Linear Stability of Schwarzschild-Anti-de Sitter spacetimes I:
The system of gravitational perturbations
Abstract
This is the main paper of a series establishing the linear stability of Schwarzschild-Anti-de Sitter (AdS) black holes to gravitational perturbations. Specifically, we prove that solutions to the linearisation of the Einstein equations with around a Schwarzschild-AdS metric arising from regular initial data and with standard Dirichlet-type boundary conditions imposed at the conformal boundary (inherited from fixing the conformal class of the non-linear metric) remain globally uniformly bounded on the black hole exterior and in fact decay inverse logarithmically in time to a linearised Kerr-AdS metric. The proof exploits a hierarchical structure of the equations of linearised gravity in double null gauge and crucially relies on boundedness and logarithmic decay results for the Teukolsky system, which are independent results proven in Part II of the series. Contrary to the asymptotically flat case, addition of a residual pure gauge solution to the original solution is not required to prove decay of all linearised null curvature and Ricci coefficients. One may however normalise the solution at the conformal boundary to be in standard AdS-form by adding such a pure gauge solution, which is constructed dynamically from the trace of the original solution at the conformal boundary and quantitatively controlled by initial data.
1 Introduction
The study of the stability of black hole solutions of the Einstein equations with cosmological constant ,
(1) |
originates in the physics literature with the pioneering work of Regge and Wheeler [RW57] for the Schwarzschild solution. In the past two decades the subject has seen a tremendous development through the introduction of modern PDE theory. As of today, satisfactory non-linear stability results are available for Schwarzschild and very slowly rotating Kerr black holes, i.e. in (1), and Kerr-de Sitter black holes, i.e. in (1); see [DHRT21, GKS22, HV16] and references therein. On the other hand, still very little is known about the non-linear evolution of perturbations of the Schwarzschild-Anti de Sitter, and more generally Kerr-Anti de Sitter, family of solutions, i.e. in (1).
The main difficulty in the analysis of asymptotically Anti-de Sitter (aAdS) spacetimes lies in their non-globally hyperbolic nature: The spacetimes possess a timelike conformal boundary at infinity, which is most easily seen for the maximally symmetric solution of (1) with , Anti-de Sitter (AdS) space [HE08]. The existence of the boundary necessitates the study of a boundary initial value problem to understand the dynamics of the hyperbolic system (1). Formulating geometric boundary conditions to establish local well-posedness is highly non-trivial. See [Fri95, EK19] for some classical well-posedness theorems for (1) with .
Turning to the global dynamics in the case, stability properties of stationary solutions are expected to depend crucially on the type of boundary condition imposed. Notably, here even the simplest case – that of perturbations around pure AdS – is still open. In the case that Dirichlet type boundary conditions are imposed at the boundary, non-linear instability of AdS has been proven for spherically symmetric toy-models [Mos18, Mos20] and is expected to hold in general.111Dirichlet conditions can be thought of as a form of reflecting boundary conditions. In particular, the asymptotic mass is held constant along the conformal boundary, hence gravitational radiation cannot escape through the conformal boundary. Geometrically, it corresponds to fixing the conformal class of the metric on the boundary to be that induced by the pure AdS metric. See [BR11] for very influential numerical study of the Dirichlet problem is the context of the spherically symmetric scalar field model. On the other hand, the (linear) results of [HLSW20] suggest non-linear stability to hold in the case of dissipative boundary conditions, where radiation is allowed to escape through the conformal boundary. Going from pure AdS to black hole spacetimes, the additional characteristic phenomena of trapped null geodesics and superradiance couple with the effect of the boundary making the analysis of the problem even more difficult. Nevertheless, the general expectation is again that of instability for reflecting boundary conditions and stability for dissipative boundary conditions.
1.1 The scalar wave equation, linear stability and non-linear stability
From a PDE perspective, non-linear stability results typically rely on a robust understanding of the underlying linearised problem including quantitative estimates on the rates of decay in the geometry under consideration. Such an analysis also seems a prerequisite for non-linear instability results in order to gain control on potential blow-up or growth mechanisms. Estimating the linearisation of the equations (1) requires choosing a gauge and, independently of the specifics of the gauge, already results in a complicated coupled system of equations. A good first intuition can often be gained from the study of the scalar wave equation on the background under consideration. This removes the problem of gauge as well as the coupled tensorial character of the problem. At the same time, the scalar equation (due to its Lagrangian structure) inherits natural coercive conservation laws from the symmetries of the background which can be exploited in the analysis.
In the case of pure AdS, one thus discovers the existence of time-periodic (i.e. non-decaying) solutions of , which lie at the heart of the non-linear instability exploited in [Mos18]. In the case of asymptotically AdS black holes, the corresponding scalar problem was studied in [HS13, HS14], where it is shown that solutions of with Dirichlet boundary conditions decay inverse logarithmically and not faster on Kerr-AdS black holes whose parameters satisfy the Hawking-Reall bound.222Beyond that bound, one has exponentially growing solutions. See [Dol17]. In view of the slow decay, the authors of [HS14] conjectured non-linear instability of these black holes. Recently, a concrete instability mechanism (related to weak turbulence and the growth of higher order Sobolev norms) has been suggested for a non-linear scalar toy-model on Schwarzschild-AdS [KM].
1.2 The main result
The goal of this series of works is to show that the logarithmic decay established for the scalar toy problem also holds for the linearised Einstein equations on Schwarzschild-AdS. More precisely we will prove the following statement:
Theorem (Informal version).
Solutions to the linearisation of the Einstein equations around a Schwarzschild-AdS metric arising from regular initial data and with standard Dirichlet boundary conditions at the conformal boundary (inherited from fixing the conformal class of the non-linear metric) remain globally uniformly bounded on the black hole exterior and in fact decay inverse logarithmically to a linearised Kerr-AdS metric.
For a precise statement of the theorem, see already Theorem 4.7 below.
Remark 1.1.
The theorem should be directly compared with the result of [DHR19] in the case. The main difference is that here only a logarithmic decay rate (as opposed to inverse polynomial in [DHR19]) can be concluded. This is characteristic of the reflective boundary conditions as explained above. In Part III of the series, we actually prove that the decay rate cannot be improved for general solutions. The other main difference compared with the asymptotically flat case is that here all quantities can be shown to decay without adding a residual pure gauge solution. This is to be constrasted with Theorem 3 in [DHR19] which establishes boundedness and Theorem 4 in [DHR19] where decay is established after having added to the solution an appropriately future normalised (dynamically determined) pure gauge solution.
Remark 1.2.
Note that despite the linear stability statement of the theorem above, one may still expect non-linear instability in view of the slow decay rate. However, the statement can still be used to establish rigidity properties concerning the Schwarzschild-AdS metric. These will be explored elsewhere.
Remark 1.3.
We finally remark that an analogous theorem is expected to hold in the Kerr-AdS case for black hole parameters satisfying the Hawking-Reall bound. For small , such a result should follow perturbatively from the techniques of this paper. We leave this for future study and refer to [GH23] for further discussion.
1.3 Overview of the paper and comments on the proof
We will only very briefly comment on the global structure of the proof of the main theorem here. Afterwards, we immediately provide the formal set-up for the problem in Section 2, which includes a derivation of the linearised Einstein equations in double null gauge. Section 3 is concerned with the construction of appropriate initial data and solutions to this linearised system, i.e. well-posedness of the linearised system. A formal version of the main theorem is then formulated in Section 4 and proven in Section 5. The impatient reader wishing to take the existence of solutions of the linearised system (Theorem 3.9) for granted may turn immediately to the main theorem in Section 4, which concerns the global properties of such solutions.
At the highest level, our strategy follows closely that of [DHR19] in the asymptotically flat () case and starts by expressing the linearised Einstein equations in a double null gauge. A first key ingredient of the analysis, carried out in our companion paper [GH24], is to prove boundedness and decay estimates for the so-called Teukolsky quantities, denoted . These are certain linearised null-curvature components of the linearised system, which (a) do not depend on the specific gauge in which the equations (1) are linearised and (b) satisfy decoupled wave equations.333For these observations go back to Bardeen–Press [BP73] and Teukolsky [Teu72] in the physics literature and are easily generalised to , see [Kha83]. In our case, the two equations couple through the boundary condition imposed at the conformal boundary. The second key ingredient is to exploit the hierarchical structure of the double null gauge to prove boundedness of all geometric quantities in a gauge normalised with respect to initial data using the bounds for the Teukolsky quantities. As already mentioned, in contrast to the asymptotically flat case, all Ricci coefficients and null curvature components can be shown to decay without adding a residual pure gauge solution. The reason can be understood as follows. As in [DHR19], one first proves boundedness and decay of the linearised shear from the estimates for . This relies on the (commuted) redshift effect for . The quantity then inherits this decay through the boundary conditions and we can hence integrate in the ingoing direction from the boundary (using the estimates for ) to establish boundedness and decay for .444This second step is not possible in the asymptotically flat case as there is no boundary and the equation cannot be integrated directly from data in the ingoing direction either because of unfavourable -weights in the integrating factor. Decay for the other Ricci-coefficients and curvature components then follows by going hierarchically through the system analogous to [DHR19], except that here the boundary condition and its consequences need to be exploited at various stages. While one does not have to add a pure gauge solution to establish decay, one can improve the radial decay of certain geometric quantities and ensure that also the metric on the double null spheres converges to the round metric in standard form by adding one. Such a pure gauge solution is constructed from the trace of the original solution at the conformal boundary and controlled uniformly by initial data. See Theorem 4.14 below.
1.4 Acknowledgements
G.H. acknowledges support by the Alexander von Humboldt Foundation in the framework of the Alexander von Humboldt Professorship endowed by the Federal Ministry of Education and Research as well as ERC Consolidator Grant 772249. Both authors acknowledge funding through Germany’s Excellence Strategy EXC 2044 390685587, Mathematics Münster: Dynamics-Geometry-Structure.
2 Preliminaries
In this section we provide the necessary background to set up the problem. We define the manifold on which the analysis takes place, introduce the double null gauge and explain the linearisation procedure leading us to the system of gravitational perturbations on the Schwarzschild-AdS manifold. The boundary conditions for the system are then also derived from the non-linear theory. The section ends with a discussion of pure gauge and linearised Kerr-AdS solutions of the system of gravitational perturbations.
2.1 The manifold with boundary
Let be the -dimensional submanifold with (piecewise smooth) boundary defined by
(2) |
We define the associated -dimensional manifold
(3) |
equipped with coordinates , which we will refer to as Kruskal coordinates on . We denote the boundary components
(4) |
which will be referred to as null infinity, the future event horizon and the initial data hypersurface respectively. Observe that all boundary components are topologically . We denote by the -spheres in .
Given a fixed parameter we can define coordinates on by
(5) |
Note on and on . Defining also and we have another coordinate system on . We observe that is a boundary defining function for in that it vanishes at and , . One may parametrise the boundary by the coordinate . Finally, we can define on a function by the relation
(6) |
where . Note that the function depends on and that we have the asymptotics
(7) |
2.2 One-parameter families of aAdS metrics in double null gauge
Let us denote
(8) |
and fix , which we may think of as the mass of a Schwarzschild-AdS background metric we are about to install on . Given , equipped with local coordinates on , we consider a -parameter family of metrics expressed in double null gauge555Note that the is on here instead of on as in [DHR19]. As is well-known, this does not change the null-structure and Bianchi equations (collected in Section 2.3.5 below). The only change is in the propagation equation for the metric component , equation (25), which is now in the outgoing direction and with an additional minus on the right. An outgoing transport equation for is desirable as it can be integrated from data, where is normalised. Note also that it is which extends regularly to the horizon as can be seen by transforming (9) to the regular Kruskal coordinates.
(9) |
such that
-
1.
The satisfy on the Einstein equations with negative cosmological constant ,
(10) -
2.
We have that is the Schwarzschild-AdS metric of mass and cosmological constant
(11) where is defined as in (6) and denotes the round metric on the unit sphere.
-
3.
The are asymptotically Anti-de Sitter in that the function as well as the conformally rescaled metric extend regularly to , i.e. in particular they can be defined on the larger manifold . Specifically, recalling , the family
(12) defines a smooth family of -dimensional Lorentzian metrics on .
-
4.
The are all conformal to the Lorentzian cylinder , the latter being the metric induced on by the conformally rescaled Schwarzschild-AdS metric . In particular, the are all locally conformally flat.
-
5.
In regular Kruskal coordinates, the and regular derivatives thereof extend smoothly to the boundary of in the sense of Section 5.1.1 of [DHR19]. In particular, arbitrary concatenations of frame vectors from the set applied to extend regularly to .
The existence of such families is of course an implicit assumption. Locally in time, i.e. for a finite -interval this can be fully justified by a local-well-posedness theorem in double null gauge. Such a theorem could either be inferred from the literature [Fri95, EK19] or be proven directly by combining the (linear) estimates obtained in this paper with an appropriate contraction mapping argument. Since our main theorem is formulated directly as a statement concerning the linearised system (for which we will prove well-posedness directly) we will not address this issue further.
2.3 The geometry of a double null gauge
For the reader’s convenience we briefly recall the basic geometric notions in a double null gauge. The familiar reader can jump immediately to Section 2.6 while the reader unfamiliar with the double null gauge can consult Section 3 of [DHR19] or the original [Chr09] for many more details.
Associated with a double null gauge (9) on is a double null frame consisting of the null vectorfields
satisfying , , which is complemented with a local coordinate frame on , for , satisfying . Note and for .
2.3.1 -tensor algebra
Let be arbitrary one-forms and be arbitrary symmetric -tensors.
We denote by and the Hodge-dual on of and , respectively, and denote by the tensor obtained from by raising an index with . We define the contractions
and denote by the one-form arising from the contraction with . We finally define
where denotes the components of the volume form associated with on .
2.3.2 -projected Lie and covariant derivates
We define the derivative operators and to act on an -tensor as the projection onto of the Lie-derivative of in the direction of and respectively. We hence have the following relations between the projected Lie-derivatives and and the -projected spacetime covariant derivatives in the direction and respectively:
(13) |
and similarly for replacing by and by .
2.3.3 Angular operators on
Let be an arbitrary one-form and an arbitrary symmetric traceless -tensor on .
-
•
denotes the covariant derivative associated with the metric on .
-
•
takes into the pair of functions where and .
-
•
, the -adjoint of , takes any pair of scalars into the -one-form .
-
•
takes into the -one-form .
-
•
, the adjoint of , takes into -tensor .
2.3.4 Ricci-coefficients and curvature components
We define the non-vanishing null-decomposed Ricci coefficients as follows:
(14) |
The above objects are scalars, one-forms and symmetric traceless tensors respectively. In particular, they transform tensorially under a choice of frame on the sphere. It is natural to decompose into its -tracefree part (a symmetric traceless 2-tensor) and its trace , and similarly for . Note also the relations
(15) |
With denoting the Weyl curvature tensor of the metric (9), the null-decomposed Weyl curvature components are defined as follows:
(16) |
with denoting the Hodge dual on of . Again the above objects are -tensors (functions, vectors, symmetric -tensors) on .
2.3.5 The null structure and Bianchi equations
In the geometric setting outline above, the Einstein equations (1) imply (via the Bianchi equations and the geometric structure equations) a complicated system of coupled hyperbolic, transport and elliptic equations for -tensors that we collect below. See again [DHR19] and in particular [Chr09] for a detailed derivation of the equations in the case. We have highlighted the additional terms arising from the non-vanishing cosmological constant in (1) as boxed terms below.
We have the first variational formulae:
(17) |
The transport equations for the second fundamental forms take the form
(18) |
(19) |
(20) | |||
(21) |
(22) | |||
(23) |
The transport equations for the torsions and the (derivative of the) lapse become:
(24) |
(25) |
Finally, we have the elliptic relations on spheres:
(26) |
(27) |
(28) |
(29) |
Equations (2.3.5)–(2.3.5) are known as the Codazzi equations, (29) is the Gauss equation on . We finally collect the Bianchi equations for the null Weyl curvature components which are formally unchanged in the presence of a cosmological constant.
2.4 Boundary regularity and boundary conditions
We collect the asymptotic behaviour towards the conformal boundary for the geometric quantities on from the assumption that (by of Section 2.2) the metric extends regularly to the larger manifold . The proof of Proposition 2.1 is postponed to Appendix A.
Proposition 2.1.
2.5 The Schwarzschild-AdS background
In this section we discuss the Schwarzschild-AdS manifold . In complete analogy with [DHR19] we use unbolded notation to indicate quantities, for instance we write , and for the metric components, for the outgoing shear etc.
Moreover, all the constructions and definitions of Sections 2.3.1–2.3.4 may be repeated in unbolded notation. As this is done in detail in Section 4.3.1 of [DHR19] we only give brief summary.
2.5.1 Ricci coefficients, curvature components
The only non-vanishing Ricci-coefficients in the case are:
(35) |
where and is defined implicitly as in (6). In particular, and are -traceless with . The only non-vanishing null-curvature component is .
2.5.2 Differential operators and commutation formulae
We have the simplified coordinate formulae for the the projected Lie-derivatives for a general -tensor of rank :
For the projected covariant derivatives one finds
We recall the (unbolded) -angular operators ,, , , , now all defined with respect to the metric on .
We define the vectorfield
which is the static Killing field of Schwarzschild-AdS. The shall employ the notation . Note that since for the background Schwarzschild-AdS metric.
We finally collect the commutation formulae holding on tensors :
(36) |
which will be used frequently. As in [DHR19] we will define the angular operators (which commute trivially with , ) acting on symmetric traceless tensors as follows:
(37) |
Elementary elliptic theory on the round sphere (see Section 2.5.5) establishes that these operators have trivial kernel. Consistent with the above, we will here also allow to act on one-forms as .
2.5.3 Norms on the spheres
Let and denote the standard spherical coordinates on . We define the pointwise norm on tensors of rank by
A weighted norm on such tensors is then defined by
(38) |
Note the absence of a factor of in the integral, which if present would make (38) the induced norm.
2.5.4 The and modes
We recall the spherical harmonics on the round sphere, where and which form a basis of . The and spherical harmonic will play a distinguished role in our problem and are given explicitly by
A function is supported for is a function whose spherical means vanishes. A function supported for is a function supported for whose projection to the also vanishes.
We can also make sense of one-forms and symmetric traceless tensors being supported on specific modes as in Section 4.2.2 of [DHR19]. To summarise this, recall that any one-form can be written uniquely as
(39) |
for functions and of vanishing mean. (Note that only constants are in the kernel of ). We say that is supported for if both and in the above representation are supported on . Furthermore, we define the projection of to by the expression (39) where and are projected to .
Similarly, an symmetric traceless tensor can be represented uniquely by functions and supported on as
Note in particular that the kernel of the operator consists precisely of functions supported for only (as shown explicitly in [DHR19]). It is in the above sense that we can say that one-forms are supported for and symmetric traceless tensors for .
2.5.5 Basic elliptic estimates
We finally collect a few elliptic estimates that are immediate consequences of Section 4.4.3 in [DHR19]:
Proposition 2.3.
Let be a pair of functions supported on , then for any
(40) |
Let be an one-form supported on , then for any
(41) |
Let be an symmetric traceless tensor, then for any
(42) |
2.6 The linearisation procedure
Recall that in our set up of Section 2.2, the members of the -parameter family of metrics (9) all live on the same underlying manifold . Moreover, the hypersurfaces and are null for any metric in the family. In other words, the notion of -tensor is independent of and we can, in particular, add and subtract -tensors associated with different . If denotes an -tensor (a null-decomposed Ricci-coefficient or curvature component associated with the metric ) and denotes the corresponding tensor for , we define its linearisation by
(43) |
The linearised Einstein equations in double null gauge are then obtained formally in the following way: One writes down the null-decomposed Bianchi and null structure equations first for general (i.e. for the metric ) and secondly for (i.e. for the Schwarzschild-AdS metric) and then subtracts the respective equations, divides by and takes the limit inserting the definition (43). This then yields the system of gravitational perturbations in double null gauge as collected in Section 2.7. It should be noted that for most of the equations deriving the linearisation is trivial because many of the Schwarzschild-AdS background quantities vanish, which trivialises a significant number of null structure and Bianchi equations for .
2.7 The system of gravitational perturbations
In summary, the system of gravitational perturbations in double null gauge is encoded by the linearised metric quantities
(44) |
where and are defined from the linearised metric , which is in turn defined by (43), the linearised connection coefficients
(45) |
and the linearised curvature components
(46) |
Depending on the number of indices, the above quantities are scalars, one-forms and symmetric traceless tensors respectively. As in [DHR19], we will speak of a solution to the system of gravitational perturbations to mean a a collection of quantities
(47) |
satisfying the system (49)–(76) below, which we call the system of linearised gravity on the Schwarzschild background. Finally, it follows just as in Section 5.1.3 of [DHR19] (from the -parameter family of metrics being smooth in the extended sense) that the following linearised quantities extend smoothly to the horizon:
(48) |
The weight for does not appear in [DHR19] as in that paper, the shift satisfies an equation in the ingoing direction. See footnote 5.
2.7.1 Equations for the linearised metric components
The following equations hold for the linearised metric components, :
(49) |
(50) |
(51) |
(52) |
2.7.2 Equations for the linearised Ricci coefficients
For we have the equations
(53) |
(54) |
(55) |
(56) |
while for we have
(57) |
(58) | ||||
(59) |
We also have the (purely elliptic) linearised Codazzi equations on the spheres , which read
(60) |
For and we have the transport equations
(61) |
together with the elliptic equations on the spheres
(62) |
We finally have the transport equations for and
(63) |
(64) |
and the linearised Gauss equation on the spheres , which reads
(65) |
We also note that , the linearised Gauss curvature of the double null spheres satisfies (see (221) of [DHR19])
(66) |
2.7.3 Equations for linearised curvature components
We finally collect the equations satisfied by the linearised curvature components, which arise from the linearisation of the Bianchi equations:
(67) | ||||
(68) | ||||
(69) | ||||
(70) | ||||
(71) | ||||
(72) | ||||
(73) | ||||
(74) | ||||
(75) | ||||
(76) |
2.7.4 Projections to the and modes
Suppose we are given a smooth solution of the above system of gravitational perturbations. Then we may project all quantities of (see (47)) to and respectively (as defined in Section 2.5.4), thereby obtaining a collection of quantities denoted by and respectively. One now readily checks that and solve the system of gravitational perturbations individually.666More abstractly, this is a consequence of the spherical symmetry of the background (in particular projection operators commuting with and ) and the linearity of the equations. We can therefore decompose
with the last term defined by the equation. This decomposition will later allow us to deal with the part of the solution independently (as far as initial data and boundary conditions are concerned), which will turn out to be convenient, as the part of the solution can be computed (more or less) explicitly.
2.8 Boundary conditions for the system of gravitational perturbations
Recall that the boundary at infinity, , is not part of our interior manifold . On the other hand, to formulate boundary conditions (on certain weighted quantities of in (47)) we will need to consider tensors on the Schwarzschild-AdS manifold , which extend smoothly to , i.e. to the larger manifold . To keep notation clean, we will often simply write or to denote the appropriate limit of such tensors on . Recall in this context from (38) that the norm is independent of the radius of the sphere .
The boundary conditions for the non-linear spacetime null-curvature components (see Proposition 2.2) can easily be linearised,777Equations (31)-(34) are all trivial to linearise since the quantities all vanish for the background Schwarzschild-AdS metric. leading to the following definition.
Definition 2.4.
We will say that a smooth solution of the system of gravitational perturbations satisfies conformal boundary conditions provided we have for any the limits
(77) | ||||
(78) | ||||
(79) | ||||
(80) |
Here the tensorial limits are to be understood componentwise in an orthonormal frame on the spheres .
Remark 2.5.
We close the section with one more definition, which translates the asymptotic behaviour of the non-linear geometric quantities collected in Proposition 2.1 to the linearised setting.
Definition 2.6.
We will say that a smooth solution of the system of gravitational perturbations is asymptotically AdS in the linearised sense if the following quantities as well as arbitrary many derivatives from the set extend to the conformal boundary :
(82) |
and
(83) |
Remark 2.7.
The bounds on (82)–(83) should be thought as having been derived by linearising the non-linear statement in (30). In fact, from the bounds on (82)–(83) we can (and will) also deduce bounds for the difference quantities , and consistent with (30) later in the paper, see (175) and Remarks 5.16 and 5.22. For simplicity, we have not included them in the above definition.
2.9 Special solutions
2.9.1 Pure gauge solutions
There are special solutions to the system of gravitational perturbations (49)–(76) corresponding to infinitesimal coordinate transformation that generate a change of double null gauge (i.e. a choice of nearby sphere and corresponding foliations of the associated ingoing and outgoing cone). In complete analogy to [DHR19] we call these pure gauge solutions. In our setting, the additional requirement that the pure gauge solutions should preserve the boundary conditions reduces the admissible pure gauge solutions and they can in fact be parametrised by a single scalar function.
Lemma 2.8.
Given an arbitrary smooth function , the corresponding functions and , interpreted as functions on independent of one of the coordinates, generate the following (pure gauge) solution of the system of gravitational perturbations on :
and
The solution satisfies the conformal boundary conditions of Definition 2.4. We will call a gauge function and denote the corresponding pure gauge solution by . Finally, if extends smoothly to then so does the associated pure gauge solution.888In particular, the quantities (48) extend smoothly to .
Proof.
There is a further pure gauge solution which only changes the linearised metric quantities but leaves linearised Ricci-coefficients and curvature components invariant:
Lemma 2.9.
For any smooth functions and the following is a pure gauge solution of the system of gravitational perturbations
(84) |
while all linearised Ricci and null curvature components vanish identically. We denote the solution by .
2.9.2 The family of linearised Kerr-AdS solutions
It is well-known that the Schwarzschild-AdS family sits as a -parameter family in the larger -parameter family of Kerr-AdS metrics. At the linear level there exists (due to the spherical symmetry of the background) a -dimensional (choosing an axis and a magnitude) family of explicit solutions that move the Schwarzschild-AdS metric to a nearby Kerr-AdS metric. Moreover, there is also the -parameter family of changing the mass. We summarise both in Lemma 2.10 below. Let us already remark that for the modes, the pure gauge solution takes a significantly more complicated form compared to the asymptotically flat case. The underlying reason is that the variable (defined in terms of the (fixed) Eddington-Finkelstein coordinates ) by (6) depends implicitly on the mass, with a dependence that is more involved than in the asymptotically flat case. See Appendix B for computational details regarding the mode.
Lemma 2.10.
For any and the following linearised metric quantities generate a smooth solution of the system of gravitational perturbations on :
(85) |
The solution has the following non-vanishing Ricci-coefficients and curvature components:
(86) |
Moreover for any , the following linearised metric quantities generate a (spherically symmetric) smooth solution of the system of gravitational perturbations on (where we have set )
(87) | ||||||
(88) |
In particular,999For convenience we collect the values of all Ricci-coefficients and curvature components in Appendix B. this solution satisfies
(89) |
We call the first type of solution a linearised Kerr-AdS solution with fixed mass and the second solution a linearised Schwarzschild-AdS solution. Together these solutions form a -parameter family of linearised Kerr-AdS solutions. Given parameters we denote by the sum of the four corresponding linearised Kerr-AdS solutions.
Remark 2.11.
2.9.3 Regularity at the horizon and at infinity
We close this section noting that our special solutions are regular at the horizon and asymptotically AdS in the linearised sense:
Proposition 2.13.
The following are smooth solutions of the system of gravitational perturbations, which are moreover asymptotically AdS in the linearised sense (Definition 2.6), satisfy the boundary conditions (77)–(80) and are such that the quantities (82), (83) extend smoothly to .
-
•
The four-parameter family of linearised Kerr-AdS solutions of Lemma 2.10.
-
•
The pure gauge solutions of Lemma 2.8, provided the function generating them is smooth on and is smooth in the extended sense at .
-
•
The pure gauge solutions of Lemma 2.9, provided the functions generating them are smooth on and are smooth in the extended sense at .
Proof.
Both linearised Kerr-AdS and pure gauge solutions are by construction smooth solutions to the system of gravitational perturbations. The validity of the boundary conditions is straightforward to check for the linearised Kerr-AdS solutions (note in particular ). For the pure gauge solutions we only need to check the boundary condition for which follows from vanishing at the conformal boundary (where ). It remains to check the extension of the quantities (82) and (83) towards the horizon and infinity. Near the horizon the regularity claims are easily verified. Near infinity, the claim on for linearised Kerr-AdS solution follows by carefully expanding the integrand in (87) and observing a cancellation of the leading order term in (88). For the pure gauge solution, the hardest to check is that extends to the boundary. For this we Taylor expand the expression of Lemma 2.8
where we have recalled and and used (7) in conjunction with . The conditions on higher derivatives implicit in Definition 2.6 are straightforward to check. The last item is immediate and and this finishes the proof of the proposition. ∎
2.10 The Teukolsky equations
Remarkably, the extremal linearised curvature components and , which by Lemma 2.8 and 2.9 vanish identically for pure gauge solutions, satisfy decoupled equations. These are the well-known Teukolsky equations. We now derive these equations and define the associated gauge invariant hierarchies and . See our companion paper [GH24] and Section 4.1 below for analytic results on the Teukolsky system.
2.10.1 Derivation of the equations
Lemma 2.14.
Given a smooth solution of the system of gravitational perturbations the quantities and satisfy the decoupled Teukolsky equations:
(90) | ||||
(91) |
2.10.2 The gauge invariant hierarchy
Given symmetric traceless tensors , we define (consistently with [DHR19]) the quantities by:
(94) | ||||
(95) |
In the above, the second equality follows from a rewriting of the Bianchi equations (67) and (76) respectively. Note that we can also rewrite (95) in a form where all terms extend smoothly to .
(96) |
Again consistent with [DHR19] we define also the higher order quantities by
(97) | ||||
(98) |
Here the last equality follows by plugging in (94) and (95) and diligently inserting the relevant null-structure and Bianchi equations produced by the extra derivative. Defining , one has from (90), (91) and the definitions (97), (98), that satisfy the Regge-Wheeler equation (see [GH24])
(99) |
where . We further define101010Recall from Proposition 4.4.4 of [DHR19] that has eigenvalues acting on symmetric traceless tensors.
(100) |
It is easy to see that, if and are regular at the horizon and at infinity (see (48) and Definition 2.6), and and hence also and extend regularly to both and . In fact, the quantities and correspond (up to an unimportant numerical constant) to the analogous quantities and defined in [GH24], where they are spin weighted functions. We refer the reader to [GH24] for further details.
3 Construction of initial data and local well-posedness
In this section we define the class of solutions of the system of gravitational perturbations that will be the relevant class for the main theorem of the paper. This is Definition 3.1. The remainder of the section is concerned with constructing such solutions from an appropriate notion of seed initial data on by solving an initial boundary value problem. The reader wishing to take for granted the existence of this class of solutions upon first reading may move directly to Section 4 after having read Definition 3.1.
3.1 The class of solutions
In the following, we will consider a class of solutions of the system of gravitational perturbations.
Definition 3.1.
As mentioned, our goal is to construct such solutions uniquely from an appropriate notion of smooth seed data on .
Remark 3.2.
The solutions we will construct (and hence the data) will have the additional regularity property of being aAdS in the linearised sense, see Definition 2.6. In fact, we will prove uniform bounds on all quantities appearing in (82) and (83) in Section 5 of the paper. However, we have not included the condition of being aAdS in the linearised sense in Definition 3.1 to make showing existence of solutions easier.
3.2 Smooth seed initial data
We will now define the notion of smooth seed initial data along the cone . Below, an -tensor is called smooth along if for all , the tensor extends smoothly to infinity and to the horizon along .111111In particular, the components in an orthonormal frame extend as smooth functions to .
Definition 3.3.
A smooth seed initial data set on for the system of gravitational perturbations consists of
-
1.
a tuple , called the gauge independent part,
-
2.
a tuple , called the gauge dependent part,
-
3.
a -dimensional vector called the Kerr-AdS part.
The gauge independent part consists of
-
•
a smooth symmetric traceless tensor prescribed along with extending smoothly to ,
-
•
smooth symmetric traceless tensors and .
The gauge dependent part consists of
-
•
scalars , , , with supported for ,
-
•
a smooth symmetric traceless tensor on ,
-
•
a smooth lapse function prescribed along with extending smoothly to ,
-
•
a smooth shift function prescribed along ,with extending smoothly to .
Some remarks are in order regarding the interpretation of the quantities prescribed. Suppose we can indeed construct, from a seed initial data set as above, a solution as in (47) of the system of gravitational perturbations satisfying the boundary conditions (as we will eventually in Theorem 3.9 below). Then we would like that solution to be related to the seed data as follows.
Definition 3.4.
Given a smooth seed initial data set
we say that a solution of the system of gravitational perturbations satisfying the boundary conditions in the sense of Definition 3.1 realises the given seed initial data if we have , and along and
(101) | ||||
(102) | ||||
(103) | ||||
(104) | ||||
(105) | ||||
(106) | ||||
(107) |
Finally,
It is clear that the quantities are gauge independent. The terminology “gauge dependent” for and “Kerr-AdS” for becomes clear from the following proposition. To state it, note that clearly any solution of the system of gravitational perturbations satisfying the boundary conditions induces a seed initial data set by restricting the solution to and taking the above limits.
Proposition 3.5.
Consider a smooth seed initial data set
and assume there exists a solution of the system of gravitational perturbations satisfying the boundary conditions and realising the given seed data in the sense of Definition 3.4. Then there exists a function , a function such that the solution
induces a seed initial data set whose gauge dependent part and Kerr-AdS part vanishes identically. Here is the pure gauge solution induced by as in Lemma 2.8, the pure gauge solution induced by as in Lemma 2.9 and the linearised Kerr-AdS solution of Lemma 2.10.
Proof.
Step 1. Subtracting the linearised Kerr-AdS solution . We note that
We choose the in the linearised Kerr-AdS solution such that for we have
Note that both these quantities are gauge invariant and hence not affected by adding and in the following steps. Clearly, by construction induces seed data with vanishing Kerr-AdS part.
Step 2. Defining . Given the solution , we define the function inducing the desired pure gauge solution (by setting and in Lemma 2.8) as follows:
(108) |
for scalars defined in turn by121212The are unique up to the spherical means of . Note that the mode of generates a trivial pure gauge solution (changing and by a constant) corresponding to the static isometry of the background.
-
•
such that ,
-
•
such that ,
-
•
such that .
We claim that with thus defined, the solution already has and , and . To verify this, we first note that
Using the expressions in Lemma 2.8 we now easily check that
(109) |
on the ingoing cone independently of the . From Lemma 2.8 we then verify
(110) |
To check the condition on , we first compute the limit
It therefore follows from Lemma 2.8 that we have
(111) |
and we see that if is chosen as above. For the last limit we note using Lemma 2.8 and the fact that we already obtained , the equalities
We want to take the limit . We compute up to terms vanishing in the limit (indicated by )
(112) |
We conclude
(113) |
Step 3. Defining . Let us define (unique) functions with vanishing spherical means by . Set to solve
with initial conditions at determined by and having vanishing spherical means and solving and . Note that this determines uniquely up to modes of on the sphere (corresponding to the rotational invariance), which generate trivial gauge transformations. It is now immediate from Lemma 2.9 that , and , while all other metric, Ricci and curvature components remain unaffected. ∎
3.3 Construction of all geometric quantities on the initial cone
Proposition 3.6.
Given a smooth seed initial data set
one can construct uniquely all geometric quantities (47) of the system of gravitational perturbations (including all their tangential and transversal derivatives) on such that
- (1)
- (2)
-
(3)
The condition holds on the sphere for any .
-
(4)
On , the relations of Proposition 3.4 between the seed data and the geometric quantities hold.
Moreover, if the seed initial data set vanishes identically then so do all geometric quantities on .
Remark 3.7.
To explain above, we observe that some condition on at is necessary to determine all quantities on the cone as the quantity only admits a null structure equation in the -direction and hence needs to be supplemented with data on . Our choice in is weaker but consistent with our desire to construct data and solutions that are being aAdS in the linearised sense (Definition 2.6).
Proof.
The logic of the proof is to construct all the geometric quantities of the solution using freely equations and relations that have to hold on by Items above. In a second step, once all geometric quantities have been constructed without contradiction, we verify that all equations and relations have been used in the construction.
First, we define consistent with (52). Note that extends smoothly to .
Since all of (49)–(76) have to hold for the geometric quantities we want to construct, it follows that also the Teukolsky equations (90)–(91) restricted to must hold. This determines (by transport along ) the regular transversal derivatives and along all of in terms of the prescribed seed data and , using (101), (102).
Next, we impose that on the boundary sphere ,
(114) |
which, since the mode of is determined by the seed data on , determines at infinity from seed data. Since the seed data require , the quantity is determined uniquely by elliptic theory. We now define the limit by
(115) |
consistently with (101). We can now obtain along all of by integrating (75) written as
Note that the right hand side is integrable in on and thus determines . Similarly, we can integrate (57) from infinity with boundary condition (115) to obtain along . This comes with the (smooth) expansion
(116) |
Note that we can now determine from (73) and the boundary condition .
From (56) we have
Since extends smoothly to the right hand side is integrable. Using the boundary condition (106) this defines on all of . Using (71), we can also determine from at infinity (set by seed data) and the bounds on and . We now write the Codazzi equation (60) as
(117) |
All four terms have a finite limit on null infinity ( actually vanishes) which therefore determines (which extends regularly to ). We also set, consistent with (52),
(118) |
One easily shows that the -equation for , (61), holds as a consequence of (117) and the -equations holding for , , and . Moreover, using (61) and the definition of by (73), one can show that . Now, both quantities in the parenthesis vanish at infinity, which shows that the elliptic equations (62) hold on . Indeed, by definition of and, by (114) and (115), one has , which, by the fact that as and (118), shows that as .
We next determine from (69) and the previously defined quantities, using the boundary condition at infinity . To determine we write (58) as
(119) |
Using (116), the right hand side is integrable and we define by imposing the condition when .
We can now determine the last curvature component from (67) and the fact that at infinity. Note that with this all Bianchi equations in the -directions hold by construction.
The definition of follows from (64) and the condition that on .
Consistently with (102) and the linearised formula for the Gauss curvature (66), we define the asymptotic linearised Gauss curvature to be
(120) |
where the second identity comes from the previous definitions of and . This determines . Now recall that we have and by construction, which in particular determines from formula (66). Conversely, formula (66) determines when read as an elliptic equation for this quantity, since is prescribed at infinity. Finally, (49) and (50) determine determine and on all of .
We next use the linearised Gauss equation (65) to determine on . Note that this shows in particular that at . Differentiating the linearised Gauss equation in we verify that (54) has to hold on . To show validity of the other Codazzi equations of (60), which we rewrite as
(121) |
on , we first use the boundary asymptotics on and , the definition of and the asymptotics of , and the validity of the underlined Codazzi equation (117), to verify that (121) holds on , and then differentiate in : Since the resulting expression vanishes after inserting the evolution equations already established, we deduce that (121) indeed holds on .
We have thus determined all quantities from seed data with quantitative estimates and obtained validity of all evolution equations in the -direction as well as elliptic relations on spheres contained in (49)–(76). Obviously, tangential derivatives to applied to these equations also hold.
Moreover, all relations of Definition 3.4 and all the boundary conditions of Definition 2.4 hold as they have been explicitly used to define quantities. To determine finally the transversal derivatives, we use that any geometric quantity (except and ) satisfies a Bianchi or null structure equation which determines its derivative in terms of angular (=tangential) derivatives. This consistently constructs all transversal derivatives and ensures validity of the relevant equations by definition. For the exceptional we can determine transversal derivatives from the Teukolsky equation (90) and the boundary conditions relating and the seed data at . For we have can obtain the -derivative from the fact that we can determine the -derivative from the boundary condition in Item and the commuted equation for , equation (64).
The last claim about the trivial data follows easily from redoing the above proof with trivial data. ∎
Corollary 3.8.
Consider a smooth seed initial data set
and assume there exists a solution of the system of gravitational perturbations satisfying the boundary conditions, realising the given seed data in the sense of Definition 3.4 and satisfying condition (3) of Proposition 3.6. If the gauge independent part of the solution vanishes (which in particular happens if the solution is supported for ), the solution is the sum of a pure gauge and a linearised Kerr-AdS solution.
Proof.
We add a pure gauge and linearised Kerr-AdS solution as in Proposition 3.5 to the solution such that also the gauge dependent and the Kerr-AdS part of the seed data vanish. By Proposition 3.6 (which applies since the data induced by both and on satisfy ) we conclude that all geometric quantities of vanish on . In particular, Lemma B.1 applies and the solution must have vanishing mode. Moreover, since and vanish identically on , it follows from the decoupled Teukolsky equations (90)–(91) and uniqueness of its solutions (see for instance Theorem 1.4 of [GH23] that and vanish identically globally in . Since and vanish on , they vanish globally by their evolution equation in the -direction. Inserting the boundary conditions we conclude vanishing of and on and by their evolution equation in the -direction, globally. Similarly, vanishes by the equation (72) and the vanishing on . From (69) and (74) and the vanishing of we now conclude . The equation (61) and the vanishing of on produces global vanishing of and hence of individually. Revisiting (69) it follows that vanishes (recall we have established in the beginning that the mode vanishes). Codazzi shows the global vanishing of and and (63), (64) that of and respectively. The proof is complete. ∎
3.4 Local Well-posedness
We can finally state the well-posedness theorem for solutions to the system of gravitational perturbations:
Theorem 3.9.
Remark 3.10.
Proof.
131313The authors would like to thank Leonhard Kehrberger for discussions and suggesting the argument with in Step 2 below.The uniqueness part follows from the proof of Corollary 3.8 where it is shown that a zero seed initial data set can only produce the zero solution. We also know from the proof of Proposition 3.5 that the given seed data restricted to agrees with the seed data induced by an appropriate pure gauge solution supported on plus the seed data of a linearised Kerr-AdS solution which establishes existence for .
In summary, we only need to construct the solution for from seed data supported for .
Step 1: Constructing geometric quantities in , Part I. We first construct from the seed data set all quantities on as in the proof of Proposition 3.6. This in particular determines smooth , on such that in particular the master energy defined in (132) is finite for all . We can hence apply the well-posedness theorem for the Teukolsky system (90)–(91) (cf. Theorem 1.4 in [GH23]) and obtain smooth and globally on . We next determine and globally from their transport equation in the -direction, i.e. integrating (57) and (68) from data. We then determine and globally by integrating the transport equations (57) and (75) from the boundary (using the boundary conditions as initial conditions for and , i.e. and on ). With this, (68) holds by definition and it is easy to see that (67) also holds because it holds on data and of this equation vanishes by the fact that the Teukolsky equation (90) holds for and the equations (57), (68) hold in the -direction by construction. Similarly (76) holds on the boundary (by the fact that (67) holds there and the boundary conditions imposed) and its -derivative is zero by the validity of the Teukolsky equation and (75) and (57). Observe also that all quantities constructed are smooth up to the boundary. This will continue to be true for the quantities constructed below.
We next define the quantity by integrating (72) from initial data. Note that with this definition, has a finite limit on the boundary (which we do not know vanishing of yet). We then set and so that (62) is satisfied. On the other hand, we determine and globally from (58), (59) by taking the of these equations and observing that with having trivial kernel on the space of functions with . This determines and uniquely by standard elliptic theory and one can also show that holds on the conformal boundary .141414To see the latter, follow the proof of Proposition 5.10 below to establish (176), from which the claim follows after subtracting (58) and (59) multiplied by .
Step 2. Verifying the -equations. With the quantities defined we can already verify some of the equations. We claim that the applied to the Codazzi equation (60) for holds. To see this, note that
by the propagation equations we have defined in the -direction. Since the quantity in round brackets also vanishes on we conclude that it is zero globally and hence applied to the Codazzi equation (60) for holds. We similarly conclude that the applied to (69) holds: Indeed, defining
we derive (using the transport equations in the -direction for and as well as (67) and the just established applied to the Codazzi equation (60) for ) an equation of the form . Since on we conclude that globally.
We next define an auxiliary quantity such that and
(122) |
Note that agrees with on and on . We verify that
and since the quantity vanishes on (by the previously imposed/derived boundary relations and the Codazzi equation for ), the expression in round brackets vanishes globally. Defining
(123) |
we first check that holds (a computation similar to the one for above) and then verify that vanishes at infinity (which follows from which is in turn a consequence of how we defined ). The two observations allow us to conclude globally. We can finally conclude that satisfies the Regge-Wheeler equation (99) by applying to (122) and inserting (123). In summary, and satisfy the same wave equation with the same data and boundary condition at and hence agree globally, i.e. on . In particular, Bianchi equation (73) holds by (122) and, adding the two Bianchi equations (72), (73), one has that . Thus, using that on the initial sphere at infinity vanishes, one has that vanishes globally on . At this point we have ensured the validity of the Bianchi equations (67), (68), (72), (73), (75), (76) as well as the of (69) and – by the vanishing of (123) and – (74) respectively. By construction also the equations (62) and the of (both of) (60) hold. One easily checks that the of (61) holds. Moreover, we check that applied to (58) and (59) respectively hold (inserting the of (60) and using the propagation equations already established). This means that (58) and (59) hold unconditionally since the part of these holds by construction.
Step 3: Constructing geometric quantities in , Part II. We now set , which is well defined as by our definition of above. Moreover, we set and . Next we define by (60) (which is well defined as we have already verified that the of this equation holds) and directly integrating the evolution equation (53) from data. We define the quantity by integrating (70). Finally, the metric quantities , , are defined by integrating their equation in the -direction from data, i.e. (49), (50) and (51).
Step 4. Verifying the remaining equations. Differentiating (58) with respect to and (59) with respect to shows that (61) must also both hold globally (recall we are on ). From our definition of and also the equations
(124) |
hold. Using (61) we can also conclude that the of (69) (and hence (69) unconditionally since we already verified the -equation) must hold globally because it holds on data and is propagated in the -direction.
Differentiating the Codazzi equation for with and shows that also (54) and (55) hold. Differentiating (55) with respect to shows that (64) holds and (63) follows from the fact that by the way we defined , . One now verifies that
(125) |
holds on the cone and , which implies in , which in turn implies that (71) holds. We next verify that (74) holds by noting that it holds on by the boundary conditions and is propagated in the -direction. Indeed, we have
on after replacing and inserting the Bianchi equations (68), (75) as well as the boundary conditions for and , . Since also and hold on we see that (74) is equivalent to (69) on . Differentiating now in one obtains after inserting the equations that have already shown to hold. As on we conclude globally.
Now the -Codazzi equation (60) and the equation (56) can be verified by noting that they hold on and are propagated in the -direction.
We have now verified that our constructed solution satisfies all equations of the system of gravitational perturbations except the Gauss equation (65) and the -equations for the metric components. But all these equations hold on and applying and inserting the already established equations one verifies they propagate to hold in all of . This finishes the proof of the proposition. ∎
3.5 The initial data normalisation
Consider a given seed data with associated solution from Theorem 3.9. The main objective of this section is to construct a pure gauge solution from the given seed data which when added to achieves a certain normalisation of the solution at the horizon. This normalisation will be crucial in the main theorem. We begin by defining the normalisation followed by a proposition establishing that it can be achieved.
Definition 3.11.
Consider a smooth seed initial data set
as in Definition 3.3 and let be the unique solution of the system of gravitational perturbations arising from Theorem 3.9. We say that is initial data normalised if the following holds for on the ingoing initial cone :
-
•
and on ,
-
•
and and ,
-
•
and on the horizon sphere .
Moreover, we call the solution initial data normalised with vanishing modes if in addition the modes of all geometric quantities of vanish.
The point is that we can always achieve the initial data normalisation:
Proposition 3.12.
Given a solution arising from a smooth seed initial data set as in Theorem 3.9 there exists a pure gauge solution (computable in terms of the seed data) and a linearised Kerr-AdS solution such that is an initial data normalised solution with vanishing modes.
Proof.
The proof is a small variation of the proof of Proposition 3.5. From Proposition 3.5 there exist and a linearised Kerr-AdS solution such that has trivial seed data and in particular vanishing modes. We now add to this another (supported for ) generated by
(inducing and in Lemma 2.8), where the are functions on the unit sphere satisfying the relation . It is clear from the proof of Proposition 3.5 that still satisfies on and on (hence by (56) on all of ). In view of
and , we now compute from Lemma 2.8 the horizon sphere relations
Recalling that we can eliminate from the relation it is an algebraic exercise to determine and such that both and hold. While might have altered we can simply repeat Step 3 of Proposition 3.5 and add a which ensures that satisfies all of the desired properties. Setting and we are done. ∎
4 The main results
We can finally give a precise formulation of our main results. In Section 4.1 we first recall the results from our companion paper [GH24] where boundedness and decay bounds on the Teukolsky system (90)–(91) have been obtained, independently of the system of gravitational perturbations. This is Theorem 4.2 below. These bounds will play a key role in proving our main theorem, which is stated in Section 4.2 as Theorem 4.7.
4.1 Estimates for the gauge invariant quantities: The Teukolsky equations
Estimates for solutions to the Teukolsky system (90)–(91) satisfying the boundary conditions (77)–(78) have been obtained in our companion paper [GH24]. We first formulate these results in a form most suitable for the present paper. We recall that in [GH24], the Teukolsky equations were expressed in an equivalent form as equations for spin weighted functions instead of symmetric traceless tensors. We briefly recall that equivalence and refer the reader for instance to Section 6 of [HS16] for more details.
4.1.1 Spin-weighted functions vs. symmetric traceless tensors
Given the tensors and and a local orthonormal frame on the sphere we define the complex scalars
(126) |
which transform like spin-weighted functions of weight under a change of orthonormal frame on . For the specific frame and one obtains the Teukolsky equation for as stated in [GH24] by expressing the equations (90) and (91) in frame components. We also note in the notation of [GH24] the relations
(127) |
Clearly the estimates on obtained in [GH24] directly translate into estimates for the (norms of the) tensors and .
4.1.2 Norms and energies for the gauge invariant quantities
To state the estimates of [GH24] in a form most useful for the present paper we first introduce certain energies on null cones. The underlying reason is that estimating quantities in the system of gravitational perturbations in a double null gauge will typically require control on fluxes on null hypersurfaces.
To keep notation concise regarding commutations we use the following shorthand notation for derivatives:
where the second sum is over all tuples with .
We first define the non-degenerate (near the horizon) outgoing and ingoing commuted energy fluxes (note that the superscript denotes the number of derivatives involved) of a general -tensor :
(128) |
as well as the outgoing degenerate energy:
(129) |
The above energies will appear for the (regular both at the horizon and the conformal boundary ) quantities
(130) |
We shall also need an auxiliary energy on spheres at the conformal boundary , which arises in the renormalised energy estimates of [GH24] and is defined only for :
(131) |
We finally define (for ) the following initial data master energy on cone :
(132) |
which contains the energy fluxes of and , the flux of and a contribution on the sphere at infinity. It is this modified energy which has been shown to propagate for the Teukolsky system in [GH24], see Theorem 4.2 below.
Remark 4.1.
One could add the terms to the energy (132). However, these terms can be shown to be controlled by the first two terms and have hence been omitted.
4.1.3 Estimates for the Teukolsky quantities
From the main theorem of our companion paper [GH24], we now easily infer the following theorem by translating the estimates on spacelike slices in [GH24] to estimates on null cones.
Theorem 4.2.
We have the following estimates for any :
-
•
Boundedness estimate: For fixed we have
(133) -
•
Decay estimates: Fix an . For fixed we denote by the -value of the intersection of and the fixed -hypersurface. We then have for
(134) with the now depending on the (the implicit constant blows up as ).
-
•
Estimates for the non-degenerate outgoing fluxes near the horizon: We have for any , and the estimates
(135) (136)
Remark 4.3.
Integrated decay estimates follow as a corollary by integrating the ingoing fluxes in . The integrated decay has again growth like .
Remark 4.4.
Note that it is the degenerate outgoing but non-degenerate ingoing energy appearing in the boundedness statement. For the decay estimate, the degenerate and the non-degenerate flux are equivalent because the outgoing flux is always restricted to . In general, the non-degenerate outgoing flux has growth as stated in (135)–(136). To see why the non-degenerate ingoing flux behaves better, we recall the (timelike) redshift vectorfield from [DR08] which generates an energy identity whose bulk term has a good sign in a region for some . We apply the energy identity in a region bounded by a -slice, the horizon and an ingoing cone emanating from the intersection of the slice with the hypersurface. This produces control on the desired ingoing flux noting that for the degenerate energy is equivalent to the non-degenerate one. Applying globally also yields (135)–(136) immediately using (• ‣ 4.2) and (• ‣ 4.2) respectively (integrated in time) to control the error in } in the corresponding vectorfield identity.
Corollary 4.5.
We have for the following estimates on spheres:
-
•
Boundedness estimates
(137) -
•
Decay estimates
(138)
Remark 4.6.
Truly pointwise estimates follow immediately from Sobolev embedding on spheres but are not stated explicitly. We also note that the above estimates are clearly not optimal, as we allow ourselves to lose one derivative for the embedding and another one for decay.
Proof.
This follows from -dimensional Sobolev embedding along the ingoing cones for which we control a non-degenerate energy by the previous proposition. ∎
4.2 The statement of the main theorem
To state the main theorem, we recall the energies involving the gauge invariant quantities introduced in Section 4.1.2. We require one additional (gauge dependent) initial data energy involving derivatives of the Ricci coefficients. For we define
(139) |
To state our main boundedness and decay theorem we define the initial master energy involving derivatives of curvature and Ricci-coefficients
(140) |
Theorem 4.7.
Given a solution of the system of gravitational perturbations satisfying the boundary conditions as arising from a smooth seed initial data set as in Theorem 3.9, let be the initial data normalised solution with vanishing modes obtained from Proposition 3.12. Let the initial energy in (140) be defined with respect to the geometric quantities of . Then the geometric quantities of the solution satisfy the following estimates. For any weighted Ricci or metric coefficient
(141) |
and any curvature component
(142) |
we have for and any
(143) | ||||
(144) |
Moreover, for the curvature components , we also obtain for any fixed, the following estimates for the top order fluxes:
(145) | ||||
(146) |
for any . Finally, for fixed, one may drop the factor of in (146) if . In this case, the in (146) will depend on .
Remark 4.8.
Remark 4.9.
The last sum in (143) and (144) expresses the fact that if we are willing to lose a derivative, we can show stronger -weighted estimates for and even stronger ones for if we are willing to lose two derivatives. A similar improved estimate with loss holds for , (see Proposition 5.15) but has not been included explicitly in the main theorem.
Remark 4.10.
The last statement after (146) can be paraphrased by saying that the top order outgoing flux is uniformly bounded provided the outgoing cone is uniformly away from the horizon.
Remark 4.11.
Remark 4.12.
Note that contrary to the asymptotically flat case, one obtains here decay of all Ricci coefficients and curvature components even without adding a residual pure gauge solution.
Finally, we note that Sobolev embedding on spheres gives the following corollary.
Corollary 4.13.
We have the following pointwise bounds:
(147) |
(148) |
(149) |
4.3 Future normalising the solution at the conformal boundary
We can improve the radial decay in our estimates on the solution if we normalise the solution with respect to the conformal boundary by adding a pure gauge solution. The precise statement is the following:
Theorem 4.14.
With the assumptions of Theorem 4.7, there exists a further pure gauge solution such that the geometric quantities associated with the corresponding solution satisfy the estimates of Theorem 4.7 but now for
Moreover, the pure gauge solutions , used in the above is uniformly bounded by initial data in the sense that the geometric quantities of the pure gauge solution , satisfy the estimates of Theorem 4.7.
Remark 4.15.
Note the improvement in -weights which is a manifestation of the fact that the solution is now normalised at the conformal boundary. In particular, the metric perturbations now vanish identically on the conformal boundary.
5 Proof of the main theorem
5.1 Brief overview
As in the asymptotically flat case, the proof exploits the hierarchical structure of the system of gravitational perturbations in the double null gauge. In Section 5.2 we prove the basic transport lemmas that will be invoked throughout the proof when integrating along null cones. Since we will always consider the geometric quantities of , which have vanishing modes, the elliptic operators have trivial kernel when acting on such a quantity and hence allow estimating the entire -Sobolev norm of angular derivatives (recall Section 2.5.5). In Section 5.3 we obtain control on certain horizon fluxes of non-gauge invariant quantities from the gauge invariant quantities. These are used in Section 5.4.1 to prove spacetimes boundedness and decay estimates for the ingoing linearised shear. The outgoing linearised shear is then estimated in Section 5.4.2 using the boundary condition and the transport equation along the ingoing direction. The estimates on the shears allow to estimate various additional components in the system, discussed in Section 5.5. However, these estimates are somewhat non-optimal in terms of regularity because estimating the ingoing shear required commutation with two transversal derivatives. The regularity is recovered in Section 5.6 applying again a hierarchy of propagation equations and the bounds already obtained. We conclude the proof of the main theorem in Section 5.8 after estimating the metric coefficients in Section 5.7.
5.2 The transport lemmas
Lemma 5.1.
Let be an tensor satisfying the propagation equation
(150) |
along the ingoing cone (which intersects in the sphere ). Assume satisfies
along the cone. Then, provided , we have
Moreover, the statement remains true replacing by everywhere.
Proof.
Direct consequence of Cauchy-Schwarz and integrability in of . ∎
Lemma 5.2.
Let be an tensor satisfying the propagation equation
(151) |
along the outgoing cone (which intersects in the sphere ). Assume satisfies
(152) |
for any along the cone and also for some fixed the bound
(153) |
Then we have
(154) |
Proof.
We write (151) as
(155) |
and hence, contracting with and applying Cauchy’s inequality with an and an absorption argument we get
(156) |
We have and therefore integrating between and yields
(157) |
An elementary calculus exercise yields the conclusion (154) without the factor of . To improve the weight near infinity we can integrate (151) directly from (where we have already proven the desired bound) using Cauchy-Schwarz and (153) only as in the proof of Lemma 5.1. ∎
Lemma 5.3.
Under the assumptions of Lemma 5.2, if satisfies in addition
(158) |
for any along the cone and also for some fixed the bound
(159) |
then satisfies along for any the decay bound
(160) |
Proof.
Simple variation of the previous proof. ∎
5.3 Estimates on the horizon
We recall that and identically on by the initial data normalisation, and that all geometric quantities have vanishing modes.
Proposition 5.4.
We have for the following flux estimates on the horizon for any
(161) |
(162) |
We have on the horizon the following estimates on spheres: For
(163) |
(164) |
Proof.
This follows as in [DHR19], so we merely sketch the argument. To obtain the bounds for we write
where we have used the Bianchi equation for the first identity. For the second identity we have integrated the cross term by parts and inserted the Codazzi equation (60) on the horizon (). Note , so the expression is indeed coercive. Angular commuted identities are obtained analogously. The result for (and by Codazzi for ) now follows from the flux (and sphere) bounds available for the quantity on the left through Theorem 4.2. The result for follows from the the fact that, by (97), on the horizon
and that we control the flux on the left from Theorem 4.2 and the flux of from the first part of the proof. ∎
Corollary 5.5.
On the event horizon , we have for the following flux estimates
In addition, we have the estimates on spheres
In addition, we may add to the list of Ricci-coefficients in the above estimates.
Proof.
In view of and on the horizon, we can clearly add the expression to the list of quantities estimated in Proposition 5.4. Since the quantities vanish by assumption, the estimates follows from standard elliptic estimates. The last claim is immediate from restricting the linearised null structure equation (58) for to the horizon and using the previous bounds. ∎
5.4 Preliminary estimates on the shears
5.4.1 The outgoing shear
We give a brief overview. One starts with the quantity , which according to (57) satisfies
(165) |
Commuting twice with the operator turns the exponentially growth factor (, a blueshift) into a decay factor (, a redshift), after which the equation can be integrated forwards in using the flux bound for and derivatives thereof on the right hand side.151515The lower order terms that arise in the commutation can be integrated by parts and produce terms of a good sign and boundary terms which are controlled on the horizon from Proposition 5.4. Roughly speaking, since the structure of the horizon does not depend on the cosmological constant, the estimates near the horizon go through exactly as in [DHR19]. Away from the horizon, where is uniformly bounded away from zero, one can of course integrate directly
(166) |
all the way to infinity. This gives in particular that is uniformly bounded on . Now let us turn to the details. We first derive the key estimate near the horizon.
Proposition 5.6.
There exists an such that the following estimate holds for any and any ,
(167) |
Proof.
We provide a sketch of the proof as the argument in entirely analogous to the proof of Proposition 13.3.2 in [DHR19]. From (165) we derive upon commutation the identity
(168) |
Of course commutation with and is trivial and is omitted.
One now proceeds as in[DHR19] multiplying (168) with and integrating over the spacetime region . The terms in the first line of (168) will produce the good desired terms in (167) (as well as the term first term on the right). The term on the right hand side of (168) can be dealt with by Cauchy-Schwarz borrowing a bit from the good spacetime term on the left. Finally, for the terms on the left in the second line of (168) we proceed as in [DHR19]: For the first term we integrate by parts, controlling the (bad-signed) boundary term on the horizon by Corollary 5.5, while the resulting spacetime term has a good sign. For the second term we use Cauchy-Schwarz and a Hardy inequality, which provides control on the terms on the left hand side of (167) in terms of the higher order quantities that have already been controlled using again the control of the horizon fluxes in Corollary 5.5. ∎
Note that we can bound the first term on the right in (167) by (4.2). A simple pigeonhole principle applied to (167) yields
Proposition 5.7.
For any and , :
(169) | ||||
(170) |
We can now easily globalise the result to the entire exterior taking care of the correct -weights and also improve to an estimate on spheres.
Proposition 5.8.
For any and ,
(171) |
Moreover, both estimates also hold replacing by .
Proof.
Note that the last claim follows immediately from the two estimates and the null structure equation (57), hence we can focus on proving the two estimates. We will also suppress the (trivial) commutation with angular and -derivatives in the algebra for the proof.
We first obtain these estimates in the region . If we restricted the sum over to run from to only, both estimates follow directly from Proposition 5.7 and the fundamental theorem of calculus (which loses one derivative, hence the restriction to ). To show it for one revisits (168), now written as
(172) |
Using the estimates already shown we obtain (after trivially commuting the above with )
from which the estimates follow also for by simple ODE theory.
Having established the estimates of the proposition for , we integrate the (appropriately commuted) linearised null structure equation from towards infinity to deduce the result also for . (Observe that near infinity .) Note that and also commute trivially on the left and that near infinity. ∎
5.4.2 The ingoing shear
The boundary condition (81) now allows us to integrate (57) written as
(173) |
directly from the boundary to produce global uniform bounds on the (regular at ) quantity :
Proposition 5.9.
For any and , :
Proof.
We apply Lemma 5.1 to (173) and the -commuted (173). The only thing which is not immediate is the initial condition for the commuted estimate. For this we note that the boundary condition on translates into on using that vanish on the boundary. Indeed,
(174) |
and the last -terms vanish on the boundary by Definition 3.1. Similarly on translates into on since up to terms vanishing in the limit on we have
and hence on
from which the claim on the boundary follows.
This means that the initial condition in the commuted estimate is always controlled from Proposition 5.8 and Corollary 4.5. Furthermore, the flux (on constant ) when integrating the transport equation from the boundary requires -derivatives of to estimate derivatives of and derivatives of if one would like to see -decay. ∎
5.4.3 Improving the weights near infinity
We now establish a few improved estimates for certain combinations (and derivatives of) and , which will be helpful in establishing estimates for the torsion later. Specifically, we claim the following:
Proposition 5.10.
We have for the following estimates for any :
(175) |
and
(176) |
and
(177) |
Moreover, we can add an additional factor of on the right, provided we replace by in the energies on the right.
Proof.
To keep the notation in the proof tidy, we ignore the trivial angular commutation by during the proof, which can be trivially inserted in all equations below. We define the shear along the boundary
with the last equality following from the boundary condition. Integrating (57) from the boundary , we deduce after an integration by parts the identities (here are the components in an orthonormal frame!161616Recall the formula for the coordinate components and hence ).)
(178) |
It follows that
and similarly
We also have
from which (175) is already immediate after using Taylor’s theorem (as well as (7))
and using (the last claim of) Proposition 5.8 for the term on the right. The bound (176) follows similarly form Propositions 5.8 and 5.9 as well as another application of Taylor’s theorem, now for the -difference.
To prove (177), the key is (besides applying (175) and estimates from Theorem 4.2) to establish
(179) |
This follows from Taylor expanding
and using (7) as well as Taylor’s theorem for the remainders. Note that Proposition 5.8 controls at most three -derivatives of on the boundary, so we cannot commute further. ∎
5.5 Some immediate consequences
In this section we obtain estimates for all curvature components and the torsions , from the preliminary estimates on the shears and the gauge invariant quantities. These estimates are not optimal in terms of regularity (caused by the loss in the estimate for the shears) and will be improved later.
5.5.1 Estimating curvature one-forms
From the Bianchi identities rewritten as (94) we see that the estimates on and in Propositions 5.8 and 5.9 respectively, will provide estimates on and using Theorem 4.2 and Corollary 4.5. For now we state these (immediate) estimates on spheres (recall that the modes are trivial by assumption), deferring top order flux bounds to a later point (namely once the regularity in Propositions 5.8 and 5.9 has been improved further).
Proposition 5.11.
We have on any sphere for the estimates
(180) | ||||
(181) |
5.5.2 Estimating curvature scalars
We now recall the equation (98) noting that the left hand side of (98) can be written as a linear combination of the regular (at both and ) , and with smooth and uniformly bounded coefficients. Combining this with fact the estimate (175) we directly obtain:
Proposition 5.12.
We have on any sphere for the estimates:
(182) | ||||
(183) |
We also have the top-order ingoing flux bound
(184) |
and the top-order outgoing flux bound
(185) |
Finally, for any fixed and the sphere lying in the region we have the uniform estimate
(186) |
5.5.3 Estimates on the torsion
Proposition 5.13.
For any and and and
(187) | ||||
(188) |
We also have
(189) | ||||
(190) |
Proof.
We show the estimate for . For , (187) and (188) for follow immediately from the equation (58) and (59) after inserting the estimates of Propositions 5.8, 5.9 and Proposition 5.10. In particular this already establishes immediately all estimates claimed in the region , so we can focus on establishing the estimates for in the region for the remainder of the proof.
Replacing by (i.e. only looking at the difference) both estimates follow after taking the difference of (58) and (59) and using the estimate (176) of Proposition 5.10.
To show the actual (187) and (188) (i.e. the estimate for and individually) we integrate (61) backwards from the boundary, where and are known to vanish by having established control on at the beginning of the proof. Inserting the estimate for already established and Proposition 5.11 to control the right hand side, the estimates (187) and (188) follow.
5.5.4 Estimates on the lapses
The following is an immediate corollary of Proposition 5.13:
Corollary 5.14.
For any and
Proof.
All estimates follows straight from the definition and using Proposition 5.13. ∎
An estimate for is easily obtained from the relations
(191) |
and previous bounds on the geometric quantities:
Proposition 5.15.
For any and and
5.6 Estimates on the expansion and improving the regularity
We write the linearised Raychaudhuri equation (55) as
(192) |
Commuting with angular derivatives and also with we deduce (note the factor of ):
Proposition 5.17.
For we have on any sphere for :
(193) | ||||
(194) |
We also have
(195) | ||||
(196) |
Proof.
With the -terms absent (for ), the estimates (193), (194) claimed are an immediate consequence of applying the transport Lemmas 5.2 and 5.3 to Equation (192) using the estimates of Corollary 5.14 with the strong weight. Estimates (195), (196) are obtained similarly, by commuting (192) with and using the estimates of Corollary 5.14 and Proposition 5.15 with the strong -weights. With the -terms present (i.e. for ), estimates (193), (194), (195), (196) follow along the same lines but using the estimates with the weaker -weights in Corollary 5.14 and Proposition 5.15. By a slight variation of the transport lemmas which we leave to the reader, this generates -terms in (193) and (194) and the claimed weights in (195), (196). The -terms will be removed immediately after the next proposition, which only uses the estimates of the current proposition with the -terms. ∎
The above estimate (with the -terms) leads to an improvement Proposition 5.8 via the Codazzi equations and previous bounds:
Proposition 5.18.
For any and ,
(197) | ||||
(198) |
Proof.
For the term in the sum on the left, the estimate is a direct consequence of Proposition 5.8 so we focus on . We write the Codazzi equation (60) as
(199) |
Applying to both sides and inserting the relevant Bianchi and null structure equations, we obtain
(200) |
Therefore, with non-optimal -weights (insert a factor of in the norms on the left), the desired estimates follow immediately from (199) and (5.6) after using the estimates of Propositions 5.11, 5.12, 5.13 and 5.17. To obtain the weights near infinity as claimed one integrates (166) from some fixed as in the proof of Proposition 5.8. ∎
Corollary 5.19.
The first two estimates of Proposition 5.17 hold without the logarithmic term.
We can now also improve the estimate on of Proposition 5.9 using that we now control more derivatives of and hence (by the boundary condition which holds with arbitrary many tangential derivatives by the smoothness of the solution) of on the boundary.
Proposition 5.20.
For any and
(201) | ||||
(202) |
A direct corollary, using (60) pointwise, is
Corollary 5.21.
For any and
(203) | ||||
(204) |
Remark 5.22.
One can prove control on by subtracting the two Codazzi equations and using previous bounds but we will not need this here. See again Remark 2.7.
With Propositions 5.20 and 5.18, Proposition 5.13 also improves by one order in regularity (at the cost of less -weights) and, as a corollary of the relation (191), also our estimate on :
Proposition 5.23.
We have for the estimates
(205) | ||||
(206) |
Moreover,
(207) | ||||
(208) |
We finally obtain the top order flux bounds for and from the improved estimates on the shear of Propositions 5.20 and 5.18:
Proposition 5.24.
We have on any sphere for the ingoing flux bounds:
(209) | ||||
(210) |
We also have the top-order outgoing flux bound
(211) |
Finally, for any fixed and the sphere lying in the region we have the uniform estimate
(212) |
5.7 Estimates on the metric components
We can now integrate the propagation equation (51) for using Lemmas 5.2, 5.3 in conjunction with Proposition 5.23 to obtain a bound on the shift:
Proposition 5.25.
For any sphere and
(213) | ||||
(214) |
The propagation equation for the linearised metric in the outgoing direction (49) immediately yields after controlling the relevant flux from Proposition 5.17
Proposition 5.26.
For any sphere and
(215) | ||||
(216) |
To estimate we cannot integrate (50) directly as is not uniformly in . We instead estimate it from the Gauss curvature.
Proposition 5.27.
For any sphere and
(217) | ||||
(218) |
Proof.
By elliptic estimates, it suffices to prove these estimates replacing by and in each sum and letting the sum start at . (We slightly abuse notation here and let also act on scalars by taking , and on one forms by taking .) For the latter part we can integrate (50) commuted with because from the Codazzi equation (60) we have
and hence
(219) |
We are in the situation of Lemmas 5.2 and 5.3 (their assumptions valid from Proposition 5.12) and we hence obtain the desired estimate for the -part. For the -part we use the linearised Gauss equation:
The estimate now follows by solving this for (which has vanishing spherical average) and inserting the estimates from Propositions 5.26, 5.18 and 5.20 as well as Propositions 5.11 and 5.12. ∎
5.8 Concluding the proof of Theorem 4.7
6 Normalising the solution at infinity: Proof of Theorem 4.14
We now prove Theorem 4.14. We consider the solution of Theorem 4.7. We define a function , supported on as follows. Define for the limit , which is the (smooth) restriction to of the weighted tensor . Similarly, define for , the limit , which is the (smooth) restriction to of the weighted tensor . By the boundary condition we have for . We finally define a function by solving for each the elliptic (since ) scalar equation171717One computes .
(220) |
The function generates a pure gauge solution according to Lemma 2.8 and using the notation of that lemma we have
(221) |
Note also that from (175) and Propositions 5.18 and 5.20 holding for the solution of Theorem 4.7, we have for the quantitative estimates
(222) |
together with the corresponding estimates with the -factor on the right-hand side.
Using (220) and (221), one can prove that satisfies on . To see this we show separately that and on , which implies the claim for by standard elliptic estimates. Indeed, this follows immediately by our choice of pure gauge solution for the part. On the other hand, it is not hard to see that is actually gauge invariant and in fact equal to zero on (use the linearised Codazzi equation (60), the decay of , and the boundary condition (80) for ). The argument for is entirely analogous.
It now immediately follows that in the new gauge we can estimate and instead of and in Theorem 4.7. Indeed, we can now integrate backwards in the - and -direction from the boundary using that and hold on the boundary in the new gauge and using the estimates on and from Theorem 4.2 and Corollary 4.5 just as in the proof of Propositions 5.8 and 5.9.
Next, since on the boundary (from holding in the new gauge), integrating again backwards from the boundary one infers estimates for and .
Estimates for and are obtained directly by (58), (59), which also imply estimates for (modulo modes). Codazzi then gives control on and (modulo modes).
Inserting the above bounds, one can infer from the linearised Gauss equation (65) that the linearised Gauss curvature behaves like .
One finally adds a pure gauge solution of Lemma 2.9 so that for , one has on the initial sphere of the boundary and along the boundary. More specifically we define
It follows from the vanishing of and , , , on and the transport equations (49) that on the whole boundary , which in turn also implies by (66) that (modulo modes). From this, we can infer bounds in on by integrating their respective transport equations backwards from . Now, using the estimates on and , and , and and , one deduces that the non-vanishing pure gauge coefficients also satisfy boundedness and logarithmic decay statements. This finishes the proof of Theorem 4.14.
Appendix A Boundary regularity and boundary conditions
A.1 Proof of Proposition 2.1
We first have the following lemma. Its proof is based on ideas of [Fri95] which we adapt and strengthen in our geometric set up.
Lemma A.1.
Assume that is a solution to the Einstein equations (1) and that extends smoothly to . Let denote the Weyl tensor of . Let denote the outgoing unit normal (for the metric ) to the -hypersurfaces, define to be the induced metric by on the -hypersurfaces and define the second fundamental forms for all tangent vectors to the -hypersurfaces. Then,
-
1.
extends smoothly to ,
-
2.
extends smoothly to (in a -normalised frame),
-
3.
extends smoothly to (in a -normalised frame).
Proof.
First note that by the double null form of , we have the relations
(223) |
The general conformal transformation formula181818See https://en.wikipedia.org/wiki/List_of_formulas_in_Riemannian_geometry. reads
which, plugging in the Einstein equation (1) and using (223), rewrites as
(224) |
with
From (224), using that extends smoothly to , we already deduce that extends smoothly to . We now want to obtain the better rate for claimed in Item 1. We first note that, from Taylor’s formula, relations (223), and the fact that extends smoothly to , the function defined by
(225) |
extends smoothly to . Thus, if we can prove that extends smoothly to then Item 1 follows from (225). To control , we take the trace in (224) and use (223), and we have
(226) |
where with the induced metric by on the hypersurfaces. Now, we want to express – at first order – in terms of . Letting , we have
where we used that and that . Hence
(227) |
with denoting the traceless part of . Plugging (225) and (227) into (226), we get
(228) |
Let us now show that vanishes at first order at . Projecting formula (224) on the -hypersurfaces, using (223), and taking the traceless part, we have
Hence, using that extends smoothly to , we have that extends smoothly to . Thus, from (228), using that , and extend smoothly at , we infer that extends smoothly to . Hence, recalling formula (225) and the definition of , Item 1 is proved. Using (227) and the regularity of obtained above, we also directly infer Item 2.
From the conformal invariance of the Bianchi equations for the Weyl tensor, we have
(229) |
which, using (223), implies
(230) |
Using that extends smoothly to , extends smoothly to . Using the symmetries of the Weyl tensor – see e.g. formulas (7.3.3) in [CK93] –, all the components of can be obtained by linear combinations of and Item 3 follows. ∎
We can now prove Proposition 2.1.
Proof of Proposition 2.1.
The regularity of is a direct consequence of Lemma A.1. The regularity of follows from the coordinate components (indices up!) extending regularly. By the conformal transformation formulas, we have
(231a) | ||||
and similarly | ||||
(231b) | ||||
(231c) | ||||
(231d) |
From (231a), and the fact that extends regularly at , one infers that extends regularly to in a orthonormal frame. The corresponding regularity for follows similarly. Moreover, from formulas (231a), (231b), one has
and from the (better) regularity for of Item 2 of Lemma A.1, we obtain the (better) regularity for the difference in (30). From (231c), (231d), we have that is regular by Item 2 of Lemma A.1, hence extends smoothly to in a orthonormal frame. Moreover, we have is regular by (the good regularity of) Item 1 of Lemma A.1, thus extends smoothly to in a orthonormal frame, and combining the above, extend smoothly to in a orthonormal frame. We have
from which, by Item 1 of Lemma A.1, we infer that and extend regularly at . The regularity of the null curvature components is a direct consequence of the last item of Lemma A.1, using the conformal invariance of the Weyl tensor. This finishes the proof of the corollary. ∎
A.2 Proof of Proposition 2.2
Lemma A.2.
Assume that extends smoothly to , satisfies the Einstein equations (1), and that the induced metric by on is conformal to the Anti-de Sitter metric at infinity . Then, the -tangent tensor extends smoothly to . Note that this is equivalent to the spacetime tensor , with ⋆ denoting the Hodge dual, extending smoothly to .
Proof.
From contractions of the second Bianchi identities and the definition of the Weyl tensor, we have the following general formula
(232) |
where is called the Cotton tensor of . The Gauss-Codazzi equations on the boundary read
(233) | ||||
(234) |
By the definition of the Weyl tensor, we have
which, plugged in the Gauss-Codazzi equation (233), using (234) to replace , gives
(235) |
with . Defining the Cotton tensor of by
where is the covariant derivative of , and applying to (235), we get
(236) |
From (236) and Lemma A.1, we deduce that extends regularly to . Combining (230) and (232), we thus deduce that
(237) |
with smoothly extending to . Now, the Cotton tensor of a 3-dimensional metric is invariant under a conformal transformation and it is easy to see from its definition that it vanishes for Lorentzian cylinders . Thus, if is conformal to such a metric, one has by (237) that extends smoothly at , and the conclusion of the lemma follows. ∎
We can now prove Proposition 2.2.
Proof of Proposition 2.2.
Using the conformal invariance of the Weyl tensor, one has
From the above formulas and the result of Lemma A.2 one directly deduces (31), (32), (33). From the Bianchi equations (67) and (76), the boundary condition (31) and the fact that holds by Proposition 2.1, one further infers (34) and this finishes the proof of the proposition. ∎
Appendix B Computation of the mode
From the linear version of the Birkhoff theorem, we already know that the space of solutions supported on can only consist of the (linearised) Schwarzschild solution and pure gauge solutions. It turns out we can parametrise the space of solutions more or less explicitly. In this section all quantities are supported on so we simply write for etc. to keep the notation clean.
We first define two quantities (supported on by the above convention):
The importance of these quantities lies (partly) in their simple propagation equations (following from (70), (71), (49), (55) and (52))
(238) |
Using the formula (66) we write the linearised Gauss equation (65) for as
or more concisely as
(239) |
We first establish that if and hold on the initial data cone, the solution is necessarily trivial.
Lemma B.1.
Let be a smooth solution of the system of gravitational perturbations supported on . If and hold on , then the solution is necessarily equal to the zero solution.
Proof.
We wish to study all radial solutions of (239) to exhaust the space of solutions for . We first note that by adding a pure gauge solution, we can restrict to the case of both and being constant on .
Lemma B.2.
Proof.
Letting generate and and a pure gauge solution as in Lemma 2.8 we achieve that satisfies . In particular, the quantity is now regular at the horizon for the solution . We next add a second pure gauge solution which does not affect but achieves the second condition. For this we define by the ODE
One now checks that is indeed constant, , and that is bounded. By Lemma 2.8, a bounded will imply . ∎
Let us denote the constants and and compute now the general regular radial solutions of (239). Setting , satisfies the ODE
(240) |
which we can write as (setting )
To make the solution regular at the horizon we require
Note that with this the right hand side is integrable near infinity and near the horizon. We finally obtain191919Note that at this point we can no longer take the limit to compare with the asymptotically flat case, since we have used that goes to zero, which it does not in the asymptotically flat case. However, in (240) we can still take the limit and check that in this case and is indeed a solution, as was obtained in [DHR19].
(241) |
which satisfies
In particular, is uniformly bounded and smooth on the exterior. All non-vanishing Ricci-coefficients and curvature components can easily be computed in terms of . We find
and from the definition of the expression
(242) |
To check that is finite at the horizon we compute
Note also
so vanishes at infinity. Finally, we obtain from the null structure equations
This concludes our derivation of the solution appearing in Lemma 2.10 of the text.
References
- [BP73] J. M. Bardeen, W. H. Press, Radiation fields in the schwarzschild background, J. Math. Phys. 14 (1973), 7–19.
- [BR11] P. Bizoń, A. Rostworowski, Weakly Turbulent Instability of Anti–de Sitter Spacetime, Phys. Rev. Lett. 107 (2011), no. 3, 031102.
- [Chr09] D. Christodoulou, The formation of black holes in general relativity, EMS Monographs in Mathematics (2009), x+589 pp.
- [CK93] D. Christodoulou, S. Klainerman, The global nonlinear stability of the Minkowski space, Princeton Univ. Press (1993), x+483 pp.
- [DHR19] M. Dafermos, G. Holzegel, I. Rodnianski, The linear stability of the Schwarzschild solution to gravitational perturbations, Acta Math. 222 (2019), no. 1, 1–214.
- [DHRT21] M. Dafermos, G. Holzegel, I. Rodnianski, M. Taylor, The non-linear stability of the Schwarzschild family of black holes, arXiv:2104.08222 (2021), 519 pp.
- [Dol17] D. Dold, Unstable mode solutions to the Klein–Gordon equation in Kerr-anti-de Sitter spacetimes, Comm. Math. Phys. 350 (2017), no. 2, 639–697.
- [DR08] M. Dafermos, I. Rodnianski, Lectures on black holes and linear waves, Clay Math. Proc. 17 (2008).
- [EK19] A. Enciso, N. Kamran, Lorentzian Einstein metrics with prescribed conformal infinity, J. Differential Geom. 112 (2019), no. 3.
- [Fri95] H. Friedrich, Einstein equations and conformal structure: Existence of anti-de Sitter-type space-times, Journal of Geometry and Physics 17 (1995), no. 2, 125–184.
- [GH23] O. Graf, G. Holzegel, Mode stability results for the Teukolsky equations on Kerr–anti-de Sitter spacetimes, Class. Quantum Grav. 40 (2023), no. 4, 045003.
- [GH24] O. Graf, G. Holzegel, Linear Stability of Schwarzschild-Anti-de Sitter spacetimes II: Logarithmic decay of solutions to the Teukolsky system, arXiv preprint (2024).
- [GKS22] E. Giorgi, S. Klainerman, J. Szeftel, Wave equations estimates and the nonlinear stability of slowly rotating Kerr black holes, arXiv:2205.14808v1 (2022), no. arXiv:2205.14808, 912 pp.
- [HE08] S. W. Hawking, G. F. R. Ellis, The Large Scale Structure of Space-Time, Cambridge Monographs on Mathematical Physics, Cambridge Univ. Press, Cambridge, 21. printing edition (2008).
- [HLSW20] G. Holzegel, J. Luk, J. Smulevici, C. Warnick, Asymptotic properties of linear field equations in anti-de Sitter space, Comm. Math. Phys. 374 (2020), no. 2, 1125–1178.
- [HS13] G. Holzegel, J. Smulevici, Decay properties of Klein-Gordon fields on Kerr-adS spacetimes, Comm. Pure Appl. Math. 66 (2013), no. 11, 1751–1802.
- [HS14] G. Holzegel, J. Smulevici, Quasimodes and a lower bound on the uniform energy decay rate for Kerr–AdS spacetimes, Anal. PDE 7 (2014), no. 5, 1057–1090.
- [HS16] G. Holzegel, A. Shao, Unique Continuation from Infinity in Asymptotically Anti-de Sitter Spacetimes, Commun. Math. Phys. 347 (2016), no. 3, 723–775.
- [HV16] P. Hintz, A. Vasy, Global Analysis of Quasilinear Wave Equations on Asymptotically Kerr-de Sitter Spaces, Int Math Res Notices 2016 (2016), no. 17, 5355–5426.
- [Kha83] U. Khanal, Rotating black hole in asymptotic de Sitter space: Perturbation of the space-time with spin fields, Phys. Rev. D 28 (1983), no. 6, 1291–1297.
- [KM] C. Kehle, G. Moschidis, in preparation .
- [Mos18] G. Moschidis, A proof of the instability of AdS for the Einstein–massless Vlasov system, arXiv:1812.04268 (2018), 132 pp.
- [Mos20] G. Moschidis, A proof of the instability of AdS for the Einstein-null dust system with an inner mirror, Analysis & PDE 13 (2020), no. 6, 1671–1754.
- [RW57] T. Regge, J. A. Wheeler, Stability of a Schwarzschild Singularity, Phys. Rev. 108 (1957), no. 4, 1063–1069.
- [Teu72] S. A. Teukolsky, Rotating Black Holes: Separable Wave Equations for Gravitational and Electromagnetic Perturbations, Phys. Rev. Lett. 29 (1972), no. 16, 1114–1118.