(Cambridge Monographs On Mathematical Physics) Alberto A. García-Díaz - Exact Solutions in Three-Dimensional Gravity-Cambridge University Press (2017)
(Cambridge Monographs On Mathematical Physics) Alberto A. García-Díaz - Exact Solutions in Three-Dimensional Gravity-Cambridge University Press (2017)
GRAVITY
A L B E RTO A . G A RC ÍA-D Í A Z
Center for Research and Advanced Studies of the
National Polytechnic Institute, Mexico
(CINVESTAV)
University Printing House, Cambridge CB2 8BS, United Kingdom
One Liberty Plaza, 20th Floor, New York, NY 10006, USA
477 Williamstown Road, Port Melbourne, VIC 3207, Australia
4843/24, 2nd Floor, Ansari Road, Daryaganj, Delhi – 110002, India
79 Anson Road, #06–04/06, Singapore 079906
www.cambridge.org
Information on this title: www.cambridge.org/9781107147898
DOI: 10.1017/9781316556566
c Alberto A. Garcı́a-Dı́az 2017
This publication is in copyright. Subject to statutory exception
and to the provisions of relevant collective licensing agreements,
no reproduction of any part may take place without the written
permission of Cambridge University Press.
First published 2017
Printed in the United Kingdom by Clays, St Ives plc
A catalogue record for this publication is available from the British Library.
Library of Congress Cataloging-in-Publication Data
Names: Garcı́a-Dı́az, Alberto A., 1942- author.
Title: Exact solutions in three-dimensional gravity / Alberto A.
Garcı́a-Dı́az (Center for Research and Advanced Studies of the National
Polytechnic Institute, Mexico (CINVESTAV).
Other titles: Cambridge monographs on mathematical physics.
Description: Cambridge, United Kingdom ; New York, NY : Cambridge University
Press, 2017. | Series: Cambridge monographs on mathematical physics
Identifiers: LCCN 2017012130 | ISBN 9781107147898 (hardback ; alk. paper) |
ISBN 1107147891 (hardback ; alk. paper)
Subjects: LCSH: Gravitation–Mathematics. | Quantum gravity–Mathematics. |
Einstein field equations. | Space and time.
Classification: LCC QC178 .G365 2017 | DDC 539.7/54–dc23
LC record available at https://lccn.loc.gov/2017012130
ISBN 978-1-107-14789-8 Hardback
Cambridge University Press has no responsibility for the persistence or accuracy of
URLs for external or third-party internet websites referred to in this publication,
and does not guarantee that any content on such websites is, or will remain,
accurate or appropriate.
Contents
1 Introduction 1
1.1 Main Features of (2 + 1) Gravity 1
1.1.1 Field Equations and Curvature Tensors 2
1.1.2 Matter Distribution Locally Curves the Spacetime 2
1.1.3 Point Particles Produce Global Effects on the Spacetime 3
1.1.4 Newtonian Limits 3
1.1.5 No Geodesic Deviation for Dust 5
1.1.6 No Dynamic Degrees of Freedom 6
1.1.7 Black Holes in (2 + 1) Gravity 7
1.1.8 Gravity in the Presence of Other Fields and Matter 7
1.2 Algebraic Classification 7
1.2.1 Classification of the Cotton–York Tensor 7
1.2.2 Classification of the Energy–Momentum Tensor 9
1.2.3 Classification of the Traceless Ricci Tensor 11
1.3 Brown–York Energy, Mass, and Momentum for Stationary
Metrics 11
1.3.1 Summary of Quasilocal Mass, Energy, and Angular
Momentum 14
1.4 Decomposition with Respect to a Frame of Reference 15
1.4.1 Kinematics of the Frame 15
1.4.2 Perfect Fluid Referred to a Frame of Reference 16
3 Dust Solutions 27
3.1 Cornish–Frankel Dust Heaviside Function Solution 27
3.2 Giddings–Abott–Kuchař Dust Solutions 28
3.2.1 Time-Dependent Class of Dust Solutions
Ω = ln (tf (x, y)) 29
3.2.2 Static Class of Dust Solutions Ω = ln g(x, y) 30
3.3 Barrow–Shaw–Tsagas Anisotropic Dust Solution; Λ = 0 30
3.4 BST Diagonal Anisotropic Dust Solutions with Λ 33
3.5 BST (t, x, y)-Dependent Cosmological Solutions with
Comoving Dust 34
3.5.1 BST Class 2 of Solutions 37
3.5.2 BST Class 1 Spacetime 38
3.5.3 BST Class 3 of Dust Solutions 39
3.6 Rooman–Spindel Dust Gödel Non-Diagonal Model 40
7 Hydrodynamic Equilibrium 80
7.1 Generalized Buchdahl’s Theorem 80
7.2 Stellar Equilibrium in (2 + 1) Dimensions with Λ 81
7.2.1 Cruz–Zanelli Existence of Hydrostatic Equilibrium for
Λ≤0 83
7.2.2 No Buchdahl’s Inequality in (2 + 1) Hydrostatics 83
7.2.3 Static Star with Constant Density μ0 and Λ = −1/l2 ≤ 0 84
7.3 Buchdahl Theorem in d Dimensions 85
7.3.1 Buchdahl’s Inequalities 86
7.3.2 Constant Density Solution 89
10 Dilaton–Inflaton Friedmann–Robertson–Walker
Cosmologies 121
10.1 Equations for a FRW Cosmology with a Perfect Fluid and a
Scalar Field 122
10.1.1 Einstein Equations for (3+1) FRW Dilaton Cosmology 122
10.1.2 Einstein Equations for (2+1) FRW Cosmology 123
10.1.3 Correspondence Between (3+1) and (2+1) Solutions 125
10.2 Single Scalar Field to Linear State Equations; Λ = 0 127
10.2.1 (2+1) Solutions for a Scalar Field 127
10.2.2 (3+1) Solutions for a Scalar Field 128
10.2.3 Slow Roll Spatially Flat FRW Solutions 129
10.3 Spatially Flat FRW Solutions for Barotropic Perfect Fluid and
Scalar Field 131
10.3.1 Spatially Flat FRW (3+1) Solutions ; γ4 = 2Γ4 131
10.3.2 Spatially Flat FRW (2+1) Solutions; γ3 = 2Γ3 133
10.3.3 Barrow–Saich Solution; γ = 2 Γ 134
10.4 Single Scalar Field Spatially Flat FRW Solutions to
pφ + ρφ = Γ ρφ β 135
10.4.1 Spatially Flat (3+1) Solutions with
V (φ) = A(αφ2/(1−β) − φ2β/(1−β) ) 135
10.4.2 Spatially Flat (2+1) Solutions with
V (φ) = A(αφ2/(1−β) − φ2β/(1−β) ) 136
10.4.3 Barrow–Burd–Lancaster (2+1) and Madsen (3+1)
Solutions 137
10.5 Scalar Field Solutions for a Given Scale Factor 139
10.5.1 Second (2+1) BBL Solution 139
10.5.2 (3+1) Generalization of the Second (2+1) BBL
Solution 141
Contents xi
References 421
Index 430
Preface
I have been working on this book project over a period of many years in order
to create a concise but comprehensive account of exact solutions in the three-
dimensional Einstein theory of gravity. This theory, although tangentially related
to the real gravity world, is a good model from which to extract some relevant
conclusions about that world, taking advantage of simplifications due to the
reduction in dimensions. Of course, after fifty years of existence, (2 + 1) gravity
has been approached from many perspectives, but, to my mind, there is still
no satisfactory account of solutions of physical interest. I therefore wrote the
present book on exact solutions in (2 + 1) gravity following closely the pattern
of the classic book, Exact solutions in Einstein’s field equations with the aim of
bringing to each class of solutions presented its theoretical framework, its general
physical and geometrical characteristics, and references to the researchers who
discovered or studied them.
I am greatly indebted to my late friends and collaborators J.F. Plebański, N.
R. Sibgatullin and S. del Campo for stimulating talks. Many thanks are due to
my collaborators E. Ayón-Beato, C. Campuzano, M. Cataldo, G. Gutierrez, F.
Hehl, A. Macı́as, V. Manko, N. Mitskievich, and C. Terrero for many enlightening
talks, and advice on various problems I faced when dealing with this task. This
work could not have come about without a fruitful sabbatical stay at the Physics
Department of the California University at Davis under the sponsorship of Pro-
fessor S. Carlip, to whom I am indebted, and to UC MEXUS-CONACYT for
financial support. I cordially thank my friends, particularly Y. Gurievich and M.
Lopez, and colleagues from the Departamento de Fı́sica, Centro de Investigación
y de Estudios Avanzados del I.P.N., for their encouragement. Many thanks to
my assistant E.Vargas-Dı́az for her daily support and typographical help in the
final preparation of the manuscript of this book. The support of the Consejo
Nacional de Ciencia y Tecnologı́a (CONACYT), México, through various grants
(at present: Grant CONACyT 178346) is acknowledged.
I really do appreciate the patience and support of my daughters, Ana Alicia
and Alexandra Sofia, and sons, Sergey, Albert, Alberto-Tiko, and Alex, over all
these years of writing this book.
Please do not hesitate to let me know about inconsistencies, unevenness, omis-
sions or errors to be found in this text. Constructive comments are welcome:
aagarcia@fis.cinvestav.mx.
1
Introduction
Given the large number of exact solutions that exist today in (2 + 1) Einstein
gravity the purpose of the present book is to present a complete and concise list
of exact solutions with emphasis on their physical and geometrical properties
from the beginnings of the field in 1963 to the present, to be useful for the
audience of experts and young researchers. Emphasis is given to solutions to
the Einstein equations in the presence of matter and fields, for instance, point
particle solutions, perfect fluids, cosmological spacetimes, dilatons, inflatons, and
stringy solutions. The second part of this book deals with solutions to vacuum
topologically massive gravity with a cosmological constant, as there exist three
big families of spacetimes: the inhomogeneous Bianchi class of solutions, the
Kundt spacetimes and the Cotton type N wave fields.
To avoid unnecessary typing, the cosmological constant is denoted by Λ, AdS
spacetime stands for an asymptotically anti-de Sitter spacetime with Λ < 0,
dS spacetime stands for an asymptotically de Sitter spacetime with Λ > 0, 3D
stands for three dimensions, while (2+1)D spacetime denotes (2+1)-dimensional
spacetime, PF stands for perfect fluid, and ρ or μ denotes the fluid energy den-
sity. Occasionally we use SL for spacelike, TL for timelike, and ST to denote
spacetime. On the other hand, when publications by various authors are cited,
an abbreviation of their family names, including their first capital initials, are
given; for instance, EEqs. and EM mean, respectively, Einstein equations and
Einstein–Maxwell, MTW stands for Misner, Thorne, and Wheeler; FRW reads
Friedmann–Robertson–Walker, and BTZ denotes Bañados–Teitelboim–Zanelli.
κ
∇2 φ = ρ. (1.15)
2
In this limit, the geodesic equation (1.10) reduces to
d2 xi 1
− ∂i h00 = 0 (1.16)
dt2 2
and, taking into account (1.13) and (1.14), becomes
d2 xi n − 3 ij
= −2 δ ∂j φ. (1.17)
dt2 n−2
Discrepancies
Comparing these results reported by GAK, Giddings et al. (1984), and Carlip
(1998), one notices that the numerical coefficients in the GAK equations for the
limits of Newton and particle motion equations correspond to those in the limit
of the geodesic motion and to the Newton equation respectively of Carlip (1998).
∇u ∇u V α = Rα βγδ uβ uγ V δ . (1.18)
Assume now that this congruence is modeled by a tube of dust with energy–
momentum tensor
T αβ = ρ uα uβ . (1.19)
Substituting this tensor into the expression of the Riemann tensor (1.3) and
contracting it with u, one gets
Rα βγδ uβ uγ = 0, (1.20)
In n dimensions, the intrinsic metric possesses n(n − 1)/2 components, and the
conjugate momentum also has n(n − 1)/2 components, and together the number
of their components is n(n − 1). On the other hand, one can fix n of them by
choosing n coordinates, and additionally n by the constraints; consequently, the
number of degree of freedom in the canonical data is: n(n − 1) − 2n = n(n − 3).
Hence, in three dimensions there is no freedom in the prescription of the initial
data in the initial hypersurface.
As a consequence of this lack of degrees of freedom, there are no gravitational
waves in 3D flat spacetime; in the terminology of Gott and Alpert (1984), there
are no gravity waves in flatland; no gravitons.
where the symmetry has been introduced explicitly. Notice that the Cotton
tensor is traceless:
C α α = 0. (1.23)
To classify the Cotton tensor with respect to its eigenvalues, one has to solve a
generalized eigenvalue problem:
αβ
C − λ g αβ Vβ = 0 , C [αβ] = 0 , C αβ gαβ = 0 . (1.24)
By lowering one index, one can reformulate this task as an ordinary eigenvalue
problem for the matrix Cα β . However, in that case, the symmetry C αβ = C βα
is no longer present:
β
Cα − λ δαβ Vβ = 0 , Cα α = 0 . (1.25)
Accordingly, the matrix Cα β is no longer symmetric and the roots of the
characteristic polynomial
det Cα β − λ δαβ = 0 (1.26)
may be complex. This point seems to have been overlooked by Barrow et al.
(1986).
I will present a classification of Cα β along the lines of Garcia–Hehl–Heineke–
Macı́as (GHHM), Garcı́a et al. (2004), where the components are referred to an
orthonormal basis:
g = gαβ dxα dxβ = ηab Θa Θb , (ηab ) = diag(−1, 1, 1). (1.27)
The trace-free condition (1.25)2 reads explicitly
C1 1 + C2 2 + C3 3 = 0 . (1.28)
Accordingly, we can eliminate C3 3 , e.g., from (1.25)1 . Then the secular determi-
nant reads
C1 1 − λ C1 2 C1 3
det −C1 2 C2 2 − λ C2 3 = 0,
(1.29)
−C1 3
C2 3
−C1 − C2 2 − λ
2
This parallels exactly the Petrov classification of the Weyl tensor in four dimen-
sions in Stephani et al. (2003). This comes about since the Weyl tensor in 4D
is equivalent to a (complex) 3 × 3 trace-free matrix, as Cα β in 3D; for a similar
classification of Cαβ , see Hall and Capocci (1999). A detailed derivation of the
Cotton tensor in any dimension together with an account of its properties is
presented here in Chapter 20; see also Garcı́a et al. (2004).
with respect to the orthonormal basis (1.27). For that reason the eigenvectors
are found by solving the matrix equation
Tab V b = λ ηab V b , → (Tab − λ ηab ) V b = 0. (1.34)
Searching the values of λ that cancel the determinant of the matrix
⎡ ⎤
T11 + λ T12 T13
⎢ ⎥
⎢ T12 T22 − λ T23 ⎥, (1.35)
⎣ ⎦
T13 T23 T33 − λ
namely the roots of the eigenvalue polynomial
λ3 + c2 λ2 + c1 λ + c0 = 0,
c0 := T11 , T22 T33 − T13 2 T22 + 2 T12 T23 T13 − T12 2 T33 − T11 T23 2 ,
c1 := −T11 T22 − T11 T33 + T12 2 + T22 T33 + T13 2 − T23 2 ,
c2 := T11 − T22 − T33 , (1.36)
which allows for three roots, with its possible degenerations,
c2 1√ √
−3
λ1 = − Rd − 6 F Rd,
3
+
3 6
c2 1 √ √
−3 1 √ √ √
−3
λ2,3 = − − Rd + 3 F Rd ±
3 3
i 3 Rd + 36 F Rd ,
3 12 12
Rd := 36 c1 c2 − 108 c0 − 8 c2 3 + 12 D,
D := 12 c1 3 − 3 c1 2 c2 2 − 54 c1 c2 c0 + 81 c0 2 + 12 c0 c2 3 ,
1 1
F = c1 − c2 2 , (1.37)
3 9
one would be able to determine the eigenvectors corresponding to each root.
The nomenclature used for eigenvectors and algebraic types of tensors is bor-
rowed from Plebański (1964): timelike, spacelike, null, and complex vectors are
denoted respectively by T, S, N, and Z. For algebraic types are used the sym-
bols: {λ1 T, λ2 S2 , λ3 S3 } ≡ {T, S, S}, meaning that the first real eigenvalue λ1
gives rise to a timelike eigenvector T, the second real eigenvalue λ2 is associ-
ated with a spacelike eigenvector S2 , and finally the third real eigenvalue λ3 is
related to a spacelike eigenvector S3 ; for the sake of simplicity I use the sym-
bols {T, S, S}. It is clear that {N, N, S} stands for the algebraic type allowing
for two different real eigenvalues giving rise to two null eigenvectors while the
third real root is associated with a spacelike eigenvector. When there are single
and double real eigenvalues giving rise correspondingly to timelike and spacelike
eigenvectors, the algebraic type is denoted by {T, 2 S}; consequently, for a triple
real eigenvalue, if that were the case, the types could be {3T }, {3N }, or {3S}.
For a complex eigenvalue λZ , in general, the related eigenvectors are complex
and are denoted by Z and Z̄ for its complex conjugate; the possible types are
{T, Z, Z̄}, {N, Z, Z̄}, or {S, Z, Z̄}.
1.3 Brown–York Energy, Mass, and Momentum Stationary Metrics 11
1 μ W μ
uμ = −N δμt , uμ = δ − δ ,
N t N φ
1 r μ
nμ = δ , n = Lδrμ . (1.40)
L μ
Therefore the projection metrics are:
kμν = −σμα hβα nλ;β hλν = hμα Θαβ hβν , Θμν = −γμβ nν; β ,
which amounts to
φ t
kμν = −LKW K, r [W δμt δνt + 2δ(μ δν) ] − LKK, r δμφ δνφ , (1.50)
while
1 1
kμ ν = − LW K, r δφμ δνt − L K, r δφμ δνφ . (1.51)
K K
Rising with σφφ = K 2 = σ φφ one of the indexes of the component kφφ =
2
−LKK, r of the extrinsic curvature k associated with the metric of B , one arrives
at
1
k := kıı = σ φφ kφφ = − LK, r . (1.52)
K
Therefore the energy density becomes
1 cl 1
= k|0 = − LK, r |R − 0 . (1.53)
κ κK
As far
as the integral characteristics are concerned, the total quasilocal energy
√
E = B dx σ = 2π K is given by
π
E = −2 LK, r |R − 2π K(R)0 , (1.54)
κ
while the total
mass
∂ μ
related to the timelike Killing vector ξ μ = ( ∂t ) = δtμ ,
√
M ( ∂ t ) = − B dx σ( uμ + jμ )ξ amounts to
∂ μ
π πL 3
M (∂/∂t) = −2 N LK, r |R − K W W, r |R − 2π N K|R 0 . (1.55)
κ κN
14 Introduction
√
Finally, the total momentum J( ∂∂φ ) = B dx σjμ ζ μ associated with the Killing
vector ζ μ = ( ∂∂φ )μ = δφμ , is given by
2π
πL 3 1 J(R)
J(∂/∂φ) = dφ jφ K = K W, r |R , jφ = . (1.56)
0 κ N 2π K(R)
Incidentally, other representations of the mass and momentum density are:
M (∂/∂t) = N (R)E(R) − W (R)J(R). (1.57)
The extrinsic curvature Θμν = −γμα ∇α nν = −nν;α γ α μ of the surface boundary
2
B reduces to
Θμν = −L(KW 2 K, r + K 2 W W, r − N N, r )δμt δνt
t φ
−LK(2W K, r + KW, r )δ(μ δν) − LKK, r δμφ δνφ , (1.58)
with trace Θ equal to
L
Θ=− (KN, r + N K, r ), (1.59)
NK
which, used in the definition of the boundary momentum
1
π μν = − − det γAB (Θ γ μν − Θμν ),
2κ
taking into account that det γAB = −K 2 N 2 , gives
L L (μ ν)
π μν = − K, r δtμ δtν + (2W K, r + KW, r )δt δφ
2κ N 2κ N
L
+ (N N, r − W 2 KK, r − K 2 W W, r )δφμ δφν . (1.60)
2κ N K
This tensor is used in the construction of the stress tensor
2
sαβ = √ σ α π μν σνβ − s0 αβ . (1.61)
σN μ
where 0 is the reference energy density, which in the case of solutions with a
negative cosmological constant Λ = −1/l2corresponds to the density of the
2
anti-de Sitter spacetime, namely 0 = − πρ
1
1 + ρl2 . The momentum density is
determined from
K(ρ)2 L(ρ) d
j(ρ) = W (ρ). (1.63)
2 π N (ρ) dρ
The integral momentum J(ρ), global energy E(ρ), and integral mass M (ρ) are
correspondingly given by
J(ρ) = 2π K(ρ) j(ρ),
E(ρ) = 2π K(ρ) (ρ),
M (ρ) = E(ρ) K(ρ) − W (ρ) J(ρ). (1.64)
Evaluation of these physical quantities for the studied classes of solutions is
presented in the corresponding sections.
where
the shear : σab = D(a ub) , σab ub = 0, (1.69a)
where μ is the energy density, and p is pressure. Substituting the above into the
Einstein field equations (1.73), the latter read
Rab ua ub = 2(κp − Λ), hca hdb Rcd = [κ(μ − p) + 2Λ]hab , hba Rbc uc = 0. (1.76)
Explicitly, it reads
The timelike and spacelike parts of this relation lead to the 3D fluid conservation
laws:
for the energy density : ua ∇a μ := μ̇ = −Θ(μ + p), (1.79)
Notice that these conservation laws of a perfect fluid have the same functional
form as their 4D counterparts.
The intrinsic curvature of the 2D space orthogonal to ua is determined by
projecting the Riemann tensor, (1.2), onto the 2D space by means of hba :
Rabcd = hqa hsb hfc hpd Rqsf p − vac vbd + vad vbc ,
1
vab = Db ua = σab + ωab + Θhab , (1.81)
2
where vab is the relative position tensor. Note that vab characterizes the extrinsic
curvature (i.e., the second fundamental form of the space.) Assuming perfect-
fluid matter, i.e., equations (1.75), the projected Riemann tensor of the 2D
(spatial) sections reads
1
Rabcd = (κ μ − Θ2 + Λ)(hac hbd − had hbc )
4
1
− Θ [(σac + ωac )hbd + (σbd + ωbd )hac − (σad + ωad )hbc − (σbc + ωbc )had ]
2
−(σac + ωac )(σbd + ωbd ) + (σad + ωad )(σbc + ωbc ), (1.82)
1
Rabcd = (κ μ − Θ2 + Λ)(hac hbd − had hbc ). (1.83)
4
Defining
Rab = Rc acb (1.84)
as the local 2D Ricci tensor, by contracting expression (1.82) one obtains the
following 3D analogue of the Gauss–Codacci formula
1
Rab = (κ μ − Θ2 + σ 2 − ω 2 ) hab , (1.85)
4
where the shear and rotation scalars are defined correspondingly as
2σ 2 = σab σ ab , 2ω 2 = ωab ω ab . (1.86)
In deriving the above result (1.85), the relations
σc[a ω c b] = 0, σc[a σ c b] = 0, ωc[a ω c b] = 0 (1.87)
1 2
have been used. The former holds because σ12 (ω 1 + ω 2 ) = 0 (i.e., the sin-
gle independent component vanishes due to the trace-free nature of the shear).
Similarly, the two independent components of
σc(a σ c b) = 0
are also identically zero. Last, the result ωc(a ω c b) = 0 is guaranteed by the
relation ωab = ωηab and the properties of ηab . The absence of a skew and also of
a symmetric and trace-free part from Eq. (1.84) agrees with the symmetries of
the Riemann tensor in 2D spaces.
Finally, the trace of (1.84) leads to the curvature scalar of the spatial sections,
which may also be seen as the generalized Friedmann equation for 3D spacetimes:
1
R = Raa = 2(κ μ − Θ2 + σ 2 − ω 2 + Λ). (1.88)
4
Besides these theoretical aspects of fluids, the paper by Barrow et al. (2006)
deals with the kinematics of fluid matter, too. However, that material is outside
the scope of the present work.
2
Point Particle Solutions
A(x, y) = A0 → 1,
2
−2η ∂ η ∂2η
0
G0 = e + = 0 → η = F (x + iy) + F ∗ (x + iy), (2.12)
∂y 2 ∂x2
because of the flat Laplace character of the partial equation for η(x, y). In
particular, for a single point particle located at the point (x1 , y1 ), one may choose
−α
η(x, y) = ln (x − x1 )2 + (y − y1 )2 . (2.13)
which explicitly gives the relation between the exponent α for the single point
particle solution in isotropic coordinates with a mass M :
−2α 2
g = −dt2 + (x − x1 )2 + (y − y1 )2 dx + dy 2 ,
κ defθ
α=1− 1− M = . (2.18)
π 2π
where (x1 , y1 ) and (x2 , y2 ) stand for the spatial location of the first and second
point masses m1 and m2 , respectively, α1 = m m2
κπ and α2 = κπ . The proper
1
hence, the energy and Euler invariant, see Eqs. (3.1–3.2) of Deser et al. (1984)
and comments therein, read
2 √ 2 √
2κ E = −2 d x −g G0 = 0
d x −gR = − d 2 xΔ ln Φ
= 2κ d 2x mn δ 2 (r − rn ) = 2κ mn . (2.26)
n n
Therefore
Tμν = Φ−1 mn δ 2 (r − rn ) δμ0 δν0 = Φ−1 Tμν DJtH (2.27)
n
h2 ik j
Rij = 0 : − f f k − h−1 h;i;j + R̄ij = 0, (2.30c)
2
where
fij = ωj;i − ωi;j , (2.31)
and R̄ij is constructed with the spatial metric ḡ ij . In 3D spacetime, for the
spatial metric one can choose the isotropic coordinates such that
1
h;i;j − h;k ;k ḡij = 0, (2.37)
2
Introducing the complex variable z = x + iy, the equation (2.37) becomes
∂2 ∂u ∂h ∂h
2
h=2 → (z, z̄) = μ(z̄) e2 u(z,z̄) , (2.38)
∂z ∂z ∂z ∂z
where μ(z̄) is an arbitrary function. The most interesting class of solutions arises
for μ(z̄) = 0; then h is a constant, which without loss of generality can be chosen
equal to 1, h = 1. Then, the equation (2.35) implies λ0 = 0. Consequently, the
Einstein equations, (2.34) and (2.36), become
Δφ = 0 = Δ u. (2.39)
On the other hand, the external region has an embedding function Z given by
κ M/π 1 M
Z(r) = r − R0 1− 1−κ ,
1 − κ M/π κ M/π 1 − κ M/π π
r ≥ R0 , (3.3)
such that
thus the corresponding Einstein equations for dust Tλν = κ m(t, x, y) δλt δνt , are
2
∂2 ∂2 ∂
Ett :→ e−2 Ω Ω + Ω − Ω = −κ m (t, x, y) , (3.5a)
∂y 2 ∂x2 ∂t
∂2
Ext = 0 = Etx :→ Ω = 0 → Ω = F2 (t, y) + F1 (x, y) , (3.5b)
∂x∂t
∂2
Eyt = 0 = Ety :→ Ω = 0 → Ω = F3 (t, x) + F1 (x, y) , (3.5c)
∂y∂t
2
∂2 ∂ ∂ ∂
Exx = 0 = Eyy , → 2 Ω + Ω = 0, → ( Ω)−1 = 1
∂t ∂t ∂t ∂t
→Ω = ln (t f (x, y) + g(x, y)) . (3.5d)
f −2 ∇ 2 ln f = 1 − κ μ0 (x, y) . (3.9)
2
−2 Ω ∂2 ∂2 ∂ 4
R3232 = − e + Ω− Ω f (x, y) t4 = κ μ0 f 4 t2 .
∂y 2 ∂x2 ∂t
(3.11)
thus, taking into account that Ω = tf (x, y), and μ = μ0 (x, y)/t2 , one gets
⎡ ⎤
∂y μ0 (x, y) − ∂x μ0 (x, y)
∂ ∂
0
κ ⎢ ⎥
(C α β ) = 2 ⎢ − e−2 Ω ∂y
∂
μ0 (x, y) 0 0 ⎥. (3.13)
2t ⎣ ⎦
e−2 Ω ∂x
∂
μ0 (x, y) 0 0
30 Dust Solutions
d2 A
Eyy = 0 → = 0 → A(t) = A1 t + A0 , (3.20)
dt2
where Ai , Bi , i = 0, 1 are integration constants.
The area expansion Θ and the shear σ scalars are
Ȧ Ḃ
2σ(t) = − , (3.21a)
A B
Ȧ Ḃ
−Θ =
+ . (3.21b)
A B
The evaluation of the dust energy density gives
B1 A1
κ ρ (t) = , (3.22)
(B1 t + B0 ) (A1 t + A0 )
while the shear and the area expansion scalars are
A1 B0 − B1 A0
2σ = , (3.23)
(A1 t + A0 ) (B1 t + B0 )
2 A1 B1 t + A1 B0 + B1 A0
−Θ = . (3.24)
(A1 t + A0 ) (B1 t + B0 )
Moreover, from the energy conservation law T αβ ;β = 0 one establishes
dρ Ȧ Ḃ
=− + ρ = Θ ρ, (3.25)
dt A B
and for σ, using (3.21a) and (3.21b) and the fact that solutions arise for the field
equations Ä = 0 and B̈ = 0, one gets
dσ Ȧ Ḃ
=− + σ = Θ σ. (3.26)
dt A B
For Θ = 2ȧ/a, where a(t) is the “geometric–mean scale factor”, the solution for
a(t) obeys
ȧ Ȧ Ḃ α0
2 = −( + ) → a(t)2 = . (3.27)
a A B A(t)B(t)
For the A(t), and B(t) above, (3.20) and (3.19), one has
α0
a(t)2 = , (3.28)
(A1 t + A0 ) (B1 t + B0 )
and for the choice
t κρ0 t + 2σ0
A(t) = A0 , B(t) = B0 . (3.29)
t0 κρ0 t0 + 2σ0
one has
α0 t0 (κρ0 t0 + 2σ0 ) 1 1
a(t)2 = =: a20 , (3.30)
A0 B0 t (κρ0 t + 2σ0 ) t (κρ0 t + 2σ0 )
32 Dust Solutions
which is the inverse of the “geometric–mean scale factor” a(t) reported in Barrow
et al. (2006).
In the cited work is given an alternative representation of the solutions
derived above following the standard integration of the Einstein equations. Using
the energy conservation law and the 3D analogue of Raychaudhuri’s equation,
the dynamical equations in the case of a perfect fluid obeying a barotropic
state equation p = wρ, where w is the barotropic index, are given in the
form of
ρ̇ = −(1 + w)Θ ρ, (3.31a)
1
Θ̇ + Θ2 = −2κ w ρ − 2σ 2 (3.31b)
2
σ̇ + Θσ = 0, (3.31c)
ω = 0, (3.31d)
with the constraint
1 2
Θ = κρ + σ 2 . (3.32)
4
For definitions of the area expansion, shear and vorticity, correspondingly Θ, σ,
and ω; see Sec. 1.4.
To my mind, there is a discrepancy in the sign at the right-hand side of
the equation (3.31a). To establish this fact, let us review the energy density
conservation law:
uα ∇α ρ = −(ρ + p)uα ;α = −(ρ + p)Θ, (3.33)
since, for the uα aligned along ∂t
d
uα ∇α = − ,
dt
therefore
d dρ dρ
− ρ = −(ρ + p) Θ → = (ρ + p) Θ, = (1 + w)ρ Θ. (3.34)
dt dt dt
The solutions of this last equation for Θ = 2ȧ/a, and of (3.31c), are
2(1+w) 2
a(t) a(t)
ρ(t) = ρ0 , σ = σ0 (3.35)
a0 a0
which certainly are inverse to the ones of Barrow et al. (2006): there, integrating
(3.31a) and (3.31c) for Θ = 2ȧ/a, one arrives at
a0 2(1+w) a0 2
energy density : ρ = ρ0 , and shear: σ = σ0 .
a a
For the specific problem under consideration, the sign of Θ in Eq. (38) of Barrow
et al. (2006) is opposite to the correct one calculated here for (3.21b).
3.4 BST Diagonal Anisotropic Dust Solutions with Λ 33
Cotton Tensor
The Cotton tensor is
α 1 B1 A1 (B1 A0 − A1 B0 ) δαx δy β δαy δxβ
C β = + . (3.36)
2 (B1 t + B0 ) (A1 t + A0 ) A(t)2 B(t)2
Searching for its eigenvectors, one establishes that they are
λ1 = 0, T α = δ α t , T α Tα = −1, (3.37a)
1 B1 A1 (B1 A0 − A1 B0 ) δ α
x δ α
y
λ2 = , S1α = + , S1α S1α = 2, (3.37b)
2 A(t)2 B(t)2 A(t) B(t)
1 B1 A1 (B1 A0 − A1 B0 ) δα x δαy
λ3 = − , S2α
= − + , S2α S2α = 2, (3.37c)
2 A(t)2 B(t)2 A(t) B(t)
where T and S are used to denote respectively timelike and spacelike vectors,
therefore, this Cotton tensor is of Type I: {T, S, S}.
√
Λ (B2 A1 − B1 A2 )
σ(t) = √ √ √ √ , (3.41b)
B1 e Λt + B2 e− Λt A1 e Λt + A2 e− Λt
√ √ √
Λ −A1 B1 e2 Λt + A2 B2 e−2 Λt
Θ=2 √ √ √ √ . (3.41c)
B1 e Λt + B2 e− Λt A1 e Λt + A2 e− Λt
34 Dust Solutions
3D Gödel Universes
One can bring the derived solutions a Gödel-like form by introducing the
trigonometric functions, namely
A(t, Λ < 0) = a1 sin |Λ|t + a2 cos |Λ|t,
B(t, Λ < 0) = b1 sin |Λ|t + b2 cos |Λ|t, (3.42a)
√ √
A(t, Λ > 0) = a1 sinh Λt + a2 cosh Λt,
√ √
B(t, Λ > 0) = b1 sinh Λt + b2 cosh Λt. (3.42b)
sin |Λ|t and cos |Λ|t functions for negative Λ = −|Λ| = −1/l2 . When
replacing the exponential functions in terms of the hyperbolic, or trigonometric
functions depending on the sign of the Λ, the coefficient functions in front of the
sines and cosines are linear combinations of the previous coefficients multiplying
the exponentials (in the negative Λ these coefficients are complex conjugated);
for them, Greek symbols will be used, and one has these possibilities: for positive
Λ: Λ > 0, one gets
√ √
A(t, x, y) = α1 (x, y) sinh Λt + α2 (x, y) cosh Λt,
√ √
B(t, x, y) = β1 (x, y) sinh Λt + β2 (x, y) cosh Λt, (3.47a)
∂ ∂
Ext : −A2 (x, y) B1 (x, y) + A1 (x, y) B2 (x, y) = 0, (3.49)
∂x ∂x
∂ ∂
Eyt : −B2 (x, y) A1 (x, y) + B1 (x, y) A2 (x, y) = 0. (3.50)
∂y ∂y
Since there are only two constraints on four unknown structural functions, then
one may consider two of them as arbitrary; let them be A1 and A2. The most
general solutions of these equations, B1 (x, y) and B2 (x, y), depending on both
spatial variables x and y, are
∂A2
∂y
B2 (x, y) = B1 (x, y) ∂A1
, (3.51)
∂y
A1 ∂ 2 A2 ∂A1
− ∂A2 ∂ 2 A1
∂y∂x ∂y ∂y ∂y∂x
B1 (x, y) = b (y) exp dx. (3.52)
∂A1
∂y A2 ∂A1
∂y − A1 ∂A2
∂y
∂ ∂
F1 (x, y) G2 (x, y) − F2 (x, y) G1 (x, y) = 0, (3.56b)
∂x ∂x
thus two functions should stay arbitrary; let them be F1 (x, y) and F2 (x, y).
From the (3.56a) equation one gets
∂
∂y F1 (x, y)
G1 (x, y) = ∂
G2 (x, y) (3.57)
∂y F2 (x, y)
which, substituted into the (3.56b) and integrating for G2, yields
F2 ∂2
F2 ∂
F1 − ∂2
F1 ∂
F2
∂y∂x ∂y ∂y∂x ∂y
G2 (x, y) = g2 (y) exp dx
∂
∂y F2 F2 ∂
∂y F1 − F1 ∂
∂y F2
(3.58)
with arbitrary functions F1 and F2 . Notice that the integration function g2 (y)
can be equated to 1 by accomplishing a coordinate transformation in y.
where F (x, y) and ν(x, y) are arbitrary functions ∂F∂y = 0, and S(y) is an
integration function, which can be set equal to unity by a redefinition of the
y-coordinate. The evaluation of the energy density can done by means of the
expression (3.45). This general dust solution corresponds to the one given in
paragraph 6.2 Class 2 in Barrow et al. (2006), through
[∂y (R(x, y, t)eν(x,y) )]2 −2α(x,y) 2
ds2 = −dt2 + e−2ν(x,y) e dy
S 2 (y)(∂y F (x, y))2
+R2 (x, y, t)e2ν(x,y) dx2 ,
38 Dust Solutions
Szekeres Solutions
Szekeres solutions arise for ∂x F = 0, and
κρ + Λ = √ √ √ √ , (3.67)
e Λt B1 + B2e− Λt e2 Λt a1 + a2e− Λt
where the dependence on x, and y has been omitted. Notice that the definitions
of the functions L(x, y) differ; L(3.65) = ∂x L(3.68) .
Since any 2D space is conformally flat, one may write the spatial sector of this
metric as
√ √
e2 φ(x,y) (α0 e Λt
+ e− Λt 2
) (dx2 + dy 2 ). (3.70)
The conservation of the energy momentum tensor Tμν = (μ + p)uμ uν + p gμν for
a perfect fluid (dust) moving along the time direction ∂t ,
uμ = −δ μ t , uμ = δμ t + Zδμ φ , (3.80)
dp
requires the constancy of the pressure, dr = 0, p(r) = p0 .
The Einstein equation
d2 Z dZ dY
Gφ t = 0 : Y − = 0, (3.81)
dr2 dr dr
integrates as
d
Y (r) = C1 Z (r) , (3.82)
dr
while Gr r = 0 = (Gφ φ ) gives the relation
1
κ p0 = Λ + = constant, (3.83)
4C12
and the equation Gt t = 0 = (Gt φ ) becomes
d3 Z 3 dZ
3
+ κ μ (r) + Λ − 2 = 0. (3.84)
dr 4C1 dr
Although the obtained solution possesses a constant pressure throughout the
entire spacetime, it is endowed with matter modeled through the energy density
subject to the equation (3.84). As pointed out in Rooman and Spindel (1998),
this equation can be thought of as defining the matter content for a given function
Z(r), or, conversely, for a given plausible matter distribution one determines the
structural function Z(r).
The evaluation of the fluid covariant derivative properties yields zero, i.e., the
area expansion Θ, shear σ, rotation ω and acceleration vector all vanish.
3.6 Rooman–Spindel Dust Gödel Non-Diagonal Model 41
As far as the algebraic type of the Cotton tensor is concerned, it is clearer and
more straightforward to establish it from the alternative representation of this
metric in term of the the structural function Y (r). Instead of (3.82), one now
uses
d
Z (r) = F0 Y (r) . (3.85)
dr
Thus, there remains a single constraint
d2 3 2
Y (r) + κ m (r) + Λ − F0 Y (r) = 0. (3.86)
dr2 4
The Cotton tensor in terms of Y (r) and M (r), 4 κ m (r) + F0 2 + 4 Λ = 8 M (r),
becomes
⎡ ⎤
−2 F0 Y M − Z dM
dr 0 −3 ZF0 Y M − Y 2 + Z 2 dM dr
⎢ ⎥
Y (C α β ) = ⎢
⎣ 0 F0 Y M 0 ⎥,
⎦
dM
dr 0 Z dM
dr + F0 Y M
(3.87)
from which it is evident that the roots for the eigenvalues could be real or complex
conjugated, fulfilling the traceless condition λ1 + λ2 + λ3 = 0, therefore the
algebraic types could be Type I, and Type IZ .
In the next paragraph, for the case of incompressible dust m(r) = m0 , the
eigenvectors of the Cotton tensor are shown explicitly.
Rooman–Spindel Dust
If the pressure is set equal to zero, p = 0, thus C21 = − 4Λ
1
, then one is dealing
with an AdS, Λ < 0, dust star, with metric
2
dZ
ds = −[dt + Z(r)dφ] + dr + C1
2 2 2
dφ2 , (3.88)
dr
and energy density subjected to
d3 Z dZ
3
+ (κ μ (r) + 4Λ) = 0. (3.89)
dr dr
or
Following Lubo et al. (1999), see Eq. (95), these solutions can be given as
r r
ds2 = −dt2 + dr2 + γ 2 sinh2 ( ) − a2 c2 (κ + cosh( ))2 dφ2
a a
r
−2acγ(κ + cosh( ))dt dφ, (3.93a)
a
2r r
ds2 = −dt2 + dr2 + γ 2 e( a ) − a2 c2 (κ + e( a ) )2 dφ2
r
−2acγ(κ + e( a ) )dt dφ, (3.93b)
r r
ds2 = −dt2 + dr2 + γ 2 cosh2 ( ) − a2 c2 (κ + sinh( ))2 dφ2
a a
r
−2acγ(κ + sinh( ))dt dφ, (3.93c)
a
Locally, these metrics are mutually equivalent; for coordinate transformations
relating to them, see Rooman and Spindel (1998); for low density stars, κμ <
−4Λ, they are determined by restricting the domain of r to r < rboundary , hence
they correspond to different non-diffeomorphic subsets of a larger space on which
r is unrestricted.
On the other hand, the metric of the high density star κμ > −4Λ is given by
r r
ds2 = −dt2 + dr2 + γ 2 sin2 ( ) − a2 c2 (κ + cos( ))2 dφ2
a a
r
−2acγ(κ + cosh( ))dt dφ. (3.94)
a
For the physical interpretation of the above solutions one has to return to the
original Gödel works.
The Cotton tensor (3.87), for this incompressible case
amounts to
⎡ ⎤
−C0 0 − 32 Z C0
⎢ ⎥
(C α β ) = ⎢
⎣ 0 1
2 C0 0 ⎥,
⎦
1
0 0 2 C0
3.6 Rooman–Spindel Dust Gödel Non-Diagonal Model 43
thus, the eigenvalues are: a single λ1 = −C0 and a double one λ2 = C0 /2 with
eigenvectors
λ1 = −C0 , V 1α = δ α t , V 1α V 1α = −1,
C0
λ2 = , V 2α = −ZV3 δ α t + V2 δ α x + V3 δ α y , V α Vα = V2 2 + V3 2 Y 2 ,
2
C0
λ3 = , V 3α = −Zv3 δ α t + v2 δ α x + v3 δ α y , V α Vα = v2 2 + v3 2 Y 2 ,
2
(3.95)
in particular, for the components
V3 2 Y 2
v1 = V1 , v2 = − , v3 = V3 ,
V2
the vectors V 2 and V 3 are orthogonal. Consequently, the Cotton tensor for these
Gödel kinds of solutions is of Type I: {T, S, S}.
4
A Shortcut to (2+1) Cyclic Symmetric
Stationary Solutions
F d2 H
Et t − Er r − W Eφ t = = 0 → H(r) = H1 r + H0 . (4.4)
2H dr2
The integration of (4.2), assuming H1 = 0, gives
J 1 J 1
W (r) = − + W0 = − + W0 . (4.5)
H1 H(r) H1 H1 r + H 0
4.1 Stationary Solutions in Canonical Coordinates 45
Without loss of generality one can set W0 = 0. Moreover, the substitution of the
above structural functions into the remaining Einstein equations gives rise to a
condition on the integration constants, which can be solved, for instance, for F0
H0 J2 4 H2
F0 = F1 + 2 − 2 02 (4.6)
H1 H1 l H1
Therefore the metric tensor components become
1 4 H0
gtt = − ( 2 r + F1 − 4 2 ),
H1 l l H1
4 J2 H0 H0
1/grr = 2 r 2 + F1 r + 2 + (F1 − 4 2 ),
l H1 H1 l H1
J
gt φ =− , gφ φ = H1 r + H0 . (4.7)
H1
energy density
f (ρ) 1 l M
(ρ, M, 0 ) = − − 0 , (ρ, M, 0) ≈ − + ,
πρ πl 2π ρ2
l M − M0
(ρ, M, 0 (M0 )) ≈ ,
2π ρ2
ρ2 / l2 − M0 1 l M0
0 (M0 ) := − , 0 (M0 ) ≈ − + , (4.10b)
πρ π l 2π ρ2
total quasi-local energy
It becomes apparent that this metric form is just another real cut of the met-
ric (4.9) when subjecting it to the complex transformations t → i φ and φ → i t.
This representation of the BTZ metric counterpart is useful when one considers,
for example, the vacuum limit of the metrics describing magnetic fields.
Moreover, by accomplishing in the above metric the transformation of the
radial coordinate
ρ 1
E(ρ, M, 0) ≈ −2 2 , E(ρ, M, 0c (M0 )) ≈ (M − M0 ),
l ρ
2 2 ρ2
M (ρ, M, 0 ) = − 2 ρ − 2π ρ 0 f (ρ), M (ρ, M, 0) = −2 2 ,
l l
M (ρ, M, 0c (M0 )) ≈ M − M0 ,
ρ
0c (M0 ) := − ,
πl 2 ρ / l 2 + M0
2
1 l M0
0c (M0 ) ≈ − + ≈ 0 (M0 ). (4.14)
π l 2π ρ2
This is one of the plausible choices for the base energy density of the
anti-de Sitter spacetime; see Brown et al. (1994) in the paragraph below
Eq. (4.12). Another possibility takes place for the naked singularity solution with
M0 = 0,
dρ2 ρ2
g = −f (ρ, 0)dt2 + + ρ2 dφ2 , f (ρ, 0) = 2 , (4.27)
f (ρ, 0) l
with energy characteristics
1 1
(ρ, 0) := − , (ρ, 0, 0 ) = − − 0 ,
πl πl
ρ
E(ρ, 0, 0 ) = 2π ρ (ρ, 0, 0 ), M (ρ, 0, 0 ) = E(ρ, 0, 0 ). (4.28)
l
For the particular referential energy density 0 = − π1l , all energy characteristics
of this naked anti-de Sitter solution vanish,
1 1 1
(ρ, 0, 0 = − ) = 0, E(ρ, 0, 0 = − ) = 0, M (ρ, 0, 0 = − ) = 0,
πl πl πl
in agreement with Brown et al. (1994), Eq. (4.12).
These results can be gathered in a more compact form by introducing the base
energy density equipped with the discrete parameter m = −1, and 0,
1 ρ2 1 lm
0 (ρ, m) =:= − − m, 0 (ρ, m) ≈ − + , (4.31)
πρ l2 πl 2π ρ2
4.2 Static AdS Black Hole 51
as
l
(ρ, M0 , 0 (ρ, m)) ≈ (M0 − m),
2πρ2
l
E(ρ, M0 , 0 (ρ, m)) ≈ (M0 − m),
ρ
M (ρ, M0 , 0 (ρ, m)) ≈ M0 − m. (4.32)
∂ ∂
2EQ13 = v1 + v3 = 0, (4.39c)
∂φ ∂t
∂ d F1 (t, φ)
EQ22 = F (r) v2 + v2 F (r) → v2 = = 0, (4.39d)
∂r dr 2lF (r)
∂ ∂ J (M l2 − r2 )
EQ23 = v2 + v3 + 2 v1 + 2 2 v3 = 0, (4.39e)
∂φ ∂r rF (r) rl2 F (r)
4.3 Symmetries of the BTZ Metrics 53
∂ 2
EQ33 = v3 + r F (r) v2 = 0, (4.39f)
∂φ
where Fi (t, φ) , i = 1, 2, 3, are integration functions.
Isolating v1 from (4.39e) in terms of v2 and v3 and their derivatives, one gets
∂ ∂
rl2 Jv1 = 2 r r2 − M l2 v3 − r2 l2 F (r)2 v2 + v3 . (4.40)
∂φ ∂r
Next, substituting v1 from above and the first integral of v2 from (4.39d) into
equation (4.39b), one arrives at a linear second-order equation for v3 with
integrals
r2 1 r l F (r) ∂ ∂
v3 = F2 (t, φ) + F3 (t, φ) − J +M F1 (t, φ)
2 2 M 2 l2 − J 2 ∂t ∂φ
(4.41)
which, substituted together with v2 from (4.39d) into Eq. (4.40) for v1, gives
2
r − M l2 1
v1 = 2 F3 (t, φ) − J F2 (t, φ)
Jl2 4
1 r l F (r) J ∂ ∂
+ +M F1 (t, φ) . (4.42)
2 (l2 M 2 − J 2 ) l2 ∂φ ∂t
The dependence of the Killing vector components on the r variable has been
established; it remains still to determine their dependence on the t and φ vari-
ables hiding in the F1 (t, φ) , F2 (t, φ) and F3 (t, φ) functions. Substituting the
expressions of v1 from (4.42), v3 from (4.41), and v2 from (4.39d), one arrives
at the independent equations
2
∂2 ∂
F1 J 2 + l2 J F1 + l 2 1 M − F1 M l = 0,
2 2
F (4.43a)
∂t∂φ ∂φ2
∂2 ∂2
2
F1 J + l J 2
F1 + M F1 l4 − F1 M 2 l2 = 0, (4.43b)
∂t∂φ ∂t2
with integral
√ √ √ √
M l−J(lφ+t) M l+J(lφ−t) M l+J(lφ−t) M l−J(lφ+t)
− −
F1 = C1 e l3/2 + C2 e l3/2 + C3 e l3/2 + C4 e l3/2 .
(4.44)
Furthermore, there have to be solved constraints on F2 (t, φ) and F3 (t, φ), namely
∂ ∂
F3 = 0, F3 = 0, F3 (t, φ) = J C6 = const.,
∂t ∂φ
∂ ∂
F2 = 0, F2 = 0, F2 (t, φ) = C5 = const. (4.45)
∂t ∂φ
54 A Shortcut to (2+1) Cyclic Symmetric Stationary Solutions
0 ≤ ρ ≤ ∞, 0 ≤ φ ≤ 2π,
4.3 Symmetries of the BTZ Metrics 57
and α is set equal to unity, α = 1 = −M , then in such case ρ and φ become polar
coordinates with φ being the angular coordinate with period 2π. This spacetime
– the (proper) anti-de Sitter space (with M = −1) – allows for six symmetries,
and as such it is maximally symmetric.
In this chapter, in the framework of the (2+1)-dimensional Einstein theory
with cosmological constant different families of exact solutions for cyclic sym-
metric stationary (static) metrics in the presence of a negative cosmological
constant have been derived. Specific branches of solutions in the general case are
determined via a straightforward integration. In this systematic approach, all
known cyclic symmetric solutions of the considered class are properly identified.
5
Perfect Fluid Static Stars; Cosmological Solutions
The purpose of this chapter is to determine the static circularly (cyclic) symmet-
ric spacetimes coupled to perfect fluids via a straightforward integration of the
Einstein equations. The structural functions of the metric depend on the energy
density, which remains in general arbitrary. Spacetimes for fluids fulfilling linear
(barotropic) and polytropic state equations are explicitly derived. By the way,
we demonstrate here that the incompressible perfect fluid solution is the only
conformally flat (with vanishing of the Cotton tensor) circularly symmetric solu-
tion. Since there is no boundary to determine a zero-pressure surface, all these
solutions fall in the cosmological category. The perfect fluid is characterized by
a fluid velocity uμ and energy–momentum tensor
Tμν = (p + μ)uμ uν + pgμν , uμ uμ = −1, (5.1)
where μ and p are the energy density and the pressure; throughout this text,
the energy density will be denoted with μ or ρ, as are commonly used in
hydrodynamics.
e−2 λ(r) = C0 − r2 κ μ0 ,
d κ p (r) r (μ0 + p (r)) B0 μ0 C0 − r2 κ μ0
p (r) = − → p(r) = ,
dr C0 − r κ μ0
2
1 − B0 C0 − r2 κ μ0
ν (r) = ν0 + ln 1 − B0 C0 − r2 κ μ0 , (5.8)
where C0 and B0 are integration constants which one fixes via boundary
conditions.
from which one determines C0 , which in the particular case of constant density
μ0 results in C0 = κμ0 r02 = 8πμ0 r02 . The pressure is expressible as
[1 + 8πμ0 (r02 − r2 )]1/2
p(r) = Bμ0 , (5.10)
[1 + 8πμ0 r02 ]1/2 − B[1 + 8πμ0 (r02 − r2 )]1/2
comparable with Eq. (40) of Collas (1977). Thus the requirement of vanishing
pressure at the boundary of the fluid ball fails to be fitted, hence the author
concludes that it is not possible to have a fluid with constant density μ and
nonzero pressure p = 0.
This dilemma found a solution in the forthcoming formulation for the constant
density μ0 and variable bounded pressure p(r) branch of metrics, but not allowing
for a match to the exterior Minkowski spacetime.
consequently
√
R2 − r 2
p(r) = pc μ0 √
R(pc + μ0 ) − pc R2 − r2
2
1 − (r/R)
= pc , (5.12a)
2
1 + μpc0 − μpc0 1 − (r/R)
For the family of polytropic fluids, the pressure and energy density can be
given as
p(r) = p0 Θ(r)n+1 , μ(r) = μ0 Θ(r)n , (5.18)
where n is the polytropic index, not necessarily an integer. Moreover, one may
impose one condition on the pressure and energy at the center of the fluid ball,
namely
p(r = 0) = p0 Θ(0)n+1 , μ(r = 0) = μ0 , Θ(r = 0) = 1. (5.19)
k (n + 1)! μ0 k 1
Skn (Θ(r)) = (−1) .
k! (n + 1 − k)! (n + 1 + k) (p0 Θ + μ0 )n+1+k
(5.22)
Hence, evaluating A0 at r = 0, Θ(r = 0) = 1, one arrives at
B0 2 1
n+1
− r = n+1 n+2 [Skn (Θ(r)) − Skn (Θ = 1)] . (5.23)
2 μ0 p0
k=0
At this stage, let us see if one can fix the value of the integration constant B0 .
The evaluation of
κ π rκ p (μ + p) κr dr
e−2λ(r) = C0 − M (r) = − dp
=− (μ0 + p0 Θ) Θn+1
π dr
n + 1 dΘ
1 κ Θ2n+2 1
=− = ,
B0 n + 1 (μ0 + p0 Θ)2n+2 grr
1 n+1 1
gΘΘ =− , (5.24)
B0 κr2 (μ0 + p0 Θ)2n+2
therefore, the positiveness of the spatial metric components requires B0 to be
negative, B0 < 0, say B0 = −2. The integration of ν(r), (5.3b) is straightforward,
−2 n−2
exp (2ν) = N02 (μ0 + p0 Θ) .
Changing Θ to a more conventional radial coordinate ρ ≡ Θ, the CF solution
in the (t, ρ, φ) coordinates can be given as
−2 n−2 n+1 dρ2
ds2 = − (μ0 + p0 ρ) dt2 + + r2 dφ2 , (5.25)
2κ r2 (μ0 + p0 ρ)2n+2
1
n+1
r(ρ)2 := [Skn (ρ) − Skn (ρ = 1)] , (5.26)
μ0 n+1 p0 n+2
k=0
k (n + 1)! μ0 k 1
Skn (ρ) = (−1) , (5.27)
k! (n + 1 − k)! (n + 1 + k) (μ0 + p0 ρ)n+1+k
p(ρ) = p0 ρn+1 , μ(ρ) = μ0 ρn . (5.28)
Since this solution does not allow for a zero-pressure circle p(ρ) = 0 except at
ρ = 0, this kind of solution is a cosmology that extends over the whole space.
64 Perfect Fluid Static Stars; Cosmological Solutions
As far as to the mass is concerned, let us return to the equations (5.24), with
B0 = −2,
κ 1 κ Θ2n+2 1
C0 − M (r) = = . (5.29)
π 2 n + 1 (μ0 + p0 Θ)2n+2 grr
From this equation one concludes that at the origin r = 0, and at the boundary
r = R one correspondingly has
κ 1 κ 1
C0 = M (r = 0) + , (5.30)
π 2 n + 1 (μ0 + p0 )2n+2
2n+2
κ 1 κ Θ(R) 1
C0 − M (r = R) = = . (5.31)
π 2 n + 1 (μ0 + p0 Θ(R))2n+2 grr (R)
From the second condition, it becomes clear that the choice Θ(R) = 0, made in
Cornish and Frankel (1991), is not an adequate one since in that case g rr (R) = 0,
grr (R) → ∞–a black fluid ball-plate. Moreover, as was been mentioned above,
the condition Θ(R) = 0 impose a dependence of the form R(μ0 , p0 ), which is also
unwelcome. Consequently, it is better to adopt as the mass within the radius r
the quantity
2 n+2
π Θ(r) π
M (Θ) = M (0) + 2 n+2 − 2 n+2 .
2 (μ0 + p0 ) (n + 1) 2 (μ0 + p0 Θ(r)) (n + 1)
In particular, the mass inside the ball with r = R, is given by
2n+2
π 1 1 π 1 Θ(R)
M (r = R) = M (0) + 2n+2 − .
2 n + 1 (μ0 + p0 ) 2 n + 1 (μ0 + p0 Θ(R))2n+2
On the other hand, at this boundary r = R, the metric components are finite if
Θ(R) = 0, and μ0 + p0 Θ(R) = 0; see (5.31). Recall that Θ(R) is a root of the
algebraic polynomial equation of degree n + k + 1 for Θ(R) arising from
1
n+1
R2 = n+1 n+2
[Skn (Θ(R)) − Skn (Θ(0) = 1)] .
μ0 p0
k=0
This is a complete analytic solution, for any polytropic index n and has no
parallel comparing with (3 + 1)-dimensional case, where only few polytropic
solutions are known.
The purpose of this chapter is to determine all static circularly (cyclic) symmet-
ric spacetimes with a cosmological constant coupled to perfect fluids with and
without zero-pressure surfaces via a straightforward integration of the Einstein
equations. The structural functions of the metrics depend on the energy den-
sity, which remains in general arbitrary. Spacetimes for fluids fulfilling linear
(barotropic) and polytropic state equations are explicitly derived; they con-
tain, among other families, stiff matter, incoherent radiation, and non-relativistic
degenerate fermions. By the way, we demonstrate here that the incompressible
perfect fluid solution is the only conformally flat – in the sense of the vanishing
of the Cotton tensor – circularly symmetric solution.
In Section 6.1, the Einstein equations for the static circularly symmetric (2+1)
metric with a cosmological constant coupled to a perfect fluid solution with vari-
able density ρ and pressure p are exhibited and integrated; in this chapter we
are using ρ to denote the energy density instead of the μ used in the previous
chapters. Section 6.2 is devoted to representing this whole class of spacetimes in
a canonical coordinate system. For a given equation of state of the form p = p(ρ),
certain particular families of perfect fluid solutions are derived; as concrete exam-
ples, the subcases of fluids obeying the barotropic law p = γ ρ in Section 6.3, and
those fluids subjected to a polytropic law p = C ργ in Section 6.4, are derived.
Moreover, the incompressible fluid, ρ = const., is the only conformally flat static
circularly symmetric solution coupled to perfect fluids. In Section 6.5, from the
Oppenheimer–Volkoff equation certain properties of the studied solutions are
established: for positive pressure p and positive density ρ, a microscopically sta-
ble fluid possesses a monotonically decreasing energy density; and conversely,
these results are in close relation with the ones reported by Cruz and Zanelli
(1995), which unfortunately contains various misprints.
In Section 6.6, to facilitate the comparison of the interior Schwarzschild (3+1)
solution with cosmological constant, the incompressible perfect fluid (2 + 1) solu-
tion is derived. With this aim in mind, we search for an adequate representation
6.1 Equations for a (2+1) Static Perfect Fluid Metric 67
The metric (6.1), with G(r) from (6.8), and N (r) from (6.10), determines the
general static circularly symmetric (2+1) solution of the Einstein equations (6.3)
with Λ for a perfect fluid, characterized by a pressure given by (6.11), and an
arbitrary density ρ(r). The fluid-velocity is aligned along the timelike Killing
direction ∂t . To deal with realistic matter distributions one has to impose pos-
itivity conditions on the density, ρ > 0, and the pressure, p > 0, requiring
additionally ρ > p.
For a finite distributed fluid, the pressure p becomes zero at the boundary,
say r = a; this value of the radial coordinate r is determined as solution of the
equation p(r) = 0.
For non-vanishing cosmological constant, assuming that the values of the struc-
tural functions at the boundary r = a are N (a) and G(a), the vanishing at r = a
of the pressure p(r), given by (6.11), requires n1 = −ΛN (a)/G(a), hence
Λ
κ p(r) = (N (r) G(a) − N (a) G(r)) . (6.12)
N (r)G(a)
If one is interested in matching the obtained perfect fluid metric with a vacuum
metric with cosmological constant Λ, the plausible choice at hand is the anti-de
Sitter metric, with Λ = −1/l2 , for which G(a) = N (a) = −M∞ + a2 /l2 at the
boundary r = a. Incidentally, for a cosmological constant different from zero,
there is no room for dust. The zero character of the pressure would yield the
vanishing of the density, and consequently the metric reduces to the (anti-)de
Sitter spacetime.
For vanishing cosmological constant, the expression of the pressure (6.11) is
G(r)
κp(r) = n1 , (6.13)
N (r)
from which it becomes apparent that the corresponding solution represents a
cosmological spacetime; there is no a surface of vanishing pressure.
For vanishing Λ and zero pressure, the situation slightly changes: the function
N becomes a constant, and the corresponding metric can be written as
dr2
ds2 = −dt2 + r + r2 dθ2 , (6.14)
C − 2κ rρ(r)dr
for any density function ρ. Of course, the choice of ρ is restricted by physically
reasonable matter distributions.
dN dr
= , (6.15)
r G(r)
hence
N
2
r = C0 + 2 G dN . (6.16)
To derive G as a function of the new variable N , one uses the Gtt –(6.3) in the
form of
G dG = −(κ ρ + Λ)r dr = −(κ ρ + Λ)GdN, (6.17)
dN 2
ds2 = −N 2 dt2 + + H(N ) dθ2 , (6.20)
H(N )
which is characterized by pressure
N
C1 1 1
p(N ) = − ρ(N ) dN (6.21)
κ N N
and an arbitrary energy density ρ(N ) depending on the variable N , both
functions p and ρ have to be positive.
The metric (6.20) together with the function H from (6.19) give an alternative
representation of our general solution. This representation will be used to derive
particular solutions for a given state equation of the form p = p(ρ). In this
approach the expression of the pressure (6.21) play a central role.
The barotropic law establishes a linear relation between pressure and energy
density of the form
p(N ) = γ ρ(N ). (6.22)
Substituting p(N ) from Eq. (6.23) into this relation, one gets
N
C1
− ρ(N ) dN = γ N ρ(N ). (6.23)
κ
Differentiating this equation with respect to the variable N , one obtains
d 1
(N ρ) + (N ρ) = 0, (6.24)
dN γN
which has as general integral
γ − 1 − γ+1
ρ(N ) = C2 N γ , (6.25)
γ2
where C2 is an integration constant. Since we arrived at the above simple linear
equation (6.24) through differentiation, then one has to replace the obtained
result into the relation (6.23), or equivalently into (6.22), to see if there arises
any constraint from it:
C1 1
p(N ) = + γ ρ(N ) = γ ρ(N ) −→ C1 = 0, (6.26)
κ N
in such manner, we establish that the constant C1 vanishes. Replacing the func-
tion ρ(N ) from (6.25) into the expression of H(N ), (6.19), and accomplishing
the integration one arrives at
Thus, the metric for a perfect fluid fulfilling a barotropic state equation in
coordinates {t, N, θ} is given by
dN 2 γ−1
ds2 = −N 2 dt2 + γ−1 +(C0 −ΛN 2 +2C2 κN γ )dθ2 . (6.28)
C0 − ΛN 2 + 2κC2 N γ
To give this solution in terms of the radial variable r, one has to be able to solve
the algebraic equation, in general a transcendent one, for N = N (r). Within this
class of solutions merits mention the pure radiation γ = 1/2, and the stiff matter
γ = 1 where the speed of sound equals the speed of light.
Using again the expression of p(N ) from (6.21), the above polytropic relation
can be written as
N
C1
− ρ(N ) dN = C N ργ (N ). (6.30)
κ
Differentiating with respect to N , one obtains
d
− ρ=C (N ργ ) (6.31)
dN
which, by introducing the auxiliary function Z := N 1/γ ρ, can be written as
1 1
d(Z γ−1 ) + d(N (γ−1)/γ ) = 0 → d (ργ−1 + ) N (γ−1)/γ = 0, (6.32)
C C
therefore, integrating
1
ργ−1 + N (γ−1)/γ = B C −(γ−1)/γ , (6.33)
C
where B is an integration constant. Consequently, the general integral of this
equation can be given as
−1 −1
−1 γ−1
γ−1
1
Entering this ρ into the equation (6.30), taking into account that the integral of
the density ρ amounts to
N N γ−1
γ
−1 γ−1
ρ(N ) dN = − d B−C γ N γ , (6.35)
one arrives at
n1 C1
p(N ) = + C ργ = C ργ −→ C1 = 0. (6.36)
κ N
Considering that the first integral of ρ is given by (6.35), the expression of the
structural function H(N ) becomes
N γ−1
γ
−1 γ−1
H(N ) = r2 = C0 − Λ N 2 + 2 κ B−C γ N γ dN. (6.37)
dN 2
ds2 = −N 2 dt2 +
N −1 γ−1
γ−1
γ
C0 − Λ N 2 + 2κ B−C γ N γ dN
N
−1 γ−1
γ−1
γ
and since the metric components grr = 1/G(r)2 has to be positive in the domain
of definition of the solution, then there exits an upper limit for the mass, namely
π
M ≤ (C − Λ a2 ). (6.45)
κ
Matter is said to be microscopically stable if dp/dρ ≥ 0, which is equivalent
to the statement that the speed of sound is less than the velocity of light. Since
(6.41) can be written as
dp r dρ
= − 2 (κ p − Λ)(ρ + p)/ , (6.46)
dρ G dr
one concludes that for a microscopically stable fluid with positive pressure p and
positive density ρ, this density is monotonically decreasing dρ/dr < 0.
For our general solution in coordinates {t, N, θ}, metric (6.20), from the
expression (6.21) for the pressure, one establishes
dp 1 dρ
= − (ρ + p)/ , (6.47)
dρ N dN
therefore the density is monotonically decreasing dρ/dN < 0 if the matter is
microscopically stable dp/dρ ≥ 0, and conversely.
Moreover, our fluids, fulfilling the barotropic state equation p = γρ, γ > 0,
as well as those ones obeying the polytropic law p = Cργ , C > 0, γ > 0, are
microscopically stable fluids.
1
κp(r) = [n1 κρ G(r) + n0 Λ(κρ + Λ)] . (6.53)
(κρ + Λ)N (r)
N (a)
N (r) = [κρ G(a) + Λ G(r)] . (6.55)
G(a)(κρ + Λ)
G(a) − G(r)
p(r) = ρ Λ . (6.56)
κρG(a) + ΛG(r)
dr2
ds2 = −N (r)2 dt2 + + r2 dθ2 (6.57)
G(r)2
with structural functions G(r) from (6.51) and N (r) from (6.55) and character-
ized by a density ρ = const. and pressure p given by (6.56).
76 Static Perfect Fluid Stars with Λ
2 dr2
ds2 = −N (r) dt2 + 2
+ r2 dθ2 + sin2 θdφ2 . (6.58)
G(r)
The Einstein equations with cosmological constant for perfect fluids for the
metric (6.58) explicitly amount to
N2 dG2
Gtt = 2 r + G2 − 1 = −N 2 (κρ + Λ),
r dr
1 dN 1
Grr = − 2 2 2rG2 − N + N G2 = − 2 (κp − Λ),
G Nr dr G
2 2
r dN 1 dG d N r dN dG2
Gθθ = − G2 + N + rG2 2 + = −r2 (κp − Λ),
N dr 2 dr dr 2 dr dr
Gφφ = sin2 θ Gθθ . (6.59)
In what follows we shall omit the dependence on r of the functions we are dealing
with, except when needed for clarity. The Einstein equations for a perfect fluid
energy–momentum tensor in four dimensions have the same form as the ones
in three dimensions, except for the modifications due to the change of dimen-
sionality; for instance, the expressions of the scalar curvatures R are different.
Because of the corresponding equations can be found in textbooks, we do not
exhibit them here explicitly.
Since we are interested in conformally flat solutions, we require the vanishing of
the conformal Weyl tensor, which for static spherically symmetric perfect fluids
fulfills the following equation
d G2 − 1
= 0 −→ G(r) = 1 + c0 r2 . (6.60)
dr r2
(3 + 1) solution (2 + 1) solution
dr 2 dr 2
ds2 = −N 2 dt2 + G2
+ r 2 dΩ2 ds2 = −N 2 dt2 + G2
+ r 2 dφ2
G2 = 1 − 13 (κρ + Λ)r 2 G2 = C − (κρ + Λ)r 2
1 N (a) 1 N (a)
N = 2(κρ+Λ) G(a)
[3κρG(a) N = (κρ+Λ) G(a)
[κρG(a) + ΛG(r)]
+(2Λ − κρ)G(r)] +ΛG(r)]
(2Λ−κρ)(G(a)−G(r)) G(a)−G(r)
p = ρ 3κρG(a)+(2Λ−κρ)G(r) p = ρΛ κρG(a)+ΛG(r)
coordinate system {t, r, θ}, and alternatively, in a system – the canonical one –
with coordinates {t, N, θ}. From the physical point of view, particularly interest-
ing are those fluids fulfilling the linear (barotropic) equation of state, p = γρ, as
well as those subjected to the polytropic law p = ργ ; both families are derived
in details from our general metric referred to the coordinate system {t, N, θ}.
Therefore, the derived solutions describe, among other things, stiff matter, pure
radiation, incoherent radiation, nonrelativistic degenerate fermions, etc. The
incompressible perfect fluid solution with cosmological constant of the (2+1)
gravity is singled out among all static circularly spacetimes as the only confor-
mally flat space – its Cotton tensor vanishes – sharing this conformally flatness
property with its (3+1) relative, the Schwarzschild interior perfect (incompress-
ible) fluid solution with Λ; a comparison table for these incompressible fluids is
included.
7
Hydrodynamic Equilibrium
Under Buchdahl’s conditions on the behavior of the density and the pressure for
regular fluid static stars in the presence of a cosmological constant, the bounds of
the mass are determined in any dimension. Here we work out the n-dimensional
case because of its parallelism with the 3D case.
For (2 + 1)-dimensional perfect fluid stars in hydrodynamic equilibrium there
are no bounds on the mass, except for their positiveness; the metric for a con-
stant density distribution is derived, and its matching with the external static
solution with a negative cosmological constant is accomplished. In the (d ≥ 4)-
dimensional case the existence of bounds for the mass is established. The metric
for a constant density is derived and its matching with the external static solu-
tion is carried out. Some mistakes in previous works on the topic are pointed
out.
explicitly
d κ
E11 = 0 : λ = r (κ μ + Λ)e2 λ → e−2 λ(r) = C0 − m (r) ,
dr
r r π
π π
m (r) := 2π r (μ + Λ/κ) dr = 2π r μ dr + Λ r2 =: M (r) + Λ r2 ,
κ κ
(7.6a)
r
dν (κp − Λ) r dr
E22 = 0 : − re2 λ(r) (κ p − Λ) = 0 → ν = , (7.6b)
dr C0 − πκ m (r)
2
d2 d d d
3
E3 = 0 : ν+ ν − ν λ + (Λ − κ p) e2 λ = 0. (7.6c)
dr2 dr dr dr
The substitution of the derivative dν
dr from (7.6b) into (7.6c) yields the same
equation arising from the energy–momentum conservation law T αβ ;β = 0,
namely
82 Hydrodynamic Equilibrium
dp dν
= − (μ + p) = −e2 λ r (κ p − Λ) (μ + p)
dr dr
dp r (κ p − Λ) (μ + p)
→ =− . (7.7)
dr C0 − κ M (r) /π − Λ r2
On the other hand, the substitution of p(r) from (7.6b) into (7.6c) gives rise to
2
d2 ν dν dν dλ 1 dν
+ − − = 0, (7.8)
dr2 dr dr dr r dr
dν
which is a first-order equation for N (r) := dr , which can be written in a very
simple form by introducing the functions
namely
r
d N N 2
+( ) ξ = 0 → dZ −1 = ξ dr → Z −1 = C1 + ξ(r)dr. (7.10)
dr ξ ξ 0
This relation determines the pressure p through the energy density μ in a func-
tional manner: if p were expressed by a state equation of the form p = p(μ), the
equation (7.15) gives rise to an integral differential equation for the energy as
function of the variable r.
The pressure has to vanish at boundary rb (perimeter) of the circle, p(rb ) = 0,
where the mass function M (r) determines the total mass of the fluid M (rb ) on
the circle. Because the metric signature has to be preserved throughout the whole
spacetime, the positiveness of grr imposes an upper bound on the value of the
total mass, namely
π
M (rb ) ≤ (C0 − Λ rb2 ). (7.16)
κ
M (0) = 0, m(0) = 0,
√
C0 − C0 − πκ m(rb )
pc = μ0 √ , (7.18)
− C0 + κ l2 μ0 C0 − πκ m(rb )
at the boundary r = rb
which is different compared with the expression determined by Eq. (23) of Cruz
and Zanelli (1995).
The external solution to which the uniform fluid solution can be matched is
the static anti-de Sitter metric with parameter M0 , known also as the static BTZ
solution
r2 r2
g = −(−M0 + 2 )dt2 + dr2 /(−M0 + 2 ) + r2 dφ2 (7.20)
l l
the continuity at the boundary rb of the metric for the fluid is achieved
rb2 κ r2
e2ν(rb ) = e−2λ(rb ) = C0 +
2
− M = −M0 + 2b
l π l
→ κ = π, M (rb ) = M = M0 − C0 . (7.21)
7.3 Buchdahl Theorem in d Dimensions 85
The static perfect fluid solution with Λ = −1/l2 exhibits an event horizon at
2
rh = κ μC0 ll2 −1 .
r2 Ω (r) r−d+3 Λ r2
e−2λ = 1 − 2 + 2C1 −2 , (7.29)
d−2 d−2 (d − 1) (d − 2)
where the function Ω(r) is defined by
1−d
Ω = κr rd−2 ρ (r) dr. (7.30)
where it has been taking into account the definitions given in (7.32). Hence, there
is no room for a restriction on the mass via a Buchdahl procedure.
where ξm.v. is the value of ξ where the mean value Lagrange theorem holds; let
2
us call it the “mean value” ξ coordinate. Moreover, since ddξζ2 ≤ 0 → dζ
dξ (ξ0 ) ≥
dζ
dξ (ξ), ξ0 < ξ, and considering that ζ(ξ) is a positive (increasing) function, then
dζ ¯ ζ(ξ) − ζ(0)
ζ(ξ0 ) < ζ(ξ), ξ0 < ξ → (ξ > ξm.v. ) ≤ , 0 < ξm.v. < ξ¯ < ξ.
dξ ξ
(7.36)
Because of the positiveness of ξ and ζ(ξ) and its increasing property implying
dξ ≥ 0, one can write
dζ
dζ ¯ ζ(ξ)
(ξ) ≤ , 0 < ξm.v. < ξ¯ < ξ. (7.37)
dξ ξ
In various publications, the inequality referred to as the one due to the mean
value theorem is
dζ ζ(ξ)
(ξ) ≤ ; (7.38)
dξ ξ
see, for instance, Straumann (1984), paragraph 6.6.3, Ponce de Leon and Cruz
(2000), Mak and Harko (2000), and Zarro (2009). This last inequality is treated
as a function of ξ in the whole range of its variability, without any comment on
the restrictions imposed by the mean value theorem. The writing of (7.38) in that
way is an abuse of typing, although it would be considered correct if one should
have in mind that ξ stands for the ξ¯ at the left-hand side of this inequality, which
should be very doubtful. Whatever the case, it is essentially misleading.
In terms of the original variables, the above inequality reads
1 −λ(r̄) ν(r̄) dν 1
e e (r̄) ≤ r λ eν(r) . (7.39)
2r̄ dr 2 0 re dr
The actual problem of handling the inequality (7.41) is hidden in the evaluation
r
of the integral ξ(r) = 0 2 reλ dr. To avoid dealing with specific models, i.e., to
give particular densities ρ(r), one can follow the Buchdahl procedure consisting
in replacing Ω(r) by Ω(R) and use the resulting inequality and the arising inte-
gral expressed in terms of elementary functions, namely, radicals. Since ρ(r) is
88 Hydrodynamic Equilibrium
r2 Λ
Δ := 1 − 2 ω(R), ω(R) := Ω (R) + ,
d−2 (d − 1)
R
1 Λ M (R) Λ
ω(R) = d−1 κ rd−2 ρ (r) dr + =: d−1 + . (7.44)
R 0 d−1 R d−1
At the boundary r = R the fluid metric has to be matched to the external vacuum
with Λ solution known as the d-dimensional Schwarzschild–Kottler–Tangherlini
(SKT) solution, which is given by the metric (7.22) with structural functions
r2 2m Λ r2
e−2λ(r) = e2ν(r) = 1 − 2 + =1−2 ωskt (r),
d−2 r d−1 d−1 d−2
2m Λ
ωskt (r) := d−1 + . (7.45)
r d−1
At the hypersphere boundary r = R one has
−λ(R) ν(R) −λ(R) ν(R) R2
eskt = eskt = epf = epf = 1−2 ω(R) = Δ1/2 (R),
d−2
ωskt (R) = ωpf (R) = ω(R). (7.46)
The algebraic equation for ω is a cubic equation with one of the roots being
zero, ω = 0; the remaining quadratic equation amounts to
2
(d − 2) + Λ R 2 (d − 1) Λ 2 (d − 2) + Λ R 2
ω2 − 2 2 ω + 2 = 0, (7.49)
R 2 (d − 1) R 2 (d − 1)
with roots ω± where ω+ ≥ ω(R) is given by
4
2
(d − 2) + Λ R (d − 1)
2 (d − 2) − 2 Λ R 2 (d − 1) (d − 2)
ω(R) ≤ 2 + 2 .
R 2 (d − 1) R 2 (d − 1)
(7.50)
Taking into account the definition of ω(R), (7.44), through Ω and M , one finally
obtains the main inequality:
4
M (R) (d − 2)
2 (d − 2) − 2 Λ R 2 (d − 1) (d − 2) 2mskt
≤ 2 + 2 ≥ d−3 . (7.51)
R d−3
(d − 1) (d − 1) R
For vanishing cosmological constant one gets
2
M (R) (d − 2) 2
≤2 2 ≥ R d−3 mskt . (7.52)
R d−3 (d − 1)
Moreover, for the fourth dimension, d = 4, this inequality reduces just to the
well-known expression R1 M (R) ≤ 89 ≥ R2 mskt .
The integral of e−2λ(r) , for metrics which are regular at the origin of the
coordinates, yields
κ ρ0 r2 Λ r2
e−2λ(r) = 1 − 2 −2 . (7.54)
(d − 1) (d − 2) (d − 1) (d − 2)
As far as the integral for ν(r) from (7.27) is concerned, that equation becomes
r
d 1 −λ deν
e = 0 → eν = C0 + C1 reλ dr (7.55)
dr r dr 0
Taking into account the expression (7.58) for C1 arising from the vanishing of
the pressure at the border rb = R, one arrives at
κ ρ0 (d − 1) −λ(R)
C0 = e ,
2(κ ρ0 + Λ)
κ ρ0 (d − 3) − 2Λ −λ(R)
C1 = e . (7.63)
(d − 2)(d − 1)
Finally, the metric components of a regular static solution in d-dimensions for a
perfect fluid with constant density ρ0 in the presence of a cosmological constant
Λ allowing for a matching on the hypersphere R with the external Schwarzschild–
Kottler–Tangherlini solution can be given as
κ ρ0 r2 Λ r2
1/grr = e−2λ(r) = 1 − 2 −2 ,
(d − 1) (d − 2) (d − 1) (d − 2)
2
−2λ(R) 1 κ ρ0 (d − 3) − 2Λ −λ(R) −λ(r)
−gtt = e 2ν(r)
=e 1+ e −e ,
2 κ ρ0 + Λ
λ(R)
[κ ρ0 (d − 3) − 2 Λ] e − eλ(r)
p = ρ0 e−λ(r) , (7.64)
(d − 1) κ ρ0 − [κ ρ0 (d − 3) − 2 Λ] eλ(R)−λ(r)
where the behavior at the frontier
R becomes apparent. The event horizon, if
there is any, is located at rh = (d−1)(d−2)
2(κ ρ0 +Λ) .
8
Stationary Circularly Symmetric Perfect Fluids
with Λ
Tαβ = (p + ρ) uα uβ + p gαβ ,
1
uα = (δt α + Ω δφ α ) . (8.2)
N − r (W + Ω)2
2 2
The function Ω(r) describes the differential rotation property of the fluid; if
dΩ(r) = 0, one is dealing with a differentially rotating perfect fluid, while if
Ω(r) = const., the fluid is rigidly rotating. κ stands for the (2+1) gravitational
constant, ρ(r) and p(r) denote respectively the perfect fluid energy density and
isotropic pressure.
The independent Einstein equations for the metric (8.1), coupled to a perfect
fluid and a cosmological constant Λ, are:
the pressure
F2
κp (r) − Λ = r3 (Ẇ )2 + 4 N Ṅ , (8.3a)
4r N 2
the energy density
2
2
W 1 W Ṅ
ρ (r) + Λ = r − F Ḟ − rF 2
N r N N
2 2
r2 F F Ḟ Ṅ
− (Ẇ )2 + r − (Ω 2 + 2 W Ω), (8.3b)
4 N N F N
F2
κp (r) − Λ = r3 (Ẇ )2 + 4 N Ṅ , (8.5a)
4r N 2
the energy density
2 2 2
W 1 W Ṅ r2 F
ρ (r) + Λ = r − F Ḟ − rF 2 − Ẇ 2 , (8.5b)
N r N N 4 N
λ1 = c22 → S1,
1 1
λ2,3 = − c22 ± (c22 + 2 c11 )2 + 4 c31 c13 , (8.7)
2 2
Depending on the radical the roots λ2,3 could be complex, hence the type of
the Cotton tensor would be Type I: {S, Z, Z̄}. This kind of metric allows for
conformally flat solutions too.
96 Stationary Circularly Symmetric Perfect Fluids with Λ
d2 W 3 dW W1
2
+ = 0, → W = 2 , W0 = 0,
dr r dr r
dW 2 W1
+ W = 0, → W = 2 , W1 = −J/2, BTZ-like rotation. (8.8)
dr r r
Hence, the stationary circularly symmetric (2+1) metric to be studied in this sub-
section for the derivation of interior solutions modeled through a rigidly rotating
perfect fluid with a unit 4 velocity
r
uα = 2 √ δt α ,
4 r2 N2 − J2
in the presence of a cosmological constant Λ amounts to
2
dr2 J
ds = −N (r) dt +
2 2 2
+ r dφ − 2 dt .
2
(8.9)
F (r)2 2r
This class of stationary (2+1) spacetimes coupled to a perfect fluid are deter-
mined completely by an arbitrary function N (r); for a given function N (r) one
determines the remaining function F (r), the energy density, and the pressure.
The equation EQN (8.5d), once substituted W = −J/(2 r2 ), can be written as
d F
EQN : 4 r3 Ṅ N + J 2 = 0, (8.10)
dr N r4
therefore
N r4 1 rF0 1 J2
F (r) = F0 → Ṅ = − , (8.11)
4 r3 N Ṅ + J 2 4 F 4 r3 N
where dots denote derivatives with respect to the radial coordinate r. Substitut-
ing W = −J/(2r2 ) and using Ṅ from (8.11) into the expression (8.5a) for the
pressure, one obtains
F0 F
κ p(r) − Λ = . (8.12)
4 N
The expression of the energy density for W = −J/(2r2 ) from (8.5b) yields
2
1 J r Ṅ + N F2 1 J 2 − 4 r2 N 2
ρ (r) κ + Λ = − + F Ḟ , (8.13)
4 r4 N 3 4 N 2 r3
which, modulo the function F (r) from (8.11), can be written entirely in terms
of N and its derivatives as
8.2 Garcia Stationary Rigidly Rotating Perfect Fluids 97
F02 r4
κρ (r) + Λ = − 3 4r4 N N̈ (J 2 − 4r2 N 2 )
4 4r3 N Ṅ + J2
+ 4r3 Ṅ J 2 r Ṅ + 4r2 N 3 + J 2 N + 16 J 2 r2 N 2 − 3J 4 . (8.14)
d( dQ d( 12 + 8 κ ρF0 2+Λ dQ
dx )
dx ) dQ
dQ
− 0
= , (8.19)
dx
1
2 + 8 κ ρF0 2+Λ dQ
dx
2Q
0
where vertical bars stand for absolute values. The above relation can be put into
the differential form
2
κ ρ0 + Λ κ ρ0 + Λ |Q|
d B0 − 4 x =d 1−8 ; (8.21)
F02 Q0 F02 |Q0 |
J2
W (r) = − , (8.23)
2r2
1 J2 2
N (r)2 = + N q02 [1 − K(r)] , (8.24)
4 r2
κ ρ0 + Λ K(r)
F (r) = 4 N (r), (8.25)
F0 1 − K(r)
1 F02
K(r) := B0 − r2 , (8.26)
16 (κ ρ0 + Λ)q02
2
F02
where q02 := Q0 8(κ ρ0 +Λ) , = ±, and N = ±.
κ ρ0 K(r) + Λ
κp(r) = . (8.27)
1 − K(r)
From this expression one establishes that at certain value r = rzp , where the
subscript zp denotes zero pressure, the pressure may vanish for = 1, Λ = −1/l2 ,
and = −1, Λ = 1/l2 ; one has to √ keep in mind these sign conditions when
extracting square roots; for instance, Λ2 = (−Λ)2 = −Λ. The vanishing of
the numerator of p(r), K(rzp ) = − Λ/(κ ρ0 ), yields
F02 2 Λ2
B0 = N 2 rzp + 2 2. (8.28)
16(κ ρ0 + Λ)q0 κ ρ0
8.2 Garcia Stationary Rigidly Rotating Perfect Fluids 99
Therefore, for vanishing pressure at the circle rzp , the structural functions N (r)
and F (r) amount to
J2 2
N (r)2 = + N q02 [1 − K(r, rzp )] , (8.29)
4r2
κ ρ0 + Λ K(r, rzp )
F (r) = 4 N (r), (8.30)
F0 1 − K(r, rzp )
Λ2 F02
K(r, rzp ) = + (r2 − r2 ). (8.31)
κ2 ρ20 N
16(κ ρ0 + Λ)q02 zp
Moreover, at the radius rzp , the functions N (r) and F (r) become
1 J2 2 (κ ρ0 + Λ)
2
N (rzp )2 = + q 0 , (8.32)
2
4 rzp N
κ2 ρ20
4Λ
F (rzp ) = − N (rzp ). (8.33)
F0
On the other hand, the vacuum solution for the metric (8.9) is determined by
the structural functions
1 J2
N (r)2 = F (r)2 = − Λ r2 − M, (8.34)
4 r2
which in the case of negative cosmological constant Λ = −1/l2 becomes the
well-known BTZ black hole solution given by the metric (8.9) with
1 J2 r2
N (r)2 = F (r)2 = 2
+ 2 − M. (8.35)
4r l
Comparing (8.34) with (8.32) at rzp , one arrives at
κ2 ρ20
q02 = N 2
(−Λ rzp − M ), (8.36)
(κ ρ0 + Λ)2
and consequently we have the following sub-branches of solutions:
A)N = 1,
2
κ2 ρ20 rzp
A1: Λ = −1/l2 , q02 = (κ ρ0 −1/l2 )2 ( l2 − M ), rzp
2
/l2 − M > 0,
2
κ2 ρ20 rzp
A2: Λ = 1/l2 , q02 = − (κ ρ0 +1/l 2 )2 ( l 2
2
+ M ), rzp /l2 + M < 0,
B) N = −1,
2
κ2 ρ2 rzp
B1:Λ = −1/l2 ,q02 = − (κ ρ0 −1/l
0
2 )2 ( l2 − M ), rzp
2
/l2 − M < 0,
2
κ2 ρ20 rzp
B2: Λ = 1/l2 , q02 = (κ ρ0 +1/l2 )2 ( l2
2
+ M ), rzp /l2 + M > 0.
Substituting F0 = −4 Λ, which in turn yields F (rzp ) = N (rzp ), and q0 from
(8.36) into (8.29–8.31), one obtains the expressions of the structural functions
100 Stationary Circularly Symmetric Perfect Fluids with Λ
κ ρ0 + Λ K(r, rzp )
F (r) = − N (r), (8.38)
Λ 1 − K(r, rzp )
Λ κ ρ0 + Λ 2
K(r, rzp ) = − 1− 2 +M
(rzp − r2 ), (8.39)
κ ρ0 Λ rzp
K(r, rzp )
F (r) = (κ ρ0 − 1/l2 ) N (r), (8.42)
κ ρ0 − l12 K(r, rzp )
κ ρ0 − 1/l2 2
K(r, rzp ) := 1− (r − r2 ), (8.43)
M − rzp
2 /l2 zp
one brings the structural functions N (r) and F (r), (8.24) and (8.25), to the form
2
2 1 J2 N Λ
N (r) = + c1 + 4 K(r) ,
4 r2 16 κ2 ρ20 − Λ2
K(r)
F (r) = 4(κ ρ0 + Λ) N (r),
c1 κ2 ρ20 − Λ2 + 4ΛK(r)
κ2 ρ20 − Λ2 16Λ2 2
K(r) := − c21 − 16c2 − N r , (8.46)
4Λ κ ρ0 + Λ
accompanied by the isotropic pressure,
c1 κ2 ρ20 − Λ2 − 4κρ0 K(r)
κ p(r) = Λ ,
c1 κ2 ρ20 − Λ2 + 4ΛK(r)
(8.47)
where = ±, and N = ± independently. As far as the behavior of the pressure
is concerned, one may establish that for r running from the value rpress↑∞ =
−N c2 (κ ρ0 + Λ)/Λ2 , at which point the denominator of the pressure p (8.47)
becomes equal to zero, up to the value rzp at which the numerator of the pressure
vanishes, K(rzp ) = c1 κ2 ρ20 − Λ2 /(4κρ0 ), the pressure is a positive function of
the radial coordinate, which decreases from a very large value at rpress↑∞ to zero
at the boundary circle rzp , at which
1 J2 c21 (κ ρ0 + Λ)2
N (rzp )2 = F (rzp )2 = + ,
2
4 rzp N
16 κ2 ρ20
κ ρ0 Λ2
c1 = 4 c2 + N r2 .
κ ρ0 − Λ
2 2 2 κ ρ0 + Λ zp
Taking into account the existence of a zero-pressure circle r = rzp , the structural
functions from (8.46) can be given as
1 J2 2
N (r)2 = + 2 2N 2 [κ ρ0 K(rzp ) + Λ K(r)] ,
4 r2 κ ρ0 − Λ
K(r)
F (r) = (κ ρ0 + Λ) N (r),
κ ρ0 K(rzp ) + ΛK(r)
N Λ2 rzp
2
K(r) := c2 + + (κ ρ0 − Λ) (rzp
2 − r 2 ), (8.48)
κ ρ0 + Λ
together with
K(r) − K(rzp )
p(r) = −ρ0 Λ , (8.49)
κ ρ0 K(rzp ) + Λ K(r)
where K(rzp ) = K(r = rzp ), c1 = 4 κ ρ0 K(rzp )/(κ2 ρ20 − Λ2 ), meaningful for
positive and negative values of the cosmological constant Λ and the parameter M .
102 Stationary Circularly Symmetric Perfect Fluids with Λ
Notice that the vacuum solution arises as a limiting case of the above interior
metric functions, (8.49), for ρ0 = 0 accompanied with the identification c2 →
−N M .
Comparing (8.34) with the expression (8.49), one establishes that the ful-
fillment of the matching conditions, Next (rzp ) = Nfluid (rzp ), together with
Fext (rzp ) = Ffluid (rzp ), is guaranteed by
κ ρ0 − Λ
K(rzp )2 = −N 2
(Λ rzp + M ).
κ ρ0 + Λ
Substituting K(rzp ) into K(r) from (8.48), one has
κ ρ0 − Λ 2
K(r) = K(rzp ) 1 − 2 +M
(rzp − r2 ), (8.50)
Λ rzp
and hence the structural functions assume the form given in (8.37)–(8.39).
et al. (1999), assuming (anti-)de Sitter exterior geometries. The metric in coor-
dinates {t, r, φ} with structural functions ρ(r), Z(r), B(r), and Y (r) is given as
2 2 2
ds2 = −(ρ(r)dt + Z(r)dφ)2 + B(r)2 dr2 + Y (r)2 dφ2 = −θ0 + θ1 + θ2 ,
θ0 = ρ(r)dt + Z(r)dφ, θ1 = B(r)dr, θ2 = Y (r)dφ, (8.55)
where θa , a = 0, 1, 2 is the orthonormal triad basis. Comparing with the corre-
sponding metric of Lubo et al. (1999), the function T (r) has been replaced by
ρ(r) here; the reason for proceeding in this way is due to the convenience of a
variable change to the spatial radial coordinate ρ(r). For the fluid energy density
here is reserved μ(r).
The Einstein tensor triad components are Ga b = Ra b − Rδ a b /2 = κT a b − δ a b Λ
are
2
1 Y 1 ρ Z
0
G 0= − , (8.56a)
BY B 4 BY Y
2
1 ρ Y 1 ρ Z
G 1 = 2 + , (8.56b)
B ρY 4 BY Y
2
2 1 ρ 1 ρ Z
G 2 = + , (8.56c)
Bρ B 4 BY Y
1 1 ρ3 Z
G 2
0 = −G 0
2 = . (8.56d)
2 Bρ2 BY ρ
∂
The fluid one-form is chosen as linear combination os the Killing vectors ∂t and
∂
∂φ , namely
T 0 0 = −μ (r) , T 2 0 = −T 0 2 = 0,
T 1 1 = p(r) = T 2 2 , (8.61)
together with
ρ
p = − (μ + p). (8.62)
ρ
Therefore, the metric for a rigidly rotating fluid in all generality can be given as
2
ρ
ds2 = −(ρ dt + Z dφ)2 + B 2 dr2 + dφ2 ,
B
β ρ
Z(r) = α ρ(r) + , Y (r) = , (8.66)
ρ(r) B
where α, β, and B are integration constants, with the energy and pressure
fulfilling
2
1 Y
1 ρ Z
−(κμ (r) + Λ) = − ,
BY B
4 BY Y
2
ρ Y 1 ρ Z
κp(r) − Λ = 2 + . (8.67)
B ρY 4 BY Y
The energy conservation p = − ρρ (μ + p) does not provide new conditions; for
the given μ and p, it becomes an identity.
8.3 Lubo–Rooman–Spindel Rotating Perfect Fluids 105
d3 Z dZ
+ (κ μ (r) + 4Λ) = 0.
dr3 dr
The incompressible dust was treated in detail in Section 3.6.
9
Friedmann–Robertson–Walker Cosmologies
This chapter is devoted to the derivation of exact solutions to the FRW model
in (3+1) and (2+1) dimensions and to establish their correspondence following
Garcı́a, Cataldo and del Campo, Garcı́a et al. (2003). It is widely known that
any (3+1) FRW perfect fluid fulfilling a barotropic equation of state of the form
p(ρ) = (γ − 1) ρ, under algebraic transformation rules of the parameters, can be
transformed into its (2+1) FRW counterpart. It is noteworthy that the physical
content of the solutions depends on the dimensionality in which they are viewed.
Using four-dimensional terminology, for a vanishing cosmological constant, the
(2+1) analog of the (3+1) dust (γ4 = 1), p4 = 0 is a radiation-dominated universe
(γ3 = 3/2), p3 = ρ3 /2. Conversely, for a (3+1) radiation-dominated universe
(γ4 = 4/3), p4 = ρ4 /3, one finds that the (2+1) counterpart is the stiff matter
(γ3 = 2), p3 = ρ3 . Moreover, the analog of the (3+1) de Sitter spacetime is the
(2+1) de Sitter spacetime, γ4 = 0 = γ3 , with equation of state p = −ρ = const.
in both cases. The family of FRW barotropic solutions are derived in details.
Moreover, the full description of collapsing dust is given for: the Mann–Ross
dust FRW solution with Λ, Section 9.4, and the Gidding–Abbott–Kuchař dust
FRW solution, Section 9.4.5.
where, as usual, a(t) is the scale factor, and k = −1, 0, 1. The FRW metric
is conformally flat, i.e., its Weyl tensor vanishes everywhere in the domain of
definition of the spacetime. The scale factor a(t) of the metric (9.1) is governed
by equations modeled in terms of the perfect fluid energy density ρ4 , the matter
isotropic pressure p4 , and a cosmological constant Λ, if present: the energy–
momentum tensor is given by
Tμν = (ρ4 (t) + p4 (t)) uμ uν + p4 (t) gμν , uμ = −δμt ,
hence, the Einstein equations
R
Rμν − gμν + Λ gμν = κ Tμν ,
2
amounts explicitly to
3 2
Ett : ȧ + k = Λ4 + κ4 ρ, (9.2a)
a2
E i i : 2 a ä + ȧ2 + k + a2 (κ p4 − Λ4 ) = 0
a
→ a ä + (κ ρ4 + 3 κ p4 − 2 Λ4 ) = 0, (9.2b)
6
d da3
ρ4 a3 + p4 = 0 ≡ T μ ν ;μ = 0. (9.2c)
dt dt
The equation (9.2c) represents the conservation of the matter content: T μ ν ;μ = 0.
It results also from the substitution of ȧ from (9.2a) into ä of (9.2b); thus only
two of the equations of this system are independent.
d 2 da2
ρa + p = 0. (9.4c)
dt dt
110 Friedmann–Robertson–Walker Cosmologies
Substituting ȧ from (9.4a) into (9.4b) one arrives at (9.4c), which is the equation
arising from the energy conservation T μ ν ;μ = 0.
For dust, the pressure vanishes, p = 0, and the conservation of the energy–
momentum tensor implies that
3 + 1 : ρ a3 = ρ0 a30 = const.,
2 + 1 : ρ a2 = ρ0 a20 = const.,
where ρ0 is the initial density and a0 the value of the scale factor at the beginning
of the time counting.
p4 = (γ4 − 1) ρ4 , (9.5)
d ln ρ4 + 3γ4 d ln a = 0, (9.6)
t − t0 = a , (9.8)
Λ4 κ4
a3γ4 − ka2( 2 γ4 −1) +
3
3 3 ρ40
where a = ±1.
From (9.8), it becomes apparent that one can not, in general, express in terms
of elementary functions t as function of a.
Nevertheless, for Λ4 = 0 and arbitrary γ4 , the above integral is given in terms
of hypergeometric functions, namely
9
2 γ4 − 2 a 2 γ4 −1
3 3 3
a a 2 γ4 1 2 γ4
t(a) = t0 + F , , ,k . (9.9)
ρ40 κ4 /3 32 γ4 2 3γ4 − 2 3γ4 − 2 ρ40 κ4 /3
9.2 Barotropic Perfect Fluid FRW Solutions 111
This kind of solution was also reported in Giddings et al. (1984), Eq.(77). In
the paragraph prior to this equation, it is pointed out “that in structure the
equation for a radiation-dominated Universe in n dimensions is identical to that
for a dust-filled Universe in n + 1 dimensions.”
Later, in 1991, this kind of cosmology was reported in Cornish and Frankel
(1994), Eq.(4.5–4.6).
(2 + 1) de Sitter Cosmological Solution
On the other hand, for γ3 = 0, consequently p3 = −ρ3 = −ρ30 = const., i.e., one
is dealing with the de Sitter metric, the scale factor amounts to
e−a C0 (t−t0 ) √
a(t) = k + e2a C0 (t−t0 ) , C0 = κ3 ρ30 . (9.16)
2C0
Saslaw (1977), in the paragraph following equation (6), wrote the sentence: “If
Λ = 0, one can similarly compare models for 2+1 and 3+1 dimensional universes.
For p = 0 – dust – one obtains a = a0 exp t/τ , which is the analogous of the de
Sitter model.”
structurally invariant function. One reaches the same conclusion when dealing
with the hypergeometric function representation of t determined by (9.9) and
(9.14). On the other hand, if one were able to express the scale factor a as a
function of the variable t, then, via parameter scaling, one would arrive at the
structurally invariant character of the function a(t).
We have established in this way that any FRW cosmology, filled with a perfect
fluid fulfilling a linear state equation, determined in (3+1) dimensions, can be
reduced to its (2+1) counterpart by using the correspondence (9.18); the converse
statement holds, too.
Moreover, considering (3+1) and (2+1) FRW cosmologies as independent enti-
ties, dominant energy conditions for fluids: ρ ≥ 0 and −ρ ≤ p ≤ ρ, have to
hold on their own account in (2+1) and (3+1) dimensions. Therefore, the (3+1)
dimensional state parameter γ4 has to fulfill the condition 0 ≤ γ4 ≤ 2, while
independently the (2+1) dimensional state parameter γ3 has to range the values
0 ≤ γ3 ≤ 2.
On the other hand, assuming that the considered spacetimes are in the corre-
spondence (9.18), one arrives at restrictions for the values one can assign to the
state parameters γ, namely:
4
0 ≤ γ3 ≤ 2, and 0 ≤ γ4 ≤ .
3
Thus, the class of (3+1) perfect fluid cosmologies which participates in the cor-
respondence with the whole family of (2+1) perfect fluid cosmologies is more
narrow compared with the whole (3+1) perfect fluid cosmology; (3+1) cosmol-
ogy with 43 < γ4 ≤ 2 are out of the comparison scheme. Hence, thinking in
terms of dimensionally reduced spaces, a perfect fluid FRW solution given in
(3+1) dimensions, which can be reduced to its (2+1) cosmological counterpart,
possesses state parameters given in the above-specified ranges. From this point
of view, using the four-dimensional terminology, for vanishing cosmological con-
stants Λ4 = Λ3 = 0, the (2+1) analog of the (3+1) dust (γ4 = 1), p4 = 0, is
a radiation-dominated universe (γ3 = 3/2), p3 = ρ3 /2. Conversely, for a (3+1)
radiation-dominated universe (γ4 = 4/3), p4 = ρ4 /3, one finds that the (2+1)
counterpart is the stiff matter (γ3 = 2), p3 = ρ3 . Moreover, the (3+1) de Sit-
ter spacetime coincides with the (2+1) de Sitter spacetime, γ4 = 0 = γ3 , with
equation of state p = −ρ = const.
p4 = α4 ρ4 γ4 , (9.19)
114 Friedmann–Robertson–Walker Cosmologies
d ln [(α4 + ρ(−γ
4
4
+1) 3(−γ4 +1)
)a ] = 0 → ρν4 4 = A40 a−3ν4 − α4 , (9.21)
where ν4 = 1 − γ4 .
Moreover, the integral of t, (9.2a), amounts to
da
t − t0 = . (9.22)
Λ4 2 1/ν4
3 a −k+ 3 a
κ4 2
A40 a−3ν4 − α4
Hence, for perfect fluids fulfilling the polytropic state equation p = αρ(1−ν) , we
have no relations between exponential factors νd alone.
The Einstein equations for a perfect fluid with a cosmological constant for the
(2 + 1) Friedmann–Robertson–Walker metric (9.3),
dr2
ds = −dt + a(t)
2 2 2 2 2
+ r dφ ,
1 − kr2
where ρ0 is the initial density and a20 is the value of the scale factor at the
beginning of the time. The remaining equations are:
2
d2 a da
+ Λmr a = 0, + k − κ ρ0 a20 + Λmr a2 = 0. (9.27)
dt2 dt
The exterior metric has to be the de Sitter one, with negative parameter M < 0
to have the correct metric signature; see (9.32) below.
ȧ0 da
a(t) = a0 cos( Λmr t) + sin( Λmr t), ȧ0 = |t=t0 ,
Λmr dt
ȧ0 = κ ρ0 a20 − k − Λmr a20 ; κ ρ0 a20 − k − Λmr a20 ≥ 0, real a(t).
(9.30)
At time of the collapse, t = tc , this solution has to vanish, a(tc ) = 0, hence
1 a0 Λmr
tc = arctan . (9.31)
Λmr κ ρ0 a20 − k − Λmr a20
[Kμν ] = 0. (9.33b)
(−) (+) (−) (+)
From g (rb )− g (R(t)) = 0, and g φφ (rb ) = g φφ (R(t)) one gets
correspondingly
2 2
dT 1 dR
− 1 = −(Λmr R(t) − M ) 2
+ ,
dt Λmr R(t)2 − M dt
rb a(t) = R(t). (9.34)
The latter condition R(t) = rb a(t) means that in the exterior coordinates the
position R of the boundary is equal to the proper distance rb a(t) from the origin
to the dust edge. The first condition yields
dT Λmr R(t)2 − M + Ṙ(t)2
= , (9.35)
dt Λmr R(t)2 − M
d
on the dust edge; the over-dot denotes dt .
9.4 Mann–Ross Collapsing Dust FRW Solutions with Λ 117
The initial conditions a0 = 1, and ȧ0 = 0 represent a ball of dust with initial
radius R(t0 ) = rb a(t0 ) = rb initially at rest Ṙ(t)|t0 = rb ȧ(t)|t0 = 0 in the
exterior coordinates.
where ei ν are the components of the tangent vector ∂i to the coordinate curve
ξ i defined on the surface, Nα is the normal vector to the surface, and as such
is orthogonal to ei ν . Operationally, it is recommendable to use the directional
description ei ν ∂x∂ ν = ∂ξ∂ i , thus
(±)
∂ α ∂
Kij = Nα ej + ei ν ej σ Γα νσ = −ej α N α − e i
ν
N σ Γ σ
να , (9.37)
∂ξ i ∂ξ i
which is close to the Israel formulation.
(−) 1 1 1 − k rb2
Kφφ = N α Γ α
φφ = Γ r
φφ = − . (9.40)
rb2 a(t)2 rb2 a(t) 1 − k rb2 rb a(t)
(+)
For Kφφ one has
(+) Λmr R2 − M
Kφφ = Nμ Γμ φφ /R2 = −T,λ , (9.44)
R
where the non-vanishing Christoffel symbols for the external metric are
Λmr R
ΓT T R = = −ΓR RR , ΓR T T = Λmr R(Λmr R2 − M ),
Λmr R2 − M
1
ΓR φφ = − R(Λmr R2 − M ), Γφ Rφ = . (9.45)
R
(+)
For Kλλ one has
(+)
Kλλ = Nμ ∂λ eλ μ + Nμ eλ ν eλ σ Γμ νσ . (9.46)
Consider first
Nμ ∂λ eλ μ = Nμ ∂λ (T,λ δTμ + R,λ δR
μ
) = (−R,λ δμT + T,λ δμR ) (T,λλ δTμ + R,λλ δR
μ
)
= −T,λλ R,λ + R,λλ T,λ . (9.47)
On the other hand,
Nμ eλ ν eλ σ Γμ νσ = −R,λ T,λ T,λ ΓT T T + 2 T,λ R,λ ΓT T R + R,λ R,λ ΓT RR
+T,λ T,λ T,λ ΓR T T + 2 T,λ R,λ ΓR T R + R,λ R,λ ΓR RR . (9.48)
Substituting the Christoffel symbols one arrives at
Λmr R
Nμ eλ ν (eλ σ Γμ νσ ) = Λmr R(Λmr R2 − M )(T,λ )3 − 3 T,λ (R,λ )2 .
Λmr R2 − M
(9.49)
Setting λ = t, and using T,t from (9.35), together with
ṘR̈
T,tt =
(Λmr R2 − M ) Λmr R2 − M + Ṙ2
Λmr R Ṙ Λmr R2 − M + 2Ṙ2
− (9.50)
(Λmr R2 − M )2 Λmr R2 − M + Ṙ2
one arrives at
(+)
∂
Ktt = Λmr R2 − M + Ṙ 2 . (9.51)
∂R
Therefore, since on the boundary r = rb the smoothness of Kφφ from (9.40) and
(9.44) yields
1 − k rb2
Λ R −M
2 Λmr R2 − M + Ṙ 2
− = −T,λ mr =− , (9.52)
rb a(t) R R
9.4 Mann–Ross Collapsing Dust FRW Solutions with Λ 119
where it has been used T,λ = T,t from (9.35). Since on the boundary
where (9.27) was used in agreement with the results reported in Mann–Ross.
The existence of an event horizon in the static BTZ black hole around the
collapsing dust, Rh = M Λmr , M > 0, requires ρ0 > 1/(κ rb2 a20 ).
This scalar factor a(t) can also be obtained from (9.30) as a limit as Λ goes to
zero. The collapse, if any, occurs at a(tc ) = 0, i.e.,
tc = −a0 /ȧ0 = a0 / κ ρ0 a20 − k,
where ȧ0 = − κ ρ0 a20 − k has been chosen to have tc > 0, assuming a0 > 0; the
collapse has to take place after the initial time t = 0.
The exterior metric (9.32) for vanishing Λ becomes a flat metric with conical
singularity
dR2
g = −C dT 2 + + R2 dφ2 , C := −M > 0. (9.57)
C
Choosing the dust edge at r = rb , with equation R(t), the matching conditions
(9.33a) and (9.33b) of the FRW metric (9.3) to the metric (9.57) are those of
(9.34), (9.35), (9.52) and (9.51) for Λ = 0, namely
(−)
(+)
d 2
Ktt = 0 = C + Ṙ(t) =Ktt , (9.58d)
dR
Using Ṫ from (9.58b) in (9.58c), taking into account that R(t) = rb a(t), one
arrives at
1 − k rb2 = C + Ṙ(t)2 , (9.59)
which, isolating C, considering that R(t) = rb a(t) and the derivative ȧ(t) from
(9.56), yields
2
C = 1 − (k + ȧ(t) ) rb2 = 1 − κ rb2 a20 ρ0 > 0, (9.60)
in agreement with Mann and Ross (1993), Eq. (32).
10
Dilaton–Inflaton Friedmann–Robertson–Walker
Cosmologies
The purpose of this chapter is to provide a new insight on (2+1) and (3+1)
Friedmann–Robertson–Walker (FRW) cosmologies by establishing a bridge
between them. In order to achieve this goal, I shall begin with a comparison
of the dynamical equations corresponding to (2+1) and (3+1) FRW spacetimes
coupled to matter perfect fluid sources, scalar field (inflaton, dilaton) fields, and
cosmological constants. A (2+1) FRW spacetime may be considered as a dimen-
sional reduction of the associated (3+1) FRW spacetime, arising as result of
the freezing (constant value assignation) of the azimuthal angle (in spherical
coordinates) of this last (3+1) space. A similar approach has been applied suc-
cessfully by Cataldo, del Campo and Garcia, Cataldo et al. (2001), to the (3+1)
Plebański–Carter[A] metric – see Plebański (1975) and Carter (1968) – to derive
the (2+1) BTZ black hole solution.
It is shown that FRW cosmological models coupled to a single scalar field
and to a perfect fluid fitting a wide class of matter perfect fluid state equations,
determined in (3+1) dimensional gravity, can be related to their (2+1) cosmo-
logical counterparts, and vice versa, by using simple algebraic transformations
relating gravitational constants, state parameters, perfect fluid and scalar field
characteristics. It should be pointed out that the demonstration of these relations
for the scalar fields and potentials does not require the fulfillment of any state
equation for the scalar field energy density and pressure. As far as the perfect
fluid is concerned, one has to demand the fulfillment of state equations of the
form p + ρ = γ f (ρ). If the considered cosmologies contain the inflaton field alone
φ, then any (3+1) scalar field cosmology possesses a (2+1) counterpart, and vice
versa.
Notice that one is tacitly assuming that coordinates remain the same ones
for both (3+1) and (2+1) FRW metrics. It is notable that these spacetimes are
both conformally flat, i.e., correspondingly their Weyl and Cotton ten-
sors vanish. By associated (corresponding) spacetimes we mean spaces that
belong to a specific family: for instance, spaces fulfilling a (linear) barotropic
state equation, or those fitting a polytropic law. Moreover, (2+1) FRW solu-
tions to a barotropic perfect fluid state equation are in correspondence with
122 Dilaton–Inflaton Friedmann–Robertson–Walker Cosmologies
where, as usual, dΩ2 := dθ2 + sin2 θdφ2 , a(t) is the scale factor, and k = −1, 0, 1
denotes the curvature index. This metric is conformally flat, i.e., its Weyl tensor
vanishes everywhere in the domain of definition of the spacetime. The scale factor
a(t) of the metric (10.1) is governed by equations modeled in terms of the perfect
fluid energy density ρ4 , the matter isotropic pressure p4 , the scalar field φ, the
self-interacting potential V (φ), and a cosmological constant Λ, if present.
ȧ
ρ˙4 + 3 (ρ4 + p4 ) = 0, (10.8)
a
ȧ
ρ̇φ + 3 (ρφ + pφ ) = 0, (10.9)
4 a 4 4
The (10.3) represents the conservation of the matter content, while (10.4)
corresponds to the energy conservation of the scalar field.
This metric is characterized by the vanishing of the Cotton tensor, and hence,
from the 3D point of view, it is conformally flat. Moreover, the metric (10.10)
can be considered as a dimensional reduction of the metric (10.1) for any fixed
value of the azimuthal angle φ.
ȧ
ρ˙3 + 2 (ρ3 + p3 ) = 0, (10.12)
a
ȧ d
φ¨3 + 2 φ̇3 + V (φ3 ) = 0. (10.13)
a dφ3
1 ˙2
pφ = φ − V3 , (10.15)
3 2 3
the (2+1) dynamical equations become
ȧ2 + k
= κ3 (ρ3 + ρφ ) + Λ3 , (10.16)
a2 3
ȧ
ρ˙3 + 2 (ρ3 + p3 ) = 0, (10.17)
a
ȧ
ρ̇φ + 2 (ρφ + pφ ) = 0. (10.18)
3 a 3 3
It is apparent that the field equations for metrics (10.1) and (10.10) are dif-
ferent because of the difference in dimensions. Nevertheless, one may assume
that the time coordinate t remains the same in both (2+1) and (3+1) dimen-
sions. Moreover, one also may assume that the scale factor a(t) is a structurally
invariant function depending on t and certain constants; by structural invari-
ance we mean that under dimensional reduction the function a(t) maintains its
form with respect to the t variable as well as its dependence on the constants
involved. The extension of this concept to functions depending on other variables
is straightforward.
The main result of this section can be formulated as a theorem.
10.1 Equations for a FRW Cosmology with a Perfect Fluid 125
Proof Considering that the time coordinate t as well the scale factor a(t) remain
unchanged, comparing (10.2) and (10.11) one has
ȧ2 + k κ4 1 2 Λ 1 2
= (ρ4 + φ˙4 + V4 ) + 4 = κ3 (ρ4 + φ˙3 + V3 ) + Λ3 ⇒
a2 3 2 3 2
κ4 Λ4
κ3 , Λ3 , ρ4 ρ3 , (10.20)
3 3
together with
1 ˙2 1 2
φ4 + V4 φ˙3 + V3 , ∼ ρφ ρφ . (10.21)
2 2 4 3
Next, assuming that in each space the state equation for matter is of the form
p + ρ = γf (ρ), where f (ρ) is a structurally invariant function, i.e., it is a form-
invariant function as viewed from the spaces under consideration, the matter
conservation equations yield
da 1 dρ4 1 dρ3
=− =− , (10.22)
a 3γ4 f (ρ4 ) 2γ3 f (ρ3 )
hence, because of by assumption f (ρ4 ) f (ρ3 ), one has
ρ4 ρ3
a 1 dρ 1 dρ
ln =− =− , (10.23)
a0 3γ4 f (ρ) 2γ3 f (ρ)
therefore
3γ4 2γ3 . (10.24)
To establish the remaining relationships on scalar fields φ and potentials V (φ)
we rewrite (10.4) and (10.13) correspondingly as:
d 1 ˙2 d ȧ 2
(3 + 1) : φ4 + V (φ4 ) + 3 φ˙4 = 0,
dt 2 dt a
d 1 ˙2 d ȧ ˙ 2
(2 + 1) : φ + V (φ3 ) + 2 φ3 = 0. (10.25)
dt 2 3 dt a
126 Dilaton–Inflaton Friedmann–Robertson–Walker Cosmologies
Starting from the (3+1) equation, assuming again that t and a(t) remain
unchanged under dimensional reduction, we shall establish the transformations
of φ4 and V4 to recover the (2+1) equation. First, one has
2 2
1 d 1 3 d ȧ 3 ˙
(1 − ) φ̇ + V (φ4 ) + 2 φ = 0, (10.26)
3 dt 2 2 4 dt a 2 4
which rewrites as
2 2
d 1 3 d 1 ˙2 ȧ 3 ˙
φ̇ + [V (φ4 ) − φ4 ] + 2 φ = 0. (10.27)
dt 2 2 4 dt 4 a 2 4
In the next sections, cosmologies with scalar fields subjected to state equa-
tions pφ + ρφ = ΓF (ρφ ), for which (10.31) holds, are derived. The advantage
of using these equations resides in the uniqueness of the derived solutions. For
such branches of solutions one tacitly assumes that the above condition (10.31)
is fulfilled.
For matter perfect fluids, dominant energy conditions require that ρ ≥ 0, and
−ρ < p < ρ, therefore determining this kind of solution, one has additionally
to take care of the fulfillment of this inequality in each spacetime, no matter its
dimension. The validity of this physical requirement is assumed to hold beyond
four dimensions.
d 2
ρ + (ρφ + pφ ) = 0, (10.32)
da φ3 a 3 3
hence
aΓ3 −1
t − t0 = a da, (10.35)
κ3 ρφ − k a2(Γ3 −1)
30
Adding (10.14) and (10.15) one gets (dφ3 /dt)2 = ρφ + pφ , which can be
3 3
written as
2
dφ3
ȧ2 = Γ3 ρφ . (10.37)
da 3
Substituting in the above equation ȧ2 from (10.34), and the expression of ρφ
3
from (10.33), one obtains
Γ3 da
dφ3 = a φ ρφ , (10.38)
− ka2Γ3 −2 a
30
κ3 ρφ
30
Γ3 −1
which, by introducing the variable z = a , amounts to
a φ Γ3
dφ3 = − √ d ln κ3 ρφ + κ3 ρφ − kz 2 − ln z . (10.39)
(Γ3 − 1) κ3 30 30
−2 ΓΓ−1
3
2 − Γ3 exp (−C3 (φ3 − φ30 )) 3
V3 = ρφ 2 κ3 ρφ . (10.42)
2 30 30 k + exp (−2C3 (φ3 − φ30 ))
t − t0 = a da, (10.44)
κ4 2( 32 Γ4 −1)
ρ
3 φ − k a
40
10.2 Single Scalar Field to Linear State Equations; Λ = 0 129
2
dln ρφ + dln a = 0, (10.50)
4 α
with solution
ρφ = ρφ (a/a0 )−2/α , (10.51)
4 40
130 Dilaton–Inflaton Friedmann–Robertson–Walker Cosmologies
where the constant a0 has been introduced for further convenience; notice that
the dynamical equations, for k = 0 and Λ4 = 0, are invariant under the change
a → a/a0 , thus without lost of generality one may set a0 = 1. Substituting this
expression of ρφ into (10.7), which yields
4
d(a/a0 )1/α = 1/α κ4 ρφ /3 dt, (10.52)
40
a(t) = a0 tα . (10.54)
Finally, the evaluation of V4 = ρφ − φ˙4 /2 yields the slow roll the self-interacting
4
potential
α −2 α κ
Vφ = (3α − 1) t = (3α − 1) e−φ 2 α4 (φ4 −φ40 )
. (10.57)
4 κ4 κ4
10.3 Spatially Flat FRW Solutions for Barotropic Perfect Fluid and
Scalar Field
The derivation of the general solutions for spatially flat (k = 0) FRW spacetimes
filled simultaneously with matter and scalar field, modeled by two perfect fluids –
one related to matter and the second one related to the scalar field – is presented
in some details.
ρ4 + p4 = γ4 p4
ρφ + pφ = Γ4 pφ (10.65)
4 4 4
1 3 2 2
a3γ4 /2 = z = γ ρ κ (t − t0 )2 − ρ40 . (10.70)
ρφ 16 4 φ40 4
40
the first scale factor, (10.69), gives rise to the slow roll power law inflationary
solution, while the last one, (10.70), yields the proper (3+1) Barrow and Saich
(1993) solution; see below.
For the general case γ4 = 2Γ4 , the integral of (10.68), is given in terms of
hypergeometric functions by
2 a 1 γ4 3γ4 − 2Γ4
t − t0 = a3γ4 /2
F , , ,
3 γ4 κ4 /3ρ40 2 2γ4 − 2Γ4 2γ4 − 2Γ4
ρφ
− 40 a3(γ4 −Γ4 ) . (10.71)
ρ40
Adding (10.5) and (10.6) one gets (dφ4 /dt)2 , which can be written as
2
dφ4
ȧ2 = Γ4 ρφ . (10.72)
da 4
Substituting above ȧ2 from (10.67), and the expression of ρφ from (10.66), one
4
obtains
Γ4 2
a 2 (γ4 −Γ4 )
3
φ4 − φ40 = a φ ln ρφ
κ4 /3 3(γ4 − Γ4 ) 40
+ ρ40 + ρφ a3(γ4 −Γ4 ) ,
40
(10.73)
√
κ /3
hence, introducing C4 = a φ √ 4 32 (γ4 − Γ4 ), the expression of a in terms of φ
Γ4
amounts to
1
a 2 (γ4 −Γ4 ) = eC4 (φ4 −φ40 ) − ρ40 e−C4 (φ4 −φ40 ) .
3
(10.74)
2 ρφ
40
−2 γ
Γ4
−Γ
2 − Γ4 1 4 4
1 Γ3
φ3 − φ30 = a φ √ ln ρφ aγ3 −Γ3
κ3 γ3 − Γ3 30
+ ρ0 + ρφ a2(γ3 −Γ3 ) , (10.78)
30
(10.79)
√
κ3
where it as been introduced the constantC3 = a φ √ (γ3 − Γ3 ). Incidentally,
Γ3
the scale factor a in terms of φ3 amounts to
1
a(γ3 −Γ3 ) = eC3 (φ3 −φ30 ) − ρ0 e−C3 (φ3 −φ30 ) . (10.80)
2 ρφ
30
1 γ32 2
aγ3
=z= ρ κ (t − t0 )2 − ρ30 , (10.82)
ρφ 4 φ30 3
30
the first scale factor corresponds to the (2+1) power law solution, and the last
scale factor gives rise to the (2+1) Barrow–Saich solution.
134 Dilaton–Inflaton Friedmann–Robertson–Walker Cosmologies
2 2 3γ4 κ4
φ4 (t) − φ40 = √ ln A (t − t0 )
3 γ4 κ4 4 φ4 3
3γ42
+ κ A2 (t − t0 )2 − A4 , (10.84)
16 4 φ4
√
e 3κ4 γ4 /2(φ4 −φ40 )
V (φ4 ) = (4 − γ4 )Aφ4 √ 2
2 . (10.85)
e 3κ4 γ4 /2(φ4 −φ40 ) − A4
Using now the correspondence (10.19) and the relation (10.31), accompanied by
the changes ρ4 → ρ3 , and ρφ → ρφ , one obtains the following (2+1) spatially
40 30
flat FRW Barrow–Saich counterpart:
κ 1/γ3
3
4 γ32 A2φ3 (t − t0 )2 − A3
a(t) = , (10.86)
Aφ3
2 γ √
φ3 (t) − φ30 = ln 3 Aφ3 κ3 (t − t0 )
γ3 κ3 2
γ32
+ κ A2 (t − t0 )2 − A3 , (10.87)
4 3 φ3
√
e 2κ3 γ3 (φ3 −φ30 )
V (φ3 ) = (4 − γ3 )Aφ3 √ 2
2 . (10.88)
e 2κ3 γ3 (φ3 −φ30 ) − A3
On the other hand, substituting the above expression into (10.91), one obtains
a(t), namely
(1−β)
1 2 (1−2β)
a(t) = a0 exp − ρφ − 3a Γ4 κ4 /3(1/2 − β) t . (10.95)
3Γ4 (1 − β) 40
√
The equation to determine φ4 (t), φ̇ = φ pφ + ρφ , amounts to
dφ4 β/2
= φ Γ4 ρφ , (10.96)
dt 4
136 Dilaton–Inflaton Friedmann–Robertson–Walker Cosmologies
consequently
1
φ4 − φ40 = −2φ a ρφ (1−β)/2 . (10.98)
3Γ4 κ4 (1 − β) 4
Finally, the evaluation of V (φ4 ) gives
2/(1−β)
1 2 1
V (φ4 ) = ρφ − φ˙4 = −a φ 3Γ4 κ4 (1 − β)(φ4 − φ40 )
4 2 2
2β/(1−β)
Γ4 1
− −a φ 3Γ4 κ4 (1 − β)(φ4 − φ40 ) . (10.99)
2 2
Recall that these families of solutions arise for β = 1/2. The case β = 1/2 gives
rise to the (3+1) Madsen cosmology.
thus, substituting this last expression into the scale factor a one has alternatively
(1−β)
1 √ 2 (1−2β)
a(t) = a0 exp − ρ − 2a Γ3 κ3 (1/2 − β) t . (10.102)
2Γ3 (1 − β) φ30
It should be mentioned that all these solutions are determined under the
condition β = 1/2 and β = 1. The case β = 1/2 yields the (2+1)
Barrow–Burd–Lancaster solution, which is treated in detail below.
pφ = Γ3 ρ1/2
φ
− ρφ , (10.105)
3 3 3
d 2
ρ + Γ3 ρ1/2 = 0. (10.106)
da φ3 a φ
3
φ̇3 = φ Γ3 ρ1/4
φ
, (10.112)
3
138 Dilaton–Inflaton Friedmann–Robertson–Walker Cosmologies
Γ2 κ 1
V3 = 3 3 κ (φ − φ30 ) − (φ3 − φ30 ) .
4 2
(10.115)
8 2 3 3
To identify the derived solution with the original Barrow–Burd–Lancaster ex-
pressions – see Barrow et al. (1986) – one has to accomplish the following changes:
1
Γ3 −→ 8μ/κ3 , ρφ −→ μκ A4 , −a −→ a = ±1, (10.116)
30 2 3
in this manner one obtains:
κ3 2 a √8μt
a = a0 exp (− A e ),
4
1 √
ρφ (t) = μκ3 A4 e2a 8μ t ,
3 2 √
φ3 − φ30 = φ a Aea 8μ t/2 ,
1
V3 = μ κ3 (φ3 − φ30 )4 − (φ3 − φ30 )2 , (10.117)
2
where a = ±1, and φ30 is an integration constant.
1 ȧ2 1 (A + 4t3 )2
ρφ = = . (10.126)
3 κ3 a2 4κ3 t2 (A + t3 )2
2
Further, since (10.18) reads ρ̇φ + 2(pφ + ρφ )ȧ/a = 0, and φ˙3 = pφ + ρφ ,
3 3 3 3 3
hence the equation for φ3 to be integrated is
ȧ 2
ρφ˙ + 2 φ˙3 = 0. (10.127)
3 a
Substituting the derivative of ρφ from (10.126), one obtains
3
(2t3 − A)
φ˙3 = √ , (10.128)
2κ3 t(A + t3 )
which has the following integral
1 A + t3
φ3 = √ ln C0 . (10.129)
2κ3 t
2
The evaluation of V (φ) = ρφ − φ˙3 /2 yields
3
3 t 3 √
V (φ3 ) = 3
= C0 e− 2κ3 φ3 . (10.130)
κ3 A + t κ3
Summarizing, the second (2+1) inflationary BBL solution is determined by
2 A
a(t) = t 1 + 3 , (10.131)
t
1 A + t3
φ3 = √ ln C0 , (10.132)
2κ3 t
3 t 3 √
− φ3 2κ3
V (φ3 ) = = C0 e , (10.133)
κ3 A + t3 κ3
where A and C0 are constants.
To get an insight into the form of the conventional state equation, i.e., on the
dependence of F (ρ) on ρ, one expresses t in terms of ρ := 4κ3 ρφ by solving
3
(10.126) with respect to t, which yields
12
√ 1 16 − 2 ρ3/2
A
t ρ=1+ 4 + ρ1/2 A1/3 Δ + t
8 −ρ A Δ+
1/2 1/3
,
2 2 4 + ρ1/2 A1/3 Δ
(10.134)
1/3
where Δ := Aρ3/2 − 16 . On the other hand, since
2 1
φ˙3 = pφ + ρφ = Γ F (ρφ ) = (2t3 − A)2 /(t2 (A + t3 )2 ),
3 3 3 2κ3
substituting t from (10.134), one obtains a very involved function F on ρ.
10.5 Scalar Field Solutions for a Given Scale Factor 141
1 40t6 + 32At3 + A2
V (φ4 ) = , (10.137)
4κ4 t2 (t3 + A)2
The expression of V (φ4 ) in terms of φ4 , which is very involved, can be achieved
by substituting the roots of t in terms of φ4 from (10.136) into (10.137).
11
Einstein–Maxwell Solutions
The purpose of this chapter is to provide a new approach on the search of electro-
magnetic–gravitational solutions to the Einstein–Maxwell fields of the (2 + 1)
gravity in the presence of a cosmological constant, allowing for stationary and
cyclic symmetries, establishing their relationship with known current solutions,
and to point out the families allowing for black hole interpretation. The search
and interpretation of this kind of solution has been the goal and realm of several
authors’ investigations, starting from quite different perspectives and using a
variety of approaches, which have sometimes brought about duplication of results
and efforts. Consequently, the completeness of the electromagnetic classes of
stationary cyclic symmetric solutions under consideration will be demonstrated
via straightforward integration of the field equations. A full characterization
of the physical content of these solutions would require considerable work; for
this reason, some short related comments are made close to those contained in
the relevant references, if there are any, and also about newly discovered families
with special emphasis on their black hole feature. Nevertheless, a full geometrical
characterization based on the algebraic classification of the physical tensors has
been produced.
From a general metric for stationary cyclic symmetric gravitational fields cou-
pled to Maxwell electromagnetic fields within the (2 + 1)-dimensional gravity
the uniqueness of wide families of exact solutions is established, including all
uniform electromagnetic solutions possessing electromagnetic fields with vanish-
ing covariant derivatives, all fields having constant electromagnetic invariants
Fμν F μν and Tμν T μν , the whole classes of hybrid electromagnetic solutions, and
also wide classes of stationary solutions, derived for third-order nonlinear key
equations. Certain of these families can be thought of as black hole solutions.
For the most general set of Einstein–Maxwell equations, reducible to three non-
linear equations for the three unknown functions, two new classes of solutions
– having anti-de Sitter spinning metric limits – are derived. The relationship
of various families with those reported by different authors’ solutions has been
11.1 Stationary Cyclic Symmetric Einstein–Maxwell Fields 143
established. Among the classes of solutions with cosmological constant are: the
electrostatic Peldan solution, the magnetostatic Peldan metric, the stationary
uniform and spinning Clément classes, the constant electromagnetic invariant
branches with the particular Kamata–Koikawa solution, the Ayón–Cataldo–
Garcı́a hybrid cyclic symmetric stationary black hole fields, and the no less
important solutions generated via SL(2, R) transformations where the Clément
spinning charged solution, the Martı́nez–Teitelboim–Zanelli black hole solution,
and Dias–Lemos metric merit mention.
k ∧ m ∧ dk = 0 = k ∧ m ∧ dm (11.1)
are identically fulfilled because of their 4-form character and hence there exists
the discrete symmetry when simultaneously t → −t and φ → −φ. One may find
a coordinate system such that the metric tensor components g(k dr) = 0 and
g(m dr) = 0, where the coordinate direction dr is orthogonal to the surface
spanned by k ∧ m. Commonly one introduces the coordinate system {t, φ, r} in
(2+1)-dimensional gravity .
The main goal of this section is to demonstrate of the following theorem:
where the constants a, b and c are subjected, by virtue of the Ricci circularity
conditions, to the equations
a c = 0 = b c, (11.3)
144 Einstein–Maxwell Solutions
Proof To establish that the field ∗F possesses the form given by (11.2) one uses
the source-free Maxwell equations
dF = 0 = d ∗ F , (11.6)
where ∗ denotes the Hodge star operation. Let us evaluate the exterior derivative
of the t-component ∗F (k) of ∗F ,
the first zero arises from the stationary character of the field F , while the second
one corresponds to the Maxwell equation. Similarly, for the φ–component ∗F (m)
one has
In this manner we have established that the t and φ components of the dual field
∗F are constants given correspondingly by a and b. The component of ∗F along
the vector direction ∂r remains to be determined. For this purpose, consider the
tφ-component F (k, m) of the field F , which can be expressed as F (k, m) =
im ik F = (−im ik ∗ ∗F = im ∗ (k ∧ ∗F ) = ∗(m ∧ k ∧ ∗F )) = − ∗ F (∗(k ∧ m)),
thus its derivative yields
Since the constant c can be written as c = − ∗ F (∗(k ∧ m)), to determine it, one
evaluates ∗(k ∧ m). Identifying the Killing vectors accordingly with k = ∂t and
m = ∂φ , then
√ √
∗ (k ∧ m) = − −gdr = − −g g rr ∂r , (11.10)
thus
√ √
c = − ∗ F (− −g g rr ∂r ) = −g g rr ∗ F (∂r ). (11.11)
11.1 Stationary Cyclic Symmetric Einstein–Maxwell Fields 145
Conversely, from the above mentioned relation one determines the r component
of the field ∗F , namely ∗F (∂r ) = √−g
c
grr . In this manner, the structure of F ,
explicitly given by (11.2), has been established.
The vanishing conditions (11.3) straightforwardly arise from the Ricci circu-
larity conditions m ∧ k ∧ R(k) = 0 and k ∧ m ∧ R(m) = 0. Correspondingly,
the vanishing conditions a c = 0 = b c can be established immediately, as we
shall see in the next section, from the Einstein equations Rμν − 12 R gμν =
κTμν − Λ gμν , where the electromagnetic energy–momentum tensor compo-
nents are defined through the electromagnetic field Fμν = −Fνμ as 4π Tμν =
Fμσ Fν σ − 14 gμν Fασ F ασ .
F (r) 2 dr2 2
g=− dt + + H(r) [dφ + W (r)dt] . (11.15)
H(r) F (r)
146 Einstein–Maxwell Solutions
F (r) 2 dr2
g=− [dt − ω(r) dφ] + h(r) dφ2 + ,
h(r) F (r)
F H ω HF
F = F, H = h − ω 2 , W = ,h = . (11.16)
h F h F − W 2 H2
Mostly we will use the metric (11.15) in the forthcoming developments, but
occasionally the metric representation (11.16) will be used . When doing so,
the derived expressions will be given in terms of the set {F (r), h(r), ω(r)} of
structural functions. Omitting the dependence of the structural functions on the
variable r, the Maxwell electromagnetic field contravariant tensor is given by
⎡ ⎤
0 b − Fc
⎢ ⎥
(F μν ) = ⎢
⎣ −b 0 a ⎥,
⎦ (11.17)
c
F −a 0
where a, b, and c are constants related with the character of the field. For
instance, if only b is different from zero, while a and c vanish, the field is called
(pure) electric field. When a = 0, b = 0 = c, one deals with a pure magnetic field;
other possibilities do not receive a particular name. The covariant components
Fμν of the field tensor are given by:
The energy–momentum tensor matrix (Tνμ ) associated with the metric (11.15) is
⎡ 2 2 2 ⎤
b (H W −F )−a2 H 2 −c2 H a[b(H 2 W 2 −F )−aH 2 W ]
ac
−
⎢ 8π F H 4π 4π F H ⎥
⎢ ⎥
⎢ cH(W b−a) −b F +H (W b−a) +c H
2 2 2 2 c[b(H W 2 −F )−aH 2 W ]
2
⎥
⎢ − ⎥
⎣ 4 πF 2 8π F H 4π F 2 H ⎦
bH(W b−a) b (H W −F )−a H +c H
2 2 2 2 2 2
4 πF
bc
4π − 8π F H
(11.22)
and possesses the trace T := Tμμ given by
2
1 c2 1 H (a − W b) 1 b2 1
T =− + − = F F, (11.23)
8π F 8π F 8π H 16π
and the electromagnetic energy momentum quadratic invariant
2
μν3 H 2 (a − b W )2 − b2 F − c2 H 3
T T = Tμν T = = F F 2 . (11.24)
64π 2 F 2H 2 256π 2
The Einstein–Maxwell equations
R
Eμν := Rμν − gμν + Λgμν − 8π Tμν = 0 (11.25)
2
for a negative cosmological constant Λ = −1/l2 explicitly read:
1 H,r,r 1 H,r F,r 1 F 1
Et t = F + − 2
H,r 2 + H 2 W W,r,r + HW W,r H,r
2 H 4 H 4H 2
1 2 F − H 2W 2 c2 a2 H 1
+ H W,r 2 + b2 + + − 2, (11.26a)
4 FH F F l
Et r = −2 ca, (11.26b)
1 H,r,r H,r F,r HW
Et Φ = W F,r,r − F W −W − 2 a2
2 H H F
F −H W 2 2
1 H,r
− 2ab − 2
F +H W 2
W,r,r + 2W,r
FH 2 H
H,r 2
+F W − H 2 W W,r 2 , (11.26c)
H2
H
Er t = −2c (W b − a) , (11.26d)
F2
1 H,r F,r 1 H,r 2 1 b2
Er r = − F 2 + H 2 W,r 2 +
4 H 4 H 4 H
c2 H 1
− − (bW − a)2 − 2 , (11.26e)
F F l
F − H 2W 2 HW
Er Φ = −2 c b 2
− 2ca 2 , (11.26f)
F H F
1 2 H
EΦ t = H W,r,r + HW,r H,r − 2 b (W b − a) , (11.26g)
2 F
148 Einstein–Maxwell Solutions
EΦ r = −2 bc, (11.26h)
one arrives at
F dr2
gc = − HW 2 dΦ2 + − Hd T 2 + 2H W d T dΦ, (11.28)
H F
which can be brought to the form
F 2 dr2 2
gc = − dT + + H (dΦ + W dT ) , (11.29)
H F
accompanied by the identification
F HW
F = F, H = − HW 2 , W = . (11.30)
H H
At the level of the field tensor F μν one has
⎡ ⎤ ⎡ ⎤
0 B − FC 0 −i a c
F
⎢ ⎥ ⎢ ⎥
(F μν ) = ⎢
⎣ −B 0 A ⎥ ⎢
⎦ = ⎣ ia 0 ib ⎥
⎦, (11.31)
C
F −A 0 − Fc −ib 0
thus the following correspondence for the field constants arises
−i a → B, i b → A, −c → C. (11.32)
Summarizing, one may say that the role of the Killingian coordinates has been
interchanged: the timelike coordinate t becomes the spacelike Φ–coordinate,
11.1 Stationary Cyclic Symmetric Einstein–Maxwell Fields 149
covariant derivatives (stationary uniform and spinning Clément classes), all fields
having constant electromagnetic field and energy–momentum tensors’ invariants
(Kamata–Koikawa solution), the whole classes of hybrid electromagnetic Ayon–
Cataldo–Garcı́a solutions, a new family of stationary electromagnetic solutions,
the electrostatic and magnetostatic solutions with Peldan limit, the Clément
spinning charged metric, the Martı́nez–Teitelboim–Zanelli black hole solution,
and Dias–Lemos electromagnetic solution.
The application of the Hayward black hole dynamics formulation, Hayward
(2008), and the Ashtekar isolated horizon approach, Ashtekar and Krishnan
(2004), to the static and stationary black hole solutions reported here is
straightforward.
Determining the eigenvalues and eigenvectors of the Cotton matrix (11.34) one
establishes the algebraic Cotton type of the spacetime one is dealing with.
Accordingly, the characteristic equation for the eigenvalue λ amounts to
(C 2 2 − λ) (C 1 1 − λ)(C 3 3 − λ) − C 3 1 C 1 3 = 0, (11.35)
11.1 Stationary Cyclic Symmetric Einstein–Maxwell Fields 151
class. For the static metric W (r) = 0, consequently the metric (11.15) becomes
F (r) 2 dr2
g=− dt + + H(r)dφ2 , (11.41)
H(r) F (r)
and the Einstein–Maxwell equations simplify drastically:
1 H,r,r 1 H,r F,r 1 H,r 2 b2 c2 a2 H 1
Et t = F + − F 2 + + + − 2,
2 H 4 H 4 H H F F l
1 H,r F,r 1 H,r 2 b2 c2 a2 H 1
Er r = − F 2 + − − − 2,
4 H 4 H H F F l
1 1 H,r,r 3 H,r F,r 3 H,r 2 b2 c2 a2 H 1
EΦ Φ = F,r,r − F − + F − + − − 2,
2 2 H 4 H 4 H2 H F F l
1
Et r = −2 ca, Et Φ = −2a b ,
H
H 1
Er t = 2a c 2 , Er Φ = −2b c ,
F FH
H
EΦ t = 2a b , EΦ r = −2 bc, (11.42)
F
Each of these Eμ ν equations has to be equated to zero, therefore one can distin-
guish the following three families of static solutions:
the electric class: b = 0, a = 0, c = 0,
the magnetic class: a = 0, b = 0, c = 0,
the hybrid class: c = 0, a = 0, b = 0.
In the next sections we proceed to integrate each class separately.
d2
H (r) = 0 ⇒ H(r) = C0 + C1 r, (11.43)
dr2
where C0 , and C1 are constants of integration; C1 at this stage is assumed to be
different from zero; the zero case deserves special attention and will be treated
separately. Substituting this structural function H into the equation Et t (a =
0 = c) one arrives at a first-order differential equation for F
2
H,r H,r b2 4
F,r − F + 4 − 2 = 0, (11.44)
H H H l
which by introducing an auxiliary function f (r) through
F (r) 2 dr2
g=− dt + + h(r)dφ2 ,
h(r) F (r)
4
F (r) = 2 2 K0 + h(r) − b2 l2 ln h(r) h(r),
C1 l
h(r) = C1 r + C0 . (11.47)
Two years later, Deser and Mazur (1985) published their version of the elec-
trostatic solution for Λ = 0. Moreover, by then, the work by Melvin (1986) was
published with the derivation of the electrostatic as well as the magnetostatic
solutions for vanishing Λ.
Melvin (1986) introduced three kinds of coordinate systems, and defined K as
the electric charge, while κ = 8π G , M as the total mass, and the scale parameter
a = (1 − 4G M )2 /(2 K), among other quantities:
physical radial coordinates, in which the electrostatic solution amounts to
dS
ds2 = −N (ρ)2 dt2 + dρ2 + S(ρ)2 dθ2 , dρ = ,
(1 − 4G M )2 − 2 K ln S
N= (1 − 4G M )2 − 2 K ln S, (11.55a)
There is a sign missprint in the function grr of Deser and Mazur (1985) , Eq. (12);
in the factor in front of (ln r)2 should be −G e2 (ln r)2 .
Kogan (1992) reported and analyzed the (electro- and magneto-) static solu-
tions of the (2+1)-dimensional Einstein–Maxwell equations for both positive and
negative signs of the gravitational constant κ; recall that in the three dimensions
there is no restriction on its sign. The r-coordinate used there was such that
grr = 1 for the signature used in the present report, consequently, the solutions
are given in the (11.55a) representation.
dρ2 ρ2
g = −F dt2 + + ρ2 dφ2 , F (ρ) = 2 − m − 2b2 ln ρ,
F l
A = b ln ρ dt. (11.56)
Comparing with the static BTZ one recognizes M as the BTZ mass. Notice that
the energy and mass include an amount of energy due to the electric field through
the logarithmical term; because of this dependence, these quantities diverge at
infinity logarithmically.
The spacelike vector V3 is aligned along the circular Killing direction ∂φ . Thus
one may have the spacetime arrangement {T1, S2, S3}, or {N1, N2, S3}, and so
on.
The Cotton tensor for electrostatic cyclic symmetric gravitational field is
given by
⎡ ⎤
b2
0 0 2ρ 2
⎢ ⎥
⎢ ⎥
(C α β ) = ⎢ 0 0 0 ⎥. (11.67)
⎣ ⎦
b2
− 2ρ 4L
2
0 0
The eigenvectors V2 and V3 are complex conjugated, or, if one wishes, one
may consider the component V 1 differently for each of the complex vectors. For
the zero eigenvalue λ1 , the vector V1 is a spacelike vector that it points along
the ρ–coordinate direction.
It is worth pointing out that the field and Cotton tensors of the solutions
generated via coordinate transformations, in particular SL(2, R) transforma-
tions, applied onto this electrostatic cyclic symmetric Peldan solution will shear
the eigenvalues λi of the corresponding field and Cotton tensors of the charged
Peldan solution; recall that eigenvalues are invariant characteristics of tensors,
although the components of the eigenvectors, in general, look different in differ-
ent coordinate systems; this remark also applies to the (eigenvalues) eigenvectors
of the seed and resulting solutions.
The subclass of solutions without Λ are explicitly given and identified, while
the magnetostatic Peldan and Hirschmann–Welch solution representation with
Λ are analyzed in detail.
where (11.70) it has been used to evaluate F (r). These structural functions
completely determine the magnetostatic solution; without any loss of generality,
by letting
dr2
g = −h(r)dt2 + + H(r)dφ2 ,
H(r)h(r)
4
H(r) = K0 + h(r) + a2 l2 ln h(r) ,
C1 2 l2
F (r) = H(r) h(r), h(r) := C1 r + C0 . (11.76)
dρ2
g = −ρ2 dt2 + + F (ρ)dφ2 , F (ρ) = k0 + 2a2 ln ρ, (11.82)
F (ρ)
162 Einstein–Maxwell Solutions
2
or by introducing a4 r2 ek0 /a = k0 + a2 ln ρ2 , and scaling the variables t and φ
one brings it to the form
2
g = er (−dt2 + dr2 ) + r2 dφ2 . (11.83)
In the paragraph devoted to stiff perfect fluid, Barrow, Burd, and Lancaster
– see Barrow et al. (1986) – pointed out that for a fluid aligned along the
time-coordinate, “in (2 + 1) dimensions the stiff fluid has an energy–momentum
tensor identical to that of a static magnetic field,” and they continued with a
statement very close to the following: if one sets the electric field components
√
F0i = 0 and magnetic components Fi j = i j 2μ in the electromagnetic energy–
momentum tensor Tμν = Fμλ Fν λ − gμν Fαλ F λα /4 reduces to the perfect fluid
energy–momentum tensor Tμν = (μ + p)uμ uν + p gμν , with energy density μ
equalling the pressure p, μ = p.
dρ2 ρ2
g = −ρ2 dt2 + + F (ρ)dφ2 , F (ρ) = k0 + 2 + 2a2 ln ρ,
F (ρ) l
A = a ln ρ dφ, (11.84)
2a 2 l a
11.3 Magnetostatic Solutions; a = 0, b = 0 163
a way similar to the electric one, through logarithmical terms; because of this
dependence, these quantities logarithmically diverge at infinity.
1 a2
λ1 = − ; T1 = [V 1 , 0, 0], V μ Vμ = −ρ2 (V 1 )2 ,
8π ρ2
1 a2
λ2 = ; S2 = [0, V 2 , V 3 ], V μ Vμ = (V 2 )2 / L2 + (V 3 )2 L2 ,
8π ρ2
1 a2
λ3 = ; S3 = [0, Ṽ 2 , Ṽ 3 ], V μ Vμ = (Ṽ 2 )2 / L2 + (Ṽ 3 )2 L2 ,
8π ρ2
Type: {T, 2S}. (11.94)
h(ρ) = (ρ2 + r+
2
− ml2 )/l2 , (11.98)
with vector potential
1
A= χ ln |(ρ2 + r+
2
)/l2 − m|dΦ, (11.99)
2
166 Einstein–Maxwell Solutions
This solution is endowed with mass, magnetic charge, and radial parameters. The
coordinate ρ ranges from zero to infinity. This magnetic solution does not allow
the existence of an event horizon since timelike geodesics can reach the origin
at finite proper time, while null geodesics approach the origin at finite affine
parameter; hence it does not describe a magnetic black hole. Moreover the Ricci
tensor, and consequently the curvature tensor, as well as the electromagnetic
field, are well behaved in this spacetime.
Cataldo et al. (2004) commented on this static circular magnetic solution of
the (2+1) Einstein–Maxwell equations, derived previously by other authors, and
came to the conclusion that this solution, considered up to that moment as a two-
parameter one, is in fact a one-parameter solution, which describes a distribution
of a radial magnetic field in a (2+1) anti-de Sitter background spacetime, and
that the mass parameter is just a pure gauge and can be rescaled to minus one.
Accomplishing in the original Peldan solution (11.86) the coordinate transfor-
mation
t → t, ρ → (ρ2 + r+ 2 − m l2 )/l2 , φ → φ l2 , χ2 := a2 l2 ,
one obtains the Hirschmann and Welch (1996) representation of the magnetos-
tatic solution, which is given by the metric functions
1
ds2 = −N (ρ)2 dt2 + dρ2 + K(ρ)2 [dφ + W (ρ)dt]2 ,
L(ρ)2
ρ2 + r+
2
− m l2 H(ρ)
H(ρ) = , L(ρ) = K(ρ),
l2 ρ
K(ρ) = ρ2 + r+ 2 + χ2 ln H(ρ),
is given by
2
r+ − ml2 + χ2 χ2 2 2
(K 2 /ρ2 )|ρ→0 → = 1 + 2 e(r+ /χ ) ,
r+ − ml
2 2 l
11.3 Magnetostatic Solutions; a = 0, b = 0 167
hence the spatial sector ( L12 dρ2 + K 2 dφ2 )|ρ→0 of the studied metric behaves as
⎡ ⎤2
2 2 2 2 2
⎣d ρ e(r+ /2χ )
⎦ + ρ2 e(r+ /χ )
2
(−r+ /2χ2 ) χ2 (r+2 /χ2 )
χ2 2 2)
dφe (1 + 2 e )
1+ χ2 2
e(r+ /χ
2)
1+ l2 e(r+ /χ l
l2
1 ρ2 + r+
2
− ml2 + χ2
(ρ, 0 ) = − − 0 , (11.103)
π l ρ2 + r2 + χ2 ln H ρ2 + r2 − l2 m
+ +
2 ρ2 + r+
2
− ml2 + χ2
E(ρ, 0 ) = − − 2π K 0 ,
l 2 − l2 m
ρ2 + r+
2 2
M (ρ, 0 ) = − 2
(ρ + r+ − ml2 + χ2 ) − 2π N K0 . (11.104)
l2
The evaluation of the above functions independent of 0 behave at infinity
according to
1 ml2 − 2χ2 χ2 ρ
(ρ → ∞, 0 = 0) ≈ − + 2
+ 2
ln ( ),
πl 2π l ρ πlρ l
ρ m l2 − r+2
− 2χ2
E(ρ → ∞, 0 = 0) ≈ −2 + ,
l lρ
2
M (ρ → ∞, 0 = 0) = − 2 (ρ2 + r+
2
− ml2 + χ2 ). (11.105)
l
Using in the expressions (11.104) the energy
density for the anti-de Sitter solu-
2
tion counterpart, namely 0 = − π l2 ρ/ M0 + ρl2 , which at the spatial infinity
1
l M0 ml2 − 2χ2 χ2 ρ
(ρ → ∞, 0|∞ (M0 )) ≈ − 2
+ 2
+ ln ( ),
2π ρ 2π lρ π l ρ2 l
l M0 ml − 2χ
2 2
χ 2
ρ
E(ρ → ∞, 0|∞ (M0 )) ≈ − + + 2 ln ( ),
ρ lρ lρ l
χ2 χ2 ρ
M (ρ → ∞, 0|∞ (M0 )) ≈ m − M0 − 2 2 + 2 2 ln ( ). (11.106)
l l l
Therefore, comparing with the energy characteristics of the BTZ solution, one
concludes that the mass logarithmically diverges at spatial infinity, and that the
role of mass is played by m.
Some data about the electro-magneto–static families of solutions can be
represented schematically in the table above.
dropping primes, one brings the static hybrid metric to the form
√
(1±√α)/2 √
2 (1∓√α)/2
ρ2 ρ
g = −ρ1∓ α
−M 2
dt + ρ1± α
−M dφ2
l2 l2
2 −1
ρ
+ 2 −M dρ2 . (11.115)
l
The electromagnetic field tensor under the above mentioned transformations
becomes
√
Fμν = M 1 − αδ[μ t δν] φ ,
M2 (1 − α)
Tν μ = × −δν t δt μ + δν r δr μ − δν φ δφ μ . (11.116)
32π ρ (ρ /l2 − M )
2 2
V μ Vμ = 0, V3 = N3,
Type : {S, N, N }. (11.123)
0 0 1
2 2
1 M l (1−α)
allows for the eigenvalues λ1 = 32 ρ2 (ρ2 −M l2 )π and the other one, of multiplicity
2 2
1 M l (1−α)
two, λ2 = λ3 = − 32 ρ2 (ρ2 −M l2 )π with the corresponding eigenvectors
1 M 2 l2 (1 − α)
λ1 = ;
32 ρ2 (ρ2 − M l2 ) π
V1 = [0, V 2 , 0], V μ Vμ = (V 2 )2 gρρ , V1 = S1,
1 M 2 l2 (1 − α)
λ2 = − ;
32 ρ2 (ρ2 − M l2 ) π
V2 = [V 1 , 0, V 3 ], V μ Vμ = (V 1 )2 gt t + (V 3 )2 gφ φ ,
V2 = T2, S2, N2,
1 M 2 l2 (1 − α)
λ3 = − ;
32 ρ2 (ρ2 − M l2 ) π
V3 = [Ṽ 1 , 0, Ṽ 3 ], V μ Vμ = (Ṽ 1 )2 gt t + (Ṽ 3 )2 gφ φ ,
V3 = T3, S3, N3. (11.125)
For V2 and V3, the character of the vector depends on the sign of its norm; for
instance, by choosing
√
V 1 = s gφ φ / |gt t | V 3 , s = const., V1μ V1μ = (1 − s2 )gφ φ (V 3 )2 ;
s > 1 → V1 = T, s = ±1 → V1 = N, s < 1 → V1 = S.
Although the solution above has been derived for Λ = −1/l2 , the branch
corresponding to Λ = 1/l2 is achieved from the above expressions by changing
l2 → −l2 .
F 2 dr2
g=− (dt − ω0 d φ) + + H0 dφ2 ,
H0 F (r)
2r2
F (r) = 2 + c1 r + c0 ,
l
1
A= √ r (dt − ω0 d φ) , (11.145)
H0 l
π02 l2
In Clément’s parametrization one adopts H0 = 4m , with m = 1/(2 κ).
F μν = 2 b δ μ [t δ ν r] ,
Fμν = −2 b/h(r) δμ [t δν r] + 2 b ω(r)/h(r) δμ [φ δν r] .
d2 F 4 r2
− = 0, → F (r) = 2 + 4 c1 r + c0 ; (11.146)
d r2 l2 l2
consequently the metric and the field vector become
F (r) 2 dr2 r
g=− 2 2
dt + + b2 l2 dφ2 , A = 2 dt. (11.147)
b l F (r) bl
11.5 Uniform Electromagnetic Solutions Fμν ;σ = 0 177
Thus, for the natural choice of a vanishing reference energy density 0 = 0 all
the energy quantities vanish: = 0, M (ρ, 0) = 0 = E(ρ, 0). On the other hand,
if the reference
energy is the one corresponding to the anti-de Sitter spacetime,
ρ2
0 = − π1ρ l2 − M0 , the energies M (ρ, 0 ) and E(ρ, 0 ) will be again expressed
through 0 .
with eigenvectors
1 2
λ1 = ; V1 = 0, 0, V 3 , Vμ V μ = V 1 , V1 = S1,
8 π l2
1
λ2,3 = − 2
; V2, 3 = V 1 , V 2 , 0 ,
8 π
2 1l
N V − V 2 N 2V 1 + V 2
Vμ V = −
μ
N2
V2 = T2, S2, N2, V3 = T3, S3, N3, (11.155)
and therefore it allows for the types {S, 2T }, {S, 2N }, {S, 2S}.
As in the previous case, F (r) and H(r) fulfill the (11.163) for γ = −1/l2 and
correspondingly their integrals are given by
2 2
F (r) = r + c1 r + c0 , H(r) = b2 l2 − β 2 l2 F (r).
l2
There are no further constraints from the field equations. Consequently the final
result can be written as
b2 l2 − H(r) 2 dr2
g=− dt + l 2 2
β
l2 β 2 H(r) b2 l2 − H(r)
2
a 1 b2 l 2 − H
+H(r) dφ + ∓ 2 dt ,
b l bβ H
H(r) = −2β 2 r2 + C1 r + C0 ,
1 ± al2 β
A = −β r [dφ − 2 dt],
l bβ
which coincides with the uniform solution (11.139). Therefore we have deter-
mined a class of uniform constant electromagnetic invariant stationary solutions
for both non-vanishing constants a = 0 = b. Although the solution mentioned
above has been derived for Λ = −1/l2 , the branch with positive Λ = 1/l2 is
achieved from the above-mentioned expressions by changing l2 → −l2 .
Substituting h(r) = h0 into the Einstein equations one obtains from Eφ t that
d2 ω 2 dF dω
2
=− , (11.169)
dr F dr dr
which when used in Et t –Er r yields
F dω 2
( ) = 0 → ω(r) = ω0 . (11.170)
h20 dr
Replacing ω = ω0 and h = h0 in the remaining equations one establishes
d2 4 2
2
F (r) = 2 → F (r) = 2 r2 + c1 r + c0 , h0 = b2 l2 . (11.171)
dr l l
Therefore we arrive at a constant electromagnetic invariant solution in the form
F dr2
g=− (dt − ω0 dφ)2 + h0 dφ2 + ,
h0 F (r)
2r2
F (r) = 2 + c1 r + c0 , h0 = b2 l2 ,
l
r
A = 2 (dt − ω0 dφ), (11.172)
bl
which in all respects is identical to the uniform electromagnetic solution (11.143)
derived in the previous section. Notice that this solution exists only for negative
cosmological constant, Λ = −1/l2 ; there is no extension to Λ = 1/l2 . It is
evident that this solution can be generated from the static one, (11.147), via the
transformations t → t − ω0 φ, φ → φ.
For ω0 = 0, the above mentioned metric and field reduce to the Matyjasek–
Zaslavskii solutions – see Section (11.5.3) – thus this class of constant electro-
magnetic invariants’ static solutions is unique with the additional property of
being a uniform static solution.
H
F F = 2a2 → H(r) = β 2 F (r). (11.173)
F
184 Einstein–Maxwell Solutions
2 d2
(r − C) H + b2 l2 = 0 → H(r) = C0 + C1 r + b2 l2 ln (r − C) .
dr2
The gravitational and electromagnetic fields of this solution can be given as
11.6 Constant Electromagnetic Invariants’ Solutions 185
F 2 dr2
g=− dt + + H(dφ + W dt)2 ,
H F
2
(r − C)
H(r) = C0 + C1 r + b2 l2 ln (r − C) , F (r) = 4 ,
√ l2
a F l
W (r) = ± , A = ∓ ln (r − C) (adt + bdφ) . (11.180)
b H 2
This solution is characterized by the field tensor
l
Fμν = ± aδ[μ t δν] r + bδ[μ φ δν] r ,
(r − C)
F μν = 2bδt [μ δr ν] − 2aδφ [μ δr ν] , (11.181)
with energy–momentum tensor
l
Tμ ν = −a bδμ t δt ν + a bδμ φ δφ ν + a2 δμ t δφ ν − b2 δμ φ δt ν . (11.182)
4 π (r − C)
Notice that the three invariants Fμν F μν , Tμμ and Tμν T μν are equal to zero.
Without any loss of generality one can always set C = 0.
This solution corresponds to a possible representation of the Kamata and
Koikawa (1995) solution to be treated in detail in Section 11.6.5. It should be
pointed out that this solution does not belong to the family of uniform solutions,
i.e., the fields possessing vanishing covariant derivatives.
Next, one restores the factor π G in the above mentioned solution through the
identifications of the physical parameters:
√ √
ρ20 = 4π G Q2 /|Λ| = 1/2
J, b = 2 π GQ3/2 , a = ± 4 π G|Λ|1/2 Q1/2 ,
2|Λ|
It should be pointed out that Clément (1993) also reported a metric expression
and electromagnetic vector field describing a solution with vanishing electromag-
netic invariants. Comments concerning the mass content of this solution can be
found in Chan (1996). This solution is horizonless and consequently does not
permit a black hole interpretation.
The Kamata and Koikawa (1995, 1997) solution, see also Garcı́a (2009),
Eq. (7.7), is defined by the metric and the structural functions
1
ds2 = −N (ρ)2 dt2 + dρ2 + K(ρ)2 [dφ + W (ρ)dt]2 ,
L(ρ)2
√
Λ 2 √
L(ρ) = (ρ − ρ20 ), Λ = 1/l,
ρ
Q2 ρ2
K(ρ) = ρ2 + ln ( 2 − 1),
Λ ρ0
11.6 Constant Electromagnetic Invariants’ Solutions 187
√ Q2 ρ2
N (ρ) = ρ L/ K = Λ(ρ2 − ρ20 )/ ρ2 + ln ( 2 − 1),
Λ ρ0
√
(ρ2 − ρ20 ) Λ √
W (ρ) = Q2 2 − Λ, (11.187)
[ρ2 + Λ ln ( ρρ2 − 1)]
0
1 Λρ0 − Q + Q ln ( ρ20 − 1)
2 2 2 ρ
j(ρ) = √ 2 2
, (11.188)
π Λ ρ2 + QΛ ln ( ρρ2 − 1)
0
ρ2 √
M (ρ, 0 ) = −2Λ ρ2 + 2(2Λρ20 − Q2 ) + 2Q2 ln ( 2 − 1) − 2π Λ (ρ2 − ρ20 )0 .
ρ0
(11.189)
The evaluation of the functions above for vanishing 0 , i.e. 0 = 0, behave at
ρ → ∞ according to
√
Λ Λρ20 − Q2 Q2 ρ
(ρ → ∞, 0 = 0) ≈ − + √ +2 √ ln ( ),
π π Λρ 2 π Λρ 2 ρ0
Λρ20 − Q2 Q2 ρ
j(ρ → ∞) ≈ √ +2 √ ln ( ),
π Λρ π Λρ ρ0
Λρ − Q
2 2
Q 2
ρ
J(ρ → ∞) ≈ 2 0√ + 4 √ ln ( ),
Λ Λ ρ0
√ Λρ20 − Q2 Q2 ρ
E(ρ → ∞, 0 = 0) ≈ −2 Λρ + 2 √ + 2√ ln ( ),
Λρ Λρ ρ0
ρ
M (ρ → ∞, 0 = 0) ≈ −2Λ ρ2 + 2(2Λρ20 − Q2 ) + 4Q2 ln ( ). (11.190)
ρ0
Using in the expressions (11.189) as reference energy density the quantity
1 ρ2
0 = − −M0 + 2 ,
πρ l
188 Einstein–Maxwell Solutions
√
which at the spatial infinity behaves as 0|∞ (M0 ) ≈ − πΛ + 2π M
√0
Λ ρ2
, the series
expansions of the corresponding quantities at ρ → ∞ result in
M Λρ20 − Q2 Q2 ρ
(ρ → ∞, 0|∞ (M0 )) ≈ − √0 + √ +2 √ ln ( ),
2π Λ ρ 2 π Λρ 2 π Λρ 2 ρ0
M0 Λρ20 − Q2 Q2 ρ
E(ρ → ∞, 0|∞ (M0 )) ≈ − √ +2 √ + 4√ ln ( ),
Λρ Λρ Λρ ρ0
ρ
M (ρ → ∞, 0|∞ (M0 )) ≈ −M0 + 2(Λρ0 − Q ) + 4Q ln ( ).
2 2 2
(11.191)
ρ0
Therefore, comparing with the energy characteristics of the BTZ solution, one
arrives at the conclusion that there is no mass parameter of the kind M present in
the BTZ solution. All characteristic functions logarithmically diverge at spatial
infinity.
allows for a triple zero eigenvalue and the following set of eigenvectors
√
λ1,2,3 = 0; V = [V 1 , V 2 , Λ V 1 ], V μ Vμ = 0, V = N,
Type N : {3N }. (11.193)
It is clear that both the Cotton and Maxwell tensors possess the same eigenvalues,
namely the triple zero eigenvalue λ = 0. Searching for the eigenvectors of these
tensors, one arrives at
11.6 Constant Electromagnetic Invariants’ Solutions 189
√ (V 2 )2 ρ2
λ1,2,3 = 0; V = [V 1 , V 2 , Λ V 1 ], V μ Vμ = 2, V = S,
Λ (ρ − ρ0 )
V(V 2 = 0) = N,
Type : {S, 2N }. (11.196)
√
11.6.7 Proper Kamata–Koikawa Solution, ρ0 = ±Q/ Λ
The proper Kamata–Koikawa solution is defined by the √ metric and structural
functions of (11.187) for Λρ20 − Q2 = 0, i.e., ρ0 = ±Q/ Λ, namely
1
ds2 = −N (ρ)2 dt2 + dρ2 + K(ρ)2 [dφ + W (ρ)dt]2 ,
L(ρ)2
√
Λ 2
L(ρ) = (ρ − ρ20 ),
ρ
Q2 ρ2
K(ρ) = ρ2 + ln ( 2 − 1),
Λ ρ0
√ Q2 ρ2
N (ρ) = ρ L/ K = Λ(ρ − ρ0 )/ ρ2 +
2 2
ln ( 2 − 1),
Λ ρ0
√
(ρ2 − ρ20 ) Λ √
W (ρ) = 2 2 − Λ, (11.197)
[ρ2 + Λ ln ( ρ2 − 1)]
Q ρ
0
1 Q2 ln ( ρρ2 − 1)
j(ρ) = √ 0
, (11.198)
π Λ ρ2 + Q2 2
Λ ln ( ρρ2 − 1)
0
ρ2 √
M (ρ, 0 ) = −2Λ ρ2 + 2Λρ20 + 2Q2 ln ( 2 − 1) − 2π Λ (ρ20 − ρ20 )0 . (11.199)
ρ0
190 Einstein–Maxwell Solutions
M Q2 ρ
(ρ → ∞, 0|∞ (M0 )) ≈ − √0 +2 √ ln ( ),
2π Λ ρ2 π Λρ 2 ρ0
M0 Q2 ρ
E(ρ → ∞, 0|∞ (M0 )) ≈ − √ + 4√ ln ( ),
Λρ Λρ ρ0
ρ
M (ρ → ∞, 0|∞ (M0 )) ≈ −M0 + 4Q ln ( ).
2
(11.201)
ρ0
In the work by Chan (1996) there are some comments addressed to the evalua-
tion of the global momentum, energy and mass of the proper Kamata–Koikawa
solution: the exact and the approximated expressions of the momentum J coin-
cide with the corresponding ones given in (Chan, 1996, Eq. 9) and (Chan, 1996,
Eq. 7). Moreover, the mass M at spatial infinity, (11.200), coincides with the
M (Chan, 1996, Eq. 10.) for a zero background energy density with the cor-
rect extra term −2Λ ρ2 . From my point of view, it is to be recommended to
accomplish series expansions of the quantities under consideration to deter-
mine how fast they approach zero or diverge at spatial infinity. From this
perspective, the evaluation of the energy density and the global energy E
yield to quantities different from zero at spatial infinity, although they both
approach faster to zero as ρ → ∞ than the momentum and mass. Compar-
ing with the energy characteristics of the BTZ solution, one concludes that
the mass, energy and momentum functions logarithmically diverge at spatial
infinity.
11.7 Ayón–Cataldo–Garcia Stationary Hybrid Solution 191
with integral
± √α/2
r − r1
ln α0 h + α0 2 h2 + J 2 = ln k1 , (11.210)
r − r2
l 1 − K12 J 2
t= T, (11.214a)
4K1 r2 − r1
l 1 − K12 J 2 K1
φ = Φ − W0 + J T, (11.214b)
4K1 r2 − r1 l
1 K1
r= r1 − r2 K1 J − 2
2 2
(r2 − r1 ) ρ ,
2
(11.214c)
1 − K12 J 2 l
where with {T, ρ, Φ} are denoted the corresponding BTZ coordinates, which
ought to be accompanied with the identification
1 + K12 J 2
J 2 K1 = −R(−) := − M l − M 2 l2 − J 2 , M = − . (11.215)
2 l K1
ρ2 f (ρ) dρ2 2
g=− dT2 + + H(ρ) [dΦ + W (ρ)d T ] ,
H(ρ) f (ρ)
ρ2 J2
f (ρ) = − M + , R± := M l ± M 2 l2 − J 2 ,
l2 4ρ2
2ρ2 − lR− 2ρ2 − lR+ √ √
H(ρ) = √ √ −(2ρ2 − lR− ) α/2 (2ρ2 − lR+ )− α/2
4 αK1 M 2 l2 − J 2
√ √
+J 2 K12 (2ρ2 − lR− )− α/2 (2ρ2 − lR+ ) α/2 ,
R− 2 √ √
W (ρ) = (2ρ − lR+ ) α (2 α M 2 l2 − J 2 + R− )R−
Jl
√ √ √ −1
−(2ρ2 − lR− ) α J 2 × (2ρ2 − lR+ ) α R− 2
− (2ρ2 − lR− ) α J 2 .
(11.216)
1 ρ2 f 2 dρ2
g = −N 2 d t2 + dρ 2
+ K 2
[dφ + W d t]2
= − dt +
L2 H f
2 2
ρ J
+H(dφ + W d t)2 , f (ρ) = 2 − M + 2 ,
l 4ρ
2ρ − lR− 2ρ − lR+
2 2 √ √
H(ρ) = √ −(2ρ2 − lR− ) α/2 (2ρ2 − lR+ )− α/2
4K1 αSM
√ √
+J K1 (2ρ2 − lR− )− α/2 (2ρ2 − lR+ ) α/2 ,
2 2
R− 2 √ √
W (ρ) = − (2ρ − lR+ ) α (2 αSM + R− )R−
Jl
√ √ √ −1
−(2ρ2 − lR− ) α J 2 × (2ρ2 − lR+ ) α R− 2
− (2ρ2 − lR− ) α J 2 ,
R−
K1 : = − , SM := M 2 l2 − J 2 . (11.218)
J2
The structural functions appearing in the definitions of the energy and momen-
tum quantities are expressed as
ρ2 f (ρ)
N (ρ) = , L(ρ) = f (ρ), K(ρ) = H(ρ), W (ρ) = W (ρ). (11.219)
H(ρ)
1 f (ρ) d
(ρ, 0 ) = − H(ρ) − 0 , (11.222)
2π H(ρ) dρ
11.7 Ayón–Cataldo–Garcia Stationary Hybrid Solution 195
f (ρ) d
E(ρ, 0 ) = − H(ρ) − 2π0 H(ρ), (11.223)
H(ρ) dρ
f (ρ) d
M (ρ, 0 ) = −ρ H(ρ) − J W (ρ) − 2ρπ0 f (ρ), (11.224)
H(ρ) dρ
Because of the involved dependence of the metric functions upon the ρ coordi-
nate, the evaluation of the energy quantities will be done in the approximation
of the spatial infinity. The momentum density at infinity becomes
J α1/4
j(ρ → ∞) ≈ , (11.225)
2π ρ
while the global momentum remains constant in the whole space
J(ρ) = J. (11.226)
It becomes apparent then that the role of the momentum parameter is played
and coincides with J.
The approximated at ρ → ∞ surface energy density, global energy and mass,
for zero base energy density 0 are given by
√
1 lM α
(ρ → ∞, 0 = 0) ≈ − + ,
πl 2π ρ2
ρ l √
E(ρ → ∞, 0 = 0) ≈ −2 1/4 + 1/4
(1 + α)M,
lα 2 ρα
√
ρ2 α−1
M (ρ → ∞, 0 = 0) ≈ −2 2 + 2M + M 2 l2 − J 2 . (11.227)
l l
Using in the expressions (11.222)–(11.224) as reference energy density
the
ρ2
energy corresponding to the anti-de Sitter metric, 0 = − πρ l2 − M0 ,
1
(r − r1) (r − r2)
F (r) = 4 ,
l2 1/2 √α
l (r − r1) (r − r2) r − r1
H(r) = √
2 K1 α (r2 − r1) r − r2
√
−1/2 α
r − r1
−J 2 K12 ,
r − r2
−1/2 √α
r − r1 (r − r1) (r − r2)
W (r) = W0 − 2 K1 J . (11.230)
r − r2 l H(r)
Fμν = 2cδμ [t δν φ] ,
c2
Tμ ν = [−δμ t δt ν + δμ r δr ν − δμ φ δφ ν ]. (11.242)
8π F0
By means of scaling transformations F0 can always be set equal to unit, F0 = 1,
hence this solution is endowed with two effective parameters c and J.
d2 F H 1
EQF = − 4 a2 − 8 2 = 0,
d r2 F l
d2 H2
EQH = 2 H + 4a2 2 = 0,
dr F
1 dH dF F dH 2 J2 H 1
Er r = − 2
( ) + − a2 − 2 = 0. (11.245)
4H d r d r 4H d r 4 H2 F l
The equation Er r can be written in the form
2 2
1 dH 1 dF 1 dF a2 H 1 J2 1
− − 2
+ 2 − + 2 = 0.
2H dr 4F dr 16 F dr F 4 F H2 l F
(11.246)
200 Einstein–Maxwell Solutions
On the other hand, using EQF one expresses H in terms of F and its derivative
2
1 d F 8
H(r) = − 2 F. (11.247)
4 a2 dr2 l
Substituting the above H(r) into EQH (11.245) one gets
2 2
d4 F d3 F dF d F 24 d2 F 64
F + 2 + 2 − + 4 = 0. (11.248)
dr4 dr3 dr dr2 l2 dr2 l
Therefore, integrating, if possible, (11.248) for F (r), substituting the solution
F (r) into (11.247) one determines H(r). The resulting functions F (r) and H(r)
ought to fulfil the (11.246) or Er r equation from (11.245). By integrating the
linear first-order (11.244) one determines W (r).
The contravariant components of electromagnetic tensor are
F μν = −2 a δ μ [φ δ ν r] . (11.249)
The equation (11.248) for F (r) can be reduced to a third-order nonlinear equa-
tion. The problem for deriving solutions in this branch actually resides in this
equation.
Another possibility arises with the introduction of the auxiliary function h(r)
by means of
H(r) = F (r)1/2 h(r), (11.250)
the (11.246) acquires the form
2 2
dF dh
EQh = −l h 2 2 2 2
+ 4l F
dr dr
√
+16 l a h F − 4 l J + 16 h2 F = 0
2 2 3 2 2
(11.251)
and one could try to determine solutions for this variant.
right-hand side of the Einstein equations for the structural functions (11.259),
for ξ02 = 1, yields
π12
Gν μ = [−δνt δtμ + δνr δrμ + δνφ δφμ + 2 ω0 δνt δφμ ],
8mρ
while the right-hand side amounts to
κ π12
4π κ Tν μ = [−δνt δtμ + δνr δrμ + δνφ δφμ + 2 ω0 δνt δφμ ]
4ρ
hence 2m = 1/κ.
If one were adopting κ = 1, then modifying the electromagnetic vector A to
be Amod = − 2√π21 m ln( ρρ0 )(dφ + ω0 dt), one would arrive at the solution in our
convention.
It is apparent that these Clément’s solutions correspond to a variant of
the solution derived in the previous Section (11.8.1), with the identification
r → ρ followed by minor scaling transformations of t and φ. Notice also that
the above generalization with W (r) = ω0 = 0 of the magnetostatic solu-
tion (11.76) can be determined applying to it SL(2, R) transformations of the
form φ → φ + ω0 t, t → t.
d2 F H 1
EQF = + 4 b2 − 8 2 = 0,
d r2 F l
d2 H4
EQH = 2 H2 − 4b2 2 = 0,
dr F
1 d H d F F dH 2 1 H 1
Err = − ( ) + J 2 + b2 − 2 = 0,
4H d r d r 2
4H d r 4H 2 F l
d J
W = 2. (11.261)
dr H
11.8 Stationary Solutions for a = 0 or b = 0 203
Continuing with the parallelism, isolating H from EQF and replacing it into
EQH one obtains
2 2
d4 F d3 F dF d F 24 d2 F 64
F + 2 + 2 − + 4 = 0. (11.262)
dr4 dr3 dr dr2 l2 dr2 l
Thus, as before, the first step in the integration of the problem depends upon
the (11.262) for F(r), structurally identical to (11.248). Substituting the solution
F(r) into EQH from (11.261) one determines H. The resulting functions F(r)
and H have to fulfill Er r from (11.261). By integrating the linear first-order
equation for W one determines W(r).
1 dρ2
g = −U (dt − ω0 dφ)2 + + 2ρdφ2 ,
ξ02 2ρU
π02 ρ
U = −2Λρ − ln( ),
4m ρ0
π0 ρ
A= ln( )(dt − ω0 dφ), (11.266)
2 ρ0
where m, π0 , ξ0 and ρ0 are constant parameters, Λ = ±1/l2 stands for the cosmo-
logical constant of both signs; for anti-de Sitter Λ = −1/l2 . The parameter ω0 is
a constant related to the angular momentum. The evaluation of the right-hand
side of the Einstein equations for the structural functions (11.266), for ξ02 = 1,
yields
π2
Gν μ = − 0 [δνt δtμ + δνr δrμ − δνφ δφμ − 2 ω0 δνφ δtμ ],
8mr
while the energy–momentum tensor in the left-hand side, for the vector A,
amounts to
π2
4π Tν μ = − 0 [δνt δtμ + δνr δrμ − δνφ δφμ − 2 ω0 δνφ δtμ ],
4r
therefore Einstein–Maxwell equations are fulfilled if 2m = 1/κ or, for κ = 1,
modifying the electromagnetic vector A to be Amod = 2√π20 m ln( ρρ0 )(dt − ω0 dφ).
Recall that additionally one has to set ξ02 = 1.
It is clear that this solution is equivalent to the one treated in Section 11.8.2 for
the identification C1 r +C0 → ρ accompanied with minor scaling transformations
of t and φ.
Notice that this branch of rotating solutions with W (r) = ω0 can be deter-
mined from the static electric field solution, i.e., metric (11.47) and vector A
(11.48), via SL(2, R) transformations: t → t − ω0 φ, φ → φ.
dB dZ 2 d2 B d2 Z
Ett ln7 = B 2 Q l2 (Z −B ) − 2 B 3 Z Q l2 (Z 2 − B 2 ),
dr dr dr dr
d2 B d2 Z
Eφφ ln7 + 3Err ln7 = 2 l2 B 3 Q(Z − B ),
dr2 dr2
hence
d2 B d2 Z dB dZ
Z 2
− B 2
= 0, Z −B = 0,
dr dr dr dr
therefore
Z(r) = c1 B(r).
After a very lengthy and time-consuming integration process we succeeded in
getting two branches of stationary electromagnetic solutions of the Einstein–
Maxwell equations. The structural functions H and W possess a multiplicative
factor a/b which can be absorbed by re–scaling of the Killingian coordinates
according to: |a/b|t → t and |b/a|φ → φ, |a| |b| = ±α.
and
3 α4 1
TT = [−F + H 2 (1 − W )2 ]2 .
64 π 2 F 2 H 2
Since this solution uses the BTZ solution as a limit for α = 0, it is natural
to search for new coordinates in which the BTZ standard structure will become
apparent. First one determines the radial transformation r = β0 (ρ2 + γ0 ); since
grr → gρρ , then
r2 r ρ4 J2
F (r) = 4 2 + 2 l w1 + l w1 − 4 w0 → F (ρ) = 2 − M ρ2 +
2 2 ,
l l l 4
hence
Ml 1
γ02 /l2 + γ0 M + J 2 /4 = 0 → γ0 /l = − ∓ l2 M 2 − J 2 ,
2 2
w0 l w1 + l2 w12 − 4 = ± 2 β l2 M 2 − J 2 .
ρ2 J2 l Q2
L(ρ)2 = − M + + (2ρ2 R− − l J 2 ) ln | Z(ρ) |,
l2 4 ρ2 2 ρ2
K(ρ)2 = H(ρ),
L(ρ)2
N (ρ)2 = ρ2 ,
H(ρ)
W (ρ) Hn = J Q4 l5 R−
3
(ln | Z(ρ) |)2 − 2 J (M 2 l2 − J 2 )(ρ2 − M l2 ),
+Q2 l2 J M 2 l2 − J 2 [J 2 l + 2 l R−
2
− 2ρ2 R− ] ln | Z(ρ) |
l l R−
Z(ρ) := ρ2 − (M l − M 2 l2 − J 2 ) = ρ2 − ,
2 2
R± := M l ± M 2 l2 − J 2 , (11.277)
J l2 JQ2 l2 JQ2 R−
j(ρ → ∞) ≈ − √ R− + √ ln (ρ)
2π ρ 2πρ M 2 l2 − J 2 π ρ M 2 l2 − J 2
l2 JQ2 2 l2 JQ2 R−
J(ρ → ∞) ≈ J − √ R− + √ ln (ρ). (11.281)
M 2 l2 − J 2 M 2 l2 − J 2
The integral energy and mass characteristic at spatial infinity can be evaluated
from the generic expressions E(ρ) = 2 π K (ρ), M (ρ) = N E(ρ) − W J(ρ).
The behavior of the corresponding functions with 0 = 0 as ρ → ∞ is
2ρ l M l2 J 2 Q2
E(ρ → ∞, 0 = 0) ≈ − + − √
l ρ ρ M 2 l2 − J 2
2 2
l Q
+ √ R− 2
ln (ρ),
ρ M 2 l2 − J 2
ρ2 l J 2 Q2
M (ρ → ∞, 0 = 0) ≈ −2 2 + 2 M − √
l M 2 l2 − J 2
2 l Q2
+√ R−2
ln (ρ). (11.282)
M 2 l2 − J 2
The series expansions of the expressions
of the energy and mass evaluated for
2
the reference energy density 0 = − πρ −M0 + ρl2 , which at the spatial infinity
1
l l2 J 2 Q2
(ρ → ∞, 0|∞ ) ≈ (M − M 0 ) − √
2πρ2 2πρ2 M 2 l2 − J 2
l3 Q2 M R−
+ √ ln (ρ),
π ρ2 M 2 l2 − J 2
l(M − M0 ) l2 J 2 Q2
E(ρ → ∞, 0|∞ ) ≈ − √
ρ ρ M 2 l2 − J 2
3 2
2l Q M
+ √ R− ln (ρ),
ρ M 2 l2 − J 2
l J 2 Q2
M (ρ → ∞, 0|∞ ) ≈ M − M0 − √
M 2 l2 − J 2
2 2
2l M Q
+√ R− ln (ρ). (11.283)
M 2 l2 − J 2
For vanishing electromagnetic field charge Q, which gives rise to the rotat-
ing BTZ black hole, the mass, and the energy–momentum quantities become
just the mass–energy–momentum characteristics of the BTZ solution; hence one
concludes that the parameters M and J are related to the mass and momentum
respectively. Moreover, in the electromagnetic case, the momentum, mass, energy
functions logarithmically diverges at spatial infinity.
210 Einstein–Maxwell Solutions
αV 2 αV2
V2 = [ , V 2, − ], V μ Vμ = 0, V2 = N2, Z,
λ2 F (r) λ2 F
F − H 2 (W − 1)2
λ3 = − √ α;
HF
αV 2 2 αV2
V3 = [ ,V ,− ], V μ Vμ = 0, V3 = N3, Z̄. (11.287)
λ3 F λ3 F
4π FH 0 8π FH
(11.288)
one has the following eigenvalues and eigenvectors
2
1 α F − H (W − 1)
2 2
Therefore one has a quite big choice of algebraic types of this tensor.
Finally, the Cotton tensor can be given as
⎡ ⎤
C 11 0 C 13
⎢ ⎥
(C α β ) = ⎢
⎣ 0 C 22 0 ⎥,
⎦ (11.290)
C 31 0 −C 1 1 − C 2 2
2
1 HW α2 (W − 1) (HF,r − F H,r ) 1 α2 (W − 1) F,r
C 11 = − +
32 π F2 32 π F
1 α2 (3 W − 2) H,r 1
− − α2 W,r , (11.291)
32 π H 16 π
212 Einstein–Maxwell Solutions
with components
2
1 α2 H (W − 1) (H F,r − F H,r ) 1 α2 H,r
C 13 = − 2
− , (11.292)
32 π F 32 π H
1 α2 (W − 1) (F,r H − 2 F H,r )
C 22 =−
16 π FH
2
1 α H (W − 1) + F W,r
2 2
+ , (11.293)
16 π F
1 α2 (W − 1) −F (1 + W ) + W 2 H 2 (W − 1) F,r
C 31 =
32 π F2
2
1 F + H 2 W 2 α2 −F + H 2 (W − 1) H,r
− ,
32 π H 3F
2
1 W α −F + H (W − 1) W,r
2 2
− , (11.294)
16 π F
2 2
1 H 2 W α2 (W − 1) F,r 1 α2 H 2 (W − 1) W,r
C 33 = 2
− ,
32 π F 16 π F
2
1 α H W (W − 1) + F (W − 2) H,r
2 2
− (11.295)
32 π FH
possesses, in general, three different eigenvalues, with the possibility of complex
conjugated roots, namely
λ1 = C 2 2 ,
λ2 = −1/2 C 2 2 + 1/2 (C 1 1 + C 2 2 )2 + 4 C 1 3 C 3 1 ,
λ3 = −1/2 C 2 2 − 1/2 (C 1 1 + C 2 2 )2 + 4 C 1 3 C 3 1 ]. (11.296)
The set of eigenvector equations reduces to
V 1 (C 1 1 − λ) + C 1 3 V 3 = 0,
2
C 2 − λ V 2 = 0,
C 3 1 V 1 − V 3 (C 1 1 + C 2 2 + λ) = 0 (11.297)
with solutions
2
λ1 = C 2 2 ; V1 = [0, V 2 , 0], Vμ V μ = V 2 /F, V1 = S1,
λ2 = −1/2 C 2 2 + 1/2 (C 1 1 + C 2 2 )2 + 4 C 1 3 C 3 1 ;
C 13 V 3
V2 = [− 1 , 0, V 3 ], V μ Vμ = {0, > 0, < 0},
C 1 − λ2
V2 = S2, N2, Z,
λ3 = −1/2 C 2 2 − 1/2 (C 1 1 + C 2 2 )2 + 4 C 1 3 C 3 1 ;
C 13 V 3
V3 = [− 1 , 0, V 3 ], V μ Vμ = {0, > 0, < 0},
C 1 − λ3
V3 = S3, N3, Z̄. (11.298)
11.9 Garcı́a Stationary Solutions for a = 0 and b = 0 213
In general, one is dealing with a Type I Cotton tensor; there is a big choice of
eigenvectors.
where C = 8παF2 H , and w0 and w1 are parameters related to mass and angu-
lar momentum, while α is an electromagnetic parameter; the electromagnetic
invariants are
α2
FF = 2 −F + H2 (1 + W)2 ,
FH
3 α4 1
TT = [−F + H2 (1 + W)2 ]2 . (11.302)
64 π 2 F 2 H2
The calligraphic capital letters have been used above to make their relationship
evident to those structural functions arising as real cuts of the complex extensions
of the studied class of metric, see (11.29). This solution of the Einstein–Maxwell
equations can be considered also as a real cut of the complex version of the
stationary electromagnetic solution with BTZ-limit given in the previous para-
graph; the structural functions F, H, and W can be constructed according to
(11.30) with F , H, and W from (11.271) accompanied by the replacement of the
214 Einstein–Maxwell Solutions
sign in front of α2 , α2 el → −α2 mg . If one searches for the anti-de Sitter limit
of this solution, one would arrive at an alternative real cut of the BTZ solution,
namely to the “BTZ solution counterpart,” or “BTZ counterpart” for short.
F dρ2 2
gc = −ρ2dT2 + + H (dΦ + W dT ) ,
H F
ρ2 J2 ρ2 J
F = 2 −M + 2
, H = 2 − M, W = . (11.303)
l 4ρ l 2H
Recall that in the above mentioned metric one can again introduce the radial
coordinate by changing
J
ρ2 → R2 + M l2 , H → R2 , W → ,
2R2
R2 J2
F(ρ) → F (R) = + M + . (11.304)
l2 4 R2
Since this solution possesses the BTZ counterpart as a limit for α = 0, it is
pertinent to search for new coordinates in which the BTZ solution counterpart
structure (11.303) will become apparent.
ρ2 J2 l α2 2
F = − M + + J l − 2 R(−) ρ2 ln |r|,
l2 4ρ2 2 J ρ2
Hn
H(ρ) = ,
Hd
Hn := 4 J 3 l3 l2 M − ρ2 l2 M 2 − J 2 R(+)
2
α2 ln |r|
2 2
−4 l2 M − ρ2 (l2 M 2 − J 2 )R(+) 4
− l6 J 6 α4 (ln |r|)
11.10 Generating Solutions via SL(2, R)–Transformations 215
Hd := −2l5 J 3 l2 M 2 − J 2 R(+)
2
α2 ln |r|
+4 l2 l2 M − ρ2 (l2 M 2 − J 2 )R(+)4
,
l2 Ω(ρ)
W (ρ) = ,
J Hn
2
Ω(ρ) = −l5 J 6 R(+) α4 (ln |r|)
+l2 J 3 l R(+)2
+ 2J 2 l2 M 2 − J 2 + J 2 − R(+)
2
ρ2 R(+)
2
α2 ln |r|
+2 J 2 l2 M 2 − J 2 ρ2 − l2 M R(+)4
,
M l2 l
r := −ρ2 + + l2 M 2 − J 2 , R(±) := M l ± l2 M 2 − J 2 . (11.306)
2 2
The correspondence of this representation with the BTZ solution counterpart in
the limit of vanishing electromagnetic parameter α is evident.
Because of the complexity of the system of equations, we have found very hard
to determine other branches, if any, of exact solutions in the general case.
gt̃t̃ = α2 gtt + 2αγ gtφ + γ 2 gφφ , gt̃φ̃ = αβ gtt + (αδ + βγ) gtφ + γδ gφφ ,
gφ̃φ̃ = β 2 gtt + 2βδ gtφ + δ 2 gφφ , grr = grr , (11.308)
where the new constant are given in terms of the original ones through
αa+ γ b
ã = , a = δã − γ b̃,
αδ − βγ
βa+δb
b̃ = , b = −βã + αb̃,
αδ − βγ
c
c̃ = . (11.310)
αδ − βγ
Notice that
hence, the expressions of the new structural functions are given in the form
F
H̃ = −β 2 + H(δ + β W )2 ,
H
F
W̃ H̃ = −α β + H(δ + β W )(γ + α W ), F̃ = F. (11.312)
H
The transformed electromagnetic field tensor F μ̃ν̃ , as it should be, exhibits its
form-invariant property
⎡ ⎤
0 b̃ − Fc̃
μ̃ν̃ ⎢ ⎥
F =⎢
⎣ −b̃ 0 ã ⎥,
⎦ (11.313)
c̃
F −ã 0
where as before the new field constant parameters are related to the old ones
according to (11.310).
Although the metrics generated via SL(2, R) transformations should be given
as subsections in this section, because of the important place that various of
these solutions occupy in the field, I have decided to report them in separate
sections together with their properties.
F 2 1
g=− dt + dr2 + Hdφ2 ,
H F
H(r)
F (r) = 4 2 2 K0 + H(r) − b2 l2 ln H(r) ,
C1 l
H(r) = C1 r + C0 , (11.314)
α β γ δ
t = √ t̃ + √ φ̃, φ = √ t̃ + √ φ̃, Δ = αδ − βγ = 0 (11.315)
Δ Δ Δ Δ
(in the general (non-normalized) case the same expressions hold except for the
absence of Δ, set simply Δ = 1), the new metric, the rotated one, acquires the
form
⎡ 2 2
⎤
− αΔ HF
+ γΔ H 0 − αβ F
+ δγ
H
⎢ Δ H Δ
⎥
⎢ ⎥
(gμ̃ν̃ ) = ⎢ 0 1
0 ⎥, (11.316)
⎣ F ⎦
2 2
− αβ F
Δ H + ΔH
δγ
0 − βΔ HF
+ δΔ H
218 Einstein–Maxwell Solutions
ν̃ ⎢ ⎥
HΔ 4π H Δ
⎢ 2 ⎥
Tμ̃ = ⎢ 0 − 8π H
1 b
0 ⎥. (11.318)
⎣ ⎦
2 2
1 (α δ+β γ)b
− 4π H Δ
1 δβb
0 8π HΔ
dr2
g = −F (r)dt2 + + r2 dφ2 ,
F (r)
K0 r2 r2 r
F (r) = 2 + 2 − b2 ln r2 = 2 − 8πGQ2 ln ,
l l l r0
√ r
A = 2 Q π G ln dt. (11.323)
r0
To establish the range of values of r0 allowing the existence of a black hole, let
us consider F (r) in the form
r2 k r2
F (r) = (1 − ln ), k = 4πGQ2 l2 , (11.324)
l2 r2 r02
2
the factor (1 − rk2 ln rr2 ) vanishes in the set of points rh determined through the
0
LambertW function, LambertW(x) exp(LambertW(x)) = x, namely
Subjecting the metric (11.323) and the vector potential A to the transformation
at uniform angular velocity
ω ω
t → t − ωφ, φ → φ − 2
t, α = 1, β = −ω, γ = − 2 , δ = 1, (11.327)
l l
one arrives at the metric
ω2 2 2 r2 dr2
g = −(F − 4
r )dt + 2ω(F − 2 )dtdφ + (r2 − ω 2 F )dφ2 + ,
l l F
r2 r2
F = 2 − 4πGQ2 ln 2 ,
l r0
√ r2
A = Q π G ln 2 (dt − ωdφ). (11.328)
r0
One could arrive at this result by using the metric components (11.319) with
transformation coefficients from (11.327) and setting C1 = 2, Δ = 1.
By choosing the axial symmetry as fundamental, the metric (11.328) can be
brought to the form
F(r) 2 dr2 2
g = −r2 (1 − ω 2 /l2 )2 dt + + H(r) (dφ + W(r)dt) ,
H(r) F(r)2
r2 r2 F 2 − r2 /l2 ω r2
F = F = 2 − 4πGQ2 ln 2 , W = ω = −4πGQ2 ln 2 ,
l r0 H H r0
2 2
ω r
H = r2 − ω 2 F = r2 (1 − 2 ) + ω 2 4πGQ2 ln 2 . (11.329)
l r0
The Clément spinning charged BTZ solution is endowed with three parameters
Q, r0 , and ω. It allows for a black hole interpretation.
Alternatively, introducing the scaling transformation r = l/¯l×r̄, the definitions
¯l2 = l2 − ω 2 , |ω| < l, and r̄0 = ¯l/l × r0 , the proper Clément (1996) solution,
dropping the bar from r → ρ, is given in its canonical representation as
1
ds2 = −N (ρ)2 dt2 + dρ2 + K(ρ)2 [dφ + W (ρ)dt]2 ,
L(ρ)2
ρ2
K(ρ)2 = H(ρ) := ρ2 + 4π G ω 2 Q2 ln ( 2 ),
ρ0
ρ 2
4π G (l − ω ) Q2
2 2
ρ2
L(ρ)2 = F (ρ) := 2 − ln ( ),
l l2 ρ2
0
F (ρ)
L(ρ) = F (ρ), K(ρ) = H(ρ), N (ρ) = ρ ,
H(ρ)
4π G Q2 ρ2
W (ρ) = −ω ln ( 2 ),
H(ρ) ρ0
√ ρ
A = 2 Q π G ln (dt − ωdφ); (11.330)
ρ0
11.11 Transformed Electrostatic b = 0 Solutions 221
see also Garcı́a (2009), Eq. (11.24). The corresponding electromagnetic fields are
given by
4Q √
Fμν = − π G(δ[μ t δν] r − ωδ[μ φ δν] ρ ),
ρ
G Q2 l2 + ω 2 t ν G Q2 ω t ν G Q2 l2 + ω 2 φ ν
Tμ ν = − 2 δμ δt − 2 2 δμ δφ + δμ δφ
2l ρ2 l ρ 2 l2 ρ2
G Q2 ω φ ν G Q2 l2 − ω 2 r ν
+ δμ δt − δμ δr . (11.331)
ρ2 2 l2 ρ2
This solution is a black hole if the condition of the form (11.326),
ρ̄20 ≤ 4πGQ2 ¯l2 /e, (11.332)
is fulfilled. It possesses two horizons, at which F (ρ) vanishes, which are roots of
the relation
ρ
ρ2 − ¯l2 8πGQ2 ln = 0, (11.333)
ρ̄0
which are given by the LambertW function, see (11.325). The largest root deter-
mines the event horizon at ρ = ρ+ = ρh , while the inner one is a Cauchy horizon
at ρ = ρ− , with ρ+ > ρ− > ρ̄0 . Since the metric function H changes sign for a
certain value ρ = ρc < ρ̄0 , similarly as the rotating BTZ solution, thus there are
closed timelike curves in the region inside the radius ρc . It is apparent that the
metric and the electromagnetic field are singular at ρ = 0.
The length rescaling was chosen in such a manner that H → ρ2 at spatial
infinity and g(11.330) → gBTZ . According to Clément, one may formally define
mass and angular momentum parameters M (ρ1 ) and J(ρ1 ) by identifying, at a
given scale ρ = ρ1 , the values of the structural functions with the corresponding
BTZ values. Nevertheless, the mass and angular momentum defined in this way
occur to be ρ1 -dependent and diverge logarithmical as ρ1 → ∞.
ρ2 π Gω 2 Q2 ρ2
M (ρ, 0 ) = −2 − 8 + 8π G Q2
ln ( )
l2 l2 ρ20
−2 π K(ρ) N (ρ) 0 . (11.335)
The evaluation of the corresponding functions for the base energy 0 = 0 yield
at spatial infinity ρ → ∞
G ω Q2 G ωQ2 ρ
j(ρ → ∞) ≈ −4 +8 ln ( ),
ρ ρ ρ0
ρ
J(ρ → ∞) ≈ −8π G ω Q2 (1 − 2 ln ( )),
ρ0
1 G ω 2 Q2 G Q2 ρ
(ρ → ∞, 0 = 0) ≈ − −4 2
+ 4 2 (l2 + ω 2 ) ln ( ),
πl lρ lρ ρ0
2ρ π G ω 2 Q2 π G Q2 l ρ
E(ρ → ∞, 0 = 0) ≈ − − 8 +8 ln ( ),
l lρ ρ ρ0
ρ2 π Gω 2 Q2 ρ
M (ρ → ∞, 0 = 0) ≈ −2 2 − 8 2
+ 16π G Q2 ln ( ). (11.336)
l l ρ0
l M0 G ω 2 Q2
(ρ → ∞, 0|∞ (M0 )) ≈ − 2
−4
2π ρ lρ2
G Q2 ρ
+ 4 2 (l2 + ω 2 ) ln ( ),
lρ ρ0
l M0 π G ω 2 Q2
E(ρ → ∞, 0|∞ (M0 )) ≈ − −8
ρ lρ
π G Q2 2 ρ
+8 (l + ω 2 ) ln ( ),
lρ ρ0
π Gω 2 Q2
M (ρ → ∞, 0|∞ (M0 )) ≈ −M0 − 8
l2
l + ω2
2
ρ
+ 8π G Q2 ln ( ). (11.337)
l2 ρ0
Comparing with the energy characteristics of the BTZ solution, one concludes
that a mass parameter M similar to the BTZ mass is absent; instead, a term
in the mass function due to the product of the rotation ω and the charge Q is
present. Notice that E(ρ) and M (ρ) logarithmically diverge at spatial infinity.
The momentum parameter is due to the product of ω Q, and hence is not a free
parameter.
11.11 Transformed Electrostatic b = 0 Solutions 223
Cotton Tensor
The Cotton characterization of this solution is given by
⎡ 2 2
⎤
1 ω (F (ρ)l +ρ ) d3 (F (ρ)ω2 +ρ2 )l2 d3
F (ρ) 0 − 18 ρ (l2 −ω2 ) dρ 3 F (ρ)
⎢ 8 ρ (l2 −ω2 ) dρ3 ⎥
α ⎢ ⎥
Cβ = ⎢ ⎢ 0 0 0 ⎥,
⎥
⎣ 4 2 2
⎦
1 (F (ρ)l +ω ρ ) d3 ω (F (ρ)l2 +ρ2 ) d3
8 l2 ρ (l2 −ω 2 ) dρ3 F (ρ) 0 − 18 ρ (l2 −ω2 ) dρ 3 F (ρ)
(11.338)
where
ρ2
F (ρ) = + 4 Q2 π G ln ρ20 − 4 Q2 π G ln ρ2 ,
l2
d3 Q2 π G
3
F (ρ) = −16 (11.339)
dρ ρ3
The eigenvalue problem yields
λ1 = 0;
2
V1 = [0, V 2 , 0], Vμ V μ = V 2 , V1 = S1,
−F (ρ)Q2 π G
λ2 = 2 ;
ρ3
V1 ωρ+ −F (ρ)l2
1
V2 = [V , 0, ] = Z,
l2 ρ + −F (ρ)ω
−F (ρ)Q2 π G
λ3 = −2 ;
ρ3
V1 ωρ− −F (ρ)l2
1
V3 = [V , 0, ] = Z̄. (11.340)
l2 ρ − −F (ρ)ω
Kamata–Koikawa Limit
It should be pointed out that Clément (1993) also reported the so-called self-dual
solution, published later in Kamata and Koikawa (1995, 1997). By accomplishing
the limiting transition ω → ±l ⇒ ¯l → 0, of the metric structural func-
tions (11.330), while the other parameters Q and r̄0 remain fixed, one arrives
then at the metric (11.330) with structural functions and vector field
r2 l r r
F = 2
, W = ∓ 8πGQ2 ln , H = r2 + l2 8πGQ2 ln ,
l H r̄0 r̄0
r
A = Q ln (dt ∓ l dφ). (11.341)
r̄0
224 Einstein–Maxwell Solutions
Notice that this solution does not possess an horizon; in the limiting transition
¯l → 0, the horizon does not survive since it disappears below ¯l = (4πGQ2 )1/2 ,
as quoted by Clément.
The proper KK representation of this one-parameter solution is achieved by
accomplishing the radial transformation and scaling of parameters
√
r2 = rKK
2
− r0KK
2
, r0KK = (4πGQ2 l2 )1/2 , r0 = r0KK / e, (11.342)
and the subscripts are self-explanatory. It is worth also noticing that a derivation
and analysis of the KK solution has been accomplished in Cataldo and Salgado
(1996), too.
The angular momentum, charge, and mass can be evaluated via quasi-local def-
initions, see below; the presence of logarithmic terms in the structural metric
functions yields to divergences at infinity of the energy–momentum quantities.
As pointed out by the authors, the divergence in the mass can be handled
by enclosing the system in a large circle of radius r0 in which will be bound
M (r0 ) – the energy within r0 – and the electrostatic energy outside r0 given
by −Q2 ln r0 /2, thus the total mass (independent of r0 and finite) is given by
M̃ = M (r0 ) − Q2 ln r0 /2.
Dropping tilde and replacing r → ρ, this metric can be brought to the canonical
form
1
ds2 = −N (ρ)2 dt2 + dρ2 + K(ρ)2 [dφ + W (ρ)dt]2 ,
L(ρ)2
l2 ω 2 Q2
K(ρ)2 = H(ρ) = ρ2 + 2 (M + ln ρ2 ),
l − ω2 4
ρ2 Q2
L(ρ)2 = F (ρ) = 2 − M − ln ρ2 ,
l 4
L(ρ) L(ρ)
N (ρ) = ρ =ρ ,
H(ρ) K(ρ)
ω l2 1 Q2
W (ρ) = − (M + ln ρ2 ). (11.347)
l2 − ω H(ρ)
2 4
ρ2 1
L2 = 2
− m − Q2 ln ρ2 , (11.352)
l 4
while its eigenvalues and the corresponding eigenvectors amount to
λ1 = 0; V1 = [V 3 ω, 0, V 3 ],
l2 − ω 2 2 3 2
Vμ V μ = ρ V , V1 = S1,
l2
√
1 Q V 1 l2 − ω 2 2 1 ω
λ2 = − ; V2 = [V 1 , L , V 2 ],
2 ρ l l
Vμ V μ = 0, V2 = N2
11.11 Transformed Electrostatic b = 0 Solutions 227
√
1 Q V1 l2 − ω 2 2 1 ω
λ3 = ; V3 = [V 1 , − L , V 2 ],
2 ρ l l
Vμ V μ = 0, V3 = N3,
Type: {S, N, N }. (11.353)
The energy–momentum tensor
⎡ 2 2 2 ⎤
1 Q (l +ω ) l2 ω Q2
− 32π ρ2 (l2 −ω 2 ) 0 1
16π ρ2 (l2 −ω 2 )
⎢ ⎥
⎢ ⎥
(T α β ) = ⎢ ⎥,
2
⎢ 0 − 32π
1 Q
ρ2 0 ⎥ (11.354)
⎣ ⎦
2 2 2
ω Q2 1 Q (l +ω )
− 16π
1
ρ2 (l2 −ω 2 ) 0 32π ρ2 (l2 −ω 2 )
1 Q2 −L2 V 1 ω ρ − −L2 l2
λ3 =− 1
; V3 = [V , 0, ] = Z̄,
8 ρ3 l2 ρ − −L2 ω
Type I: {S, Z, Z̄}. (11.357)
228 Einstein–Maxwell Solutions
obtaining
ω2 F dr2
g = −(r2 − 4
F )dt2 + (F − ω 2 r2 )dφ2 − 2ω( 2 − r2 )dtdφ + ,
l l F (r)
K0 r2
F = 2
+ 2 + a2 ln r2 . (11.365)
l l
The proper Dias–Lemos representation uses a more involved definition of the
transformed r-coordinate and parameterizations of the SL(2, R) transforma-
tions, namely
h(r) = C1 r + C0 → (ρ2 + r+ 2
− ml2 )/l2 , C1 = 2/l2 , χ2 = a2 l4 ,
ω
t = 1 + ω 2 t̃ − l ω φ̃ φ = − t̃ + 1 + ω 2 φ̃;
l
dropping tildes, one has
ω2
g = − h − 2 m l2 + χ2 ln |h| dt2
l
ω
−2 ω 2 + 1 m l2 + χ2 ln |h| dt dφ
l
+[h l2 + (ω 2 + 1) m l2 + χ2 ln |h| ]dφ2
l2 ρ2 dρ2
+ 2 − ml2 )[ρ2 + r 2 + χ2 ln |h|] ,
(ρ2 + r+ +
h = (ρ2 + r+
2
− ml2 )/l2 , (11.366)
used in the Eq. (3.2) of Dias and Lemos (2002) was not used here. According to
these authors, this rotating magnetic spacetime is null and timelike geodesically
complete, and as such horizonless.
It it noteworthy to point out that by subjecting the static Hirschman–Welch
metric (11.101) to the SL(2, R) transformation
ω
t→ 1 + ω 2 t − ω lφ, ρ → ρ, φ → − t+ 1 + ω 2 φ,
l
one arrives at the Dias and Lemos (2002) solution too, see also Garcı́a (2009),
Eq. (11.41), given in the canonical representation by metric
1
ds2 = −N (ρ)2 dt2 + dρ2 + K(ρ)2 [dφ + W (ρ)dt]2 ,
L(ρ)2
H(ρ) := (ρ2 + r+
2
− ml2 )/l2 ,
11.12 Transformed Magnetostatic a = 0 Solutions 231
H(ρ) 2 2 + χ2 ln H(ρ),
L(ρ) = ρ + r+
ρ
K(ρ) = 2 + ω 2 l2 m + (1 + ω 2 )χ2 ln H(ρ),
ρ2 + r+
√
L(ρ) ω 1 + ω 2 [ml2 + χ2 ln H(ρ)]
N (ρ) = ρ , W (ρ) = − . (11.369)
K(ρ) l K(ρ)2
Notice that this metric, in the case of vanishing charge χ = 0, yields to an alter-
native coordinate representation of the rotating BTZ solution, with parameter
ω, namely
2
ρ + r+ 2 ρ2 l2
ds2 = − 2
− 1 + ω 2
m l 2
dt2 + 2 2 − m l2
dρ2
l (ρ + r+ ) ρ2 + r+
2
rotating BTZ solution counterpart, and consequently one may think of it as the
reference vacuum solution in the evaluation of the quasi-local energy, momentum
and mass.
The evaluation of the main parts of above functions, i.e., the corresponding
functions independent of 0 behave at infinity according to
ω ρ
j(ρ → ∞) ≈ 1 + ω 2 [m l2 − χ2 + 2χ2 ln ( )],
lπρ l
ω ρ
J(ρ → ∞) ≈ 2 1 + ω 2 [m l2 − χ2 + 2χ2 ln ( )],
l l
1 ml2 − 2χ2 χ2 ρ
(ρ → ∞, 0 = 0) ≈ − + + ln ( )
πl 2π l ρ2 π l ρ2 l
m l 2 − χ2 χ2 ρ
+ ω2 [ 2
+2 ln ( )],
πl ρ π l ρ2 l
2ρ ml2 − r+
2
− 2χ2
E(ρ → ∞, 0 = 0) ≈ − +
l lρ
ω2 ρ
+[m l2 − 2χ2 + 2χ2 ln ( )],
lρ l
1 2
M (ρ → ∞, 0 = 0) ≈ 2m − 2 2 (ρ + r+ 2
+ χ2 )
l
ω2 ρ
+ 2 2 [m l2 − χ2 + 2χ2 ln ( )]. (11.374)
l l
11.12 Transformed Magnetostatic a = 0 Solutions 233
where
χ2 ln (h) + ρ2 ρ2 + Mg
L2 = , h = 2
, Mg = r+ − ml2 . (11.377)
l2 l2
In this representation, the electromagnetic field tensor becomes
⎡ ⎤
0 − lρ3 Lχ2ωh 0
⎢ √ ⎥
⎢ 2
χ 1+ω 2 L2 ⎥
(F α β ) = ⎢ − ω χlρL 0 ⎥, (11.378)
⎣ √
ρ
⎦
2
0 − ρ χL2 1+ω
l4 h
0
ω h 2
λ1 = 0; V1 = [V 1 , 0, √ V 1 ], Vμ V μ = − 2
V 1 , V1 = T1,
l 1+ω 2 1 + ω
√
χ ωρ 1 + ω2 ρ
λ2 = −i √ ; V2 = [i 2 √ V , V , i 2 √ V 2 ], V2 = Z,
2 2
hl 2 lL h L l2 h
√
χ ωρ 1 + ω2 ρ
λ3 = i √ ; V3 = [−i 2 √ V , V , −i 2 √ V 2 ], V3 = Z̄,
2 2
hl2 lL h L l2 h
Type : {T, Z, Z̄}. (11.379)
1 χ2 ω
λ1 = − 2 2
; V1 = [V 1 , 0, √ V 1 ],
8 π l (ρ + Mg) l 1 + ω2
h 2
Vμ V μ = − V 1 , V1 = T1,
1 + ω2
1 χ2 ωl
λ2 = 2 2
; V2 = [ √ V 3 , V 2 , V 3 ],
8 π l (ρ + Mg) 1 + ω2
l2 L2 3 2 ρ2 2
V μ Vμ = V + V 2 , V2 = S2,
1 + ω2 h L2 l 2
1 χ2 ωl
λ3 = ; V3 = [ √ Ṽ 3 , Ṽ 2 , Ṽ 3 ],
8 π l2 (ρ2 + Mg) 1 + ω2
l2 L2 ρ2
V μ Vμ = ( Ṽ 3 2
) + (Ṽ 2 )2 , V3 = S3,
1 + ω2 h L2 l 2
Type : {T, 2 S}. (11.381)
α2 γ2 αβ
g = −( H(−) − H(+) )dt2 + 2(− H(−)
Δ Δ Δ
2 −1
γδ β2 δ2 ρ
+ H(+) )dtdφ + (− H(−) + H(+) )dφ + 2 − M2
dr2 ,
Δ Δ Δ l
2 (1−√α0 )/2 2 (1+√α0 )/2
√ ρ √ ρ
H(+) := ρ1+ α0 −M , H(−) := ρ1− α0 −M .
l2 l2
(11.384)
In the seed hybrid static metric α0 has been used instead of the original α to
avoid confusion with the parameters appearing in the SL(2, R) transformations.
Therefore the transformed electromagnetic tensors stay unchanged. Hence, by
means of an SL(2, R) transformation applied to the hybrid static cyclic symmet-
ric (2+1) Einstein–Maxwell solution, one generates a family of hybrid, stationary
cyclic symmetric solutions with structurally unique field tensors Fμν and Tμ ν .
In particular, one could choose the rotation boost transformation
1 1 ω
t→ (t − ωφ) , φ → φ− t ,
1− ω 2 /l2 1− ω 2 /l2 l2
1
ds2 = −N (ρ)2 dt2 + dρ2 + K(ρ)2 [dφ + W (ρ)dt]2 ,
L(ρ)2
ρ(ρ2 − M l2 )1/2 2 √α0 √α0 2 √
K(ρ)2 = δ l ρ (ρ − M l2 )− α0 /2
lΔ
√ √ √
−β 2 l− α0 − α0
ρ (ρ2 − M l2 ) α0 /2
,
ρ2 ρ2 − M l2
N (ρ)2 = , Δ := (αδ − βγ) = 0,
l2 K(ρ)2
1 ρ ρ2 − M l2 √ √ √
W (ρ) = − 2
αβ l− α0 ρ− α0 (ρ2 − M l2 ) α0 /2
Δl K(ρ)
√ √ √
− γδ l α0
ρ α0 (ρ2 − M l2 )− α0 /2 ,
ρ2
L(ρ)2 = − M. (11.386)
l2
11.13 Transformed Cataldo Hybrid Static Solution 237
ρ2
M (ρ → ∞, 0 = 0) = M − 2
l2
√ √
√ δ 2 l2 α0 + β 2 √ βδ αβ − γδ l2 α0
+M α0 2 2√α + 2 α0 √ . (11.390)
δ l 0 − β2 Δ δ 2 l2 α0 − β 2
Using in expressions (11.387)–(11.389) as reference energy density at the spatial
infinity the quantity which behaves as 0|∞ ≈ − π1l + 2πl M0
ρ2 , the series expansions
of the corresponding quantities at ρ → infinity result in
√
l √ l α0 M β2
(ρ → ∞, 0|∞ ) ≈ ( α 0 M − M 0 ) − √ ,
2π ρ2 πρ 2 δ 2 l α0 − β 2
2
238 Einstein–Maxwell Solutions
√
√ δ 2 l2 α0 − β 2 √
E(ρ → ∞, 0|∞ ) ≈ l1/2− α0 /2
√ ( α0 M − M0 )
ρ Δ
√
√
1/2− α0 /2 β 2 α0 M
+ 2l √ ,
ρ δ 2 l2 α0 − β 2
√
M (ρ → ∞, 0|∞ ) ≈ α0 M − M0
√ √
α0 M β γδ 2 l2 α0 − 2αβδ + β 2 γ
− √ . (11.391)
Δ δ 2 l2 α0 − β 2
Therefore, comparing with the energy characteristics of the BTZ solution, one
√
concludes that role of the mass parameter is played by the product α0 M . At
spatial infinity the mass function is finite, and the energy density and global
energy approach infinity as fast as 1/ρ2 and 1/ρ correspondingly.
J(ρ → ∞) ≈ αJ J + βJ ln ρ,
l Q2
(ρ → ∞, 0|∞ (M0 )) ≈ (αM M − α M 0
M 0 ) + αQ
2π ρ2 2π ρ2
a A
+ J+ ln ρ,
2π ρ π ρ2
l Q2 b B
E(ρ → ∞, 0|∞ (M0 )) ≈ (βM M − βM0 M0 ) + β2 + J + 2 ln ρ,
ρ ρ ρ ρ
M (ρ → ∞, 0|∞ (M0 )) ≈ γM M − γM0 M0 + γQ Q + c J + C ln ρ,
2
11.14 Summary on Electromagnetic Maxwell Solutions 239
Cotton : {0}.
Recall that algebraic structures {T, 2S} are thought of as perfect fluids.
12
Black Holes Coupled To Nonlinear
Electrodynamics
In (3+1) gravity it is well known that the vacuum plus a cosmological constant Λ,
i.e., the (anti) de Sitter–Kottler (1918) solution, is a regular non-asymptotically
flat solution (the scalar curvature is equal to 4Λ and all the invariants of the
conformal Weyl tensor are zero.) On the other hand, Einstein–Maxwell elec-
trovacuum asymptotically flat metrics endowed with timelike and spacelike
symmetries do not allow for the existence of regular black hole solutions. In
order to be able to derive regular (black hole) solutions one has to enlarge the
class of electrodynamics to nonlinear ones; as an example, the regular solution
by Ayón–Beato and Garcı́a (1998), which is a solution to gravitational–nonlinear
electromagnetic fields; the first examples belonging to this class are the Borde
(1997) model and the Ayón–Garcı́a regular charged static black hole.
In (2+1) gravity, in the vacuum case, all solutions are locally Minkowski (the
Riemann tensor is zero); the extension to the vacuum plus cosmological con-
stant allows for the existence of the static and the rotating anti-de Sitter regular
black holes; see Bañados et al. (1992). The static (2+1)-charged black hole with
cosmological constant is singular (when radial coordinate goes to zero the curva-
ture and the Ricci square invariants blow up). Similarly, as in (3+1) gravity, one
may search for regular solutions in (2+1) gravity incorporating nonlinear elec-
tromagnetic fields to which one imposes the weak energy conditions in order
to have physically plausible matter-field distributions. One may look for re-
gular solutions with nonlinear electromagnetic fields of the Born–Infeld type
(Born and Infeld, 1934); Salazar et al. (1987); Salazar et. al (1984); Gibbons
and Rasheed (1995); Fradkin and Tseytlin (1985); Deser and Gibbons (1998),
and/or electrodynamics of wider spectra (Cataldo and Garcı́a, 1999, 2000;
Cataldo et al., 2000).
The Born–Infeld electrodynamics (Born and Infeld, 1934) is free from certain
singularities appearing in the classical Maxwell electromagnetic field theory. Non-
linear electromagnetic Lagrangians, in particular the Born–Infeld Lagrangian,
arise in open string theory (the low-energy effective action for an electromagnetic
12.1 Nonlinear Electrodynamics in (2 + 1) Dimensions 241
field is precisely the Born–Infeld action; see Fradkin and Tseytlin, 1985); string
theory has emerged as the most promising candidate for the consistent quanti-
zation of gravity. In particular, the open string theory has Born–Infeld coupled
vector fields, but it is not clear that this remains the case after compactification
to a 3D space with negative cosmological constant Λ. It is worth mentioning
that there is renewed interest in Born–Infeld theory in various contexts (Frolov
et al., 1996 and Salazar et al., 1987).
Using our experience in the determination of exact solutions in the standard
Einstein gravity with nonlinear electromagnetic fields, in this chapter we deal
with (2 + 1) solutions, mostly static, having nonlinear electrodynamics sources
depending on a single electromagnetic invariant. Stationary generalizations of
the derived static solutions can be constructed using SL(2, R) transformations.
L(F ) −→ −F/4π,
the corresponding energy–momentum tensor has to fulfill the weak energy con-
ditions: for any timelike vector ua , ua ua = −1 (we are using signature – + +)
one requires
Tab ua ub ≥ 0, qa q a ≤ 0,
where q a = Tba ub . This invariant F can be expressed in terms of the electric (vec-
tor) and magnetic (scalar) fields: in a Lorentzian frame, for an observer moving
with the 3-velocity v a , the electric and the magnetic fields are correspondingly
defined as
1
Ea = −Fab v b , B = − abc Fbc v a , (12.1)
2
where Latin indices, to keep the notation used in the original publications, run
the values 0, 1, 2 and abc is the totally anti-symmetric Levi–Civita symbol with
012 = 1, usually the v a is oriented along the time coordinate, i.e., v a = δta , with
such a choice
Ea = F0a , B = −F12 . (12.2)
while the evaluation of the same relation using the electromagnetic energy-
momentum gives
f
A = −8πL,F ( B)2 . (12.8)
r
Therefore, the scalar magnetic field B should be equated to zero, B = 0, thus
the only case to be treated is just the one with the electric field E,
Fab = E(r) δat δbr − δar δbt . (12.9)
12.2 General Nonlinear Electrostatic Solution 243
2F = −E 2 (r), (12.12)
permits to express the electric field E in term of F . Thus, using (12.12), the
derivative, LF can be expressed as a function of r as follows
q
L,r = E,r . (12.13)
4πr
The Einstein’s equations equivalently can be written as
As was pointed out above, the Lagrangian L(F ) must satisfy: (i) correspondence
to Maxwell theory, i.e. L(F ) −→ −L/4π, and (ii) the weak energy conditions:
Tab ua ub ≥ 0 and qa q a ≤ 0, where q a = Tba ub for any timelike vector ua ; in our
case the first inequality requires
− (L + E 2 L,F ) ≥ 0, (12.16)
If one replaces f,r from (12.19) and its derivative f,rr into (12.18) one arrives,
taking into account the equation (12.13), at an identity. Therefore one can forget
the equation (12.18) and integrate the relevant Einstein equation (12.19):
f (r) = −M − Λr2 + 16π r L(F (r)) + E 2 L,F dr. (12.20)
= 1 d
divE (rE(r)) = 2πρe , (12.31)
r dr
substituting here E(r) from (12.30) one obtains
qr02
ρe = , (12.32)
2πr(r2+ r02 )3/2
where r0 = q/b. It is easy to verify that the surface integral of ρe is equal to q,
in fact ∞ ∞
dr
ρe dA = qr02 2 + r 2 )3/2
= q. (12.33)
0 0 (r 0
246 Black Holes Coupled To Nonlinear Electrodynamics
It is worth pointing out the regular behavior of the vector E of the electric field
and the surface charge distribution ρe ; the same regular behavior one encounters
for the static spherically symmetric electric field in (3+1) Born–Infeld theory.
From this last expression one sees that there is a contribution of the Born–Infeld
field to the term with the cosmological constant. The Lagrangian and the electric
field are given by
b2 F b2 r
L(F ) = − 1+2 2 −1 =− −1 ,
4π b 4π r2 + q 2 /b2
q
E(r) = . (12.37)
r + q 2 /b2
2
Invariants
Having the explicit metric, one can easily calculate the curvature tensor
components:
2b2 r
R0110 = 2b2 − Λ − , (12.38)
r2 + q 2 /b2
q2
R0202 = f (r) Λr − 2b r + 2b r
2 2 2 2
r2 + 2 , (12.39)
b
and
−1 q2
R1221 = f (r) Λr − 2b r + 2b r
2 2 2 2
r2 + 2 . (12.40)
b
characteristics – Ricci and Kretschmann scalars – such as the Ricci scalar and
the Ricci square blow up at r = 0, (the Riemann square does not need to be
evaluated since in (2+1) dimensions the Riemann tensor is given in terms of
the Ricci tensor, curvature scalar and the metric tensor). Additionally, in (2+1)
dimensions one considers the behavior of the invariant det(Rab )/ det(gab ), thus
one has to evaluate the invariants (Weinberg, 1972)
det(Rab )
R, Rab Rab ,
det(gab )
at critical points. In our case these three invariants are given as
4(2q 4 + 5q 2 b2 r2 + 3b4 r4 )
R = 6Λ − 12b2 + 3/2
, (12.41)
b2 r (r2 + q 2 /b2 )
8b2 (4b2 − 2Λ) 3r2 + 2q 2 /b2
Rab Rab = 3(4b − 2Λ) −
2 2
r r2 + q 2 /b2
r2 3(r2 + q 2 /b2 )2
+8b4 2 + 2 + , (12.42)
r + q 2 /b2 r2
and
2
det Rab 2Λ − 4b2 2b2 (2r2 + q 2 /b2 )
= 4b2
−2Λ + 4b −
2
.
det gab r2 + r3 r2 + q 2 /b2 r r2 + q 2 /b2
(12.43)
We have an extreme black hole if Λ < 0, Mextr > 0 and q 2 < b2 ; this last con-
straint arises from Λ < 4b2 q 2 /(q 2 − b2 ) when one demands Mextr > 0. Fixing
the values of the Mextr for given values of q, b and Λ, one has a black hole
solution with inner and outer horizons when M > Mextr . For M < Mextr one
has a soliton solution, i.e., there are no horizons at all and we have a naked
singularity. It may occur that for certain values of the parameters there is
only one positive root of the equation f (r) = 0: the horizon rh > 0. In such
a case one has also a black hole solution. A similar analysis can be carried
out for other rextr which arises for Λ > 0. At infinity, for weak electromag-
netic field this BI solution asymptotically behaves as the static charged Peldan
solution.
Moreover, for Λ = −1/l2 this BI solution at infinity behaves as anti-de Sitter
spacetime. At the origin r = 0, this BI solution is singular. If one requires
additionally the vanishing of the cosmological constant, one arrives at a solution
reported in Gott et al. (1986).
As regards the analytical extension of our solution, one may follow step by
step the procedure presented in standard textbooks (for instance, Wald, 1984)
to determine the Kruskal–Szekeres coordinates. First one has to integrate for
the tortoise r∗ coordinate: r∗ = 1/f (r)dr, which in our case has no expression
in terms of elementary functions; next one defines the null coordinate u and v
by u = t − r∗ , v = t + r∗ ; in these coordinates the studied metric acquires the
form
ds2 = −f (r)dudv + r2 dΩ2 , (12.46)
kB TH = k. (12.47)
2π
For a spherically symmetric (and for circularly symmetric in (2+1) dimensions)
system the surface gravity can be computed via (for our signature)
1 ∂r gtt
k = − lim √ , (12.48)
r→r+ 2 −gtt grr
12.4 Regular Black Hole Solution 249
where r+ is the outermost horizon. For our solution we have from (12.22), (12.47
and (12.48) that
q2
kB T = −2(Λ − 2b )r+ − 4b r+ + 2 .
2 2 2 (12.49)
4π b
where now dΩ2 = dθ2 + sin2 θdφ2 . As in the (2+1) case, there is also a contribu-
tion to the cosmological constant term of the nonlinear field. The corresponding
electric field is given by
q
E(r) = . (12.51)
r + q 2 /b2
4
Notice that the electric field in this case is regular everywhere. This gravitational
field asymptotically behaves as the Kottler charged solution, with the structural
function and electromagnetic field of the form
2M Λ 2 q2 1
f (r) = 1 − − r + 2 +O ,
r 3 r r6
q 1
E(r) = 2 + O .
r r3
By canceling Λ one obtains an asymptotically flat solution.
12.4.1 Regularity
To establish that this solution is regular one has to evaluate the curvature Ricci
and Kretschmann scalars. The non-vanishing curvature components, which are
regular at r = 0, are given by:
q 2 (a2 − r2 )
R0110 = + Λ, (12.54)
(r2 + a2 )2
2 2
q r
R0202 = −f (r) + Λr 2
, (12.55)
r2 + a2
2 2
q r
R1212 = f (r)−1 + Λr 2
, (12.56)
r2 + a2
where 0,1,2 stand respectively for t, r and Ω.
Evaluating the invariants R, and Rab Rab one has
2q 2 (r2 + 3a2 )
R= + 6Λ (12.57)
(r2 + a2 )2
r4 + 2r2 a2 + 3a4
Rab Rab = 12Λ2 + 4q 4
(r2 + a2 )4
8Λq 2 (3a2 + r2 )
+ . (12.58)
(r2 + a2 )2
Since the metric, the electric field and these invariants behave regularly for all
values of r, we conclude that this solution is curvature regular everywhere. Nev-
ertheless, for solutions without any horizon or black hole solutions with an inner
and outer horizons, at r = 0 a conical singularity may arise. At r = 0 the function
f (r) becomes f (0) = −M − q 2 ln(a2 ). Thus for M positive, M > 0, and a in the
range 0 < a < 1, the value of f (0) will be f (0) = −M + q 2 ln(1/a)2 , which will
be positive, say f (0) := β 2 , if ln(1/a)2 > M/q 2 . In such a case, for 0 < β < 1 the
solutions will show angular deficit since the angular variable Ω, which originally
runs 0 ≤ Ω < 2π will now run 0 ≤ Ω < 2βπ; the parameter a can be expressed in
terms of β, q and M as a2 = exp[−(β 2 +M )/q 2 ]. For β = 1, there will be no angu-
lar deficit, the ratio of the perimeter of a small circle around r = 0 to its radius,
12.4 Regular Black Hole Solution 251
as this last tends to zero, will be 2π. If one allows M to be negative, M < 0,
and a to take values in the interval 0 < a < 1, then f (0) will be always posi-
tive; in this case one can adopt the following parametrization: −M = β 2 cos2 α,
q 2 ln(1/a)2 = β 2 sin2 α, therefore f (0) = β 2 . One will have angular deficit if
0 < β < 1, and for β = 1 the resulting (2+1) spacetime will be free of singular-
ities. Another possibility with positive f (0) = β 2 arises for M < 0, and a > 1;
f (0) can be parameterized as −M = β 2 cosh2 α, q 2 ln(1/a)2 = β 2 sinh2 α. Again
the values taken by β will govern the existence of angular deficit, for β = 1
the solutions will be regular. If f (0) is negative, f (0) =: −β 2 , the character of
the coordinates t and r changes, the coordinate t becomes spacelike, while r is
now timelike and one could think of the singularities, if any, as causal structure
singularities because they could arise at the “time” r = 0. In what follows we
shall treat the parameter a as a free one, having in mind the above restrictions
to have solutions free of conical singularities.
12.4.2 Horizons
To establish that this solution represents a black hole, one has to demonstrate
the existence of horizons, which require the vanishing of the gtt component, i.e.,
f (r) = 0. The roots of this equation give the location of the horizons (inner and
outer in our case). The roots – at most four – of the equation f (r) = 0 can be
expressed in terms of the Lambert W (r) function
2
Λa2 − M Λ Λa − M 1
r1,2,3,4 = ±[exp( − LW [ exp ]) − a2 ] 2 .
q2 q2 q2
There arise various cases which depend upon the values of the parameters: four
real roots (two positive and two negative roots: the negative roots have to be
ignored), two complex and two real roots, two complex and one real positive
root (the extreme case), and four complex roots (no black hole solutions). This
analytical expression for the Lambert function can be used in all calculations,
recall that Lambert function fulfills the following equation
Analytically one can completely treat the extreme black hole case; for it, the
derivative of f (r) has to be zero, ∂r (f (r)) = 0 , at the rextr , this gives
q2
rextr = −a2 − >0 (12.59)
Λ
for Λ < 0. From this expression one concludes that the following inequality holds:
a2 < −q 2 /Λ. Entering now rextr into f (r) = 0 one obtains a relation between the
parameters involved, which can be solved explicitly for the mass (the extreme
one):
252 Black Holes Coupled To Nonlinear Electrodynamics
2 2 −Λ
Mextr = a Λ + q 1 + ln , (12.60)
q2
this Mextr varies its values depending on the values given to the parameters a,
q and Λ. We have an extreme black hole characterized by negative cosmological
constant, Λ < 0, and positive extreme mass, Mextr > 0, if the parameter a is
restricted by the inequality a2 < −(q 2 (1 + ln(−Λ/q 2 )))/Λ.
For other values of the mass M , one distinguishes the following branches:
if M > Mextr one has a black hole solution, and if M < Mextr there are no
horizons.
12.4.3 Thermodynamics
Similarly, as this question was treated in the previous section, here one can also
evaluate the temperature (12.47), of the black hole through its surface gravity
(12.48). For our present solution we have from (12.22), (12.47) and (12.48) that
q 2 r+
kB T = −Λr+ − 2 . (12.61)
2π r+ + a2
The entropy can be trivially obtained using the entropy formula S = 4πr+ .
To achieve the maximal extension of our regular black solutions one has to
proceed step by step through the procedure presented above, determining first
the Kruskal–Szekeres coordinates.
one obtains
L = C |F |3/4 , (12.63)
where C is a constant of integration, and bars denote moduli. Taking into account
(12.12) one gets that
L = C E 3/2 . (12.64)
Entering this L into (12.13) one gets that E = (q 2 /6πC)2 1/r2 . By choosing now
C = |q|/6π we arrive at the electric field
q
E(r) = , (12.65)
r2
12.5 Coulomb-Like Black Hole Solution 253
which coincides with the standard Coulomb field for a point charge of the
Maxwell theory in the Minkowski space.
The Lagrangian in this case is given by
q2 |q| 3/2
L= = E . (12.66)
6πr3 6π
It is easy to check that this Lagrangian satisfies the weak energy conditions:
qa q a ≤ 0, where q a = Tba ub for any timelike vector ua and on the other hand
q2
− (L + E 2 L,F ) = ≥ 0. (12.67)
12πr3
We rewrite the Einstein’s equations equivalently as
where it has been considered that the trace of Tab is equal to zero, T = 0. The
Einstein equations for Rtt (= −f 2 Rrr ) and RΩΩ components yield respectively
the equations
f,r 2q 2
f,rr + = −2Λ + 3 , (12.69)
r 3r
4q 2
f,r = −2Λr − 2 . (12.70)
3r
It is easy to show that equation (12.69), by virtue equation the Maxwell equa-
tions is just an identity. Therefore the only Einstein equation to be integrated
is (12.70), which gives
4q 2
f (r) = C0 − Λr2 + , (12.71)
3r
where C0 is a constant of integration. We will see now that the constant C0
can be expressed in terms of asymptotic values of the mass. For the circularly
symmetric metric (12.6) the quasilocal energy E(r) and the quasilocal mass M (r)
at a radial boundary r can be shown to be respectively
where f0 (r) = g0rr (r) is a background metric function which determines the zero
of the energy. The function f0 (r) can be obtained simply by setting constants of
integration of our solution (12.71) to some special values that then specify the
reference spacetime. We set in (12.71) q = C0 = 0 as the background and as
a consequence it is the vacuum anti-de Sitter spacetime. The same background
function was used in Brown et al. (1994); Chan (1996) for analogous calcula-
tions. Now for Λ = −1/l2 < 0 we have that f0 (r) = r/l (for a de Sitter
254 Black Holes Coupled To Nonlinear Electrodynamics
spacetime Λ > 0) and then the quasi-local energy and the quasi-local mass are
given respectively by
r r2 4 q2
E(r) = − C0 + 2 + , (12.74)
l l 3 r
r r2 4 q2 r2 4 q2
M (r) = 2 C0 + 2 + − 2 C0 + 2 + . (12.75)
l l 3 r l 3 r
These equations give the location of the horizons (if there are any). Since the
coordinate r ranges over positive values from 0 to infinity, we exclude the roots
which are negative. Because the positive character of M > 0, the complex or
real character of these roots depends crucially on Λ values. For Λ > 0 there is
only one real root. For Λ < 0 there are two possibilities: two complex roots and
one real; or three real roots. To obtain only
√ real roots we must cancel the term
h/3Λ + M/h. Then we have that h = ± −3ΛM . If M /Λ + 12q 4 = 0 then one
3
12.5 Coulomb-Like Black Hole Solution 255
has an extreme black hole, in such case M = Mextr = −(12q 4 Λ)1/3 and the roots
become
1/3 1/3
2q 2 2q 2
r1 = 2 , r2 = r3 = rextr = − . (12.81)
3Λ 3Λ
From these expressions we see that for Λ > 0 there is not a positive rextr ,
although the horizon r1 > 0 the behavior of the f function is quite peculiar, for
r > r1 , f (r) < 0, while for r < r1 , f (r) > 0. For Λ < 0, we have a positive
extreme horizon (since r1 < 0 it does no represent an horizon).
For Λ > 0 (and M > 0) we have that α > 0 and then we always have one
real root r1 > 0 and two complex roots r2 and r3 . From (12.71), hence f (r) is a
decreasing function for r > 0.
For Λ < 0 the roots r2 and r3 can be real or complex roots (in this case always
r1 < 0). For M 3 + 12q 4 Λ > 0 or q 2 > −M/3Λ/2 we have that these roots are
complex. For M 3 + 12q 4 Λ < 0 or 0 ≤ q 2 < −M/3Λ/2 we can write the real
roots as
⎛ ⎡ ⎤ ⎞
−M 1 2q 2
2π
r± = −2 cos ⎝ arccos ⎣ ⎦± ⎠. (12.82)
3Λ 3 −M
3 3
3Λ
In this case these roots represent the horizons of a black hole; the outer horizon
(the event horizon) is r+ , while the inner horizon is r− . When q = 0 we have
that r− = 0 and r+ = −M/Λ, which is the horizon of the static BTZ metric.
Using the negative cosmological constant as Λ = −1/l2 one gets
⎛ ⎛ ⎡ ⎤ ⎞⎞
⎝ 16M 1 2q 2
2π
kB T = − cos ⎝ arccos ⎣ ⎦+ ⎠⎠
4π 3l2 3 M 3 l2 3
3
⎛ ⎡ ⎤ ⎞
q2 1 2q 2
2π
− cos−2 ⎝ arccos ⎣ ⎦+ ⎠. (12.83)
M l2 3 M 3 l2 3
3
4q 2 3 M 2
rH = ,T =− < 0. (12.84)
3M 16π q 4
256 Black Holes Coupled To Nonlinear Electrodynamics
Using the Schwarzschild coordinate frame for a static cyclic symmetric met-
ric in (2 + 1) gravity coupled minimally to a dilaton logarithmically depending
on the radial coordinate in the presence of an exponential potential together
with a Maxwell electric field, by solving first-order linear Einstein equations,
the general solution is derived and identified with the Chan–Mann dilaton solu-
tion, and, in the charged case, with the Chan–Mann charged dilaton solution.
In these coordinates, a new stationary dilaton solution is obtained; it does not
allow for a de Sitter–Anti-de Sitter limit at spatial infinity, where its structural
functions increase indefinitely. On the other hand, it is horizonless and allows for
a naked singularity at the origin of coordinates; moreover, one can identify at a
large radial coordinate a (quasi-local) mass parameter and in the whole space a
constant angular momentum.
Via a general SL(2, R) transformation, applied on the static cyclic symmetric
metric, a family of stationary (charged) dilaton solutions has been generated. A
particular SL(2, R) transformation is identified, which gives rise to the rotat-
ing Chan–Mann (charged) dilaton solution. All the exhibited solutions have
been characterized by their quasi-local energy, mass, and momentum through
their series expansions at spatial infinity. The algebraic structure of the Ricci–
energy-momentum, and Cotton tensors is given explicitly. A summary of results
is presented in section 13.9.
leaving still a freedom in the choice of the r–coordinate, which can be used to
fix the metric structure, for instance:
the Schwarzschild coordinate frame: ds2 = −A(r)2 dt2 + B(r)2 dr2 + r2 dt2 , or
the gtt = −1/grr coordinate frame: ds2 = −F (r)2 dt2 + dr2 /F (r)2 + H(r)2 dt2 . In
one of the well-known works on dilaton Chan and Mann (1994), this last metric
was used to argue that its use simplifies the calculations; this is partially true
when it is also assumed that H(r) = rN/2 ; nevertheless, the field equations to be
integrated are of the second order. The r-gauge freedom allows one to fix only one
structural function and leave two structural metric functions undetermined in
the class of static circularly symmetric metric in (2+1) gravity; the Einstein field
equations yield further constraints fixing the dependence on r of the remaining
structural functions.
In the Schwarzschild coordinate frame, gθθ = r2 , the equations to be inte-
grated, for a dilaton logarithmically depending on r and associated to an
exponential potential, are first-order equations for each of the (different) struc-
tural functions. Thus, one can claim on its generality and uniqueness under the
restrictions imposed on the dilaton field. It should be pointed out that in the
Chan–Mann papers the solutions were derived under the ansatz gθθ = γ 2 rN . In
this section, the integration of the field equations is done for the static cyclic
symmetric metric in the r-gauge fixed according to gtt = −1/grr = F (r), and
gθθ = H(r); it happens that the resulting two possible solutions can be identified,
via parameter choice, with the static Chan and Mann (1994) solution.
On the other hand, dealing with stationary cyclic symmetric metrics for
logarithmically dependent dilaton field, there exists a single solution for the
Schwarzschild coordinate gθθ = r2 –gauge.
The action to be considered in this work dealing with (2+1)-dimensional
gravity is given by
√
S = d3 x −g R − 4 ∇μ Ψ ∇μ Ψ + 2 ebΨ Λ , (13.1)
where Ψ is the massless minimally coupled scalar field, and R is the scalar
curvature. The variations of this action yield the dynamical equations
EQμν := Rμν = 4 ∇μ Ψ ∇ν Ψ − 2gμν e b Ψ Λ,
b
∇μ ∇μ Ψ + Λ e b Ψ = 0, (13.2)
4
where at this stage Λ and b are arbitrary parameters. Details about the derivation
of the Einstein–dilaton equations and analysis of the known solutions can be
found in Garcia–Diaz and Gutierrez–Cano (2014b).
dr2
g = −N (r)2 dt2 + + r2 dφ2 . (13.3)
L(r)2
The scalar field equation for the dilaton Ψ(r) = k ln(r) becomes
L2 d N L dL 1
F EQ := + + bΛ rbk = 0. (13.4)
r N dr r dr 4k
The simplest Einstein equation is EQ3 3
L2 d N L dL
EQ3 3 = − − + 2 Λ rbk = 0, (13.5)
N r dr r dr
thus, from EQ3 3 + F EQ, one gets the constant relation
On the other hand, equation EQ1 1 − EQ2 2 − EQ3 3 gives a first-order equation
for L2 , namely
d 2 L2 2
L + 4 k2 − 2 Λr1−8 k = 0 (13.7)
dr r
with integral
Λ
L(r)2 = r−4 k C1 + r2−8 k
2 2
1 − 2 k2
Λ
4 k2
r r−8 k .
2
2
= r C1 + (13.8)
1−2 k 2
The remaining Einstein equation arises from −(EQ1 1 − EQ2 2 + EQ3 3 ) r/(2 L2 ),
and yields
1 dN Λ 2 k2
− 2 r(1−8 k ) − 2 = 0. (13.9)
N dr L r
Finally, substituting L(r)2 from (13.8) into the above equation (13.9) for N (r),
and integrating one obtains
2 4 k2 Λ 2
2
N (r) = CN r C1 + r , N (r)2 = CN 2 r8 k L(r)2 . (13.10)
2
1−2 k 2
2 2
2 N , −EQ1
d 1 L d
The remaining equation with second derivatives of dr = N r dr N +
2 2
N dr dr + N dr 2 −2 Λ r
L dL dN L d N bk
= 0, is fulfilled for the determined structural functions.
Thus, the general static solution can be given as
2 dr2
g = −CN 2 r8 k L(r)2 dt2 + + r2 dφ2 ,
L(r)2
Λ
4 k2
r r−8 k ,
2
2 2
L(r) = r C1 +
1 − 2 k2
Ψ(r) = k ln (r). (13.11)
260 Dilaton Field Minimally Coupled to (2 + 1) Gravity
1 2
α , r → rα1/(4 k ) , φ → φ α−1/(4 k ) ,
2 2
Λ=± 2
l
C1 → C1 α1+1/(2 k ) , CN → CN α−(1+1/(4 k )) ,
2 2
(13.12)
one arrives at the metric (13.11) with structural functions with Λ = ± l12 .
Notice that the Λ used by in Chan–Mann, when considered as a cosmological
constant, differs from the standard cosmological constant Λs = ± l12 = −Λ, where
+ and − stand correspondingly for de Sitter and anti-de Sitter (AdS).
If Λ (1 − 2 k 2 ) > 0 then one has
Notice that for vanishing dilaton parameter k = 0, the derived solution reduces
to the BTZ one for C1 = −M , CN = 1, and Λ = 1/l2 , on the other hand, for
opposite in sign constants one gets the cosmological dS solution.
Accomplishing in the general above solution (13.11) the transformations and
constant parameterizations:
1 2−N
r→r N/2
, φ → β θ, k = ± , CN = 2β/N, 2 β 2 C1 = −M N,
2 N
one gets the Chan–Mann static solution: Chan (1997); Chan and Mann (1996)
dr2
ds2 = −U (r)dt2 + β 2 + β 2 rN dθ2 ,
U (r)
2M 1− N 8Λβ 2
U (r) = β 2 F (r) = − r 2 + rN ,
N (3 N − 2)N
1
Ψ(r) = ± N (2 − N ) ln (r), 0 < N < 2, N = 2/3. (13.13)
4
13.2 Static Black Hole Coupled to a Scalar Ψ(r) = k ln(r) 261
δV1N 2
V3 = , Vμ V μ = (V 1 ) N 2 δ 2 − 1 ;
r
δ 2 > 1, V1 = S1, δ = 1, V1 = N1, δ 2 < 1, V1 = T1
2
2 −2−4 k2 −8 k2 4 k − 1
λ3 = 4 k r C1 − 2Λ r = T 2 2 : V3 = [0, V 2 , 0],
(2 k 2 − 1)
2
V2
Vμ V μ = , V3 = S3. (13.17)
L2
Hence, depending on the sign of the norm, one will have spacelike, δ 2 > 1, null,
δ 2 = 1, or timelike, δ 2 < 1, eigenvectors: V1, 2 = {S, N, T}. Therefore, in the
case of a double root one may choose different vector components determining,
for instance, one spacelike vector S and the other timelike T or null N one.
Therefore, one may have the algebraic Ricci types: {S, 2S}, {S, 2N }, {S, 2T } and
{S, (S, T )}, {S, (T, T )}, . . . , {S, (T, N )}.
The matrix of the Cotton tensor is given by
⎡ ⎤
0 0 C 13
⎢ ⎥
(C μ ν ) = ⎢ ⎣ 0 0 0 ⎥
⎦,
C 31 0 0
C1 2
k 1 − 2 k 2 r−2−8 k ,
2
C 13 =
CN
C 31 = −C1 k 2 1 − 2 k 2 r−4 L(r)2 ,
C1 2 4 2
k 1 − 2 k 2 r−6−8 k L(r)2 < 0.
2
C 13C 31 = − (13.18)
CN
The characteristic equation allows for one real and a pure imaginary eigenvalues.
Therefore, the Cotton tensor is characterized algebraically by
V2
λ1 = 0 : V1 = [0, V 2 , 0], Vμ V μ = , V1 = S1,
L2
λ2 = λ̄3 = C 13C 31 :
13.3 General Static Chan–Mann Solution 263
√
C 31
V2 = Z = [V , 0, V = √ 1 V 1 ], V3 = Z̄.
1 3
(13.19)
C 3
Thus, the algebraic type of the Cotton tensor is Type I : {S, Z, Z̄}.
dr2
g = −F (r)dt2 + + H(r)dφ2 .
F (r)
The Einstein-field equations (13.2) for this metric are
2
F d d 1 d
EQtt = −2 Λ ebΨ F + H F+ F F ,
4 H dr dr 2 dr2
d 2 d2 d2
Λ ebΨ 1 dr H 1 dr 2H 1 dr 2F
EQrr = 2 + − −
F 4 (H)2 2 H 2 F
d d 2
1 dr F dr H d
− −4 Ψ ,
4 HF dr
d 2 2
1 d d 1 F dr H 1 d
EQφφ = 2 Λ e H −
bΨ
F H+ − H F
2 dr dr 4 H 2 dr2
(13.20)
The field
√ equation (13.21), introducing a new function Y (r) through F =
Y (r)/ H and replacing Ψ = kln(r), becomes
d Y (r) √
4k ( ) + bΛ rb k H = 0, (13.25)
dr r
with first integral
√
4 k Y (r) = −r bΛ rb k Hdr + C0 r, (13.26)
N =1− 1 − 16 k 2 → k = ± N (2 − N )/4,
266 Dilaton Field Minimally Coupled to (2 + 1) Gravity
this set of functions. On the other hand, the equation for W (13.39) becomes
d
W = W0 CN r4 k −3
2
(13.47)
dr
with general integral
W0 CN
r4 k −2 + W1
2
W =
2(2 k − 1)
2
Λ C
√ N r4 k −2 + W1 .
2
→ W (r) = (13.48)
2 k2 − 1 2 k
268 Dilaton Field Minimally Coupled to (2 + 1) Gravity
2 dr2 2
g = −CN 2 r8 k L(r)2 dt2 + + r2 (dφ + W (r)d t) ,
L(r)2
Λ 1
L2 = r−4 k C1 +
2
,
2 k 2 (2 k 2 − 1) r2
Λ C
√ N r4 k −2 ,
2
W =
2 k − 1 2k
2
Recall
that in the derivation of this solution √one has to fit the condition
Λ 2 k 2 − 1 > 0, together with k = 0, k = 1/ 2. Moreover, having in mind
2
that gtt = −CN 2 C1 r4 k , the signature {−, +, +} imposes some constraints on
the range of values of the constant:
assuming that Λ is related with the standard cosmological constant Λs then
Λs = ± l12 = −Λ, where + and − stand correspondingly for de Sitter and anti-de
Sitter (AdS) cases.
If Λ < 0, then 2 k 2 − 1 < 0 with
0 0 − 2rΛ
4
In the case of a double root, one may choose different vector components
determining, for instance, one spacelike vector S and the other timelike T
or null N one, and other possible combinations. Therefore, one may have
270 Dilaton Field Minimally Coupled to (2 + 1) Gravity
C 3 1 = k 2 2 k 2 − 1 CN C1 2 r−4 k −4 .
2
(13.55)
The characteristic equation allows for one real and pure imaginary eigenvalues.
Consequently, the Cotton tensor is algebraically characterized by
(V 2 )2
λ1 = 0 : V1 = [0, V 2 , 0], Vμ V μ = , V1 = S1,
L2
= i k 2 C1 2 k 2 − 1 r−4 k −3 L(r) :
2
λ2 = λ̄3 = (C 1 1 )2 + C 1 3 C 3 1
V2 = Z = [V 1 , 0, V 3 ], V3 = Z̄ = [V̄ 1 , 0, V̄ 3 ]. (13.56)
Hence, the algebraic type of the Cotton tensor is Type I: {S, Z, Z̄}.
t = α T + β Φ,
φ = γ T + δ Φ, Δ := αδ − βγ. (13.57)
or, in the standard representation used to evaluate energy, mass, and momentum,
one has
dr2
g = −N(r)2 dT 2 + + K(r)2 (dΦ + W(r)dT )2 ,
L(r)2
N 2 r2 2
N(r)2 = (α δ − β γ) ,
− β2N 2
δ 2 r2
K(r)2 = δ 2 r2 − β 2 N 2 ,
Λ
L(r)2 = r−4 k C1 + r−8 k +2 ,
2 2
1 − 2 k2
δ γ r2 − β α N 2
W(r) = , β = 0 = δ,
δ 2 r2 − β 2 N 2
2 Λ 2 8 k2
N 2 := CN2
r4 k C1 + r 2
= CN r L(r)2 . (13.59)
1−2 k 2
N2B −4+4 B 2
N (r)2 = NB2 = r2 r−2 B C1 B 2 + Λ , L(r)2 = L2B =
2
r . (13.60)
B2
The quasi-local momentum J(r) is constant in the whole spacetime, namely
δ 2 − β 2 Λ − β 2 C1 B 2 r−2 B
2
E(r) = −2
B δ 2 − β 2 Λ − β 2 C1 B 2 r−2 B 2
+β 2 C1 r−2 B B 4 ,
2
(13.62)
2Δ Λ + C1 B 2 r−2 B
2
M (r) = − β 2 C1 B 2 (B 2 − 1)
B δ 2 − Λβ 2 − β 2 C1 B 2 r−2 B 2
2
β α Λ − δ γ + β α C1 B 2 r−2 B
2
+r2 B (δ 2 − Λβ 2 ) + J0
δ 2 − β 2 Λ − β 2 C1 B 2 r−2 B 2
−2 0 π Δ r2 Λ + C1 B 2 r−2 B 2 . (13.63)
When the dilaton field and the rotation vanish, B = 1, β = 0, the generated
solution reduces to the AdS metric and one identifies C1 = −M , or com-
paring with the quasi-local BTZ mass MBT Z (r → ∞) = M − M0 , the same
conclusion is achieved. Frequently, in the literature one encounters the SL(2, R)
transformation
T Φ ω T Φ
t= −ω ,φ=− + , (13.64)
1− ω2
1− ω2 l2 1 − ω2
1− ω2
l2 l2 l2 l2
dr2 2
+ 2 + δ 2 r2 − β 2 r8 k CN 2 L2 dφ2 ,
L
Λ
r−8 k , Ψ(r) = k ln (r),
2 2
L(r) = r4 k C1 +
2
r 2
(13.67)
1 − 2 k2
13.5 Stationary Dilaton Solutions Generated 273
β 2 C1 CN 2 + ρ2 2 M + M2 − J0 2 Λ ρ2 − J0 2
r2 = = ,
δ 2 − β 2 CN 2 Λ 4 δ 2 M 2 − J0 2 Λ
δ J0 M+ M2 − J 0 2 Λ
α= ,β= , C1 = − , (13.68)
CN 2 δ CN C1 2 δ2
this sub-branch becomes the rotating BTZ metric
The transformations above (13.68) may be used in the metric (13.67) to obtain
the studied case in terms of the physical BTZ parameters M and J0 and the
dilaton parameter k. Nevertheless, the expressions are quite involved due to the
r → ρ transformation.
one arrives at the Chan and Mann (1996) metric, Eq. (3.1a), namely
ρN Λ B 2
g = − Aρ1− N/2 + 8 dT 2
(3 N − 2) N
−1
ρN Λ ρ1− N/2 Λ ω2
+ 8 + A − 2 dρ2
(3 N − 2) N B2 A (3 N − 2) N
ω 2 ρ1− N/2
−ω ρ1− N/2 dT dΦ + ρN B 2 − dΦ2 (13.71)
4A
274 Dilaton Field Minimally Coupled to (2 + 1) Gravity
where r of Eq. (3.1a) is replaced by ρ, and B stands for β of Eq. (3.1a) of Chan
and Mann (1996), to avoid confusion.
It calls one’s attention to the sophisticated modification of the function gρρ ,
which under the SL(2, R) remains invariant and could be used (the original in
relation to the static solution) for the adequate evaluation of the energy and
mass.
one gets, modulo minor constants arrangements, the Chan and Mann (1994,
1996) static solution:
N2
ds2 = −C2N U (r)dt2 + dr2 + rN dφ2 ,
4 U (r)
2N Λ N
U (r) = C1 r1−N/2 + rN + 4 Q2 ,
(3 N − 2) 2−N
N
δν] , Ftr = QCN rN/2−2 = −At,r
t r
Fμν = 2Ftr δ[μ
2
N QCN N/2−1
→ At = r ,
2−N
N (2 − N )
Ψ(r) = ± ln (r), 0 < N < 2, N = 2/3. (13.85)
2B
2
λ3 = 0 : V3 = [0, 0, V 3 ], Vμ V μ = V 3 r2 , V3 = S3, (13.88)
278 Dilaton Field Minimally Coupled to (2 + 1) Gravity
therefore, its type is {N, N, S}. For this class of charged dilatons, the energy–
momentum tensor is described by
T μ ν = 2Q2 r−2(1+ Bk ) δ μ φ δ φ ν ,
2
(13.89)
and is algebraically characterized by the following eigenvectors:
2
12 2 B k2 V2
λ1,2 = 0 : V1, 2 = [V , V , 0], N := Vμ V
1 2 μ
= −V CN r 2
L(r) + 2 ,
L
V1, 2 = T1, 2, N1, 2, S1, 2,
2
λ3 = 2 Q2 r−2−2 Bk : V3 = [0, 0, V 3 ], Vμ V μ = V 3 r2 , V3 = S3.
2
(13.90)
Hence depending on the sign of the norm N one will have spacelike, N > 0, null,
N = 0, or timelike, N < 0, eigenvectors: V1, 2 = {S, N, T}. Therefore, in the
case of a double root one may choose different vector components determining,
for instance, one spacelike vector S and the other timelike T or null N one.
Therefore, for the Maxwell energy tensor one may have the algebraic types:
{2S, S}, {2N, S}, {2T, S} and {(S, T ), S},{(T, T ), S},. . .
The Cotton tensor amounts to
C μ ν = C 1 3δμ tδφ ν + C 3 1δμ φδtν ,
1
B C1 k 2 4 − Bk 2 r−Bk + 16 Q2 2 + Bk 2 r−3/2 Bk
2 2
C 13 = 2
32 CN r
CN
r−Bk /2 C1 2 B 2 k 4 4 − Bk 2
2
C 31 = −
32 r4 Bk 2
+12 r−Bk Q2 C1 Bk 2 4 + Bk 2
2
+4 B 2 r2−Bk Λ C1 k 4 + 64 r−3 Bk /2 Q4 2 + Bk 2
2 2
2
2 2 + Bk
+64 Br2−3 Bk /2 Λ Q 2 k 2 . (13.91)
(4 − Bk 2 )
Depending on the sign of the norm N one has spacelike, N > 0, null, N = 0,
or timelike, N < 0, eigenvectors, denoted by: V1, 2 = {S, N, T}. There-
fore, when the eigenvalues are real, the Cotton tensor allows for types I :
{2S, S},{2N, S},{2T, S} and {(S, T ), S}, {(T, T ), S},. . .
In the spacetime region where C 1 3 and C 3 1 are of opposite signs, one of the
eigenvalues becomes imaginary and the following scheme arises
V2
λ1 = 0 : V1 = [0, V, 0], Vμ V μ = , V1 = S1,
L2 √
C 31
λ2 = λ̄3 = C 13C 31 : V2 = Z = [V , 0, √ 1 V 1 ], V3 = Z̄,
1
(13.93)
C 3
thus, the algebraic type of the Cotton tensor is {S, Z, Z̄}.
t = α T + β Φ,
φ = γ T + δ Φ, Δ := αδ − βγ,
2 dr2
g = − α2 CN 2 rBk L2 − γ 2 r2 d T 2 +
L2
2 2
−2 β α CN 2 rBk L2 − δ γ r2 d T d Φ + δ 2 r2 − β 2 CN 2 rBk L2 dΦ2 ,
r2 Λ Q2
r−B k .
2 2
L(r)2 = rB k /2 C1 + 4 + 4 (13.94)
4−Bk 2 Bk 2
Using the standard notation, this charged rotating dilaton solution is given as
dr2
g = −N(r)2 d T 2 + + K(r)2 (dΦ + W(r)d T )2 ,
L(r)2
N 2 r2 2
N(r)2 = (α δ − β γ) ,
− β2N 2
δ 2 r2
2 r2 Λ Q2 2 B k2
N 2 := CN2
rB k /2 C1 + 4 + 4 = CN r L(r)2 ,
4 − B k2 Bk 2
r2 Λ Q2
B k2 /2
r−B k ,
2
L(r) = r
2
C1 + 4 +4
4 − B k2 Bk 2
K(r)2 = δ 2 r2 − β 2 N 2 ,
δ γ r2 − β α N 2
W(r) = , β = 0 = δ,
δ 2 r2 − β 2 N 2
Ψ = k ln(r),
T r Φ r
Fμν = 2FT r δ[μ δν] + 2FΦ r δ[μ δν] ,
280 Dilaton Field Minimally Coupled to (2 + 1) Gravity
d QCN −B k2 /2
FT r = −AT,r = α Ftr = −α 2 r ,
dr B k2
Ftr = QCN r−1/2 Bk −1
2
,
d QCN −B k2 /2
FΦ r = −AΦ,r = β Ftr = −β 2 r ,
dr B k2
QCN −B k2 /2
At = 2 r . (13.95)
B k2
It is worth noticing that the structure of the electromagnetic energy tensor
amounts to
⎡ 2 ⎤
γ 0 γδ
2 ⎢ ⎥
(Tμν ) = 2 Q2 r−2 Bk ⎢⎣ 0 0 0 ⎦.
⎥ (13.96)
γδ 0 δ2
For the AdS (Λ = 1/l2 )–black hole branch, the
√ constants appearing√in the struc-
√
tural functions
√ will be replaced by C N → 1 − 2 k 2 = B, k = 1 − B 2 / 2,
0 < k < 1/ 2, 1 > B > 0, thus
N2B −4+4 B 2
N (r)2 = NB2 = r2 r−2 B C1 B 2 + Λ , L(r)2 = L2B =
2
r . (13.97)
B2
+4 Λ CN 2 α β Bk 2 − 4 r2
2 2
−64β α r−1/2 Bk Q4 CN 2 Bk 2 −4 +β αB 2 k 4 C1 2 CN 2 r1/2 Bk Bk 2 −4
2
13.8 Stationary Charged Dilaton Generated via SL(2, R) 281
+4 β α Bk2 C1 CN 2 Q2 Bk 2 − 4 Bk 2 − 8 ,
β 2 CN 2 Q2 Bk 2 − 4
− β 2 CN 2 r( Bk −4)/2 C1 Bk 2 − 4 ,
2
D = P −4
Br2 k 2
(13.100)
Λ β 2 CN 2 Bk 2 + δ 2 Bk 2 − 4 δ 2
M0 = 2 M Δ
δ 2 Bk 2 − 4 δ 2 + 4 β 2 CN 2 Λ
Bk 2 δ γ + 4 β α CN 2 Λ − 4 δ γ
−J0 2 2 ,
δ Bk − 4 δ 2 + 4 β 2 CN 2 Λ
(13.101)
2
],
Δ L CN Δ L2 CN
Vμ V μ = 0, V2 = N2,
δV2 − Bk2 /2 γ V
2
λ3 = −Q r−1−Bk : V3 = [r− Bk /2
2 2
, V 2
, −r ],
Δ L2 CN Δ L2 CN
μ
Vμ V = 0, V3 = N3. (13.103)
(T μ ν ) = −2 ⎢ 0 0 ⎥
Δ ⎣ 0 ⎦, (13.104)
−α γ 0 −α δ
and is algebraically characterized by:
where parenthesis is used to stand out the multiplicity of the root under con-
sideration. The matrix of the Cotton tensor for the rotating charged solution is
given by
⎡ ⎤
−α β C 3 1 +γ δ C 1 3 −β 2 C 3 1 +δ 2 C 1 3
0
⎢ Δ Δ
⎥
⎢ ⎥
(C μ ν ) = ⎢ 0 0 0 ⎥, (13.106)
⎣ ⎦
− −α C 1Δ+γ C 3 0 − −α β C Δ
2 3 2 1 3 1
1 +γ δ C 3
where
1
B C1 k 2 4 − Bk 2 r−Bk
2
C 13 =
32 CN r2
+16 Q2 2 + Bk 2 r−3/2 Bk ,
2
CN
r−Bk /2 C1 2 B 2 k 4 4 − Bk 2
2
C 31 =− 4 2
32 r Bk
+12 r−Bk Q2 C1 Bk 2 4 + Bk 2
2
+4 B 2 r2−Bk Λ C1 k 4 + 64 r−3 Bk /2 Q4 2 + Bk 2
2 2
2
2−3 Bk2 /2 2 2 2 + Bk
+64 Br ΛQ k . (13.107)
(4 − Bk 2 )
In this case the Cotton tensor is Type I with particular eigenvector subclasses:
{T, T, S}, {T, S, S}, . . . , {S, S, S}.
For one real and two complex conjugate eigenvalues, one has
V2
λ1 = 0 : V1 = [0, V, 0], Vμ V μ = , V1 = S1,
L2
λ2 = λ̄3 = C 13C 31 :
√
V 1 −α β C 3 1 + γ δ C 1 3 − C 3 1 C 1 3 Δ
V 2 = Z = [V , 0, V ], V = −
1 3 3
,
δ2C 1 3 − β 2 C 3 1
V 3 = Z̄. (13.109)
Thus, the algebraic type of the Cotton tensor is Type I: {S, Z, Z̄}.
T r Φ r
Fμν = 2FT r δ[μ δν] + 2FΦ r δ[μ δν] , FT r = α Ftr , FΦ r = β Ftr ,
Ftr = QCN r−1−B k
2
/2
. (13.111)
Switching the rotation β = 0, the above-presented metric reduces to the Chan–
Mann charged dilaton metric.
A stationary cyclic symmetric black hole solution for a scalar field non-minimally
coupled to (2+1)-dimensional gravity is derived and analyzed; its quasi-local
momentum, energy and mass are evaluated. They are asymptotically similar to
the ones of the rotating anti–de Sitter black hole solution with a contribution
due to the angular momentum parameter. The algebraic structure of the Ricci,
energy-momentum, and Cotton tensors is established.
Ψ2 = ζ R Ψ, (14.2)
where = ∇ ∇μ stands for the Laplace–Beltrami operator.
μ
lω l
φ = −√ T+√ Φ,
l −ω
2 2 l − ω2
2
ρ = ρ, (14.4)
one arrives at the metric
2
l − ω 2 ρ3 − 3 l2 ρ m2 − 2 l2 m3 2 ω m2 (3 ρ + 2 m)
g=− dt − 2 dt dφ
(l − ω ) ρ l
2 2 2 (l2 − ω 2 ) ρ
3 ω 2 ρ m2 + 2 ω 2 m3 + l2 − ω 2 ρ3 2 ρ l2 2
+ dφ + 2 dρ . (14.5)
(l2 − ω 2 ) ρ (ρ − 2 m) (ρ + m)
Choosing the parameters as
lR− lJ M 1/2
ω=− =− , R± = M l ± M 2 l2 + J 2 , m = l, (14.6)
J R+ 31/2
one brings (14.5) to a representation close to the one of the BTZ solution at
spatial infinity ρ → ∞, namely
dr2
g = −N (ρ)2 dt2 + + K(ρ)2 (dφ + W (ρ)dt)2 ,
L(ρ)2
ρ2 M 3/2 l
L(ρ)2 = − M − 2 ,
l2 33/2 ρ
288 Scalar Field Non-Minimally Coupled to (2+1) Gravity
1 M 1/2 l2 R− L(ρ)
K(ρ)2 = ρ2 − lR− − 3/2 , N (ρ) = ρ ,
2 3 ρ K(ρ)
J M 1/2 l
W (ρ) = − 1 + 2 ,
2 K(ρ)2 33/2 ρ
√
2 Ml
Ψ =8 √ √ . (14.7)
κ 3ρ + M l
1/2
This solution represents a black hole with horizon at ρ = 2 m = 2 M
31/2
l.
(14.12)
and is algebraically characterized by the following eigenvectors:
2
ρ3 + m3 3 3 μ (V 3 ) ρ2 l2 − ω 2
λ1 = 2 2 3 : V1 = [ω V , 0, V ], Vμ V = , V1 = S1,
l ρ l2
2 ρ3 − m3 V1ω
λ2,3 = 2 3
: V2, 3 = [V 1 , V 2 , 2 ],
l ρ l
2 4 2 2
μ −(V ) (ρ − 2 m) (ρ + m) l − ω 2 + (V 2 ) ρ2 l6
1 2
Vμ V = 2 , (14.13)
l4 ρ (ρ − 2 m) (ρ + m)
290 Scalar Field Non-Minimally Coupled to (2+1) Gravity
hence depending√on the sign of this norm one will have spacelike (V 1 = 0),
2
null (V 2 = ± l2 − ω 2 (ρ + m) (ρ − 2 m) V 1 /(l3 ρ)), or timelike (V 2 = 0)
eigenvectors, and also when
√ 2
1 δ l − ω (ρ + m) (ρ − 2 m) V
1
2 2 V1ω
V2, 3 = [V , , ] = {S, N, T}, (14.14)
l3 ρ l2
where δ may assume the values greater than unit δ > 1, equal to unit δ = 1
and less than unit δ < 1, determining correspondingly spacelike, null, or timelike
eigenvectors. Thus, in the case of a double root one may choose different vec-
tor components determining, for instance, one spacelike vector S and the other
timelike T or null N one. Therefore, one may have the algebraic Ricci types:
{S, 2S}, {S, 2N }, {S, 2T } and {S, (S, T )}, {S, (T, T )}, . . . , {S, (T, N )}.
For the energy–momentum tensor matrix (T μ ν )
⎡ ⎤
(l2 m2 +2 ω2 m2 +ω2 ρ2 −l2 ρ2 )m 3
κ l2 ρ2 (ρ+m)(l2 −ω 2 ) 0 − κ ρ2 (ρ+m)(l
3ω m
2 −ω 2 )
⎢ ⎥
⎢ ⎥
⎢ 0 (m−ρ)m
0 ⎥
⎢ ρ2 κ l2 ⎥
⎣ ⎦
3ω m3 (2 l 2
m +ω m2 −ω 2 ρ2 +l2 ρ2 )m
2 2
κ l2 ρ2 (ρ+m)(l2 −ω 2 ) 0 − κ l2 ρ2 (ρ+m)(l2 −ω 2 )
Therefore, as in the case of the algebraic structure of the Ricci tensor, one may
distinguishes spacelike (V1 = 0), null
2
V2 = ± l2 − ω 2 (ρ + m) (ρ − 2 m) V1 /(l3 ρ),
for δ > 1, = 1, < 1. The same comment about the double root and its vectors
applies in this case. Consequently, one may have the algebraic energy–momentum
types: {S, 2S}, {S, 2T }, {S, 2N } and {S, (S, T )}, . . . , {S, (T, N )}.
14.2 Stationary Black Hole for a Non-Minimally Coupled Scalar Field 291
The existence of this relationship between the eigenvectors of the Ricci and
energy–momentum tensors is due to the tensor relation
κ ν (2 ρ − m) ν
Rμ ν + Tμ − δμ = 0.
ρ l2 ρ
For the Cotton tensor matrix (C μ ν )
⎡ ⎤
3ω m3 (−2 ρ3 +3 ρ m2 +2 m3 ) 3m3 (−l2 ρ3 −ω 2 ρ3 +3 ω 2 ρ m2 +2 ω 2 m3 )
− 0
⎢ 2l2 ρ6 (l2 −ω 2 ) 2l2 ρ6 (l2 −ω 2 ) ⎥
⎢ ⎥
⎢ 0 0 0 ⎥,
⎢ ⎥
⎣ ⎦
3m (−l ρ +3 l ρ m +2 l m −ω ρ )
3 2 3 2 2 2 3 2 3
3ω m 3
(−2 ρ3
+3 ρ m2 +2 m3 )
− 2l4 ρ6 (l2 −ω 2 ) 0 2l2 ρ6 (l2 −ω 2 )
the characteristic equation allow for one real and a pure imaginary eigenvalues,
therefore, the Cotton tensor is characterized algebraically by
(V2 )2 ρ l2
λ1 = 0 : V1 = [0, V2 , 0], Vμ V μ = 2 , V1 = S1,
(ρ − 2 m) (ρ + m)
ρ (ρ − 2 m) (ρ + m) m3
3i
λ2 = λ̄3 = :
2ρ5 l3
1 m3 ω 6 ρ3 − 9 m2 ρ − 6 m3 − 2 λ l2 ρ6 (l2 − ω 2 ) 1
V2 = Z = [V , 0, V ],
3 [(l2 + ω 2 )ρ3 − ω 2 m2 (3 ρ + 2 m)] m3
V3 = Z̄. (14.17)
Thus, the algebraic type of the Cotton tensor is given by Type I: {S, Z, Z̄}.
15
Low-Energy (2+1) String Gravity
1
Gμν − gμν (−2Λ + U ) + 2∇μ ∇ν Φ − 2gμν ∇2 Φ + 2gμν (∇ Φ)2
2
1 1 1
−2 (Fμα Fνβ g αβ − gμν F 2 ) − (Hμαβ Hν αβ − gμν H 2 ) = 0, (15.2a)
4 4 6
1 1 dU
4∇2 Φ − 4(∇ Φ)2 + R − 2 Λ − F 2 − H 2 + U (Φ) − = 0, (15.2b)
6 2 dΦ
1 n − 2 dU
2∇2 Φ − 4(∇Φ)2 − 2Λ + F 2 + H 2 + U + = 0, (15.4b)
6 4 dΦ
and choose
2 n−2
n σ − 2 σ − 2Φ = 0 → σ = Φ, Φ = σ. (15.8)
n−2 2
Substituting these relations in the action above (15.6), one arrives at
n √ 4
S= d x −g R − (∇Φ)2 − 2e4 Φ/(n−2) Λ + V (Φ)
n−2
−4 Φ/(n−2) 2 1 −8 Φ/(n−2) 2
−e F − e H , (15.9)
12
where the divergence has been dropped from this action
√ √
−2(n − 1) −gg μν ∇ν ∇μ σ = −2(n − 1)( −gσ ,ν ),ν ,
and denoted
e2σ U (Φ) = V (Φ).
2 −8Φ/(n−2) 2 n − 2 d V
8 ∇ν ∇ν Φ − 8 Λe4Φ/(n−2) + 4 e−4Φ/(n−2) F 2 + e H + = 0,
3 2 dΦ
(15.10b)
a −8Φ/(n−2)
∇λ e−4Φ/(n−2) F λ + e Fαβ H αβ = 0. (15.10d)
12
Replacing in (15.10a) the scalar curvature R,
4 n − 4 −4Φ/(n−2) 2 1 n − 6 −8Φ/(n−2) 2
R= (∇Φ)2 + e F + e H
n−2 n−2 12 n − 2
n
− (−2Λ e4Φ/(n−2) + V ),
n−2
one rewrites (15.10a) as
15.2 Dynamical Equations in (2+1) String Gravity 295
2Λ 1 4
Rμν = gμν e4Φ/(n−2) − gμν V (Φ) + ∇μ Φ ∇ν Φ
n−2 n−2 n−2
1
+ e−4Φ/(n−2) 2 Fμσ Fν σ − gμν F 2
n−2
1 2 1
+ e−8Φ/(n−2) 2 Hμαβ Hν αβ − gμν H 2 . (15.11)
4 3n−2
2 −8Φ 2 1 d V
8 ∇2 Φ − 8 Λe4Φ + 4 e−4Φ F 2 + e H + = 0, (15.13b)
3 2 dΦ
∇σ e−8Φ H αβσ = 0, (15.13c)
−4Φ λ a −8Φ
∇λ e F + e Fαβ H αβ = 0. (15.13d)
12
One can recover the string dynamical equations by using the conformal inverse
relations (see Eisenhart, 1966),
˜ 2 , e−2 σ σ̃ ,k = σ ,k ;k − (n − 2)(∇σ)
σ,i;j = σ̃,i;j + 2σ,i σ,j − g̃ij (∇σ) ˜ 2,
;k
W W̃ ,k ˜ 2 ,
Rij = R ij + g̃ij σ̃;k + (n − 2) σ̃,i;j + σ,i σ,j − g̃ij (∇σ)
W W̃
e−2 σ R = R + 2(n − 1)σ̃;k
,k ˜ 2,
− (n − 1)(n − 2)(∇σ) (15.14)
where tilde is used to denote that covariant differentials are constructed with Γ̃s
or contravariant tensor components are built with g̃ μν . For σ = 2Φ and n = 3
one gets
˜ 2,
Φ,μ;ν = Φ̃,μ;ν + 4Φ,μ Φ,ν − 2g̃μν (∇Φ)
Φ;α ;α = e4Φ Φ̃;α
;α ˜ 2 ,
− 2(∇Φ)
˜ 2
;α + 4Φ,μ Φ,ν − 4g̃μν (∇Φ) ,
Rμν = R̃μν + 2Φ̃,μ;ν + 2g̃μν Φ̃;α
˜ 2 ,
;α − 8 (∇Φ)
R = e4Φ R̃ + 8 Φ̃;α (15.15)
296 Low-Energy (2+1 ) String Gravity
and conformally transforming the above dynamical equations (15.13), using rela-
tions (15.15), one gets the barred dynamical equations of the (2+1) string theory
under consideration
−2Φ ˜ 1 2
S = d x −g̃e
3
R̃ − 2Λ + 4(∇Φ) − F̃ − H̃ + U (φ) ,
2 2
12
(15.16a)
1
R̃μν + 2∇μ ∇ν Φ − 2Fμα Fνβ g̃ αβ − Hμαβ Hνγλ g̃ γα g̃ λβ
4
H̃ 2 ˜ ˜
+ g̃μν −2Λ + U + F̃ +
2
+ 2∇ Φ − 4(∇Φ)
2 2
= 0, (15.16b)
6
2∇ ˜ 2 + F̃ 2 − 2Λ + 1 H̃ 2 + U (Φ) + 1 d U = 0,
˜ 2 Φ − 4(∇Φ) (15.16c)
6 4 dΦ
˜ σ (e−2Φ H̃ μνσ ) = 0,
∇ (15.16d)
t = α T + β φ, Δ := αδ − βγ.
x = γ T + δ φ,
which yields
(r − M ) rM − Q2 rk l2 − ω 2
g=− 2 dT 2
2 l (−ω 2 rM + ω 2 M 2 + rM − Q2 )
M −1 Q2 −1 k 2 dr2
+ (1 − ) (1 − )
r Mr 16
l2 kr −ω 2 rM + ω 2 M 2 + rM − Q2
+
2M (l2 − ω 2 )
2 2
ω −l M r + l2 M 2 + rM − Q2
× dφ − 2 dT ,
l (−ω 2 rM + ω 2 M 2 + rM − Q2 )
Q 1 1 k
Hrtx = , Φ = − ln r − ln . (15.20)
r2 2 4 2
The evaluation of the quasilocal mass, energy and momentum is done using the
Brown–York approach; this yields
2
2 M −Q
2
J(r → ∞) = 2ω l 2 , (15.21a)
(l − ω 2 ) M
1 (M − ω Q) (M + ω Q)
(r → ∞) = −1/2 − 1/4 − 0 , (15.21b)
rπ l2 M l2 π (ω − 1) (ω + 1) r2
√
2 1 − ω2
E(r → ∞) = √ − 2π K(r)0 , (15.21c)
l2 − ω 2
298 Low-Energy (2+1 ) String Gravity
1 1 M0 l
0 = − + . (15.21e)
π l 2 π r2
yields the string solution derived in Horne and Horowitz (1992); see also Section
15.3. Dropping primes, it becomes
S M 2 Q2 M −1 Q2 −1 l2 dr2
g = −(1 − )dt + (1 − )dx2 + (1 − ) (1 − ) ,
r Mr r Mr 4r2
Q 1
Bxt = , φ = − ln (r l), (15.36)
r 2
where M = r+ 2
/l and Q = J/2.
The identification of the functions appearing in the action (15.1) and equations
(15.2) with the ones of Horne and Horowitz (1992),
√ 1
S= d 3 x −geφ R − 2Λ + (∇Φ)2 − H 2 , Λ = −4/kHH , (15.37)
12
requires that φ = −2ΦHW , and kHH = 2 kHW , kHW = l2 , Λ = −2/l2 , where the
subscripts are in correspondence with the initial of the author’s family name.
where the parameters ΛCM , b are arbitrary at this stage, Ψ is the massless
minimally coupled scalar field, R is the scalar curvature, and F 2 = Fμ ν F μ ν
the electromagnetic invariant. The variations of this action yield the dynamical
equations
B
Rμν = ∇μ Ψ ∇ν Ψ − 2gμν e b Ψ ΛCM + 2 e−4 a Ψ Fμ α Fν α − gμ ν F 2 ,
2
B μ
∇ ∇μ Ψ + b e b Ψ ΛCM + 2 a e−4 a Ψ F 2 = 0,
2
∇μ e−4 a Ψ Fμ ν = 0.
15.5 Chan–Mann String Solution 301
where ω stands for the rotation parameter, this metric can be written as
L2 1 − ω 2 /l2
gS = − dτ 2
r2 1 − ω 2 + L2 /r2
2
1 − ω 2 L2 /r2 ω 1 − ω 2 /l2 dr2
+ dθ − dτ + 2, (15.43)
1 − ω /l
2 2 l 1 − ω L /r
2 2 2 2 L
and is endowed with four parameters: the mass M = −C1 , charge Q, rotation
ω, and cosmological constant Λs = ±1/l2 . It is worth pointing out that string
solutions can be found in various dimensions; see, for instance, the works of
Witten (1991); Mandal et al. (1991); Maki and Shiraishi (1993), among others.
16
Topologically Massive Gravity
In this part of the book we deal with exact solutions to the Einstein topologically
massive gravity equations. However, since the material to be included only rep-
resents twenty per cent of the whole book’s subject matter, we prefer to present
this content in the form of chapters devoted to a concise but, we hope, complete
(in the range of the possibilities) exposition of the exact solutions in topologically
massive gravity (TMG) in three dimensions in the case of vacuum in the presence
of a cosmological constant Λ of both signs. Thus, this chapter has an introduc-
tory character, while the next three chapters deal with very specific families of
the existing Petrov-type Cotton solutions in TMG.
The extension of the 3D Einstein gravity to other field theories to provide
them with certain degrees of freedom (a massive spin 2 graviton) was proposed
more than thirty-five years ago by Deser, Jackiw and Templeton; see Deser et al.
(1982a): “Three-dimensional massive gauge theories,” which is known as the
TMG. It includes a Chern–Simons term constructed from connections (Cot-
ton tensor) with broken parity invariance; see also its extended version, with
a detailed analysis, in Deser et al. (1982b). A cosmological constant was intro-
duced in these three-dimensional theories by Deser (1984). The sign in front of
the curvature scalar has been chosen opposite to the standard one of the 4D
Einstein gravity to yield, in the limit of the linearized theory, to the existence of
a spin 2 graviton with positive energy. A modern treatment of these aspects in
cosmological massive gravity appeared recently in Carlip et al. (2009).
As far as the determination of exact solutions to vacuum equations with a
cosmological constant in TMG is concerned, through this long period, various
classes of solutions have been reported apart from the trivial Minkowski flat
spacetime, the de Sitter (A)dS3 cosmology, and the BTZ black hole – confor-
mally flat (zero Cotton tensor) solutions – for any value of the coupling mass
parameter μ. Among these classes one can mention homogeneous spaces, pp-
waves metrics, cosmological solutions, and Kerr–Schild metrics. Fortunately, a
systematic up-to-date classification of the existing families of vacuum solutions
304 Topologically Massive Gravity
in TMG has been done recently by Chow, Pope, and Sezgin (CPS): see Chow
et al. (2010a), referred to as CPSa, and Chow et al. (2010b), referred to as CPSb,
for Kundt spacetimes. Thus, I shall take advantage of this progress and incor-
porate their results in this book; I shall work out the derivation of the various
classes reported in the literature using a presentation pattern similar to CPS and
exhibit their characterization by means of the Garcı́a et al. (2004) Petrov–Segré
classification of the Cotton tensor and other relevant features.
There are other kinds of generalizations of the TMG that pursue other
purposes. For instance, implications in the AdS3 /CF T2 correspondence. The
construction of a new massive gravity theory, NMG, preserving the parity invari-
ance and possessing a single massive spin 2 field has been achieved by Bergshoeff,
Hohm and Townsend (see Bergshoeff et al., 2009); this theory is constructed
with combinations of quadratic invariants F (R, Rμν Rμν ). The generalization of
this last theory took place with the publication of “minimal massive gravity,”
which has positive energy spin 2 graviton and positive central charges for the
“asymptotic AdS-boundary algebra.” These theories experience nowadays a lot
of activity; many researchers are engaged in their developments, and the list of
references on these topics is quite vast: a Living review concerned these devel-
opments not only in three dimensions but also in higher dimensions, “Massive
gravity,” has been published recently by de Rham (2014).
R = 6 Λ, (16.3a)
1
Sα β + Cα β = 0, (16.3b)
μ
16.2 Exact Vacuum Solutions of TMG with Λ 305
Later Nutku (1993), in the same Bianchi-type VIII framework but now
with a cosmological constant, derived timelike biaxially squashed and spacelike
spacetimes.
All these squashed spacetimes are Petrov type D, and exhibit the constant
scalar invariant (CSI) property, i.e., all scalar polynomial curvature invariants
are constant.
2. Details in Section 17.5. Timelike biaxially squashed AdS3 spacetime can
be regarded as a timelike fibration over H, or as a stationary cyclic symmetric
spacetime decomposed with respect to the timelike Killing vector. This spacetime
is Petrov type D.
3. Details in Section 17.6. Spacelike biaxially squashed AdS3 spacetime can
be regarded as a spacelike fibration over AdS2 , or as a stationary cyclic sym-
metric spacetime decomposed with respect to the spacelike Killing vector. This
spacetime is Petrov type D.
4. Details in Chapter 18. In the spirit of CPS, the AdS3 pp-wave solutions are
generalizations of pp-waves to include a negative cosmological constant; these
solutions are Petrov type N. Nevertheless, we prefer to restrict the name of
pp-waves to those solutions possessing a covariantly constant kμ;ν = 0 null con-
gruence k as they are defined in the standard 4D Einstein gravity. In TMG there
is room for a wider class of solutions with a non-covariantly constant kμ;ν = 0
null congruence k of Cotton type N, to be denoted as TN-wave solutions.
5. Details in Chapter 19. In general, Kundt spacetimes are defined by metrics
admitting a null geodesic vector field that is shear-free, twist-free, and expansion-
free. Therefore, in three dimensions, because the twist and shear trivially vanish,
a “Kundt spacetime is simply one that admits an expansion-free null geodesic
congruence.” These spacetimes allow for Petrov type II, D, III, N, and O solu-
tions. Moreover, all the existing Kundt solutions of TMG are CSI – constant
scalar invariants – spacetimes for which all polynomial scalar (Ricci) curvature
invariants are constant; these late metric structures were studied for the first time
by Coley, Hervik, and Pelavas (see Coley et al., 2006 and Coley et al., 2008) in
three dimensions in a wider context.
Many of these solutions “have been independently rediscovered as solutions of
TMG several times. The literature is rather fragmented, using different coordi-
nate systems that are not obviously related.” Thus, adopting the aim of CPS,
a review establishing the relationships, if any, between families of solutions
becomes pertinent; the metrics are presented in the same coordinates as the orig-
inal literature, although certain rearrangements have been made. Each metric is
then transformed to a canonical form.
17
Bianchi-Type (BT) Spacetimes in TMG; Petrov
Type D
[KA , KB ] = −C N A B KN , C N A B = −C N B A ,
where T L̃ D is inverse to T A M̃ .
Returning to the classification of 3D homogeneous spaces, to describe these
spaces one uses the left invariant 1–forms σ α , α = 1, 2, 3 which fulfill the Maurer–
Cartan structure equations,
1 α
dσ α = C βγ σ β ∧ σ γ (17.1)
2
where the structure constants C α βγ = −C α γβ obey the Lie identity
C α β[μ C β νγ] = 0. (17.2)
Each Bianchi type is characterized by a specific set of structure constants,
which in turn determines the corresponding group. The classification of homo-
geneous 3D Bianchi-type spaces can be found also in the textbook by Stephani
(1990).
Following Nutku and Baekler (1989), the triad description of the homogeneous,
anisotropic cosmological models in TMG keep a close parallelism to the Maurer–
Cartan structure description. By means of a linear transformation between the
1-forms σ a and the triad 1-forms ω a
ω a = λa α σ α ,
λa α = λa δ a α no sum in a, λa = constants, (17.3)
one determines the 3D spacetime metric with Lorentzian signature defined
through
ds2 = ηab ω a ω b , ω a = λa σ a , no sum in a,
ηab = diag(−1, 1, 1). (17.4)
From the first Cartan structure equations
dω a + Γa b ∧ ω b = 0, (17.5)
one determines the connection 1-forms, Γa b = Γa bs ω s , which, used in the second
Cartan structure equations
1 a
Θa b := dω a b + ω a s ∧ ω s b = R bcd ω c ∧ ω d , (17.6)
2
17.2 Nutku–Baekler–Ortiz “Timelike” BT VIII Spacetime 309
In vacuum TMG, the vanishing of the curvature scalar R requires the fulfillment
of the relations
λ0 ± λ1 ∓ λ2 = 0 → λ0 + λ1 + λ2 = 0; (17.12)
this last relation can be always adopted because of the triad {ω (a) } is defined up
the signs of λa . The equations E (a) (b) = 0 lead, taking into account the relation
above, to the condition
λ0 2 + λ1 2 + λ2 2 λ1 2 + λ2 λ1 + λ2 2
μ=− =2 . (17.13)
λ0 λ1 λ2 λ1 λ2 (λ1 + λ2 )
These Bianchi type VIII solutions are called triaxially squashed spacetimes of
the TMG theory; it is apparent that they are unique to their class; they were
derived in Nutku and Baekler (1989), Eq. (4.1), and Ortiz (1990) (type(a) with
a = 0). They are characterized by a Petrov type D Cotton tensor: its eigenvectors
are given by
L1 = −L+ λ1 ; VI α = ω (1) α , spacelike,
L2 = −L+ λ2 ; VII α = ω (2) α , spacelike,
L0 = −L+ λ0 ; VIII α = ω (0) α , timelike, (17.14)
where the positive L+ is given by
μ λ0 2 + λ1 2 + λ2 2 λ1 2 + λ2 λ1 + λ2 2
L+ = −2 =2 = 4 . (17.15)
λ0 λ1 λ2 λ0 2 λ2 2 λ1 2 2
(λ1 + λ2 ) λ2 2 λ1 2
In terms of its eigenvectors, the Cotton tensor is expressed as
Cα β = L+ λ0 ω (0) α ω (0) β − L+ λ1 ω (1) α ω (1) β − L+ λ2 ω (2) α ω (2) β , (17.16)
hence, its constant triad components C(a) (b) = ω α (a) Cα β ω β (b) are
C(a) (b) = L+ λ0 δ (0) (a) δ (0) (b) − L+ λ1 δ (1) (a) δ (1) (b) − L+ λ2 δ (2) (a) δ (2) (b) . (17.17)
where the left invariant 1-forms σ (a) fulfill the Maurer–Cartan equations
λ0 + λ1 + λ2 = 0, (17.30)
C α β V β A − L VA α = 0, det(C α β − L δ α β ) = 0,
which lead to
hence, its constant triad components C(a) (b) = ω α (a) Cα β ω β (b) are
C(a) (b) = λ0 L− δ (0) (a) δ (0) (b) − λ1 L− δ (1) (a) δ (1) (b) − λ2 L− δ (2) (a) δ (2) (b) . (17.35)
Its substitution into the remaining TMG equations leads to a single constraint,
namely
μ 1
λ2 μ2 − 27 λ2 Λ + 6 μ = 0 → λ2 = −6 , λ1 = 3 , (17.38)
μ2 − 27 Λ μ2 − 27 Λ
hence, the Nutku solution counterpart can be given by the metric
9 4μ2
ds2 = 2 −dθ2 + cos2 θdφ2 + 2 (dψ + sin θdφ)2 . (17.39)
μ − 27Λ μ − 27Λ
Although this solution is not explicitly given in Nutku (1993), it can straightfor-
wardly be derived by complex coordinate transformations from Nutku (1993),
Eq. (18). This solution is characterized by a Petrov type D Cotton tensor.
scalar curvature (17.29); as the first step one uses the equality λ2 = λ1 , next its
substitution in R gives
where σ (a) are Bianchi III left invariant 1-forms that satisfy
dσ (1) = ασ (3) ∧ σ (1) + σ (3) ∧ σ (2) , dσ (2) = ασ (3) ∧ σ (2) + σ (3) ∧ σ (1) ,
dσ (3) = 0, C (1) (3)(1) = α = C (2) (3)(2) , C (1) (3)(2) = 1 = C (2) (3)(1) , (17.44)
which corrects the relations in the text line following Eqs. (A.28) and (A.56)
of Chow et al. (2010a), denoted for short CPSa. The scalar curvature for the
vacuum TMG is given by
2 λ1 2 λ2 2 + 12 λ1 2 α2 λ2 2 − λ1 4 − λ2 4
R=− =0
2 λ1 2 λ2 2 λ3 2
from which one isolates α2
2 2
1 (λ1 − λ2 ) (λ1 + λ2 )
α2 =
12 λ1 2 λ2 2
with two possible branches
√
1 3 (λ1 + λ2 ) (λ1 − λ2 )
α=± .
6 λ2 λ1
17.4 Nutku–Baekler–Ortiz Solutions of Bianchi Type III 315
3 λ1 2 + λ2 2
μ= .
2 λ3 λ1 λ2
Substituting μ into E z y one gets
15 λ1 2 λ2 4 − 15 λ1 4 λ2 2 − λ2 6 + λ1 6 = 0,
with solutions
√
λ1 = 2 + 3 λ2 , (17.45a)
or
√
λ1 = 2 − 3 λ2 . (17.45b)
one obtains the metric given in Chow et al. (2010a) as (A.28), namely
√ 2 2
36 2 36
ds2 = λ2 −(2 + 3)2 σ (1) + σ (2) + 2 σ (3) = 2 dθ2
μ μ
√
+ λ2 e2θ −(2 + 3)2 (cosh θ dx + sinh θ dy)2 + (sinh θ dx + cosh θ dy)2
√
3 + 2 3 2 √ 2
=− λ 3(dx − dy) + 2e2θ (dx + dy)
6
√
3 + 2 3 2 4θ 36
+ λ e (dx + dy)2 + 2 dθ2 . (17.47b)
6 μ
316 Bianchi-Type (BT ) Spacetimes in TMG; Petrov Type D
which coincides with the solution given by Eq. (A.11) of Chow et al. (2010a)
with Λ = 0, CPSa(A.11).
√
2 3 − 3 2 √ 2
= λ 3(dx − dy) − 2e2θ (dx + dy)
6
√
2 3 − 3 2 4θ 36
− λ e (dx + dy)2 + 2 dθ2 , (17.48b)
6 μ
Table 17.4.1 Bianchi Type (BT) VIII and BT III TMG solutions
with inverse matrix h(a) α , ∂(a) = ∂(a) α ∂ ∂xα = h(a) α ∂ ∂xα , given by
⎡ √ 2 ⎤
μ −27 Λ
0 0
⎢ 3 ⎥
⎢ √ 2 √ 2 ⎥
h(a) = ⎢
α μ −27 Λ
⎢ 3 sinh θ 0 −
μ −27 Λ cosh θ ⎥ .
⎥ (17.68)
⎣ 3 sinh θ ⎦
μ −27 Λ
2
0 0 6μ
These tensors are used to construct the triad tensor (constant) components. In
the case of the Cotton tensor given by coordinate components (17.63), using
Proceeding in the standard way in the search of the eigenvectors of the Cotton
tensor certainly one should arrive at the same result.
cosh θ = x → θ = arccoshx
x = i y, τ = i z, , φ = t, (17.74)
17.5 Timelike Biaxially Squashed Metrics 321
4μ2
λ2 := , (17.82a)
μ2 − 27Λ
explicitly
9
ds2tl = −λ2 (dτ + cosh θ dφ)2 + dθ2 + sinh2 θdφ2 , (17.82b)
μ2 − 27Λ
9 dx 2 dx2 + dy 2
ds2tl = 2 −λ2 (dτ + ) + , (17.82c)
μ − 27Λ y y2
9
ds2tl = 2 −λ2 (dτ + sinh θ dφ)2 + dθ2 + cosh2 θdφ2 . (17.82d)
μ − 27Λ
Another possibility is achieved by transformations of the type cosh = Ar2 + B,
which yields CPSa(A.16), namely
μ 2 r2
ds2tl = −(dt − r dφ)2 + dr2 + F (r)dφ2 ,
3 F (r)
1 2
F (r) = (μ − 27Λ)r4 + k1 r2 + k0 . (17.82e)
36
In the forthcoming subsections the identification of the solutions reported by
different authors is carried out following the pattern given in Chow et al. (2010a),
citing their equation for reference.
Vuorio Solution; Λ = 0
Vuorio (1985) reported the first stationary rotationally symmetric solu-
tion (2.21), CPSa(A.25), to the E-TMG equations, namely:
9 2
ds2 = 2 − (dt + 2(1 − cosh σ) dθ) + dσ 2 + sinh2 σ dθ2 . (17.83)
μ
Carrying out the coordinates transformations and constants redefinitions
t → 2τ + 2φ, θ → −φ, σ → θ, Λ → 0,
Percacci–Sodano–Vuorio Solution; Λ = 0
Percacci et al. (1987) considered stationary solutions possessing a timelike Killing
vector with constant scalar twist; their solution (3.20), CPSa(A.26), is given by
2
2 1 1 1
ds2 = −3 dx2 + e(μ x /3) dx0 + (d x1 )2 + e(2 μ x /3) (d x0 )2 . (17.84a)
3 3
324 Bianchi-Type (BT ) Spacetimes in TMG; Petrov Type D
Clément Solutions; Λ = 0
Clément (1992a), and Clément (1992b) too, considered stationary rotationally
symmetric solutions. His solution (4.4), CPSa(A.30), is given by the metric
μ 2 μ
ds2 = − dt + 2cdθ − 2c cosh( r)dθ + dr2 + c2 sinh2 ( r) dθ2 . (17.85a)
3 3
Making the coordinates transformations and constants redefinitions
μ μ μc
(t + 2c θ) → τ, r → θ , − θ → φ, (17.85b)
6 3 3
dropping prime, one arrives at (17.82b), CPSaA(24), with Λ = 0.
The Clément’s solution (4.5), CPSa(A.31),
μ 2 μ
ds2 = − dt − 2c sinh( r)dθ + dr2 + c2 cosh2 ( r) dθ2 , (17.86a)
3 3
subjected to the coordinates transformations
μ μ μc
t → τ, r → θ , − θ → φ, (17.86b)
6 3 3
dropping prime, gives rise to (17.82d), CPSa (A.12), for Λ = 0.
On the other hand, the solution (4.7) by Clément (1992a) has various possible
choices of signs:
Two choices give
c 2
ds2 = − dt ∓ 2d e±μ r/3 dθ + dr2 + c d e±2μ r/3 dθ2 , (17.87a)
d
which is CPSa(A.32). Subjecting (17.87a) to the coordinates transformations
CPSa(A.32)
c μ√
t → τ, ∓ c dθ → x, e∓ μ r/3 → y, (17.87b)
d 3
one arrives at (17.82c), CPSa(A.11), with Λ = 0.
17.5 Timelike Biaxially Squashed Metrics 325
Nutku Solutions
Nutku (1993) published a black hole solution (25), CPSa(A.21), to the E–TMG
equations given by
2
2J − M 2μ r2 − 3J/μ r2 6
ds2 = − dt − dθ + dr2 + F (r)dθ2 ,
6 2J − M F (r) 2J − M
1 2 1 1 J2
F (r) = (μ − 27Λ)r4 − M r2 + . (17.89a)
36 6 4 μ2
Making the coordinates transformations CPSa(A.22)
√
2J − M 3 1 J 6
t+ √ θ→t, √ θ → φ,
6 2 2J − M μ 2J − M
M 1 J2
− → k1 , → k0 , (17.89b)
6 4 μ2
dropping prime, one arrives at the solution (17.82e), CPSa(A.16).
Moreover, Nutku generalized the Λ zero Vuorio (1985) solution (see 17.5.4)
to include the cosmological constant; the Nutku (1993) solution (3.11), after
relabelling ψN → τ corresponds to (17.82d), CPSa(A.12). On the other hand,
Nutku succeeded in deriving his solution (3.17), which corresponds to (17.82e),
CPSa(A.16), for k1 = 1 and k0 = 0 and θN → φ.
Gürses Solution
In the work by Gürses (1994) dealing with a perfect fluid (5) of constant pressure
and density as solution of the Einstein plus Λ equations, its interpretation is
established as a vacuum TMG plus Λ solution given by CPSa(A.18):
2
√ 3c0 + μ e0 r2 e2 1
ds2 = − a0 dt − √ dθ + 0 r2 ψ(r)dθ2 + dr2 ,
3 a0 a0 ψ(r)
μ2 − 27Λ 2 b1
ψ(r) = r + b0 + 2 . (17.91a)
36 r
Accomplishing the coordinates transformations and constants redefinitions
√ c0 e0
a0 t − √ θ → t , √ θ → φ, b0 → k1 , b1 → k0 , (17.91b)
a0 a0
dropping prime, one arrives at the solution (17.82e), CPSa(A.16).
Clément Solution
Clément (1994) considered a “Killing symmetry reduction procedure to
obtain stationary rotationally symmetric solutions”. Clément’s solution (18),
CPSa(A.23), is given as
2
√ 1 3b 2 μ dρ2 1
ds = −
2
2 adt + √ − ρ dθ + + F (ρ) dθ2 ,
2a μ 3 F (r) 2 a
1 b2
F (ρ) = (μ2 − 27Λ)ρ2 + 4(a − b) ρ + 9 2 . (17.92a)
9 μ
Accomplishing ing the coordinates transformations and constants redefinitions
CPSa(A.24)
√
√ 3b 2 9 b2
2 a t + √ θ → t , ρ → r2 , √ θ → φ, a − b → k1 , → k0 , (17.92b)
μ 2a a 4 μ2
dropping prime, one arrives at (17.82e), CPSa(A.16).
17.6 Spacelike Biaxially Squashed Metrics 327
Anninos–Li–Padi–Song–Strominger Solution
Anninos et al. (2009) reported “warped” AdS3 black hole solutions of TMG; their
solution (3.4), after accomplishing the redefinitions σ → θ, u → φ, becomes the
solution (17.82d), CPSa(A.12).
The solutions reported in this subsection are gathered in the following table:
4 λ2 + 3 Λ λ2 2 − λ1
Eμμ = = 0 → λ1 = 4 λ2 + 3 Λ λ2 2 , (17.94a)
λ2 2
the equation component
μ2 + 9Λ α 1 μ2 − 27 Λ α
Sβα = (δ β − 3 K α Kβ ), K α = ± δ z, (17.97)
9 6 μ
where K α is a spacelike unit vector.
cosh (ρ)
Θ(3) α = [0, 0, 3 ], timelike, (17.99)
μ2 − 27 Λ
hence, the triad tensor components h(a) α can be represented by the matrix
⎡ ⎤
μ
6 μ2 −27 Λ 0 6 μμ2sinh(ρ)
−27 Λ
⎢ ⎥
⎢ √ 21 ⎥
(a)
h β =⎢ ⎢ 0 3 0 ⎥. (17.100)
μ −27 Λ ⎥
⎣ ⎦
0 0 3 √cosh(ρ)
2 μ −27 Λ
Cαβ = C(a)(b) Θ(a) α Θ(b) β = λs Θ(1) α Θ(1) β + λd Θ(2) α Θ(2) β − λd Θ(3) α Θ(3) β ,
μ 2 μ 2
λs = 2 μ + 9 Λ , λd = − μ + 9Λ . (17.104)
9 9
−(X 0 )2 + (X 1 )2 − (X 2 )2 = −L2
one gets ds2 2 in conformally flat coordinates that cover the Poincaré patch
CPSa(A.40):
L2 2
ds2 2 = 2
dρ − dt2 . (17.105d)
ρ
The choice CPSa(A.41):
To facilitate comparison with the solutions reported in the literature, the space-
like squashed (SLS) solution, following Chow et al. (2010a), is presented in
various alternative forms with minor changes from our side, namely, restoring
the negative cosmological constant Λ and the λ constant, explicitly
4μ2
m2 = −Λ, λ2 = . (17.106)
μ2 − 27Λ
These solutions, corresponding respectively to CPSa(A.45), CPSa(A.47), and
CPSa(A.46), are given by
9
ds2sl = λ2 (dz + sinh ρ dτ )2 + dρ2 − cosh2 ρdτ 2 , (17.107a)
μ2 − 27Λ
9
ds2sl = λ2 (dz + cosh ρ dτ )2 + dρ2 − sinh2 ρdτ 2 , (17.107b)
μ2 − 27Λ
9 dt 2 − dt2 + dx2
ds2sl = 2 2
λ (dz + ) + , (17.107c)
μ − 27Λ x x2
one sees that f (u) is redundant, and one can equate it to zero, f (u) → 0. Under
such a choice, the second coordinate transformations
μ2 1 μ2
u → û, + u → v̂ (17.108c)
18 v 18
bring the metric to the form CPSa(A.52)
2
36 dû dv̂ 12 dû
ds2 = − + dy + . (17.108d)
μ2 (û − v̂)2 μ û − v̂
Finally, the coordinate changes
1 1 μ 1
(û + v̂) → t, (v̂ − û) → x, y + log| (v̂ − û)| → −z, (17.108e)
2 2 6 2
leads to (17.107c), CPSa(A.46), with Λ = 0.
are called plane-fronted gravitational waves with parallel propagated null rays
and denoted as pp-waves. They have been known in 4D Einstein relativity since
long ago; Brinkmann (1923) discovered them in 1923; see Kramer et al. (1980),
§21.5. The condition on the Ricci tensor point on the existence of solutions for
null electromagnetic fields, radiation and vacuum, excluding other field solutions,
among them dS–AdS cosmological constant solutions. This fact establishes a
difference with the 3D wave solutions in TMG.
To get an insight of how the wave solutions with a non-vanishing cosmological
constant arise in the framework of the TMG, let us consider the Brinkmann-like
3D Killing wave metric
but if this were the case, the metric could not allow for a cosmological con-
stant even in TMG; this subclass of solutions with vanishing cosmological
constant can properly be named pp-wave solutions; see Section 18.3 for specific
solutions.
For the class of wave solutions with a Killingian null congruence k, such that
kμ;ν = 0, (18.4), all the scalar invariants representing respectively the twist,
shear, and expansion:
1 μ
ω 2 = k μ;ν k[μ;ν] , σ 2 = k μ;ν kμ;ν − (k μ ;μ )2 , θ = k ;μ , (18.6)
2
vanish; these wave solutions can be called type N Killing waves and denoted as
TN-waves instead of pp-waves. This class of TN-solutions is different from type
N Kundt family of metrics, arising as limit of type III Kundt metric, to be treated
in detail in Chapter 19; it is worth pointing out that the null vector for this Kundt
class is a non-Killingian one; thus this class of wave solutions with a shear-free,
twist-free, expansionless, and non-Killingian null congruence can be called type
N non-Killingian waves and denoted TNnK-wave solutions. Having made these
remarks, we can proceed with the search for the corresponding solutions.
The factor F0 (u) can be set equal to the unit, F0 (u) → 1, by a transformation
of the coordinate u, moreover, it is enough to treat the case with m, the branch
of solutions with −m is derivable from the first one simply by replacing there
m → −m. The E v v component yields a dependent equation, which, for the
solutions of (18.8), identically holds.
The only remaining equation is the E v u one, which for the F (ρ) solution (18.8)
becomes
∂3 ∂2 ∂
3
H (ρ, u) + (μ − 3 m) 2 H (ρ, u) − 2 m (μ − m) H (ρ, u) = 0. (18.9)
∂ρ ∂ρ ∂ρ
There arise three different cases depending on the μ(m) relation.
18.2 AdS3 Non-Covariantly Constant TN-Waves 335
Case μ = ±m:
The general solution of (18.9) for H(ρ, u) is given by
thus one concludes that the shear tensor and the divergence vanish, k(α;β) = 0
and k α ;α = 0, while the rotation tensor (twist) is different from zero k[α;β] =
0, although the shear, rotation, and expansion invariant scalars are all zero.
Consequently, this geeodesic null vector is also a Killing vector k(α;β) = 0. With
the help of this null Killing vector k, one may write the Cotton tensor (18.11)
as
1
Cαβ = μ μ2 − m2 e−(μ+3 m)ρ f3 (u) kα kβ , k α = δvα , (18.13)
2
which manifestly shows the Petrov type N character of this Cotton tensor pos-
sessing a triple zero eigenvalue, Cαβ k α = 0.
Case μ = m:
For this branch of solution, the E v u equation (18.9) becomes
∂3 ∂2
H (ρ, u) − 2 m H (ρ, u) = 0 (18.14)
∂ρ3 ∂ρ2
hence it general solution can be given as
or, using the null vector k from (18.3), one can represent it as
Case μ = −m:
For this branch of solutions, the E v u equation (18.9) amounts to
∂3 ∂2 ∂
H (ρ, u) − 4 m H (ρ, u) + 4 m2 H (ρ, u) = 0, (18.18)
∂ρ3 ∂ρ2 ∂ρ
336 Petrov Type N Wave Metrics
This type N-wave gravitational field is characterized by the Petrov type N Cotton
tensor
CPSa(A.64), where {β0 (v), β1 (v), β2 (v)} are arbitrary functions. Carrying out
the coordinates transformations
u → v , v → u , ∓ρ → ρ ,
dropping primes, one arrives at (18.21a), CPSa(4.1), with
f1 = β2 , f2 = β1 , f3 = β0 .
r2 κ2 M
−2
κ2 M
ds2 = m2 r2 − 2m2 dr2 − m2 r2 − 2m2 dt2 − m12 dφ2
2
κ2 M r2 κ2 M 2
+ k r2 − 2m2 log r02
− 2m2 r02
+ κ 2M dt + dφ m , (18.29)
h(u − v) = 2) − α
μ tanh 4 (u − v)α − β ,
1
⎪
⎩ 3) α 1 (u − v)α − β −1 ,
μ 4
f1 = −B, f2 = A, f3 = −C.
344 Petrov Type N Wave Metrics
Table 18.3.1 Type N solutions with a covariantly constant null vector; pp-wave
solutions
In general, Kundt spacetimes are defined as those spacetimes that admit a null
geodesic vector field which is shear-free, twist-free, and expansion-free. In three
dimensions, because of this dimensionality, the twist and shear trivially vanish,
therefore a Kundt spacetime is simply one that admits an expansion-free null
geodesic congruence. These spacetimes allow for Petrov type II, D, III, N, and O
solutions as we shall see in the next paragraphs. Moreover, except for a general
type II branch of metrics, all the existing Kundt solutions are CSI-constant
scalar invariants – spacetimes for which all polynomial scalar (Ricci) curvature
invariants are constant.
1 ¯
H = H − f ,u f,u /P 2
+ W f ,u
¯ + W̄ f¯,u f,ζ ,
f, ζ̄ (19.7a)
f,ζ f¯ ,ζ̄
v = v + g(ζ, ζ̄, u) :
P = P, W = W − g,ζ , H = H − g,u , (19.7b)
348 Kundt Spacetimes in TMG
v
u = h(u), v = :
h,u
W 1 h,u,u
P = P, W = , H = H + . (19.7c)
h,u h,u 2 h,u
A special subclass of metrics – Kundt vacuum type III and N metrics (R12 =
∂ζ ∂ζ̄ ln P + Ψ2 + Ψ̄2 = 0) – arises for
∂f (ζ, u)
e2F (ζ,u) f,ζ → 1, = e−2F (ζ,u) ,
∂ζ
ζ
f (ζ, u) ≡ ζ (ζ, u) = e−2F (ζ,u) dζ,
ζ0
r = r (r, u) :
2
P = P r,r , W = W/r,r
− r ,u /(P r,r ) ,
f = f + (r ,u ) /(P r,r
)2 − W r ,u /(r,r
2
). (19.10)
19.2 General Kundt Metrics 349
r
In particular, the change r (u, r) = 1/P (u, r) dr, can be used to bring,
dropping primes, the Kundt metric to
k μ = δ μ v , kμ = δ u μ (19.11b)
S μ ν = S 1 1 (δ μ v δ v ν + δ μ u δ u ν − 2δ μ r δ r ν ) + S 1 2 δ μ v δ u ν + S 1 3 δ μ v δ r ν + S 3 2 δ μ r δ u ν
(19.14a)
1 1 ∂ 1 2
S 1 1 = S 2 2 = − EVC = − W1 − (W1 ) − 2 Λ ,
2 2 ∂r 2
S 3 3 = EV C = −2S 1 1 ,
1 1 ∂ ∂
S 1 3 = − EVC v (3 W1 − 2 μ) + F1 − W1 ,
2 2 ∂r ∂u
1 ∂ ∂
S 3 2 = vμ EVC + F1 − W1 ,
2 ∂r ∂u
352 Kundt Spacetimes in TMG
1 1 ∂ ∂
S 2 = − EVC μ (3 W1 − 2 μ) v −
1 2
F1 − W1 (W1 − 2 μ) v
4 4 ∂r ∂u
2
1 ∂ ∂ ∂
− 2
F0 + W1 F0 + F0 W1 ,
2 ∂r ∂r ∂r
∂ 1 2
EVC = W1 − (W1 ) − 2 Λ. (19.14b)
∂r 2
S = ∂r W1 − W1 2 /2 − 2Λ
S α β :{S 1 1 = S 2 2 = S, S 3 3 = −2S, S 1 2 , S 1 3 , S 3 2 }
adds rows multiplied by certain specific factors and permute the resulting second
by the third row. It is assumed that S 1 2 = 0. If S̃ 23 := −S 3 2 S 1 3 /S 1 2 is different
from zero, S̃ 23 = 0 then the metric is of Petrov type III. On the contrary, if
S̃ 23 = 0, then the spacetime is of Petrov type N,
⎡ ⎤ ⎡ ⎤
0 S12 0 0 1 0
⎢ ⎥ ⎢ ⎥
JSN = T S = ⎢ ⎣ 0 0 0 ⎥ ⎢
⎦ ∼ ⎣ 0 0 0 ⎦.
⎥ (19.20)
0 0 0 0 0 0
3
Rαβ Rαβ = EVC2 + 12 Λ2 , (19.21b)
2
5
Rαβ;γ Rαβ;γ = (3 W1 (u, r) − 2μ) − μ EV C 2
2 2
(19.21c)
8
traceless Ricci tensor S α β invariants
3
Sαβ S αβ = EVC2 , (19.22a)
2
S σ α Sσβ S αβ = 3/4 EV C 3 , (19.22b)
9
S σ α Sσβ S ν α Sν β = EVC4 , (19.22c)
8
3
Sαβ;γ S αβ;γ = W1 (u, r) (3 W1 (u, r) − 4 μ) EVC2 , (19.22d)
4
3
Sαβ;γ S αβ; S γ = − W1 (u, r) (3 W1 (u, r) − 4 μ) EVC3 (19.22e)
8
a Kundt spacetime, and all 3D CSI spacetimes were determined. The proof of
this assertion is based on a theorem stating, roughly speaking, “that on an open
neighborhood in which the Segré type does not change and all Ricci invariants
are constant (consequently its eigenvalues are constant too) there exits a frame
such that all the components of the Ricci tensor are constants and of the canon-
ical (Segré) form in any dimension.” An extensive use of the Segré forms and of
triad frames in which the Ricci tensor components are constant, together with
the constancy (vanishing) of the invariants constructed from covariant derivatives
of the Ricci tensor allow to establish this powerful result in three dimensions.
Extensions to higher dimensions have been carried out by Coley et al. (2009a
and 2009b).
The invariants of Sαβ , equivalently, of the Cotton tensor Cαβ , involv-
ing covariant differentiations, contain, as factor, the function W1 (u, r) EV C
to some power. Thus, there are type II metrics, unknown at present, with
EV C = 0 and with a function W1 (u, r) depending on its variables and con-
strained to fulfill the key equation (19.12b). Moreover, type II metrics with
constant EV C fulfilling (19.12b) arise for W 1(u, r) = 2μ/3; the invariants
become constants and the solutions become constant scalar invariant (CSI)
spacetimes.
If the Cotton tensor is of type III, its eigenvalues are zero, EV C = 0, together
with its invariants. Moreover, the Cotton tensor invariants, or the curvature Ricci
invariants, become constants, hence all Kundt type III metrics are constant scalar
invariant spacetimes.
namely
5 ∂W1 ∂ 2 W1 dr
F1 (u, r) = { μ − W1 +3 }EI(r)dr
2 ∂u ∂u∂r EI(r)
1
+f11 (u) EI(r)dr + f10 (u) , EI(r) := e (−μ+ 2 W1 )dr , (19.23)
where the function W1 (u, r) participates actively in the integrals, and, in general
for W1 (u, r) depending on r and u too, it should contribute substantially with
nonzero terms to the function F1 (u, r).
Therefore, extending these considerations to the case of the key equation
(19.12b) for W1 (u, r), a second-order equation, from the two parameters of inte-
gration depending on u one would be able, by transformations, to exclude one
of them, but not both at the same time, thus the remaining integration func-
tion will depend on the coordinate u. The general solution of the key equation
(19.12b) still remains unknown; it is a challenge to overcome.
The case W1 (r) is a sub-branch of Petrov type II Kundt metrics; even in this
case the general solution is unknown, although the families of solutions of type
II and D have been completely integrated in the case of constant W1 = 2μ/3
which are CSI spacetimes.
where the structural functions W1 (r), and Fi (u, r), i = 0, 1, 2, ought to satisfy
the following equations:
d d 1 2 3 d 1
W1 − W1 − 2Λ + (μ − W1 )( W1 − W12 − 2Λ) = 0, (19.24b)
dr dr 2 2 dr 2
3 2 d
F2 (r) = W1 (r) − W1 (r) + 3 Λ, (19.24c)
4 dr
∂2 1 ∂
F 1 + μ − W1 F1 = 0, (19.24d)
∂r2 2 ∂r
2
∂3 3 ∂ 7 ∂ 1 ∂
3
F 0 + μ + W1 2
F0 − 3 Λ − W1 + W1 − W1 μ
2
F0
∂r 2 ∂r 2 ∂r 4 ∂r
3 9 ∂ 3
+ 4 Λ μ − W1 − μ − W1 W1 + W1 2 μ − W1 F 0
2 2 ∂r 2
∂2 1 ∂
=− F1 − F 1 F1 . (19.24e)
∂u∂r 2 ∂r
19.4 Type II CSI Kundt Metric; W1 = 2μ/3 357
which coincides with the the homogeneous solution of (19.23) for W1 (r). By a
coordinate transformation, one can send f12 (u) → 0, as stated in Chow et al.
(2010b).
4μ
ds2 = dr2 +2du dv+ v 2 F2 (r) + vF1 (u, r) + F0 (u, r) du2 + v du d r, (19.27a)
3
see Chow et al. (2010b), Eqn. (A.32), or more compactly using for this kind of
citation the notation CPSb(A.32). The structural functions ought to satisfy the
following equations
1
F2 (r) = μ2 + 3 Λ, (19.27b)
9
∂2 2 ∂
F1 + μ F1 = 0, (19.27c)
∂r2 3 ∂r
∂3 ∂2 5 ∂ ∂2 1 ∂
3
F 0 + 2μ 2
F 0 + μ − 3Λ F0 = − F1 − F 1 F1 . (19.27d)
∂r ∂r 9 ∂r ∂u∂r 2 ∂r
The solution of (19.27c) is
S μ ν = S v v (δ μ v δ v ν + δ μ u δ u ν − 2δ μ r δ r ν ) + S v u δ μ v δ u ν + S r u δ μ r δ u ν , (19.32a)
1 2 1 2
Svv = μ + 9 Λ = 0, S r u = −2 vμ μ + 9Λ ,
9 9
2
1 ∂ 2 ∂
Svu = − F0 + μ F0 . (19.32b)
2 ∂r2 3 ∂r
In general, within this branch of solutions with constant W1 = 2μ/3, the
spacetime is of type II for S v u = 0,
⎡ v ⎤
S v 1 0
⎢ ⎥
(JII ) = ⎢
⎣ 0 Svv 0 ⎥ , F0 (u, r) = 0.
⎦
0 0 −2S v v
where the constants of integration f0i (u) depending on the variable u are denoted
by f0i := f0i (u), i = 1, 2, 3. Via a coordinate transformation the arbitrary
function f03 (u) can be set zero, f03 (u) = 0.
the function f03 (u) can be equated to zero, f03 (u) → 0, by coordinate
transformations. This solution corresponds to the CPSb(3.12) one.
where f0i = f0i (u). The function f03 can be equated to zero, f03 → 0, by
coordinate transformations. This solution is just the CPSb(3.11) one.
S μ ν = S v v (δ μ v δ v ν + δ μ u δ u ν − 2δ μ r δ r ν ) + S r u δ μ r δ u ν , (19.39b)
360 Kundt Spacetimes in TMG
1 2 2
Svv = μ + 9 Λ = 0, S r u = − vμ μ2 + 9 Λ ,
9 9
⎡ v ⎤
S v 0 0
⎢ ⎥
JD =⎢
⎣ 0 Svv 0 ⎥.
⎦ (19.39c)
0 0 −2S v v
Making the coordinate transformation
18 1
u= ũ, v = ,
(μ2 − 27 Λ) ṽ − ũ
one brings (19.39a) to the form
2
36 dṽ dũ 12μ dũ
ds2 = − + dr − , (19.39d)
μ2 − 27 Λ (ũ − ṽ)2 μ2 − 27 Λ ũ − ṽ
which, by the subsequent transformations
1 1 μ2 − 27 Λ
t= ũ + ṽ, x = ṽ − ũ, z = r − ln(ṽ − ũ), (19.39e)
2 2 6μ
becomes the: biaxially squashed spacelike limit
2
2 9 dx2 − dt2 4μ2 dt
ds = 2 + 2 dz + , (19.39f)
μ − 27 Λ x2 μ − 27 Λ x
where the 2D sector AdS2 is written using Poincaré coordinates (t, x).
In a concise way all these solutions can be visualized in the table above.
∂ 1 2
W1 (u, r) − (W1 ) − 2 Λ = 0, (19.40)
∂r 2
which guarantees the fulfillment of the key equation (19.24b). The solutions of
this equation (19.40) depend on r − f (u), thus, by a shift of the coordinate
r → r + f (u) one brings the dependence to r. One distinguishes the following
branches of solution depending on the sign of the cosmological constant Λ:
Case Λ = 0. For zero cosmological constant Λ = 0, there are two cases
Λ = 0 : W1 (u, r) = 0, (19.41a)
2
Λ = 0 : W1 (u, r) = − . (19.41b)
r
Case Λ = −m2 . For negative cosmological constant Λ = −m2 , one distinguishes
three different classes
F1 (u, r) = f12 (u) + f11 (u) e−μ r → F1 (u, r) = f11 (u) e−μ r , (19.44)
362 Kundt Spacetimes in TMG
2 1
W1 (u, r) = − → F2 = 2 . (19.51)
r r
Substituting this function into the F1 (u, r)–equation (19.24d) one gets
∂ μ r ∂F1
e r = 0, (19.52)
∂r ∂r
therefore, integrating
e−μ r
F1 (u, r) = − dr f11 (u) =: Ei (1, μ r) f11 (u), (19.53)
r
where Ei(x) is the exponential integral,
r −μ x
e d e−μ r
Ei (1, μ r) = − dx, Ei (1, μ r) = − . (19.54)
x dr r
For the determined structural functions, the F0 (u, r)–equation (19.24e) becomes
2
∂3 3 ∂ 2 ∂ 2
F0 + μ − F0 − F0 + 2 F0 = Rhs(u, r),
∂r3 r ∂r2 r ∂r r
1 1 df
f11 2 Ei (1, μ r)e−μ r +
11 −μ r
Rhs(u, r) = e . (19.55)
2r r du
Since the solution of the homogeneous equation is proportional to r, explicitly
F0 (u, r) = r Φ(19.60) + re−μ r f01 (u) + r2 f02 (u) + r f03 (u) . (19.61)
XN H = C1 X1 + C2 X2 + C3 X3
Via the variation of the parameters Ci (u, r), their derivatives are
∂C1 rhsΦ (1 + μ r) ∂C2 rhsΦ ∂C3 rhsΦ eμ r
=− , = , = .
∂r μ2 ∂r μ ∂r μ2
19.8 Type III Kundt Solution; Λ = 0, W1 = −2/r 365
W1 = −2m → F2 = 0, (19.71a)
e−2 (m+μ)r
− f11 2 . (19.72c)
8(μ + 2m)(μ + 3m)
Finally one may write the corresponding Kundt solution as
ds2 = dr2 + 2dudv − 4m vdu dr + v e−(μ+m)r f11 + F0 (u, r) du2 ,
e−(m+μ)r df11 e−2 (m+μ)r f11 2
F0 = e(m−μ)r f01 − − , (19.72d)
2m(μ + 3m) du 8(μ + 2m)(μ + 3m)
where f02 and f03 have been canceled by means of coordinate transformations.
This solution is characterized by a Petrov type III Cotton tensor
C αβ = C v u δαv δu β + C r u δαr δu β + C v r δαv δr β ,
1 μ (μ + m) −2 (μ+m)r v
f11 2 + μ (μ + m) f11 e−(μ+m)r
2
Cvu = − e
4 3m + μ 2
1 μ (μ + m) −(μ+m)r df11 1
− e + μ μ2 − m2 f01 e(−μ+m)r
4 m du 2
v r 1 −(μ+m)r
C r = C u = μ (μ + m) f11 e . (19.72e)
2
Type N Limit
For f11 = 0 this solution becomes a Petrov type N spacetime with metric
ds2 = dr2 + 2dudv − 4m vdu dr + e(m−μ)r f01 (u) du2 , (19.73a)
and Cotton tensor
1
C α β = μ μ2 − m2 f01 e(−μ+m)r k α kβ , k α = δ α v . (19.73b)
2
where f02 and f03 have been canceled by coordinate transformations. This solu-
tion is characterized by the type III Cotton and traceless Ricci tensor of the
form
Type N Limit
Nevertheless, for f11 = 0 this spacetime becomes a type N solution given by
Moreover, for f01 = 0 the metric describes an AdS spacetime – a type O metric
field – given by
where the functions f02 (u) and f03 (u) have been eliminated via coordinate
transformations.
This solution is characterized by the Cotton and traceless Ricci tensor of the
form
Type N Limit
Again, as in the previous branch of solutions, one finds the type N subcase arising
for f11 = 0, namely
ΦH (u, r) = f01 (u) e−rμ + f02 (u) em r + f03 (u) e−m r , (19.86e)
Type N Limit
The solution presented above (19.86f) allows for a type N Kundt subcase with
metric
ds2 = dr2 + 2du dv + (v 2 F2 + F0 )du2 − 4 m v coth(m r) du dr,
m2
F2 (u, r) = , F0 (u, r) = sinh(m r) f03 (u) e−rm , (19.87a)
sinh2 (m r)
with a type N Cotton tensor, JCN ,
μ 2
C α β = C v u k α kβ , , C v u = μ − m2 f03 (u) e−μ r sinh (m r) , k α = δ α v .
2
(19.87b)
where without loss of generality one may assume the highest coefficient an = 1,
beside the VP procedure based on the determination of the roots of the char-
acteristic equation, one can search for an expression of the equation of the
form
d d d d α1 r
eαn r eαn−1 r · · · (e φ(r)) = f (r)e(α1 +α2 +···+αn )r , (19.89)
dr dr dr dr
d
with n in number derivative operators dr . Comparing the coefficients ai of (19.88)
di
with those of (19.89) multiplying derivatives dr i φ of the same order in both
1 = 1,
nα1 + (n − 1)α2 + · · · + 1 αn = an−1 ,
........................................................
α1 (α1 + α2 ) · · · (α1 + α2 + · · · + αn−1 + αn ) = a0 , (19.90)
whose solutions determine the exponents. Finally, integrating (19.89), one arrives
at a compact expression for the main part of the non-homogeneous solution
r y
−α1 r −α2 t
φN H (r) = e e ··· f (z)e (α1 +α2 +···+αn ) z
dz dy · · · dt.
(19.91)
The general solution for the equation (19.88) with constant coefficients is just the
sum of φN H (r) and the n-independent solutions to the homogeneous equation,
which are of the exponential form eαi r multiplied by powers of r in the case of
degeneracy.
This procedure will be used extensively in the derivation of the solutions to
the non-homogeneous equations, thus we shall refer to this subsection as 19.10.1.
Of course, this expression can be simplified in various manners; for instance one
dξ
can partially integrate the first term in Φ(u, r) using the derivative dr from
(19.81).
Finally, the metric becomes
∂3 ∂2 ∂ RHS
3
Φ − m 2 Φ − m 2 Φ + m3 Φ = = Rhs
∂r ∂r ∂r sinh (m r)
2 d
1 mem r f11 ξ (r) + 2 du f11
=− 2 . (19.95a)
2 sinh (m r)
Since the characteristic equation (k + m)(k − m)2 = 0 allows for a double root,
then the homogeneous solution is
Φ( H) (u r) = C1 e−m r + C2 em r + C3 r em r . (19.95b)
γ + 3 α + 2 β = −m,
2γ α + β 2 + 3 α2 + 4 β α + γ β = −m2 ,
γ β α + γ α 2 + 2 β α 2 + β 2 α + α 3 = m3 (19.96a)
with solutions
{α = m, β = −2 m, γ = 0} , {α = −m, β = 2 m, γ = −2 m, } ,
{α = −m, β = 0, γ = 2 m} . (19.96b)
thus, for each set of exponential constants one gets its corresponding exponent–
integral representation.
∂3 ∂2 2 ∂ f11 2 ξ (r) + 2 du
d
f11
φ − 4m φ + 4 m φ = −m 2 = Rhs. (19.97b)
∂r 3 ∂r 2 ∂r (1 − e−2 mr )
The solution to the homogeneous equation is given by
via the variation of parameters Ci (u, r), where u plays the role of a parameter.
The first derivatives of Ci (u, r) are
∂C1 1 Rhs ∂C2 1 (2 m r + 1) e−2 m r ∂C3 1 e−2 m r
= , = − Rhs, = Rhs.
∂r 4 m2 ∂r 4 m2 ∂r 2 m
19.10 Type III Kundt Metric; Λ = −m2 , W1 = −2 m coth(m r) 379
for the dilog function, which is presented in details in the next equations (19.99).
The above-integrated solution (19.97e) can be given in the CPS form,
F0 (u, r) (f11 )2 2 rm 2
= e −1 1 + ln e2 rm − 1 −1
1 − e2 rm 16m2
1 df11 2 rm
+2e2m r L2 e2m r + 2
e − 1 ln e2 rm − 1 .
4m du
(19.97f)
Finally, this family μ = −m of Petrov type III CSI spacetimes is given by the
metric
ds2 = dr2 + 2du dv + (v 2 F2 + vF1 + F0 )du2 − 4 v m coth(m r) du dr,
m2
F2 (u, r) = ,
sinh2 (m r)
F1 (u, r) = f11 (u)ξ(r), ξ (r) = ln e2 m r − 1 = m r + ln (sinh (m r)) ,
F0 (u, r) = (1 − e−2 rm ) [φH (u, r) + φN H (u, r)(19.97e)] . (19.97g)
This solution is characterized by the Cotton:
C αβ = C v u δαv δuβ + C v r δαv δr β + C r u δαr δuβ ,
m3 e2 mr 1 2 mr df11
C v u = −m2 e2 mr − 1 f03 + 2 v f11 2 − 2 me
(e 2 mr − 1) du
1
− m f11 2 ln e2 mr − 1 + e2 mr ,
4
e2 mr
C r u = C v r = m2 2 mr f11 . (19.97h)
e −1
Type N Limit
It allows for a type N Kundt solution with metric
ds2 = dr2 + 2du dv + (v 2 F2 + F0 )du2 − 4 v m coth(m, r) du dr,
m2
F2 = , F0 = 2f03 r sinh(m r) e−rm (19.98a)
sinh2 (m r)
380 Kundt Spacetimes in TMG
hence
r
dilog 1 + e± 2 m r = ∓ 2m ln 1 + e± 2 mt dt,
d
dilog e±2 m r + 1 = ∓2 m ln e±2 m r + 1 , (19.99b)
dr
dilog e2 m r + dilog e−2 m r + 2 m2 r2 = 0,
dilog 1 + e−2 m r + dilog 1 + e2 m r + 2 m2 r2 = 0, (19.99c)
2 m ln e2 mr − 1 dr = dilog e2 mr + 2 m r ln e2 mr − 1 . (19.99d)
● for x ≤ 1
Li(x) = dilog(1 − x), x ≤ 1; Li(e−2 m r ) = dilog 1 − e−2 mr ,
Li(−e2 m r ) = dilog 1 + e2 mr , (19.99f)
∂
Consequently, integrating ∂r Ci , the non-homogeneous Φ solution amounts to
1
ΦN H = − 2
f11 2 −2 dilog e−2 rm + 4 rm ln 1 − e−2 rm
8m
2
− e2 rm − 1 1 + ln 1 − e−2 rm − 1 e−rm
e−rm df11 2 rm
+ 2
e − 1 ln e2 rm − 1 − 2 m r e2 rm . (19.101d)
2m du
γ + 2 β + 3 α = m,
γ β + 2 γ α + β 2 + 4 β α + 3 α2 = −m2 ,
γ β α + γ α2 + β 2 α + 2 β α2 + α3 = −m3 (19.102b)
{α = −m, β = 2 m, γ = 0} ,
{α = m, β = 0, γ = −2 m} ,
{α = m, β = −2 m, γ = 2 m} . (19.102c)
via the variation of parameters Ci (u, r), where u plays the role of a parameter.
The derivatives of the Ci are
∂C1 1 em r Rhs (2 m r − 1) ∂C2 1 em r Rhs ∂C3 1 Rhs
= 2
, = − , = ,
∂r 8 m ∂r 4 m ∂r 8 e m r m2
19.10 Type III Kundt Metric; Λ = −m2 , W1 = −2 m coth(m r) 383
f11 2 , 2 m r
−2 m r 2
φN H = e − 1 1 + ln 1 − e − 1
16m2 +
+2L2 (e2 m r ) + 4 m2 r2
1 df11 2m r
+ 2 e − 1 ln 1 − e−2 m r , (19.103f)
4m du
which coincides with the CPS Chow et al. (2010b) form of the solution:
1 df11 1
+ me−2 mr + m f11 2 ln 1 − e−2 mr + e−2 mr ,
2 du 4
m2
Cru = C v r = − 2 mr f11 . (19.103h)
e −1
Type N Limit
It allows for a type N Kundt solution with metric
We follow here the same pattern used in the previous sections: first the general
case of μ = ± m is treated; next, the derived results and key equations are
specialized to handle the sub-cases μ = ± m.
Evaluating the equation (19.24c) for F2 one gets
m2
F2 = − . (19.106)
cosh2 (m r)
Replacing these functions into the equation (19.24d) for F1 , one arrives at
∂2 ∂
F1 (u, r) + [μ + m tanh (m r)] F1 (u, r) = 0, (19.107)
∂r2 ∂r
with solution
m e−μ r
F1 (u, r) = f11 (u) dr = f11 ξ(r), (19.108)
cosh (m r)
where it has been introduced the auxiliary function ξ(r),
dξ me−μ r e−μ r
= → ξ (r) = m dr. (19.109)
dr cosh (m r) cosh (m r)
The equation (19.24e) for F0 (u, r) becomes
∂3 ∂2
F 0 + [μ − 3 m tanh (m r)] F0
∂r3 ∂r2
∂
+2 m −2 m + 3 m tanh2 (m r) − μ tanh (m r) F0
∂r
+2 m2 tanh2 (m r) − 1 (μ − 3 m tanh (m r)) F0
1 d d
=− 2
2 f11 + f11 ξ (r) ξ (r) =: RHS . (19.110)
2 du dr
Following the standard procedure for solving non-homogeneous linear ordinary
differential equations by means of the variation of parameters of the homogeneous
solution
F0H = cosh( m r) C1 (u) e−μ r + C2 (u) em r + C3 (u) e−m r , (19.111)
∂3 ∂2 ∂ RHS
3
Φ + μ 2 Φ − m2 Φ − m2 μ Φ = = Rhs, (19.113)
∂r ∂r ∂r cosh (m r)
386 Kundt Spacetimes in TMG
2 df11
1 f11 ξ + 2 du dξd me−μ r
Rhs = − ,
ξ (r) = . (19.114)
2 cosh (m r) dr dr cosh (m r)
In what follows we proceed to determine the three different branches of solutions
in the specific cases of μ = ± m, μ = − m, and μ = m separately.
k 3 + μ k 2 − m2 k − m2 μ = 0 = (k + μ)(k − m)(k + m)
where u enters here as a parameter. The variation of the constants Ci (u, r),
within the variation of parameters procedure, yields
∂ eμ r RHS ∂ 1 1 RHS
C1 = 2 , C2 = e−m r ,
∂r (μ − m2 ) cosh (rm) ∂r 2 m(μ + m) cosh (m r)
∂ 1 1 RHS
C3 = − em r .
∂r 2 m(μ − m) cosh (m r)
Notice the presence of μ ± m in the denominators of the above expressions,
which clearly point to the degeneracy of these quantities when dealing with
the sub-branches μ = ±m; this is the reason why the quoted cases have to be
integrated separately for the equation (19.113).
Thus, the non-homogeneous solution amounts to
−μ r
df11 e m
ΦN H (u, r) = dr
du m − μ
2 2
cosh2 (mr)
−(m+μ)r
1 emr e 1 e−mr e(m−μ)r
− dr − dr
2 (m + μ) cosh2 (mr) 2 (m − μ) cosh2 (mr)
19.11 Type III Kundt Metric; Λ = −m2 ; W1 = −2 m tanh(m r) 387
1 m e−μr
2 ξ (r)
+ f11 dr
2m −μ
2 2
cosh2 (mr)
−(m+μ)r (m−μ)r
1 emr e ξ (r) 1 e−m r e ξ (r)
− dr − dr , (19.115b)
4 (m + μ) cosh2 (mr) 4 (m − μ) cosh2 (mr)
ΦH (u, r) = f01 (u) e−rμ + f02 (u) em r + f03 (u) e−m r , (19.115c)
Type N Limit
The type N limit Kundt solution is given by the metric
{α (β + α) = −2 m (μ − m) , β + 2 α = μ − 3 m} , (19.119c)
with solutions
Consequently, φ integrates as
r t y
φ= e−α t e−β y Rhs (z) e(α+β)z dzdydt,
Type N Limit
For vanishing function f11 one gets a type N metric
ds2 = dr2 + 2du dv + (v 2 F2 + F0 )du2 − 4m v tanh(m r) du dr,
F0 (u, r) = cosh(m r) f01 (u) e−m r + f02 (u) em r + f03 (u) r em r ,
m2
F2 (r) = − , (19.121a)
cosh2 (m r)
characterized by a Petrov type N Cotton tensor
1
C α β = C v u k α kβ , C v u = m2 e2 mr + 1 f03 , k α = δ α v . (19.121b)
2
via the variation of parameters Ci (u, r), where u plays the role of a parameter.
The first derivatives are
∂ 1 Rhs ∂C2 1 (2 m r + 1) e−2 m r Rhs
C1 = , = − ,
∂r 4 m2 ∂r 4 m2
∂ 1 e−2 m r Rhs
C3 = .
∂r 2 m
Therefore the general φ(u, r) solution to the equation (19.123b) via the variation
of parameters method is
F0 (u, r)
φ(u, r) = = f01 (u) + f02 (u) e2 mr + f03 (u) r e2 m r
(1 + e−2 mr )
+C1 (u, r) + C2 (u, r) e2 mr + C3 (u, r) r e2 m r , (19.123e)
Multi-Exponent–Integral Representation of Φ
In this approach the search of the solution to the key equation (19.125a),
following 19.10.1, gives rise to the algebraic system
γ + 2 β + 3 α = m,
γ β + 2 γ α + β 2 + 4 β α + 3 α2 = −m2 ,
γ β α + γ α2 + β 2 α + 2 β α2 + α3 = −m3 (19.126a)
with the following set of solutions
{α = −m, β = 2 m, γ = 0} ,
{α = m, β = 0, γ = −2 m} ,
{α = m, β = −2 m, γ = 2 m} , (19.126b)
396 Kundt Spacetimes in TMG
with the corresponding solution representations. For instance, for the first set of
exponents one has
mr −2 m r
Φ1 = e e em r Rhs (r) dr drdr.
1 ,
−2 m r 2
− f11
2
e 2mr
+ 1 1 + ln 1 + e − 1 + 4 m r + 2 e2 m r
16 m2 +
+2 dilog 1 + e−2 m r . (19.127f)
One can always eliminate the terms proportional to the independent homoge-
neous solutions {1, r, e2mr }, and use the relation (19.99c) between the dilog and
Li2 (19.99f) and write, without loss of generality, the function φN H (u, r)
1 , 2
φN H = − 2
f11 2 e2 m r + 1 1 + ln 1 + e−2 m r −1
16 m +
−2 Li2 −e2m r − 4 m2 r2
1 df11 2 mr
+ e + 1 ln 1 + e−2 m r , (19.127g)
4 m2 du
i.e., the formula CPSb(4.56). Notice that the relation Φ = 2 e− m r φ holds, as
indeed it should.
The metric of this solution can be written as
Type N Limit
Its Petrov type N limit is given by the metric
Replacing these functions into the equation (19.24d) for F1 one arrives at
∂2 ∂
2
F1 (u, r) + (μ + m cot (m r)) F1 = 0 (19.132)
∂r ∂r
with solution
e−μ r
F1 (u, r) = f11 (u) m dr = f11 ξ(r); (19.133)
sin (m r)
where it has been introduced the auxiliary function ξ(r),
dξ me−μ r e−μ r
= → ξ (r) = m dr. (19.134)
dr sin (m r) sin (m r)
The key equation (19.24e) for F0 (u, r) becomes
∂3 ∂2
F 0 (u, r) + (μ − 3m cot (m r)) F0
∂r3 ∂r2
∂
+2 m 2 m − μ cot (m r) + 3 m cot2 (rm) F0
∂r
+2 m2 F0 1 + cot2 (m r) (μ − 3m cot (m r))
1 d 2 d
=− 2 f11 + (f11 ) ξ (r) ξ (r) = RHS. (19.135)
2 du dr
Following the standard VP procedure for solving non-homogeneous linear
ordinary differential equations the solution F0 N H is searched as
F0 = C1 (u, r) e−μ r + C2 (u, r) sin (m r) + C3 (u, r) cos (m r) sin (m r) .
(19.136)
Nevertheless, the structure of the solution above points to a better function
choice, namely
Φ (u, r) = F0 (u, r)/sin (m r), F0 (u, r) = Φ (u, r) sin (m r) . (19.137)
The key equation (19.135) in terms of Φ becomes
∂3 ∂2 ∂ RHS
3
Φ + μ 2
Φ + m2 Φ + Φm2 μ = . (19.138)
∂r ∂r ∂r sin (m r)
In what follows we proceed to determine the single solution of this class.
First the derivation of the solution via the variation of parameters procedure is
done. Next, the solution in a multi-exponent–integral representation is curried
out.
The variation of the parameters Ci (u, r), within the VP procedure, yields
∂ eμ r RHS
C1 = 2 ,
∂r μ + m2 sin (m r)
∂ (m sin (m r) − μ cos (m r)) RHS
C2 = − ,
∂r m (μ2 + m2 ) sin (m r)
∂ (μ sin (m r) + m cos (m r)) RHS
C3 = − . (19.139c)
∂r m (μ2 + m2 ) sin (m r)
Some simplifications arise by replacing ξ(r), or more precisely, its derivative from
(19.134), to get the non-homogeneous solution as
ΦH = f01 (u) e−rμ + f02 (u) sin(m r) + f03 (u) cos(m r), (19.139f)
19.12 Type III Kundt Metric; Λ = m2 , W1 = −2m cot(m r), μ 401
which together determine the general solution. The metric of this solution can
be written as
Type N Limit
Its Petrov type N limit is given by the metric
df11
interestingly, the factor of du can be simplified up to
sin (m r) ξ (r) df11
,
m2 du
as was done in the previous VP paragraph, coinciding with the expression
of CPS Chow et al. (2010b), which determines in turn the non-homogeneous
F0 (u, r) function. Finally, the metric becomes
ds2 = dr2 + 2du dv + m2 v 2 cosec2 (m r)du2 − 4mv cot(m r)du dr
+[vξ(r)f11 (u) + F0 (u, r)]du2 ,
r t y
F0 (u, r) m cos[m (t − 2 y + z)]
= − f11 2 dt dy eμ (t−z−r) ξdz
sin(m r) 2 sin2 (m z)
sin (mr) ξ (r) df11
+ + e−μr f01 (u) + sin(m r)f02 (u) + cos(m r)f03 (u),
m2 du
e−μ r
ξ (r) = m dr. (19.141g)
sin (m r)
For other choice of exponents, one arrives at other representations of the sought
inhomogeneous solution.
All the references of the solutions of the type III and N Kundt families can be
represented schematically by the table:
S = ∂r W1 − W1 2 /2 − 2Λ = 0
0 μ 19.7 (4.15) – 0
−2/r μ 19.8 (4.39) (19.70) 0
μ = ±m 19.9.1 (4.3) (19.73) <0
-2m μ=m 19.9.2 (4.9) (19.75) <0
μ = −m 19.9.3 (4.13) (19.77) <0
μ = ±m 19.10.1 (4.26) (19.87) <0
−2m coth(2mr) μ = −m 19.10.3 (4.34) (19.98) <0
μ=m 19.10.4 (4.30) (19.104) <0
μ = ±m 19.11.1 (4.53) (19.116) <0
−2m tanh(2mr) μ = −m 19.11.2 (4.57) (19.121) <0
μ=m 19.11.3 (4.55) (19.128) <0
−2m cot(2mr) μ 19.12 (4.44) (19.140) >0
20
Cotton Tensor in Riemannian Spacetimes
modern language. These results seem presently not to be too well known in the
community.
In the theory of conformal spaces the main geometrical objects to be ana-
lyzed are the Weyl (1918) and the Cotton (1899) tensors. It is well known that
for conformally flat spaces the Weyl tensor has to vanish. However, the Cotton
tensor is only conformally invariant in three dimensions. Any 3D space is con-
formally flat if the Cotton tensor vanishes. If matter is present, the Ricci tensor
is related to the energy–momentum tensor of matter by means of the Einstein
equations. Then the vanishing of the Cotton tensor imposes severe restrictions on
the energy–momentum tensor. The Cotton tensor also plays a role in the context
of the Hamiltonian formulation of general relativity; see Arnowitt et al. (1962).
First the Cotton 2-form in the context of the Bianchi identities is derived.
Subsequently its characteristic properties are described and an irreducible
decomposition with respect to the (pseudo-)orthogonal group is performed. This
allows ones to determine the number of irreducible components in any dimen-
sion. Moreover, in four dimensions, the relation of the Cotton to the Bach tensor
is exhibited. After that it is shown how to derive the Cotton 2-form in three
dimensions by means of a variational procedure. The classification of the Cotton
2-form in three dimensions is accomplished by means of its eigenvalues.
Our notation and conventions for this chapter are taken from Schouten
(1954) and for exterior calculus we refer to Hehl et al. (1995). For quick and
easy reference, we display our conventions for index positions and signs of the
Christoffel symbol, the Riemann tensor, and the Ricci tensor (holonomic indices
i, j, · · · = 0, 1, 2, 3). The sign of the Ricci tensor is the same as those of the Lij
tensor and the Cotton tensor. In particular, the Ricij sign introduces a relative
sign between the Lij tensor and the Weyl tensor in the decomposition of the
curvature:
where ind , the number of negative eigenvalues, denotes the index of the metric.
Then dg = 0 and, according to (20.10), the connection is antisymmetric. In turn,
from (20.11) we can infer the antisymmetry of Rαβ .
The exterior covariant derivative of a p-form Xα β is given by
DXα β = dXα β − Γα γ ∧ Xγ β + Γγ β ∧ Xα γ . (20.13)
The Bianchi identities can be formulated concisely with the help of this definition.
In a Riemannian space the torsion T α = Dϑα vanishes. Thus, the first Bianchi
identity reads
0 = DT α = DDϑα = Rβ α ∧ ϑβ , (20.14)
or, in components,
R[αβγ] δ = 0 . (20.15)
20.1 Bianchi Identities and the Irreducible Decomposition 407
DD Cα = −Rα β ∧ Cβ
= −Weylα β ∧ Cβ − Ricci
α β ∧ Cβ − Scalarα β ∧ Cβ .
(20.58)
Next, we use the “double duality relations” for the irreducible pieces of the
curvature,
1 μν
Weylαβ = Weylμν η αβ , (20.60)
2
1
αβ
Ricci = −Ricci
μν η μν αβ , (20.61)
2
1 μν
Scalarαβ = Scalarμν η αβ . (20.62)
2
Together with eqs.(20.37) and (20.31), we obtain
1 1
D Weylα β ∧ Lβ = η μν α β DWeylμν ∧ Lβ = − η μν α β ϑ[μ ∧ Cν] ∧ Lβ
2 2
1 μβν
=− η α ϑμ ∧ Lβ ∧ Cν
2
= − Ricci
ν α ∧ Cν + Scalarν α ∧ Cν
α ν ∧ Cν .
= Ricci (20.63)
Thus,
Bα := D Cα + Weylα β ∧ Lβ =: Bα β ηβ (20.65)
or, in components,
Bαβ = ∇μ Cαμβ + Lμν Weylαμβν , (20.66)
One recognizes the Bach tensor Bαβ , Bach (1921); Schouten (1954); Penrose and
Rindler (1986); Tsantilis et al. (1996) . From the symmetry properties of Cα , Lα ,
and Weylαβ it follows that
Bα ∧ ϑα = 0 (Bα α = 0) , eα B α = 0 (B[αβ] = 0) . (20.68)
Moreover, it transforms as a conformal density and can be derived from a vari-
ational principle. Since in a conformally flat space the Weyl and the Cotton
tensors vanish, the vanishing of the Bach tensor is also a necessary (but not
sufficient) condition for a 4D space to be conformally flat.
Cα as a Variational Derivative
It is well known Deser et al. (1984); Baekler et al. (1992) that Cα can be obtained
by means of varying the 3D Chern–Simons action
1 2
CRR = − Γα β ∧ dΓβ α − Γα β ∧ Γβ γ ∧ Γγ α (20.69)
2 3
with respect to the metric keeping the connection fixed. In order to enforce
vanishing torsion
T α := Dϑα = dϑα − Γβ α ∧ ϑβ (20.70)
and vanishing non-metricity
Qαβ := −Dgαβ = −dgαβ + Γα γ gγβ + Γβ γ gαγ , (20.71)
we have to apply Lagrange multipliers. Then the total Lagrangian reads
L = CRR + λα ∧ T α + λαβ ∧ Qαβ , (20.72)
where λα is a 1-form and λαβ = λβα a symmetric 2-form.
The variation of this Chern–Simons Lagrangian (20.69), which only depends
on the connection, turns out to be
1
δCRR = −δΓα β ∧ Rβ α + d Γα β ∧ δΓβ α . (20.73)
2
In the next step, we enforce vanishing torsion and non-metricity by means of
respective Lagrange multiplier terms:
L = CRR + λα ∧ T α + λαβ ∧ Qαβ . (20.74)
The variation then yields
δL = δCRR + δλα ∧ T α + λα ∧ δT α + δλαβ ∧ Qαβ + λαβ ∧ δQαβ
= −δΓα β ∧ Rβ α + δλα ∧ T α + λα ∧ dδϑα + δΓβ α ∧ ϑβ + Γβ α ∧ δϑβ
1
+ d Γα β ∧ δΓβ α + δλαβ ∧ Qαβ
2
+λαβ ∧ (−dδgαβ + δΓα γ gγβ + Γα γ δgγβ + δΓβ γ gαγ + Γβ γ δgαγ )
20.2 Cotton 2-Form 413
1 β
= −δΓα β ∧ Rβ α + δλα ∧ T α + δλαβ ∧ Qαβ + d Γα ∧ δΓβ α
2
+λα ∧ Dδϑα − δΓβ α ∧ λα ∧ ϑβ − λαβ ∧ Dδgαβ + δΓα β ∧ (λα β + λβ α )
= δλα ∧ T α + δλαβ ∧ Qαβ + δϑα ∧ Dλα + δgαβ Dλαβ
−δΓα β ∧ (Rβ α + λβ ∧ ϑα − λα β − λβ α )
1 β
−d −λα ∧ δϑ + Γα ∧ δΓβ − λ δgαβ .
α α αβ
(20.75)
2
The corresponding field equations read
δL
= Tα = 0, (20.76)
δλα
δL
= Qαβ = 0 , (20.77)
δλαβ
δL
= −Rβ α − λβ ∧ ϑα + 2λβ α = 0 , (20.78)
δΓα β
δL
= Dλα = 0 , (20.79)
δϑα
δL
= Dλαβ = 0 . (20.80)
δgαβ
We can solve (20.78) for its symmetric and its antisymmetric parts,
Eventually,
1 δL δL 2 δL
= Cα , = −ϑ(α ∧ C β) , − eβ = Cα . (20.86)
2 δϑα δgαβ n−1 δgαβ
In the presence of matter, the gravitational field equation is given by δ(L +
Lmat )/δϑα = 0. Hence, the Cotton 2-form can be coupled to the energy–
momentum 2-form of matter
δLmat
Σα := . (20.87)
δϑα
414 Cotton Tensor in Riemannian Spacetimes
This is carried out in the topologically massive gravity model of Deser, Jackiw
and Templeton (DJT); see Deser et al. (1982a, 1982b), where the Lagrangian
(20.72) is enriched by a Hilbert–Einstein term and a cosmological term ( is the
gravitational constant and θ a dimensionless coupling constant):
Gα = Lβ ∧ ηβα ; (20.90)
α β = Weylα β ,
Weyl (20.99)
2
R = exp(−2σ) [R − 2(n − 1) σ ;α − (n − 1)(n − 2)σ,α σ ] . (20.100)
,α ,α
Thus, in n = 3, where the Weyl 2-form vanishes, the Cotton 2-form becomes
conformally invariant.
Thus,
2=0 R
R ⇐⇒ σ ,i ;i = . (20.103)
2
This is a scalar wave equation for the conformal factor σ with R as source. Since
the wave equation always has a solution, we conclude that all 2D spaces are
conformally flat.
n≥3
For more than two dimensions we start from (20.35), namely
2
Rαβ = Weylαβ − ϑ[α ∧ Lβ] . (20.104)
n−2
Since the Weyl 2-form is conformally invariant it cannot be transformed to zero
by means of a conformal transformation. Consequently, the vanishing of the Weyl
2-form is a necessary condition for conformal flatness.
The Lα 1-form transforms according to
2 α = Lα − (n − 2) Sα .
L (20.105)
Lα = (n − 2) Sα . (20.106)
20.5 Classification of the Cotton 2-Form in 3D 417
or, in components,
1
Cα = ∇μ
β
Ricνα − Rgνα η μνβ . (20.112)
4
This alternative representation of the Cotton 2-form, often called Cotton–York
tensor – see York Jr. (1971) (even though it was already discussed explicitly by
ADM Arnowitt et al., 1962) – can only be defined in three dimensions. Sometimes
it appears under the name Bach tensor in the literature; see Christodoulou and
Klainerman (1993), for example This seems to be a misnomer.
The Cotton tensor is tracefree
Cα α = eα C α = (C α ∧ ϑα ) = 0 . (20.113)
In the three dimensions, the second Bianchi identity (20.37) amounts to ϑ[α ∧
Cβ] = 0. In view of the definition (20.111), we infer that the Cotton tensor is
symmetric Cαβ = Cβα . Introducing this symmetry explicitly into (20.112), we
obtain the alternative representation
C αβ = C βα = η μν(α ∇μ Ricν β) . (20.114)
We now perform a classification of the Cotton tensor with respect to its
eigenvalues. The corresponding generalized eigenvalue problem reads:
αβ
C − λ g αβ Vβ = 0 , C [αβ] = 0 , C αβ gαβ = 0 . (20.115)
By lowering one index, we can reformulate this as ordinary eigenvalue problem
for the matrix Cα β . However, in that case, the symmetry C αβ = C βα is no longer
manifest:
β
Cα − λ δαβ Vβ = 0 , Cα α = 0 . (20.116)
● Class A
Three distinct eigenvalues: λ1 = λ2 and λ3 = −(λ1 + λ2 ).
● Class B
Two distinct eigenvalues: λ1 = λ2 = 0, λ3 = −2λ1 .
● Class C
One distinct eigenvalue: λ1 = λ2 = λ3 = 0.
In the present context of Euclidean signature, this implies Cαβ = 0.
20.5 Classification of the Cotton 2-Form in 3D 419
complex eigenvalues occur, too. This point seems to have been overlooked by the
authors of Barrow et al. (1986). Consequently, the classification will not be as
simple as was the case for the Euclidean metric.
In the following, we will present a classification of Cα β . The tracefree condition
(20.116)2 , in orthonormal coordinates, reads explicitly
C1 1 + C2 2 + C3 3 = 0 . (20.119)
λ3 + b λ + c = 0 . (20.121)
A cubic polynomial with real coefficients has at least one real root and the
complex roots have to be complex conjugates. Since one eigenvalue is real, types
D and II with only one independent eigenvalue λ1 = λ2 = −2λ3 are always real.
For class I, besides the real eigenvalue, two complex eigenvalues may occur. In
that case, they are complex conjugated. Therefore, class I can be subdivided
into class I with three real eigenvalues, [111], and class I with one real and two
complex conjugated eigenvalues, [1z z̄].
We can now specify simple criteria for deciding to which of these classes the
Cotton tensor Cα β belongs. First determine the eigenvalues.
This parallels exactly the Petrov classification of the Weyl tensor in four dimen-
sions; see Stephani et al. (2003). This comes about since the Weyl tensor in 4D is
equivalent to a (complex) 3 × 3 tracefree matrix, as Cα β in 3D; for a similar clas-
sification of Cαβ , see Hall and Capocci (1999). A curvature classification in terms
of spinors has been done by Torres del Castillo and Gómez-Ceballos (2003).
References
Anninos, D., Li, W., Padi, M., Song, W., and Strominger, A. (2009). “Warped AdS3
black holes,” JHEP 0903, 130, arXiv:0807.3040.
Aragone, C. (1987). “Topologically massive gravity in the dreibein light-front gauge,”
Class. Quant. Grav. 4, L1.
Ayón-Beato, E., Cataldo, M., and Garcia, A. A. (2005). “Electromagnetic fields in
stationary cyclic symmetric 2+1 gravity”, In Proceedings of the 10th. PASCOS04
and Pran Nath Fest, Eds. G. Alverson, E. Barberi, P. Nath, M.T. Vaughn (World
Scientific, 2005), p.359.
Ayón-Beato, E., Garcia, A. A., Macias, A., and Perez-Sanchez, J. M. (2000). “Note
on scalar fields nonminimally coupled to (2+1) gravity,” Phys. Lett. B495, 164
[gr-qc/0101079].
Ayón-Beato, E. and Hassaı̈ne, M. (2005). “pp–waves of conformal gravity with self-
interacting source,” Annals Phys. 317, 175. hep-th/0409150.
Ayón-Beato, E. and Hassaı̈ne, M. (2006). “Exploring AdS waves via nonminimal
coupling,” Phys. Rev. D 73, 104001. hep-th/0512074.
Ayón-Beato, E., Martinez, C., and Zanelli, J. (2004). “Birkhoff’s theorem for three-
dimensional AdS gravity,” Phys. Rev. D70, 044027. [hep-th/0403227].
Bañados, M., Henneaux, M. Teitelboim, C., and Zanelli, J. (1993). “Geometry of the
(2+1) black hole,” Phys. Rev. D48, 1506. [gr-qc/9302012].
Bañados, M., Teitelboim, C., and Zanelli, J. (1992). “The black hole in three-
dimensional spacetime,” Phys. Rev. Lett. 69, 1849. [hep-th/9204099].
Barrow, J. D., Burd, A. B., and Lancaster, D. (1986). “Three-dimensional classical
spacetimes,” Class. Quant. Grav. 3, 551–567.
Barrow, J. D., Shaw, D. J., and Tsagas, C. G. (2006). “Cosmology in three dimensions:
Steps towards the general solution,” Class. Quant. Grav. 23, 5291. [gr-qc/0606025].
Bouchareb A. and Clément, G. (2007). “Black hole mass and angular momentum in
topologically massive gravity,” Class. Quant. Grav. 24, 5581. arXiv:0706.0263.
Brown, J. D., Creighton, J., and Mann, R. B. (1994). “Temperature, energy and heat
capacity of asymptotically anti-de Sitter black holes,” Phys. Rev. D50, 6394. [gr-
qc/9405007].
Brown, J. D. and York Jr., J. W. (1993). “Quasilocal energy and conserved charges
derived from the gravitational action”, Phys. Rev. D47, 1407.
Carlip, S. (1998). Quantum Gravity in 2+1 Dimensions, Cambridge, UK: Cambridge
University Press.
Carlip, S., Deser, S., Waldron, A., and Wise, D. K. (2008). “Topologically massive AdS
gravity,” Phys. Lett. B666, 272. arXiv:0807.0486.
Carlip, S., Deser, S., Waldron, A., and Wise, D. K. (2009). “Cosmological topologically
massive gravitons and photons,” Class. Quant. Grav. 26, 075008.
422 References
Coley, A., Hervik, S., and Pelavas, N. (2008). “Lorentzian spacetimes with con-
stant curvature invariants in three dimensions,” Class. Quant. Grav. 25, 025008.
arXiv:0710.3903.
Coley, A., Hervik, S., and Pelavas, N. (2009). “Lorentzian spacetimes with con-
stant curvature invariants in four dimensions,” Class. Quant. Grav. 26, 125011.
arXiv:0904.4877.
Coley, A., Hervik, S., Papadopoulos, G. O., and Pelavas, N. (2009). “Kundt space-
times,” Class. Quant. Grav. 26, 105016. arXiv:0901.0394.
Collas, P. (1977). “General relativity in two-and three-dimensional space-times,” Am.
J. Phys. 45, 833.
Cornish, N. J. and Frankel, N. E. (1991). “Gravitation in (2+1)–dimensions,” Phys.
Rev. D43, 2555.
Cornish, N. J. and Frankel, N. E. (1994). “Gravitation versus rotation in (2+1)–
dimensions,” Class. Quant. Grav. 11, 723. [gr-qc/9306012].
Coussaert, O. and Henneaux, M. (1994b). “Self-dual solutions of the 2 + 1 Einstein
gravity with a negative cosmological constant,” [hep-th9407181].
Cruz, N. and Martı́nez, C. (2000). “Cosmological scaling solutions of minimally coupled
scalar fields in three dimensions” Class. Quant. Grav. 17, 2867.
Cruz, N. and Zanelli, J. (1995). “Stellar equilibrium in (2+1)-dimensions,” Class.
Quant. Grav. 12, 975 [gr-qc/9411032].
Dereli, T. and Sarioğlu, O. (2001). “Supersymmetric solutions to topologically mas-
sive gravity and black holes in three dimensions,” Phys. Rev. D 64, 027501.
gr-qc/0009082.
Dereli, T. and Tucker, R. W. (1988). “Gravitational interactions in 2 + 1 dimensions,”
Class. Quant. Grav. 5, 951.
Deser, S. (1984).“Cosmological topological supergravity,” in Quantum Theory of
Gravity, ed. S.M. Christensen, Adam Hilger, London.
Deser, S. and Jackiw, R. (1989). “String sources in (2+1)-dimensional gravity,” Ann.
of Phys. 192, 352.
Deser, S., Jackiw, R., and Pi, S. Y. (2005). “Cotton blend gravity pp waves,” Acta
Phys. Polon. B36, 27. [gr-qc/0409011].
Deser, S., Jackiw, R. and Templeton, S. (1982a). “Three-Dimensional Massive Gauge
Theories,” Phys. Rev. Lett. 48, 975.
Deser, S., Jackiw, R., and Templeton, S. (1982b). “Topological massive gauge theories,”
Ann. of Phys. 140, 372.
Deser, S., Jackiw, R. and ’t Hooft, G. (1984). “Three-dimensional Einstein gravity:
dynamics of flat space,” Ann. of Phys. 152, 220–235.
Deser, S. and Mazur, P. O. (1985). “Static solutions in D = 3 Einstein–Maxwell theory,”
Class. Quant. Grav. 2 L51–L56.
Deser, S. and Steif, A.R. (1992). “Gravity theories with lightlike sources in D = 3,”
Class. Quant. Grav. 9, L153. hep-th/9208018.
Dias, O. J. C. and Lemos, J. P. S. (2002b). “Rotating magnetic solution in three-
dimensional Einstein gravity,” JHEP 01(2002), 006. [hep-th/0201058].
Garbarz, A., Giribet, G., and Vásquez, Y. (2009). “Asymptotically AdS 3 solutions to
topologically massive gravity at special values of the coupling constants,” Phys.
Rev. D 79, 044036. arXiv:0811.4464.
Garcı́a, A. (1999). “On the rotating charged BTZ metric,” hep-th/9909111.
Garcı́a, A. A. (2004). “Stationary circularly symmetric 2+1 rigidly rotating perfect
fluid,” Phys. Rev. D69, 124024.
424 References
Kamata, M. and Koikawa, T. (1997). “2+1-dimensional charged black hole with (anti-)
self dual Maxwell field,” Phys. Lett. B391, 196.
Kogan, I. I. (1992). “About some exact solutions for (2+1) gravity coupled to gauge
fields,” Mod. Phys. Lett. A7, 2341. [hep-th/9205095].
Lubo, M., Rooman, M. and Spindel, P. (1999). “(2+1)-dimensional stars,” Phys. Rev.
D59, 044012 [gr-qc/9806104].
Macı́as, A. and Camacho, A. (2005). “Kerr–Schild metric in topological massive (2+1)
gravity,” Gen. Rel. Grav. 37, 759.
Mann, R. B. and Ross, S. F. (1993). “Gravitationally collapsing dust in 2 + 1
dimensions”, Phys. Rev. D47, 3319.
Martinez, E. A. and Shepley, L. (1986). preprint, University of Texas.
Martı́nez, C., Teitelboim, C. and Zanelli, J. (2000). “Charged rotating black hole in
three space-time dimensions,” Phys. Rev. D61, 104013 [hep-th/9912259].
Martı́nez, C. and Zanelli, J. (1996). “Conformally dressed black hole in (2+1)-
dimensions,” Phys. Rev. D 54, 3830. [gr-qc/9604021].
Matyjasek, J. and Zaslavskii, O. B. (2004). “Extremal limit for charged and rotating
(2+1)–dimensional black holes and Bertotti-Robinson geometry,” Class. Quant.
Grav. 21, 4283 [gr-qc/0404090].
Melvin, M. A. (1986). “Exterior solutions for electric and magnetic stars in 2+1
dimensions,” Class. Quant. Grav. 3, 117-131.
Moussa, K. A., Clément, G., and Leygnac, C. (2003). “The black holes of topologically
massive gravity,” Class. Quant. Grav. 20, L277, gr-qc/0303042.
Nutku, Y. (1993). “Exact solutions of topologically massive gravity with a cosmological
constant,” Class. Quant. Grav. 10, 2657.
Nutku, Y. and Baekler, P. (1989). “Homogeneous, anisotropic three-manifolds of
topologically massive gravity,” Annals Phys. 195, 16.
Ortiz, M. E. (1990). “Homogeneous solutions to topologically massive gravity,” Annals
Phys. 200, 345.
Obukhov, Y. N. (2003). “New solutions in 3-D gravity,” Phys. Rev. D68, 124015.
[gr-qc/0310069].
Ölmez, S., Sarioğlu, O., and Tekin, B. (2005). “Mass and angular momentum of asymp-
totically AdS or flat solutions in the topologically massive gravity,” Class. Quant.
Grav. 22, 4355, gr-qc/0507003.
Peldan, P. (1993). “Unification of gravity and Yang-Mills theory in (2+1)-dimensions,”
Nucl. Phys. B395, 239 [gr-qc/9211014].
Percacci, R., Sodano, P., and Vuorio, I. (1987). “Topologically massive planar universes
with constant twist,” Ann. of Phys. 176, 344.
Rooman, M. and Spindel, P. (1998). “Gödel metric as a squashed anti-de Sitter
geometry,” Class. Quant. Grav. 15, 3241 [gr-qc/9804027].
Saslaw, W. C. (1977). “A relation between the homogeneity of the Universe and the
dimensional of space” Mon. Not. R. Astr. Soc. 179, 659.
Staruszkiewicz, A. (1963). “Gravitation theory in three-dimensional space,” Acta Phys.
Polon. 24, 735.
Torres del Castillo, G. F. and Gómez-Ceballos, L. F. (2003). “Algebraic classification
of the curvature of three-dimensional manifolds with indefinite metric,” J. Math.
Phys. 44, 4374 .
Vuorio, I. (1985). “Topologically massive planar universe,” Phys. Lett. B163, 91.
Gen. Rel. Grav. 23, 181–187.
Zaslavskii, O. B. (1994). “Thermodynamics of 2+1 black holes,” Class. Quant. Grav.
11, L33–L38.
426 References
Additional Literature
Abrahams, A. M. and Evans, C. R. (1993). “Critical behavior and scaling in vacuum
axisymmetric gravitational collapse,” Phys. Rev. Lett. 70, 2980.
Abramowitz, M. and Stegun, I. (1965). Handbook of Mathematical Functions. New
York: Dover Publications Inc.
Adler, R., Bazin, M., and Schiffer, M. (1965). Introduction to General Relativity,
McGraw–Hill.
Arnowitt, R., Deser, S., and Misner, C. W. (1962). “The dynamics of General Rela-
tivity.” In: Gravitation: An Introduction to Current Research, L. Witten (Ed.).
Wiley, New York.
Ashtekar, A., and Krishnan, B. (2004). “Isolated and dynamical horizons and their
applications,” Living Rev. Rel. 7, 10.
Ayón-Beato, E. and Garcı́a, A. (1998). “Regular black hole in general relativity coupled
to a nonlinear electrodynamics,” Phys. Rev. Lett. 80, 5056.
Bach, R. (1921). “Zur Weylschen Relativitätstheorie und der Weylschen Erweiterung
des Krümmungsbegriffs,” Math. Zeitschr. 9, 110–135.
Baekler, P., Mielke, E. W., and Hehl, F. W. (1992). “Dynamical symmetries in
topological 3D gravity with torsion,” Nuovo Cimento. B107, 91–110.
Barrow, J. D., and Saich, P. (1993). “Scalar–field cosmologies,” Class. Quant. Grav.
10, 279.
Bergshoeff, E. A., Hohm, O., and Townsend, P. K. (2009) “Massive gravity in three
dimensions,” Phys. Rev. Lett.102, 201301.
Bertotti, I. (1959). “Uniform electromagnetic field in the theory of general relativity,”
Phys. Rev. 116, 1331–1333.
Bini, D., Jantzen, R. T., and Minutti, G. (2001). “The Cotton, Simon–Mars and
Cotton–York tensors in stationary spacetimes,” Class. Quant. Grav. 18, 4969.
Borde, A. (1997). “Regular black holes an topological change,” Phys. Rev. D50, 7615.
Born, M., and Infeld, L. (1934). “On the quantum theory of the electromagnmetic
field,” Proc. Roy. Soc. (London) A144, 425.
Brinkmann, M. W. (1923). “On Riemann spaces conformal to Euclidean spaces,” Proc.
Natl. Acad. Sci. U.S. 9, 1.
Buchdahl, H. A. (1959). “General relativistic fluid spheres,” Phys. Rev. 116, 1027.
Buscher, T. (1993). “Path integral derivation of quantum duality in nonlinear sigma
models,” Phy. Lett.B201, 466.
Carter, B. (1968). “Hamilton–Jacobi and Schrodinger separable solutions of Ein-
stein’s equations,”Commun. Math. Phys. 10, 280 (1968), (1968). “A new family of
Einstein spaces,” Phys. Lett. A26, 399.
Cotton, E. (1899). “Sur les variétés a trois dimensions,” Ann. Fac. d. Sc. Toulouse (II)
1, 385.
Choptuik, M. W. (1993). “Universality and scaling in gravitational collapse of a
massless scalar field,” Phys. Rev. Lett. 70, 9.
Christodoulou, D. and Klainerman, S. (1993). The Global Nonlinear Stability of the
Minkowski Space. Princeton University Press, Princeton, NJ.
de Rham, C. (2014). “Massive Gravity,” Living Rev.Rel. 17, 7. arXiv:1401.4173
[hep-th].
Deser, S. and Gibbons, G. W. (1998). “Born–Infeld–Einstein actions,” Class. Quant.
Grav. 15, L35.
Ehlers, J. (1961). “Beiträge zur realativistischen Mechanik kontinuierlich Medien,” Abh.
Mainzer Akad. Wiss., Math.–natuwiss,” No, 11, English Translation: (1993) “Con-
tributions to relativistic mechanics of continuous media,” Gen. Rel. Grav. 25,
References 427
1225–1266. See also, (2002). “Relativistic hydrodynamics,” Gen. Rel. Grav. 34,
2171.
Eisenhart, L. P. (1966). Riemannian Geometry, Princeton University Press. Princeton,
NJ.
Ellis, G. F. R. and Elst, H. (1999). Theoretical and Observational Cosmology, Ed. ML
Lachieze-Rey (Dordrecht: Kluwer) p.1.
Ferrando, J. J. and Sáez, A. J. (2002). “On the classification of Type D spacetimes,”
[gr-qc/0212086].
Fradkin, E. and Tseytlin, A. (1985). “Nonlinear electrodynamics from quantized
strings,” Phys. Lett. B163, 123.
Frolov, V., Hendy, S., and Larsen, A. L. (1996). “Stationary strings and principal Killing
triads in 2+1 gravity,” Nucl. Phys. B 468, 336.
Garcı́a, A. A. (1988). “Comments on a paper by Collinson,” Gen. Rel. Grav. 20, 589.
Gibbons, G. W. and Rasheed, D. A. (1995). “Electric–magnetic duality rotations in
non-linear electrodynamics,” Nucl. Phys. B454, 185.
Gubser, S. S., Klebanov, I. R., and Polyakov, A. M. (1998). “Gauge Theory Correlators
from Non-Critical String Theory,” Phys. Lett. B428, 105 (arXiv: hep-th/9802109).
Gundlach, C. (1999). “Critical Phenomena in Gravitational Collapse,” Living Rev. Rel.
2, 4 (arXiv: gr-qc/0001046).
Guralnik, G., Iorio, A., Jackiw, R., and Pi, S.-Y. (2003). “Dimensionally reduced gravi-
tational Chern–Simons term and its kink,” Ann. Phys. (NY) 308, 222–236. (arXiv:
hep-th/0305117).
Gürses, M. and Gürsey, Y. (1975). “Conformal uniqueness and various forms of the
Schwarzschild interior metric,” Nuovo Cimento 25 B, 786.
Hayward, S. A. (2008). “Dynamics of black holes,” arXiv:0810.0923 [gr-qc].
Hehl, F. W., McCrea, J. D., Mielke, E. W., and Ne’eman, Y. (1995). “Metric–affine
gauge theory of gravity: Field equations, Noether identities, world spinors, and
breaking of dilation invariance,” Phys. Repts. 258, 1.
Heinicke, C. (2001). “The Einstein 3-form Gα and its equivalent 1–form Lα in Riemann–
Cartan spacetime,” Gen. Relat. Grav. 33, 1115.
Israel, R. B. (1966). “Singular hypersurfaces and thin shells in General Relativity,”
Nuovo Cimento. XLIV B, 1.
Infeld, L. and Plebański, J. F. (1960). Motion and Relativity, Pergamon Press, New
York.
Jackiw, R. S. and Pi, Y. (2003). “Chern–Simons Modification of General Relativity,”
arXiv: gr-qc/0308071.
Korn, G. A. and Korn, T. M. (1961). Mathematics Handbook for Scientists and
Engineers: Definitions, Theorems, and Formulas for References and Review,
McGraw–Hill, New York.
Kottler, F. (1918). “Über die physikalischen Grudlagen der Einsteischen Gravitations-
theorie,” Annalen Physik 56, 410.
Kramer, D., Stephani, H., MacCallum, M., and Hertl, E. (1980). Exact Solutions of
the Einstein’s Field Equations, Deutsch. Ver. der Wiss., Berlin.
Landau, L. and Lifshitz, E. (1970). Théorie des Champs, Mir, Moscow,
Liddle, A. R. and Lyth, D. H. (2000). Cosmological Inflation and Large-Scale Structure,
Cambridge University Press.
Lucchin, F. and Matarrese, S. (1985). “Power–law inflation,” Phys. Rev. D32, 1316.
Madsen, M. S. (1986). “An unusual cosmological solutions for Λφ4 theory with broken
symmetry,” Gen. Rel. and Grav. 18, 879.
428 References