Geometric Relations of Black Hole Thermodynamics: MSC Physics
Geometric Relations of Black Hole Thermodynamics: MSC Physics
Theoretical Physics
Author: Supervisor:
Vasilis Lefkos Dr. Alejandra Castro
Examiner:
Dr. B.W. Freivogel
1
Contents
1 Introduction 5
2
5.3 Kerr/CFT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
5.3.1 Three dimensional gravity as a conformal field theory . . . . 67
5.3.2 Near horizon geometry . . . . . . . . . . . . . . . . . . . . . 69
5.3.3 NHEK as a conformal field theory . . . . . . . . . . . . . . . 70
5.3.4 Temperature and entropy . . . . . . . . . . . . . . . . . . . 71
5.3.5 Other extremal black objects and their dual CFTs . . . . . . 72
5.4 Hidden conformal symmetry . . . . . . . . . . . . . . . . . . . . . . 73
7 Conclusions 98
3
B Useful tools 137
B.1 Diffeomorphism invariance . . . . . . . . . . . . . . . . . . . . . . . 137
B.2 Energy conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
B.3 Frobenius’ theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
H Kerr-Newman-CFT 153
References 161
4
The law that entropy always increases, holds, I think, the supreme position
among the laws of Nature. If someone points out to you that your pet theory of the
universe is in disagreement with Maxwell’s equations — then so much the worse
for Maxwell’s equations. If it is found to be contradicted by observation — well,
these experimentalists do bungle things sometimes. But if your theory is found
to be against the second law of thermodynamics I can give you no hope; there is
nothing for it but to collapse in deepest humiliation.
-Arthur Eddington
1 Introduction
General relativity is a non-linear theory and non renormalizable perturbatively.
The fact that spacetime it self, instead of being a fixed background as in normal
field theories, is dynamical, makes it really difficult to find a correct method for
quantization. One would need to quantize spacetime itself. The main symmetry of
quantum gravity, diffeomorphisms, cause problems with locality in contrast with
other gauge field theories, since physical observables are non local. Moreover,
causality and unitarity problems arise when one attempts to formulate a quantum
field theory near black holes - one example is the black hole information paradox,
where pure states passing the horizon come out as mixed states through thermal
radiation.
Unfortunately or not, the supreme agreement of general relativity with experi-
ment renders it the most popular low energy effective theory. Black holes however,
belong to the regime of general relativity where one cannot rely solely upon the
classical theory, or even semi-classical. They serve as the perfect playground for
quantum gravity and it is a rich arena for theoretical physics.
The attempt of quantizing gravity with normal quantum field theory and the
discovery of black hole thermodynamics, a celebrated work by Bekenstein, Bardeen,
Carter and Hawking [103], was evidence towards a deep connection between ther-
modynamics and quantum gravity. Later, the nourishing of string theory and
holography with the greatest conjecture of contemporary physics, the AdS/CFT
correspondence, established the use of a thermal field theory.
These innovations led to the exploration and better understanding of conformal
field theories, as dual theories to a gravitational theory, even for the case of black
holes. Hence the thermal behavior of black holes could be studied both macroscop-
ically and microscopically. A very interesting endeavor is examining gravitational
behaviors from a conformal field theory point of view and vice versa.
One of the latest research subjects, however, is the role of inner horizons in
black hole thermodynamics and the dual holographic description of a black hole.
5
Despite the fact that inner horizons can have problems in the classical theory, such
as divergences for observers who cross them, they turn out to have an impact on
the system. For instance, they seem important for the holographic dual as well
as the string theory description. Inner horizons have thermodynamical properties
similarly with the outer event horizons. The same applies for multiple horizon
black holes, with different asymptotics or dimensionalities.
The aim of this project is to make use of these horizons and their thermal
properties, entropies and potentials, in order to have a clearer view of their role
in black hole physics, as well as their dual conformal theories. There are two
geometrical potential relations reviewed; one is the entropy product, which seems
to be mass independent for a large class of black holes and theories, and which
was studied at some level in the literature [58, 89, 174, 190, 196]. The second type
of relation is one that does not necessarily involve entropies, rather than geometric
potentials, namely the surface gravity and the angular velocity of the horizon.
These relations are merely observed for some cases in the literature [183, 184, 203]
and not deeply analyzed.
The structure of the project is as follows: We first introduce black hole physics
in General Relativity in Chapter 2, as well as black hole thermodynamics in Chap-
ter 3. We especially review the first law, which we will use extensively in the
rest of the project. Then we preview cosmological gravity in Chapter 4, together
with gravity in various dimensions. What follows is Chapter 5, the introduction
to conformal field theories and their correlation to black hole physics, with many
modern theories, such as Kerr/CFT and hidden conformal symmetry. We end up
with inner horizons and their thermal properties, as well as the geometric potential
relations and the work done on them in Chapter 6. Finally, there is an appendix
with a presentation of all the relations, as well as some computations on various
subjects used throughout.
6
2 Black hole physics preliminaries
a definition due to Penrose in 1967 [5]. It is the set of points from which there
exist future and past directed timelike curves to arbitrarily large distances.
A black hole is now defined as
B = M − I − (J + ) . (2.3)
7
The region M − D is called eternal black hole and it is the maximal analytical
extension of the black hole (see fig.2.2 for example). This region has of course a
boundary H which can be thought as a union of two subsets H+ ∪H− , where H+ is
the boundary of I − (J + ), and similarly for H− . There is an equivalent definition.
Let us have a set of points U ⊂ M . Then the topological closure J¯− (U )of J − (U )
is the causal past of U but including the limit points (infinities). The boundary of
this closure will be naturally
Then the future horizon(future boundary of D will be the boundary of the closure
of the causal past of J + [6]
H+ = J˙− (J + ) (2.5)
and similarly for H− . So we have finally a region defined as a black hole. It should
be clear from the previous that no light can escape from this region to an external
observer, which serves as a reason for naming it black. For a black object that
was formed from a gravitational collapse, H− would not exist. Theoretically, the
region behind this subset cannot be reached by any signal and it is called a white
hole, but it will not be relevant for this thesis. Similarly to the black hole, it is
defined as W = M − I + (J − ).
8
for an asymptotically flat spacetime. These points, forming the conformal boundary
of the spacetime, also include infinity. An easy illustration is the flat 4d Minkowski
spacetime conformal diagram. In order to obtain the diagram, we start with the
4d Minkowski metric in polar coordinates
with dΩ2 being the unit two sphere metric and −∞ < t < ∞ , 0 ≤ r < ∞. We
then transform into null coordinates u = t − r , v = t + r with −∞ < u ≤ v < ∞.
The element becomes
1 1
ds2 = − (dudv + dvdu) + (u − v)2 dΩ2 . (2.9)
2 4
The next transformation is needed so that we can have infinity at a finite coordi-
nate. We use U = arctan(u) , V = arctan(v) which means −π/2 < U ≤ V < π/2
leading to the metric
2 1 2 2
ds = − 2(dV dU + dU dV ) + sin (V − U )dΩ . (2.10)
4cos2 U cos2 V
The final step is to go back to time/radial coordinates (T, R) which now have
different ranges. If we set R = V − U , T = V + U then the metric becomes
1
ds2 = (−dT 2 + dR2 + sin2 RdΩ2 ) = Ω−2 (T, R)ds2M ink . (2.11)
(cosT + cosR)2
with ranges 0 ≤ R < π , |T | + R < π. It is obvious now that the two metrics are
related with the scaling parameter. As we will see in the diagram, there are some
special points that need to be defined:
9
Figure 2.1: Minkowski Penrose diagram [9]
The first three are represented as points in the diagram , which means they are
S spheres. The latter are null hypersurfaces R × S 2 . Null radial geodesics are at
2
2.4.1 Schwarzschild
The first exact vacuum solution to Einstein’s equations was found in 1916 by
Karl Schwarzschild. It is a spherically symmetric solution in four dimensional
asymptotically flat spacetime. The metric reads (in coordinates (t, r, θ, φ)
−1
2 2M 2 2M
ds = − 1 − dt + 1 − dr2 + r2 dΩ2 , (2.12)
r r
where M will be the mass of the black hole. We will discuss what this mass is
and how it is computed later, in section 2.6.1. One can already observe that there
seems to be a problem at r = 2M , what is called the Schwarzschild radius. It
is also the radius at which the boundary of the black hole is formed, known as
10
the event horizon. The metric there is singular; however, one can get rid of this
singularity with a change in coordinates. It is just a coordinate singularity. On the
other hand, the r = 0 point cannot be remedied. It is a curvature singularity and
these kind of singularities appear in most black hole solutions. In order to see if a
singularity is physical, one needs to check if the quantity
Rµνρσ Rµνρσ ,
which is called the Kretschmann scalar, diverges- this is at least a sufficient con-
dition, and usually the case. Another way to define a curvature singularity is
geodesic incompleteness; if there is a singularity of this type, geodesics cannot be
extended to all values of their affine parameters. That would mean there is a
geodesic congruence that converges to that very point.
As promised, a coordinate transformation should make the r = 2M singularity
disappear. For, instance, one could do the following: Define a coordinate
r − 2M
v = t + r∗ where r∗ = r + 2M ln
2M
which are both null (`2 = n2 = 0) and they are normalized as ` · n = −1. Assum-
ing r → ∞ is outwards, then `a is outward pointing and na is inward pointing.
Similarly there are the outgoing coordinates with
v = t − r∗ .
where dΩ2 = dθ2 + sinθ dφ2 . We go through all this trouble to finaly introduce the
11
very useful Kruskal-Szekeres coordinates. With a final transformation
r − 2M r/2M
U = −e−u/4M V = ev/4M with U V = − e (2.16)
2M
we get
−32M 2 −r/2M
ds2 = e dU dV + r2 dΩ2 , (2.17)
r
which has the privilege of analytic continuation to all values of U, V ∈ R. So we
are ready to have a diagrammatic picture of the Schwarzschild solution:
12
2.4.2 Reissner-Nordström
The equation of motion for the strength field is ∇µ F µν = 0. The solution is also
spherically symmetric like Schwarzschild and its metric reads
Q2 dr2
2 2M 2 2 2
ds = − 1 − + 2 dt + 2M Q 2 + r dΩ . (2.20)
r r 1 − r + r2
The Maxwell one-form is A = Qr dt. We can observe that now there is an ex-
tra quantity in the metric, namely the electric charge Q. One could also have a
generalization of this [206], the dyonic black hole with e2 = Q2 + P 2 the dyonic
charge, where Q, P would be the electric and magnetic charge respectfully, assum-
ing that magnetic monopoles exist. The Maxwell one-form would get an extra
term A = Qr dt + P cosθdφ.
The horizons of the Reissner-Nordström black hole can be found by solving the
equation
2M Q2
∆=1− + 2 =0. (2.21)
r r
We will later see (section 2.5) that this is related to the Killing vector being null on
that surface. Hence, the Reissner-Nordström metric admits two horizons sitting at
p
r± = M ± M 2 − Q2 .
The outer most one is always the event horizon and we name the other one
inner horizon. The radicand will determine the reality of the event horizon radius
and hence existence of the horizon. The limit where M = Q is called extremal.
An extremal black hole is a black hole with the minimum possible mass in relation
with the rest of its charges. At that limit the two horizons also coincide. The
cosmic censorship conjecture [13] prevents M acquiring values below Q.
Performing some coordinate transformations as in the Schwarzschild case, we
13
first take ingoing EF coordinates leading to
∆ 2
ds2 = − dv + 2dvdr + r2 dΩ2 . (2.22)
r2
We can see now that the singularities of r± in the previous metric were just a
coordinate singularities, justifying their identifications as horizons. There is only
one singularity in these coordinates, at r = 0.
We do similarly for outgoing EF coordinates. The difference comes at the
Kruskal type coordinates. We have two sets, since we have two horizons. Specifi-
cally the KS coordinates will be
The parameter κ± is called surface gravity and we will explain its role and meaning
in chapter 2.5. The metric for the plus sign takes the form
κκ+ −1
r+ r− e−2κ+ r
2 r− −
ds = − 2 2
dV + dU + + r2 dΩ2 (2.24)
κ+ r r − r−
with
κκ+
+ + 2κ+ r r − r+ r − r− −
V U = −e . (2.25)
r+ r−
This way we get regions I-IV as in the Schwarzschild. These coordinates are sin-
gular for r = r− and we need different ones to cover the inner region. That is why
we defined V − , U − . The metric there is
κκ− −1
r+ r− e−2κ− r
r+ +
2
ds = − 2 2
dV − dU − + r2 dΩ2 (2.26)
κ− r r+ − r
with
κκ−
− − 2κ− r r− − r r+ − r +
V U = −e . (2.27)
r− r+
This gives us another four regions, with one being common with region II of the
previous coordinates (see fig 2.3). Similarly we can patch a different set of ”plus”
coordinates on top of these four regions, leading to another exterior spacetime,
another black hole etc. Following this method the diagram can be extended both
ways infinitely.
14
Figure 2.3: The RN Penrose diagram [6]
Roy P.Kerr found a rotating solution to Einstein’s vacuum field equations, general-
ized in 1965 by Ezra Newman to include electric/magnetic charge as well, solving
the Einstein-Maxwell field equations. The metric is (in Boyer-Lindquist coordi-
nates):
∆ − a2 sin2 θ 2 2 2
2 r +a −∆
ds2 = − dt − 2asin θ dtdφ+
ρ2 ρ2
(2.28)
(r2 + a2 )2 − ∆a2 sin2 θ 2 2 ρ2 2
sin θdφ + dr + ρ2 dθ2 ,
ρ2 ∆
15
with
and the new parameter a is essentially the rotating parameter, containing the
angular momentum J. The Maxwell one-form is
M ≥ e2 + a2 . (2.31)
The quantity
gtt a
Ω± = − = 2
gtφ H± r± + a2
is called the angular velocity of the horizon. That is because if we consider an
observer with zero angular momentum (also known as ZAMO- zero angular mo-
mentum observer), then if they try to stay at the horizon, they will rotate with
this velocity.
The horizons are generated by the Killing vectors (they become null on the
horizons):
µ
ξ± = k µ + Ω± mµ ,
16
Figure 2.4: The Kerr Penrose diagram [6]
r 2 + a2
Z
0
z = rcosθ t = dt + dr − r . (2.33)
∆
Then the metric becomes
2
2M r3
2 02 2 2 2 r(xdx + ydy) − a(xdy − ydx) zdz 0
ds = −dt +dx +dy +dz + 4 + +dt ,
r + a2 z 2 r 2 + a2 r
(2.34)
17
which gives us
x2 + y 2 z2
− = a2 (2.36)
sin2 θ cos2 θ
or
x2 + y 2
− r 2 = a2 . (2.37)
sin2 θ
Taking now the singular point of our initial metric r = 0, θ = π/2, we end up with
the equation
x 2 + y 2 = a2 , (2.38)
Ergosurface
We will digress here to elaborate on this special feature of the Kerr geometry due to
its rotation. In the Schwarzschild solution, as we mentioned there is only the time
translation Killing vector that becomes null on the horizon. That is essentially
when the component gtt of the metric vanishes. Specifically looking at the metric
2.12 we can see that it happens for r = 2M .
This Killing vector exists in the Kerr solution as well; but it becomes null on
a surface outside the event horizon, called the ergosurface. The region between
this surface and the horizon is called ergosphere and no observer can be stationary
in this region. They will experience the effect of frame dragging, which means
the black hole forces everything beyond this limit, the ergosurface (also called the
stationary limit), to rotate with it. This can be seen by requiring constant r, then
geodesics move along φ. Suppose there is an observer that wants to stay still in a
coordinate sense. The tangent vector to the observer’s world line will be
dxµ
Pµ = = (1, 0, 0, 0) . (2.39)
dt
In order for his trajectory to be timelike, one also needs
18
where we have defined the ergosurfaces as
√
rE± = M ± M 2 − a2 cos2 θ . (2.42)
k µ kµ = 0 =⇒ gtt k t k t = 0 ,
which leads to the same result. There is actually an ergosurface inside the inner
horizon, but it is never considered due to the instability of r− . The outer one is
important due to the fact that we can extract energy from the black hole ergoregion.
This is called Penrose process, discovered as a gedanken experiment from Penrose
[156]. This is because the time translation Killing vector is spacelike, so one can
have a particle with negative energy
as we saw earlier. Let us elaborate a little bit more on this process. Suppose there
is an observer (or particle) that follows a geodesic towards the ergosphere. Suppose
also that this observer is a system consisting of two parts (1 and 2). The total
momentum and energy of the system will be
since ξ µ is the Killing vector of a rotating black hole. Then (2.45) becomes, if L2
is the angular momentum of the particle (we will assume this happens outside the
outer event horizon for simplicity)
which means the particle must have opposite sign angular momentum than the
black hole (E2 is negative). In terms of the black hole charges, since they changed
19
by
δM = E2 δJ = L2 , (2.47)
Ω+ δJ ≤ δM , (2.48)
where we added an equal sign as the ideal process. This process is also called
superradiance when referring to waves with frequency ω and momentum m, that
get amplified when they enter this region in a band
All these generalize to the Kerr-Newman case, for instance the superradiant band
becomes
0 < ω < mΩ± + qΦ± , (2.50)
where q is the charge of the wave or particle and Φ± is the electric potential of the
horizon, defined as
Φ± = −Aµ ξ µ . (2.51)
r±
As argued by Bekenstein [186], superradiance follows from the second law of black
hole mechanics, which we will present in the next chapter.
20
Since the Killing vector ξ µ is normal to the the Killing horizon H, it will be
proportional to its unit normal aµ . Then we have
aν ∇ ν aµ =0 ξ µ = f aµ =⇒ ξ ν ∇ν ξ µ = κξ µ , (2.52)
H H
where f is a function and κ = ξ µ ∇µ lnf is what we will call the surface gravity.
Using Frobenius’ theorem (B.7) and Killing’s equation (B.2), we can get another
expression for κ:
1
κ2 = − (∇µ ξ ν )(∇µ ξν ) . (2.53)
2 N
It turns out that κ is constant on orbits of ξ. The importance of surface gravity
will be demonstrated in the next chapter. In terms of geodesics, eq. (2.52) means
that the Killing vector is a geodesic on a horizon that is not parameterized affinely,
and that is measured by the surface gravity. In order to demonstrate that, suppose
we have an affine parameter λ and a non affine parameter ν with which we can
express the Killing vector as ξ µ = ∂ν . Then on an orbit of the Killing vector, we
can parameterize ν as ν(λ), so
dλ d
ξµ = . (2.54)
dν dλ
This gives us the expression for f , being f = dλ
dν
. Using then its relation with the
surface gravity
∂ν lnf = κ . (2.55)
we can see that the surface gravity indeed measures the non affine parameterization.
Suppose now ξ µ vanishes on a spacelike surface Σ. Then near Σ the orbits
of ξ µ will look like Lorentz boosts in Minkowski or accelerations in Rindler. The
acceleration horizon is formed by two intersecting lines, which are null surfaces
generated by null geodesics normal to Σ. These are called bifurcate Killing
horizons and the intersecting point is called bifurcation surface. The Killing vector
is of course null on the bifurcation surface (in four dimensions, it is a two sphere)
S. Conversely, if the group of isometries generated by the Killing vector leaves
a surface S fixed, then the null geodesics orthogonal to S comprise a bifurcate
Killing horizon [112]. In the Penrose diagrams we illustrated before in Kruskal
coordinates, the bifurcation sphere is at the point U = V = 0.
21
The vacuum action for GR is the Einstein-Hilbert action, namely
√
Z
1
IEH = R −g d4 x ,
16π V
where g is the determinant of the metric, R is the Ricci scalar and V is the volume
of the spacetime. Adding matter fields φ and some extra terms results in the full
action:
√
Z I
1 4 1
I[g; φ] = R + 16πLM −g d x + ε (K − K0 )|h|1/2 d3 y , (2.56)
16π V 8π ∂V
where LM (φ, φ,µ ; gµν , gµν,ρ ...) is the matter Lagrangian density and hab is the in-
duced metric on ∂V, the boundary of V, with normal nµ . The action is a scalar
with respect to spacetime coordinate transformations, but behaves as a tensor for
the hypersurface coordinate transformations.
K is the extrinsic curvature of the submanifold ∂V, a geometrical quantity that
depends on the embedding, and is symmetric. We will see below how it is defined
in terms of vectors. This term is added so that after varying the action with
respect to the metric, we get the correct equations of motion, Einstein’s equations.
It cancels out the boundary term that arises when one varies the Einstein Hilbert
term. The reason it is needed is that the Ricci tensor depends on higher (second)
derivatives of the field (the metric). The factor ε = nµ nµ .
K0 is the constant extrinsic curvature of ∂V embedded in flat spacetime, and is
added to cancel the divergence of the action for an asymptotically flat spacetime.
This is a term that is non-dynamical and does not affect the equations of motion.
Specifically, it is equal to the gravitational action of flat spacetime
I
1
If lat = εK|h|1/2 d3 y . (2.57)
8π ∂V
Subtracting this term gives a zero gravitational action for flat spacetime and reg-
ularizes the gravitational action for curved asymptotically flat spacetime.
Defining a stress-energy tensor as
δLM
Tµν = −2 + LM gµν ,
δg µν
we have :
δ(IG + IM )
= 0 =⇒ Gµν = 8πTµν , (2.58)
δg µν
with Gµν being of course the Einstein tensor.
Now, if one wants to quantize a theory, one needs to proceed to the equivalent
Hamiltonian formulation, find canonical commutation relations, so that he can
promote fields to operators etc. The most famous work done in the Hamiltonian
22
formulation of General relativity was by Arnowitt, Deser and Misner (ADM) [23].
They introduced a 3 + 1 decomposition of spacetime, foliating it with spacelike
hypersurfaces Σt . A scalar field t(xµ ) is introduced so that constant t describes a
family of hypersurfaces Σt [15]. Each hypersurface has a coordinate set y a , with
the Roman indices taking values excluding time. The normal to the hypersur-
faces nµ will be parallel to the derivative of t. Let us have a congruence of curves
parametrized by t, intersecting non-orthogonally the hypersurfaces Σt , with tan-
gent vectors υ µ = ∂t xµ . The metric takes the form
where N (t, y a ) is called the lapse funtion, a normalization of the normal on the
hypersurfaces, i.e. nµ = −N ∂µ t, and N a (t, y b ) is the shift three-vector, which
measures how the normal is not parallel to the vector υ µ , i.e. υ µ = N nµ + N a eµa ,
where eµa = ∂a xµ are tangent vectors on Σt . The extrinsic curvature can now be
defined as
Kab = n(µ;ν) eµa eνb K = hab Kab = nµ;µ (2.60)
The relations between the full metric and the three metric are:
∂L
where pab is the conjugate momentum to the three metric field pab = ∂£t hab
, then
p = pii = hab pab , the quantities
√ 1
R0 ≡ − h 3 R + h−1 ( p2 − pab pab ) Ra ≡ −2pab |b
2
and the semicolon ( | ) denotes covariant differentation with respect to the three
metric. We also have the relations between the fields and the extrinsic curvature
√
1 ab h ab
Kab = (£t hab − Na|b − Nb|a ) p = (K − Khab ) . (2.64)
2N 16π
23
The shift and the lapse work as Lagrange multipliers. The graviational Hamiltonian
takes the form
Z
1 −1/2 ab 1 2 1/2 3 1/2 −1/2 ab
HG = Nh (p pab − p ) − N h R − 2Na h (h p )|b d3 y
16π Σt 2
√
I
1 −1/2 ab
− N (k − k0 ) − Na h p rb σd2 θ
8π St
(2.65)
√
I
1
− ab ab
N (k − k0 ) − Na (K − Kh )rb σd2 θ , (2.66)
8π St
where σ is the determinant of the induced metric of the boundary St of Σt and ra is
√
the normal, with the surface one form being dSa = ra σd2 θ. The dot corresponds
to the Lie differentation £t hab = ḣab . We have also defined k = σ IJ kIJ as the trace
of the extrinsic curvature of St embedded in Σt , which is defined as kIJ = rµ;ν eµI eνJ .
The capital roman indices are for the coordinates in the two surface St , similarly
with the lower case indices for the three surfaces Σt . The normal to the two surfaces
µ ∂y a
is rµ and the tangent vectors are the eµI = ∂x∂y a ∂z I
.
We have also added a boundary term like before, that does not affect the
equations of motion, namely the term with the constant k0 . This is the extrinsic
curvature of St in flat space and serves the same purpose: it ensures the action is
zero for flat space and regularizes the action for asymptotically flat space.
The variation of the action in Hamiltonian form will be
Z Z
δIG = dt ab ab ab 0 a
(ḣab − Hab )δp − (ṗ + P )δhab + R δN + 2R δNa d3 y , (2.67)
Σt
where Hab , P ab are given below. Imposing stationarity gives Hamilton’s equations
for GR,
1
ḣab = Hab = 2N h−1/2 (pab − hab p) + Na|b + Nb|a
2
√ 1 1 1
ṗab = −P ab = −N h3 Gab + N hab (pcd pcd − π 2 ) − 2N h−1/2 (pac pcb − ppab )+
√ 2 2 2
|ab ab |m ab c a cb b ca
h(N − h N |m ) + (p N )|c − N |c p − N |c p
R0 = 0 Ra = 0 . (2.68)
24
2.6.1 ADM charges
The first two of equations (2.68) are called evolution equations and the latter two
are the constraint equations, Hamiltonian and momentum respectively. When the
equations of motion are satisfied, we are only left with the boundary term. This
serves as a definition for the mass or the angular momentum of an asymptotically
flat stationary spacetime, quantities that will be named ADM charges [23, 207].
The asymptotic behavior of the spacetime will play a big role in this definition.
Because the variation of the Hamiltonian must vanish when the equations of motion
are satisfied, only the boundary terms will survive the variation. The gravitational
Hamiltonian on shell is
√
I
os 1
HG = − N (k − k0 ) − Na (K − Kh )rb σd2 θ .
ab ab
(2.69)
8π St
Then the choices of our lapse and shift functions will yield different values. In
particular, the ADM mass is defined as
√
I
1
MADM = − lim (k − k0 ) σd2 θ , (2.70)
8π St →∞ St
which came about with the choice of an asymptotic flow vector υ µ = ∂t xµ , or the
choice of N = 1, N a = 0, with St taken to (spatial) infinity as well. The choice
of this vector is of course an asymptotic time translation, so we have a connec-
tion between time translation and a conserved charge - the total energy. Another
equivalent definition of the ADM energy for asymptotically flat spacetimes, given
in the original paper [23] using the metric is
I
1
EADM = (hab,b − h,a )dSa , (2.71)
16π S∞
with hab being the three metric on the spacelike hypersurfaces and h = haa . We
can prove that these definitions are equivalent by using the fact that
Then substituting in 2.70 and integrating by parts, we get 2.71. Taking a choice
for asymptotic rotations, one would expect to get the total angular momentum.
Suppose there is a rotational symmetry around the φ axis. Indeed, with N =
0, N a = ∂φ y a = ma , we get contribution from the other term
√
I
1
JADM = − lim (Kab − Khab )ma rb σd2 θ , (2.73)
8π St →∞ St
25
form. Using equation 2.64, we can express this in terms of the conjugate momenta
I
1
JADM = − pab mb dSa . (2.74)
8π S∞
An equivalent and handy way of defining mass and angular momentum for
axially symmetric stationary spacetimes is using the Komar integrals:
I I I
1 ν µ 1 µ;ν 1
MKomar = − lim ∇ k dSµν = − k dSµν = − ∗dk
8π St →∞ St 8π S∞ 8π S∞
I I I
1 ν µ 1 µ;ν 1
JKomar = lim ∇ m dSµν = m dSµν = ∗dm (2.75)
16π St →∞ St 16π S∞ 16π S∞
√
where dSµν = 2n[µ rν] σd2 θ is the surface element and k µ , mµ are the Killing
vectors generating the time translations and rotations respectively. Using Killing’s
equation it is easily proven they are equivalent to the ADM definitions. Suppose we
have an axially symmetric stationary spacetime in four dimensions (for simplicity
lets consider the Kerr metric 2.28). Then taking the asymptotic limit (for r M )
the terms become (keeping leading and second to leading order terms)
r2 − 2M r 2M
gtt ' − 2
= −1 + (2.76)
r r
2M r 4Jsin2 θ
gtφ ' −2asin2 θ = − (2.77)
r2 r
2 2
gφφ ' r sin θ (2.78)
−1
2M 2M
grr ' 1 − '1+ gθθ ' r2 . (2.79)
r r
So we can rewrite the metric (for convenience we add a small term in the angular
part)
4Jsin2 θ
2 2M 2 2M
dsasympt ' − 1 − dt + 1 + (dr2 + r2 dΩ2 ) − dtdφ . (2.80)
r r r
We will now prove that the Komar expressions are equal to the (M, J) quantities
of the metric. But first lets see how they are equal to the ADM expressions. In
order to do that, we will need the inverse asymptotic metric components. These
will be (see appendix F for the original ones):
−1
r2
tt 2M 2M
g '− 2 = −1+ ' −1 − (2.81)
r − 2M r r r
2J r2 − 2M r 2M
g tφ ' − 3 g rr ' = 1 − (2.82)
r r2 r
2M 1 2M 1
g θθ ' 1− 2
g φφ ' 1 − . (2.83)
r r r r sin2 θ
2
26
We will also need the explicit expression for the normals. We foliate with
spacelike hypsersurfaces Σt , which have a normal
1/2
−∂µ t 2M M
nµ = tt 1/2 ' − 1 + ∂µ t ' − 1 + ∂µ t . (2.84)
|g | r r
Then we can also find the normal to the two sphere St , boundary of Σt :
−1/2
∂a r 2M M
ra = rr 1/2 ' 1− ∂a r ' 1 + ∂a r . (2.85)
|g | r r
Now we are ready to calculate the extrinsic curvatures. For the St sphere, it is
a
k = r|a = hab ra|b = hab ra,b − hab Γcab rc . (2.86)
where r̃ = r(1 + M/r) so that r̃2 ' r2 (1 + 2M/r). Our new extrinsic curvature is
now
2 2 M
k0 = = (1 − ) . (2.96)
r̃ r r
27
Their difference will be
2M
k − k0 = − . (2.97)
r2
We also have the two-metric determinant
σ = r4 sin2 θ . (2.98)
It is straightforward to show that the mass parameter coincides with the ADM
mass:
Z π Z 2π
√ 2
I
1 1 2M
MADM = − (k − k0 ) σd θ = − (− 2 )r2 sin2 θdφdθ = M .
8π S∞ 8π 0 0 r
(2.99)
Now we can prove that the Komar expression is also equal to this quantity. We
have
∇ν k µ nµ rν = k;ν
µ
nµ rν = k,ν
µ
nµ rν − Γµνρ k ρ nµ rσ g νσ . (2.100)
2M M M 2M 2M
− 2Γttr nt rr g rr ' 2
(−1 + )(1 + )(1 − ) ' − 2 = k − k0 . (2.105)
r r r r r
So we have proven that the ADM and Komar definitions are equivalent for the
mass.
Let us see the angular momentum. The second term in 2.73 vanishes since
there is no φ − r component in the metric. As for the first term
M
Kab ma rb = Kφr rr = Kφr g rr rr ' Kφr (1 − ). (2.106)
r
28
We need to calculate Kφr . Using the definition of the extrinsic curvature we have
M
Kφr = nφ;r = nφ,r − Γµφr nµ ' Γtφr (1 − ). (2.107)
r
But
1
Γtφr = g tµ (gµφ,r + gµr,φ − gφr,µ ) . (2.108)
2
The last term vanishes because there is no φ − r component. The second term
vanishes because there is no r − t component as well. We are left with
1 1 2M 4Jsin2 θ 2Jsin2 θ 2M
Γtφr = (g tt gtφ,r + g tφ gφφ,r ) ' − (1 + 2
+ 3
r(1 + ))
2 2 r r r r
3Jsin2 θ −3Jsin2 θ
'− =⇒ K φr ' . (2.109)
r2 r2
Plugging this in the ADM definition 2.73 we get
Z π Z 2π
1
JADM = − (−3Jsin3 θ)dθdφ = J , (2.110)
8π 0 0
where we used the fact that sin3 xdx = −cosx + 13 cos3 x + C. What is finally left
R
is to prove that the Komar definition of angular momentum is also equal to this
parameter. Once again, we start with the Komar integrand
ν µ µ ν µ ν
∇ m nµ rν = m ;ν nµ r = m nµ r − mµ nµ rν;ν − mµ nµ;ν rν . (2.111)
;ν
Putting again the factor of two from the surface element, we end up with the ADM
expression 2.73.
The nice property of the Komar integrals is that they contain Killing vectors.
This way if we use Stokes and eq.(B.2-B.6) we get expressions involving the energy
tensor. Explicitly I Z
∇ν ξ µ dSµν = 2 (∇ν ξ µ );ν dΣµ
S Σ
Z Z
µ
=2 ξ dΣµ = −2 Rµν ξ ν dΣµ . (2.113)
Σ Σ
√
For spacelike hypersurfaces Σ we have dΣµ = −nµ hd3 y. So by virtue of Einstein’s
equations we finally arrive to the results for the charges
Z √
1
MKomar = −2 Tµν − T gµν nµ k ν hd3 y
Σ 2
29
Z √
1
JKomar = Tµν − T gµν nµ mν hd3 y . (2.114)
Σ 2
Again, these hold for stationary spacetimes. We therefore have an expression for
conserved charges involving geometrical quantities and the stress energy tensor.
As for the electric charge, suppose we have a charge distribution on Σ, with
current density j µ . Then the charge will be
Z Z I
µ 1 µν 1
Q= j dΣµ = F ;ν dΣµ = F µν dSµν , (2.115)
Σ 4π Σ 8π S
where we used Maxwell’s equation and Stokes theorem.
30
3 Black hole thermodynamics
The renowned work done by Hawking, Bardeen and Carter [103] linking the
usual thermodynamics to black holes, formulating the ”Black hole mechanics”
seems to be strong indication for the deep connection between quantum physics
and gravity [4].In this project, the first law will be of outmost importance, but we
will also make a sketch for the other laws.
It began when Christodoulou and Ruffini in 1971 [32], when examining the
gedanken Penrose process of extracting energy from black holes, found a mass
formula for the Kerr Newman black hole
2
Q2 J2
2
M = Mirr + + 2
, (3.1)
4Mirr 4Mirr
p
where Mirr = 12 r+ 2
+ a2 is an integration constant called irreducible mass and it
turns out it obeys the inequality
δMirr ≥ 0 , (3.2)
with the equality holding for reversible transformations (changes in the black hole
charges through processes like superradiance etc) and the inequality for irreversible
ones. In order to prove this relation we solve 3.1 with respect to Mirr
1 1
q q
p 2 p
Mirr = 2
M + M −Q −a +a =2 2 2 2M 2 − Q2 + 2M M 2 − Q2 − a2
2 2
(3.3)
and then we vary. We will prove the relation for the Kerr black hole for simplicity
(non-charged), so that we can also do the parallelism with the Penrose process we
showed in chapter 2.
√ √
So we have (r+ = M + M 2 − a2 = M + M1 M 4 − J 2 )
2 2
√
4Mirr = r+ + a2 = 2M 2 + M 4 − J 2 =⇒
1 1
4Mirr δMirr = 2M δM + √ (4M 3 δM − 2JδJ) =
2 M4 − J2
δM √ JδJ 2
δM (r+ + a2 ) − aδJ
√ (2M 2 M 2 − a2 + 2M 3 ) − √ = √
M M 2 − a2 M M 2 − a2 M 2 − a2
2
(δM − Ω+ δJ)(r+ + a2 )
=⇒ δMirr = √ , (3.4)
4Mirr M 2 − a2
a
where we used the facts that J/M = a and Ω+ = r2 +a2 for a Kerr black hole.
+
We can see now where the Penrose process comes in play; we saw in (2.48) that
31
there is a limit in energy extraction from a black hole. The constraint was
δJΩ+ ≤ δM , (3.5)
For solar mass black holes, it is much bigger than the life of the universe.
32
It was proved in [112] that there is a one to one correspondence between the
constancy of the surface gravity and the Killing horizon being part of a bifurcate
Killing horizon, under the condition that all orbits of ξ µ on the horizon H diffeo-
morphic to R and H has a cross section intersected only once by these orbits.
- First law: Any perturbation to a stationary black hole parametrized by (M, J, Q),
with horizon surface gravity κ, horizon angular velocity ΩH , electric potential ΦH ,
and horizon surface A, will obey the equation
κ
dM = dA + ΩH dJ + ΦH dQ , (3.9)
8π
where we put the subscript H implying this holds for any horizon, as we will see in
chapter 6. Equation (3.9) is reminiscent of the first law of thermodynamics, with
energy on the left side, work terms on the right and the entropy being proportional
to the area, something proposed by Bekenstein and Hawking (BH) and called the
BH entropy SBH = A4 . This was motivated by Hawking’s area law (1971), which
is the next one.
- Second law: If the null EC is satisfied, the surface area of the black hole
cannot decrease dA ≥ 0. It is a fact that this area theorem is strikingly similar
to the entropy law in thermodynamics. It was generalized by Bekenstein, what is
known as the Generalized second law (GSL): the entropy of the black hole and the
surrounding matter can never decrease.
- Third law: If the weak EC is satisfied, one cannot reduce the surface of a black
hole to zero in finite steps.
A J2 Q4 Q2
= 64π 2 M 4 − 64π 2 J 2 − 64π 2 M 2 Q2 =⇒ M 2 = − 4π + π + . (3.11)
16π A A 2
A function is homogeneous of order n in a variable x if
33
Then deriving with respect to k and using the chain rule
∂f (kx) ∂(kx) ∂f
nk n−1 f (x) = =⇒ nf (x) = x , (3.13)
∂(kx) ∂k ∂x
where in the last step we set the parameter k = 1. This is easily generalized for
an arbitrary number of variables
∂f
nf (x0 , x1 , ...xr ) = xi , (3.14)
∂xi
where summation is implied. We can see that the mass is homogeneous of degree
1
2
in (J, A, Q2 ) so one gets the simple bilinear form (we will see how the factors
come about below):
κ
M = 2 A + 2ΩH J + ΦH Q . (3.15)
8π
Note that this is not the integration of the differential formula (3.9). In fact there
is a more general expression of the Smarr formula. Let us start from the Komar
expression of the angular momentum. Since we have two boundaries (one at infinity
and one at the horizon) Gauss’ law gives two terms, so we have
Z I I
1 µ;ν 1 µ;ν 1
m dΣµ = m dSµν − mµ;ν dSµν . (3.16)
8π Σ ;ν 16π S∞ 16π H
Hence the total angular momentum is (the integral calculated at the infinity
boundary, i.e. the first term in the right hand side of eq 3.16)
Z I Z
1 1 1
J= µ;ν
m dΣµ + µ;ν
m dSµν = − Rµ mν dΣµ + JH , (3.17)
8π Σ ;ν 16π H 8π Σ ν
where we let JH be the angular momentum evaluated at the horizon. Then manip-
ulating the Komar mass expression, essentially following the exact same procedure
we have
Z Z I
1 µ;ν 1 µ ν 1
M =− k;ν dΣµ + MH = R ν k dΣµ − k µ;ν dSµν =
4π Σ 4π Σ 8π H
Z I
1 µ ν 1
= R ν ξ dΣµ − ξ µ;ν dSµν + 2ΩH J , (3.18)
4π Σ 8π H
where we used the fact that k µ = ξ µ − Ωmµ as well as the previous equation.
√
Now, we said before that dSµν = 2n[µ rν] σd2 θ. On the horizon the normal nµ
can be identified with the Killing vector that generates the horizon, ξ µ , and the
normalization is such that nµ rµ = −1. Then the second term becomes (we will
√
denote σd2 θ = dA since it is the area element):
I I I
1 ν;µ 1 ν;µ 1 1
− ξ ξ[µ rν] dA = − ξµ ξ rν dA = − κξ ν rν dA = κA , (3.19)
4π H 4π H 4π H 4π
34
where on the second step we used Killing’s equation antisymmetry, on the third
step the expression for the surface gravity on the horizon ,and on the last step
the fact that it is constant over the horizon (zeroth law). Then the first term on
eq.(3.18) can be transformed using Einstein’s equation as
Z Z
1 1
R ν ξ dΣµ = 2 (T µν − T δνµ )ξ ν dΣµ .
µ ν
(3.20)
4π Σ Σ 2
So the generalized Smarr formula becomes
Z
κA 1
M= + 2ΩH J + 2 (T µν − T δ µν )ξ ν dΣµ . (3.21)
4π Σ 2
Of course, for an isolated black hole, one has M = MH , J = JH since the total
charges of the spacetime will be equal to the black hole charges, so the Smarr
formula quantities are those of the black hole. In order to obtain (3.15) in four
dimensional electrovacuum, one takes the energy momentum tensor of the Maxwell
theory, namely
1
F µρ F νρ − g µν Fρσ F ρσ , (3.22)
4
which is traceless
1
4πT = 4πgµν T µν = F µρ Fµρ − 4Fρσ F ρσ = 0 , (3.23)
4
where the strength of the field Fµν = ∂µ Āν − ∂ν µ = Āν,µ − µ,ν (or F = dĀ
in differential geometry language) is antisymmetric and traceless and obeys the
Bianchi identity. The quantity µ is the gauge 1-form of electromagnetism (we
are using a bar to distinguish it from the horizon area A).
Now using the fact that the Lie derivative of the gauge potential is
£ξ µ = ξ ν ∂ν µ + Āν ∂µ ξ ν = ξ ν Fνµ + ξ ν ∂µ Āν + Āν ∂µ ξ ν = ξ ν Fνµ + (ξ ν Āν ),µ . (3.24)
Since the Killing vector field is also a symmetry for the electromagnetic field we
have
£ξ µ = 0 (3.25)
There is also a quantity that is proven to be constant over the horizon, like the
surface gravity, associated with the electromagnetic field, namely
ΦH = −(µ ξ µ ) , (3.26)
H
which is the electric potential on the horizon. The potential is of course defined
everywhere, and we will assume it is zero at infinity, as well as the gauge potential
35
µ . Hence, equation (3.24) becomes
Then using the fact that E µ is proportional to ξ µ on the horizon (since ξ µ ξ ν Fνµ = 0
due to the strength’s antisymmetry) with a proportionality σE , we can attempt to
Lie transport the electric potential along the horizon, with a tangent vector field
tµ as
H H H
£t Φ = tµ Φ,µ = tµ Eµ = σE tµ ξµ = 0 =⇒ Φ = const. (3.30)
(3.31)
(ξ ρ F µν Āν − ξ µ F ρν Āν );ρ + ξ µ F ρν Āν;ρ = 0 =⇒
where we used Killing’s equation and Maxwell’s equations F µν;µ = 0. They are
source free since the charge is in the singularity, while we integrate from the horizon
to infinity. Then substituting the energy tensor we have
Z Z
µ ν 1 µρ 1 µ 2 ν
2 T ν ξ dΣµ = F Fνρ + δ ν F ξ dΣµ =
Σ 2π Σ 4
Z Z
1 µρ 1
F Φ;ρ dΣµ − ξ µ F 2 dΣµ =
2π Σ 8π Σ
36
Z Z
1 µρ 1
(ξ ρ F µν Āν − ξ µ F ρν Āν );ρ dΣµ .
− F Φ ;ρ dΣµ + (3.33)
2π Σ 4π Σ
Then we can use Gauss law with two boundaries, since both tensors in the integrals
are antisymmetric. The first term will be
Z I I
1 µρ Φ∞ µρ ΦH
F µρ dSµρ = −ΦH Q ,
F Φ ;ρ dΣµ = F dSµρ − (3.34)
2π Σ 4π S∞ 4π H
since the electric potentials are constant at infinity (zero) and at the horizon. As
for the second term, we use the convention that the gauge potential vanishes at
infinity so only the horizon boundary contributes
Z I
1 ρ µν µ ρν 1
(ξ F Āν − ξ F Āν );ρ dΣµ = ξ [ρ F µ]ν Āν dSµρ = 2ΦH Q , (3.35)
4π Σ 4π H
for which we used the fact that E µ = F µν ξν is proportional to ξ µ on the horizon and
the null condition ξ 2 = 0. So for the Einstein Maxwell theory, one finally arrives to
the required result (3.15). If there is external matter or current contribution [105]
one obtains
κA
M − 2ΩH J − ΦH Q − = Mmatter − 2ΩH Jmatter − ΦH Qcurrent . (3.36)
4π
The original first law by Bardeen, Carter and Hawking [103] was a slightly dif-
ferent statement: two neighboring solutions were considered, both stationary and
axisymmetric, and the variation formula was describing the difference of these two
solutions. Specifically, the way it is stated is that if there are two stationary solu-
tions (M, g) and (M+δM, g+δg) then the difference in masses, angular momenta,
areas and charges is eq. (3.9).
The proof begins by assuming that the Killing fields are conserved, they remain
unchanged from one solution to the other i.e.
δk µ = δmµ = 0 . (3.37)
Then denoting the metric variation δgµν = hµν , the Killing vector generating the
horizon will obey
δξ µ = mµ δΩH =⇒ δξµ = δ(gµν ξ ν ) =
37
Varying the formula (3.21) yields
Z
1 1
δM = (δκA + κδA) + 2(ΩH δJ + δΩH J) + δ Rµν ξ ν dΣµ , (3.39)
4π 4π Σ
where we reinserted the Ricci tensor for simplicity. Let’s start by manipulating
the first term. Following Hawking-Carter-Bardeen’s paper [103], we have the Lie
derivatives vanishing
£ξ δξν = ξ µ δξν;µ + δξµ ξ µ;ν = 0 (3.40)
since as we said the Killing fields remain unchanged, and hence these are essen-
tially Lie transporting the Killing vectors along their orbits. Afterwards, using the
definition for the surface gravity κ = 21 (ξ 2 ),µ nµ and varying (we are dropping the
indication that equations hold on the horizon but it is implied) we get
1 1
δκ = (ξ 2 ),ν δnν + nν (ξ µ δξµ + ξµ δξ µ ),ν =
2 2
1
δnν ξ µ;ν ξµ + nν ξ µ;ν δξµ + (nµ ξ ν + nν ξ µ )δξµ;ν + δξ µ;ν ξµ nν
2
1 ν
+ n (ξµ;ν δξ µ − ξµ δξ µ;ν − ξ µ;ν δξµ − ξ µ δξν;µ ) . (3.42)
2
The first two terms in the last line vanish because of (3.41) and Killing’s equation
(in order to write ξµ;ν δξ µ = −ξν;µ δξ µ ) and the last two terms vanish because of
(3.40). For the first two terms in the second line, we use a null tetrad [103],also
known as the Newman-Penrose formalism, on the horizon to express the metric as
where ηµ , η̄µ are complex conjugate null vectors lying on the horizon, and they
are orthogonal to the two normals. The convergence and shear of the horizon
generators vanish [103], i.e.
and the same will apply for the new Killing vector δξ µ , which on the horizon is
proportional to the old one. The first one of (3.44) vanishes since if there was a
convergence, there would be a caustic (singular point) at the horizon in finite affine
distance for the null generator. The second one vanishes due to the fact that ξµ;ν
is antisymmetric. Furthermore, on the horizon we can decompose the derivative
of the Killing vector as [106]
38
which comes after doing a decomposition of the derivative on the horizon with
respect to the null tetrad, and taking into account Frobenius’ theorem, Killing’s
equation and the surface gravity expression. Then we can express these two terms
as
ξ µ;ν (δξµ nν + ξµ δnν ) = ξµ;ν ξ µ δη ν , (3.46)
1 1 dSµν
δΩH mµ;ν ξµ nν = δΩH (mµ;ν ξµ nν + mν;µ ξν nµ ) = − δΩH mµ;ν . (3.49)
2 2 dA
As for the other term, exploiting the orthogonality and normalization conditions
ξ 2 = 0, ξµ nν = −1, one has
1 1 1 1 1
− hµν ;µ ξ ν = − ξµ hµλ;λ = ξµ nν ξ ν hµλ;λ − ξν ξ ν nµ hµλ;λ = − (nµ ξν −nν ξµ )ξ ν hµλ;λ =
2 2 2 2 2
1 1
(nµ ξν − nν ξµ )ξ µ hν λ;λ − (nµ ξν − nν ξµ )ξ µ hλλ;ν , (3.50)
2 2
where we added a term that is identically zero due to £ξ h = ξ ν h;ν = 0 and ξ 2 = 0.
Moving on this is equal to
ν] ν] dSµν
(nµ ξν − nν ξµ )ξ µ ∇[λ hλ = ξ µ ∇[λ hλ . (3.51)
dA
Plugging these back to eq (3.47)
1 ν]
δκdA = dSµν (2ξ µ ∇[λ h λ − mµ;ν δΩH ) . (3.52)
2
Integrating over the horizon and applying the zeroth law to take surface gravity
39
out of the integral we get
I
ν]
Aδκ = −8πJH δΩH + ξ µ ∇[λ h λ dSµν . (3.53)
H
Then using the fact that the asymptotic metric variation is related to the mass
variation by [105] (see 2.80)
δM
hµν = 2 δµν + O(1/r2 ) , (3.54)
r
we have
I I
µ ν]
dSµν k ∇[λ h λ = −δM k µ (nν r,ν rµ + 2nµ )dA = 4πδM . (3.55)
S∞ S∞
Then, postulating that the horizon surface is invariant under rotations, i.e. mν dSµν =
0, we have using Gauss’ law
I I Z
µ [λ ν] µ [λ ν]
ξ ∇ h λ dSµν = k ∇ h λ dSµν = k µ (∇[ρ hν]ν );ρ dΣµ − 4πδM . (3.56)
H H Σ
Before proving this, lets see what the left side is exactly:
2h[µ;ν] µ;ν
µ ;ν = hµ ;ν − hν;µ
µ ;ν = gµν δg µν − ∇µ ∇ν δg µν , (3.60)
where we can pull out the metric tensor on the first term of the last step since it
is compatible. We can already see where this is going; the left side of eq.(3.61)
is the variation of the Einstein Hilbert action and the right side is the Einstein
tensor, which should vanish when the equations of motion are satisfied (looking at
the asymptotic metric we can see that it is a vacuum solution). We will prove that
their sum is equal to the variation of the Einstein Hilbert action.
40
In order to prove (3.61) we start from
√ √
Z Z
4
SEH = −gRd x = −gRµν g µν d4 x . (3.62)
As for the second term of (3.63), one needs to vary the Riemann tensor since
Ricci comes from its contraction. The variation of the Riemann tensor turns out
to be [209]
δRρµλν = ∇λ δΓρµν − ∇ν δΓρµλ , (3.68)
which is precisely the last term of (3.61). So (3.59) follows from the variation of
the Einstein Hilbert action.
One also has
1 √ 1
δ(dΣµ ) = √ (δ −g)dΣµ = − h dΣµ (3.70)
−g 2
with h = hµµ . Again, we can denote this also as δ(∗k) = (−g)−1/2 δ(−g)1/2 ∗ k.
Using this identity, we also have (−g)−1/2 δ(R(−g)1/2 ) ∗ k = δ(R ∗ k). Hence we
41
acquire the result
Z Z
1 µν 1 ρ
Aδκ + 8πJH δΩH + 4πδM = − G hµν k dΣρ − δ Rk µ dΣµ (3.71)
2 Σ 2 Σ
or equivalently
Z Z
1 µν 1
Aδκ + 8πJH δΩH + 4πδM = − G hµν ∗k− δ R∗k . (3.72)
2 Σ 2 Σ
Then we can finally plug this back to (3.39) and make the simplifications
Z Z Z
1 1 µν 1 1
δM = κδA+ΩH δJ − G hµν ∗k − δ R∗k + δ ∗R(k) , (3.73)
8π 16π Σ 16π Σ 8π Σ
with G(k) = R(k) − 12 Rk. This is the most general form of the first law for
stationary black holes in Einstein gravity. Having then a specific matter field and
a metric it changes accordingly will the appropriate substitutions be applied. For
an Einstein-Maxwell theory, one has
κ
δM = δA + ΩH δJ + ΦδQ . (3.75)
8π
An even stronger version of the first law was derived by Wald [111], one that holds
for arbitrary n-dimensional gravitational theories with a diffeomorphism invariant
Lagrangian. The main innovation is treating entropy as a local Noether charge
on the horizon, associated with the Killing vector that generates the horizon, nor-
malized as to have unit surface gravity. The horizon also has to be bifurcate. As
he mentions, the strength of this derivation lies at the fact that this is a local
geometrical quantity. We will briefly present the idea in the following paragraphs.
Suppose we have a diffeomorphism Lagrangian L on a manifold M, treating
it as an n-form instead of a scalar density, that depends on a metric and a finite
number of its derivatives (assuming there is a derivative operator ∇µ ), as well as
other matter fields with their respective derivatives. Diffeomorphism invariance
means that
for ψ : M → M, L[ψ ∗ (φ)] = ψ ∗ L[φ] , (3.76)
where φ are all the dynamical fields of the theory, and ψ ∗ is applied to them (star
denotes the cotangent bundle). The next step is varying the Lagrangian. This
will yield a term that vanishes when the equations of motion are satisfied and a
42
surface term, which will be an (n-1)-form in our case (summation over the fields is
implied)
δL = Eδφ + dθ . (3.77)
j =θ−ξ·L. (3.79)
δj = δθ − ξ · δL = δθ − ξ · (Eδφ + dθ) =
where in the last step we imposed the equations of motion and denoted the quantity
ω = δθ − £ξ θ, which is called the presymplectic current and is a local function
of the fields, their derivatives and two linearized perturbations (in this case one of
them is the Lie derivative). Foliating the spacetime by spacelike hypersurfaces Σ
as slices of ”time” we can define the presymplectic form
Z
ΩΣ = ω. (3.82)
Σ
43
a surface term
where δ1 , δ2 are the two linearized perturbations. As shown in [114], one can define
the Hamiltonian conjugate of a vector field ξ µ if its variation converges on the
constraint submanifold, and its variation is
Z
δHξ = ΩΣ (φ, δφ, £ξ φ) = ω(φ, δφ, £ξ φ) . (3.84)
Σ
In [114], they also explore whether this quantity exists, and if not, how one can
proceed. Leaving that aside, we can now integrate equation (3.81) over the hyper-
surfaces Σ Z Z
δ j = δHξ + δ d(ξ · θ) . (3.85)
Σ Σ
If we had a compact support, the second term would not exist and the Noether
current would be purely a Hamiltonian density [108]. Proceeding, if the linearized
equations of motion are satisfied from the fields δφ the Hamiltonian becomes a
surface term, since we can substitute the current as a derivative of the Noether
charge. We can do the same trick as before and define the total or canonical
energy and angular momentum of the spacetime as the surface contributions of the
Hamiltonian at infinity, by setting the corresponding vector as the time translation
k µ or rotation vector mµ . So their variations after using Stokes will be
Z
δEtot = (δQ(k) − k · θ) (3.86)
∞
Z
δJtot = − δQ(m) (3.87)
∞
and then the corresponding ADM mass and angular momentum can be defined
accordingly. Note that we take mµ to be tangent to the sphere at infinity.
If now we have an extra boundary, i.e. a stationary black hole with a bifurcate
Killing horizon and a bifurcation (n-2)-surface S, there exists a Killing vector that
generates the horizon and vanishes on S, namely ξ µ = k µ + ΩaH mµa where a counts
the number of independent rotations and summation is implied. Since ξ µ generates
an isometry group, we have £ξ φ = 0 and hence
Z Z Z Z
δ j = δ d(ξ · θ) =⇒ δ dQ = δ d(ξ · θ) (3.88)
Σ Σ Σ Σ
44
and then using Stokes with two boundaries
Z Z Z Z
δ Q−δ Q=δ ξ·θ− ξ·θ . (3.89)
∞ S ∞ S
But the last term on the right hand side vanishes since it is a bifurcation surface.
Also, the first term on the left can be analyzed as
Z Z Z Z
a
δ Q(ξ) = δ Q(k) + ΩH δ Qa (m) = δEtot + δ k · θ − ΩaH δJtot
a
. (3.90)
∞ ∞ ∞ ∞
So finally plugging it back in (3.88) and taking into account that mµ is tangent to
the sphere at infinity, Z
a
δE − ΩH δJa = δ Q , (3.91)
S
where now we have on the right hand side as a local quantity that depends on
the Killing fields, their derivatives and the fields of the theory. We can although
express the derivatives of the Killing fields in terms of geometrical objects such as
the binormal to S [111]. Then we construct a Killing vector normalized to have
unit surface gravity as ξˆµ = κ−1 ξ µ . Then we assume we have a perturbation from
a stationary black hole to another black hole (which can be non-stationary under
some conditions) with the Killing vectors staying the same close to the horizon, as
in the original first law derivation. Then the entropy can just be defined as the
Noether charge associated to ξˆµ
Z Z
−1
S = 2π Q̂ = 2πκ Q (3.92)
S S
45
isfies 3.76) that takes the form
L = L gµν , ∇µ1 Rνρστ , . . . , ∇(µ1 ... ∇µm ) Rνρστ , φ, ∇µ1 φ, ∇(µ1 ... ∇µl ) , (3.95)
with
(X ρσ )ρ1 ...ρn = E µνρσ µνρ1 ...ρn (3.97)
∂L ∂L ∂L
E µνρσ = −∇µ1 + . . . +(−1)∇(µ1 ... ∇µm ) ,
∂Rµνρσ ∂∇µ1 Rµνρσ ∂∇(µ1 ... ∇µm ) Rµνρσ
(3.98)
which are the equations of motion, and then as argued in [121], the ambiguities
that were in θ before, go on and get transferred to Q in the Y and Z quantities.
Then specializing for an Einstein theory in four dimensions, where the Lagrangian
as a four-form will be
1
Lµνρσ = µνρσ R , (3.99)
16π
the Noether current is obtained as
1
j µνρ = − µνρσ ∇τ ∇[τ ξ σ] (3.100)
8π
and the Noether charge
1
Qµν = −µνρσ ∇ρ ξ σ (3.101)
16π
and then go on by giving a two dimensional example, as well as proving the first
law for not necessarily stationary perturbations with the entropy given by
Z
S = 2π X ρσ µν , (3.102)
S
where the integral is taken on the bifurcation (n-2)-surface and is the binormal
to the surface. We will not though provide more details on this derivation. Finally,
the dynamical entropy formula is proposed, which means that the black hole does
not have to be stationary, i.e. it can depend on time- hence the dynamical. Some
months later, in [122], it was proven that Iyer-Wald’s method for computing the
Noether charge does not need to be evaluated on a bifurcation surface, but it can
be equivalently evaluated on an arbitrary cross section C of the horizon. Albeit,
46
they state that the entropy will be given by
Z
S = −2π (Rµνρσ − ∇τ P τ µνρσ )µν ρσ , (3.103)
C
where
∂L ∂L
Rµνρσ = P τ µνρσ = (3.104)
∂Rµνρσ ∂∇τ Rµνρσ
and L = L, so they do not include higher order derivatives of the curvature,
although that does not mean it is impossible to add them.
47
4 Cosmological and diverse dimensional gravity
In this chapter we will take a look in gravity beyond the four dimensional asymp-
totically flat case, which was well studied up until the 80’s. There has been a great
interest in (Anti)deSitter black holes, especially after the AdS/CFT correspondence
conjecture by Juan Maldacena [24], one of the most used tools of contemporary
physics, as well as lower and higher dimensional gravity.
with τ ∈ [0, 2π), ρ ≥ 0 and i Ω2i = 1. The metric becomes then in these global
P
coordinates:
ds2 = l2 (−cosh2 ρ dτ 2 + dρ2 + sinh2 ρ dΩ2D−2 ) . (4.3)
The isometry group is SO(2, D − 1). Proceeding further, one can perform the
transformation
tan θ = cosh ρ with θ ∈ [0, π/2)
48
resulting in a metric
l2
ds2 = (−dτ 2 + dθ2 + sin2 θ dΩ2D−2 ) , (4.4)
cos2 θ
which is coformally equivalent to R × S D−1 . Although there is a caveat; not all
values of θ are covered. The equator of S D−1 is a boundary. One needs to specify
boundary conditions to solve the initial value problem. This conformal equivalence
was one of the motivations for the AdS/CFT correspondence.
4.1.1 Schwarzschild-AdS
Lets restrict to the four dimensional case to take a look in black holes and how
they behave differently. In order to make a good comparison of empty space with
one containing a black hole, lets make a final coordinate transformation t = l τ, r =
l sinh ρ to have the empty AdS4 metric:
−1
r2 r2
2 2
dsAdS4 = − 1 + 2 dt + 1 + 2 dr2 + r2 dΩ2 =
l l
4.1.2 Kerr-AdS
Asymptotically AdS rotating black holes were obtained first by Carter [27]. The
metric reads:
2 2
a sin2 θ ρ2 2 ρ2 2 ∆θ sin2 θ r 2 + a2 2
2 ∆r
ds = − 2 dt− dφ + dr + dθ + a dt− dφ ,
ρ Ξ ∆r ∆θ ρ2 Ξ
where
r2 a2
∆r = (r2 + a2 )(1 + ) − 2M r Ξ=1−
l2 l2
a2 2
∆θ = 1 − cos θ ρ2 = r2 + a2 cos2 θ .
l2
49
The solution admits four horizons, two real and two imaginary, as roots of the
equation ∆r = 0. Lets call the largest root r+ . In order to avoid naked singularities,
we must have a < l [28]. More details on Kerr-AdS are in the appendix A.
Note that the rotational velocity measured by zero angular momentum ob-
servers (ZAMO) at infinity has to be corrected by their non-zero velocity, since we
are in AdS space, so their angular velocity is non-vanishing asymptotically. This
later will be important for the first law [31]. An analysis on the dS black holes is
made in [120].
50
The nice feature of this action is that it can be expressed as a pure gauge action
Z
κ 2
I= tr(A ∧ dA + A ∧ A ∧ A) , (4.8)
2π 3
known as the Chern-Simons action for a gauge field A. We will not go in further
detail about this though.
In the ADM formalism the metric of the rotating BTZ reads
2
ds2 = −N 2 dt2 + N −2 dr2 + r2 dφ + N φ dt , (4.9)
r2 J2 J
N 2 = −M + + Nφ = − , (4.10)
l2 4r2 2r2
where now we have set the cosmological constant to be Λ = −1/l2 . Currently there
is no black hole solution for Λ > 0 in Einstein gravity and a positive cosmological
constant quantum gravity is thought to maybe not exist nonperturbatively [210].
As for the case of Λ = 0, black holes do not exist, since as we said in three
dimensions there are no gravitational waves.
There are two horizons for the rotating case at
M l2
J 1/2
r± = 1± 1−
2 Ml
and therefore the expressions for the mass and angular momentum are
2 2
r+ + r− 2r+ r−
M= J= ,
l2 l
which are the ADM charges, defined by flux integrals at infinity (we can also use
Komar integrals, modified for AdS space, see [120]). A very interesting feature is
the mass gap that appears in the spectrum. The chargeless state M = J = 0 gives
the ”massless BTZ” geometry,
r2 2 l2 2
ds2 = − dt + 2 dr + r2 dφ2
l2 r
and then all negative masses leave a naked conical singularity (and hence they
are not allowed) at r = 0 except the M = −1, J = 0 case (vacuum); there the
singularity disappears, there is no horizon and we get back empty AdS.
By finding the free energy from the Euclidean action one can find the entropy
and then the temperature of BTZ. The resulting entropy is a quarter of the perime-
ter of the horizon, and the temperature is
κ+ r2 − r−
2
TBT Z = = + .
2π 2πr+
51
This means that at M = 0, there is no temperature or entropy, but for the extreme
rotating BTZ (at J = M l and r+ = r− ) there is non zero entropy like the four
dimensional case. The killing vector generating the horizon is again ξ µ = ∂t +ΩH ∂φ
and its norm is
r2 J2
ξ 2 = −N 2 = M − 2 − 2 , (4.11)
l 4r
which is obviously timelike from infinity to the horizon. As for the ergosurface, it
is located at
gtt = 0 =⇒ r = ±M 1/2 l . (4.12)
4.3.1 Non-rotating
Tangherlini
52
action
√
Z
1
I= dD x −g(R − 2Λ) . (4.13)
16π
It’s horizon topology is S D−2 , when it is asymptoticaly flat, similar to the four
dimensional case. The metric is
m 16πM
N2 = 1 − m= , (4.14)
rD−3 (D − 2)ÃD−2
where M is the ADM mass and ÃD is the D-dimensional unit sphere area. The
horizon is located at r = m1/(D−3) . If we add N − D flat dimensions to a flat
Tangherlini metric, the full metric is Ricci flat, so it is still a solution to Einstein’s
equations. One can also add the cosmological constant to have a Tangherlini-
Schwarzschild-(A)dS ,when
m 2Λ
N2 = 1 − − r2 .
rD−3 (D − 1)(D − 2)
Higher dimensional RN
One can add charge to the Schwarzschild-Tangherlini with now the black hole being
a solution to the Einstein-Maxwell theory with
m q2 8πQ2
N2 = 1 − + , q2 = ,
rD−3 r2 D − 3 (D − 2)(D − 3)
1/2
1 D−2 D
A= dt (4.15)
8π D − 3 rD−3
with a horizon (for q 2 < m2 /4) at rD−3 = 21 m − ( 14 m2 − q 2 )1/2 . It also seems that
the RN black hole cannot be a source of both electric and magnetic fields in D
dimensions, and that four dimensions is a coincidence [43].
To my knowledge, no higher dimensional Kerr-Newman solution has been found
so far, because the Kerr-Schild ansatz does not yield a solution.
4.3.2 Rotating
Myers-Perry
The rotating cases are the most interesting, since we can have rotations in differ-
ent planes. We will start by presenting the D-dimensional case and give the five
dimensional example.
Using the Kerr-Schild ansatz, Myers and Perry [43] [44] generated a rotat-
ing black solution for odd and even D-dimensional spacetimes. For simplicity,
53
lets make a redefinition of the spacetime dimensions D = N + 1 and work with
N + 1 dimensions instead. The symmetry group of N + 1 dimensional spacetime
is translations and SO(N, 1) Lorentz group. The stable black hole will be char-
acterized by the mass and N/2 angular momenta (this later was proven to not
be the only case, see the black ring discussion below). For even D (or odd N ), in
Boyer-Lindquist-like coordinates the metric is
mr ΠF
ds2 = −dt2 + (r2 + a2i )(dµ2i + µ2i dφ2i ) + (dt + ai µ2i dφi )2 + dr2 , (4.16)
ΠF Π − mr
where
(N −1)/2
a2 µ 2 Y
F = 1 − 2i i 2 Π= (r2 + a2i ) .
r + ai i=1
The indices for the angular variables run from 1 to (N − 1)/2 in general (so for
P 2
the even case they run up to ((N − 2)/2) and there is the constraint µi = 1.
Some modifications are needed for even N . Except the correct range of indices, the
P 2
constraint changes to µi + α2 = 1, with α being a separate angular coordinate
and the metric becomes:
mr ΠF
ds2 = −dt2 + r2 dα2 + (r2 + a2i )(dµ2i + µ2i dφ2i ) +
(dt + ai µ2i dφi )2 + dr2 .
ΠF Π − mr2
(4.17)
The mass parameter is the same as the non rotating case (with D replaced by
N + 1)
16πM
m= . (4.18)
(N − 1)ÃN −1
The angular momenta are J i = N2−1 M ai .
The equation that determines the horizons is :
For odd N , in general, there are no analytical solutions for the horizons, except
the cases N = 3, 5, when all spin paramters are not zero, by Galois’ theorem that
roots of an n-th order polynomial are only analytically soluble for these odd powers.
For instance, for N = 7 one has the horizon equation
which is obviously not analytically solvable. For even N, again there are no ana-
lytic solutions except the cases 4, 6, 8. The topology of the horizons is S N −1 . Also,
a sufficient condition for the existence of a horizon is the vanishing of a spin pa-
rameter ai [43], while for the case that all spin parameters are non-zero there are
bounds for naked singularities.
54
A very significant solution (since we will use it for crucial relations in this
project) is the five dimensional one (N = 4) with horizons at
p
2r± = m − a2 − b2 ± (m − a2 − b2 )2 − 4a2 b2 ,
where a, b are the spin parameters, and the singularity bound is m ≥ a2 +b2 +2|ab|.
In order to find the asymptotic charges, one can use Komar integrals generalized
for N + 1 dimensional spacetime, with an asymptotically timelike Killing vector
k µ and the N/2 spacelike rotation Killing vectors mµ ,
1 N −1
I I
µ ν 1
M =− lim ∇ k dSµν Ji = − lim ∇µ mνi dSµν ,
16π N − 2 St →∞ St 16π St →∞ St
(4.20)
with St being an N −1 sphere at spacial infinity (for more details see appendix A.4).
Then using Stokes and equation (B.6) as before we can transform these into horizon
valued quantities and the Ricci tensor evaluated on a spacelike hypersurface
Z
1 N −1
I
µ ν µ ν
M =− 2 Rν k dΣµ + ∇ k dSµν
16π N − 2 Σ H
Z I
1 µ ν µ ν
Ji = − 2 Rν mi dΣµ + ∇ mi dSµν . (4.21)
16π Σ H
The generator of the horizon is the killing vector ξ µ = k µ −Ωi mµi , with a summation
over i implied. Combining (4.21) and defining a surface gravity from the equation
κξ ν = ξ µ ∇µ ξ ν ,
∂r Π − 2mr ∂r Π − m
even N : κ= odd N : κ= ,
2mr2 rH 2mr rH
ÃN −1 m a2
A± = (N − 2 − 2 2 i 2 )
2κ± r ± + ai
and again ÃD is the D dimensional unit sphere area. As for the angular velocities
Ωi , they are simply the generalization from the four dimensional case and i numbers
the spin parameter, i.e.
ai
Ωi± = 2 . (4.23)
r± + a2i
55
In appendix A.4, where the five dimensional case is previewed, we denote the two
angles with φ, ψ, and one can notice that because of the relations between the spin
parameters a, b and the horizon, one can express the angular velocities as
a ar∓
Ωφ± = 2
Ωψ± = . (4.24)
r± + a2 r± (r ∓2 +a2 )
In D > 4 dimensions are objects with non spherical topology called black rings,
with horizon topology S 1 ×S D−3 . Found by Emparan and Reall [48] [49], they were
the shining example of non uniqueness in more than four spacetime dimensions.
Emparan and Reall proved that for a certain range in angular momentum and
mass, there exist one black hole and two black ring solutions(one with larger area-
entropy) in five dimensions.
The spectacular observation is that not only these solutions exist, but they
are preferred states above certain spin values. There is a phase transition from a
black hole to a black ring with larger area for fixed values of mass and angular
momentum. However, it is not clear if a black hole can evolve into a black ring,
changing its horizon topology. Regarding the bounds of angular momentum, for
black rings in five dimensions there is a lower bound and, like Myers-Perry in six
or more dimensions, they can have arbitrarily big spin. So an ultraspinning black
object in five dimensions can only be a black ring.
56
The metric reads
2
R2 G(y) 2 dy 2 dx2 G(x) 2
2 F (y) 1+y
ds = − dt−CR dψ + F (x) − dψ − + + dφ ,
F (x) F (y) (x − y)2 F (y) G(y) G(x) F (x)
where
r
2 1+λ
F (x) = 1 + λx G(x) = (1 − x )(1 + νx) C= λ(λ − ν)
1−λ
with the coordinates in the ranges −∞ ≤ y ≤ −1 and −1 ≤ x ≤ 1, and the
dimensionless parameters 0 < ν ≤ λ < 1. Asymptotic infinity is at x = y = −1
and the horizon lies at y = −1/ν [40]. In order to avoid the conical defect at the
plane of the ring, one has to identify
p √
F (−1) 1−λ
∆ψ = ∆φ = 2π 0 = 2π (4.28)
G (−1) 1−ν
57
singularity at ξ = 0 we have to identify φ with a period
p
[(G2 (x))0 ]2 F (−1)
∆φ = 2π =⇒ ∆φ = 2π 0 (4.33)
4G2 (x)F (x) x=−1 |G (−1)|
and similarly with ψ. Now we can also prove that this metric is asymptotically
flat. The asymptotic form of the metric is (plugging values for x = y = −1
√
where possible, and also making the redefinitions φ̃ = φ(1 − ν)/ 1 − λ , ψ̃ =
√ p p
ψ(1 − ν) 1 − λ and R̃ = R (1 − λ)/ (1 − ν)):
R̃2 dψ 2 dx2
ds2as 2
∼ −dt + 2 2
(y − 1)dψ̃ + 2 + 2 2
+ (1 − x )dφ̃ . (4.34)
(x − y)2 y − 1 1 − x2
performing a transformation
p √
R̃ y2 − 1 R̃ 1 − x2
ξ= η= , (4.35)
x−y x−y
we end up in the flat form of the metric
The five dimensional black ring also satisfies the same Smarr relation with
Myers-Perry
3 κA
M= + ΩH J
2 8π
with the interesting relation between κ and M ,
3πR
κ= .
8M
One can also take a particular limit (R → 0, λ, ν → 1 and fix the parameters
a, m [49]) and perform a coordinate transformation to retrieve the Myers-Perry five
dimensional black hole with rotation in one plane. These rings have a rotation along
the S 1 . Solutions with rotation only along the S 2 were constructed [70] [71] but
there is no balancing condition like (4.28,4.29), so they have conical singularities
[40].
Some years later Pomeransky and Sen’kov [62] generated an asymptotically flat
black ring solution with rotation in both along the S 1 and S 2 plane (details on the
metric are in appendix A.15). This doubly spinning black ring has two horizons,
the roots of the equation
1 + λy + νy 2 = 0 , (4.37)
58
which in turn gives constraints for the parameters if one imposes the reality of the
metric, existence of the horizons and positivity of the mass. The extremal black
√
ring is retrieved when λ = 2 ν and the black ring with one angular momentum
when ν = 0. The doubly spinning black ring has a bound on the rotation in the
√
S 2 plane, namely λ ≥ 2 ν. One can also get the extremal Myers Perry as a limit
of the extremal black ring with the additional limit ν → 1, λ → 2. Spatial infinity
is again at x = y = −1.
√
Z
1 5 1 2 1 −αφ µν
I= d x −g R − (∂φ) − e Fµν F , (4.38)
16π 2 4
where φ is the dilaton with coupling constant α and F is the Maxwell field. The
dilaton coupling can be expressed as
4 4
α2 = − ,
N 3
with the values N = 1, 2, 3 having different relevance for string theory. We focus
on the zero dilaton coupling theory N = 3 (specifics of the metric are in appendix
A.16). One however need not have an EMD action for a dipole ring; for instance,
one can start with a minimal supergravity action [213].
For a fixed dipole charge, again there are two rings with different entropies. The
difference with the neutral black ring is that there is an upper bound in angular
momentum, which is acquired at extremality. The solution has two horizons, an
event horizon at y = −1/ν and an inner horizon at y = −∞. The extremal limit
is ν = 0. The dipole charge also enters the Smarr formula and the first law (see
appendix A.16). Regarding this fact, there is a derivation [116] similar to the
one done by Wald (Hamiltonian method) and a nice discussion because there is a
contradiction on the fact that dipole charges arise from the surface integrals. They
also prove the first law for black rings in various theories, such as supergravity.
59
Multiple black objects
Unlike four dimensions, one can have multiple black object stationary systems in
higher dimensions. Some examples are black saturns [211], which are Myers-Perry
black holes surrounded by black rings and black bi-rings or bicycle rings [212]
rotating on orthogonal planes.
Making a coordinate transformation it can be seen that the solution has two Killing
horizons at
√
r± = M ± M 2 + n2 (4.40)
associated with the time translation Killing vector. The region r < r− and r > r+
is stationary but the Taub region r− < r < r+ is time dependent. Null geodesics
become twisted close to the horizons and both regions are geodesically incomplete.
The Kerr-NUT generalization was made [128] in 1966 with explicit metric solu-
tions by Frolov [131] and then the more general Kerr-Newman-Taub-NUT-(A)dS
(KNTN-(A)dS) [129].
We present the KNTN solution in the appendix, and use it to check if our
geometric relations hold and how. One can see there that the NUT charge enters
the first law and the Smarr relation. Kerr-NUT-AdS metrics in various dimensions
are analyzed in [132].
60
5 Black holes and conformal field theories
Conformal field theories (CFT’s) have proven to be very useful, especially in the
last two decades with Maldacena’s conjecture [24] in the context of stringy or non-
stringy physics. Besides that, there has been interest in black hole CFT duals,
such as Kerr/CFT. We will review them in this chapter.
Conformal transformations
0 ∂x0ρ ∂x0σ
gρσ (x0 ) = Λ(x)gµν (x) , (5.1)
∂xµ ∂xν
where Λ is called the scale factor. We previously denoted it by Ω2 . The conformal
group by definition is the group of transformations that preserves angles [140];
specifically, the angle between two arbitrary vectors v · u/(v 2 u2 )1/2 where · denotes
the inner product. The conformal group has the Poincaré group as a subgroup.
Viewing it as an infinitesimal transformation xµ → x0µ + µ (x) the metric changes
as [141]
gµν → gµν − (∂µ ν + ∂ν µ ) (5.2)
where we can already see that the dimensions d = 1, 2 are special. For d = 1 there
is no point in talking about angles, so this case should not be considered. For d ≥ 3
61
the transformations that belong to the conformal group are the Poincaré and the
dilations and Special conformal transformations(SCT). These are expressed as
x µ − bµ x 2
x0µ = ωxµ x0µ = . (5.5)
1 − 2b · x + b2 x2
The SCT translates to a scale factor Λ(x) = (1 − 2b · x + b2 x2 )2 as it can be easily
checked. Equivalently it can be seen as an inversion followed by a translation
followed by an inversion.
The generators of the conformal group are the usual translational and rotational
from the Poincaré group, but also dilation D and the SCT generator K:
[Lµν , Lρσ ] = i(ηνρ Lµσ + ηµσ Lνρ − ηµρ Lνσ − ηνσ Lµρ ) . (5.7)
∂1 1 = ∂2 2 ∂1 2 = −∂2 1 (5.8)
and we transform the parameters into complex ones, by also going to complex
coordinates z, z̄ = x1 ± ix2 . The metric transformation is now
2
∂f
ds2 = dzdz̄ → dzdz̄ . (5.9)
∂z
In fact, any two dimensional metric is conformally flat [142]. The infinitesimal
transformations are
with
n = −z n+1 ¯n = −z̄ n+1 (5.11)
62
and the generators
ln = −z n+1 ∂z ¯ln = −z̄ n+1 ∂z̄ (5.12)
The linear combinations that generate dilatations and rotations are l0 + ¯l0 and
i(l0 − ¯l0 ) respectively. Naming their eigenvalues as h, h̄, we can characterize a
state by the physical quantities scaling dimension ∆ = h + h̄ and spin s = h − h̄.
The eigenvalues are called conformal weights of the state. Under a conformal map
z, z̄ → w(z), w̄(z̄) if a field of the theory transforms as
−h −h̄
0 dw dw̄
φ (w, w̄) = φ(z, z̄) , (5.15)
dz dz̄
it is called primary, with conformal weights (h, h̄). If this holds only for global
transformations, it is called quasi-primary. The fields which are not primary are
called secondary.
5.1.3 Quantization
Two dimensional CFTs are quantized by radial quantization, by first taking light
cone coordinates in Euclidian space, then compactifying the space coordinate by
imposing a periodicity, and then mapping the cylinder to a plane. This way the
Hamiltonian is the dilatation generator. We can find charges by the standard
Noether procedure. These charges will generate the symmetries, and can be con-
structed by the energy momentum tensor Tµν , which is traceless for conformal the-
ories. This also allows us to split it into holomorphic and antiholomorphic pieces,
factorizing many properties of the theory into left and right moving sectors.
63
Of course, quantization requires promoting every field into an operator. A very
useful tool in quantum CFTs is the operator product expansion (OPE). This is an
expression that shows what happens when two operators get close to each other.
Denoting an operator with O, the OPE will be
X
Oi (z, z̄)Oj (w, w̄) = Cijk (z − w, z̄ − w̄)Ok (w, w̄) . (5.16)
k
A primary operator is one for which the OPE with the energy momentum tensor
is
h 1
T (z)O(w, w̄) = 2
O(w, w̄) + ∂w O(w, w̄) + . . .
(z − w) z−w
h̄ 1
T̄ (z̄)O(w, w̄) = O(w, w̄) + ∂w̄ O(w, w̄) + . . . . (5.17)
(z̄ − w)2 z̄ − w
In other words, the OPE must truncate at order (z − w)−2 regarding the poles.
All operators are either primary or primary descendants. The stress energy tensor
is not primary. Unitarity though [145] implies that the OPE of the energy tensor
with it self must truncate at inverse fourth order, i.e.
If now we Fourier expand the energy tensor on the cylinder it has the form
∞
X c
Tcyl (w) = − Lm eimw + , (5.20)
m=−∞
24
64
which mapping it to the plane for the radial quantization becomes
∞
X Lm
T (z) = , (5.21)
m=−∞
z m+1
and similarly for the right moving part. Inverting these lead to the contour integrals
I I
1 n+1 1
Ln = dzz T (z) L̄n = dz̄ z̄ n+1 T̄ (z̄) . (5.22)
2πi 2πi
These are the conserved charges associated to δz = z n+1 and δz̄ = z̄ n+1 , which in
the quantum theory becomes the generators for these conformal transformations,
named Virasoro generators. The algebra they obey is the Virasoro algebra
c
[Lm , Ln ] = (m − n)Lm+n + m(m2 − 1)δm+n,0
12
c̄
[L̄m , L̄n ] = (m − n)L̄m+n + m(m2 − 1)δm+n,0
12
[Lm , L̄n ] = 0 , (5.23)
which is like the Witt algebra but with a central extension, associated with the cen-
tral charges. The subalgebra comprised by {L−1 , L0 , L1 } ∪ {L̄−1 , L̄0 , L̄1 } generates
the global conformal group and has no extension. The Hamiltonian is expressed
as L0 + L̄0 .
aτ + b
τ→ ad − bc = 1 . (5.24)
cτ + d
Some useful modular transformations are :
• T : τ →τ +1
65
τ
• U:τ→ τ +1
• S : τ → − τ1 .
Now, the partition function is defined as the sum over all configurations weighted
appropriately. In the path integral formulation, choosing Re(w) to be the space
direction and Im(w) to be the time direction, the partition function reads [139]
where H is the Hamiltonian and P the momentum of the system, and we trace
over all states in the Hilbert space. In order to express this in terms of Virasoro
operators, we use the cylinder expressions for these quantities, namely
c + c̄
H = L0 + L̄0 − P = i(L0 − L̄0 ) (5.26)
12
and arrive to
Z(τ, τ̄ ) = T rH (q L0 −c/24 q̄ L̄0 −c̄/24 ) , (5.27)
where (q, q̄) = (e2πiτ , e−2πiτ̄ ) and now the trace means we sum over all eigen-
values of L0 , L̄0 , or all scaling fields [146]. According to John Cardy [147] this
partition fnction is modular invariant. Modular invariance under T means that
all fields must have integer conformal spin [146]. The S invariance though is re-
ally different; it means that we can express the partition function in terms of
¯ = (e−2πi/τ , e2πi/τ̄ ).
(q̃, q̃)
¯ respectively, then the partition func-
Suppose the eigenvalues of L0 , L̄0 are ∆, ∆
tion can be written in terms of the density of states ρ(∆, ∆) ¯ due to degeneracy
Then we can invert this relation to extract the density through contour integration
[1]
Z(q, q̄)
Z
¯ 1 dq dq̄
ρ(∆, ∆) = ¯ −c/24
, (5.29)
2
(2πi) C q ∆+1 q̄ ∆+1 q q̄ −c̄/24
where the contour includes the poles q, q̄ = 0 and is defined for τ, τ̄ in the upper
complex plane. Then using the modular invariance of Z we have
66
Treating τ, τ̄ as independent variables and knowing that the partition function
factorizes for the limit of low temperature Im(τ ) → i∞, we use the saddle point
approximation for large ∆. The extremum of both exponents is at
r
c c c
∆− + 2
= 0 =⇒ τ ≈ i (5.32)
24 24τ 24∆
¯ τ̄ ). Then plugging these back in the density we get
and similarly for (c̄, ∆,
r r
c∆ ¯
c̄∆
ρ(∆, ∆)¯ ≈ exp 2π + 2π Z̃(i∞)Z̃(−i∞) , (5.33)
6 6
yielding the celebrated Cardy formula for the statistical entropy of the system
r r
¯ c ∆
L L cR ∆R
SCardy = log[ρ(∆, ∆)] ≈ 2π + = SL + SR , (5.34)
6 6
where we skipped a lot of details but this is the main idea. We also changed in the
last equation from barred and unbarred quantities to right and left. An important
point to make is that if the lowest eigenvalue of the theory ∆0 is not zero, then
we have to modify the central charges [1] as c → cef f = c − 24∆0 . Of course this
formula has space for logarithmic [2] and beyond logarithmic [3] corrections.
Using definitions for left and right temperatures [152]
−1
π2
∂SCardy 2
TL,R = =⇒ ∆L,R = cL,R TL,R (5.35)
∂∆L,R ∆R,L =const 6
π2
SC = (cL TL + cR TR ) . (5.36)
3
5.3 Kerr/CFT
The Bekenstein-Hawking formula for the black hole entropy comes from macro-
scopic analysis. However, one would expect that there is a microscopic derivation
for the same entropy, by counting microstates, since an entropy is inevitably re-
lated to thermodynamics, which have a statistical interpretation. Of course, the
weird part in this entropy is its relation with the horizon; it seems that all different
configurations will happen strictly there.
The existence of the statistical origin of entropy in the context of string theory was
explored by Strominger and Vafa [153] for a large class of five dimensional black
67
holes, and later in the context of ”early AdS/CFT correspondence” without the
use of string theory [154]. We use quotation marks because the conjecture was
only recently made at that time by Maldacena [24] and the motivation was taken
by the work of Brown and Hennaux [155]. They proved, even before Maldacena’s
conjecture, that any consistent theory of quantum gravity in AdS3 is a conformal
3l
field theory with central charge (we restore the units) c = 2G√3 −Λ = 2G , where Λ is
the cosmological constant.
What they did is define boundary conditions at spatial infinity obeyed by the
fields of the theory, and then try to find the symmetries that preserve the behavior
of the asymptotics. These symmetries are the gauge transformations that leave the
fields asymptotically invariant [155]. The AdS3 has an SL(2, R)L ×SL(2, R)R isom-
etry group. The allowed diffeomorphisms preserving the symmetries are generated
by vector fields given in [155]. Out of these diffeomorphisms, some are consid-
ered ”pure” or ”proper” gauge transformations, they act trivially in the sense that
they cannot be associated with charges, since they vanish weakly giving unphys-
ical effects. The group comprised of the allowed diffeomorphisms modulo these
pure gauge transformations, named trivial diffeomorphisms, is called Asymptotic
symmetry group (ASG).
Suppose the generators of these symmetries are Ln , L̄n with n ∈ Z. They
showed they obey an algebra with a central extension, which would vanish if they
were exact symmetries. The algebra is
c
[Lm , Ln ] = (m − n)Lm+n + m(m2 − 1)δm+n,0
12
c
[L̄m , L̄n ] = (m − n)L̄m+n + m(m2 − 1)δm+n,0
12
[Lm , L̄n ] = 0 (5.37)
3l
with c = 2G , which is of course the Virasoro algebra with equal right and left
charges.
But there is no need to stop here. One can go further and compute the entropy
of a BTZ black hole! Associating our Hamiltonian and momentum operators of a
conformal field theory with the mass and momentum of the black hole, we have
[154]
1
M = (L0 + L̄0 ) J = L0 − L̄0 . (5.38)
l
Then counting the excitations of the vacuum for large M, the number of states will
be given by Cardy’s formula
r r
c∆L c∆R
S = 2π + 2π , (5.39)
6 6
68
where ∆R,L are the eigenvalues of L0 , L̄0 . Using the mass and angular momentum
we arrive to r r
l(lM + J) l(lM − J)
S=π +π , (5.40)
2G 2G
which is precisely the Bekenstein-Hawking entropy of the BTZ black hole found
by Banãdos, Teitelboim and Zanelli!
This, in turn, many years later [157] motivated Guica et.al. to use the Kerr throat
geometry studied by Bardeen and Horowitz in 1999 [158], where they showed that
the near horizon limit of an extreme Kerr black hole in four dimensions has an
SL(2, R) × U (1) symmetry, and arises as an AdS2 × S 2 geometry. This limit is
now referred as Near Horizon Extreme Kerr or NHEK. They also showed that the
Myers Perry generalization of Kerr in five dimensions has an SL(2, R) × U (1)2
symmetry group. An extensive study of near horizon geometries is given in the
more recent paper [160] .
The extreme Kerr solution has the properties
1
r± = a = M S = 2πM 2 = 2πJ TH = 0 ΩH = , (5.41)
2M
with the usual metric
∆ ρ2 2 sin2 θ 2
ds2 = − (dt − asin 2
θ dφ)2
+ dr + 2 [(r + a2 )dφ − adt]2 + ρ2 dθ2 . (5.42)
ρ2 ∆ ρ
The near horizon limit is obtained by introducing a parameter λ, making coordinate
transformations and ”zooming in” with λ → 0. One sets [159]
r−M λt t
r̃ = t̃ = φ̃ = φ − (5.43)
λM 2M 2M
and then taking the variations and substituting in the metric, removing tildes
afterwards, one gets
2
2 2 dr 2 2 2 2 2
ds = 2Ω J 2 + dθ − r dr + Λ (dφ + rdt) , (5.44)
r
where
2 1 + cos2 θ 2sinθ
Ω = Λ= . (5.45)
2 1 + cos2 θ
This is no longer asymptotically flat. For θ = 0, π we get AdS2 along the axis.
This metric, in addition to the original Kerr symmetries (translation and rotation),
has a dilation symmetry, and we say its symmetry is enhanced. The full isometry
group is SL(2, R) × U (1). The U (1) comes from the rotation ∂φ and the original
translational symmetry becomes part of the enhanced SL(2, R). Slices of angle
69
θ give quotients of warped AdS3 , since the angle φ is periodic. These quotients
are black holes, the same way as BTZ comes from the periodic identification of
coordinate φ in AdS3 . In the appendix we have the specifics of a warped AdS3
black hole, and an extensive review is given in [102].
After achieving this geometry, we can excite it. Following the steps of Brown
and Hennaux, we can find the asymptotic symmetry group by imposing boundary
conditions preserving the asymptotics. One of the most difficult parts of this is
finding the proper boundary conditions for this, which are not uniquely determined
(for instance, for the AdS3 case, different asymptotics where chosen in [172] giving
more degrees of freedom). The generators ζm are found [157] to obey
with ζ0 generating the U (1). We then find the conserved charges Qζ associated
with the generators, whose Dirac brackets will obey an algebra similar to (5.46),
with a possible central extension. Indeed, it turns out that the Dirac bracket of
two charges is
i
(m3 + 2m)δm+n J
Qζm , Qζn = Q[ζm ,ζn ] + − (5.47)
8πG
and defining quantum operators, and taking Dirac brackets to commutators in
order to quantize the theory ,we have
3J
~Ln = Qζn + δn , (5.48)
2
with the algebra
J
[Lm , Ln ] = (m − n)Lm+n + m(m2 − 1)δm+n,0 , (5.49)
~
which is a Virasoro algebra with central charge c = 12J ~
. From now on we will use
units where ~ = 1 so c = 12J. We see that there is only one copy of Virasoro
algebra. We usually say that it is the chiral half of a 2D CFT. The other set of
Virasoro generators is ”frozen”.
The other way it can be viewed is the discrete light cone quantized (DLCQ) full
2D CFT [161], which was proved to be equivalent for the extremal BTZ case, and
also serves as support for the validity of Kerr/CFT. The way DLCQ is performed
is a coordinate transformation, followed by the near horizon BTZ limit, ending up
with a null cylinder boundary with a compact null direction. Then one boosts the
cylinder with a certain rapidity, and gets a Hilbert space as a product of a left and
70
right part, where only one will survive.
An important thing to note is that the SL(2, R) isometry of the AdS2 geometry
is contained in the frozen sector, while the surviving SL(2, R) with the Virasoro
algebra is the enhancement of the U (1) part of the geometry.
One can already notice where this is heading; since we managed to associate the
black hole with a CFT, we can have a temperature and find the entropy through
the Cardy formula. Of course we cannot the use the Hawking temperature, since
it is zero for the extremal case.
Now, a free field in a Schwarzschild spacetime yields the well known Hartle-
Hawking vacuum in the path integral formulation [162]. The field expanded in
modes of frequency ω can be written as
X
Φ= e−iωt χl (r, θ) . (5.50)
ω,l
Tracing out the black hole region yields a density matrix weighted by a Boltz-
mann factor e−ω/TH , with TH being the Hawking temperature and that is the
temperature for the black hole radiation as a quantum field theory computation.
Similarly Frolov and Thorne did the same for a quantum field near the horizon of
a rotating black hole [163], expanding it in eigenmodes
X
Φ= e−iωt+imφ χl (r, θ) , (5.51)
ω,m,l
with m being the angular momentum part. This vacuum is not as well defined as
the Hartle-Hawking one, since there is no global timelike Killing vector. Neverthe-
less, it works fine near the horizon. Taking the NHEK limit (5.43) and inverting
the relations, we have
2M t t̃
t= t̃ φ = φ̃ + = φ̃ + . (5.53)
λ 2M λ
So the exponential in (5.51) will become
2M ω t̃
exp{−iωt + imφ} = exp{−i t̃ + im(φ̃ + )} = exp{−ihR t + ihR φ} , (5.54)
λ λ
2M ω−m
with hR = λ
and hL = m.
71
The Boltzmann factor (5.52) becomes exp{− ThLL − hR
TR
} after defining left and
right temperatures
r+ − M r+ − M
TL = TR = , (5.55)
2π(r+ − a) 2πλr+
1
which at extremality r+ → M = a become (TL , TR ) → ( 2π , 0). So we have achieved
to get a mixed thermal state with temperature TL . Using the central charge ac-
quired earlier and assuming unitarity, Cardy’s formula reads
π2 π2 1
SC = cL TL = 12J = 2πJ , (5.56)
3 3 2π
which is miraculously the Bekenstein Hawking entropy of the extremal Kerr solu-
tion! The caveat is that we do not know if the Cardy formula is applicable [159].
Similar analysis has been done for Reissner-Nördstrom black holes [165] that also
have an AdS2 × S 2 throat, admitting a dual CFT with central charge, temperature
and entropy:
1
cRN = 6Q3 TRN = SRN = πQ2 , (5.57)
2πQ
again reproducing the macroscopic entropy. The generalization can be done to
the Kerr-Newman case, where one can use either of the central charges with the
corresponding temperature [152]
a2 + r+
2
a2 + r+
2
TJ = TQ = 2
(5.58)
4πar+ 2π(r+ − a2 )3/2
and then also in five dimensional extremal black holes [170]. Inclusion of the
cosmological constant for rotating or charged black holes was achieved in The
more general Kerr-(A)dS/CFT correspondence in four,five and more dimensions
was done in [168]. For example, the Kerr-AdS4 , using a Frolov-Thorne vacuum
again appears to have an extremal limit chiral CFT dual with central charge,
temperature p
2 2 −2 2 −2
12r+ (1 − r+ l )(1 + 3r+ l )
cL = 2 −2 4 −4
1 + 6r+ l − 3r+ l
2 −2 4 −4
1 + 6r+ l − 3r+ l
TL = 2 −2
p
2 −2 2 −2
(5.59)
2π(1 − 3r+ l ) (1 − r+ l )(1 + 3r+ l )
and entropy a la Cardy
2
2πr+
SKN AdS4 = 2 −2
, (5.60)
1 − 3r+ l
which agrees with the macroscopical one.
72
A slightly different technique was used in [166] to obtain the entropy from a
2D CFT of Schwarzschild, RN, Kerr, KN, (A)dS3,4 and BTZ. They use a Liouville
theory, and their results are more general than the extremal case, but we will not
elaborate more on this.
Lately and since the black rings were discovered, there have been CFTs con-
jectured to be dual to the doubly spinning and the dipole solutions [54] [168].
Finally, the Kerr/CFT correspondence was tested in the context of string theory
[173], with stringy black holes and new insights on the validity of the theory.
Suppose we have a massless scalar field in the Kerr background. Its equation of
motion will be the Klein-Gordon (KG) equation, which in a curved background
73
takes the form
1 √
Φ = 0 =⇒ √ ( −gg µν Φ,ν ),µ = 0 , (5.61)
−g
where Φ is the scalar field, and we can again expand it as
In the appendix we calculate explicitly the Klein Gordon operator for the Kerr
spacetime and we arrive to
where ∇S 2 is the Laplacian on the two sphere, with only angular dependence
1 1
∇S 2 = ∂θ (sinθ∂θ ) + ∂2 . (5.64)
sinθ sin2 θ φ
Equation (5.63), as well its Dirac analogue, can be separated. This is due to the
existence of a Killing-Stäckel tensor, sometimes called Killing-Yano [179]. These
are generators of hidden background symmetries, they are antisymmetric 2-tensors
and they satisfy
∇(λ fµ)ν = 0 . (5.65)
Its existence allows the construction of a conserved charge along the geodesics
Q = Kµν ẋµ ẋν and allows the construction of an operator that commutes with the
box operator, therefore making the separation of the wave equation possible.
Back to our scalar, we separate it in a radial and an angular part (for the χ
since the exponents cancel)
where the new equations will have a separation constant S labeled by l and will
read
α(r+ )2 α(r− )2
2 2
∂r (∆∂r ) + − + (r + 2M (r + 2M ))ω − Sl R(r) = 0 (5.69)
(r − r+ ) (r − r− )
74
m2
1 2 2 2
∂θ (sinθ∂θ ) − + ω a cos θ + Sl Θ(θ) = 0 , (5.70)
sinθ sin2 θ
where α(r) = 2M rω − am. These are solved by Heun functions, solutions to the
Heun equation, which is a second order linear differential equation. We need to
make an approximation to get the eigenvalues analytically.
Approximation
In order to make an approximation, one needs to notice which terms are ”annoy-
ing”, in the sense that if they vanish we can get an analytic solution. We can
observe that if the third term in equation (5.70) was neglected, a simple Lapla-
cian is left. A way of achieving this, is going to low energy, or large wavelength
compared to the radius
λ r+ =⇒ ωr+ 1 =⇒ ωM 1 (5.71)
ω 2 a2 ∼ ω 2 /M 2 ω 2 M 2 1 , (5.72)
so we can neglect the third term in (5.70) in the near region, obtaining a Legendre
equation with spherical harmonic solutions, as in the Schrodinger problem, so
Sl = l(l + 1) with l ∈ (−m, m). Plugging it back in (5.69) gives a hypergeometric
differential equation, after we neglect the third term, since
r 1/ω =⇒ r2 ω 2 1 , (5.73)
75
Conformal symmetry
Motivated by the change of BTZ coordinates to Poincare so that it looks like AdS3
locally we perform a coordinate transformation (t, r, φ) → (w+ , w− , y)
q
+ r−r+ 2πTR φ
w = r−r−
e
q
r−r+ 2πTL φ− t
w− = r−r −
e 2M (5.75)
q
y = r+ −r− eπ(TL +TR )φ− 4M
t
r−r−
with √
r+ + r− M r+ − r− M 2 − a2
TL = = TR = = . (5.76)
4πa 2πa 4πa 2πa
Then in this quest of finding the conformal algebra we define the local vector fields
1
L1 = i∂+ L0 = i(w+ ∂+ + y∂y ) L−1 = i[(w+ )2 ∂+ + w+ y∂y − y 2 ∂− ] , (5.77)
2
which will constitute the left conformal algebra, so we define the corresponding
barred quantities for which + ↔ − which will generate the right algebra. It is a
trivial task to prove that
for both, confirming this point. The center of the algebra is the Casimir operator
(it commutes with everything), which quadratically reads
1
C 2 = C̄ 2 = −L20 + (L1 L−1 + L−1 L1 ) . (5.79)
2
It turns out [178] that this operator coincides with the Klein Gordon operator in
the near region, when changing back to the original coordinates. The eigenvalue
problem becomes
C 2 χ = C̄ 2 χ = l(l + 1)χ . (5.80)
76
coordinates (5.75) get identified as
2
w+ ∼ w+ e4π TR
2
w− ∼ w− e4π TL (5.81)
y ∼ ye4π2 (TR +TL ) .
2 (T L +T L̄ )
generated by e−i4π R 0 L 0
, since we have
1 i
L0 w+ = (iw+ ∂+ + y∂y )w+ = iw+ L̄0 w− = iw− L0 y = y
2 2
i
L̄0 w+ = L0 w− = 0 L̄0 y = L0 y = y . (5.82)
2
So acting with this generator
2 (T L +T L̄ )
e−i4π R 0 L 0
w+ =
1
w+ − i4π 2 (TR L0 w+ + TL L̄0 w+ ) + (i4π 2 )2 (TR L0 + TL L̄0 )2 w+ + . . . =
2
1 2
w+ + 4π 2 TR w+ + (4π 2 )2 TR2 w+ + . . . = e4π TR w+ (5.83)
2
and similarly for w− , since L̄0 will act non trivially and we will obtain the left
temperature. As for y,
2
e−i4π (TR L0 +TL L̄0 ) y =
i i 1 i
y − i4π 2 (TR y + TL y) + (i4π 2 )2 (TR + TL )2 ( )2 y + . . . =
2 2 2! 2
2 (T +T )
e2π R L
y, (5.84)
Entropy
Can we find a microscopical entropy for this process? Assuming there is a dual
CFT and the validity of Cardy’s formula, we can use the central charges obtained
with the Brown-Hennaux method cR = cL = 12J for Kerr, together with the
temperatures we obtained from the conformal algebra. Plugging everything in
yields (using a = J/M
π2 r+
SCF T = (cL TL + cR TR ) = 4π 2 = 2πM r+ , (5.85)
3 2πa
which is the macroscopic entropy given by the Bekenstein Hawking area law. Fi-
nally, in [178] they check the agreement of scattering amplitudes in the two regimes.
77
6 Inner horizons and geometric relations
The inner horizon is thought as being unstable classically, first argued by
Simpson and Penrose [99] and then by Poisson and Israel’s mass inflation [100] that
happens during an observers trip to this surface, becoming singular (specifically
transforms into a null singularity). It is although suggested [101] that for quantum
black holes this instability may not occur. However there are indications in three
dimensional gravity and the divergence of the energy momentum tensor that it
is also quantum mechanically unstable [182]. But why would we be interested in
inner horizons?
Lets digress here and mention a crucial fact. The above form of the inner
horizon first law is not universal in literature. That is because of the definition of
the temperature of the inner horizon. Usually it is the case that this temperature
turns out to be negative. In order to remedy this ”unphysical” property, one can
define the temperature to be instead the absolute value of this quantity i.e.
T− = T+ r+ ↔r−
. (6.3)
or equivalently
78
The latter version makes it clear that the mass in the inner horizon is negative,
since the time translation Killing vector is spacelike there. Unless stated otherwise,
the former version (6.2) will be used, here and in the appendix.
There are several interesting (and conflicting) views of how the inner horizon
radiates. One view is that the same way a stationary observer will see a flux of
Hawking radiation from the outer horizon, he will see a flux of absorption from the
inner horizon [192]. The other view is that it actually radiates particles towards
the singularity and antiparticles towards the other direction (which will end up in
another universe) [193]
Inner horizon thermodynamics were studied in the case of a Kerr black hole
first by Curir [183] [184] [185]. She names the inner horizon quantities as ”spin
temperature” and ”spin entropy” for various reasons. First of all, it is related
to the spinning of the black hole. Also, it being negative can be related to its
interaction with particles rotating or counter-rotating with respect to the black
hole, the spin-spin interaction also studied by Bekenstein [186]. Furthermore, as
pointed out in [187], for the Kerr case we have
S+ + S−
M2 = , (6.6)
4π
which resembles classical thermodynamical systems with spin, with their total
energy being the sum of internal and spin energy.
As for multiple horizon solutions, we indeed have a first law being satisfied for
each one of them, again defining the potentials by substituting the outer horizon
with another, i.e.
79
tions [89], for any charged and/or rotating black hole the relation
Y Y
Ai = Ai,ext = F (Qj , Jk ) (6.9)
i i
always holds. The Ai,ext is the value of the areas for zero Hawking temperature.
For the cases with two horizons (all four and five dimensional asymptotically flat
black holes fall in this category) this collapses to
A+ A− = (Aext )2 , (6.10)
which is essentially the Cardy formula, and NL,R are the left and right movers
excitations, in the two dimensional conformal field theory.
But what about the inner horizon entropy? It can also be expressed as a linear
combination of the excitations, specifically [190] [174]
p p
S− = 2π( NL − NR ) . (6.12)
How can we get this formula as a Cardy entropy? Looking back at the derivation
of the outer horizon entropy as a statistical entropy, when taking the saddle point
approximation we picked the positive solution for the modular parameter (equation
5.32) ; that is because we defined it to be on the positive plane. However, one can
analytically continue it so that we can have yet another density and hence another
linear combination which will yield the inner horizon entropy as
r r
¯ 0 c ∆
L L cR ∆R
SCardy = log[ρ(∆, ∆) ] ≈ 2π − = SL − SR . (6.13)
6 6
80
Now, as argued in [190], there should only be one quantum number in the
macroscopic theory, and we apparently have two; the left and right moving levels.
This is remedied by the constraint
NL − NR ∈ Z . (6.14)
It is clear now how the entropy product gets into the game. In terms of the
excitation levels
S+ S−
S+ S− = 4π 2 (NL − NR ) =⇒ ∈Z, (6.15)
4π 2
so indeed, we expect the entropy product to be quantized, vis. mass independent
(if no other moduli are present). This condition is called level-matching and it is
predicted by the CFT [63], due to modular invariance.
This can be seen as an indication for an underlying holographic description
of black holes and points out the peculiar role the inner horizon plays- of course,
assuming all previous definitions hold for the inner horizon as well. Despite the
fact that the behavior of inner horizons is ill, and not clear in the quantum case, it
serves as a playground for understanding the conformal field theory that is possibly
dual to the black holes.
In the latest paper about this universality of the area product [196], Page
claims (as a conjecture) that for any black hole for which this product is indeed
independent of the moduli, it will continue to be even after the distortion of the
solution with the addition of external matter.
Exploiting this right-left moving breaking, one can also break the first law into
parts, with left and right entities instead of outer and inner (see appendix B). There
is multiple ways to do this depending on the definitions of course. Regardless of
this technical detail, in [194], the use this formulation that they name conformal
thermodynamics, and in [192] they prove that for the Kerr-Newman black hole this
leads to two systems that are quite different. One (lets say the left) describes a
black hole with three hair (M, J, Q) and the other a two-hair one (M, Q). We will
see later that this is due to a very specific geometric relation. What is useful to
note here is the relation of this asymmetry with the fact that in four dimensions
the supersymmetry is broken, due to the angular momentum [195]. The residual
symmetry is (0, 4) superconformal. This can also be seen in Kaluza-Klein reduced
black holes [56], which we include in the appendix.
An important implication that follows directly from the mass independence of
the entropy product, is one that concerns the central charges of the the conformal
81
field theory; specifically, using the first law, the equality of left and right charges
follows [54]. Let us sketch the proof. The differential of the product of the areas,
using the inner and outer horizon first laws is
and solving the first laws for the entropies yields (lets assume one rotation and one
charge for simplicity)
1 Ω± Φ±
dS± = dM − dJ − dQ , (6.17)
T± T± T±
∂(S+ S− )
T+ S+ + T− S− = 0 ⇐⇒ =0 . (6.19)
∂M
we can indeed use Cardy’s formula to obtain a left and a right entropy, precicely
because these two sectors decouple [148]. We then instead have the relation
SR SL ∂(S+ S− )
= ⇐⇒ =0 . (6.21)
TR TL ∂M
Again, let us note here that these definitions vary in the literature, and there can
be factors of two or left and right defined oppositely than here. For instance, if one
takes the inner first law as in (6.4), the mass independence of the entropy becomes
an equality instead of a sum :T+ S+ = T− S− . As for (6.21) it can remain the same
if one defines entropies with reversed left and right. Unless stated otherwise, we
will use the conventions presented in our equations.
Proceeding, we know from Cardy’s formula that the right and left sector en-
tropies can be written in the form
π2
SR,L = cR,L TR,L . (6.22)
3
82
Connecting the two results to
∂(S+ S− )
= 0 =⇒ cR = cL . (6.23)
∂M
Ω+ Ω−
Kerr-Newman: =± , (6.24)
T+ T−
where the sign depends on the definition of the inner horizon temperature (mi-
nus if we use the definition T− = Tr+ ↔r− . As we can see this relation does not
involve entropies - it is between potentials on the outer and inner horizons. This
means that we expect it to hold regardless of the gravitational theory. There is
no reason either for the surface gravity (and hence the temperature) or the co-
efficient of the Killing vector that generates a horizon (the angular velocity in
ξiµ = ∂t + Ωi ∂φ ) to be dependent on the theory. However, unlike the previous
relation, it is not a product, so its generalization to more rotational parameters or
more dimensions is not so clear, since more possibilities arise. Moreover, it seems
to be heavily metric dependent, and it is expected to change with some coordinate
redefinition.
The immediate generalization comes from the Myers-Perry black hole in five
dimensions, with two rotational parameters, related to the coordinates ψ and φ.
83
We observe (see appendix A.4) that there is a mixing of the rotational velocities
when we try to find such a relation, namely
ΩR
− ΩR ΩL− ΩL+
Myers-Perry: ± + =0 ∓ =0 , (6.27)
T− T+ T− T+
where again the plus or minus sign depends on the definition of the inner temper-
ature. It is clear of course that the definition does not play a significant role, since
there are ambiguities in the definitions of right and left as well. However, these
relations keep having an opposite sign (except if we do a trick as shown in A.4).
The last piece of information at hand is the relations for the BTZ black hole.
For this one, there are two types of relations that we will use. The first one involves
the entropy
Ω+ Ω−
BTZ a: S+ = ± S− , (6.28)
T+ T−
where once again, the sign depends on the temperature convention (as always, the
upper sign is if we take the inner temperature to be positive). As we will see
later, this relation has the upside that it can be generalized to higher dimensions.
However, it involves Wald entropies, which means it will be theory dependent.
Even with the mere addition of a Chern Simons term in the theory, this relation
does not hold (although different ones arise, see appendix A.1). Of course we have
again the same expectation that it will hold for areas, like for the area product,
but we cannot prove that (and we cannot even test it since black hole solutions
in Lovelock theories have not been found) and we do not know how useful it is,
since we need the entropies for the conformal field theory as well. The second one
involves only the potentials, but it takes a peculiar form
Ω+ 1 Ω− 1
BTZ b: =± =± , (6.29)
T+ lT− T− lT+
where again the same sign conventions apply and l is the Ads radius. We do not
know which one of this relations is more relevant or more useful, so we will try to
explore them both.
84
What could these relations mean or where do they come from? Do they imply
something for the gravitational theory and/or the quantum conformal field theory?
These are some of the questions we are trying to investigate through different
routes which we will present here. Before proceeding, it is crucial to state that all
of these and the following relations are obtained at a frame that is not rotating
at infinity.
6.3.1 Generalization
The first route is a useful one in terms of generalization, if one wants to consider
multiple horizon solutions. We observed, and as far as we know this does not
exist in current literature, that for even dimensional (A)dS solutions to EM
gravity, using all horizons we obtain
X Ωa X Ωa X 1
i i
(A)dSd : =0 Si = 0 =0 , (6.30)
i
Ti i
Ti i
Ti
where the i index runs over the horizons and the a index over the different rotational
parameters. Specifically, we observed that it holds for the four and six dimensional
cases, so we expect it to hold for even dimensions.
The first one of (6.30) looks like a generalization of the Kerr-Newman relation.
Of course this is in conjunction with the 4d asymptotically flat case, which we also
checked for various solutions, even string theory or KK reduced (see appendix A),
in which cases the sum runs over two horizons.
The second one of (6.30) is a generalization of the BTZ relation (6.28), and it
seems to be a relation that holds for any (A)dS black hole.
The third one of (6.30) seems to appear only for the even dimensional case of
the (A)dS solutions.
The other observation that does not appear either in the literature, is for odd
dimensional (A)dS solutions to EM gravity, which have double horizon roots.
If we once again attempt to use all horizons, we obtain the same relation
X Ωa
i
=0. (6.31)
i
Ti
BUT is it really meaningful? The caveat here is the fact that we have double
horizon roots; This means that mainly these equations are identities. Since we
would also like to use the first law, it does not really make sense if we have first
laws for negative horizons. However, it seems all horizons are not needed. In fact,
85
including the BTZ black hole, using only half of the horizons, we get
X Ωa
i
(A)dSd+1 : Si = 0 . (6.32)
i
Ti
The cute feature that comes with that is the similarity to the entropy product.
We can always use the same amount of horizons used in [58] (where they show the
mass independence of the area product for many different solutions) to have our
relations. This also means there is possibly some correspondence between the two
relations. Finally, as seen in the appendix, these relations seem to generalize for
the supergravity solutions as well, again in conjunction with the area product [89].
For this part, let us start with the Kerr type relation (6.24). We will also use the
plus convention for simplicity.
So the relation can be rewritten as
Kerr: β+ Ω+ = β− Ω− , (6.33)
where β is the inverse temperature. The observation here is that this product
appears as the period of the angular coordinate when one goes to the euclidian
sector of a black hole. Specifically, euclidianizing a black hole transforms the
Killing vector as
ξEµ = ∂τ + iΩH ∂φ . (6.34)
Regularity on the boundary, lets say the outer event horizon r+ requires periodic
identification of the Wick rotated time with the period being equal to the inverse
temperature, otherwise we get a conical singularity. Since the boundary is a fixed
point of ξEµ [204], we need to impose periodicity on the angular coordinate as well.
This is a well known procedure in the path integral formulation of gravity [205].
So the identification becomes
on the outer horizon, and, we can easily generalize to each horizon. Of course all
of the potentials will be evaluated on each horizon separately, since they are fixed
points of different Killing vectors. That said, we can see now where this is leading:
relation (6.33) implies that the periods of the angular coordinates will be equal at
the inner and outer horizons. That means we can see these geometric potential
relations as smoothness conditions. Naming the periods of time and angle T
86
and Φ respectively, we have
Kerr: Φ+ = Φ− . (6.36)
However there is a problem arising. For the BTZ black hole, as we saw, what
is true instead is
BTZ: β+ Ω+ = β− β− Ω− = β+ , (6.37)
where we have set l = 1 for simplicity (we can always rescale the time coordinate
so that we get the prefactor). So for the BTZ, it is instead the period of the time
coordinate in the inner horizon that is equal to the angle coordinate in the outer
horizon and vice versa. In the same notation as before we have
BTZ: T± = Φ∓ . (6.38)
We do not really know why this condition should hold, or why it is violated by the
BTZ. However, it is worth noting since there has to be some connection.
As an aside, looking back at the Kerr Penrose diagram2.4, we can see that there
are some points in the diagram where the inner and outer horizons meet. Maybe
this relation ensures a smooth patching at that point. However, once again, the
reason BTZ violates it seems elusive.
As for the Myers-Perry black hole, the relations in terms of angular periodicities
become
Myers-Perry: Φ± = Ψ± , (6.39)
A last problem that occurs is the difficulty of interpreting this condition for
multi horizon black holes. We would have to go to the sum convention, instead of
equalities, and we would have for the even dimensional Kerr-Ads
X
Kerr − (A)dSd : Φi = 0 , (6.41)
i
87
6.3.3 Conformal field theory
We saw that for the entropy product there is a clear conformal field theory output:
the mass independence ensures the equality of left and right charges. Could we
say something similar for this relation?
Since the conformal field theory favors the left-right sector breakup of our
system, we are going to reside to that. But lets first take a look on what these
relations imply through the first law.
So the first law as we mentioned, has the form which holds for each horizon
separately
dM = Ti dSi + Ωai dJa + Φbi dQb , (6.42)
where i runs over distinct horizons and a, b run over different rotations and charges
respectively (summation is implied). One can easily see that the ratios of potentials
are equal to
∂Si Ωa ∂Si2 Ωa ∂Si 1
=− i = − i Si = . (6.43)
∂Ja Ti ∂Ja Ti ∂M Ti
That means we can express our relations through the first law, in order to make
connections to the entropy and potentially the conformal theory. A nice feature is
that, independently of our initial sign conventions for temperatures, these relations
are invariant. Let us rewrite the previous relations in this way first, making the
distinction between the different types of black holes we presented. So for our four
dimensional asymptotically flat black holes we will have
∂(S+ + S− )
4dAF: =0 , (6.44)
∂J
∂(S+2 + S−2 )
BTZ: =0 . (6.45)
∂J
If we can furthermore define left and right sectors, with left and right first laws,
then again our relations will transform accordingly. The only two solutions we can
do this are the former since they have only two horizons (the Myers Perry also has
two horizons but it can be treated as two independent Kerr-type relations). In this
setting the relations become
4dAF: ∂J SR = 0 or ΩR = 0 (6.46)
88
Before going into more details about this, let us present for completeness the other
two types of black holes
X ∂Si X ∂S 2 X ∂Si
i
(A)dSd : =0 =0 =0 , (6.48)
i
∂J i
∂J i
∂M
X ∂S 2
i
(A)dSd+1 : =0 . (6.49)
i
∂J
In this subsection we will analyze the BTZ relation and how it implies once again,
the equality of left and right central charges. An prior essential note to make is that
we use the conventions of [148], which means that we have the inner temperature
taken as an absolute value, and as a result the left and right sectors are interchanged
(with respect to the rest of the project). The reason we do this is in order to have
a meaningful comparison.
In [148](the discussion part) the mass independence of the entropy product is
used in order to obtain a Cardy type of relation for a dual two dimensional CFT ,
and an expression for those central charges is found. Specifically, they show that
if
S+ S−
= F (J) , (6.50)
4π 2
then the central charges are equal to
∂F
cR = cL = c = 6 . (6.51)
∂J
Essentially we aim to do the same, but using the other relation we have for the
BTZ. We will set l = 1 for simplicity. Following the method of [148] we have for
the entropies
δM Ω± 4π 4πΩ±
± δS± = − δJ = δM − δJ = 4πα± . (6.52)
T± T± 2κ± 2κ±
Now, instead of solving in terms of angular momentum, we solve in terms of mass,
by eliminating J:
δM κ± δS±
=± + δJ =⇒
Ω± 2πΩ±
1 1 1 κ+ κ+
δM − = δS+ + δS− =⇒
Ω+ Ω− 2π Ω+ Ω+
2
δM = (κ+ Ω− α+ − κ− Ω+ α− )
Ω− − Ω+
89
" #
κ+ Ω− + κ− Ω+ κ+ Ω− − κ− Ω+
= 2π ωL TL + ωR TR , (6.53)
κ+ + κ− κ− − κ+
which can be rewritten as
Ω+ Ω−
S+ = S− (6.56)
T+ T−
∂(S+2 + S−2 )
= 0 =⇒ S+2 + S−2 = 8π 2 G(M ) (6.57)
∂J
and varying
∂G
8π 2 δM = δ(S+2 + S−2 ) = 2S+ δS+ + 2S− δS− = 4(SR δSR + SL δSL ) ,
∂M
we get
∂G
2π 2
δM = 2πSL ωL + 2πSR ωR , (6.58)
∂M
where we similarly introduced
S± = SL ± SR δSL,R = 2πωL,R .
90
Comparing with (6.54) we get:
∂G 0
SL,R = 2π 2 T . (6.59)
∂M L,R
So this is a Cardy formula
π2 0
SL,R = c T0 (6.60)
3 L,R L,R
with central charges
∂G
c0 = c0R = c0L = 6 . (6.61)
∂M
This is very important. It shows that essentially the two geometric relations
(6.19 and 6.56) are equivalent for the case of BTZ in Einstein gravity .
The central charge in [148] will be equal to this one if
∂G ∂F
c0 = c ⇐⇒ = , (6.62)
∂M ∂J
where
S+ S−
F (J) = ,
4π 2
which holds for BTZ (see A.1).
Here we will try to simulate the approach made in [149] ( in the section ”Digression
on Inner horizons”) for the central charges, by trying to make use of our new
relation. Our starting point is the BTZ, but in fact, this analysis holds for any
(M, J) black hole.
Essentially, one arrives to expressions for the central charges that depend only
on inner and outer horizon quantities and their entropy derivatives. This way the
geometric relations can be exploited, as we will see. However, these expressions,
general as may be, lack certainty of applicability, since they do not guarantee the
constancy of the central charges.
Using solely the assumption that the entropies are functions of (M, J) we have
91
Now, using the exact same methods and quantities as before and as in [148] we get
again:
δM = 2πTL0 ωL + 2πTR0 ωR δJ = 2πTL ωL − 2πTR ωR (6.64)
and , as in (6.58)
2πSL ωL + 2πSR ωR = 2π 2 (AδM + BδJ) = 4π 3 (ATL0 + BTL )ωL + (ATR0 − BTR )ωR ,
(6.65)
which yields
0
SL,R = 2π 2 (ATL,R ± BTL,R ) . (6.66)
Similarly, from the monodromy paper [148] there is the analogue of (6.58)(eq (4.6)
in [148]) :
∂F
2π 2 δJ = 2πSL ωL − 2πSR ωR . (6.67)
∂J
So the generalization of this in this case will be (using (6.63))
resulting in
0
SL,R = 2π 2 (±CTL,R + DTL,R ) . (6.69)
where all A,B,C,D are functions of (M,J) and defined in (6.63). This shows that
we can equivalently use either one of these CFT pictures.
In order to carry on this analysis, we will be using TL,R only (and not the
primed ones) , putting the extra factors in the central charges.
So equations (6.66) and (6.69) can be rewritten
2
L,R = 2π (AbL,R ± B)TL,R
S
, (6.71)
S = 2π 2 (±Cb + D)T
L,R L,R L,R
where
κ+ Ω− ± κ− Ω+
bL,R = . (6.72)
κ− ± κ+
Here we take a small leap. As argued in [149], the ratio SL,R /TL,R is usually
constant and related to the central charge. That means of course that we assume
the applicability of the Cardy formula, the existence of a dual two dimensional
92
CFT etc, as done in Kerr/CFT. So the central charges will be using the second
one of (6.71) :
δ(S+ S− ) δ(S+ S− ) κ+ Ω− ± κ− Ω+
cL,R = 6 D ± CbL,R = 6 ± =
δJ δM κ− ± κ+
δ(S+ S− ) δ(S+ S− ) κ− Ω+ ± κ+ Ω−
=6 + , (6.73)
δQ∂φ δQ∂t κ− ± κ+
which is equation (84) in [149] (normalization only is changed due to initial nor-
malization of our equations but the net result is the same).
If instead we use the first one of eq (6.71) we have
which appears to be general for any (M,J) black hole, and is a direct consequence
of the first law. In the appendix H, we do the same type of analysis for a Kerr-
Newman black hole, meaning that three charges are involved.
This approach relies on a different observation for the three dimensional gravity.
Specifically, let us start from a warped AdS3 black hole [102] (appendices A.2,I),
which is a quotient of the warped AdS3 space, exactly the same way the BTZ is
a quotient of usual AdS3 . The difference of the vacuum space is a warp factor
ν. If this warp factor is set to 1, then one gets AdS3 . Similarly, setting ν = 1
transmogrifies the Warped black hole into a BTZ.
The observation here is that the Warped AdS3 black hole obeys the Kerr re-
lation (6.24). The contradiction is obvious; if we can get the BTZ as a limit of
the other black hole by changing the warp factor, how can the geometric relation
transform into (6.28) or (6.29)? The answer relies on the fact that we do not get
the BTZ in the same frame as the original solution. Instead, we have a rotating
frame at infinity with rotational velocity 1/l (see appendix I).
This gives rise to the idea that there is always a frame where we can have
a Kerr-type relation hold, if we need one. We can always perform such a frame
93
change, by means of
φ → φ − st , (6.76)
which only changes the shift vector N φ → N φ − s and leaves the lapse unchanged.
This means the rotational velocities will get shifted as
where i denotes the horizon root. This can of course be generalized for more angles
or rotational parameters. Lets consider the case where we have one rotation and
two horizons. Then it turns out that
Ω+ + s κ+
= =⇒ s = ΩR , (6.78)
Ω− + s κ−
depending on the definition we use for right or left. For BTZ, it turns out that
sB = 1/l as expected, while for Kerr sK = 0.
The CFT repercussion of finding a frame where the Kerr-relation holds is that
there is always a frame where the right entropy is independent of the angular
momentum.
Unfortunately, there are problems with this method. As discussed in chapter
4.1.2 and [31], the first law as a differential formula is very strict on the potentials
and charges that need to be used. However, one can still have a first law as long as
the appropriate transformations are performed. Adding a rotation at infinity will
result in a new first law, with potentially new charges, depending on the constancy
of the frame shift. They entropy will also change, as well as the rest of the charges.
It is not clear whether this is meaningful or physical, however it is worthwhile
examining merely as a method, in order to view the emergent relations.
But lets take a look at each situation separately with some analysis.
BTZ
Ω+ + s κ+
= (6.79)
Ω− + s κ−
yields
r+ + r−
s = 1/l Ω̂+ = . (6.80)
lr+
What we can try is to make sense of this shift in the gtt component. So for the
BTZ we have
r2
gtt = −N 2 + r2 (N φ )2 = − 2 + M , (6.81)
l
94
which diverges at infinity. If now we make the shift N φ → N φ − s, it will become
r2
gtt = − + M + r2 c2 + Jc , (6.82)
l2
which, plugging in the value s = 1/l cancels the divergent terms, yielding specifi-
cally
gtt = M + J/l . (6.83)
J → J − s−1 M . (6.86)
But inserting this in the first law is effectively a change in mass, i.e.
Ω̂
dM = T dS + Ω̂dJ − Ω̂s−1 dM =⇒ dM (1 + ) = T dS + Ω̂dJ . (6.87)
s
We can view it conversely as starting with a fist law of the form
dM = T dS + ΩdJ (6.88)
M → M + sJ . (6.90)
The reason for doing this is to show that the right sector indeed gets angular
momentum independent
p √
SR = M − J/l → M . (6.91)
On the other hand, if we used the transformation (6.86), we would get one sector
95
independent of the mass
p √
SL = M + J/l → J . (6.92)
So for the BTZ there seems to be some equivalence in this frame between mass
and angular momentum and the left and right sectors. Furthermore, with this new
mass the gtt component becomes
gtt = M 0 , (6.93)
making the whole space an ergosphere like the warped AdS3 black hole, and defines
a global Killing vector that does not diverge at infinity.
It is exactly this ambiguity of the Killing vector as a time translation, rendering
the BTZ ergosphere free, remedied by this coordinate transformation. This could
possibly imply a connection between the existence of the ergosphere and a Kerr-
type relation.
The problem with the BTZ black hole is that this coordinate transformation
is not that unique; it achieves too many things. First of all, it is the only frame
at which we can have a static observer at infinity. Also, it is the same coordinate
transformation performed for the near horizon geometry. Furthermore, it creates
an ergoregion as discussed earlier. Finally, it is constant which is pretty special
with this method.
Kerr
For Kerr, this equation yields s = 0. One could argue that this is expected since
the Killing vector is a time translation at infinity and any shift would change that
property. We will see though that in five dimensions this is no longer true.
Myers-Perry
We can see the complication here. Instead of retrieving the usual Myers-Perry
geometric potential relations (6.27) we get something totally different. Moving up
dimension-wise makes the relation non-unique, which is problematic.
96
Kerr-AdS
but also
X 1
=0. (6.98)
i
Ti
So adding any kind of parameter would still yield the same result. Of course if
we try to find a relation of the type
Ω+ + s κ+
= , (6.99)
Ω− + s κ−
where we use only the real horizons, we cannot conclude to any meaningful result
since Vieta’s theorem needs all roots to be applied. Also, for all the previous cases
the ratio of temperatures was just r− /r+ but now its much more complicated and
in order to do simplifications, we need Vieta’s theorem.
It is not clear whether this shift is constant or not, and as argued earlier, how
physical or relevant it is since it changes the meaning of the time translation Killing
vector at infinity for an asymptotically flat space.
97
7 Conclusions
We introduced general relativity, black holes and their thermodynamics and then
conformal field theories. We saw how two dimensional CFT duals are found for
higher dimensional black holes with Kerr/CFT, although not proven to be valid.
Finally we examined inner horizon thermodynamics and the two geometric re-
lations that seem enticing for the underlying CFT, as well as the gravitational
theory.
We presented a number of observations involving these potentials and at-
tempted some interpretations. The two geometric relations presented are quite
different, but seem to share some characteristics, such as the number of horizons
used. It would be useful to understand their connection if existent.
Specifically for the main relation issued by this project, the one involving tem-
peratures and angular velocities, as discussed, it is not clear if it is significant or
not, since it has repercussions on the entropy dependence on a quantized charge.
However, it can be further explored since (see appendix E) it can be viewed as a
Killing vector condition, or as a smoothness and patching condition as discussed.
Another reason for studying this type of relation further is to possibly create
CFT duals for multiple horizon solutions, which seems elusive without using the
near horizon approach. It seems interesting to explore and generalize this relation,
even maybe classify spacetimes depending on how it behaves. Unfortunately, re-
garding the CFT, we saw how it does not have a clear interpretation, except the
BTZ case, where it once again leads to the left-right sector charge equality.
Last but not least, it is alluring to go into more depth in the changing frame
procedure we used in the last subsection, since it produces relations and creates
further questions, including ergoregions and Killing vectors. We strongly believe
that these relations will play an important role in the understanding of black hole
physics and are worthwhile attending.
98
A Metrics and relations
In this section of the appendix we present all useful relations for various black hole
solutions (old and new relations) in a concise manner, as well as references for more
studying. Wherever it is needed, there is some analysis on some aspects, relevant
to the issue of this project. We hope that this is going to help the reader, since all
useful equations are concentrated in this appendix, as well as inspire him/her to
work more on these geometric relations. The following appendices are smaller and
cover various other issues, such as proofs, useful tools and some further analysis.
A.1 BTZ
For more details, see [50].
1 2
ds2BT Z = −N 2 dt2 + dr + r2 (N φ dt + dφ)2 , (A.1)
N2
where
J2 r2 J
N2 = 2 + 2 − M Nφ = − 2
4r l 2r
r r r r
1 2 1 2 1 2 J2
r± = M l (1 ± δ) = (l M + lJ) ± (l M − lJ), δ = 1− 2 2
2 4 4 M l
1 2 2
r+ r− = lJ r+ + r− = M l2
2
r + r2
2
2r+ r−
M= + 2 − J=
l l
4πµr∓ J r∓
S± = 4πr± − Ω± = 2 =
l 2r± lr±
P 2
Y 1 X Ai 1 2M X 1 2l2 M
Ωi = 2 Ωi = Q± = =
±
l ±
± Ai l J ± Ωi J
2 2
1 dN 2 r± − r∓ dN 2 J2 2r
T± = |r = , =− 3 + 2 .
4π dr ± 2πr± l2 dr 2r l
κA
Smarr [51] : + ΩJ = M .
16π
where the second term in the entropy formula is the Chern Simons term. Relations
99
we find for BTZ are (these hold for entropies if the Chern Simons term is ignored):
A+ Ω+ A− Ω− A+ A− Ω+ Ω−
+ =0 T+ A+ +T− A− = 0 + = −2l A− + A+
T+ T− T+ T− T+ T−
(A.3)
which lead to (through the first law for the three first and the last is inferred by
the first two):
Specifically, the area product is A+ A− = 32π 2 Jl and defining left and right areas
p
AR,L = A+ ± A− = 8πl M ± J/l .
Lets ignore the Chern Simons term and work with entropies now, assuming that it is
proportional to the area. We are trying to see if we can arrive to the correct entropy
by using the geometric relations (A.3). Or equivalently, see which assumptions are
2
needed to get the correct formula. The previous relation means SR,L have the
same (explicit) dependence on mass and opposite on angular momentum. So we
can write
2
p
SR,L = S+2 + S−2 ± 2S+ S− = af (M ) ± bg(J) =⇒ SR,L = af (M ) ± bg(J) ,
100
We also have from (A.3): ΩR + ΩL = 0 (see appendix C). We have now
∂SR,L a df ∂SR,L b dg
= =± (A.7)
∂M 2SR,L dM ∂J 2SR,L dJ
and (see appendix C for first laws with right and left quantities) using ΩR +ΩL = 0
which lead to
∂SR ∂SL ∂SR ∂SL
TR = 2 − TL TR = −2ΩR + TL .
∂M ∂M ∂J ∂J
SR ΩR SR
= A(M ) = B(J) =⇒ ΩR A(M ) = B(J) .
TR TR
TR SR SR 2J
dΩR = (dSR − dTR ) d = − 2 dΩR . (A.8)
2J TR TR TR
But
SR ∂ΩR F (M )TR2 ∂ΩR
d = dA = F (M )dM =⇒ = , =0.
TR ∂M −2J ∂J
So B does not depend on J, so its a constant.Hence g is a linear function of J. Then,
since ΩR does not depend on J, that means its derivative will also not depend on
J. So
∂ ∂ΩR ∂ F (M )TR2
= = 0 =⇒ F (M ) = 0. .
∂J ∂M ∂J −2J
Hence A is also a constant, and f is a linear function of M. Which is true, with
a = 64π 2 l2 and b = 64π 2 l. So we get the required result assuming that the first
law and Smarr is true, as well as using the two first equations of (A.3).
101
A second attempt
Here we will try to be more general and see what is the impact on the central
charges. Now, taking into account only the first equation from (A.3) , we have
2
SR,L = S+2 +S−2 ±2S+ S− = af (M )±bg(M, J) = aM ±b[cM +dJ] = (a±bc)M ±(bd)J ,
2
where in the third step we use the fact that SR,L is homogeneous of scale 1
with respect to M and J, so we can have only linear polynomial dependence.
Then we get (using the first law) the following relations:
which imply
SR SL SR SL
a= + bc = − bd = ∓(a ± bc)ΩR,L .
TR TL TR TL
Since their difference and their addition are independent of M and J, each one of
SR,L
TR,L
are independent of M and J. So using (A.8) , ΩR,L are also constants.
Using equations (A.3) fix some of the coefficients. The first one fixes cR2 = −cL2
, the second one cR1 = cL1 , and the third one cR1 + cL1 = cR2 − cL2 (taking l = 1
units) . Knowing that the coefficient of M and J is the same (that amounts to
ΩR = −ΩL = −1) :
2
SR,L = cR,L (M ± J)
leads to the mass independence, using only the first one of (A.3), since
102
and (A.3) imply cR = cL , which then means the mass independence through the
last equation.
N.B.: Adding the Chern Simons term in the entropy , relations (A.4) become
∂M S+ − ∂J S− = ∂M S− − ∂J S+ = 0 . (A.12)
103
A.2 Asymptotically Warped AdS3
For more details, see [102] [151]. The streched black holes ν 2 > 1 have a metric
l2 dr2
2 2 2 2
p
dsW AdS3 = l dt + 2 2
+ l 2νr − r+ r− (ν + 3) dtdθ
(ν + 3)(r − r+ )(r − r− )
l2 r
p
+ 3(ν − 1)r + (ν + 3)(r+ + r− ) − 4ν r+ r− (ν 2 + 3) dθ2
2 2
(A.13)
4
in Schwarzschild coordinates or
l4 dr2
ds2 = −N 2 (r)dt2 + l2 R2 (r)(dθ + N θ (r)dt)2 + , (A.14)
4R2 (r)N 2 (r)
where
2 r 2 2
p
2
R (r) = 3(ν − 1)r + (ν + 3)(r+ + r− ) − 4ν r+ r− (ν + 3) (A.15)
4
p
2 l2 (ν 2 + 3)(r − r+ )(r − r− ) θ 2νr − r+ r− (ν 2 + 3)
N (r) = N (r) = (A.16)
4R2 (r) 2R2 (r)
and ν is the Chern-Simons coupling related to the graviton mass by ν = µl/3. The
coordinates range as 0 ≤ r ≤ ∞, −∞ ≤ t ≤ ∞ and θ ∼ θ + 2π. Then we have
πl 2 2
p
S± = (9ν + 3)r± − (ν + 3)r∓ − 4ν (ν 2 + 3)r+ r− , (A.17)
24ν
π2 l
which can be written as a Cardy entropy S± = 3
(cL TL ± cR TR ) with
p
(ν 2 + 3)(r+ − r− ) ν2 + 3
(ν 2 + 3)r+ r−
TR = TL = r+ + r− − (A.18)
8πl 8πl ν
ν2 + 3 r± − r∓ 2
T± = p Ω± = − p .
4πl 2νr± − (ν 2 + 3)r+ r− (2νr± − (ν 2 + 3)r+ r−
(A.20)
N.B.: We employed two differences from [102]. First, we do not divide the angular
potentials by a factor of l as done there. Second, we take it to be Ω = −N θ instead
of plus, in consistency with the usual angular velocity definition we use throughout
the project.
The charges (ADT mass and angular momentum) are
ν2 + 3
1p 2
M= r+ + r− − (ν + 3)r+ r− (A.21)
24 ν
104
2
νl(ν 2 + 3) 5ν 2 + 3
1p 2 2
J= r+ + r− − (ν + 3)r+ r− − (r+ − r− ) (A.22)
96 ν 4ν 2
are used to form the first law
dM = T dS + ΩdJ (A.23)
M = T S + 2ΩJ . (A.24)
Notice the change in the Smarr formula. Furthermore, we obviously have the
relation
Ω+ Ω−
+ =0, (A.25)
T+ T−
which is reminiscent of the four dimensional case, and not the BTZ.
Note that the WadS3 black hole spacetime is rotating always, even with the
absence of angular momentum of the black hole . That means the whole space is
an ergoregion and there are no static observers, except of at infinity [198] [199].
105
A.3 KNTN
For more details, see [52] [133] [134] [135] [136].
∆ sin2 θ 2 2 dr
2
ds2KN T N = − (dt−P dφ)2
+ [(r +a 2
+n 2
)dφ−adt] 2
+ρ ( +dθ2 ) (A.26)
ρ2 ρ2 ∆
in BL or [53]
where n is the NUT parameter. Taking various limits leads to different solutions:
Q, n → 0 Kerr Q, a → 0Schwarzschild-NUT
Q, n, a → 0 Schwarzschild .
We have
2
∆ − a2 sin2 θ
2
gtφ gtφ
N = + , Nφ = (A.28)
ρ2 gφφ gφφ
∆ = r2 − 2M r + a2 + Q2 − n2 , a = J/M , ρ2 = r2 + (n + acosθ)2
P = asin2 θ − 2ncosθ ,
r+ r− = a2 + Q2 − n2 r+ + r− = 2M
a Qr±
A± = 4π(2M r± −Q2 +2n2 ) Ω± = Φ± =
2M r± + 2n2 − Q2 2M r± + 2n2 − Q2
Y Q4
Ai = (8π)2 [J 2 + + n2 (M 2 + n2 − Q2 )] mass indep for KN (n → 0)
±
4
X Y a2
Ai = 8π(2M 2 + 2n2 − Q2 ) Ωi =
±
±
4[J 2 + Q4 /4 + n2 (M 2 + n2 − Q2 )]
P
X ± Ai a(2M 2 + 2n2 − Q2 ) X X 4πa
Ωi = 4πa Q = Ai =
±
± Ai 2[J 2 + Q4 /4 + n2 (M 2 + n2 − Q2 ] ± ± Ωi
1 (N 2 )0 r± − r∓
T± = p |r± = 2
.
4π N 2 grr 4π(r± + a2 + n2 )
106
Smarr [52] : M = 2Γ± A± + 2Ω± J + Φq± Q + Φn± n (KNTN)
κ±
M= A± + 2Ω± J + Φ± Q (KN) . (A.29)
4π
κ±
dM = dA± + Ω± dJ + Φq± dQ (KN) , (A.31)
8π
A± 2
where Γ± = 2M (A±1−4πn2 )2 ( 16π − 4πJ 2 − πQ4 − n 2A± + 2πn2 Q2 ) is referred to as
the ”surface tension”. Parthapratim [52]claims that this is not equal to the usual
surface gravity factor, and hence the first law ”does not hold” for Kerr Newman
Taub NUT. Indeed if one solves the expression of area with respect to the mass
2
A± Q2
2 1 A± 2 2 2 2 2
M = + 4πJ + + πQ − n (A± − 4πn + 4πQ ) (A.32)
A± − 4πn2 16π 2
it is apparent that this does not equal the surface gravity over 8π, because there is
an n2 missing in the denominator. And there is obviously a contradiction: despite
the fact that T+ A+ + T− A− = 0, the product of areas is not mass independent.
There has to be a better way though to resolve this contradiction without giving
up on the first law; since Wald proved that the entropy is the charge that enters the
first law, it has to be that the entropy is not proportional to the area in this case,
so that the first law still holds. In [137] though, it is stressed how for non-zero NUT
charge the fixed points of the bifurcate horizons cannot be covered universally, and
one has to construct a maximal analytic manifold which is non-Hausdorff. These
wierd properties of the KNTN spacetime could have an impact on the first law,
since the existence of a bifurcate Killing horizon was crucial. Later we present
a the KNTN solution from other authors, in a way that it does not present this
problem. Relations for KN and KNTN are:
Ω+ Ω−
T+ A+ + T− A− = 0 + =0. (A.34)
T+ T−
107
For KN, where the first law is the usual one, these lead to mass independence of
the area product and
∂(A+ + A− )
=0. (A.35)
∂J
Defining
1 1 1 ΩR,L Ω+ Ω−
= ± = ± AR,L = A+ ± A−
TR,L T+ T− TR,L T+ T−
We get
∂AR
ΩR = 0 =0.
∂J
NB: These relations hold for Kerr as well.
Regarding the NUT parameter, in [136], the potentials agree with [52]. In [133]
[134] they use a slightly different metric
2
sin2 θ (n − acosθ)2
−1 2 ∆
ds2KN T N 2 2 2 2 2
= ρ (∆ dr +dθ )+ 2 [adt−(r +a )dφ] − 2 dt− a− dφ ,
ρ ρ a
(A.36)
2 2 2
where now ρ = r + (n − acosθ) and the electromagnetic field form is
(n − acosθ)2
Q 2 2
F = 4 [r − (n − acosθ) ]dr ∧ dt − a − dφ
ρ a
2Q
+ (n − acosθ)r sinθ dθ ∧ [adt − (r2 + a2 )dφ] . (A.37)
ρ4
The reason I am including this is because they reach to a different potentials and
area [134] [135]
r± − r∓ a 2
T± = 2
Ω± = 2
A± = 4π(r± + a2 ) (A.38)
4π(r± + a2 ) r± + a2
Qr± nr±
Φq± = 2
Φn± = 2
, (A.39)
r± + a2 r±+ a2
where Φn± are the potentials associated with the NUT charge, and arrive to Smarr
formulas and first laws [134]
κ±
M= A± + 2Ω± J + Φq± Q + Φn± (A.40)
4π
κ±
δM = δA± + Ω± δJ + Φq± δQ + Φn± δn . (A.41)
8π
108
They also compute the entropy [134] as
1
S± = [A± − A∓ ] . (A.42)
4
Thankfully, even if the expressions change, the relations
Ω+ Ω−
+ =0 T+ A+ + T− A− = 0 (A.43)
T+ T−
still holds. We can see the difference now, since the product of the areas
A+ A−
= n4 + 4J 2 − 2Q2 n2 (A.44)
16π 2
is indeed mass independent, in conjunction with the first law and the previous
equation.
109
A.4 5d Myers-Perry with two angular momenta
For more details, see [43] [44] [54].
r 2 ρ2 2 µ
ds2M P = −dt2 + dr + ρ2 dθ2 + 2 (dt − asin2 θdφ − bcos2 θdψ)2 +
∆ ρ (A.45)
2 2 2 2 2 2 2 2
(r + a )sin θdφ + (r + b )cos θdψ
in BL or [64]
where
gtψ gφψ − gψψ gtφ
N 2 = −gtt + gφφ (N φ )2 + gψψ (N ψ )2 + 2gφψ N φ N ψ , Nφ = 2
gφψ − gφφ gψψ
(this formalism also for doubly spinning BR later) and same with ψ ↔ φ for N ψ
3 Jφ 3 Jψ
a= b=
2M 2M
2
p
2r± = µ − a2 − b 2 ± (µ − a2 − b2 )2 − 4a2 b2
2 2 2 2
r+ r 3π(r+ + a)(r− + a2 )
b = 2− 2
M=
a 8a2
2π r± (r± + a )(r∓ + a2 )
2 2 2 2
1 (N 2 )0 a2 (r±
2 2
− r∓ )
A± = T± = |p |r± = 2 2
a2 4π grr N 2 2πr± (r± + a )(r∓ + a2 )
2
2
a π(r+ + a2 )(r−
2
+ a2 )
Ωφ± = 2 Jφ =
r± + a2 4a
2
ar∓ πr+ r− (r+ + a2 )(r−
2
+ a2 )
Ωψ± = 2
Jψ =
r± (r∓ + a2 ) 4a3
Y 4π 4 r+ r− (r+
2
+ a2 )2 (r−
2
+ a2 )2
Ai = = 16π 2 Jφ Jψ mass indep
±
a4
X 2π 2 (r+
2
+ a2 )(r−
2
+ a2 )(r+ + r− )
Ai =
± a2
Y a2
Ωφ,ψ i = Ωφ+ Ωφ− = Ωψ+ Ωψ− = 2 2
±
(r+ + a2 )(r− + a2 )
2
X a[(r+ + a2 ) + (r−
2
+ a2 )] X 2
a[r− 2
(r+ + a2 ) + r+
2 2
(r− + a2 )]
Ωφi = 2 2
Ωψi = 2 2
.
± (r+ + a2 )(r− + a2 ) ± r+ r− (r+ + a2 )(r− + a2 )
2 κ±
Smarr : M = Ωφ± Jφ + Ωψ± Jψ + A± .
3 8πG
110
κ±
First law : dM = dA± + Ωφ± dJφ + Ωψ± dJψ . (A.47)
8πG
where of course the last equation holds also when we swap plus with minus (or φ
with ψ). Defining
1 1 1 ΩR,L Ω+ Ω−
= ± = ± AR,L = A+ ± A− .
TR,L T+ T− TR,L T+ T−
We get
ΩφR ΩψR ΩφL ΩψL
+ =0 − =0,
TR TR TL TL
or
ΩφR = −ΩψR ΩφL = ΩψL , (A.49)
ΩR,L ψ φ
+ = Ω+ ± Ω+ ΩR,L ψ φ
− = Ω− ± Ω− ,
leading to
ΩR
− ΩR ΩL+ ΩL−
+ + =0 − =0. (A.50)
T− T+ T+ T−
Either one of (A.49,A.50) gives us
As an aside, we can also have different definitions for the plus and minus velocities
leading to two systems that independently remind us of the four dimensional case,
i.e.
ΩR,L ψ
+ = Ω+ ± Ω+
φ
ΩR,L φ
− = Ω− ± Ω− ,
ψ
(A.52)
where we just added a ± factor in the minus velocities. Then the equations become
more familiar and symmetric
ΩR ΩR ΩL+ ΩL−
+
+ − =0 + =0, (A.53)
T+ T− T+ T−
which essentially means that we have our potential relation satisfied independently
111
for two rotations, with the exception that we would have to use a coordinate
transformation to see the rotation planes since in these coordinates we need to mix
φ and ψ.
NB: Second eq from (A.48) has the same information if T+ r+ + T− r− = 0.
NB2: Equations still hold for a=b, but it all breaks down for b=0 since we have
one horizon.
Other interesting relations:
Ωφ+ Ωφ−
∂ A2+ + A2− a ∂ A2+ + A2−
a A+ A−
A+ + A− = + =⇒ = (A.54)
T+ T− µ T+ T− ∂Jφ µ ∂M
We also have,
And finally
Ωφ+ Ωφ−
1 1 1 ∂ A+ + A− 1 ∂ A+ + A−
+ = + =⇒ = (A.57)
T+ T− 4aµ T+ T− ∂Jφ 4aµ ∂M
112
A.5 KN-(A)dS4
For more details, see [55] [69].
∆r asin2 θ ρ2 2 ρ2 2 ∆θ sin2 θ r 2 + a2
ds2KN AdS4 = − (dt − dφ)2
+ dr + dθ + (adt − dφ)2
ρ2 Ξ ∆r ∆θ ρ2 Ξ
(A.59)
in BL or
ds2M P = −N 2 dt2 + γij (dxi + N i dt)(dxj + N j dt) , (A.60)
where
2 ρ2 ∆r ∆θ aΞ
N = 2
, N φ = − 2 [∆θ (r2 + a2 ) − ∆r ]
Σ Σ
Σ = (r + a ) ∆θ − a ∆r sin2 θ
2 2 2 2 2
Λ = −3/l2
r2 a2 2
∆r = (r2 + a2 )(1 + ) − 2mr + z 2 , ∆θ = 1 − cos θ .
l2 l2
a2
ρ2 = r2 + a2 cos2 θ Ξ=1− z 2 = qe2 + qm
2
l2
qe r qm cosθ 3
potential Ae = − √ e0 − √ e
ρ ∆r ρ ∆θ sinθ
m am qe qm
M= J= 2 Qe = Qm =
Ξ2 Ξ Ξ Ξ
2 r2 2 2
4π(ri2 + a2 ) 1 (N 2 )0 ri (1 + al2 + 3 l2i − a r+z
2 )
Ai = Ti = |p |ri = i
Ξ 4π grr N 2 4π(ri2 + a2 )
A π Q2 J 2 A A A2
Gen.Smarr : M2 = + (4J 2 + Q4 ) + + 2 + (Q 2
+ + ).
16π A 2 l 8πl2 4π 32π 2 l2
113
1 1
Smarr(if Λ variable): M = T S + ΩJ + ΦQ − ΘΛ
2 2
1 S 2 ΦS ΩJ
= T S + ΩJ + ΦQ − l ( + ).
2 2π πQ 1 + S/πl2
A2 A3 A4 A1 A3 A4 A2 A1 A4 A2 A3 A1
+ + + =0, (A.62)
T1 T2 T3 T4
which implies the mass independence of the area product(see appendix C for n
horizons) i.e. Q
∂ i Ai
=0. (A.63)
∂M
We also have
X A X Ω
( )i = 0 ( A)i = 0 , (A.64)
i
T i
T
which imply the mass and angular momentum independence of the sum of the area
squares respectively
∂ i A2i ∂ i A2i
P P
= =0 (A.65)
∂M ∂J
and finally
X 1 X Ω X Φ
( )i = 0 ( )i = 0 ( )i = 0 , (A.66)
i
T i
T i
T
which imply the mass, angular momentum and charge independence of the sum of
the areas respectively, i.e.
P P P
∂ i Ai ∂ i Ai ∂ i Ai
= = =0, (A.67)
∂M ∂J ∂Q
which means it only depends on the cosmological constant.
NB: The real and virtual horizons contribute exactly the opposite. The type of
the horizons is
where α, β, γ ∈ <.
Same relations apply for KNdS4, sending l → il.
114
Using Vieta’s theorem, we have
∆r = c4 r4 + c3 r3 + c2 r2 + c1 r + c0 , (A.68)
with
115
A.6 KN-(A)dS5
For more details, see [55] [69].
∆r asin2 θ ρ2 2 ρ2 2 ∆θ sin2 θ r 2 + a2
ds2KN AdS4 = − (dt − dφ)2
+ dr + dθ + (adt − dφ)2
ρ2 Ξ ∆r ∆θ ρ2 Ξ
(A.74)
in BL or
ds2M P = −N 2 dt2 + γij (dxi + N i dt)(dxj + N j dt) , (A.75)
where
2 ρ2 ∆r ∆θ aΞ
N = 2
, N φ = − 2 [∆θ (r2 + a2 ) − ∆r ]
Σ Σ
Σ = (r + a ) ∆θ − a ∆r sin2 θ
2 2 2 2 2
Λ = −3/l2
r2 a2 2
∆r = (r2 + a2 )(1 + ) − 2mr + z 2 , ∆θ = 1 − cos θ .
l2 l2
a2
ρ2 = r2 + a2 cos2 θ Ξ=1− z 2 = qe2 + qm
2
l2
qe r qm cosθ 3
potential Ae = − √ e0 − √ e
ρ ∆r ρ ∆θ sinθ
m am qe qm
M= J= 2 Qe = Qm =
Ξ2 Ξ Ξ Ξ
2 r2 2 2
4π(ri2 + a2 ) 1 (N 2 )0 ri (1 + al2 + 3 l2i − a r+z
2 )
Ai = Ti = |p |ri = i
Ξ 4π grr N 2 4π(ri2 + a2 )
A π Q2 J 2 A A A2
Gen.Smarr : M2 = + (4J 2 + Q4 ) + + 2 + (Q 2
+ + ).
16π A 2 l 8πl2 4π 32π 2 l2
116
Relations for KNAdS5 are(confirmed for various values on parameters):
6 Qn 3 Qn
X j6=i Aj X j6=i Aj
=0 =0, (A.76)
i=1
Ti i=1
Ti
which implies the mass independence of the area product(see appendix B for n
horizons),but also the mass independence of the product of the three areas, not
taking into account the opposite horizons, i.e.
∂ 3i Ai
Q
=0. (A.77)
∂M
So from now on, we will present only independences that involve these horizons,
since the ones that include all of them are true by identity.
Ωa,b
1 Ωa,b
2 Ωa,b
A1 + A2 + 3 A3 = 0 , (A.78)
T1 T2 T3
which implies angular momenta independence of the sum of the area squares
∂ 3i Ai ∂ 3i Ai
P P
= =0. (A.79)
∂Ja ∂Jb
117
A.7 KK-reduced 4d
For more details, see [56].
H3 p dr2 ∆
ds2KKred = − √ (dt + B)2 + H1 H2 ( + dθ2 + sin2 θdφ2 ) , (A.80)
H1 H2 ∆ H3
where r
H2
dilaton: e−2Φ4 =
H1
√
horizons: r± = m ± m2 − a2
p (p − 2m)(q − 2m)
H1 = r2 + a2 cos2 θ + r(p − 2m) +
p+q 2
p p
− (q 2 − 4m2 )(p2 − 4m2 )acosθ (A.81)
2m(p + q)
q (p − 2m)(q − 2m)
H2 = r2 + a2 cos2 θ + r(q − 2m) +
p+q 2
q p
+ (q 2 − 4m2 )(p2 − 4m2 )acosθ (A.82)
2m(p + q)
Z Z q
√
A= gθθ gφφ dθdφ = −H3 Bφ2 dθdφ (A.86)
√ s
A± m pq pq pq + 4m2 2
S± = = 2π[± + ( ) − J 2] (A.87)
4 2G4 16G24 p + q
118
4m(p + q)
Ω± = √ a (A.88)
pq(4m2 (p + q) + (pq + 4m2 )(r± − r∓ ))
pq + 4m2 4m2
1
β± = = + . (A.89)
T± p+q r+ − r−
Quantization of the electric and magnetic charges:
2P Q
Q = 2G4 M0 nQ P = 2G4 M6 nP = nQ nP since 8G4 M0 M6 = 1
G4
(A.90)
Y
2 2 2
n2Q n2P
Si = P Q − J = − J 2 = NL − NR mass indep , (A.91)
i
4
where
pq pq + 4m2 2 2 m2 pq
NL = ( ) − J N R = (A.92)
16G24 p + q 4G24
s
X pq pq + 4m2 2
Si = 4π ( ) − J2 (A.93)
i 16G24 p + q
X πa √
= 2πm pq . (A.94)
i Ωi
NB: In [56] it is noted that NL and NR are taken like this by convention. If we
P √
change them then the sum of entropies is i Si = 2πm pq which equals to the
sum of inverse angular velocities.
Smarr : 2T± S± = M − ΦE M
± Q − Φ± P − 2Ω± J . (A.95)
119
A.8 4d pairwise equal
For more details, see [57] [58] [59].
where
r12 + a2 cos2 θ r1 (r1 − r2 )
eφ1 = =1+ ri = r + 2ms2i
W W
∆θ = 1 − g 2 a2 cos2 θ W = r1 r2 + a2 cos2 θ Ξ = 1 − a2 g 2 ,
which has four distinct horizons. Taking non rotating at infinity coords φ0 =
φ + ag 2 t
∆r ∆θ 2 dr2 dθ2
ds2pairweq = − dt + Bsin 2
θ(dφ + f dt) 2
+ W ( + )
BΞ2 ∆r ∆θ
2ms1 c1 r2 2ms2 c2 r1
Φ1 = Φ2 = Φ3 = Φ4 = ,
r1 r2 + a2 r1 r2 + a2
whereci = coshδi . Charges:
m m
E= 2
(1 + s21 + s22 ) = (cosh2δ1 + cosh2δ2 )
Ξ 2Ξ2
ma ma
J=2
(1 + s21 + s22 ) = (cosh2δ1 + cosh2δ2 )
Ξ 2Ξ2
ms1 c1 m ms2 c2 m
Q1 = Q2 = = sinh2δ1 Q3 = Q4 = = sinh2δ2
2Ξ 4Ξ 2Ξ 4Ξ
Y
Si = (4π 2 )g −4 (8πJ)2 +4g −4 (4πQ1 )2 (4πQ2 )2 mass indep- 4 horizon solutions .
i
X
First law : dE = T dS + ΩdJ + Φi dQi .
i
120
Relations for 4d pairwise equal charges are again:
X Ω
( )i = 0 (tested for various values of parameters) , (A.99)
i T
121
A.9 5d minimal gauged supergravity
For more details, see [58] [143].
2
∆θ [(1 + g 2 r2 )ρ2 dt + 2qν]dt 2qνω
f ∆θ dt
ds25dM G =− + 2 + 4 −ω
Ξa Ξb ρ2 ρ ρ Ξa Ξb
ρ2 dr2 ρ2 dθ2 r2 + a2 2 2 r 2 + b2 2
+ + + sin θdφ + cos θdψ 2 , (A.100)
∆r ∆θ Ξa Ξb
where
dφ dψ
ν = bsin2 θdφ + acos2 θdψ ω = asin2 θ + bcos2 θ
Ξa Ξb
122
I also found that
3
X Ωia,b
Ai = 0 , (A.106)
i
Ti
which means P3
∂ i A2i
=0, (A.107)
∂Ja,b
which is the same as the Kerr-AdS5 case(and a generalization of the BTZ).
123
A.10 6d gauged supergravity
For more details, see [58] [144].
2 2 2 2
2 1/2 (r + y )(r + z ) (r2 + y 2 )(y 2 − z 2 ) 2 (r2 + z 2 )(z 2 − y 2 ) 2
ds6dGS = H dr2 + dy + dz
R Y Z
RB 2
−
H 2 (r2 + y 2 )(r2 + z 2 )
2
Y 0 2 2 2 2 qrB
+ 2 dt + (z − r )dψ1 − r z dψ2 −
(r + y 2 )(y 2 − z 2 ) H(r2 + y 2 )(r2 + z 2 )
2
Z 0 2 2 2 2 qrB
+ 2 dt + (y − r )dψ1 − r y dψ2 − ,
(r + z 2 )(z 2 − y 2 ) H(r2 + y 2 )(r2 + z 2 )
(A.108)
where
and we have again two independent rotations, and now six horizons. The thermo-
dynamic quantites are
2πma(1 + Ξb s2 ) 2πmsc
Ja = 2
Jb = Ja↔b Q=
3Ξa Ξb Ξa Ξb
π 1 1 Ξa Ξb
M= 2m + +q 1+ + . (A.111)
3Ξa Ξb Ξa Ξb Ξb Ξa
The product of the areas is independent of the mass [58]
6
Y 8π 2 2 8πQ 6
Ai = g −8 (8πJa )2 (8πJb )2 + g −6 . (A.112)
i
3 3
124
The first law is satisfied [144]
δM = T δS + Ωa δJ a + Ωb δJ b + ΦδQ . (A.113)
The fact that the product of the areas is mass independent is confirmed by the
fact that
6
X 1 Y
Aj = 0 . (A.114)
i
T i
j6=i
which translate to P6
∂ 6i Ai ∂ 6i Ai
P P
∂ Aii
= = =0 (A.116)
∂M ∂Ja,b ∂Q
and furthermore
6 6
X Ai X Ωia,b
= Ai = 0 , (A.117)
i
Ti i
Ti
which also mean P6
A2i ∂ 6i A2i
P
∂ i
= =0, (A.118)
∂M ∂Ja,b
which are exactly the same relations with Kerr-Newman AdS4 generalized for six
horizons.
125
A.11 7d gauged supergravity
For more details, see [175] [58].
RB 2
−
H 2 (r2 + y 2 )(r2 + z 2 )
2
Y 0 2 2 2 2 qB
+ 2 dt + (z − r )dψ1 − r z dψ2 −
(r + y 2 )(y 2 − z 2 ) H(r2 + y 2 )(r2 + z 2 )
2
Z 0 2 2 2 2 qB
+ 2 dt + (y − r )dψ1 − r y dψ2 −
(r + z 2 )(z 2 − y 2 ) H(r2 + y 2 )(r2 + z 2 )
a21 a22 a23
+ 2 2 2 dt0 + (y 2 + z 2 − r2 )dψ1 + (y 2 z 2 − r2 y 2 − r2 z 2 )dψ2 − r2 y 2 z 2 dψ3
r y z
2
q gy 2 z 2
− 1+ B , (A.119)
H(r2 + y 2 )(r2 + z 2 ) a1 a2 a3
where
3
1 + g 2 r2 Y 2 2 2 2 2 2 2 2gqa1 a2 a3 q 2 g 2
R= (r + a ) + qg (2r + a 1 + a 2 + a 3 ) − + 2 − 2m
r2 i=1
r2 r
3 3
1 − g2y2 Y 2 1 − g2z2 Y 2
Y = (ai − y 2 ) Z= (ai − z 2 )
y2 i=1
z2 i=1
q
H =1+ q = 2ms2 s = sinh δ c = cosh δ
(r2 + y 2 )(r2 + z2)
B = dt0 + (y 2 + z 2 )dψ1 + y 2 z 2 dψ2 Ξi = 1 − a2i g 2 (A.120)
and there are three independent rotations. Solving R = 0 gives eight horizons,
pairwise opposite (as in the 3d and 5d case). So we want to use only four of them.
The thermodynamical quantities are
π 3 [(rH
2
+ a21 )(rH
2
+ a22 )(rH
2
+ a23 ) + q(rH
2
− a1 a2 a3 g)]
SH =
4Ξ1 Ξ2 Ξ3
2
2mscrH
ΦH = 2
(rH + a21 )(rH 2
+ a22 )(rH2
+ a23 ) + q(rH 2
− a1 a2 a3 g)
(1 + g 2 rH
2 2 2 2 2 2 2 4 2 2
P Q Q
)rH i j6=i (rH + aj ) − i (rH + ai ) + 2q(g rH + ga1 a2 a3 ) − q g
THi = 2
2πrH [(rH + a21 )(rH2
+ a22 )(rH 2
+ a23 ) + q(rH2
− a1 a2 a3 g)]
ai [(1 + g 2 rH
2 2
+ a2j ) + qg 2 rH 2
Q Q
i
) j6=i (rH ] − q j6=i aj g
ΩH = 2
, (A.121)
(rH + a21 )(rH2
+ a22 )(rH2
+ a23 ) + q(rH 2
− a1 a2 a3 g)
126
where the H takes values from the set of horizons. The largest value is the outer
event horizon. Not to be confused with the i index that takes values from the set
of the three independent angular momenta. The charges are
π2 2(1 + 2a1 a2 a3 g 3 )
X
2m 5q q X X 2Ξj
M= −m+ + − Ξi −
8Ξ1 Ξ2 Ξ3 i Ξi 2 2 i j6=i
Ξi Ξi
Unfortunately, lack of time and computational power did not let us find any useful
relations for this one. It is obvious though, that the mass independence of the area
product will reflect on the first law.
127
A.12 EMDA
For more details, see [60].
∆ − a2 sin2 θ 2 2asin2 θ 2
ds2EM DA = − dt − [(r + a2 − 2Dr) − ρ2 ]dtdφ
ρ2 ρ2
(A.125)
ρ2 sin2 θ 2
+ dr2 + ρ2 dθ2 + [(r + a2 − 2Dr)2 − ∆a2 sin2 θ]dφ2 ,
∆ ρ2
where
W ω
e2Φ = = (r2 + a2 cos2 θ) ω = e2Φ0
∆ ∆
∆ = r2 − 2mr + a2 ρ2 = r2 − 2Dr + a2 cos2 θ
2 a
S± = π(r± + a2 − 2Dr± ) Ω± = 2
r± + a2 − 2Dr±
2
Qr± r± − a2
ΦQ
± = 2
T± = 2
.
ω(r± + a2 − 2Dr± ) 4πr± (r± + a2 − 2Dr± )
√
Horizons at r± = m ± m2 − a2 . Charges:
M =m−D J = a(m − D)
p
Q= 2ωD(D − m) P =g
Y
Si = 4πa2 (m − D)2 = 4π 2 J 2 mass indep
i
X 4πmJ
Si =
i a
X πa 4πmJ
= .
i Ωi a
1 1
Smarr : M = T S + ΦQ + ΩJ .
2 2
First law : dE = T dS + ΦQ dQ + ΩdJ . (A.126)
128
A.13 Kaluza Klein stationary
For more details, see [60].
1 − Z 2 2aZsin2 θ Bρ2 2
ds2KKstat = − dt − √ dtdφ + dr + Bρ2 dθ2
B B 1 − v2 ∆r (A.128)
Z
+ B(r2 + a2 ) + a2 sin2 θ sin2 θdφ2 ,
B
where
2mr v 2 Z 2
Z= 2 B = 1+
ρ 1 − v2
ρ2 = r2 + a2 cos2 θ ∆r = r2 + a2 − 2mr ,
2
√
(r± − a2 ) 1 − v 2 Q Qr± (1 − v 2 )
T± = 2
Φ± = 2
.
4πr± (r± + a2 ) r± + a2
√
Horizons at r± = m ± m2 − a2 . Charges:
v2 ma mv
M = 1+ 2
m J=√ Q=
2(1 − v ) 1 − v2 1 − v2
Y 4π 2 a2 m2
Si = = 4π 2 J 2 mass indep
i
1 − v2
X 2πmJ
Si =
i a
X πa 2πmJ
= .
i Ωi a
Relations for KK are:
X Ω
( )i = 0 T+ S+ + T− S− = 0 , (A.129)
i T
129
A.14 Sen black hole in heterotic string
For more details, see [60] [181].
ds2Sen = ∆1/2 − ∆−1 (r2 + a2 cos2 θ − 2mr)dt2 + (r2 + a2 − 2mr)−1 dr2 + dθ2
where
r± − r∓
T± = .
2π(r± + r∓ )(coshα + coshβ)r±
√
Horizons at r± = m ± m2 − a2 . Charges:
1 1
M = m(1 + coshα coshβ) J = ma(coshα + coshβ)
2 2
m
Qa = √ sinhα coshβ na 1 ≤ a ≤ 22
2
Y
Si = 4πa2 (m − D)2 = 4π 2 J 2 mass indep
i
X 4πmJ
Si =
i a
X πa 4πmJ
=
i Ωi a
Relations for SEN are:
X Ω
( )i = 0 T+ S+ + T− S− = 0 , (A.131)
i T
130
A.15 Doubly spinning neutral black ring
For more details, see [54] [62] [63] [64].
with p
√
2kλ (1 + ν)2 − λ2 1+y
Ω=− (1 − x2 )y νdψ +
H(y, x) 1−λ+ν
2
×[1 + λ − ν + x yν(1 − λ − ν) + 2νx(1 − y)]dφ
√
−λ± λ2 −4ν
Horizons: 1 + λy + νy 2 = 0 =⇒ y± = 2ν
√
2 (1 + λ + ν)λ
3 (1/y± − y± )(1 − ν) λ2 − 4ν
A± = 32π R T± = ±
(1 − ν)2 (1/y± − y± ) 8πRλ(1 + ν + λ)
√ r r
λ(1 + ν) ∓ (1 − ν) λ2 − 4ν 1 + ν − λ 1 1+ν−λ
Ωφ± = √ Ωψ± = .
4Rλ ν 1+ν+λ 2R 1+ν+λ
(A.133)
Charges:
√ r
3πR2 λ 4πR3 λ ν 1+ν+λ
M= Jφ = 2
G5 1 − λ + ν G5 (1 − ν) 1+ν−λ
r
2πR3 λ(1 + λ − 6ν + λν + ν 2 ) 1 + ν + λ
Jψ = . (A.134)
G5 (1 − ν)2 (1 + ν − λ) 1+ν−λ
131
NB: Jφ and Jψ are swapped in [62]. I am consistent with [63].
Y
Ai = 4π 2 (4G5 )2 Jφ2 mass indep (A.135)
i
X 32π 2 R3 λ2 32π 2 R3 λ2
Ai = = . (A.136)
i 1 − λ + λν − ν 2 (1 − ν)(1 + ν − λ)
2 A±
Smarr: M = T± + Ωφ± Jφ + Ωψ± Jψ .
3 4G5
1
First law : dM = T± dA± + Ωφ± dJφ + Ωψ± dJψ . (A.137)
4G5
Relations for doubly spinning BR are:
Ωψ+ Ωψ− AR ∂ψ AL
+ = 0 =⇒ = , (A.139)
T+ A+ T− A− AL ∂ψ AR
or
∂(A2R − A2L ) ∂(A+ A− )
= 0 =⇒ =0, (A.140)
∂Jψ ∂Jψ
which obviously holds (A.135).
(1 − ν)(λ2 − 4ν)
T+ y+ + T− y− = (A.141)
8πRν 2 (1 + ν + λ)
1 −AR TL
ΩφL Jφ = − T+ A+ Jφ = . (A.142)
2 2ΩφL
132
A.16 Dipole black ring
For more details, see [63] [65].
N/3 2
F (y) H(x) 1+y
ds2dipoleBR
=− dt + C(ν, λ)R dψ
F (x) H(y) F (y)
N/3
R2 dy 2 dx2
2
G(y) 2 G(x) 2
+ F (x) H(x) H(y) − dψ − + + dφ ,
(x − y)2 F (y)H(y)N G(y) G(x) F (x)H(x)N
(A.143)
where
F (ξ) = 1 + λξ G(ξ) = (1 − ξ 2 )(1 + νξ)
r
1−µ
H(ξ) = 1 − µξ C(ν, −µ) = −µ(−µ − ν) .
1+µ
4 4
N is related to the dilaton coupling a2 = N
− 3
, 0 < N ≤ 3. From now on we
set N=3 (dilation free).
which result in
(1 + µ)3 p (1 + µ)3 p
A+ = 8π 2 R3 λ(µ + ν)3 (1 − λ2 ) A− = 8π 2 R3 (λ − ν)µ3 (1 − λ2 ) .
(1 + ν)(1 − ν)2 (1 − ν)2
133
NB: One extra λ in [63] in the inner horizon area.
s s
ν(1 + ν) 1−λ ν 1−λ
T+ = T− = −
4πR (1 + λ)λ(ν + µ)3 4πR (1 + λ)(λ − ν)µ3
s s
1 λ−ν 1−ν λ
Ω+ = Ω − =
R (1 + λ)λ(1 + µ)3 R (1 + λ)(λ − ν)(1 + µ)3
p
3 2πR3 2/3 (1 + µ) µ(1 − µ)(1 − λ)
Φ+ = √
4R G5 µ+ν
s
3 2πR3 2/3 1 + µ (µ + ν)(1 − µ)(1 − λ)
Φ− = . (A.145)
4R G5 1−ν µ
Charges:
p
3πR2 (1 + µ)2 πR3 (1 + µ)9/2 λ(λ − ν)(1 + λ)
M= (λ + µ) J=
4G 1 − ν 2G (1 − ν)2
s
2πR3 1/3 1 + µ µ(µ + ν)(1 − λ)
Q= . (A.146)
G5 1−ν 1−µ
NB: Some differences in coefficients between [63] and [65] but it is a matter of
definition. I follow [63].I also add a minus sign to the inner horizon temperature(so
that (A.148) comes with the same sign as always).
Y
Ai = (4G5 )2 (4π 2 )JQ3 mass indep
i
2 3 3
p p
X 8π R (1 + µ) (ν + 1) µ3 (1 − λ2 )(λ − ν) + λ(1 − λ2 )(µ + ν)3
Ai = .
i (ν − 1)2 (ν + 1)
3 A± 1
Smarr: M = T± + Ω± J + Φ± Q .
2 4G5 2
dA±
First law : dM = T± + Ω± dJ + Φ± dQ . (A.147)
4G5
Relations for dipole BR are:
T+ A+ + T− A− = 0 . (A.148)
134
A.17 5d sugra CS topological BH
For more details, see [66] [67].
where
l2
dΩ3 = −cos2 θdt2 + 2
(dθ2 + sin2 θdχ2 ) N 2 = (r − r+
2 2 2
) /l
r+
Y Y
Ai = Si = 16π 2 r+ r− = 8π 2 M l2 J indep (A.150)
i i
X X
Ai = Si = 4π(r+ + r− ) = 4πl(J − lM )
i i
X X
A2i = Si2 = 16π 2 lJ mass indep (A.151)
i i
We can see that there is a swapping between J and M , which is expected since the
corresponding Killing vectors also swap regarding the charges [68] (if considered
as a generalization of 3d topological gravity).
135
Relations that they obey are:
2 2 2
(r+ − r− )
T+ A+ + T− A− = 0 BUT T+ S+ + T− S− = − 2 6 0
= (A.152)
2l r+ r−
Ω+ Ω− Ω+ Ω−
+ =0 A+ + A− = 0 . (A.153)
T+ S+ T− S− T+ T−
So these imply through the ”first laws” :
136
B Useful tools
Some examples that are used in proofs are the Lie derivatives of scalars, vectors
and dual vectors
£ξ φ = ξ µ ∂µ φ = ξ µ φ,µ , (B.3)
where in all of the above the partial derivative can be changed with a covariant
one, due to antisymmetry of the connection. Killing vectors in GR also satisfy
∇ν ∇ρ ξ µ = Rρνσ
µ
ξσ , ξ µ = ∇ν ∇ν ξ µ = −Rµν ξ ν . (B.6)
The important take home feature is that Killing vectors and fields are associated
with symmetries. Linear combinations of Killing vectors are also Killing vectors.
• Weak energy condition: For any timelike vector v µ , it is true that Tµν v µ v ν ≥
0, which means that the energy density is non-negative in any frame . If the
vector is null, it is called null energy condition.
1
• Strong energy condition: For any timelike vector, Tµν v µ v ν ≥ 2
T v µ vµ or
(Tµν − 12 gµν T )v µ v ν ≥ 0.
137
• Generic energy condition: If the strong EC holds, and every causal geodesic
contains a point where l[a Rb]cd[e lf ] lc ld 6= 0 , which shows some nonalignment
of the curvature with geodesics, so implies some kind of convergence or di-
vergence.
• Dominant energy condition: For each future pointing causal vector v µ , the
vector Tµν v µ is causal and past pointing.
138
C First law relations and byproducts
For two horizons:
i
Ω+ S− Ωi− S+
S+ S− Φ+ S− Φ− S+
d(S+ S− ) = + dM − + dJi − + dQ (C.1)
T− T+ T+ T− T+ T−
i
Ω+ S+ Ωi− S−
1 S+ S−
d S+2 + S−2 =
+ dM − + dJi −
2 T+ T− T+ T−
j (C.2)
Φ+ S+ Φj− S−
+ dQj
T+ T−
j
Ωi+ Ωi− Φ+ Φj−
1 1
dM + = d(S+ + S− ) + + dJi + + dQj (C.3)
T+ T− T+ T− T+ T−
SR SL
T+ S+ + T− S− = 0 =⇒ = (C.6)
TR TL
S + Ω+ S − Ω− SR ΩR SL ΩL
+ = 0 =⇒ + =0. (C.8)
T+ T− TR TL
139
Generalization for n horizons
Y n
S2 S3 ...Sn S1 S3 ...Sn
d Si = d(S1 S2 ...Sn ) = (n − 1) dM + + ... −
i
T 1 T 2
X Ωl Ωl2
1
dJl (S2 S3 ...Sn ) + (S1 S3 ...Sn ) + ... −
l
T1 T2
l
Φl2
X Φ1
dQl (S2 S3 ...Sn ) + (S1 S3 ...Sn ) + ...
l
T1 T2
( n " n n n
#)
Y Sj XY Ωli Φli
X XY
= (n − 1) dM − Sj dJl − Sj dQl .
i=1 j6=i
Ti l j6=i
Ti l j6=i
Ti
(C.10)
Xn n
1 2 1 2 2 2
X
d Sn = d(S1 + S2 + ...Sn ) = S1 dS1 + S2 dS2 + ... + Sn dSn = Sn dSn =
2 i
2 i
X l
Ωl2
l
Φl2
X
S1 S2 Ω1 Φ1
= dM + + ... − dJl S1 + S2 + ... − dQl S1 + S2 + ... =
T1 T2 l
T 1 T2
l
T1 T2
n l l
X Sn X Ωn X Φn
= dM − Sn dJl − Sn dQl .
i
T n
l
Tn
l
Tn
(C.11)
n
X
d Sn = d(S1 + S2 + ...Sn ) = dS1 + dS2 + ...dSn =
i
X l
Ω1 Ωl2
l
Φ1 Φl2
X
1 1
= dM + + ... − dJl + + ... − dQl + + ...
T1 T2 l
T1 T2
l
T1 T2
n l l
X 1 X Ωn X Φn
= dM − dJl − dQl .
i
Tn
l
T n
l
T n
(C.12)
140
D KNTN temperature calculation procedure
Here we show how to calculate the temperature of the Kerr-Newman-Taub-NUT
black hole, in the metric given in [52].
After foliating space time ADM-like, we have
and yields
gtφ
Nφ = gtφ N φ gφφ = gtφ =⇒ N φ =
gφφ
2
gtφ
gtt = −N 2 + N φ Nφ = −N 2 + ,
gφφ
These obey
2
∆ − a2 sin2 θ gtφ
2 a
N |r± = + =0 N φ |r± = − = −Ω± .
ρ2 gφφ r± 2M r± + 2n2 − Q2
(D.3)
Since this holds, we have
2 2
[gtφ ρ − gφφ a2 sin2 θ]|r± = 0 .
All together
(N 2 )0 x0 ∆1/2 (∆xgφφ )1/2 (ρ2 gφφ )0
p = 1/2 1/2
− .
N 2 grr ρ2 gφφ x (ρ2 gφφ )2
The second term will vanish on the two horizons and only the first will survive ,
since there it is indeterminate. Using L’ Hospital’s rule we get
1/2
1 (N 2 )0 1 x0 ∆0
r± − r∓
= = 2
. (D.4)
1/2 x0
p
4π N 2 grr r± 4π ρ2 gφφ r± r± 4π(r± + a2 + n2 )
141
E Killing vector computations
The norm of the Killing vector generating a horizon when we have one angular
momentum is
ξ 2 = gµν k µ k ν + 2gµν ΩH k µ mν + gµν Ω2H
Notable relations
In 3d and 4d, we have on the horizon
H
ξ 2 = gtt + 2ΩH gφt + Ω2H gφφ = −N 2 , (E.2)
or, this will hold everywhere if we have a vector(which is not Killing, but coincides
with it on the horizon)
χµ = k µ − N φ mµ . (E.3)
(N 2 )0
(ξ 2 ),µ nµ = p = κH , (E.4)
N 2 grr
E.1 Kerr
The norm of the Killing vector in Kerr is
which is obviously spacelike outside the horizon, timelike between outer and inner
and again spacelike inside the inner.
Using the linear combination of the Killing vectors that generate the horizons,
normalized as to have unit surface gravity (I will name them ζ)
1 µ Ω+ µ 1 µ Ω− µ
lµ = ζ+µ + ζ−µ = k + m + k + m (E.6)
κ+ κ+ κ− κ−
and using the geometric relation with a plus sign, the angular terms will vanish
κ+ + κ− µ
lµ = k = fKerr k µ . (E.7)
κ+ κ−
I will call this the surface parameter f . This Killing vector has a norm at infinity
2
l∞ = −fKerr = −(r+ + r− ) < 0 (E.8)
142
and of course vanishes on the ergosurfaces, where gtt = 0, or
√
re = M ± M 2 − a2 cos2 θ . (E.9)
but we cannot connect it with angular velocities or temperatures since the ergo-
surface radius cannot be expressed in terms of inner/outer horizons(on par with
the BTZ).
A covariant way of writing the geometric potential relation is
Ω+ Ω−
=± =⇒ ζ+µ pµ = ±ζ−µ pµ , (E.11)
T+ T−
depending on the sign we choose for the temperature (the plus sign if we take the
absolute value for the inner horizon temperature), and where pµ is defined as
pµ kµ = 0 pµ mµ 6= 0 . (E.12)
In coordinates where k µ = ∂t :
pr gtt pθ
pµ = pt (1, , − , ) (E.13)
pt gφt pt
and pt , pr , pθ are arbitrary. The problem in these coordinates is that the second
condition means, if mµ = ∂φ
gφt
pφ 6= −pt , (E.14)
gφφ
2
which cannot be true on the horizon, where gφt = gtt gφφ .
E.2 BTZ
The norm of the Killing vector generating the horizon in BTZ is
r2 J2
ξ2 = M − − = −N 2 , (E.15)
l2 4r2
which now is timelike outside the horizon, spacelike in between and again timelike
inside the inner, which are the opposite than Kerr.
For the BTZ the same linear combination of the Killing vectors, normalized to
have unit surface gravity
κ+ + κ− µ mµ
lµ = (k − ) (E.16)
κ+ κ− l
143
has a norm at infinity
2 l
l∞ = (M − J/l)fBT Z = (lM − J) , (E.17)
r+ + r−
which is not always timelike.
The ergosurface is at gtt = 0 or
re2 = M l2 = r+
2 2
+ r− . (E.18)
ξ2 e+
= 2gφt ΩH + gφφ Ω2H . (E.19)
2 κ2+
ξ+ e
= −lΩ+ κ+ r− = r− (E.20)
κ−
and the inner horizon one
2 κ2−
ξ− e+
= −lΩ− κ− r+ = r+ . (E.21)
κ+
So if we normalize them to have unit surface gravity they will be
r− r+
ζ+2 e+
= ζ−2 e+
= . (E.22)
κ− κ+
A covariant way of writing the geometric potential relations is
Ω+ 1 ζ+µ pµ ζ−µ qµ
=± =⇒ =± (E.23)
T+ T− m·p k·q
Ω− 1 ζ µ pµ ζ µ qµ
=± =⇒ − = ± + , (E.24)
T− T+ m·p k·q
depending on the sign we choose for the temperature (the plus sign if we take
the absolute value for the inner horizon temperature), and where pµ has the same
definition with the Kerr case and now q µ is defined as
q µ mµ = 0 q µ kµ 6= 0 . (E.25)
We can see that now the inner products enter the equations, since there is a mixing
between these two vectors, due to the time- angle mixing. In coordinates where
mµ = ∂φ we have
qr gφt
q µ = q t (1, t , − ) (E.26)
q gφφ
144
and q t , q r arbitrary. We have the same problem as before, if k µ = ∂t
gtt
q φ 6= −q t , (E.27)
gφt
E.3 5d Myers-Perry
For the Myers Perry the Killing vectors generating the horizons are
µ
ξ± = k µ + Ωφ± φµ + Ωψ± ψ µ , (E.28)
where I denote φµ , ψ µ the Killing vectors associated with the rotations along φ, ψ
respectively. Normalized as to have unit surface gravity they become
1 µ Ωφ± µ Ωψ± µ
ζ±µ = k + φ + ψ . (E.29)
κ± κ± κ±
The geometric potential relations are
ΩR,L ΩR,L
+
± − =0, (E.30)
κ+ κ−
where we take the definition of the surface gravity as κ− = κr+ →r− and
ψ φ
ΩR
± = Ω± + Ω± ΩL± = Ωψ± − Ωφ± . (E.31)
pµ kµ = 0 pµ φµ = pµ ψµ 6= 0 (E.32)
q µ kµ = 0 q µ φµ = −pµ ψµ 6= 0 , (E.33)
we have
Ωψ± + Ωφ± ΩR
ζ±µ pµ = (p · ψ) = ± (p · ψ) (E.34)
κ± κ±
Ωψ± − Ωφ± ΩL
ζ±µ qµ = (p · ψ) = ± (p · ψ) , (E.35)
κ± κ±
so that
ζ+ · p + ζ− · p = 0 ζ+ · q − ζ− · q = 0 . (E.36)
1
pµ kµ = 0 =⇒ pt = − (gtφ pφ + gtψ pψ ) . (E.37)
gtt
145
From now on, I am going to use plus minus signs, where the bottom symbol will
hold for p and the top for q.
2
gψψ gtt − gψt ± (gφψ gtt − gtφ gtψ ) A±C
pµ φµ = ∓pµ ψµ =⇒ pφ = ∓pψ 2
= ∓pψ
gφφ gtt − gφt ± (gφψ gtt − gtφ gtψ ) B±C
and so
ψ φ
t ψ 2 ±(gψt gφψ − gφt gψψ ) + (gψt gφφ − gφt gφψ ) ψN ∓ N
p = −p (gφψ − gφφ gψψ ) 2
= Dp ,
gφφ gtt − gφt ± (gφψ gtt − gtφ gtψ ) B±C
where i used the definition for the shift as shown in (A.4) labeled these quantities
A, B, C, D for simplicity. Specifically, I defined them as
2 2
A = gψψ gtt − gψt B = gφφ gtt − gφt
2
C = gφψ gtt − gtφ gtψ D = gφψ − gφφ gψψ . (E.38)
N ψ ∓ N φ pr p θ
µ ψ A±C
p =p D , , ,∓ ,1 , (E.39)
B ± C pψ pψ B ± C
Ωφ± Ωψ∓
+ =0 (E.40)
κ± κ∓
(if the surface gravity is defined as above) and we can define different vectors such
that
u·k =u·φ=0 u · ψ 6= 0 (E.41)
so that
Ωψ± Ωφ±
ζ± · u = (u · ψ) ζ± · v = (v · φ) . (E.43)
κ± κ±
Getting the covariant expressions
ζ± · u ζ± · v
=± , (E.44)
u·ψ v·φ
146
again, depending on the definition of the surface gravity for the plus minus.
E.4 KN(A)dS4
For these the things change a bit. That is because, not only we have the relation
X Ωi
=0, (E.45)
i
Ti
but also
X 1
=0, (E.46)
i
Ti
with i denoting each of the four different horizons. This means that we don’t even
need to define any new vector, since the covariant expression is just
X
ζiµ = 0 . (E.47)
i
Of course if we want our expression to look like the previous ones, we can again
define a vector such that
pµ kµ = 0 pµ mµ 6= 0 , (E.48)
hence getting
X
ζi · p = 0 . (E.49)
i
147
F Hidden conformal symmetry: The KG equa-
tion
In the Boyer-Lindquist coordinates of the Kerr spacetime (2.28) the metric com-
ponents are
2M r ρ2
gtt = 2 − 1 grr = gθθ = ρ2
ρ ∆
a2 ∆sin4 θ sin2 θ 2 2M ra2
2 2 2 2 2 2
gφφ = − + (r + a ) = sin θ sin θ + r + a
ρ2 ρ2 ρ2
∆ 2 sin2 θ 2 2M ra
gφt = gtφ = 2
asin θ − 2
a(r + a2 ) = − 2 sin2 θ . (F.1)
ρ ρ ρ
In order to find the inverse metric, we only need to invert the (t, φ) block, since
the rest is diagonal. So we take the reduced metric
!
gtt gtφ
g̃ab = , (F.2)
gφt gφφ
= −∆sin2 θ . (F.3)
1 2M ra2
tt 1 2 2 2
g =− gφφ = − sin θ + r + a (F.6)
∆sin2 θ ∆ ρ2
1 2M ra
g tφ = g φt = g tφ = − (F.7)
∆sin2 θ ρ2 ∆
1 −2M r + ρ2 ∆ − a2 sin2 θ
g φφ = − g tt = = , (F.8)
∆sin2 θ ρ2 ∆sin2 θ ρ2 ∆sin2 θ
√
with determinant g = ρ4 sin2 θ =⇒ −g = ρ2 sinθ. Let us calculate each term of
(5.61) separately.
148
√ ω 2 2M ra2 sin2 θ
1 tt 2 2
=⇒ √ −gg Φ,t = +r +a Φ. (F.9)
−g ,t ∆ ρ2
Then
√ √
rr 1 rr 1
−gg Φ,r = ∆sinθΦ,r =⇒ √ −gg Φ,r = 2 ∆Φ,r ,r . (F.10)
−g ,r ρ
Also
√ √
θθ 1 θθ 1
−gg Φ,θ = sinθΦ,θ =⇒ √ −gg Φ,θ = 2 2 sinθΦ,θ ,θ . (F.11)
−g ,θ ρ sin θ
Procceding,
√ √
tφ 2sinθM ra 1 tφ 2M ra
−gg Φ,φ = − (im)Φ =⇒ √ −gg Φ,φ =− mωΦ
∆ −g ,t ∆ρ2
(F.12)
and the φt part
√ √
φt 2sinθM ra 1 φt 2M ra
−gg Φ,t = (iω)Φ =⇒ √ −gg Φ,t =− mωΦ .
∆ −g ,φ ∆ρ2
(F.13)
Finally
√ ∆ − a2 sinθ √ −∆ + a2 sin2 θ 2
φφ 1 φφ
−gg Φ,φ = (im)Φ =⇒ √ −gg Φ,φ = m Φ.
∆ −g ,φ ∆ρ2 sin2 θ
(F.14)
So the KG equation reads
ω 2 ρ2 2M ra2
2 2 2 1
sin θ + r + a Φ + ∆Φ,r + sinθΦ,θ
∆ ρ2 ,r sinθ ,θ
2 2
4M ra a sin θ − ∆ 2
− mω Φ + m Φ=0. (F.15)
∆ ∆sin2 θ
Now lets bring it to a more appropriate form, the one that is used at [178]. The
last term of (F.15) can be rewritten as
2 2
a2 m2 m2
a sin θ − ∆ 2
m Φ = Φ − Φ. (F.16)
∆sin2 θ ∆ sin2 θ
(r − r+ )(r − r− ) = ∆ r+ r− = a2 r+ + r− = 2M , (F.17)
we have
(2M r+ ω − am)2 (2M r− ω − am)2
− =
(r − r+ )(r − r− ) (r − r+ )(r − r− )
149
−4M aωmr a2 m2 4M 2 ω 2 r2
= + − 4M 2 ω 2 + . (F.18)
∆ ∆ ∆
The first term is already in eq. (F.15), as well as the second (eq F.16). So adding
and subtracting the last two terms we have
150
G KG equation in BTZ background
The BTZ metric components are
r2 − M l2
gtt = −N 2 + r2 (N φ )2 =
l2
gφφ = r2
J
gφt = gtφ = r2 N φ = −. (G.1)
2
In order to find the inverse metric, we only need to invert the (t, φ) block, since
the rest is diagonal. So we take the reduced metric
!
gtt gtφ
g̃ab = , (G.2)
gφt gφφ
151
So the KG equation reads
2
m2
2 00 N 2 0 0 1 φ 2
N Φ + + (N ) Φ + (ω + N m) − 2 Φ = 0 . (G.10)
r2 N2 r
R(r)
Now, if we choose the radial part of the field to be χ(r) = √ ,
r
and define a
coordinate [200]
l2
∗ r − r+ r − r−
r = 2 2
r+ ln − r− ln , (G.11)
2(r+ − r− ) r + r+ r + r−
the equation becomes
d2 R
2
R 1 d N2 N2
φ 2 2 m
+ (ω + mN ) − N + + + 5/2 R = 0 . (G.12)
dr∗2 r2 8 2 dr r3/2 2r
out 1
< 0|in Nωm |0 >in = , (G.15)
exp[2π(ω − mΩH )β] − 1
152
H Kerr-Newman-CFT
2J(r+ − r− ) 2(2M J − Q2 )
Ω− −Ω+ = Ω− +Ω+ =
(2M r− − Q2 )(2M r+ − Q2 ) (2M r− − Q2 )(2M r+ − Q2 )
Q3 (r+ − r− ) 2M Q(2J + Q2 )
Φ− −Φ+ = Φ− +Φ+ =
(2M r− − Q2 )(2M r+ − Q2 ) (2M r− − Q2 )(2M r+ − Q2 )
M Q(r+ − r− )
κ+ Φ− + κ− Φ+ =
(2M r− − Q2 )(2M r+ − Q2 )
−Q(r+ − r− )2
κ+ Φ− − κ− Φ+ =
2(2M r− − Q2 )(2M r+ − Q2 )
2JQ
Ω+ Φ− + Ω− Φ+ =
(2M r− − Q2 )(2M r+ − Q2 )
aQ(r− − r+ )
Ω+ Φ− − Ω− Φ+ =
2(2M r− − Q2 )(2M r+ − Q2 )
a
κ+ Ω− + κ− Ω+ = κ+ Ω− − κ− Ω+ = 0 . (H.1)
(2M r− − Q2 )(2M r + − Q2 )
Then we can define temperatures for Kerr, RN and KN as:
√
1 κ− + κ+ M M2 M4 − J2
r+ − r−
TLK = = = TRK = =
2π Ω− − Ω+ 2πa 2πJ 4πa 2πJ
p
1 κ− + κ+ 2M 2 − Q2 M (r+ − r− ) 2 M 4 − M 2 Q2
TLRN = = TRRN = =
2π Φ− − Φ+ 2πQ3 2πQ3 2πQ3
2M 2 − Q2 2M 2 − Q2
0
TLKN = or TLKN =
4πJ 2πQ3
p p
r+ − r− M 4 − J 2 − Q2 M 2 KN 0 2 M 4 − J 2 − Q2 M 2
TRKN = = or TR = .
4πa 2πJ 2πQ3
(H.2)
Using the fact that
Q4
S+ S− = (2π)2 [J 2 +], (H.3)
4
for the Kerr and the RN black holes, we can get central charges from
K
SL,R
2 1 δ(S+ S− )
K
= 2π = 4π 2 J (H.4)
TL,R 4π 2 δJ
153
RN
SL,R
2 1 δ(S+ S− )
RN
= 2π = 2π 2 Q3 (H.5)
TL,R 4π 2 δQ
and then
K
SL,R π2 K
= c =⇒ cK
L,R = 12J (H.6)
K
TL,R 3 L,R
RN
SL,R π 2 RN
= c =⇒ cRN 3
L,R = 6Q , (H.7)
RN
TL,R 3 L,R
which indeed reproduce the correct BH entropies
π2 K K
S+K = SLK + SRK = cL (TL + TRK ) = 2πM r+ = SBH
K
(H.8)
3
π 2 RN RN
S+RN = SLRN + SRRN = c (TL + TRRN ) = 2πM r+ − πQ2 = SBH
RN
. (H.9)
3 L
Kerr-Newman-CFT 2d dual
Here we attempt the same analysis we did for the (M,J) black hole. Now for the
Kerr Newman black hole there is a caveat. Since there are three different charges,
one will always have two charges involved in each thermodynamic relation, after
taking one out. The method I used involves all different possible expressions of
the first law. I will show the computation for the mass for instance and then I will
present the rest.
First I will start by expressing as usual the variation of entropies as
ω − Ω± m − Φ± q
δS± = ±4πα± where α± = (H.10)
2κ±
δ(S+ S− )
= AδM + BδJ + CδQ . (H.11)
4π 2
I will also use the temperatures
1 κ− ± κ+
TL,R = , (H.12)
2π Ω− − Ω+
154
2 (1) (1) (1) (1)
= Ω− α+ κ+ − Ω+ α− κ− = 2πbL TL ωL + 2πbR TR ωR , (H.13)
Ω− − Ω+
where bL,R are the same as before (eq (6.72)) and we have defined
(1) Φ± q ω − Ω± m
α± = α± + = . (H.14)
2κ± 2κ±
Similarly we define
So now the other way to express the mass is excluding the electric charge:
2 1 1
δM = Φ− (κ+ α+ + Ω+ m) − Φ+ (κ− α− + Ω− m)
Φ− − Φ+ 2 2
2 (2) (2) (2) (2)
= Φ− α+ κ+ − Φ+ α− κ− = 2πkq bQ Q
L TL ωL + 2πkq bR TR ωR , (H.17)
Φ− − Φ+
where bQ
L,R is the analog of bL,R for the electric charge and kq is the coefficient that
makes the temperature go from ”RN-like” to ”Kerr-like”
κ+ Φ− ± κ− Φ+ Ω− − Ω+
bQ
L,R = kq = . (H.18)
κ− ± κ+ Φ− − Φ+
We can do a similar procedure for all the charges. We get
(2) (2) (3) (2)
δQ = 2πkq TL ωL − 2πkq TR ωR = 2πkq0 bL TL ωL + 2πkq0 bR TR ωR (H.19)
(1) (1) (3) (3)
δJ = 2πTL ωL − 2πTR ωR = 2π(−kq0 )bQ 0 Q
L TL ωL + 2π(−kq )bR TR ωR , (H.20)
where
Ω− − Ω+
kq0 = (H.21)
Φ− Ω+ − Φ+ Ω−
is the quantity that is needed for the new ”mixing” of angular momentum and
electric charge.
Now , we can observe that
X (i) ω − Ω± m − Φ± q X (i)
α± = 3α± − = 2α± =⇒ ωL,R = 2ωL,R . (H.22)
i
2κ± i
155
So we can express
1 (1) (2) (3) (1) (2) (3)
δ(S+ S− ) = 2πSL ωL −2πSR ωR = 2πSL (ωL +ωL +ωL )−2πSR (ωR +ωR +ωR )
2
(H.23)
and furthermore, we can express the variation of the charges as sum of their two
”halves” i.e.:
1 (1) (1) 1 Q (2) Q (2)
δM = 2πbL TL ωL + 2πbR TR ωR + 2πkq bL TL ωL + 2πkq bR TR ωR =⇒
2 2
1 (1) Q (2) (1) Q (2)
δM = 2π TL (bL ωL + kq bL ωL ) − TR (−bR ωR − kq bR ωR )
2
1 (1) 0 Q (3) (1) 0 Q (3)
δJ = 2π TL (ωL − kq bL ωL ) − TR (ωR + kq bR ωR )
2
1 (2) 0 (3) (2) 0 (3)
δQ = 2π TL (kq ωL + kq bL ωL ) − TR (kq ωR − kq bR ωR ) . (H.24)
2
So then using (H.11) and (H.23) and taking equalities for each frequency separately
we get three equivalent Cardy formulas:
SL,R
= B ± AbL,R = kq (C ± AbQ 0 Q
L,R ) = ±kq (CbL,R − BbL,R ) . (H.25)
2π 2 TL,R
and
2J 2M a
kq = kq0 = − bL = bR = 0
Q3 Q 2M 2 − Q2
MQ Q
bQ
L = bQ
R = − .
2M 2 − Q2 2M
Indeed, using any one of them we get equal left and right central charges c = 12J
, reproducing again the correct BH entropy.
On the other hand, we could have equivalently used the ”RN-like ” temperatures
and the result would be the same with the roles of angular momentum and charge
interchanged. Namely
SL,R
Q
= C ± AbQ 0 Q
L,R = kj (B ± AbL,R ) = ±kj (BbL,R − CbL,R ) , (H.27)
2
2π TL,R
156
where
Q κ− ± κ+
TL,R =
Φ− ± Φ+
Φ − −Φ+ 1 Q3 Φ− − Φ+ kq0 M Q2
kj = = = kj0 = =− = .
Ω− − Ω+ kq 2J Φ+ Ω− − Φ− Ω+ kq J
Indeed, we get equal left and right central charges c = 6Q3 , reproducing the cor-
rect BH entropy.
A note on 3D Cardy
Following [150] we have in 3 dimensions
M = EL + ER J = EL − ER , (H.28)
π2 2
EL,R = cL,R TL,R (H.29)
6
and hence the entropy product can be written as (using the corresponding Cardy
formulas)
π4 π2
S+ S− = (c2L TL2 − c2R TR2 ) = 2 (cL EL − cR ER )
9 3
2
π
= [(cL − cR )(EL + ER ) + (cL + cR )(EL − ER )] , (H.30)
3
which shows that the entropy product being independent of mass shows no diffeo-
morphism anomaly (left and right charges are equal). Equivalently, we have
π4 2 2 π2
S+2 + S−2 = (cL TL + c2R TR2 ) = 2 (cL EL + cR ER )
9 3
π2
= [(cL − cR )(EL − ER ) + (cL + cR )(EL + ER )] , (H.31)
3
which indeed means that if the sum of the entropy squares are independent of J ,
we have the same result.
Moving to 4D
In 4D we have
cL M 4 cR M 4
EL = ER = −1 , (H.32)
24 J 2 24 J 2
157
so the new relations are (if we use the correct central charges)
Nevertheless, the relation for the sum of the entropies this time is
2π 2 4π 4 2 2 8π 2
S+ + S− = cL TL =⇒ (S+ + S− )2 = c T = cL EL
3 9 L L 3
π2 2 M 4
(S+ + S− )2 = c . (H.34)
9 L J2
If this is independnt of J, it implies that the left charge must be of order J. While
the usual entropy product
2π 2 c2L c2R
4 4 2
S+ S− = M − (M − J )
3 24J 2 24J 2
π2
2 2 4 2 2
= (c − cR )M − cR J , (H.35)
36J 2 L
still shows that if there is no diffeo anomaly there is a mass independence.
158
I BTZ from Warped Ads3 black hole
According to [102], one can get the BTZ black hole in a rotating frame by setting
the warp factor ν 2 = 1, and we will pick the plus sing here. We will denote the
Warped coordinate ρ and the BTZ coordinate r.
Setting ν = 1 makes the WAdS3 black hole metric (we use coordinates (τ, ρ, θ))
2
l2 dρ2
2 2 2 2 √ 2 √ √
ds = l dτ + + 2l ρ − ρ+ ρ− dτ dθ + l ρ ρ+ − ρ− dθ2 ,
4(ρ − ρ+ )(ρ − ρ− )
(I.1)
with potentials
1 1 ρ+ − ρ−
ΩW
+ = − √ T+W = √ (I.2)
ρ+ − ρ+ ρ− 2π ρ+ − ρ+ ρ−
and entropy
πl √ πl √ √ 2
S+W = 3ρ+ − ρ− − 2 ρ+ ρ− = 2(ρ+ − ρ− ) + ρ+ − ρ− . (I.3)
6 6
We can get the usual BTZ metric (in (t, r, φ) coordinates) by making the trans-
formations
r+ − r− 1 r2
lτ = t lθ = φ − t ρ= (I.4)
l l r+ − r−
and we can get the potentials as well with the transformations
ΩW
+ (r+ − r− ) 1 r+ − r− 1
ΩBT
+
Z
= + =− √ + =
l l l(ρ+ − ρ+ ρ− ) l
(r+ − r− )2 1 r+ − r− 1 r−
− 2 2
+ =− + = (I.5)
l (r+ − r+ r− ) l lr+ l lr+
and 2 2 2 2
r+ − r− W r+ − r− r+ − r− r+ − r−
T+BT Z = T+ = 2
= . (I.6)
l 2πl2 r+ − r+ r− 2πl2 r+
So it is clear that they are connected with the relation
l 1 l
ΩW
+ = (ΩBT
+
Z
− ) T+W = T BT Z , (I.7)
r+ − r− l r+ − r− +
which means that we essentially subtract a potential from infinity for the angular
velocity and then multiply both velocity and temperature with a factor. Also,
this factor correctly gives extremal limits for the BTZ. Of course that means that
taking one system of coordinates or the other, we get the corresponding relation
Ω+ 1
(t, r, φ) → =± (I.8)
T+ lT−
159
Ω+ Ω−
(τ, ρ, θ) → =± , (I.9)
T+ T−
where the plus or minus signs depend on the definition of the spin temperature.
160
References
[1] S. Carlip, “What we don’t know about BTZ black hole entropy,” Class. Quant.
Grav. 15 (1998) 3609 arXiv:hep-th/9806026v3
[4] Carter B 1973, Black Holes ed: DeWitt C & DeWitt B S (Gordon & Breach)
[5] Carter B 1972, Black hole equilibrium states Part II, General theory of station-
ary black hole states,- 2010. Gen.Rel.Grav.,42,653
[17] W. Israel, “Event horizons in static vacuum space-times,” Phys. Rev. 164
(1967) 1776.
161
[18] S. W. Hawking, “Black holes in general relativity,” Commun. Math. Phys. 25
(1972) 152.
[20] R. Schon and S. T. Yau, “On the Proof of the positive mass conjecture in
general relativity,” Commun. Math. Phys. 65 (1979) 45.
[22] L. Smarr, “Mass formula for Kerr black holes,” Phys. Rev. Lett. 30 (1973) 71
Phys. Rev. Lett. 30 (1973) 521
[29] X. Wang and A. Ritz, “Kerr-AdS Black Holes and Force-Free Magneto-
spheres,” Phys. Rev. D 89 (2014) 10, 106011 arXiv:1402.1452v2 [hep-th]
[30] S. W. Hawking and H. S. Reall, “Charged and rotating AdS black holes and
their CFT duals,” Phys. Rev. D 61 (2000) 024014 arXiv:hep-th/9908109v2
162
[33] D. R. Brill, J. Louko and P. Peldan, “Thermodynamics of (3+1)-dimensional
black holes with toroidal or higher genus horizons,” Phys. Rev. D 56 (1997)
3600 arXiv:gr-qc/9705012v2
[36] S. Carlip, “Quantum gravity in 2+1 dimensions: The Case of a closed uni-
verse,” Living Rev. Rel. 8 (2005) 1 arXiv:gr-qc/0409039v2
[40] R. Emparan and H. S. Reall, “Black Holes in Higher Dimensions,” Living Rev.
Rel. 11 (2008) 6 arXiv:0801.3471v2 [hep-th]
[41] V.P. Frolov and A. Zelnikov. 2011. Introduction to Black Hole Physics, Oxford
University Press.
163
[47] O. J. C. Dias, P. Figueras, R. Monteiro and J. E. Santos, “Ultraspin-
ning instability of rotating black holes,” Phys. Rev. D 82 (2010) 104025
arXiv:1006.1904v1 [hep-th]
[49] R. Emparan and H. S. Reall, “Black Rings,” Class. Quant. Grav. 23 (2006)
R169 arXiv:hep-th/0608012v2
[50] T. Sarkar, G. Sengupta and B. Nath Tiwari, “On the thermodynamic geom-
etry of BTZ black holes,” JHEP 0611 (2006) 015 arXiv:hep-th/0606084v1
[51] S. Wang, S. Q. Wu, F. Xie and L. Dan, “The First laws of thermodynamics of
the (2+1)-dimensional BTZ black holes and Kerr-de Sitter spacetimes,” Chin.
Phys. Lett. 23 (2006) 1096 arXiv:hep-th/0601147v3
[56] F. Larsen, “Rotating Kaluza-Klein black holes,” Nucl. Phys. B 575 (2000)
211 arXiv:hep-th/9909102v3
[57] Z.-W. Chong, M. Cvetic, H. Lu and C. N. Pope, “Charged rotating black holes
in four-dimensional gauged and ungauged supergravities,” Nucl. Phys. B 717
(2005) 246 arXiv:hep-th/0411045v1
164
[59] M. Cvetic, G. W. Gibbons, H. Lu and C. N. Pope, “Rotating black holes in
gauged supergravities: Thermodynamics, supersymmetric limits, topological
solitons and time machines,” arXiv:hep-th/0504080v3
[62] A. A. Pomeransky and R. A. Sen’kov, “Black ring with two angular momenta,”
arXiv:hep-th/0612005v1
[63] A. Castro and M. J. Rodriguez, “Universal properties and the first law of black
hole inner mechanics,” Phys. Rev. D 86 (2012) 024008 arXiv:1204.1284v3 [hep-
th]
[67] M. Banados, “Constant curvature black holes,” Phys. Rev. D 57 (1998) 1068
arXiv:gr-qc/9703040v2
[71] P. Figueras, “A Black ring with a rotating 2-sphere,” JHEP 0507 (2005) 039
arXiv:hep-th/0505244v1
[72] V. P. Frolov, “Mass-gap for black hole formation in higher derivative and
ghost free gravity,” arXiv:1505.00492v1 [hep-th]
165
[73] K. S. Stelle, “Renormalization of Higher Derivative Quantum Gravity,” Phys.
Rev. D 16 (1977) 953.
[77] C. Brans and R. H. Dicke, “Mach’s principle and a relativistic theory of grav-
itation,” Phys. Rev. 124 (1961) 925.
[78] T. P. Sotiriou, “Gravity and Scalar Fields,” Lect. Notes Phys. 892 (2015) 3
arXiv:1404.2955v1 [gr-qc]
[80] T. P. Sotiriou and V. Faraoni, “f(R) Theories Of Gravity,” Rev. Mod. Phys.
82 (2010) 451 arXiv:0805.1726v4 [gr-qc]
[81] Weyl, Hermann, Sitz. Königlich Preußischen Akademie Wiss. (1918) 465
[86] D. Lovelock, “The Einstein tensor and its generalizations,” J. Math. Phys. 12
(1971) 498.
166
[88] S. S. Chern, “A simple intrinsic proof of the Gauss-Bonnet formula for closed
Riemannian manifolds,” Ann. Math. 45 (1944) 747–752
[93] C. Garraffo and G. Giribet, “The Lovelock Black Holes,” Mod. Phys. Lett. A
23 (2008) 1801 arXiv:0805.3575v4 [gr-qc]
[98] A. Salam and J. A. Strathdee, “On Kaluza-Klein Theory,” Annals Phys. 141
(1982) 316.
[100] E. Poisson and W. Israel, “Internal structure of black holes,” Phys. Rev. D
41 (1990) 1796.
167
[102] D. Anninos, W. Li, M. Padi, W. Song and A. Strominger, “Warped AdS(3)
Black Holes,” JHEP 0903 (2009) 130 arXiv:0807.3040v1 [hep-th]
[103] J. M. Bardeen, B. Carter and S. W. Hawking, “The Four laws of black hole
mechanics,” Commun. Math. Phys. 31, 161 (1973).
[107] R. M. Wald, “The First law of black hole mechanics,” In *College Park 1993,
Directions in general relativity, vol. 1* 358-366 arXiv:gr-qc/9305022v1
[108] J. Lee and R. M. Wald, “Local symmetries and constraints,” J. Math. Phys.
31 (1990) 725.
[110] D. Sudarsky and R. M. Wald, “Mass formulas for stationary Einstein Yang-
Mills black holes and a simple proof of two staticity theorems,” Phys. Rev. D
47 (1993) 5209 arXiv:gr-qc/9305023v1
[111] R. M. Wald, “Black hole entropy is the Noether charge,” Phys. Rev. D 48
(1993) 3427 arXiv:gr-qc/9307038v1
[113] R. M. Wald, “On identically closed forms locally constructed from a field ,”
J. Math. Phys. 31 (1990) 2378.
168
[116] K. Copsey and G. T. Horowitz, “The Role of dipole charges in black hole
thermodynamics,” Phys. Rev. D 73 (2006) 024015 arXiv:hep-th/0505278v3
[117] M. Rogatko, “Extrema of mass, first law of black hole mechanics and staticity
theorem in Einstein-Maxwell axion dilaton gravity,” Phys. Rev. D 58 (1998)
044011 arXiv:hep-th/9807012v2
[118] S. Gao, “The First law of black hole mechanics in Einstein-Maxwell and
Einstein-Yang-Mills theories,” Phys. Rev. D 68 (2003) 044016 arXiv:gr-
qc/0304094v2
[122] T. Jacobson, G. Kang and R. C. Myers, “On black hole entropy,” Phys. Rev.
D 49 (1994) 6587 arXiv:gr-qc/9312023v2
[125] C. W. Misner, “The Flatter regions of Newman, Unti and Tamburino’s gen-
eralized Schwarzschild space,” J. Math. Phys. 4 (1963) 924.
169
[130] J. Bicak, “Selected solutions of Einstein’s field equations: Their role in gen-
eral relativity and astrophysics,” Lect. Notes Phys. 540 (2000) 1 arXiv:gr-
qc/0004016v1
[136] J. H. Yue, G. H. Yang, L. J. Tian and S. Zhu, “Euler characteristic and topo-
logical phase transition of NUT-Kerr-Newman black hole,” Commun. Theor.
Phys. 49 (2008) 941.
170
[144] D. D. K. Chow, “Charged rotating black holes in six-dimensional gauged
supergravity,” Class. Quant. Grav. 27 (2010) 065004 arXiv:0808.2728v2 [hep-
th]
[146] J. L. Cardy, “Conformal Field theory and Statistical Mechanics,” Cond. Mat.
Stat. Mech. arXiv:0807.3472v1 [cond-mat.stat-mech]
[150] S. Detournay, “Inner Mechanics of 3d Black Holes,” Phys. Rev. Lett. 109
(2012) 031101 arXiv:1204.6088v1 [hep-th]
[154] A. Strominger, “Black hole entropy from near horizon microstates,” JHEP
9802 (1998) 009 arXiv:hep-th/9712251v3
171
[158] J. M. Bardeen and G. T. Horowitz, “The Extreme Kerr throat geometry: A
Vacuum analog of AdS(2) x S**2,” Phys. Rev. D 60 (1999) 104030 arXiv:hep-
th/9905099v1
[164] A. Castro and F. Larsen, “Near Extremal Kerr Entropy from AdS(2) Quan-
tum Gravity,” JHEP 0912 (2009) 037 arXiv:0908.1121v2 [hep-th]
172
[171] T. Hartman, K. Murata, T. Nishioka and A. Strominger, “CFT Duals for
Extreme Black Holes,” JHEP 0904 (2009) 019 arXiv:0811.4393v2 [hep-th]
[174] M. Cvetic and F. Larsen, “Grey body factors for rotating black holes in
four-dimensions,” Nucl. Phys. B 506 (1997) 107 arXiv:hep-th/9706071v2
[175] D. D. K. Chow, “Equal charge black holes and seven dimensional gauged
supergravity,” Class. Quant. Grav. 25 (2008) 175010 arXiv:0711.1975v3 [hep-
th]
[176] M. Cvetic and F. Larsen, JHEP 1202 (2012) 122 arXiv:1106.3341v1 [hep-th]
[181] A. Sen, “Rotating charged black hole solution in heterotic string theory,”
Phys. Rev. Lett. 69 (1992) 1006 arXiv:hep-th/9204046v1
[183] A. Curir, “Spin entropy of a rotating black hole,” Nuovo Cimento B 51B,
262 (1979)
173
[186] J. D. Bekenstein, “Extraction of energy and charge from a black hole,” Phys.
Rev. D 7 (1973) 949.
[188] M. Ansorg and J. Hennig, “The Inner Cauchy horizon of axisymmetric and
stationary black holes with surrounding matter,” Class. Quant. Grav. 25 (2008)
222001 arXiv:0810.3998v1 [gr-qc]
[189] M. Ansorg and J. Hennig, “The Inner Cauchy horizon of axisymmetric and
stationary black holes with surrounding matter in Einstein-Maxwell theory,”
Phys. Rev. Lett. 102 (2009) 221102 arXiv:0903.5405v2 [gr-qc]
[193] A. Peltola and J. Makela, “Radiation of the inner horizon of the Reissner-
Nordstrom black hole,” Int. J. Mod. Phys. D 15 (2006) 817 arXiv:gr-
qc/0508095v2
[198] W. Kim and E. J. Son, “Thermodynamics of warped AdS(3) black hole in the
brick wall method,” Phys. Lett. B 673 (2009) 90 arXiv:0812.0876v3 [hep-th]
174
[199] J. J. Oh and W. Kim, “Absorption cross section in the topologically massive
gravity at the critical point,” Eur. Phys. J. C 65 (2010) 275 arXiv:0904.0085v1
[hep-th]
[201] B. Carter, “Global structure of the Kerr family of gravitational fields,” Phys.
Rev. 174 (1968) 1559.
[204] S. W. Hawking, C. J. Hunter and M. Taylor, “Rotation and the AdS / CFT
correspondence,” Phys. Rev. D 59 (1999) 064005 arXiv:hep-th/9811056v2
[210] N. Goheer, M. Kleban and L. Susskind, “The Trouble with de Sitter space,”
JHEP 0307 (2003) 056 arXiv:hep-th/0212209v3
[211] H. Elvang and P. Figueras, “Black Saturn,” JHEP 0705 (2007) 050
arXiv:hep-th/0701035v2
175
[213] S. Tomizawa, Y. Yasui and A. Ishibashi, “Uniqueness theorem for charged
dipole rings in five-dimensional minimal supergravity,” Phys. Rev. D 81 (2010)
084037 arXiv:0911.4309v3 [hep-th]
176