Weak null singularities in general relativity

By Jonathan Luk

Abstract

We construct a class of spacetimes (without symmetry assumptions) satisfying the vacuum Einstein equations with singular boundaries on two null hypersurfaces intersecting in the future on a 2-sphere. The metric of these spacetimes extends continuously beyond the singularities while the Christoffel symbols fail to be square integrable in a neighborhood of any point on the singular boundaries. The construction shows moreover that the singularities are stable in a suitable sense. These singularities are stronger than the impulsive gravitational spacetimes considered by Luk and Rodnianski, and conjecturally they are present in the interior of generic black holes arising from gravitational collapse.

1. Introduction

In this paper we study the existence and stability of weak null singularities in general relativity without symmetry assumptions. More precisely, a weak null singularity is a singular null boundary of a spacetime solving the Einstein equations

such that the Christoffel symbols blow up and are not square integrable while the metric is continuous up to the boundary. This can be interpreted as a terminal singularity of the spacetime as it cannot be made sense of as a weak solution⁠Footnote1 to the Einstein equations along the singular boundary. While the singularity is sufficiently strong to be terminal, it is at the same time sufficiently weak such that the metric in an appropriate coordinate system is continuous up to the boundary.

1

One can define a weak solution to the Einstein equations by requiring in the weak sense for all compactly supported smooth vector fields and . After integration by parts, the minimal regularity required for the spacetime for this to be defined is that the Christoffel symbols are square integrable; see the discussion in Reference 5, p.13.

The study of weak null singularities began with the attempts to understand the (in)stability of the Cauchy horizon in the black hole interior of Reissner–Nordström spacetimes. Reissner–Nordström spacetimes are the unique two-parameter family of asymptotically flat (with two ends), spherically symmetric, static solutions to the Einstein–Maxwell equations. Their Penrose diagrams⁠Footnote2 are given by Figure 1. As seen in the Penrose diagram, the Reissner–Nordström solution possesses a smooth Cauchy horizon in the interior of the black hole such that the spacetime can be extended nonuniquely as a smooth solution to the Einstein–Maxwell system. This feature is also shared⁠Footnote3 by the Kerr family of solutions to the vacuum Einstein equations, which can also be depicted by a Penrose diagram given by Figure 1. According to the strong cosmic censorship conjecture (see section 1.1 below), the Reissner–Nordström and Kerr spacetimes are expected to be nongeneric and the smooth Cauchy horizons are expected to be unstable.

2

for

3

for

In a seminal work Dafermos Reference 7Reference 8 showed that for a spacetime solution to the spherically symmetric Einstein–Maxwell–real scalar field system, if an appropriate upper and lower bound for the scalar field is assumed on the event horizon, then in a neighborhood of timelike infinity, the black hole terminates in a weak null singularity. The necessary upper bound was shown to hold for nonextremal black hole spacetimes arising from asymptotically flat initial data by Dafermos and Rodnianski Reference 10. In particular this implies that near timelike infinity, the terminal boundary of the Cauchy development does not contain a spacelike portion.

In a more recent work Reference 9, Dafermos showed that if, in addition to assuming the two black hole exterior regions settle to Reissner–Nordström with appropriate rates, the initial data are moreover globally close to that of Reissner–Nordström, then the maximal Cauchy development of the data possesses the same Penrose diagram as Reissner–Nordström. In particular the spacetime terminates in a global bifurcate weak null singularity and the singular boundary does not contain any spacelike portion.

The works Reference 7Reference 8Reference 9 were in part motivated by the physics literature on the instability of Cauchy horizons, weak null singularities and the strong cosmic censorship conjecture. It will be discussed below in section 1.1.

While the works of Dafermos Reference 7Reference 8Reference 9 are restricted to the class of spherically symmetric spacetimes, they nonetheless suggest the genericity of weak null singularities in the black hole interior, at least “in a neighborhood of timelike infinity”. In particular they motivate the following conjecture for the vacuum Einstein equations,

Conjecture 1.
(1)

Consider the characteristic initial value problem with smooth characteristic initial data on a pair of null hypersurfaces and intersecting on a -sphere. Suppose that is an affine complete null hypersurface on which the data approach that of the event horizon of a Kerr solution (with at a sufficiently fast polynomial rate.⁠Footnote4 Then the development of the initial data possesses a null boundary “emanating from timelike infinity through which the spacetime is extendible with a continuous metric (see shaded region in Figure 2). Moreover, given an appropriate “lower bound” on the , this piece of null boundary is generically a weak null singularity with nonsquare-integrable Christoffel symbols.

4

In particular this applies if an asymptotically flat spacetime has an exterior region which approaches a subextremal Kerr solution at a sufficiently fast polynomial rate. This also holds in the case where the Cauchy hypersurface has only one asymptotically flat end. In that case, numerical work in spherical symmetry Reference 13 suggests that the singular boundary may also contain a nonempty spacelike portion, in addition to the null portion.

(2)

(Ori, see discussion in Reference 9) If the data for on a complete two-ended asymptotically flat Cauchy hypersurface are globally a small perturbation of two-ended Kerr initial data (with ), then the maximal Cauchy development possesses a global bifurcate future null boundary . Moreover, for generic such perturbations of Kerr, is a global bifurcate weak null singularity which intersects every futurely causally incomplete geodesic.

If Conjecture 1 is true, then in particular there exist local stable weak null singularities for the vacuum Einstein equations without symmetry assumptions. We show in this paper that there is in fact a large class of such singularities, parameterized by singular initial data. More specifically, we solve a characteristic initial value problem with singular initial data and construct a class of stable bifurcate weak null singularities.

To motivate the strength of the singularity considered in this paper, we first recall the strength of the spherically symmetric weak null singularities in a neighborhood of Reissner–Nordström studied in Reference 8. The instability of the Reissner–Nordström Cauchy horizon is in fact already suggested by a linear analysis (see Reference 4Reference 20Reference 23). For a spherically symmetric solution to the linear wave equation which has a polynomially decaying (in the Eddington–Finkelstein coordinates) tail⁠Footnote5 along the event horizon, there is a singularity in a ()-regular coordinate system near the Cauchy horizon of the strength⁠Footnote6

5

with upper and lower bounds

6

This statement regarding the linear wave equation can be inferred using the methods in Reference 7 for the nonlinear coupled Einstein–Maxwell–scalar field system.

for some as . In particular along an outgoing null curve, is integrable but not -integrable for any . In the spacetimes constructed by Dafermos Reference 7Reference 8, it was shown moreover that even in the nonlinear setting, is also singular but remains integrable. A more precise analysis will show that in fact the spherically symmetric scalar field in the nonlinear setting of Reference 8 also blows up at a rate given by Equation 2.

Returning to the problem of constructing stable weak null singularities in vacuum, our construction is based on solving a characteristic initial value problem with singular data. We will in fact construct spacetimes not only with one weak null singularity, but instead they will contain two weak null singularities terminating at a bifurcate sphere. More precisely, the data on the initial characteristic hypersurface (resp. ) is determined by the traceless part of the null second fundamental form (resp. ). We consider singular initial data satisfying in particular

and

This singularity is consistent with the strength of the weak null singularities in Equation 2.

The following is a first version of the main result of this paper (see Figure 3). We refer the readers to the statement of Theorems 2, 3 and 4 for a more precise formulation of the theorem.

Theorem 1 (Main theorem, first version).

For a class of singular characteristic initial data without any symmetry assumptions for the vacuum Einstein equations

with the singular profile as above (see precise requirements on the data in section 1.3) and for sufficiently small and , , there exists a unique smooth spacetime endowed with a double null foliation in , , which satisfies the vacuum Einstein equations with the given data. Associated to , there exists a coordinate system such that the metric extends continuously to the boundary but the Christoffel symbols are not in .

Remark 1.

This class of stable local weak null singularities that we construct in particular provides the first construction of weak null singularities of such strength for the vacuum Einstein equations.⁠Footnote7

7

We recall Birkhoff’s theorem which states that the only spherically symmetric vacuum spacetimes are the Minkowski and Schwarzschild solutions. Thus to construct stable examples of weak null singularities in vacuum, one necessarily works outside the class of spherically symmetric spacetimes.

Theorem 1 allows singularities on both initial null hypersurface and is valid in the region where and are sufficiently small. In the context of the interior of black holes, this corresponds to the darker shaded region in Figure 4. The existence theorem clearly implies an existence result when the data are only singular on one of the initial null hypersurfaces. In that context, we can in fact combine the methods in this paper with that in Reference 17 to show that the domain of existence can be extended so that only one of the characteristic length scales is required to be small. More precisely, we allow that data on such that

on and the data on are smooth on . Then for sufficiently small, the spacetime remains smooth in , (see for example the lightly shaded region in Figure 4). We will omit the details of the proof of this result.

Theorem 1, which proves the existence and stability of the conjecturally generic weak null singularities, can be viewed as a first step toward Conjecture 1. A next step is an analogue of Reference 8 for the vacuum Einstein equations without symmetry assumptions, i.e., to solve the characteristic initial value problem inside the black hole with data prescribed on the event horizon that is approaching Kerr at appropriate rates. This requires an understanding of the formation of weak null singularities from smooth data on the event horizon (see part (1) of Conjecture 1). A full resolution of Conjecture 1, part (2), however, requires in addition an understanding of the decay rates of gravitational radiation along the event horizon for generic perturbations of Kerr spacetime. This latter problem is intimately tied to the problem of the nonlinear stability of Kerr spacetimes, which continues to be one of the most important and challenging open problems in mathematical general relativity. Nevertheless, significant progress has been made for the corresponding linear problem in the past decade. We refer the readers to the survey of Dafermos and Rodnianski Reference 11 for more about this linear problem.

The approach for the main theorem applies equally well to the Einstein–Maxwell–scalar field system without symmetry assumptions.⁠Footnote8 Thus, we show that the weak null singularity of Dafermos Reference 8, which arises from appropriately decaying data on the event horizon, is stable against nonspherically symmetric perturbations on the hypersurface sufficiently far within the black hole region (see Figure 5).

8

This can be easily seen by decomposing the Maxwell field and the gradient of the scalar field in terms of the null frame below. The components in this decomposition obey equations that can be put in the same schematic form as in section 2.4. Therefore, the Maxwell field and the scalar field and their derivatives satisfy estimates similar to those for the Ricci coefficients and curvature components.

1.1. Weak null singularities and strong cosmic censorship conjecture

The study of weak null singularities can be viewed in the larger context of Penrose’s celebrated strong cosmic censorship conjecture in general relativity. The conjecture states that for generic asymptotically flat initial data for “reasonable” Einstein-matter systems, the maximal Cauchy development is future inextendible as a suitably regular Lorentzian manifold. This would guarantee general relativity to be a deterministic theory.

As pointed out above, the Kerr and Reissner–Nordström families of solutions (of the Einstein vacuum and Einstein–Maxwell equations, respectively) have maximal Cauchy developments that are extendible as larger smooth spacetimes unless the angular momentum or the charge vanishes. This is connected with the existence of a smooth Cauchy horizon in the black hole interior such that the spacetime can be extended beyond as a smooth solution. According to the strong cosmic censorship conjecture, this is expected to be nongeneric.

On the other hand, the situation for the Schwarzschild spacetime is more preferable from the point of view of the deterministic nature of the theory. The maximal development of the Schwarzschild spacetime terminates with a spacelike singularity at which the Hawking mass and the curvature scalar invariants blow up. In particular the spacetime cannot be extended in .

The early motivation for the strong cosmic censorship conjecture, besides the desirability of a deterministic theory, is a linear heuristic argument by Penrose Reference 23 suggesting that the Reissner–Nordström Cauchy horizon is unstable. This was also confirmed by the numerical work by Simpson and Penrose Reference 27. It is thus conjectured that a small global perturbation would lead to a singularity in the interior of the black hole in such a way that the maximal Cauchy development is future inextendible.

However, the nature of the singular boundary in the interior of black holes was not well understood⁠Footnote9 until the first study of weak null singularity carried out by Hiscock Reference 12. In an attempt to understand the instability of the Reissner–Nordström Cauchy horizon, he considered the Vaidya model allowing for a self-gravitating ingoing null dust. In this model, an explicit solution can be found, and he showed that various components of the Christoffel symbols blow up. This, however, was called a whimper singularity as the Hawking mass and the curvature scalar invariants remain bounded.

9

In particular it was believed that a perturbation of the Reissner–Nordström Cauchy horizon would lead to a Schwarzschild type singularity.

In subsequent works, Poisson and Israel Reference 25Reference 26 added an outgoing null dust to the model considered by Hiscock. While explicit solutions were not available, they were able to deduce that the second outgoing null dust would cause the Hawking mass to blow up at the null singularity. It was then thought of as a stronger singularity than that of Hiscock.

However, from the point of view of partial differential equations, it is more natural to view this singularity at the level of the nonsquare-integrability of the Christoffel symbols, which is exactly the threshold such that the spacetime cannot be defined as a weak solution to the Einstein equations. From this perspective, the singularity of Poisson and Israel is as strong as that of Hiscock, and both singularities can be viewed as terminal boundaries for the spacetimes in question.

While the Christoffel symbols blow up at the Cauchy horizon, one can also think that the Cauchy horizon is “stable” in the sense that no singularity arises before the “original Cauchy horizon”. In particular there is no spacelike portion of the singular boundary in a neighborhood of timelike infinity. Thus, this is contrary to the case of the Schwarzschild spacetime. This weak null singularity picture has been further explored and justified in many numerical works (see Reference 1Reference 2Reference 3).

As we described before, the aforementioned picture of the interior of black holes was finally established by Dafermos in the context of the spherically symmetric Einstein–Maxwell–scalar field system Reference 7. This is the main motivation for our present work in which we initiate the study of weak null singularities of similar strength in vacuum without any symmetry assumptions.

Finally, we note that a class of analytic spacetimes with slightly weaker singularities have been previously constructed in Reference 22. While this class of spacetime is more restrictive, as discussed in Reference 22, it nonetheless admits the full “functional degrees of freedom” of the Einstein equations.

1.2. Comparison with impulsive gravitational waves

As pointed out by Dafermos Reference 9, the weak null singularities that we consider in this paper share many similarities with impulsive gravitational waves. The latter are vacuum spacetimes admitting null hypersurfaces which support delta function singularities in the Riemann curvature tensor. Explicit examples were first constructed by Penrose Reference 24, Khan and Penrose Reference 14, and Szekeres Reference 28. In these spacetimes, while the Christoffel symbols are not continuous, they remain bounded. Therefore, in contrast with the weak null singularities that we consider here, these impulsive gravitational waves are not terminal singularities. In fact, the solution to the vacuum Einstein equation extends beyond the singularity and is smooth except across the singular hypersurface. Nevertheless, both scenarios represent singularities propagating along null hypersurfaces and from a mathematical point of view, the proofs of the existence theory for these singularities share many common features.

In recent joint works with Rodnianski Reference 18Reference 19, we initiated the rigorous mathematical study for general impulsive gravitational waves without symmetry assumptions. We constructed the impulsive gravitational waves via solving the characteristic initial problem such that the initial data admit curvature delta singularities supported on an embedded 2-sphere. One of the new ideas in the proof is the use of renormalized energy estimates for the curvature components; i.e., instead of controlling the spacetime curvature components in , we subtract off an correction from some curvature components. This allowed us to derive a closed system of estimates which is completely independent of the singular curvature components.

In Reference 18, when the interaction of impulsive gravitational waves was studied, we also extended the analysis to include a class of spacetimes such that when measured in the worst direction, the Christoffel symbols are merely in . We proved an existence and uniqueness theorem for spacetimes with such low regularity and showed that the spacetime solution can be extended beyond the singularities. Notice that this result is in fact sharp: this is because if the Christoffel symbols fail to be square integrable, the spacetime cannot be extended as a weak solution to the Einstein equations (see footnote 1).

By contrast, the spacetimes considered in this paper have Christoffel symbols which are⁠Footnote10 not in . Even though the weak null singularities are terminal singularities in the sense that there cannot be an existence theory beyond them, the theory developed in Reference 18Reference 19 can be extended to control the spacetime up to the singularity. Moreover, our main theorem, which allows for two weak null singularities terminating at their intersection, can be viewed as an extension of the result in Reference 18 on the interaction of two impulsive gravitational waves. In particular the renormalized energy of Reference 18Reference 19 plays an important role in the proof of our main theorem. However, even after renormalization, the renormalized curvature is still singular (i.e., not in ) and has to be dealt with using an additional weighted estimate.

10

In fact, we allow initial data to be in only for , but not for any .

1.3. Description of the main results

Our setup is the characteristic initial value problem with initial data given on two null hypersurfaces and intersecting at a 2-sphere (see Figure 6). We will follow the general notations in Reference 5Reference 15Reference 16.

We introduce a null frame adapted to a double null foliation (see section 2.1). Denote the constant hypersurfaces by , the constant hypersurfaces by and their intersections by . Decompose the Riemann curvature tensor with respect to the null frame :

We also define the Gauss curvature of the 2-spheres associated to the double null foliation to be . Define also the following Ricci coefficients with respect to the null frame:

Let (resp. ) be the traceless part of (resp. ).

The data on are given on such that becomes singular as . Similarly, the data on is given on such that becomes singular as .

More precisely, let be a smooth function such that is decreasing and

(resp. let be a smooth function such that is decreasing and

For example, can be taken to be for .

Our main theorem shows local existence for a class of singular initial data with⁠Footnote11

11

We assume also bounds for the angular derivatives that are consistent with this singular profile (see the precise statement in Theorem 2).

We construct a (unique) solution to the vacuum Einstein equations in the region , , where , , and

Here, is a double null foliation for and the metric takes the form

in the coordinate system (to be defined in section 2.2). Define also to be the induced Levi-Cevita connection on the -spheres of constant and , i.e., , and , to be the projections of the covariant derivatves , to the tangent space of . Our main theorem (Theorem 1) can be stated precisely as a combination of Theorems 2, 3 and 4. The first main result is the following theorem, which shows the existence of a spacetime up to the (potentially singular) null boundaries:

Theorem 2.

Consider the characteristic initial value problem for

with data that are smooth on and such that the following hold.

There exists an atlas such that in each coordinate chart with local coordinates , the initial metric on obeys

and

The metric on and satisfies the gauge conditions

and

The Ricci coefficients on the initial hypersurface verify

The Ricci coefficients on the initial hypersurface verify

Then for sufficiently small (depending only on and ) and

there exists a unique spacetime endowed with a double null foliation in and , which is a solution to the vacuum Einstein equations Equation 4 with the given data. Moreover, the spacetime remains smooth in and .

Remark 2.

In the following, we will only prove a priori estimates for spacetimes arising from these initial data (see Theorem 5). The existence of a spacetime and the propagation of regularity follow from standard arguments. (For an example of this argument in low regularity, see Reference 19, Sections 4 and 5. See also Reference 5, Chapter 16.)

Remark 3.

In order to simplify notations, we will omit the subscripts and in the weight functions and . They can be inferred from whether is a function of or .

Remark 4.

In section 4, we will construct a class of characteristic initial data which satisfies the assumptions of Theorem 2.

While the weight in the spacetime norms allows the spacetime to be singular, the spacetime metric can be extended beyond the singular hypersurfaces and continuously.

Theorem 3.

Under the assumptions of Theorem 2, the spacetime can be extended continuously up to and beyond the singular boundaries , . Moreover, the induced metric and null second fundamental form on the interior of the limiting hypersurfaces and are regular. More precisely, for any coordinate chart on , the metric components , , satisfy the following estimates in the coordinate chart given by , where and are the diffeomorphisms generated by and respectively:⁠Footnote12

12

See definition of and in section 2.1.

Moreover, for any fixed , we have the following bounds for the Ricci coefficients :

Similar regularity statements hold on .

Remark 5.

If we assume in addition that the higher angular derivatives of are bounded in , then the metric and the second fundamental form also inherit higher regularity in the interior of . In particular if all angular derivatives of are bounded in , then the metric restricted to is smooth along the directions tangential to . Similar statements hold on . We will omit the details.

Moreover, we show that if initially the data are indeed singular, then and are terminal singularities of the spacetime in the following sense:

Theorem 4.

If, in addition to the assumptions of Theorem 2, we also have the following for the initial data,

along Lebesgue-almost every null generator on , then the Christoffel symbols in the coordinate system do not belong to in a neighborhood of any point on .

Similarly if the initial data satisfy

along Lebesgue-almost every null generator on , then the Christoffel symbols in the coordinate system do not belong to in a neighborhood of any point on .

Remark 6.

Theorem 4 guarantees that if we extend the spacetime metric continuously in the obvious differentiable structure given by the coordinate system , then the Christoffel symbols are nonsquare-integrable in the extension. However, it is an open problem whether the spacetime admits any continuous extensions with square integrable Christoffel symbols.

1.4. Main ideas of the proof

All the known proofs of regularity for the Einstein equations without symmetry assumptions rely on estimates on the metric and its derivatives or the Riemann curvature tensor and its derivatives. Let us denote schematically by a general Ricci coefficient and by a general curvature component decomposed with respect to a null frame adapted to the double null foliation. In the double null foliation gauge (see, for example, Reference 5Reference 15), the standard approach to obtain a priori bounds is to couple the estimates for the curvature components

with the estimates for the Ricci coefficients obtained using the transport equations

However, in the setting of two weak null singularities, none of the spacetime curvature components are in !

Nevertheless, while these curvature components are singular, the nature of their singularity is specific. More precisely, while the spacetime curvature components and are not in , they can be written as a sum of some regular intrinsic curvature components and (see further discussion in section 1.4.1) which belong to and terms which are quadratic in . We therefore prove estimates for and , which we will call the renormalized curvature components (see Reference 18Reference 19). Moreover, by considering instead of , we remove all appearances of and in the estimates and so that we do not have to deal with the singularities of and ! It still remains to control the singular curvature components and . Here, we make use of the fact that and are singular in a specific manner toward the singular boundary and respectively. We therefore introduce degenerate norms that incorporate these singularities. We will explain the renormalization and the degenerate estimates in more detail below.

1.4.1. Renormalized energy estimates

As described above, a main ingredient of the proof of the main theorem is the renormalized energy estimates introduced in Reference 18Reference 19 in the study of impulsive gravitational waves. This can be seen as follows. For the class of weak null singularities that we consider, while the derivative of the spacetime metric blows up, the metric restricted to the -sphere remains regular in the angular directions. Since the Gauss curvature is intrinsic to the -spheres, it remains bounded. On the other hand, by the Gauss equation,

and the fact that and blow up at , also blows up at . In view of this, we estimate the Gauss curvature instead of the spacetime curvature component .

Indeed, we see that the Gauss curvature satisfies equations such that the right-hand side contains terms that are less singular than the terms in the corresponding equation for . More precisely, for the curvature component , we have (up to lower-order terms) the Bianchi equation

which contains the nonintegrable curvature component . On the other hand, the Gauss curvature obeys the equation (see Equation 12)

where there are no terms containing or that are quadratic in , and , i.e., every term on the right-hand side of the equation is integrable in the direction.⁠Footnote13

13

The can be compared with the renormalization introduced in Reference 19 and Reference 18, where we estimated instead of . Whereas the renormalization using allows one to eliminate in the estimates, it nonetheless introduces a term , which is not integrable in the direction in the setting of the present paper. Instead, by studying the equation for , we see none of these terms which are quadratic in , or ! This fact can also be derived directly by considering the equations for using the intrinsic definition of the Gauss curvature.

In a similar fashion, by considering the renormalized curvature component⁠Footnote14

14

This is in fact related to the intrinsic curvature of the normal bundle to .

instead of , we see that it satisfies an equation such that all the terms on the right-hand side are integrable in the direction.

One consequence of the renormalization is that we have completely removed the appearances of the curvature component in the equations. In fact, as in Reference 18Reference 19, this allows us to derive a set of estimates for the renormalized curvature component without requiring any information on the curvature component .

Moreover, when considering the equations for and for the renormalized curvature components, one sees that does not appear and all the terms are integrable in the direction. Therefore, although or can be very singular near one of the singular boundaries, we do not need to derive any estimates for them!

1.4.2. Degenerate estimates

Since the renormalization above deals with the singularity in the and components and avoids any information on and , it remains to derive appropriate estimates for and .

The main observation is that while and are both singular and fail to be in , their singularities can be captured quantitatively. Consider the curvature component . Since the blow-up rate of and can be bounded above by , in view of the Codazzi equations in Equation 10, is also bounded above by . In particular while is only in but not in for any , the assumptions on the initial data allow us to control in . We will thus incorporate this blowup in the norms and will be able to still use an based estimate.

The energy estimates will be obtained directly from two sets of Bianchi equations instead of using the Bel–Robinson tensor. Notice that since the energy estimates for are obtained either together with that for or that for , even though and are regular, their energy estimates degenerate. Therefore, at the highest level of derivatives, we have to be content with the weaker estimates for these curvature components.

A potentially more serious challenge is that the introduction of the degenerate weights in and would create terms that cannot be estimated by the energy estimates themselves. Nevertheless, since the weights are chosen to be decreasing toward the future, these uncontrollable terms in fact possess a good sign.

1.4.3. Estimates for the Ricci coefficients

As indicated above, the Ricci coefficients enter as error terms in the energy estimates. Thus, to close all the estimates, we need to control the Ricci coefficients by using the transport equations which in turn have the curvature components in the source terms. Since the various Ricci coefficients have different singular behavior, we separate them according to the bounds that they obey. More precisely, denote by the components that behave like as , by the components that behave like as , and by the components that are bounded.

For the singular Ricci coefficients , we have the following schematic transport equations:

The first three terms on the right-hand side of this equation are bounded while the last term is singular. Nevertheless, the singularity of still allows it to be controlled in along the direction. Thus, this equation can be integrated to show that the initial (singular) bounds for can be propagated. It is important that the terms of the form and do not appear in the equations. A similar structure can also be seen in the equation for the other singular Ricci coefficients , which takes the form

For the regular Ricci coefficients , we have transport equations of the form

The bounds that we prove show that the right-hand side is integrable, and therefore remains bounded. For example, in the equation, it is important that we do not have terms of the form , , , and , which are not uniformly bounded after integrating along the direction.

1.4.4. Null structure in the energy estimates

A priori, the degenerate estimates that we introduce may not be sufficient to control the error terms. Nevertheless, the vacuum Einstein equations possess a remarkable null structure which allows one to close the estimates using only the degenerate estimates.

For example, in the energy estimates for the singular component , we have

To estimate the first term, it suffices to note that , while singular, can be shown to be small after integrating along the direction. Thus, the first term can be controlled using Gronwall’s inequality. For the second term, since the singularity for has the same strength as that for (and similarly the singularity for has the same strength as that for ), the singularity in this term is similar to that in the first term and can also be bounded. The final term is less singular since and are both uniformly bounded.⁠Footnote15 Notice that if other combinations of curvature terms and Ricci coefficients such as , , or appear in the error terms, the degenerate energy will not be strong enough to close the bounds!

15

Although, as pointed out before, the highest derivative estimates for in the energy norm suffer a loss as one approaches the singular boundaries, this term can nevertheless be controlled.

In order to close all the estimates, we need to commute also with higher derivatives. As in Reference 18Reference 19, we will only commute with angular covariant derivatives. These commutations will not introduce terms that are more singular. Moreover, the null structure of the estimates indicated above is also preserved under these commutations.

Similar to Reference 18Reference 19, the renormalization introduces error terms in the energy estimates such that the Ricci coefficients have one more derivative compared to the curvature components. These terms cannot be estimated via transport equations alone but are controlled using also elliptic estimates on the spheres. A form of null structure similar to that described above also makes an appearance in these elliptic estimates, allowing all the bounds to be closed.

1.5. Outline of the paper

We end the introduction with an outline of the remainder of the paper. In section 2, we introduce the basic setup of the paper, including the double null foliation, the coordinate system, and the Einstein vacuum equations recast in terms of the geometric quantities associated to the double null foliation. In section 3, we introduce the norms used in the paper and state a theorem on a priori estimates (Theorem 5) which imply our main existence theorem (Theorem 2). In section 4, we construct a class of characteristic initial data satisfying the assumptions of Theorem 2. In sections 58, we prove Theorem 5. In section 5, we obtain the estimates for the metric components and derive functional inequalities useful in our setting. Then in sections 6 and 7, we prove bounds for the Ricci coefficients assuming control of the curvature components. In section 8, we close all the estimates by obtaining bounds for the curvature components. Finally, in section 9, we discuss the nature of the singular boundary and prove Theorems 3 and 4.

2. Basic setup

2.1. Double null foliation

For a smooth⁠Footnote16 spacetime in a neighborhood of , we define a double null foliation as follows: Let and be solutions to the eikonal equation

16

The spacetimes considered in this paper are not smooth at or . However, since we first construct the spacetime in the region in which the spacetime is smooth (see Theorem 2), it suffices to define the double null foliation for smooth spacetimes.

such that on and on . Let

These are null and geodesic vector fields. Let

Define

to be the normalized null pair such that

and

to be the so-called equivariant vector fields.

In this paper, we will consider spacetime solutions to the vacuum Einstein equations Equation 1 in the gauge such that

The level sets of (resp. ) are denoted by (resp. ). The eikonal equations imply that and are null hypersurfaces. The intersections of the hypersurfaces and are topologically 2-spheres, which we denote by . Note that the integral flows of and respect the foliation .

2.2. The coordinate system

We define a coordinate system in a neighborhood of as follows. On the sphere , we have an atlas such that in the local coordinate system in each coordinate chart, the metric is smooth, bounded, and positive definite. Recall that in a neighborhood of , and are solutions to the eikonal equations,

We then require the coordinates to satisfy

on the initial hypersurface and

in the spacetime region. Here, and denote the restriction of the Lie derivative to (See Reference 5, Chapter 1.) and and are defined as in section 2.1. Relative to the coordinate system , the null pair and can be expressed as

for some such that on , while the metric takes the form

2.3. Equations

We will recast the Einstein equations as a system for Ricci coefficients and curvature components associated to a null frame , defined above and an orthonormal frame⁠Footnote17 tangent to the 2-spheres . We define the Ricci coefficients relative to the null fame,

17

Of course the orthonormal frame is only defined locally. Alternatively, the capital Latin indices can be understood as abstract indices.

where . We also introduce the null curvature components,

Here denotes the Hodge dual of . We denote by the induced covariant derivative operator on and by , the projections to of the covariant derivatives , (see precise definitions in Reference 15, Chapter 3.1).

Observe that

Define the following contractions of the tensor product and with respect to the metric :

where is the volume form associated to the metric . We also define by for -forms and symmetric -tensors, respectively, as follows (note that on -forms this is the Hodge dual on ):

Define the operator on a -form by

For totally symmetric tensors, define the and operators as follows

Define also the trace of totally symmetric tensors to be

We separate the trace and traceless part of and . Let and be the traceless parts of and , respectively. Then and satisfy the following null structure equations:

The other Ricci coefficients satisfy the following null structure equations:

The Ricci coefficients also satisfy the following constraint equations:

with the Gauss curvature of the spheres . The null curvature components satisfy the following null Bianchi equations:

where denotes the Hodge dual on .

We now rewrite the Bianchi equations in terms of the Gauss curvature of the spheres and the renormalized curvature component defined by

The Bianchi equations take the following form:

Notice that we have obtained a system for the renormalized curvature components in which the curvature components and do not appear.⁠Footnote18

18

Moreover, compared to the renormalization in Reference 19, this system does not contain the terms and , which would be uncontrollable in the context of this paper.

From now on, we will use capital Latin letters for indices on the spheres and Greek letters for indices in the whole spacetime.

2.4. Schematic notation

We define a schematic notation for the Ricci coefficients according to the estimates that they obey. Introduce the following conventions:⁠Footnote19

19

Notice that this definition is different form that in Reference 19, since in the context of the present paper and verify different bounds compared to Reference 19.

We will use this schematic notation only in the situations where the exact constant in front of the term is irrelevant to the argument. We will denote by (or , etc.) an arbitrary contraction with respect to the metric and by an arbitrary angular covariant derivative. will be used to denote the sum of all terms which are products of factors, such that each factor takes the form and that the sum of all ’s is , i.e.,

We will use brackets to denote terms with one of the components in the brackets. For instance, the notation denotes the sum of all terms of the form or .

In this schematic notation, the Ricci coefficients satisfy

The Ricci coefficients similarly obey

The Ricci coefficients obey either one of the following equations:

or

We also rewrite the Bianchi equations in the following schematic notation:

3. Norms

In this section we define the norms that we will use to control the geometric quantities. We will in particular use the schematic notation defined in section 2.4. Our norms will be of the form , where and are defined with respect to the measures and , respectively, and is defined for any tensors on by

where the integral is with respect to the volume form induced by .

We define the following norms for the Ricci coefficients for , :

Define the following norms for the Ricci coefficients for , :

Similarly, we define the following norms for the Ricci coefficients for , :

As a shorthand, we define the following norm combining all of the norms above:

We make two remarks concerning these norms.

Remark 7.

While the norms for and are based on in and , respectively, by virtue of the weights and , they actually control the norms. More precisely, since and , by the Cauchy–Schwarz inequality we have

and

Remark 8.

The norm (resp. ) allows us to first take along the direction (resp. direction) before the norm in (resp. ) is taken. This is stronger than the norms such that the order is reversed; i.e., we have

and

In addition to the above norms, we need to define norms for the highest derivatives for the Ricci coefficients. Let

Remark 9.

Here, note that for the norms for , , , , , and , in (or ) is taken after in (or ). According to Remark 8, this is weaker than the norms defined above.

Remark 10.

Notice that the norms for the fourth derivatives of and come with a weight or . This is in contrast to the lower-order derivatives for and , which can be estimated in without any degeneration. The degeneration here arises from the fact that these higher-order derivatives are recovered from the energy estimates for . These energy estimates for , which are derived simultaneously with the estimates for the singular components or , have a degeneration either in or .

We also define the curvature norms for the curvature components. For , let

As a shorthand, we also let

Finally, let and denote the corresponding norms for the initial data, i.e.,

and

In order to prove Theorem 2, we will establish a priori estimates for the geometric quantities in the above norms:

Theorem 5.

Assume that the initial data for the characteristic initial value problem satisfy the assumptions of Theorem 2 with sufficiently small. Then there exists depending only on and such that

In the remainder of the paper, we will focus on the proof of Theorem 5 (after constructing initial data sets in the next section). Standard methods show that Theorem 5 implies Theorem 2. We will omit the details and refer the readers to Reference 5Reference 19 for a proof that the a priori estimates imply the existence theorem.

Remark 11.

The assumptions of Theorem 2 imply the boundedness of the following weighted norms of the curvature components:

and

for some depending only on and . These estimates for , , and follow immediately from the constraint equations on the 2-spheres (see Equation 10). The bound for follows after integrating the null Bianchi equations for on each of the initial null hypersurfaces (see Equation 12).⁠Footnote20 In particular the assumptions of Theorem 2 imply that

20

Notice that it is precisely for the initial bound for that we require an extra derivative for on (and on ) in the assumptions of the theorem. This is related to the intrinsic loss of derivatives for the characteristic initial value problem for second-order hyperbolic systems (see Reference 21).

4. Construction of initial data set

In this section we construct initial data sets satisfying the assumptions of Theorems 2 and 4. In particular we show that the constraint equations can be solved for and . Our approach in this section follows closely that of Christodoulou in Reference 5, Chapter 2.

Assume for simplicity that is a standard sphere of radius . Introduce⁠Footnote21 the standard stereographic coordinates such that the standard metric on the sphere takes the form

21

While we only write down one coordinate chart, it is implicit that we have two stereographic charts—the north pole chart and the south pole chart. In the following, when we derive the estimates for the geometric quantities, we only prove the bounds in a sufficiently large ball in each of these charts.

Clearly, it suffices to construct initial data on (with for ). The construction on is similar. On , we set and therefore . We will construct a metric on in the coordinates taking the form

and and . In order to ensure that satisfies , we write

with , where denotes the set of all matrices taking the form

We will impose upper and lower bounds on . Since there are no smooth globally non-vanishing on the 2-sphere, we use the convention that denotes that the quantity is bounded above by a uniform constant, while denotes that the quantity is bounded above by a uniform constant, and is bounded below at every by a constant depending on (where the constant is moreover allowed to vanish at finitely many isolated points). We require to satisfy⁠Footnote22

22

Here and in the rest of this section, we use the notation that is a multi-index and . We moreover denote .

for some sufficiently large integer . Following Reference 5, we have

We can also derive that

Thus by Equation 20, we have

In particular this implies the requirement in Theorem 4 is satisfied if . By the equation

can be solved from the ODE

We prescribe on to obey the initial conditions

Finally, we prescribe on such that

We check that these initial data obey all the estimates required by Theorem 2:

Estimates for and the metric

To satisfy the upper bounds in Theorem 2, we need to show that

We will show the estimates separately for and . By Equation 22, Equation 26 holds for when . To derive this bound for , notice that by the ODE Equation 23 for , the initial conditions Equation 24, and the bound Equation 22 for , we have

and

for sufficiently small. In the above estimate, we have used . By Equation 21, we thus have

We now move on to control the angular derivatives of . By Equation 20,

Using this bound and commuting the ODE Equation 23 with , we also have that for up to coordinate angular derivatives ,

This implies via Equation 19 and Equation 20 that the metric obeys the bounds

Together with Equation 20 and Equation 21, Equation 28 implies

By Equation 21, we also have

Finally, we notice that by Equation 29, the angular covariant derivatives of and can be controlled by the angular coordinate derivatives of and . Therefore, Equation 26 follows from Equation 30 and Equation 31.

Estimates for

To control , we simply notice that by Equation 29, we have

Estimates for

On , since , . Thus combining the transport equation for in Equation 9 and the Codazzi equation for in Equation 10, and rewriting in (instead of ), we have

Recall from Equation 25 that the initial data for and its angular derivatives are bounded. Therefore, by the estimates for and (and their angular derivatives) above, we have

The bounds for the metric and Christoffel symbols on the sphere imply

as desired.

Estimates for

Similarly to , obeys a transport equations along the null generators of . More precisely, Equation 9 and the Gauss equation in Equation 10 imply that

Thus, the previous estimates imply

Now, combining all the estimates that we have obtained so far, requiring to satisfy

and taking to sufficiently large, we have thus constructed initial data set on that obeys the assumptions of Theorems 2 and 4 on . As mentioned above, it is easy to construct initial data set analogously on so that the full set of assumptions of Theorems 2 and 4 are satisfied.

5. The preliminary estimates

We now turn to the proof Theorem 5, which will form the content of sections 58. In this section we derive the necessary preliminary estimates. In section 6 (see Proposition 15), we will prove the bound

in section 7 (see Proposition 25), we will prove

and in section 8 (see Proposition 32), we will derive the estimate

Combining these estimates then implies the conclusion of Theorem 5.

We now begin with the preliminary estimates. All estimates in this section will be proved under the bootstrap assumption

where is a constant that will be chosen later.

5.1. Estimates for metric components

We first show that we can control under the bootstrap assumption (Equation A1):

Proposition 1.

There exists such that for every ,

Moreover, is continuous up to and .

Proof.

Consider the equation

Fix . Notice that both and are scalars and therefore the norm is independent of the metric. We can integrate equation (Equation 32) using the fact that on to obtain

This implies both the upper and lower bounds for for sufficiently small . To show continuity, take a sequence of points such , , , and . Then

Since by the bootstrap assumption Equation A1, is uniformly bounded, is uniformly integrable in for all and is uniformly integrable in for all , the right-hand side can be made arbitrarily small by taking for sufficiently large. The conclusion thus follows.

We then show that we can control under the bootstrap assumption (Equation A1):

Proposition 2.

There exists such that for , in the coordinate system, we have

where the constants depend only on and . Moreover, remains continuous up to and .

Proof.

We first prove the bound for on the initial hypersurface . Using

we get⁠Footnote23

23

Note that on .

on . We therefore have

This implies that the is bounded above and below. Let be the larger eigenvalue of . Clearly,

Then

Using the first upper bound in (Equation 34), we thus obtain the upper bound for after choosing to be sufficiently small. The upper bound for follows from the upper bound for and the lower bound for .

Now, in order to obtain the bounds for in the spacetime, we argue similarly but use the propagation equation in the direction and compare with . Here, we use bootstrap assumption Equation A1 instead of the assumptions on the initial data. More precisely, we have

We then derive as above that

where is the larger eigenvalue for . As before, we thus obtain the upper bounds for and . Finally, the continuity of up to the boundary follows as in the proof of continuity for in Proposition 1.

With the estimates on , it follows that the norms defined with respect to the metric and the norms defined with respect to the coordinate system are equivalent.

Proposition 3.

Given a covariant tensor on , we have

We can also bound under the bootstrap assumption, thus controlling the full spacetime metric:

Proposition 4.

In the coordinate system ,

Moreover, is continuous up to and .

Proof.

satisfies the equation

This can be derived from

Now, integrating Equation 37 and using Proposition 3 gives the bound on . Continuity of up to the boundary follows as in the proof of Proposition 1.

5.2. Estimates for transport equations

In this subsection, we prove general propositions for obtaining bounds from the covariant null transport equations. Such estimates require the integrability of and , which is consistent with our bootstrap assumption (Equation A1). This will be used in the following sections to derive some estimates for the Ricci coefficients and the null curvature components from the null structure equations and the null Bianchi equations, respectively. Below, we state two propositions which provide estimates for general quantities satisfying transport equations either in the or direction.

Proposition 5.

There exists such that for all and for every , we have

for any tensor tangential to the spheres .

Proof.

The following identity⁠Footnote24 holds for any scalar :

24

Here, on the left-hand side is to be understood as the coordinate vector field in the -plane. Similarly for below.

Similarly, we have

Hence, taking , we have

The bootstrap assumption (Equation A1) implies that and are integrable (and in fact it also implies that and are small after choosing to be small depending on ). Thus the proposition can be proved by using Hölder’s inequality and Gronwall’s inequality, together with the bound for given in Proposition 1.

We also have the following bounds for the case by integrating along the integral curves of and :

Proposition 6.

There exists such that for all , we have

for any tensor tangential to the spheres .

Proof.

This follows simply from integrating along the integral curves of and , and the estimate on in Proposition 1.

5.3. Sobolev embedding

Using the estimates for the metric in Proposition 2, we have the following Sobolev embedding theorem:

Proposition 7.

There exists such that as long as , we have

and

for any tensor tangential to the spheres . Combining the above estimates, we also have

Proof.

By Equation 35 in the proof of Proposition 2, can be made arbitrarily small by choosing to be small. Therefore, the isoperimetric constant

on every sphere is controlled⁠Footnote25 up to a constant factor by the corresponding isoperimetric constant on . Once the isoperimetric constants are uniformly controlled, the Sobolev embedding theorem follows from Reference 5, Lemmas 5.1 and 5.2 and the fact that the volume of is bounded uniformly above and below.

25

This argument is standard. We refer the readers for instance to Reference 5, Lemma 5.4.

5.4. Commutation formulae

We have the following formula from Reference 6, Lemma 7.3.3:

Proposition 8.

The commutator acting on a rank tensor tangential to the spheres is given by

Similarly, the commutator is given by

Recall the schematic notation

By induction and the schematic Codazzi equations

we get the following schematic formula for repeated commutations (see Reference 19):

Proposition 9.

Suppose for some tensors and . Let be the tensor defined by . Then

Similarly, suppose for some tensors and . Let be the tensor defined by . Then

5.5. General elliptic estimates for Hodge systems

We recall the definition of the divergence and curl of a symmetric covariant tensor of rank :

where is the volume form associated to the metric . Recall also that the trace is defined to be

The following elliptic estimate is standard (see, for example, Reference 6, Lemmas 2.2.2, 2.2.3 or Reference 5, Lemmas 7.1, 7.2, 7.3):

Proposition 10.

Let be a symmetric covariant tensor on a 2-sphere satisfying

Suppose also that

Then for , there exists a constant depending only on such that

For the special case that is a symmetric traceless 2-tensor, we only need to know its divergence:

Proposition 11.

Suppose is a symmetric traceless 2-tensor satisfying

Suppose moreover that

Then for , there exists a constant depending only on such that

Proof.

This follows from Proposition 10 and the fact that

6. Estimates for the Ricci coefficients via transport equations

In this section we prove estimates for the Ricci coefficients and their first, second, and third derivatives. We will assume bounds for and and show that for chosen to be sufficiently small, is likewise bounded. In order to achieve this, we continue to work under the bootstrap assumption (Equation A1) and will show that the constant in Equation A1 can in fact be improved (see Proposition 15).

Recall that we will use the following notation: , and .

We first show bounds for .

Proposition 12.

Assume

Then there exists such that whenever ,

i.e., the bounds depends only on the initial data norm . In particular is independent of .

Proof.

We first estimate ; the estimates for are similar after we replace with and with . Using the null structure equations, we have a schematic equation of the type

We also commute the null structure equations with angular derivatives to get

By Proposition 5 in order to estimate , it suffices to estimate the initial data and the norm of the right-hand side Equation 39. Using the bootstrap assumption, we will show that the right-hand side is bounded in a weighted norm. This in turn implies via an application of the Cauchy–Schwarz inequality that the norm is also bounded. We now turn to the details.

We first estimate the curvature term

For the terms such that at most derivative falling on , the bootstrap assumption Equation A1 allows us to control by . We then need to control in . By the Cauchy–Schwarz inequality, since the norm of is smaller than , we can bound this by in the weighted norms. More precisely, we have

For the term where exactly two derivatives fall on (notice that this is the highest number of derivatives that can fall on ), we control in by (using Equation A1). Thus we are left with in . By Sobolev embedding (Proposition 7), this can be bounded by , which in turn can be controlled by after applying the Cauchy–Schwarz inequality as in Equation 40. More precisely,

Combining Equation 40 and Equation 41, we have

We then estimate the second term in Equation 39. We separate the terms where more derivatives fall on and those where more derivatives fall on :

Hence, by Proposition 5, we have

after choosing to be sufficiently small. Similarly, we consider the equation for to get

We now move to the terms that we denote by , i.e., , , and . All of them obey a equation. Unlike the previous estimates for , the initial data for the quantities are not in . We will therefore prove only a bound for in the weighted norm .

Proposition 13.

Assume

Then there exists such that whenever ,

In particular as before, this estimate is independent of .

Proof.

According to the definition of the norm, we need to control the weighted norm of . Using the null structure equations, for each , we have an equation of the type

We also use the null structure equations commuted with angular derivatives:

We estimate the curvature term using the curvature norm. Recall that the curvature norm for along the is weighted with . Using the Sobolev embedding theorem in Proposition 7, we have

The term linear in can be estimated analogously but using the norms instead of the norms:

We now move to control the terms that are nonlinear in the Ricci coefficients. First, we estimate the terms without or :

We then control the term with both and :

Therefore, by the bounds Equation 43, Equation 44, Equation 45, and Equation 46, we have that for every fixed ,

We now multiply this inequality by and take the norm in to get

for sufficiently small.

Using instead the equation for , we obtain the following estimates in a completely analogous manner:

Proposition 14.

Assume

Then there exists such that whenever ,

In particular this estimate is independent of .

By the Sobolev embedding theorems given by Proposition 7, we have thus closed our bootstrap assumption (Equation A1) after choosing to be sufficiently large depending on the initial data norm . We have therefore proved the desired estimates for the Ricci coefficients and their first three angular covariant derivatives. We summarize this in the following proposition.

Proposition 15.

Assume

Then there exists such that whenever ,

7. Elliptic estimates for fourth derivatives of the Ricci coefficients

We now estimate the fourth derivative of the Ricci coefficients. We introduce the following bootstrap assumption:

where is a constant to be chosen later.

The estimates for the fourth derivative of the Ricci coefficients cannot be achieved only by the transport equations since there would be a loss in derivatives. We can however use the transport equation—the Hodge system type estimates as in Reference 5Reference 15Reference 16. We will first derive estimates for some chosen combination of by using transport equations. We will then show that the estimates for all the fourth derivatives of the Ricci coefficients can be proved via elliptic estimates.

In order to apply the elliptic estimates in section 5.5, we need to first control the Gauss curvature and its first and second derivatives in .

Proposition 16.

Assume

Then there exists such that whenever ,

Proof.

obeys the following Bianchi equation:

Commuting with angular derivatives, we have, for ,

By Proposition 5, in order to control in , we need to bound the right hand side in . We first control the term containing :

where we have used the estimates for given by Proposition 15. The term containing can be controlled by

The remaining term has been bounded in the previous section. By Equation 42 and Proposition 15,

Therefore, by Proposition 5,

Gronwall’s inequality implies

since by Proposition 15, for sufficiently small.

It is easy to see that since satisfies a similar schematic Bianchi equation as , we also have the following estimates for and its derivative.

Proposition 17.

Assume

Then there exists such that whenever ,

Using Proposition 16, we now control the fourth derivatives of the Ricci coefficients. We first bound using the transport equation.

Proposition 18.

There exists such that whenever ,

Proof.

Consider the following equation:

After commuting with angular derivatives, we have

By Proposition 5, in order to control in , we need to bound the right-hand side in . Using the fact that is decreasing, this can be achieved using Sobolev embedding (Proposition 7) by

By Proposition 5, we have

Multiplying Equation 47 by and taking first the norm in and then the norm in , we have

where we have used

Thus, the conclusion follows by choosing to be sufficiently small depending on .

Once we have the estimates for , we can control using elliptic estimates:

Proposition 19.

Assume

Then there exists such that whenever ,

Proof.

We now use the Codazzi equation

and apply elliptic estimates from Proposition 11 to get

Notice that we can apply elliptic estimates using Proposition 11, since we have the estimates for the Gauss curvature from Proposition 16. Multiply (Equation 48) by and take the norm to get

By Proposition 15 and Sobolev embedding theorem in Proposition 7, we have

Therefore,

The estimates for and follow identically as that for and :

Proposition 20.

Assume

Then there exists such that whenever ,

and

We then prove estimates for . To do so, we first prove estimates for third derivatives of and recover the control for via elliptic estimates.

Proposition 21.

Assume

Then there exists such that whenever ,

and

Proof.

Recall that

Then satisfies the following equation:⁠Footnote26

26

It is important to note that the potentially harmful term is absent in this equation. This required structure is the reason that we perform this renormalization instead of using as in Reference 18Reference 19.

After commuting with angular derivatives, we get

We now control each of the terms on the right-hand side in . The first term, which contains curvature components, can be estimated by

using the bounds obtained in Proposition 15. The second term can be controlled using Sobolev embedding in Proposition 7 by

using the estimates in Proposition 15. Therefore, by Proposition 5, we have

Recall that the norm of is bounded by . Thus, multiplying Equation 49 by and taking the norm in , we get

for sufficiently small. Similarly, multiplying Equation 49 by and taking the norm in , we get

We can obtain bounds for from the control of using elliptic estimates as follows. By the div-curl systems

and the elliptic estimates given by Propositions 10 and 16, we have

Therefore,

Similarly,

A similar proof shows that the conclusion of Proposition 21 holds also for :

Proposition 22.

Assume

Then there exists such that whenever ,

and

We now move to the estimates for :

Proposition 23.

Assume

Then there exists such that whenever ,

Proof.

Let be defined as the solution to

with zero data on and

By the definition of , it is easy to see that using Proposition 5,

In other words, satisfies much better estimates⁠Footnote27 than for . With this in mind, in the proof of this proposition, we will also use to denote (in addition to , and ).

27

We recall that for we only have the degenerate estimate

With this convention, then obeys the schematic equation

After commuting with angular derivatives, we get

Therefore,

Multiplying by and taking the norm in , we get

The first term is an initial data term and it is bounded by a constant depending only on . We estimate each of the nonlinear terms. The second term can be controlled by

The third term can be bounded by

The fourth term can be estimated by

Therefore,

after choosing to be sufficiently small. Finally, we retrieve the estimates for and from the bounds for . To this end, consider the div-curl system

By elliptic estimates given by Propositions 10 and 16, we have

Therefore, using Proposition 12, Equation 50, and the curvature norm,

By switching and as well as and , we also have the following estimates for :

Proposition 24.

Assume

Then there exists such that whenever ,

We have thus controlled the fourth angular derivatives of all Ricci coefficients and have closed the bootstrap assumption (Equation A2) after choosing to be sufficiently large depending on and . We summarize this in the following proposition:

Proposition 25.

Assume

There exists such that whenever ,

8. Estimates for curvature

In this section, we derive and prove the energy estimates. To this end, we introduce the following bootstrap assumptions:

where is a constant to be chosen later.

In order to derive the energy estimates, we need the following integration by parts formula, which can be proved by direct computation:

Proposition 26.

Let be defined as the spacetime region whose coordinates satisfy and . Suppose and are tensors of rank , then

Proposition 27.

Suppose we have a tensor of rank and a tensor of rank . Then

With these we are now ready to derive energy estimates for in and for in . The most important observation is that the two uncontrollable terms have favorable signs. This in turn is due to the choice of which is decreasing toward the future.

Proposition 28.

The following estimates for the curvature hold:

Proof.

Consider the following schematic Bianchi equations:

Commute the first equation with angular derivatives for . We get the equation for ,

Notice that in the above equation, there are terms arising from the commutator , which can be expressed in terms of the Gauss curvature. After substituting also the Codazzi equations for , we get that these terms have the form of the first term in the above expression. The equation for has a similar structure:

Finally, we have the following structure for :

As a shorthand, we denote by the terms of the form

and by the terms of the form

Contracting Equation 53 with , integrating in the region , applying Proposition 27 and using equations Equation 51 and Equation 52 yield the following identity on the derivatives of the curvature:

Using Proposition 26, since , we have

For the terms with and , we similarly apply Proposition 26, but noting that there is an extra contribution coming from :

Similarly,

Combining Equation 54Equation 57, we thus have the identity

The terms

on the left-hand side, which cannot be controlled⁠Footnote28 by the curvature flux (i.e., the integrals of of the curvature components along or ), have a favorable sign! This is because the weight function satisfies . Therefore, we get an inequality for every :

28

In fact, if we do not drop this term, we can control the spacetime integral

where the weight can be singular. For weights such as for or for , this bound is logarithmically stronger than simply taking the bound for and integrating in .

We now add the above inequalities for . One can easily check that the terms

and

have the form of one of the terms in the statement of the proposition. After applying the Codazzi equation

to one of the ’s, we note that the term

is also one of the terms in the statement of the proposition.

To close the energy estimates, we also need to control in and in . It is not difficult to see, by virtue of the structure of the Einstein equations, that Proposition 28 also holds when all the barred and unbarred quantities are exchanged. The proof is exactly analogous to that of Proposition 28.

Proposition 29.

The following estimates for the curvature components hold:

We now show that we can control all the nonlinear error terms in the energy estimates. We show this for and in and in . The other case can be dealt with in a similar fashion (see Proposition 31).

Proposition 30.

There exists sufficiently small such that whenever ,

Proof.

To prove the curvature estimates, we use Proposition 28. By assumptions of Theorem 2 (see also Remark 11), the two terms corresponding to the initial data are bounded by a constant depending only on initial data. Therefore, we need to control the remaining five error terms in Proposition 28. We first look at the term

Using Propositions 15 and 25, together with the bootstrap assumption (Equation A3), we have

The term

similarly as in the previous estimate since by Propositions 16 and 17, satisfies exactly the same estimates as . We then consider the third nonlinear term

Using Propositions 15 and 25 and the bootstrap assumptions (Equation A3), we have

The fourth nonlinear term can be estimated analogously as the third nonlinear term by

As before, this is because by Propositions 16 and 17, satisfies exactly the same estimates as . Thus it remains to control

This term can be bounded as follows:

Therefore, gathering all the above estimates, we have

which implies the conclusion of the proposition after taking to be sufficiently small.

Notice that the schematic equations are symmetric under the change , and . Since the conditions for the initial data are also symmetric, we also have the following analogous energy estimates for on and on :

Proposition 31.

There exists sufficiently small such that whenever ,

Propositions 30 and 31 together imply

Proposition 32.

There exists such that whenever ,

Proof.

Let

where is taken to be the maximum of the upper bounds in Propositions 30 and 31. Hence, the choice of depends only on and . Thus, by Propositions 30 and 31, the bootstrap assumption (Equation A3) can be improved by choosing sufficiently small depending on and .

Combining Propositions 15, 25, and 32, we conclude the proof of Theorem 5. As mentioned previously, standard methods then imply Theorem 2.

9. Nature of the singular boundary

As described by Theorems 3 and 4, we will also prove the regularity and singularity of the boundary and . We first prove the regularity of the boundary asserted in Theorem 3.

Proof of Theorem 3.

The fact that can be extended continuously up to and beyond and simply follows from the continuity of the metric components , and proved in Propositions 14. To obtain the higher regularity for , we recall the equations Equation 32, Equation 36, and Equation 37:

Commuting these equations with and using the bounds⁠Footnote29 for the Ricci coefficients obtained in the proof of Theorem 5, we conclude that

29

Notice that by controlling and its coordinate angular derivatives , we can show also that and are comparable up to lower-order terms, which allows us to apply the estimates for , , and to bound the coordinate angular derivatives of the metric components.

The boundedness of and its angular derivatives

are already proved in Theorem 5. To control and its angular derivatives on the singular boundary , we first note that by the smoothness assumption on the interior of the initial hypersurface , we have that for every fixed ,

for some finite . We now revisit the proof of Proposition 13 to bound up to for . Restricting to , is bounded. Therefore, the estimates in Equation 43, Equation 44, and Equation 45 are bounded uniformly in . Finally, Equation 46 can be replaced by the estimate

Putting these bounds together, we have

which implies

after applying Gronwall’s inequality.

To conclude the proof, it remains to control and for . Since obeys a equation (see Equation 9), by directly controlling the right-hand side of the null structure equation (commuted with angular derivatives) and using the bounds in Theorem 5, we get

To control the term , notice that combining the equation in Equation 9 and the equations in Equation 7, we have

Upon expressing in terms of using the Codazzi equation in Equation 10, commuting the equation with and using the bound Equation 58, we get

Finally, we control the terms . Commuting the null structure equations for in Equation 8 and Equation 9 with , we have

Estimating directly the right-hand side of the null structure equations or the Bianchi equations, we can easily show that

Using also Equation 59, we thus have

Using Gronwall’s inequality, we get

In particular combining the above estimates, we obtain

on , as desired.

Finally, we move to the proof of Theorem 4. First, we prove

Proposition 33.

Suppose, in addition to the assumptions in Theorem 2, initially obeys

along an outgoing null generator of . Let be the image of under the one-parameter family of diffeomorphisms generated by . Then

holds for every .

Similarly suppose, in addition to the assumptions in Theorem 2, initially obeys

along an outgoing null generator of . Let be the image of under the one-parameter family of diffeomorphisms generated by . Then

holds for every .

Proof.

Fix . Suppose

We want to show that under the assumption Equation 60, we have

which will then imply the desired conclusion.

Using Equation 60, define by

such that

Consider the following null structure equation for :

Along the integral curve of emanating from , we thus have

By the estimates derived in the proof of Theorem 5, , , are bounded and , are in . Therefore,

Consider the following null structure equation for :

Contract this equation with to get

which implies

This implies that along the integral curve of , we have

Using again the fact that , , , , , are bounded for , as well as the estimate Equation 61, we have

Notice that is bounded above and below uniformly in . Taking the norm implies that for , we have

by the assumption of the proposition. The blowup for can be proved in a similar manner.

This implies

Proposition 34.

Suppose the assumptions of Theorem 4 hold. Then, in a neighborhood of any point on , is not integrable with respect to the spacetime volume form. Similarly, in a neighborhood of any point on , is not integrable with respect to the spacetime volume form.

Proof.

We begin with near . By definition, the image of the initial incoming null generator under the map defined in Proposition 33 has constant , and values. Also, by Propositions 1 and 2, the spacetime volume element is bounded uniformly above and below. Therefore, for any neighborhood of , we have

by Proposition 33.

To prove the corresponding statement for near , we first change to the coordinate system such that . This coordinate system can be constructed by solving the ordinary differential equations

with initial condition⁠Footnote30

30

We note that since we do not have a global coordinate chart on , the above ODE only makes sense in , where , are coordinate charts on and and are as defined in Proposition 33. Nevertheless, since and are both diffeomorphisms between and , for every point , there exists and such that , where this change of coordinates makes sense.

By Equation 37, as well as the estimates for , and their derivatives, and the following first derivatives of are uniformly bounded:

Therefore,

In the new coordinate system, we apply the same argument as in the case for near and have the estimate

for any neighborhood of any point , as desired.

Finally, this allows us to conclude that the Christoffel symbols do not belong to :

Proposition 35.

Suppose the assumptions of Theorem 4 hold. Then, the Christoffel symbols in the coordinate system are not in in a neighborhood of any point on or .

Proof.

Recall that the metric in the coordinates takes the form

Note that

One computes that

Since and is uniformly bounded and positive definite, is not in in a neighborhood of any point on the singular boundary in the coordinate system.

To show that the incoming hypersurface is singular, first notice that

We then compute

where the regular terms denote metric components and their derivatives that are uniformly bounded by the estimates proved in the previous sections. By the same reasoning as in the case near , is not in in a neighborhood of any point on the singular boundary in the coordinate system.

This concludes the proof of Theorem 4.

10. Acknowledgments

The author thanks Mihalis Dafermos for suggesting the problem and sharing many insights from the works Reference 7Reference 8Reference 9 as well as offering valuable comments on an earlier version of the manuscript. He thanks Igor Rodnianski for very helpful suggestions. He also thanks Spyros Alexakis, Amos Ori, and Yakov Shlapentokh-Rothman for stimulating discussions. Finally, he is grateful for the suggestions given by the anonymous referees.

Most of the work was carried out when the author was at Princeton University and University of Pennsylvania.

Table of Contents

  1. Abstract
  2. 1. Introduction
    1. Conjecture 1.
    2. Theorem 1 (Main theorem, first version).
    3. 1.1. Weak null singularities and strong cosmic censorship conjecture
    4. 1.2. Comparison with impulsive gravitational waves
    5. 1.3. Description of the main results
    6. Theorem 2.
    7. Theorem 3.
    8. Theorem 4.
    9. 1.4. Main ideas of the proof
    10. 1.5. Outline of the paper
  3. 2. Basic setup
    1. 2.1. Double null foliation
    2. 2.2. The coordinate system
    3. 2.3. Equations
    4. 2.4. Schematic notation
  4. 3. Norms
    1. Theorem 5.
  5. 4. Construction of initial data set
    1. Estimates for and the metric
    2. Estimates for
    3. Estimates for
    4. Estimates for
  6. 5. The preliminary estimates
    1. 5.1. Estimates for metric components
    2. Proposition 1.
    3. Proposition 2.
    4. Proposition 3.
    5. Proposition 4.
    6. 5.2. Estimates for transport equations
    7. Proposition 5.
    8. Proposition 6.
    9. 5.3. Sobolev embedding
    10. Proposition 7.
    11. 5.4. Commutation formulae
    12. Proposition 8.
    13. Proposition 9.
    14. 5.5. General elliptic estimates for Hodge systems
    15. Proposition 10.
    16. Proposition 11.
  7. 6. Estimates for the Ricci coefficients via transport equations
    1. Proposition 12.
    2. Proposition 13.
    3. Proposition 14.
    4. Proposition 15.
  8. 7. Elliptic estimates for fourth derivatives of the Ricci coefficients
    1. Proposition 16.
    2. Proposition 17.
    3. Proposition 18.
    4. Proposition 19.
    5. Proposition 20.
    6. Proposition 21.
    7. Proposition 22.
    8. Proposition 23.
    9. Proposition 24.
    10. Proposition 25.
  9. 8. Estimates for curvature
    1. Proposition 26.
    2. Proposition 27.
    3. Proposition 28.
    4. Proposition 29.
    5. Proposition 30.
    6. Proposition 31.
    7. Proposition 32.
  10. 9. Nature of the singular boundary
    1. Proposition 33.
    2. Proposition 34.
    3. Proposition 35.
  11. 10. Acknowledgments

Figures

Figure 1.

The Penrose diagram of Reissner–Nordström spacetimes

Graphic without alt text
Figure 2.

Region of existence in Conjecture 1

Graphic without alt text
Figure 3.

Region of existence in Theorem 1

Graphic without alt text
Figure 4.

Domains of existence

Graphic without alt text
Figure 5.

Perturbations in the black hole interior of Dafermos spacetimes

Graphic without alt text
Figure 6.

The basic setup

Graphic without alt text

Mathematical Fragments

Equation (1)
Conjecture 1.
(1)

Consider the characteristic initial value problem with smooth characteristic initial data on a pair of null hypersurfaces and intersecting on a -sphere. Suppose that is an affine complete null hypersurface on which the data approach that of the event horizon of a Kerr solution (with at a sufficiently fast polynomial rate.⁠Footnote4 Then the development of the initial data possesses a null boundary “emanating from timelike infinity through which the spacetime is extendible with a continuous metric (see shaded region in Figure 2). Moreover, given an appropriate “lower bound” on the , this piece of null boundary is generically a weak null singularity with nonsquare-integrable Christoffel symbols.

4

In particular this applies if an asymptotically flat spacetime has an exterior region which approaches a subextremal Kerr solution at a sufficiently fast polynomial rate. This also holds in the case where the Cauchy hypersurface has only one asymptotically flat end. In that case, numerical work in spherical symmetry Reference 13 suggests that the singular boundary may also contain a nonempty spacelike portion, in addition to the null portion.

(2)

(Ori, see discussion in Reference 9) If the data for on a complete two-ended asymptotically flat Cauchy hypersurface are globally a small perturbation of two-ended Kerr initial data (with ), then the maximal Cauchy development possesses a global bifurcate future null boundary . Moreover, for generic such perturbations of Kerr, is a global bifurcate weak null singularity which intersects every futurely causally incomplete geodesic.

Equation (2)
Theorem 1 (Main theorem, first version).

For a class of singular characteristic initial data without any symmetry assumptions for the vacuum Einstein equations

with the singular profile as above (see precise requirements on the data in section 1.3) and for sufficiently small and , , there exists a unique smooth spacetime endowed with a double null foliation in , , which satisfies the vacuum Einstein equations with the given data. Associated to , there exists a coordinate system such that the metric extends continuously to the boundary but the Christoffel symbols are not in .

Theorem 2.

Consider the characteristic initial value problem for

with data that are smooth on and such that the following hold.

There exists an atlas such that in each coordinate chart with local coordinates , the initial metric on obeys

and

The metric on and satisfies the gauge conditions

and

The Ricci coefficients on the initial hypersurface verify

The Ricci coefficients on the initial hypersurface verify

Then for sufficiently small (depending only on and ) and

there exists a unique spacetime endowed with a double null foliation in and , which is a solution to the vacuum Einstein equations 4 with the given data. Moreover, the spacetime remains smooth in and .

Theorem 3.

Under the assumptions of Theorem 2, the spacetime can be extended continuously up to and beyond the singular boundaries , . Moreover, the induced metric and null second fundamental form on the interior of the limiting hypersurfaces and are regular. More precisely, for any coordinate chart on , the metric components , , satisfy the following estimates in the coordinate chart given by , where and are the diffeomorphisms generated by and respectively:⁠Footnote12

12

See definition of and in section 2.1.

Moreover, for any fixed , we have the following bounds for the Ricci coefficients :

Similar regularity statements hold on .

Theorem 4.

If, in addition to the assumptions of Theorem 2, we also have the following for the initial data,

along Lebesgue-almost every null generator on , then the Christoffel symbols in the coordinate system do not belong to in a neighborhood of any point on .

Similarly if the initial data satisfy

along Lebesgue-almost every null generator on , then the Christoffel symbols in the coordinate system do not belong to in a neighborhood of any point on .

Equation (7)
Equation (8)
Equation (9)
Equation (10)
Equation (12)
Remark 8.

The norm (resp. ) allows us to first take along the direction (resp. direction) before the norm in (resp. ) is taken. This is stronger than the norms such that the order is reversed; i.e., we have

and

Theorem 5.

Assume that the initial data for the characteristic initial value problem satisfy the assumptions of Theorem 2 with sufficiently small. Then there exists depending only on and such that

Remark 11.

The assumptions of Theorem 2 imply the boundedness of the following weighted norms of the curvature components:

and

for some depending only on and . These estimates for , , and follow immediately from the constraint equations on the 2-spheres (see Equation 10). The bound for follows after integrating the null Bianchi equations for on each of the initial null hypersurfaces (see Equation 12).⁠Footnote20 In particular the assumptions of Theorem 2 imply that

20

Notice that it is precisely for the initial bound for that we require an extra derivative for on (and on ) in the assumptions of the theorem. This is related to the intrinsic loss of derivatives for the characteristic initial value problem for second-order hyperbolic systems (see Reference 21).

Equation (19)
Equation (20)
Equation (21)
Equation (22)
Equation (23)
Equation (24)
Equation (25)
Equation (26)
Equation (28)
Equation (29)
Equation (30)
Equation (31)
Equation (A1)
Proposition 1.

There exists such that for every ,

Moreover, is continuous up to and .

Equation (32)
Proposition 2.

There exists such that for , in the coordinate system, we have

where the constants depend only on and . Moreover, remains continuous up to and .

Equation (34)
Equation (35)
Equation (36)
Proposition 3.

Given a covariant tensor on , we have

Proposition 4.

In the coordinate system ,

Moreover, is continuous up to and .

Equation (37)
Proposition 5.

There exists such that for all and for every , we have

for any tensor tangential to the spheres .

Proposition 7.

There exists such that as long as , we have

and

for any tensor tangential to the spheres . Combining the above estimates, we also have

Proposition 10.

Let be a symmetric covariant tensor on a 2-sphere satisfying

Suppose also that

Then for , there exists a constant depending only on such that

Proposition 11.

Suppose is a symmetric traceless 2-tensor satisfying

Suppose moreover that

Then for , there exists a constant depending only on such that

Proposition 12.

Assume

Then there exists such that whenever ,

i.e., the bounds depends only on the initial data norm . In particular is independent of .

Equation (39)
Equation (40)
Equation (41)
Equation (42)
Proposition 13.

Assume

Then there exists such that whenever ,

In particular as before, this estimate is independent of .

Equation (43)
Equation (44)
Equation (45)
Equation (46)
Proposition 15.

Assume

Then there exists such that whenever ,

Equation (A2)
Proposition 16.

Assume

Then there exists such that whenever ,

Proposition 17.

Assume

Then there exists such that whenever ,

Equation (47)
Equation (48)
Proposition 21.

Assume

Then there exists such that whenever ,

and

Equation (49)
Equation (50)
Proposition 25.

Assume

There exists such that whenever ,

Equation (A3)
Proposition 26.

Let be defined as the spacetime region whose coordinates satisfy and . Suppose and are tensors of rank , then

Proposition 27.

Suppose we have a tensor of rank and a tensor of rank . Then

Proposition 28.

The following estimates for the curvature hold:

Equation (51)
Equation (52)
Equation (53)
Equation (54)
Equation (57)
Proposition 30.

There exists sufficiently small such that whenever ,

Proposition 31.

There exists sufficiently small such that whenever ,

Proposition 32.

There exists such that whenever ,

Equation (58)
Equation (59)
Proposition 33.

Suppose, in addition to the assumptions in Theorem 2, initially obeys

along an outgoing null generator of . Let be the image of under the one-parameter family of diffeomorphisms generated by . Then

holds for every .

Similarly suppose, in addition to the assumptions in Theorem 2, initially obeys

along an outgoing null generator of . Let be the image of under the one-parameter family of diffeomorphisms generated by . Then

holds for every .

Equation (60)
Equation (61)

References

Reference [1]
A. Bonanno, S. Droz, W. Israel, and S. M. Morsink, Structure of the charged spherical black hole interior, Proc. Roy. Soc. London Ser. A 450 (1995), no. 1940, 553–567, DOI 10.1098/rspa.1995.0100. MR1356176,
Show rawAMSref \bib{BDIM}{article}{ author={Bonanno, A.}, author={Droz, S.}, author={Israel, W.}, author={Morsink, S. M.}, title={Structure of the charged spherical black hole interior}, journal={Proc. Roy. Soc. London Ser. A}, volume={450}, date={1995}, number={1940}, pages={553--567}, issn={0962-8444}, review={\MR {1356176}}, doi={10.1098/rspa.1995.0100}, }
Reference [2]
P. R. Brady and J. D. Smith, Black hole singularities: a numerical approach, Phys. Rev. Lett. 75 (1995), no. 7, 1256–1259, DOI 10.1103/PhysRevLett.75.1256. MR1343439,
Show rawAMSref \bib{BS}{article}{ author={Brady, Patrick R.}, author={Smith, John D.}, title={Black hole singularities: a numerical approach}, journal={Phys. Rev. Lett.}, volume={75}, date={1995}, number={7}, pages={1256--1259}, issn={0031-9007}, review={\MR {1343439}}, doi={10.1103/PhysRevLett.75.1256}, }
Reference [3]
L. M. Burko, Structure of the black hole’s Cauchy-horizon singularity, Phys. Rev. Lett. 79 (1997), no. 25, 4958–4961, DOI 10.1103/PhysRevLett.79.4958. MR1487881,
Show rawAMSref \bib{Burko}{article}{ author={Burko, Lior M.}, title={Structure of the black hole's Cauchy-horizon singularity}, journal={Phys. Rev. Lett.}, volume={79}, date={1997}, number={25}, pages={4958--4961}, issn={0031-9007}, review={\MR {1487881}}, doi={10.1103/PhysRevLett.79.4958}, }
Reference [4]
S. Chandrasekhar and J. B. Hartle, On crossing the Cauchy horizon of a Reissner–Nordström black-hole, Proc. Roy. Soc. London Ser. A 384 (1982), no. 1787, 301–315, DOI 10.1098/rspa.1982.0160. MR684313,
Show rawAMSref \bib{CH}{article}{ author={Chandrasekhar, S.}, author={Hartle, J. B.}, title={On crossing the Cauchy horizon of a Reissner--Nordstr\"om black-hole}, journal={Proc. Roy. Soc. London Ser. A}, volume={384}, date={1982}, number={1787}, pages={301--315}, issn={0080-4630}, review={\MR {684313}}, doi={10.1098/rspa.1982.0160}, }
Reference [5]
D. Christodoulou, The formation of black holes in general relativity, EMS Monographs in Mathematics, European Mathematical Society (EMS), Zürich, 2009, DOI 10.4171/068. MR2488976,
Show rawAMSref \bib{Chr}{book}{ author={Christodoulou, Demetrios}, title={The formation of black holes in general relativity}, series={EMS Monographs in Mathematics}, publisher={European Mathematical Society (EMS), Z\"urich}, date={2009}, pages={x+589}, isbn={978-3-03719-068-5}, review={\MR {2488976}}, doi={10.4171/068}, }
Reference [6]
D. Christodoulou and S. Klainerman, The global nonlinear stability of the Minkowski space, Princeton Mathematical Series, vol. 41, Princeton University Press, Princeton, NJ, 1993. MR1316662,
Show rawAMSref \bib{CK}{book}{ author={Christodoulou, Demetrios}, author={Klainerman, Sergiu}, title={The global nonlinear stability of the Minkowski space}, series={Princeton Mathematical Series}, volume={41}, publisher={Princeton University Press, Princeton, NJ}, date={1993}, pages={x+514}, isbn={0-691-08777-6}, review={\MR {1316662}}, }
Reference [7]
M. Dafermos, Stability and instability of the Cauchy horizon for the spherically symmetric Einstein-Maxwell-scalar field equations, Ann. of Math. (2) 158 (2003), no. 3, 875–928, DOI 10.4007/annals.2003.158.875. MR2031855,
Show rawAMSref \bib{D1}{article}{ author={Dafermos, Mihalis}, title={Stability and instability of the Cauchy horizon for the spherically symmetric Einstein-Maxwell-scalar field equations}, journal={Ann. of Math. (2)}, volume={158}, date={2003}, number={3}, pages={875--928}, issn={0003-486X}, review={\MR {2031855}}, doi={10.4007/annals.2003.158.875}, }
Reference [8]
M. Dafermos, The interior of charged black holes and the problem of uniqueness in general relativity, Comm. Pure Appl. Math. 58 (2005), no. 4, 445–504, DOI 10.1002/cpa.20071. MR2119866,
Show rawAMSref \bib{D2}{article}{ author={Dafermos, Mihalis}, title={The interior of charged black holes and the problem of uniqueness in general relativity}, journal={Comm. Pure Appl. Math.}, volume={58}, date={2005}, number={4}, pages={445--504}, issn={0010-3640}, review={\MR {2119866}}, doi={10.1002/cpa.20071}, }
Reference [9]
M. Dafermos, Black holes without spacelike singularities, Comm. Math. Phys. 332 (2014), no. 2, 729–757, DOI 10.1007/s00220-014-2063-4. MR3257661,
Show rawAMSref \bib{D3}{article}{ author={Dafermos, Mihalis}, title={Black holes without spacelike singularities}, journal={Comm. Math. Phys.}, volume={332}, date={2014}, number={2}, pages={729--757}, issn={0010-3616}, review={\MR {3257661}}, doi={10.1007/s00220-014-2063-4}, }
Reference [10]
M. Dafermos and I. Rodnianski, A proof of Price’s law for the collapse of a self-gravitating scalar field, Invent. Math. 162 (2005), no. 2, 381–457, DOI 10.1007/s00222-005-0450-3. MR2199010,
Show rawAMSref \bib{DRPrice}{article}{ author={Dafermos, Mihalis}, author={Rodnianski, Igor}, title={A proof of Price's law for the collapse of a self-gravitating scalar field}, journal={Invent. Math.}, volume={162}, date={2005}, number={2}, pages={381--457}, issn={0020-9910}, review={\MR {2199010}}, doi={10.1007/s00222-005-0450-3}, }
Reference [11]
M. Dafermos and I. Rodnianski, The black hole stability problem for linear scalar perturbations, in Proceedings of the Twelfth Marcel Grossmann Meeting on General Relativity (T. Damour et al., Eds.), World Scientific, Singapore, 2011, pp. 132–189.
Reference [12]
W. A. Hiscock, Evolution of the interior of a charged black hole, Phys. Lett. A 83 (1981), no. 3, 110–112, DOI 10.1016/0375-9601(81)90508-9. MR617171,
Show rawAMSref \bib{Hiscock}{article}{ author={Hiscock, William A.}, title={Evolution of the interior of a charged black hole}, journal={Phys. Lett. A}, volume={83}, date={1981}, number={3}, pages={110--112}, issn={0031-9163}, review={\MR {617171}}, doi={10.1016/0375-9601(81)90508-9}, }
Reference [13]
S. Hod and T. Piran, Mass inflation in dynamical gravitational collapse of a charged scalar field, Phys. Rev. Lett. 81 (1998), 1554–1557.
Reference [14]
K. A. Khan and R. Penrose, Scattering of two impulsive gravitational plane waves, Nature 229 (1971), 185–186.
Reference [15]
S. Klainerman and F. Nicolò, The evolution problem in general relativity, Progress in Mathematical Physics, vol. 25, Birkhäuser Boston, Inc., Boston, MA, 2003, DOI 10.1007/978-1-4612-2084-8. MR1946854,
Show rawAMSref \bib{KN}{book}{ author={Klainerman, Sergiu}, author={Nicol\`o, Francesco}, title={The evolution problem in general relativity}, series={Progress in Mathematical Physics}, volume={25}, publisher={Birkh\"auser Boston, Inc., Boston, MA}, date={2003}, pages={xiv+385}, isbn={0-8176-4254-4}, review={\MR {1946854}}, doi={10.1007/978-1-4612-2084-8}, }
Reference [16]
S. Klainerman and I. Rodnianski, On the formation of trapped surfaces, Acta Math. 208 (2012), no. 2, 211–333, DOI 10.1007/s11511-012-0077-3. MR2931382,
Show rawAMSref \bib{KlRo}{article}{ author={Klainerman, Sergiu}, author={Rodnianski, Igor}, title={On the formation of trapped surfaces}, journal={Acta Math.}, volume={208}, date={2012}, number={2}, pages={211--333}, issn={0001-5962}, review={\MR {2931382}}, doi={10.1007/s11511-012-0077-3}, }
Reference [17]
J. Luk, On the local existence for the characteristic initial value problem in general relativity, Int. Math. Res. Not. IMRN 20 (2012), 4625–4678, DOI 10.1093/imrn/rnr201. MR2989616,
Show rawAMSref \bib{L}{article}{ author={Luk, Jonathan}, title={On the local existence for the characteristic initial value problem in general relativity}, journal={Int. Math. Res. Not. IMRN}, date={2012}, number={20}, pages={4625--4678}, issn={1073-7928}, review={\MR {2989616}}, doi={10.1093/imrn/rnr201}, }
Reference [18]
J. Luk and I. Rodnianski, Nonlinear interaction of impulsive gravitational waves for the vacuum Einstein equations, Cambridge J. Math., to appear.
Reference [19]
J. Luk and I. Rodnianski, Local propagation of impulsive gravitational waves, Comm. Pure Appl. Math. 68 (2015), no. 4, 511–624, DOI 10.1002/cpa.21531. MR3318018,
Show rawAMSref \bib{LR}{article}{ author={Luk, Jonathan}, author={Rodnianski, Igor}, title={Local propagation of impulsive gravitational waves}, journal={Comm. Pure Appl. Math.}, volume={68}, date={2015}, number={4}, pages={511--624}, issn={0010-3640}, review={\MR {3318018}}, doi={10.1002/cpa.21531}, }
Reference [20]
J. M. McNamara, Instability of black hole inner horizons, Proc. Roy. Soc. London Ser. A 358 (1978), no. 1695, 499–517, DOI 10.1098/rspa.1978.0024. MR0489678,
Show rawAMSref \bib{McN}{article}{ author={McNamara, J. M.}, title={Instability of black hole inner horizons}, journal={Proc. Roy. Soc. London Ser. A}, volume={358}, date={1978}, number={1695}, pages={499--517}, issn={0962-8444}, review={\MR {0489678}}, doi={10.1098/rspa.1978.0024}, }
Reference [21]
H. Müller zum Hagen, Characteristic initial value problem for hyperbolic systems of second order differential equations (English, with French summary), Ann. Inst. H. Poincaré Phys. Théor. 53 (1990), no. 2, 159–216. MR1079777,
Show rawAMSref \bib{M}{article}{ author={M\"uller zum Hagen, H.}, title={Characteristic initial value problem for hyperbolic systems of second order differential equations}, language={English, with French summary}, journal={Ann. Inst. H. Poincar\'e Phys. Th\'eor.}, volume={53}, date={1990}, number={2}, pages={159--216}, issn={0246-0211}, review={\MR {1079777}}, }
Reference [22]
A. Ori and É. É. Flanagan, How generic are null spacetime singularities?, Phys. Rev. D (3) 53 (1996), no. 4, R1754–R1758, DOI 10.1103/PhysRevD.53.R1754. MR1380002,
Show rawAMSref \bib{FO}{article}{ author={Ori, Amos}, author={Flanagan, \'Eanna \'E.}, title={How generic are null spacetime singularities?}, journal={Phys. Rev. D (3)}, volume={53}, date={1996}, number={4}, pages={R1754--R1758}, issn={0556-2821}, review={\MR {1380002}}, doi={10.1103/PhysRevD.53.R1754}, }
Reference [23]
R. Penrose, Gravitational collapse and space-time singularities, Phys. Rev. Lett. 14 (1965), 57–59, DOI 10.1103/PhysRevLett.14.57. MR0172678,
Show rawAMSref \bib{Penrose.2}{article}{ author={Penrose, Roger}, title={Gravitational collapse and space-time singularities}, journal={Phys. Rev. Lett.}, volume={14}, date={1965}, pages={57--59}, issn={0031-9007}, review={\MR {0172678}}, doi={10.1103/PhysRevLett.14.57}, }
Reference [24]
R. Penrose, The geometry of impulsive gravitational waves, General relativity (papers in honour of J. L. Synge), Clarendon Press, Oxford, 1972, pp. 101–115. MR0503490,
Show rawAMSref \bib{Penrose72}{article}{ author={Penrose, Roger}, title={The geometry of impulsive gravitational waves}, conference={ title={General relativity (papers in honour of J. L. Synge)}, }, book={ publisher={Clarendon Press, Oxford}, }, date={1972}, pages={101--115}, review={\MR {0503490}}, }
Reference [25]
E. Poisson and W. Israel, Inner-horizon instability and mass inflation in black holes, Phys. Rev. Lett. 63 (1989), no. 16, 1663–1666, DOI 10.1103/PhysRevLett.63.1663. MR1018317,
Show rawAMSref \bib{PI1}{article}{ author={Poisson, E.}, author={Israel, W.}, title={Inner-horizon instability and mass inflation in black holes}, journal={Phys. Rev. Lett.}, volume={63}, date={1989}, number={16}, pages={1663--1666}, issn={0031-9007}, review={\MR {1018317}}, doi={10.1103/PhysRevLett.63.1663}, }
Reference [26]
E. Poisson and W. Israel, Internal structure of black holes, Phys. Rev. D (3) 41 (1990), no. 6, 1796–1809, DOI 10.1103/PhysRevD.41.1796. MR1048877,
Show rawAMSref \bib{PI2}{article}{ author={Poisson, Eric}, author={Israel, Werner}, title={Internal structure of black holes}, journal={Phys. Rev. D (3)}, volume={41}, date={1990}, number={6}, pages={1796--1809}, issn={0556-2821}, review={\MR {1048877}}, doi={10.1103/PhysRevD.41.1796}, }
Reference [27]
M. Simpson and R. Penrose, Internal instability in a Reissner-Nordström black hole, Internat. J. Theoret. Phys. 7 (1973), 183–197.
Reference [28]
P. Szekeres, Colliding gravitational waves, Nature 228 (1970), 1183–1184.

Article Information

MSC 2010
Primary: 83C75 (Space-time singularities, cosmic censorship, etc.), 35L67 (Shocks and singularities)
Author Information
Jonathan Luk
Department of Mathematics, Stanford University, Stanford, California 94305-2125
jluk@stanford.edu
MathSciNet
Additional Notes

This work is supported by the NSF Postdoctoral Fellowship DMS-1204493 and the NSF grant DMS-1709458.

Journal Information
Journal of the American Mathematical Society, Volume 31, Issue 1, ISSN 1088-6834, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2017 American Mathematical Society
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/jams/888
  • MathSciNet Review: 3718450
  • Show rawAMSref \bib{3718450}{article}{ author={Luk, Jonathan}, title={Weak null singularities in general relativity}, journal={J. Amer. Math. Soc.}, volume={31}, number={1}, date={2018-01}, pages={1-63}, issn={0894-0347}, review={3718450}, doi={10.1090/jams/888}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.