We show that linear scalar waves are bounded and continuous up to the Cauchy horizon of Reissner–Nordström–de Sitter and Kerr–de Sitter spacetimes and in fact decay exponentially fast to a constant along the Cauchy horizon. We obtain our results by modifying the spacetime beyond the Cauchy horizon in a suitable manner, which puts the wave equation into a framework in which a number of standard as well as more recent microlocal regularity and scattering theory results apply. In particular, the conormal regularity of waves at the Cauchy horizon—which yields the boundedness statement—is a consequence of radial point estimates, which are microlocal manifestations of the blue-shift and red-shift effects.
I. INTRODUCTION
We present a detailed analysis of the regularity and decay properties of linear scalar waves near the Cauchy horizon of cosmological black hole spacetimes. Concretely, we study charged and non-rotating (Reissner–Nordström–de Sitter) as well as uncharged and rotating (Kerr–de Sitter) black hole spacetimes for which the cosmological constant is positive. See Fig. 1 for their Penrose diagrams. These spacetimes, in the region of interest for us, have the topology , where is an interval and are equipped with a Lorentzian metric g of signature (1, 3). The spacetimes have three horizons located at different values of the radial coordinate r, namely, the Cauchy horizon at r = r1, the event horizon at r = r2, and the cosmological horizon at r = r3, with . In order to measure decay, we use a time function t0, which is equivalent to the Boyer–Lindquist coordinate t away from the cosmological event and Cauchy horizons, i.e., t0 differs from t by a smooth function of the radial coordinate r, and t0 is equivalent to the Eddington–Finkelstein coordinate u near the Cauchy and cosmological horizons and to the Eddington–Finkelstein coordinate near the event horizon. We consider the Cauchy problem for the linear wave equation with Cauchy data posed on a surface HI as indicated in Fig. 1.
The study of asymptotics and decay for linear scalar (and non-scalar) wave equations in a neighborhood of the exterior region of such spacetimes has a long history. Methods of scattering theory have proven very useful in this context, see Refs. 8, 25, 26, 38, 51, 55, 67, and 70 and references therein (we point out that near the black hole exterior, Reissner–Nordström–de Sitter space can be studied using exactly the same methods as Schwarzschild–de Sitter space); see Ref. 21 for a different approach using vector field commutators. There is also a substantial amount of literature on the case of the asymptotically flat Reissner–Nordström and Kerr spacetimes; we refer the reader to Refs. 2, 5, 17, 20, 22–24, 44, 47, 59, 60, and 63 and references therein.
The purpose of the present work is to show how a uniform analysis of linear waves up to the Cauchy horizon can be accomplished using methods from scattering theory and microlocal analysis. Our main result is as follows:
More precisely, as well as all such derivatives of lie in the weighted spacetime Sobolev space in , where is the surface gravity of the Cauchy horizon.
For the massive Klein–Gordon equation , small, the same result holds true without the constant term u0.
Here, the spacetime Sobolev space Hs, for , consists of functions that remain in L2 under the application of up to s stationary vector fields; for general , Hs is defined using duality and interpolation. The final part of Theorem 1.1 in particular implies that lies in near the Cauchy horizon on any surface of fixed t0. After introducing the Reissner–Nordström–de Sitter and Kerr–de Sitter metrics at the beginning of Secs. II and III, we will prove Theorem 1.1 in Secs. II F and III C, see Theorems 2.24 and 3.5. Our analysis carries over directly to non-scalar wave equations as well, as we discuss for differential forms in Sec. II H; however, we do not obtain uniform boundedness near the Cauchy horizon in this case. Furthermore, a substantial number of ideas in the present paper can be adapted to the study of asymptotically flat () spacetimes; corresponding boundedness, regularity, and (polynomial) decay results on Reissner–Nordström and Kerr spacetimes are discussed in Ref. 36.
Let us also mention that a minor extension of our arguments yields analogous boundedness, decay, and regularity results for the Cauchy problem with a “two-ended” Cauchy surface HI up to the bifurcation sphere B, see Fig. 2.
Theorem 1.1 is the first result known to the authors establishing asymptotics and regularity near the Cauchy horizon of rotating black holes. (However, we point out that Dafermos and Luk have recently announced the C0 stability of the Cauchy horizon of the Kerr spacetime for Einstein’s vacuum equations.1) In the case of and in spherical symmetry (Reissner–Nordström), Franzen31 proved the uniform boundedness of waves in the black hole interior and regularity up to , while Luk and Oh45 showed that linear waves generically do not lie in at . (We also mention the more recent in Ref. 46). There is also ongoing work by Franzen on the analogue of her result for Kerr spacetimes.30 Gajic,32 based on previous work by Aretakis,3,4 showed that for extremal Reissner–Nordström spacetimes, waves do lie in . We do not present a microlocal study of the event horizon of extremal black holes here; however, we remark that our analysis reveals certain high regularity phenomena at the Cauchy horizon of near-extremal black holes, which we will discuss below. Closely related to this, the study of Costa et al.14–16 of the nonlinear Einstein–Maxwell–scalar field system in spherical symmetry shows that, close to extremality, rather weak assumptions on initial data on a null hypersurface transversal to the event horizon guarantee regularity of the metric at ; however, they assume exact Reissner–Nordström–de Sitter data on the event horizon, while in the present work, we link non-trivial decay rates of waves along the event horizon to the regularity of waves at . Compare this also with the discussions in Sec. II G and Remark 2.25.
One could combine the treatment of Reissner–Nordström–de Sitter and Kerr–de Sitter spacetimes by studying the more general Kerr–Newman–de Sitter family of charged and rotating black hole spacetimes, discovered by Carter,12 which can be analyzed in a way that is entirely analogous to the Kerr–de Sitter case. However, in order to prevent cumbersome algebraic manipulations from obstructing the flow of our analysis, we give all details for Reissner–Nordström–de Sitter black holes, where the algebra is straightforward and where moreover mode stability can easily be shown to hold for subextremal spacetimes; we then indicate rather briefly the (mostly algebraic) changes for Kerr–de Sitter black holes and leave the similar general case of Kerr–Newman–de Sitter black holes to the reader. In fact, our analysis is stable under suitable perturbations and one can thus obtain results entirely analogous to Theorem 1.1 for Kerr–Newman–de Sitter metrics with small non-zero angular momentum a and small charge Q (depending on a) or for small charge Q and small non-zero angular momentum a (depending on Q), by perturbative arguments: indeed, in these two cases, the Kerr–Newman–de Sitter metric is a small stationary perturbation of the Kerr–de Sitter, respectively, Reissner–Nordström–de Sitter metric, with the same structure at .
In the statement of Theorem 1.1, we point out that the amount of regularity of the remainder term at the Cauchy horizon is directly linked to the amount of exponential decay of : the more decay, the higher the regularity. This can intuitively be understood in terms of the blue-shift effect:58 the more a priori decay has along the Cauchy horizon (approaching i+), the less energy can accumulate at the horizon. The precise microlocal statement capturing this is a radial point estimate at the intersection of with the boundary at infinity of a compactification of the spacetime at , which we will discuss in Sec. I A.
Now, can be any real number less than the spectral gap of the operator , which is the infimum of over all non-zero resonances (or quasi-normal modes) ; the resonance at gives rise to the constant u0 term. (We refer to Refs. 8, 26, 55, and 67 for the discussion of resonances for black hole spacetimes.) Due to the presence of a trapped region in the black hole spacetimes considered here, is bounded from above by a quantity associated with the null-geodesic dynamics near the trapped set, as proved by Dyatlov28,29 in the present context following breakthrough work by Wunsch and Zworski70 and by Nonnenmacher and Zworski:54 below (respectively, above) any line , , there are infinitely (respectively, finitely) many resonances. In principle however, one expects that there indeed exist a non-zero number of resonances above this line, and correspondingly, expansion (1.1) can be refined to take these into account. (In fact, one can obtain a full resonance expansion due to the complete integrability of the null-geodesic flow near the trapped set, see Refs. 8 and 27.) Since for the mode solution corresponding to a resonance at , , we obtain the regularity at , shallow resonances, i.e., those with small , give the dominant contribution to the solution u both in terms of decay and regularity at . The authors are not aware of any rigorous results on shallow resonances, so we shall only discuss this briefly in Remark 2.25, taking into account insights from numerical results: these suggest the existence of resonant states with imaginary parts roughly equal to and , and hence the relative sizes of the surface gravities play a crucial role in determining the regularity at .
Whether resonant states are in fact no better than and the existence of shallow resonances, which, if true, would yield a linear instability result for cosmological black hole spacetimes with Cauchy horizons analogous to Ref. 45, will be studied in future work. Once these questions have been addressed, one can conclude that the lack of, say, H1 regularity at is caused precisely by shallow quasinormal modes. Thus, somewhat surprisingly, the mechanism for the linear instability of the Cauchy horizon of cosmological spacetimes is more subtle than for asymptotically flat spacetimes in that the presence of a cosmological horizon, which ultimately allows for a resonance expansion of linear waves u, leads to a much more precise structure of u at , with the regularity of u directly tied to quasinormal modes of the black hole exterior.
The interest in understanding the behavior of waves near the Cauchy horizon has its roots in Penrose’s Strong Cosmic Censorship conjecture, which asserts that maximally globally hyperbolic developments for the Einstein–Maxwell or Einstein vacuum equations (depending on whether one considers charged or uncharged solutions) with generic initial data (and a complete initial surface, and/or under further conditions) are inextendible as suitably regular Lorentzian manifolds. In particular, the smooth, even analytic, extendability of the Reissner–Nordström(–de Sitter) and Kerr(–de Sitter) solutions past their Cauchy horizons is conjectured to be an unstable phenomenon. It turns out that the question what should be meant by “suitable regularity” is very subtle; we refer to studies by Christodoulou,13 Dafermos,18,19 and Costa et al.14–16 in the spherically symmetric setting for positive and negative results for various notions of regularity. There is also work in progress by Dafermos and Luk on the C0 stability of the Kerr Cauchy horizon, assuming a quantitative version of the non-linear stability of the exterior region. We refer to these studies, as well as to the excellent introductions of Refs. 17, 19, and 45, for a discussion of heuristic arguments and numerical experiments that sparked this line of investigation.
Here, however, we only consider linear equations motivated by similar studies in the asymptotically flat case by Dafermos17 [see Ref. 45 (Footnote 11)], Franzen,31 Sbierski,56 and Luk and Oh.45 The main insight of the present paper is that a uniform analysis up to can be achieved by using methods of scattering theory and geometric microlocal analysis which are by now standard, in the spirit of recent studies by Vasy,67 Baskin, Vasy, and Wunsch,6,39 the core of the precise estimates of Theorem 1.1 is microlocal propagation results at (generalized) radial sets, as we will discuss in Sec. I B. From this geometric microlocal perspective however, i.e., taking into account merely the phase space properties of the operator , it is both unnatural and technically inconvenient to view the Cauchy horizon as a boundary; after all, the metric g is a non-degenerate Lorentzian metric up to and beyond. Thus, the most subtle step in our analysis is the formulation of a suitable extended problem (in a neighborhood of ), which reduces to the equation of interest, namely, the wave equation, in .
A. Geometric setup
The Penrose diagram is rather singular at future timelike infinity i+, yet all relevant phenomena, in particular trapping and red-/blue-shift effects, should be thought of as taking place there, as we will see shortly; therefore, we work instead with a compactification of the region of interest, the domain of dependence of HI in Fig. 1, in which the horizons as well as the trapped region remain separated, and the metric remains smooth, as . Concretely, using the coordinate t0 employed in Theorem 1.1, the radial variable r and the spherical variable , we consider a region
i.e., we add the ideal boundary at future infinity, , to the spacetime, and equip M with the natural smooth structure in which vanishes simply and non-degenerately at . (It is tempting, and useful for purposes of intuition, to think of M as being a submanifold of the blow-up of the compactification suggested by the Penrose diagram—adding an “ideal sphere at infinity” at i+—at i+. However, the details are somewhat subtle; see Ref. 51.)
Due to the stationary nature of the metric g, the (null-)geodesic flow should be studied in a version of phase space, which has a built-in uniformity as . A clean way of describing this uses the language of b-geometry (and b-analysis); we refer the reader to Melrose48 for a detailed introduction, and Ref. 67, Sec. 3 and Ref. 39, Sec. 2 for brief overviews. We recall the most important features here: on M, the metric g is a non-degenerate Lorentzian b-metric, i.e., a linear combination with smooth (on M) coefficients of
where (x1, x2, x3) are coordinates in X; in fact, the coefficients are independent of . Then, g is a section of the symmetric second tensor power of a natural vector bundle on M, the b-cotangent bundle, which is spanned by the sections . We stress that is a smooth non-degenerate section of up to and including the boundary . Likewise, the dual metric G is a section of the second symmetric tensor power of the b-tangent bundle, which is the dual bundle of and thus spanned by . The dual metric function, which we also denote by by a slight abuse of notation, associates with , the squared length .
Over the interior of M, denoted , the b-cotangent bundle is naturally isomorphic to the standard cotangent bundle. The geodesic flow, lifted to the cotangent bundle, is generated by the Hamilton vector field , which extends to a smooth vector field tangent to . Now, HG is homogeneous of degree 1 with respect to dilations in the fiber, and it is often convenient to rescale it by multiplication with a homogeneous degree −1 function , obtaining the homogeneous degree 0 vector field . As such, it extends smoothly to a vector field on the radial (or projective) compactification of , which is a ball bundle over M, with fiber over given by the union of with the “sphere at fiber infinity” . The b-cosphere bundle is then conveniently viewed as the boundary of the compactified b-cotangent bundle at fiber infinity.
The projection to the base M of integral curves of HG or with null initial direction, i.e., starting at a point in , yields (reparameterizations of) null-geodesics on (M, g); this is clear in the interior of M, and the important observation is that this gives a well-defined notion of null-geodesics, or null-bicharacteristics, at the boundary at infinity, X. We remark that the characteristic set has two components, the union of the future null cones and of the past null cones .
The red-shift or blue-shift effect manifests itself in a special structure of the flow near the b-conormal bundles of the horizons r = rj, j = 1, 2, 3. (Here, for a boundary submanifold and is the annihilator of the space of all vectors in tangent to Z; is naturally isomorphic to the conormal bundle of Z in X.) Indeed, in the case of the Reissner–Nordström–de Sitter metric, Lj, more precisely, its boundary at fiber infinity is a saddle point for the flow, with stable (or unstable, depending on which of the two components one is working on) manifold contained in and an unstable (or stable) manifold transversal to . In the Kerr–de Sitter case, does not vanish everywhere on but rather is non-zero and tangent to it, so there are non-trivial dynamics within , but the dynamics in the directions normal to still has the same saddle point structure. See Fig. 3.
B. Strategy of the proof
In order to take full advantage of the saddle point structure of the null-geodesic flow near the Cauchy horizon, one would like to set up an initial value problem or equivalently a forced forward problem , with vanishing initial data but non-trivial right hand side f, on a domain which extends a bit past . Because of the finite speed of propagation for the wave equation, one is free to modify the problem beyond in whichever way is technically most convenient; waves in the region of interest are unaffected by the choice of extension.
A natural idea then is to simply add a boundary , , which one could use to cap the problem off beyond ; now is timelike, hence, to obtain a well-posed problem, one needs to impose boundary conditions there. While perfectly feasible, the resulting analysis is technically rather involved as it necessitates studying the reflection of singularities at quantitatively in a uniform manner as . (Near , one does not need the precise, microlocal, control as in Refs. 52, 53, and 61 however.)
A technically much easier modification involves the use of a complex absorbing “potential” in the spirit of Refs. 54 and 70; here is a second order b-pseudodifferential operator on M, which is elliptic in a large subset of near . (Without b-language, one can take for large t0 to be a time translation-invariant, properly supported ps.d.o. on M.) One then considers the operator
The point is that a suitable choice of the sign of on the two components of the characteristic set leads to an absorption of high frequencies along the future-directed null-geodesic flow over the support of , which allows one to control a solution u of in terms of the right hand side f there. However, since we are forced to work on a domain with boundary in order to study the forward problem, the pseudodifferential complex absorption does not make sense near the relevant boundary component, which is the extension of the left boundary in Fig. 1 past r = r1.
A doubling construction as in Ref. 67, Sec. 4 on the other hand, doubling the spacetime across the timelike surface , say, amounts to gluing an “artificial exterior region” to our spacetime, with one of the horizons identified with the original Cauchy horizon; this in particular creates another trapped region, which we can however easily hide using a complex absorbing potential! We then cap off the thus extended spacetime beyond the cosmological horizon of the artificial exterior region, located at r = r0, by a spacelike hypersurface HI,0 at , , at which the analysis is straightforward (Ref. 39, Sec. 2), see Fig. 4. (In the spherically symmetric setting, one could also replace the region beyond the Cauchy horizon by a static de Sitter type space, thus not generating any further trapping or horizons and obviating the need for complex absorption; but for Kerr–de Sitter, this gluing procedure is less straightforward to implement; hence we use the above doubling-type procedure for Reissner–Nordström already.) The construction of the extension is detailed in Sec. II A.
We thus study the forcing problem
with f and u supported in the future of the “Cauchy” surface HI in and in the future of HI,0 in . The natural function spaces are weighted b-Sobolev spaces
where the spacetime Sobolev space Hs(M) measures regularity relative to L2 with respect to stationary vector fields, as defined after the statement of Theorem 1.1. More invariantly, , for integer s, consists of L2 functions that remain in L2 upon applying up to s b-vector fields; the space of b-vector fields consists of all smooth vector fields on M, which are tangent to , and is equal to the space of smooth sections of the b-tangent bundle .
Now, (1.2) is an equation on a compact space , which degenerates at the boundary: the operator is a b-differential operator, i.e., a sum of products of b-vector fields, and is a b-ps.d.o. [Note that this point of view is much more precise than merely stating that (1.2) is an equation on a noncompact space !] Thus, the analysis of the operator consists of two parts: first, the regularity analysis, in which one obtains precise regularity estimates for u using microlocal elliptic regularity, propagation of singularities, and radial point results, see Sec. II C, which relies on the precise global structure of the null-geodesic flow discussed in Sec. II B and second, the asymptotic analysis of Secs. II D and II E, which relies on the analysis of the Mellin transformed into (equivalently: Fourier transformed in −t0) operator family , its high energy estimates as , and the structure of poles of , which are known as resonances or quasi-normal modes; this last part, in which we use the shallow resonances to deduce asymptotic expansions of waves, is the only low frequency part of the analysis.
The regularity one obtains for u solving (1.2) with, say, smooth compactly supported (in ) forcing f is determined by the behavior of the null-geodesic flow near the trapping and near the horizons Lj, j = 1, 2, 3. Near the trapping, we use the aforementioned results,29,37,54,70 while near Lj, we use radial point estimates, originating in work by Melrose,49 and proved in the context relevant for us in Ref. 39, Sec. 2; we recall these in Sec. II C and refer the reader to Refs. 72 and 65 for further details. Concretely, Eq. (1.2) combines a forward problem for the wave equation near the black hole exterior region with a backward problem near the artificial exterior region , with hyperbolic propagation in the region between these two (called “no-shift region” in Ref. 31). Near r = r2 and r = r3 then and by propagation estimates in any region , , the radial point estimate, encapsulating the red-shift effect, yields smoothness of u relative to a b-Sobolev space with weight , i.e., allowing for exponential growth (in which case trapping is not an issue), while near r = r1, one is solving the equation away from the boundary X at infinity, and hence the radial point estimate, encapsulating the blue-shift effect, there, yields an amount of regularity that is bounded from above by , where is the surface gravity of . In the extended region , the regularity analysis is very simple, since the complex absorption makes the problem elliptic at the trapping there and at , and one then only needs to use real principal type propagation together with standard energy estimates.
Combined with the analysis of , which relies on the same dynamical and geometric properties of the extended problem as the b-analysis, we deduce in Sec. II D that is Fredholm on suitable weighted b-Sobolev spaces (and in fact solvable for any right hand side f if one modifies f in the unphysical region ). In order to capture the high, respectively, low, regularity near [r2, r3], respectively, r1, these spaces have variable orders of differentiability depending on the location in M. (Such spaces were used already by Unterberger64 and in a context closely related to the present paper in Ref. 6. We present results adapted to our needs in Appendix A.)
In Sec. II E then, we show how the properties of the meromorphic family yield a partial asymptotic expansion of u as in (1.1). Using more refined regularity statements at L1, we show in Sec. II F that the terms in this expansion are in fact conormal to r = r1, i.e., they do not become more singular upon applying vector fields tangent to the Cauchy horizon.
We stress that the analysis is conceptually very simple and close to the analysis in Refs. 6, 33, 39, and 67, in that it relies on tools in microlocal analysis and scattering theory that have been frequently used in recent years.
As a side note, we point out that one could have analyzed in only by proving very precise estimates for the operator , which is a hyperbolic (wave-type) operator in , near r = r1; while this would have removed the necessity to construct and analyze an extended problem, the mechanism underlying our regularity, and decay estimates, namely, the radial point estimate at the Cauchy horizon, would not have been apparent from this. Moreover, the radial point estimate is very robust; it works for Kerr–de Sitter spaces just as it does for the spherically symmetric Reissner–Nordström–de Sitter solutions.
A more interesting modification of our argument relies on the observation that it is not necessary for us to incorporate the exterior region in our global analysis, since this has already been studied in detail before; instead, one could start assuming asymptotics for a wave u in the exterior region and then relate u to a solution of a global, extended problem, for which one has good regularity results and deduce them for u by restriction. Such a strategy, implemented in Ref. 36, is in particular appealing in the study of spacetimes with vanishing cosmological constant using the analytic framework of the present paper, since the precise structure of the “resolvent” has not been analyzed so far, whereas boundedness and decay for scalar waves on the exterior regions of Reissner–Nordström and Kerr spacetimes are known by other methods; see the references at the beginning of Sec. I.
In the remaining parts of Sec. II, we analyze the essential spectral gap for near-extremal black holes in Sec. II G; we find that for any desired level of regularity, one can choose near-extremal parameters of the black hole such that solutions u to (1.2) with f in a finite-codimensional space achieve this level of regularity at . However, as explained in the discussion of Theorem 1.1, it is very likely that shallow resonances cause the codimension to increase as the desired regularity increases. Lastly, in Sec. II H, we indicate the simple changes to our analysis needed to accommodate wave equations on natural tensor bundles.
II. REISSNER–NORDSTRÖM–DE SITTER SPACE
We focus on the case of 4 spacetime dimensions; the analysis in more than 4 dimensions is completely analogous. In the domain of outer communications of the 4-dimensional Reissner–Nordström–de Sitter black hole, given by , with described below, the metric takes the form
where and are the mass and the charge of the black hole, and , with the cosmological constant. Setting Q = 0, this reduces to the Schwarzschild–de Sitter metric. We assume that the spacetime is non-degenerate:
We say that the Reissner–Nordström–de Sitter spacetime with parameters is non-degenerate if has 3 simple positive roots .
Suppose satisfy the non-degeneracy condition (2.2) and denote the three non-negative roots of by . Then for small , the function has three positive roots rj(Q), j = 1, 2, 3, with rj(0) = rj,0, depending smoothly on Q, and .
The existence of the functions rj(Q) follows from the implicit function theorem, taking into account the simplicity of the roots rj,0 of . Let us write ; these are smooth functions of q. Differentiating with respect to q gives , hence , which yields the analogous expansion for r1(Q).
A. Construction of the compactified spacetime
We now discuss the extension of the metric (2.1) beyond the event and cosmological horizon, as well as beyond the Cauchy horizon; the purpose of the present section is to define the manifold on which our analysis of linear waves will take place. See Proposition 2.4 for the final result. We begin by describing the extension of the metric (2.1) beyond the event and the cosmological horizon, thereby repeating the arguments of Ref. 67, Sec. 6; see Fig. 5.
Write , so
We denote by a smooth function such that
small, with cj, smooth near r = rj, to be specified momentarily. (Thus, as and .) We then put
and compute
which is a non-degenerate Lorentzian metric up to r = r2, r3, with dual metric
We can choose cj so as to make dt23 timelike, i.e., : indeed, choosing [which undoes the coordinate change (2.5), up to an additive constant] accomplishes this trivially in [r2, r3] away from ; however, we need cj to be smooth at as well. Now, dt23 is timelike in if and only if , which holds for any . Therefore, we can choose c2 smooth near r2, with for , and c3 smooth near r3, with for , and thus a function such that in the new coordinate system , the metric g extends smoothly to r = r2, r3, and dt23 is timelike for ; and furthermore we can arrange that t23 = t in by possibly changing F23 by an additive constant.
Extending cj smoothly beyond rj in an arbitrary manner, expression (2.6) makes sense for as well as for . We first notice that we can choose the extension cj such that dt23 is timelike also for : indeed, for such r, we have , and the timelike condition becomes , which is satisfied as long as there. In particular, we can take for and for , in which case we get
for with j = 2, and for with j = 3. We define beyond r2 and r3 by the same formula (2.4), using the extensions of c2 and c3 just described; in particular in . We define a time orientation in by declaring dt23 to be future timelike.
We introduce spacelike hypersurfaces in the thus extended spacetime as indicated in Fig. 5, namely,
and
Here and below, the subscript “I” (initial), respectively, “F” (final), indicates that outward pointing timelike vectors are past, respectively, future, oriented. The number in the subscript denotes the horizon near which the surface is located.
We extend the metric to (r0, r1) by defining . We then extend g beyond r = r0 as in (2.6), put
with , when , j = 0, 1, where we set ; further let for , so t01 = t12 in (up to redefining F01 by an additive constant). Then, in -coordinates, the metric g takes form (2.6) near r0, with t23 replaced by t01 and sj = s0 = −1; hence g extends across r = r0 as a non-degenerate stationary Lorentzian metric, and we can choose c0 to be smooth across r = r0 so that dt01 is timelike in and such that moreover in , thus ensuring the (2.7) of the metric (replacing t23 and sj by t01 and −1, respectively).
We can glue the functions t01 and t12 together by defining the smooth function in to be equal to t01 in and equal to t12 in . Define
note here that dt01 is past timelike in . Lastly, we put
Note that in the region , we have produced an artificial horizon at r = r0. Again, the notation is incorrect from the perspective of but is consistent with the notation used in the discussion of .
Let us summarize our construction as follows:
and possess natural partial compactifications M and , respectively, obtained by introducing and adding to them their ideal boundary at infinity, ; the metric g is a non-degenerate Lorentzian b-metric on M and .
The extensions described above amount to a direct construction of a manifold , where we obtained the function by gluing t01 and t12 in , and similarly t12 and t23 in ; we then extend the metric g non-degenerately to a stationary metric in and , thus obtaining a metric g on with the listed properties.
B. Global behavior of the null-geodesic flow
One reason for constructing the compactification step by step is that the null-geodesic dynamics almost decouple in the subdomains , , and , see Figs. 7, 6, and 5.
We denote by G the dual metric of g. We recall that we can glue in , −dr in and in together using a non-negative partition of unity and obtain a 1-form
which is everywhere future timelike in . Thus, the characteristic set of ,
with the dual metric function, globally splits into two connected components
(Indeed, if , then , which is spacelike, so shows that .) Thus, , respectively, , is the union of the past, respectively, future, causal cones. We note that and are smooth codimension 1 submanifolds of in view of the Lorentzian nature of the dual metric G. Moreover, is transversal to , in fact the differentials dG and ( lifted to a function on ) are linearly independent everywhere in .
We begin by analyzing the null-geodesic flow (in the b-cotangent bundle) near the horizons: we will see that the Hamilton vector field HG has critical points where the horizons intersect the ideal boundary Y of ; more precisely, HG is radial there. In order to simplify the calculations of the behavior of HG nearby, we observe that the smooth structure of the compactification , which is determined by the function , is unaffected by the choice of the functions cj in Proposition 2.4, since changing cj merely multiplies by a positive function that only depends on r hence is smooth on our initial compactification . Now, the intersections are smooth boundary submanifolds of M, and we define
which is well-defined given merely the smooth structure on . The point of our observation then is that we can study the Hamilton flow near Lj using any choice of cj. Thus, introducing t0 = t − F(r), with near rj, we find from (2.19) that
Let . Then, with and writing b-covectors as
the dual metric function near Lj is then given by
Correspondingly, the Hamilton vector field is
To study the HG-flow in the radially compactified b-cotangent bundle near , we introduce rescaled coordinates
We then compute the rescaled Hamilton vector field in to be
writing in a local coordinate chart on , we have . Thus, at . In particular,
have opposite signs (by definition of sj), and the quantity that will control regularity and decay thresholds at the radial set Lj is the quotient
see Definition 2.6 and the proof of Proposition 2.9 for their role. We remark that the reciprocal
is equal to the surface gravity of the horizon at r = rj, see, e.g., Ref. 22.
We proceed to verify that is a source/sink for the -flow within by constructing a quadratic defining function of within for which
modulo terms that vanish cubically at Lj; note that has the same relative sign. Now, is defined within by the vanishing of and , and we have , likewise for ; therefore
satisfies (2.29). (One can in fact easily diagonalize the linearization of at its critical set by observing that
modulo quadratically vanishing terms.)
Further studying the flow at r = rj, we note that dr is null there, and writing
a covector is in the orthocomplement of dr if and only if [using the form (2.19) of the metric], which then implies in view of . Since , we deduce that at , where we let
we note that this set is invariant under the Hamilton flow. More precisely, we have , so for j = 3, i.e., at r = r3, dr is in the same causal cone as , hence in the future null cone; thus, letting and taking , we find that lies in the same causal cone as dr, but is not orthogonal to dr, hence we obtain ; more generally,
It follows that forward null-bicharacteristics in can only cross r = r3 in the inward direction (r decreasing), while those in can only cross in the outward direction (r increasing). At r = r0, there is a sign switch both in the definition of (because there is past timelike) and in s0 = −1, so the same statement holds there. At r = r2, there is a single sign switch in the calculation because of s2 = −1, and at r = r1 there is a single sign switch because of the definition of there, so forward null-bicharacteristics in can only cross r = r1 or r = r2 in the inward direction (r decreasing), and forward bicharacteristics in only in the outward direction (r increasing).
Next, we locate the radial sets Lj within the two components of the characteristic set, i.e., determining the components
of the radial sets. The calculations verifying the initial/final character of the artificial hypersurfaces appearing in the arguments of Sec. II A show that at r1 and r3, while at r0 and r2, so since , respectively, , is the union of the future, respectively, past, null cones, we have
In view of (2.26) and taking into account that differs from by an r-dependent factor, while HGr = 0 at Lj, we thus have
We connect this with Fig. 9. Namely, if we let , then is the unstable manifold at Lj,− for j = 0, 1 and the stable manifold at Lj,− for j = 2, 3, and the other way around for . In view of (2.26), Lj,− is a sink for the flow within for j = 0, 1, while it is a source for j = 2, 3, with sink/source switched for the “+” sign. See Fig. 10.
We next shift our attention to the two domains of outer communications, in and in , where we study the behavior of the radius function along the flow using the form (2.18) of the metric: thus, at a point , we have , so HGr = 0 necessitates , hence , and thus we get
Now for ,
vanishes at the radius of the photon sphere, and for ; likewise, for , by construction (2.15), we have only at and for . Therefore, if HGr = 0, then unless , in which case lies in the trapped set
Restricting to bicharacteristics within (which is invariant under the HG-flow since there) and defining
we can conclude that all critical points of along null-geodesics in (r2, r3) [or (r0, r1)] are strict local minima: indeed, if at , then either, in which case unless HGr = 0, hence , or HGr = 0, in which case unless , hence again . As in Ref. 67, Sec. 6.4, this implies that within X, forward null-bicharacteristics in (r2, r3) [respectively, (r0, r1)] either tend to (respectively, ) or they reach r = r2 or r = r3 (respectively, r = r0 or r = r1) in finite time, while backward null-bicharacteristics either tend to (respectively, ) or they reach r = r2 or r = r3 (respectively, r = r0 or r = r1) in finite time. (For this argument, we make use of the source/sink dynamics at .) Further, they cannot tend to , respectively, , in both the forward and backward direction while remaining in (r0, r1), respectively, (r2, r3), unless they are trapped, i.e., contained in , respectively,, since otherwise would attain a local maximum along them. Lastly, bicharacteristics reaching a horizon r = rj in finite time in fact cross the horizon by our earlier observation. The trapping at is in fact r-normally hyperbolic for every r.70
Next, in , we recall that dr is future, respectively, past, timelike in and , respectively, ; therefore, if lies in one of these three regions, implies
[This is consistent with (2.31) and the paragraph following it.]
In order to describe the global structure of the null-bicharacteristic flow, we define the connected components of the trapped set in the exterior domain of the spacetime,
then have stable/unstable manifolds , with the convention that , while is transversal to . Concretely, is the union of forward trapped bicharacteristics, i.e., bicharacteristics that tend to in the forward direction, while is the union of backward trapped bicharacteristics, tending to in the backward direction; further is the union of backward trapped bicharacteristics, and is the union of forward trapped bicharacteristics, tending to , see Fig. 11.
The structure of the flow in the neighborhood of the artificial exterior region is the same as that in the neighborhood of the exterior domain, except the time orientation and thus the two components of the characteristic set are reversed. Write a denote by , the forward and backward trapped sets, with the same sign convention as for above. We note that backward, respectively, forward, trapped null-bicharacteristics in , respectively, may be forward, respectively, backward, trapped in the artificial exterior region, i.e., they may lie in , respectively, , but this is the only additional trapping present in our setup. To state this succinctly, we write
Then we have the following:
The null-bicharacteristic flow in has the following properties:
-
Let be a null-bicharacteristic at infinity, , where . Then in the backward direction, either crosses HI,0 in finite time or tends to , while in the forward direction, either crosses HF,3 in finite time or tends to . The curve can tend to in at most one direction and likewise for .
-
Let be a null-bicharacteristic in . Then in the backward direction, either crosses in finite time or tends to , while in the forward direction, either crosses in finite time or tends to .
-
In both cases, in the region where , is strictly decreasing, respectively, increasing, in the forward, respectively, backward, direction in , while in the regions where or , is strictly increasing, respectively, decreasing, in the forward, respectively, backward, direction in .
-
, as well as and are invariant under the flow.
Here HI,0, etc., is a shorthand notation for .
Statement (3) follows from (2.35), and (4) holds by the definition of the radial and trapped sets. To prove the “backward” part of (1) note that if on , then crosses HI,0 by (2.35); if r = r0 on , then crosses into since . If remains in in the backward direction, it either tends to or it crosses r = r1 since it cannot tend to because of the sink nature of this set. Once crosses into , it must tend to r = r2 by (3) and hence either tend to the source L2,− or cross into . In , must tend to , as it cannot cross r = r2 or r = r3 into or in the backward direction. The analogous statement for , now in the forward direction, is immediate, since reflecting pointwise across the origin in the b-cotangent bundle but keeping the affine parameter the same gives a bijection between backward bicharacteristics in and forward bicharacteristics in . The “forward” part of (1) is completely analogous.
It remains to prove (2). Note that at ; thus in , where is future timelike, is strictly decreasing in the backward direction along bicharacteristics , hence the arguments for part (1) show that crosses HI,0 or tends to if it lies in , otherwise it crosses into in the backward direction. In the latter case, recall that in , is monotonically increasing in the backward direction; we claim that cannot cross HF,2: with the defining function of HF,2, we arranged for df to be past timelike, so for , i.e., f is increasing in the backward direction along the HG-integral curve near HF,2, which proves our claim. This now implies that enters in the backward direction, from which point on is strictly increasing, hence either crosses HI in or it crosses into . In the latter case, it in fact crosses HI by the arguments proving (1). The “forward” part is proved in a similar fashion.
C. Global regularity analysis
Forward solutions to the wave equation in the domain of dependence of HI, i.e., in , are not affected by any modifications of the operator outside, i.e., in . As indicated in Sec. I, we are therefore free to place complex absorbing operators at and L0, which obviate the need for delicate estimates at normally hyperbolic trapping (see the proof of Proposition 2.9) and for a treatment of regularity issues at the artificial horizon [related to in (2.27), see also Definition 2.6].
Concretely, let be a small neighborhood of , with the projection so that
in the notation of Proposition 2.4; thus, stays away from . Choose with Schwartz kernel supported in and real principal principal symbol satisfying
with the inequality strict at , thus is elliptic at . We then study the operator
the convention for the sign of is such that . We will use weighted, variable order b-Sobolev spaces, with weight and the order given by a function ; in fact, the regularity will vary only in the base not in the fibers of the b-cotangent bundle. We refer the reader to [Ref. 6, Appendix A] and Appendix A for details on variable order spaces. We define the function space
as the space of restrictions to of elements of , which are supported in the causal future of ; thus, distributions in are supported distributions at and extendible distributions at (and at ), see Ref. 42, Appendix B; in fact, on manifolds with corners, there are some subtleties concerning such mixed supported/extendible spaces and their duals, which we discuss in Appendix B. The supported character at the initial surfaces, encoding vanishing Cauchy data, is the reason for the subscript “fw” (“forward”). The norm on is the quotient norm induced by the restriction map, which takes elements of with the stated support property to their restriction to . Dually, we also consider the space
consisting of restrictions to of distributions in , which are supported in the causal past of .
Concretely, for the analysis of , we will work on slightly growing function spaces, i.e., allowing exponential growth of solutions in ; we will obtain precise asymptotics (in particular, boundedness) in Secs. II E–II F. In the present section, the stationary nature of the metric g and of near X is irrelevant; only the dynamical structure of the null-geodesic flow and the spacelike nature of the artificial boundaries are used.
Fix a weight
The Sobolev regularity is dictated by the radial sets L1, L2, and L3, as captured by the following definition:
Backward order functions will be used for the analysis of the dual problem.
Note that is a forward order function for the weight if and only if is a backward order function for the weight . The lower, respectively, upper, bounds on the order functions at the radial sets are forced by the propagation estimate (Ref. 39, Proposition 2.1), which will we use at the radial sets: one can propagate high regularity from into the radial set and into the boundary (“red-shift effect”), while there is an upper limit on the regularity one can propagate out of the radial set and the boundary into the interior of the spacetime (“blue-shift effect”); the definition of order functions here reflects the precise relationship of the a priori decay or growth rate and the regularity (i.e., the “strength” of the red- or blue-shift effect depending on a priori decay or growth along the horizon). We recall the radial point propagation result in a qualitative form (the quantitative version of this, yielding estimates, follows from the proof of this result or can be recovered from the qualitative statement using the closed graph theorem).
(Ref. 39, Proposition 2.1). Suppose is as above and let . Let j = 1, 2, 3.
If , , and if , then (and thus a neighborhood of ) is disjoint from provided , , and in a neighborhood of , is disjoint from .
On the other hand, if and if , then (and thus a neighborhood of ) is disjoint from provided and a punctured neighborhood of , with removed, in is disjoint from .
We then have the following:
Both estimates hold in the sense that if the quantities on the right hand side are finite, then so is the left hand side and the inequality is valid.
The arguments are very similar to the ones used in Ref. 39, Sec. 2.1. The proof relies on standard energy estimates near the artificial hypersurfaces, various microlocal propagation estimates, and crucially relies on the description of the null-bicharacteristic flow given in Proposition 2.5.
Then, in , we use the propagation of singularities (forwards in , backwards in ) to obtain local -regularity away from the boundary at infinity, . At the radial sets L2 and L3, the radial point estimate, Proposition 2.8, allows us, using the a priori -regularity of u, to propagate -regularity into ; propagation within then shows that we have -control on u on . Since , we can then use [Ref. 37, Theorem 3.2] to control u in microlocally at and propagate this control along . Near HF,3, the microlocal propagation of singularities only gives local control away from HF,3, but we can get uniform regularity up to HF,3 by standard energy estimates, using a cutoff near HF,3 and the propagation of singularities for an extended problem (solving the forward wave equation with forcing , cut off near HF,3, plus an error term coming from the cutoff), see Ref. 39, Proposition 2.13 and the similar discussion around (2.42) below in the present proof. We thus obtain an estimate for the -norm of u in .
Next, we propagate regularity in , using part (3) of Proposition 2.5 and our assumption ; the only technical issue is now at HF,2, where the microlocal propagation only gives local regularity away from HF,2; this will be resolved shortly.
Focusing on the remaining region , we start with the control on u near HI,0, which we propagate forwards in and backwards in either up to HF or into the complex absorption hiding ; see Ref. 67, Sec. 2 for the propagation of singularities with complex absorption. (This is a purely symbolic argument; hence the present b-setting is handled in exactly the same way as the standard ps.d.o. setting discussed in the reference.) Moreover, at the elliptic set of the complex absorbing operator , we get -control on u, and we can propagate -estimates from there. The result is that we get -estimates of u in a punctured neighborhood of L1 within ; thus, the low regularity part of Proposition 2.8 applies. We can then propagate regularity from a neighborhood L1 along . This gives us local regularity away from , where the microlocal propagation results do not directly give uniform estimates.
Putting all these estimates together, we obtain an estimate for in terms of .
The proof of the dual estimate is completely analogous: we now obtain initial regularity (that we can then propagate as above) by solving the backward problem for near and HF,3.
D. Fredholm analysis and solvability
The estimates in Proposition 2.9 do not yet yield the Fredholm property of . As explained in Ref. 39, Sec. 2, we therefore study the Mellin-transformed normal operator family , see (Ref. 48, Sec. 5.2), which in the present (dilation-invariant in , or translation-invariant in ) setting is simply obtained by conjugating by the Mellin transform in or equivalently the Fourier transform in , i.e., , acting on functions on the boundary at infinity . Concretely, we need to show that is invertible between suitable function spaces on for a weight , since this will allow us to improve the error term in (2.40) by a space with an improved weight, so injects compactly into it; an analogous procedure for the dual problem gives the full Fredholm property for ; see Ref. 39 and below for details. As in Sec. II C, only dynamical and geometric properties of the metric g and the operator are used here; in fact, only their properties at infinity matter for the analysis of , which is in general defined by conjugating the normal operator , obtained by freezing coefficients of at the boundary X [i.e., can be thought of as the stationary part of ], by the Mellin transform in .
For any finite value of , we can analyze the operator , , using standard microlocal analysis (and energy estimates near and ). The natural function spaces are variable order Sobolev spaces
which we define to be the restrictions to of elements of with support in , and dually on , the restrictions to Y of elements of with support in , obtaining Fredholm mapping properties between suitable function spaces. However, in order to obtain useful estimates for our global b-problem, we need uniform estimates for as in strips of bounded , on function spaces that are related to the variable order b-Sobolev spaces on which we analyze .
Thus, let , , and consider the semiclassical rescaling (Ref. 67, Sec. 2)
We refer to Ref. 40, Sec. 4 for details on the relationship between the b-operator and its semiclassical rescaling; in particular, we recall that the Hamilton vector field of the semiclassical principal symbol of for is naturally identified with the Hamilton vector field of the b-principal symbol of restricted to , where we use the coordinates (2.30) in the b-cotangent bundle. For any Sobolev order function and a weight , the Mellin transform in gives an isomorphism
where [for ] is a semiclassical variable order Sobolev space with a non-constant weighting in h; see Appendix A for definitions and properties of such spaces.
The analysis of , , acting on -type spaces is now straightforward, given the properties of the Hamilton flow of . Indeed, in view of the supported/extendible nature of the b-spaces Hb,fw and Hb,bw into account, we are led to define the corresponding semiclassical space
to be the space of restrictions to Y of elements of with support in , respectively, . Then, in the region where is not constant (recall that this is a subset of ), is a (semiclassical) real principal type operator, as follows from (2.35), and hence the only microlocal estimates we need there are elliptic regularity and the real principal type propagation for variable order semiclassical Sobolev spaces; these estimates are proved in Propositions A.4 and A.5. The more delicate estimates take place in standard semiclassical function spaces; these are the radial point estimates near r = rj, in the present context proved in Ref. 67, Sec. 2, and the semiclassical estimates of Wunsch–Zworski70 and Dyatlov29 (microlocalized in Ref. 40, Sec. 4) at the normally hyperbolic trapping. Near the artificial hypersurfaces , intersected with , the operator is a (semiclassical) wave operator, and we use standard energy estimates there similar to the proof of Proposition 2.9, but keeping track of powers of h; see Ref. 67, Sec. 3 for details.
We thus obtain the following:
Notice here that if were constant, the estimate (2.46) would read , which is the usual hyperbolic loss of one derivative and one power of h. The estimate (2.46) is conceptually the same, but in addition takes care of the variable orders. Trapping causes no additional losses here, since .
We have and ; the change of sign in when going from (2.46) to the dual estimate (2.47) is analogous to the change of sign in the weight in Proposition 2.9.
Using the above notation, the (uniform) estimates (2.46) and (2.47) hold with replaced by on the right hand sides, provided , where .
The effect of replacing by is that this adds an additional h−1 to the right hand side, i.e., we get a weaker estimate (which in the presence of trapping cannot be avoided by Ref. 7); the strengthening of the norm in the regularity sense is unnecessary, but does not affect our arguments later.
Let be an order function satisfying both (2.50) and (2.53), so by the above discussion, is Fredholm. Since , we a forteriori get the finite-dimensionality of . On the other hand, if annihilates , it also annihilates , hence we can find solving . The propagation of singularities, Proposition 2.9, implies , and the proof is complete.
To obtain a better result, we need to study the structure of resonances. Notice that for the purpose of dealing with a single resonance, one can simplify the notation by working with the space , see (2.43), rather than , since the semiclassical (high energy) parameter is irrelevant then.
(1) Every resonant state corresponding to a resonance with is supported in the artificial exterior region ; more precisely, every element in the range of the singular part of the Laurent series expansion of at such a resonance is supported in . In fact, this holds more generally for any , which is not a resonance of the forward problem for the wave equation in a neighborhood of the black hole exterior.
(2) If R denotes the restriction of distributions on Y to , then the only pole of with is at , has rank 1, and the space of resonant states consists of constant functions.
Since u has supported character at , we obtain in , since u solves the wave equation there. On the other hand, the forward problem for the wave equation in the neighborhood of the black hole exterior does not have any resonances with positive imaginary part; this is well-known for Schwarzschild–de Sitter spacetimes8,55 and for slowly rotating Kerr–de Sitter spacetimes, either by direct computation26 or by a perturbation argument.38,67 For the convenience of the reader, we recall the argument for the Schwarzschild–de Sitter case, which applies without change in the present setting as well: a simple integration by parts argument, see, e.g., Ref. 26 or (Ref. 38, Sec. 2), shows that u must vanish in . Now the propagation of singularities at radial points implies that u is smooth at r = r2 and r = r3 (where the a priori regularity exceeds the threshold value), and hence in , u is a solution to the homogeneous wave equation on an asymptotically de Sitter space that decays rapidly at the conformal boundary (which is r = r3) hence must vanish identically in [see (Ref. 67, Footnote 58) for details]; the same argument applies in , yielding there. Therefore, , as claimed. An iterative argument, similar to (Ref. 6, Proof of Lemma 8.3), yields the more precise result.
The more general statement follows along the same lines (and is in fact much easier to prove, since it does not entail a mode stability statement): suppose is not a resonance of the forward wave equation on , then a resonant state must vanish in , and we obtain as before; likewise for the more precise result. This proves (1).
For the proof of (2), it remains to study the resonance at 0, since the only resonance in the closed upper half plane is 0. Note that an element in the range of the most singular Laurent coefficient of at lies in ; but elements in , which vanish near r = r1 vanish identically in and hence are annihilated by R, while elements that are not identically 0 near r = r1 are not identically 0 in as well but the only non-trivial elements of (which are smooth at r2 and r3) are constant in , and since in , we deduce (by unique continuation) that indeed consists of constant functions. But then the order of the pole of at equals the order of the 0-resonance of the forward problem for in , which is known to be equal to 1, see the references above. The 1-dimensionality of then implies that the rank of the pole of at 0 indeed equals 1.
Since we are dealing with an extended global problem here, involving (pseudodifferential!) complex absorption, solvability is not automatic but it holds in the region of interest ; to show this, we first need the following:
We can then conclude:
Under the assumptions of Theorem 2.13, all elements in the kernel of in (2.54) are supported in the artificial exterior domain .
Moreover, for all with support in , there exists such that in .
If lies in , then the supported character of u at together with uniqueness for the wave equation in and implies that u vanishes identically there, giving the first statement.
For the second statement, we use Lemma 2.15 and solve the equation , which gives the desired u.
In particular, solutions of the equation exist and are unique in , which we of course already knew from standard hyperbolic theory in the region on “our” side of the Cauchy horizon; the point is that we now understand the regularity of u up to the Cauchy horizon. We can refine this result substantially for better-behaved forcing terms, e.g., for with support in ; we will discuss this in Secs. II E and II F.
E. Partial asymptotics and decay
The only resonance of the forward problem in in is a simple resonance at , with resonant states equal to constants, see the references given in the proof of Lemma 2.14, and there exists such that 0 is the only resonance in . (This does not mean that the global problem for does not have other resonances in this half space!) In the notation of Proposition 2.12, we may assume so that we have high energy estimates in .
Our proof uses the stationarity of g (and ) near X, which allows us to pass freely between and the Mellin-transformed normal operator family ; see also Remark 2.18.
Thus, we have shown regularity in the region where , i.e., where we did not cut off; however, considering (2.56) on an enlarged domain and running the argument there, with the cutoff supported in the enlarged domain and identically 1 on , we obtain the full regularity result upon restricting to .
Now, by Lemma 2.14 (1), all resonant states of , which are not resonant states of the forward problem in must in fact vanish in , and by part (2) of Lemma 2.14, the only term in (2.58) that survives upon restriction to is the constant term.
Thus, we obtain a partial expansion with a remainder that decays exponentially in in an L2 sense; we will improve this in particular to decay in Sec. II F.
If g and were not dilation-invariant, then in the partial expansion (2.58), one would not be able to show improved regularity for at r = r1 in general because r = r1 (or rather ) no longer has a geometric meaning as the stable/unstable manifold of the radial set L1. (See also the setup leading to Proposition 2.23 below.) Concretely, has an error term that in general loses two derivatives, which cannot be recouped by Proposition 2.9. On the other hand, assuming that is characteristic for (or choosing the dilation orbits of more carefully), and tracking the singular nature of the resonant states more precisely should allow for the above proposition to generalize to the non-dilation-invariant setting; however, we do not pursue this further here.
F. Conormal regularity at the Cauchy horizon
Suppose u solves (2.56), hence it has an expansion (2.57). For any Killing vector field V, we then have ; now if solves the global problem (using the extension operator from Lemma 2.15), then in by the uniqueness for the Cauchy problem in this region. But by Proposition 2.17, has an expansion like (2.57), with constant term vanishing because X annihilates the constant term in the expansion of u, and therefore lies in space near the Cauchy horizon {r = r1} as well. More generally, we can take V to be any (finite) product of Killing vector fields, and therefore obtain
where is arbitrary, and the vector fields Vj, , are equal to or rotation vector fields on the -factor of the spacetime, independent of . (This uses that g is stationary!) These vector fields are all tangent to the Cauchy horizon. We obtain for any small open interval containing r1 that
A posteriori, by Sobolev embedding, this gives
Using the notation of Proposition 2.17, the solution u of (2.56) has an asymptotic expansion with , and there exists a constant such that . In particular, u is uniformly bounded in and extends continuously to .
Translated back to , the estimate on the remainder states that for scalar waves, one has exponentially fast pointwise decay to a constant. This recovers Franzen’s boundedness result31 for linear scalar waves on the Reissner–Nordström spacetime near the Cauchy horizon in the cosmological setting.
The above argument is unsatisfactory in two ways: first, they are not robust and in particular do not quite apply in the Kerr–de Sitter setting discussed in Sec. III; however, see Remark 3.4, which shows that using a “hidden symmetry” of Kerr–de Sitter space related to the completely integrable nature of the geodesic equation, one can still conclude boundedness in this case. Second, the regularity statement (2.59) is somewhat unnatural from a PDE perspective; thus, we now give a more robust microlocal proof of the conormality of , i.e., iterative regularity under application of vector fields tangent to r = r1, which relies on the propagation of conormal regularity at the radial set L1, see Proposition 2.23.
Since L1 is Lagrangian and thus in particular coisotropic, the first statement follows from the symbol calculus.
In particular:
If u is a resonant state of , i.e., , then u is conormal to r = r1 relative to , i.e., for any number of vector fields on X which are tangent to r = r1, we have .
Indeed, by the propagation of singularities, u is smooth away from , and then Lemma 2.21 implies the stated conormality property.
Suppose is as above, and let , .
If and if , then (and thus a neighborhood of ) is disjoint from for all provided for , and provided a punctured neighborhood of , with removed, in is disjoint from .
Thus, if is conormal to , i.e., remains in microlocally under iterative applications of elements of —this in particular holds if —then u is conormal relative to , provided u lies in in a punctured neighborhood of L1. Using Proposition 2.23 at the radial set L1 in the part of the proof of Proposition 2.17 where the regularity of is established, we obtain the following:
The same result holds true, without the constant term u0, for the forward solution of the massive Klein–Gordon equation , small.
For the massive Klein–Gordon equation, the only change in the analysis is that the simple resonance at 0 moves into the lower half plane, see, e.g., the perturbation computation in Ref. 39, Lemma 3.5; this leads to the constant term u0, which was caused by the resonance at 0, being absent.
This implies estimate (2.59) and thus yields Corollary 2.19 as well.
G. Existence of high regularity solutions at the Cauchy horizon for near-extremal black holes
The amount of decay (and thus the amount of regularity we obtain) in Theorem 2.24 is directly linked to the size of the spectral gap, i.e., the size of the resonance-free strip below the real axis, as explained in Sec. II E. Due to the work of Sá Barreto–Zworski55 in the spherically symmetric case and general results by Dyatlov28 at (r-)normally hyperbolic trapping (for every r), the size of the essential spectral gap is given in terms of dynamical quantities associated with the trapping, see Proposition 2.12; we recall that the essential spectral gap is the supremum of all such that there are only finitely many resonances above the line . Thus, the essential spectral gap only concerns the high energy regime, i.e., it does not give any information about low energy resonances. In this section, we compute the size of the essential spectral gap in some limiting cases; the possibly remaining finitely many resonances between 0 and the resonances caused by the trapping will be studied separately in future work. We give some indications of the expected results in Remark 2.25.
In order to calculate the relevant dynamical quantities at the trapped set, we compute the linearization of the flow in the variables at the trapped set : we have
modulo functions vanishing quadratically at , and in the same sense
which in view of [see also (2.34)] gives
Therefore, the expansion rate of the flow in the normal direction at is equal to
To find the size of the essential spectral gap for the forward problem of , we need to compute the size of the imaginary part of the subprincipal symbol of the semiclassical rescaling of at the semiclassical trapped set. Put , , then
With , , we thus obtain
The essential spectral gap thus has size at least provided , so
We compute the quantity on the right for near-extremal Reissner–Nordström–de Sitter black holes with very small cosmological constant; first, using the radius of the photon sphere for the Reissner–Nordström black hole with ,
and the radius of the Cauchy horizon
we obtain
for the size of the essential spectral gap for resonances caused by the trapping in the case . [For Q = 0, one finds , which agrees with [Ref. 27, Eq. (0.3)] for .] In the extremal case , we find . Furthermore, we have
Thus, ; therefore,
which blows up as ; this corresponds to the fact the surface gravity of extremal black holes vanishes. Given , we can thus choose small enough so that , and then taking to be small, the same relation holds for the -dependent quantities and . Since there are only finitely many resonances in any strip , we conclude by Theorem 2.24, taking close to that for forcing terms f, which are orthogonal to a finite-dimensional space of dual resonant states (corresponding to resonances in ), the solution u has regularity at the Cauchy horizon. Put differently, for near-extremal Reissner–Nordström–de Sitter black holes with very small cosmological constant , waves with initial data in a finite codimensional space (within the space of smooth functions) achieve any fixed order of regularity at the Cauchy horizon, in particular better than .
Numerical investigations of linear scalar waves9–11 suggest that there are indeed resonances roughly at , j = 2, 3, where and are the surface gravities of the cosmological horizon, see (2.28); as , we have , and for extremal black holes with , we have . (On the static de Sitter spacetime, there is a resonance exactly at , as a rescaling shows: for , one has decay to constants away from the cosmological horizon, t0 the static time coordinate, see, e.g., Ref. 66; now static de Sitter space with cosmological constant can be mapped to dS3 via , , where is the surface gravity of the cosmological horizon, and t0, r0, respectively, t, r, are static coordinates on dS3, respectively, . Under this map, the metric on dS3 is pulled back to a constant multiple of the metric on . Thus, waves on decay to constants with the speed , which corresponds to a resonance at .)
Our analysis is consistent with the numerical results, assuming the existence of these resonances: we expect linear waves in this case to be generically no smoother than at the Cauchy horizon, which highlights the importance of the relative sizes of the surface gravities for understanding the regularity at the Cauchy horizon. For near-extremal black holes, where , this gives , thus the local energy measured by an observer crossing the Cauchy horizon is of the order , which diverges in view of ; this agrees with [Ref. 11, Eq. (9)]. We point out however that the waves are still in if , which is satisfied for near-extremal black holes. This is analogous to Sbierski’s criterion (Ref. 56, Sec. 4.4) for ensuring the finite energy of waves at the Cauchy horizon of linear waves with fast decay along the event horizon.
The rigorous study of resonances associated with the event and cosmological horizons will be subject of future work.
H. Tensor-valued waves
The analysis presented in Secs. II A–II G goes through with only minor modifications if we consider the wave equation on natural vector bundles.
For definiteness, we focus on the wave equation, more precisely the Hodge d’Alembertian, on differential k-forms, . In this case, mode stability and asymptotic expansions up to decaying remainder terms in the region , a neighborhood of the black hole exterior region, were proved in Ref. 38. The previous arguments apply to ; the only difference is that the threshold regularity at the radial points at the horizons shifts. At the event horizon and the cosmological horizon, this is inconsequential, as we may work in spaces of arbitrary high regularity there; at the Cauchy horizon however, one has, fixing a time-independent positive definite inner product on the fibers of the k-form bundle with respect to which one computes adjoints,
at , with , and and endomorphism on the k-form bundle; and one can compute that the lowest eigenvalue of (which is self-adjoint with respect to the chosen inner product) is equal to −k. But then the regularity one can propagate into for , , solving , compactly supported and smooth, is , as follows from Ref. 39, Proposition 2.1 and Footnote 5. Thus, in the partial asymptotic expansion in Theorem 2.24 (which has a different leading order term now, coming from stationary k-form solutions of the wave equation), we can only establish conormal regularity of the remainder term at the Cauchy horizon relative to the space , which for small gives Sobolev regularity , for small . Assuming that the leading order term is smooth at the Cauchy horizon (which is the case, for example, for 2-forms, see Ref. 38, Theorem 4.3), we therefore conclude that, as soon as we consider k-forms u with , our methods do not yield uniform boundedness of u up to the Cauchy horizon; however, we remark that the conormality does imply uniform bounds as of the form , small.
A finer analysis would likely yield more precise results, in particular boundedness for certain components of u, and, as in the scalar setting, a converse result, namely, showing that such a blow-up does happen, is much more subtle. We do not pursue these issues in the present work.
III. KERR–DE SITTER SPACE
We recall from Ref. 67, Sec. 6, the form of the Kerr–de Sitter metric with parameters (cosmological constant), (black hole mass) and a (angular momentum),
where
[Our are denoted in Ref. 67, while our are denoted there.] In order to guarantee the existence of a Cauchy horizon, we need to assume . Analogous to Definition 2.1, we make a non-degeneracy assumption:
We say that the Kerr–de Sitter spacetime with parameters is non-degenerate if has 3 simple positive roots .
We consider a simple case in which non-degeneracy can be checked immediately.
Suppose and denote the three non-negative roots of by . Then for small , has three positive roots rj(a), j = 1, 2, 3, with rj(0) = rj,0, depending smoothly on a2, and .
We recall that the condition (2.2) ensures the existence of the roots rj,0 as stated. One then computes for that , giving the first statement.
In order to state unconditional results later on, we in fact from now on assume to be in the setting of this lemma, i.e., we consider slowly rotating Kerr–de Sitter black holes; see Remark 3.6 for further details.
A. Construction of the compactified spacetime
As in Sec. II A, we discuss the smooth extension of the metric g across the horizons and construct the manifold on which the linear analysis will take place; all steps required for this construction are slightly more complicated algebraically but otherwise very similar to the ones in the Reissner–Nordström–de Sitter setting, so we shall be brief.
Thus, with
we will take
for r near rj, where
Using