1. Introduction
The Kerr two-parameter family of metrics is arguably the most important solution of Einstein’s equations. These metrics represent compact objects in an asymptotically Minkowskian, de Sitter or anti de Sitter background, according to the value of the cosmological constant
. In Boyer-Lindquist coordinates
, the Kerr metric is given by:
where
are the standard angular coordinates of the unit sphere,
is the cosmological constant and we introduced the notation
For large
r, Equation (
1) approaches the metric of Minkowski spacetime if
, de Sitter if
is positive and anti-de Sitter if negative. The parameters
appear as constants of integration and thus can take any value. They can be seen to correspond, respectively, to the mass and the angular momentum per unit mass of these objects. If
, we may change coordinates
and get the same form (
1) with
. So we may assume without loss of generality that
. All these spacetimes contain a curvature singularity. This is milder in the rotating,
case: it lies at
,
and it is possible to go across it to negative
r values [
1].
In the non-rotating
case, Kerr’s metric reduces to Schwarzschild’s:
for which the singularity is at
and signals a boundary of the spacetime.
For certain combinations of and , one or two horizons keep the singularity causally isolated from the external, asymptotically simple region where the metric approaches that of Minkowski or (anti) de Sitter space, called the Domain of Outer Communications (DOC): these are the black hole (BH) solutions. For other combinations the singularity is visible from the DOC: these are called naked singularities (NSs).
Consider, for example, the asymptotically Minkowskian case
. The horizons correspond to the hypersurfaces given by the roots of
in (
2) [
1]:
We will focus on the following cases:
Rotating Kerr metric (): sub extreme BH (, two horizons, at ); extreme BH (, a single horizon at ); super-extreme NS (, no horizon).
Non-rotating Schwarzschild metric (): Schwarzschild BH (, one horizon at ), Schwarzschild NS (, no horizon).
When analyzing geodesics we find that the Schwarzschild NS effectively behaves as a central object with a negative mass: it is a totally unphysical solution of Einstein’s vacuum equations. The rotating NS
has a causal behavior that is completely pathological: through any point there passes a closed timelike curve [
1]. The sub-extreme Kerr BH, on the other hand, is among the physically most relevant solutions of Einstein’s equation. This is so in view of the uniqueness theorems [
2,
3], that state that the
Kerr’s BH is
the only asymptotically Minkoswkian rotating BH solution of Einstein’s equations. Black holes are the most extraordinary prediction of Einstein’s gravity, and if sixty years ago we were beginning to
mathematically understand them, in 2015, with the first detection of gravitational waves from BH collisions, we entered an era of
direct observation of these objects. We are now potentially able to test whether or not the
piece of the
Kerr metric (
1) correctly models the spacetime outside a rotating BH.
One may wonder if the existence of unphysical solutions such as the Kerr NS is a flaw of General Relativity. The answer is in the negative. A close analogy can be found in Classical Mechanics: Kerr’s solution (
1) is
stationary, roughly meaning that the space geometry is the same at all times (technically: the metric is invariant under the flow of the vector field
), so it is analogous to time-independent solutions in Classical Mechanics, which can be characterized as the system sitting at a critical point of the potential energy function. Critical points can be local minima or not, and it is only in the first case that the time-independent solution is feasible. A regular cone “resting” vertical on its tip is a solution of Newton’s equations, but we never encounter these configurations. The reason is that these are
unstable time-independent solutions, they require fine tuning of the initial condition and under the slightest perturbations the system evolves into a very different configuration. In the following sections we will show that this is exactly the case of the NSs in the Kerr
family of solutions.
For sub-extreme rotating BHs, the inner horizon is also a Cauchy horizon: a null hypersurface beyond which the uniqueness of the evolution is lost. Although there is a unique analytic continuation of the metric beyond the Cauchy horizon, there are infinitely many possible smooth continuations satisfying Einstein’s vacuum equation. This is an annoying feature, contrary to the central idea in Classical Physics that we can predict uniquely the evolution of a system from initial data. These inner BH regions also exhibit severe causal pathologies, such as closed timelike curves. We will prove below that these spacetime regions are unstable under perturbations, that is, unphysical extensions of the BH metric.
Having found an explanation for the unphysical solutions, we would like to make sure that the solutions that model BH exteriors, that is the piece of the Kerr and Schwarzschild metrics, are stable under perturbations. Otherwise they would be irrelevant solutions of General Relativity. We have entered an era where the strong gravity regime of BH can be tested by means of gravitational waves, multi-messanger Astronomy and direct observations of BH horizons. General Relativity is the available theory of gravity, and thus Kerr solution is used to model the BH exterior region. If a mismatch is observed between theory and experiment, in view of the uniqueness theorems, corrections to General Relativity will have to be considered.
A complete proof of the stability of Kerr BH using the full, non-linear Einstein’s equation is and will be lacking for a time. We should recall at this point that such a proof only exists for: (i) de Sitter spacetime (proved by H. Friedrich in 1986 [
4,
5]), (ii) Minkowski spacetime (proved by Christdoulou and Klainerman in 1993 [
6]), (iii) Schwarzschild de Sitter BH, (proved by Hintz and Vasy in 2016 [
7]). A preprint is available since a few months ago with a proof of non-linear stability of the
Schwarzschild BH (see [
8]), the proof takes five hundred pages. For the Kerr BH, at the moment, we must content ourselves with a proof of
linear stability.
Linear perturbations around a metric can be organized in modes by taking advantage of the symmetries of the background: perturbations are classified according to the way they transform under the action of the isometry group of the unperturbed metric. Single modes are found to either oscillate or grow/damp exponentially in time. If a single unstable—that is, exponentially growing—mode is found, the background metric is certainly unstable. The existence of such modes is what allows us to rule out NS and pathological BH interiors, as proved below. The absence of unstable modes on BH exteriors is a solid signal of stability: we call this modal linear stability. There is still the possibility that oscillating modes add up to a locally growing perturbation. Ruling out this possibility is what we call nonmodal linear stability.
This review is organized as follows: in
Section 2 we briefly introduce the ideas of mode decomposition of linear perturbations based on background symmetries; in
Section 3 we review the proofs of instability of the Schwarzschild and Kerr NS instabilities, and also of the instability of the region beyond the Cauchy horizon of Kerr BHs; in
Section 4 we review the proof of nonmodal stability of the Schwarzschild (Schwarzschild de Sitter) BH. In doing this we show that there are gauge invariant scalar fields, constructed out of the perturbed curvature scalars (CSs), that is, contractions of the Weyl tensor and its first covariant derivative, that contain the exact same information as the gauge class of the metric perturbation, and use them to show that, for the most general perturbation, no transient growths are possible, and that a perturbed Schwarzschild BH always decays into a slowly rotating Kerr BH. The discussion is centered on the effect of perturbations on the background geometry.
4. Nonmodal Stability of the Schwarzschild Black Hole
Prior to [
36] all notions of stability of the exterior region of a Schwarzschild BH were concerned with finding bounds to the Regge-Wheeler and Zerilli fields
. These fields are defined in the two dimensional
orbit space and parametrize time dependent metric perturbations in the Regge-Wheeler (RW) gauge as shown in the set of equations, (
12)–(
27), that we will write more concisely as
where
and
is a second order differential operator.
Some key results were:
In [
37], it was shown that separable solutions
that do not diverge as
require
, ruling out exponentially growing solutions in the odd sector.
In [
38], it was shown that separable solutions
that do not diverge as
require
, ruling out exponentially growing solutions in the even sector.
In [
39], it was shown that, for large
t and fixed
r,
decays as
an effect known as “Price tails”.
In [
40], the conserved energy
was used to rule out uniform exponential growth in time.
Furthermore, in [
40], a pointwise bound
was proved, where the constants
are given in terms of the initial data
To understand the limitations of these results it is important to keep in mind that the
are an infinite set of fields defined on the
orbit space,
whose first and second order derivatives enter the terms in the series (
72) through (
73),
together with the sheperical harmonics and their derivatives. Two extra derivatives are then required to calculate the perturbed Riemann tensor from the metric perturbation, as a first step to measure the effects of the perturbation on the curvature. Thus, the relation of the
to geometrically meaningful quantities is remote, and the usefulness of the bounds (
76) to measure the strength of the perturbation is not obvious at all.
Even if we knew the impact of these bounds on the components of , we would face the unavoidable problem of the lack of a natural measure of the “size” of tensor fields on a Lorentzian manifold. On a Riemannian manifold, where the metric is positive definite, the pointwise size of a tensor , could be measured by , which is positive if is nonzero, and an norm of the field given by . None of these notions is available for a Lorentzian .
Besides the problem, inherent to Lorentzian geometry, of measuring the “size” of tensors, there is the often overlooked fact that controlling the size of time dependent series terms of a quantity does not entirely control the quantity itself. This is what led to the notion of
nonmodal stability in fluid dynamics, where the limitations of the mode analysis were realized in experiments involving shear flows bounded by walls [
41]. In this case, the linearized Navier–Stokes operator is non normal, so their eigenfunctions are non orthogonal. As a consequence, even if the individual modes decay as
—a condition that assures large
t stability—there may be important transient growths [
41]. Take, for example, the following toy model (from Section 2.3 in [
41]) of a system with two degrees of freedoms:
obeying the equation
, with
a matrix with (non orthogonal!) eigenvectors
, say,
Consider the case
. Note that, although
, that is, normal modes decay exponentially, if
, then
reaches a maximum norm
at a finite time before decaying to zero. The operators involved in the LEE are normal, so the above situation of non orthogonal eigenfunctions does not arise. However, it is easy to construct examples of, e.g., time dependent scalars on the sphere, say
where the
oscillate (as the pure modes (
28) were shown to do) and yet the function
develops arbitrarily high localized transient growths, or even grows without bound as
in continuously narrowing areas. This is so because (recall that our
are an orthonormal basis of
real spherical harmonics on the unit sphere
)
and the integral above can be kept bounded in
t while at
grows high in narrowing areas. In fact, this was precisely the motivation in [
40]: to “undo” the separation of variables in
and show that the sum (
81) of oscillating modes does not allow unbounded transient growths in
, and that we can place pointwise bounds based on the initial data of the form (
76) and (
77). This could be done exploiting the form of the partial differential Equation (
24) obeyed by the fields
. Although finding an exponentially growing mode is a definite signal of instability, as we can see, there are different possible degrees of stability when all modes are oscillating (real
’s). The symmetries of the background geometry is what allowed us to decompose metric perturbations as in (
73) and (
28). The results in [
40] can be regarded as a way to “undo” the
variable separation (
28): they prove (
76) and (
77) for
arbitrary (that is, non necessarily separable) solutions of the 1 + 1 wave Equations (
24). A natural question after the work [
40] is: Can we also “undo” the
separation of variables? Can we find bounds for arbitrary solutions of the LEE (
7)? The question is very tricky since it omits a difficult issue: which
spacetime scalar field should we try to place bounds on?
The strategy in [
36] was to measure the pointwise intensity of a perturbation by its effect on curvature related scalar fields (CSs, for short, not to be confused with the Newman-Penrose
spin weighted scalars (
33) and (
34), which depend point by point on a selected null tetrad). These scalars are full contractions of the Weyl tensor
(which equals the Riemann tensor in vacuum), its covariant derivatives, the metric and its inverse, and the volume form
. Some examples are
Working with scalar fields avoids the issue discussed above of “measuring the size of tensors” when the background metric is Lorentzian. There is, however, the misconception that the first order variation of any CS is gauge invariant (that is, coordinate independent). This is not the case: under the transformation (
9), the first order perturbation of a CS field
Z changes as
, so
is gauge invariant if and only if the background CS vanishes:
. As an example, in the Schwarzschid de Sitter (SdS) background, the CSs (
82) take the values
so in the linearized theory only
is a gauge invariant quantity. However,
combinations of first order variations of CSs whose background values are nonzero can be gauge invariant. As an example, the field
is gauge invariant since, under a gauge transformation along
,
In [
36], the fields
and
were proposed to measure the strength of perturbations. It was shown that:
depends only on the odd piece of the perturbation, whereas depends only on
There is a one to one relation between the gauge class
of a metric perturbation (that is, the set of perturbations obtained from
by a transformation (
9)) and the
. More precisely, the maps
are bijections. In particular, the gauge perturbation in, say, the RW gauge, can be recovered from the
fields.
For the odd sector of the LEE, a four dimensional approach relating the metric perturbation with a scalar potential
defined on the spacetime
, instead of the
orbit manifold, was found in [
36]. It was noticed that the sum over
of (
73) simplifies to
where
is the dual of the Weyl tensor, and
is a field assembled using spherical harmonics and the
:
The odd sector LEE equations for
, combined with the spherical harmonic Equation (
17) for the
(see (
24) and (
27)), turn out to be equivalent to what we call the
four dimensional Regge-Wheeler equation (4DRWE), which reads
Note, however, that
is no more than the collection of fields
: its connection to geometrically relevant fields such as CSs is, a priori, loose. Note also that
any field of the form
satisfies, for arbitrary constants
, the 4DRWE (
89). This is a consequence of the form of the potential
(see (
27)), which contains an “angular momentum” term
(contrast with
)
Much more important is the fact (proved in [
36] for
, generalized to nonzero
in [
9]) that the LEE implies that the field
can be written, after reiterated use of the LEE and equations derived from these, as
The first term contains the
contribution which, as we commented above, is time independent and irrelevant to the stability problem (there is no
odd contribution). This time independent contribution amounts to a shift of the Schwarzschild BH to a “slowly rotating Kerr black hole”. Technically, by “slowly rotating Kerr black hole” we mean the metric we get if we Taylor expand Kerr’s metric (
1) and (
2) in the rotation parameter
a and keep only first order terms: for such a perturbation,
looks exactly like the
(rotation around the
axis) term in the first sum in (
91). The second term of
, being of the form (
90), satisfies the 4DRWE (
89). What is not obvious, but can be checked by direct calculation, is that the
piece of
(and thus
)
also satisfies (
89).
in (
91) contains all the information we need to reconstruct the metric perturbation: the
can be obtained as the
harmonic coefficients of
, and the
(which are all we need to reconstruct the
piece of the metric perturbation, see [
9]) can be obtained from the
coefficients. This proves that the gauge invariant, curvature scalar
contains all the information encoded in an odd metric perturbation, while being a meaningful scalar to measure the strength of the perturbation. If we managed to place a pointwise bound on this quantity we would have a proof of nonmodal linear stability of the Schwarzschild BH under odd perturbations. A pointwise bound can be placed on
using the fact that
satisfies (
89), by adapting a technique from [
42] (see [
9,
36] for details). The result is the following: for all
in the
BH (or
r between the event and cosmological horizons if
), and all
t, there is a constant
that depends on the initial datum such that
This equation settles the issue of nonmodal stability under odd perturbations.
To treat even perturbations, we must face the problem that, as becomes obvious when inspecting the even potential
in (
27), particularly the factors involving
ℓ in the denominator, the even Equation (
24) cannot be used together with (
17) to construct a scalar field
that satisfies a covariant equation such as the (
89). A remarkable result by Chandrasekhar [
27,
28] comes to our help: there exists operators
of the form
where
such that, if
is a solution of the odd Equation (
24), then
is a solution of the even equation, and similarly replacing
. In [
9], Lemma 7, it was furthermore proved that, although the operators (
93) clearly have a non trivial kernel
when acting on arbitrary functions, they are 1-1 when restricted to
solutions of the 1 + 1 wave Equations (24). As a consequence, every solution
of the even 1 + 1 wave equation in the
domain of a
Schwarzschild BH (
r between the event and cosmological horizons if
), can be written as
.
Using the even LEE and equations derived from those,
can be reduced to [
36]
where
The—time-independent—first term in (
95) comes from the
even perturbation, which amounts to a mass shift of the background metric [
9] (there is no
even contribution, see [
9]). Further use of the Chandrasekhar operators described above allows to rewrite
entirely in terms of functions obeying (
89). The details, which are quite involved, as well as the details of how to use this fact to set a bound on
, can be found in [
9]. We only quote the result obtained in [
9,
36]: for all
in the
BH (
r between the event and cosmological horizons if
) and all
t, there is a constant
that depends on the initial datum such that
Together with (
92), this equation proves the nonmodal linear stability of
Schwarzschid BHs.
We close this sections with two observations made in [
9]. Combining (
91) and (
95) with the Price tail decay at fixed
r, Equation (
74), we find that, at large
t and fixed
,
which corresponds to a stationary BH in the Kerr (or Kerr de Sitter) family with a mass
and angular momentum components
. For example, if only
, the perturbation (
98) corresponds (in some gauge) to the metric perturbation obtained by applying the operator
to the metric (
1) and (
2). Equation (
98) indicates that, after a long time, the perturbation settles into a “slowly rotating” Kerr (or Kerr dS) BH. For
, this fact was rigorously proved in [
43], where it was shown, working in an specific gauge, that the metric perturbation decays at large
t into that of a slowly rotating BH.
The second observation is that there is a much simpler CS connected to even perturbations, but this CS is not gauge invariant. This is
, defined in (
82) which,
in the Regge-Wheeler gauge, has a first order variation [
9]
Thus,
, being of the form (
90), also satisfies the 4DRW equation. We might think of using
to measure the strength of even perturbations, but this field, as we said, is gauge dependent, due to the fact that
for the background Schwarzschild or S(A)dS black hole and, as a consequence, under the gauge transformation (
9),
A gauge invariant field
was found in [
9] for the even perturbations which satisfies the 4DRWE (
89) (see the discussion in Section 5.2 of this reference). This has the drawback of not having a direct geometric interpretation in terms of CSs. In any case, we learn from the existence of
, or just from Equation (
99), that the degrees of freedom of the linearized even perturbations can be encoded in a scalar field
satisfying the 4DRWE (
89), as is the case for the odd perturbations. This field is independent from the odd field (
88), that also satisfies this equation. In other words:
the most general linear perturbation can be encoded in two independent scalar fields which satisfy (89). Thus, proving the stability and the large
t decay of perturbations of a Schwarzschild or Schwarzschild-de Sitter BH into slowly rotating Kerr (de Sitter) BH amounts to proving the decay of the
components of generic solutions of (
89). This is at the heart of the proof in [
43] of the decay of perturbations for
, where the two scalars satisfying (
89) were integrated into an
symmetric tensor, as explained in Remark 7.1 in [
43].
5. Conclusions and Current Developments
The proof of the instabilities of the naked singularities and the regions beyond the black hole Cauchy horizon within the Kerr–Newman family of metrics, conciliates General Relativity with basic Physics principles, such as the uniqueness of evolution from initial data, and the requirement that causal pathologies such as closed timelike and null curves do not occur. A modal analysis of the LEE is enough to rule out these solutions, given that exponentially growing modes are found. It is an interesting fact that, in the non rotating case, the geometrical effects of the unstable modes show up in
differential curvature scalar invariants and in the splitting of one of the background degenerate Petrov null directions (see Equations (
47) and (
49)). They do not leave traces on
algebraic curvature scalars, in contrast to what happens for generic black hole perturbations, (see, for example, Equations (
91) and (
95)). It is also interesting that the unstable modes in all the analyzed static cases (negative mass Schwarzschild NS, super extreme Reissner-Nordström NS, inner region of the Reissner-Nordström BH) are even under
P, as defined in (
14), and that there is exactly one unstable mode in each harmonic sector.
An analysis of these cases suggests that the singularity plays no role in the instability. As an example, for the Schwarzschild naked singularity, a negative sector of the potential
in the 1 + 1 wave Equations (
24)–(
27) is responsible for the existence of the unstable modes. If the singularity were replaced by a small spherically symmetric negative mass distribution, away from
this potential would not change substantially, and an unstable mode could be found anyway. This is also the case of the super-extreme Reisner-Nordström NS, treated in [
19,
20], for which the singularity could be replaced by a spherically symmetric charge distribution with
leading to a non singular metric which would, anyway, be unstable. Similarly, for the Kerr NS, the ring singularity plays no role in its instability. This leaves the impression that General Relativity simply “dislikes” unusual matter or over-rotating and over-charged compact objects.
The analysis of the stability of black hole outer regions, having passed decades ago the test of modal linear stability, has had, after a long period of little activity, a remarkable progress in the last few years. An incomplete list of recent advances related to the stability of the Schwarschild and Kerr black holes follow: (i) for
Schwarzschild black holes the nonmodal linear stability was established in [
9,
36]; (ii) in the
case, the
decay in time of generic linear perturbations of the Schwarzschild black hole, leaving a “slowly” rotating Kerr black hole was proved in [
43]; (iii) the conditional stability of the
Schwarzschild black hole, and the breaking of the even/odd symmetry mediated by the Chandrasekhar operators (
93) and (
94) was studied in [
29]; (iv) the
non-linear stability of the Schwarzschild de Sitter black hole was proved in [
7]; (v) a preprint is now available with a proof of the
non-linear stability of the
Schwarzschild black hole [
8]; (vi) pointwise decay estimates for solutions of the linearized Einstein’s equations on the outer region of a Kerr black hole were obtained in [
13]; (vii) the role of hidden symmetries (see the review [
44]) in type D spacetimes, and the reconstruction of (gravitational, Maxwell and spinor) perturbation fields from “Debye potentials” (first introduced in [
45,
46]), was studied in depth and made clear in the series of papers [
47,
48,
49,
50].
The ultimate challenge in Black Hole perturbation theory remains open: proving the non-linear stability of the outer region of a Kerr black hole.