Equality in the spacetime positive mass theorem II
Abstract.
We provide a new proof of the equality case of the spacetime positive mass theorem, which states that if a complete asymptotically flat initial data set satisfying the dominant energy condition has null ADM energy-momentum (that is, ), then must isometrically embed into Minkowski space with as its second fundamental form. Previous proofs either used spinor methods [Wit81, BC96, CM06], relied on the Jang equation [HL20, Eic13], or assumed three spatial dimensions [HZ23]. In contrast, our new proof only requires knowing that for all complete initial data sets near on satisfying the dominant energy condition.
1. Introduction
Let us first explicitly state the definition of asymptotic flatness we will be using.
Definition 1.1.
Let , let be an -dimensional connected manifold, possibly with boundary, and let . We say that an initial data set is complete asymptotically flat with decay rate if the following hold:
-
•
Each noncompact end of (of which there is at least one) is diffeomorphic to , and for some ,
(1.1) where denotes a background metric that is equal to the standard Euclidean metric on each end, and and denote weighted Hölder spaces.
-
•
The energy density and current density defined by
(1.2) are integrable over , where denotes the vector field dual to the one-form .
Note that asymptotic flatness of implies that the ADM energy-momentum of each end is well-defined. Our main result characterizes the equality case of the spacetime positive mass theorem, but we would like to emphasize that, unlike all previous results characterizing , ours follows from the positive mass inequality itself and does not rely on any particular proof of this inequality. Consequently, we introduce the following definition.
Definition 1.2.
Let be a complete asymptotically flat initial data set with decay rate , possibly with boundary. We say that the positive mass theorem is true near if there is an open ball around in , for some , such that for each complete asymptotically flat initial data set with decay rate in that open ball, satisfying the dominant energy condition and having weakly outer trapped boundary (if there is a boundary), we have , where denotes the ADM energy-momentum of any asymptotically flat end of .
We say the boundary of an asymptotically flat manifold is weakly outer trapped if its outward null expansion satisfies , with respect to the normal that points toward .111Our convention is that the mean curvature is the tangential divergence of the chosen unit normal vector. For example, the mean curvature of a round sphere in Euclidean space with respect to the outward normal is positive.
It follows from E. Witten’s proof of the positive mass theorem [Wit81] (and [GHHP83, Her98] for the boundary case) that the positive mass theorem is true near any on a spin manifold. Building on the classic works of R. Schoen and S.-T. Yau [SY79, SY81], in joint work with M. Eichmair and Schoen we proved that the positive mass theorem is true222The definition of asymptotic flatness used in [EHLS16] is slightly different from the one stated here. By going through the proofs there with more care, one sees that Definition 1.1 is adequate for proving the positive mass theorem. See [LLU22] for details. near any on a manifold of dimension less than 8 [EHLS16], with the boundary case proved by the second author with M. Lesourd and R. Unger [LLU22]. J. Lohkamp has also announced a proof in all dimensions [Loh16].
Understanding the case of the spacetime positive mass theorem breaks down into two separate steps: The first step is to show that , and the second step is to use the fact that to find an embedding of the initial data into Minkowski space. In the first paper in this series [HL20], we built on the work of R. Beig and P. Chruściel [BC96] (see also [CM06]) in the spin case to construct an argument that proves the first step without appealing to any particular proof of the inequality. We state this result below and note that it was generalized to include the possibility of a boundary in [LLU22].
Theorem 1.3.
Let be a complete asymptotically flat initial data set with decay rate , possibly with boundary, and assume that the positive mass theorem is true near . Furthermore, we make the stronger asymptotic assumption that with this decay rate , there exists satisfying (1.1) such that
(1.3) |
and
(1.4) |
for some , where is the dimension. If satisfies the dominant energy condition and the boundary (if there is one) is weakly outer trapped, then for each asymptotically flat end, implies that .
Although the stronger asymptotic assumption (1.3) may seem undesirable in higher dimensions (note that (1.3) can be dropped for if in (1.1) is close to ), we found pp-wave counterexamples showing that the theorem is false without it for . See [HL, Example 7] for details.
Once one knows that , the “second step” described above follows from work of Eichmair [Eic13] if . More specifically, following Schoen and Yau’s proof for the case333When , one must make an extra decay assumption on . [SY81], Eichmair used the Jang equation to prove the inequality , and then by analyzing the case of equality in that proof, he showed that the case can only occur when the Jang solution provides a graphical isometric embedding of into Minkowski space with as its second fundamental form. The main result of this article is to replace this argument for the “second step” with one that naturally extends the methods used to prove Theorem 1.3 and is self-contained in the sense that it does not depend on how one proves that (nor on how one proves ).
Theorem 1.4.
Let be a complete asymptotically flat initial data set with decay rate , possibly with boundary, and assume that the positive mass theorem is true near . Further assume that is locally . If satisfies the dominant energy condition and the boundary (if there is one) is weakly outer trapped, then in each asymptotically flat end, unless has no boundary and isometrically embeds into Minkowski space with as its second fundamental form, in which case .
Combining the two previous theorems gives the following result.
Corollary 1.5.
Let be a complete asymptotically flat initial data set with decay rate , possibly with boundary, and assume that the positive mass theorem is true near . Further assume that (1.3) and (1.4) hold, and that is locally . If satisfies the dominant energy condition and the boundary (if there is one) is weakly outer trapped, then in each asymptotically flat end, unless has no boundary and embeds into Minkowski space with as its second fundamental form, in which case .
This fact was essentially already proved in dimension 3 by Beig and Chruściel [BC96] and for general spin manifolds by Chruściel and D. Maerten [CM06]. These results required a slightly stronger assumption than (1.3) but do not require regularity. More recently, Sven Hirsch and Yiyue Zhang [HZ23] gave a new proof in dimension that avoids assuming any of (1.3), (1.4), or regularity.
The basic outline of the proof of Theorem 1.4 is the following: As in the proof of Theorem 1.3, we use the fact that minimizes a modified Regge-Teitelboim Hamiltonian subject to a constraint, and then we invoke Lagrange multipliers to construct lapse-shift pairs that satisfy a Killing initial data type of equation. The difference is that when , minimizes many different Regge-Teitelboim Hamiltonians, so we can actually construct an entire -dimensional space of these lapse-shift pairs that can be thought of as being asymptotic to the -dimensional space of translational Killing fields on Minkowski space as we approach spatial infinity. The essential new ingredient is that by invoking our recent work [HL], we can also make sure that these lapse-shift pairs satisfy what we call the “-null-vector equation.” Moreover, having so many solutions implies that is vacuum. All of this is explained in Section 2, culminating in the statement of Lemma 2.2. Observe that proving that is vacuum is itself a difficult task. Note that the vacuum condition is a much stronger condition than the “borderline” case of the dominant energy condition. The fact that implies is a previously unstated direct consequence of [HL20, Theorem 5.2] and [HL, Theorem 6], but the aforementioned pp-wave counterexamples from [HL, Example 7] demonstrate that does not imply vacuum.
Once we know that is vacuum, it follows that each of these lapse-shift pairs is actually vacuum Killing initial data for , and we can extend them to become actual Killing fields on the vacuum Killing development of . The next step is to show that on an asymptotically flat Lorentzian manifold, having such a space of Killing fields that are asymptotic to the translational directions implies that the Lorentzian manifold must be Minkowski space. This observation is described in Theorem 1.6 below, which requires no curvature assumptions.
Theorem 1.6.
Let be an -dimensional connected Lorentzian spacetime, possibly with boundary, that is asymptotically flat in the sense that a subset of is diffeomorphic to , and in these coordinates,
for some , where is the standard Minkwoski metric and denotes the norm of the spatial coordinates only. More specifically, the notation means that there exists a positive number such that
where denotes the partial derivatives with respect to all coordinates.
Assume that is equipped with Killing fields that are asymptotic to the coordinate translation directions in the sense that
Then form a covariant constant global Lorentzian orthonormal frame, and in particular, must be flat.
As a direct consequence, the above theorem also implies a version for asymptotically flat Riemannian manifolds, which is included in Corollary 3.2 below.
Note that we do not assume is complete in Theorem 1.6. If is complete without boundary, then it must be Minkowski space by the Killing-Hopf theorem. To rule out the possibility of having a weakly outer trapped boundary, we prove the following general result.
Theorem 1.7.
Let be a compact connected manifold with boundary such that where and are both closed and nonempty. Let , and let be a Lorentzian metric on such that is spacelike. Suppose the following holds:
-
(1)
is a global timelike Killing vector field, where denotes the coordinate for the factor. Thus is strictly stationary.
-
(2)
satisfies the null energy condition.
-
(3)
is a strictly outer untrapped surface (that is, ) with respect to the “outward” future null normal pointing out of .
Then cannot be a weakly outer trapped surface (that is, it cannot have on all of ) with respect to the “outward” future null normal pointing toward the interior of .
When the spatial dimension is less than , Theorem 1.7 follows from an elegant argument of A. Carrasco and M. Mars (used to prove [CM08, Corollary 1]). The dimension restriction is needed because their proof relies on the existence theory of smooth MOTS by Eichmair [Eic09]. Our proof is valid in all dimensions and also applies to more general situations (e.g. the stationary assumption can be significantly relaxed). See Theorem 4.1.
For the special case of static spacetime metrics, which are of the form where , the null energy condition becomes the assumption
(1.5) |
Corollary 1.8.
Let be a compact connected Riemannian manifold with boundary such that where and are both closed and nonempty. Suppose there is a function on such that (1.5) holds.
Suppose has mean curvature with respect to the unit normal pointing out of . Then cannot have everywhere, with respect to the unit normal pointing into .
Specializing even more to the case where reduces to a well-known classical theorem for manifolds with nonnegative Ricci curvature, stated as Theorem 4.2. Note that many examples, such as static vacuum metrics and electro-vacuum metrics, with any cosmological constant, satisfy condition (1.5).
The paper is organized as follows. In Section 2, we explore the Lagrange multiplier method carried out in [HL20] for the case using the new results from [HL] to construct an -dimensional space of asymptotically translational vacuum Killing initial data. In Section 3, we prove Theorem 1.6 and then complete the proof of the main result, Theorem 1.4, for the case of no boundary. Finally, the possibility of having a nonempty weakly outer trapped boundary is ruled out by Theorem 1.7, whose proof is presented in Section 4 and can be read independently from all prior sections.
2. Lagrange multipliers
We follow closely the definitions and notations in [HL20]. In particular, we refer the reader to [HL20, Section 2] for the definitions of the weighted Hölder space and the weighted Sobolev space . Assume , and let denote the set of symmetric -tensors such that and is positive definite at each point. By the inclusion , for any , whenever we work on the weighted Sobolev spaces, we shall assume the decay rate in the weighted Sobolev spaces to be slightly smaller than the assumed asymptotic decay rate appearing in Definition 1.1.
Initial data can be equivalently described by a pair where is a symmetric -tensor called the conjugate momentum tensor, which is related to the -tensor via the equation
(2.1) |
By slight abuse of vocabulary, we will refer to as initial data. We define the constraint map on initial data by
where and are functions of , defined by (1.2) combined with (2.1). Recall that the dominant energy condition (or DEC) is the condition .
As in [HL], given an initial data set and a smooth function on , we introduce the -modified constraint operator, which is defined on initial data on by
(2.2) |
where denotes the current density of , and denotes the contraction of and with respect to . The crucial property of is that for any , if satisfies the DEC, , and
then also satisfies the DEC (see Lemma 3.4 of [HL]). Note that this is a property that the ordinary constraint operator does not have, but since only differs from by an affine function of , it behaves similarly from a PDE perspective.
Throughout this article, a lapse-shift pair on will simply refer to a function and a vector field on . Assuming that is locally , Theorem 4.1 of [HL], which was the main technical result of [HL], tells us444Take in the statement of that theorem. that there exists a dense subset such that if , then any lapse-shift pair solving
(2.3) |
in the interior of must also satisfy the pair of equations
(2.4) |
where the second equation is what we call the -null-vector equation. Choosing any bounded ensures that the terms in (2.2) involving have the appropriate fall-off rates so that
We will use the method of [HL20] to construct solutions to (2.3) and hence also (2.4). The following definition simply generalizes Definition 5.1 of [HL20] from the case to allow for general .
Definition 2.1.
Let be a complete asymptotically flat manifold, possibly with boundary, and choose to be slightly smaller than the asymptotic decay rate of . Fix an end of and constants and . Let be a smooth lapse-shift pair on such that is equal to in the specified end and vanishes on all other ends of (if any) and in a neighborhood of (if nonempty). Let be a smooth bounded function on .
We define the -modified Regge-Teitelboim Hamiltonian corresponding to and to be the function given by
(2.5) |
where and denote the ADM energy-momentum of the specified end, and the volume measure and the inner product in the integral are both with respect to .
As explained in [HL20], although the terms in (2.5) are not individually well-defined for arbitrary , the functional extends to all of in a natural way.
In what follows, we will assume that and satisfy . Continuing to follow [HL20] (with input from [LLU22] for the boundary case), we define
We claim that that if satisfies the DEC with weakly outer trapped (if nonempty) and has , then is a local minimizer of in . To see why, observe that for all nearby , also satisfies the DEC (as mentioned above, due to [HL, Lemma 3.4]) and has weakly outer trapped boundary, so the assumption that the positive mass theorem is true near tells us that . (Technically, since only has Sobolev regularity, we must invoke [HL20, Theorem 4.1] and [LLU22, Lemma 4.2].) Therefore
where we used the assumptions that and has . Hence the claim is true.
Thus locally minimizes the functional subject to the constraint . In the finite dimensional setting, as long as the constraint carves out a smooth submanifold of a linear space, we can use the method of Lagrange multipliers to gain information about the minimizer. In our more general Banach space setting, in order to construct Lagrange multipliers, all we need is for to be the level set of a map on whose linearization is surjective (see [HL20, Theorem D.1] for details). For , this surjectivity was established in [HL20, Lemma 2.10] for the case of no boundary, and in [LLU22, Proposition 3.9] when boundary is present. Meanwhile, the introduction of does not change anything in these arguments.
Invoking Lagrange multipliers exactly as in [HL20] (or [LLU22] for the boundary case) leads to the existence of a lapse-shift pair such that
where denotes an object that decays in . Although we will not use this fact elsewhere in the paper, we remark that the existence of even one such lapse-shift pair implies that can only have one asymptotically flat end.555The same argument also applies to the case and leads to the following fact. If is a complete asymptotically flat initial data set satisfying the dominant energy condition with , and if the positive mass theorem is true near , then has only one end. Note that this is consistent with the pp-waves counterexamples having from [HL, Example 7], which all have only one end. This is because [CÓM81, Theorem 3.3] implies that it is impossible for a nontrivial solution of (2.3) to asymptotically vanish in an end.
As described above, as long as we select so that Theorem 4.1 of [HL] applies, this also satisfies (2.4) on . Since we can construct such a solution for any satisfying , and the space of all satisfying (2.4) forms a vector space , it follows that for any , we obtain a lapse-shift pair which is both asymptotic to in the desired end and also satisfies (2.4). Moreover, the dimension of is at least .
But Corollary 6.6 of [HL] states that if the dimension of is greater than 1, then must be vacuum initial data. Since is vacuum, . The intuition behind Corollary 6.6 of [HL] is the following: lying in the kernel of the overdetermined Hessian-type operator already means that is determined by its -jet at any point. If there is a point where , we can use (2.4) to write in terms of , and hence the -jet of at determines the -jet of at , which then determines the -jet of at by the -null-vector equation, which determines everywhere. For details, see [HL].
Putting together the arguments described in this section, we have proved the following lemma.
Lemma 2.2.
Let satisfy the hypotheses of Theorem 1.4, but suppose that . Then is vacuum, and for any , there exists a lapse-shift pair such that
Definition 2.3.
Let be a vacuum initial data set. Then any lapse-shift pair satisfying is called vacuum Killing initial data (or a vacuum KID) on .
3. Killing fields
In this section, we prove Theorem 1.6, whose proof is completely unrelated to the material in the previous section. After that, we will explain how to combine Lemma 2.2 with Theorem 1.6 to prove our main theorem, Theorem 1.4.
But first, we recall how the Killing condition for vector fields on a spacetime restricts to a spacelike slice. This computation was made in [BC97] (written in different notation).666We provided details of this computation in lapse-shift coordinates in the proof of [HL, Corollary B.3], but this required the assumption that is transverse to . Performing the computation in Gaussian coordinates as in [BC97] shows that this assumption is unnecessary.
Lemma 3.1.
Let be a spacetime admitting a Killing vector field . Let be a smooth spacelike hypersurface with the induced data and future unit normal . Write along . Then satisfies the following system along , for ,
(3.1) |
where denotes the Einstein tensor of restricted on the tangent bundle of , and is the Einstein tensor of . Or equivalently, in terms of the constraint operator,
(3.2) |
where .
We recall the statement of Theorem 1.6.
Theorem 1.6.
Let be an -dimensional connected Lorentzian spacetime, possibly with boundary, such that a subset of is diffeomorphic to , and in these coordinates,
for some . Assume that is equipped with Killing fields such that
Then form a covariant constant global Lorentzian orthonormal frame, and in particular, must be flat.
Proof.
We first claim that for any , . To see why, let . Since Killing fields form a Lie algebra under Lie bracket, is also a Killing field. Let be any constant slice of the asymptotically flat end of . If is the future unit normal to , then we can decompose and along , and routine computation shows that
By our asymptotic assumptions, it follows that . By Lemma 3.1, satisfies the system (3.1), which in turn implies that satisfies a homogeneous system of Hessian-type equations. By [CÓM81, Theorem 3.3], any satisfying such a system of equations with decay in must vanish identically.
At any point , define to be the maximum absolute value of all sectional curvatures of at . Thus if and only the full Riemann curvature tensor of vanishes at . Define the set
We will prove that , which implies the desired result on all of by continuity.
Our first goal is to prove that is non-empty, and here is where we must use our asymptotic assumptions. The fact that is asymptotic to guarantees that there exists a ball in the asymptotically flat end of with the property that following any integral curve of from will lead out to spatial infinity in . Since is Killing, must be constant along an integral curve of . But our asymptotic flatness assumptions tells us that as approaches spatial infinity, so it follows that vanishes on . Meanwhile, since and is Killing, it follows that for any , the product must be constant along integral curves of . Since we know that the ’s are asymptotically an orthonormal frame at infinity, it follows that they form an orthonormal frame at every point of . Finally, since the Riemann curvature vanishes on , it is locally isometric to Minkowski space. Recall that we know exactly what every Killing field on Minkowski space looks like (even locally). Specifically, since the ’s have constant length and the only Killing fields of Minkowski with constant length are the covariant constant ones (corresponding to translation symmetries), it follows that in . Hence .
It is clear that is relatively closed in . Therefore we can conclude that if we can show that is also relatively open in . Let , and consider the set of all points that can be reached by following integral curves of elements of originating at . The key point is that since the ’s are orthonormal at , this set must include a small ball around in . But since and must be constant along all of these integral curves for Killing fields, it follows that vanishes identically on . Once again, using our knowledge of the Killing fields of Minkowski space, we know that since and , it follows that must be covariant constant on . Consequently, the ’s are orthonormal at every point of . Thus , which is what we wanted to prove.
∎
By taking to be a trivial product, we can easily obtain a corresponding theorem in Riemannian geometry.
Corollary 3.2.
Let be a connected Riemannian manifold, possibly with boundary, that is “weakly777In other words, this result uses a much weaker definition of asymptotic flatness than the one given in Definition 1.1. asymptotically flat” in the sense that a subset of is diffeomorphic to , and in these coordinates,
for some . Assume that is equipped with Killing fields that are asymptotic to the coordinate translation directions in the sense that
Then form a covariant constant global orthonormal frame, and in particular, must be flat.
Again, note that there is no completeness assumption here, but when is complete (without boundary), it follows from the Killing-Hopf Theorem (and a simple covering space argument as in the proof of Theorem 1.4 below) that is isometric to Euclidean space.
In order to connect Lemma 2.2 to Theorem 1.6, we will use the concept of a Killing development, first introduced by Beig and Chruściel [BC96].
Definition 3.3.
Let be a Riemannian manifold, and let be a lapse-shift pair on with on . The Killing development of with respect to is the Lorentzian manifold where
(3.3) |
where denotes a new coordinate on the factor and are just any coordinates on the factor.
Observe that is a Killing field on .
Now suppose that is a vacuum initial data set equipped with a vacuum KID with (recall Definition 2.3). In this case, one can verify that the Killing development is vacuum and contains the initial data set in the sense that the slice is just the Riemannian manifold with second fundamental form (corresponding to via (2.1)). For details, see [BC96] or [HL, Appendix B]. Furthermore, by (3.2), one can see that every Killing field on restricts to a vacuum KID on . It turns out that this provides a one-to-one correspondence between Killing fields on and vacuum KIDs on . This fact is a simple special case of a theorem of Beig and Chruściel [BC97] that describes more general non-vacuum circumstances under which this one-to-one correspondence holds. A similar correspondence also holds for the maximal vacuum development (rather than the Killing development) by a classical result of V. Moncrief [Mon75].
Proof of Theorem 1.4.
Assume that satisfies the hypotheses of Theorem 1.4, but suppose that . We want to show that embeds into Minkowski space. Invoking Lemma 2.2, we can construct vacuum KIDs on such that on the specified end where
(3.4) | ||||
(3.5) |
for , where is the asymptotic decay rate of .
Consider the open subset of where , and let be the connected component of that set that contains our specified end. We define to be the Killing development of with respect to . Note that asymptotic flatness of together with (3.4) implies that the metric given by (3.3) is asymptotically flat in the sense of Theorem 1.6.
As discussed above, is vacuum and contains as the slice. Moreover, each gives rise to a Killing field . Note that the same argument used in the proof of Theorem 1.6 proves that for any , we have . (This is because that part of the proof only used asymptotics of along , which we know because of (3.4) and (3.5).) Recall that . Since we also have along , we have , and consequently, we also have
for . Since , this formula for holds on all rather than merely along , where are all extended to to be independent of the variable. In particular, it follows that is asymptotic to in the sense required for application of Theorem 1.6.
We can now invoke Theorem 1.6 to see that is flat and is a covariant constant global orthonormal frame. In particular, has constant length everywhere in . Or in other words, on all of . This means that on , and consequently, must be all of . So lies inside .
The last step is to prove that is globally isometric to Minkowski space. We will first consider the case where has no boundary, and then we will explain why having a nontrivial weakly outer trapped boundary is impossible.
Suppose has no boundary. Let be the universal cover of , and let be its Killing development with respect to the lift of . So is simply connected, and by what we proved above, is flat. Moreover, the lack of boundary implies that is geodesically complete. (This is a nontrivial fact since is Lorentzian, but it was previously explained in the proof of [BC96, Theorem 4.1].) By the Killing–Hopf Theorem,888Technically, this is a Lorentzian version of the Killing–Hopf Theorem for flat manifolds, but the proof is the same. must be isometric to Minkowski space. In particular, it follows that has only one asymptotically flat end, and we claim that this implies that the covering map is trivial, and the result follows. To see why the claim is true, let be an asymptotically flat end of , and consider the covering map from to obtained by restriction of . Since is simply connected, must be a disjoint union of isometric copies of . Since has only one asymptotically flat end, this implies that is one-to-one.
We show that cannot have a nontrivial weakly outer trapped boundary. Of course, it is well-known that it is impossible to have a weakly outer trapped boundary in Minkowski space, but we cannot apply this fact directly because at the moment we only know that our spacetime is both locally exactly Minkowski and asymptotically flat in the sense of Theorem 1.6. To complete the proof, we apply Theorem 1.7 where we take “” to be the part of bounded by and a large enough coordinate sphere so that it is strictly outer untrapped.
∎
Remark 3.4.
In the proof of Theorem 1.4, we primarily used the fact that is vacuum in order to conclude that each extends to a Killing field on the Killing development of with respect to . However, as mentioned earlier, a fascinating paper by Beig and Chruściel [BC97] explores when this is possible in non-vacuum scenarios, and it turns out that even without knowing that is vacuum, simply knowing that solves (2.4) turns out to be sufficient to conclude that each extends to a Killing field on the Killing development of with respect to , and also that this Killing development satisfies the dominant energy condition [HL]. This argument provides an alternative to invoking Corollary 6.6 of [HL] when proving Theorem 1.4, but the argument that we presented above is simpler.
4. Trapped and untrapped boundaries
We prove Theorem 1.7 in this section, which is a direct consequence of the following more general result.
Theorem 4.1.
Let be a compact connected manifold with boundary such that where and are both closed and nonempty, and let be a Riemannian metric on (which will only serve as a fixed Riemannian background metric). Define and , where denotes the coordinate function for the factor.
Let be a Lorentzian metric on that is in the sense that it extends to a Lorentzian metric on some manifold containing , and assume that is uniformly bounded on . We also assume that is a time function such that on all of for some uniform constant , and that each slice is spacelike.
Now suppose that satisfies the null energy condition, and that for all , is a strictly outer untrapped surface (that is, ) with respect to the “outward” future null normal pointing out of . Then cannot be a weakly outer trapped surface (that is, it cannot have on all of ) with respect to the “outward” future null normal pointing toward the interior of .
Our proof of Theorem 4.1 is elementary and does not depend on dimension. It is a Lorentzian analog of the following well-known theorem in Riemannian geometry, which follows from a classical argument of T. Frankel [Fra66]. We thank Jose Espinar for pointing us toward Frankel’s argument.
Theorem 4.2.
Let be a compact connected Riemannian manifold with boundary such that where and are both closed and nonempty. Suppose that has nonnegative Ricci curvature, and that satisfies everywhere with respect to the normal pointing out of . Then cannot satisfy with respect to the normal pointing into .
R. Ichida [Ich81] proved the stronger statement that if one instead assumes that and , then splits as a product.
Proof of Theorem 4.1.
By assumption, there is a larger spacetime manifold without boundary, called , which contains , and all of our causal theoretic definitions below take place within the spacetime , though it will be apparent that the geometry of outside is irrelevant in the proof. We also extend the Riemannian background metric on to a complete Riemannian metric on .
Our first claim is that the timelike future of , , intersects . Let be any curve with unit speed with respect to that connects to in . Let be a constant, and consider the curve in connecting to . Then
(4.1) |
where is the constant assumed in the hypotheses, and is interpreted to lie in . Since is uniformly bounded, the right side of (4.1) will be negative for sufficiently large . Thus is future-timelike for sufficiently large , proving the claim.
Let be the infimum over all such that intersects . So there is some point such that . In fact, can be written as the limit of ending points of future timelike curves that begin in and stay inside . (If the timelike curve crosses through the part of the boundary, then the first part of the curve up to the last time it intersects can be replaced by an integral curve of , and if the timelike curve crosses through the part, then the last part of the timelike curve after the first time it intersects can be discarded.) By a standard limit curve argument, there exists a limit curve of (the time-reversed version of) the above-described family of timelike curves such that is a past null geodesic starting at and lying in such that either ends at or is inextendible. See, for example, [Min19, Theorem 2.56] or [Gal14, Proposition 3.4]. Moreover, since is a limit curve of curves lying in the compact set , itself must lie in as well. In fact, lies in because it begins at and is past null.
We claim that is not inextendible, and consequently, it ends at . To prove the claim, recall that inextendible curves must have infinite -length, where is the complete Riemannian background metric on the spacetime . See, e.g. [Min19, Lemma 2.17]. Note that since each is assumed to be spacelike, in the compact space there must be a uniform constant with the property that for all tangent vectors that are tangent to the directions in the splitting . Writing , the null condition says that
(4.2) |
where is interpreted to lie in . Solving this quadratic inequality implies that , where depends on and the uniform bound on . Therefore the -length of can be bounded in terms of , which is bounded by since is a time function, is past pointing, and lies in . Hence has finite -length, establishing the claim.

With slight abuse of notation, let denote the same geodesic described above, but with parameter reversed so that it is now a future null geodesic starting at and ending at the point , lying entirely in .
By a standard argument (see, e.g. [O’N83, Lemma 50, p. 298]), must leave in an outward future null normal direction and not have any focal points in its interior (with respect to the surface ). Moreover, the construction of guarantees that is pointing in the outward future null normal direction when it hits at the point and that there are no focal points of along . (For otherwise, in either case there is a timelike curve from to , which contradicts the minimality of .)
It follows that the normal exponential map on the normal bundle of is nonsingular along [O’N83, Proposition 30, p. 283], and thus is contained in a smooth null hypersurface, denoted by , generated by outward future null normal geodesics starting from a small neighborhood of in . Similarly, there is a smooth null hypersurface, denoted by , generated by past inward null normal geodesics starting from a neighborhood of in . Note that lies in both and .
Since the spacelike -level sets must intersect the null hypersurfaces and transversely, the intersections are all smooth. For sufficiently small , let be the smooth slice of with (because of our assumption that ), and let be the smooth slice of . By construction, must intersect both and at some point . Moreover, the minimality of guarantees that is “outside” of as hypersurfaces of .
Now assume that is weakly outer trapped, that is, . Using a standard argument involving the Raychaudhuri equation and the null energy condition, it follows that at . Since we already saw that at and encloses , this contradicts the strong comparison principle for (see, for example, [AG05, Proposition 3.1]). ∎
Proof of Theorem 1.7.
Let be the Lorentzian metric and be the global timelike Killing vector from the hypotheses of Theorem 1.7. We verify that they satisfy the assumptions of Theorem 4.1. Let where is the induced Riemannian metric of on . Since is a global timelike Killing vector field, the metric is invariant in , and thus is uniformly bounded on and the -slices stay spacelike. The compactness of implies that for a uniform constant on and hence on all of . ∎
References
- [AG05] Abhay Ashtekar and Gregory J. Galloway. Some uniqueness results for dynamical horizons. Adv. Theor. Math. Phys., 9(1):1–30, 2005.
- [BC96] Robert Beig and Piotr T. Chruściel. Killing vectors in asymptotically flat space-times. I. Asymptotically translational Killing vectors and the rigid positive energy theorem. J. Math. Phys., 37(4):1939–1961, 1996.
- [BC97] Robert Beig and Piotr T. Chruściel. Killing initial data. volume 14, pages A83–A92. 1997. Geometry and physics.
- [CM06] Piotr T. Chruściel and Daniel Maerten. Killing vectors in asymptotically flat space-times. II. Asymptotically translational Killing vectors and the rigid positive energy theorem in higher dimensions. J. Math. Phys., 47(2):022502, 10, 2006.
- [CM08] Alberto Carrasco and Marc Mars. On marginally outer trapped surfaces in stationary and static spacetimes. Classical Quantum Gravity, 25(5):055011, 19, 2008.
- [CÓM81] D. Christodoulou and N. Ó Murchadha. The boost problem in general relativity. Comm. Math. Phys., 80(2):271–300, 1981.
- [EHLS16] Michael Eichmair, Lan-Hsuan Huang, Dan Lee, and Richard Schoen. The spacetime positive mass theorem in dimensions less than eight. J. Eur. Math. Soc. (JEMS), 18(1):83–121, 2016.
- [Eic09] Michael Eichmair. The Plateau problem for marginally outer trapped surfaces. J. Differential Geom., 83(3):551–583, 2009.
- [Eic13] Michael Eichmair. The Jang equation reduction of the spacetime positive energy theorem in dimensions less than eight. Comm. Math. Phys., 319(3):575–593, 2013.
- [Fra66] T. Frankel. On the fundamental group of a compact minimal submanifold. Ann. of Math. (2), 83:68–73, 1966.
- [Gal14] Gregory Galloway. Notes on Lorentzian causality. 2014.
- [GHHP83] Gary W. Gibbons, Stephen W. Hawking, Gary T. Horowitz, and Malcolm J. Perry. Positive mass theorems for black holes. Comm. Math. Phys., 88(3):295–308, 1983.
- [Her98] Marc Herzlich. The positive mass theorem for black holes revisited. J. Geom. Phys., 26(1-2):97–111, 1998.
- [HL] Lan-Hsuan Huang and Dan A. Lee. Bartnik mass minimizing initial data sets and improvability of the dominant energy scalar. J. Differential Geom. to appear.
- [HL20] Lan-Hsuan Huang and Dan A. Lee. Equality in the spacetime positive mass theorem. Comm. Math. Phys., 376(3):2379–2407, 2020.
- [HZ23] Sven Hirsch and Yiyue Zhang. The case of equality for the spacetime positive mass theorem. J. Geom. Anal., 33(1):Paper No. 30, 13, 2023.
- [Ich81] Ryosuke Ichida. Riemannian manifolds with compact boundary. Yokohama Math. J., 29(2):169–177, 1981.
- [LLU22] Dan A. Lee, Martin Lesourd, and Ryan Unger. Density and positive mass theorems for initial data sets with boundary. Comm. Math. Phys., 395(2):643–677, 2022.
- [Loh16] Joachim Lohkamp. The higher dimensional positive mass theorem II. arXiv:1612.07505, 2016.
- [Min19] Ettore Minguzzi. Lorentzian causality theory. Living Rev Relativ, 22(3), 2019.
- [Mon75] Vincent Moncrief. Spacetime symmetries and linearization stability of the Einstein equations. I. J. Mathematical Phys., 16:493–498, 1975.
- [O’N83] Barrett O’Neill. Semi-Riemannian geometry, volume 103 of Pure and Applied Mathematics. Academic Press, Inc. [Harcourt Brace Jovanovich, Publishers], New York, 1983. With applications to relativity.
- [SY79] Richard Schoen and Shing-Tung Yau. On the proof of the positive mass conjecture in general relativity. Comm. Math. Phys., 65(1):45–76, 1979.
- [SY81] Richard Schoen and Shing-Tung Yau. Proof of the positive mass theorem. II. Comm. Math. Phys., 79(2):231–260, 1981.
- [Wit81] Edward Witten. A new proof of the positive energy theorem. Comm. Math. Phys., 80(3):381–402, 1981.