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

Equality in the spacetime positive mass theorem II

Lan-Hsuan Huang Department of Mathematics, University of Connecticut, Storrs, CT 06269, USA [email protected]  and  Dan A. Lee CUNY Graduate Center and Queens College [email protected]
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 (M,g,k)(M,g,k) satisfying the dominant energy condition has null ADM energy-momentum (that is, |E|=|P||E|=|P|), then (M,g)(M,g) must isometrically embed into Minkowski space with kk 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 E|P|E\geq|P| for all complete initial data sets near (g,k)(g,k) on MM satisfying the dominant energy condition.

The first named author was partially supported by the NSF CAREER Award DMS-1452477 and NSF DMS-2005588. The authors thank Harvard’s Center of Mathematical Sciences and Applications for their hospitality while some of this research was being carried out.

1. Introduction

Let us first explicitly state the definition of asymptotic flatness we will be using.

Definition 1.1.

Let n3n\geq 3, let MM be an nn-dimensional connected manifold, possibly with boundary, and let n22<q<n2\frac{n-2}{2}<q<n-2. We say that an initial data set (M,g,k)(M,g,k) is complete asymptotically flat with decay rate qq if the following hold:

  • Each noncompact end of MM (of which there is at least one) is diffeomorphic to nB¯1(0)\mathbb{R}^{n}\setminus\bar{B}_{1}(0), and for some 0<α<10<\alpha<1,

    (1.1) (gg𝔼,k)Cq2,α×Cq11,α,(g-g_{\mathbb{E}},k)\in C^{2,\alpha}_{-q}\times C^{1,\alpha}_{-q-1},

    where g𝔼g_{\mathbb{E}} denotes a background metric that is equal to the standard Euclidean metric on each nB¯1(0)\mathbb{R}^{n}\setminus\bar{B}_{1}(0) end, and Cq2,αC^{2,\alpha}_{-q} and Cq11,αC^{1,\alpha}_{-q-1} denote weighted Hölder spaces.

  • The energy density μ\mu and current density JJ defined by

    (1.2) (μ,J):=(12(Rg|k|g2(trgk)2),(divgk)(trgk))(\mu,J):=\left(\,\tfrac{1}{2}(R_{g}-|k|^{2}_{g}-(\operatorname{tr}_{g}k)^{2})\,,\,(\operatorname{div}_{g}k)^{\sharp}-\nabla(\operatorname{tr}_{g}k)\,\right)

    are integrable over MM, where (divgk)(\operatorname{div}_{g}k)^{\sharp} denotes the vector field dual to the one-form divgk\operatorname{div}_{g}k.

Note that asymptotic flatness of (g,k)(g,k) implies that the ADM energy-momentum (E,P)(E,P) of each end is well-defined. Our main result characterizes the equality case E=|P|E=|P| of the spacetime positive mass theorem, but we would like to emphasize that, unlike all previous results characterizing E=|P|E=|P|, ours follows from the positive mass inequality E|P|E\geq|P| itself and does not rely on any particular proof of this inequality. Consequently, we introduce the following definition.

Definition 1.2.

Let (M,g,k)(M,g,k) be a complete asymptotically flat initial data set with decay rate qq, possibly with boundary. We say that the positive mass theorem is true near (g,k)(g,k) if there is an open ball around (g,k)(g,k) in (g𝔼+Cq2,α)×Cq11,α\left(g_{\mathbb{E}}+C^{2,\alpha}_{-q}\right)\times C^{1,\alpha}_{-q-1}, for some α\alpha, such that for each complete asymptotically flat initial data set (g¯,k¯)(\bar{g},\bar{k}) with decay rate qq in that open ball, satisfying the dominant energy condition and having weakly outer trapped boundary (if there is a boundary), we have E¯|P¯|\bar{E}\geq|\bar{P}|, where (E¯,P¯)(\bar{E},\bar{P}) denotes the ADM energy-momentum of any asymptotically flat end of (M,g¯,k¯)(M,\bar{g},\bar{k}).

We say the boundary M\partial M of an asymptotically flat manifold MM is weakly outer trapped if its outward null expansion θ+:=H+trMk\theta^{+}:=H+\operatorname{tr}_{\partial M}k satisfies θ+0\theta^{+}\leq 0, with respect to the normal that points toward MM.111Our convention is that the mean curvature HH 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 (g,k)(g,k) 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 (g,k)(g,k) 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 E=|P|E=|P| case of the spacetime positive mass theorem breaks down into two separate steps: The first step is to show that E=|P|=0E=|P|=0, and the second step is to use the fact that E=0E=0 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 E|P|E\geq|P| 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 (M,g,k)(M,g,k) be a complete asymptotically flat initial data set with decay rate qq, possibly with boundary, and assume that the positive mass theorem is true near (g,k)(g,k). Furthermore, we make the stronger asymptotic assumption that with this decay rate qq, there exists 0<α<10<\alpha<1 satisfying (1.1) such that

(1.3) q+α>n2,\displaystyle q+\alpha>n-2,

and

(1.4) (μ,J)Cnδ2,α,(\mu,J)\in C^{2,\alpha}_{-n-\delta},

for some δ>0\delta>0, where nn is the dimension. If (g,k)(g,k) satisfies the dominant energy condition and the boundary (if there is one) is weakly outer trapped, then for each asymptotically flat end, E=|P|E=|P| implies that E=|P|=0E=|P|=0.

Although the stronger asymptotic assumption (1.3) may seem undesirable in higher dimensions (note that (1.3) can be dropped for n=3,4n=3,4 if α\alpha in (1.1) is close to 11), we found pp-wave counterexamples showing that the theorem is false without it for n>8n>8. See [HL, Example 7] for details.

Once one knows that E=0E=0, the “second step” described above follows from work of Eichmair [Eic13] if n<8n<8. More specifically, following Schoen and Yau’s proof for the n=3n=3 case333When n=3n=3, one must make an extra decay assumption on trgk\operatorname{tr}_{g}k. [SY81], Eichmair used the Jang equation to prove the inequality E0E\geq 0, and then by analyzing the case of equality in that proof, he showed that the E=0E=0 case can only occur when the Jang solution provides a graphical isometric embedding of (M,g)(M,g) into Minkowski space with kk 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 E|P|E\geq|P| (nor on how one proves E0E\geq 0).

Theorem 1.4.

Let (M,g,k)(M,g,k) be a complete asymptotically flat initial data set with decay rate qq, possibly with boundary, and assume that the positive mass theorem is true near (g,k)(g,k). Further assume that (g,k)(g,k) is locally C5×C4C^{5}\times C^{4}. If (g,k)(g,k) satisfies the dominant energy condition and the boundary (if there is one) is weakly outer trapped, then E>0E>0 in each asymptotically flat end, unless (M,g)(M,g) has no boundary and isometrically embeds into Minkowski space with kk as its second fundamental form, in which case E=|P|=0E=|P|=0.

Combining the two previous theorems gives the following result.

Corollary 1.5.

Let (M,g,k)(M,g,k) be a complete asymptotically flat initial data set with decay rate qq, possibly with boundary, and assume that the positive mass theorem is true near (g,k)(g,k). Further assume that (1.3) and (1.4) hold, and that (g,k)(g,k) is locally C5×C4C^{5}\times C^{4}. If (g,k)(g,k) satisfies the dominant energy condition and the boundary (if there is one) is weakly outer trapped, then E>|P|E>|P| in each asymptotically flat end, unless (M,g)(M,g) has no boundary and embeds into Minkowski space with kk as its second fundamental form, in which case E=|P|=0E=|P|=0.

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 C5×C4C^{5}\times C^{4} regularity. More recently, Sven Hirsch and Yiyue Zhang [HZ23] gave a new proof in dimension 33 that avoids assuming any of (1.3), (1.4), or C5×C4C^{5}\times C^{4} 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 (g,k)(g,k) 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 E=0E=0, (g,k)(g,k) minimizes many different Regge-Teitelboim Hamiltonians, so we can actually construct an entire (n+1)(n+1)-dimensional space of these lapse-shift pairs that can be thought of as being asymptotic to the (n+1)(n+1)-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 “JJ-null-vector equation.” Moreover, having so many solutions implies that (g,k)(g,k) is vacuum. All of this is explained in Section 2, culminating in the statement of Lemma 2.2. Observe that proving that (g,k)(g,k) is vacuum is itself a difficult task. Note that the vacuum condition μ=|J|g=0\mu=|J|_{g}=0 is a much stronger condition than the “borderline” μ=|J|g\mu=|J|_{g} case of the dominant energy condition. The fact that E=|P|E=|P| implies μ=|J|g\mu=|J|_{g} 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 E=|P|E=|P| does not imply vacuum.

Once we know that (g,k)(g,k) is vacuum, it follows that each of these lapse-shift pairs is actually vacuum Killing initial data for (g,k)(g,k), and we can extend them to become actual Killing fields on the vacuum Killing development of (g,k)(g,k). 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 (n+1)(n+1) 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 (𝐍,𝐠)(\mathbf{N},\mathbf{g}) be an (n+1)(n+1)-dimensional connected C2C^{2} Lorentzian spacetime, possibly with boundary, that is asymptotically flat in the sense that a subset of 𝐍\mathbf{N} is diffeomorphic to ×(nB¯1(0))\mathbb{R}\times(\mathbb{R}^{n}\setminus\bar{B}_{1}(0)), and in these x0,,xnx^{0},\ldots,x^{n} coordinates,

𝐠μν(x0,x)=ημν+O2(|x|β)\mathbf{g}_{\mu\nu}(x^{0},x)=\eta_{\mu\nu}+O_{2}(|x|^{-\beta})

for some β>0\beta>0, where ημν\eta_{\mu\nu} is the standard Minkwoski metric and |x||x| denotes the norm of the spatial coordinates x:=(x1,,xn)x:=(x^{1},\ldots,x^{n}) only. More specifically, the O2O_{2} notation means that there exists a positive number CC such that

|I|=02||x|β+|I|I(𝐠μνημν)(x0,x)|C,\sum_{|I|=0}^{2}\big{|}|x|^{\beta+|I|}\partial^{I}(\mathbf{g}_{\mu\nu}-\eta_{\mu\nu})(x^{0},x)\big{|}\leq C,

where I\partial^{I} denotes the partial derivatives with respect to all (n+1)(n+1) coordinates.

Assume that (𝐍,𝐠)(\mathbf{N},\mathbf{g}) is equipped with Killing fields 𝐘0,,𝐘n\mathbf{Y}_{0},\ldots,\mathbf{Y}_{n} that are asymptotic to the coordinate translation directions in the sense that

𝐘μ=xμ+O2(|x|β).\mathbf{Y}_{\mu}=\frac{\partial}{\partial x^{\mu}}+O_{2}(|x|^{-\beta}).

Then 𝐘0,,𝐘n\mathbf{Y}_{0},\ldots,\mathbf{Y}_{n} form a covariant constant global Lorentzian orthonormal frame, and in particular, 𝐠\mathbf{g} 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 (𝐍,𝐠)(\mathbf{N},\mathbf{g}) is complete in Theorem 1.6. If (𝐍,𝐠)(\mathbf{N},\mathbf{g}) 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 Ω\Omega be a compact connected manifold with boundary such that Ω=ΣinΣout\partial\Omega=\Sigma_{\mathrm{in}}\sqcup\Sigma_{\mathrm{out}} where Σin\Sigma_{\mathrm{in}} and Σout\Sigma_{\mathrm{out}} are both closed and nonempty. Let 𝐍=[0,)×Ω\mathbf{N}=[0,\infty)\times\Omega, and let 𝐠\mathbf{g} be a C2C^{2} Lorentzian metric on 𝐍\mathbf{N} such that {0}×Ω\{0\}\times\Omega is spacelike. Suppose the following holds:

  1. (1)

    t\frac{\partial}{\partial t} is a global timelike Killing vector field, where tt denotes the coordinate for the [0,)[0,\infty) factor. Thus (𝐍,𝐠)(\bf N,\bf g) is strictly stationary.

  2. (2)

    𝐠\mathbf{g} satisfies the null energy condition.

  3. (3)

    {0}×Σout\{0\}\times\Sigma_{\mathrm{out}} is a strictly outer untrapped surface (that is, θ+>0\theta^{+}>0) with respect to the “outward” future null normal pointing out of 𝐍\mathbf{N}.

Then {0}×Σin\{0\}\times\Sigma_{\mathrm{in}} cannot be a weakly outer trapped surface (that is, it cannot have θ+0\theta^{+}\leq 0 on all of {0}×Σin\{0\}\times\Sigma_{\mathrm{in}}) with respect to the “outward” future null normal pointing toward the interior of 𝐍\mathbf{N}.

When the spatial dimension is less than 88, 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 𝐠=f2dt2+g\mathbf{g}=-f^{2}dt^{2}+g where f>0f>0, the null energy condition becomes the assumption

(1.5) fRicg2f+(Δgf)g0.\displaystyle f\mathrm{Ric}_{g}-\nabla^{2}f+(\Delta_{g}f)g\geq 0.
Corollary 1.8.

Let (Ω,g)(\Omega,g) be a compact connected C2C^{2} Riemannian manifold with boundary such that Ω=ΣinΣout\partial\Omega=\Sigma_{\mathrm{in}}\sqcup\Sigma_{\mathrm{out}} where Σin\Sigma_{\mathrm{in}} and Σout\Sigma_{\mathrm{out}} are both closed and nonempty. Suppose there is a C2C^{2} function f>0f>0 on Ω\Omega such that (1.5) holds.

Suppose Σout\Sigma_{\mathrm{out}} has mean curvature H>0H>0 with respect to the unit normal pointing out of Ω\Omega. Then Σin\Sigma_{\mathrm{in}} cannot have H0H\leq 0 everywhere, with respect to the unit normal pointing into Ω\Omega.

Specializing even more to the case where f=1f=1 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 E=0E=0 case using the new results from [HL] to construct an (n+1)(n+1)-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 Cqk,α(M)C^{k,\alpha}_{-q}(M) and the weighted Sobolev space Wqk,p(M)W^{k,p}_{-q}(M). Assume p>np>n, and let q2,p\mathscr{M}^{2,p}_{-q} denote the set of symmetric (0,2)(0,2)-tensors γ\gamma such that γg𝔼Wq2,p(M)\gamma-g_{\mathbb{E}}\in W^{2,p}_{-q}(M) and γ\gamma is positive definite at each point. By the inclusion Cqk,αWqk,pC^{k,\alpha}_{-q}\subset W^{k,p}_{-q^{\prime}}, for any q<qq^{\prime}<q, whenever we work on the weighted Sobolev spaces, we shall assume the decay rate qq in the weighted Sobolev spaces to be slightly smaller than the assumed asymptotic decay rate appearing in Definition 1.1.

Initial data (g,k)(g,k) can be equivalently described by a pair (g,π)(g,\pi) where π\pi is a symmetric (2,0)(2,0)-tensor called the conjugate momentum tensor, which is related to the (0,2)(0,2)-tensor kk via the equation

(2.1) kij:=gigjmπm1n1(trgπ)gij.\displaystyle k_{ij}:=g_{i\ell}g_{jm}\pi^{\ell m}-\tfrac{1}{n-1}(\operatorname{tr}_{g}\pi)g_{ij}.

By slight abuse of vocabulary, we will refer to (g,π)(g,\pi) as initial data. We define the constraint map Φ\Phi on initial data by

Φ(g,π)=(2μ,J),\Phi(g,\pi)=(2\mu,J),

where μ\mu and JJ are functions of (g,π)(g,\pi), defined by (1.2) combined with (2.1). Recall that the dominant energy condition (or DEC) is the condition μ|J|g\mu\geq|J|_{g}.

As in [HL], given an initial data set (M,g,π)(M,g,\pi) and a smooth function φ\varphi on MM, we introduce the φ\varphi-modified constraint operator, which is defined on initial data (γ,τ)(\gamma,\tau) on MM by

(2.2) Φ(g,π)φ(γ,τ)\displaystyle\Phi^{\varphi}_{(g,\pi)}(\gamma,\tau) :=Φ(γ,τ)+(2φ|J|γ2,(12+φ|J|g)γJ),\displaystyle:=\Phi(\gamma,\tau)+\left(2\varphi|J|_{\gamma}^{2},(\tfrac{1}{2}+\varphi|J|_{g})\gamma\cdot J\right),

where JJ denotes the current density of (g,π)(g,\pi), and (γJ)i:=gijγjkJk(\gamma\cdot J)^{i}:=g^{ij}\gamma_{jk}J^{k} denotes the contraction of γ\gamma and JJ with respect to gg. The crucial property of Φ(g,π)φ\Phi^{\varphi}_{(g,\pi)} is that for any (γ,τ)(\gamma,\tau), if (g,π)(g,\pi) satisfies the DEC, |γg|g<1|\gamma-g|_{g}<1, and

Φ(g,π)φ(γ,τ)=Φ(g,π)φ(g,π),\Phi^{\varphi}_{(g,\pi)}(\gamma,\tau)=\Phi^{\varphi}_{(g,\pi)}(g,\pi),

then (γ,τ)(\gamma,\tau) also satisfies the DEC (see Lemma 3.4 of [HL]). Note that this is a property that the ordinary constraint operator Φ\Phi does not have, but since Φ(g,π)φ\Phi^{\varphi}_{(g,\pi)} only differs from Φ\Phi by an affine function of (γ,τ)(\gamma,\tau), it behaves similarly from a PDE perspective.

Throughout this article, a lapse-shift pair (f,X)(f,X) on MM will simply refer to a function ff and a vector field XX on MM. Assuming that (g,π)(g,\pi) is locally C5×C4C^{5}\times C^{4}, Theorem 4.1 of [HL], which was the main technical result of [HL], tells us444Take Z=JZ=J in the statement of that theorem. that there exists a dense subset 𝒟C3(IntM)\mathcal{D}\subset C^{3}(\operatorname{Int}M) such that if φ𝒟\varphi\in\mathcal{D}, then any lapse-shift pair (f,X)(f,X) solving

(2.3) (DΦ(g,π)φ|(g,π))(f,X)=0\left(\left.D\Phi_{(g,\pi)}^{\varphi}\right|_{(g,\pi)}\right)^{*}(f,X)=0

in the interior of MM must also satisfy the pair of equations

(2.4) (DΦ(g,π)0|(g,π))(f,X)=02fJ+|J|gX=0,\displaystyle\begin{split}\left(\left.D\Phi_{(g,\pi)}^{0}\right|_{(g,\pi)}\right)^{*}(f,X)&=0\\ 2fJ+|J|_{g}X&=0,\end{split}

where the second equation is what we call the JJ-null-vector equation. Choosing any bounded φ𝒟\varphi\in\mathcal{D} ensures that the terms in (2.2) involving φ\varphi have the appropriate fall-off rates so that

Φ(g,π)φ:q2,p×W1q1,pL2qp.\Phi^{\varphi}_{(g,\pi)}:\mathscr{M}^{2,p}_{-q}\times W^{1,p}_{-1-q}\longrightarrow L^{p}_{-2-q}.

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 φ=0\varphi=0 to allow for general φ\varphi.

Definition 2.1.

Let (M,g,π)(M,g,\pi) be a complete asymptotically flat manifold, possibly with boundary, and choose qq to be slightly smaller than the asymptotic decay rate of (g,π)(g,\pi). Fix an end of MM and constants aa\in\mathbb{R} and bnb\in\mathbb{R}^{n}. Let (f¯,X¯)(\bar{f},\bar{X}) be a smooth lapse-shift pair on MM such that (f¯,X¯)(\bar{f},\bar{X}) is equal to (a,bixi)\left(a,\sum b^{i}\frac{\partial}{\partial x^{i}}\right) in the specified end and vanishes on all other ends of MM (if any) and in a neighborhood of M\partial M (if nonempty). Let φ\varphi be a smooth bounded function on MM.

We define the φ\varphi-modified Regge-Teitelboim Hamiltonian corresponding to (g,π)(g,\pi) and (f¯,X¯)(\bar{f},\bar{X}) to be the function :q2,p×W1q1,p\mathcal{H}:\mathscr{M}^{2,p}_{-q}\times W^{1,p}_{-1-q}\longrightarrow\mathbb{R} given by

(2.5) (γ,τ)=(n1)ωn1[2aE(γ,τ)+bP(γ,τ)]MΦ(g,π)φ(γ,τ)(f¯,X¯)𝑑μg\displaystyle\mathcal{H}(\gamma,\tau)=(n-1)\omega_{n-1}\left[2aE(\gamma,\tau)+b\cdot P(\gamma,\tau)\right]-\int_{M}\Phi^{\varphi}_{(g,\pi)}(\gamma,\tau)\cdot(\bar{f},\bar{X})\,d\mu_{g}

where E(γ,τ)E(\gamma,\tau) and P(γ,τ)P(\gamma,\tau) denote the ADM energy-momentum of the specified end, and the volume measure dμgd\mu_{g} and the inner product in the integral are both with respect to gg.

As explained in [HL20], although the terms in (2.5) are not individually well-defined for arbitrary (γ,τ)q2,p×W1q1,p(\gamma,\tau)\in\mathscr{M}^{2,p}_{-q}\times W^{1,p}_{-1-q}, the functional \mathcal{H} extends to all of q2,p×W1q1,p\mathscr{M}^{2,p}_{-q}\times W^{1,p}_{-1-q} in a natural way.

In what follows, we will assume that aa and bb satisfy 2a|b|2a\geq|b|. Continuing to follow [HL20] (with input from [LLU22] for the boundary case), we define

𝒞(g,π)={(γ,τ)q2,p×W1q1,p|Φ(g,π)φ(γ,τ)=Φ(g,π)φ(g,π) and M (if nonempty) has the same outward null expansion θ+ with respect to both (γ,τ) and (g,π)}.\mathcal{C}_{(g,\pi)}=\big{\{}(\gamma,\tau)\in\mathscr{M}^{2,p}_{-q}\times W^{1,p}_{-1-q}\,\big{|}\,\Phi^{\varphi}_{(g,\pi)}(\gamma,\tau)=\Phi^{\varphi}_{(g,\pi)}(g,\pi)\text{ and $\partial M$ (if nonempty) has }\\ \text{the same outward null expansion $\theta^{+}$ with respect to both }(\gamma,\tau)\text{ and }(g,\pi)\big{\}}.

We claim that that if (g,π)(g,\pi) satisfies the DEC with M\partial M weakly outer trapped (if nonempty) and has E=|P|=0E=|P|=0, then (g,π)(g,\pi) is a local minimizer of \mathcal{H} in 𝒞(g,π)\mathcal{C}_{(g,\pi)}. To see why, observe that for all nearby (γ,τ)𝒞(g,π)(\gamma,\tau)\in\mathcal{C}_{(g,\pi)}, (γ,τ)(\gamma,\tau) 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 (g,π)(g,\pi) tells us that E(γ,τ)|P(γ,τ)|E(\gamma,\tau)\geq|P(\gamma,\tau)|. (Technically, since (γ,τ)(\gamma,\tau) only has Sobolev regularity, we must invoke [HL20, Theorem 4.1] and [LLU22, Lemma 4.2].) Therefore

1(n1)ωn1((γ,τ)(g,π))=2aE(γ,τ)+bP(γ,τ)2aE(γ,τ)2a|P(γ,τ)|=2a(E(γ,τ)|P(γ,τ)|)0,\tfrac{1}{(n-1)\omega_{n-1}}\big{(}\mathcal{H}(\gamma,\tau)-\mathcal{H}(g,\pi)\big{)}=2aE(\gamma,\tau)+b\cdot P(\gamma,\tau)\\ \geq 2aE(\gamma,\tau)-2a|P(\gamma,\tau)|=2a\big{(}E(\gamma,\tau)-|P(\gamma,\tau)|\big{)}\geq 0,

where we used the assumptions that 2a|b|2a\geq|b| and (g,π)(g,\pi) has E=|P|=0E=|P|=0. Hence the claim is true.

Thus (g,π)(g,\pi) locally minimizes the functional \mathcal{H} subject to the constraint 𝒞(g,π)\mathcal{C}_{(g,\pi)}. 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 𝒞(g,π)\mathcal{C}_{(g,\pi)} to be the level set of a map on q2,p×W1q1,p\mathscr{M}^{2,p}_{-q}\times W^{1,p}_{-1-q} whose linearization is surjective (see [HL20, Theorem D.1] for details). For φ=0\varphi=0, 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 φ\varphi 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 (f,X)Cloc2,α(IntM)(f,X)\in C^{2,\alpha}_{\mathrm{loc}}(\operatorname{Int}M) such that

(DΦ(g,π)φ|(g,π))(f,X)=0 in IntM,\displaystyle\left(\left.D\Phi^{\varphi}_{(g,\pi)}\right|_{(g,\pi)}\right)^{*}(f,X)=0\quad\mbox{ in }\operatorname{Int}M,
(f,X)=(a,bixi)+O2+α(|x|q) in the specified end where E=0, and\displaystyle(f,X)=\left(a,\sum b^{i}\frac{\partial}{\partial x^{i}}\right)+O_{2+\alpha}(|x|^{-q})\text{ in the specified end where $E=0$, and}
(f,X)=O2+α(|x|q) in each other end,\displaystyle(f,X)=O_{2+\alpha}(|x|^{-q})\text{ in each other end,}

where O2+α(|x|q)O_{2+\alpha}(|x|^{-q}) denotes an object that decays in Cq2,αC^{2,\alpha}_{-q}. 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 MM can only have one asymptotically flat end.555The same argument also applies to the E=|P|E=|P| case and leads to the following fact. If (M,g,π)(M,g,\pi) is a complete asymptotically flat initial data set satisfying the dominant energy condition with E=|P|E=|P|, and if the positive mass theorem is true near (g,k)(g,k), then MM has only one end. Note that this is consistent with the pp-waves counterexamples having E=|P|0E=|P|\neq 0 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 φ\varphi so that Theorem 4.1 of [HL] applies, this (f,X)(f,X) also satisfies (2.4) on IntM\operatorname{Int}M. Since we can construct such a solution for any (a,b)n+1(a,b)\in\mathbb{R}^{n+1} satisfying 2a|b|2a\geq|b|, and the space of all (f,X)(f,X) satisfying (2.4) forms a vector space 𝒩\mathcal{N}, it follows that for any (a,b)n+1(a,b)\in\mathbb{R}^{n+1}, we obtain a lapse-shift pair (f,X)(f,X) which is both asymptotic to (a,bixi)\left(a,\sum b^{i}\frac{\partial}{\partial x^{i}}\right) in the desired end and also satisfies (2.4). Moreover, the dimension of 𝒩\mathcal{N} is at least n+1n+1.

But Corollary 6.6 of [HL] states that if the dimension of 𝒩\mathcal{N} is greater than 1, then (g,π)(g,\pi) must be vacuum initial data. Since (g,π)(g,\pi) is vacuum, Φ(g,π)φ=Φ(g,π)0=Φ\Phi^{\varphi}_{(g,\pi)}=\Phi^{0}_{(g,\pi)}=\Phi. The intuition behind Corollary 6.6 of [HL] is the following: (f,X)(f,X) lying in the kernel of the overdetermined Hessian-type operator (DΦ(g,π)0|(g,π))\left(\left.D\Phi^{0}_{(g,\pi)}\right|_{(g,\pi)}\right)^{*} already means that (f,X)(f,X) is determined by its 11-jet at any point. If there is a point pp where J0J\neq 0, we can use (2.4) to write f\nabla f in terms of ff, and hence the 0-jet of ff at pp determines the 11-jet of ff at pp, which then determines the 11-jet of (f,X)(f,X) at pp by the JJ-null-vector equation, which determines (f,X)(f,X) everywhere. For details, see [HL].

Putting together the arguments described in this section, we have proved the following lemma.

Lemma 2.2.

Let (M,g,π)(M,g,\pi) satisfy the hypotheses of Theorem 1.4, but suppose that E=0E=0. Then (g,π)(g,\pi) is vacuum, and for any (a,b)n+1(a,b)\in\mathbb{R}^{n+1}, there exists a lapse-shift pair (f,X)Cloc2,α(IntM)(f,X)\in C^{2,\alpha}_{\mathrm{loc}}(\operatorname{Int}M) such that

(DΦ|(g,π))(f,X)=0 in IntM\displaystyle\left(\left.D\Phi\right|_{(g,\pi)}\right)^{*}(f,X)=0\quad\mbox{ in }\operatorname{Int}M
(f,X)=(a,bixi)+O2+α(|x|q) in the specified end where E=0.\displaystyle(f,X)=\left(a,\sum b^{i}\frac{\partial}{\partial x^{i}}\right)+O_{2+\alpha}(|x|^{-q})\text{ in the specified end where $E=0$}.
Definition 2.3.

Let (M,g,π)(M,g,\pi) be a vacuum initial data set. Then any lapse-shift pair (f,X)(f,X) satisfying (DΦ|(g,π))(f,X)=0\left(\left.D\Phi\right|_{(g,\pi)}\right)^{*}(f,X)=0 is called vacuum Killing initial data (or a vacuum KID) on (M,g,π)(M,g,\pi).

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 UU is transverse to 𝐘\mathbf{Y}. Performing the computation in Gaussian coordinates as in [BC97] shows that this assumption is unnecessary.

Lemma 3.1.

Let (𝐍,𝐠)(\mathbf{N},\mathbf{g}) be a spacetime admitting a Killing vector field 𝐘\mathbf{Y}. Let UU be a smooth spacelike hypersurface with the induced data (g,π)(g,\pi) and future unit normal 𝐧\mathbf{n}. Write 𝐘=2f𝐧+Y\mathbf{Y}=2f\mathbf{n}+Y along UU. Then (f,X)(f,X) satisfies the following system along UU, for i,j,=1,,ni,j,\ell=1,\dots,n,

(3.1) 12(LXg)ij=(2n1(trgπ)gij2πij)fij2f+(Δgf)gij=12(LXπ)ij+(𝐆ijGij)f[3n1(trgπ)πij+2πiπj]f[(12(n1)(trgπ)212|π|g2)gij]f\displaystyle\begin{split}\tfrac{1}{2}(L_{X}g)_{ij}&=\left(\tfrac{2}{n-1}(\operatorname{tr}_{g}\pi)g_{ij}-2\pi_{ij}\right)f\\ -\nabla^{2}_{ij}f+(\Delta_{g}f)g_{ij}&=\tfrac{1}{2}(L_{X}\pi)_{ij}+\left(\mathbf{G}_{ij}-G_{ij}\right)f\\ &\quad-\left[-\tfrac{3}{n-1}(\operatorname{tr}_{g}\pi)\pi_{ij}+2\pi_{i\ell}\pi^{\ell j}\right]f\\ &\quad-\left[\left(\tfrac{1}{2(n-1)}(\operatorname{tr}_{g}\pi)^{2}-\tfrac{1}{2}|\pi|_{g}^{2}\right)g_{ij}\right]f\end{split}

where 𝐆ij\mathbf{G}_{ij} denotes the Einstein tensor of 𝐠\mathbf{g} restricted on the tangent bundle of UU, and GijG_{ij} is the Einstein tensor of gg. Or equivalently, in terms of the constraint operator,

(3.2) (DΦ|(g,π))(f,X)\displaystyle\left(\left.D\Phi\right|_{(g,\pi)}\right)^{*}(f,X) =((𝐆ij+μgij)f12gijJ,Xg(JX)ij,0)\displaystyle=\Big{(}-\left(\mathbf{G}_{ij}+\mu g_{ij}\right)f-\tfrac{1}{2}g_{ij}\langle J,X\rangle_{g}-(J\odot X)_{ij},0\Big{)}

where (JX)ij:=12(JiXj+JjXi)(J\odot X)_{ij}:=\tfrac{1}{2}(J_{i}X_{j}+J_{j}X_{i}).

We recall the statement of Theorem 1.6.

Theorem 1.6.

Let (𝐍,𝐠)(\mathbf{N},\mathbf{g}) be an (n+1)(n+1)-dimensional connected C2C^{2} Lorentzian spacetime, possibly with boundary, such that a subset of 𝐍\mathbf{N} is diffeomorphic to ×(nB¯1(0))\mathbb{R}\times(\mathbb{R}^{n}\setminus\bar{B}_{1}(0)), and in these x0,,xnx^{0},\ldots,x^{n} coordinates,

𝐠μν(x0,x)=ημν+O2(|x|β)\mathbf{g}_{\mu\nu}(x^{0},x)=\eta_{\mu\nu}+O_{2}(|x|^{-\beta})

for some β>0\beta>0. Assume that (𝐍,𝐠)(\mathbf{N},\mathbf{g}) is equipped with Killing fields 𝐘0,,𝐘n\mathbf{Y}_{0},\ldots,\mathbf{Y}_{n} such that

𝐘μ=xμ+O2(|x|β).\mathbf{Y}_{\mu}=\frac{\partial}{\partial x^{\mu}}+O_{2}(|x|^{-\beta}).

Then 𝐘0,,𝐘n\mathbf{Y}_{0},\ldots,\mathbf{Y}_{n} form a covariant constant global Lorentzian orthonormal frame, and in particular, 𝐠\mathbf{g} must be flat.

Proof.

We first claim that for any μ,ν=0,,n\mu,\nu=0,\ldots,n, [𝐘μ,𝐘ν]=0[\mathbf{Y}_{\mu},\mathbf{Y}_{\nu}]=0. To see why, let 𝐘=[𝐘μ,𝐘ν]\mathbf{Y}=[\mathbf{Y}_{\mu},\mathbf{Y}_{\nu}]. Since Killing fields form a Lie algebra under Lie bracket, 𝐘\mathbf{Y} is also a Killing field. Let U𝐍U\subset\mathbf{N} be any constant x0x^{0} slice of the asymptotically flat end of 𝐍\mathbf{N}. If 𝐧\mathbf{n} is the future unit normal to UU, then we can decompose 𝐘=2f𝐧+X\mathbf{Y}=2f\mathbf{n}+X and 𝐘μ=2fμ𝐧+Xμ\mathbf{Y}_{\mu}=2f_{\mu}\mathbf{n}+X_{\mu} along UU, and routine computation shows that

f\displaystyle f =Xμ(fν)Xν(fμ)\displaystyle=X_{\mu}(f_{\nu})-X_{\nu}(f_{\mu})
X\displaystyle X =[Xμ,Xν]+4(fμfνfνfμ).\displaystyle=[X_{\mu},X_{\nu}]+4(f_{\mu}\nabla f_{\nu}-f_{\nu}\nabla f_{\mu}).

By our asymptotic assumptions, it follows that (f,X)=O1(|x|β1)(f,X)=O_{1}(|x|^{-\beta-1}). By Lemma 3.1, (f,X)(f,X) satisfies the system (3.1), which in turn implies that (f,X)(f,X) satisfies a homogeneous system of Hessian-type equations. By [CÓM81, Theorem 3.3], any (f,X)(f,X) satisfying such a system of equations with decay in O1(|x|β1)O_{1}(|x|^{-\beta-1}) must vanish identically.

At any point p𝐍p\in\mathbf{N}, define K(p)K(p) to be the maximum absolute value of all sectional curvatures of 𝐠\mathbf{g} at pp. Thus K(p)=0K(p)=0 if and only the full Riemann curvature tensor of 𝐠\mathbf{g} vanishes at pp. Define the set

𝐕:={pInt𝐍|K(p)=0,{𝐘0,,𝐘n} is a Lorentzian orthonormal frame at p, and 𝐘0(p)==𝐘n(p)=0}.\displaystyle\begin{split}\mathbf{V}:=\big{\{}p\in\operatorname{Int}\mathbf{N}\,\big{|}\,K(p)=0,\,\{\mathbf{Y}_{0},\ldots,\mathbf{Y}_{n}\}\text{ is a Lorentzian orthonormal frame at }p,\\ \text{ and }\nabla\mathbf{Y}_{0}(p)=\cdots=\nabla\mathbf{Y}_{n}(p)=0\big{\}}.\end{split}

We will prove that 𝐕=Int𝐍\mathbf{V}=\operatorname{Int}\mathbf{N}, which implies the desired result on all of 𝐍\mathbf{N} by continuity.

Our first goal is to prove that 𝐕\mathbf{V} is non-empty, and here is where we must use our asymptotic assumptions. The fact that 𝐘1\mathbf{Y}_{1} is asymptotic to x1\frac{\partial}{\partial x^{1}} guarantees that there exists a ball BB in the asymptotically flat end of 𝐍\mathbf{N} with the property that following any integral curve of 𝐘1\mathbf{Y}_{1} from BB will lead out to spatial infinity in 𝐍\mathbf{N}. Since 𝐘1\mathbf{Y}_{1} is Killing, KK must be constant along an integral curve of 𝐘1\mathbf{Y}_{1}. But our asymptotic flatness assumptions tells us that K(p)0K(p)\to 0 as pp approaches spatial infinity, so it follows that KK vanishes on BB. Meanwhile, since [𝐘1,𝐘μ]=0[\mathbf{Y}_{1},\mathbf{Y}_{\mu}]=0 and 𝐘1\mathbf{Y}_{1} is Killing, it follows that for any μ,ν\mu,\nu, the product 𝐠(𝐘μ,𝐘ν)\mathbf{g}(\mathbf{Y}_{\mu},\mathbf{Y}_{\nu}) must be constant along integral curves of 𝐘1\mathbf{Y}_{1}. Since we know that the 𝐘μ\mathbf{Y}_{\mu}’s are asymptotically an orthonormal frame at infinity, it follows that they form an orthonormal frame at every point of BB. Finally, since the Riemann curvature vanishes on BB, 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 𝐘μ\mathbf{Y}_{\mu}’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 𝐘μ=0\nabla\mathbf{Y}_{\mu}=0 in BB. Hence B𝐕B\subset\mathbf{V}.

It is clear that 𝐕\mathbf{V} is relatively closed in Int𝐍\operatorname{Int}\mathbf{N}. Therefore we can conclude that 𝐕=Int𝐍\mathbf{V}=\operatorname{Int}\mathbf{N} if we can show that 𝐕\mathbf{V} is also relatively open in Int𝐍\operatorname{Int}\mathbf{N}. Let p𝐕p\in\mathbf{V}, and consider the set of all points that can be reached by following integral curves of elements of span(𝐘0,,𝐘n)\mathrm{span}(\mathbf{Y}_{0},\ldots,\mathbf{Y}_{n}) originating at pp. The key point is that since the 𝐘μ\mathbf{Y}_{\mu}’s are orthonormal at pp, this set must include a small ball BB around pp in 𝐍\mathbf{N}. But since K(p)=0K(p)=0 and KK must be constant along all of these integral curves for Killing fields, it follows that KK vanishes identically on BB. Once again, using our knowledge of the Killing fields of Minkowski space, we know that since 𝐘μ(p)0\mathbf{Y}_{\mu}(p)\neq 0 and 𝐘μ(p)=0\nabla\mathbf{Y}_{\mu}(p)=0, it follows that 𝐘μ\mathbf{Y}_{\mu} must be covariant constant on BB. Consequently, the 𝐘μ\mathbf{Y}_{\mu}’s are orthonormal at every point of BB. Thus B𝐕B\subset\mathbf{V}, which is what we wanted to prove.

By taking (𝐍,𝐠)(\mathbf{N},\mathbf{g}) to be a trivial product, we can easily obtain a corresponding theorem in Riemannian geometry.

Corollary 3.2.

Let (M,g)(M,g) be a connected C2C^{2} 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 MM is diffeomorphic to nB¯1(0)\mathbb{R}^{n}\setminus\bar{B}_{1}(0), and in these x1,,xnx^{1},\ldots,x^{n} coordinates,

gij=δij+O2(|x|β),g_{ij}=\delta_{ij}+O_{2}(|x|^{-\beta}),

for some β>0\beta>0. Assume that (M,g)(M,g) is equipped with Killing fields X1,,XnX_{1},\ldots,X_{n} that are asymptotic to the coordinate translation directions in the sense that

Xi=xi+O2(|x|β).X_{i}=\frac{\partial}{\partial x^{i}}+O_{2}(|x|^{-\beta}).

Then X1,,XnX_{1},\ldots,X_{n} form a covariant constant global orthonormal frame, and in particular, gg must be flat.

Again, note that there is no completeness assumption here, but when (M,g)(M,g) 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 (M,g)(M,g) 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 (M,g)(M,g) be a Riemannian manifold, and let (f0,X0)(f_{0},X_{0}) be a lapse-shift pair on MM with f0>0f_{0}>0 on MM. The Killing development of (M,g)(M,g) with respect to (f0,X0)(f_{0},X_{0}) is the Lorentzian manifold (𝐍:=×M,𝐠)(\mathbf{N}:=\mathbb{R}\times M,\mathbf{g}) where

(3.3) 𝐠:=4f02(dx0)2+gij(dxi+X0idx0)(dxj+X0jdx0),\mathbf{g}:=-4f_{0}^{2}(dx^{0})^{2}+g_{ij}(dx^{i}+X_{0}^{i}dx^{0})(dx^{j}+X_{0}^{j}dx^{0}),

where x0x^{0} denotes a new coordinate on the \mathbb{R} factor and xix^{i} are just any coordinates on the MM factor.

Observe that 𝐘0=x0\mathbf{Y}_{0}=\frac{\partial}{\partial x^{0}} is a Killing field on (𝐍,𝐠)(\mathbf{N},\mathbf{g}).

Now suppose that (M,g,π)(M,g,\pi) is a vacuum initial data set equipped with a vacuum KID (f0,X0)(f_{0},X_{0}) with f0>0f_{0}>0 (recall Definition 2.3). In this case, one can verify that the Killing development (𝐍,𝐠)(\mathbf{N},\mathbf{g}) is vacuum and contains the initial data set (M,g,π)(M,g,\pi) in the sense that the x0=0x^{0}=0 slice is just the Riemannian manifold (M,g)(M,g) with second fundamental form kk (corresponding to π\pi via (2.1)). For details, see [BC96] or [HL, Appendix B]. Furthermore, by (3.2), one can see that every Killing field 𝐘\mathbf{Y} on (𝐍,𝐠)(\mathbf{N},\mathbf{g}) restricts to a vacuum KID (f,X)(f,X) on (M,g,π)(M,g,\pi). It turns out that this provides a one-to-one correspondence between Killing fields on (𝐍,𝐠)(\mathbf{N},\mathbf{g}) and vacuum KIDs on (M,g,π)(M,g,\pi). 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 (M,g,π)(M,g,\pi) satisfies the hypotheses of Theorem 1.4, but suppose that E=0E=0. We want to show that (M,g,π)(M,g,\pi) embeds into Minkowski space. Invoking Lemma 2.2, we can construct vacuum KIDs (f0,X0),(f1,X1),,(fn,Xn)(f_{0},X_{0}),(f_{1},X_{1}),\ldots,(f_{n},X_{n}) on (M,g,π)(M,g,\pi) such that on the specified end where E=0E=0

(3.4) (2f0,X0)\displaystyle(2f_{0},X_{0}) =(1,0)+O2(|x|q)\displaystyle=(1,0)+O_{2}(|x|^{-q})
(3.5) (2fi,Xi)\displaystyle(2f_{i},X_{i}) =(0,xi)+O2(|x|q),\displaystyle=\left(0,\frac{\partial}{\partial x^{i}}\right)+O_{2}(|x|^{-q}),

for i=1,,ni=1,\ldots,n, where qq is the asymptotic decay rate of (M,g,π)(M,g,\pi).

Consider the open subset of MM where f0>0f_{0}>0, and let UU be the connected component of that set that contains our specified end. We define (𝐍,𝐠)(\mathbf{N},\mathbf{g}) to be the Killing development of (U,g)(U,g) with respect to (f0,X0)(f_{0},X_{0}). Note that asymptotic flatness of (g,π)(g,\pi) together with (3.4) implies that the metric 𝐠\mathbf{g} given by (3.3) is asymptotically flat in the sense of Theorem 1.6.

As discussed above, (𝐍,𝐠)(\mathbf{N},\mathbf{g}) is vacuum and contains (U,g,π)(U,g,\pi) as the x0=0x^{0}=0 slice. Moreover, each (fμ,Xμ)(f_{\mu},X_{\mu}) gives rise to a Killing field 𝐘μ\mathbf{Y}_{\mu}. Note that the same argument used in the proof of Theorem 1.6 proves that for any μ,ν=0,,n\mu,\nu=0,\ldots,n, we have [𝐘μ,𝐘ν]=0[\mathbf{Y}_{\mu},\mathbf{Y}_{\nu}]=0. (This is because that part of the proof only used asymptotics of 𝐘μ\mathbf{Y}_{\mu} along {0}×U\{0\}\times U, which we know because of (3.4) and (3.5).) Recall that 𝐘0=x0\mathbf{Y}_{0}=\frac{\partial}{\partial x^{0}}. Since we also have 𝐘μ=2fμ𝐧+Xμ\mathbf{Y}_{\mu}=2f_{\mu}\mathbf{n}+X_{\mu} along {0}×U\{0\}\times U, we have 𝐧=12f0(x0X0)\mathbf{n}=\frac{1}{2f_{0}}(\frac{\partial}{\partial x^{0}}-X_{0}), and consequently, we also have

𝐘i=fif0(x0X0)+Xi,\mathbf{Y}_{i}=\frac{f_{i}}{f_{0}}\left(\frac{\partial}{\partial x^{0}}-X_{0}\right)+X_{i},

for i=1,,ni=1,\ldots,n. Since [𝐘0,𝐘i]=0[\mathbf{Y}_{0},\mathbf{Y}_{i}]=0, this formula for 𝐘i\mathbf{Y}_{i} holds on all ×U\mathbb{R}\times U rather than merely along {0}×U\{0\}\times U, where f0,X0,fi,Xif_{0},X_{0},f_{i},X_{i} are all extended to ×U\mathbb{R}\times U to be independent of the x0x^{0} variable. In particular, it follows that 𝐘μ\mathbf{Y}_{\mu} is asymptotic to xμ\frac{\partial}{\partial x^{\mu}} in the sense required for application of Theorem 1.6.

We can now invoke Theorem 1.6 to see that 𝐠\mathbf{g} is flat and {𝐘0,,𝐘n}\{\mathbf{Y}_{0},\ldots,\mathbf{Y}_{n}\} is a covariant constant global orthonormal frame. In particular, 𝐘0\mathbf{Y}_{0} has constant length 1-1 everywhere in 𝐍\mathbf{N}. Or in other words, 4f02+|X0|2=1-4f_{0}^{2}+|X_{0}|^{2}=-1 on all of UU. This means that 2f012f_{0}\geq 1 on UU, and consequently, UU must be all of MM. So (M,g,π)(M,g,\pi) lies inside (𝐍,𝐠)(\mathbf{N},\mathbf{g}).

The last step is to prove that (𝐍,𝐠)(\mathbf{N},\mathbf{g}) is globally isometric to Minkowski space. We will first consider the case where MM has no boundary, and then we will explain why MM having a nontrivial weakly outer trapped boundary is impossible.

Suppose MM has no boundary. Let (M~,g~,π~)(\tilde{M},\tilde{g},\tilde{\pi}) be the universal cover of (M,g,π)(M,g,\pi), and let (𝐍~,𝐠~)(\tilde{\mathbf{N}},\tilde{\mathbf{g}}) be its Killing development with respect to the lift of (f0,X0)(f_{0},X_{0}). So 𝐍~\tilde{\mathbf{N}} is simply connected, and by what we proved above, 𝐠~\tilde{\mathbf{g}} is flat. Moreover, the lack of boundary implies that (𝐍~,𝐠~)(\tilde{\mathbf{N}},\tilde{\mathbf{g}}) is geodesically complete. (This is a nontrivial fact since 𝐠\mathbf{g} 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. (𝐍~,𝐠~)(\tilde{\mathbf{N}},\tilde{\mathbf{g}}) must be isometric to Minkowski space. In particular, it follows that (M~,g~,π~)(\tilde{M},\tilde{g},\tilde{\pi}) has only one asymptotically flat end, and we claim that this implies that the covering map Π:M~M\Pi:\tilde{M}\longrightarrow M is trivial, and the result follows. To see why the claim is true, let \mathcal{E} be an asymptotically flat end of MM, and consider the covering map from Π1()\Pi^{-1}(\mathcal{E}) to \mathcal{E} obtained by restriction of Π\Pi. Since \mathcal{E} is simply connected, Π1()\Pi^{-1}(\mathcal{E}) must be a disjoint union of isometric copies of \mathcal{E}. Since M~\tilde{M} has only one asymptotically flat end, this implies that Π\Pi is one-to-one.

We show that MM 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 “Ω\Omega” to be the part of MM bounded by M\partial M 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 (g,π)(g,\pi) is vacuum in order to conclude that each (fi,Xi)(f_{i},X_{i}) extends to a Killing field on the Killing development of (U,g)(U,g) with respect to (f0,X0)(f_{0},X_{0}). 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 (g,π)(g,\pi) is vacuum, simply knowing that (fμ,Xμ)(f_{\mu},X_{\mu}) solves (2.4) turns out to be sufficient to conclude that each (fi,Xi)(f_{i},X_{i}) extends to a Killing field on the Killing development of (U,g)(U,g) with respect to (f0,X0)(f_{0},X_{0}), 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 Ω\Omega be a compact connected manifold with boundary such that Ω=ΣinΣout\partial\Omega=\Sigma_{\mathrm{in}}\sqcup\Sigma_{\mathrm{out}} where Σin\Sigma_{\mathrm{in}} and Σout\Sigma_{\mathrm{out}} are both closed and nonempty, and let g0g_{0} be a C2C^{2} Riemannian metric on Ω\Omega (which will only serve as a fixed Riemannian background metric). Define 𝐍=[0,)×Ω\mathbf{N}=[0,\infty)\times\Omega and h0=dt2+g0h_{0}=dt^{2}+g_{0}, where tt denotes the coordinate function for the [0,)[0,\infty) factor.

Let 𝐠\mathbf{g} be a Lorentzian metric on 𝐍\mathbf{N} that is C2C^{2} in the sense that it extends to a C2C^{2} Lorentzian metric on some manifold 𝐍\mathbf{N}^{\prime} containing 𝐍\mathbf{N}, and assume that |𝐠|h0|\mathbf{g}|_{h_{0}} is uniformly bounded on 𝐍\mathbf{N}. We also assume that tt is a time function such that 𝐠(t,t)ϵ<0\mathbf{g}\left(\frac{\partial}{\partial t},\frac{\partial}{\partial t}\right)\leq-\epsilon<0 on all of 𝐍\mathbf{N} for some uniform constant ϵ\epsilon, and that each {t}×Ω\{t\}\times\Omega slice is spacelike.

Now suppose that 𝐠\mathbf{g} satisfies the null energy condition, and that for all t>0t>0, {t}×Σout\{t\}\times\Sigma_{\mathrm{out}} is a strictly outer untrapped surface (that is, θ+>0\theta^{+}>0) with respect to the “outward” future null normal pointing out of 𝐍\mathbf{N}. Then {0}×Σin\{0\}\times\Sigma_{\mathrm{in}} cannot be a weakly outer trapped surface (that is, it cannot have θ+0\theta^{+}\leq 0 on all of {0}×Σin\{0\}\times\Sigma_{\mathrm{in}}) with respect to the “outward” future null normal pointing toward the interior of 𝐍\mathbf{N}.

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 (Ω,g)(\Omega,g) be a compact connected C2C^{2} Riemannian manifold with boundary such that Ω=ΣinΣout\partial\Omega=\Sigma_{\mathrm{in}}\sqcup\Sigma_{\mathrm{out}} where Σin\Sigma_{\mathrm{in}} and Σout\Sigma_{\mathrm{out}} are both closed and nonempty. Suppose that gg has nonnegative Ricci curvature, and that Σout\Sigma_{\mathrm{out}} satisfies H>0H>0 everywhere with respect to the normal pointing out of Ω\Omega. Then Σin\Sigma_{\mathrm{in}} cannot satisfy H0H\leq 0 with respect to the normal pointing into Ω\Omega.

R. Ichida [Ich81] proved the stronger statement that if one instead assumes that HΣout0H_{\Sigma_{\mathrm{out}}}\geq 0 and HΣin0H_{\Sigma_{\mathrm{in}}}\leq 0, then (Ω,g)(\Omega,g) splits as a product.

Proof of Theorem 4.1.

By assumption, there is a larger spacetime manifold without boundary, called 𝐍\mathbf{N}^{\prime}, which contains 𝐍\mathbf{N}, and all of our causal theoretic definitions below take place within the spacetime 𝐍\mathbf{N}^{\prime}, though it will be apparent that the geometry of 𝐍\mathbf{N}^{\prime} outside 𝐍\mathbf{N} is irrelevant in the proof. We also extend the Riemannian background metric h0h_{0} on 𝐍\mathbf{N} to a complete Riemannian metric on 𝐍\mathbf{N}^{\prime}.

Our first claim is that the timelike future of {0}×Σin\{0\}\times\Sigma_{\mathrm{in}}, I+({0}×Σin)I^{+}\big{(}\{0\}\times\Sigma_{\mathrm{in}}\big{)}, intersects (0,)×Σout(0,\infty)\times\Sigma_{\mathrm{out}}. Let σ\sigma be any curve with unit speed with respect to g0g_{0} that connects Σin\Sigma_{\mathrm{in}} to Σout\Sigma_{\mathrm{out}} in Ω\Omega. Let c>0c>0 be a constant, and consider the curve σ~(s):=(cs,σ(s))\tilde{\sigma}(s):=(cs,\sigma(s)) in 𝐍\mathbf{N} connecting {0}×Σin\{0\}\times\Sigma_{\mathrm{in}} to (0,)×Σout(0,\infty)\times\Sigma_{\mathrm{out}}. Then

(4.1) 𝐠(σ~(s),σ~(s))=c2𝐠(t,t)+2c𝐠(t,σ(s))+𝐠(σ(s),σ(s))c2ϵ+2c|𝐠|h0+|𝐠|h0,\mathbf{g}(\tilde{\sigma}^{\prime}(s),\tilde{\sigma}^{\prime}(s))=c^{2}\mathbf{g}\left(\tfrac{\partial}{\partial t},\tfrac{\partial}{\partial t}\right)+2c\mathbf{g}\left(\tfrac{\partial}{\partial t},\sigma^{\prime}(s)\right)+\mathbf{g}(\sigma^{\prime}(s),\sigma^{\prime}(s))\leq-c^{2}\epsilon+2c|\mathbf{g}|_{h_{0}}+|\mathbf{g}|_{h_{0}},

where ϵ>0\epsilon>0 is the constant assumed in the hypotheses, and σ(s)\sigma^{\prime}(s) is interpreted to lie in Tσ~(s)𝐍T_{\tilde{\sigma}(s)}\mathbf{N}. Since |𝐠|h0|\mathbf{g}|_{h_{0}} is uniformly bounded, the right side of (4.1) will be negative for sufficiently large cc. Thus σ~\tilde{\sigma} is future-timelike for sufficiently large cc, proving the claim.

Let TT be the infimum over all tt such that {t}×Σout\{t\}\times\Sigma_{\mathrm{out}} intersects I+({0}×Σin)I^{+}\big{(}\{0\}\times\Sigma_{\mathrm{in}}\big{)}. So there is some point pΣoutp\in\Sigma_{\mathrm{out}} such that (T,p)I+({0}×Σin)(T,p)\in\partial I^{+}\big{(}\{0\}\times\Sigma_{\mathrm{in}}\big{)}. In fact, (T,p)(T,p) can be written as the limit of ending points of future timelike curves that begin in {0}×Σin\{0\}\times\Sigma_{\mathrm{in}} and stay inside [0,T+1]×Ω[0,T+1]\times\Omega. (If the timelike curve crosses through the [0,)×Σin[0,\infty)\times\Sigma_{\mathrm{in}} part of the boundary, then the first part of the curve up to the last time it intersects [0,)×Σin[0,\infty)\times\Sigma_{\mathrm{in}} can be replaced by an integral curve of t\frac{\partial}{\partial t}, and if the timelike curve crosses through the [0,)×Σout[0,\infty)\times\Sigma_{\mathrm{out}} part, then the last part of the timelike curve after the first time it intersects [0,)×Σout[0,\infty)\times\Sigma_{\mathrm{out}} can be discarded.) By a standard limit curve argument, there exists a limit curve γ\gamma of (the time-reversed version of) the above-described family of timelike curves such that γ\gamma is a past null geodesic starting at (T,p)(T,p) and lying in I+({0}×Σin)\partial I^{+}(\{0\}\times\Sigma_{\mathrm{in}}) such that γ\gamma either ends at {0}×Σin\{0\}\times\Sigma_{\mathrm{in}} or is inextendible. See, for example, [Min19, Theorem 2.56] or [Gal14, Proposition 3.4]. Moreover, since γ\gamma is a limit curve of curves lying in the compact set [0,T+1]×Ω[0,T+1]\times\Omega, γ\gamma itself must lie in [0,T+1]×Ω[0,T+1]\times\Omega as well. In fact, γ\gamma lies in [0,T]×Ω[0,T]\times\Omega because it begins at (T,p)(T,p) and is past null.

We claim that γ\gamma is not inextendible, and consequently, it ends at {0}×Σin\{0\}\times\Sigma_{\mathrm{in}}. To prove the claim, recall that inextendible curves must have infinite h0h_{0}-length, where h0h_{0} is the complete Riemannian background metric on the spacetime 𝐍\mathbf{N}^{\prime}. See, e.g. [Min19, Lemma 2.17]. Note that since each {t}×Ω\{t\}\times\Omega is assumed to be spacelike, in the compact space [0,T]×Ω[0,T]\times\Omega there must be a uniform constant δ>0\delta>0 with the property that 𝐠(v,v)δg0(v,v)\mathbf{g}(v,v)\geq\delta g_{0}(v,v) for all tangent vectors vT𝐍v\in T\mathbf{N} that are tangent to the Ω\Omega directions in the splitting [0,)×Ω[0,\infty)\times\Omega. Writing γ(s)=(t(s),σ(s))\gamma(s)=(t(s),\sigma(s)), the null condition says that

(4.2) 0=𝐠(γ(s),γ(s))=(dtds)2𝐠(t,t)+2(dtds)𝐠(t,σ(s))+𝐠(σ(s),σ(s))(dtds)2|𝐠|h02|dtds||𝐠|h0|σ(s)|g0+δ|σ(s)|g02,0=\mathbf{g}(\gamma^{\prime}(s),\gamma^{\prime}(s))=\left(\tfrac{dt}{ds}\right)^{2}\mathbf{g}\left(\tfrac{\partial}{\partial t},\tfrac{\partial}{\partial t}\right)+2\left(\tfrac{dt}{ds}\right)\mathbf{g}\left(\tfrac{\partial}{\partial t},\sigma^{\prime}(s)\right)+\mathbf{g}(\sigma^{\prime}(s),\sigma^{\prime}(s))\\ \geq-\left(\tfrac{dt}{ds}\right)^{2}|\mathbf{g}|_{h_{0}}-2\left|\tfrac{dt}{ds}\right||\mathbf{g}|_{h_{0}}|\sigma^{\prime}(s)|_{g_{0}}+\delta|\sigma^{\prime}(s)|^{2}_{g_{0}},

where σ(s)\sigma^{\prime}(s) is interpreted to lie in Tγ(s)𝐍T_{\gamma(s)}\mathbf{N}. Solving this quadratic inequality implies that |σ(s)|g0C|dtds||\sigma^{\prime}(s)|_{g_{0}}\leq C\left|\frac{dt}{ds}\right|, where CC depends on δ\delta and the uniform bound on |𝐠|h0|\mathbf{g}|_{h_{0}}. Therefore the h0h_{0}-length of γ\gamma can be bounded in terms of γ|dtds|\int_{\gamma}\left|\frac{dt}{ds}\right|, which is bounded by TT since tt is a time function, γ\gamma is past pointing, and γ\gamma lies in [0,T]×Ω[0,T]\times\Omega. Hence γ\gamma has finite h0h_{0}-length, establishing the claim.

Refer to caption
Figure 1. An illustration for the proof of Theorem 4.1. We first obtain a null geodesic γ\gamma from {0}×Σin\{0\}\times\Sigma_{\mathrm{in}} to (T,p){T}×Σout(T,p)\in\{T\}\times\Sigma_{\mathrm{out}}. Suppose on the contrary that θ{0}×Σin+0\theta_{\{0\}\times\Sigma_{\mathrm{in}}}^{+}\leq 0. Then the Raychaudhuri equation and the null energy condition implies θΣ1+0\theta_{\Sigma_{1}}^{+}\leq 0. On the other hand, by construction, θΣ2+>0\theta^{+}_{\Sigma_{2}}>0, and Σ2\Sigma_{2} encloses Σ1\Sigma_{1} and is tangent to Σ1\Sigma_{1} at (Tϵ,z)(T-\epsilon,z). This contradicts the strong comparison principle for θ+\theta^{+}.

With slight abuse of notation, let γ:[0,a]𝐍\gamma:[0,a]\to\mathbf{N} denote the same geodesic described above, but with parameter reversed so that it is now a future null geodesic starting at γ(0){0}×Σin\gamma(0)\in\{0\}\times\Sigma_{\mathrm{in}} and ending at the point γ(a)=(T,p){T}×Σout\gamma(a)=(T,p)\in\{T\}\times\Sigma_{\mathrm{out}}, lying entirely in I+({0}×Σin)\partial I^{+}(\{0\}\times\Sigma_{\mathrm{in}}).

By a standard argument (see, e.g. [O’N83, Lemma 50, p. 298]), γ\gamma must leave {0}×Σin\{0\}\times\Sigma_{\mathrm{in}} in an outward future null normal direction and not have any focal points in its interior (with respect to the surface {0}×Σin\{0\}\times\Sigma_{\mathrm{in}}). Moreover, the construction of TT guarantees that γ\gamma is pointing in the outward future null normal direction when it hits {T}×Σout\{T\}\times\Sigma_{\mathrm{out}} at the point (T,p)(T,p) and that there are no focal points of {T}×Σout\{T\}\times\Sigma_{\mathrm{out}} along γ|(0,a]\gamma|_{(0,a]}. (For otherwise, in either case there is a timelike curve from {T}×Σout\{T\}\times\Sigma_{\mathrm{out}} to γ(0)\gamma(0), which contradicts the minimality of TT.)

It follows that the normal exponential map on the normal bundle of {0}×Σin\{0\}\times\Sigma_{\mathrm{in}} is nonsingular along γ|[0,a)\gamma|_{[0,a)} [O’N83, Proposition 30, p. 283], and thus γ|[0,a)\gamma|_{[0,a)} is contained in a smooth null hypersurface, denoted by 𝒩1\mathcal{N}_{1}, generated by outward future null normal geodesics starting from a small neighborhood of γ(0)\gamma(0) in {0}×Σin\{0\}\times\Sigma_{\mathrm{in}}. Similarly, there is a smooth null hypersurface, denoted by 𝒩2\mathcal{N}_{2}, generated by past inward null normal geodesics starting from a neighborhood of γ(a)\gamma(a) in {T}×Σout\{T\}\times\Sigma_{\mathrm{out}}. Note that γ\gamma lies in both 𝒩1\mathcal{N}_{1} and 𝒩2\mathcal{N}_{2}.

Since the spacelike tt-level sets must intersect the null hypersurfaces 𝒩1\mathcal{N}_{1} and 𝒩2\mathcal{N}_{2} transversely, the intersections are all smooth. For sufficiently small ϵ>0\epsilon>0, let Σ2\Sigma_{2} be the smooth t=Tϵt=T-\epsilon slice of 𝒩2\mathcal{N}_{2} with θΣ2+>0\theta^{+}_{\Sigma_{2}}>0 (because of our assumption that θ{T}×Σout+>0\theta^{+}_{\{T\}\times\Sigma_{\mathrm{out}}}>0), and let Σ1\Sigma_{1} be the smooth t=Tϵt=T-\epsilon slice of 𝒩1\mathcal{N}_{1}. By construction, γ\gamma must intersect both Σ1\Sigma_{1} and Σ2\Sigma_{2} at some point (Tϵ,z)(T-\epsilon,z). Moreover, the minimality of TT guarantees that Σ2\Sigma_{2} is “outside” of Σ1\Sigma_{1} as hypersurfaces of {Tϵ}×Ω\{T-\epsilon\}\times\Omega.

Now assume that {0}×Σin\{0\}\times\Sigma_{\mathrm{in}} is weakly outer trapped, that is, θ{0}×Σin+0\theta^{+}_{\{0\}\times\Sigma_{\mathrm{in}}}\leq 0. Using a standard argument involving the Raychaudhuri equation and the null energy condition, it follows that θΣ1+0\theta^{+}_{\Sigma_{1}}\leq 0 at (Tϵ,z)(T-\epsilon,z). Since we already saw that θΣ2+>0\theta^{+}_{\Sigma_{2}}>0 at (Tϵ,z)(T-\epsilon,z) and Σ2\Sigma_{2} encloses Σ1\Sigma_{1}, this contradicts the strong comparison principle for θ+\theta^{+} (see, for example, [AG05, Proposition 3.1]). ∎

Proof of Theorem 1.7.

Let 𝐠\mathbf{g} be the Lorentzian metric and t\frac{\partial}{\partial t} 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 h0:=dt2+g0h_{0}:=dt^{2}+g_{0} where g0g_{0} is the induced Riemannian metric of 𝐠\mathbf{g} on {0}×Ω\{0\}\times\Omega. Since t\frac{\partial}{\partial t} is a global timelike Killing vector field, the metric 𝐠\mathbf{g} is invariant in tt, and thus |𝐠|h0|\mathbf{g}|_{h_{0}} is uniformly bounded on 𝐍\mathbf{N} and the tt-slices stay spacelike. The compactness of Ω\Omega implies that 𝐠(t,t)ϵ<0\mathbf{g}(\tfrac{\partial}{\partial t},\tfrac{\partial}{\partial t})\leq-\epsilon<0 for a uniform constant ϵ\epsilon on {0}×Ω\{0\}\times\Omega and hence on all of 𝐍\mathbf{N}. ∎

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.