Janis-Newman Algorithm: Generating Rotating and NUT Charged Black Holes
Janis-Newman Algorithm: Generating Rotating and NUT Charged Black Holes
Abstract
In this review we present the most general form of the Janis–Newman algorithm.
This extension allows to generate configurations which contain all bosonic fields with
spin less than or equal to two (real and complex scalar fields, gauge fields, metric field)
and with five of the six parameters of the Plebański–Demiański metric (mass, electric
charge, magnetic charge, NUT charge and angular momentum). Several examples are
included to illustrate the algorithm. We also discuss the extension of the algorithm to
other dimensions.
1
Contents
1 Introduction 5
1.1 Motivations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2 The Janis–Newman algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.4 Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.5 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
4 Complete algorithm 19
4.1 Seed configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4.2 Janis–Newman algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.2.1 Complex transformation . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.2.2 Function transformation . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.2.3 Null coordinates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.2.4 Boyer–Lindquist coordinates . . . . . . . . . . . . . . . . . . . . . . . 22
4.3 Open questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2
6 Examples 30
6.1 Kerr–Newman–NUT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
6.2 Charged (a)dS–BBMB–NUT . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
6.3 Ungauged N = 2 BPS solutions . . . . . . . . . . . . . . . . . . . . . . . . . . 32
6.3.1 Pure supergravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
6.3.2 STU model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
6.4 Non-extremal rotating solution in T 3 model . . . . . . . . . . . . . . . . . . . 35
6.5 SWIP solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
6.6 Gauged N = 2 non-extremal solution . . . . . . . . . . . . . . . . . . . . . . . 38
A Coordinate systems 53
A.1 d-dimensional . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
A.1.1 Cartesian system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
A.1.2 Spherical . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
A.1.3 Spherical with direction cosines . . . . . . . . . . . . . . . . . . . . . . 54
A.1.4 Spheroidal with direction cosines . . . . . . . . . . . . . . . . . . . . . 55
A.1.5 Mixed spherical–spheroidal . . . . . . . . . . . . . . . . . . . . . . . . 55
A.2 4-dimensional . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
A.2.1 Cartesian system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
A.2.2 Spherical . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
A.2.3 Spherical with direction cosines . . . . . . . . . . . . . . . . . . . . . . 56
A.2.4 Spheroidal with direction cosines . . . . . . . . . . . . . . . . . . . . . 56
A.3 5-dimensional . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
A.3.1 Spherical with direction cosines . . . . . . . . . . . . . . . . . . . . . . 56
A.3.2 Hopf coordinates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
C Technical properties 58
C.1 Group properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
C.2 Chaining transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
C.3 Arbitrariness of the transformation . . . . . . . . . . . . . . . . . . . . . . . . 60
3
References 60
4
1 Introduction
1.1 Motivations
General relativity is the theory of gravitational phenomena. It describes the dynamical evol-
ution of spacetime through the Einstein–Hilbert action that leads to Einstein equations. The
latter are highly non-linear differential equations and finding exact solutions is a notoriously
difficult problem.
There are different types of solutions but this review will cover only black-hole-like solu-
tions (type-D in the Petrov classification) which can be described as particle-like objects
that carry some charges, such as a mass or an electric charge.
Black holes are important objects in any theory of gravity for the insight they provide
into the quantum gravity realm. For this reason it is a key step, in any theory, to obtain
all possible black holes solutions. Rotating black holes are the most relevant subcases for
astrophysics as it is believed that most astrophysical black holes are rotating. These solutions
may also provide exterior metric for rotating stars.
The most general solution of this type in pure Einstein–Maxwell gravity with a cosmolo-
gical constant Λ is the Plebański–Demiański metric [1, 2]: it possesses six charges: mass m,
NUT charge n, electric charge q, magnetic charge p, spin a and acceleration α. A challenging
work is to generalize this solution to more complex Lagrangians, involving scalar fields and
other gauge fields with non-minimal interactions, as is typically the case in supergravity. As
the complexity of the equations of motion increase, it is harder to find exact analytical solu-
tions, and one often consider specific types of solutions (extremal, BPS), truncations (some
fields are constant, equal or vanishing) or solutions with restricted number of charges. For
this reason it is interesting to find solution generating algorithms – procedures which trans-
form a seed configuration to another configuration with a greater complexity (for example
with a higher number of charges).
An on-shell algorithm is very precious because one is sure to obtain a solution if one starts
with a seed configuration which solves the equations of motion. On the other hand off-shell
algorithms do not necessarily preserve the equations of motion but they are nonetheless very
useful: they provide a motivated ansatz, and it is always easier to check if an ansatz satisfy
the equations than solving them from scratch. Even if in practice this kind of solution
generating technique does not provide so many new solutions, it can help to understand
better the underlying theory (which can be general relativity, modified gravities or even
supergravity) and it may shed light on the structure of gravitational solutions.
5
the literature. We stress that all results are totally equivalent in both approaches, and every
computation that can be done with the Giampieri prescription can be done with the other.
Finally [6] provides an alternative view on the algorithm.
In order for the metric to be still real, the radial functions inside the metric must be
transformed such that reality is preserved.1 Despite that there is no rigorous statement
concerning the possible complexification of these functions, some general features have been
worked out in the last decades and a set of rules has been established. Note that this
step is the same in both prescriptions. In particular these rules can be obtained by solving
the equations of motion for some examples and by identifying the terms in the solution [7].
Another approach consists in expressing the metric functions in terms of the Boyer–Lindquist
functions – that appear in the change of coordinates and which are real –, the latter being
then determined from the equations of motion [8, 9].
It is widely believed that the JN algorithm is just a trick without any physical or math-
ematical basis, which is not accurate. Indeed it was proved by Talbot [10] shortly after its
discovery why this transformation was well-defined, and he characterizes under which con-
ditions the algorithm is on-shell for a subclass of Kerr–Schild (KS) metrics (see also [11]).2
KS metrics admit a very natural formulation in terms of complex functions for which (some)
complex change of coordinates can be defined. Note that KS metrics are physically inter-
esting as they contain solutions of Petrov type II and D. Another way to understand this
algorithm has been provided by Schiffer et al. [12] (see also [13]) who showed that some KS
metrics can be written in terms of a unique complex generating function, from which other
solutions can be obtained through a complex change of coordinates. In various papers, New-
man shows that the imaginary part of complex coordinates may be interpreted as an angular
momentum, and there are similar correspondences for other charges (magnetic. . . ) [14–16].
More recently Ferraro shed a new light on the JN algorithm using Cartan formalism [17].
Uniqueness results for the case of pure Einstein theory have been derived in [8]. A recent
account on these different points can be found in [18].
In its current form the algorithm is independent of the gravity theory under considera-
tion since it operates independently at the level of each field in order to generate an ansatz,
and the equations of motion are introduced only at the end to check if the configuration
is a genuine solution. We believe that a better understanding of the algorithm would lead
to an on-shell formulation where the algorithm would be interpreted as some kind of sym-
metry or geometric property. One intuition is that every configuration found with the JN
algorithm and solving the equations of motion is derived from a seed that also solves the
equations of motion (in particular no useful ansatz has been generated from an off-shell seed
configuration).
Other solution generating algorithms rely on a complex formulation of general relativity
which allows complex changes of coordinates. This is the case of the Ernst potential formu-
lation [19, 20] or of Quevedo’s formalism who decomposes the Riemann tensor in irreducible
representations of SO(3, C) ∼ SO(3, 1) and then uses the symmetry group to generate new
solutions [21, 22].
Despite its long history the Janis–Newman algorithm has not produced any new rotating
solution for non-fluid configurations (which excludes radiating and interior solutions) beside
the Kerr–Newman metric [4], and very few known examples have been reproduced [3, 23–
26]. Generically the application the Janis–Newman algorithm to interior and radiating
systems [9, 27–31] consist in deriving a configuration that do not solve the equations of
motion by itself and to interpret the mismatch as a fluid (whose properties can be studied)
– in this review we will not be interested by this kind of applications. Moreover the only
1 For simplifying, we will say that we complexify the functions inside the metric when we perform this
transformation, even if in practice we "realify" them.
2 It has not been proved that the KS condition is necessary, but all known examples seem to fit in this
category.
6
solutions that have been fully derived using the algorithm are the original Kerr metric [3],
the d = 3 BTZ black hole [24, 25] and the d-dimensional Myers–Perry metric with a single
angular momentum [23]: only the metric was found in the other cases [4, 26] and the other
fields had to be obtained using the equations of motion.
A first explanation is that there is no real understanding of the algorithm in its most
general form (as reviewed above it is understood in some cases): there is no geometric or
symmetry-related interpretation. Another reason is that the algorithm has been defined only
for the metric (and real scalar fields) and no extension to the other types of fields was known
until recently. It has also been understood that the algorithm could not be applied in the
presence of a cosmological constant [7]: in particular the (a)dS–Kerr(–Newman) metrics [32]
(see also [1, 2, 33, 34]) cannot be derived in this way despite various erroneous claims [30,
35]. Finally many works [36–44] (to cite only few) are (at least partly) incorrect or not
reliable because they do not check the equations of motion or they perform non-integrable
Boyer–Lindquist changes of coordinates [31, 45, 46].
The algorithm was later shown to be generalizable by Demiański and Newman who
demonstrated by writing a general ansatz and solving the equations of motion that other
parameters can be added [7, 47], even in the presence of a cosmological constant. While one
parameter corresponds to the NUT charge, the other one did not receive any interpretation
until now.3 Unfortunately Demiański did not express his result to a concrete algorithm
(the normal prescription fails in the presence of the NUT charge and of the cosmological
constant) which may explain why this work did not receive any further attention. Note that
the algorithm also failed in the presence of magnetic charges.
A way to avoid problems in defining the changes of coordinates to the Boyer–Lindquist
system and to find the complexification of the functions has been proposed in [8] and exten-
ded in [31]: the method consists in writing the unknown complexified function in terms of
the functions of the coordinate transformation. This philosophy is particularly well-suited
for providing an ansatz which does not relies on a static seed solution.
More recently it has been investigated whether the JN algorithm can be applied in mod-
ified theories of gravity. Pirogov put forward that rotating metrics obtained from the JN
algorithm in Brans–Dicke theory are not solutions if α 6= 1 [51]. Similarly Hansen and
Yunes have shown a similar result in quadratic modified gravity (which includes Gauss–
Bonnet) [52].4 These do not include Sen’s dilaton–axion black hole for which α = 1 (sec-
tion 6.4), nor the BBMB black hole from conformal gravity (section 6.2). Finally it was
proved in [54] that it does not work either for Einstein–Born–Infled theories.5 We note that
all these no-go theorem have been found by assuming a transformation with only rotation.
Previous reviews of the JN algorithm can be found in [18, 55, chap. 19, 8, 38, sec. 5.4]
(see also [56]).
1.3 Summary
The goal of the current work is to review a series of recent papers [57–60] in which the
JN algorithm has been extended in several directions, opening the doors to many new
applications. This review evolved from the thesis of the author [61], which presented the
material from a slightly different perspective, and from lectures given at Hri (Allahabad,
India).
3 Demiański’s metric has been generalized in [48–50].
4 There are some errors in the introduction of [52]: they report incorrectly that the result from [51]
implies that Sen’s black hole cannot be derived from the JN algorithm, as was done by Yazadjiev [26]. But
this black hole corresponds to α = 1 and as reported above there is no problem in this case (see [53] for
comparison). Moreover they argue that several works published before 2013 did not take into account the
results of Pirogov [51], published in 2013. . .
5 It may be possible to circumvent the result of [54] by using the results described in this review since
7
As explained in the previous section, the JN algorithm was formulated only for the metric
and all other fields had to be found using the equations of motion (with or without using an
ansatz). For example neither the Kerr–Newman gauge field or its associated field strength
could be derived in [4]. The solution to this problem is to perform a gauge transformation
in order to remove the radial component of the gauge field in null coordinates [57]. It is
then straightforward to apply the JN algorithm in either prescription.6 Another problem was
exemplified by the derivation of Sen’s axion–dilaton rotating black hole [63] by Yazadjiev [26],
who could find the metric and the dilaton, but not the axion (nor the gauge field). The reason
is that while the JN algorithm applies directly to real scalar fields, it does not for complex
scalar fields (or for a pair of real fields that can naturally be gathered into a complex scalar).
Then it is necessary to consider the complex scalar as a unique object and to perform the
transformation without trying to keep it real [60]. Hence this completes the JN algorithm
for all bosonic fields with spin less than or equal to two.
A second aspect for which the original form of the algorithm was deficient is configuration
with magnetic and NUT charges and in presence of a cosmological constant. The issue
corresponds to finding how one should complexify the functions: the usual rules do not work
and if there were no way to obtain the functions by complexification then the JN algorithm
would be of limited interest as it could not be exported to other cases (except if one is willing
to solve equations of motion, which is not the goal of a solution generating technique). We
have found that to reproduce Demiański’s result [7] it is necessary to complexify also the
mass and to consider the complex parameter m + in [59, 60] and to shift the curvature
of the spherical horizon. Similarly for configurations with magnetic charges one needs to
consider the complex charge q + ip [60]. Such complex combinations are quite natural from
the point of view of the Plebański–Demiański solution [1, 2] described previously. It is
to notice that the appearance of complex coordinate transformations mixed with complex
parameter transformations was a feature of Quevedo’s solution generating technique [21, 22],
yet it is unclear what the link with our approach really is. Hence the final metric obtained
from the JN algorithm may contain (for vanishing cosmological constant) five of the six
Plebański–Demiański parameters [1, 2] along with Demiański’s parameter.
An interesting fact is that the previous argument works in the presence of the cosmolo-
gical constant only if one considers the possibility of having a generic topological horizons
(flat, hyperbolic or spherical) and for this reason we have provided an extension of the
formalism to this case [59].
We also propose a generalization of the algorithm to any dimension [58], but while new
examples could be found for d = 5 the program could not be carried to the end for d > 5.
All these results provide a complete framework for most of the theories of gravity that
are commonly used. As a conclusion we summarize the features of our new results:
• all bosonic fields with spin ≤ 2;
• topological horizons;
• charges m, n, q, p, a (with a only for Λ = 0);
• extend to d = 3, 5 dimensions (and proposal for higher).
We have written a general Mathematica package for the JN algorithm in Einstein–Maxwell
theory.7 Here is a list of new examples that have been completely derived using the previous
results (all in 4d except when said explicitly):
6 Another solution has been proposed by Keane [62] but it is applicable only to the Newman–Penrose
coefficients of the field strength. Our proposal requires less computations and yields directly the gauge field
from which all relevant quantities can easily be derived.
7 Available at http://www.lpthe.jussieu.fr/~erbin/.
8
• Kerr–Newman–NUT;
• dyonic Kerr–Newman;
• Yang–Mills Kerr–Newman black hole [64];
• adS–NUT Schwarzschild;
• Demiański’s solution [7];
• ungauged N = 2 BPS solutions [65];
• non-extremal solution in T 3 model [63] (partly derived in [26]);
• 5d BMPV [69];
• NUT charged black hole8 in gauged N = 2 sugra with F = −i X 0 X 1 [70].
We also found a more direct derivation of the rotating BTZ black hole (derived in another
way by Kim [24, 25]).
1.4 Outlook
A major playground for this modified Janis–Newman (JN) algorithm is (gauged) super-
gravity – where many interesting solutions remain to be discovered – since all the neces-
sary ingredients are now present. Moreover important solutions are still missing in higher-
dimensional Einstein–Maxwell (in particular the charged Myers–Perry solution) and one can
hope that understanding the JN algorithm in higher dimensions would shed light on this
problem. Another open case is whether black rings can also be derived using the algorithm.
A major question about the JN algorithm is whether it is possible to include rotation for
non-vanishing cosmological constant. A possible related problem concerns the addition of
acceleration α, which is the only missing parameter when Λ = 0. It is indeed puzzling that
one could get all Plebański–Demiański parameters but the acceleration, which appears in
the combination a + iα. Both problems are linked to the fact that the JN algorithm – in its
current form – does not take into account various couplings between the parameters (such
as the spin with the cosmological constant or the acceleration with the mass in the simplest
cases). On the other hand it does not mean that it is impossible to find a generalization
of the algorithm: philosophically the problem is identical to the ones of adding NUT and
magnetic charges.
In any case the meaning and a rigorous derivation of the JN algorithm – perhaps elevating
it to the status of a true solution generating algorithm – are still to be found. It is also
interesting to note that almost all of the examples quoted in the previous section can be
embedded into N = 2 supergravity. This calls for a possible interpretation of the algorithm
in terms of some hidden symmetry of supergravity, or even of string theory.
We hope that our new extension of the algorithm will help to bring it outside the shadow
where it stayed since its creation and to establish it as a standard tool for deriving new
solutions in the various theories of gravity.
8 Derived by D. Klemm and M. Rabbiosi, unpublished work.
9
1.5 Overview
In section 2 we review the original Janis–Newman algorithm and its alternative form due
to Giampieri before illustrating it with some examples. Section 3 shows how to extend the
algorithm to more complicated set of fields (complex scalars, gauge fields) and parameters
(magnetic and NUT charges, topological horizon). Then section 4 provides a general de-
scription of the algorithm in its most general form. The complex transformation described
in the previous section are derived in section 5. Section 6 describes several examples. Finally
section 7 extends first the algorithm to five dimensions and section 8 generalizes these ideas
to any dimension.
Appendix A gathers useful formulas on coordinate systems in various numbers of di-
mensions. Appendix B reviews briefly the main properties of N = 2 supergravity. Finally
appendix C discusses some additional properties of the JN algorithm.
In our conventions the spacetime signature is mostly plus.
2.1 Summary
The general procedure for the Janis–Newman algorithm can be summarized as follows:
1. Perform a change of coordinates (t, r) to (u, r) and a gauge transformation such that
grr = 0 and Ar = 0.
2. Take u, r ∈ C and replace the functions fi (r) inside the real fields by new real-valued
functions f˜i (r, r̄) (there is a set of “empirical” rules).
3. Perform a complex change of coordinates and transform accordingly:
(a) the tensor structure, i.e. the dxµ (two prescriptions: Janis–Newman [3] and
Giampieri [5]);
(b) the functions f˜i (r, r̄).
4. Perform a change of coordinates to simplify the metric (for example to Boyer–Lindquist
system). If the transformation is infinitesimal then one should check that it is a valid
diffeomorphism, i.e. that it is integrable.
Note that in the last point the operations (a) and (b) are independent. In practice one
is performing the algorithm for a generic class of configurations with unspecified fi (r) in
order to obtain general formulas. One leaves point 2 and (3b) implicit since the other steps
are independent of the form of the functions. Then given a specific configuration one can
perform 2 and (3b).
10
Throughout the review we will not be interested in showing that the examples discussed
are indeed solutions but merely to explain how to extend the algorithm. All examples we
are discussing have been shown to be solutions of the theory under concerned and we refer
the reader to the original literature for more details. For this reason we will rarely mention
the action or the equations of motion and just discussed the fields and their expressions.
One could add a fifth point to the list: checking the equations of motion. We stress again
that the algorithm is off-shell and there is no guarantee (except in some specific cases [18])
that a solution is mapped to a solution.
2.2 Algorithm
We present the algorithm for a metric gµν and a gauge field Aµ associated with a U(1) gauge
symmetry. This simple case is sufficient to illustrate the main features of the algorithm.
As already mentioned in the introduction, the authors of [4] failed to derive the field
strength of the Kerr–Newman black hole from the Reissner–Nordström one. In the null
tetrad formalism it is natural to write the field strength in terms of its Newman–Penrose
coefficients, but a problem arises when one tries to generate the rotating solution since
one of the coefficients is zero in the case of Reissner–Nordström, but non-zero for Kerr–
Newman. Three different prescriptions have been proposed: two works in the Newman–
Penrose formalism – one with the field strength [62] and one with the gauge field [57] –
while the third extends Giampieri’s approach to the gauge field [57]. Since the proposals
from [57] fit more directly (and parallel each other) inside the prescriptions of Janis–Newman
and Giampieri, we will focus on them. It is also more convenient to work with the gauge
fields since any other quantity can be easily computed from them.
ds2 = −f (r) dt2 + f (r)−1 dr2 + r2 dΩ2 , dΩ2 = dθ2 + H(θ)2 dφ2 , (2.1a)
A = fA (r) dt. (2.1b)
The normalized curvature of the (θ, φ) sections (or equivalently of the horizon) is denoted
by κ (
+1 S 2 ,
κ= (2.2)
−1 H 2
where S 2 and H 2 are respectively the sphere and the hyperboloid,9 and one has
(
sin θ κ = 1,
H(θ) = (2.3)
sinh θ κ = −1.
In all this section we will consider the case of spherical horizon with κ = 1.
Introduce Eddington–Finkelstein coordinates (u, r)
du = dt − f −1 dr (2.4)
in order to remove the grr term of the metric [3]. Under this transformation the gauge field
becomes
A = fA (du + f −1 dr). (2.5)
9 We leave aside the case of the plane R2 with κ = 0. The formulas can easily be extended to this case.
11
The changes of coordinate has introduced an Ar component but since it depends only on
the radial coordinate Ar = Ar (r) it can be removed by a gauge transformation.
At the end the metric and gauge fields are
This last step was missing in [4] and explains why they could not derive the full solution
from the algorithm. The lesson to draw is that the validity of the algorithm depends a lot
on the coordinate basis10 and of the parametrization of the fields, although guiding principle
founded on all known examples seems that one needs to have
grr = 0, Ar = 0. (2.7)
0 0 1 0
such that
g µν = η ab Zaµ Zbν = −`µ nν − `ν nµ + mµ m̄ν + mν m̄µ . (2.11)
The explicit tetrad expressions are
1
f i
` =
µ
δrµ , n =
µ
δuµ − δrµ , m =√
µ
δθµ + δµ (2.12)
2 2r̄ sin θ φ
12
Consistency implies that one recovers the seed for r̄ = r and ū = u.
Finally one can perform a complex change of coordinates
where a is a parameter (to be interpreted as the angular momentum per unit of mass) and
r0 , u0 ∈ R. While this transformation seems arbitrary we will show later (sections 4 and 5)
how to extend it and that general consistency limits severely the possibilities. The tetrads
transform as vectors
∂x0µ ν
Za0µ = Z (2.16)
∂xν a
and this lead to the expressions
f˜ µ
`0µ = δrµ , n0µ = δuµ − δ ,
2 r
(2.17)
1
i
m 0µ
=√ δθµ + µ
δ − ia sin θ (δu − δr ) .
µ µ
2(r + ia cos θ)
0 sin θ φ
After inverting the contravariant form of the metric and the gauge field one is lead to the
final expressions
ds02 = −f˜ (du0 − a sin2 θ dφ)2 − 2 (du0 − a sin2 θ dφ)(dr0 + a sin2 θ dφ) + ρ2 dΩ2 , (2.18a)
A0 = f˜A (du0 − a sin2 θ dφ). (2.18b)
where
2
ρ2 = |r| = r02 + a2 cos2 θ. (2.19)
The coordinate dependence of the functions can be written as
in the new coordinates (and similarly for f˜A ), but note that the θ dependence is not arbitrary
and comes solely from Im r.
Note that it should be done only in the infinitesimal transformation and not elsewhere in
the metric. Although some authors [17, 30] mentioned the equivalence between the tetrad
computation and (2.21), it is surprising that this direction has not been followed further.
While this new prescription is not rigorous and there is no known way to derive (2.23),
it continues to hold for the most general seed (section 4) and it gives systematically the
13
same results as the Janis–Newman prescription, as can be seen by simple inspection. In
particular this approach is not adding nor removing any of the ambiguities due to the
function transformations that are already present and well-known in JN algorithm. Since
this prescription is much simpler we will continue to use it throughout the rest of this review
(we will show in section 4 how it is modified for topological horizons).
Finally the comparison of the two prescriptions clearly shows that the r2 factor in front
of dΩ2 should be considered as a function instead of a part of the tensor structure: the
replacement r2 → ρ2 is dictated by the rules given in the next section. We did not want to
enter into these subtleties here but this will become evident in section 4.
The idea is to use geometric or arithmetic means. All other functions can be reduced to a
2
combination of them, for example 1/r2 is complexified as 1/|r| .
Every known configuration which does not involve a magnetic, a NUT charge, complex
scalar fields or powers higher of rthan quadratic can be derived with these rules (these
cases will be dealt with in sections 3 and 4). Hence despite the fact that there is some
arbitrariness, it is ultimately quite limited and very few options are possible in most cases.
r 2 + a2 a
g(r) = , h(r) = (2.26)
∆ ∆
where we have defined
∆(r) = f˜ρ2 + a2 sin2 θ. (2.27)
As indicated by the r-dependence this change of variables is integrable provided that g and
h are functions of r only. However ∆ as given in (2.27) for a generic configuration contains a
θ dependence: one should check that this dependence cancels once restricted to the example
of interest. Otherwise one is not allowed to perform this change of coordinates (but other
systems may still be found).
14
Given (2.26) one gets the metric and gauge fields (deleting the prime)
ρ2 2 Σ2
ds2 = −f˜ dt2 + dr + ρ2 dθ2 + 2 sin2 θ dφ2 + 2a(f˜ − 1) sin2 θ dtdφ, (2.28a)
∆ ρ
ρ2
A = f˜A dt − dr − a sin2 θ dφ (2.28b)
∆
with
Σ2
= r2 + a2 + agtφ . (2.29)
ρ2
The rr-term has been computed from
ρ2
g − a sin2 θ h = . (2.30)
∆
Generically the radial component of the gauge field depends only on radial coordinate Ar =
Ar (r) (θ-dependence of the function f˜A sits in a factor 1/ρ2 which cancels the one in front
of dr) and one can perform a gauge transformation in order to set it to zero, leaving
2.3 Examples
2.3.1 Flat space
It is straightforward to check that the algorithm applied to the Minkowski metric – which
has f = 1, leading to f˜ = 1 – in spherical coordinates
gives again the Minkowski metric but in spheroidal coordinates (A.31) (after a Boyer–
Lindquist transformation)
ρ2
ds2 = −dt2 + dr2 + ρ2 dθ2 + (r2 + a2 ) sin2 θ dφ2 , (2.33)
r2 + a2
recalling that ρ2 = r2 + a2 cos2 θ. The metric is exactly diagonal because gtφ = 0 for f˜ = 1
from (2.28a).
Hence for flat space the JN algorithm reduces to a change of coordinates, from spherical
to (oblate) spheroidal coordinates: the 2-spheres foliating the space in the radial direction
are deformed to ellipses with semi-major axis a.
This fact is an important consistency check that will be useful when extending the
algorithm to higher dimensions (section 8) or to other coordinate systems (such as one with
direction cosines). Moreover in this case one can forget about the time direction and consider
only the transformation of the radial coordinate.
2.3.2 Kerr–Newman
The seed function is the Reissner–Nordström for which the metric and gauge field are
2m q 2 q
f (r) = 1 − + 2, fA = . (2.34)
r r r
15
Applications of the rules (2.24) leads to
2m Re r q2 q 2 − 2mr0
f˜ = 1 − 2 + 2 =1+ , (2.35a)
|r| |r| ρ2
q Re r qr0
f˜A = 2 = ρ2 . (2.35b)
|r|
These functions together with (2.28) describe correctly the Kerr–Newman solution [18, 71].
For completeness we spell out the expressions of the quantities appearing in the metric
Σ2 q 2 − 2mr 2 2
= r 2
+ a2
− a sin θ, (2.36a)
ρ2 ρ2
∆ = r2 − 2mr + a2 + q 2 . (2.36b)
In particular ∆ does not contain any θ dependence and the BL transformation is well defined.
Moreover the radial component of the gauge field is
f˜A ρ2 qr
Ar = − = (2.37)
∆ ∆
and it is independent of θ.
Z = q + ip. (3.2)
16
On the other hand transforming directly the r inside fA according to (2.24) does not yield
the correct result. Instead one needs to first rewrite the gauge field function as
Z
fA = Re (3.7)
r
Note that it not useful to replace p by Im Z in (3.4) since it is not accompanied by any
2
r dependence. Moreover it is natural that the factor |Z| appears in the metric and this
explains why the charges there do not mix with the coordinates.
The gauge field in BL coordinates is finally
qr − pa cos θ p(r2 + a2 )
qr
A= dt + − 2 a sin θ +
2
cos θ dφ (3.9a)
ρ2 ρ ρ2
qr p cos θ
= 2 (dt − a sin2 θdφ) + a dt + (r2 + a2 ) dφ . (3.9b)
ρ ρ2
The radial component has been removed thanks to a gauge transformation since it depends
only on r
qr − pa cos θ 2
∆ × Ar = − ρ − pa cos θ = −qr. (3.10)
ρ2
There is a coupling between the parameters a and p which can be interpreted from the
fact that a rotating magnetic charge has an electric quadrupole moment. This coupling is
taken into account from the product of the imaginary parts which yield a real term. In view
of the form of the algorithm such contribution could not arise from any other place. Moreover
the combination Z = q + ip appears naturally in the Plebański–Demiański solution [1, 2].
The Yang–Mills Kerr–Newman black hole found by Perry [64] can also be derived in this
way, starting from the seed
qI 2
AI = dt + pI cos θ dφ, |Z| = q I q I + pI pI (3.11)
r
where q I and pI are constant elements of the Lie algebra.
The cosmological constant is denoted by Λ. We give only the main formulas to motivate the
modification of the algorithm, leaving the details of the transformation for section 4.
The complex transformation that adds a NUT charge is
17
Note that it is κ and not κ0 that appears in m. After having shown
The metric derived from the seed (2.1a) is
ds2 = −f˜ (dt − 2κnH 0 (θ) dφ)2 + f˜−1 dr2 + ρ2 dΩ2 , (3.14)
see (4.33), where
ρ2 = r02 + n2 . (3.15)
The function corresponding to the (a)dS–Schwarzschild metric is
2m Λ 2 m Λ
f =κ− − r = κ − 2 Re − r2 . (3.16)
r 3 r 3
The transformation is
2 Re(mr̄) Λ 2 4Λ 2 2 m0 r0 + κ0 − 4Λ
Λ 2
˜ n2 n2
f =κ− − |r| = κ −
0
n − 3
− ρ (3.17)
|r|
2 3 3 ρ2 3
which after simplification gives
2m0 r0 + 2κ0 n2 Λ 02 8Λ n4
f˜ = κ0 − − (r + 5n 2
) + (3.18)
ρ2 3 3 ρ2
which corresponds correctly to the function of (a)dS–Schwarzschild–NUT [73].
Note that it is necessary to consider the general case of massive black hole with topological
horizon (if Λ 6= 0 for the latter) even if one is ultimately interested in the m = 0 or κ = 1
cases.
The transformation (3.13) can be interpreted as follows. In similarity with the case of
the magnetic charge, writing the mass as a complex parameter is needed to take into account
some couplings between the parameters that would not be found otherwise. Moreover the
shift of κ is required because the curvature of the (θ, φ) section should be normalized to
κ = ±1 but the coupling of the NUT charge with the cosmological constant modifies the
curvature: the new shift is necessary to balance this effect and to normalize the (θ, φ)
curvature to κ0 = ±1 in the new metric. The NUT charge in the Plebański–Demiański
solution [1, 2] is
4Λ 2
`=n 1− n (3.19)
3
so the natural complex combination is m + i` and not m + iκn from this point of view,
and similarly for the curvature [74, sec. 5.3] (such relations appear when taking limit of the
Plebański–Demiański solution to recover subcases).
Finally we conclude this section with two remarks to quote different contexts where the
above expression appear naturally :
• Embedding Einstein–Maxwell into N = 2 supergravity with a negative cosmological
constant Λ = −3g 2 , the solution is BPS if [73]
1
κ0 = −1, n=± , (3.20)
2g
in which case κ0 = κ.
• The Euclidean NUT solution is obtained from the Wick rotation
t = −iτ, n = iν. (3.21)
The condition for regularity is [75, 76]
4Λ 2
m=m −ν κ+
0
ν = 0. (3.22)
3
18
3.3 Complex scalar fields
For a complex scalar field, or any pair of real fields that can be naturally gathered as a
complex field, one should treat the full field as a single entity instead of looking at the
real and imaginary parts independently. In particular one should not impose any reality
condition. A typical case of such system is the axion–dilaton pair
In order to demonstrate this principle consider the seed (for a complete example see
section 6.4)
µ
τ =1+ (3.24)
r
where only the dilaton is non-zero. Then the transformation (2.15) gives
µ µ µr0 µa cos θ
τ0 = 1 + =1+ 0 =1+ 2 +i . (3.25)
r r − ia cos θ ρ ρ2
The transformation generates an imaginary part which cannot be obtained if Im τ is treated
separately: the algorithm does not change fields that vanish except if they are components
of a larger field. Note that both τ and τ 0 are harmonic functions.
4 Complete algorithm
In this section we gather all the facts on the Janis–Newman algorithm and we explain how
to apply it to a general setting. We write the formulas corresponding to the most general
configurations that can be obtained. We insist again on the fact that all these results can
also be derived from the tetrad formalism.
Z I = q I + ipI . (4.1)
where (
sin θ κ = 1 (S 2 ),
dΩ = dθ + H(θ) dφ ,
2 2 2 2
H(θ) = (4.3)
sinh θ κ = −1 (H 2 ).
Note that only two functions in the metric are relevant since the last one can be fixed through
a diffeomorphism. All the real functions are denoted collectively by
fi = {ft , fr , fΩ , f I , q u }. (4.4)
19
The transformation to null coordinates is
s
fr
dt = du − dr (4.5)
ft
and yields
A = f du + p H dφ
I I I 0
(4.6b)
i dθ = H(θ) dφ (4.9)
The ansatz (4.9) is a direct consequence of the fact that the 2-dimensional slice (θ, φ) is q
given
by dΩ2 = dθ2 + H(θ)2 dφ2 , such that the function in the RHS of (4.9) corresponds to Ω
gφφ
(where g is the static metric), as can be seen by doing the computation with i dθ = H(θ)dφ
and identifying H = H at the end.
The most general known transformation is
H(θ/2)
F (θ) = n − a H 0 (θ) + c 1 + H 0 (θ) ln 0 , (4.11a)
H (θ/2)
H(θ/2)
G(θ) = κa H 0 (θ) − 2κn ln H(θ) − κc H 0 (θ) ln 0 , (4.11b)
H (θ/2)
m = m0 + iκn, (4.11c)
4Λ
κ = κ0 − n2 , (4.11d)
3
where a, c 6= 0 only if Λ = 0 (see section 5 for the derivation). The mass that is transformed
is the physical mass: even if it written in terms of other parameters one should identify it
and transform it.
The parameters a and n correspond respectively to the angular momentum and to the
NUT charge. On the other hand the constant c did not receive any clear interpretation (see
for example [7, 18, 77, sec. 5.3]). It can be noted that the solution is of type II in Petrov
classification (and thus the JN algorithm can change the Petrov type) and it corresponds to
a wire singularity on the rotation axis. Moreover the BL transformation is not well-defined.
20
4.2.2 Function transformation
All the real functions fi = fi (r) must be modified to be kept real once r ∈ C
f˜i = f˜i (r, r̄) = f˜i r0 , F (θ) ∈ R. (4.12)
The last equality means that f˜i can depend on θ only through Im r = F (θ). The condition
that one recovers the seed for r̄ = r = r0 is
f˜i (r0 , 0) = fi (r0 ). (4.13)
If all magnetic charges are vanishing or in terms without electromagnetic charges the
rules for finding the f˜i are
1
r −→ (r + r̄) = Re r, (4.14a)
2
1 1 1 1 Re r
−→ + = 2 , (4.14b)
r 2 r r̄ |r|
2
r2 −→ |r| . (4.14c)
Up to quadratic powers of r and r−1 these rules determine almost uniquely the result. This
is not anymore the case when the configurations involve higher power. These can be dealt
with by splitting it in lower powers: generically one should try to factorize the expression
into at most quadratic pieces. Some examples of this with natural guesses are
b
r − b = (r + b)(r − b),
4 2 2 2
r +b=r r + 2 .
4 2 2
(4.15)
r
Moreover the same power of r can be transformed differently, for example
1 1 1
−→ n−2 2 . (4.16)
rn r |r|
Denoting by Q(r) and P (r) collectively all functions that multiply q I and pI respectively,
all such terms should be rewritten as
q I Q(r), pI P (r) = Re Z I Q(r) , Im Z I P (r) (4.17)
before performing the transformation (4.8). Note that in this case one does not use the rules
(4.14).
Finally the transformed complex scalars are obtained by simply plugging (4.8)
τ 0i (r0 , θ) = τ i r + iF (θ) . (4.18)
A = f˜ (du + G H dφ) + p H dφ
I I 0 0 I 0
(4.19b)
where (one should not confuse the primes to indicate derivatives from the primes on the
coordinates)
s
f˜r 0 f˜r 02
s
f˜r
ω=G + 0
F , σ =1+
2
F , α= , β = f˜r F 0 H. (4.20)
f˜t f˜Ω f˜t
11 We stress that at this stage these formula do not satisfy Einstein equations, they are just proxies to
21
4.2.4 Boyer–Lindquist coordinates
The Boyer–Lindquist transformation
f˜Ω 02 ˜
ds2 = −f˜t (dt0 + ωH dφ0 )2 + dr + fΩ dθ2 + σ 2 H 2 dφ02 , (4.25a)
∆
f˜
!
AI = f˜I dt0 − pΩ dr0 + G0 H dφ0 + pI H 0 dφ0 (4.25b)
˜ ˜
∆ ft fr
ft = f −1 , fr = f, fΩ = r2 f. (4.28)
22
Constant F The expressions simplify greatly if F 0 = 0 (for example when Λ 6= 0). First
all functions depend only on r since F (θ) = cst
is always well-defined. For the same reason it is always possible to perform a gauge trans-
formation. Finally the metric and gauge fields (4.25) becomes
23
5.1 Setting up the ansatz
We first recall the action and equations of motion before describing the ansatz for the metric
and gauge fields. We refer to section 4 for the general formulas from which the expressions
in this section are derived.
4 √ 1 1 2
Z
S = d x −g (R − 2Λ) − F , (5.1)
2κ 2 4
where κ 2 = 8πG is the Einstein coupling constant, gµν is the metric with Ricci scalar R and
F = dA is the field strength of the Maxwell field Aµ . In the rest of this section we will set
κ = 1. The corresponding equations of motion (respectively Einstein and Maxwell) are
ft = f, fr = f −1 , fΩ = r2 . (5.4)
24
where
F 02
ρ2 = r 2 + F 2 , ω = G0 + f˜−1 F 0 , σ2 = 1 + , α = f˜−1 , β = f˜−1 F 0 H. (5.10)
f˜ρ2
with functions
ρ2 − F 0 G0 F0
g(r) = , h(r) = , ∆ = f˜ρ2 σ 2 (5.12)
∆ H∆
leads to (omitting the primes on the coordinates)
ρ2 2
ds2 = −f˜t (dt + ωH dφ)2 + dr + ρ2 dθ2 + σ 2 H 2 dφ2 , (5.13a)
∆
2
˜ ρ
A = fA dt − dr + G H dφ .
0
(5.13b)
∆
q2
− κ + r2 Λ + f + rf 0 = 0 (5.16)
r2
whose solution reads
2m q 2 Λ
f (r) = κ − + 2 − r2 , (5.17)
r r 3
m being a constant of integration that is identified to the mass.
We stress that we are just looking for solutions of Einstein equations and we are not
concerned with regularity (in particular it is well-known that only κ = 1 is well-defined for
Λ = 0).
The solution we will find in the next section should reduce to this one upon setting
F, G = 0.
25
5.3.1 Simplifying the equations
The components (rr) and (rθ) give respectively the equation
H0 0
G00 + G = ±2F, (5.18a)
H
0
H 0
F 0 G00 + G = 2F F 0 . (5.18b)
H
If F 0 = 0 then F is an arbitrary constant and the sign of the first equation can be absorbed
into its definition.12 On the other hand if F 0 6= 0 one can simplify by the latter in the
second equation and this fixes the sign of the first equation. Then in both cases the relevant
equation reduces to
H0 0
G00 + G = 2F, (5.19)
H
which depends only on θ and allows to solve for G in terms of F .
Integrating the r-component of the Maxwell equation gives
qr r2 − F 2
f˜A = +α 2 . (5.20)
r2 +F 2 r + F2
The θ-equation reads
αF0 = 0 (5.21)
which implies α = 0 if F 0 6= 0. The φ- and t-equations follow from these two equations. As
seen above, α can be removed in the static limit F → 0 and in the rest of this section we
consider only the case13
α = 0. (5.22)
˜
The (tr) equation contains only r-derivatives of f and can be integrated, giving 14
2mr − q 2 + 2F (κ F + K) Λ 2 4Λ 2 8Λ F 4
f˜ = κ − − (r + F 2
) − F + (5.23)
r2 + F 2 3 3 3 r2 + F 2
where again m is a constant of integration interpreted as the mass and the function K is
defined by
H0 0
2K = F 00 + F . (5.24)
H
This implies the equations (rφ) and (θθ).
As explained below (4.12) the θ-dependence should be contain in F (θ) only. The second
term of the function f˜ contains some lonely θ from the H(θ) in the function K: this means
that they should be compensated by the F , and we therefore ask that the sum κF + K be
constant15
κ F 0 + K 0 = 0 =⇒ κ F + K = κn. (5.25)
The parameter n is interpreted as the NUT charge.
The components (tθ) and (θφ) give the same equation
Λ F 0 = 0. (5.26)
Finally one can check that the last three equations (tt), (tφ) and (φφ) are satisfied.
12 In particular all expressions are quadratic in F , but only linear in F 0 .
13 We relax this assumption in section 5.4.2.
14 In [7] the last term of f˜ is missing as pointed out in [21].
15 In section 5.4.1 we relax this last assumption by allowing non-constant κF + K. In this context the
equations and the function f˜ are modified and this provides an explanation for the Demiański’s error in f˜
in [7].
26
5.3.2 Summary of the equations
The equations to be solved are
H0 0
2F = G00 + G, (5.27a)
H
κ n = κ F + K, (5.27b)
0 = ΛF 0
(5.27c)
2mr − q 2 + 2F (κ F + K) Λ 2 4Λ 2 8Λ F 4
f˜ = κ − − (r + F 2
) − F + . (5.27d)
r2 + F 2 3 3 3 r2 + F 2
We also defined
H0 0
2K = F 00 + F . (5.27e)
H
As explained in the introduction the second step will be to explain (5.27d) in terms of new
rules for the algorithm: they have been found in [59] and this was the topic of section 4.2.
In the next subsections we solve explicitly the equations (5.27) in both cases Λ 6= 0 and
Λ = 0.
H(θ/2)
G(θ) = c1 − 2κ n ln H(θ) + c2 ln (5.29)
H 0 (θ/2)
where c1 and c2 are two constants of integration. Since only G0 appears in the metric we
can set c1 = 0. On the other hand the constant c2 can be removed by the transformation
du = du0 − c2 dφ (5.30)
Λ 4
∆ = κr2 − 2mr + q 2 + Λn4 − r − n2 (κ + 2Λr2 ). (5.34)
3
27
As noted by Demiański the only parameters that appear are the mass and the NUT
charge, and it is not possible to add angular momentum for non-vanishing cosmological
constant.16 As a consequence the JN algorithm cannot provide a derivation of the (a)dS–
Kerr–Newman solution.
mass and the NUT charge appear when Λ 6= 0. We would like to thank D. Klemm for this remark.
28
5.4.1 Metric function F -dependence
In section 5.3.1 we obtained the equation (5.27b)
H0 0
κ F + K = κ n, 2K = F 00 + F (5.41)
H
by requiring that the function (5.27d)
2mr − q 2 + 2F (κ F + K) Λ 2 4Λ 2 8Λ F 4
f˜ = κ − − (r + F 2 ) − F + (5.42)
r +F
2 2 3 3 3 r2 + F 2
depends on θ only through F (θ). A more general assumption would be that κF + K is some
function χ = χ(F )
κ F + K = κ χ(F ). (5.43)
First if F 0 = 0 then K = 0 and the definition of K implies
χ = F = n. (5.44)
4Λ F 2 F 0 = F 0 ∂F χ. (5.45)
If Λ = 0 we find that
∂F χ = 0 =⇒ χ = n (5.46)
which reduces to the case studied in section 5.3.1, while if F 0 = 0 this equation does not
provide anything.
On the other hand if F 0 6= 0 and Λ 6= 0 then the previous equation becomes
∂F χ = 4ΛF 2 (5.47)
2mr − q 2 + 2κ n F Λ 4Λ 2
f˜ = κ − − (r2 + F 2 ) − F . (5.50)
r2 + F 2 3 3
One can recognize the function given by Demiański [7] and may explain his error.
29
5.4.2 Gauge field integration constant
In section 5.3.1 we obtained a second integration constant α in the expression of the gauge
field
qr r2 − F 2
f˜A = 2 + α . (5.51)
r + F2 r2 + F 2
One of the Maxwell equation gives α = 0 if F 0 6= 0, but otherwise no equation fixes its value.
For this reason we focus on the case F 0 = 0 or equivalently Λ 6= 0 through equation (5.27c).
In this case the function f˜ is modified to
2mr − q 2 + 2F (κ F + K) + 4α2 F 2 Λ 4Λ 2 8Λ F 4
f˜ = κ − − (r2 + F 2 ) − F + . (5.52)
r +F
2 2 3 3 3 r2 + F 2
Equation (5.27c) is modified but it is still solved by F 0 = 0 and all other equations are left
unchanged (in particular κF + K is still given by the function χ(F ) (5.48)). For χ(F ) = n
the configuration with α 6= 0 provides another solution when Λ 6= 0 but it is not clear how
to get it from a complexification of the function.
6 Examples
In this section we list several examples that can be derived from the JN algorithm described
in section 4. Other examples were described previously: Kerr–Newman in section 2.3.2,
dyonic Kerr–Newman and Yang–Mills Kerr–Newman in section 3.1. For simplicity we will
always consider the case κ = 1 except when Λ 6= 0.
The first two examples are the Kerr–Newmann–NUT solution (already derived by an-
other path in section 5.3.4) and the charged (a)dS–BBMB–NUT solution in conformal grav-
ity. We will also give examples from ungauged N = 2 supergravity coupled to nv = 0, 1, 3
vector multiplets (pure supergravity, T3 model and STU model): this theory is reviewed in
appendix B.
6.1 Kerr–Newman–NUT
The Reissner–Nordström metric and gauge fields are given by
2m q 2
ds2 = −f dt2 + f −1 dr2 + r2 dΩ2 , f =1− + 2, (6.1a)
r r
q
A = fA dt, fA = , (6.1b)
r
m and q being the mass and the electric charge.
The two functions are complexified as
2 Re(mr̄) + q 2 q Re r
f˜ = 1 − 2 , f˜A = 2 . (6.2)
|r| |r|
30
The metric and the gauge fields in BL coordinates are
ρ2 2
ds2 = −f˜ (dt + Ω dφ)2 + dr + ρ2 (dθ2 + σ 2 H 2 dφ2 ), (6.5a)
∆
qr
A = 2 dt − (a sin2 θ + 2n cos θ)dφ + Ar dr (6.5b)
ρ
where
∆
Ω = −2n cos θ − (1 − f˜−1 ) a sin2 θ, σ2 = ,
˜
f ρ2 (6.6)
∆ = r2 − 2mr + a2 + q 2 − n2 .
This corresponds to the Kerr–Newman–NUT solution [73].
One can check that Ar is a function of r only
qr
Ar = − (6.7)
∆
and it can be removed by a gauge transformation.
1 4 √ 1
Z
S= d x −g R − 2Λ − Rφ − (∂φ) − 2αφ − F ,
2 2 4 2
(6.8)
2 6
where α is a coupling constant, and we have set 8πG = 1.
For F, α, Λ = 0, the Bocharova–Bronnikov–Melnikov–Bekenstein (BBMB) solution [79,
80] is static and spherically symmetric – it can be seen as the equivalent of the Schwarzschild
black hole in conformal gravity.
The general static charged solution with cosmological constant and quartic coupling
reads
31
where the function f˜ is
Λ 4Λ 2 (r − m)2
f˜ = − (r2 + n2 ) + κ − n . (6.13)
3 3 r 2 + n2
Note that the term (r − m) is invariant. Similarly one obtains the scalar field
r √
Λ m2 + n2
φ= − (6.14)
6α r − m
where the m in the numerator as been complexified as |m|. Finally it is trivial to find the
gauge field
q
A= 2 dt − 2n cos θ dφ (6.15)
r +n 2
4Λ 2 Λ
q2 = κ − n (m2 + n2 ) 1 + . (6.16)
3 36α
An interesting point is that the radial coordinate is redefined in [67] when obtaining the
stationary solution from the static one.
Note that the BBMB solution and its NUT version are obtained from the limit
Λ
Λ, α −→ 0, with − −→ 1, (6.17)
36α
which also implies q = 0 from the constraint (6.10). Since no other modifications are needed,
the derivation from the JN algorithm also holds in this case.
32
{H Λ , HΛ }
The object Ωi is the connection of the line bundle corresponding to the fibration of time over
the spatial manifold (its curl is related to the Kähler connection). Its only non-vanishing
component is Ωφ ≡ Ω = ωH.
Starting from the metric (6.18) in spherical coordinates with Ω = 0, one can use the JN
algorithm of section 4 with
ft = f −1 , fr = f, fΩ = r2 f, (6.21)
Then one needs only to find the complexification of f and to check that it gives the
correct ω, as would be found from the equations (6.20). However it appears that one cannot
complexify directly f since it should be viewed as a composite object made of complex
functions. Therefore one needs to complexify first the harmonic functions HΛ and H Λ (or
equivalently X Λ ), and then to reconstruct the other quantities. Nonetheless, equations
(6.20) ensure that finding the correct harmonic functions gives a solution, thus it is not
necessary to check these equations for all the other quantities.
In the next subsections we provide two examples,19 one for pure supergravity as an
appetizer, and then one with nv = 3 multiplets (STU model).
f = e−K = X 0 X̄ 0 . (6.25)
by the JNA. No confusion is possible since the index position will always indicate which function we are
using.
19 They correspond to singular solutions, but we are not concerned with regularity here.
33
Performing the JN transformation for the angular momentum gives
m(r + ia cos θ)
X̃ 0 = 1 + . (6.27)
ρ2
This corresponds to the second solution of which is stationary with
m(2r + m)
ω= a sin2 θ. (6.28)
ρ2
Alternatively one can use the JN algorithm to add a NUT charge. In this case using the
rule
1
r −→ (r + r̄) = Re r = r0 (6.29)
2
must be use for transforming f and r2 (in front of dΩ), leading to
m + in
X0 = 1 + . (6.30)
r
Note that it gives
m 2 n 2
f˜ = 1 +
+ 2. (6.31)
r r
It is slightly puzzling that the above rule should be used instead of the two others in (4.14).
One possible explanation is the following: in the seed solution shift the radial coordinate
such that r = R − m and apply the JN transformation in this coordinate system. It is clear
that every function of r is left unchanged while the tensor structure transforms identically
since dr = dR. After the transformation one can undo the coordinate transformation. As
we mentioned earlier the algorithm is very sensible to the coordinate system and to the
parametrization (but it is still not clear why the R-coordinate is the natural one). This kind
of difficulty will reappear in the SWIP solution (section 6.5).
X 1X 2X 3
F =− . (6.32)
X0
The expressions for the Kähler potential and the scalar fields in terms of the harmonic
functions are complicated and will not be needed (see [65, sec. 3] for the expressions).
Various choices for the functions will give different solutions.
A class of static black hole-like solutions are given by the harmonic functions [65, sec. 4.4]
q0 pi
H0 = h0 + , H i = hi + , H 0 = Hi = 0. (6.33)
r r
These solutions carry three magnetic pi and one electric q0 charges.
Let’s form the complex harmonic functions
H0 = H0 + i H 0 , Hi = H i + i Hi . (6.34)
q0 (r + ia cos θ) pi (r + ia cos θ)
H0 = h0 + , Hi = hi + , (6.35)
ρ2 ρ2
34
for which the various harmonic functions read explicitly
q0 r pi r q0 a cos θ pi a cos θ
H0 = h0 + , H i = hi + , H0 = , Hi = . (6.36)
ρ2 ρ2 ρ2 ρ2
This set of functions corresponds to the stationary solution of [65, sec. 4.4] where the mag-
netic and electric dipole momenta are not independent parameters but obtained from the
magnetic and electric charges instead.
F = −i X 0 X 1 , (6.37)
The dilaton and the axion corresponds to the complex scalar field
τ = e−2φ + i σ. (6.38)
Sen derived the rotating black hole for this theory using the fact that it can be embedded
in heterotic string theory [63].
The static metric, gauge field and the complex field read respectively
f1
ds2 = − dt2 + f2 f1−1 dr2 + r2 dΩ2 , (6.39a)
f2
fA
A= dt, (6.39b)
f2
τ = e−2φ = f2 (6.39c)
where
r1 r2 q
f1 = 1 − , f2 = 1 + , fA = . (6.40)
r r r
The radii r1 and r2 are related to the mass m and the charge q by
q2
r1 + r2 = 2m, r2 = . (6.41)
m
Applying the Janis–Newman algorithm with rotation, the two functions f1 and f2 are
complexified as
r1 r r2 r
f˜1 = 1 − 2 , f˜2 = 1 + 2 . (6.42)
ρ ρ
The final metric in BL coordinates is given by
f˜1 f˜2
2
ρ dr ∆
2 2
ds2 = − dt − a 1 − sin2 θ dφ + f˜2 + ρ2 dθ2 + sin2 θ dφ2 (6.43)
f˜2 f˜1 ∆ f˜1
for which the BL functions are
ˆ
∆ a
g(r) = , h(r) = (6.44)
∆ ∆
20 This model can be obtained from the STU model by setting the sections pairwise equal X 2 = X 0 and
35
with
∆ = f˜1 ρ2 + a2 sin2 θ, ˆ = f˜2 ρ2 + a2 sin2 θ.
∆ (6.45)
Once fA has been complexified as
qr
f˜A = 2 (6.46)
ρ
the transformation of the gauge field is straightforward
f˜A qr
A= (dt − a sin2 θ dφ) − dr. (6.47)
˜
f2 ∆
The Ar depending solely on r can again be removed thanks to a gauge transformation.
Finally the scalar field is complex and is transformed as
r2 r̄
τ =1+ . (6.48)
ρ2
The explicit values for the dilaton and axion are then
r2 a cos θ
e−2φ = f˜2 , σ= . (6.49)
ρ2
This reproduces Sen’s solution and it completes the computation from [26] which could
not derive the gauge field nor the axion. It is interesting to note that for another value of
the dilaton coupling we cannot use the transformation [51, 53].21
τ = σ + ie−2φ . (6.51)
In order to avoid redundancy we first provide the general metric with a, n 6= 0, and we
explain how to find it from the restricted case a = n = 0. The stationary Israel–Wilson–
Perjés (SWIP) solutions correspond to
36
The spatial 3-dimensional metric dΣ2 reads
ρ2 − r02
dΣ2 = hij dxi dxj = dr2 + (ρ2 − r02 )dθ2 + (r2 + a2 − r02 ) sin2 θ dφ2 . (6.54)
r2 + a2 − r02
Finally, r0 corresponds to
2 2
X 2
r02 = |M| + |Υ| − Γi (6.55)
i
37
6.6 Gauged N = 2 non-extremal solution
The simplest deformation of N = 2 supergravity with nv vector multiplets consists in the
so-called Fayet–Iliopoulos (FI) gauging. It amounts to gauging (nv + 1) times the diagonal
U(1) group of the SU(2) part of the R-symmetry group (automorphism of the supersym-
metry algebra). The potential can be entirely written in terms of the quantities defined in
appendix B and of the (nv + 1) coupling constants gI , where I = 0, . . . , nv .
We consider the model with prepotential (see also section 6.4)
F = −i X 0 X 1 . (6.63)
Q0
δ = −∆ . (6.66b)
P0
The independent parameters are given by QI (electric charges), P I (magnetic charges), gΛ
(FI gaugings), ∆ (scalar charge) and Λ = −3/`2 (the cosmological constant).
In order to perform the complexification the functions are first rewritten as
2
2 Re(mr̄) − 2`2 I gI Z I
P
fΩ
ft = κ − + 2, (6.67a)
fΩ `
2
2 2 ∆2 Z 1
fΩ = |r| − ∆ − δ = |r| −
2 2
, (6.67b)
Im(Z 1 )2
Re(QI r̄) Im Z 1 − ∆ Im(Z I Z 1 )
fI = , (6.67c)
Im Z 1 fΩ
g0 r̄ + ∆ − iδ
τ= . (6.67d)
g1 r̄ − ∆ + iδ
22 The original derivation is due to D. Klemm and M. Rabbiosi and has not been published. I am grateful
38
Applying the transformations (4.8) with (4.11a) gives (omitting the primes)
f˜Ω
2
4n2 2mr + 2 κ + 4n2 /`2 n2 − 2`2 I gI Z I
P
˜
ft = κ + 2 − + 2, (6.68a)
` f˜Ω `
f˜Ω = r + n − ∆ − δ ,
2 2 2 2
(6.68b)
(Q r + P n) Im Z − ∆ Im(Z Z )
I I 1 I 1
f˜I = , (6.68c)
Im Z 1 f˜Ω
g0 r + ∆ − i(δ + n)
τ̃ = . (6.68d)
g1 r − ∆ + i(δ − n)
The last step is to simplify these expressions
f˜Ω
2
4n2 2mr + 2κn2 + 8n4 /`2 − 2`2 gI Z I
P
f˜t = κ + 2 − I
+ , (6.69a)
` f˜Ω `2
f˜Ω = r + n − ∆2 − δ 2 ,
2 2
(6.69b)
Q (r − ∆) + P (n − δ)
I I
f˜I = , (6.69c)
f˜Ω
g0 r + ∆ − i(δ + n)
τ̃ = . (6.69d)
g1 r − ∆ + i(δ − n)
It is straightforward to check that the form of the metric and gauge fields are correctly repro-
duced by the algorithm given in section 4 for the tensor structure. In total this reproduces
the eq. (4.22) and formulas below in [70] with j = 0.
An important thing that we learn here is that the mass parameter needs to be transformed
as if it was not composed of other parameters.
39
A major application of our work would be to find the charged solution with two angular
momenta of the 5d Einstein–Maxwell gravity. This problem is highly non-trivial and there
is few chances that this technique would work directly [85], but one can imagine that a
generalization of Demiański’s approach [7] (see section 5) could lead to new interesting
solutions in five dimensions. An intermediate step is represented by the CCLP metric [87]
which is a solution of the Einstein–Maxwell theory with a Chern–Simons term, but it cannot
be derived from the JN algorithm and we give some intuition about this fact in the last
subsection.
Finally one could seek for an extension of the algorithm to the derivation of black
rings [84, 88]. Similarly it may be possible that such techniques could be used in d = 4 to
derive multicentre solutions (for instance one could imagine adding rotation to both centres
successively, changing coordinate system in-between to place the origin of the coordinates
at each centre).
where dΩ23 is the metric on S 3 , which can be expressed in Hopf coordinates (see ap-
pendix A.3.2)
dΩ23 = dθ2 + sin2 θ dφ2 + cos2 θ dψ 2 , (7.2)
and the function f (r) is given by
m
f (r) = 1 − . (7.3)
r2
An important feature of the JN algorithm is the fact that a given set of transformations
in the (r, φ)-plane generates rotation in the latter. Generating several angular momenta in
different 2-planes would then require successive applications of the JN algorithm on different
hypersurfaces. In order to do so, one has to identify what are the 2-planes which will be
submitted to the algorithm. In five dimensions, the two different planes that can be made
rotating are the planes (r, φ) and (r, ψ). We claim that it is necessary to dissociate the radii
of these 2-planes in order to apply separately the JN algorithm on each plane and hence to
generate two distinct angular momenta. In order to dissociate the parts of the metric that
correspond to the rotating and non-rotating 2-planes, one can protect the function r2 to
be transformed under complex transformations in the part of the metric defining the plane
which will stay static. We thus introduce the function
R(r) = r (7.4)
ds2 = −du (du + 2dr) + (1 − f ) du2 + r2 (dθ2 + sin2 θ dφ2 ) + R2 cos2 θ dψ 2 . (7.5)
u = u0 + ia cos χ1 , r = r0 − ia cos χ1 ,
i dχ1 = sin χ1 dφ, with χ1 = θ, (7.6)
du = du − a sin θ dφ,
0 2
dr = dr + a sin θ dφ,
0 2
40
and f is replaced by f˜{1} = f˜{1} (r, θ). Indeed one needs to keep track of the order of the
transformation, since the function f will be complexified twice consecutively. On the other
hand R(r) = Re(r) is transformed23 into R0 = r0 and one finds (omitting the primes)
ds2u,r = (1 − f˜{1} ) (du − a sin2 θ dφ)2 − du(du + 2dr) + 2a sin2 θ drdφ + a2 sin4 θ dφ2 ,
(7.9a)
ds2θ,φ = (r + a cos θ)dθ + r + a (1 − sin θ) sin θ dφ .
2 2 2 2 2 2 2 2 2
(7.9b)
In addition to the terms present in (7.5) one obtains new components corresponding to
the rotation of the first plane (r, φ). Since the structure is very similar one can perform a
transformation24 in the second plane (r, ψ)
u = u0 + ib sin χ2 , r = r0 − ib sin χ2 ,
i dχ2 = − cos χ2 dψ, with χ2 = θ, (7.10)
du = du − b cos θ dψ,
0 2
dr = dr + b cos θ dψ,
0 2
where we introduced once again the function R(r) = Re(r) to protect the geometry of the
first plane to be transformed under complex transformations.
The final result (using again R = r0 and omitting the primes) becomes
where
ρ2 = r2 + a2 cos2 θ + b2 sin2 θ. (7.13)
Furthermore, the function f˜{1} has been complexified as
m m m
f˜{1,2} = 1 − =1− = 1 − 2. (7.14)
2
|r| + a2 cos2 θ r02 + a2 cos2 θ + b2 sin2 θ ρ
The metric can then be transformed into the Boyer–Lindquist (BL) using
direction cosines, see section 8.2.3. Otherwise this term can be guessed by looking at Myers–Perry non-
diagonal terms.
41
Defining the parameters25
Π = (r2 + a2 )(r2 + b2 ), ∆ = r4 + r2 (a2 + b2 − m) + a2 b2 , (7.16)
the functions can be written
Π Π a Π b
g(r) = , hφ (r) = , hψ (r) = . (7.17)
∆ ∆ r2 + a2 ∆ r 2 + b2
Finally one gets
r2 ρ2 2
ds2 = −dt2 + 1 − f˜{1,2} (dt − a sin2 θ dφ − b cos2 θ dψ)2 + dr
∆ (7.18)
+ ρ2 dθ2 + (r2 + a2 ) sin2 θ dφ2 + (r2 + b2 ) cos2 θ dψ 2 .
One recovers here the five dimensional Myers–Perry black hole with two angular momenta [68].
42
where dΩ23 is the metric of the 3-sphere written in (7.2). The function H is harmonic
m
H(r) = 1 + , (7.21)
r2
and the electromagnetic field reads
√
3 m
A= dt = (H − 1) dt. (7.22)
2λ r2
In the next subsections we apply successively the transformations (7.6) and (7.10) with
a = b in the case λ = 1.
dt = du + H 3/2 dr (7.23)
gives
For transforming the above metric one should follow the recipe of the previous section:
the transformations (7.6)
and (7.10)
u = u0 + ia sin θ, du = du0 − a cos2 θ dψ (7.26)
are performed one after another, transforming each time only the terms that pertain to the
corresponding rotation plane.30 In order to preserve the isotropic form of the metric the
function H is complexified everywhere (even when it multiplies terms that belong to the
other plane).
Since the procedure is exactly similar to the Myers–Perry case we give only the final
result in (u, r) coordinates
2
ds2 = − H̃ −2 du − a(1 − H̃ 3/2 )(sin2 θ dφ + cos2 θ dψ)
− 2H̃ −1/2 du − a(1 − H̃ 3/2 ) (sin2 θ dφ + cos2 θ dψ) dr
(7.27)
+ 2aH̃ (sin2 θ dφ + cos2 θ dψ) dr − 2a2 H̃ cos2 θ sin2 θ dφdψ
+ H̃ (r2 + a2 )(dθ2 + sin2 θ dφ2 + cos2 θ dψ 2 ) + a2 (sin2 θ dφ + cos2 θ dψ)2 .
43
is ill-defined because the functions depend on θ. The way out is to take the extremal limit
alluded above.
Following the prescription of [69, 89] and taking the extremal limit
m
a, m −→ 0, imposing = cst, (7.30)
a2
one gets at leading order
m 3 ma
H̃(r) = 1 + = H(r), a (1 − H̃ 3/2 ) = − (7.31)
r2 2 r2
which translate into the metric
2
3 ma
ds2 = −H −2 du + (sin 2
θ dφ + cos 2
θ dψ)
2 r2
3 ma (7.32)
− 2H −1/2 du + (sin 2
θ dφ + cos 2
θ dψ) dr
2 r2
+ H r2 (dθ2 + sin2 θ dφ2 + cos2 θ dψ 2 ).
One can recognize the BMPV solution [69, p. 4, 89, p. 16]. The fact that this solution has
only one rotation parameter can be seen more easily in Euler angle coordinates [89, sec. 3,
95, sec. 2] or by looking at the conserved charges in the φ- and ψ-planes [69, sec. 3].
44
which is again the result presented in [69, p. 5].
Despite the fact that the seed metric (7.20) together with the gauge field (7.22) solves
the equations of motion for any value of λ, the resulting rotating metric solves the equations
only for λ = 1 (see [89, sec. 7] for a discussion). An explanation in this reduction can be
found in the limit (7.30) that was needed for transforming the metric to Boyer–Lindquist
coordinates and which gives a supersymmetric black hole – which necessarily has λ = 1.
a2 + (r2 + a2 )H̃(r) a
g(r) = , hφ (r) = hψ (r) = (7.42)
r2 + 2a2 r2 + 2a2
which do not depend on θ. They lead to the metric
2
ds2 = − H̃ −2 dt − a(1 − H̃ 3/2 )(sin2 θ dφ + cos2 θ dψ)
dr2
+ H̃ (r + a )
2 2
+ dθ + sin θ dφ + cos θ dψ
2 2 2 2 2
(7.43)
r2 + 2a2
+ a (sin θ dφ + cos θ dψ) .
2 2 2 2
At this point it is straightforward to check that this solution does not satisfy Einstein
equations and we need to take the extremal limit (7.30)
m
a, m −→ 0, imposing = cst (7.44)
a2
31 In opposition to our initial recipe, but this is done in a controlled way.
45
in order to get the BMPV solution (7.34)
2
3 ma
ds2 = − H̃ −2 dt + (sin 2
θ dφ + cos 2
θ dψ)
2 r2 (7.45)
+ H̃ dr2 + r2 dθ2 + sin2 θ dφ2 + cos2 θ dψ 2 .
It is surprising that the BL transformation is simpler in this case. Another point that is
worth stressing is that we did not need to take the extremal limit at an intermediate stage,
whereas in section 7.2 we had to in order to get a well-defined BL transformation.
46
the metric can be done along the same line, the – major – obstacle comes from the function
f that cannot be transformed as expected. Finding the correct complexification seems very
challenging and it may be necessary to use a different complex coordinate transformation in
order to perform a completely general transformation in any dimension. It might be possible
to gain insight into this problem by computing the transformation within the framework of
the tetrad formalism. One may think that a possible solution would be to replace complex
numbers by quaternions, assigning one angular momentum to each complex direction but it
is straightforward to check that this approach is not working.
The key element to perform the algorithm on the metric is to parametrize the metric
on the sphere by direction cosines since these coordinates are totally symmetric under per-
mutation of angular momenta (at the opposite of the spherical coordinates). We are able to
derive the general form of a rotating metric with the maximal number of angular momenta
it can have in d dimensions, but we are not able to apply this result to any specific example
for d ≥ 6, except if all momenta but one are vanishing. Nonetheless this provides a unified
view of the JN algorithm in any d ≥ 3. We conclude this section by few examples, including
the singly-rotating Myers–Perry solution in any dimension and the rotating BTZ black hole.
It would be very desirable to derive the general d-dimensional Myers–Perry solution [68],
or at least to understand why only the metric can be found, and not the function inside.
The metric looks like a 2-dimensional space (t, r) with a certain number of additional
2-spheres (µi , φi ) which are independent from one another. Then we can consider only the
piece (u, r, µi , φi ) (for fixed i) which will transform like a 4-dimensional spacetime, while the
other part of the metric (µj , φj ) for all j 6= i will be unchanged. After the first transformation
we can move to another 2-sphere. We can thus imagine to put in rotation only one of these
spheres. Then we will apply again and again the algorithm until all the spheres have angular
momentum: the whole complexification will thus be a n-steps process. Moreover if these 2-
spheres are taken to be independent this implies that we should not complexify the functions
that are not associated with the plane we are putting in rotation.
To match these demands the metric is rewritten as
X
ds2 = (1 − f ) du2 − du (du + 2dri1 ) + ri21 (dµ2i1 + µ2i1 dφ2i1 ) + ri21 dµ2i + R2 µ2i dφ2i . (8.4)
i6=i1
47
where we introduced the following two functions of r
This allows to choose different complexifications for the different terms in the metric. It
may be surprising to note that the factors in front of dµ2i have been chosen to be ri21 and
not R2 , but the reason is that the µi are all linked by the constraint
X
µ2i = 1 (8.6)
i
and the transformation of one i1 -th 2-sphere will change the corresponding µi1 , but also all
the others, as it is clear from the formula (A.14) with all the ai vanishing but one (this can
also be observed in 5d where both µi are gathered into θ).
dri1 = dri01 + ai1 µ2i1 dφi1 , du = du0 − ai1 µ2i1 dφi1 . (8.7c)
It is easy to check that this transformation reproduces the one given in four and five dimen-
sions. The complexified version of f is written as f˜{i1 } : we need to keep track of the order
in which we gave angular momentum since the function f˜ will be transformed at each step.
We consider separately the transformation of the (u, r) and {µi , φi } parts. Inserting the
transformations (8.7) in (8.3) results in
2
ds2u,r = (1 − f˜{i1 } ) du − ai1 µ2i1 dφi1 − du (du + 2dri1 ) + 2ai1 µ2i1 dri1 dφi1 + a2i1 µ4i1 dφ2i1 ,
X
ds2µ,φ = ri21 + a2i1 (dµ2i1 + µ2i1 dφ2i1 ) + ri21 dµ2i + R2 µ2i dφ2i − a2i1 µ4i1 dφ2i1
i6=i1
X
+ a2i1 − µ2i1 dµ2i1 + (1 − µ2i1 ) dµ2i .
i6=i1
The term in the last bracket vanishes as can be seen by using the differential of the
constraint X X
µ2i = 1 =⇒ µi dµi = 0. (8.9)
i i
Since this step is very important and non-trivial we expose the details
2
X X X X
[· · · ] = µ2i1 dµ2i1 − (1 − µ2i1 ) dµ2i = µi dµi − µ2j dµ2i
i6=i1 i6=i1 j6=i1 i6=i1
X X
= µi µj dµi dµj − µ2j dµ2i = µj µi dµj − µj dµi dµi = 0
i,j6=i1 i,j6=i1
48
by antisymmetry.
Setting ri1 = R = r one obtains the metric
2
ds2 = (1 − f˜{i1 } ) du − ai1 µ2i1 dφi1 − du (du + 2dr) + 2ai1 µ2i1 drdφi1
X (8.10)
+ r2 + a2i1 (dµ2i1 + µ2i1 dφ2i1 ) + r2 dµ2i + µ2i dφ2i .
i6=i1
+ ri22 + a2i1 dµ2i1 + R2 + a2i1 µ2i1 dφ2i1 + ri22 (dµ2i2 + µ2i2 dφ2i2 )
(8.11)
X
+ ri2 dµi + R µi dφi .
2 2 2 2 2
i6=i1 ,i2
The steps are exactly the same as before, except that we have some inert terms. The
complexified functions is now f˜{i1 ,i2 } .
Repeating the procedure n times we arrive at
X X
ds2 = − du2 − 2dudr + (r2 + a2i )(dµ2i + µ2i dφ2i ) − 2 ai µ2i drdφi
i i
!2 (8.13)
+ 1 − f˜{i1 ,...,in }
X
du + ai µ2i dφi .
i
One recognizes the general form of the d-dimensional metric with n angular momenta [68].
Let’s quote the metric in Boyer–Lindquist coordinates (omitting the indices on f˜) [68]
!2
˜
X r 2 ρ2 2 X 2
ds = −dt + (1 − f ) dt −
2 2
ai µi dφi +
2
dr + (r + a2i ) dµ2i + µ2i dφ2i (8.14)
i
∆ i
with functions
Π 1 Π ai
g= = , hi = , (8.16)
∆ 1 − F (1 − f˜) ∆ r + a2i
2
and where the various quantities involved are (see appendix A.1.4)
Y X a2 µ2 X µ2
Π= (r2 + a2i ), F =1− i i
2 + a2
= r 2 i
2 + a2
,
i i
r i i
r i (8.17)
r ρ = ΠF,
2 2
∆ = f˜ r2 ρ2 + Π(1 − F ).
49
Before ending this section, we comment the case of even dimensions: the term ε0 r2 dα2
is complexified as ε0 ri21 dα2 , since it contributes to the sum
X
µ2i + α2 = 1. (8.18)
i
This can be seen more clearly by defining µn+1 = α (we can also define φn+1 = 0), in which
case the index i runs from 1 to n + ε, and all the previous computations are still valid.
In this case the JN algorithm is equivalent to a (true) change of coordinates and there is no
intrinsic rotation. The presence of a non-trivial function f then deforms the algorithm.
r = r0 − ia cos θ. (8.22)
ρ2 = r2 + a2 (1 − µ2 ) = r2 + a2 cos2 θ. (8.23)
X (8.25)
+ r2 + a2 (dµ2 + µ2 dφ2 ) + r2 dµ2i + µ2i dφ2i .
i6=1
50
In the mixed coordinate system one has [23, 85]
rd−3 ρ2 2
ds2 = − f˜ dt2 + 2a(1 − f˜) sin2 θ dtdφ + dr + ρ2 dθ2
∆
(8.27)
Σ2
+ sin2 θ dφ2 + r2 cos2 θ2 dΩ2d−4 .
ρ2
where we defined as usual
Σ2
∆ = f˜ρ2 + a2 sin2 θ, = r2 + a2 + agtφ . (8.28)
ρ2
This last expression explains why the transformation is straightforward with one angular
momentum: the transformation is exactly the one for d = 4 and the extraneous dimensions
are just spectators.
We have not been able to generalize this result for several non-vanishing momenta for
d ≥ 6, even for the case with equal momenta .
For µ = sin θ and ν = cos θ one recovers the transformations from sections 7.1 and 7.2.
Let’s denote the denominator by ρ2 and compute
ρ2
= r2 + a2 (1 − µ2 ) + b2 (1 − ν 2 ) = (µ2 + ν 2 )r2 + ν 2 a2 + µ2 b2
r2
µ2 ν2
= µ2 (r2 + b2 ) + ν 2 (r2 + a2 ) = (r2 + b2 )(r2 + a2 ) + .
r2 + a2 r 2 + b2
and thus
r2 ρ2 = ΠF. (8.33)
Plugging this into f˜{1,2} we have [68]
mr2
1 − f˜{1,2} = . (8.34)
ΠF
51
8.2.4 Three dimensions: BTZ black hole
As another application we show how to derive the d = 3 rotating BTZ black hole from its
static version [86]
r2
ds2 = −f dt2 + f −1 dr2 + r2 dφ2 , f (r) = −M + . (8.35)
`2
In three dimensions the metric on S 1 in spherical coordinates is given by
The transformation of f is
ρ2
f˜ = −m + 2 , ρ2 = r2 + a2 (1 − µ2 ). (8.41)
`
The transformation (8.16)
ρ2 (1 − f˜) a
g= , h= , ∆ = r2 + a2 + (f˜ − 1)ρ2 (8.42)
∆ ∆
to Boyer–Lindquist coordinates leads to the metric (8.14)
ρ2 2
ds2 = −dt2 + (1 − f˜)(dt + aµ2 dφ)2 + dr + (r2 + a2 )(dµ2 + µ2 dφ2 ). (8.43)
∆
Finally the constraint µ2 = 1 can be used to remove the µ. In this case one finds
ρ2 = r 2 , ∆ = a2 + f˜r2 (8.44)
t = t0 − aφ (8.47)
52
we get (omitting the prime)
Acknowledgments
I am particularly grateful and indebted to Lucien Heurtier for our collaboration and our
many discussions on this project. I thank also Nick Halmagyi and Dietmar Klemm for
interesting discussions, and I am grateful to the latter and Marco Rabbiosi for allowing me
to reproduce an unpublished example of application. Finally I wish to thank the members of
the Harish–Chandra Research Institute (Allahabad, India) for organizing the set of lectures
that helped me to transform my thesis in the current review.
A Coordinate systems
This appendix is partly based on [68, 99, 104]. We present formulas for any dimension before
summarizing them for 4 and 5 dimensions.
A.1 d-dimensional
Let’s consider d = N + 1 dimensional Minkowski space whose metric is denoted by
In all the following coordinates systems the time direction can separated from the spatial
(positive definite) metric as
where x0 = t.
One defines by n the number of independent 2-planes of rotation
N
n= (A.3)
2
53
such that
d + ε = 2n + 2, N + ε = 2n + 1, ε0 = 1 − ε (A.4)
where (
1 0 d even (or N odd)
ε = (1 − (−1)d ) = (A.5)
2 1 d odd (or N even),
and conversely for ε0 .
A.1.2 Spherical
Introducing a radial coordinate r, the flat space metric can be written as a (N − 1)-sphere
of radius r
dΣ2 = dr2 + r2 dΩ2N −1 . (A.7)
The term dΩ2N −1 corresponds the metric on the unit (N − 1)-sphere S N −1 , which is parame-
trized by (N − 1) angles θi and is defined recursively as
dΩ2N −1 = dθN
2
−1 + sin θN −1 dΩN −2 .
2 2
(A.8)
This surface can be embedded in N -dimensional flat space with coordinates X a con-
strained by
X a X a = 1. (A.9)
X N = α. (A.11)
Each pair parametrizes a 2-sphere of radius µi . The µi are called the direction cosines
and satisfy X
µ2i + ε0 α2 = 1 (A.12)
i
since there is one superfluous coordinate from the embedding. Finally the metric is
X
dΩ2N −1 = dµ2i + µ2i dφ2i + ε0 dα2 . (A.13)
i
The interest of these coordinates is that all rotational directions are symmetric.
32 Note that this is linked to the fact that the little group of massive representation in D dimension is
54
A.1.4 Spheroidal with direction cosines
From the previous system we can define the spheroidal (r̄, µ̄i , φ̄i ) system – adapted when
some of the 2-spheres are deformed to ellipses – by introducing parameters ai such that (for
d odd) X
r2 µ2i = (r̄2 + a2i )µ̄2i , µ̄2i = 1. (A.14)
i
and we defined
X a2 µ̄2 X r̄2 µ̄2
F =1− i i
2+a 2 = i
2 + a2
. (A.17)
i
r̄ i i
r̄ i
Here the ai are just introduced as parameters in the transformation, but in the main
text they are interpreted as "true" rotation parameters, i.e. angular momenta (per unit of
mass) of a black hole. They all appear on the same footing.
Another quantity of interest is
Y
Π= (r̄2 + a2i ). (A.18)
i
θ = θN −1 , φ = θN −2 . (A.19)
A.2 4-dimensional
In this section one considers
d = 4, N = 3, n = 1. (A.23)
55
A.2.1 Cartesian system
dΣ2 = dx2 + dy 2 + dz 2 . (A.24)
A.2.2 Spherical
A.3 5-dimensional
In this section one considers
d = 4, N = 3, n = 1. (A.33)
56
A.3.2 Hopf coordinates
The constraint (A.34) can be solved by
{gµν , A0 } (B.1)
while each of the vector multiplets contains a gauge field and a complex scalar field
{Ai , τ i }, i = 1, . . . , nv . (B.2)
The scalar fields τ i (the conjugate fields (τ i )∗ are denoted by τ̄ ı̄ ) parametrize a special
Kähler manifold with metric gī . This manifold is uniquely determined by an holomorphic
function called the prepotential F . The latter is better defined using the homogeneous (or
projective) coordinates X Λ such that
Xi
τi = . (B.3)
X0
The first derivative of the prepotential with respect to X Λ is denoted by
∂F
FΛ = . (B.4)
∂X Λ
Finally it makes sense to regroup the gauge fields into one single vector
AΛ = (A0 , Ai ). (B.5)
One needs to introduce two more quantities, respectively the Kähler potential and the
Kähler connection
i
K = − ln i(X̄ Λ FΛ − X Λ F̄ Λ ), Aµ = − (∂i K ∂µ τ i − ∂ı̄ K ∂µ τ̄ ı̄ ). (B.6)
2
The Lagrangian for the theory without gauge group is given by
R
L=− + gī (τ, τ̄ ) ∂µ τ i ∂ ν τ̄ ı̄ + IΛΣ (τ, τ̄ ) Fµν
Λ
F Σ µν − RΛΣ (τ, τ̄ ) Fµν
Λ
?F Σ µν (B.7)
2
where R is the Ricci scalar and ?F Λ is the Hodge dual of F Λ . The matrix
N = R + iI (B.8)
can be expressed in terms of F . From this Lagrangian one can introduce the symplectic
dual of F Λ
δL
GΛ = = RΛΣ F Σ − IΛΣ ?F Σ . (B.9)
δF Λ
57
C Technical properties
In this chapter we describe few technical properties of the algorithm. We comment on the
group properties that some of the JN transformations possess [59]. Another useful prop-
erty of Giampieri’s prescription is to allow to chain all coordinate transformation, making
computations easier [57]. Then finally we discuss the fact that not all the rules (4.14) are
independent and several choices of complexification are equivalent [57], contrary to what is
commonly believed.
where
f˜i = f˜i (r, F1 ).
{1} {1}
(C.4)
Performing a second transformation
r = r0 + i F2 , u = u0 + i G2 (C.5)
the previous metric becomes (omitting the primes)
58
where
f˜i = f˜i
{1,2} {1,2}
(r, F1 , F2 ). (C.7)
This function is required to satisfy the following conditions (omitting the primes)
The second condition means that the order of the transformations should not matter because
we want to obtain the same solution given identical seed metric and parameters.
The metric (C.6) is obviously equivalent to the one we would get with a unique trans-
formation34
r = r0 + i (F1 + F2 ), u = u0 + i (G1 + G2 ). (C.9)
Then, for the transformations which are such that
the DJN transformations form an Abelian group thanks to the group properties of the
function space. This structure implies that we can first add one parameter, and later another
one (say first the NUT charge, and then an angular momentum). Said another way this
group preserves Einstein equations when the seed metric is a known (stationary) solution.
But note that it may be very difficult to do it as soon as one begins to replace the F in the
functions by their expression, because it obscures the original function – in one word we can
not find f˜i (r, F ) from f˜i (r, θ).
Another point worth to mention is that not all DJN transformation are in this group
since the condition (C.10) may not satisfied: we recall that imposing or not the latter is a
choice that one is doing when performing the algorithm. A simple example is provided by
f (r) = r2 , (C.11)
in two ways:
f˜{1} = |r| = r02 + F12 , f˜{1,2} = |r0 | + F12 = r002 + F12 + F22 ,
2 2
1. (C.13a)
f˜{1,2} = |r00 + i(F1 + F2 )|
2
f˜{1} = |r| = |r0 + iF1 | ,
2 2
2. (C.13b)
= r002 + F12 + F22 + 2F1 F2 .
Only the second option satisfy the property (C.10) that leads to a group. Such an example
is provided in 5d where the function fΩ (r) = r2 is successively transformed as [58]
2 2
r2 −→ |r| = r2 + a2 cos2 θ −→ |r| + a2 cos2 θ = r2 + a2 cos2 θ + b2 sin2 θ, (C.14)
dimensions [58].
59
C.2 Chaining transformations
The JN algorithm is summarized by the following table
t → u → u∈C → u0 → t0
r → r∈C → r0
(C.16)
φ → φ0
f → f˜
where the complexification of the metric function f can be made at the end. It is impressive
that the algorithm from section 2 can be written in such a compact way.
using respectively the rules (4.14b) and (4.14c) (in the denominator).
But it is possible to arrive at the same result with a different combinations of rules. In
fact the functions can be rewritten as
r 1 1
f1 (r) = , f2 (r) = . (C.20)
r2 r r
The following set of rules results again in (C.19):
• f1 : (4.14a) (numerator) and (4.14c) (denominator);
• f2 : (4.14a) (first fraction) and (4.14b) (second fraction).
References
[1] J. F. Plebański. ‘A Class of Solutions of Einstein-Maxwell Equations’. Annals of
Physics 90.1 (Mar. 1975), pp. 196–255.
doi: 10.1016/0003-4916(75)90145-1.
60
[2] J. F. Plebański and M. Demiański. ‘Rotating, Charged, and Uniformly Accelerating
Mass in General Relativity’. Annals of Physics 98.1 (May 1976), pp. 98–127.
doi: 10.1016/0003-4916(76)90240-2.
[3] E. T. Newman and A. I. Janis. ‘Note on the Kerr Spinning-Particle Metric’. Journal
of Mathematical Physics 6.6 (June 1965), pp. 915–917.
doi: 10.1063/1.1704350.
[4] E. T. Newman, E. Couch, K. Chinnapared, A. Exton, A. Prakash and R. Torrence.
‘Metric of a Rotating, Charged Mass’. Journal of Mathematical Physics 6.6 (June
1965), pp. 918–919.
doi: 10.1063/1.1704351.
[5] G. Giampieri. ‘Introducing Angular Momentum into a Black Hole Using Complex
Variables’. Gravity Research Foundation (1990).
[6] D. Nawarajan and M. Visser. ‘Cartesian Kerr-Schild Variation on the Newman-Janis
Ansatz’ (Jan. 2016).
arXiv: 1601.03532 [gr-qc].
[7] M. Demiański. ‘New Kerr-like Space-Time’. Physics Letters A 42.2 (Nov. 1972),
pp. 157–159.
doi: 10.1016/0375-9601(72)90752-9.
[8] S. P. Drake and P. Szekeres. ‘Uniqueness of the Newman-Janis Algorithm in Gen-
erating the Kerr-Newman Metric’. General Relativity and Gravitation 32.3 (2000),
pp. 445–457.
doi: 10.1023/A:1001920232180.
arXiv: gr-qc/9807001.
[9] M. Azreg-Aïnou. ‘From Static to Rotating to Conformal Static Solutions: Rotating
Imperfect Fluid Wormholes with(out) Electric or Magnetic Field’. The European
Physical Journal C 74.5 (May 2014).
doi: 10.1140/epjc/s10052-014-2865-8.
arXiv: 1401.4292.
[10] C. J. Talbot. ‘Newman-Penrose Approach to Twisting Degenerate Metrics’. Commu-
nications in Mathematical Physics 13.1 (Mar. 1969), pp. 45–61.
doi: 10.1007/BF01645269.
[11] M. Gürses and F. Gürsey. ‘Lorentz Covariant Treatment of the Kerr–Schild Geo-
metry’. Journal of Mathematical Physics 16.12 (Dec. 1975), pp. 2385–2390.
doi: 10.1063/1.522480.
[12] M. M. Schiffer, R. J. Adler, J. Mark and C. Sheffield. ‘Kerr Geometry as Complexified
Schwarzschild Geometry’. Journal of Mathematical Physics 14.1 (Jan. 1973), pp. 52–
56.
doi: 10.1063/1.1666171.
[13] R. J. Finkelstein. ‘The General Relativistic Fields of a Charged Rotating Source’.
Journal of Mathematical Physics 16.6 (June 1975), pp. 1271–1277.
doi: 10.1063/1.522667.
[14] E. T. Newman. ‘Complex Coordinate Transformations and the Schwarzschild-Kerr
Metrics’. Journal of Mathematical Physics 14.6 (June 1973), pp. 774–776.
doi: doi:10.1063/1.1666393.
[15] E. T. Newman and J. Winicour. ‘A Curiosity Concerning Angular Momentum’.
Journal of Mathematical Physics 15.7 (July 1974), pp. 1113–1115.
doi: doi:10.1063/1.1666761.
61
[16] E. T. Newman. ‘Heaven and Its Properties’. en. General Relativity and Gravitation
7.1 (Jan. 1976), pp. 107–111.
doi: 10.1007/BF00762018.
[17] R. Ferraro. ‘Untangling the Newman-Janis Algorithm’. General Relativity and Grav-
itation 46.4 (Apr. 2014).
doi: 10.1007/s10714-014-1705-3.
arXiv: 1311.3946,.
[18] T. Adamo and E. T. Newman. ‘The Kerr-Newman Metric: A Review’. Scholarpedia
9 (Oct. 2014), p. 31791.
doi: 10.4249/scholarpedia.31791.
arXiv: 1410.6626.
[19] F. J. Ernst. ‘New Formulation of the Axially Symmetric Gravitational Field Problem’.
Physical Review 167.5 (Mar. 1968), pp. 1175–1178.
doi: 10.1103/PhysRev.167.1175.
[20] F. J. Ernst. ‘New Formulation of the Axially Symmetric Gravitational Field Problem.
II’. Physical Review 168.5 (Apr. 1968), pp. 1415–1417.
doi: 10.1103/PhysRev.168.1415.
[21] H. Quevedo. ‘Complex Transformations of the Curvature Tensor’. en. General Re-
lativity and Gravitation 24.7 (July 1992), pp. 693–703.
doi: 10.1007/BF00760076.
[22] H. Quevedo. ‘Determination of the Metric from the Curvature’. en. General Relativity
and Gravitation 24.8 (Aug. 1992), pp. 799–819.
doi: 10.1007/BF00759087.
[23] D.-Y. Xu. ‘Exact Solutions of Einstein and Einstein-Maxwell Equations in Higher-
Dimensional Spacetime’. en. Classical and Quantum Gravity 5.6 (June 1988), p. 871.
doi: 10.1088/0264-9381/5/6/008.
[24] H. Kim. ‘Notes on Spinning AdS_3 Black Hole Solution’ (June 1997).
[25] H. Kim. ‘Spinning BTZ Black Hole versus Kerr Black Hole: A Closer Look’. Physical
Review D 59.6 (Feb. 1999), p. 064002.
doi: 10.1103/PhysRevD.59.064002.
arXiv: gr-qc/9809047.
[26] S. Yazadjiev. ‘Newman-Janis Method and Rotating Dilaton-Axion Black Hole’. Gen-
eral Relativity and Gravitation 32.12 (2000), pp. 2345–2352.
doi: 10.1023/A:1002080003862.
arXiv: gr-qc/9907092.
[27] L. Herrera and J. Jiménez. ‘The Complexification of a Nonrotating Sphere: An Exten-
sion of the Newman–Janis Algorithm’. Journal of Mathematical Physics 23.12 (Dec.
1982), pp. 2339–2345.
doi: 10.1063/1.525325.
[28] S. P. Drake and R. Turolla. ‘The Application of the Newman-Janis Algorithm in
Obtaining Interior Solutions of the Kerr Metric’. Classical and Quantum Gravity
14.7 (July 1997), pp. 1883–1897.
doi: 10.1088/0264-9381/14/7/021.
arXiv: gr-qc/9703084.
[29] E. N. Glass and J. P. Krisch. ‘Kottler-Lambda-Kerr Spacetime’ (May 2004).
arXiv: gr-qc/0405143.
62
[30] N. Ibohal. ‘Rotating Metrics Admitting Non-Perfect Fluids in General Relativity’.
General Relativity and Gravitation 37.1 (Jan. 2005), pp. 19–51.
doi: 10.1007/s10714-005-0002-6.
arXiv: gr-qc/0403098.
[31] M. Azreg-Aïnou. ‘Generating Rotating Regular Black Hole Solutions without Com-
plexification’. Physical Review D 90.6 (Sept. 2014).
doi: 10.1103/PhysRevD.90.064041.
arXiv: 1405.2569.
[32] B. Carter. ‘Hamilton-Jacobi and Schrödinger Separable Solutions of Einstein’s Equa-
tions’. Communications in Mathematical Physics (1965-1997) 10.4 (1968), pp. 280–
310.
[33] G. W. Gibbons and S. W. Hawking. ‘Cosmological Event Horizons, Thermodynamics,
and Particle Creation’. Physical Review D 15.10 (May 1977), pp. 2738–2751.
doi: 10.1103/PhysRevD.15.2738.
[34] D. Klemm, V. Moretti and L. Vanzo. ‘Rotating Topological Black Holes’ (Oct. 1997).
[35] E. J. G. de Urreta and M. Socolovsky. ‘Extended Newman-Janis Algorithm and
Rotating and Kerr-Newman de Sitter (Anti de Sitter) Metrics’ (Apr. 2015).
arXiv: 1504.01728 [gr-qc, physics:math-ph].
[36] R. Mallett. ‘Metric of a Rotating Radiating Charged Mass in a de Sitter Space’.
Physics Letters A 126.4 (Jan. 1988), pp. 226–228.
doi: 10.1016/0375-9601(88)90750-5.
[37] S. Viaggiu. ‘Interior Kerr Solutions with the Newman-Janis Algorithm Starting with
Static Physically Reasonable Space-Times’. International Journal of Modern Physics
D 15.09 (Sept. 2006), pp. 1441–1453.
doi: 10.1142/S0218271806009169.
arXiv: gr-qc/0603036.
[38] R. Whisker. ‘Braneworld Black Holes’. PhD thesis. University of Durham, Oct. 2008.
[39] G. Lessner. ‘The “complex Trick” in Five-Dimensional Relativity’. en. General Re-
lativity and Gravitation 40.10 (Mar. 2008), pp. 2177–2184.
doi: 10.1007/s10714-008-0625-5.
[40] S. Capozziello, M. De Laurentis and A. Stabile. ‘Axially Symmetric Solutions in
f (R)-Gravity’. Class.Quant.Grav. 27 (2010), p. 165008.
doi: 10.1088/0264-9381/27/16/165008.
[41] F. Caravelli and L. Modesto. ‘Spinning Loop Black Holes’. Classical and Quantum
Gravity 27.24 (Dec. 2010), p. 245022.
doi: 10.1088/0264-9381/27/24/245022.
arXiv: 1006.0232.
[42] N. Dadhich and S. G. Ghosh. ‘Rotating Black Hole in Einstein and Pure Lovelock
Gravity’ (July 2013).
arXiv: 1307.6166 [astro-ph, physics:gr-qc, physics:hep-th].
[43] S. G. Ghosh and U. Papnoi. ‘Spinning Higher Dimensional Einstein-Yang-Mills Black
Holes’ (Sept. 2013).
arXiv: 1309.4231 [gr-qc].
[44] S. G. Ghosh. ‘Rotating Black Hole and Quintessence’ (Dec. 2015).
arXiv: 1512.05476 [gr-qc].
63
[45] M. Azreg-Aïnou. ‘Comment on "Spinning Loop Black holes" [arXiv:1006.0232]’. Clas-
sical and Quantum Gravity 28.14 (July 2011), p. 148001.
doi: 10.1088/0264-9381/28/14/148001.
arXiv: 1106.0970.
[46] D. Xu. ‘Radiating Metric, Retarded Time Coordinates of Kerr-Newman-de Sitter
Black Holes and Related Energy-Momentum Tensor’. Science in China Series A:
Mathematics 41.6 (June 1998), pp. 663–672.
doi: 10.1007/BF02876237.
[47] M. Demiański and E. T. Newman. ‘Combined Kerr-NUT Solution of the Einstein
Field Equations’. Bull. Acad. Pol. Sci., Ser. Sci. Math. Astron. Phys. 14 (1966),
pp. 653–657.
[48] L. K. Patel. ‘Radiating Demianski-Type Space-Times’. Indian J. Pure Appl. Math 9
(1978), p. 1019.
[49] K. D. Krori, T. Chaudhury and R. Bhattacharjee. ‘Charged Demianski Metric’.
Journal of Mathematical Physics 22.10 (Oct. 1981), pp. 2235–2236.
doi: 10.1063/1.524792.
[50] L. K. Patel, R. P. Akabari and U. K. Dave. ‘Radiating Demianski-Type Metrics and
the Einstein-Maxwell Fields’. The ANZIAM Journal 30.01 (July 1988), pp. 120–126.
doi: 10.1017/S0334270000006081.
[51] Y. F. Pirogov. ‘Towards the Rotating Scalar-Vacuum Black Holes’ (June 2013).
arXiv: 1306.4866 [gr-qc, physics:hep-ph, physics:math-ph].
[52] D. Hansen and N. Yunes. ‘Applicability of the Newman-Janis Algorithm to Black
Hole Solutions of Modified Gravity Theories’. Physical Review D 88.10 (Nov. 2013),
p. 104020.
doi: 10.1103/PhysRevD.88.104020.
arXiv: 1308.6631.
[53] J. H. Horne and G. T. Horowitz. ‘Rotating Dilaton Black Holes’. Physical Review D
46.4 (Aug. 1992), pp. 1340–1346.
doi: 10.1103/PhysRevD.46.1340.
arXiv: hep-th/9203083.
[54] D. Cirilo-Lombardo. ‘The Newman-Janis Algorithm, Rotating Solutions and Einstein-
Born-Infeld Black Holes’ (Dec. 2006).
[55] R. D’Inverno. Introducing Einstein’s Relativity. Anglais. Clarendon Press, Aug. 1992.
[56] J. F. Reed. ‘Some Imaginary Tetrad-Transformations of Einstein Spaces’. PhD thesis.
Rice University, 1974.
[57] H. Erbin. ‘Janis-Newman Algorithm: Simplifications and Gauge Field Transforma-
tion’. General Relativity and Gravitation 47.3 (Mar. 2015), p. 19.
doi: 10.1007/s10714-015-1860-1.
arXiv: 1410.2602.
[58] H. Erbin and L. Heurtier. ‘Five-Dimensional Janis-Newman Algorithm’. Classical and
Quantum Gravity 32.16 (Aug. 2015), p. 165004.
doi: 10.1088/0264-9381/32/16/165004.
arXiv: 1411.2030.
[59] H. Erbin. ‘Deciphering and Generalizing Demianski-Janis-Newman Algorithm’. Gen-
eral Relativity and Gravitation 48.5 (May 2016).
doi: 10.1007/s10714-016-2054-1.
arXiv: 1411.2909.
64
[60] H. Erbin and L. Heurtier. ‘Supergravity, Complex Parameters and the Janis-Newman
Algorithm’. Classical and Quantum Gravity 32.16 (Aug. 2015), p. 165005.
doi: 10.1088/0264-9381/32/16/165005.
arXiv: 1501.02188.
[61] H. Erbin. ‘Black Holes in N = 2 Supergravity’. PhD thesis. Université Pierre et Marie
Curie – Paris VI, Sept. 2015.
[62] A. J. Keane. ‘An Extension of the Newman-Janis Algorithm’. Classical and Quantum
Gravity 31.15 (Aug. 2014), p. 155003.
doi: 10.1088/0264-9381/31/15/155003.
arXiv: 1407.4478.
[63] A. Sen. ‘Rotating Charged Black Hole Solution in Heterotic String Theory’. Physical
Review Letters 69.7 (Aug. 1992), pp. 1006–1009.
doi: 10.1103/PhysRevLett.69.1006.
arXiv: hep-th/9204046.
[64] M. J. Perry. ‘Black Holes Are Coloured’. Physics Letters B 71.1 (Nov. 1977), pp. 234–
236.
doi: 10.1016/0370-2693(77)90786-9.
[65] K. Behrndt, D. Lüst and W. A. Sabra. ‘Stationary Solutions of N = 2 Supergravity’.
Nuclear Physics B 510.1-2 (Jan. 1998), pp. 264–288.
doi: 10.1016/S0550-3213(97)00633-0.
arXiv: hep-th/9705169.
[66] E. Bergshoeff, R. Kallosh and T. Ortín. ‘Stationary Axion/Dilaton Solutions and
Supersymmetry’. Nuclear Physics B 478.1-2 (Oct. 1996), pp. 156–180.
doi: 10.1016/0550-3213(96)00408-7.
arXiv: hep-th/9605059.
[67] Y. Bardoux, M. M. Caldarelli and C. Charmousis. ‘Integrability in Conformally
Coupled Gravity: Taub-NUT Spacetimes and Rotating Black Holes’ (Nov. 2013).
arXiv: 1311.1192 [gr-qc, physics:hep-th].
[68] R. Myers and M. Perry. ‘Black Holes in Higher Dimensional Space-Times’. Annals of
Physics 172.2 (Dec. 1986), pp. 304–347.
doi: 10.1016/0003-4916(86)90186-7.
[69] J. C. Breckenridge, R. C. Myers, A. W. Peet and C. Vafa. ‘D-Branes and Spinning
Black Holes’. Physics Letters B 391.1-2 (Jan. 1997), pp. 93–98.
doi: 10.1016/S0370-2693(96)01460-8.
arXiv: hep-th/9602065.
[70] A. Gnecchi, K. Hristov, D. Klemm, C. Toldo and O. Vaughan. ‘Rotating Black Holes
in 4d Gauged Supergravity’. Journal of High Energy Physics 2014.1 (Jan. 2014).
doi: 10.1007/JHEP01(2014)127.
arXiv: 1311.1795.
[71] M. Visser. ‘The Kerr Spacetime: A Brief Introduction’. The Kerr Spacetime. Rotating
Black Holes in General Relativity. Ed. by D. L. Wiltshire, M. Visser and S. M. Scott.
Cambridge University Press, Feb. 2009.
[72] S. M. Carroll. Spacetime and Geometry: An Introduction to General Relativity. Eng-
lish. Addison Wesley, 2004.
65
[73] N. Alonso-Alberca, P. Meessen and T. Ortín. ‘Supersymmetry of Topological Kerr-
Newmann-Taub-NUT-aDS Spacetimes’. Classical and Quantum Gravity 17.14 (July
2000), pp. 2783–2797.
doi: 10.1088/0264-9381/17/14/312.
arXiv: hep-th/0003071.
[74] J. B. Griffiths and J. Podolsky. ‘A New Look at the Plebanski-Demianski Family of
Solutions’. International Journal of Modern Physics D 15.03 (Mar. 2006), pp. 335–
369.
doi: 10.1142/S0218271806007742.
arXiv: gr-qc/0511091.
[75] A. Chamblin, R. Emparan, C. V. Johnson and R. C. Myers. ‘Large N Phases, Grav-
itational Instantons and the Nuts and Bolts of AdS Holography’. Physical Review D
59.6 (Feb. 1999).
doi: 10.1103/PhysRevD.59.064010.
arXiv: hep-th/9808177.
[76] C. V. Johnson. ‘Thermodynamic Volumes for AdS-Taub-NUT and AdS-Taub-Bolt’.
Class.Quant.Grav. 31 (Nov. 2014), p. 235003.
doi: 10.1088/0264-9381/31/23/235003.
arXiv: 1405.5941.
[77] A. Krasiński. Inhomogeneous Cosmological Models. English. Cambridge University
Press, Nov. 2006.
[78] R. G. Leigh, A. C. Petkou, P. M. Petropoulos and P. K. Tripathy. ‘The Geroch Group
in Einstein Spaces’. Classical and Quantum Gravity 31.22 (Nov. 2014), p. 225006.
doi: 10.1088/0264-9381/31/22/225006.
arXiv: 1403.6511.
[79] J. D. Bekenstein. ‘Exact Solutions of Einstein-Conformal Scalar Equations’. Annals
of Physics 82.2 (Feb. 1974), pp. 535–547.
doi: 10.1016/0003-4916(74)90124-9.
[80] N. M. Bocharova, K. A. Bronnikov and V. N. Melnikov. ‘An Exact Solution of the
System of Einstein Equations and Mass-Free Scalar Field’. Vestn. Mosk. Univ. Fiz.
Astro. 6 (1970), p. 706.
[81] K. Hristov, H. Looyestijn and S. Vandoren. ‘BPS Black Holes in N=2 D=4 Gauged
Supergravities’. Journal of High Energy Physics 2010.8 (Aug. 2010).
doi: 10.1007/JHEP08(2010)103.
arXiv: 1005.3650.
[82] D. D. K. Chow and G. Compère. ‘Black Holes in N=8 Supergravity from SO(4,4)
Hidden Symmetries’. Physical Review D 90.2 (July 2014), p. 025029.
doi: 10.1103/PhysRevD.90.025029.
arXiv: 1404.2602.
[83] T. Ortín. Gravity and Strings. English. Cambridge University Press, 2004.
[84] R. Emparan and H. S. Reall. ‘Black Holes in Higher Dimensions’. Living Rev.Rel. 11
(Jan. 2008), p. 6.
[85] A. N. Aliev. ‘Rotating Black Holes in Higher Dimensional Einstein-Maxwell Gravity’.
Physical Review D 74.2 (July 2006), p. 024011.
doi: 10.1103/PhysRevD.74.024011.
66
[86] M. Bañados, C. Teitelboim and J. Zanelli. ‘The Black Hole in Three Dimensional
Space Time’. Physical Review Letters 69.13 (Sept. 1992), pp. 1849–1851.
doi: 10.1103/PhysRevLett.69.1849.
arXiv: hep-th/9204099.
[87] Z.-W. Chong, M. Cvetic, H. Lu and C. N. Pope. ‘General Non-Extremal Rotating
Black Holes in Minimal Five-Dimensional Gauged Supergravity’. Physical Review
Letters 95.16 (Oct. 2005), p. 161301.
doi: 10.1103/PhysRevLett.95.161301.
arXiv: hep-th/0506029.
[88] R. Emparan and H. S. Reall. ‘A Rotating Black Ring in Five Dimensions’. Physical
Review Letters 88.10 (Feb. 2002), p. 101101.
doi: 10.1103/PhysRevLett.88.101101.
arXiv: hep-th/0110260.
[89] J. P. Gauntlett, R. C. Myers and P. K. Townsend. ‘Black Holes of D=5 Supergravity’.
Classical and Quantum Gravity 16.1 (Jan. 1999), pp. 1–21.
doi: 10.1088/0264-9381/16/1/001.
arXiv: hep-th/9810204.
[90] H. S. Reall. ‘Higher Dimensional Black Holes and Supersymmetry’. Physical Review
D 68.2 (July 2003), p. 024024.
doi: 10.1103/PhysRevD.68.024024.
[91] A. N. Aliev. ‘Superradiance and Black Hole Bomb in Five-Dimensional Minimal Un-
gauged Supergravity’ (Aug. 2014).
arXiv: 1408.4269 [gr-qc, physics:hep-th].
[92] J. P. Gauntlett, J. B. Gutowski, C. M. Hull, S. Pakis and H. S. Reall. ‘All Su-
persymmetric Solutions of Minimal Supergravity in Five Dimensions’. Classical and
Quantum Gravity 20.21 (Nov. 2003), pp. 4587–4634.
doi: 10.1088/0264-9381/20/21/005.
arXiv: hep-th/0209114.
[93] G. W. Gibbons, D. Kastor, L. A. J. London, P. K. Townsend and J. Traschen. ‘Super-
symmetric Self-Gravitating Solitons’. Nuclear Physics B 416.3 (Apr. 1994), pp. 850–
880.
doi: 10.1016/0550-3213(94)90558-4.
arXiv: hep-th/9310118.
[94] A. Puhm. ‘Black Holes in String Theory: Guides to Quantum Gravity’. PhD thesis.
Université Pierre et Marie Curie - Paris VI, 2013.
[95] G. W. Gibbons and C. A. R. Herdeiro. ‘Supersymmetric Rotating Black Holes and
Causality Violation’. Classical and Quantum Gravity 16.11 (Nov. 1999), pp. 3619–
3652.
doi: 10.1088/0264-9381/16/11/311.
arXiv: hep-th/9906098.
[96] A. N. Aliev and D. K. Ciftci. ‘Note on Rotating Charged Black Holes in Einstein-
Maxwell-Chern-Simons Theory’. Physical Review D 79.4 (Feb. 2009), p. 044004.
doi: 10.1103/PhysRevD.79.044004.
arXiv: 0811.3948.
[97] B. Ett and D. Kastor. ‘An Extended Kerr-Schild Ansatz’. Classical and Quantum
Gravity 27.18 (Sept. 2010), p. 185024.
doi: 10.1088/0264-9381/27/18/185024.
arXiv: 1002.4378.
67
[98] T. Málek. ‘Extended Kerr-Schild Spacetimes: General Properties and Some Explicit
Examples’. Classical and Quantum Gravity 31.18 (Sept. 2014), p. 185013.
doi: 10.1088/0264-9381/31/18/185013.
arXiv: 1401.1060.
[99] F. R. Tangherlini. ‘Schwarzschild Field in Dimensions and the Dimensionality of
Space Problem’. en. Il Nuovo Cimento 27.3 (Feb. 1963), pp. 636–651.
doi: 10.1007/BF02784569.
[100] G. Clément. ‘Classical Solutions in Three-Dimensional Einstein-Maxwell Cosmolo-
gical Gravity’. en. Classical and Quantum Gravity 10.5 (1993), p. L49.
doi: 10.1088/0264-9381/10/5/002.
[101] G. Clément. ‘Spinning Charged BTZ Black Holes and Self-Dual Particle-like Solu-
tions’. Physics Letters B 367.1-4 (Jan. 1996), pp. 70–74.
doi: 10.1016/0370-2693(95)01464-0.
arXiv: gr-qc/9510025.
[102] C. Martinez, C. Teitelboim and J. Zanelli. ‘Charged Rotating Black Hole in Three
Spacetime Dimensions’. Physical Review D 61.10 (Apr. 2000), p. 104013.
doi: 10.1103/PhysRevD.61.104013.
arXiv: hep-th/9912259.
[103] P.-H. Lambert. ‘Conformal Symmetries of Gravity from Asymptotic Methods: Fur-
ther Developments’ (Sept. 2014).
arXiv: 1409.4693 [gr-qc, physics:hep-th].
[104] G. W. Gibbons, H. Lu, D. N. Page and C. N. Pope. ‘The General Kerr-de Sitter
Metrics in All Dimensions’. Journal of Geometry and Physics 53.1 (Jan. 2005), pp. 49–
73.
doi: 10.1016/j.geomphys.2004.05.001.
arXiv: hep-th/0404008.
[105] D. Z. Freedman and A. Van Proeyen. Supergravity. English. Cambridge University
Press, May 2012.
[106] L. Andrianopoli, M. Bertolini, A. Ceresole, R. D’Auria, S. Ferrara and P. Fré. ‘General
Matter Coupled N=2 Supergravity’. Nuclear Physics B 476.3 (Sept. 1996), pp. 397–
417.
doi: 10.1016/0550-3213(96)00344-6.
arXiv: hep-th/9603004.
[107] L. Andrianopoli, M. Bertolini, A. Ceresole, R. D’Auria, S. Ferrara, P. Fré and T.
Magri. ‘N=2 Supergravity and N=2 Super Yang-Mills Theory on General Scalar
Manifolds: Symplectic Covariance, Gaugings and the Momentum Map’. Journal of
Geometry and Physics 23.2 (Sept. 1997), pp. 111–189.
doi: 10.1016/S0393-0440(97)00002-8.
arXiv: hep-th/9605032.
68