cd

(T
c
(1)d
)
2
, (B.1.27)
where 1 is the D-dimensional Ricci scalar.
The equations of motion of the D-dimensional metric, the Kaluza-Klein poten-
tials /
a
(1)
, the axions /
a
(0)
m
and the dilatons

are obtained by substituting (B.1.24)-
(B.1.27) into the 11-dimensional Einstein equation (B.1.14) and evaluating the 11-
dimensional energy-momentum tensor (B.1.17) for

F
(4)
given in (B.1.11). We nd
that the (A, B) components of (B.1.14) yield the D-dimensional Einstein equation
1
AB

1
2

AB
1 =
1
2


1
4

AB
(

)
2
(B.1.28)
+e
a

T
AB
(F
(4)
) +

d
e
a
d

T
AB
(F
(3)d
)
+

c<d
e
a
cd

T
AB
(F
(2)cd
) +

b<c<d
e
a
bcd

T
AB
(F
(1)bcd
)
+

d
e

b
d

T
AB
(T
d
(2)
) +

c<d
e

b
cd

T
AB
(T
c
(1)d
),
where the energy-momentum tensors of the D-dimensional gauge elds are dened
similarly to (B.1.17).
The (A, a) components of (B.1.17) give rise to the equations of motion for the
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 190
Kaluza-Klein potentials /
a
(1)
. In a dierential form notation, they can be written as
()
D2
d
_
e

b
a

T
(2)a
_
(B.1.29)
= e
a

F
(4)
F
(3)a
+e
a
d

F
d
(3)
F
(2)da

1
2
e
a
cd

F
cd
(2)
F
(1)cda
e

b
d

T
(2)d
T
d
(1)
a
.
The equations of motion for the axions /
a
(0)
m
come from the (a, b) components
of (B.1.14). In a dierential form notation, they can be written as
()
D1
d
_
e

b
ab

T
(1)a
b
_
(B.1.30)
= e
a
b

F
b
(3)
F
(3)a
e
a
bc

F
bc
(2)
F
(2)ac
+e

b
a

T
(2)a
T
b
(2)

1
2
e
a
bcd

F
bcd
(1)
F
(1)acd
+e

b
ac

T
(1)a
c
T
b
(1)
c
e

b
cb

T
(1)c
b
T
c
(1)
a
,
where we have assumed that a < b.
Finally, the (a, a) component of (B.1.14) can be interpreted as the equation of
motion for the dilaton
a

f
a

. Using the denitions (B.1.13) and the dilaton


vector products (B.1.6), one can show that the equations of motion for the
a
are
equivalent to

2

=
1
2
a e
a

(F
(4)
)
2
+
1
2

a
a
a
e
a
a

(F
(3)a
)
2
(B.1.31)
+
1
2

a<b
a
ab
e
a
ab

(F
(2)ab
)
2
+
1
2

b
a
e

b
a

(T
a
(2)
)
2
+
1
2

a<b<c
a
abc
e
a
abc

(F
(1)abc
)
2
+
1
2

a<b

b
ab
e

b
ab

(T
a
(1)b
)
2
.
The equations of motion for the D-dimensional potentials A
(3)
, A
(2)m
and A
(1)mn
and the axions A
(0)mnp
that come from the reduction of the 11-dimensional 3-form
potential

A
(3)
are obtained by reducing the 11-dimensional gauge eld equation
(B.1.15). Recall that the 11-dimensional eld strength

F
(4)
is given in term of the
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 191
D-dimensional eld strengths by (B.1.11). It is not dicult to show that the 11-
dimensional dual of

F
(4)
is therefore given by

F
(4)
= e
a

F
(4)
v +e
a
a

F
a
(3)
v
a
+
1
2
e
a
ab

F
ab
(2)
v
ab
+
1
6
e
a
abc

F
abc
(1)
v
abc
,
(B.1.32)
where v
a
1
a
n
is the (11 D n) form dened by
v
a
1
a
n
=
1
(11Dn)!

a
1
a
n
a
n+1
a
11D
h
a
n+1
a
11D
. (B.1.33)
The equations of motion for the D-dimensional potentials are obtained by sub-
stituting (B.1.11) and (B.1.32) into (B.1.15) and equating the coecients of v
a
1
a
n
in the resulting expression. We nd that they are given by
()
D4
d
_
e
a

F
(4)
_
= X
(D3)
, (B.1.34)
()
D3
d
_
e
a
a

F
a
(3)
_
= e
a

F
(4)
T
a
(2)
+e
a
b

F
b
(3)
T
a
(1)b
+X
a
(D2)
,
()
D2
d
_
e
a
ab

F
ab
(2)
_
= 2 e
a
a

F
[a
(3)
T
b]
(2)
+ 2 e
a
ac

F
[a|c|
(2)
T
b]
(1)c
+X
ab
(D1)
,
()
D1
d
_
e
a
abc

F
abc
(1)
_
= 3 e
a
ab

F
[ab
(2)
T
c]
(2)
+ 3 e
a
abd

F
[ab|d|
(1)
T
c]
(1)d
+X
abc
(D)
.
The X
(D3)
, X
a
(D2)
, X
ab
(D1)
and X
abc
(D)
contributions to the D-dimensional equa-
tions of motion come from the reduction of the
1
2

F
(4)


F
(4)
term in (B.1.15). These
contributions depend on the dimension under consideration. In 3 D 10, they
are given by
3
D = 10 : X
(D3)
= F
(4)
F
(3)
, (B.1.35)
X
(D2)
=
1
2
F
(4)
F
(4)
,
D = 9 : X
(D3)
=
1
2

ab
(F
(4)
F
(2)ab
F
(3)a
F
(3)b
) , (B.1.36)
3
In D = 10, the internal indices range over one value only, and hence have been dropped.
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 192
X
a
(D2)
=
ab
F
(4)
F
(3)b
,
X
ab
(D1)
=
1
2

ab
F
(4)
F
(4)
,
D = 8 : X
(D3)
=
abc
_
1
6
F
(4)
F
(1)abc

1
2
F
(3)a
F
(2)bc
_
, (B.1.37)
X
a
(D2)
=
1
2

abc
(F
(4)
F
(2)bc
F
(3)b
F
(3)c
) ,
X
ab
(D1)
=
abc
F
(4)
F
(3)c
,
X
abc
(D)
=
1
2

abc
F
(4)
F
(4)
,
D = 7 : X
(D3)
=
abcd
_
1
8
F
(2)ab
F
(2)cd

1
6
F
(3)a
F
(1)bcd
_
, (B.1.38)
X
a
(D2)
=
abcd
_
1
6
F
(4)
F
(1)bcd

1
2
F
(3)b
F
(2)cd
_
,
X
ab
(D1)
=
1
2

abcd
(F
(4)
F
(2)cd
F
(3)c
F
(3)d
) ,
X
abc
(D)
=
abcd
F
(4)
F
(3)d
,
D = 6 : X
(D3)
=
1
12

abcde
F
(2)ab
F
(1)cde
, (B.1.39)
X
a
(D2)
=
abcde
_
1
8
F
(2)bc
F
(2)de

1
6
F
(3)b
F
(1)cde
_
,
X
ab
(D1)
=
abcde
_
1
6
F
(4)
F
(1)cde

1
2
F
(3)c
F
(2)de
_
,
X
abc
(D)
=
1
2

abcde
(F
(4)
F
(2)de
F
(3)d
F
(3)e
) ,
D = 5 : X
(D3)
=
1
72

abcdef
F
(1)abc
F
(1)def
, (B.1.40)
X
a
(D2)
=
1
12

abcdef
F
(2)bc
F
(1)def
,
X
ab
(D1)
=
abcdef
_
1
8
F
(2)cd
F
(2)ef

1
6
F
(3)c
F
(1)def
_
,
X
abc
(D)
=
abcdef
_
1
6
F
(4)
F
(1)def

1
2
F
(3)d
F
(2)ef
_
,
D = 4 : X
(D3)
= 0, (B.1.41)
X
a
(D2)
=
1
72

abcdefg
F
(1)bcd
F
(1)efg
,
X
ab
(D1)
=
1
12

abcdefg
F
(2)cd
F
(1)efg
,
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 193
X
abc
(D)
=
abcdefg
_
1
8
F
(2)de
F
(2)fg

1
6
F
(3)d
F
(1)efg
_
,
D = 3 : X
a
(D2)
= 0, (B.1.42)
X
ab
(D1)
=
1
72

abcdefgh
F
(1)cde
F
(1)fgh
,
X
abc
(D)
=
1
12

abcdefgh
F
(2)de
F
(1)fgh
.
We restrict our attention to massless supergravities only, and therefore consistently
truncate the 4-form eld strength in D = 4 and the 3-form eld strengths in D = 3.
In addition to the equations of motion of the D-dimensional potentials we also
need the Bianchi identities satised by the D-dimensional eld strengths. They
are easily derived from dd

A
(3)
= 0 and ddh
a
= 0, where d

A
(3)
and dh
a
are given
in (B.1.11) and (B.1.12) respectively, by equating the coecients of h
a
1
a
n
in the
resulting expressions. We nd that they are given by
dF
(4)
= F
(3)a
T
a
(2)
, (B.1.43)
dF
(3)a
= F
(2)ab
T
b
(2)
+F
(3)b
T
b
(1)
a
,
dF
(2)ab
= F
(1)abc
T
c
(2)
+ 2 F
(2)c[a
T
c
(1)b]
,
dF
(1)abc
= 3 F
(1)d[ab
T
d
(1)c]
,
dT
a
(2)
= T
b
(2)
T
a
(1)b
,
dT
a
(1)b
= T
a
(1)
c
T
c
(1)b
.
Killing spinor equations in 3 D 11 supergravity
A purely bosonic solution of the 11-dimensional eld equations preserves a residual
amount of supersymmetry provided that there exist supersymmetry transformations
that leave the 11-dimensional gravitino


M
invariant. The variation of the gravitino
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 194
under a supersymmetry transformation with spinorial parameter

is [4]


M
=

T
M

(B.1.44)
where the supercovariant derivative

T
M
in a bosonic background is given by

T
M

=

D
M

+
1
12
1
24

F
N
1


N
4
_

N
1


N
4

M
+ 8

[

N
1

N
2

N
3

N
4
]

M
_

(B.1.45)
and

D
M

=
M

+
1
4

A

B

M

A

(B.1.46)
is the ordinary covariant derivative of

.
The -matrices with tangent space indices are constant (32 32) matrices that
form an irreducible representation of the Cliord algebra in D = (1, 10)

A
,
B
= 2
A

B
. (B.1.47)
In a Majorana representation, the -matrices are real, with
0
antisymmetric and

I
symmetric, where I = 1, , 10. We also adopt the convention that

=
0

1

9
(B.1.48)
where denotes the number 10. Antisymmetrised products of -matrices are dened
by

A
1


A
p

[

A
1

A
p
]
. (B.1.49)
and the elfbein e

M
and its inverse e
A

M
are used to convert tangent space indices
into spacetime indices and vice versa.
The condition for a bosonic supergravity background to preserve some residual
supersymmetries is


M
= 0, which from (B.1.44) translates into the existence
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 195
of a Killing spinor, i.e. a non-trivial solution of the 11-dimensional Killing spinor
equation

T
M

= 0. (B.1.50)
evaluated in the bosonic supergravity background under consideration.
If the bosonic supergravity background is of the Kaluza-Klein form (B.1.2) and
(B.1.3), the condition for the existence of residual supersymmetries is no longer
equivalent to the existence of a non-trivial Killing spinor satisfying (B.1.50). In
addition, we need to impose that the Killing spinor be independent of the coordi-
nates that have been compactied in the reduction from D = 11 to D dimensions.
Since the Kaluza-Klein supergravity background is naturally interpreted as (the ox-
idation of) a solution of D-dimensional maximal supergravity, it is convenient to
develop a formalism that allows us to determine its residual supersymmetries from
the D-dimensional perspective. Hence, we are lead to consider the toroidal reduc-
tion of the 11-dimensional gravitino, or, more precisely, the toroidal reduction of its
supersymmetry variation.
For this purpose, it most convenient to work in the tangent space basis. As
before, we split the 11-dimensional tangent space indices according to

A = (A, a).
The representation of the D = (1, 10) Cliord algebra in (B.1.47) splits naturally
into (reducible) representations the Cliord algebras in D Lorentzian and (11 D)
Euclidean dimensions

A
,
B
= 2
AB
,
a
,
b
= 2
ab
,
A
,
b
= 0. (B.1.51)
The Kaluza-Klein ansatz for the 11-dimensional vielbein and the 4-form eld strength
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 196
is augmented by the Kaluza-Klein ansatz for the 11-dimensional gravitino

A
= e

1
4
s

1
4(D2)

A

d

d
_
, (B.1.52)

a
=
1
4
e

1
4
s

a
(B.1.53)
and the 11-dimensional supersymmetry parameter is dimensionally reduced accord-
ing to

= e
1
4
s

(B.1.54)
All D-dimensional elds, i.e. all unhatted elds, are assumed to be independent of
the coordinates of the internal (11 D)-torus.
It is straightforward to evaluate the supercovariant derivative (B.1.45) in the
Kaluza-Klein background (B.1.20), (B.1.21), (B.1.22) and (B.1.11). We nd that

T
A

splits up into
e
1
4
s


T
A

= D
A

1
4
s
C

(
C
A

C
A
) (B.1.55)

1
4

a
e
1
2

b
a

T
a
CA

C

a

1
4

a<b
e
1
2

b
ab

T
ab
A

ab

+
1
12
1
24
e
1
2
a

F
C
1
C
4
_

C
1
C
4
A
+ 8
[C
1
C
2
C
3

C
4
]
A
_

1
12
1
6

a
e
1
2
a
a

F
a
C
1
C
2
C
3
_

C
1
C
2
C
3
A
+ 6
[C
1
C
2

C
3
]
A
_

+
1
12
1
2

a<b
e
1
2
a
ab

F
ab
C
1
C
2
_

C
1
C
2
A
+ 4
[C
1

C
2
]
A
_

ab

1
12

a<b<c
e
1
2
a
abc

F
abc
C
(
C
A
+ 2
C
A
)
abc
,
where D
A
is the D-dimensional covariant derivative of , and
e
1
4
s


T
a

=
1
2

a

a

1
4
1
2
e
1
2

b
a

T
C
1
C
2
a

C
1
C
2
(B.1.56)

1
4

d
_
e
1
2

b
ad

T
C
ad
+e
1
2

b
da

T
C
da
_

C

d

+
1
12
1
24
e
1
2
a

F
C
1
C
4

C
1
C
4

APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 197


+
1
12
1
6

d
e
1
2
a
d

F
d
C
1
C
2
C
3

C
1
C
2
C
2
(
da
+ 2
da
)
+
1
12
1
4

c,d
e
1
2
a
cd

F
cd
C
1
C
2

C
1
C
2
(
cda
+ 4
c

da
)
+
1
12
1
6

b,c,d
e
1
2
a
bcd

F
bcd
C

C
(
bcda
+ 6
bc

da
) .
In a purely bosonic D-dimensional supergravity background, the variations of

A
and
a
under a supersymmetry transformation with parameter are dened by

A
= e

1
4
s

1
4(D2)

A

d


d
_
=

T
A

, (B.1.57)

a
=
1
4
e

1
4
s


a
=

T
a

. (B.1.58)
It follows immediately from (B.1.56) that


a
= /
a
, (B.1.59)
where
/
a
2
a

1
2
e
1
2

b
a

T
C
1
C
2
a

C
1
C
2
(B.1.60)

d
_
e
1
2

b
ad

T
C
ad
+e
1
2

b
da

T
C
da
_

C

d
+
1
3
1
24
e
1
2
a

F
C
1
C
4

C
1
C
4

a
+
1
3
1
6

d
e
1
2
a
d

F
d
C
1
C
2
C
3

C
1
C
2
C
2
(
da
+ 2
da
)
+
1
3
1
4

c,d
e
1
2
a
cd

F
cd
C
1
C
2

C
1
C
2
(
cda
+ 4
c

da
)
+
1
3
1
6

b,c,d
e
1
2
a
bcd

F
bcd
C

C
(
bcda
+ 6
bc

da
) .
A little bit more eort is required to deduce the supersymmetry variation of
A
.
We nd that it can be written as


A
= T
A
, (B.1.61)
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 198
where the D-dimensional supercovariant derivative T
A
is dened by
T
A
= D
A

1
4

a<b
e
1
2

b
ab

T
ab
A

ab

1
4

a<b<c
e
1
2
a
abc

F
abc
A

abc
(B.1.62)
+
1
4(D2)
1
24
e
1
2
a

F
C
1
C
4
_
3
C
1
C
4
A
+ 4(D 5)
[C
1
C
2
C
3

C
4
]
A
_

1
4(D2)
1
6

a
e
1
2
a
a

F
a
C
1
C
2
C
3
_
2
C
1
C
2
C
3
A
+ 3(D 4)
[C
1
C
2

C
3
]
A
_

+
1
4(D2)
1
2

a<b
e
1
2
a
ab

F
ab
C
1
C
2
_

C
1
C
2
A
+ 2(D 3)
[C
1

C
2
]
A
_

ab

1
4(D2)

a
e
1
2

b
a

T
a
C
1
C
2
_

C
1
C
2
A
+ 2(D 3)
[C
1

C
2
]
A
_

a
.
The condition for a D-dimensional bosonic supergravity background to preserve
some residual supersymmetries is


A
= 0 and


a
= 0, which from (B.1.59) and
(B.1.61) translates into
T
A
= 0, /
a
= 0, a = 1, , (11 D), (B.1.63)
where T
A
and M
a
are evaluated in the supergravity background under consider-
ation. In other words, the D-dimensional bosonic supergravity background is su-
persymmetric provided there exists a Killing spinor, i.e. a non-trivial solution of
the D-dimensional Killing spinor equation T
A
= 0, that is also a simultaneous
eigenvector of the (11 D) matrices /
a
with eigenvalue zero.
B.2 IIB supergravity
Bosonic eld equations and T-duality in D = 9
The bosonic sector of IIB supergravity in D = 10 consists of the metric, a dilatonic
scalar , an axionic scalar , a pair of 2-form gauge potentials A
NS
(2)
and A
R
(2)
with
eld strengths F
NS
(3)
and F
R
(3)
and a 4-form potential B
(4)
whose eld strength H
(5)
is
self-dual.
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 199
The self-duality of the 5-form eld strength H
(5)
makes it dicult to write down
a manifestly covariant Lagrangian for the IIB supergravity elds. This shall not
concern us here, as we are primarily interested in the eld equations, which of
course can be expressed in a manifestly covariant way. The equations of motion for
the gauge potentials are
H
(5)
= H
(5)
, dH
(5)
= F
NS
(3)
F
R
(3)
,
d
_
e

F
NS
(3)
_
= e

F
R
(3)
d H
(5)
F
R
(3)
, dF
NS
(3)
= 0,
d
_
e

F
R
(3)
_
= H
(5)
F
NS
(3)
, dF
R
(3)
= F
NS
(3)
d.
(B.2.1)
The eld equations for the dilatonic and the axionic scalar can be written as
d
_
e
2
d
_
= e

F
R
(3)
F
NS
(3)
,
d d = e
2
d d +
1
2
e

F
NS
(3)
F
NS
(3)

1
2
e

F
R
(3)
F
R
(3)
,
(B.2.2)
and the equation of motion for the metric is given by
1
MN
=
1
2

m

N
+
1
2
e
2

N
+
1
4
(H
(5)
)
2
MN
(B.2.3)
+
1
2
e

_
(F
NS
(3)
)
2
MN

1
4
g
MN
(F
NS
(3)
)
2
_
+
1
2
e

_
(F
R
(3)
)
2
MN

1
4
g
MN
(F
R
(3)
)
2
_
.
We should mention that it is possible to write down a covariant Lagrangian whose
variation leads to the above eld equations [89], albeit only after a (consistent)
truncation. This is achieved by doubling the degrees of freedom of B
(4)
such that
its eld strength is no longer self-dual. Since the self-duality constraint on H
(5)
is
lifted, the Lagrangian for the enlarged set of elds can be expressed in a manifestly
covariant way. The eld equations for the degrees of freedom of IIB supergravity are
recovered by consistently truncating the anti-self-dual part of H
(5)
from the resulting
equations of motion.
As is well known, the 9-dimensional massless supergravities obtained by Kaluza-
Klein reduction of the 10-dimensional type IIA and IIB supergravities on a circle
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 200
are equivalent, up to local eld redenitions. This is just the eld-theoretic mani-
festation of the T-duality between type IIA and IIB string theory compactied on
a circle. Even though the equivalence of the IIA and IIB supergravities in D = 9 is
by now extensively documented in the literature, we want to review briey how this
T-duality is realised on the level of the 9-dimensional equations of motion.
To this end, consider the Kaluza-Klein ansatz for the 10-dimensional IIB elds,
which in the following will be distinguished from the 9-dimensional IIB elds by a
hat:
d s
2
IIB
= e
2
ds
2
9
+e
14
h
2
,

H
(5)
= H
(4)
h + (H
(4)
h),

F
NS
(3)
= F
NS
(3)
+F
NS
(2)
h,

F
R
(3)
= F
R
(3)
+F
R
(2)
h,

= , = ,
(B.2.4)
where h = dz + /
(1)
and =
1
4

7
. The Kaluza-Klein eld strength is dened as
usual by T
(2)
= d/
(1)
. Note that the Kaluza-Klein ansatz for

H
(5)
is manifestly
consistent with the self-duality constraint

H
(5)
=

H
(5)
. It can also be written as

H
(5)
= e
8
H
(4)
+H
(4)
h, (B.2.5)
where is the Hodge dual with respect to the 9-dimensional metric.
The eld equations of 9-dimensional IIB supergravity are obtained by substitut-
ing the Kaluza-Klein ansatz (B.2.4) directly into the 10-dimensional eld equations
(B.2.1), (B.2.2) and (B.2.4). We have veried that they are equivalent to the eld
equations of 9-dimensional IIA supergravity, as given in the previous section, pro-
vided that we make the following identications within the scalar sector of the two
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 201
9-dimensional theories:

1
=
1
4
(3

7),
2
=
1
4
(

7 + 3), /
1
(0)
2
= , (B.2.6)
where

= (
1
,
2
) is the IIA vector of dilatonic scalars, and /
1
(0)
2
the IIA axion.
The 9-dimensional IIA and IIB metrics are also identied.
In addition, the various p-form eld strengths of IIA and IIB supergravity in
D = 9 are related to one another as depicted in Table B.2. Note in particular that
2-forms 3-forms 4-forms
IIA T-duality IIB IIA T-duality IIB IIA T-duality IIB
F
(2)12
T
(2)
F
(3)1
F
NS
(3)
F
(4)
H
(4)
T
1
(2)
F
R
(2)
F
(3)2
F
R
(3)
T
2
(2)
F
NS
(2)
Table B.2: T-duality of IIA and IIB p-form eld strengths in D = 9.
T-duality interchanges the IIA eld strength F
(2)12
, which comes from the reduction
of the 4-form eld strength in D = 11, with the Kaluza-Klein eld strength T
(2)
whose potential arises out of the reduction of the IIB metric in D = 10.
It is now not dicult to establish that after performing a T-duality transforma-
tion, a 9-dimensional IIA eld conguration oxidises to the following 10-dimensional
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 202
IIB eld conguration:
d s
2
IIB
= e
1
4
a

ds
2
9
+e

7
4
a

h
2
,

H
(5)
= F
(4)
h +e
a

F
(4)
,

F
NS
(3)
= F
(3)1
T
2
(2)
h,

F
R
(3)
= F
(3)2
+T
1
(2)
h,

=
1
2

b
12

= /
1
(0)
2
,
(B.2.7)
where the internal vielbein is now given by h = dz+A
(1)12
and a and

b
12
are (linearly
independent) 2-component dilaton vectors dened in (B.1.13).
Killing spinor equation
A purely bosonic solution of the IIB eld equations preserves a residual amount of
supersymmetry provided that there exist supersymmetry variations that leave the
fermionic elds of IIB supergravity invariant. In IIB supergravity, the fermions are
complex Weyl spinors, with the spin 3/2 gravitino
M
and the spin 1/2 dilatino
having opposite chirality. In a purely bosonic IIB supergravity background, the
variations of
M
and under a supersymmetry transformation with complex Weyl
spinor parameter (of the same chirality as ) are [90]


M
= D
M
+
i
8
1
120
H
N
1
N
5

N
1
N
5
M
(B.2.8)
+
1
16
1
6
e

1
2

F
NS
N
1
N
2
N
3
_

N
1
N
2
N
3
M
+ 9
[N
1
N
2

N
3
]
M
_

+
i
16
1
6
e
1
2

F
RR
N
1
N
2
N
3
_

N
1
N
2
N
3
M
+ 9
[N
1
N
2

N
3
]
M
_

=
1
2
_

M
+i
M
e

(B.2.9)
+
1
4
1
6
_
e

1
2

F
NS
N
1
N
2
N
3
+i e
1
2

F
RR
N
1
N
2
N
3
_

N
1
N
2
N
3
.
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 203
where
D
M
=
M
+
1
4

AB
M

AB
(B.2.10)
is the ordinary covariant derivative of .
The -matrices with tangent space indices are constant (32 32) matrices that
form an irreducible representation of the Cliord algebra in D = (1, 9)

A
,
B
= 2
AB
. (B.2.11)
In a Majorana representation, the -matrices are real, with
0
antisymmetric and

I
symmetric, where I = 1, , 9. The chirality -matrix is dened by

=
0

1

9
(

)
2
= 1,
A
,

= 0. (B.2.12)
It is real and symmetric in a Majorana representation. In our conventions, the IIB
spinors satisfy


M
=
M
,

= ,

= . (B.2.13)
Antisymmetrised products of -matrices are dened as before and the zehnbein e
A
M
and its inverse e
A
M
are used to convert tangent space indices into spacetime indices
and vice versa.
In the main body of this thesis, we are primarily concerned with vacuum so-
lutions of IIB supergravity that are supported only by the 5-form eld strength.
For this class of IIB supergravity backgrounds, the supersymmetry variations of the
fermionic elds simplify considerably. Indeed, we see immediately from (B.2.8) and
(B.2.9) that the IIB vacua supported only by H
(5)
preserve residual supersymmetries
provided that there exist non-trivial solutions to the Killing spinor equation
T
M
= 0. (B.2.14)
APPENDIX B. MAXIMAL MASSLESS SUPERGRAVITIES IN 3 D 11 204
where the supercovariant derivative T
M
for this particular class of IIB supergravity
backgrounds is given by
T
M
= D
M
+
i
8
1
120
H
N
1
N
5

N
1
N
5
M
. (B.2.15)
Appendix C
U(1) brations and Hopf
reduction
C.1 Hopf reduction of U(1) brations
In this appendix, we review a particular class of solutions of the the (D + 1)-
dimensional Einstein equation

1
M

N
= D g
M

N
(C.1.1)
for which the (D+1)-dimensional manifold can be regarded as a U(1) bration over
a D-dimensional base manifold. We shall follow closely the treatment in reference
[77], where the interested reader will also nd the necessary conditions for these
U(1) brations to be non-trivial and well-dened.
We start by writing the line element d s
2
of the U(1) bration in Kaluza-Klein
form as
d s
2
= ds
2
+ (dz +/
(1)
)
2
, (C.1.2)
205
APPENDIX C. U(1) FIBRATIONS AND HOPF REDUCTION 206
where z is the (periodic) bre coordinate, A
(1)
is the U(1) connection and ds
2
the
line element of the base manifold. The tangent space components of the Ricci tensor
for d s
2
are

1
AB
= 1
AB

1
2
(T
(2)
)
2
AB
,

1
Az
=
1
2

B
T
BA
,

1
zz
=
1
2
(T
(2)
)
2
, (C.1.3)
where T
(2)
= d/
(1)
is the U(1) eld strength and 1
AB
are the tangent space com-
ponents of the Ricci tensor for ds
2
. The Einstein equation (C.1.1) therefore admits
U(1) bration solutions provided that
1
MN
=
1
2
(T
(2)
)
2
MN
+D g
MN
, d T
(2)
= 0, (T
(2)
)
2
= 2D . (C.1.4)
We are principally interested in the case when D = 2n and the base manifold
is the n-fold product of Einstein 2-manifolds. Other choices of base manifold that
have been discussed in the literature [77, 91] include Einstein-Kahler manifolds and
products of Einstein-Kahler manifolds and Einstein 2-manifolds.
Accordingly, we make the following product ansatz for the base metric and the
U(1) eld strength
ds
2
2n
=
n

i=1
(ds
i
2
)
2
, T
(2)
=
n

i=1
v
i

(2)i
, (C.1.5)
where (ds
i
2
)
2
and
(2)i
are the line element and the volume form of the i
th
submani-
fold. The v
i
are constants and characterise the topologically non-trivial gauge eld
conguration. They are related to the winding numbers of the U(1) gauge eld over
the 2-manifolds in the base [77].
It is not dicult to see that this ansatz solves (C.1.4), provided that the scalar
curvature of the i
th
submanifolds is equal to

i
=
1
2

i
v
2
i
+D (C.1.6)
APPENDIX C. U(1) FIBRATIONS AND HOPF REDUCTION 207
and the constants v
i
satisfy
n

i=1

i
v
2
i
= 2D , (C.1.7)
where
i
= 1 is the signature parameter of the i
th
submanifold ( = 1 and
= +1 corresponding to Lorentzian, respectively Euclidean, signature). Since a
2-dimensional Einstein manifold is maximally symmetric, the sign of the scalar cur-
vature
i
and the signature parameter
i
determines the 2-manifolds according to
the Table C.1.
> 0 < 0 = 0
= +1 S
2
H
2
T
2
= 1 dS
2
AdS
2
T
(1,1)
Table C.1: 2-dimensional Einstein manifolds.
We have therefore shown that there exist (2n+1)-dimensional Einstein manifolds
that can be interpreted as U(1) brations over a base manifold that is the n-fold
product of 2-dimensional Einstein manifolds. These U(1) brations will be denoted
by N
2n+1
(v
1
, . . . , v
n
) in the following. Before we proceed, we want to draw attention
to two classes of U(1) brations that are frequently encountered in the main body
of this paper:

Q
2n+1
(v
1
, . . . , v
n
) : (2n+1)-dimensional Lorentzian signature U(1) brations with
negative scalar curvature =
2
and base manifold AdS
2
K
2
K
2
,
where K
2
is S
2
, or H
2
, or T
2
.
Q
2n+1
(v
1
, . . . , v
n
) : (2n+1)-dimensional Euclidean signature U(1) brations with
positive scalar curvature =
2
and base manifold S
2
S
2
.
APPENDIX C. U(1) FIBRATIONS AND HOPF REDUCTION 208
Examples of these U(1) brations that have featured in the literature are the
Hopf brations of S
3
= Q
3
(v) and AdS
3

=

Q
3
(v), see e.g. Ref. [91], and the coset
spaces Q
5
(v
1
, v
2
) and Q
7
(v
1
, v
2
, v
3
), the latter two predominantly in the context of
spontaneous compactications of IIB supergravity and D = 11 supergravity, see e.g.
Refs. [92, 93].
The Q
2n+1
(v
1
, . . . , v
n
) spaces are in fact the only U(1) bration solutions (of
the class considered here) of the (2n + 1)-dimensional Euclidean signature Einstein
equations, since in 2n+1 Euclidean dimensions the LHS of (C.1.7) is positive denite,
implying that the U(1) bration must have positive scalar curvature. In contrast,
in 2n + 1 Lorentzian dimensions, the LHS of (C.1.7) is indenite, and thus does
not restrict the sign of the scalar curvature of the U(1) bration. It is not even
necessary to assume that the U(1) bration has non-zero scalar curvature. Indeed, it
is not dicult to see from (C.1.7) that the (2n+1)-dimensional Lorentzian signature
Einstein equation admits Ricci at ( = 0) U(1) bration solutions, provided that

n
i=1

i
v
2
i
= 0. The scalar curvatures of the 2-manifolds in the base are according
to (C.1.6) given by
i
=
1
2

i
v
2
i
. It follows that in 2n+1 5 Lorentzian dimensions,
there exit Ricci at U(1) brations over AdS
2
S
2
S
2
.
Since the ansatz (C.1.2) for the metric d s
2
is written in Kaluza-Klein form,
and the coordinate z along the bre is naturally periodic, it is possible to consider
the N
2n+1
(v
1
, . . . , v
n
) U(1) bration in the context of Kaluza-Klein dimensional re-
duction. The metric ds
2
on the base manifold and the U(1) connection /
(1)
are
interpreted as the (D = 2n)-dimensional metric and Kaluza-Klein potential coming
from the dimensional reduction of the (D+1)-dimensional metric d s
2
. This partic-
ular reduction scheme where the compact direction is along a U(1) bre has been
APPENDIX C. U(1) FIBRATIONS AND HOPF REDUCTION 209
dubbed Hopf reduction [91], and we shall continue to use this terminology here.
A comparison of the Hopf reduction ansatz (C.1.2) with the usual Kaluza-Klein
ansatz for dimensional reduction from D + 1 to D dimensions on a circle
ds
2
D+1
= e
2
ds
2
D
+e
2
(dz +/
(1)
)
2
, (C.1.8)
reveals that the dening characteristic of the Hopf reduction ansatz is that the
D-dimensional dilaton has been truncated. It is well known that in general the
truncation = 0 is inconsistent, in the sense that is does not commute with the
variation of the D + 1-dimensional Lagrangian that gives the Einstein equation
(C.1.1). Note that in the case of the N
2n+1
(v
1
, . . . , v
n
) solution, consistency has
been achieved by substituting the Hopf reduction ansatz directly in the (D + 1)-
dimensional equations of motion. It follows that the base metric ds
2
and the U(1)
potential /
(1)
in the N
2n+1
(v
1
, . . . , v
n
) solution are interpreted as a constant scalar
solution of the D-dimensional theory from the Kaluza-Klein perspective.
C.2 Supersymmetry of U(1) brations
If we consider the U(1) bration solutions of the Einstein equation (C.1.1) within
the context of a (D+1)-dimensional supergravity theory, whose bosonic Lagrangian
is given by

L
(D+1)
=

1D(D 1) , (C.2.1)
it is of interest to determine whether these solutions preserve any residual supersym-
metry. The condition for preserved supersymmetry for a purely bosonic solution is
the existence of a Killing spinor , i.e. a solution of the (D+1)-dimensional Killing
APPENDIX C. U(1) FIBRATIONS AND HOPF REDUCTION 210
spinor equation

D
A
=
1
2

A
, (C.2.2)
where
2
= . The integrability conditions for the Killing spinor equation are
[

D
A
,

D
B
] =
1
2

A

B

1
4

1
A

C

D

C

D
. (C.2.3)
For the N
2n+1
(v
1
, . . . , v
n
) solution, the non-zero tangent space components of
the Riemann tensor are

1
AB
CD
= 1
AB
CD

1
2
T
AB
T
CD

1
2
T
A
[C
T
B
D]
,

1
AzBz
=
1
4
T
AC
T
B
C
, (C.2.4)
where 1
AB
CD
is the Riemann tensor for the base metric ds
2
. Note that within each
Einstein 2-manifold in the base, with scalar curvature
i
, 1
AB
CD
is simply given by
1
ab
cd
= 2
i

a
[c

b
d]
, (C.2.5)
where a, b, . . . = 1, 2 are tangent space indices for the 2-manifold in question. The
U(1) eld strength T
(2)
is given in (C.1.5).
We nd that the integrability conditions (C.2.3) for the Killing spinor equation
are satised, provided that the constants v
i
satisfy

i
v
2
i
= 4 , i 1, . . . , n (C.2.6)
and the Killing spinor satises the
1
2
n(n 1) commuting projection conditions
P
ij
= 0, P
ij
=
1
2
_
1 +

i

j
v
i
v
j
4

12

2
_
, i ,= j 1, . . . , n, (C.2.7)
where 1, 2 and

1,

2 are tangent space labels for the directions in i


th
, respectively
j
th
2-manifold. It follows that if we can satisfy the condition (C.2.6), the 2n + 1
APPENDIX C. U(1) FIBRATIONS AND HOPF REDUCTION 211
dimensional U(1) bration N
2n+1
(v
1
, . . . , v
n
) preserves a fraction of 2

1
2
n(n1)
of
the maximal number of supersymmetries.
Combining equations (C.1.6) and (C.2.6), we see that for supersymmetric solu-
tions the scalar curvatures of the base submanifolds are equal to

i
= 2(n + 1) =
1
2

i
(n + 1)v
2
i
, i 1, . . . , n. (C.2.8)
Note also that the other condition (C.1.7), which follows from the eld equations, is
already implied by (C.2.6). This observation should not come as a surprise, since the
supersymmetry integrability condition (C.2.3) implies the Einstein equation (C.1.1).
It is now easy to list the supersymmetric N
2n+1
(v
1
, . . . , v
n
) solutions of the
(2n + 1)-dimensional Einstein equation. For n=1, there are two supersymmetric
solutions. If the 2-dimensional base manifold has Lorentzian signature, it has nega-
tive curvature and hence is AdS
2
. In this case the U(1) bration is

Q
3
(v)

= AdS
3
,
which preserves the full amount of supersymmetry. If, one the other hand, the 2-
dimensional base manifold has Euclidean signature, it has positive curvature and
hence is S
2
. The U(1) bration is Q
3
(v)

= S
3
, which also preserves the maximum
number of supersymmetries. For n 2, we see from (C.2.6) that the n 2-manifolds
that form the base must all have the same signature. As we are not considering
many time theories here, they all have Euclidean signature, and hence positive
curvature. The base manifold is thus the n-fold product S
2
S
2
, and the U(1)
brations are the spaces Q
2n+1
(v, . . . , v). They preserve a fraction of 2

1
2
n(n1)
of
the maximal number of supersymmetries.
We have shown above that the U(1) brations in 2n + 1 dimensions lend them-
selves naturally to Hopf dimensional reduction to 2n dimensions. In this context,
APPENDIX C. U(1) FIBRATIONS AND HOPF REDUCTION 212
it is important to determine how many of the supersymmetries survive the Hopf
reduction
1
. To this end, it is sucient to consider the z-component of the Killing
spinor equation (C.2.2), which, for the metric (C.1.2), is given by

z
=
1
2
_
1
4
T
AB

AB
+
z
_
(C.2.9)
Note that for the metric (C.1.2) we do not need to distinguish between the tan-
gent space index and the spacetime index along the direction of the U(1) bre. In
particular, the derivative on the right hand side of (C.2.9) is an ordinary partial
derivative.
Consider rst the

Q
3
(v)

= AdS
3
solution. It follows from (C.2.6) that v = 2
and hence that (C.2.9) becomes

z
=
1
2
(
01
+
z
) , (C.2.10)
where 0, 1 label the directions in the AdS
2
base. Now recall that the Lorentzian
signature D = 3 Cliord algebra implies that
z
=
01
. Therefore, it is always pos-
sible to correlate the sign choices in such a way that the Killing spinor is independent
of the U(1) bre coordinate. In that case, the maximum number of supersymmetries
survive the Hopf reduction. Otherwise, the supersymmetry is fully broken in the
reduction process.
A similar situation obtains for the Q
3
(v)

= S
3
solution. Equation (C.2.6) now
implies that v = i2 and, therefore, (C.2.9) becomes

z
=
1
2
(i
12
+
z
) , (C.2.11)
1
Similar results have previously been obtained in Ref. [91], in parts by solving the Killing spinor
equation explicitly.
APPENDIX C. U(1) FIBRATIONS AND HOPF REDUCTION 213
where 1, 2 label the directions in the S
2
base. By correlating the sign convention
for
z
= i
12
with the sign choice in the Killing spinor equation, it is again
possible to have a z-independent Killing spinor. In that case, the maximum number
of supersymmetries survive the Hopf reduction. Otherwise, the supersymmetry is
fully broken in the reduction process.
In contrast, we nd that the Killing spinor for the Q
2n+1
(v, . . . , v) solutions with
n 2 depends invariably on the reduction coordinate. Without loss of generality, we
may x the v
i
according to (C.2.6) by v
i
= 2i . It then follows from the projection
conditions (C.2.7) that

z
=
1
2
(i n
12
+
z
) , (C.2.12)
where 1, 2 label the directions on, say, the rst of the S
2
in the base manifold.
The matrix
z
is now represented as
z
=
122n
, where is such that (
z
)
2
= 1.
The projection conditions (C.2.7) allow us to simplify the Killing spinor equation
further to

z
=
1
2
(i n
12
+ (
12
)
n
) , (C.2.13)
It is not dicult to realise that for n 2 it is impossible to restrict the Killing
spinor such that
z
= 0 admits a non-trivial solution. The Killing spinor for the
Q
2n+1
(v, . . . , v) solution depends necessarily on the U(1) bre coordinate that is
being compactied in the Hopf reduction scheme. Therefore, the 2n-dimensional
constant scalar S
2
S
2
solution that comes from the Hopf reduction of the
Q
2n+1
(v, . . . , v) solution is non-supersymmetric.
Bibliography
[1] P.K. Townsend, Four lectures on M-theory, hep-th/9612121.
[2] M.J. Du and K.S. Stelle, Multi-membrane solution of D = 11 supergravity,
Phys.Lett. B253 (1991) 113.
[3] R. G uven, Black p-brane solutions of D = 11 supergravity theory, Phys.Lett.
B276 (1992) 49.
[4] E. Cremmer, B. Julia and J. Scherk, Supergravity theory in eleven dimensions,
Phys.Lett. B76 (1978) 409.
[5] J.P. Gauntlett, Intersecting branes, hep-th/9705011.
[6] H. Nicolai, P.K. Townsend and P. van Nieuwenhuizen, Comments on eleven-
dimensional supergravity, Lett.Nuov.Cim. 30 (1981) 315.
[7] E. Bergshoe, M.J. Du, C.N. Pope and E. Sezgin, Supersymmetric superme-
mbrane vacua and singletons, Phys.Lett. B199 (1987) 69; Compactications
of the eleven-dimensional supermembrane, Phys.Lett. B224 (1989) 71.
214
BIBLIOGRAPHY 215
[8] E. Bergshoe, E. Sezgin and P.K. Townsend, Supermembranes and eleven-
dimensional supergravity, Phys.Lett. B189 (1987) 75; Properties of the eleven-
dimensional supermembrane theory, Ann.Phys. 185 (1988) 330.
[9] G.W. Gibbons and P.K. Townsend, Vacuum interpolation in supergravity via
super p-branes, Phys.Rev.Lett. 71 (1993) 3754, hep-th/9307049.
[10] M. Aganagic, J. Park, C. Popescu and J.H. Schwarz, World-volume action of
the M-theory ve-brane, Nucl.Phys. B496 (1997) 191, hep-th/9701166.
[11] I. Bandos, K. Lechner, A. Nurmagambetov, P. Pasti, D. Sorokin and M.
Tonin, Covariant action for the super-ve-brane of M-theory, Phys.Rev.Lett.
78 (1997) 4332, hep-th/9701149.
[12] S.P. de Alwis, Coupling of branes and normalisation of eective actions in
string/M-theory, Phys.Rev. D56 (1997) 7963, hep-th/9705139.
[13] I. Bandos, N. Berkovits and D. Sorokin, Duality-symmetric eleven-dimensional
supergravity and its coupling to M-branes, Nucl.Phys. B522 (1998) 214, hep-
th/9711055.
[14] M.J. Du, R.R. Khuri, J.X. Lu, String solitons, Phys.Rep. 259 (1994) 213,
hep-th/9412184.
[15] M.J. Du, Supermembranes, hep-th/9611203.
[16] K.S. Stelle, Lectures on supergravity p-branes, hep-th/9701088; BPS branes
in supergravity, hep-th/9803116.
BIBLIOGRAPHY 216
[17] P. Pasti, D. Sorokin and M. Tonin, On Lorentz invariant actions for chiral
p-forms, Phys.Rev. D55 (1997) 6292, hep-th/9611100; Covariant action for
a D=11 ve-brane with the chiral eld, Phys.Lett. B398 (1997) 41, hep-
th/9701037.
[18] P.S. Howe, E. Sezgin and P.C. West, Covariant eld equations of the M theory
ve-brane, Phys.Lett. B399 (1997) 49, hep-th/9702008.
[19] I. Bandos, K. Lechner, A. Nurmagambetov, P. Pasti, D. Sorokin and M.
Tonin, On the equivalence of dierent formulations of the M theory ve-brane,
Phys.Lett. B408 (1997) 135, hep-th/9703127.
[20] W. Thirring, A Course in Mathematical Physics, Vols. 1 and 2, (New York:
Springer, 1978).
[21] O. Aharony, String dualities from M theory, Nucl.Phys. B474 (1996) 309,
hep-th/9604103.
[22] E. Bergshoe, M. de Roo and T. Ortin, The eleven-dimensional ve-brane,
Phys.Lett. B386 (1996) 85, hep-th/9606118.
[23] P.A.M. Dirac, Theory of magnetic poles, Phys.Rev. 74 (1948) 817.
[24] C. Teitelboim, Gauge invariances for extended objects, Phys.Lett. B167
(1986) 63; Monopoles of higher rank, Phys.Lett B167 (1986) 69.
[25] P. Goddard and D. Olive, New developments in the theory of magnetic
monopoles, Rep.Prog.Phys. 41 (1978) 1357.
BIBLIOGRAPHY 217
[26] M. Henneaux and C. Teitelboim, P-form electrodynamics, Found.Phys. 16
(1986) 593.
[27] P.K. Townsend, D-branes from M-branes, Phys.Lett. B373 (1996) 68, hep-
th/9512062.
[28] A. Strominger, Open p-branes, Phys.Lett. B383 (1996) 44, hep-th/9612059.
[29] A. Dabholkar, G. Gibbons, J.A. Harvey and F. Ruiz Ruiz, Superstrings and
solitons, Nucl.Phys. B340 (1990) 33.
[30] M.J. Du, G.W. Gibbons and P.K. Townsend, Macroscopic superstrings as
interpolating solitons, Phys.Lett B332 (1994) 321, hep-th/9405124.
[31] G.W. Gibbons, G.T. Horowitz and P.K. Townsend, Higher-dimensional reso-
lution of dilatonic black hole singularities, Class.Quant.Grav. 12 (1995) 297,
hep-th/9410073.
[32] J.X. Lu, ADM masses for black strings and p-branes, Phys.Lett. B313 (1993)
29, hep-th/9304159.
[33] D.N. Page, Classical stability of round and squashed seven spheres in eleven-
dimensional supergravity, Phys.Rev. D28 (1983) 2976.
[34] M.J. Du, Problems in the classical and quantum theories of gravity, Ph.D.
Thesis, Imperial College, London, 1972.
[35] M.J. Du, J.M. Evans, R.R. Khuri and J.X. Lu, The octionic membrane,
Phys.Lett. B412 (1997) 281, hep-th/9706124.
BIBLIOGRAPHY 218
[36] M.J. Du, R.R. Khuri and J.X. Lu, String and ve-brane solitons: Singular
or non-singular?, Nucl.Phys. B377 (1992) 281, hep-th/9112023.
[37] S. Hyun, U-duality between Three and Higher Dimensional Black Holes, hep-
th/9704005.
[38] H.J. Boonstra, B. Peters and K. Skenderis, Duality and asymptotic geome-
tries, Phys.Lett. B411 (1997) 59, hep-th/9706192; Branes and anti-de Sitter
spacetimes, Fortschr.Phys. 47 (1999) 109, hep-th/9801076; Brane intersec-
tions, anti-de-Sitter spacetimes and dual superconformal theories, Nucl.Phys.
B533 (1998) 127, hep-th/9803231.
[39] K. Sfetsos and K. Skenderis, Microscopic derivation of the Beckenstein-
Hawking entropy formula for non-extremal Black holes, Nucl.Phys. B517
(1998) 179, hep-th/9711138.
[40] E. Bergshoe and K. Behrndt, D-Instantons and asymptotic geometries,
Class.Quant.Grav. 15 (1998) 1801, hep-th/9803090.
[41] E. Cremmer, I.V. Lavrinenko, H. L u, C.N. Pope, K.S. Stelle and T.A. Tran,
Euclidean signature supergravities, dualities and instantons, Nucl.Phys. B534
(1998) 40, hep-th/9803259.
[42] P. Claus, R. Kallosh, J. Kumar, P.K. Townsend and A. van Proeyen, Con-
formal Theory of M2, D3, M5 and D1+D5 Branes, JHEP 06 (1998) 004,
hep-th/9801206.
[43] G.T. Horowitz and A. Strominger, Black Strings and p-Branes, Nucl.Phys.
B360 (1991) 197.
BIBLIOGRAPHY 219
[44] M.J. Du and J.X. Lu, Black and Super p-branes in Diverse Dimensions,
Nucl.Phys. B416 (1994) 301, hep-th/9306052.
[45] R. Arnowitt, S. Deser, C.W. Misner, The dynamics of General Relativity, in
L. Witten (ed.), Gravitation: an introduction to current research, (New York:
Wiley, 1962).
[46] M.J. Du, Anti-de Sitter space, branes, singletons, superconformal eld theo-
ries and all that, Int.J.Mod.Phys. A14 (1999) 815, hep-th/9808100.
[47] M.J. Du and J.X. Lu, The self-dual type IIB superthreebrane, Phys.Lett.
B273 (1991) 409.
[48] J. Rahmfeld, Extreme black holes as bound states, Phys.Lett. B372 (1996)
198, hep-th/9512089.
[49] N. Khviengia Z. Khviengia, H. L u and C.N. Pope, Intersecting Mp-branes and
bound states, Phys.Lett. B388 (1996) 21, hep-th/9605077.
[50] M.J. Du and J. Rahmfeld, Bound states of black holes and other p-branes,
Nucl.Phys. B481 (1996) 332, hep-th/9605085.
[51] M.J. Du, S. Ferrara, R. Khuri and J. Rahmfeld, Supersymmetry and dual
string solitons, Phys.Lett. B356 (1995) 479, hep-th/9506957.
[52] A.A. Tseytlin, Extreme dyonic black holes in string theory, Mod.Phys.Lett.
A11 (1996) 689, hep-th/9601177.
[53] M. Cvetic and A.A. Tseytlin, General class of BPS saturated balck holes as
exact superstring solutions, Phys.Lett. B366 (1996) 95, hep-th/9510097; Soli-
BIBLIOGRAPHY 220
tonic strings and BPS saturated dyonic black holes, Phys.Rev. D53 (1996)
5619, hep-th/9512031.
[54] R. Kallosh and J. Kumar, Supersymmetry enhancement of D-p-branes and
M-branes, Phys.Rev. D56 (1997) 4934, hep-th/9704189.
[55] G. Papadopoulos and P.K. Towsend, Intersecting M-branes, Phys.Lett. B380
(1996) 273, hep-th/9603087.
[56] J.P. Gauntlett, D.A. Kastor and J. Traschen, Overlapping Branes in M-
Theory, Nucl.Phys. B478 (1996) 544, hep-th/9604179.
[57] A.A. Tseytlin, Harmonic superpositions of M-branes, Nucl.Phys. B475 (1996)
149, hep-th/9604035.
[58] R. Kallosh and A. Peet, Dilaton black foles near the horizon, Phys.Rev. D46
(1992) 5223, hep-th/9209112.
[59] A. Chamseddine, S. Ferrara, G.W. Gibbons and R. Kallosh, Enhancement
of supersymmetry near 5d black hole horizon, Phys.Rev. D55 (1997) 3647,
hep-th/9610155.
[60] P.M. Cowdall and P.K. Townsend, Gauged supergravity vacua from inter-
secting branes, Phys.Lett. B429 (1998) 218; Erratum-ibid. B434 (1998) 456,
hep-th/9801165.
[61] A.A. Teytlin, Composite BPS congurations of p-branes in 10 and 11 dimen-
sions, Class.Quant.Grav. 14 (1997) 2085, hep-th/9702163.
BIBLIOGRAPHY 221
[62] J.P. Gauntlett, G.W. Gibbons, G. Papadopoulos and P.K. Townsend, Hyper-
Kahler manifolds and multiply intersecting branes, Nucl.Phys. B500 (1997)
133, hep-th/9707232.
[63] E. Cremmer and J. Scherk, Spontaneous compactication of extra space di-
mensions, Nucl.Phys. B118 (1977) 61.
[64] M.J. Du and C.N. Pope, Consistent truncations in Kaluza-Klein theories,
Nucl.Phys. B255 (1985) 355.
[65] P.G.O. Freund and M.A. Rubin, Dynamics of dimensional reduction,
Phys.Lett. B97 (1980) 233.
[66] M.J. Du, B.E.W. Nilsson and C.N. Pope, Kaluza-Klein supergravity,
Phys.Rep. 130 (1986) 1.
[67] K. Pilch, P.K. Townsend and P. van Nieuwenhuizen, Compactication of d =
11 supergravity and S
4
(or 11 = 7 + 4, too), Nucl.Phys. B242 (1984) 377.
[68] B. de Wit and H. Nicolai, The consistency of the S
7
truncation in D = 11
supergravity, Nucl.Phys. B281 (1987) 211.
[69] M. Pernici, K. Pilch and P. van Nieuwenhuizen, Gauged maximally extended
supergravity in seven dimensions, Phys.Lett. B143 (1984) 103.
[70] B. de Wit and H. Nicolai, N=8 supergravity, Nucl.Phys. B208 (1982) 323.
[71] H. Nastase, D. Vaman and P. van Nieuwenhuizen, Consistent non-linear KK
reduction of 11d supergravity on AdS
7
S
4
and self-duality in odd dimensions,
hep-th/9905075.
BIBLIOGRAPHY 222
[72] M.S. Bremer, M.J. Du, H. L u, C.N. Pope and K.S. Stelle, Instanton cos-
mology and domain walls from M-theory and string theory, Nucl.Phys. B543
(1999) 321, hep-th/9807051.
[73] C.M. Hull and B. Julia, Duality and moduli spaces for time-like reductions,
Nucl.Phys. B534 (1998) 250, hep-th/9803239.
[74] J. Scherk and J.H. Schwarz, Spontaneous breaking of supersymmetry through
dimensional reduction, Phys.Lett. B82 (1979) 60; How to get masses from
extra dimensions, Nucl.Phys. B153 (1979) 61.
[75] E. Bergshoe, M. de Roo, M.B. Green, G. Papadopoulos and P.K. Townsend,
Duality of Type II 7-branes and 8-branes, Nucl.Phys. B470 (1996) 113, hep-
th/9601150.
P.M. Cowdall, H. L u, C.N. Pope, K.S. Stelle and P.K. Townsend, Domain
walls in massive supergravities, Nucl.Phys. B486 (1997) 49, hep-th/9608173.
N. Kaloper, R.R. Khuri and R.C. Meyers, On generalized axion reductions,
Phys.Lett. B428 (1998) 297, hep-th/9803066.
[76] I.C.G. Campbell and P.C. West, N=2 D=10 non-chiral supergravity and its
spontaneous compactications, Nucl.Phys. B243 (1984) 112.
[77] B.E.W. Nilsson and C.N. Pope, Hopf bration of eleven dimensional super-
gravity, Class.Quant.Grav. 1 (1984) 499.
[78] H. L u, C.N. Pope and K.S. Stelle, Weyl group invariance and p-brane multi-
plets, Nucl.Phys. B476 (1996) 89, hep-th/9602140.
BIBLIOGRAPHY 223
[79] J.H. Schwarz, Covariant eld equations of chiral N = 2 D = 10 supergravity,
Nucl.Phys. B226 (1983) 269.
[80] E. Cremmer, B. Julia, H. L u and C.N. Pope, Dualisation of Dualities. I.,
Nucl.Phys. B523 (1998) 73, hep-th/9710119.
[81] E. Cremmer, The N = 8 supergravity, in S. Ferrara, J. Ellis and P. van
Nieuwenhuizen (eds.), Unication of the fundamental particle interactions,
(New York: Plenum, 1980).
[82] J.H. Schwarz, N = 8 supergravity in various dimensions and the implication
for four dimensions, Phys.Lett. B95 (1980) 219.
[83] H. L u, C.N. Pope, T.A. Tran and K.-W. Xu, Classication of p-branes, NUTs,
waves and intersection, Nucl.Phys. B511 (1998) 98, hep-th/9708055.
[84] H. L u, C.N. Pope and P.K. Townsend, Domain walls from anti-de Sitter space-
time, Phys.Lett. B391 (1997) 39, hep-th/9607164.
[85] M. Huq and M. Namazie, Kaluza-Klein supergravity in ten dimensions,
Class.Quant.Grav. 2 (1985) 293;.Erratum-ibid. 2 (1985) 597.
[86] H. L u, C.N. Pope, p-brane solitons in maximal supergravities, Nucl.Phys.
B465 (1996) 127, hep-th/9512012.
[87] E. Cremmer, B. Julia, H. L u and C.N. Pope, Dualisation of Dualities II:
Twisted self-duality of doubled eld and superdualities, Nucl.Phys. B535
(1998) 242, hep-th/9806106.
BIBLIOGRAPHY 224
[88] P. Howe and P. West, The complete N = 2 D = 10 supergravity, Nucl.Phys.
B238 (1984) 181.
[89] E. Bergshoe, H.J. Boonstra and T. Ortin, S-duality and dyonic p-brane so-
lutions in type II string theory, Phys.Rev. D53 (1996) 89, hep-th/9602140.
[90] J.H. Schwarz and P. West, Symmetries and transformations of chiral N = 2
and D = 10 supergravity, Phys.Lett. 126B (1983) 301.
[91] M.J. Du, H. L u and C.N. Pope, Supersymmetry without supersymmetry,
Phys.Lett. B409 (1997) 136, hep-th/9704186; AdS
5
S
5
untwisted, Nucl.Phys.
B532 (1998) 181, hep-th/9803061; AdS
3
S
3
(un)twisted and squashed, and
an O(2, 2, ZZ) multiplet of dyonic strings, Nucl.Phys. B544 (1999) 145, hep-
th/9807173.
[92] L. Castellani, L.J. Romans and N.P. Warner, A classication of compactifying
solutions for D = 11 supergravity, Nucl.Phys. B241 (1984) 429.
[93] L.J. Romans, New compactications of chiral N = 2, D = 10 supergravity,
Phys.Lett. B153 (1985) 392.

Potrebbero piacerti anche