This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Linear Stability of Schwarzschild-Anti-de Sitter spacetimes I:
The system of gravitational perturbations

Olivier Graf [email protected] Univ. Grenoble Alpes, CNRS, IF, 38000 Grenoble, France   Gustav Holzegel [email protected] Universität Münster, Mathematisches Institut, Einsteinstrasse 62 48149 Münster, Bundesrepublik Deutschland   Imperial College London, Department of Mathematics, South Kensington Campus, London SW7 2AZ, United Kingdom
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 Ric(g)=Λg\textrm{Ric}(g)=\Lambda g with Λ<0\Lambda<0 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 Λ\Lambda,

Ric(g)=Λg,\displaystyle\mathrm{Ric}(g)=\Lambda g\,, (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. Λ=0\Lambda=0 in (1), and Kerr-de Sitter black holes, i.e. Λ>0\Lambda>0 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. Λ<0\Lambda<0 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 Λ<0\Lambda<0, 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 Λ<0\Lambda<0.

Turning to the global dynamics in the Λ<0\Lambda<0 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 gψ=0\Box_{g}\psi=0 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 gAdSψ=0\Box_{g_{AdS}}\psi=0, 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 gKAdSψ+μψ=0\Box_{g_{KAdS}\psi}+\mu\psi=0 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 Ric=3k2gRic=-3k^{2}g 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 Λ=0\Lambda=0 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 aa, 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 (Λ=0\Lambda=0) 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 α(1),α¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}. 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 Λ=0\Lambda=0 these observations go back to Bardeen–Press [BP73] and Teukolsky [Teu72] in the physics literature and are easily generalised to Λ0\Lambda\neq 0, 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 χ^(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}} from the estimates for α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}. This relies on the (commuted) redshift effect for χ^(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}. The quantity χ^¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}} then inherits this decay through the boundary conditions and we can hence integrate χ^¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}} in the ingoing direction from the boundary (using the estimates for α¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}) to establish boundedness and decay for χ^¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}.444This second step is not possible in the asymptotically flat case as there is no boundary and the χ^¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}} equation cannot be integrated directly from data in the ingoing direction either because of unfavourable rr-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 𝒬U,V2\mathcal{Q}\subset\mathbb{R}^{2}_{U,V} be the 22-dimensional submanifold with (piecewise smooth) boundary defined by

𝒬:=([1,0]U×[1,)V){VU1}.\displaystyle\mathcal{Q}:=\left(\left[-1,0\right]_{U}\times\left[1,\infty\right)_{V}\right)\cap\{-VU\leq 1\}\,. (2)

We define the associated 44-dimensional manifold

:=𝒬×S2,\displaystyle\mathcal{M}:=\mathcal{Q}\times S^{2}\,, (3)

equipped with coordinates (U,V,θ1,θ2)(U,V,\theta^{1},\theta^{2}), which we will refer to as Kruskal coordinates on \mathcal{M}. We denote the boundary components

:={VU=1},+={U=0},Ndata={V=1},\displaystyle\mathcal{I}:=\mathcal{M}\cap\{-VU=1\}\ \ \ ,\ \ \ \mathcal{H}^{+}=\mathcal{M}\cap\{U=0\}\ \ \ ,\ \ \ N_{data}=\mathcal{M}\cap\{V=1\}\,, (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 ×𝕊2\mathbb{R}\times\mathbb{S}^{2}. We denote by SU,V2S^{2}_{U,V} the 22-spheres (U,V)×S2(U,V)\times S^{2} in \mathcal{M}.

Given a fixed parameter M>0M>0 we can define coordinates (u,v,θ1,θ2)(u,v,\theta^{1},\theta^{2}) on +\mathcal{M}\setminus\mathcal{H}^{+} by

u=2Mlog(U),v=2MlogV.\displaystyle u=-2M\log(-U)\ \ \ ,\ \ \ v=2M\log V\,. (5)

Note v0v\geq 0 on +\mathcal{M}\setminus\mathcal{H}^{+} and v=uv=u on \mathcal{I}. Defining also r(u,v)=vur^{\star}(u,v)=v-u and t(u,v)=v+ut(u,v)=v+u we have another coordinate system (t,r,θ1,θ2)(t,r^{\star},\theta^{1},\theta^{2}) on +\mathcal{M}\setminus\mathcal{H}^{+}. We observe that rr^{\star} is a boundary defining function for \mathcal{I} in that it vanishes at \mathcal{I} and ur=1\partial_{u}r^{\star}=-1, vr=1\partial_{v}r^{\star}=1. One may parametrise the boundary =[0,)t×St,t2\mathcal{I}=[0,\infty)_{t}\times S^{2}_{t,t} by the coordinate tt. Finally, we can define on +\mathcal{M}\setminus\mathcal{H}^{+} a function r(u,v)=r(r=vu)r(u,v)=r(r^{\star}=v-u) by the relation

drdr=12Mr+k2r2,r(0)=,\displaystyle\frac{dr}{dr^{\star}}=1-\frac{2M}{r}+k^{2}r^{2}\ \ \ ,\ \ \ r(0)=\infty\,, (6)

where 3k2=Λ-3k^{2}=\Lambda. Note that the function rr depends on MM and that we have the asymptotics

|(vu)r|=|rr|=1k2+O(r2).\displaystyle|(v-u)r|=|r^{\star}r|=\frac{1}{k^{2}}+O(r^{-2}). (7)

2.2 One-parameter families of aAdS metrics in double null gauge

Let us denote

int=\displaystyle\mathcal{M}_{int}=\mathcal{M}\setminus\mathcal{I} (8)

and fix M>0M>0, which we may think of as the mass of a Schwarzschild-AdS background metric we are about to install on int\mathcal{M}_{int}. Given int\mathcal{M}_{int}, equipped with local coordinates (u,v,θ1,θ2)(u,v,{\theta}^{1},{\theta}^{2}) on int+\mathcal{M}_{int}\setminus\mathcal{H}^{+}, we consider a 11-parameter family of metrics 𝒈(ϵ)\boldsymbol{g}(\epsilon) expressed in double null gauge555Note that the 𝒃\boldsymbol{b} is on dudu here instead of on dvdv 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 𝒃\boldsymbol{b}, equation (25), which is now in the outgoing direction and with an additional minus on the right. An outgoing transport equation for 𝒃\boldsymbol{b} is desirable as it can be integrated from data, where 𝒃\boldsymbol{b} is normalised. Note also that it is 𝛀2𝒃\boldsymbol{\Omega}^{-2}\boldsymbol{b} which extends regularly to the horizon +\mathcal{H}^{+} as can be seen by transforming (9) to the regular Kruskal coordinates.

𝒈(ϵ)=4𝛀2(ϵ)dudv+AB(ϵ)(dθA𝒃A(ϵ)du)(dθB𝒃B(ϵ)du)\displaystyle\boldsymbol{g}\left(\epsilon\right)=-4\boldsymbol{\Omega}^{2}\left(\epsilon\right)d{u}d{v}+\not{\boldsymbol{g}}_{AB}\left(\epsilon\right)\left(d\theta^{A}-\boldsymbol{b}^{A}\left(\epsilon\right)d{u}\right)\left(d\theta^{B}-\boldsymbol{b}^{B}\left(\epsilon\right)d{u}\right)\, (9)

such that

  1. 1.

    The 𝒈(ϵ)\boldsymbol{g}\left(\epsilon\right) satisfy on int+\mathcal{M}_{int}\setminus\mathcal{H}^{+} the Einstein equations with negative cosmological constant Λ=3k2\Lambda=-3k^{2},

    Ric(𝒈)=3k2𝒈.\displaystyle\mathrm{Ric}(\boldsymbol{g})=-3k^{2}\boldsymbol{g}\,. (10)
  2. 2.

    We have that 𝒈(0)\boldsymbol{g}\left(0\right) is the Schwarzschild-AdS metric of mass MM and cosmological constant Λ=3k2\Lambda=-3k^{2}

    𝒈(0):=4(12Mr(u,v)+k2r2(u,v))dudv+r2(u,v)γABdθAdθB,\displaystyle\boldsymbol{g}\left(0\right):=-4\left(1-\frac{2M}{r(u,v)}+k^{2}\cdot r^{2}({u},{v})\right)d{u}d{v}+r^{2}({u},{v}){\gamma}_{AB}d{\theta}^{A}d{\theta}^{B}\,, (11)

    where r(u,v)r(u,v) is defined as in (6) and γ{\gamma} denotes the round metric on the unit sphere.

  3. 3.

    The 𝒈(ϵ)\boldsymbol{g}\left(\epsilon\right) are asymptotically Anti-de Sitter in that the function (vu)2𝛀2(v-u)^{2}\boldsymbol{\Omega}^{2} as well as the conformally rescaled metric 𝛀2𝒈\boldsymbol{\Omega}^{-2}\boldsymbol{g} extend regularly to \mathcal{I}, i.e. in particular they can be defined on the larger manifold +\mathcal{M}\setminus\mathcal{H}^{+}. Specifically, recalling t=u+vt=u+v, the family

    𝒈(3)(ϵ)=dt2+AB(ϵ)𝛀2(ϵ)(dθA𝒃A(ϵ)dt2)(dθB𝒃B(ϵ)dt2)\displaystyle\boldsymbol{g}^{(3)}_{\mathcal{I}}(\epsilon)=-d{t}^{2}+\frac{\not{\boldsymbol{g}}_{AB}(\epsilon)}{\boldsymbol{\Omega}^{2}(\epsilon)}\left(d\theta^{A}-\boldsymbol{b}^{A}(\epsilon)\frac{d{t}}{2}\right)\left(d\theta^{B}-\boldsymbol{b}^{B}(\epsilon)\frac{d{t}}{2}\right)\, (12)

    defines a smooth family of 33-dimensional Lorentzian metrics on \mathcal{I}.

  4. 4.

    The 𝒈(3)(ϵ)\boldsymbol{g}^{(3)}_{\mathcal{I}}(\epsilon) are all conformal to the Lorentzian cylinder (×S2,k2dt2+γ)(\mathbb{R}\times S^{2},-k^{2}\mathrm{d}t^{2}+\gamma), the latter being the metric induced on \mathcal{I} by the conformally rescaled Schwarzschild-AdS metric 𝛀2𝒈(0)\boldsymbol{\Omega}^{-2}\boldsymbol{g}(0). In particular, the 𝒈(3)(ϵ)\boldsymbol{g}^{(3)}_{\mathcal{I}}(\epsilon) are all locally conformally flat.

  5. 5.

    In regular Kruskal coordinates, the 𝒈(ϵ)\boldsymbol{g}\left(\epsilon\right) and regular derivatives thereof extend smoothly to the boundary +\mathcal{H}^{+} of int\mathcal{M}_{int} in the sense of Section 5.1.1 of [DHR19]. In particular, arbitrary concatenations of frame vectors from the set {𝛀2(ϵ)(u+𝒃AA),v,θ1,θ2}\{\boldsymbol{\Omega}^{-2}(\epsilon)(\partial_{u}+\boldsymbol{b}^{A}\partial_{A}),\partial_{v},\partial_{\theta^{1}},\partial_{\theta^{2}}\} applied to 𝒈(ϵ)\boldsymbol{g}\left(\epsilon\right) extend regularly to +\mathcal{H}^{+}.

The existence of such families is of course an implicit assumption. Locally in time, i.e. for a finite vv-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 int+\mathcal{M}_{int}\setminus\mathcal{H}^{+} is a double null frame consisting of the null vectorfields

e3=1𝛀(u+𝒃AA),e4=1𝛀v,e_{3}=\frac{1}{\boldsymbol{\Omega}}\left(\partial_{u}+\boldsymbol{b}^{A}\partial_{A}\right)\ \ ,\ \ e_{4}=\frac{1}{\boldsymbol{\Omega}}\partial_{v}\,,

satisfying 𝐠(e3,e4)=2{\bf g}(e_{3},e_{4})=-2, 𝐠(e3,e3)=𝐠(e4,e4)=0{\bf g}(e_{3},e_{3})={\bf g}(e_{4},e_{4})=0, which is complemented with a local coordinate frame on Su,v2S^{2}_{u,v}, eA=Ae_{A}=\partial_{A} for A{1,2}A\in\{1,2\}, satisfying 𝐠(eA,eB)=AB{\bf g}(e_{A},e_{B})=\not{\boldsymbol{g}}_{AB}. Note 𝐠(e3,eA)=0{\bf g}(e_{3},e_{A})=0 and 𝐠(e4,eA){\bf g}(e_{4},e_{A}) for A{1,2}A\in\{1,2\}.

2.3.1 Su,v2S^{2}_{u,v}-tensor algebra

Let 𝝃,𝝃~\boldsymbol{\xi},\boldsymbol{\tilde{\xi}} be arbitrary Su,v2S^{2}_{u,v} one-forms and 𝜽,𝜽~\boldsymbol{\theta},\boldsymbol{\tilde{\theta}} be arbitrary symmetric Su,v2S^{2}_{u,v} 22-tensors.

We denote by 𝝃{}^{\star}\boldsymbol{\xi} and 𝜽{}^{\star}\boldsymbol{\theta} the Hodge-dual on (Su,v2,)\left(S^{2}_{u,v},\boldsymbol{\not{g}}\right) of 𝝃\boldsymbol{\xi} and 𝜽\boldsymbol{\theta}, respectively, and denote by 𝜽\boldsymbol{\theta}^{\sharp} the tensor obtained from 𝜽\boldsymbol{\theta} by raising an index with \not{\boldsymbol{g}}. We define the contractions

(𝝃,𝝃~):=AB𝝃A𝝃~B and (𝜽,𝜽~):=ABCD𝜽AC𝜽~BD,\left(\boldsymbol{\xi},\boldsymbol{\tilde{\xi}}\right):=\boldsymbol{\not{g}}^{AB}\boldsymbol{\xi}_{A}\boldsymbol{\tilde{\xi}}_{B}\textrm{ \ \ \ and \ \ \ }\left(\boldsymbol{\theta},\boldsymbol{\tilde{\theta}}\right):=\boldsymbol{\not{g}}^{AB}\boldsymbol{\not{g}}^{CD}\boldsymbol{\theta}_{AC}\boldsymbol{\tilde{\theta}}_{BD},

and denote by 𝜽𝝃\boldsymbol{\theta}^{\sharp}\cdot\boldsymbol{\xi} the one-form 𝜽AB𝝃B\boldsymbol{\theta}_{A}^{\phantom{A}B}\boldsymbol{\xi}_{B} arising from the contraction with \not{\boldsymbol{g}}. We finally define

(𝜽×𝜽~)BC\displaystyle\left(\boldsymbol{\theta}\times\boldsymbol{\tilde{\theta}}\right)_{BC} :=AD𝜽AB𝜽~DC,\displaystyle:=\not{\boldsymbol{g}}^{AD}\boldsymbol{\theta}_{AB}\boldsymbol{\tilde{\theta}}_{DC}\,,
(𝝃^𝝃~)AB\displaystyle\left(\boldsymbol{\xi}\widehat{\otimes}\boldsymbol{\tilde{\xi}}\right)_{AB} :=𝝃A𝝃~B+𝝃B𝝃~AAB𝝃A𝝃~B,\displaystyle:=\boldsymbol{\xi}_{A}\boldsymbol{\tilde{\xi}}_{B}+\boldsymbol{\xi}_{B}\boldsymbol{\tilde{\xi}}_{A}-\not{\boldsymbol{g}}^{AB}\boldsymbol{\xi}_{A}\boldsymbol{\tilde{\xi}}_{B}\,,
𝜽𝜽~\displaystyle\boldsymbol{\theta}\wedge\boldsymbol{\tilde{\theta}} :=ϵ̸ABCD𝜽AC𝜽~BD,\displaystyle:=\boldsymbol{\not{\epsilon}}^{AB}\not{\boldsymbol{g}}^{CD}\boldsymbol{\theta}_{AC}\boldsymbol{\tilde{\theta}}_{BD}\,,

where ϵ̸AB\boldsymbol{\not{\epsilon}}_{AB} denotes the components of the volume form associated with \not{\boldsymbol{g}} on Su,v2S^{2}_{u,v}.

2.3.2 Su,v2S^{2}_{u,v}-projected Lie and covariant derivates

We define the derivative operators 𝑫¯\underline{\boldsymbol{D}} and 𝑫{\boldsymbol{D}} to act on an Su,v2S^{2}_{u,v}-tensor ϕ\boldsymbol{\phi} as the projection onto Su,v2S^{2}_{u,v} of the Lie-derivative of ϕ\boldsymbol{\phi} in the direction of 𝛀𝒆3\boldsymbol{\Omega}{\boldsymbol{e}}_{3} and 𝛀𝒆4\boldsymbol{\Omega}{\boldsymbol{e}}_{4} respectively. We hence have the following relations between the projected Lie-derivatives 𝑫¯\underline{\boldsymbol{D}} and 𝑫{\boldsymbol{D}} and the Su,v2S^{2}_{u,v}-projected spacetime covariant derivatives ∇̸𝟑=∇̸𝒆3,∇̸𝟒=∇̸𝒆4\not{\nabla}_{\boldsymbol{3}}=\not{\nabla}_{{\boldsymbol{e}}_{3}},\not{\nabla}_{\boldsymbol{4}}=\not{\nabla}_{{\boldsymbol{e}}_{4}} in the direction 𝒆3{\boldsymbol{e}}_{3} and 𝒆4{\boldsymbol{e}}_{4} respectively:

𝑫𝒇=𝛀∇̸𝟒𝒇 on functions 𝒇,𝑫𝝃=𝛀∇̸𝟒𝝃+𝛀𝝌𝝃 on one-forms 𝝃,𝑫𝜽=𝛀∇̸𝟒𝜽+𝛀𝝌×𝜽+𝛀𝜽×𝝌 on symmetric 2-tensors 𝜽,\begin{split}\boldsymbol{D}\boldsymbol{f}&=\boldsymbol{\Omega}{\not{\nabla}}_{\boldsymbol{4}}\boldsymbol{f}\textrm{ \ \ \ on functions $\boldsymbol{f}$,}\\ \boldsymbol{D}\boldsymbol{\xi}&=\boldsymbol{\Omega}{\not{\nabla}}_{\boldsymbol{4}}\boldsymbol{\xi}+\boldsymbol{\Omega}\boldsymbol{\chi}^{\sharp}\cdot\boldsymbol{\xi}\textrm{ \ \ \ on one-forms $\boldsymbol{\xi}$,}\\ \boldsymbol{D}\boldsymbol{\theta}&=\boldsymbol{\Omega}{\not{\nabla}}_{\boldsymbol{4}}\boldsymbol{\theta}+\boldsymbol{\Omega}\boldsymbol{\chi}\times\boldsymbol{\theta}+\boldsymbol{\Omega}\boldsymbol{\theta}\times\boldsymbol{\chi}\textrm{ \ \ \ on symmetric $2$-tensors $\boldsymbol{\theta}$,}\end{split} (13)

and similarly for ∇̸𝟑{\not{\nabla}}_{\boldsymbol{3}} replacing 𝝌\boldsymbol{\chi} by 𝝌¯\underline{\boldsymbol{\chi}} and 𝑫{\boldsymbol{D}} by 𝑫¯\underline{\boldsymbol{D}}.

2.3.3 Angular operators on Su,v2S^{2}_{u,v}

Let 𝝃\boldsymbol{\xi} be an arbitrary one-form and 𝜽\boldsymbol{\theta} an arbitrary symmetric traceless 22-tensor on Su,v2S^{2}_{u,v}.

  • ∇̸\boldsymbol{\not{\nabla}} denotes the covariant derivative associated with the metric AB\not{\boldsymbol{g}}_{AB} on Su,v2S^{2}_{u,v}.

  • 1\boldsymbol{\not{\mathcal{D}}}_{1} takes 𝝃\boldsymbol{\xi} into the pair of functions (𝐝𝐢𝐯𝝃,𝐜𝐮𝐫𝐥𝝃)\left(\not{\bf{div}}\boldsymbol{\xi},\not{\bf{curl}}\boldsymbol{\xi}\right) where 𝐝𝐢𝐯𝝃=AB∇̸A𝝃B\not{\bf{div}}\boldsymbol{\xi}=\not{\boldsymbol{g}}^{AB}\boldsymbol{\not{\nabla}}_{A}\boldsymbol{\xi}_{B} and 𝐜𝐮𝐫𝐥𝝃=ϵ̸AB∇̸A𝝃B\not{\bf{curl}}\boldsymbol{\xi}=\boldsymbol{\not{\epsilon}}^{AB}\boldsymbol{\not{\nabla}}_{A}\boldsymbol{\xi}_{B}.

  • 𝒟̸1\not{\mathbfcal{D}}_{1}^{\star}, the L2L^{2}-adjoint of 1\boldsymbol{\not{\mathcal{D}}}_{1}, takes any pair of scalars 𝝆,𝝈\boldsymbol{\rho},\boldsymbol{\sigma} into the Su,v2S^{2}_{u,v}-one-form ∇̸A𝝆+ϵ̸AB∇̸B𝝈-\boldsymbol{\not{\nabla}}_{A}\boldsymbol{\rho}+\boldsymbol{\not{\epsilon}}_{AB}\boldsymbol{\not{\nabla}}^{B}\boldsymbol{\sigma}.

  • 𝒟̸2\not{\mathbfcal{D}}_{2} takes 𝜽\boldsymbol{\theta} into the Su,v2S^{2}_{u,v}-one-form (𝐝𝐢𝐯𝜽)C=AB∇̸A𝜽BC\left(\not{\bf{div}}\boldsymbol{\theta}\right)_{C}=\not{\boldsymbol{g}}^{AB}\boldsymbol{\not{\nabla}}_{A}\boldsymbol{\theta}_{BC}.

  • 𝒟̸2\not{\mathbfcal{D}}_{2}^{\star}, the L2L^{2} adjoint of 𝒟̸2\not{\mathbfcal{D}}_{2}, takes 𝝃\boldsymbol{\xi} into 22-tensor (𝒟̸2𝝃)AB=12(∇̸B𝝃A+∇̸A𝝃B(𝐝𝐢𝐯𝝃)AB)\left(\not{\mathbfcal{D}}^{\star}_{2}\boldsymbol{\xi}\right)_{AB}=-\frac{1}{2}\left(\boldsymbol{\not{\nabla}}_{B}\boldsymbol{\xi}_{A}+\boldsymbol{\not{\nabla}}_{A}\boldsymbol{\xi}_{B}-\left(\not{\bf{div}}\boldsymbol{\xi}\right)\not{\boldsymbol{g}}_{AB}\right).

2.3.4 Ricci-coefficients and curvature components

We define the non-vanishing null-decomposed Ricci coefficients as follows:

𝝌AB=𝒈(A𝒆4,𝒆B) , 𝝌¯AB=𝒈(A𝒆3,𝒆B),𝜼A=12𝒈(𝒆3𝒆A,𝒆4) , 𝜼¯A=12𝒈(𝒆4𝒆A,𝒆3),𝜻=12𝒈(A𝒆4,𝒆3),𝝎=12𝛀𝒈(𝒆4𝒆3,𝒆4) , 𝝎¯=12𝛀𝒈(𝒆3𝒆4,𝒆3).\begin{split}\boldsymbol{\chi}_{AB}&=\boldsymbol{g}\left(\boldsymbol{\nabla}_{A}{\boldsymbol{e}}_{4},{\boldsymbol{e}}_{B}\right)\textrm{ \ \ , \ \ \ \ }\underline{\boldsymbol{\chi}}_{AB}=\boldsymbol{g}\left(\boldsymbol{\nabla}_{A}{\boldsymbol{e}}_{3},{\boldsymbol{e}}_{B}\right)\,,\\ \boldsymbol{\eta}_{A}&=-\frac{1}{2}\boldsymbol{g}\left(\boldsymbol{\nabla}_{{\boldsymbol{e}}_{3}}{\boldsymbol{e}}_{A},{\boldsymbol{e}}_{4}\right)\textrm{ \ \ , \ \ }\underline{\boldsymbol{\eta}}_{A}=-\frac{1}{2}\boldsymbol{g}\left(\boldsymbol{\nabla}_{{\boldsymbol{e}}_{4}}{\boldsymbol{e}}_{A},{\boldsymbol{e}}_{3}\right)\ \ ,\ \ \ \boldsymbol{\zeta}=\frac{1}{2}\boldsymbol{g}\left(\boldsymbol{\nabla}_{A}{\boldsymbol{e}}_{4},{\boldsymbol{e}}_{3}\right)\,,\\ \boldsymbol{\omega}&=\frac{1}{2}\boldsymbol{\Omega}\boldsymbol{g}\left(\boldsymbol{\nabla}_{{\boldsymbol{e}}_{4}}{\boldsymbol{e}}_{3},{\boldsymbol{e}}_{4}\right)\textrm{ \ \ \ \ , \ \ \ \ \ }\underline{\boldsymbol{\omega}}=\frac{1}{2}\boldsymbol{\Omega}\boldsymbol{g}\left(\boldsymbol{\nabla}_{{\boldsymbol{e}}_{3}}{\boldsymbol{e}}_{4},{\boldsymbol{e}}_{3}\right)\,.\end{split} (14)

The above objects are Su,v2S^{2}_{u,v} 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 𝝌{\boldsymbol{\chi}} into its \not{\boldsymbol{g}}-tracefree part 𝝌^\boldsymbol{\hat{\chi}} (a symmetric traceless Su,v2S^{2}_{u,v} 2-tensor) and its trace 𝒕𝒓𝝌\boldsymbol{tr}\boldsymbol{\chi}, and similarly for 𝝌¯\underline{\boldsymbol{\chi}}. Note also the relations

𝝎=v𝛀𝛀,𝝎¯=(𝒖+𝒃A𝜽A)𝛀𝛀,𝜼A=𝜻A+∇̸Alog𝛀 , 𝜼¯A=𝜻A+∇̸Alog𝛀.\displaystyle{\boldsymbol{\omega}}=\frac{\boldsymbol{\partial}_{v}{\boldsymbol{\Omega}}}{\boldsymbol{\Omega}}\ \ ,\ \ \underline{{\boldsymbol{\omega}}}=\frac{(\boldsymbol{\partial}_{\boldsymbol{u}}+\boldsymbol{b}^{A}\boldsymbol{\partial}_{\boldsymbol{\theta}^{A}}){\boldsymbol{\Omega}}}{\boldsymbol{\Omega}}\ \ ,\ \ \boldsymbol{\eta}_{A}=\boldsymbol{\zeta}_{A}+\boldsymbol{\not{\nabla}}_{A}\log\boldsymbol{\Omega}\textrm{ \ \ , \ \ }\underline{\boldsymbol{\eta}}_{A}=-\boldsymbol{\zeta}_{A}+\boldsymbol{\not{\nabla}}_{A}\log\boldsymbol{\Omega}. (15)

With 𝑾\boldsymbol{W} denoting the Weyl curvature tensor of the metric (9), the null-decomposed Weyl curvature components are defined as follows:

𝜶AB=𝑾(𝒆A,𝒆4,𝒆B,𝒆4) , 𝜶¯AB=𝑾(𝒆A,𝒆3,𝒆B,𝒆3),𝜷A=12𝑾(𝒆A,𝒆4,𝒆3,𝒆4) , 𝜷¯A=12𝑾(𝒆A,𝒆3,𝒆3,𝒆4),𝝆=14𝑾(𝒆4,𝒆3,𝒆4,𝒆3) , 𝝈=14𝑾(𝒆4,𝒆3,𝒆4,𝒆3),\begin{split}\boldsymbol{\alpha}_{AB}&=\boldsymbol{W}\left({\boldsymbol{e}}_{A},{\boldsymbol{e}}_{4},{\boldsymbol{e}}_{B},{\boldsymbol{e}}_{4}\right)\textrm{ \ \ \ , \ \ }\underline{\boldsymbol{\alpha}}_{AB}=\boldsymbol{W}\left({\boldsymbol{e}}_{A},{\boldsymbol{e}}_{3},{\boldsymbol{e}}_{B},{\boldsymbol{e}}_{3}\right)\,,\\ \boldsymbol{\beta}_{A}&=\frac{1}{2}\boldsymbol{W}\left({\boldsymbol{e}}_{A},{\boldsymbol{e}}_{4},{\boldsymbol{e}}_{3},{\boldsymbol{e}}_{4}\right)\textrm{ \ \ \ , \ \ \ }\underline{\boldsymbol{\beta}}_{A}=\frac{1}{2}\boldsymbol{W}\left({\boldsymbol{e}}_{A},{\boldsymbol{e}}_{3},{\boldsymbol{e}}_{3},{\boldsymbol{e}}_{4}\right)\,,\ \\ \boldsymbol{\rho}&=\frac{1}{4}\boldsymbol{W}\left({\boldsymbol{e}}_{4},{\boldsymbol{e}}_{3},{\boldsymbol{e}}_{4},{\boldsymbol{e}}_{3}\right)\textrm{ \ \ \ , \ \ \ \ \ \ }\boldsymbol{\sigma}=\frac{1}{4}{}^{\star}\boldsymbol{W}\left({\boldsymbol{e}}_{4},{\boldsymbol{e}}_{3},{\boldsymbol{e}}_{4},{\boldsymbol{e}}_{3}\right)\,,\end{split} (16)

with 𝑾{}^{\star}\boldsymbol{W} denoting the Hodge dual on (,𝒈)({\mathcal{M},\boldsymbol{g}}) of 𝑾\boldsymbol{W}. Again the above objects are Su,v2S^{2}_{u,v}-tensors (functions, vectors, symmetric 22-tensors) on (,𝒈)(\mathcal{M},\boldsymbol{g}).

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 Su,v2S^{2}_{u,v}-tensors that we collect below. See again [DHR19] and in particular [Chr09] for a detailed derivation of the equations in the Λ=0\Lambda=0 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:

𝑫=2𝛀𝝌=2𝛀𝝌^+𝛀𝐭𝐫𝝌 and 𝑫¯=2𝛀𝝌¯=2𝛀𝝌¯^+𝛀𝐭𝐫𝝌.\displaystyle\boldsymbol{D}\not{\boldsymbol{g}}=2\boldsymbol{\Omega}\boldsymbol{\chi}=2\boldsymbol{\Omega}\hat{\boldsymbol{\chi}}+\boldsymbol{\Omega}{\bf tr}\boldsymbol{\chi}\not{g}\textrm{ \ \ \ \ \ and \ \ \ \ \ }\underline{\boldsymbol{D}}\not{\boldsymbol{g}}=2\boldsymbol{\Omega}\underline{\boldsymbol{\chi}}=2\boldsymbol{\Omega}\underline{\hat{\boldsymbol{\chi}}}+\boldsymbol{\Omega}{\bf tr}{\boldsymbol{\chi}}\not{\boldsymbol{g}}\,. (17)

The transport equations for the second fundamental forms take the form

∇̸3𝝌¯^+𝐭𝐫𝝌¯𝝌¯^𝛀1𝝎¯𝝌¯^=𝜶¯,∇̸4𝝌^+𝐭𝐫𝝌𝝌^𝛀1𝝎𝝌^=𝜶,\displaystyle\boldsymbol{\not{\nabla}}_{3}\underline{\hat{\boldsymbol{\chi}}}+{\bf tr}\underline{\boldsymbol{\chi}}\ \hat{\underline{\boldsymbol{\chi}}}-\boldsymbol{\Omega}^{-1}{\underline{\boldsymbol{\omega}}}\ \underline{\hat{\boldsymbol{\chi}}}=-\underline{\boldsymbol{\alpha}}\ \ \ ,\ \ \ \boldsymbol{\not{\nabla}}_{4}{\hat{\boldsymbol{\chi}}}+{\bf tr}{\boldsymbol{\chi}}\ \hat{\boldsymbol{\chi}}-\boldsymbol{\Omega}^{-1}{{\boldsymbol{\omega}}}\ {\hat{\boldsymbol{\chi}}}=-{\boldsymbol{\alpha}}, (18)
∇̸3(𝐭𝐫𝝌¯)+12(𝐭𝐫𝝌¯)2𝛀1𝝎¯𝐭𝐫𝝌¯=(𝝌¯^,𝝌¯^),∇̸4(𝐭𝐫𝝌)+12(𝐭𝐫𝝌)2𝛀1𝝎𝐭𝐫𝝌=(𝝌^,𝝌^),\displaystyle\boldsymbol{\not{\nabla}}_{3}\left({\bf tr}\underline{\boldsymbol{\chi}}\right)+\frac{1}{2}\left({\bf tr}\underline{\boldsymbol{\chi}}\right)^{2}-\boldsymbol{\Omega}^{-1}\underline{{\boldsymbol{\omega}}}{\bf tr}\underline{{\boldsymbol{\chi}}}=-\left(\underline{\hat{\boldsymbol{\chi}}},\underline{\hat{\boldsymbol{\chi}}}\right)\ \ ,\ \ \boldsymbol{\not{\nabla}}_{4}\left({\bf tr}\boldsymbol{\chi}\right)+\frac{1}{2}\left({\bf tr}\boldsymbol{\chi}\right)^{2}-\boldsymbol{\Omega}^{-1}{\boldsymbol{\omega}}{\bf tr}{\boldsymbol{\chi}}=-\left({\hat{\boldsymbol{\chi}}},{\hat{\boldsymbol{\chi}}}\right), (19)
∇̸3𝝌^+12𝐭𝐫𝝌¯𝝌^+𝛀1𝝎¯𝝌^=2𝒟̸2𝜼12𝐭𝐫𝝌𝝌¯^+(𝜼^𝜼),\displaystyle\boldsymbol{\not{\nabla}}_{3}{\hat{\boldsymbol{\chi}}}+\frac{1}{2}{\bf tr}\underline{\boldsymbol{\chi}}\ \hat{{\boldsymbol{\chi}}}+\boldsymbol{\Omega}^{-1}{\underline{\boldsymbol{\omega}}}\ {\hat{\boldsymbol{\chi}}}=-2\not{\mathbfcal{D}}_{2}^{\star}\boldsymbol{\eta}-\frac{1}{2}{\bf tr}\boldsymbol{\chi}\ \underline{\hat{\boldsymbol{\chi}}}+\left({\boldsymbol{\eta}}\widehat{\otimes}{\boldsymbol{\eta}}\right)\,, (20)
∇̸4𝝌¯^+12𝐭𝐫𝝌𝝌¯^+𝛀1𝝎𝝌¯^=2𝒟̸2𝜼¯12𝐭𝐫𝝌¯𝝌^+(𝜼¯^𝜼¯),\displaystyle\boldsymbol{\not{\nabla}}_{4}\underline{\hat{\boldsymbol{\chi}}}+\frac{1}{2}{\bf tr}{\boldsymbol{\chi}}\ \hat{\underline{\boldsymbol{\chi}}}+\boldsymbol{\Omega}^{-1}{{\boldsymbol{\omega}}}\ \underline{\hat{\boldsymbol{\chi}}}=-2\not{\mathbfcal{D}}_{2}^{\star}\underline{\boldsymbol{\eta}}-\frac{1}{2}{\bf tr}\underline{\boldsymbol{\chi}}\ \hat{\boldsymbol{\chi}}+\left(\underline{\boldsymbol{\eta}}\widehat{\otimes}\underline{\boldsymbol{\eta}}\right)\,, (21)
∇̸3(𝐭𝐫𝝌)+12(𝐭𝐫𝝌¯)(𝐭𝐫𝝌)+𝛀1𝝎¯𝐭𝐫𝝌=(𝝌¯^,𝝌^)+2(𝜼,𝜼)+2𝝆+2𝐝𝐢𝐯𝜼4k2,\displaystyle\boldsymbol{\not{\nabla}}_{3}\left({\bf tr}{\boldsymbol{\chi}}\right)+\frac{1}{2}\left({\bf tr}\underline{\boldsymbol{\chi}}\right)\left({\bf tr}{\boldsymbol{\chi}}\right)+\boldsymbol{\Omega}^{-1}\underline{{\boldsymbol{\omega}}}{\bf tr}{{\boldsymbol{\chi}}}=-\left(\underline{\hat{\boldsymbol{\chi}}},{\hat{\boldsymbol{\chi}}}\right)+2\left(\boldsymbol{\eta},\boldsymbol{\eta}\right)+2\boldsymbol{\rho}+2\not{\bf{div}}\boldsymbol{\eta}\boxed{-4k^{2}}\,, (22)
∇̸4(𝐭𝐫𝝌¯)+12(𝐭𝐫𝝌)(𝐭𝐫𝝌¯)+𝛀1𝝎𝐭𝐫𝝌¯=(𝝌¯^,𝝌^)+2(𝜼¯,𝜼¯)+2𝝆+2𝐝𝐢𝐯𝜼¯4k2.\displaystyle\boldsymbol{\not{\nabla}}_{4}\left({\bf tr}\underline{\boldsymbol{\chi}}\right)+\frac{1}{2}\left({\bf tr}{\boldsymbol{\chi}}\right)\left({\bf tr}\underline{\boldsymbol{\chi}}\right)+\boldsymbol{\Omega}^{-1}{{\boldsymbol{\omega}}}{\bf tr}\underline{{\boldsymbol{\chi}}}=-\left(\underline{\hat{\boldsymbol{\chi}}},{\hat{\boldsymbol{\chi}}}\right)+2\left(\underline{\boldsymbol{\eta}},\underline{\boldsymbol{\eta}}\right)+2\boldsymbol{\rho}+2\not{\bf{div}}\underline{\boldsymbol{\eta}}\boxed{-4k^{2}}\,. (23)

The transport equations for the torsions and the (derivative of the) lapse become:

∇̸3𝜼¯=𝝌¯(𝜼𝜼¯)+𝜷¯,∇̸4𝜼=𝝌(𝜼𝜼¯)𝜷,\displaystyle\boldsymbol{\not{\nabla}}_{3}\underline{\boldsymbol{\eta}}=\underline{\boldsymbol{\chi}}^{\sharp}\cdot\left(\boldsymbol{\eta}-\underline{\boldsymbol{\eta}}\right)+\underline{\boldsymbol{\beta}}\ \ \ ,\ \ \ \boldsymbol{\not{\nabla}}_{4}{\boldsymbol{\eta}}=-{\boldsymbol{\chi}}^{\sharp}\cdot\left(\boldsymbol{\eta}-\underline{\boldsymbol{\eta}}\right)-{\boldsymbol{\beta}}\,, (24)
𝑫(𝝎¯)=𝛀2[2(𝜼,𝜼¯)|𝜼|2𝝆k2] , 𝑫¯(𝝎)=𝛀2[2(𝜼,𝜼¯)|𝜼¯|2𝝆k2],\displaystyle\boldsymbol{D}\left({\underline{\boldsymbol{\omega}}}\right)=\boldsymbol{\Omega}^{2}\left[2\left(\boldsymbol{\eta},\underline{\boldsymbol{\eta}}\right)-|\boldsymbol{\eta}|^{2}-\boldsymbol{\rho}\boxed{-k^{2}}\right]\textrm{ \ , \ }\underline{\boldsymbol{D}}\left({\boldsymbol{\omega}}\right)=\boldsymbol{\Omega}^{2}\left[2\left(\boldsymbol{\eta},\underline{\boldsymbol{\eta}}\right)-|\underline{\boldsymbol{\eta}}|^{2}-\boldsymbol{\rho}\boxed{-k^{2}}\right]\,,
v𝒃A=2𝛀2(𝜼A𝜼¯A).\displaystyle\boldsymbol{\partial}_{v}\boldsymbol{b}^{A}=-2\boldsymbol{\Omega}^{2}\left(\boldsymbol{\eta}^{A}-\underline{\boldsymbol{\eta}}^{A}\right)\,. (25)

Finally, we have the elliptic relations on spheres:

𝐜𝐮𝐫𝐥𝜼=12𝝌𝝌¯+𝝈 and 𝐜𝐮𝐫𝐥𝜼¯=+12𝝌𝝌¯𝝈,\displaystyle\not{\bf{curl}}\boldsymbol{\eta}=-\frac{1}{2}\boldsymbol{\chi}\wedge\underline{\boldsymbol{\chi}}+\boldsymbol{\sigma}\textrm{ \ \ \ \ \ and \ \ \ \ \ }\not{\bf{curl}}\underline{\boldsymbol{\eta}}=+\frac{1}{2}\boldsymbol{\chi}\wedge\underline{\boldsymbol{\chi}}-\boldsymbol{\sigma}\,, (26)
𝐝𝐢𝐯𝝌^\displaystyle\not{\bf{div}}\hat{\boldsymbol{\chi}} =12𝝌^(𝜼𝜼¯)+14𝐭𝐫𝝌(𝜼𝜼¯)+12∇̸𝐭𝐫𝝌𝜷\displaystyle=-\frac{1}{2}\hat{\boldsymbol{\chi}}^{\sharp}\cdot\left(\boldsymbol{\eta}-\underline{\boldsymbol{\eta}}\right)+\frac{1}{4}{\bf tr}\boldsymbol{\chi}\left(\boldsymbol{\eta}-\underline{\boldsymbol{\eta}}\right)+\frac{1}{2}\boldsymbol{\not{\nabla}}{\bf tr}{\boldsymbol{\chi}}-\boldsymbol{\beta}
=12𝝌^(𝜼𝜼¯)12𝐭𝐫𝝌𝜼¯+12𝛀∇̸(𝛀𝐭𝐫𝝌)𝜷,\displaystyle=-\frac{1}{2}\hat{\boldsymbol{\chi}}^{\sharp}\cdot\left(\boldsymbol{\eta}-\underline{\boldsymbol{\eta}}\right)-\frac{1}{2}{\bf tr}\boldsymbol{\chi}\underline{\boldsymbol{\eta}}+\frac{1}{2\boldsymbol{\Omega}}\boldsymbol{\not{\nabla}}\left({\bf\Omega}{\bf tr}{\boldsymbol{\chi}}\right)-\boldsymbol{\beta}\,, (27)
𝐝𝐢𝐯𝝌¯^\displaystyle\not{\bf{div}}\underline{\hat{\boldsymbol{\chi}}} =12𝝌¯^(𝜼𝜼¯)14𝐭𝐫𝝌¯(𝜼𝜼¯)+12∇̸𝐭𝐫𝝌¯+𝜷¯\displaystyle=\frac{1}{2}\underline{\hat{\boldsymbol{\chi}}}^{\sharp}\cdot\left(\boldsymbol{\eta}-\underline{\boldsymbol{\eta}}\right)-\frac{1}{4}{\bf tr}\underline{\boldsymbol{\chi}}\left(\boldsymbol{\eta}-\underline{\boldsymbol{\eta}}\right)+\frac{1}{2}\boldsymbol{\not{\nabla}}{\bf tr}\underline{\boldsymbol{\chi}}+\underline{\boldsymbol{\beta}}
=12𝝌¯^(𝜼𝜼¯)12𝐭𝐫𝝌¯𝜼+12𝛀∇̸(𝛀𝐭𝐫𝝌¯)+𝜷¯,\displaystyle=\frac{1}{2}\underline{\hat{\boldsymbol{\chi}}}^{\sharp}\cdot\left(\boldsymbol{\eta}-\underline{\boldsymbol{\eta}}\right)-\frac{1}{2}{\bf tr}\underline{\boldsymbol{\chi}}{\boldsymbol{\eta}}+\frac{1}{2\boldsymbol{\Omega}}\boldsymbol{\not{\nabla}}\left({\bf\Omega}{\bf tr}\underline{\boldsymbol{\chi}}\right)+\underline{\boldsymbol{\beta}}\,, (28)
𝐊=14𝐭𝐫𝝌𝐭𝐫𝝌¯+12(𝝌^,𝝌¯^)𝝆k2.\displaystyle\mathbf{K}=-\frac{1}{4}{\bf tr}\boldsymbol{\chi}{\bf tr}\underline{\boldsymbol{\chi}}+\frac{1}{2}\left(\hat{\boldsymbol{\chi}},\hat{\underline{\boldsymbol{\chi}}}\right)-\boldsymbol{\rho}\boxed{-k^{2}}\,. (29)

Equations (2.3.5)–(2.3.5) are known as the Codazzi equations, (29) is the Gauss equation on Su,v2S^{2}_{u,v}. We finally collect the Bianchi equations for the null Weyl curvature components which are formally unchanged in the presence of a cosmological constant.

∇̸3𝜶+12𝐭𝐫𝝌¯𝜶+2𝛀1𝝎¯𝜶\displaystyle\boldsymbol{\not{\nabla}}_{3}\boldsymbol{\alpha}+\frac{1}{2}{\bf tr}\underline{\boldsymbol{\chi}}\boldsymbol{\alpha}+2\boldsymbol{\Omega}^{-1}\underline{{\boldsymbol{\omega}}}\boldsymbol{\alpha} =2𝒟̸𝟐𝜷3𝝌^𝝆3𝝌^𝝈+(4𝜼+𝜻)^𝜷,\displaystyle=-2{\bf\not{\mathbfcal{D}}_{2}^{\star}}\boldsymbol{\beta}-3\hat{\boldsymbol{\chi}}\boldsymbol{\rho}-3{}^{\star}\hat{\boldsymbol{\chi}}\boldsymbol{\sigma}+\left(4\boldsymbol{\eta}+\boldsymbol{\zeta}\right)\hat{\otimes}\boldsymbol{\beta}\,,
∇̸4𝜷+2𝐭𝐫𝝌𝜷𝛀1𝝎𝜷\displaystyle\boldsymbol{\not{\nabla}}_{4}\boldsymbol{\beta}+2{\bf tr}\boldsymbol{\chi}\boldsymbol{\beta}-\boldsymbol{\Omega}^{-1}{\boldsymbol{\omega}}\boldsymbol{\beta} =𝐝𝐢𝐯𝜶+(𝜼¯+2𝜻)𝜶,\displaystyle=\not{\bf{div}}\boldsymbol{\alpha}+\left(\underline{\boldsymbol{\eta}}^{\sharp}+2\boldsymbol{\zeta}^{\sharp}\right)\cdot\boldsymbol{\alpha}\,,
∇̸3𝜷+𝐭𝐫𝝌¯+𝛀1𝝎¯𝜷\displaystyle\boldsymbol{\not{\nabla}}_{3}\boldsymbol{\beta}+{\bf tr}\underline{\boldsymbol{\chi}}+\boldsymbol{\Omega}^{-1}\underline{{\boldsymbol{\omega}}}\boldsymbol{\beta} =𝒟̸1(𝝆,𝝈)+3𝜼𝝆+3𝜼𝝈+2𝝌^𝜷¯,\displaystyle=\not{\mathbfcal{D}}_{1}^{\star}\left(-\boldsymbol{\rho},\boldsymbol{\sigma}\right)+3\boldsymbol{\eta}\boldsymbol{\rho}+3{}^{\star}\boldsymbol{\eta}\boldsymbol{\sigma}+2\hat{\boldsymbol{\chi}}^{\sharp}\cdot\underline{\boldsymbol{\beta}}\,,
∇̸4𝝆+32𝐭𝐫𝝌𝝆\displaystyle\boldsymbol{\not{\nabla}}_{4}\boldsymbol{\rho}+\frac{3}{2}{\bf tr}\boldsymbol{\chi}\boldsymbol{\rho} =𝐝𝐢𝐯𝜷+(2𝜼¯+𝜻,𝜷)12(𝝌¯^,𝜶),\displaystyle=\not{\bf{div}}\boldsymbol{\beta}+\left(2\underline{\boldsymbol{\eta}}+\boldsymbol{\zeta},\boldsymbol{\beta}\right)-\frac{1}{2}\left(\underline{\hat{\boldsymbol{\chi}}},\boldsymbol{\alpha}\right)\,,
∇̸4𝝈+32𝐭𝐫𝝌𝝈\displaystyle\boldsymbol{\not{\nabla}}_{4}\boldsymbol{\sigma}+\frac{3}{2}{\bf tr}\boldsymbol{\chi}\boldsymbol{\sigma} =𝐜𝐮𝐫𝐥𝜷(2𝜼¯+𝜻)𝜷+12𝝌¯^𝜶,\displaystyle=-\not{\bf{curl}}\boldsymbol{\beta}-\left(2\underline{\boldsymbol{\eta}}+\boldsymbol{\zeta}\right)\wedge\boldsymbol{\beta}+\frac{1}{2}\underline{\hat{\boldsymbol{\chi}}}\wedge\boldsymbol{\alpha}\,,
∇̸3𝝆+32𝐭𝐫𝝌¯𝝆\displaystyle\boldsymbol{\not{\nabla}}_{3}\boldsymbol{\rho}+\frac{3}{2}{\bf tr}\underline{\boldsymbol{\chi}}\boldsymbol{\rho} =𝐝𝐢𝐯𝜷¯(2𝜼𝜻,𝜷¯)12(𝝌^,𝜶¯),\displaystyle=-\not{\bf{div}}\underline{\boldsymbol{\beta}}-\left(2\boldsymbol{\eta}-\boldsymbol{\zeta},\underline{\boldsymbol{\beta}}\right)-\frac{1}{2}\left(\hat{\boldsymbol{\chi}},\underline{\boldsymbol{\alpha}}\right)\,,
∇̸3𝝈+32𝐭𝐫𝝌¯𝝈\displaystyle\boldsymbol{\not{\nabla}}_{3}\boldsymbol{\sigma}+\frac{3}{2}{\bf tr}\underline{\boldsymbol{\chi}}\boldsymbol{\sigma} =𝐜𝐮𝐫𝐥𝜷¯(2𝜼𝜻)𝜷¯12𝝌^𝜶¯,\displaystyle=-\not{\bf{curl}}\underline{\boldsymbol{\beta}}-\left(2\boldsymbol{\eta}-\boldsymbol{\zeta}\right)\wedge\underline{\boldsymbol{\beta}}-\frac{1}{2}\hat{\boldsymbol{\chi}}\wedge\underline{\boldsymbol{\alpha}}\,,
∇̸4𝜷¯+𝐭𝐫𝝌𝜷¯+𝛀1𝝎𝜷¯\displaystyle\boldsymbol{\not{\nabla}}_{4}\underline{\boldsymbol{\beta}}+{\bf tr}\boldsymbol{\chi}\underline{\boldsymbol{\beta}}+\boldsymbol{\Omega}^{-1}{\boldsymbol{\omega}}\underline{\boldsymbol{\beta}} =𝒟̸1(𝝆,𝝈)3𝜼¯𝝆+3𝜼¯𝝈+2𝝌¯^𝜷,\displaystyle=\not{\mathbfcal{D}}_{1}^{\star}\left(\boldsymbol{\rho},\boldsymbol{\sigma}\right)-3\underline{\boldsymbol{\eta}}\boldsymbol{\rho}+3{}^{\star}\underline{\boldsymbol{\eta}}\boldsymbol{\sigma}+2\underline{\hat{\boldsymbol{\chi}}}^{\sharp}\cdot\boldsymbol{\beta}\,,
∇̸3𝜷¯+2𝐭𝐫𝝌¯𝜷¯𝛀1𝝎¯𝜷¯\displaystyle\boldsymbol{\not{\nabla}}_{3}\underline{\boldsymbol{\beta}}+2{\bf tr}\underline{\boldsymbol{\chi}}\underline{\boldsymbol{\beta}}-\boldsymbol{\Omega}^{-1}{\underline{\boldsymbol{\omega}}}\underline{\boldsymbol{\beta}} =𝐝𝐢𝐯𝜶¯(𝜼2𝜻)𝜶¯,\displaystyle=-\not{\bf{div}}\underline{\boldsymbol{\alpha}}-\left(\boldsymbol{\eta}^{\sharp}-2\boldsymbol{\zeta}^{\sharp}\right)\cdot\underline{\boldsymbol{\alpha}}\,,
∇̸4𝜶¯+12𝐭𝐫𝝌𝜶¯+2𝛀1𝝎𝜶¯\displaystyle\boldsymbol{\not{\nabla}}_{4}\underline{\boldsymbol{\alpha}}+\frac{1}{2}{\bf tr}\boldsymbol{\chi}\underline{\boldsymbol{\alpha}}+2\boldsymbol{\Omega}^{-1}{\boldsymbol{\omega}}\underline{\boldsymbol{\alpha}} =2𝒟̸2𝜷¯3𝝌¯^𝝆+3𝝌¯^𝝈(4𝜼¯𝜻)^𝜷¯.\displaystyle=2\not{\mathbfcal{D}}_{2}^{\star}\underline{\boldsymbol{\beta}}-3\underline{\hat{\boldsymbol{\chi}}}\boldsymbol{\rho}+3{}^{\star}\underline{\hat{\boldsymbol{\chi}}}\boldsymbol{\sigma}-\left(4\underline{\boldsymbol{\eta}}-\boldsymbol{\zeta}\right)\hat{\otimes}\underline{\boldsymbol{\beta}}\,.

2.4 Boundary regularity and boundary conditions

We collect the asymptotic behaviour towards the conformal boundary for the geometric quantities on int\mathcal{M}_{int} from the assumption that (by 3.3. of Section 2.2) the metric (uv)2𝒈(u-v)^{2}{\boldsymbol{g}} extends regularly to the larger manifold \mathcal{M}. The proof of Proposition 2.1 is postponed to Appendix A.

Proposition 2.1.

Assume that 𝐠\boldsymbol{g} is asymptotically Anti-de Sitter (3.3. of Section 2.2). The following geometric quantities associated with 𝐠\boldsymbol{g} defined in Section 2.2 extend regularly to the conformal boundary \mathcal{I} in the sense that the components with respect to a 𝐠\boldsymbol{g}-orthonormal frame extend smoothly:

r2(k2(uv)2𝛀21),r1𝒃,𝛀𝐭𝐫𝝌2uv,𝛀𝐭𝐫𝝌¯+2uv,r𝝌^,r𝝌¯^,r2𝜼,r2𝜼¯,r(𝝎1uv),r(𝝎¯+1uv),r(𝛀𝐭𝐫𝝌𝛀𝐭𝐫𝝌¯4uv),r2(𝝌^𝝌¯^),r3(𝜼+𝜼¯),r2(𝝎+𝝎¯),r3𝜶,r3𝜶¯,r3𝜷,r3𝜷¯,r3𝝆,r3𝝈.\displaystyle\begin{aligned} r^{2}\left(k^{2}(u-v)^{2}\boldsymbol{\Omega}^{2}-1\right),r^{-1}\boldsymbol{b},\\ \boldsymbol{\Omega}{\bf tr}{\boldsymbol{\chi}}-\frac{2}{u-v},\boldsymbol{\Omega}{\bf tr}\underline{\boldsymbol{\chi}}+\frac{2}{u-v},r\hat{\boldsymbol{\chi}},r\hat{\underline{\boldsymbol{\chi}}},r^{2}\boldsymbol{\eta},r^{2}\underline{\boldsymbol{\eta}},r\Big{(}\boldsymbol{\omega}-\frac{1}{u-v}\Big{)},r\Big{(}\boldsymbol{\underline{\omega}}+\frac{1}{u-v}\Big{)},\\ r\Big{(}\boldsymbol{\Omega}{\bf tr}{\boldsymbol{\chi}}-\boldsymbol{\Omega}{\bf tr}\underline{\boldsymbol{\chi}}-\frac{4}{u-v}\Big{)},r^{2}\left(\hat{\boldsymbol{\chi}}-\hat{\underline{\boldsymbol{\chi}}}\right),r^{3}\left(\boldsymbol{\eta}+\boldsymbol{\underline{\eta}}\right),r^{2}\left(\boldsymbol{\omega}+\boldsymbol{\underline{\omega}}\right),\\ r^{3}\boldsymbol{\alpha},r^{3}\underline{\boldsymbol{\alpha}},r^{3}\boldsymbol{\beta},r^{3}\underline{\boldsymbol{\beta}},r^{3}\boldsymbol{\rho},r^{3}\boldsymbol{\sigma}\,.\end{aligned} (30)

From the assumption that the metric is conformal to the Anti-de Sitter metric on the boundary at infinity (4.4. of Section 2.2), one has the following boundary conditions for the null curvature components. The proof is postponed to Appendix A.

Proposition 2.2.

Assume that 𝐠\boldsymbol{g} is asymptotically Anti-de Sitter (3.3. of Section 2.2) and its conformally induced metric on \mathcal{I} is conformal to the conformally induced metric of Anti-de Sitter (4.4. of Section 2.2). Then, the following boundary conditions hold

limvu(r)3𝜶\displaystyle\lim_{v\rightarrow u}(r^{\star})^{-3}\boldsymbol{\alpha} =limvu(r)3𝜶¯,\displaystyle=\lim_{v\rightarrow u}(r^{\star})^{-3}\underline{\boldsymbol{\alpha}}\,, (31)
limvu(r)3𝜷\displaystyle\lim_{v\rightarrow u}(r^{\star})^{-3}\boldsymbol{\beta} =limvu(r)3𝜷¯,\displaystyle=-\lim_{v\rightarrow u}(r^{\star})^{-3}\underline{\boldsymbol{\beta}}\,, (32)
limvu(r)3𝝈\displaystyle\lim_{v\rightarrow u}(r^{\star})^{-3}\boldsymbol{\sigma} =0,\displaystyle=0,\, (33)

as well as

limvu[𝛀∇̸4𝛀∇̸3]((r)3𝜶)\displaystyle\lim_{v\rightarrow u}[\boldsymbol{\Omega}\boldsymbol{\not{\nabla}}_{4}-\boldsymbol{\Omega}\boldsymbol{\not{\nabla}}_{3}]((r^{\star})^{-3}\boldsymbol{\alpha}) =limvu[𝛀∇̸4𝛀∇̸3]((r)3𝜶¯).\displaystyle=-\lim_{v\rightarrow u}[\boldsymbol{\Omega}\boldsymbol{\not{\nabla}}_{4}-\boldsymbol{\Omega}\boldsymbol{\not{\nabla}}_{3}]((r^{\star})^{-3}\underline{\boldsymbol{\alpha}})\,. (34)

2.5 The Schwarzschild-AdS background

In this section we discuss the Schwarzschild-AdS manifold (int,g=𝒈(0))(\mathcal{M}_{int},g=\boldsymbol{g}(0)). In complete analogy with [DHR19] we use unbolded notation to indicate ϵ=0\epsilon=0 quantities, for instance we write Ω\Omega, \not{g} and bb for the metric components, χ^\hat{\chi} for the outgoing shear etc.

Moreover, all the constructions and definitions of Sections 2.3.12.3.4 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 ϵ=0\epsilon=0 case are:

ΩχAB=Ωχ¯AB=Ω2rr2γAB,ω=ω¯=Mr2+k2r,\displaystyle\Omega\chi_{AB}=-\Omega\underline{\chi}_{AB}=\frac{\Omega^{2}}{r}r^{2}\gamma_{AB}\ \ \ ,\ \ \ \omega=-\underline{\omega}=\frac{M}{r^{2}}+k^{2}\,r\,, (35)

where Ω2=12Mr+k2r2\Omega^{2}=1-\frac{2M}{r}+k^{2}r^{2} and rr is defined implicitly as in (6). In particular, Ωχ\Omega\chi and Ωχ¯\Omega\underline{\chi} are \not{g}-traceless with Ωtrχ=Ωtrχ¯=2Ω2r\Omega tr\chi=-\Omega tr\underline{\chi}=\frac{2\Omega^{2}}{r}. The only non-vanishing null-curvature component is ρ=2Mr3\rho=-\frac{2M}{r^{3}}.

2.5.2 Differential operators and commutation formulae

We have the simplified coordinate formulae for the the projected Lie-derivatives for a general Su,v2S^{2}_{u,v}-tensor ξ\xi of rank NN:

(Dξ)A1AN=v(ξA1AN)and(D¯ξ)A1AN=u(ξA1AN)(D\xi)_{A_{1}...A_{N}}=\partial_{v}(\xi_{A_{1}...A_{N}})\ \ \textrm{and}\ \ (\underline{D}\xi)_{A_{1}...A_{N}}=\partial_{u}(\xi_{A_{1}...A_{N}})

For the projected covariant derivatives one finds

Ω(∇̸3ξ)A1AN=u(ξA1AN)N2Ωtrχ¯(ξA1AN)andΩ(∇̸4ξ)A1AN=v(ξA1AN)N2Ωtrχ(ξA1AN).\Omega(\not{\nabla}_{3}\xi)_{A_{1}...A_{N}}=\partial_{u}(\xi_{A_{1}...A_{N}})-\frac{N}{2}\Omega tr\underline{\chi}(\xi_{A_{1}...A_{N}})\ \ \textrm{and}\ \ \Omega(\not{\nabla}_{4}\xi)_{A_{1}...A_{N}}=\partial_{v}(\xi_{A_{1}...A_{N}})-\frac{N}{2}\Omega tr{\chi}(\xi_{A_{1}...A_{N}}).

We recall the (unbolded) Su,v2S^{2}_{u,v}-angular operators ∇̸\not{\nabla},1\not{\mathcal{D}}_{1}, 2\not{\mathcal{D}}_{2}, 1\not{\mathcal{D}}_{1}^{\star}, 2\not{\mathcal{D}}_{2}^{\star}, now all defined with respect to the metric =r2γ\not{g}=r^{2}\gamma on Su,v2S^{2}_{u,v}.

We define the vectorfield

T=12u+12v,T=\frac{1}{2}\partial_{u}+\frac{1}{2}\partial_{v}\,,

which is the static Killing field of Schwarzschild-AdS. The shall employ the notation 2∇̸Tξ=Ω∇̸3ξ+Ω∇̸4ξ2\not{\nabla}_{T}\xi=\Omega\not{\nabla}_{3}\xi+\Omega\not{\nabla}_{4}\xi. Note that (∇̸Tξ)A1AN=t(ξA1An)(\not{\nabla}_{T}\xi)_{A_{1}\ldots A_{N}}=\partial_{t}(\xi_{A_{1}\ldots A_{n}}) since Ωtrχ+Ωtrχ¯=0\Omega tr\chi+\Omega tr\underline{\chi}=0 for the background Schwarzschild-AdS metric.

We finally collect the commutation formulae holding on Su,v2S^{2}_{u,v} tensors ξ\xi:

[Ω∇̸3,r∇̸A]ξ=0,[Ω∇̸4,r∇̸A]ξ=0,[Ω∇̸3,Ω∇̸4]ξ=0,\displaystyle[\Omega\not{\nabla}_{3},r\not{\nabla}_{A}]\xi=0\ \ \ ,\ \ \ [\Omega\not{\nabla}_{4},r\not{\nabla}_{A}]\xi=0\ \ \ ,\ \ \ [\Omega\not{\nabla}_{3},\Omega\not{\nabla}_{4}]\xi=0\,, (36)

which will be used frequently. As in [DHR19] we will define the angular operators 𝒜[i]\mathcal{A}^{[i]} (which commute trivially with ∇̸3\not{\nabla}_{3}, ∇̸4\not{\nabla}_{4}) acting on symmetric traceless tensors as follows:

𝒜[0]=1,and then inductively𝒜[2i+1]=r2𝒜[2i],𝒜[2i]=r222𝒜[2i2].\displaystyle\mathcal{A}^{[0]}=1\ \ ,\textrm{and then inductively}\ \ \mathcal{A}^{[2i+1]}=r\not{\mathcal{D}}_{2}\mathcal{A}^{[2i]}\ \ ,\ \ \mathcal{A}^{[2i]}=r^{2}\not{\mathcal{D}}^{\star}_{2}\not{\mathcal{D}}_{2}\mathcal{A}^{[2i-2]}\,. (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 𝒜[i]\mathcal{A}^{[i]} to act on one-forms as 𝒜[i]ξ=𝒜[i1]r2ξ\mathcal{A}^{[i]}\xi=\mathcal{A}^{[i-1]}r\not{\mathcal{D}}^{\star}_{2}\xi.

2.5.3 Norms on the spheres Su,v2S^{2}_{u,v}

Let θ\theta and ϕ\phi denote the standard spherical coordinates on Su,v2S^{2}_{u,v}. We define the pointwise norm on Su,v2S^{2}_{u,v} tensors ξ\xi of rank NN by

|ξ|2=A1ANB1,,BNξA1ANξB1BN.|\xi|^{2}=\not{g}^{A_{1}...A_{N}}\not{g}^{B_{1},...,B_{N}}\xi_{A_{1}...A_{N}}\xi_{B_{1}...B_{N}}\,.

A weighted L2(Su,v2)L^{2}(S^{2}_{u,v}) norm on such tensors is then defined by

ξu,v2:=Su,v2|ξ|2sinθdθdϕ.\displaystyle\|\xi\|^{2}_{u,v}:=\int_{S^{2}_{u,v}}|\xi|^{2}\sin\theta d\theta d\phi\,. (38)

Note the absence of a factor of r2(u,v)r^{2}(u,v) in the integral, which if present would make (38) the induced norm.

2.5.4 The =0\ell=0 and =1\ell=1 modes

We recall the spherical harmonics YmY^{\ell}_{m} on the round sphere, where 0\ell\in\mathbb{N}^{0} and m{,}m\in\{-\ell,...\ell\} which form a basis of L2(S2)L^{2}(S^{2}). The =0\ell=0 and =1\ell=1 spherical harmonic will play a distinguished role in our problem and are given explicitly by

Y00=14π,Y01=34πcosθ,Y11=34πsinθcosϕ,Y11=34πsinθsinϕ.Y^{0}_{0}=\frac{1}{\sqrt{4\pi}}\ \ \ ,\ \ \ Y^{1}_{0}=\sqrt{\frac{3}{4\pi}}\cos\theta\ \ \ ,\ \ \ Y^{1}_{1}=\sqrt{\frac{3}{4\pi}}\sin\theta\cos\phi\ \ \ ,\ \ \ Y^{1}_{-1}=\sqrt{\frac{3}{4\pi}}\sin\theta\sin\phi\,.

A function is supported for 1\ell\geq 1 is a function whose spherical means vanishes. A function supported for 2\ell\geq 2 is a function supported for 1\ell\geq 1 whose projection to the Ym1Y^{1}_{m} also vanishes.

We can also make sense of Su,v2S^{2}_{u,v} one-forms and symmetric traceless tensors being supported on specific \ell modes as in Section 4.2.2 of [DHR19]. To summarise this, recall that any Su,v2S^{2}_{u,v} one-form ξ\xi can be written uniquely as

ξ=r1(f,g)\displaystyle\xi=r\not{\mathcal{D}}_{1}^{\star}(f,g) (39)

for functions ff and gg of vanishing mean. (Note that only constants are in the kernel of 1\not{\mathcal{D}}_{1}^{\star}). We say that ξ\xi is supported for 2\ell\geq 2 if both ff and gg in the above representation are supported on 2\ell\geq 2. Furthermore, we define the projection of ξ\xi to =1\ell=1 by the expression (39) where ff and gg are projected to =1\ell=1.

Similarly, an Su,v2S^{2}_{u,v} symmetric traceless tensor ξ\xi can be represented uniquely by functions ff and gg supported on 2\ell\geq 2 as

ξ=r221(f,g).\xi=r^{2}\not{\mathcal{D}}_{2}^{\star}\not{\mathcal{D}}_{1}^{\star}(f,g)\,.

Note in particular that the kernel of the operator r2r\not{\mathcal{D}}_{2}^{\star} consists precisely of functions supported for 1\ell\geq 1 only (as shown explicitly in [DHR19]). It is in the above sense that we can say that one-forms are supported for 1\ell\geq 1 and symmetric traceless tensors for 2\ell\geq 2.

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 (f,g)(f,g) be a pair of functions supported on 2\ell\geq 2, then for any j0j\geq 0

i=0j+2[r∇̸]i(f,g)u,v2𝒜[j]r221(f,g)u,v2.\displaystyle\sum_{i=0}^{j+2}\|[r\not{\nabla}]^{i}(f,g)\|_{u,v}^{2}\lesssim\|\mathcal{A}^{[j]}r^{2}\not{\mathcal{D}}_{2}^{\star}\not{\mathcal{D}}_{1}^{\star}(f,g)\|^{2}_{u,v}\,. (40)

Let η\eta be an Su,v2S^{2}_{u,v} one-form supported on 2\ell\geq 2, then for any j0j\geq 0

i=0j+1[r∇̸]iηu,v2𝒜[j]r2ηu,v2.\displaystyle\sum_{i=0}^{j+1}\|[r\not{\nabla}]^{i}\eta\|_{u,v}^{2}\lesssim\|\mathcal{A}^{[j]}r\not{\mathcal{D}}_{2}^{\star}\eta\|^{2}_{u,v}\,. (41)

Let ξ\xi be an Su,v2S^{2}_{u,v} symmetric traceless tensor, then for any j0j\geq 0

i=0j[r∇̸]jξu,v2𝒜[j]ξu,v2.\displaystyle\sum_{i=0}^{j}\|[r\not{\nabla}]^{j}\xi\|_{u,v}^{2}\lesssim\|\mathcal{A}^{[j]}\xi\|^{2}_{u,v}\,. (42)

2.6 The linearisation procedure

Recall that in our set up of Section 2.2, the members of the 11-parameter family of metrics (9) all live on the same underlying manifold \mathcal{M}. Moreover, the hypersurfaces u=const.u=\textrm{const.} and v=const.v=\textrm{const.} are null for any metric in the family. In other words, the notion of Su,v2S^{2}_{u,v}-tensor is independent of ϵ\epsilon and we can, in particular, add and subtract Su,v2S^{2}_{u,v}-tensors associated with different 𝒈(ϵ)\boldsymbol{g}(\epsilon). If 𝝃\boldsymbol{\xi} denotes an Su,v2S^{2}_{u,v}-tensor (a null-decomposed Ricci-coefficient or curvature component associated with the metric 𝒈(ϵ)\boldsymbol{g}(\epsilon)) and ξ\xi denotes the corresponding tensor for ϵ=0\epsilon=0, we define its linearisation by

ξ(1):=ddϵ𝝃|ϵ=0=limϵ0𝝃ξϵ.\displaystyle\overset{(1)}{\xi}:=\frac{d}{d\epsilon}\boldsymbol{\xi}|_{\epsilon=0}=\lim_{\epsilon\rightarrow 0}\frac{\boldsymbol{\xi}-\xi}{\epsilon}\,. (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 ϵ\epsilon (i.e. for the metric 𝒈(ϵ)\boldsymbol{g}(\epsilon)) and secondly for ϵ=0\epsilon=0 (i.e. for the Schwarzschild-AdS metric) and then subtracts the respective equations, divides by ϵ\epsilon and takes the limit ϵ0\epsilon\rightarrow 0 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 ϵ=0\epsilon=0.

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

(1),^(1)AB,b(1)A,Ω1Ω(1),\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}_{AB},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{A},\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}, (44)

where (1)=12tr(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}=\frac{1}{2}\sqrt{\not{g}}\cdot\mathrm{tr}_{\not{g}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\not{g}} and ^(1)AB=(1)AB12ABtr(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}_{AB}=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\not{g}}_{AB}-\frac{1}{2}\not{g}_{AB}\mathrm{tr}_{\not{g}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\not{g}} are defined from the linearised metric (1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\not{g}}, which is in turn defined by (43), the linearised connection coefficients

(Ωtrχ)(1),(Ωtrχ¯)(1),ω(1),ω¯(1),η(1)A,η¯(1)A,χ^(1)AB,χ^¯(1)AB,\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}_{A},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}_{A},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}_{AB},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}_{AB}, (45)

and the linearised curvature components

ρ(1),σ(1),β(1)A,β¯(1)A,α(1)AB,α¯(1)AB.\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}_{A},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}_{A},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}_{AB},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}_{AB}. (46)

Depending on the number of indices, the above quantities are Su,v2S^{2}_{u,v} scalars, one-forms and symmetric traceless tensors respectively. As in [DHR19], we will speak of a solution 𝒮\mathscr{S} to the system of gravitational perturbations to mean a a collection of quantities

𝒮=(^(1),(1),Ω(1),b(1),(Ωtrχ)(1),(Ωtrχ¯)(1),χ^(1),χ^¯(1),η¯(1),η(1),ω(1),ω¯(1),α(1),β(1),ρ(1),σ(1),β¯(1),α¯(1),K(1))\displaystyle\mathscr{S}=\left(\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{K}\right) (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 11-parameter family of metrics being smooth in the extended sense) that the following linearised quantities extend smoothly to the horizon:

(^(1),(1),Ω1Ω(1),b(1)Ω2,(Ωtrχ)(1),Ω2(Ωtrχ¯)(1),Ωχ^(1),Ω1χ^¯(1),η¯(1),η(1),ω(1),Ω2ω¯(1),Ω2α(1),Ωβ(1),ρ(1),σ(1),Ω1β¯(1),Ω2α¯(1),K(1)).\displaystyle\big{(}\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}},\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}},\,\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega},\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}\Omega^{-2},\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\,,\,\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}\,,\,\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\,,\,\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}\,,\,\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}\,,\,\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}\,,\,\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho},\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}\,,\,\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}\,,\,\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{K}\big{)}. (48)

The Ω2\Omega^{-2} weight for b(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b} 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, (1),^(1),b(1),Ω1Ω(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}\,,\,\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}:

u((1))=(Ωtrχ¯)(1)divb(1),v((1))=(Ωtrχ)(1),\displaystyle\partial_{u}\Big{(}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}\Big{)}=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}-\not{\mathrm{div}}\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}\qquad,\qquad\partial_{v}\Big{(}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}\Big{)}=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\,, (49)
Ω∇̸3^(1)=2Ωχ^¯(1)+22b(1),Ω∇̸4^(1)\displaystyle\Omega\not{\nabla}_{3}\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}=2\Omega\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}+2\not{\mathcal{D}}_{2}^{\star}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}\ \ \ ,\ \ \ \Omega\not{\nabla}_{4}\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}} =2Ωχ^(1),\displaystyle=2\Omega\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}, (50)
vb(1)A=2Ω2(η(1)Aη¯(1)A),\displaystyle\partial_{v}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}^{A}=-2\Omega^{2}\left(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}^{A}-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}^{A}\right)\,, (51)
v(Ω1Ω(1))=ω(1) , u(Ω1Ω(1))=ω¯(1) , 2∇̸A(Ω1Ω(1))=η(1)A+η¯(1)A.\displaystyle\partial_{v}\left(\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}\right)=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}\textrm{ \ , \ }\partial_{u}\left(\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}\right)=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}\textrm{ \ , \ }2\not{\nabla}_{A}\left(\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}\right)=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}_{A}+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}_{A}. (52)

2.7.2 Equations for the linearised Ricci coefficients

For (Ωtrχ)(1),(Ωtrχ¯)(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)} we have the equations

v(Ωtrχ¯)(1)=Ω2(2divη¯(1)+2ρ(1)+4(ρ2k2)Ω1Ω(1))12Ωtrχ((Ωtrχ¯)(1)(Ωtrχ)(1)),\displaystyle\partial_{v}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}=\Omega^{2}\left(2\not{\mathrm{div}}\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}+2\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}+4\left(\rho-2k^{2}\right)\,\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}\right)-\frac{1}{2}\Omega tr\chi\Big{(}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\Big{)}, (53)
u(Ωtrχ)(1)=Ω2(2divη(1)+2ρ(1)+4(ρ2k2)Ω1Ω(1))12Ωtrχ((Ωtrχ¯)(1)(Ωtrχ)(1)),\displaystyle\partial_{u}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}=\Omega^{2}\left(2\not{\mathrm{div}}\,{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}}+2\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}+4\left(\rho-2k^{2}\right)\,\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}\right)-\frac{1}{2}\Omega tr\chi\Big{(}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\Big{)}, (54)
v(Ωtrχ)(1)=(Ωtrχ)(Ωtrχ)(1)+2ω(Ωtrχ)(1)+2(Ωtrχ)ω(1),\displaystyle\partial_{v}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}=-\left(\Omega tr\chi\right)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}+2\omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}+2\left(\Omega tr\chi\right)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}, (55)
u(Ωtrχ¯)(1)=(Ωtrχ¯)(Ωtrχ¯)(1)+2ω¯(Ωtrχ¯)(1)+2(Ωtrχ¯)ω¯(1),\displaystyle\partial_{u}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}=-\left(\Omega tr\underline{\chi}\right)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}+2\underline{\omega}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}+2\left(\Omega tr\underline{\chi}\right)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}, (56)

while for χ^(1),χ^¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}} we have

∇̸3(Ω1χ^¯(1))+Ω1(trχ¯)χ^¯(1)=Ω1α¯(1),∇̸4(Ω1χ^(1))+Ω1(trχ)χ^(1)=Ω1α(1),\begin{split}\not{\nabla}_{3}\Big{(}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}\Big{)}+\Omega^{-1}\left(tr\underline{\chi}\right)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}=-\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}\,,\\ \not{\nabla}_{4}\Big{(}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\Big{)}+\Omega^{-1}\left(tr{\chi}\right)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}=-\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}\,,\end{split} (57)
∇̸3(Ωχ^(1))+12(Ωtrχ¯)χ^(1)+12(Ωtrχ)χ^¯(1)\displaystyle\not{\nabla}_{3}\left(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\right)+\frac{1}{2}\left(\Omega tr\underline{\chi}\right)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}+\frac{1}{2}\left(\Omega tr\chi\right)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}} =2Ω2η(1),\displaystyle=-2\Omega\not{\mathcal{D}}_{2}^{\star}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}\,, (58)
∇̸4(Ωχ^¯(1))+12(Ωtrχ)χ^¯(1)+12(Ωtrχ¯)χ^(1)\displaystyle\not{\nabla}_{4}\left(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}\right)+\frac{1}{2}\left(\Omega tr\chi\right)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}+\frac{1}{2}\left(\Omega tr\underline{\chi}\right)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}} =2Ω2η¯(1).\displaystyle=-2\Omega\not{\mathcal{D}}_{2}^{\star}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}\,. (59)

We also have the (purely elliptic) linearised Codazzi equations on the spheres Su,v2S^{2}_{u,v}, which read

ivχ^¯(1)=12(trχ¯)η(1)+β¯(1)+12Ω∇̸A(Ωtrχ¯)(1),ivχ^(1)=12(trχ)η¯(1)β(1)+12Ω∇̸A(Ωtrχ)(1).\begin{split}\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}=-\frac{1}{2}\left(tr\underline{\chi}\right)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}+\frac{1}{2\Omega}\not{\nabla}_{A}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)},\\ \not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}=-\frac{1}{2}\left(tr{\chi}\right)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}+\frac{1}{2\Omega}\not{\nabla}_{A}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\,.\end{split} (60)

For η(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}} and η¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}} we have the transport equations

∇̸3η¯(1)=12(trχ¯)(η(1)η¯(1))+β¯(1) , ∇̸4η(1)=12(trχ)(η(1)η¯(1))β(1),\displaystyle\not{\nabla}_{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}=\frac{1}{2}\left(tr\underline{\chi}\right)\left(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}\right)+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}\textrm{ \ \ \ \ , \ \ \ \ }\not{\nabla}_{4}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}=-\frac{1}{2}\left(tr{\chi}\right)\left(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}\right)-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}, (61)

together with the elliptic equations on the spheres Su,v2S^{2}_{u,v}

urlη(1)=σ(1),urlη¯(1)=σ(1).\displaystyle\not{curl}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}\ \ \ ,\ \ \ \not{curl}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}=-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}\,. (62)

We finally have the transport equations for ω(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega} and ω¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}

vω¯(1)=Ω2(ρ(1)+2(ρ+k2)Ω1Ω(1)),\displaystyle\partial_{v}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}=-\Omega^{2}\left(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}+2\left(\rho+k^{2}\right)\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}\right)\,, (63)
uω(1)=Ω2(ρ(1)+2(ρ+k2)Ω1Ω(1)),\displaystyle\partial_{u}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}=-\Omega^{2}\left(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}+2\left(\rho+k^{2}\right)\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}\right)\,, (64)

and the linearised Gauss equation on the spheres Su,v2S^{2}_{u,v}, which reads

K(1)=ρ(1)14trχΩ((Ωtrχ¯)(1)(Ωtrχ)(1))+12Ω1Ω(1)(trχtrχ¯).\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{K}=-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}-\frac{1}{4}\frac{tr{\chi}}{\Omega}\left(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\right)+\frac{1}{2}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}\left(tr\chi tr\underline{\chi}\right)\,. (65)

We also note that K(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{K}, the linearised Gauss curvature of the double null spheres satisfies (see (221) of [DHR19])

2K(1)=12Δ̸tr(1)+divdiv^(1)1r2tr(1)where (1)=12tr(1).\displaystyle 2\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{K}=-\frac{1}{2}\not{\Delta}\mathrm{tr}_{\not{g}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\not{g}}+\not{\mathrm{div}}\not{\mathrm{div}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}-\frac{1}{r^{2}}\mathrm{tr}_{\not{g}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\not{g}}\ \ \ \ \textrm{where $\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}=\frac{1}{2}\sqrt{\not{g}}\cdot\mathrm{tr}_{\not{g}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\not{g}}$.} (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:

∇̸3α(1)+12trχ¯α(1)+2Ω1ω¯α(1)\displaystyle\not{\nabla}_{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}+\frac{1}{2}tr\underline{\chi}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}+2\Omega^{-1}\underline{{\omega}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} =22β(1)3ρχ^(1),\displaystyle=-2\not{\mathcal{D}}_{2}^{\star}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}-3\rho\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\,, (67)
∇̸4β(1)+2(trχ)β(1)Ω1ωβ(1)\displaystyle\not{\nabla}_{4}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}+2(tr\chi)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}-\Omega^{-1}{\omega}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}} =ivα(1),\displaystyle=\not{div}\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}\,, (68)
∇̸3β(1)+(trχ¯)β(1)+Ω1ω¯β(1)\displaystyle\not{\nabla}_{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}+(tr\underline{\chi})\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}+\Omega^{-1}\underline{{\omega}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}} =1(ρ(1),σ(1))+3ρη(1),\displaystyle=\not{\mathcal{D}}_{1}^{\star}\left(-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}\,\right)+3\rho\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}\,, (69)
∇̸4ρ(1)+32(trχ)ρ(1)\displaystyle\not{\nabla}_{4}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}+\frac{3}{2}(tr\chi)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho} =ivβ(1)32ρΩ(Ωtrχ)(1),\displaystyle=\not{div}\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}-\frac{3}{2}\frac{\rho}{\Omega}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\,, (70)
∇̸3ρ(1)+32(trχ¯)ρ(1)\displaystyle\not{\nabla}_{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}+\frac{3}{2}(tr\underline{\chi})\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho} =ivβ¯(1)32ρΩ(Ωtrχ¯)(1),\displaystyle=-\not{div}\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}-\frac{3}{2}\frac{\rho}{\Omega}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}\,, (71)
∇̸4σ(1)+32(trχ)σ(1)\displaystyle\not{\nabla}_{4}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}+\frac{3}{2}(tr\chi)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}} =urlβ(1),\displaystyle=-\not{curl}\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}\,, (72)
∇̸3σ(1)+32(trχ¯)σ(1)\displaystyle\not{\nabla}_{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}+\frac{3}{2}(tr\underline{\chi})\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}} =urlβ¯(1),\displaystyle=-\not{curl}\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}\,, (73)
∇̸4β¯(1)+(trχ)β¯(1)+Ω1ωβ¯(1)\displaystyle\not{\nabla}_{4}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}+(tr\chi)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}+\Omega^{-1}{\omega}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}} =1(ρ(1),σ(1))3ρη¯(1),\displaystyle=\not{\mathcal{D}}_{1}^{\star}\left(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}\,\right)-3\rho\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}\,, (74)
∇̸3β¯(1)+2(trχ¯)β¯(1)Ω1ω¯β¯(1)\displaystyle\not{\nabla}_{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}+2(tr\underline{\chi})\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}-\Omega^{-1}{\underline{\omega}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}} =ivα¯(1),\displaystyle=-\not{div}\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}\,, (75)
∇̸4α¯(1)+12(trχ)α¯(1)+2Ω1ωα¯(1)\displaystyle\not{\nabla}_{4}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}+\frac{1}{2}(tr\chi)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}+2\Omega^{-1}{\omega}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}} =22β¯(1)3ρχ^¯(1).\displaystyle=2\not{\mathcal{D}}_{2}^{\star}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}-3\rho\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}\,. (76)

2.7.4 Projections to the =0\ell=0 and =1\ell=1 modes

Suppose we are given a smooth solution 𝒮\mathscr{S} of the above system of gravitational perturbations. Then we may project all quantities of 𝒮\mathscr{S} (see (47)) to =0\ell=0 and =1\ell=1 respectively (as defined in Section 2.5.4), thereby obtaining a collection of quantities denoted by 𝒮=0\mathscr{S}_{\ell=0} and 𝒮=1\mathscr{S}_{\ell=1} respectively. One now readily checks that 𝒮=0\mathscr{S}_{\ell=0} and 𝒮=1\mathscr{S}_{\ell=1} 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 ∇̸3\not{\nabla}_{3} and ∇̸4\not{\nabla}_{4}) and the linearity of the equations. We can therefore decompose

𝒮=𝒮=0+𝒮=1+𝒮2,\mathscr{S}=\mathscr{S}_{\ell=0}+\mathscr{S}_{\ell=1}+\mathscr{S}_{\ell\geq 2}\,,

with the last term defined by the equation. This decomposition will later allow us to deal with the =0,1\ell=0,1 part of the solution independently (as far as initial data and boundary conditions are concerned), which will turn out to be convenient, as the =0,1\ell=0,1 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, \mathcal{I}, is not part of our interior manifold int\mathcal{M}_{int}. On the other hand, to formulate boundary conditions (on certain weighted quantities of 𝒮\mathscr{S} in (47)) we will need to consider Su,v2S^{2}_{u,v} tensors ξ\xi on the Schwarzschild-AdS manifold (int,g)(\mathcal{M}_{int},g), which extend smoothly to \mathcal{I}, i.e. to the larger manifold \mathcal{M}. To keep notation clean, we will often simply write ξ(u,u,θ,ϕ)\xi(u,u,\theta,\phi) or ξu0,u0\|\xi\|_{u_{0},u_{0}} to denote the appropriate limit of such tensors on \mathcal{I}. Recall in this context from (38) that the u,v\|\cdot\|_{u,v} norm is independent of the radius of the sphere Su,v2S^{2}_{u,v}.

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 𝜶,𝜷,𝜶¯,𝜷¯,𝝈\boldsymbol{\alpha},\boldsymbol{\beta},\underline{\boldsymbol{\alpha}},\underline{\boldsymbol{\beta}},\boldsymbol{\sigma} all vanish for the background Schwarzschild-AdS metric. leading to the following definition.

Definition 2.4.

We will say that a smooth solution 𝒮\mathscr{S} of the system of gravitational perturbations satisfies conformal boundary conditions provided we have for any uu0=0u\geq u_{0}=0 the limits

limvur3α(1)\displaystyle\lim_{v\rightarrow u}r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} =limvur3α¯(1),\displaystyle=\lim_{v\rightarrow u}r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}\,, (77)
limvu[Ω∇̸4Ω∇̸3](r3α(1))\displaystyle\lim_{v\rightarrow u}[\Omega\not{\nabla}_{4}-\Omega\not{\nabla}_{3}](r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}) =limvu[Ω∇̸4Ω∇̸3](r3α¯(1)),\displaystyle=-\lim_{v\rightarrow u}[\Omega\not{\nabla}_{4}-\Omega\not{\nabla}_{3}](r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}})\,, (78)
limvur3β(1)\displaystyle\lim_{v\rightarrow u}r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}} =limvur3β¯(1),\displaystyle=-\lim_{v\rightarrow u}r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}\,, (79)
limvur3σ(1)\displaystyle\lim_{v\rightarrow u}r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}} =0.\displaystyle=0\,. (80)

Here the tensorial limits are to be understood componentwise in an orthonormal frame on the spheres Su,v2S^{2}_{u,v}.

Remark 2.5.

Note that if the solution 𝒮\mathscr{S} satisfies conformal boundary conditions then also

limvurχ^(1)=limvurχ^¯(1).\displaystyle\lim_{v\rightarrow u}r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}=\lim_{v\rightarrow u}r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}\,. (81)

using the linearised Bianchi and null-structure equations. The bound (81) could of course also be deduced directly from the fact that limvur𝛘^=limvur𝛘¯^\lim_{v\rightarrow u}r\hat{\boldsymbol{\chi}}=\lim_{v\rightarrow u}r\underline{\hat{\boldsymbol{\chi}}} holds by Proposition 2.1 and trivially linearises to (81).

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 𝒮\mathscr{S} 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 {r2Ω1∇̸3,Ω∇̸4,[r∇̸]}\{r^{2}\Omega^{-1}\not{\nabla}_{3},\Omega\not{\nabla}_{4},[r\not{\nabla}]\} extend to the conformal boundary \mathcal{I}:

^(1),(1),r2Ω(1)Ω,rΩ2b(1),(Ωtrχ)(1),r2Ω2(Ωtrχ¯)(1),Ωχ^(1),r2Ω1χ^¯(1),r2η¯(1),r2η(1),rω(1),Ω2r3ω¯(1),r2K(1)\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}\,,\,\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}\,,\,r^{2}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}}{\Omega}\,,\,r\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\,,\,r^{2}\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}\,,\,\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\,,\,r^{2}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}\,,\,r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}\,,\,r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}\,,\,r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}\,,\,\Omega^{-2}r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}\,,\,r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{K} (82)

and

rΩ2α(1),r2Ωβ(1),r3ρ(1),r3σ(1),r4Ω1β¯(1),r5Ω2α¯(1).\displaystyle r\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}\,,\,r^{2}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}\,,\,r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}\,,\,r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}\,,\,r^{4}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}\,,\,r^{5}\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}. (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 r2χ^(1)r2χ^¯(1)r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}-r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}, r2ω(1)+r2ω¯(1)r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}+r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}} and r(Ωtrχ)(1)r2(Ωtrχ¯)(1)r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}-r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)} 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 f:0+×S2f:\mathbb{R}_{0}^{+}\times S^{2}\rightarrow\mathbb{R}, the corresponding functions fu=f(u,θ,ϕ)f_{u}=f(u,\theta,\phi) and fv=f(v,θ,ϕ)f_{v}=f(v,\theta,\phi), interpreted as functions on +\mathcal{M}\setminus\mathcal{H}^{+} independent of one of the coordinates, generate the following (pure gauge) solution of the system of gravitational perturbations on +\mathcal{M}\setminus\mathcal{H}^{+}:

Ω(1)Ω\displaystyle\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}}{\Omega} =12Ω2v(Ω2fv)+12Ω2u(Ω2fu),\displaystyle=\frac{1}{2\Omega^{2}}\partial_{v}\left(\Omega^{2}f_{v}\right)+\frac{1}{2\Omega^{2}}\partial_{u}\big{(}\Omega^{2}f_{u}\big{)}, ^(1)\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}} =+4rr22∇̸fu\displaystyle=+\frac{4}{r}r^{2}\not{\mathcal{D}}_{2}^{\star}\not{\nabla}{f_{u}}
(1)\displaystyle\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}} =2Ω2(fvfu)r2rr2Δ̸fu,\displaystyle=\frac{2\Omega^{2}({f_{v}-f_{u}})}{r}-\frac{2}{r}r^{2}\not{\Delta}{f_{u}}, b(1)\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b} =+2r2∇̸[u(fur)]+2Ω2∇̸fv\displaystyle=+2r^{2}\not{\nabla}\left[{\partial_{u}}\left(\frac{{f_{u}}}{r}\right)\right]+2\Omega^{2}\not{\nabla}{f_{v}}
η(1)\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}} =Ω2r2[r∇̸]fv+rΩ2∇̸[u(Ω2rfu)],\displaystyle=\frac{\Omega^{2}}{r^{2}}[r\not{\nabla}]f_{v}+\frac{r}{\Omega^{2}}\not{\nabla}\Big{[}\partial_{u}\Big{(}\frac{\Omega^{2}}{r}f_{u}\Big{)}\Big{]}, η¯(1)\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}} =rΩ2∇̸[v(Ω2rfv)]Ω2r∇̸fu,\displaystyle=\frac{r}{\Omega^{2}}\not{\nabla}\left[\partial_{v}\left(\frac{\Omega^{2}}{r}f_{v}\right)\right]-\frac{\Omega^{2}}{r}\not{\nabla}f_{u}\,,
χ^(1)\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}} =2Ωr2r22(∇̸fu),\displaystyle=-2\frac{\Omega}{r^{2}}r^{2}\not{\mathcal{D}}_{2}^{\star}(\not{\nabla}f_{u}), (Ωtrχ)(1)\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)} =2v(fvΩ2r)+2Ω2r2[(14Mrk2r2)fu+ΔS2fu],\displaystyle=2\partial_{v}\left(\frac{f_{v}\Omega^{2}}{r}\right)+\frac{2\Omega^{2}}{r^{2}}\Big{[}\Big{(}1-\frac{4M}{r}-k^{2}r^{2}\Big{)}f_{u}+\Delta_{S^{2}}f_{u}\Big{]},
χ^¯(1)\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}} =2Ωr2r22(∇̸fv),\displaystyle=-2\frac{\Omega}{r^{2}}r^{2}\not{\mathcal{D}}_{2}^{\star}(\not{\nabla}f_{v}), (Ωtrχ¯)(1)\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)} =2Ω2r2[Δ𝕊2fv+(14Mrk2r2)fv]u(2Ω2rfu),\displaystyle=\frac{2\Omega^{2}}{r^{2}}\left[\Delta_{\mathbb{S}^{2}}f_{v}+\Big{(}1-\frac{4M}{r}-k^{2}r^{2}\Big{)}f_{v}\right]-\partial_{u}\Big{(}\frac{2\Omega^{2}}{r}f_{u}\Big{)},
β(1)\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}} =6MΩr4[r∇̸]fu,\displaystyle=-\frac{6M\Omega}{r^{4}}[r\not{\nabla}]f_{u}, β¯(1)\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}} =6MΩr4r∇̸fv,\displaystyle=\frac{6M\Omega}{r^{4}}r\not{\nabla}f_{v},
ρ(1)\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho} =6MΩ2r4(fvfu),\displaystyle=\frac{6M\Omega^{2}}{r^{4}}(f_{v}-f_{u}), K(1)\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{K} =Ω2r3(Δ𝕊2(fvfu)+2(fvfu))\displaystyle=-\frac{\Omega^{2}}{r^{3}}\left(\Delta_{\mathbb{S}^{2}}(f_{v}-f_{u})+2(f_{v}-f_{u})\right)

and

α(1)=α¯(1)=0,σ(1)=0.\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}=0\ \ \ ,\ \ \ \accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}=0\,.

The solution satisfies the conformal boundary conditions of Definition 2.4. We will call ff a gauge function and denote the corresponding pure gauge solution by 𝒢f\mathscr{G}_{f}. Finally, if Ω2(u,v)f(u,θ,ϕ)\Omega^{2}(u,v)f(u,\theta,\phi) extends smoothly to +\mathcal{H}^{+} then so does the associated pure gauge solution.888In particular, the quantities (48) extend smoothly to +\mathcal{H}^{+}.

Proof.

This is verified exactly as in [DHR19] by direct computation. Since fvfuf_{v}-f_{u} vanishes on \mathcal{I}, the linearised boundary conditions (77)–(80) are indeed satisfied. ∎

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 q1(u,θ,ϕ)q_{1}(u,\theta,\phi) and q2(u,θ,ϕ)q_{2}(u,\theta,\phi) the following is a pure gauge solution of the system of gravitational perturbations

Ω1Ω(1)=0,^(1)=2r221(q1,q2),(1)=r2Δ̸q1,b(1)=r21(uq1,uq2),\displaystyle\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}=0\ \ \ ,\ \ \ \accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}=2r^{2}\not{\mathcal{D}}_{2}^{\star}\not{\mathcal{D}}_{1}^{\star}(q_{1},q_{2})\ \ ,\ \ \ \ \frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}=r^{2}\not{\Delta}q_{1}\ \ ,\ \ \ \ \accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}=r^{2}\not{\mathcal{D}}_{1}^{\star}(\partial_{u}q_{1},\partial_{u}q_{2})\,, (84)

while all linearised Ricci and null curvature components vanish identically. We denote the solution by 𝒢q\mathscr{G}_{q}.

2.9.2 The family of linearised Kerr-AdS solutions

It is well-known that the Schwarzschild-AdS family sits as a 11-parameter family in the larger 22-parameter family of Kerr-AdS metrics. At the linear level there exists (due to the spherical symmetry of the background) a 33-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 11-parameter family of changing the mass. We summarise both in Lemma 2.10 below. Let us already remark that for the =0\ell=0 modes, the pure gauge solution takes a significantly more complicated form compared to the asymptotically flat case. The underlying reason is that the variable rr (defined in terms of the (fixed) Eddington-Finkelstein coordinates (u,v)(u,v)) 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 =0\ell=0 mode.

Lemma 2.10.

For any 𝔞\mathfrak{a}\in\mathbb{R} and m{1,0,1}m\in\{-1,0,1\} the following linearised metric quantities generate a smooth solution of the system of gravitational perturbations on \mathcal{M}:

Ω1Ω(1)=0,^(1)=0,(1)=0,b(1)A=(bKAdS,m)A=2(2Mrk2r2)𝔞ϵ̸ABBYm=1.\displaystyle\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}=0\ \ \ ,\ \ \ \accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}=0\ \ \ ,\ \ \ \accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}=0\ \ \ ,\ \ \ \accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}^{A}=\left(b^{KAdS,m}\right)^{A}=-2\left(\frac{2M}{r}-k^{2}r^{2}\right)\mathfrak{a}\not{\epsilon}^{AB}\partial_{B}Y^{\ell=1}_{m}\,. (85)

The solution has the following non-vanishing Ricci-coefficients and curvature components:

η(1)A=η¯(1)A=(ηKAdS,m)A:=3M𝔞r2ϵ̸ABBYm=1,β(1)=β¯(1)=ΩrηKAdS,m,σ(1)=6r4𝔞MYm=1.\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}^{A}=-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}^{A}=\left(\eta^{KAdS,m}\right)^{A}:=\frac{3M\mathfrak{a}}{r^{2}}\not{\epsilon}^{AB}\partial_{B}Y^{\ell=1}_{m}\ \ \ ,\ \ \ \accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}=-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}=\frac{\Omega}{r}\eta^{KAdS,m}\ \ \ ,\ \ \ \accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}=\frac{6}{r^{4}}\mathfrak{a}MY^{\ell=1}_{m}\,. (86)

Moreover for any 𝔪\mathfrak{m}\in\mathbb{R}, the following linearised metric quantities generate a (spherically symmetric) smooth solution of the system of gravitational perturbations on \mathcal{M} (where we have set l2=k2=13Λl^{2}=k^{-2}=-\frac{1}{3\Lambda})

^(1)\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}} =0,\displaystyle=0\,, (1)\displaystyle\qquad\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}} =𝔪(2Mr(1+3r2l2)+MΩ2rr2Ω2(l2(l23r+2)(l2+3r+2)21r+l2(l23r~2)(l2+3r~2)21r~)𝑑r~),\displaystyle=\mathfrak{m}\left(-\frac{2M}{r\left(1+\frac{3r^{2}}{l^{2}}\right)}+\frac{M\Omega^{2}}{r}\int_{r}^{\infty}\frac{2}{\Omega^{2}}\left(\frac{l^{2}(l^{2}-3r_{+}^{2})}{(l^{2}+3r_{+}^{2})^{2}}\frac{1}{r_{+}}-\frac{l^{2}(l^{2}-3\tilde{r}^{2})}{(l^{2}+3\tilde{r}^{2})^{2}}\frac{1}{\tilde{r}}\right)d\tilde{r}\right)\,, (87)
b(1)\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b} =0,\displaystyle=0\,, Ω1Ω(1)\displaystyle\qquad\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega} =14(2𝔪MrΩ24𝔪Ml2(l23r+2)r+(l2+3r+2)2)+14(2Mr2+2rl2)(1)rΩ2.\displaystyle=\frac{1}{4}\left(\frac{2\mathfrak{m}M}{r\Omega^{2}}-4\mathfrak{m}M\frac{l^{2}(l^{2}-3r_{+}^{2})}{r_{+}(l^{2}+3r_{+}^{2})^{2}}\right)+\frac{1}{4}\left(\frac{2M}{r^{2}}+\frac{2r}{l^{2}}\right)\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}\frac{r}{\Omega^{2}}\,. (88)

In particular,999For convenience we collect the values of all Ricci-coefficients and curvature components in Appendix B. this solution satisfies

r3ρ(1)3M(1)=𝔪M.\displaystyle r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}-3M\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}=\mathfrak{m}\cdot M\,. (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 44-parameter family of linearised Kerr-AdS solutions. Given parameters (𝔪,𝔞1,𝔞0,𝔞1)(\mathfrak{m},\mathfrak{a}_{-1},\mathfrak{a}_{0},\mathfrak{a}_{1}) we denote by 𝒦m,𝔞\mathscr{K}_{m,\vec{\mathfrak{a}}} the sum of the four corresponding linearised Kerr-AdS solutions.

Remark 2.11.

Note in particular that the quantity on the left hand side of (89) is gauge invariant: Any pure gauge solution supported on =0\ell=0 leaves the quantity invariant. This follows directly from Lemma 2.8.

Remark 2.12.

The solution (85) actually remains a solution without the k2r2k^{2}r^{2} term in the definition of b(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b} as this term corresponds to a pure gauge solution from Lemma 2.9 (with q1=0q_{1}=0 and q2q_{2} supported on =1\ell=1).

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 +\mathcal{H}^{+}.

  • The four-parameter family of linearised Kerr-AdS solutions of Lemma 2.10.

  • The pure gauge solutions of Lemma 2.8, provided the function f(u)f(u) generating them is smooth on +\mathbb{R}^{+} and f(u)Ω2(u,v)r2(u,v)f(u)\frac{\Omega^{2}(u,v)}{r^{2}(u,v)} is smooth in the extended sense at +\mathcal{H}^{+}.

  • The pure gauge solutions of Lemma 2.9, provided the functions q1(u),q2(u)q_{1}(u),q_{2}(u) generating them are smooth on +\mathbb{R}^{+} and q1(u),q2(u)q_{1}(u),q_{2}(u) are smooth in the extended sense at +\mathcal{H}^{+}.

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 σ(1)r4\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}\sim r^{-4}). For the pure gauge solutions we only need to check the boundary condition for r3β(1)+r3β¯(1)0r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}+r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}\rightarrow 0 which follows from [r∇̸](fu(u)fv(v))[r\not{\nabla}](f_{u}(u)-f_{v}(v)) vanishing at the conformal boundary (where v=uv=u). 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 Ω1Ω(1)\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega} 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 r2Ω1Ω(1)r^{2}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega} extends to the boundary. For this we Taylor expand the expression of Lemma 2.8

Ω1Ω(1)\displaystyle\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega} =12(v(fv)+u(fu))+(k2r+Mr2)(fvfu)\displaystyle=\frac{1}{2}(\partial_{v}(f_{v})+\partial_{u}(f_{u}))+\left(k^{2}r+\frac{M}{r^{2}}\right)(f_{v}-f_{u})
=f(v)+f′′(v)2(vu)+(k2r+Mr2)(f(v)(uv)+f′′(v)2(vu)2)+𝒪(r2)=𝒪(r2),\displaystyle=f^{\prime}(v)+\frac{f^{\prime\prime}(v)}{2}(v-u)+\left(k^{2}r+\frac{M}{r^{2}}\right)\left(-f^{\prime}(v)(u-v)+\frac{f^{\prime\prime}(v)}{2}(v-u)^{2}\right)+\mathcal{O}(r^{-2})=\mathcal{O}(r^{-2}),

where we have recalled fv=f(v)f_{v}=f(v) and fu=f(u)f_{u}=f(u) and used (7) in conjunction with uv0u-v\geq 0. 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 α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} and α¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}, 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 (α(1),ψ(1),Ψ(1))(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\psi}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}) and (α¯(1),ψ¯(1),Ψ¯(1))(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\psi}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}}). 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 𝒮\mathscr{S} of the system of gravitational perturbations the quantities α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} and α¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}} satisfy the decoupled Teukolsky equations:

(Ω∇̸4)(Ω∇̸3)(rΩ2α(1))2(2r+6Mr2)Ω∇̸3(rΩ2α(1))Ω2r2(2r22ivΩ2rα(1)6MrrΩ2α(1))\displaystyle(\Omega\not{\nabla}_{4})(\Omega\not{\nabla}_{3})(r\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}})-2\left(-\frac{2}{r}+\frac{6M}{r^{2}}\right)\Omega\not{\nabla}_{3}(r\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}})-\frac{\Omega^{2}}{r^{2}}\left(-2r^{2}\not{\mathcal{D}}_{2}^{\star}\not{div}\Omega^{2}r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}-\frac{6M}{r}r\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}\right) =0,\displaystyle=0\,, (90)
(Ω∇̸3)(Ω∇̸4)(rΩ2α¯(1))+2(2r+6Mr2)Ω∇̸4(rΩ2α¯(1))Ω2r2(2r22ivΩ2rα(1)6MrrΩ2α¯(1))\displaystyle(\Omega\not{\nabla}_{3})(\Omega\not{\nabla}_{4})(r\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}})+2\left(-\frac{2}{r}+\frac{6M}{r^{2}}\right)\Omega\not{\nabla}_{4}(r\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}})-\frac{\Omega^{2}}{r^{2}}\left(-2r^{2}\not{\mathcal{D}}_{2}^{\star}\not{div}\Omega^{2}r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}-\frac{6M}{r}r\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}\right) =0.\displaystyle=0\,. (91)
Proof.

We give the argument for α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}, the one for α¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}} being entirely analogous. Write (67) as

(Ω∇̸3)(rΩ2α(1))=Ω4r4(2r2(Ω1β(1)r4)+6MΩ1r2χ^(1)).\displaystyle(\Omega\not{\nabla}_{3})(r\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}})=\frac{\Omega^{4}}{r^{4}}\left(-2r\not{\mathcal{D}}_{2}^{\star}(\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}r^{4})+6M\Omega^{-1}r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\right)\,. (92)

Apply Ω∇̸4\Omega\not{\nabla}_{4}, commute derivatives using (36) and insert (67) and (57) to deduce

(Ω∇̸4)(Ω∇̸3)(rΩ2α(1))=2(2r+6Mr2)Ω∇̸3(rΩ2α(1))+Ω4r4(2r52ivα(1)6Mr2α(1)).\displaystyle(\Omega\not{\nabla}_{4})(\Omega\not{\nabla}_{3})(r\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}})=2\left(-\frac{2}{r}+\frac{6M}{r^{2}}\right)\Omega\not{\nabla}_{3}(r\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}})+\frac{\Omega^{4}}{r^{4}}\left(-2r^{5}\not{\mathcal{D}}_{2}^{\star}\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}-6Mr^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}\right)\,. (93)

2.10.2 The gauge invariant hierarchy

Given symmetric traceless tensors α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}, α¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}} we define (consistently with [DHR19]) the quantities (ψ(1),ψ¯(1))(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\psi}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\psi}}) by:

2r3Ωψ(1):=r2Ω2Ω∇̸3(rΩ2α(1))\displaystyle-2r^{3}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\psi}}:=\frac{r^{2}}{\Omega^{2}}\Omega\not{\nabla}_{3}(r\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}) =2r2(Ωr2β(1))3ρr3Ωχ^(1),\displaystyle=-2r\not{\mathcal{D}}_{2}^{\star}\left(\Omega r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}\right)-3\rho r^{3}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\,, (94)
2r3Ωψ¯(1):=r2Ω2Ω∇̸4(rΩ2α¯(1))\displaystyle 2r^{3}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\psi}}:=\frac{r^{2}}{\Omega^{2}}\Omega\not{\nabla}_{4}(r\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}) =2r2(Ωr2β¯(1))3ρr3Ωχ^¯(1).\displaystyle=2r\not{\mathcal{D}}_{2}^{\star}\left(\Omega r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}\right)-3\rho r^{3}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}. (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 +\mathcal{H}^{+}.

Ω∇̸4(r5Ω2α¯(1))4r5Ω2α¯(1)(1r3Mr2)\displaystyle\Omega\not{\nabla}_{4}(r^{5}\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}})-4r^{5}\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}\left(\frac{1}{r}-\frac{3M}{r^{2}}\right) =2r52Ω1β¯(1)3ρr5Ω1χ^¯(1).\displaystyle=2r^{5}\not{\mathcal{D}}_{2}^{\star}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}-3\rho r^{5}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}\,. (96)

Again consistent with [DHR19] we define also the higher order quantities (Ψ(1),Ψ¯(1))(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}}) by

Ψ(1):=r2Ω2Ω∇̸3(r3Ωψ(1))=r2Ω1∇̸3(12r2Ω1∇̸3(rΩ2α(1)))\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}:=\frac{r^{2}}{\Omega^{2}}\Omega\not{\nabla}_{3}(r^{3}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\psi}})=r^{2}\Omega^{-1}\not{\nabla}_{3}\left(-\frac{1}{2}r^{2}\Omega^{-1}\not{\nabla}_{3}(r\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}})\right) =r521(ρ(1),σ(1))+34r5ρtrχ(χ^(1)χ^¯(1)),\displaystyle=r^{5}\not{\mathcal{D}}_{2}^{\star}\not{\mathcal{D}}_{1}^{\star}\left(-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}\right)+\frac{3}{4}r^{5}\rho tr\chi(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}})\,, (97)
Ψ¯(1):=+r2Ω2Ω∇̸4(r3Ωψ¯(1))=r2Ω1∇̸4(12r2Ω1∇̸4(rΩ2α¯(1)))\displaystyle-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}}:=+\frac{r^{2}}{\Omega^{2}}\Omega\not{\nabla}_{4}(r^{3}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\psi}})=r^{2}\Omega^{-1}\not{\nabla}_{4}\left(\frac{1}{2}r^{2}\Omega^{-1}\not{\nabla}_{4}(r\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}})\right) =r521(ρ(1),σ(1))34r5ρtrχ¯(χ^(1)χ^¯(1)).\displaystyle=r^{5}\not{\mathcal{D}}_{2}^{\star}\not{\mathcal{D}}_{1}^{\star}\left(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}\right)-\frac{3}{4}r^{5}\rho tr\underline{\chi}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}). (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 :=r2Δ̸+4\mathcal{L}:=-r^{2}\not{\Delta}+4, one has from (90), (91) and the definitions (97), (98), that Ψ(1),Ψ¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}} satisfy the Regge-Wheeler equation (see [GH24])

(Ω∇̸4)(Ω∇̸3)Φ+Ω2r2(6Mr)Φ\displaystyle(\Omega\not{\nabla}_{4})(\Omega\not{\nabla}_{3})\Phi+\frac{\Omega^{2}}{r^{2}}\left(\mathcal{L}-\frac{6M}{r}\right)\Phi =0,\displaystyle=0, (99)

where Φ=Ψ(1),Ψ¯(1)\Phi=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}}. We further define101010Recall from Proposition 4.4.4 of [DHR19] that \mathcal{L} has eigenvalues (+1)6\ell(\ell+1)\geq 6 acting on symmetric traceless tensors.

Ψ(1)D:=Ψ(1)Ψ¯(1),Ψ(1)R:=Ψ(1)+Ψ¯(1)+12M1(2)1t(Ψ(1)Ψ¯(1)).\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{D}:=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}}\ \ \ ,\ \ \ \accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{R}:=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}}+12M\mathcal{L}^{-1}(\mathcal{L}-2)^{-1}\partial_{t}\left(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}}\right). (100)

It is easy to see that, if α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} and α¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}} are regular at the horizon and at infinity (see (48) and Definition 2.6), Ψ(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi} and Ψ¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}} and hence also Ψ(1)D\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{D} and Ψ(1)R\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{R} extend regularly to both +\mathcal{H}^{+} and \mathcal{I}. In fact, the quantities Ψ(1)D\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{D} and Ψ(1)R\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{R} correspond (up to an unimportant numerical constant) to the analogous quantities ΨD\Psi^{D} and ΨR\Psi^{R} 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 C¯v0\underline{C}_{v_{0}} 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 𝒮\mathscr{S} of the system of gravitational perturbations.

Definition 3.1.

We will say that 𝒮\mathscr{S} (as in (47)) is a smooth solution of the system of gravitational perturbations satisfying the boundary conditions if

  • 𝒮\mathscr{S} satisfies the equations (49)–(76) on int+\mathcal{M}_{int}\setminus\mathcal{H}^{+}, with all dynamical quantities of 𝒮\mathscr{S} being smooth functions on int+\mathcal{M}_{int}\setminus\mathcal{H}^{+} and

  • the boundary conditions (77)–(80) hold on \mathcal{I}.

As mentioned, our goal is to construct such solutions uniquely from an appropriate notion of smooth seed data on C¯v0\underline{C}_{v_{0}}.

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 C¯v0=0\underline{C}_{v_{0}=0}. Below, an Su,v2S^{2}_{u,v}-tensor ξ\xi is called smooth along C¯v0\underline{C}_{v_{0}} if for all i,j{0}i,j\in\mathbb{N}\cup\{0\}, the tensor [r2Ω2Ω∇̸3]i[r∇̸]jξ\left[\frac{r^{2}}{\Omega^{2}}\Omega\not{\nabla}_{3}\right]^{i}[r\not{\nabla}]^{j}\xi extends smoothly to infinity and to the horizon along C¯v0\underline{C}_{v_{0}}.111111In particular, the components in an orthonormal frame extend as smooth functions to \mathcal{I}.

Definition 3.3.

A smooth seed initial data set on C¯v0=0\underline{C}_{v_{0}=0} for the system of gravitational perturbations consists of

  1. 1.

    a tuple (α¯(1)0;ψ¯(1)0,Ψ¯(1)0)(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}_{0};\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\psi}}_{0},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}}_{0}), called the gauge independent part,

  2. 2.

    a tuple (B¯0,R0;H¯0;Ω1Ω(1)0,b(1)0,G0=1,G^0)(\underline{B}_{0},R_{0};\underline{H}_{0};\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}_{0},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{0},G^{\ell=1}_{0},\hat{G}_{0}), called the gauge dependent part,

  3. 3.

    a 44-dimensional vector (𝔪,𝔞1,𝔞0,𝔞1)4(\mathfrak{m},\mathfrak{a}_{-1},\mathfrak{a}_{0},\mathfrak{a}_{1})\in\mathbb{R}^{4} called the Kerr-AdS part.

The gauge independent part (α¯(1)0;ψ¯(1)0,Ψ¯(1)0)(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}_{0};\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\psi}}_{0},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}}_{0}) consists of

  • a smooth symmetric traceless tensor α¯(1)0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}_{0} prescribed along C¯v0\underline{C}_{v_{0}} with r3α¯(1)0r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}_{0} extending smoothly to Su0=v0,v02S^{2}_{u_{0}=v_{0},v_{0}},

  • smooth symmetric traceless Su0=v0,v02S^{2}_{u_{0}=v_{0},v_{0}} tensors ψ¯(1)0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\psi}}_{0} and Ψ¯(1)0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}}_{0}.

The gauge dependent part (B¯0,R0;H¯0;Ω1Ω(1)0,b(1)0,G0=1,G^0)(\underline{B}_{0},R_{0};\underline{H}_{0};\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}_{0},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{0},G^{\ell=1}_{0},\hat{G}_{0}) consists of

  • Su0=v0,v02S^{2}_{u_{0}=v_{0},v_{0}} scalars B¯0\underline{B}_{0}, R0R_{0}, H¯0\underline{H}_{0}, G0=1G^{\ell=1}_{0} with G0=1G^{\ell=1}_{0} supported for =1\ell=1,

  • a smooth symmetric traceless tensor G^0\hat{G}_{0} on Su0=v0,v02S^{2}_{u_{0}=v_{0},v_{0}},

  • a smooth lapse function Ω1Ω(1)0\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}_{0} prescribed along C¯v0\underline{C}_{v_{0}} with r2Ω1Ω(1)0r^{2}\cdot\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}_{0} extending smoothly to Su0=v0,v02S^{2}_{u_{0}=v_{0},v_{0}},

  • a smooth shift function b(1)0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{0} prescribed along C¯v0\underline{C}_{v_{0}},with 1rb(1)0\frac{1}{r}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{0} extending smoothly to Su0=v0,v02S^{2}_{u_{0}=v_{0},v_{0}}.

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 𝒮\mathscr{S} 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

(α¯(1)0;ψ¯(1)0,Ψ¯(1)0),(B¯0,R0;H¯0;Ω1Ω(1)0,b(1)0,G0=1,G^0),(𝔪,𝔞1,𝔞0,𝔞1),(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}_{0};\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\psi}}_{0},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}}_{0})\ \ ,\ \ (\underline{B}_{0},R_{0};\underline{H}_{0};\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}_{0},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{0},G^{\ell=1}_{0},\hat{G}_{0})\ \ ,\ \ (\mathfrak{m},\mathfrak{a}_{-1},\mathfrak{a}_{0},\mathfrak{a}_{1}),

we say that a solution 𝒮\mathscr{S} 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 α¯(1)(u,v0,θ)=α¯(1)0(u,θ)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}(u,v_{0},\theta)=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}_{0}(u,\theta), Ω1Ω(1)(u,v0,θ)=Ω1Ω(1)0(u,θ)\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}(u,v_{0},\theta)=\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}_{0}(u,\theta) and b(1)(u,v0,θ)=b(1)0(u,θ)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}(u,v_{0},\theta)=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{0}(u,\theta) along C¯v0\underline{C}_{v_{0}} and

ψ¯(1)0\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\psi}}_{0} =limuu0(Ω∇̸4(r3α¯(1)))(u,v0,θ)=limuu0k(2r42β¯(1)+6Mrχ^¯(1))(u,v0,θ),\displaystyle=\lim_{u\rightarrow u_{0}}\left(\Omega\not{\nabla}_{4}(r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}})\right)(u,v_{0},\theta)=\lim_{u\rightarrow u_{0}}k\left(2r^{4}\not{\mathcal{D}}_{2}^{\star}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}+6Mr\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}\right)(u,v_{0},\theta)\,, (101)
Ψ¯(1)0\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}}_{0} =limuu0(Ω∇̸4(Ω∇̸4(r3α¯(1))))(u,v0,θ)=2k2r22∇̸R0+6Mk3limuu0r2(χ^(1)χ^¯(1))(u,v0,θ),\displaystyle=\lim_{u\rightarrow u_{0}}\left(\Omega\not{\nabla}_{4}(\Omega\not{\nabla}_{4}(r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}))\right)(u,v_{0},\theta)=-2k^{2}r^{2}\not{\mathcal{D}}_{2}^{\star}\not{\nabla}R_{0}+6Mk^{3}\lim_{u\rightarrow u_{0}}r^{2}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}})(u,v_{0},\theta)\,, (102)
B¯0\displaystyle\underline{B}_{0} =limuu0[riv]β¯(1)r3(u,v0,θ),\displaystyle=\lim_{u\rightarrow u_{0}}[r\not{div}]\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}r^{3}(u,v_{0},\theta)\,, (103)
R0\displaystyle R_{0} =limuu0r3ρ(1)(u,v0,θ),\displaystyle=\lim_{u\rightarrow u_{0}}r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}(u,v_{0},\theta)\,, (104)
H¯0\displaystyle\underline{H}_{0} =limuu0r2Ω2(Ωtrχ¯)(1)(u,v0,θ),\displaystyle=\lim_{u\rightarrow u_{0}}r^{2}\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}(u,v_{0},\theta)\,, (105)
G0=1\displaystyle G^{\ell=1}_{0} =limuu0((1))=1(u,v0,θ),\displaystyle=\lim_{u\rightarrow u_{0}}\Big{(}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}\Big{)}_{\ell=1}(u,v_{0},\theta)\,, (106)
G^0\displaystyle\hat{G}_{0} =limuu0^(1)(u,v0,θ).\displaystyle=\lim_{u\rightarrow u_{0}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}(u,v_{0},\theta)\,. (107)

Finally,

𝔞m=16Mlimuu0(r4urlβ(1))=1,m(u,v0,θ)and𝔪=1Mlimuu0(r3ρ(1)3M(1))=0(u,v0,θ).\mathfrak{a}_{m}=\frac{1}{6M}\lim_{u\rightarrow u_{0}}(r^{4}\not{curl}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}})_{\ell=1,m}(u,v_{0},\theta)\ \ \ \textrm{and}\ \ \ \mathfrak{m}=\frac{1}{M}\lim_{u\rightarrow u_{0}}\bigg{(}r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}-3M\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}\bigg{)}_{\ell=0}(u,v_{0},\theta)\,.

It is clear that the quantities (α¯(1)0;ψ¯(1)0,Ψ¯(1)0)(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}_{0};\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\psi}}_{0},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}}_{0}) are gauge independent. The terminology “gauge dependent” for (Ω1Ω(1)0,b(1)0;B¯0,H¯0,R0,G0=1,G^0)(\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}_{0},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{0};\underline{B}_{0},\underline{H}_{0},R_{0},G^{\ell=1}_{0},\hat{G}_{0}) and “Kerr-AdS” for (𝔪,𝔞1,𝔞0,𝔞1)(\mathfrak{m},\mathfrak{a}_{-1},\mathfrak{a}_{0},\mathfrak{a}_{1}) becomes clear from the following proposition. To state it, note that clearly any solution 𝒮\mathscr{S} of the system of gravitational perturbations satisfying the boundary conditions induces a seed initial data set by restricting the solution to C¯v0\underline{C}_{v_{0}} and taking the above limits.

Proposition 3.5.

Consider a smooth seed initial data set

(α¯(1)0;ψ¯(1)0,Ψ¯(1)0),(B¯0,R0;H¯0;Ω1Ω(1)0,b(1)0,G0=1,G^0),(𝔪,𝔞1,𝔞0,𝔞1)(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}_{0};\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\psi}}_{0},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}}_{0})\ \ ,\ \ (\underline{B}_{0},R_{0};\underline{H}_{0};\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}_{0},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{0},G^{\ell=1}_{0},\hat{G}_{0})\ \ ,\ \ ({\mathfrak{m}},{\mathfrak{a}}_{-1},{\mathfrak{a}}_{0},{\mathfrak{a}}_{1})

and assume there exists a solution 𝒮\mathscr{S} 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 f:[u0,)×𝕊2f:[u_{0},\infty)\times\mathbb{S}^{2}\rightarrow\mathbb{R}, a function q:[u0,)×𝕊2q:[u_{0},\infty)\times\mathbb{S}^{2}\rightarrow\mathbb{R} such that the solution

𝒮+𝒢f+𝒢q𝒦𝔪,𝔞\mathscr{S}+\mathscr{G}_{f}+\mathscr{G}_{q}-\mathscr{K}_{\mathfrak{m},\vec{\mathfrak{a}}}

induces a seed initial data set whose gauge dependent part and Kerr-AdS part vanishes identically. Here 𝒢f\mathscr{G}_{f} is the pure gauge solution induced by ff as in Lemma 2.8, 𝒢q\mathscr{G}_{q} the pure gauge solution induced by qq as in Lemma 2.9 and 𝒦𝔪,𝔞\mathscr{K}_{\mathfrak{m},\vec{\mathfrak{a}}} the linearised Kerr-AdS solution of Lemma 2.10.

Proof.

Step 1. Subtracting the linearised Kerr-AdS solution 𝒦𝔪,𝔞\mathscr{K}_{\mathfrak{m},\vec{\mathfrak{a}}}. We note that

(r3ρ(1)3M(1))=0|𝒮𝒦𝔪,𝔞(u0,v0)=(r3ρ(1)3M(1))=0|𝒮(u0,v0)𝔪M=0.\bigg{(}r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}-3M\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}\bigg{)}_{\ell=0}\Bigg{|}_{\mathscr{S}-\mathcal{K}_{\mathfrak{m},\vec{\mathfrak{a}}}}(u_{0},v_{0})=\bigg{(}r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}-3M\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}\bigg{)}_{\ell=0}\Bigg{|}_{\mathscr{S}}(u_{0},v_{0})-\mathfrak{m}\cdot M=0\,.

We choose the 𝔞=(𝔞1,𝔞0,𝔞1)\vec{\mathfrak{a}}=(\mathfrak{a}_{1},\mathfrak{a}_{0},\mathfrak{a}_{-1}) in the linearised Kerr-AdS solution such that for m{1,0,1}m\in\{-1,0,1\} we have

(rurl(β(1)r3))=1,m|𝒮𝒦𝔪,𝔞(u0,v0)=(rurl(β(1)r3))=1,m|𝒮(u0,v0)6M𝔞m=0.\left(r\not{curl}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}r^{3})\right)_{\ell=1,m}\Big{|}_{\mathscr{S}-\mathcal{K}_{\mathfrak{m},\vec{\mathfrak{a}}}}(u_{0},v_{0})=\left(r\not{curl}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}r^{3})\right)_{\ell=1,m}\Big{|}_{\mathscr{S}}(u_{0},v_{0})-6M\mathfrak{a}_{m}=0\,.

Note that both these quantities are gauge invariant and hence not affected by adding 𝒢f\mathscr{G}_{f} and 𝒢q\mathscr{G}_{q} in the following steps. Clearly, by construction 𝒮~:=𝒮𝒦𝔪,𝔞\tilde{\mathscr{S}}:=\mathscr{S}-\mathscr{K}_{\mathfrak{m},\vec{\mathfrak{a}}} induces seed data with vanishing Kerr-AdS part.

Step 2. Defining 𝒢f\mathscr{G}_{f}. Given the solution 𝒮~=𝒮𝒦𝔪,𝔞\tilde{\mathscr{S}}=\mathscr{S}-\mathscr{K}_{\mathfrak{m},\vec{\mathfrak{a}}}, we define the function ff inducing the desired pure gauge solution 𝒢f\mathscr{G}_{f} (by setting fu=f(u,θ)f_{u}=f(u,\theta) and fv=f(v,θ)f_{v}=f(v,\theta) in Lemma 2.8) as follows:

f(u,θ)\displaystyle f(u,\theta) =λ1r(u,v0)+λ0Ω2(u,v0)+λ2Ω2(u,v0)1Ω2(u,v0)u0u2Ω2Ω1Ω(1)(u^,v0,θ)𝑑u^\displaystyle=\frac{\lambda_{1}r(u,v_{0})+\lambda_{0}\Omega^{2}(u,v_{0})+\lambda_{2}}{\Omega^{2}(u,v_{0})}-\frac{1}{\Omega^{2}(u,v_{0})}\int_{u_{0}}^{u}2\Omega^{2}\cdot\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}(\hat{u},v_{0},\theta)d\hat{u} (108)

for scalars λ0,λ1,λ2\lambda_{0},\lambda_{1},\lambda_{2} defined in turn by121212The λ0,λ1,λ2\lambda_{0},\lambda_{1},\lambda_{2} are unique up to the spherical means of λ0\lambda_{0}. Note that the =0\ell=0 mode of λ0\lambda_{0} generates a trivial pure gauge solution (changing uu and vv by a constant) corresponding to the static isometry of the background.

  • λ0\lambda_{0} such that (B¯0)𝒮~(θ)+6Mkr2Δ̸λ0=0(\underline{B}_{0})_{\tilde{\mathscr{S}}}(\theta)+6Mkr^{2}\not{\Delta}\lambda_{0}=0,

  • λ1\lambda_{1} such that (R0)𝒮~(θ)+6Mλ1=0(R_{0})_{\tilde{\mathscr{S}}}(\theta)+6M\lambda_{1}=0,

  • λ2\lambda_{2} such that (H¯0)𝒮~(θ)+2Δ𝕊2λ02λ2=0(\underline{H}_{0})_{\tilde{\mathscr{S}}}(\theta)+2\Delta_{\mathbb{S}^{2}}\lambda_{0}-2\lambda_{2}=0.

We claim that with 𝒢f\mathscr{G}_{f} thus defined, the solution 𝒮~+𝒢f\tilde{\mathscr{S}}+\mathscr{G}_{f} already has (Ω1Ω(1)0)𝒮~+𝒢f=0(\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}_{0})_{\tilde{\mathscr{S}}+\mathscr{G}_{f}}=0 and (B¯0)𝒮~+𝒢f=0(\underline{B}_{0})_{\tilde{\mathscr{S}}+\mathscr{G}_{f}}=0, (R0)𝒮~+𝒢f=0(R_{0})_{\tilde{\mathscr{S}}+\mathscr{G}_{f}}=0 and (H¯0)𝒮~+𝒢f=0(\underline{H}_{0})_{\tilde{\mathscr{S}}+\mathscr{G}_{f}}=0. To verify this, we first note that

limuu0fu(u,θ)=limvv0fv(v,θ)=λ0andlimuu0(ufu)(u,θ)=limvv0(vfv)(v,θ)=λ1.\lim_{u\rightarrow u_{0}}f_{u}(u,\theta)=\lim_{v\rightarrow v_{0}}f_{v}(v,\theta)=\lambda_{0}\ \ \ \textrm{and}\ \ \ \ \lim_{u\rightarrow u_{0}}(\partial_{u}f_{u})(u,\theta)=\lim_{v\rightarrow v_{0}}(\partial_{v}f_{v})(v,\theta)=\lambda_{1}\,.

Using the expressions in Lemma 2.8 we now easily check that

(Ω1Ω(1))𝒮~+𝒢f(u,v0,θ)\displaystyle(\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega})_{\tilde{\mathscr{S}}+\mathscr{G}_{f}}(u,v_{0},\theta)
=(Ω1Ω(1))(u,v0,θ)𝒮~+12(vfv)(v0,θ)+12(rΩ2)(u,v0)fv(v0,θ)+12Ω2u(Ω2fu)(u,v0,θ)\displaystyle=(\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega})(u,v_{0},\theta)_{\tilde{\mathscr{S}}}+\frac{1}{2}(\partial_{v}f_{v})(v_{0},\theta)+\frac{1}{2}(\partial_{r}\Omega^{2})(u,v_{0})f_{v}(v_{0},\theta)+\frac{1}{2\Omega^{2}}\partial_{u}\big{(}\Omega^{2}f_{u}\big{)}(u,v_{0},\theta)
=(Ω1Ω(1))𝒮~(u,v0,θ)+12λ1+12λ0(rΩ2)(u,v0)+12(λ1λ0r(Ω2)2Ω1Ω(1)𝒮~)(u,v0,θ)=0\displaystyle=(\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega})_{\tilde{\mathscr{S}}}(u,v_{0},\theta)+\frac{1}{2}\lambda_{1}+\frac{1}{2}\lambda_{0}(\partial_{r}\Omega^{2})(u,v_{0})+\frac{1}{2}\left(-\lambda_{1}-\lambda_{0}\partial_{r}(\Omega^{2})-2\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}_{\tilde{\mathscr{S}}}\right)(u,v_{0},\theta)=0 (109)

on the ingoing cone C¯v0\underline{C}_{v_{0}} independently of the λi\lambda_{i}. From Lemma 2.8 we then verify

limuu0(rivr4Ω1β¯(1))𝒮~+𝒢f(u,v0,θ)\displaystyle\lim_{u\rightarrow u_{0}}(r\not{div}r^{4}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}})_{\tilde{\mathscr{S}}+\mathscr{G}_{f}}(u,v_{0},\theta) =limuu0(rivr4Ω1β¯(1))𝒮~(u,v0,θ)+6Mr2Δ̸λ0=0.\displaystyle=\lim_{u\rightarrow u_{0}}(r\not{div}r^{4}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}})_{\tilde{\mathscr{S}}}(u,v_{0},\theta)+6Mr^{2}\not{\Delta}\lambda_{0}=0\,. (110)

To check the condition on ρ(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}, we first compute the limit

limuv((r(u,v)(fv(v,θ)fu(u,θ)))=limvu(fvfuvu)=f(u).\lim_{u\rightarrow v}\left((r(u,v)(f_{v}(v,\theta)-f_{u}(u,\theta))\right)=\lim_{v\rightarrow u}\left(\frac{f_{v}-f_{u}}{v-u}\right)=f^{\prime}(u)\,.

It therefore follows from Lemma 2.8 that we have

(r3ρ(1))𝒮~+𝒢f(v0,v0,θ)\displaystyle(r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho})_{\tilde{\mathscr{S}}+\mathscr{G}_{f}}(v_{0},v_{0},\theta) =(r3ρ(1))𝒮~(v0,v0,θ)+6Mlimuv0((r(u,v0)(fv(v0,θ)fu(u,θ)))\displaystyle=(r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho})_{\tilde{\mathscr{S}}}(v_{0},v_{0},\theta)+6M\lim_{u\rightarrow v_{0}}\left((r(u,v_{0})(f_{v}(v_{0},\theta)-f_{u}(u,\theta))\right)
=(r3ρ(1))𝒮~(v0,v0,θ)+6Mlimuv0(f(v0,θ)f(u,θ)v0u)\displaystyle=(r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho})_{\tilde{\mathscr{S}}}(v_{0},v_{0},\theta)+6M\lim_{u\rightarrow v_{0}}\left(\frac{f(v_{0},\theta)-f(u,\theta)}{v_{0}-u}\right)
=(r3ρ(1))𝒮~(v0,v0,θ)+6Mf(v0),\displaystyle=(r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho})_{\tilde{\mathscr{S}}}(v_{0},v_{0},\theta)+6Mf^{\prime}(v_{0})\,, (111)

and we see that (R0)𝒮~+𝒢f=0(R_{0})_{\tilde{\mathscr{S}}+\mathscr{G}_{f}}=0 if λ1\lambda_{1} is chosen as above. For the last limit we note using Lemma 2.8 and the fact that we already obtained ω¯(1)𝒮~+𝒢f=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}_{\tilde{\mathscr{S}}+\mathscr{G}_{f}}=0, the equalities

r2Ω2(Ωtrχ¯)(1)𝒮~+𝒢f(u,v0,θ)\displaystyle\frac{r^{2}}{\Omega^{2}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}_{\tilde{\mathscr{S}}+\mathscr{G}_{f}}(u,v_{0},\theta)
=r2Ω2(Ωtrχ¯)(1)𝒮~(u,v0,θ)+2[Δ𝕊2λ0+λ0(14Mrk2r2)(u,v0)]r2Ω2u(2Ω2rfu)(u,v0,θ).\displaystyle=\frac{r^{2}}{\Omega^{2}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}_{\tilde{\mathscr{S}}}(u,v_{0},\theta)+2\left[\Delta_{\mathbb{S}^{2}}\lambda_{0}+\lambda_{0}\Big{(}1-\frac{4M}{r}-k^{2}r^{2}\Big{)}(u,v_{0})\right]-\frac{r^{2}}{\Omega^{2}}\partial_{u}\Big{(}\frac{2\Omega^{2}}{r}f_{u}\Big{)}(u,v_{0},\theta)\,.

We want to take the limit uu0=v0u\rightarrow u_{0}=v_{0}. We compute up to terms vanishing in the limit uu0u\rightarrow u_{0} (indicated by \equiv)

r2Ω2u(2Ω2rfu)(u,v0,θ)\displaystyle-\frac{r^{2}}{\Omega^{2}}\partial_{u}\Big{(}\frac{2\Omega^{2}}{r}f_{u}\Big{)}(u,v_{0},\theta) 2Ω2fu(u,v0,θ)2r(λ1(Ω2)rλ02Ω1Ω(1)𝒮~)(u,v0,θ)\displaystyle\equiv-2\Omega^{2}f_{u}(u,v_{0},\theta)-2r\left(-\lambda_{1}-(\Omega^{2})_{r}\lambda_{0}-2\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}_{\tilde{\mathscr{S}}}\right)(u,v_{0},\theta)
2(+λ0Ω2(u,v0)r(Ω2)rλ0+λ2)\displaystyle\equiv-2\left(+\lambda_{0}\Omega^{2}(u,v_{0})-r(\Omega^{2})_{r}\lambda_{0}+\lambda_{2}\right)
2(λ0k2r2(u,v0)+λ0+λ2).\displaystyle\equiv-2\left(-\lambda_{0}k^{2}r^{2}(u,v_{0})+\lambda_{0}+\lambda_{2}\right)\,. (112)

We conclude

(H¯0)𝒮~+𝒢f=(H¯0)𝒮~+2Δ𝕊2λ02λ2=0.\displaystyle(\underline{H}_{0})_{\tilde{\mathscr{S}}+\mathscr{G}_{f}}=(\underline{H}_{0})_{\tilde{\mathscr{S}}}+2\Delta_{\mathbb{S}^{2}}\lambda_{0}-2\lambda_{2}=0\,. (113)

Step 3. Defining 𝒢q\mathscr{G}_{q}. Let us define (unique) functions ϕ1,ϕ1\phi_{1},\phi_{1} with vanishing spherical means by b(1)𝒮~+𝒢f=r1(ϕ1,ϕ2)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{\tilde{\mathscr{S}}+\mathscr{G}_{f}}=r\not{\mathcal{D}}_{1}^{\star}(\phi_{1},\phi_{2}). Set q1,q2q_{1},q_{2} to solve

uq1+ϕ1r=0,uq2+ϕ2r=0,\partial_{u}q_{1}+\frac{\phi_{1}}{r}=0\ \ \ ,\ \ \ \partial_{u}q_{2}+\frac{\phi_{2}}{r}=0\,,

with initial conditions at u0u_{0} determined by q1(u0,θ)q_{1}(u_{0},\theta) and q2(u0,θ)q_{2}(u_{0},\theta) having vanishing spherical means and solving ^(1)(u0,θ)=2r221(q1(u0,θ),q2(u0,θ))\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}(u_{0},\theta)=2r^{2}\not{\mathcal{D}}_{2}^{\star}\not{\mathcal{D}}_{1}^{\star}(q_{1}(u_{0},\theta),q_{2}(u_{0},\theta)) and r2Δ̸q1=G0=1(u0,θ)r^{2}\not{\Delta}q_{1}=G^{\ell=1}_{0}(u_{0},\theta). Note that this determines q1,q2q_{1},q_{2} uniquely up to =1\ell=1 modes of q2q_{2} on the sphere Su0,v02S^{2}_{u_{0},v_{0}} (corresponding to the rotational invariance), which generate trivial gauge transformations. It is now immediate from Lemma 2.9 that (G0=1)𝒮~+𝒢f+𝒢q(θ)=0(G^{\ell=1}_{0})_{\tilde{\mathscr{S}}+\mathscr{G}_{f}+\mathscr{G}_{q}}(\theta)=0, (G^)𝒮~+𝒢f+𝒢q(θ)=0(\hat{G})_{\tilde{\mathscr{S}}+\mathscr{G}_{f}+\mathscr{G}_{q}}(\theta)=0 and b(1)𝒮~+𝒢f+𝒢q(u,v0,θ)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{\tilde{\mathscr{S}}+\mathscr{G}_{f}+\mathscr{G}_{q}}(u,v_{0},\theta)=0, 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

(α¯(1)0;ψ¯(1)0,Ψ¯(1)0),(B¯0,R0;H¯0;Ω1Ω(1)0,b(1)0,G0=1,G^0),(𝔪,𝔞1,𝔞0,𝔞1)(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}_{0};\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\psi}}_{0},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}}_{0})\ \ ,\ \ (\underline{B}_{0},R_{0};\underline{H}_{0};\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}_{0},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{0},G^{\ell=1}_{0},\hat{G}_{0})\ \ ,\ \ (\mathfrak{m},\mathfrak{a}_{-1},\mathfrak{a}_{0},\mathfrak{a}_{1})

one can construct uniquely all geometric quantities (47) of the system of gravitational perturbations (including all their tangential and transversal derivatives) on C¯v0\underline{C}_{v_{0}} such that

  1. (1)

    The equations (49)–(76) and all tangential and transversal derivatives thereof, hold on the cone C¯v0\underline{C}_{v_{0}}.

  2. (2)

    The boundary conditions (77)–(81) hold on the sphere Su0,v02S^{2}_{u_{0},v_{0}} of the conformal boundary.

  3. (3)

    The condition [∇̸T]p[r∇̸]qω(1)=0[\not{\nabla}_{T}]^{p}[r\not{\nabla}]^{q}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}=0 holds on the sphere Su0,v02S^{2}_{u_{0},v_{0}} for any p,q0p,q\in\mathbb{N}^{0}.

  4. (4)

    On C¯v0\underline{C}_{v_{0}}, 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 C¯v0\underline{C}_{v_{0}}.

Remark 3.7.

To explain (3)(3) above, we observe that some condition on ω(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega} at Su0,v02S^{2}_{u_{0},v_{0}} is necessary to determine all quantities on the cone as the quantity ω(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega} only admits a null structure equation in the 33-direction and hence needs to be supplemented with data on Su0,v02S^{2}_{u_{0},v_{0}}. Our choice in (3)(3) 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 C¯v0\underline{C}_{v_{0}} by Items (1)(4)(1)-(4) 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 ω¯(1)=Ω∇̸3(Ω1Ω(1)0)(u,v0,θ)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}=\Omega\not{\nabla}_{3}(\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}_{0})(u,v_{0},\theta) consistent with (52). Note that rω¯(1)(u0,v0,θ)r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}(u_{0},v_{0},\theta) extends smoothly to u=u0u=u_{0}.

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 C¯v0\underline{C}_{v_{0}} must hold. This determines (by transport along C¯v0\underline{C}_{v_{0}}) the regular transversal derivatives Ω∇̸4(r5Ω2α¯(1))\Omega\not{\nabla}_{4}(r^{5}\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}) and Ω∇̸4(Ω∇̸4(r5Ω2α¯(1)))\Omega\not{\nabla}_{4}(\Omega\not{\nabla}_{4}(r^{5}\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}})) along all of C¯v0\underline{C}_{v_{0}} in terms of the prescribed seed data α¯(1)0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}_{0} and ψ¯(1)0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\psi}}_{0}, Ψ¯(1)0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}}_{0} using (101), (102).

Next, we impose that on the boundary sphere Su0,v02S^{2}_{u_{0},v_{0}},

(r2Δ̸+2K)rurl(r3β¯(1))=2r3urliv2(r3β¯(1))=r2urlivψ¯(1)0,\displaystyle-(r^{2}\not{\Delta}+2K)r\not{curl}(r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}})=2r^{3}\not{curl}\not{div}\not{\mathcal{D}}_{2}^{\star}(r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}})=r^{2}\not{curl}\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\psi}}_{0}, (114)

which, since the =1\ell=1 mode of r4urlβ¯(1)r^{4}\not{curl}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}} is determined by the seed data on Su0,v02S^{2}_{u_{0},v_{0}}, determines r4urlβ¯(1)(u0,v0,θ)r^{4}\not{curl}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}(u_{0},v_{0},\theta) at infinity from seed data. Since the seed data require limuu0r4ivβ¯(1)(u,v0,θ)=B¯0\lim_{u\rightarrow u_{0}}r^{4}\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}(u,v_{0},\theta)=\underline{B}_{0}, the quantity r3β¯(1)(u0,v0,θ)r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}(u_{0},v_{0},\theta) is determined uniquely by elliptic theory. We now define the limit X¯:=limuu0rχ¯^\underline{X}:=\lim_{u\rightarrow u_{0}}r\underline{\hat{\chi}} by

6MkX¯\displaystyle 6Mk\underline{X} =ψ¯(1)0limuu02kr42β¯(1),\displaystyle=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\psi}}_{0}-\lim_{u\to u_{0}}2kr^{4}\not{\mathcal{D}}^{\star}_{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}, (115)

consistently with (101). We can now obtain β¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}} along all of C¯v0\underline{C}_{v_{0}} by integrating (75) written as

Ω∇̸3(r4Ω1β¯(1))=Ω2r4ivα¯(1)Ω2.\displaystyle\Omega\not{\nabla}_{3}(r^{4}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}})=\Omega^{2}r^{4}\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}\Omega^{-2}\,.

Note that the right hand side is integrable in uu on C¯v0\underline{C}_{v_{0}} and thus determines r4Ω1β¯(1)r^{4}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}. Similarly, we can integrate (57) from infinity with boundary condition (115) to obtain r2Ω1χ^¯(1)r^{2}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}} along C¯v0\underline{C}_{v_{0}}. This comes with the (smooth) expansion

r2Ω1χ^¯(1)=X¯+𝒪(r2).\displaystyle r^{2}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}=\underline{X}+\mathcal{O}(r^{-2})\,. (116)

Note that we can now determine r3σ(1)r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}} from (73) and the boundary condition r3σ(1)(u0,v0,θ)=0r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}(u_{0},v_{0},\theta)=0.

From (56) we have

u((Ω2(Ωtrχ¯)(1)r2)=4rω¯(1).\displaystyle\partial_{u}((\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}r^{2})=-4r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}\,.

Since rω¯(1)r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}} extends smoothly to Su0,v02S^{2}_{u_{0},v_{0}} the right hand side is integrable. Using the boundary condition (106) this defines (Ωtrχ¯)(1)Ω2r2\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}\Omega^{-2}r^{2} on all of C¯v0\underline{C}_{v_{0}}. Using (71), we can also determine ρ(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho} from R0R_{0} at infinity (set by seed data) and the bounds on (Ωtrχ¯)(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)} and β¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}. We now write the Codazzi equation (60) as

(r2η(1))+12r∇̸((Ωtrχ¯)(1)r2Ω2)=r3ivΩ1χ^¯(1)r3Ω1β¯(1).\displaystyle\left(r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}\right)+\frac{1}{2}r\not{\nabla}\Big{(}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}r^{2}\Omega^{-2}\Big{)}=r^{3}\not{div}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}-r^{3}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}\,. (117)

All four terms have a finite limit on null infinity (r3Ω1β¯(1)r^{3}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}} actually vanishes) which therefore determines r2η(1)r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}} (which extends regularly to u=u0u=u_{0}). We also set, consistent with (52),

η¯(1)=η(1)+2∇̸(Ω1Ω(1)0).\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}=-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}+2\not{\nabla}(\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}_{0})\,. (118)

One easily shows that the 33-equation for η¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}, (61), holds as a consequence of (117) and the 33-equations holding for χ^¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}, (Ωtrχ¯)(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}, Ω1Ω(1)\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega} and β(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}. Moreover, using (61) and the definition of σ(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}} by (73), one can show that Ω∇̸3(r3urlη¯(1)+r3σ(1))=0\Omega\not{\nabla}_{3}\left(r^{3}\not{curl}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}+r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}\right)=0. Now, both quantities in the parenthesis vanish at infinity, which shows that the elliptic equations (62) hold on C¯v0\underline{C}_{v_{0}}. Indeed, r3σ(1)0r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}\to 0 by definition of σ(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}} and, by (114) and (115), one has limuu0r3urlivχ^¯(1)=0\lim_{u\rightarrow u_{0}}r^{3}\not{curl}\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}=0, which, by the fact that r2β¯(1)=0r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}=0 as uu0u\rightarrow u_{0} and (118), shows that r3urlη¯(1)=r3urlη(1)=0r^{3}\not{curl}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}=r^{3}\not{curl}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}=0 as uu0u\rightarrow u_{0}.

We next determine r2Ωβ(1)r^{2}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}} from (69) and the previously defined quantities, using the boundary condition at infinity r3β(1)=r3β¯(1)r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}=-r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}. To determine χ^(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}} we write (58) as

Ω∇̸3(rΩχ^(1)rX¯)\displaystyle\Omega\not{\nabla}_{3}\left(r\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}-r\underline{X}\right) =2Ω2r2η(1)Ω2(Ωχ^¯(1)X¯).\displaystyle=-2\Omega^{2}r\not{\mathcal{D}}_{2}^{\star}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}-\Omega^{2}\left(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}-\underline{X}\right)\,. (119)

Using (116), the right hand side is integrable and we define Ωχ^(1)\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}} by imposing the condition rΩχ^(1)rX¯0r\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}-r\underline{X}\to 0 when uu0u\to u_{0}.

We can now determine the last curvature component α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} from (67) and the fact that r3α(1)=r3α¯(1)r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}=r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}} at infinity. Note that with this all Bianchi equations in the 33-directions hold by construction.

The definition of ω(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega} follows from (64) and the condition that ω(1)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}=0 on Su0,v02S^{2}_{u_{0},v_{0}}.

Consistently with (102) and the linearised formula for the Gauss curvature (66), we define the asymptotic linearised Gauss curvature K(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{K} to be

limuu0r3∇̸K(1):=13M(rivΨ¯(1)0+2r3iv2∇̸R0)=limuu0(r3ivχ^¯(1)+r3ivχ^(1)),\displaystyle\lim_{u\rightarrow u_{0}}r^{3}\not{\nabla}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{K}:=\frac{1}{3M}\left(r\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}}_{0}+2r^{3}\not{div}\not{\mathcal{D}}_{2}^{\star}\not{\nabla}R_{0}\right)=\lim_{u\rightarrow u_{0}}\left(-r^{3}\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}+r^{3}\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\right)\,, (120)

where the second identity comes from the previous definitions of χ^¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}} and χ^(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}. This determines (K(1))2(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{K})_{\ell\geq 2}. Now recall that we have ((1))=0(u0,v0)=13M(R0)=013𝔪(\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}})_{\ell=0}(u_{0},v_{0})=\frac{1}{3M}(R_{0})_{\ell=0}-\frac{1}{3}\mathfrak{m} and ((1))=1(u0,v0)=0(\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}})_{\ell=1}(u_{0},v_{0})=0 by construction, which in particular determines (K(1))=0,1(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{K})_{\ell=0,1} from formula (66). Conversely, formula (66) determines ((1))2(u0,v0)(\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}})_{\ell\geq 2}(u_{0},v_{0}) when read as an elliptic equation for this quantity, since ^(1)(u0,v0,θ)=G^0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}(u_{0},v_{0},\theta)=\hat{G}_{0} is prescribed at infinity. Finally, (49) and (50) determine determine ((1))(\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}) and ^(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}} on all of C¯v0\underline{C}_{v_{0}}.

We next use the linearised Gauss equation (65) to determine (Ωtrχ)(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)} on C¯v0\underline{C}_{v_{0}}. Note that this shows in particular that (Ωtrχ)(1)(Ωtrχ¯)(1)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}=0 at (u0,v0)(u_{0},v_{0}). Differentiating the linearised Gauss equation in ∇̸3\not{\nabla}_{3} we verify that (54) has to hold on C¯v0\underline{C}_{v_{0}}. To show validity of the other Codazzi equations of (60), which we rewrite as

rivΩχ^(1)=Ω2η¯(1)rΩβ(1)+12r∇̸(Ωtrχ)(1),\displaystyle r\not{div}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}=-\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}-r\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}+\frac{1}{2}r\not{\nabla}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\,, (121)

on C¯v0\underline{C}_{v_{0}}, we first use the boundary asymptotics on β(1),β¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}} and (Ωtrχ)(1),(Ωtrχ¯)(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}, the definition of η¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}} and the asymptotics of Ω1Ω(1)0\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}_{0}, and the validity of the underlined Codazzi equation (117), to verify that (121) holds on Su0,v02S^{2}_{u_{0},v_{0}}, and then differentiate in ∇̸3\not{\nabla}_{3}: Since the resulting expression vanishes after inserting the evolution equations already established, we deduce that (121) indeed holds on C¯v0\underline{C}_{v_{0}}.

We have thus determined all quantities from seed data with quantitative estimates and obtained validity of all evolution equations in the 33-direction as well as elliptic relations on spheres contained in (49)–(76). Obviously, tangential derivatives to C¯v0\underline{C}_{v_{0}} 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 ω(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega} and α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}) satisfies a Bianchi or null structure equation which determines its Ω∇̸4\Omega\not{\nabla}_{4} 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 α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} we can determine transversal derivatives from the Teukolsky equation (90) and the boundary conditions relating α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} and the seed data α¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}} at Su0,v02S^{2}_{u_{0},v_{0}}. For ω(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega} we have can obtain the Ω∇̸4\Omega\not{\nabla}_{4}-derivative from the fact that we can determine the ∇̸T\not{\nabla}_{T}-derivative from the boundary condition in Item 22 and the commuted equation for ω(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}, 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

(α¯(1)0;ψ¯(1)0,Ψ¯(1)0),(B¯0,R0;H¯0;Ω1Ω(1)0,b(1)0,G0=1,G^0),(𝔪,𝔞1,𝔞0,𝔞1)(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}_{0};\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\psi}}_{0},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}}_{0})\ \ ,\ \ (\underline{B}_{0},R_{0};\underline{H}_{0};\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}_{0},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{0},G^{\ell=1}_{0},\hat{G}_{0})\ \ ,\ \ ({\mathfrak{m}},{\mathfrak{a}}_{-1},{\mathfrak{a}}_{0},{\mathfrak{a}}_{1})

and assume there exists a solution 𝒮\mathscr{S} 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 𝒮\mathscr{S} vanishes (which in particular happens if the solution is supported for 1\ell\leq 1), the solution 𝒮\mathscr{S} is the sum of a pure gauge and a linearised Kerr-AdS solution.

Proof.

We add a pure gauge and linearised Kerr-AdS solution 𝒢𝒻+𝒢q𝒦𝔪,𝔞\mathscr{G_{f}}+\mathscr{G}_{q}-\mathscr{K}_{\mathfrak{m},\vec{\mathfrak{a}}} as in Proposition 3.5 to the solution 𝒮\mathscr{S} 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 𝒮\mathscr{S} and 𝒮=𝒮+𝒢𝒻+𝒢q𝒦𝔪,𝔞\mathscr{S}^{\prime}=\mathscr{S}+\mathscr{G_{f}}+\mathscr{G}_{q}-\mathscr{K}_{\mathfrak{m},\vec{\mathfrak{a}}} on C¯v0\underline{C}_{v_{0}} satisfy (1)(4)(1)-(4)) we conclude that all geometric quantities of 𝒮\mathscr{S}^{\prime} vanish on C¯v0\underline{C}_{v_{0}}. In particular, Lemma B.1 applies and the solution 𝒮\mathscr{S}^{\prime} must have vanishing =0\ell=0 mode. Moreover, since α¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}} and α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} vanish identically on C¯v0\underline{C}_{v_{0}}, it follows from the decoupled Teukolsky equations (90)–(91) and uniqueness of its solutions (see for instance Theorem 1.4 of [GH23] that α¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}} and α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} vanish identically globally in \mathcal{M}. Since χ^(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}} and β(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}} vanish on C¯v0\underline{C}_{v_{0}}, they vanish globally by their evolution equation in the 44-direction. Inserting the boundary conditions we conclude vanishing of χ^¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}} and β¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}} on \mathcal{I} and by their evolution equation in the 33-direction, globally. Similarly, σ(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}} vanishes by the equation (72) and the vanishing on C¯v0\underline{C}_{v_{0}}. From (69) and (74) and the vanishing of σ(1),β(1),β¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}} we now conclude η(1)η¯(1)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}=0. The equation (61) and the vanishing of η(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}} on C¯v0\underline{C}_{v_{0}} produces global vanishing of η(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}} and hence of η¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}} individually. Revisiting (69) it follows that ρ(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho} vanishes (recall we have established in the beginning that the =0\ell=0 mode vanishes). Codazzi shows the global vanishing of (Ωtrχ)(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)} and (Ωtrχ¯)(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)} and (63), (64) that of ω(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega} and ω¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}} 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.

Given a smooth seed initial data set

(α¯(1)0;ψ¯(1)0,Ψ¯(1)0),(Ω1Ω(1)0,b(1)0;B¯0,H¯0,R0,G0=1,G^0),(𝔪,𝔞1,𝔞0,𝔞1),(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}_{0};\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\psi}}_{0},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}}_{0})\ \ ,\ \ (\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}_{0},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{0};\underline{B}_{0},\underline{H}_{0},R_{0},G^{\ell=1}_{0},\hat{G}_{0})\ \ ,\ \ (\mathfrak{m},\mathfrak{a}_{-1},\mathfrak{a}_{0},\mathfrak{a}_{1})\,,

there exists a unique solution 𝒮\mathscr{S} of the system of gravitational perturbations on \mathcal{M} satisfying the boundary conditions (Definition 3.1), realising the given seed data in the sense of Definition 3.4 and satisfying that [∇̸T]p[r∇̸]qω(1)=0[\not{\nabla}_{T}]^{p}[r\not{\nabla}]^{q}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}=0 holds on the sphere Su0,v02S^{2}_{u_{0},v_{0}} for any p,q0p,q\in\mathbb{N}^{0}.

Remark 3.10.

With more work one can show that the solution constructed in the above theorem is aAdS in the linearised sense according to Definition 2.6. In particular, we will establish uniform bounds on all quantities in (82) and (83) as part of our main theorem. See also Remarks 2.7 and 3.2 above.

Proof.
131313The authors would like to thank Leonhard Kehrberger for discussions and suggesting the argument with σ\sigma^{\prime} 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 1\ell\leq 1 agrees with the seed data induced by an appropriate pure gauge solution supported on =1\ell=1 plus the seed data of a linearised Kerr-AdS solution which establishes existence for 1\ell\leq 1.

In summary, we only need to construct the solution for 2\ell\geq 2 from seed data supported for 2\ell\geq 2.

Step 1: Constructing geometric quantities in \mathcal{M}, Part I. We first construct from the seed data set all quantities on C¯v0\underline{C}_{v_{0}} as in the proof of Proposition 3.6. This in particular determines smooth α¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}, α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} on C¯v0\underline{C}_{v_{0}} such that in particular the master energy defined in (132) is finite for all nn. We can hence apply the well-posedness theorem for the Teukolsky system (90)–(91) (cf. Theorem 1.4 in [GH23]) and obtain smooth α¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}} and α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} globally on \mathcal{M}. We next determine χ^(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}} and β(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}} globally from their transport equation in the 44-direction, i.e. integrating (57) and (68) from data. We then determine χ^¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}} and β¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}} globally by integrating the transport equations (57) and (75) from the boundary (using the boundary conditions as initial conditions for χ^¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}} and β¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}, i.e. r3β¯(1)=r3β(1)r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}=-r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}} and rχ^(1)=rχ^¯(1)r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}=r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}} on \mathcal{I}). With this, (68) holds by definition and it is easy to see that (67) also holds because it holds on data and Ω∇̸4\Omega\not{\nabla}_{4} of this equation vanishes by the fact that the Teukolsky equation (90) holds for α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} and the equations (57), (68) hold in the 44-direction by construction. Similarly (76) holds on the boundary (by the fact that (67) holds there and the boundary conditions imposed) and its Ω∇̸3\Omega\not{\nabla}_{3}-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 σ\sigma by integrating (72) from initial data. Note that with this definition, r3σ(1)r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}} has a finite limit on the boundary (which we do not know vanishing of yet). We then set urlη(1):=σ(1)\not{curl}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}:=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}} and urlη¯(1):=σ(1)\not{curl}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}:=-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}} so that (62) is satisfied. On the other hand, we determine ivη(1)\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}} and ivη¯(1)\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}} globally from (58), (59) by taking the iviv\not{div}\not{div} of these equations and observing that iviv2ξ=(12Δ̸K)ivξ\not{div}\not{div}\not{\mathcal{D}}_{2}^{\star}\xi=\left(-\frac{1}{2}\not{\Delta}-K\right)\not{div}\xi with (Δ̸+2K)\left(\not{\Delta}+2K\right) having trivial kernel on the space of functions with 2\ell\geq 2. This determines η(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}} and η¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}} uniquely by standard elliptic theory and one can also show that r(η(1)+η¯(1))=0r(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}})=0 holds on the conformal boundary \mathcal{I}.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 rr.

Step 2. Verifying the url\not{curl}-equations. With the quantities defined we can already verify some of the equations. We claim that the url\not{curl} applied to the Codazzi equation (60) for χ^\hat{\chi} holds. To see this, note that

Ω∇̸4(r2urlivΩ1r2χ^(1)+1r(rurl)Ω1r4β(1)r3σ(1))=0.\Omega\not{\nabla}_{4}\left(r^{2}\not{curl}\not{div}\Omega^{-1}r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}+\frac{1}{r}(r\not{curl})\Omega^{-1}r^{4}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}-r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}\right)=0\,.

by the propagation equations we have defined in the 44-direction. Since the quantity in round brackets also vanishes on C¯v0\underline{C}_{v_{0}} we conclude that it is zero globally and hence url\not{curl} applied to the Codazzi equation (60) for χ^\hat{\chi} holds. We similarly conclude that the url\not{curl} applied to (69) holds: Indeed, defining

B:=Ω∇̸3(Ω2r2rurlr4β(1)Ω)+r3Ω2Δ̸σ(1)3ρr3Ω2σ(1)B:=\Omega\not{\nabla}_{3}\Big{(}\frac{\Omega^{2}}{r^{2}}r\not{curl}\frac{r^{4}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}}{\Omega}\Big{)}+r^{3}\Omega^{2}\not{\Delta}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}-3\rho r^{3}\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}

we derive (using the transport equations in the 44-direction for β(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}} and σ(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}} as well as (67) and the just established url\not{curl} applied to the Codazzi equation (60) for χ^\hat{\chi}) an equation of the form Ω∇̸4B=(2r6Mr2)B\Omega\not{\nabla}_{4}B=-\left(\frac{2}{r}-\frac{6M}{r^{2}}\right)B. Since B=0B=0 on C¯v0\underline{C}_{v_{0}} we conclude that B=0B=0 globally.

We next define an auxiliary quantity σ(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}^{\prime} such that limuvr3σ(1)=limuvr3σ(1)\lim_{u\to v}r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}^{\prime}=\lim_{u\to v}r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}} and

Ω∇̸3(r3σ(1))=urlr3β¯(1)Ω.\displaystyle\Omega\not{\nabla}_{3}(r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}^{\prime})=-\not{curl}r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}\Omega\,. (122)

Note that σ(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}^{\prime} agrees with σ(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}} on C¯v0\underline{C}_{v_{0}} and on \mathcal{I}. We verify that

Ω∇̸3(r2urlivΩ1χ¯^r21r(rurl)Ω1r4β¯r3σ(1))=0,\Omega\not{\nabla}_{3}\Big{(}r^{2}\not{curl}\not{div}\Omega^{-1}\hat{\underline{\chi}}r^{2}-\frac{1}{r}(r\not{curl})\Omega^{-1}r^{4}\underline{\beta}-r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}^{\prime}\Big{)}=0\,,

and since the quantity vanishes on \mathcal{I} (by the previously imposed/derived boundary relations and the Codazzi equation for χ^(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}), the expression in round brackets vanishes globally. Defining

B¯:=Ω∇̸4(Ω2r2rurlr4β¯(1)Ω)+r3Ω2Δ̸σ(1)3ρr3Ω2σ(1),\displaystyle\underline{B}:=\Omega\not{\nabla}_{4}\left(\frac{\Omega^{2}}{r^{2}}r\not{curl}\frac{r^{4}\underline{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}}}{\Omega}\right)+r^{3}\Omega^{2}\not{\Delta}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}^{\prime}-3\rho r^{3}\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}^{\prime}\,, (123)

we first check that Ω∇̸3(r2Ω2B¯)=0\Omega\not{\nabla}_{3}\left(\frac{r^{2}}{\Omega^{2}}\underline{B}\right)=0 holds (a computation similar to the one for BB above) and then verify that B¯\underline{B} vanishes at infinity (which follows from T(r4urlβ(1)+r4urlβ¯(1))=0T(r^{4}\not{curl}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}+r^{4}\not{curl}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}})=0 which is in turn a consequence of how we defined β¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}). The two observations allow us to conclude B¯=0\underline{B}=0 globally. We can finally conclude that σ(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}^{\prime} satisfies the Regge-Wheeler equation (99) by applying Ω∇̸4\Omega\not{\nabla}_{4} to (122) and inserting (123). In summary, σ(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}} and σ(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}^{\prime} satisfy the same wave equation with the same data and boundary condition at \mathcal{I} and hence agree globally, i.e. σ(1)=σ(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}^{\prime} on \mathcal{M}. In particular, Bianchi equation (73) holds by (122) and, adding the two Bianchi equations (72), (73), one has that limuvT(r3σ(1))=0\lim_{u\to v}T(r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}})=0. Thus, using that on the initial sphere at infinity Su0,v0S_{u_{0},v_{0}} r3σ(1)r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}} vanishes, one has that r3σ(1)r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}} vanishes globally on \mathcal{I}. At this point we have ensured the validity of the Bianchi equations (67), (68), (72), (73), (75), (76) as well as the url\not{curl} of (69) and – by the vanishing of (123) and σ(1)=σ(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}^{\prime}=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}} – (74) respectively. By construction also the equations (62) and the url\not{curl} of (both of) (60) hold. One easily checks that the url\not{curl} of (61) holds. Moreover, we check that urliv\not{curl}\not{div} applied to (58) and (59) respectively hold (inserting the url\not{curl} of (60) and using the propagation equations already established). This means that (58) and (59) hold unconditionally since the iviv\not{div}\not{div} part of these holds by construction.

Step 3: Constructing geometric quantities in \mathcal{M}, Part II. We now set 2∇̸(Ω1Ω(1)):=η(1)+η¯(1)2\not{\nabla}(\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}):=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}, which is well defined as url(η(1)+η¯(1))=0\not{curl}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}})=0 by our definition of η(1),η¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}} above. Moreover, we set ω(1)=Ω∇̸4(Ω1Ω(1))\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}=\Omega\not{\nabla}_{4}(\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}) and ω¯(1)=Ω∇̸3(Ω1Ω(1))\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}=\Omega\not{\nabla}_{3}(\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}). Next we define (Ωtrχ)(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)} by (60) (which is well defined as we have already verified that the url\not{curl} of this equation holds) and (Ωtrχ¯)(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)} directly integrating the evolution equation (53) from data. We define the quantity ρ\rho by integrating (70). Finally, the metric quantities (1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}, ^(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}, b(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b} are defined by integrating their equation in the 44-direction from data, i.e. (49), (50) and (51).

Step 4. Verifying the remaining equations. Differentiating (58) with respect to Ω∇̸4\Omega\not{\nabla}_{4} and (59) with respect to Ω∇̸3\Omega\not{\nabla}_{3} shows that (61) must also both hold globally (recall we are on 2\ell\geq 2). From our definition of ω(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega} and ω¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}} also the equations

Ω∇̸4(r2η¯(1))=2r2∇̸ω(1)+r2Ωβ(1),Ω∇̸3(r2η(1))=2r2∇̸ω¯(1)r2Ωβ¯(1)\displaystyle\Omega\not{\nabla}_{4}\left(r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}\right)=2r^{2}\not{\nabla}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}+r^{2}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}\ \ \ ,\ \ \ \Omega\not{\nabla}_{3}\left(r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}\right)=2r^{2}\not{\nabla}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}-r^{2}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}} (124)

hold. Using (61) we can also conclude that the iv\not{div} of (69) (and hence (69) unconditionally since we already verified the url\not{curl}-equation) must hold globally because it holds on data and is propagated in the 44-direction.

Differentiating the Codazzi equation for χ^(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}} with Ω∇̸3\Omega\not{\nabla}_{3} and Ω∇̸4\Omega\not{\nabla}_{4} shows that also (54) and (55) hold. Differentiating (55) with respect to Ω∇̸3\Omega\not{\nabla}_{3} shows that (64) holds and (63) follows from the fact that Ω4∇̸ω¯(1)=Ω∇̸3ω(1)\Omega_{4}\not{\nabla}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}=\Omega\not{\nabla}_{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega} by the way we defined ω(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}, ω¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}. One now verifies that

D:=Ω∇̸3(r3ρ+r3ivη(1))+2r2ΩΔ̸ω¯(1)3M(Ωtrχ¯)(1)=0\displaystyle D:=\Omega\not{\nabla}_{3}(r^{3}\rho+r^{3}\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}})+2r^{2}\Omega\not{\Delta}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}-3M\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}=0 (125)

holds on the cone C¯v0\underline{C}_{v_{0}} and Ω∇̸4D=0\Omega\not{\nabla}_{4}D=0, which implies D=0D=0 in \mathcal{M}, which in turn implies that (71) holds. We next verify that (74) holds by noting that it holds on \mathcal{I} by the boundary conditions and is propagated in the 33-direction. Indeed, we have

Ω∇̸4(r2Ωβ¯(1))=Ω∇̸3(r2Ωβ(1))\Omega\not{\nabla}_{4}(r^{2}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}})=-\Omega\not{\nabla}_{3}(r^{2}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}})

on \mathcal{I} after replacing Ω∇̸4=2∇̸TΩ∇̸3\Omega\not{\nabla}_{4}=2\not{\nabla}_{T}-\Omega\not{\nabla}_{3} and inserting the Bianchi equations (68), (75) as well as the boundary conditions for α(1),α¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}} and β(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}, β¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}. Since also rη(1)=rη¯(1)r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}=-r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}} and r3σ(1)=0r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}=0 hold on \mathcal{I} we see that (74) is equivalent to (69) on \mathcal{I}. Differentiating now E:=Ω∇̸4(r2Ωβ¯(1))r2Ω21(ρ(1),σ(1))3ρr2Ω2η¯(1)E:=\Omega\not{\nabla}_{4}(r^{2}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}})-r^{2}\Omega^{2}\not{\mathcal{D}}_{1}^{\star}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}})-3\rho r^{2}\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}} in Ω∇̸3\Omega\not{\nabla}_{3} one obtains Ω∇̸3(r2Ω2E)=0\Omega\not{\nabla}_{3}\left(\frac{r^{2}}{\Omega^{2}}E\right)=0 after inserting the equations that have already shown to hold. As E=0E=0 on \mathcal{I} we conclude E=0E=0 globally.

Now the χ^¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}-Codazzi equation (60) and the equation (56) can be verified by noting that they hold on C¯v0\underline{C}_{v_{0}} and are propagated in the 44-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 33-equations for the metric components. But all these equations hold on C¯v0\underline{C}_{v_{0}} and applying Ω∇̸4\Omega\not{\nabla}_{4} and inserting the already established equations one verifies they propagate to hold in all of \mathcal{M}. This finishes the proof of the proposition. ∎

3.5 The initial data normalisation

Consider a given seed data with associated solution 𝒮\mathscr{S} from Theorem 3.9. The main objective of this section is to construct a pure gauge solution 𝒢\mathscr{G} from the given seed data which when added to 𝒮\mathscr{S} achieves a certain normalisation of the solution 𝒮+𝒢\mathscr{S}+\mathscr{G} 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

(α¯(1)0;ψ¯(1)0,Ψ¯(1)0),(B¯0,R0;H¯0;Ω1Ω(1)0,b(1)0,G0=1,G^0),(𝔪,𝔞1,𝔞0,𝔞1),(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}_{0};\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\psi}}_{0},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\Psi}}_{0})\ \ ,\ \ (\underline{B}_{0},R_{0};\underline{H}_{0};\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}_{0},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{0},G^{\ell=1}_{0},\hat{G}_{0})\ \ ,\ \ (\mathfrak{m},\mathfrak{a}_{-1},\mathfrak{a}_{0},\mathfrak{a}_{1}),

as in Definition 3.3 and let 𝒮\mathscr{S} be the unique solution of the system of gravitational perturbations arising from Theorem 3.9. We say that 𝒮\mathscr{S} is initial data normalised if the following holds for 𝒮\mathscr{S} on the ingoing initial cone C¯v0=0\underline{C}_{v_{0}=0}:

  • b(1)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}=0 and Ω1Ω(1)=0\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}=0 on v=v0=0v=v_{0}=0,

  • G^0\hat{G}_{0} and G0=1=0G_{0}^{\ell=1}=0 and H¯0=0\underline{H}_{0}=0,

  • ivη(1)+ρ(1)=0\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}=0 and (Ωtrχ)(1)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}=0 on the horizon sphere S,v02S^{2}_{\infty,v_{0}}.

Moreover, we call the solution initial data normalised with vanishing =0,1\ell=0,1 modes if in addition the =0,1\ell=0,1 modes of all geometric quantities of 𝒮\mathscr{S} vanish.

The point is that we can always achieve the initial data normalisation:

Proposition 3.12.

Given a solution 𝒮\mathscr{S} arising from a smooth seed initial data set as in Theorem 3.9 there exists a pure gauge solution 𝒢f+𝒢q\mathscr{G}_{f}+\mathscr{G}_{q} (computable in terms of the seed data) and a linearised Kerr-AdS solution 𝒦𝔪,𝔞\mathscr{K}_{\mathfrak{m},\vec{\mathfrak{a}}} such that 𝒮+𝒢f+𝒢q𝒦𝔪,𝔞\mathscr{S}+\mathscr{G}_{f}+\mathscr{G}_{q}-\mathscr{K}_{\mathfrak{m},\vec{\mathfrak{a}}} is an initial data normalised solution with vanishing =0,1\ell=0,1 modes.

Proof.

The proof is a small variation of the proof of Proposition 3.5. From Proposition 3.5 there exist 𝒢f~+𝒢q\mathscr{G}_{\tilde{f}}+\mathscr{G}_{q} and a linearised Kerr-AdS solution 𝒦𝔪,𝔞\mathscr{K}_{\mathfrak{m},\vec{\mathfrak{a}}} such that 𝒮:=𝒮+𝒢f~+𝒢q~𝒦𝔪,𝔞\mathscr{S}^{\prime}:=\mathscr{S}+\mathscr{G}_{\tilde{f}}+\mathscr{G}_{\tilde{q}}-\mathscr{K}_{\mathfrak{m},\vec{\mathfrak{a}}} has trivial seed data and in particular vanishing =0,1\ell=0,1 modes. We now add to this another 𝒢f^\mathscr{G}_{\hat{f}} (supported for 2\ell\geq 2) generated by

f^(u,θ)=λ^1r(u,v0)+λ^0Ω2(u,v0)+λ^2Ω2(u,v0)\hat{f}(u,\theta)=\frac{\hat{\lambda}_{1}r(u,v_{0})+\hat{\lambda}_{0}\Omega^{2}(u,v_{0})+\hat{\lambda}_{2}}{\Omega^{2}(u,v_{0})}

(inducing f^u(u,θ)=f^(u,θ)\hat{f}_{u}(u,\theta)=\hat{f}(u,\theta) and f^v(v,θ)=f^(v,θ)\hat{f}_{v}(v,\theta)=\hat{f}(v,\theta) in Lemma 2.8), where the λ^i\hat{\lambda}_{i} are functions on the unit sphere satisfying the relation Δ𝕊2λ^0λ^2=0\Delta_{\mathbb{S}^{2}}\hat{\lambda}_{0}-\hat{\lambda}_{2}=0. It is clear from the proof of Proposition 3.5 that 𝒮+𝒢f^\mathscr{S}^{\prime}+\mathscr{G}_{\hat{f}} still satisfies Ω1Ω(1)=0\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}=0 on v=v0v=v_{0} and H¯=0\underline{H}=0 on Su0,v02S^{2}_{u_{0},v_{0}} (hence by (56) on all of C¯v0\underline{C}_{v_{0}}). In view of

Ω2f^u(,θ)=λ^1r++λ^2,1Ω2ru(Ω2f^ur)(,θ)=λ^01r+[k24Mr3+1r2]\Omega^{2}\hat{f}_{u}(\infty,\theta)=\hat{\lambda}_{1}r_{+}+\hat{\lambda}_{2}\ \ ,\ \ \frac{1}{\Omega^{2}r}\partial_{u}\left(\frac{\Omega^{2}\hat{f}_{u}}{r}\right)(\infty,\theta)=\hat{\lambda}_{0}\frac{1}{r_{+}}\left[-k^{2}-\frac{4M}{r^{3}}+\frac{1}{r^{2}}\right]

and fv(v0,v0)=λ^0f_{v}(v_{0},v_{0})=\hat{\lambda}_{0}, we now compute from Lemma 2.8 the horizon sphere relations

(Ωtrχ)(1)|𝒮+𝒢f^(,v0,θ)=(Ωtrχ)(1)|𝒮(,v0,θ)+2r+2[(14Mr+k2r+2)(λ^1r++λ^2)+ΔS2(λ^1r++λ^2)],\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\big{|}_{\mathscr{S}^{\prime}+\mathscr{G}_{\hat{f}}}(\infty,v_{0},\theta)=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\big{|}_{\mathscr{S}^{\prime}}(\infty,v_{0},\theta)+\frac{2}{r_{+}^{2}}\left[\left(1-\frac{4M}{r_{+}}-k^{2}r_{+}^{2}\right)(\hat{\lambda}_{1}r_{+}+\hat{\lambda}_{2})+{\Delta}_{S^{2}}(\hat{\lambda}_{1}r_{+}+\hat{\lambda}_{2})\right]\,,
(ivη(1)+ρ(1))|𝒮+𝒢f^(,v0,θ)=(ivη(1)+ρ(1))|𝒮(,v0,θ)6Mr+4(λ^1r++λ^2)+(ΔS2λ^0)1r+[k24Mr+3+1r+2].\displaystyle(\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho})\big{|}_{\mathscr{S}^{\prime}+\mathscr{G}_{\hat{f}}}(\infty,v_{0},\theta)=(\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho})\big{|}_{\mathscr{S}^{\prime}}(\infty,v_{0},\theta)-\frac{6M}{r_{+}^{4}}(\hat{\lambda}_{1}r_{+}+\hat{\lambda}_{2})+(\Delta_{S^{2}}\hat{\lambda}_{0})\frac{1}{r_{+}}\left[-k^{2}-\frac{4M}{r_{+}^{3}}+\frac{1}{r_{+}^{2}}\right]\,.

Recalling that we can eliminate λ^0\hat{\lambda}_{0} from the relation ΔS2λ^0λ^2=0\Delta_{S^{2}}\hat{\lambda}_{0}-\hat{\lambda}_{2}=0 it is an algebraic exercise to determine λ^1\hat{\lambda}_{1} and λ^2\hat{\lambda}_{2} such that both (ivη(1)+ρ(1))|𝒮+𝒢f^(,v0,θ)=0(\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho})\big{|}_{\mathscr{S}^{\prime}+\mathscr{G}_{\hat{f}}}(\infty,v_{0},\theta)=0 and (Ωtrχ)(1)|𝒮+𝒢f^(,v0,θ)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\big{|}_{\mathscr{S}^{\prime}+\mathscr{G}_{\hat{f}}}(\infty,v_{0},\theta)=0 hold. While 𝒢f^\mathscr{G}_{\hat{f}} might have altered G^0{\hat{G}}_{0} we can simply repeat Step 3 of Proposition 3.5 and add a 𝒢q^\mathscr{G}_{\hat{q}} which ensures that 𝒮+𝒢f^+𝒢q^\mathscr{S}^{\prime}+\mathscr{G}_{\hat{f}}+\mathscr{G}_{\hat{q}} satisfies all of the desired properties. Setting 𝒢f:=𝒢f~+𝒢f^\mathscr{G}_{f}:=\mathscr{G}_{\tilde{f}}+\mathscr{G}_{\hat{f}} and 𝒢q:=𝒢q~+𝒢q^\mathscr{G}_{q}:=\mathscr{G}_{\tilde{q}}+\mathscr{G}_{\hat{q}} 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 α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} and α¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}} and a local orthonormal frame e1,e2e_{1},e_{2} on the sphere we define the complex scalars

α[2]=2r4Ω2(α¯(1)(e1,e1)iα¯(1)(e1,e2)),α[+2]=2Ω2(α(1)(e1,e1)+iα(1)(e1,e2)),\displaystyle\alpha^{[-2]}=2r^{4}\Omega^{2}\left(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}(e_{1},e_{1})-i\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}(e_{1},e_{2})\right)\ \ \ ,\ \ \ {\alpha}^{[+2]}=2\Omega^{-2}\left(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}(e_{1},e_{1}){+}i\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}(e_{1},e_{2})\right)\,, (126)

which transform like spin-weighted functions of weight ±2\pm 2 under a change of orthonormal frame on S2S^{2}. For the specific frame e1=1rθe_{1}=\frac{1}{r}\partial_{\theta} and e2=1rsinθφe_{2}=\frac{1}{r\sin\theta}\partial_{\varphi} one obtains the Teukolsky equation for α[±2]\alpha^{[\pm 2]} as stated in [GH24] by expressing the equations (90) and (91) in frame components. We also note in the notation of [GH24] the relations

α~[2]=2rΩ2(α¯(1)(e1,e1)iα¯(1)(e1,e2)),α~[+2]=2rΩ2(α(1)(e1,e1)+iα(1)(e1,e2)).\displaystyle\widetilde{\alpha}^{[-2]}=2r\Omega^{2}\left(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}(e_{1},e_{1})-i\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}(e_{1},e_{2})\right)\ \ \ ,\ \ \ \widetilde{\alpha}^{[+2]}=2r\Omega^{2}\left(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}(e_{1},e_{1}){+}i\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}(e_{1},e_{2})\right)\,. (127)

Clearly the estimates on α~[±2]\widetilde{\alpha}^{[\pm 2]} obtained in [GH24] directly translate into estimates for the (norms of the) tensors rΩ2α(1)r\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} and rΩ2α¯(1)r\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}.

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:

j=0n|𝔇jξ|2=j=0n|i|j|Ti1[r2Ω2Ω∇̸3]i2[Ω∇̸4]i3[r∇̸]i4ξ|2,\sum_{j=0}^{n}|\mathfrak{D}^{j}\xi|^{2}=\sum_{j=0}^{n}\sum_{|i|\leq j}\Big{|}T^{i_{1}}\left[\frac{r^{2}}{\Omega^{2}}\Omega\not{\nabla}_{3}\right]^{i_{2}}\left[\Omega\not{\nabla}_{4}\right]^{i_{3}}\left[r\not{\nabla}\right]^{i_{4}}\xi\Big{|}^{2}\,,

where the second sum is over all tuples i=(i1,i2,i3,i4)i=(i_{1},i_{2},i_{3},i_{4}) with i1+i2+i3+i4ji_{1}+i_{2}+i_{3}+i_{4}\leq j.

We first define the non-degenerate (near the horizon) outgoing and ingoing commuted energy fluxes (note that the superscript n+1n+1 denotes the number of derivatives involved) of a general Su,v2S^{2}_{u,v}-tensor ξ\xi:

𝔼¯un+1[ξ](v1,v2)\displaystyle\overline{\mathbb{E}}^{n+1}_{u}[\xi]\left(v_{1},v_{2}\right) =i=0nv1v2S2[|Ω∇̸4𝔇iξ|2+|r∇̸𝔇iξ|2](u,v¯)𝑑v¯sinθdθdϕ\displaystyle=\sum_{i=0}^{n}\int_{v_{1}}^{v_{2}}\int_{S^{2}}\left[|\Omega\not{\nabla}_{4}\mathfrak{D}^{i}\xi|^{2}+|r\not{\nabla}\mathfrak{D}^{i}\xi|^{2}\right]\left(u,\bar{v}\right)d\bar{v}\sin\theta d\theta d\phi
𝔼¯vn+1[ξ](u1,u2)\displaystyle\overline{\mathbb{E}}^{n+1}_{v}[\xi]\left(u_{1},u_{2}\right) =i=0nu1u2S2[|r2Ω1∇̸3𝔇iξ|2+|r∇̸𝔇iξ|2](u,v¯)Ω2𝑑u¯sinθdθdϕ\displaystyle=\sum_{i=0}^{n}\int_{u_{1}}^{u_{2}}\int_{S^{2}}\left[|r^{2}\Omega^{-1}\not{\nabla}_{3}\mathfrak{D}^{i}\xi|^{2}+|r\not{\nabla}\mathfrak{D}^{i}\xi|^{2}\right]\left(u,\bar{v}\right)\Omega^{2}d\bar{u}\sin\theta d\theta d\phi (128)

as well as the outgoing degenerate energy:

𝔼un+1[ξ](v1,v2)\displaystyle{\mathbb{E}}^{n+1}_{u}[\xi]\left(v_{1},v_{2}\right) =i=0nv1v2S2[|Ω∇̸4𝔇iξ|2+Ω2r2|r∇̸𝔇iξ|2](u,v¯)𝑑v¯sinθdθdϕ\displaystyle=\sum_{i=0}^{n}\int_{v_{1}}^{v_{2}}\int_{S^{2}}\left[|\Omega\not{\nabla}_{4}\mathfrak{D}^{i}\xi|^{2}+\frac{\Omega^{2}}{r^{2}}|r\not{\nabla}\mathfrak{D}^{i}\xi|^{2}\right]\left(u,\bar{v}\right)d\bar{v}\sin\theta d\theta d\phi (129)

The above energies will appear for the (regular both at the horizon +\mathcal{H}^{+} and the conformal boundary \mathcal{I}) quantities

ξ{rΩ2α(1),Ω2r5α¯(1),Ψ(1)R}.\displaystyle\xi\in\{r\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\Omega^{-2}r^{5}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{R}\}\,. (130)

We shall also need an auxiliary energy on spheres at the conformal boundary \mathcal{I}, which arises in the renormalised energy estimates of [GH24] and is defined only for n2n\geq 2:

𝔼¯n1[Ψ(1)R](v)=i=0n2S2[|t𝔇i12(2)12Ψ(1)R|2+|𝔇i(2)12Ψ(1)R|2](v,v)sinθdθdϕ.\displaystyle\overline{\mathbb{E}}^{n-1}_{\mathcal{I}}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{R}](v)=\sum_{i=0}^{n-2}\int_{S^{2}}\left[|\partial_{t}\mathfrak{D}^{i}\mathcal{L}^{-\frac{1}{2}}(\mathcal{L}-2)^{-\frac{1}{2}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{R}|^{2}+|\mathfrak{D}^{i}(\mathcal{L}-2)^{-\frac{1}{2}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{R}|^{2}\right](v,v)\sin\theta d\theta d\phi\,. (131)

We finally define (for n2n\geq 2) the following initial data master energy on cone C¯v0=0\underline{C}_{v_{0}=0}:

𝔼¯datan+1[α(1),α¯(1)]:=𝔼¯0n+1[Ω2rα(1)](0,)+𝔼¯0n+1[Ω2r5α¯(1)](0,)+𝔼¯0n1[Ψ(1)R](0,)+𝔼¯n1[Ψ(1)R](0),\displaystyle\overline{\mathbb{E}}^{n+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]:=\overline{\mathbb{E}}^{n+1}_{0}[\Omega^{2}r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}]\left(0,\infty\right)+\overline{\mathbb{E}}^{n+1}_{0}[\Omega^{-2}r^{5}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]\left(0,\infty\right)+\overline{\mathbb{E}}^{n-1}_{0}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{R}]\left(0,\infty\right)+\overline{\mathbb{E}}^{n-1}_{\mathcal{I}}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{R}]\left(0\right)\,, (132)

which contains the energy fluxes of α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} and α¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}, the flux of Ψ(1)R\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{R} 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 𝔼¯0n1[Ψ(1)D](0,)+𝔼¯n1[Ψ(1)D](0)\overline{\mathbb{E}}^{n-1}_{0}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{D}]\left(0,\infty\right)+\overline{\mathbb{E}}^{n-1}_{\mathcal{I}}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{D}]\left(0\right) 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 Σt\Sigma_{t^{\star}} in [GH24] to estimates on null cones.

Theorem 4.2.

We have the following estimates for any n3n\geq 3:

  • Boundedness estimate: For fixed (u,v)(u,v)\in\mathcal{M} we have

    𝔼¯vn[Ω2rα(1)](v,u)+𝔼¯vn[Ω2r5α¯(1)](v,u)+𝔼un[Ω2rα(1)](0,v)+𝔼un[Ω2r5α¯(1)](0,v)\displaystyle\overline{\mathbb{E}}^{n}_{v}[\Omega^{2}r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}]\left(v,u\right)+\overline{\mathbb{E}}^{n}_{v}[\Omega^{-2}r^{5}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]\left(v,u\right)+{\mathbb{E}}^{n}_{u}[\Omega^{2}r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}]\left(0,v\right)+{\mathbb{E}}^{n}_{u}[\Omega^{-2}r^{5}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]\left(0,v\right)
    +𝔼¯vn2[Ψ(1)R](v,u)+𝔼un2[Ψ(1)R](0,v)+𝔼¯n2[Ψ(1)R](v)\displaystyle+\overline{\mathbb{E}}^{n-2}_{v}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{R}]\left(v,u\right)+{\mathbb{E}}^{n-2}_{u}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{R}]\left(0,v\right)+\overline{\mathbb{E}}^{n-2}_{\mathcal{I}}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{R}]\left(v\right) 𝔼¯datan[α(1),α¯(1)].\displaystyle\,\lesssim\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]\,. (133)
  • Decay estimates: Fix an r=r0>r+r=r_{0}>r_{+}. For fixed v2v\geq 2 we denote by u(r0,v)u(r_{0},v) the uu-value of the intersection of {r=r0}\{r=r_{0}\} and the fixed vv-hypersurface. We then have for n4n\geq 4

    𝔼¯vn1[Ω2rα(1)](u(r0,v),)+𝔼¯u(r0,v)n1[Ω2rα(1)](v,u(r0,v))\displaystyle\overline{\mathbb{E}}^{n-1}_{v}[\Omega^{2}r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}]\left(u(r_{0},v),\infty\right)+\overline{\mathbb{E}}^{n-1}_{u(r_{0},v)}[\Omega^{2}r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}]\left(v,u(r_{0},v)\right)
    +\displaystyle+ 𝔼¯vn1[Ω2r5α¯(1)](u(r0,v),)+𝔼¯u(r0,v)n1[Ω2r5α¯(1)](v,u(r0,v))\displaystyle\overline{\mathbb{E}}^{n-1}_{v}[\Omega^{-2}r^{5}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]\left(u(r_{0},v),\infty\right)+\overline{\mathbb{E}}^{n-1}_{u(r_{0},v)}[\Omega^{-2}r^{5}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]\left(v,u(r_{0},v)\right)
    +\displaystyle+ 𝔼¯vn3[Ψ(1)R](u(r0,v),)+𝔼¯u(r0,v)n3[Ψ(1)R](v,u(r0,v))+𝔼¯n2[Ψ(1)R](u(r0,v))𝔼¯datan[α(1),α¯(1)](logv)2,\displaystyle\overline{\mathbb{E}}^{n-3}_{v}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{R}]\left(u(r_{0},v),\infty\right)+\overline{\mathbb{E}}^{n-3}_{u(r_{0},v)}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{R}]\left(v,u(r_{0},v)\right)+\overline{\mathbb{E}}^{n-2}_{\mathcal{I}}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{R}]\left(u(r_{0},v)\right)\lesssim\frac{\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]}{(\log v)^{2}}\,, (134)

    with the \lesssim now depending on the r0r_{0} (the implicit constant blows up as r0r+r_{0}\rightarrow r_{+}).

  • Estimates for the non-degenerate outgoing fluxes near the horizon: We have for any 0u0\leq u\leq\infty, v2v1+1v0v_{2}\geq v_{1}+1\geq v_{0} and n3n\geq 3 the estimates

    𝔼¯un[Ω2rα(1)](v1,v2)+𝔼¯un2[Ψ(1)R](v1,v2)(v2v1)𝔼¯datan[α(1),α¯(1)],\displaystyle\overline{\mathbb{E}}^{n}_{u}[\Omega^{2}r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}]\left(v_{1},v_{2}\right)+\overline{\mathbb{E}}^{n-2}_{u}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{R}]\left(v_{1},v_{2}\right)\lesssim(v_{2}-v_{1})\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]\,, (135)
    𝔼¯un[Ω2rα(1)](v1,v2)+𝔼¯un2[Ψ(1)R](v1,v2)(v2v1)log(v1)2𝔼¯datan+1[α(1),α¯(1)].\displaystyle\overline{\mathbb{E}}^{n}_{u}[\Omega^{2}r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}]\left(v_{1},v_{2}\right)+\overline{\mathbb{E}}^{n-2}_{u}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}^{R}]\left(v_{1},v_{2}\right)\lesssim\frac{(v_{2}-v_{1})}{\log(v_{1})^{2}}\overline{\mathbb{E}}^{n+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]\,. (136)
Remark 4.3.

Integrated decay estimates follow as a corollary by integrating the ingoing fluxes in vv. The integrated decay has again growth like (v2v1)log(v1)2\frac{(v_{2}-v_{1})}{\log(v_{1})^{2}}.

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 rr0r\geq r_{0}. 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 NN from [DR08] which generates an energy identity whose bulk term has a good sign in a region rr1r\leq r_{1} for some r1>r+r_{1}>r_{+}. We apply the energy identity in a region bounded by a Σt\Sigma_{t^{\star}}-slice, the horizon and an ingoing cone emanating from the intersection of the Σt\Sigma_{t^{\star}} slice with the r=r1r=r_{1} hypersurface. This produces control on the desired ingoing flux noting that for rr1r\geq r_{1} the degenerate energy is equivalent to the non-degenerate one. Applying NN globally also yields (135)–(136) immediately using (4.2) and (4.2) respectively (integrated in time) to control the error in {rr1\{r\geq r_{1}} in the corresponding vectorfield identity.

Corollary 4.5.

We have for n3n\geq 3 the following estimates on spheres:

  • Boundedness estimates

    i=0n1supu,v𝔇irΩ2α(1)(u,v)2+i=0n1supu,v𝔇iΩ2r5α¯(1)(u,v)2\displaystyle\sum_{i=0}^{n-1}\sup_{u,v}\|\mathfrak{D}^{i}r\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}\|^{2}_{(u,v)}+\sum_{i=0}^{n-1}\sup_{u,v}\|\mathfrak{D}^{i}\Omega^{-2}r^{5}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}\|^{2}_{(u,v)} 𝔼¯datan[α(1),α¯(1)].\displaystyle\lesssim\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]\,. (137)
  • Decay estimates

    i=0n1supu,v𝔇irΩ2α(1)(u,v)2+i=0n1supu,v𝔇iΩ2r5α¯(1)(u,v)2\displaystyle\sum_{i=0}^{n-1}\sup_{u,v}\|\mathfrak{D}^{i}r\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}\|^{2}_{(u,v)}+\sum_{i=0}^{n-1}\sup_{u,v}\|\mathfrak{D}^{i}\Omega^{-2}r^{5}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}\|^{2}_{(u,v)} 1(logv)2𝔼¯datan+1[α(1),α¯(1)].\displaystyle\lesssim\frac{1}{(\log v)^{2}}\overline{\mathbb{E}}^{n+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]\,. (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 11-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 nn derivatives of the Ricci coefficients. For n3n\geq 3 we define

𝔻0n:=\displaystyle\mathbb{D}^{n}_{0}:= supu[0,)i+j+k=0i2,k1n1𝒜[j][∇̸T]k[riv][r2Ω1∇̸3]i(Ωχ^(1)))2u,0\displaystyle\sup_{u\in[0,\infty)}\sum_{\begin{subarray}{c}i+j+k=0\\ i\leq 2,k\leq 1\end{subarray}}^{n-1}\big{\|}\mathcal{A}^{[j]}[\not{\nabla}_{T}]^{k}[r\not{div}]\left[r^{2}\Omega^{-1}\not{\nabla}_{3}\right]^{i}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}))\big{\|}^{2}_{u,0}
+\displaystyle+ supu[0,)i+j=0i1n1𝒜[j][r2Ω1∇̸3]i[r∇̸](Ωtrχ)(1)Ω2r2u,02.\displaystyle\sup_{u\in[0,\infty)}\sum_{\begin{subarray}{c}i+j=0\\ i\leq 1\end{subarray}}^{n-1}\big{\|}\mathcal{A}^{[j]}\left[r^{2}\Omega^{-1}\not{\nabla}_{3}\right]^{i}[r\not{\nabla}]\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\Omega^{-2}r^{2}\big{\|}^{2}_{u,0}. (139)

To state our main boundedness and decay theorem we define the initial master energy involving nn derivatives of curvature and Ricci-coefficients

𝔼:=n𝔻0n+𝔼¯datan[α(1),α¯(1)].\displaystyle\overset{\circ}{\mathbb{E}}{}^{n}:=\mathbb{D}^{n}_{0}+\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]\,. (140)
Theorem 4.7.

Given a solution 𝒮\mathscr{S} 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 𝒮=𝒮+𝒢f+𝒢q𝒦𝔪,𝔞\underaccent{\lor}{\mathscr{S}}=\mathscr{S}+\mathscr{G}_{f}+\mathscr{G}_{q}-\mathscr{K}_{\mathfrak{m},\vec{\mathfrak{a}}} be the initial data normalised solution with vanishing =0,1\ell=0,1 modes obtained from Proposition 3.12. Let the initial energy 𝔼̊n\mathring{\mathbb{E}}^{n} in (140) be defined with respect to the geometric quantities of 𝒮\underaccent{\lor}{\mathscr{S}}. Then the geometric quantities of the solution 𝒮\underaccent{\lor}{\mathscr{S}} satisfy the following estimates. For any weighted Ricci or metric coefficient

ξ{Ωχ^(1),Ω1r2χ^¯(1),r2Ω2(Ωtrχ)(1),r2Ω2(Ωtrχ¯)(1),rη(1),rη¯(1),,ω(1),ω¯(1),rΩ1Ω(1),(1),^(1),rΩ2b(1)},\displaystyle\xi\in\{\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}},\Omega^{-1}r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}},\frac{r^{2}}{\Omega^{2}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)},\frac{r^{2}}{\Omega^{2}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)},r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}},r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}},,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}},r\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega},\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}},r\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}\}\,, (141)

and any curvature component

Ξ{rΩ2α(1),r2Ωβ(1),r3ρ(1),r3σ(1),r4Ω1β¯(1),r5Ω2α¯(1)},\displaystyle\Xi\in\{r\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},r^{2}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}},r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho},r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}},r^{4}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}},r^{5}\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}\}\,, (142)

we have for n3n\geq 3 and any u,v0u,v\geq 0

j=0n[r∇̸]jξu,v2+j=0n1[r∇̸]jΞu,v2+j=0n1[r∇̸]j(r2η(1),r2η¯(1),r2Ω1Ω(1))u,v2\displaystyle\sum_{j=0}^{n}\bigg{\|}[r\not{\nabla}]^{j}\xi\bigg{\|}^{2}_{u,v}+\sum_{j=0}^{n-1}\bigg{\|}[r\not{\nabla}]^{j}\Xi\bigg{\|}^{2}_{u,v}+\sum_{j=0}^{n-1}\bigg{\|}[r\not{\nabla}]^{j}(r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}},r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}},r^{2}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega})\bigg{\|}^{2}_{u,v} 𝔼,n\displaystyle\lesssim\overset{\circ}{\mathbb{E}}{}^{n}\,, (143)
j=0n[r∇̸]jξu,v2+j=0n1[r∇̸]jΞu,v2+j=0n1[r∇̸]j(r2η(1),r2η¯(1),r2Ω1Ω(1))u,v2\displaystyle\sum_{j=0}^{n}\bigg{\|}[r\not{\nabla}]^{j}\xi\bigg{\|}^{2}_{u,v}+\sum_{j=0}^{n-1}\bigg{\|}[r\not{\nabla}]^{j}\Xi\bigg{\|}^{2}_{u,v}+\sum_{j=0}^{n-1}\bigg{\|}[r\not{\nabla}]^{j}(r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}},r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}},r^{2}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega})\bigg{\|}^{2}_{u,v} 𝔼n+1(logv)2.\displaystyle\lesssim\frac{\overset{\circ}{\mathbb{E}}{}^{n+1}}{(\log v)^{2}}\,. (144)

Moreover, for the curvature components Ξ\Xi, we also obtain for any vv0=0v\geq v_{0}=0 fixed, the following estimates for the top order fluxes:

u=v𝑑u¯Ω2r2([r2Ω1∇̸3][r∇̸]nΞu¯,v2+[r∇̸]n+1Ξu¯,v2)\displaystyle\int_{u=v}^{\infty}d\bar{u}\frac{\Omega^{2}}{r^{2}}\left(\big{\|}[r^{2}\Omega^{-1}\not{\nabla}_{3}][r\not{\nabla}]^{n}\Xi\big{\|}^{2}_{\bar{u},v}+\big{\|}[r\not{\nabla}]^{n+1}\Xi\big{\|}^{2}_{\bar{u},v}\right) 𝔼,n+1\displaystyle\lesssim\overset{\circ}{\mathbb{E}}{}^{n+1}\,, (145)
vvf𝑑v¯([Ω∇̸4][r∇̸]nΞu¯,v2+[r∇̸]n+1Ξu,v¯2)\displaystyle\int_{v}^{v_{f}}d\bar{v}\left(\big{\|}[\Omega\not{\nabla}_{4}][r\not{\nabla}]^{n}\Xi\big{\|}^{2}_{\bar{u},v}+\big{\|}[r\not{\nabla}]^{n+1}\Xi\big{\|}^{2}_{{u},\bar{v}}\right) 𝔼(vfv)n+1,\displaystyle\lesssim\overset{\circ}{\mathbb{E}}{}^{n+1}\left(v_{f}-v\right)\,, (146)

for any uvfvu\geq v_{f}\geq v. Finally, for r0>r+r_{0}>r_{+} fixed, one may drop the factor of (vfv)(v_{f}-v) in (146) if r(u,v)r0r(u,v)\geq r_{0}. In this case, the \lesssim in (146) will depend on r0r_{0}.

Theorem 4.7 will be proven in Section 5. For now we provide some additional remarks.

Remark 4.8.

The fact that the (formally irregular at the horizon) quantity Ω2(Ωtrχ)(1)\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)} appears in (141) is due to the fact that the solution is initial data normalised (see Definition 3.11) and hence (Ωtrχ)(1)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}=0 on S,v02S^{2}_{\infty,v_{0}} and, as a consequence of the linearised Raychaudhuri equation (55), on the entire event horizon.

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 rr-weighted estimates for η(1),η¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}} and even stronger ones for η(1)+η¯(1)=2[r∇̸]Ω1Ω(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}=2[r\not{\nabla}]\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega} if we are willing to lose two derivatives. A similar improved estimate with loss holds for ω(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}, ω¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}} (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.

In the proof of Theorem 4.7 we will also show the estimate 𝔻n𝔼̊n\mathbb{D}^{n}\lesssim\mathring{\mathbb{E}}^{n} for n4n\geq 4, where 𝔻n\mathbb{D}^{n} is defined as (4.2) but replacing the supu[u0,)\sup_{u\in[u_{0},\infty)} by sup\sup_{\mathcal{M}}. In other words, there is no loss of derivatives in the boundedness statement that we obtain.

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:

|r1b(1)|+|^(1)AB|+|(1)|(𝔼)412logv,\displaystyle|r^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}|+\Big{|}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}_{AB}}{\sqrt{\not{g}}}\Big{|}+\Big{|}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}\Big{|}\leq\frac{(\overset{\circ}{\mathbb{E}}{}^{4})^{\frac{1}{2}}}{\log v}\,, (147)
|r2Ω1Ω(1)|+|Ωχ^(1)|+|r2Ω1χ^¯(1)|+|(Ωtrχ)(1)|+|r2Ω2(Ωtrχ¯)(1)|+|r2η(1)|+|r2η¯(1)|+|ω(1)|+|r2Ω2ω¯(1)|(𝔼)412logv,\displaystyle|r^{2}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}|+|\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}|+|r^{2}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}|+|\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}|+|r^{2}\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}|+|r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}|+|r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}|+|\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}|+|r^{2}\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}|\leq\frac{(\overset{\circ}{\mathbb{E}}{}^{4})^{\frac{1}{2}}}{\log v}\,, (148)
|rΩ2α(1)|+|r2Ωβ(1)|+|r3ρ(1)|+|r3σ(1)|+|r4Ω1β(1)|+|r5Ω2α¯(1)|(𝔼)412logv.\displaystyle|r\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}|+|r^{2}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}|+|r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}|+|r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}|+|r^{4}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}|+|r^{5}\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}|\leq\frac{(\overset{\circ}{\mathbb{E}}{}^{4})^{\frac{1}{2}}}{\log v}\,. (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 𝒢f+𝒢q\mathscr{G}_{f}+\mathscr{G}_{q} such that the geometric quantities associated with the corresponding solution 𝒮=𝒮+𝒢f+𝒢q\accentset{\land}{\mathscr{S}}=\underaccent{\lor}{\mathscr{S}}+\mathscr{G}_{f}+\mathscr{G}_{q} satisfy the estimates of Theorem 4.7 but now for

ξ{r2Ωχ^(1),Ω1r4χ^¯(1),r(Ωtrχ)(1),r3Ω2(Ωtrχ¯)(1),r3η(1),r3η¯(1),,rω(1),rω¯(1),r2Ω1Ω(1),r(1),r^(1),b(1)}.\xi\in\{r^{2}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}},\Omega^{-1}r^{4}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}},r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)},r^{3}\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)},r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}},r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}},,r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega},r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}},r^{2}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega},r\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}},r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}\}\,.

Moreover, the pure gauge solutions 𝒢f\mathscr{G}_{f}, 𝒢q\mathscr{G}_{q} used in the above is uniformly bounded by initial data in the sense that the geometric quantities of the pure gauge solution 𝒢f\mathscr{G}_{f}, 𝒢q\mathscr{G}_{q} satisfy the estimates of Theorem 4.7.

Remark 4.15.

Note the improvement in rr-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.

As the proof involves repeating many of the same estimates of the proof of Theorem 4.7, we only provide a sketch of Theorem 4.14 in Section 6.

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 𝒮\underaccent{\lor}{\mathscr{S}}, which have vanishing =0,1\ell=0,1 modes, the elliptic operators 𝒜[i]\mathcal{A}^{[i]} have trivial kernel when acting on such a quantity and hence allow estimating the entire HiH^{i}-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 ξ\xi be an Su,v2S^{2}_{u,v} tensor satisfying the propagation equation

Ω∇̸3ξ=Ω2r2B\displaystyle\Omega\not{\nabla}_{3}\xi=\frac{\Omega^{2}}{r^{2}}B (150)

along the ingoing cone C¯v\underline{C}_{v} (which intersects \mathcal{I} in the sphere Sv,v2S^{2}_{v,v}). Assume BB satisfies

vΩ2r2Bu¯,v2𝑑u¯𝔻\int_{v}^{\infty}\frac{\Omega^{2}}{r^{2}}\|B\|^{2}_{\bar{u},v}d\bar{u}\lesssim\mathbb{D}\,

along the cone. Then, provided ξv,v<\|\xi\|_{v,v}<\infty, we have

supuξu,vξv,v+𝔻r.\sup_{u}\|\xi\|_{u,v}\lesssim\|\xi\|_{v,v}+\sqrt{\frac{\mathbb{D}}{r}}\,.

Moreover, the statement remains true replacing 𝔻\mathbb{D} by 𝔻(logv)2\frac{\mathbb{D}}{(\log v)^{2}} everywhere.

Proof.

Direct consequence of Cauchy-Schwarz and integrability in uu of Ω2r2=urr2=u(1r)\frac{\Omega^{2}}{r^{2}}=-\frac{\partial_{u}r}{r^{2}}=\partial_{u}\left(\frac{1}{r}\right). ∎

Lemma 5.2.

Let ξ\xi be an Su,v2S^{2}_{u,v} tensor satisfying the propagation equation

Ω∇̸4ξ=Ω2r2B\displaystyle\Omega\not{\nabla}_{4}\xi=\frac{\Omega^{2}}{r^{2}}B (151)

along the outgoing cone Cu{C}_{u} (which intersects \mathcal{I} in the sphere Su,u2S^{2}_{u,u}). Assume BB satisfies

v1v2Bu,v¯2𝑑v¯𝔻(v2v1)\displaystyle\int_{v_{1}}^{v_{2}}\|B\|^{2}_{{u},\bar{v}}d\bar{v}\lesssim\mathbb{D}\cdot\left(v_{2}-v_{1}\right)\, (152)

for any v2v10v_{2}\geq v_{1}\geq 0 along the cone and also for some fixed r0>r+r_{0}>r_{+} the bound

v(r0,u)uBu,v¯2𝑑v¯r0𝔻.\displaystyle\int_{v(r_{0},u)}^{u}\|B\|^{2}_{{u},\bar{v}}d\bar{v}\lesssim_{r_{0}}\mathbb{D}.\, (153)

Then we have

supvΩ2r2ξu,vΩ2r2ξu,v0+𝔻.\displaystyle\sup_{v}\|\Omega^{-2}r^{2}\xi\|_{u,v}\lesssim\|\Omega^{-2}r^{2}\xi\|_{u,v_{0}}+\sqrt{\mathbb{D}}\,. (154)
Proof.

We write (151) as

Ω∇̸4(Ω2ξ)+2ω(Ω2ξ)=1r2B\displaystyle\Omega\not{\nabla}_{4}(\Omega^{-2}\xi)+2\omega(\Omega^{-2}\xi)=\frac{1}{r^{2}}B (155)

and hence, contracting with (Ω2ξ)(\Omega^{-2}\xi) and applying Cauchy’s inequality with an ϵ\epsilon and an absorption argument we get

v((Ω2ξ)u,v2)+(k2r+Mr2)(Ω2ξ)u,v24r4Bu,v2.\displaystyle\partial_{v}\left(\|(\Omega^{-2}\xi)\|_{u,v}^{2}\right)+\left(k^{2}r+\frac{M}{r^{2}}\right)\|(\Omega^{-2}\xi)\|_{u,v}^{2}\lesssim\frac{4}{r^{4}}\|B\|_{u,v}^{2}\,. (156)

We have (k2r+Mr2)c1\left(k^{2}r+\frac{M}{r^{2}}\right)\geq c_{1} and therefore integrating between v1v_{1} and v2v_{2} yields

(Ω2ξ)u,v22+c1v1v2(Ω2ξ)u,v2𝑑v(Ω2ξ)u,v12+𝔻(v2v1).\displaystyle\|(\Omega^{-2}\xi)\|_{u,v_{2}}^{2}+c_{1}\int_{v_{1}}^{v_{2}}\|(\Omega^{-2}\xi)\|_{u,v}^{2}dv\lesssim\|(\Omega^{-2}\xi)\|_{u,v_{1}}^{2}+\mathbb{D}\cdot\left(v_{2}-v_{1}\right)\,. (157)

An elementary calculus exercise yields the conclusion (154) without the factor of r2r^{2}. To improve the weight near infinity we can integrate (151) directly from Su,v(r0,u)2S^{2}_{u,v(r_{0},u)} (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 BB satisfies in addition

v1v2Bu,v¯2𝑑v¯𝔻(v2v1)(logv1)2\displaystyle\int_{v_{1}}^{v_{2}}\|B\|^{2}_{{u},\bar{v}}d\bar{v}\lesssim\mathbb{D}\cdot\frac{\left(v_{2}-v_{1}\right)}{(\log v_{1})^{2}}\, (158)

for any v2v10v_{2}\geq v_{1}\geq 0 along the cone and also for some fixed r0>r+r_{0}>r_{+} the bound

v(r0,u)uBu,v¯2𝑑v¯r0𝔻(logv(r0,u)))2,\displaystyle\int_{v(r_{0},u)}^{u}\|B\|^{2}_{{u},\bar{v}}d\bar{v}\lesssim_{r_{0}}\frac{\mathbb{D}}{(\log v(r_{0},u)))^{2}},\, (159)

then ξ\xi satisfies along CuC_{u} for any vv the decay bound

Ω2r2ξu,vΩ2r2ξu,v0+𝔻logv.\displaystyle\|\Omega^{-2}r^{2}\xi\|_{u,v}\lesssim\frac{\|\Omega^{-2}r^{2}\xi\|_{u,v_{0}}+\sqrt{\mathbb{D}}}{\log v}\,. (160)
Proof.

Simple variation of the previous proof. ∎

5.3 Estimates on the horizon

We recall that (Ωtrχ)(1)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}=0 and ivη(1)+ρ(1)=0\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}=0 identically on +\mathcal{H}^{+} by the initial data normalisation, and that all geometric quantities have vanishing =0,1\ell=0,1 modes.

Proposition 5.4.

We have for n3n\geq 3 the following flux estimates on the horizon for any v2v1v0v_{2}\geq v_{1}\geq v_{0}

i=1n1v1v2dv¯𝒜[i]2(Ωβ(1)),𝒜[i]2iv(Ωχ^(1)),𝒜[i1]21(r3ρ(1),r3σ(1)),v¯2(v2v1)𝔼¯datan[α(1),α¯(1)],\displaystyle\sum_{i=1}^{n-1}\int_{v_{1}}^{v_{2}}d\bar{v}\|\mathcal{A}^{[i]}\not{\mathcal{D}}_{2}^{\star}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}),\mathcal{A}^{[i]}\not{\mathcal{D}}_{2}^{\star}\not{div}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}),\mathcal{A}^{[i-1]}\not{\mathcal{D}}_{2}^{\star}\not{\mathcal{D}}_{1}^{\star}(r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho},r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}})\|_{\infty,\bar{v}}^{2}\lesssim\left(v_{2}-v_{1}\right)\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]\,, (161)
i=1n1v1v2dv¯𝒜[i]2(Ωβ(1)),𝒜[i]2iv(Ωχ^(1)),𝒜[i1]21(r3ρ(1),r3σ(1)),v¯2(v2v1)𝔼¯datan+1[α(1),α¯(1)](logv1)2.\displaystyle\sum_{i=1}^{n-1}\int_{v_{1}}^{v_{2}}d\bar{v}\|\mathcal{A}^{[i]}\not{\mathcal{D}}_{2}^{\star}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}),\mathcal{A}^{[i]}\not{\mathcal{D}}_{2}^{\star}\not{div}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}),\mathcal{A}^{[i-1]}\not{\mathcal{D}}_{2}^{\star}\not{\mathcal{D}}_{1}^{\star}(r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho},r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}})\|_{\infty,\bar{v}}^{2}\lesssim\frac{\left(v_{2}-v_{1}\right)\overline{\mathbb{E}}^{n+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]}{(\log v_{1})^{2}}\,. (162)

We have on the horizon the following estimates on spheres: For n3n\geq 3

i=1n2𝒜[i]2(Ωβ(1)),𝒜[i]2iv(Ωχ^(1)),𝒜[i1]21(r3ρ(1),r3σ(1)),v2𝔼¯datan[α(1),α¯(1)],\displaystyle\sum_{i=1}^{n-2}\|\mathcal{A}^{[i]}\not{\mathcal{D}}_{2}^{\star}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}),\mathcal{A}^{[i]}\not{\mathcal{D}}_{2}^{\star}\not{div}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}),\mathcal{A}^{[i-1]}\not{\mathcal{D}}_{2}^{\star}\not{\mathcal{D}}_{1}^{\star}(r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho},r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}})\|_{\infty,v}^{2}\lesssim\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]\,, (163)
i=1n2𝒜[i]2(Ωβ(1)),𝒜[i]2iv(Ωχ^(1)),𝒜[i1]21(r3ρ(1),r3σ(1)),v2𝔼¯datan+1[α(1),α¯(1)](logv)2.\displaystyle\sum_{i=1}^{n-2}\|\mathcal{A}^{[i]}\not{\mathcal{D}}_{2}^{\star}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}),\mathcal{A}^{[i]}\not{\mathcal{D}}_{2}^{\star}\not{div}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}),\mathcal{A}^{[i-1]}\not{\mathcal{D}}_{2}^{\star}\not{\mathcal{D}}_{1}^{\star}(r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho},r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}})\|_{\infty,v}^{2}\lesssim\frac{\overline{\mathbb{E}}^{n+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]}{(\log v)^{2}}\,. (164)
Proof.

This follows as in [DHR19], so we merely sketch the argument. To obtain the bounds for β(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}} we write

1Ω∇̸3(Ω2rα(1)),v=S,v2sinθdθdϕ|22Ωβ(1)+3ρΩχ^(1)|2=S,v2sinθdθdϕ(|22Ωβ(1)|26ρ|β(1)|2+9ρ2|Ωχ^(1)|2),\displaystyle\Big{\|}\frac{1}{\Omega}\not{\nabla}_{3}(\Omega^{2}r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}})\Big{\|}_{\infty,v}=\int_{S^{2}_{\infty,v}}\sin\theta d\theta d\phi|2\not{\mathcal{D}}_{2}^{\star}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}+3\rho\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}|^{2}=\int_{S^{2}_{\infty,v}}\sin\theta d\theta d\phi\left(|2\not{\mathcal{D}}_{2}^{\star}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}|^{2}-6\rho|\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}|^{2}+9\rho^{2}|\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}|^{2}\right)\,,

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 (ivΩχ^(1)=β(1)\not{div}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}=-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}). Note ρ=2Mr3\rho=-\frac{2M}{r^{3}}, so the expression is indeed coercive. Angular commuted identities are obtained analogously. The result for β(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}} (and by Codazzi for χ^(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}) now follows from the flux (and sphere) bounds available for the quantity on the left through Theorem 4.2. The result for (ρ(1),σ(1))(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}) follows from the the fact that, by (97), on the horizon

ivΨ(1)=r5iv21(ρ(1),σ(1))+34r5ρtrχΩivΩχ^(1)\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Psi}=r^{5}\not{div}\not{\mathcal{D}}_{2}^{\star}\not{\mathcal{D}}_{1}^{\star}(-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}})+\frac{3}{4}r^{5}\rho\frac{tr\chi}{\Omega}\not{div}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}

and that we control the flux on the left from Theorem 4.2 and the flux of χ^(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}} from the first part of the proof. ∎

Corollary 5.5.

On the event horizon +\mathcal{H}^{+}, we have for n3n\geq 3 the following flux estimates

i=0nv1v2dv¯[r∇̸]i(Ωβ(1)),[r∇̸]i(ρ(1),σ(1)),v¯2+i=0n+1v1v2dv¯[[r∇̸]i(Ωχ^(1)),[r∇̸]iη(1),v¯2(v2v1)𝔼¯datan[α(1),α¯(1)],\displaystyle\sum_{i=0}^{n}\int_{v_{1}}^{v_{2}}d\bar{v}\|[r\not{\nabla}]^{i}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}),[r\not{\nabla}]^{i}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}})\|_{\infty,\bar{v}}^{2}+\sum_{i=0}^{n+1}\int_{v_{1}}^{v_{2}}d\bar{v}\|[[r\not{\nabla}]^{i}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}),[r\not{\nabla}]^{i}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}\|_{\infty,\bar{v}}^{2}\lesssim\left(v_{2}-v_{1}\right)\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]\,,
i=0nv1v2dv¯[r∇̸]i(Ωβ(1)),[r∇̸]i(ρ(1),σ(1)),v¯2+i=0n+1v1v2dv¯[[r∇̸]i(Ωχ^(1)),[r∇̸]iη(1),v¯2(v2v1)𝔼¯datan+1[α(1),α¯(1)](logv1)2.\displaystyle\sum_{i=0}^{n}\int_{v_{1}}^{v_{2}}d\bar{v}\|[r\not{\nabla}]^{i}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}),[r\not{\nabla}]^{i}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}})\|_{\infty,\bar{v}}^{2}+\sum_{i=0}^{n+1}\int_{v_{1}}^{v_{2}}d\bar{v}\|[[r\not{\nabla}]^{i}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}),[r\not{\nabla}]^{i}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}\|_{\infty,\bar{v}}^{2}\lesssim\frac{\left(v_{2}-v_{1}\right)\overline{\mathbb{E}}^{n+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]}{(\log v_{1})^{2}}\,.

In addition, we have the estimates on spheres

i=0n1[r∇̸]i(Ωβ(1)),[r∇̸]i(ρ(1),σ(1)),v2+i=0n[[r∇̸]i(Ωχ^(1)),[r∇̸]iη(1),v2𝔼¯datan[α(1),α¯(1)],\displaystyle\sum_{i=0}^{n-1}\|[r\not{\nabla}]^{i}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}),[r\not{\nabla}]^{i}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}})\|_{\infty,{v}}^{2}+\sum_{i=0}^{n}\|[[r\not{\nabla}]^{i}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}),[r\not{\nabla}]^{i}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}\|_{\infty,{v}}^{2}\lesssim\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]\,,
i=0n1[r∇̸]i(Ωβ(1)),[r∇̸]i(ρ(1),σ(1)),v2+i=0n[[r∇̸]i(Ωχ^(1)),[r∇̸]iη(1),v2𝔼¯datan+1[α(1),α¯(1)](logv)2.\displaystyle\sum_{i=0}^{n-1}\|[r\not{\nabla}]^{i}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}),[r\not{\nabla}]^{i}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}})\|_{\infty,{v}}^{2}+\sum_{i=0}^{n}\|[[r\not{\nabla}]^{i}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}),[r\not{\nabla}]^{i}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}\|_{\infty,{v}}^{2}\lesssim\frac{\overline{\mathbb{E}}^{n+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]}{(\log v)^{2}}\,.

In addition, we may add [r∇̸]i11Ω∇̸3(Ωχ^(1))[r\not{\nabla}]^{i-1}\frac{1}{\Omega}\not{\nabla}_{3}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}) to the list of Ricci-coefficients in the above estimates.

Proof.

In view of ivη(1)+ρ(1)=0\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}=0 and urlησ(1)=0\not{curl}\eta-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}=0 on the horizon, we can clearly add the expression 𝒜[i1]21(r3ivη(1),r3urlη(1))\mathcal{A}^{[i-1]}\not{\mathcal{D}}_{2}^{\star}\not{\mathcal{D}}_{1}^{\star}(-r^{3}\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}},r^{3}\not{curl}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}) to the list of quantities estimated in Proposition 5.4. Since the 1\ell\geq 1 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 Ω1∇̸3(Ωχ^(1))\Omega^{-1}\not{\nabla}_{3}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}) 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 χ^(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}, which according to (57) satisfies

Ω∇̸4(χ^(1)Ωr2)2ω(χ^(1)Ωr2)=α(1)Ω2r2.\displaystyle\Omega\not{\nabla}_{4}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\Omega r^{2})-2\omega(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\Omega r^{2})=-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}\Omega^{2}r^{2}\,. (165)

Commuting twice with the operator Ω1∇̸3\Omega^{-1}\not{\nabla}_{3} turns the exponentially growth factor (2ω-2\omega, a blueshift) into a decay factor (+2ω+2\omega, a redshift), after which the equation can be integrated forwards in vv using the flux bound for α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} 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 Ω2\Omega^{2} is uniformly bounded away from zero, one can of course integrate directly

Ω∇̸4(χ^(1)Ω1r2)=α(1)r2\displaystyle\Omega\not{\nabla}_{4}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\Omega^{-1}r^{2})=-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}r^{2} (166)

all the way to infinity. This gives in particular that r2χ^(1)Ωrχ^(1)\frac{r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}}{\Omega}\sim r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}} is uniformly bounded on \mathcal{I}. Now let us turn to the details. We first derive the key estimate near the horizon.

Proposition 5.6.

There exists an 2r+>r1>r+2r_{+}>r_{1}>r_{+} such that the following estimate holds for any vv0v\geq v_{0} and any j0j\in\mathbb{N}_{0}, k{0,1}k\in\{0,1\}

i=02supv(v1,v2)u(r1,v)duΩ2𝒜[j][∇̸T]k[Ω1∇̸3]i(r2Ωχ^(1)))2u,v\displaystyle\sum_{i=0}^{2}\sup_{v\in(v_{1},v_{2})}\int_{u(r_{1},v)}^{\infty}du\Omega^{2}\big{\|}\mathcal{A}^{[j]}[\not{\nabla}_{T}]^{k}\left[\Omega^{-1}\not{\nabla}_{3}\right]^{i}(r^{2}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}))\big{\|}^{2}_{u,v}
+i=02v1v2dvu(r1,v)duΩ2𝒜[j][∇̸T]k[Ω1∇̸3]i(r2Ωχ^(1)))2u,v\displaystyle+\sum_{i=0}^{2}\int_{v_{1}}^{v_{2}}dv\int_{u(r_{1},v)}^{\infty}du\Omega^{2}\big{\|}\mathcal{A}^{[j]}[\not{\nabla}_{T}]^{k}\left[\Omega^{-1}\not{\nabla}_{3}\right]^{i}(r^{2}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}))\big{\|}^{2}_{u,v}
u(r1,v1)duΩ2𝒜[j][∇̸T]k[Ω1∇̸3]2(r2Ωχ^(1)))2u,v1+(v2v1)supv(v1,v2)𝔼¯vj+k+2[Ω2rα(1)](u(v,r1),).\displaystyle\lesssim\int_{u(r_{1},v_{1})}^{\infty}du\Omega^{2}\big{\|}\mathcal{A}^{[j]}[\not{\nabla}_{T}]^{k}\left[\Omega^{-1}\not{\nabla}_{3}\right]^{2}(r^{2}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}))\big{\|}^{2}_{u,v_{1}}+(v_{2}-v_{1})\sup_{v\in(v_{1},v_{2})}\overline{\mathbb{E}}_{v}^{j+k+2}[\Omega^{2}r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}](u(v,r_{1}),\infty)\,. (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

Ω∇̸4(Ω[Ω1∇̸3]2χ^(1)Ωr2)+ω(Ω[Ω1∇̸3]2χ^(1)Ωr2)+(4Mr3+1l2)Ω1Ω∇̸3(χ^(1)Ωr2)Ω12Mr4(χ^(1)Ωr2)=Ω[Ω1∇̸3]2(α(1)Ω2r2).\displaystyle\begin{aligned} \Omega\not{\nabla}_{4}(\Omega[\Omega^{-1}\not{\nabla}_{3}]^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\Omega r^{2})+\omega(\Omega[\Omega^{-1}\not{\nabla}_{3}]^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\Omega r^{2})&\\ +\left(-\frac{4M}{r^{3}}+\frac{1}{l^{2}}\right)\Omega\frac{1}{\Omega}\not{\nabla}_{3}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\Omega r^{2})-\Omega\frac{12M}{r^{4}}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\Omega r^{2})=&-\Omega[\Omega^{-1}\not{\nabla}_{3}]^{2}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}\Omega^{2}r^{2})\,.\end{aligned} (168)

Of course commutation with 𝒜[j]\mathcal{A}^{[j]} and ∇̸Tk\not{\nabla}_{T}^{k} is trivial and is omitted.

One now proceeds as in[DHR19] multiplying (168) with (Ω[Ω1∇̸3]2χ^(1)Ωr2)(\Omega[\Omega^{-1}\not{\nabla}_{3}]^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\Omega r^{2}) and integrating over the spacetime region {rr0}{v1vv2}\mathcal{M}\cap\{r\leq r_{0}\}\cap\{v_{1}\leq v\leq v_{2}\}. 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 i=0i=0 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 v0v\geq 0 and n3n\geq 3, k{0,1}k\in\{0,1\}:

j=0n2i=02u(r1,v)duΩ2𝒜[j][∇̸T]k[Ω1∇̸3]i(r2Ωχ^(1)))2u,v\displaystyle\sum_{j=0}^{n-2}\sum_{i=0}^{2}\int_{u(r_{1},v)}^{\infty}du\Omega^{2}\big{\|}\mathcal{A}^{[j]}[\not{\nabla}_{T}]^{k}\left[\Omega^{-1}\not{\nabla}_{3}\right]^{i}(r^{2}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}))\big{\|}^{2}_{u,v} 𝔼¯datan+k[α(1),α¯(1)]+𝔻0n+k,\displaystyle\lesssim\overline{\mathbb{E}}^{n+k}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+k}_{0}\,, (169)
j=0n2i=02u(r1,v)duΩ2𝒜[j][∇̸T]k[Ω1∇̸3]i(r2Ωχ^(1)))2u,v\displaystyle\sum_{j=0}^{n-2}\sum_{i=0}^{2}\int_{u(r_{1},v)}^{\infty}du\Omega^{2}\big{\|}\mathcal{A}^{[j]}[\not{\nabla}_{T}]^{k}\left[\Omega^{-1}\not{\nabla}_{3}\right]^{i}(r^{2}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}))\big{\|}^{2}_{u,v} 𝔼¯datan+k+1[α(1),α¯(1)]+𝔻0n+k+1(logv)2.\displaystyle\lesssim\frac{\overline{\mathbb{E}}^{n+k+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+k+1}_{0}}{(\log v)^{2}}\,. (170)

We can now easily globalise the result to the entire exterior taking care of the correct rr-weights and also improve to an estimate on spheres.

Proposition 5.8.

For any v0v\geq 0 and n3n\geq 3, k{0,1}k\in\{0,1\}

j=0n2i=02𝒜[j][∇̸T]k[r2Ω1∇̸3]i(Ωχ^(1)))2u,v\displaystyle\sum_{j=0}^{n-2}\sum_{i=0}^{2}\big{\|}\mathcal{A}^{[j]}[\not{\nabla}_{T}]^{k}\left[r^{2}\Omega^{-1}\not{\nabla}_{3}\right]^{i}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}))\big{\|}^{2}_{u,v} 𝔼¯datan+k[α(1),α¯(1)]+𝔻0n+k,\displaystyle\lesssim\overline{\mathbb{E}}^{n+k}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+k}_{0}\,,
j=0n2i=02𝒜[j][∇̸T]k[r2Ω1∇̸3]i(Ωχ^(1)))2u,v\displaystyle\sum_{j=0}^{n-2}\sum_{i=0}^{2}\big{\|}\mathcal{A}^{[j]}[\not{\nabla}_{T}]^{k}\left[r^{2}\Omega^{-1}\not{\nabla}_{3}\right]^{i}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}))\big{\|}^{2}_{u,v} 𝔼¯datan+k+1[α(1),α¯(1)]+𝔻0n+k+1(logv)2.\displaystyle\lesssim\frac{\overline{\mathbb{E}}^{n+k+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+k+1}_{0}}{(\log v)^{2}}\,. (171)

Moreover, both estimates also hold replacing [∇̸T]k[r2Ω1∇̸3]i[\not{\nabla}_{T}]^{k}\left[r^{2}\Omega^{-1}\not{\nabla}_{3}\right]^{i} by [∇̸T]k+2[\not{\nabla}_{T}]^{k+2}.

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 TT-derivatives in the algebra for the proof.

We first obtain these estimates in the region rr1r\leq r_{1}. If we restricted the sum over ii to run from 0 to 11 only, both estimates follow directly from Proposition 5.7 and the fundamental theorem of calculus (which loses one Ω1∇̸3\Omega^{-1}\not{\nabla}_{3} derivative, hence the restriction to i1i\leq 1). To show it for i=2i=2 one revisits (168), now written as

Ω∇̸4([Ω1∇̸3]2χ^(1)Ωr2)+2ω(Ω[Ω1∇̸3]2χ^(1)Ωr2)\displaystyle\Omega\not{\nabla}_{4}([\Omega^{-1}\not{\nabla}_{3}]^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\Omega r^{2})+2\omega(\Omega[\Omega^{-1}\not{\nabla}_{3}]^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\Omega r^{2})
+(4Mr3+1l2)1Ω∇̸3(χ^(1)Ωr2)12Mr4(χ^(1)Ωr2)=\displaystyle+\left(-\frac{4M}{r^{3}}+\frac{1}{l^{2}}\right)\frac{1}{\Omega}\not{\nabla}_{3}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\Omega r^{2})-\frac{12M}{r^{4}}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\Omega r^{2})= [Ω1∇̸3]2(α(1)Ω2r2).\displaystyle-[\Omega^{-1}\not{\nabla}_{3}]^{2}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}\Omega^{2}r^{2})\,. (172)

Using the estimates already shown we obtain (after trivially commuting the above with 𝒜[j][∇̸T]k\mathcal{A}^{[j]}[\not{\nabla}_{T}]^{k} )

𝒜[j][∇̸T]k[Ω1∇̸3]2χ^(1)Ωr2u,v22+v1v2𝑑v𝒜[j][∇̸T]k[Ω1∇̸3]2χ^(1)Ωr2u,v¯2\displaystyle\|\mathcal{A}^{[j]}[\not{\nabla}_{T}]^{k}[\Omega^{-1}\not{\nabla}_{3}]^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\Omega r^{2}\|_{u,v_{2}}^{2}+\int_{v_{1}}^{v_{2}}dv\|\mathcal{A}^{[j]}[\not{\nabla}_{T}]^{k}[\Omega^{-1}\not{\nabla}_{3}]^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\Omega r^{2}\|_{u,\bar{v}}^{2}
𝒜[j][∇̸T]k[Ω1∇̸3]2χ^(1)Ωr2u,v12+(𝔼¯dataj+k+2[α(1),α¯(1)]+𝔻0j+k)(v2v1),\displaystyle\lesssim\|\mathcal{A}^{[j]}[\not{\nabla}_{T}]^{k}[\Omega^{-1}\not{\nabla}_{3}]^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\Omega r^{2}\|_{u,v_{1}}^{2}+\left(\overline{\mathbb{E}}^{j+k+2}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{j+k}_{0}\right)(v_{2}-v_{1})\,,

from which the estimates follow also for i=2i=2 by simple ODE theory.

Having established the estimates of the proposition for rr1r\leq r_{1}, we integrate the (appropriately commuted) linearised null structure equation Ω∇̸4(Ω1r2χ^(1))=r2α(1)\Omega\not{\nabla}_{4}(\Omega^{-1}r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}})=-r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} from r=r1r=r_{1} towards infinity to deduce the result also for rr1r\geq r_{1}. (Observe that near infinity Ω1r2Ωr\Omega^{-1}r^{2}\sim\Omega\sim r.) Note that Ω∇̸3\Omega\not{\nabla}_{3} and also 𝒜[j]\mathcal{A}^{[j]} commute trivially on the left and that Ω∇̸3r2Ω1∇̸3\Omega\not{\nabla}_{3}\sim r^{2}\Omega^{-1}\not{\nabla}_{3} near infinity. ∎

5.4.2 The ingoing shear

The boundary condition (81) now allows us to integrate (57) written as

Ω∇̸3(χ^¯(1)Ω1r2)=Ω2r5α¯(1)Ω2r3\displaystyle\Omega\not{\nabla}_{3}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}\Omega^{-1}r^{2})=-\Omega^{-2}r^{5}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}\frac{\Omega^{2}}{r^{3}} (173)

directly from the boundary to produce global uniform bounds on the (regular at +\mathcal{H}^{+}) quantity r2Ω1χ^¯(1)r^{2}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}:

Proposition 5.9.

For any v0v\geq 0 and n3n\geq 3, k{0,1}k\in\{0,1\}:

j=0n2i=02𝒜[j][∇̸T]k[Ω∇̸4]i(r2Ω1χ^¯(1)))2u,v\displaystyle\sum_{j=0}^{n-2}\sum_{i=0}^{2}\big{\|}\mathcal{A}^{[j]}[\not{\nabla}_{T}]^{k}\left[\Omega\not{\nabla}_{4}\right]^{i}(r^{2}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}))\big{\|}^{2}_{u,v} 𝔼¯datan+k[α(1),α¯(1)]+𝔻0n+k,\displaystyle\lesssim\overline{\mathbb{E}}^{n+k}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+k}_{0}\,,
j=0n2i=02𝒜[j][∇̸T]k[Ω∇̸4]i(r2Ω1χ^¯(1)))2u,v\displaystyle\sum_{j=0}^{n-2}\sum_{i=0}^{2}\big{\|}\mathcal{A}^{[j]}[\not{\nabla}_{T}]^{k}\left[\Omega\not{\nabla}_{4}\right]^{i}(r^{2}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}))\big{\|}^{2}_{u,v} 𝔼¯datan+k+1[α(1),α¯(1)]+𝔻0n+k+1(logv)2.\displaystyle\lesssim\frac{\overline{\mathbb{E}}^{n+k+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+k+1}_{0}}{(\log v)^{2}}\,.
Proof.

We apply Lemma 5.1 to (173) and the Ω∇̸4\Omega\not{\nabla}_{4}-commuted (173). The only thing which is not immediate is the initial condition for the Ω∇̸4\Omega\not{\nabla}_{4} commuted estimate. For this we note that the boundary condition ∇̸T(rχ^(1)rχ^¯(1))=0\not{\nabla}_{T}(r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}-r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}})=0 on \mathcal{I} translates into Ω∇̸3(rχ^(1))Ω∇̸4(rχ^¯(1))=0\Omega\not{\nabla}_{3}(r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}})-\Omega\not{\nabla}_{4}(r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}})=0 on \mathcal{I} using that α(1)r2,α¯(1)r2\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}r^{2},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}r^{2} vanish on the boundary. Indeed,

0=∇̸T(rχ^(1)rχ^¯(1))\displaystyle 0=\not{\nabla}_{T}(r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}-r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}) =Ω∇̸3(rχ^(1))Ω∇̸4(rχ^¯(1))+Ω∇̸4(r2Ω2Ωχ^(1)Ωr)Ω∇̸3(r2Ω2Ωχ^¯(1)Ωr)\displaystyle=\Omega\not{\nabla}_{3}(r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}})-\Omega\not{\nabla}_{4}(r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}})+\Omega\not{\nabla}_{4}\left(\frac{r^{2}}{\Omega^{2}}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\frac{\Omega}{r}\right)-\Omega\not{\nabla}_{3}\left(\frac{r^{2}}{\Omega^{2}}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}\frac{\Omega}{r}\right)
=Ω∇̸3(rχ^(1))Ω∇̸4(rχ^¯(1))Ωrα(1)r2+Ωrα¯(1)r2+(r2rΩr)(Ωχ^(1)+Ωχ^¯(1)),\displaystyle=\Omega\not{\nabla}_{3}(r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}})-\Omega\not{\nabla}_{4}(r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}})-\frac{\Omega}{r}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}r^{2}+\frac{\Omega}{r}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}r^{2}+\left(r^{2}\partial_{r}\frac{\Omega}{r}\right)\left(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}+\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}\right), (174)

and the last 44-terms vanish on the boundary by Definition 3.1. Similarly ∇̸T2(rχ^(1)rχ^¯(1))=0\not{\nabla}_{T}^{2}(r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}-r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}})=0 on \mathcal{I} translates into [Ω∇̸3]2(rχ^(1))[Ω∇̸4]2(rχ^¯(1))=2k3α(1)r3[\Omega\not{\nabla}_{3}]^{2}(r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}})-[\Omega\not{\nabla}_{4}]^{2}(r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}})=2k^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}r^{3} on \mathcal{I} since up to terms vanishing in the limit on \mathcal{I} we have

0=∇̸T2(rχ^(1)rχ^¯(1))\displaystyle 0=\not{\nabla}^{2}_{T}(r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}-r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}) =[Ω∇̸3]2(rχ^(1))[Ω∇̸4]2χ^¯(1)+Ω∇̸3(Ω∇̸4(r2Ω2Ωχ^(1)Ωr))Ω∇̸4(Ω∇̸3(r2Ω2Ωχ^¯(1)Ωr))\displaystyle=[\Omega\not{\nabla}_{3}]^{2}(r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}})-[\Omega\not{\nabla}_{4}]^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}+\Omega\not{\nabla}_{3}\left(\Omega\not{\nabla}_{4}\left(\frac{r^{2}}{\Omega^{2}}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\frac{\Omega}{r}\right)\right)-\Omega\not{\nabla}_{4}\left(\Omega\not{\nabla}_{3}\left(\frac{r^{2}}{\Omega^{2}}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}\frac{\Omega}{r}\right)\right)

and hence on \mathcal{I}

0=[Ω∇̸3]2(rχ^(1))[Ω∇̸4]2χ^¯(1)+Ω∇̸3(Ωrα(1)r2+r2Ω2Ωχ^(1)(Ω∇̸4Ωr))Ω∇̸4(Ωrα¯(1)r2+r2Ω2Ωχ^¯(1)(Ω∇̸3Ωr)),\displaystyle 0=[\Omega\not{\nabla}_{3}]^{2}(r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}})-[\Omega\not{\nabla}_{4}]^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}+\Omega\not{\nabla}_{3}\left(-\frac{\Omega}{r}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}r^{2}+\frac{r^{2}}{\Omega^{2}}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\left(\Omega\not{\nabla}_{4}\frac{\Omega}{r}\right)\right)-\Omega\not{\nabla}_{4}\left(-\frac{\Omega}{r}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}r^{2}+\frac{r^{2}}{\Omega^{2}}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}\left(\Omega\not{\nabla}_{3}\frac{\Omega}{r}\right)\right)\,,

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 vv) when integrating the transport equation from the boundary requires mm-derivatives of α¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}} to estimate mm derivatives of χ^¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}} and m+1m+1 derivatives of α¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}} if one would like to see log\log-decay. ∎

5.4.3 Improving the weights near infinity

We now establish a few improved estimates for certain combinations (and derivatives of) χ^(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}} and χ^¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}, which will be helpful in establishing estimates for the torsion later. Specifically, we claim the following:

Proposition 5.10.

We have for n3n\geq 3 the following estimates for any k{0,1}k\in\{0,1\}:

j=0n2supr8M[∇̸T]k𝒜[j]r(Ωχ^(1)Ωχ^¯(1))u,v2𝔼¯datan+k[α(1),α¯(1)]+𝔻0n+k1\displaystyle\sum_{j=0}^{n-2}\sup_{r\geq 8M}\Big{\|}[\not{\nabla}_{T}]^{k}\mathcal{A}^{[j]}r\left(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}-\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}\right)\Big{\|}^{2}_{u,v}\lesssim\overline{\mathbb{E}}^{n+k}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+k-1}_{0} (175)

and

j=0n2supr8M[∇̸T]k𝒜[j]r(Ω∇̸3(Ωχ^(1))Ω∇̸4(Ωχ^¯(1)))u,v2𝔼¯datan+k[α(1),α¯(1)]+𝔻0n+k\displaystyle\sum_{j=0}^{n-2}\sup_{r\geq 8M}\Big{\|}[\not{\nabla}_{T}]^{k}\mathcal{A}^{[j]}r\left(\Omega\not{\nabla}_{3}\left(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\right)-\Omega\not{\nabla}_{4}\left(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}\right)\right)\Big{\|}^{2}_{u,v}\lesssim\overline{\mathbb{E}}^{n+k}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+k}_{0} (176)

and

j=0n3supr8M𝒜[j]r2(Ω∇̸3(Ωχ^(1))+Ω∇̸4(Ωχ^¯(1))+(Ωtrχ¯)Ωχ^(1)+(Ωtrχ)Ωχ^¯(1))u,v2𝔼¯datan[α(1),α¯(1)]+𝔻0n.\displaystyle\sum_{j=0}^{n-3}\sup_{r\geq 8M}\Big{\|}\mathcal{A}^{[j]}r^{2}\left(\Omega\not{\nabla}_{3}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}})+\Omega\not{\nabla}_{4}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}})+\left(\Omega tr\underline{\chi}\right)\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}+\left(\Omega tr\chi\right)\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}\right)\Big{\|}^{2}_{u,v}\lesssim\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}. (177)

Moreover, we can add an additional factor of (logv)2(\log v)^{-2} on the right, provided we replace nn by n+1n+1 in the 𝔼¯data\overline{\mathbb{E}}_{data} energies on the right.

Proof.

To keep the notation in the proof tidy, we ignore the trivial angular commutation by [∇̸T]k𝒜[j][\not{\nabla}_{T}]^{k}\mathcal{A}^{[j]} during the proof, which can be trivially inserted in all equations below. We define the shear along the boundary

X(t,θ):=rχ^(1)(t,t,θ)=rχ^¯(1)(t,t,θ),X(t,\theta):=r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}(t,t,\theta)=r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}(t,t,\theta)\,,

with the last equality following from the boundary condition. Integrating (57) from the boundary \mathcal{I}, we deduce after an integration by parts the identities (here A^,B^\hat{A},\hat{B} are the components in an orthonormal frame!161616Recall the formula for the coordinate components (Ω∇̸3ξ)AB=uξABΩtrχ¯ξAB=r2u(r2ξAB)=r2u(ξA^B^)(\Omega\not{\nabla}_{3}\xi)_{AB}=\partial_{u}\xi_{AB}-\Omega tr\underline{\chi}\xi_{AB}=r^{2}\partial_{u}(r^{-2}\xi_{AB})=r^{2}\partial_{u}(\xi_{\hat{A}\hat{B}}) and hence (Ω∇̸3ξ)A^B^=u(ξA^B^(\Omega\not{\nabla}_{3}\xi)_{\hat{A}\hat{B}}=\partial_{u}(\xi_{\hat{A}\hat{B}}).)

r2Ωχ^(1)A^B^(u,v,θ)\displaystyle\frac{r^{2}}{\Omega}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}_{\hat{A}\hat{B}}(u,v,\theta) =XA^B^(u,θ)+12α(1)A^B^r5Ω21r2(u,v,θ)+vu[12r2vα(1)A^B^r5Ω2](u,v¯,θ)𝑑v¯,\displaystyle=X_{\hat{A}\hat{B}}(u,\theta)+\frac{1}{2}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}_{\hat{A}\hat{B}}r^{5}}{\Omega^{2}}\frac{1}{r^{2}}(u,v,\theta)+\int_{v}^{u}\left[\frac{1}{2r^{2}}\partial_{v}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}_{\hat{A}\hat{B}}r^{5}}{\Omega^{2}}\right](u,\bar{v},\theta)d\bar{v}\,,
r2Ωχ^¯(1)A^B^(u,v,θ)\displaystyle\frac{r^{2}}{\Omega}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}_{\hat{A}\hat{B}}(u,v,\theta) =XA^B^(v,θ)12α¯(1)A^B^r5Ω21r2(u,v,θ)+vu[12r2uα¯(1)A^B^r5Ω2](u¯,v,θ)𝑑u¯.\displaystyle=X_{\hat{A}\hat{B}}(v,\theta)-\frac{1}{2}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}_{\hat{A}\hat{B}}r^{5}}{\Omega^{2}}\frac{1}{r^{2}}(u,v,\theta)+\int_{v}^{u}\left[\frac{1}{2r^{2}}\partial_{u}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}_{\hat{A}\hat{B}}r^{5}}{\Omega^{2}}\right](\bar{u},v,\theta)d\bar{u}. (178)

It follows that

Ω∇̸3(Ωχ^(1))A^B^\displaystyle\Omega\not{\nabla}_{3}\left(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}\right)_{\hat{A}\hat{B}} =(2r6Mr2)Ωχ^(1)A^B^+Ω2r2(X˙A^B^(u,θ)+α(1)A^B^r2+12r2T(α(1)A^B^r5Ω2)+vuu[12r2vα(1)A^B^r5Ω2](u,v¯,θ)dv¯)\displaystyle=\left(\frac{2}{r}-\frac{6M}{r^{2}}\right)\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}_{\hat{A}\hat{B}}+\frac{\Omega^{2}}{r^{2}}\left(\dot{X}_{\hat{A}\hat{B}}(u,\theta)+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}_{\hat{A}\hat{B}}r^{2}+\frac{1}{2r^{2}}T\left(\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}_{\hat{A}\hat{B}}r^{5}}{\Omega^{2}}\right)+\int_{v}^{u}\partial_{u}\left[\frac{1}{2r^{2}}\partial_{v}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}_{\hat{A}\hat{B}}r^{5}}{\Omega^{2}}\right](u,\bar{v},\theta)d\bar{v}\right)

and similarly

Ω∇̸4(Ωχ^¯(1))A^B^\displaystyle\Omega\not{\nabla}_{4}\left(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}\right)_{\hat{A}\hat{B}} =(2r6Mr2)Ωχ^¯(1)A^B^+Ω2r2(X˙A^B^(v,θ)+α¯(1)A^B^r212r2T(α¯(1)A^B^r5Ω2)+vuv[12r2uα¯(1)A^B^r5Ω2](u¯,v,θ)du¯).\displaystyle=-\left(\frac{2}{r}-\frac{6M}{r^{2}}\right)\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}_{\hat{A}\hat{B}}+\frac{\Omega^{2}}{r^{2}}\left(\dot{X}_{\hat{A}\hat{B}}(v,\theta)+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}_{\hat{A}\hat{B}}r^{2}-\frac{1}{2r^{2}}T\left(\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}_{\hat{A}\hat{B}}r^{5}}{\Omega^{2}}\right)+\int_{v}^{u}\partial_{v}\left[\frac{1}{2r^{2}}\partial_{u}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}_{\hat{A}\hat{B}}r^{5}}{\Omega^{2}}\right](\bar{u},v,\theta)d\bar{u}\right)\,.

We also have

(Ωtrχ¯)Ωχ^(1)A^B^+(Ωtrχ)Ωχ^¯(1)A^B^=+2Ω4r3(XA^B^(v,θ)\displaystyle\left(\Omega tr\underline{\chi}\right)\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}_{\hat{A}\hat{B}}+\left(\Omega tr\chi\right)\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}_{\hat{A}\hat{B}}=+\frac{2\Omega^{4}}{r^{3}}\Big{(}X_{\hat{A}\hat{B}}(v,\theta) 12α¯(1)A^B^r5Ω21r2(u,v,θ)+vu[12r2uα¯(1)A^B^r5Ω2](u¯,v,θ)𝑑u¯\displaystyle-\frac{1}{2}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}_{\hat{A}\hat{B}}r^{5}}{\Omega^{2}}\frac{1}{r^{2}}(u,v,\theta)+\int_{v}^{u}\left[\frac{1}{2r^{2}}\partial_{u}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}_{\hat{A}\hat{B}}r^{5}}{\Omega^{2}}\right](\bar{u},v,\theta)d\bar{u}
XA^B^(u,θ)\displaystyle-X_{\hat{A}\hat{B}}(u,\theta) 12α(1)A^B^r5Ω21r2(u,v,θ)vu[12r2vα(1)A^B^r5Ω2](u,v¯,θ)dv¯),\displaystyle-\frac{1}{2}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}_{\hat{A}\hat{B}}r^{5}}{\Omega^{2}}\frac{1}{r^{2}}(u,v,\theta)-\int_{v}^{u}\left[\frac{1}{2r^{2}}\partial_{v}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}_{\hat{A}\hat{B}}r^{5}}{\Omega^{2}}\right](u,\bar{v},\theta)d\bar{v}\,\Big{)},

from which (175) is already immediate after using Taylor’s theorem (as well as (7))

r(u,v)[X(v,θ)X(u,θ)]u,v|r(u,v)(vu)|supuX˙(u,θ)u,usupuT(rχ^(1))u,u\|r(u,v)\left[X(v,\theta)-X(u,\theta)\right]\|_{u,v}\lesssim\big{|}r(u,v)(v-u)\big{|}\sup_{u}\|\dot{X}(u,\theta)\|_{u,u}\lesssim\sup_{u}\|T(r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}})\|_{u,u}\

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 X˙\dot{X}-difference.

To prove (177), the key is (besides applying (175) and estimates from Theorem 4.2) to establish

supr8Mr2(X˙(u,θ)+α¯(1)r2+X˙(v,θ)+α(1)r2)+2r3(X(v,θ)12α¯(1)r5Ω21r2X(u,θ)12α(1)r5Ω21r2)\displaystyle\sup_{r\geq 8M}\Big{\|}r^{2}\left(\dot{X}(u,\theta)+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}r^{2}+\dot{X}(v,\theta)+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}r^{2}\right)+2r^{3}\left(X(v,\theta)-\frac{1}{2}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}r^{5}}{\Omega^{2}}\frac{1}{r^{2}}-X(u,\theta)-\frac{1}{2}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}r^{5}}{\Omega^{2}}\frac{1}{r^{2}}\right)\Big{\|}
sup{r8M}r3α(1)u,v+sup{r8M}r3α¯(1)u,v+supuT3(rχ^(1))u,u.\displaystyle\lesssim\sup_{\mathcal{M}\cap\{r\geq 8M\}}\|r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}}\|_{u,v}+\sup_{\mathcal{M}\cap\{r\geq 8M\}}\|r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}\|_{u,v}+\sup_{u}\|T^{3}(r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}})\|_{u,u}\,. (179)

This follows from Taylor expanding

X(u,θ)\displaystyle X(u,\theta) =X(v,θ)+X˙(v,θ)(vu)+12X¨(v,θ)(vu)2+RX,\displaystyle=X(v,\theta)+\dot{X}(v,\theta)(v-u)+\frac{1}{2}\ddot{X}(v,\theta)(v-u)^{2}+R_{X}\,,
X˙(u,θ)\displaystyle\dot{X}(u,\theta) =X˙(v,θ)+X¨(v,θ)(vu)+R~X\displaystyle=\dot{X}(v,\theta)+\ddot{X}(v,\theta)(v-u)+\tilde{R}_{X}

and using (7) as well as Taylor’s theorem for the remainders. Note that Proposition 5.8 controls at most three TT-derivatives of χ^(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}} 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 η(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}, η¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}} 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 χ^(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}} and χ^¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}} in Propositions 5.8 and 5.9 respectively, will provide estimates on β(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}} and β¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}} using Theorem 4.2 and Corollary 4.5. For now we state these (immediate) estimates on spheres (recall that the 1\ell\leq 1 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 Su,v2S^{2}_{u,v} for n3n\geq 3 the estimates

i=0n1𝒜[i](Ωr2β(1))u,v2+i=0n1𝒜[i](Ω1r4β¯(1))u,v2\displaystyle\sum_{i=0}^{n-1}\|\mathcal{A}^{[i]}(\Omega r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}})\|_{u,v}^{2}+\sum_{i=0}^{n-1}\|\mathcal{A}^{[i]}(\Omega^{-1}r^{4}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}})\|_{u,v}^{2} 𝔼¯datan[α(1),α¯(1)]+𝔻0n,\displaystyle\lesssim\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}\,, (180)
i=0n1𝒜[i](Ωr2β(1))u,v2+i=0n1𝒜[i](Ω1r4β¯(1))u,v2\displaystyle\sum_{i=0}^{n-1}\|\mathcal{A}^{[i]}(\Omega r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}})\|_{u,v}^{2}+\sum_{i=0}^{n-1}\|\mathcal{A}^{[i]}(\Omega^{-1}r^{4}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}})\|_{u,v}^{2} 𝔼¯datan+1[α(1),α¯(1)]+𝔻0n+1(logv)2.\displaystyle\lesssim\frac{\overline{\mathbb{E}}^{n+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+1}_{0}}{(\log v)^{2}}\,. (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 +\mathcal{H}^{+} and \mathcal{I}) Ω2r5α¯(1)\Omega^{-2}r^{5}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}, Ω∇̸4(Ω2r5α¯(1))\Omega\not{\nabla}_{4}(\Omega^{-2}r^{5}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}) and [Ω∇̸4]2(Ω2r5α¯(1))[\Omega\not{\nabla}_{4}]^{2}(\Omega^{-2}r^{5}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}) with smooth and uniformly bounded coefficients. Combining this with fact the estimate (175) we directly obtain:

Proposition 5.12.

We have on any sphere Su,v2S^{2}_{u,v} for n3n\geq 3 the estimates:

i=0n3𝒜[i]r521(ρ(1),σ(1))u,v2\displaystyle\sum_{i=0}^{n-3}\|\mathcal{A}^{[i]}r^{5}\not{\mathcal{D}}_{2}^{\star}\not{\mathcal{D}}_{1}^{\star}\left(-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}\right)\|_{u,v}^{2} 𝔼¯datan[α(1),α¯(1)]+𝔻0n,\displaystyle\lesssim\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}\,, (182)
i=0n3𝒜[i]r521(ρ(1),σ(1))u,v2\displaystyle\sum_{i=0}^{n-3}\|\mathcal{A}^{[i]}r^{5}\not{\mathcal{D}}_{2}^{\star}\not{\mathcal{D}}_{1}^{\star}\left(-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}\right)\|_{u,v}^{2} 𝔼¯datan+1[α(1),α¯(1)]+𝔻0n+1(logv)2.\displaystyle\lesssim\frac{\overline{\mathbb{E}}^{n+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+1}_{0}}{(\log v)^{2}}\,. (183)

We also have the top-order ingoing flux bound

i=0n2vu𝑑u¯Ω2r2𝒜[i]r521(ρ(1),σ(1))u,v¯2𝔼¯datan[α(1),α¯(1)]+𝔻0n\displaystyle\sum_{i=0}^{n-2}\int_{v}^{u}d\bar{u}\frac{\Omega^{2}}{r^{2}}\|\mathcal{A}^{[i]}r^{5}\not{\mathcal{D}}_{2}^{\star}\not{\mathcal{D}}_{1}^{\star}\left(-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}\right)\|_{u,\bar{v}}^{2}\lesssim\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}\, (184)

and the top-order outgoing flux bound

i=0n2v1v2𝑑v¯𝒜[i]r521(ρ(1),σ(1))u,v¯2(v2v1)[𝔼¯datan[α(1),α¯(1)]+𝔻0n].\displaystyle\sum_{i=0}^{n-2}\int_{v_{1}}^{v_{2}}d\bar{v}\|\mathcal{A}^{[i]}r^{5}\not{\mathcal{D}}_{2}^{\star}\not{\mathcal{D}}_{1}^{\star}\left(-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}\right)\|_{u,\bar{v}}^{2}\lesssim\left(v_{2}-v_{1}\right)\left[\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}\right]\,. (185)

Finally, for any fixed r0>r+r_{0}>r_{+} and the sphere Su,v12S^{2}_{u,v_{1}} lying in the region rr0r\geq r_{0} we have the uniform estimate

i=0n2v1v2𝑑v¯𝒜[i]r521(ρ(1),σ(1))u,v¯2r0𝔼¯datan[α(1),α¯(1)]+𝔻0n.\displaystyle\sum_{i=0}^{n-2}\int_{v_{1}}^{v_{2}}d\bar{v}\|\mathcal{A}^{[i]}r^{5}\not{\mathcal{D}}_{2}^{\star}\not{\mathcal{D}}_{1}^{\star}\left(-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}\right)\|_{u,\bar{v}}^{2}\lesssim_{r_{0}}\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}\,. (186)
Proof.

The bounds on spheres follow directly from the identity (97), estimate (175) of Proposition 5.10 and Corollary 4.5. For the fluxes one uses Theorem 4.2 instead of Corollary 4.5. ∎

5.5.3 Estimates on the torsion

Proposition 5.13.

For any v0v\geq 0 and n3n\geq 3 and k{0,1}k\in\{0,1\} and i{0,1}i\in\{0,1\}

j=0n1+i[∇̸T]k𝒜[j](r2iη(1),r2iη¯(1))u,v2\displaystyle\sum_{j=0}^{n-1+i}\big{\|}[\not{\nabla}_{T}]^{k}\mathcal{A}^{[j]}(r^{2-i}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}},r^{2-i}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}})\big{\|}^{2}_{u,v} 𝔼¯datan+k[α(1),α¯(1)]+𝔻0n+k,\displaystyle\lesssim\overline{\mathbb{E}}^{n+k}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+k}_{0}\,, (187)
j=0n1i[∇̸T]k𝒜[j](r2iη(1),r2iη¯(1))u,v2\displaystyle\sum_{j=0}^{n-1-i}\big{\|}[\not{\nabla}_{T}]^{k}\mathcal{A}^{[j]}(r^{2-i}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}},r^{2-i}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}})\big{\|}^{2}_{u,v} 𝔼¯datan+k+1[α(1),α¯(1)]+𝔻0n+k+1(logv)2.\displaystyle\lesssim\frac{\overline{\mathbb{E}}^{n+k+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+k+1}_{0}}{(\log v)^{2}}\,. (188)

We also have

j=0n2𝒜[j](r3η(1)+r3η¯(1))u,v2\displaystyle\sum_{j=0}^{n-2}\big{\|}\mathcal{A}^{[j]}(r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}+r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}})\big{\|}^{2}_{u,v} 𝔼¯datan[α(1),α¯(1)]+𝔻0n,\displaystyle\lesssim\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}\,, (189)
j=0n2𝒜[j](r3η(1)+r3η¯(1))u,v2\displaystyle\sum_{j=0}^{n-2}\big{\|}\mathcal{A}^{[j]}(r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}+r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}})\big{\|}^{2}_{u,v} 𝔼¯datan+1[α(1),α¯(1)]+𝔻0n+1(logv)2.\displaystyle\lesssim\frac{\overline{\mathbb{E}}^{n+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+1}_{0}}{(\log v)^{2}}\,. (190)
Proof.

We show the estimate for k=0k=0. For i=1i=1, (187) and (188) for (rη(1),rη¯(1))(r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}},r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}) 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 r8Mr\leq 8M, so we can focus on establishing the estimates for i=0i=0 in the region r8Mr\geq 8M for the remainder of the proof.

Replacing (r2η(1),r2η¯(1))(r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}},r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}) by (r2η(1)r2η¯(1))(r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}-r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}) (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 r2η(1)r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}} and r2η¯(1)r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}} individually) we integrate (61) backwards from the boundary, where η(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}} and η¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}} are known to vanish by having established control on (rη(1),rη¯(1))(r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}},r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}) at the beginning of the proof. Inserting the estimate for (r2η(1)r2η¯(1))(r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}-r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}) already established and Proposition 5.11 to control the right hand side, the estimates (187) and (188) follow.

The last two estimates follows from adding (58) and (59) and using the estimate (177). ∎

5.5.4 Estimates on the lapses

The following is an immediate corollary of Proposition 5.13:

Corollary 5.14.

For any v0v\geq 0 and n3n\geq 3

j=0n1r𝒜[j][r∇̸](Ω1Ω(1))u,v2+j=0n2r2𝒜[j][r∇̸](Ω1Ω(1))u,v2\displaystyle\sum_{j=0}^{n-1}\big{\|}r\mathcal{A}^{[j]}[r\not{\nabla}](\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega})\big{\|}^{2}_{u,v}+\sum_{j=0}^{n-2}\big{\|}r^{2}\mathcal{A}^{[j]}[r\not{\nabla}](\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega})\big{\|}^{2}_{u,v} 𝔼¯datan[α(1),α¯(1)]+𝔻0n,\displaystyle\lesssim\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}\,,
j=0n1r𝒜[j][r∇̸](Ω1Ω(1))u,v2+j=0n2r2𝒜[j][r∇̸](Ω1Ω(1))u,v2\displaystyle\sum_{j=0}^{n-1}\big{\|}r\mathcal{A}^{[j]}[r\not{\nabla}](\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega})\big{\|}^{2}_{u,v}+\sum_{j=0}^{n-2}\big{\|}r^{2}\mathcal{A}^{[j]}[r\not{\nabla}](\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega})\big{\|}^{2}_{u,v} 𝔼¯datan+1[α(1),α¯(1)]+𝔻0n+1(logv)2.\displaystyle\lesssim\frac{\overline{\mathbb{E}}^{n+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+1}_{0}}{(\log v)^{2}}\,.
Proof.

All estimates follows straight from the definition 2∇̸(Ω1Ω(1))=η(1)+η¯(1)2\not{\nabla}(\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega})=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}} and using Proposition 5.13. ∎

An estimate for ω¯(1),ω(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega} is easily obtained from the relations

2∇̸Tη(1)=Ωβ(1)Ωβ¯(1)+Ω2r(η(1)+η¯(1))+2∇̸ω¯(1), 2∇̸Tη¯(1)=Ωβ(1)+Ωβ¯(1)Ω2r(η(1)+η¯(1))+2∇̸ω(1)\displaystyle 2\not{\nabla}_{T}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}=-\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}-\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}+\frac{\Omega^{2}}{r}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}})+2\not{\nabla}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}\ \ \ ,\ \ \ 2\not{\nabla}_{T}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}=\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}+\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}-\frac{\Omega^{2}}{r}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}})+2\not{\nabla}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}\, (191)

and previous bounds on the geometric quantities:

Proposition 5.15.

For any v0v\geq 0 and n3n\geq 3 and i{0,1}i\in\{0,1\}

j=0n2ri𝒜[j][r∇̸]ω¯(1)r2Ω2u,v2+j=0n2ri𝒜[j][r∇̸]ω(1)u,v2\displaystyle\sum_{j=0}^{n-2}\big{\|}r^{i}\mathcal{A}^{[j]}[r\not{\nabla}]\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}r^{2}\Omega^{-2}\big{\|}^{2}_{u,v}+\sum_{j=0}^{n-2}\big{\|}r^{i}\mathcal{A}^{[j]}[r\not{\nabla}]\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}\big{\|}^{2}_{u,v} 𝔼¯datan+i[α(1),α¯(1)]+𝔻0n+i,\displaystyle\lesssim\overline{\mathbb{E}}^{n+i}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+i}_{0}\,,
j=0n3ri𝒜[j][r∇̸]ω¯(1)r2Ω2u,v2+j=0n3ri𝒜[j][r∇̸]ω(1)u,v2\displaystyle\sum_{j=0}^{n-3}\big{\|}r^{i}\mathcal{A}^{[j]}[r\not{\nabla}]\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}r^{2}\Omega^{-2}\big{\|}^{2}_{u,v}+\sum_{j=0}^{n-3}\big{\|}r^{i}\mathcal{A}^{[j]}[r\not{\nabla}]\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}\big{\|}^{2}_{u,v} 𝔼¯datan+i[α(1),α¯(1)]+𝔻0n+i(logv)2.\displaystyle\lesssim\frac{\overline{\mathbb{E}}^{n+i}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+i}_{0}}{(\log v)^{2}}\,.
Remark 5.16.

In particular ω(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega} and ω¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}} vanish on \mathcal{I}. Summing the relations (191) one can also show that rω(1)+rω¯(1)r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}+r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}} vanishes on \mathcal{I} as follows from r2∇̸T(η(1)+η¯(1))r^{2}\not{\nabla}_{T}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}) vanishing on \mathcal{I}. However, this requires controlling one more TT-derivative on data and is hence omitted. See also Remark 2.7.

5.6 Estimates on the expansion and improving the regularity

We write the linearised Raychaudhuri equation (55) as

Ω∇̸4((Ωtrχ)(1)r2Ω24rΩ1Ω(1))=4Ω2(Ω1Ω(1)).\displaystyle\Omega\not{\nabla}_{4}\left(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}r^{2}\Omega^{-2}-4r\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}\right)=-4\Omega^{2}(\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega})\,. (192)

Commuting with nn angular derivatives and also with Ω1∇̸3\Omega^{-1}\not{\nabla}_{3} we deduce (note the factor of Ω2\Omega^{-2}):

Proposition 5.17.

For n3n\geq 3 we have on any sphere Su,v2S^{2}_{u,v} for i{0,1}i\in\{0,1\}:

j=0n1i𝒜[j][r∇̸](Ωtrχ)(1)Ω2r2(logr)1+iu,v2\displaystyle\sum_{j=0}^{n-1-i}\big{\|}\mathcal{A}^{[j]}[r\not{\nabla}]\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\Omega^{-2}r^{2}\xcancel{(\log r)^{-1+i}}\big{\|}^{2}_{u,v} 𝔼¯datan[α(1),α¯(1)]+𝔻0n,\displaystyle\lesssim\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}\,, (193)
j=0n2i𝒜[j][r∇̸](Ωtrχ)(1)Ω2r2(logr)1+iu,v2\displaystyle\sum_{j=0}^{n-2-i}\big{\|}\mathcal{A}^{[j]}[r\not{\nabla}]\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\Omega^{-2}r^{2}\xcancel{(\log r)^{-1+i}}\big{\|}^{2}_{u,v} 𝔼¯datan[α(1),α¯(1)]+𝔻0n(logv)2.\displaystyle\lesssim\frac{\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}}{(\log v)^{2}}\,. (194)

We also have

j=0n2iri+1Ω2[Ω∇̸3]𝒜[j][r∇̸](Ωtrχ)(1)u,v2\displaystyle\sum_{j=0}^{n-2-i}\big{\|}\frac{r^{i+1}}{\Omega^{2}}[\Omega\not{\nabla}_{3}]\mathcal{A}^{[j]}[r\not{\nabla}]\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\big{\|}^{2}_{u,v} 𝔼¯datan[α(1),α¯(1)]+𝔻0n,\displaystyle\lesssim\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}\,, (195)
j=0n3iri+1Ω2[Ω∇̸3]𝒜[j][r∇̸](Ωtrχ)(1)u,v2\displaystyle\sum_{j=0}^{n-3-i}\big{\|}\frac{r^{i+1}}{\Omega^{2}}[\Omega\not{\nabla}_{3}]\mathcal{A}^{[j]}[r\not{\nabla}]\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\big{\|}^{2}_{u,v} 𝔼¯datan[α(1),α¯(1)]+𝔻0n(logv)2.\displaystyle\lesssim\frac{\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}}{(\log v)^{2}}\,. (196)
Proof.

With the log\log-terms absent (for i=1i=1), 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 r2r^{2} weight. Estimates (195), (196) are obtained similarly, by commuting (192) with Ω∇̸3\Omega\not{\nabla}_{3} and using the estimates of Corollary 5.14 and Proposition 5.15 with the strong rr-weights. With the log\log-terms present (i.e. for i=0i=0), estimates (193), (194), (195), (196) follow along the same lines but using the estimates with the weaker rr-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 log\log-terms in (193) and (194) and the claimed weights in (195), (196). The log\log-terms will be removed immediately after the next proposition, which only uses the estimates of the current proposition with the log\log-terms. ∎

The above estimate (with the log\log-terms) leads to an improvement Proposition 5.8 via the Codazzi equations and previous bounds:

Proposition 5.18.

For any v0v\geq 0 and n3n\geq 3, k{0,1}k\in\{0,1\}

i+jn,i2𝒜[j][∇̸T]k[r2Ω1∇̸3]i(Ωχ^(1)))2u,v\displaystyle\sum_{i+j\leq n,i\leq 2}\big{\|}\mathcal{A}^{[j]}[\not{\nabla}_{T}]^{k}\left[r^{2}\Omega^{-1}\not{\nabla}_{3}\right]^{i}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}))\big{\|}^{2}_{u,v} 𝔼¯datan+k[α(1),α¯(1)]+𝔻0n+k,\displaystyle\lesssim\overline{\mathbb{E}}^{n+k}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+k}_{0}\,, (197)
i+jn,i2𝒜[j][∇̸T]k[r2Ω1∇̸3]i(Ωχ^(1)))2u,v\displaystyle\sum_{i+j\leq n,i\leq 2}\big{\|}\mathcal{A}^{[j]}[\not{\nabla}_{T}]^{k}\left[r^{2}\Omega^{-1}\not{\nabla}_{3}\right]^{i}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}))\big{\|}^{2}_{u,v} 𝔼¯datan+k+1[α(1),α¯(1)]+𝔻0n+k+1(logv)2.\displaystyle\lesssim\frac{\overline{\mathbb{E}}^{n+k+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+k+1}_{0}}{(\log v)^{2}}\,. (198)
Proof.

For the term i=2i=2 in the sum on the left, the estimate is a direct consequence of Proposition 5.8 so we focus on i1i\leq 1. We write the Codazzi equation (60) as

rivΩχ^(1)=Ω2η¯(1)Ωrβ(1)+r2∇̸(Ωtrχ)(1).\displaystyle r\not{div}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}=-\Omega^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}-\Omega r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}+\frac{r}{2}\not{\nabla}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\,. (199)

Applying Ω1∇̸3\Omega^{-1}\not{\nabla}_{3} to both sides and inserting the relevant Bianchi and null structure equations, we obtain

riv(Ω1∇̸3(Ωχ^(1)))=(2k2r+2Mr2)η¯(1)+(Ω2r(η(1)η¯(1))Ωβ¯(1))+Ω1∇̸3(r2∇̸(Ωtrχ)(1))\displaystyle r\not{div}\left(\Omega^{-1}\not{\nabla}_{3}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}})\right)=\left(2k^{2}r+\frac{2M}{r^{2}}\right)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}+\left(\frac{\Omega^{2}}{r}\left(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}\right)-\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}\right)+\Omega^{-1}\not{\nabla}_{3}\left(\frac{r}{2}\not{\nabla}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\right)
(r1(ρ(1),σ(1))+3ρrη(1))+β(1)Ω.\displaystyle-\left(r\not{\mathcal{D}}_{1}^{\star}\left(-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}\,,\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}\,\right)+3\rho r\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}\right)+\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}\Omega\,. (200)

Therefore, with non-optimal rr-weights (insert a factor of 1r2\frac{1}{r^{2}} 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 r0r_{0} as in the proof of Proposition 5.8. ∎

Corollary 5.19.

The first two estimates of Proposition 5.17 hold without the logarithmic term.

Proof.

Estimate (Ωtrχ)(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)} from (199) now using the improved estimate on Ωχ^(1)\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}} from Proposition 5.18. ∎

We can now also improve the estimate on χ^¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}} of Proposition 5.9 using that we now control more derivatives of χ^(1)r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}r and hence (by the boundary condition χ^(1)r=χ^¯(1)r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}r=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}r which holds with arbitrary many tangential derivatives by the smoothness of the solution) of χ^¯(1)r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}r on the boundary.

Proposition 5.20.

For any v0v\geq 0 and n3n\geq 3

i+jn,i2𝒜[j][Ω∇̸4]i(r2Ω1χ^¯(1)))2u,v\displaystyle\sum_{i+j\leq n,i\leq 2}\big{\|}\mathcal{A}^{[j]}\left[\Omega\not{\nabla}_{4}\right]^{i}(r^{2}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}))\big{\|}^{2}_{u,v} 𝔼¯datan[α(1),α¯(1)]+𝔻0n,\displaystyle\lesssim\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}\,, (201)
i+jn,i2𝒜[j][Ω∇̸4]i(r2Ω1χ^¯(1)))2u,v\displaystyle\sum_{i+j\leq n,i\leq 2}\big{\|}\mathcal{A}^{[j]}\left[\Omega\not{\nabla}_{4}\right]^{i}(r^{2}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}))\big{\|}^{2}_{u,v} 𝔼¯datan+1[α(1),α¯(1)]+𝔻0n+1(logv)2.\displaystyle\lesssim\frac{\overline{\mathbb{E}}^{n+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+1}_{0}}{(\log v)^{2}}\,. (202)
Proof.

Revisit the proof of Proposition 5.9 using that we control the higher order “initial” term on the boundary \mathcal{I} by the boundary condition and Proposition 5.18. ∎

A direct corollary, using (60) pointwise, is

Corollary 5.21.

For any v0v\geq 0 and n3n\geq 3

j=1n1𝒜[j](r∇̸Ω2r2(Ωtrχ¯)(1))u,v2\displaystyle\sum_{j=1}^{n-1}\big{\|}\mathcal{A}^{[j]}(r\not{\nabla}\Omega^{-2}r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)})\big{\|}^{2}_{u,v} 𝔼¯datan[α(1),α¯(1)]+𝔻0n,\displaystyle\lesssim\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}\,, (203)
j=1n2𝒜[j](r∇̸Ω2r2(Ωtrχ¯)(1))u,v2\displaystyle\sum_{j=1}^{n-2}\big{\|}\mathcal{A}^{[j]}(r\not{\nabla}\Omega^{-2}r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)})\big{\|}^{2}_{u,v} 𝔼¯datan[α(1),α¯(1)]+𝔻0n(logv)2.\displaystyle\lesssim\frac{\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}}{(\log v)^{2}}\,. (204)
Remark 5.22.

One can prove control on |r(Ωtrχ)(1)r(Ωtrχ¯)(1)||r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}-r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}| 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 rr-weights) and, as a corollary of the relation (191), also our estimate on ω(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}:

Proposition 5.23.

We have for k{0,1}k\in\{0,1\} the estimates

j=0nk𝒜[j][∇̸T]k(rη(1),rη¯(1))u,v2\displaystyle\sum_{j=0}^{n-k}\big{\|}\mathcal{A}^{[j]}[\not{\nabla}_{T}]^{k}(r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}},r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}})\big{\|}^{2}_{u,v} 𝔼¯datan[α(1),α¯(1)]+𝔻0n,\displaystyle\lesssim\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}\,, (205)
j=0nk𝒜[j][∇̸T]k(rη(1),rη¯(1))u,v2\displaystyle\sum_{j=0}^{n-k}\big{\|}\mathcal{A}^{[j]}[\not{\nabla}_{T}]^{k}(r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}},r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}})\big{\|}^{2}_{u,v} 𝔼¯datan+1[α(1),α¯(1)]+𝔻0n+1(logv)2.\displaystyle\lesssim\frac{\overline{\mathbb{E}}^{n+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+1}_{0}}{(\log v)^{2}}\,. (206)

Moreover,

j=0n𝒜[j](ω(1),ω¯(1))u,v2\displaystyle\sum_{j=0}^{n}\big{\|}\mathcal{A}^{[j]}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}})\big{\|}^{2}_{u,v} 𝔼¯datan[α(1),α¯(1)]+𝔻0n,\displaystyle\lesssim\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}\,, (207)
j=0n𝒜[j](ω(1),ω¯(1))u,v2\displaystyle\sum_{j=0}^{n}\big{\|}\mathcal{A}^{[j]}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}})\big{\|}^{2}_{u,v} 𝔼¯datan+1[α(1),α¯(1)]+𝔻0n+1(logv)2.\displaystyle\lesssim\frac{\overline{\mathbb{E}}^{n+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+1}_{0}}{(\log v)^{2}}\,. (208)

We finally obtain the top order flux bounds for β(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}} and β¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}} from the improved estimates on the shear of Propositions 5.20 and 5.18:

Proposition 5.24.

We have on any sphere Su,v2S^{2}_{u,v} for n3n\geq 3 the ingoing flux bounds:

i=0nvu𝑑u¯Ω2r2𝒜[i](Ωr2β(1),Ω1r4β¯(1))u¯,v2\displaystyle\sum_{i=0}^{n}\int_{v}^{u}d\bar{u}\frac{\Omega^{2}}{r^{2}}\|\mathcal{A}^{[i]}(\Omega r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}},\Omega^{-1}r^{4}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}})\|_{\bar{u},{v}}^{2} 𝔼¯datan[α(1),α¯(1)]+𝔻0n,\displaystyle\lesssim\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}\,, (209)
i=0nvu𝑑u¯Ω2r2𝒜[i](Ωr2β(1),Ω1r4β¯(1))u¯,v2\displaystyle\sum_{i=0}^{n}\int_{v}^{u}d\bar{u}\frac{\Omega^{2}}{r^{2}}\|\mathcal{A}^{[i]}(\Omega r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}},\Omega^{-1}r^{4}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}})\|_{\bar{u},{v}}^{2} 𝔼¯datan+1[α(1),α¯(1)]+𝔻0n+1(logv)2.\displaystyle\lesssim\frac{\overline{\mathbb{E}}^{n+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+1}_{0}}{(\log v)^{2}}\,. (210)

We also have the top-order outgoing flux bound

i=0nv1v2𝑑v¯𝒜[i](Ωr2β(1),Ω1r4β¯(1))u,v¯2(v2v1)[𝔼¯datan[α(1),α¯(1)]+𝔻0n].\displaystyle\sum_{i=0}^{n}\int_{v_{1}}^{v_{2}}d\bar{v}\|\mathcal{A}^{[i]}(\Omega r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}},\Omega^{-1}r^{4}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}})\|_{u,\bar{v}}^{2}\lesssim\left(v_{2}-v_{1}\right)\left[\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}\right]\,. (211)

Finally, for any fixed r0>r+r_{0}>r_{+} and the sphere Su,v12S^{2}_{u,v_{1}} lying in the region rr0r\geq r_{0} we have the uniform estimate

i=0nv1v2𝑑v¯𝒜[i](Ωr2β(1),Ω1r4β¯(1))u,v¯2r0𝔼¯datan[α(1),α¯(1)]+𝔻0n.\displaystyle\sum_{i=0}^{n}\int_{v_{1}}^{v_{2}}d\bar{v}\|\mathcal{A}^{[i]}(\Omega r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}},\Omega^{-1}r^{4}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}})\|_{u,\bar{v}}^{2}\lesssim_{r_{0}}\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}\,. (212)
Proof.

This is a direct consequence of the relations (94), the estimates on α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} and α¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}} of Theorem 4.2 and the estimates on the shear of Propositions 5.20 and 5.18. For the last statement recall that Ω2cr0>0\Omega^{2}\geq c_{r_{0}}>0 for r0>r+r_{0}>r_{+}. ∎

5.7 Estimates on the metric components

We can now integrate the propagation equation (51) for b(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b} 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 Su,v2S^{2}_{u,v} and n3n\geq 3

j=0n𝒜[j](rΩ2b(1)))2u,v\displaystyle\sum_{j=0}^{n}\big{\|}\mathcal{A}^{[j]}(r\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}))\big{\|}^{2}_{u,v} 𝔼¯datan[α(1),α¯(1)]+𝔻0n,\displaystyle\lesssim\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}\,, (213)
j=0n𝒜[j](rΩ2b(1)))2u,v\displaystyle\sum_{j=0}^{n}\big{\|}\mathcal{A}^{[j]}(r\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}))\big{\|}^{2}_{u,v} 𝔼¯datan+1[α(1),α¯(1)]+𝔻0n+1(logv)2.\displaystyle\lesssim\frac{\overline{\mathbb{E}}^{n+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+1}_{0}}{(\log v)^{2}}\,. (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 Su,v2S^{2}_{u,v} and n3n\geq 3

j=0n1𝒜[j][r∇̸](1)u,v2\displaystyle\sum_{j=0}^{n-1}\bigg{\|}\mathcal{A}^{[j]}[r\not{\nabla}]\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}\bigg{\|}^{2}_{u,v} 𝔼¯datan[α(1),α¯(1)]+𝔻0n,\displaystyle\lesssim\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}\,, (215)
j=0n1𝒜[j][r∇̸](1)u,v2\displaystyle\sum_{j=0}^{n-1}\bigg{\|}\mathcal{A}^{[j]}[r\not{\nabla}]\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}\bigg{\|}^{2}_{u,v} 𝔼¯datan+1[α(1),α¯(1)]+𝔻0n+1(logv)2.\displaystyle\lesssim\frac{\overline{\mathbb{E}}^{n+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+1}_{0}}{(\log v)^{2}}\,. (216)

To estimate ^(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}} we cannot integrate (50) directly as Ωχ^(1)\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}} is not uniformly in Lv1L^{1}_{v}. We instead estimate it from the Gauss curvature.

Proposition 5.27.

For any sphere Su,v2S^{2}_{u,v} and n3n\geq 3

j=0n𝒜[j]^(1)u,v2\displaystyle\sum_{j=0}^{n}\bigg{\|}\mathcal{A}^{[j]}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}\bigg{\|}^{2}_{u,v} 𝔼¯datan[α(1),α¯(1)]+𝔻0n,\displaystyle\lesssim\overline{\mathbb{E}}^{n}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n}_{0}\,, (217)
j=0n𝒜[j]^(1)u,v2\displaystyle\sum_{j=0}^{n}\bigg{\|}\mathcal{A}^{[j]}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}\bigg{\|}^{2}_{u,v} 𝔼¯datan+1[α(1),α¯(1)]+𝔻0n+1(logv)2.\displaystyle\lesssim\frac{\overline{\mathbb{E}}^{n+1}_{data}[\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}]+\mathbb{D}^{n+1}_{0}}{(\log v)^{2}}\,. (218)
Proof.

By elliptic estimates, it suffices to prove these estimates replacing 𝒜[j]\mathcal{A}^{[j]} by 𝒜[j2]r2iviv\mathcal{A}^{[j-2]}r^{2}\not{div}\not{div} and 𝒜[j2]r2urliv\mathcal{A}^{[j-2]}r^{2}\not{curl}\not{div} in each sum and letting the sum start at j=2j=2. (We slightly abuse notation here and let 𝒜\mathcal{A} also act on scalars by taking r1r\not{\mathcal{D}}_{1}^{\star}, and on one forms by taking r2r\not{\mathcal{D}}_{2}^{\star}.) For the latter part we can integrate (50) commuted with 𝒜[j2]r2urliv\mathcal{A}^{[j-2]}r^{2}\not{curl}\not{div} because from the Codazzi equation (60) we have

𝒜[j2]r2urliv(Ωχ^(1))=𝒜[j2](Ω2rσ(1)r2urlΩβ(1))\mathcal{A}^{[j-2]}r^{2}\not{curl}\not{div}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}})=\mathcal{A}^{[j-2]}\left(\Omega^{2}r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}-r^{2}\not{curl}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}\right)

and hence

Ω∇̸4(𝒜[j2]r2urliv^(1)2𝒜[j2]r3σ(1))=2Ω2𝒜[j2]rσ(1).\displaystyle\Omega\not{\nabla}_{4}\left(\mathcal{A}^{[j-2]}r^{2}\not{curl}\not{div}\,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}-2\mathcal{A}^{[j-2]}r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}\right)=2\Omega^{2}\mathcal{A}^{[j-2]}r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}\,. (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 urliv\not{curl}\not{div}-part. For the iviv\not{div}\not{div}-part we use the linearised Gauss equation:

r3∇̸(12Δ̸tr(1)+iviv^(1)1r2tr(1))=r3∇̸K(1)=r3∇̸ρ(1)r2iv(Ωχ^¯(1))+r2Ωβ¯(1)+r2iv(Ωχ^(1))+r2Ωβ(1).\displaystyle r^{3}\not{\nabla}\left(-\frac{1}{2}\not{\Delta}tr_{\not{g}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\not{g}}+\not{div}\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}-\frac{1}{r^{2}}tr_{\not{g}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\not{g}}\right)=r^{3}\not{\nabla}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{K}=-r^{3}\not{\nabla}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}-r^{2}\not{div}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}})+r^{2}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}+r^{2}\not{div}(\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}})+r^{2}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}\,.

The estimate now follows by solving this for r3∇̸iviv^(1)r^{3}\not{\nabla}\not{div}\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}} (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

The estimates (143) and (144) claimed in Theorem 4.7 for the Ricci-coefficients ξ\xi are implied

  • for χ^(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}} and χ^¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}} by Propositions 5.18 and 5.20,

  • for (Ωtrχ)(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)} and (Ωtrχ¯)(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)} by Proposition 5.17 and Corollary 5.21,

  • for η(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}} and η¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}} by Propositions 5.13 and 5.23, and for ω(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega} and ω¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}} by 5.15 and 5.23,

  • for the metric quantities Ω1Ω(1)\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega} by Corollary 5.14, for b(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b} by Proposition 5.25 and for ^(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}, (1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}} by Propositions 5.26 and 5.27 respectively.

The estimates (143) and (144) claimed in Theorem 4.7 for the curvature components Ξ\Xi are implied by Corollary 4.5 for α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} and α¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}, by Proposition 5.12 for ρ(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho} and σ(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}} and by Proposition 5.24 for β(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}} and β¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}.

Finally, the top order curvature bounds claimed in (145) and (146) are implied by Theorem 4.2 for α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} and α¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}}, by Proposition 5.12 for ρ(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho} and σ(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}} and by Proposition 5.24 for β(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}} and β¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\beta}}.

6 Normalising the solution at infinity: Proof of Theorem 4.14

We now prove Theorem 4.14. We consider the solution 𝒮\underaccent{\lor}{\mathscr{S}} of Theorem 4.7. We define a function f:0+×S2f:\mathbb{R}_{0}^{+}\times S^{2}\rightarrow\mathbb{R}, supported on 2\ell\geq 2 as follows. Define for u0u\geq 0 the limit rχ^(1)𝒮(u,θ,ϕ)=limvurχ^(1)𝒮(u,v,θ,ϕ)r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}_{\underaccent{\lor}{\mathscr{S}}}^{\infty}(u,\theta,\phi)=\lim_{v\rightarrow u}r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}_{\underaccent{\lor}{\mathscr{S}}}(u,v,\theta,\phi), which is the (smooth) restriction to \mathcal{I} of the weighted tensor rχ^(1)𝒮r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}_{\underaccent{\lor}{\mathscr{S}}}. Similarly, define for v0v\geq 0, the limit rχ^¯(1)𝒮(v,θ,ϕ)=limuvrχ^¯(1)𝒮(u,v,θ,ϕ)r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}_{\underaccent{\lor}{\mathscr{S}}}^{\infty}(v,\theta,\phi)=\lim_{u\rightarrow v}r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}_{\underaccent{\lor}{\mathscr{S}}}(u,v,\theta,\phi), which is the (smooth) restriction to \mathcal{I} of the weighted tensor rχ^¯(1)𝒮r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}_{\underaccent{\lor}{\mathscr{S}}}. By the boundary condition we have rχ^(1)𝒮(t,θ,ϕ)=rχ^¯(1)𝒮(t,θ,ϕ)r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}_{\underaccent{\lor}{\mathscr{S}}}^{\infty}(t,\theta,\phi)=r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}_{\underaccent{\lor}{\mathscr{S}}}^{\infty}(t,\theta,\phi) for t0t\geq 0. We finally define a function ff by solving for each tt the elliptic (since 2\ell\geq 2) scalar equation171717One computes r2ivivr22∇̸=r4iv(12Δ̸12K)∇̸=r4iv(12(11K)12K)∇̸=r4(12Δ̸2KΔ̸)r^{2}\not{div}\not{div}r^{2}\not{\mathcal{D}}_{2}^{\star}\not{\nabla}=r^{4}\not{div}\left(-\frac{1}{2}\not{\Delta}-\frac{1}{2}K\right)\not{\nabla}=r^{4}\not{div}\left(\frac{1}{2}\left(\not{\mathcal{D}}_{1}^{\star}\not{\mathcal{D}}_{1}-K\right)-\frac{1}{2}K\right)\not{\nabla}=r^{4}\left(\frac{1}{2}\not{\Delta}^{2}-K\not{\Delta}\right).

r2ivivrχ^(1)𝒮(t,θ,ϕ)=2kr2ivivr22∇̸f(t,θ,ϕ).\displaystyle r^{2}\not{div}\not{div}\,r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}_{\underaccent{\lor}{\mathscr{S}}}^{\infty}(t,\theta,\phi)=-2kr^{2}\not{div}\not{div}r^{2}\not{\mathcal{D}}_{2}^{\star}\not{\nabla}f(t,\theta,\phi). (220)

The function ff generates a pure gauge solution 𝒢f\mathscr{G}_{f} according to Lemma 2.8 and using the notation of that lemma we have

r2ivivrχ^(1)𝒢f(u,θ,ϕ)=2Ωrivivr22∇̸fu,r2ivivrχ^¯(1)𝒢f(v,θ,ϕ)=2Ωrivivr22∇̸fv.\displaystyle r^{2}\not{div}\not{div}\,r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}_{\mathscr{G}_{f}}(u,\theta,\phi)=-2\Omega r\not{div}\not{div}r^{2}\not{\mathcal{D}}_{2}^{\star}\not{\nabla}f_{u}\ \ \ \ ,\ \ \ r^{2}\not{div}\not{div}\,r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}_{\mathscr{G}_{f}}(v,\theta,\phi)=-2\Omega r\not{div}\not{div}r^{2}\not{\mathcal{D}}_{2}^{\star}\not{\nabla}f_{v}\,. (221)

Note also that from (175) and Propositions 5.18 and 5.20 holding for the solution 𝒮\underaccent{\lor}{\mathscr{S}} of Theorem 4.7, we have for n3n\geq 3 the quantitative estimates

i=0n+2|[r∇̸]ifu|+i=0n+2|[r∇̸]ifv|𝔼n,i=0n|r[r∇̸]i(fufv)|𝔼,n\displaystyle\sum_{i=0}^{n+2}|[r\not{\nabla}]^{i}f_{u}|+\sum_{i=0}^{n+2}|[r\not{\nabla}]^{i}f_{v}|\lesssim\overset{\circ}{\mathbb{E}}{}^{n}\ \ \ ,\ \ \ \sum_{i=0}^{n}|r[r\not{\nabla}]^{i}(f_{u}-f_{v})|\lesssim\overset{\circ}{\mathbb{E}}{}^{n}\,, (222)

together with the corresponding estimates with the 1(logv)2\frac{1}{(\log v)^{2}}-factor on the right-hand side.

Using (220) and (221), one can prove that 𝒮=𝒮+𝒢f{\accentset{\land}{\mathscr{S}}^{\prime}}=\underaccent{\lor}{\mathscr{S}}+\mathscr{G}_{f} satisfies rχ^(1)𝒮=rχ^¯(1)𝒮=0r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}_{{\accentset{\land}{\mathscr{S}}^{\prime}}}=r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}_{{\accentset{\land}{\mathscr{S}}^{\prime}}}=0 on \mathcal{I}. To see this we show separately that r2urlivrχ^(1)𝒮=0r^{2}\not{curl}\not{div}r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}_{{\accentset{\land}{\mathscr{S}}^{\prime}}}=0 and r2ivivrχ^(1)𝒮=0r^{2}\not{div}\not{div}r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}_{{\accentset{\land}{\mathscr{S}}^{\prime}}}=0 on \mathcal{I}, which implies the claim for χ^(1)𝒮\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}_{{\accentset{\land}{\mathscr{S}}^{\prime}}} by standard elliptic estimates. Indeed, this follows immediately by our choice of pure gauge solution for the iviv\not{div}\not{div} part. On the other hand, it is not hard to see that r2urlivχ^(1)r^{2}\not{curl}\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}} is actually gauge invariant and in fact equal to zero on \mathcal{I} (use the linearised Codazzi equation (60), the decay of β(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\beta}}, urlη(1)=σ(1)\not{curl}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}} and the boundary condition (80) for σ(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\sigma}}). The argument for χ^¯(1)𝒮\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}_{{\accentset{\land}{\mathscr{S}}^{\prime}}} is entirely analogous.

It now immediately follows that in the new gauge we can estimate r2Ωχ^(1)r^{2}\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}} and r2Ω1r2χ^¯(1)r^{2}\Omega^{-1}r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}} instead of Ωχ^(1)\Omega\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}} and Ω1r2χ^¯(1)\Omega^{-1}r^{2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}} in Theorem 4.7. Indeed, we can now integrate backwards in the 44- and 33-direction from the boundary \mathcal{I} using that χ^(1)r𝒮=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}r_{{\accentset{\land}{\mathscr{S}}^{\prime}}}=0 and χ^¯(1)r𝒮=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}r_{{\accentset{\land}{\mathscr{S}}^{\prime}}}=0 hold on the boundary in the new gauge and using the estimates on α(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\alpha}} and α¯(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\alpha}} from Theorem 4.2 and Corollary 4.5 just as in the proof of Propositions 5.8 and 5.9.

Next, since Ω∇̸3(χ^(1)r)𝒮=0\Omega\not{\nabla}_{3}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}r)_{{\accentset{\land}{\mathscr{S}}^{\prime}}}=0 on the boundary (from T(χ^(1)r)𝒮=0T(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}r)_{{\accentset{\land}{\mathscr{S}}^{\prime}}}=0 holding in the new gauge), integrating again backwards from the boundary one infers estimates for r2Ω∇̸3(χ^(1)r)𝒮r^{2}\Omega\not{\nabla}_{3}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}r)_{{\accentset{\land}{\mathscr{S}}^{\prime}}} and r2Ω∇̸4(χ^¯(1)r)𝒮r^{2}\Omega\not{\nabla}_{4}(\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}r)_{{\accentset{\land}{\mathscr{S}}^{\prime}}}.

Estimates for r3η(1)𝒮r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\eta}}_{{\accentset{\land}{\mathscr{S}}^{\prime}}} and r3η¯(1)𝒮r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\eta}}_{{\accentset{\land}{\mathscr{S}}^{\prime}}} are obtained directly by (58), (59), which also imply estimates for r2Ω1Ω(1)𝒮r^{2}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}_{{\accentset{\land}{\mathscr{S}}^{\prime}}} (modulo =0,1\ell=0,1 modes). Codazzi then gives control on r(Ωtrχ)(1)𝒮r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}_{{\accentset{\land}{\mathscr{S}}^{\prime}}} and r3Ω2(Ωtrχ¯)(1)𝒮r^{3}\Omega^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}_{{\accentset{\land}{\mathscr{S}}^{\prime}}} (modulo =0,1\ell=0,1 modes).

Inserting the above bounds, one can infer from the linearised Gauss equation (65) that the linearised Gauss curvature behaves like K(1)𝒮r3\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{K}_{{\accentset{\land}{\mathscr{S}}^{\prime}}}\sim r^{-3}.

One finally adds a 𝒢q\mathscr{G}_{q} pure gauge solution of Lemma 2.9 so that for 𝒮=𝒮+𝒢q\accentset{\land}{\mathscr{S}}={\accentset{\land}{\mathscr{S}}^{\prime}}+\mathscr{G}_{q}, one has ^(1)𝒮=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}_{\accentset{\land}{\mathscr{S}}}=0 on the initial sphere of the boundary and r1b(1)𝒮=0r^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{\accentset{\land}{\mathscr{S}}}=0 along the boundary. More specifically we define

(q1,q2)=0u(ΔS21ivb(1)𝒮,ΔS21urlb(1)𝒮)(u¯,u¯)𝑑u¯+(q¯1,q¯2)with 2r221(q¯1,q¯2)=^(1)𝒮(0,0).\displaystyle(q_{1},q_{2})=-\int_{0}^{u}\left(\Delta_{S^{2}}^{-1}\not{div}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{{\accentset{\land}{\mathscr{S}}^{\prime}}},-\Delta_{S^{2}}^{-1}\not{curl}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{{\accentset{\land}{\mathscr{S}}^{\prime}}}\right)\left(\bar{u},\bar{u}\right)d\bar{u}+(\bar{q}_{1},\bar{q}_{2})\ \ \ \textrm{with}\ \ 2r^{2}\not{\mathcal{D}}_{2}^{\star}\not{\mathcal{D}}_{1}^{\star}(\bar{q}_{1},\bar{q}_{2})=-\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}_{{\accentset{\land}{\mathscr{S}}^{\prime}}}(0,0)\,.

It follows from the vanishing of b(1)𝒮\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{\accentset{\land}{\mathscr{S}}} and (Ωtrχ¯)(1)𝒮\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}_{\accentset{\land}{\mathscr{S}}}, (Ωtrχ)(1)𝒮\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}_{\accentset{\land}{\mathscr{S}}}, χ^(1)𝒮\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{{\hat{\chi}}}_{\accentset{\land}{\mathscr{S}}}, χ^¯(1)𝒮\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\hat{\chi}}}_{\accentset{\land}{\mathscr{S}}} on \mathcal{I} and the transport equations (49) that ^(1)𝒮=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}_{\accentset{\land}{\mathscr{S}}}=0 on the whole boundary \mathcal{I}, which in turn also implies by (66) that r2(1)𝒮=0r^{-2}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}_{\accentset{\land}{\mathscr{S}}}=0 (modulo =1\ell=1 modes). From this, we can infer bounds in int\mathcal{M}_{int} on b(1)𝒮,(1)𝒮,^(1)𝒮\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{\accentset{\land}{\mathscr{S}}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}_{\accentset{\land}{\mathscr{S}}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}_{\accentset{\land}{\mathscr{S}}} by integrating their respective transport equations backwards from \mathcal{I}. Now, using the estimates on b(1)𝒮\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{\accentset{\land}{\mathscr{S}}} and b(1)𝒮\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{{\accentset{\land}{\mathscr{S}}^{\prime}}}, (1)𝒮\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}_{\accentset{\land}{\mathscr{S}}} and (1)𝒮\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}_{{\accentset{\land}{\mathscr{S}}^{\prime}}}, and ^(1)𝒮\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}_{\accentset{\land}{\mathscr{S}}} and (1)𝒮\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}_{{\accentset{\land}{\mathscr{S}}^{\prime}}}, one deduces that the non-vanishing pure gauge coefficients b(1)𝒢f,(1)𝒢f,^(1)𝒢f\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{b}_{\mathscr{G}_{f}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}_{\mathscr{G}_{f}},\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\hat{\not{g}}}_{\mathscr{G}_{f}} also satisfy boundedness and logarithmic decay statements. This finishes the proof of Theorem 4.14.

Appendix A Boundary regularity and boundary conditions

This section is dedicated to the proof of Proposition 2.1 and Proposition 2.2.

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 𝐠\boldsymbol{g} is a solution to the Einstein equations (1) and that 𝐠~:=(uv)2𝐠\boldsymbol{\widetilde{g}}:=(u-v)^{2}\boldsymbol{g} extends smoothly to ={uv=0}\mathcal{I}=\{u-v=0\}. Let 𝐖~:=W(𝐠~)\boldsymbol{\widetilde{W}}:=\mathrm{W}(\boldsymbol{\widetilde{g}}) denote the Weyl tensor of 𝐠~\boldsymbol{\widetilde{g}}. Let 𝐍~=12(uv)1(e4e3)\boldsymbol{\widetilde{N}}=\frac{1}{2}(u-v)^{-1}\left(e_{4}-e_{3}\right) denote the outgoing unit normal (for the metric 𝐠~\boldsymbol{\widetilde{g}}) to the {uv=cst}\{u-v=cst\}-hypersurfaces, define 𝐡\boldsymbol{h} to be the induced metric by 𝐠~\boldsymbol{\widetilde{g}} on the {uv=cst}\{u-v=cst\}-hypersurfaces and define the second fundamental forms 𝚯(X,Y)=𝐠~(~X𝐍~,Y)\boldsymbol{\Theta}(X,Y)=\boldsymbol{\widetilde{g}}(\boldsymbol{\widetilde{\nabla}}_{X}\boldsymbol{\widetilde{N}},Y) for all tangent vectors X,YX,Y to the {uv=cst}\{u-v=cst\}-hypersurfaces. Then,

  1. 1.

    (uv)2(k2(uv)2𝛀21)(u-v)^{-2}\left(k^{2}(u-v)^{2}\boldsymbol{\Omega}^{2}-1\right) extends smoothly to \mathcal{I},

  2. 2.

    (uv)1𝚯(u-v)^{-1}\boldsymbol{\Theta} extends smoothly to \mathcal{I} (in a 𝒉\boldsymbol{h}-normalised frame),

  3. 3.

    (uv)1𝑾~(u-v)^{-1}\boldsymbol{\widetilde{W}} extends smoothly to \mathcal{I} (in a 𝒈~\boldsymbol{\widetilde{g}}-normalised frame).

Proof.

First note that by the double null form of 𝒈~\boldsymbol{\widetilde{g}}, we have the relations

|~(uv)|𝒈~2\displaystyle\big{|}\boldsymbol{\widetilde{\nabla}}(u-v)\big{|}^{2}_{\boldsymbol{\widetilde{g}}} =1(uv)2𝛀2,\displaystyle=\frac{1}{(u-v)^{2}\boldsymbol{\Omega}^{2}}, 𝑵~(uv)\displaystyle\boldsymbol{\widetilde{N}}(u-v) =1(uv)𝛀,\displaystyle=-\frac{1}{(u-v)\boldsymbol{\Omega}}, 𝑵~\displaystyle\boldsymbol{\widetilde{N}} =(uv)𝛀~(uv).\displaystyle=-(u-v)\boldsymbol{\Omega}\boldsymbol{\widetilde{\nabla}}(u-v). (223)

The general conformal transformation formula181818See https://en.wikipedia.org/wiki/List_of_formulas_in_Riemannian_geometry. reads

(uv)2Ric(𝒈~)\displaystyle(u-v)^{2}\mathrm{Ric}(\boldsymbol{\widetilde{g}}) =(uv)2Ric(𝒈)2(uv)~2(uv)((uv)~μ~μ(uv)3|~(uv)|𝒈~2)𝒈~,\displaystyle=(u-v)^{2}\mathrm{Ric}(\boldsymbol{g})-2(u-v)\boldsymbol{\widetilde{\nabla}}^{2}(u-v)-\left((u-v)\boldsymbol{\widetilde{\nabla}}^{\mu}\boldsymbol{\widetilde{\nabla}}_{\mu}(u-v)-3|\boldsymbol{\widetilde{\nabla}}(u-v)|^{2}_{\boldsymbol{\widetilde{g}}}\right)\boldsymbol{\widetilde{g}},

which, plugging in the Einstein equation (1) and using (223), rewrites as

(uv)2Ric(𝒈~)\displaystyle(u-v)^{2}\mathrm{Ric}(\boldsymbol{\widetilde{g}}) =2(uv)~2(uv)(uv)~μ~μ(uv)𝒈~𝔰𝒈~,\displaystyle=-2(u-v)\boldsymbol{\widetilde{\nabla}}^{2}(u-v)-(u-v)\boldsymbol{\widetilde{\nabla}}^{\mu}\boldsymbol{\widetilde{\nabla}}_{\mu}(u-v)\boldsymbol{\widetilde{g}}-\mathfrak{s}\boldsymbol{\widetilde{g}}, (224)

with

𝔰\displaystyle\mathfrak{s} :=3𝛀2(uv)2(k2(uv)2𝛀21).\displaystyle:=3\boldsymbol{\Omega}^{-2}(u-v)^{-2}\left(k^{2}(u-v)^{2}\boldsymbol{\Omega}^{2}-1\right).

From (224), using that 𝒈~\boldsymbol{\widetilde{g}} extends smoothly to \mathcal{I}, we already deduce that (uv)1𝔰(u-v)^{-1}\mathfrak{s} extends smoothly to \mathcal{I}. We now want to obtain the better rate for 𝔰\mathfrak{s} claimed in Item 1. We first note that, from Taylor’s formula, relations (223), and the fact that (uv)1𝔰(u-v)^{-1}\mathfrak{s} extends smoothly to \mathcal{I}, the function 𝔰~\widetilde{\mathfrak{s}} defined by

𝔰~:=(uv)2(𝔰+(uv)((uv)𝛀)𝑵~𝔰)=(uv)2𝔰3𝛀𝑵~((uv)2𝛀2)=(uv)2𝔰+6(uv)2𝛀1𝑵~(log((uv)𝛀)),\displaystyle\begin{aligned} \widetilde{\mathfrak{s}}&:=(u-v)^{-2}\left(\mathfrak{s}+(u-v)\left((u-v)\boldsymbol{\Omega}\right)\boldsymbol{\widetilde{N}}\mathfrak{s}\right)\\ &\,=(u-v)^{-2}\mathfrak{s}-3\boldsymbol{\Omega}\boldsymbol{\widetilde{N}}\left((u-v)^{-2}\boldsymbol{\Omega}^{-2}\right)\\ &\,=(u-v)^{-2}\mathfrak{s}+6(u-v)^{-2}\boldsymbol{\Omega}^{-1}\boldsymbol{\widetilde{N}}\left(\log((u-v)\boldsymbol{\Omega})\right),\end{aligned} (225)

extends smoothly to \mathcal{I}. Thus, if we can prove that (uv)1𝑵~(log((uv)𝛀))(u-v)^{-1}\boldsymbol{\widetilde{N}}\left(\log((u-v)\boldsymbol{\Omega})\right) extends smoothly to \mathcal{I} then Item 1 follows from (225). To control 𝑵~(log((uv)𝛀))\boldsymbol{\widetilde{N}}\left(\log((u-v)\boldsymbol{\Omega})\right), we take the trace in (224) and use (223), and we have

(uv)R(𝒈~)=6~μ~μ(uv)4(uv)1𝔰=6(𝑵~(𝑵~(uv))+𝑵~(uv)tr𝚯)4(uv)1𝔰=6𝛀1(uv)1(𝑵~(log((uv)𝛀))+tr𝚯)4(uv)1𝔰,\displaystyle\begin{aligned} (u-v)\mathrm{R}(\boldsymbol{\widetilde{g}})&=-6\boldsymbol{\widetilde{\nabla}}^{\mu}\boldsymbol{\widetilde{\nabla}}_{\mu}(u-v)-4(u-v)^{-1}\mathfrak{s}\\ &=-6\left(\boldsymbol{\widetilde{N}}(\boldsymbol{\widetilde{N}}(u-v))+\boldsymbol{\widetilde{N}}(u-v)\mathrm{tr}\boldsymbol{\Theta}\right)-4(u-v)^{-1}\mathfrak{s}\\ &=-6\boldsymbol{\Omega}^{-1}(u-v)^{-1}\left(-\boldsymbol{\widetilde{N}}(\log((u-v)\boldsymbol{\Omega}))+\mathrm{tr}\boldsymbol{\Theta}\right)-4(u-v)^{-1}\mathfrak{s},\end{aligned} (226)

where tr𝚯:=𝒉ij𝚯ij\mathrm{tr}\boldsymbol{\Theta}:=\boldsymbol{h}^{ij}\boldsymbol{\Theta}_{ij} with 𝒉\boldsymbol{h} the induced metric by 𝒈~\boldsymbol{\widetilde{g}} on the {uv=cst}\{u-v=cst\} hypersurfaces. Now, we want to express tr𝚯\mathrm{tr}\boldsymbol{\Theta} – at first order – in terms of 𝑵~(log((uv)𝛀))\boldsymbol{\widetilde{N}}\left(\log((u-v)\boldsymbol{\Omega})\right). Letting 𝑻~:=12(uv)1(e4+e3)\boldsymbol{\widetilde{T}}:=\frac{1}{2}(u-v)^{-1}(e_{4}+e_{3}), we have

𝚯(𝑻~,𝑻~)\displaystyle\boldsymbol{\Theta}(\boldsymbol{\widetilde{T}},\boldsymbol{\widetilde{T}}) =𝒈~([𝑻~,𝑵~],𝑻~)=𝑵~log((uv)𝛀).\displaystyle=\boldsymbol{\widetilde{g}}\left(\left[\boldsymbol{\widetilde{T}},\boldsymbol{\widetilde{N}}\right],\boldsymbol{\widetilde{T}}\right)=-\boldsymbol{\widetilde{N}}\log((u-v)\boldsymbol{\Omega}).

where we used that 𝑻~=12𝛀1(uv)1(u+v+bAA)\boldsymbol{\widetilde{T}}=\frac{1}{2}\boldsymbol{\Omega}^{-1}(u-v)^{-1}\left(\partial_{u}+\partial_{v}+b^{A}\partial_{A}\right) and that 𝒈~(𝑻~,𝑵~)=𝒈~(𝑻~,A)=0\boldsymbol{\widetilde{g}}(\boldsymbol{\widetilde{T}},\boldsymbol{\widetilde{N}})=\boldsymbol{\widetilde{g}}(\boldsymbol{\widetilde{T}},\partial_{A})=0. Hence

tr𝚯=3𝚯(𝑻~,𝑻~)+3𝑵~log((uv)𝛀)+tr𝚯=3𝑵~log((uv)𝛀)+3𝚯^(𝑻~,𝑻~),\displaystyle\begin{aligned} \mathrm{tr}\boldsymbol{\Theta}&=3\boldsymbol{\Theta}(\boldsymbol{\widetilde{T}},\boldsymbol{\widetilde{T}})+3\boldsymbol{\widetilde{N}}\log((u-v)\boldsymbol{\Omega})+\mathrm{tr}\boldsymbol{\Theta}=3\boldsymbol{\widetilde{N}}\log((u-v)\boldsymbol{\Omega})+3\boldsymbol{\hat{\Theta}}(\boldsymbol{\widetilde{T}},\boldsymbol{\widetilde{T}}),\end{aligned} (227)

with 𝚯^:=𝚯13tr𝚯𝒉\boldsymbol{\hat{\Theta}}:=\boldsymbol{\Theta}-\frac{1}{3}\mathrm{tr}\boldsymbol{\Theta}\boldsymbol{h} denoting the traceless part of 𝚯\boldsymbol{\Theta}. Plugging (225) and (227) into (226), we get

(uv)R(𝒈~)=6𝛀1(uv)1(1+3+4)𝑵~(log(uv)𝛀)18𝛀1(uv)1𝚯^(𝑻~,𝑻~)4(uv)𝔰~.\displaystyle\begin{aligned} (u-v)\mathrm{R}(\boldsymbol{\widetilde{g}})&=-6\boldsymbol{\Omega}^{-1}(u-v)^{-1}(-1+3+4)\boldsymbol{\widetilde{N}}(\log(u-v)\boldsymbol{\Omega})\\ &\quad-18\boldsymbol{\Omega}^{-1}(u-v)^{-1}\boldsymbol{\hat{\Theta}}(\boldsymbol{\widetilde{T}},\boldsymbol{\widetilde{T}})-4(u-v)\widetilde{\mathfrak{s}}.\end{aligned} (228)

Let us now show that 𝚯^\boldsymbol{\hat{\Theta}} vanishes at first order at \mathcal{I}. Projecting formula (224) on the {uv=cst}\{u-v=cst\}-hypersurfaces, using (223), and taking the traceless part, we have

𝛀(uv)(Ric(𝒈~)ij𝒉ijRic(𝒈~)ij𝒉ij)\displaystyle\boldsymbol{\Omega}(u-v)\left(\mathrm{Ric}(\boldsymbol{\widetilde{g}})_{ij}-\boldsymbol{h}^{i^{\prime}j^{\prime}}\mathrm{Ric}(\boldsymbol{\widetilde{g}})_{i^{\prime}j^{\prime}}\boldsymbol{h}_{ij}\right) =2(uv)1𝚯^ij,\displaystyle=-2(u-v)^{-1}\boldsymbol{\hat{\Theta}}_{ij},

Hence, using that 𝒈~\boldsymbol{\widetilde{g}} extends smoothly to \mathcal{I}, we have that (uv)1𝚯^(u-v)^{-1}\boldsymbol{\hat{\Theta}} extends smoothly to \mathcal{I}. Thus, from (228), using that 𝒈~\boldsymbol{\widetilde{g}}, (uv)1𝚯^(u-v)^{-1}\boldsymbol{\hat{\Theta}} and 𝔰~\widetilde{\mathfrak{s}} extend smoothly at \mathcal{I}, we infer that (uv)1𝑵~(log(uv)𝛀)(u-v)^{-1}\boldsymbol{\widetilde{N}}(\log(u-v)\boldsymbol{\Omega}) extends smoothly to \mathcal{I}. Hence, recalling formula (225) and the definition of 𝔰\mathfrak{s}, Item 1 is proved. Using (227) and the regularity of 𝚯^\boldsymbol{\hat{\Theta}} obtained above, we also directly infer Item 2.
From the conformal invariance of the Bianchi equations for the Weyl tensor, we have

~α𝒅~βγδα\displaystyle\boldsymbol{\widetilde{\nabla}}_{\alpha}\boldsymbol{\widetilde{d}}^{\alpha}_{\beta\gamma\delta} =0,\displaystyle=0, with𝒅~\displaystyle\text{with}\quad\boldsymbol{\widetilde{d}} :=(uv)1𝑾~.\displaystyle:=(u-v)^{-1}\boldsymbol{\widetilde{W}}. (229)

which, using (223), implies

(uv)~α𝑾~αβγδ(uv)1𝛀1𝑵~α𝑾~αβγδ=0.\displaystyle(u-v)\boldsymbol{\widetilde{\nabla}}^{\alpha}\boldsymbol{\widetilde{W}}_{\alpha\beta\gamma\delta}-(u-v)^{-1}\boldsymbol{\Omega}^{-1}\boldsymbol{\widetilde{N}}^{\alpha}\boldsymbol{\widetilde{W}}_{\alpha\beta\gamma\delta}=0. (230)

Using that 𝒈~\boldsymbol{\widetilde{g}} extends smoothly to \mathcal{I}, (uv)1𝑵~α𝑾~αβγδ(u-v)^{-1}\boldsymbol{\widetilde{N}}^{\alpha}\boldsymbol{\widetilde{W}}_{\alpha\beta\gamma\delta} extends smoothly to \mathcal{I}. Using the symmetries of the Weyl tensor – see e.g. formulas (7.3.3) in [CK93] –, all the components of 𝑾~\boldsymbol{\widetilde{W}} can be obtained by linear combinations of 𝑵~α𝑾~αβγδ\boldsymbol{\widetilde{N}}^{\alpha}\boldsymbol{\widetilde{W}}_{\alpha\beta\gamma\delta} and Item 3 follows. ∎

We can now prove Proposition 2.1.

Proof of Proposition 2.1.

The regularity of 𝛀\boldsymbol{\Omega} is a direct consequence of Lemma A.1. The regularity of 𝒃\boldsymbol{b} follows from the coordinate components 𝒃A\boldsymbol{b}^{A} (indices up!) extending regularly. By the conformal transformation formulas, we have

𝝌AB=𝒈(Ae4,B)=(uv)2𝒈~(~Ae4,B)e4(log(uv))𝒈AB=(uv)1𝒈~(~A(𝑻~+𝑵~),B)+𝒈AB(uv)𝛀,\displaystyle\begin{aligned} \boldsymbol{\chi}_{AB}=\boldsymbol{g}\left(\boldsymbol{\nabla}_{\partial_{A}}e_{4},\partial_{B}\right)&=(u-v)^{-2}\boldsymbol{\widetilde{g}}\left(\boldsymbol{\widetilde{\nabla}}_{\partial_{A}}e_{4},\partial_{B}\right)-e_{4}(\log(u-v))\boldsymbol{g}_{AB}\\ &=(u-v)^{-1}\boldsymbol{\widetilde{g}}\left(\boldsymbol{\widetilde{\nabla}}_{\partial_{A}}(\boldsymbol{\widetilde{T}}+\boldsymbol{\widetilde{N}}),\partial_{B}\right)+\frac{\boldsymbol{g}_{AB}}{(u-v)\boldsymbol{\Omega}},\end{aligned} (231a)
and similarly
𝝌¯AB\displaystyle\boldsymbol{\underline{\chi}}_{AB} =(uv)1𝒈~(~A(𝑻~𝑵~),B)𝒈AB(uv)𝛀,\displaystyle=(u-v)^{-1}\boldsymbol{\widetilde{g}}\left(\boldsymbol{\widetilde{\nabla}}_{\partial_{A}}(\boldsymbol{\widetilde{T}}-\boldsymbol{\widetilde{N}}),\partial_{B}\right)-\frac{\boldsymbol{g}_{AB}}{(u-v)\boldsymbol{\Omega}}, (231b)
𝜼A\displaystyle\boldsymbol{\eta}_{A} =𝒈~(~A𝑵~,𝑻~)+∇̸Alog𝛀,\displaystyle=\boldsymbol{\widetilde{g}}\left(\boldsymbol{\widetilde{\nabla}}_{\partial_{A}}\boldsymbol{\widetilde{N}},\boldsymbol{\widetilde{T}}\right)+\boldsymbol{\not{\nabla}}_{A}\log\boldsymbol{\Omega}, (231c)
𝜼¯A\displaystyle\boldsymbol{\underline{\eta}}_{A} =𝒈~(~A𝑵~,𝑻~)+∇̸Alog𝛀,\displaystyle=-\boldsymbol{\widetilde{g}}\left(\boldsymbol{\widetilde{\nabla}}_{\partial_{A}}\boldsymbol{\widetilde{N}},\boldsymbol{\widetilde{T}}\right)+\boldsymbol{\not{\nabla}}_{A}\log\boldsymbol{\Omega}, (231d)

From (231a), and the fact that 𝒈~\boldsymbol{\widetilde{g}} extends regularly at \mathcal{I}, one infers that 𝛀𝝌𝒈(uv)𝛀\boldsymbol{\Omega}\boldsymbol{\chi}-\frac{\boldsymbol{g}}{(u-v)\boldsymbol{\Omega}} extends regularly to \mathcal{I} in a 𝒈\boldsymbol{g} orthonormal frame. The corresponding regularity for 𝝌¯\boldsymbol{\underline{\chi}} follows similarly. Moreover, from formulas (231a), (231b), one has

𝝌AB𝝌¯AB2𝒈AB(uv)𝛀\displaystyle\boldsymbol{\chi}_{AB}-\boldsymbol{\underline{\chi}}_{AB}-2\frac{\boldsymbol{g}_{AB}}{(u-v)\boldsymbol{\Omega}} =2(uv)1𝚯AB,\displaystyle=2(u-v)^{-1}\boldsymbol{\Theta}_{AB},

and from the (better) regularity for 𝚯\boldsymbol{\Theta} of Item 2 of Lemma A.1, we obtain the (better) regularity for the difference 𝝌𝝌¯\boldsymbol{\chi}-\boldsymbol{\underline{\chi}} in (30). From (231c), (231d), we have that r(𝜼𝜼¯)A=2r𝚯(𝑻~,A)r(\boldsymbol{\eta}-\underline{\boldsymbol{\eta}})_{A}=2r\boldsymbol{\Theta}(\boldsymbol{\widetilde{T}},\partial_{A}) is regular by Item 2 of Lemma A.1, hence r2(𝜼𝜼¯)r^{2}(\boldsymbol{\eta}-\underline{\boldsymbol{\eta}}) extends smoothly to \mathcal{I} in a 𝒈\boldsymbol{g} orthonormal frame. Moreover, we have r2(𝜼+𝜼¯)A=r2𝛀1A𝛀r^{2}(\boldsymbol{\eta}+\underline{\boldsymbol{\eta}})_{A}=r^{2}\boldsymbol{\Omega}^{-1}\partial_{A}\boldsymbol{\Omega} is regular by (the good regularity of) Item 1 of Lemma A.1, thus r3(𝜼+𝜼¯)r^{3}(\boldsymbol{\eta}+\underline{\boldsymbol{\eta}}) extends smoothly to \mathcal{I} in a 𝒈\boldsymbol{g} orthonormal frame, and combining the above, r2𝜼,r2𝜼¯r^{2}\boldsymbol{\eta},r^{2}\underline{\boldsymbol{\eta}} extend smoothly to \mathcal{I} in a 𝒈\boldsymbol{g} orthonormal frame. We have

𝝎\displaystyle\boldsymbol{\omega} =12vlog((uv)2𝛀2)+1uv,\displaystyle=\frac{1}{2}\partial_{v}\log\left((u-v)^{2}\boldsymbol{\Omega}^{2}\right)+\frac{1}{u-v}, 𝝎¯\displaystyle\boldsymbol{\underline{\omega}} =12ulog((uv)2𝛀2)1uv,\displaystyle=\frac{1}{2}\partial_{u}\log\left((u-v)^{2}\boldsymbol{\Omega}^{2}\right)-\frac{1}{u-v},

from which, by Item 1 of Lemma A.1, we infer that r(𝝎𝝎¯)r(\boldsymbol{\omega}-\boldsymbol{\underline{\omega}}) and r2(𝝎+𝝎¯)r^{2}(\boldsymbol{\omega}+\boldsymbol{\underline{\omega}}) extend regularly at \mathcal{I}. 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

We first have the following lemma. See also [Fri95] and Section 6.2 in [HLSW20].

Lemma A.2.

Assume that 𝐠~\boldsymbol{\widetilde{g}} extends smoothly to \mathcal{I}, satisfies the Einstein equations (1), and that the induced metric 𝐡\boldsymbol{h} by 𝐠~\boldsymbol{\widetilde{g}} on \mathcal{I} is conformal to the Anti-de Sitter metric at infinity t2+k2𝛄-\d{t}^{2}+k^{-2}\boldsymbol{\gamma}. Then, the {uv}\{u-v\}-tangent tensor (uv)2𝐍~α𝐖~αijk(u-v)^{-2}\boldsymbol{\widetilde{N}}^{\alpha}\boldsymbol{\widetilde{W}}_{\alpha ijk} extends smoothly to \mathcal{I}. Note that this is equivalent to the spacetime tensor (uv)2𝐍~α𝐍~γ𝐖~αμγν(u-v)^{-2}\boldsymbol{\widetilde{N}}^{\alpha}\boldsymbol{\widetilde{N}}^{\gamma}{{}^{\star}}\boldsymbol{\widetilde{W}}_{\alpha\mu\gamma\nu}, with denoting the Hodge dual, extending smoothly to \mathcal{I}.

Proof.

From contractions of the second Bianchi identities and the definition of the Weyl tensor, we have the following general formula

~αW(𝒈~)αβγδ\displaystyle\boldsymbol{\widetilde{\nabla}}^{\alpha}\mathrm{W}(\boldsymbol{\widetilde{g}})_{\alpha\beta\gamma\delta} =12~γ(Ric(𝒈~)βδ16R(𝒈~)𝒈~βδ)12~δ(Ric(𝒈~)βγ16R(𝒈~)𝒈~βγ)=:12C(𝒈~)βδγ,\displaystyle=\frac{1}{2}\boldsymbol{\widetilde{\nabla}}_{\gamma}\left(\mathrm{Ric}(\boldsymbol{\widetilde{g}})_{\beta\delta}-\frac{1}{6}\mathrm{R}(\boldsymbol{\widetilde{g}})\boldsymbol{\widetilde{g}}_{\beta\delta}\right)-\frac{1}{2}\boldsymbol{\widetilde{\nabla}}_{\delta}\left(\mathrm{Ric}(\boldsymbol{\widetilde{g}})_{\beta\gamma}-\frac{1}{6}\mathrm{R}(\boldsymbol{\widetilde{g}})\boldsymbol{\widetilde{g}}_{\beta\gamma}\right)=:\frac{1}{2}\mathrm{C}(\boldsymbol{\widetilde{g}})_{\beta\delta\gamma}, (232)

where C(𝒈~)\mathrm{C}(\boldsymbol{\widetilde{g}}) is called the Cotton tensor of 𝒈~\boldsymbol{\widetilde{g}}. The Gauss-Codazzi equations on the boundary \mathcal{I} read

Ric(𝒈~)ijRm(𝒈~)i𝑵~j𝑵~\displaystyle\mathrm{Ric}(\boldsymbol{\widetilde{g}})_{ij}-\mathrm{Rm}(\boldsymbol{\widetilde{g}})_{i\boldsymbol{\widetilde{N}}j\boldsymbol{\widetilde{N}}} =Ric(𝒉)ij𝚯ik𝚯jk+tr𝚯𝚯ij,\displaystyle=\mathrm{Ric}(\boldsymbol{h})_{ij}-\boldsymbol{\Theta}_{ik}\boldsymbol{\Theta}^{k}_{j}+\mathrm{tr}\boldsymbol{\Theta}\boldsymbol{\Theta}_{ij}, (233)
R(𝒈~)2Ric(𝒈~)𝑵~𝑵~\displaystyle\mathrm{R}(\boldsymbol{\widetilde{g}})-2\mathrm{Ric}(\boldsymbol{\widetilde{g}})_{\boldsymbol{\widetilde{N}}\boldsymbol{\widetilde{N}}} =R(𝒉)|𝚯|2+(tr𝚯)2.\displaystyle=\mathrm{R}(\boldsymbol{h})-|\boldsymbol{\Theta}|^{2}+(\mathrm{tr}\boldsymbol{\Theta})^{2}. (234)

By the definition of the Weyl tensor, we have

W(𝒈~)i𝑵~j𝑵~\displaystyle\mathrm{W}(\boldsymbol{\widetilde{g}})_{i\boldsymbol{\widetilde{N}}j\boldsymbol{\widetilde{N}}} =Rm(𝒈~)i𝑵~j𝑵~12Ric(𝒈~)ij12Ric(𝒈~)𝑵~𝑵~𝒈~ij+16R(𝒈~)𝒈~ij=:w(𝒈~)ij,\displaystyle=\mathrm{Rm}(\boldsymbol{\widetilde{g}})_{i\boldsymbol{\widetilde{N}}j\boldsymbol{\widetilde{N}}}-\frac{1}{2}\mathrm{Ric}(\boldsymbol{\widetilde{g}})_{ij}-\frac{1}{2}\mathrm{Ric}(\boldsymbol{\widetilde{g}})_{\boldsymbol{\widetilde{N}}\boldsymbol{\widetilde{N}}}\boldsymbol{\widetilde{g}}_{ij}+\frac{1}{6}\mathrm{R}(\boldsymbol{\widetilde{g}})\boldsymbol{\widetilde{g}}_{ij}=:\mathrm{w}(\boldsymbol{\widetilde{g}})_{ij},

which, plugged in the Gauss-Codazzi equation (233), using (234) to replace Ric(𝒈~)𝑵~𝑵~\mathrm{Ric}(\boldsymbol{\widetilde{g}})_{\boldsymbol{\widetilde{N}}\boldsymbol{\widetilde{N}}}, gives

12(Ric(𝒈~)ij16R(𝒈~)𝒈~ij)w(𝒈~)ij=Ric(𝒉)ij14R(𝒉)𝒉ij+θij,\displaystyle\begin{aligned} \frac{1}{2}\left(\mathrm{Ric}(\boldsymbol{\widetilde{g}})_{ij}-\frac{1}{6}\mathrm{R}(\boldsymbol{\widetilde{g}})\boldsymbol{\widetilde{g}}_{ij}\right)-\mathrm{w}(\boldsymbol{\widetilde{g}})_{ij}&=\mathrm{Ric}(\boldsymbol{h})_{ij}-\frac{1}{4}\mathrm{R}(\boldsymbol{h})\boldsymbol{h}_{ij}+\theta_{ij},\end{aligned} (235)

with θij:=𝚯ik𝚯jk+tr𝚯𝚯ij+14(|𝚯|2(tr𝚯)2)𝒉ij\theta_{ij}:=-\boldsymbol{\Theta}_{ik}\boldsymbol{\Theta}^{k}_{j}+\mathrm{tr}\boldsymbol{\Theta}\boldsymbol{\Theta}_{ij}+\frac{1}{4}\left(|\boldsymbol{\Theta}|^{2}-(\mathrm{tr}\boldsymbol{\Theta})^{2}\right)\boldsymbol{h}_{ij}. Defining the Cotton tensor of 𝒉\boldsymbol{h} by

C(𝒉)ijk:=¯k(Ric(𝒉)ij14R(𝒉)𝒉ij)¯j(Ric(𝒉)ik14R(𝒉)𝒉ik),\displaystyle\mathrm{C}(\boldsymbol{h})_{ijk}:=\boldsymbol{\overline{\nabla}}_{k}\left(\mathrm{Ric}(\boldsymbol{h})_{ij}-\frac{1}{4}\mathrm{R}(\boldsymbol{h})\boldsymbol{h}_{ij}\right)-\boldsymbol{\overline{\nabla}}_{j}\left(\mathrm{Ric}(\boldsymbol{h})_{ik}-\frac{1}{4}\mathrm{R}(\boldsymbol{h})\boldsymbol{h}_{ik}\right),

where ¯\boldsymbol{\overline{\nabla}} is the covariant derivative of 𝒉\boldsymbol{h}, and applying ¯\boldsymbol{\overline{\nabla}} to (235), we get

12C(𝒈~)ijk=C(𝒉)ijk+¯kθij¯jθik+¯kw(𝒈~)ij¯kw(𝒈~)ij.\displaystyle\begin{aligned} \frac{1}{2}\mathrm{C}(\boldsymbol{\widetilde{g}})_{ijk}&=\mathrm{C}(\boldsymbol{h})_{ijk}+\boldsymbol{\overline{\nabla}}_{k}\theta_{ij}-\boldsymbol{\overline{\nabla}}_{j}\theta_{ik}+\boldsymbol{\overline{\nabla}}_{k}\mathrm{w}(\boldsymbol{\widetilde{g}})_{ij}-\boldsymbol{\overline{\nabla}}_{k}\mathrm{w}(\boldsymbol{\widetilde{g}})_{ij}.\end{aligned} (236)

From (236) and Lemma A.1, we deduce that (uv)1(12C(𝒈~)ijkC(𝒉)ijk)(u-v)^{-1}\left(\frac{1}{2}\mathrm{C}(\boldsymbol{\widetilde{g}})_{ijk}-\mathrm{C}(\boldsymbol{h})_{ijk}\right) extends regularly to \mathcal{I}. Combining (230) and (232), we thus deduce that

(uv)1W(𝒈~)𝑵~ijkC(𝒉)ikj\displaystyle(u-v)^{-1}\mathrm{W}(\boldsymbol{\widetilde{g}})_{\boldsymbol{\widetilde{N}}ijk}-\mathrm{C}(\boldsymbol{h})_{ikj} =(uv)Eijk,\displaystyle=(u-v)E_{ijk}, (237)

with EE smoothly extending to \mathcal{I}. 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 dt2+k2γ-\mathrm{d}t^{2}+k^{-2}\gamma. Thus, if 𝒉\boldsymbol{h} is conformal to such a metric, one has by (237) that (uv)2W(𝒈~)𝑵~ijk(u-v)^{-2}\mathrm{W}(\boldsymbol{\widetilde{g}})_{\boldsymbol{\widetilde{N}}ijk} extends smoothly at \mathcal{I}, 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

(r)1(𝜶AB𝜶¯AB)\displaystyle(r^{\star})^{-1}\left(\boldsymbol{\alpha}_{AB}-\underline{\boldsymbol{\alpha}}_{AB}\right) =2(uv)2𝑾~(𝑵~,A,𝑻~,B)+2(uv)2𝑾~(𝑵~,B,𝑻~,A),\displaystyle=2(u-v)^{-2}\boldsymbol{\widetilde{W}}(\boldsymbol{\widetilde{N}},\partial_{A},\boldsymbol{\widetilde{T}},\partial_{B})+2(u-v)^{-2}\boldsymbol{\widetilde{W}}(\boldsymbol{\widetilde{N}},\partial_{B},\boldsymbol{\widetilde{T}},\partial_{A}),
(r)2(𝜷A+𝜷¯A)\displaystyle(r^{\star})^{-2}\left(\boldsymbol{\beta}_{A}+\underline{\boldsymbol{\beta}}_{A}\right) =2(uv)2𝑾~(𝑵~,𝑻~,𝑻~,A),\displaystyle=2(u-v)^{-2}\boldsymbol{\widetilde{W}}(\boldsymbol{\widetilde{N}},\boldsymbol{\widetilde{T}},\boldsymbol{\widetilde{T}},\partial_{A}),
(r)3𝝈\displaystyle(r^{\star})^{-3}\boldsymbol{\sigma} =(uv)2𝑾~(𝑵~,𝑻~,𝑵~,𝑻~).\displaystyle=(u-v)^{-2}{{}^{\star}}\boldsymbol{\widetilde{W}}(\boldsymbol{\widetilde{N}},\boldsymbol{\widetilde{T}},\boldsymbol{\widetilde{N}},\boldsymbol{\widetilde{T}}).

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 limvur𝝌^=limvur𝝌¯^\lim_{v\rightarrow u}r\hat{\boldsymbol{\chi}}=\lim_{v\rightarrow u}r\underline{\hat{\boldsymbol{\chi}}} holds by Proposition 2.1, one further infers (34) and this finishes the proof of the proposition. ∎

Appendix B Computation of the =0\ell=0 mode

From the linear version of the Birkhoff theorem, we already know that the space of solutions supported on =0\ell=0 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 =0\ell=0 so we simply write (Ωtrχ)(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)} for (Ωtrχ)(1)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}_{\ell=0} etc. to keep the notation clean.

We first define two quantities (supported on =0\ell=0 by the above convention):

𝒫(1):=r3ρ(1)3M(1),𝒬(1):=r(Ωtrχ)(1)Ω24Ω1Ω(1)+(1).\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{P}}:=r^{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}-3M\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}\ \ \ ,\ \ \ \accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{Q}}:=\frac{r\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}}{\Omega^{2}}-4\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}+\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}\,.

The importance of these quantities lies (partly) in their simple propagation equations (following from (70), (71), (49), (55) and (52))

Ω∇̸3𝒫(1)=0,Ω∇̸4𝒫(1)=0andΩ∇̸4𝒬(1)=0.\displaystyle\Omega\not{\nabla}_{3}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{P}}=0\ ,\ \Omega\not{\nabla}_{4}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{P}}=0\ \ \ \textrm{and}\ \ \ \Omega\not{\nabla}_{4}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{Q}}=0\,. (238)

Using the formula (66) we write the linearised Gauss equation (65) for =0\ell=0 as

1r2(1)=1r3𝒫(1)3Mr3(1)12rΩ∇̸3((1))+Ω22r2𝒬(1)Ω22r2(1),-\frac{1}{r^{2}}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}=-\frac{1}{r^{3}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{P}}-\frac{3M}{r^{3}}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}-\frac{1}{2r}\Omega\not{\nabla}_{3}\left(\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}\right)+\frac{\Omega^{2}}{2r^{2}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{Q}}-\frac{\Omega^{2}}{2r^{2}}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}\,,

or more concisely as

Ω2rΩ∇̸3(rΩ2(1))=2r2𝒫(1)+Ω2r𝒬(1).\displaystyle\frac{\Omega^{2}}{r}\Omega\not{\nabla}_{3}\left(\frac{r}{\Omega^{2}}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}\right)=-\frac{2}{r^{2}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{P}}+\frac{\Omega^{2}}{r}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{Q}}\,. (239)

We first establish that if 𝒫(1)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{P}}=0 and 𝒬(1)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{Q}}=0 hold on the initial data cone, the solution is necessarily trivial.

Lemma B.1.

Let 𝒮\mathscr{S} be a smooth solution of the system of gravitational perturbations supported on =0\ell=0. If 𝒫(1)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{P}}=0 and 𝒬(1)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{Q}}=0 hold on C¯v0\underline{C}_{v_{0}}, then the solution is necessarily equal to the zero solution.

Proof.

The quantities 𝒫(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{P}} and 𝒬(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{Q}} are conserved in the 44-direction by (238), hence zero on int\mathcal{M}_{int}. It follows that 𝒫(1)=𝒬(1)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{P}}=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{Q}}=0 in (239), which since rΩ2(1)\frac{r}{\Omega^{2}}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}} vanishes at \mathcal{I} implies that (1)=0\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}=0 on int\mathcal{M}_{int}. It now immediately follows that K(1)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{K}=0, ρ(1)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}=0, (Ωtrχ)(1)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}=0, (Ωtrχ¯)(1)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}=0, Ω1Ω(1)=0\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}=0, ω(1)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}=0, ω¯(1)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}=0. ∎

We wish to study all radial solutions of (239) to exhaust the space of solutions for =0\ell=0. We first note that by adding a pure gauge solution, we can restrict to the case of both 𝒫(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{P}} and 𝒬(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{Q}} being constant on int\mathcal{M}_{int}.

Lemma B.2.

Let 𝒮\mathscr{S} be a solution of the system of gravitational perturbations supported on =0\ell=0. Then we can add a pure gauge solution 𝒢f\mathscr{G}_{f} from Lemma 2.8 such that the solution 𝒮=𝒮+𝒢f\mathscr{S}^{\prime}=\mathscr{S}+\mathscr{G}_{f} satisfies

(Ωtrχ)(1)|𝒮(,v0)=0and𝒬(1)|𝒮(u,v0)is constant in u.\displaystyle\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\big{|}_{\mathscr{S}^{\prime}}(\infty,v_{0})=0\ \ \textrm{and}\ \ \accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{Q}}|_{\mathscr{S}^{\prime}}(u,v_{0})\ \ \textrm{is constant in $u$.}

As a consequence of (238), 𝒫(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{P}} and 𝒬(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{Q}} are both constant on int\mathcal{M}_{int} for 𝒮\mathscr{S}^{\prime}.

Proof.

Letting f~(u)=1Ω2(u,v0)12r+22Mr++k2r+2(Ωtrχ)(1)𝒮(,v0)\tilde{f}(u)=\frac{1}{\Omega^{2}(u,v_{0})}\frac{1}{2}\frac{r_{+}^{2}}{\frac{2M}{r_{+}}+k^{2}r_{+}^{2}}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}_{\mathscr{S}}(\infty,v_{0}) generate f~u\tilde{f}_{u} and f~v\tilde{f}_{v} and a pure gauge solution 𝒢f~\mathscr{G}_{\tilde{f}} as in Lemma 2.8 we achieve that 𝒮+𝒢f~\mathscr{S}+\mathscr{G}_{\tilde{f}} satisfies (Ωtrχ)(1)|𝒮+𝒢f~(,v0)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\big{|}_{\mathscr{S}+\mathscr{G}_{\tilde{f}}}(\infty,v_{0})=0. In particular, the quantity 𝒬(1)(,v0)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{Q}}(\infty,v_{0}) is now regular at the horizon for the solution 𝒮+𝒢f~\mathscr{S}+\mathscr{G}_{\tilde{f}}. We next add a second pure gauge solution 𝒢f^\mathscr{G}_{\hat{f}} which does not affect (Ωtrχ)(1)(,v0)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}(\infty,v_{0}) but achieves the second condition. For this we define f^u\hat{f}_{u} by the ODE

𝒬(1)𝒮+𝒢f~(,v0)=𝒬(1)𝒮+𝒢f~(u,v0)2uf^u(u,v0),f^u(u0,v0)=0.\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{Q}}_{\mathscr{S}+\mathscr{G}_{\tilde{f}}}(\infty,v_{0})=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{Q}}_{\mathscr{S}+\mathscr{G}_{\tilde{f}}}(u,v_{0})-2\partial_{u}\hat{f}_{u}(u,v_{0})\,\ \ ,\ \ \hat{f}_{u}(u_{0},v_{0})=0\,.

One now checks that 𝒬(1)𝒮:=𝒮+𝒢f~+𝒢f^\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{Q}}_{\mathscr{S}^{\prime}:=\mathscr{S}+\mathscr{G}_{\tilde{f}}+\mathscr{G}_{\hat{f}}} is indeed constant, 𝒬(1)𝒮(u,v0)=𝒬(1)𝒮+𝒢f~(,v0)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{Q}}_{\mathscr{S}^{\prime}}(u,v_{0})=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{Q}}_{\mathscr{S}+\mathscr{G}_{\tilde{f}}}(\infty,v_{0}), and that f^u\hat{f}_{u} is bounded. By Lemma 2.8, a bounded f^u\hat{f}_{u} will imply (Ωtrχ)(1)|𝒮(,v0)=(Ωtrχ)(1)|𝒮+𝒢f~(,v0)=0\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}|_{\mathscr{S}^{\prime}}(\infty,v_{0})=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}\big{|}_{\mathscr{S}+\mathscr{G}_{\tilde{f}}}(\infty,v_{0})=0. ∎

Let us denote the constants d:=𝒫(1)d:=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{P}} and c:=𝒬(1)c:=\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{Q}} and compute now the general regular radial solutions of (239). Setting (1)=f(r)\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}=f(r), ff satisfies the ODE

r(frΩ2)=+2dMrΩ4cΩ2,\displaystyle\partial_{r}\left(f\frac{r}{\Omega^{2}}\right)=+\frac{2d\cdot M}{r\Omega^{4}}-\frac{c}{\Omega^{2}}\,, (240)

which we can write as (setting 1l2=k2\frac{1}{l^{2}}=k^{2})

r(frΩ2+2dM1+3r2l21Ω2)=2dMl2(l23r2)(l2+3r2)21Ω2rcΩ2.\partial_{r}\left(f\frac{r}{\Omega^{2}}+\frac{2dM}{1+3\frac{r^{2}}{l^{2}}}\frac{1}{\Omega^{2}}\right)=2dM\frac{l^{2}(l^{2}-3r^{2})}{(l^{2}+3r^{2})^{2}}\frac{1}{\Omega^{2}r}-\frac{c}{\Omega^{2}}\,.

To make the solution regular at the horizon we require

c=2dMl2(l23r+2)r+(l2+3r+2)2.c=2dM\frac{l^{2}(l^{2}-3r_{+}^{2})}{r_{+}(l^{2}+3r_{+}^{2})^{2}}\,.

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 ll\rightarrow\infty to compare with the asymptotically flat case, since we have used that rΩ2\frac{r}{\Omega^{2}} goes to zero, which it does not in the asymptotically flat case. However, in (240) we can still take the limit \ell\rightarrow\infty and check that in this case c=dc=d and f=df=-d is indeed a solution, as was obtained in [DHR19].

f(r)=2dMr(1+3r2l2)+Ω2rr2dMΩ2(l2(l23r+2)(l2+3r+2)21r+l2(l23r2)(l2+3r2)21r)𝑑r,\displaystyle f(r)=-\frac{2dM}{r\left(1+\frac{3r^{2}}{l^{2}}\right)}+\frac{\Omega^{2}}{r}\int_{r}^{\infty}\frac{2dM}{\Omega^{2}}\left(\frac{l^{2}(l^{2}-3r_{+}^{2})}{(l^{2}+3r_{+}^{2})^{2}}\frac{1}{r_{+}}-\frac{l^{2}(l^{2}-3r^{2})}{(l^{2}+3r^{2})^{2}}\frac{1}{r}\right)dr\,, (241)

which satisfies

f(r+)=2dMr+(1+3r+2l2)andf()=2dMl2(l23r+2)(l2+3r+2)21r+.f(r_{+})=-\frac{2dM}{r_{+}\left(1+\frac{3r_{+}^{2}}{l^{2}}\right)}\ \ \ \textrm{and}\ \ \ \ f(\infty)=2dM\frac{l^{2}(l^{2}-3r_{+}^{2})}{(l^{2}+3r_{+}^{2})^{2}}\frac{1}{r_{+}}\,.

In particular, ff is uniformly bounded and smooth on the exterior. All non-vanishing Ricci-coefficients and curvature components can easily be computed in terms of ff. We find

(Ωtrχ)(1)=(rf)Ω2,(Ωtrχ¯)(1)=(rf)Ω2,\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\chi\right)}=(\partial_{r}f)\Omega^{2}\ \ \ ,\ \ \ \accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\left(\Omega\mathrm{tr}\underline{\chi}\right)}=-(\partial_{r}f)\Omega^{2}\,,

and from the definition of 𝒬(1)\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\mathcal{Q}} the expression

Ω1Ω(1)\displaystyle\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega} =14(r(Ω2Ω2fr)c)=14(2dMrΩ24dMl2(l23r+2)r+(l2+3r+2)2)+14(2Mr2+2rl2)frΩ2.\displaystyle=\frac{1}{4}\left(\partial_{r}(\Omega^{2}\Omega^{-2}f\cdot r)-c\right)=\frac{1}{4}\left(\frac{2dM}{r\Omega^{2}}-4dM\frac{l^{2}(l^{2}-3r_{+}^{2})}{r_{+}(l^{2}+3r_{+}^{2})^{2}}\right)+\frac{1}{4}\left(\frac{2M}{r^{2}}+\frac{2r}{l^{2}}\right)\frac{fr}{\Omega^{2}}\,. (242)

To check that Ω1Ω(1)\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega} is finite at the horizon we compute

limrr+Ω1Ω(1)(r)=dM2(1+3r+2l2).\lim_{r\rightarrow r_{+}}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}(r)=\frac{dM}{2\left(1+\frac{3r_{+}^{2}}{l^{2}}\right)}\,.

Note also

limrΩ1Ω(1)(r)=dMl2(l23r+2)r+(l2+3r+2)2+dMl2(l23r+2)r+(l2+3r+2)2=0,\lim_{r\rightarrow\infty}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}(r)=-dM\frac{l^{2}(l^{2}-3r_{+}^{2})}{r_{+}(l^{2}+3r_{+}^{2})^{2}}+dM\frac{l^{2}(l^{2}-3r_{+}^{2})}{r_{+}(l^{2}+3r_{+}^{2})^{2}}=0\,,

so Ω1Ω(1)\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega} vanishes at infinity. Finally, we obtain from the null structure equations

ω(1)=Ω∇̸4Ω1Ω(1),ω¯(1)=Ω∇̸3Ω1Ω(1),K(1)=1r2(1),ρ(1)=dMr3+3Mr3(1).\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\omega}=\Omega\not{\nabla}_{4}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}\ \ ,\ \ \accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\underline{\omega}}=\Omega\not{\nabla}_{3}\Omega^{-1}\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\Omega}\ \ ,\ \ \accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{K}=-\frac{1}{r^{2}}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}\ \ ,\ \ \accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\rho}=\frac{dM}{r^{3}}+\frac{3M}{r^{3}}\frac{\accentset{\scalebox{0.6}{\mbox{\tiny(1)}}}{\sqrt{\not{g}}}}{\sqrt{\not{g}}}\,.

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.