Paper The following article is Open access

Green operators in low regularity spacetimes and quantum field theory

, , and

Published 3 August 2020 © 2020 The Author(s). Published by IOP Publishing Ltd
, , Citation G Hörmann et al 2020 Class. Quantum Grav. 37 175009 DOI 10.1088/1361-6382/ab839a

0264-9381/37/17/175009

Abstract

In this paper we develop the mathematics required in order to provide a description of the observables for quantum fields on low-regularity spacetimes. In particular we consider the case of a massless scalar field ϕ on a globally hyperbolic spacetime M with C1,1 metric g. This first entails showing that the (classical) Cauchy problem for the wave equation is well-posed for initial data and sources in Sobolev spaces and then constructing low-regularity advanced and retarded Green operators as maps between suitable function spaces. In specifying the relevant function spaces we need to control the norms of both ϕ and □gϕ in order to ensure that □gG± and G±◦□g are the identity maps on those spaces. The causal propagator G = G+G is then used to define a symplectic form ω on a normed space V(M) which is shown to be isomorphic to ker(□g). This enables one to provide a locally covariant description of the quantum fields in terms of the elements of quasi-local C*-algebras.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

This paper is concerned with developing the theory of quantum fields on low regularity spacetimes. We will follow the algebraic approach to quantisation as described in [1, 2]. In particular we will draw heavily on the detailed scheme given in the book [3]. The starting point is a smooth Lorentzian manifold (M, g) and a field equation = f where P is a normally hyperbolic differential operator P acting on a vector bundle F. In this paper we will consider the scalar operator P = □g as there are no significant additional mathematical issues in dealing with the general case. Adding lower order terms to □g would require their coefficients to be globally bounded and Lipschitz continuous (as in theorem 3.7), while going from the case of a scalar operator to one acting on a vector bundle would require a change to vector valued Sobolev spaces. In terms of the proofs, the main change would be in the proof of uniqueness. For fields which satisfy the dominant energy condition, the proof of lemma 4.1 would be essentially unchanged, cf [4, lemma 7.4.4], while in the general case one would need to follow the lines of the proof of [5, lemma 12.8]. Thus the results obtained here would also apply to the case of a massive scalar field or an Abelian gauge field.

Almost all the literature dealing with quantum field theory on a curved spacetime assumes that the background metric is smooth. However from the physical point of view it is desirable to be able to deal with a much wider class of spacetimes. For example a spacetime with some kind of interface (such as at the surface of a star) or matching (e.g. at a domain wall) will only have finite differentiability. Other examples with finite differentiability include the Oppenheimer–Snyder model of a collapsing star and spacetimes with branes or cosmic strings. In braneworld scenarios cosmological perturbations should be described by QFT on corresponding backgrounds describing the collision of the branes [6] while recent discussions on entanglement entropy, compute this via the replica method which involves putting QFTs on backgrounds with conical singularities [7]. There is thus a strong physical motivation to establish a mathematical framework to deal with quantum field theories on low regularity spacetimes. In generalising the smooth results to the low regularity setting we need to choose an appropriate class of metrics. In the present paper we have chosen to work with metrics that are C1,1. In many ways this choice is natural since it allows one to deal with spacetimes where the curvature remains finite while allowing for discontinuities in the energy–momentum tensor at, for example, an interface or the surface of a star. From the mathematical point of view C1,1 is the minimal condition which ensures existence and uniqueness of geodesics and for which the standard results from smooth causality theory go through more or less unchanged [8]. It also guarantees that the solutions to the wave equation are in ${H}_{\mathrm{loc}}^{2}\left(M\right)$ (see appendix B for details of the function spaces we use) as shown in theorem 4.7 below, which ensures we have enough regularity to define the quantisation functors we need. Although one can define solutions to the wave equation for metrics of lower regularity [9, 10] (including the case of cosmic strings) there are additional technical difficulties that need addressing, so we will return to these cases in a later paper.

The essence of the algebraic approach as outlined in [3] is to first construct the advanced and retarded Green operators for P and use these to construct the causal propagator G = G+G. Note that in order for the Green operators to be unique we require the spacetime to be globally hyperbolic. The causal propagator is then used to construct a skew symmetric bilinear map on the space of smooth functions of compact support by $\tilde {\omega }\left(\phi ,\psi \right){:=}{\langle G\left(\phi \right),\psi \rangle }_{{L}^{2}\left(M,g\right)}$. This form is degenerate but gives rise to a symplectic form ω on the quotient space $\mathcal{D}\left(M\right)/\mathrm{ker}\left(G\right)$ of the test function space. The next step in the process is to use ω to construct representations of the canonical commutation relations (CCRs) as quasi-local C*-algebras [3, theorem 4.4.11] which satisfy the Haag–Kastler axioms [11]. Each of these steps may be described in terms of a functor so that we have a functorial description of how to go to from the category of globally hyperbolic manifolds equipped with (formally self-adjoint) normally hyperbolic operators to the space of quasi-local C*-algebras whose elements describe the observables of the field (see details below). Indeed this scheme gives rise to a locally covariant quantum field in the sense of [12]. Going from the rather abstract quantisation procedure described here to the more familiar Fock space representation requires one to pick out the physically relevant states. For the smooth case a mathematically appealing criterion (which corresponds to the standard answer in Minkowski space) is the micro-local spectrum condition of Radzikowski [13]. However this is not directly applicable in the low-regularity case. We return to this point and suggest a suitable modification in the discussion section at the end.

In section 3 we establish the results we need to prove existence and uniqueness of solutions to the forward (and backward) initial value problem for the wave equation on ${\mathbb{R}}^{n+1}$ for C1,1 metrics. Rather than rework the entire theory of the wave equation for metrics of low-regularity we use a method of regularising the coefficients [14], using the smooth theory to obtain the corresponding solutions of the Cauchy problem and then using a compactness argument to show that this converges to a weak ${H}_{\mathrm{loc}}^{2}\left({\mathbb{R}}^{n+1}\right)$ solution of the original equation. This proceeds via the theory of Colombeau generalised functions [15] and is related to the work of [16] on very weak solutions. Furthermore by controlling the causal structure of the regularisation gɛ in such a way that ${J}_{\varepsilon }^{+}\left(U\right)\subset {J}^{+}\left(U\right)$ holds for any relatively compact subset U (cf appendix A) we obtain in lemma 3.5 that the forward solution u+ with zero initial data satisfies the causal support condition supp(u+) ⊂ J+(supp(f)) (cf [5, theorem 2.6.4]).

In section 4 we introduce the notion of global hyperbolicity [17] and temporal functions for non-smooth metrics [18] as well as the other results from C1,1 causality theory that we require [8]. We use the fact that even for C1,1 spacetimes the temporal function can be chosen to be smooth so we can write M as $\mathbb{R}{\times}{\Sigma}$, where Σ is a Cauchy surface, and define function spaces where we make a split between space and time. However, as far as possible we formulate our final results in a way that is independent of the choice of temporal function, so that the particular choice of space-time split is not important. The remainder of the section shows how to go from existence and uniqueness results on ${\mathbb{R}}^{n+1}$ to global results on a globally hyperbolic C1,1 spacetime. Our approach to this closely follows Ringström [5] and the causality results for C1,1 metrics [8] ensure that the existence proof remains similar to the smooth case.

The next step is to define appropriate Green operators. In the smooth case the Green operator takes (compactly supported) smooth functions to smooth functions. However in the C1,1 case it is crucial that the map is between suitable Sobolev type spaces. In fact, we find precise conditions on the regularity of the solutions and the causal support in order to define unique Green operators in globally hyperbolic spacetimes of limited differentiability. We need to control the (local) Sobolev norms of both ϕ and □gϕ in order to ensure that □gG± and G±◦□g are the identity maps on the corresponding spaces. The choice of function space is also relevant in the definition of ker(G) which is used to construct the factor space for the symplectic map ω. We end the section by considering the dependence of the construction on the choice of temporal function.

The passage from the symplectic space defined by ω to the canonical commutation relations proceeds almost identically to that of the smooth case [3] so we only sketch out the details although the analogue of [3, theorem 4.5.1] requires some work.

We end the paper with a summary of the results we have obtained and a discussion of the outstanding issues, including the choice of a Sobolev micro-local spectrum condition to single out the physical states in the low-regularity setting.

Appendix A briefly describes the basic properties of regularisation methods we use while appendix B gives a brief description of various Sobolev spaces we use in the paper.

Notation. The space of smooth functions of compact support on a manifold M will be denoted by $\mathcal{D}\left(M\right)$. A function f on an open subset $\mathcal{U}$ of ${\mathbb{R}}^{n}$ is said to be Lipschitz if there is some constant K such that for each pair of points $p,q\in \mathcal{U}$, |f(p) − f(q)| ⩽ K|pq|. We denote by Ck,1 those Ck functions where the kth derivative is a Lipschitz continuous function. A function on a smooth manifold is said to be Lipschitz or Ck,1 if it has this property upon composition with any smooth coordinate chart. We gather basic notions and results concerning Sobolev and related function spaces in appendix B.

2. The smooth setting

In this section we briefly review the results in the smooth setting. The starting point is an existence and uniqueness result for solutions to a smooth second order hyperbolic equation on ${\mathbb{R}}^{n+1}$ (see e.g. [5, theorem 8.6]). This is used to obtain a corresponding result showing the existence of a unique smooth solution uC(M) to the Cauchy initial value problem for a smooth normally hyperbolic operator P on a smooth globally hyperbolic manifold with smooth spacelike Cauchy surface Σ and normal vector field n given by

which satisfies the causal support condition

Note that the globally hyperbolic condition is essential in order to ensure that the solution is unique.

The next step is to show the existence of advanced and retarded Green operators.

Definition 2.1. Let M be a time-oriented connected Lorentzian manifold and let P be a normally hyperbolic operator. An advanced Green operator G+ is a linear map ${G}^{+}:\mathcal{D}\left(M\right)\to {C}^{\infty }\left(M\right)$ such that

  • (a)  
    $P{\circ}{G}^{+}={\mathrm{i}\mathrm{d}}_{\mathcal{D}\left(M\right)}$
  • (b)  
    ${G}^{+}{\circ}P{\vert }_{\mathcal{D}\left(M\right)}={\mathrm{i}\mathrm{d}}_{\mathcal{D}\left(M\right)}$
  • (c)  
    supp(G+ϕ) ⊂ J+(supp(ϕ)) for all $\phi \in \mathcal{D}\left(M\right)$.

A retarded Green operator G satisfies (a) and (b) but (c) is replaced by supp(Gϕ) ⊂ J(supp(ϕ)) for all $\phi \in \mathcal{D}\left(M\right)$.

Corollary 3.4.3 of [3] shows that these exist and are unique on a globally hyperbolic manifold.

The advanced and retarded Green operators are then used to define the causal propagator G := G+G which maps $\mathcal{D}\left(M\right)$ to the space of spatially compact maps ${C}_{\text{sc}}^{\infty }\left(M\right)$ i.e. the smooth maps ϕ such that there exists a compact subset KM with supp(ϕ) ⊂ J(K). If M is globally hyperbolic then one has the following exact sequence [3, theorem 3.4.7]

in particular, $\text{ker}\left(G\right)=P\left(\mathcal{D}\left(M\right)\right)$.

We want to use G to construct a symplectic vector space to which one can apply the CCR functor. We first define a skew symmetric bilinear form on $\mathcal{D}\left(M\right)$ by $\tilde {\omega }\left(\phi ,\psi \right){:=}{\langle G\left(\phi \right),\psi \rangle }_{{L}^{2}\left(M,g\right)}$. Unfortunately the bilinear form is degenerate so it fails to provide the required symplectic form. However we can rectify this by passing to the quotient space $V{:=}\mathcal{D}\left(M\right)/\text{ker}\left(G\right)$, which by the above is just $\mathcal{D}\left(M\right)/P\left(\mathcal{D}\left(M\right)\right)$. Hence $\tilde {\omega }$ induces a symplectic form ω on V. One can go on to use this to construct representations of the canonical commutation relations (CCRs) on the space of quasi-local C*-algebras [3, theorem 4.4.11] which satisfy the Haag–Kastler axioms [3, theorem 4.5.1].

3. The Cauchy problem on ${\mathbb{R}}^{n+1}$ for C1,1 metrics

In this section we establish theorem 3.7 which gives the existence, uniqueness and causal support results we need concerning solutions to the Cauchy problem for the wave equation on ${\mathbb{R}}^{n+1}$ for a C1,1 metric.

The proof follows from lemmas 3.1, 3.2, 3.3 below, which cover a slightly more general version of the Cauchy problem. The basic technique is to employ a Chruściel–Grant regularisation of the metric [19] (see details in appendix A) to obtain a family of smooth metrics ${\left({g}_{\varepsilon }\right)}_{\varepsilon \in \left(0,1\right]}$ which converge to g in the C1 topology, have uniformly bounded second derivatives on compact sets and satisfy ${J}_{\varepsilon }^{+}\left(K\right)\subset {J}^{+}\left(K\right)$ for every compact subset KM and ɛ > 0. This enables us to obtain the required solution of the wave equation by taking a suitable limit of solutions to the smooth equation as ɛ → 0 while preserving the causal support properties. Lemma 3.1 provides a detailed form of the energy estimates which proves crucial in the transition to the non-smooth setting. The causal support properties follow from lemma 3.5. Note that although lemma 3.2 shows the existence of a generalised Colombeau solution [15] to a corresponding generalised Cauchy problem the main result of this section, theorem 3.7, concerns a classical weak solution and does not require explicitly referring to the Colombeau solution used in the proof.

As a basic setup, we consider a Lorentzian metric g of signature (+, −, ..., −) on ${\mathbb{R}}^{n+1}$ with spatial components ${\left(-{h}_{ij}\right)}_{1{\leqslant}i,j{\leqslant}n}$ and the corresponding wave operator

Equation (3.1)

where all coefficients are supposed to be real-valued, smooth, and bounded in all orders of derivatives. Moreover we assume that there exist positive constants τmin, τmax, λmin, λmax such that

Equation (3.2)

Equation (3.3)

for all $\left(t,x\right)\in {\mathbb{R}}^{n+1}$ and for all $\xi \in {\mathbb{C}}^{n}$. It follows from theorem 23.2.2 in [20] or theorem 2.6 in [21] that for initial data ${u}_{0},{u}_{1}\in {H}^{\infty }\left({\mathbb{R}}^{n}\right)$ (see appendix A) and right-hand side $f\in {H}^{\infty }\left(\left(0,T\right){\times}{\mathbb{R}}^{n}\right)\subset {C}^{\infty }\left(\left[0,T\right]{\times}{\mathbb{R}}^{n}\right)$ the Cauchy problem

Equation (3.4)

has a unique solution u which belongs to $C\left(\left[0,T\right],{H}^{s+1}\left({\mathbb{R}}^{n}\right)\right)\cap {C}^{1}\left(\left[0,T\right],{H}^{s}\left({\mathbb{R}}^{n}\right)\right)\cap {H}^{s+1}\left(\left(0,T\right){\times}{\mathbb{R}}^{n}\right)$ for every $s\in \mathbb{R}$. In the following lemma we provide an explicit energy estimate with s = 1.

Lemma 3.1. Let $u\in C\left(\left[0,T\right],{H}^{2}\left({\mathbb{R}}^{n}\right)\right)\cap {C}^{1}\left(\left[0,T\right],{H}^{1}\left({\mathbb{R}}^{n}\right)\right)\cap {H}^{2}\left(\left(0,T\right){\times}{\mathbb{R}}^{n}\right)$ such that $Pu\in {L}^{2}\left(\left[0,T\right],{H}^{1}\left({\mathbb{R}}^{n}\right)\right)$ and denote tilde u0 := u(0, ⋅) and tilde u1 := ∂tu(0, ⋅). Then we have

where β = C/min(τmin, λmin) with the constant C ⩾ 0 depending only on the LT)-norms of the coefficients g, a, b and their first-order derivatives.

Proof. We show how to derive the energy estimate in terms of application of P to any function $u\in {H}^{\infty }\left({\mathbb{R}}^{n+1}\right)$: we may write

Equation (3.5)

where

and

Here we have used the notation fi,μ for ${\partial }_{{x}^{\mu }}{f}_{i}$, where Greek indices range from 0 to n and x0 := t. Considering time t as a parameter and integrating equation (3.5) over the spatial domain ${\mathbb{R}}^{n}$, we obtain

Equation (3.6)

where we have used that ${\int }_{{\mathbb{R}}^{n}}{\partial }_{{x}_{j}}{X}_{j}\left(u\left(t,x\right);t,x\right)\mathrm{d}\left({x}_{1},\dots ,\;{x}_{n}\right)=0$ for all j = 1, ..., n and for all t ∈ [0, T], since xu(t, x) belongs to ${H}^{1}\left({\mathbb{R}}^{n}\right)$ and the latter possesses ${C}_{c}^{\infty }\left({\mathbb{R}}^{n}\right)$ as a dense subspace. Using the notation gμ for the vector ${\left({g}_{\mu j}\right)}_{j=1}^{n}$ and div gμ for its divergence, we may estimate Y(u(t, x); t, x) as follows:

We note that

with Cg',a,b := max(C1, C2, C3) depending only on the LT)-norms of the coefficients g, a, b and of the first-order derivatives of g. (See appendix B for the definition of ${\tilde {H}}^{1}\left(M\right)$.) Moreover we have

where Cmin := min(τmin, λmin) and Cmax := max(τmax, λmax). We observe that $\vert 2\;\text{Re}\left(Pu{\partial }_{t}\overline{u}\right)\left(t,x\right)\vert {\leqslant}{\vert Pu\left(t,x\right)\vert }^{2}+{\vert {\partial }_{t}u\left(t,x\right)\vert }^{2}.$ Using the fundamental theorem of calculus we may write

and integration over the spatial variables yields

Equation (3.7)

where we have deliberately multiplied by CminCmax. Integrating equation (3.6) with respect to the time variable from 0 to t we obtain

We will from now on write tilde u0 := u(0, ⋅) and tilde u1 := ∂tu(0, ⋅). Upon adding inequality (3.7) we get

Employing the bounds ${C}_{\mathrm{min}}\left({{\Vert}{\partial }_{t}u\left(t,\cdot \right){\Vert}}_{{L}^{2}\left({\mathbb{R}}^{n}\right)}^{2}+{{\Vert}\nabla u\left(t,\cdot \right){\Vert}}_{{L}^{2}\left({\mathbb{R}}^{n}\right)}^{2}\right){\leqslant}{\int }_{{\mathbb{R}}^{n}}\mathcal{T}\left(u\left(t,\cdot \right);t,\cdot \right)\mathrm{d}V$ and ${\int }_{{\mathbb{R}}^{n}}\mathcal{T}\left(u\left(0,\cdot \right);0,\cdot \right)\mathrm{d}V{\leqslant}{C}_{\mathrm{max}}\left({{\Vert}{\tilde{u}}_{1}{\Vert}}_{{L}^{2}\left({\mathbb{R}}^{n}\right)}^{2}+{{\Vert}\nabla {\tilde{u}}_{0}{\Vert}}_{{L}^{2}\left({\mathbb{R}}^{n}\right)}^{2}\right)$ as well as the other estimates and rearranging terms, we arrive at

Equation (3.8)

where

Gronwall's inequality [22, chapter XVIII, sections 5 and 2.2] then yields the a priori energy estimate

where $\beta {:=}{C}_{\mathrm{min}}^{-1}\left(1+\mathrm{max}\left({C}_{1},{C}_{2},{C}_{3}\right)\right)$.

To obtain similar estimates for spatial derivatives $\nabla u={\left({\partial }_{{x}_{1}}u,\dots ,\;{\partial }_{{x}_{n}}u\right)}^{\mathsf{T}}$ of u, we first divide the equation by g00:

Differentiating this equation with respect to xk yields

Equation (3.9)

Multiplying by g00 we may write this as

Equation (3.10)

where we have brought all terms which do not contain spatial derivatives ${\partial }_{{x}_{k}}u$ to the right-hand side. Multiplying by ${\partial }_{t}{\partial }_{{x}_{k}}\overline{u}$ and summing over 1 ⩽ kn, we get

Equation (3.11)

where the whole expression depends on (t, x) and X and Y are defined exactly as before. The additional term Z collects all terms from (3.10) which are of lower order with respect to ${\partial }_{{x}_{k}}u$ (that is, they contain spatial derivatives of order 2 at most). We have

and it is easy to see that Z can be estimated as follows:

and thus

where ${{\Vert}\nabla u\left(t,x\right){\Vert}}^{2}{:=}{\sum }_{k=1}^{n}{\vert {\partial }_{{x}_{k}}u\left(t,x\right)\vert }^{2}$ and Cg',a' depends only on the LT)-norms of the coefficients g, a and their first-order derivatives. Moreover, we have

and therefore

We may now proceed as in the proof for the basic energy estimate (3.8), following the steps from (3.6) to (3.8). Upon integrating equation (3.11) over the spatial domain ${\mathbb{R}}^{n}$, we get

and by employing the estimates for Y, Z, and F, we arrive at an energy estimate for ∇u (recall that ${\int }_{{\mathbb{R}}^{n}}\vert Y\left({\partial }_{{x}_{k}}u;t,\cdot \right)\vert \mathrm{d}V{\leqslant}{C}_{{g}^{\prime },a,b}{{\Vert}{\partial }_{{x}_{k}}u\left(t,\cdot \right){\Vert}}_{{\tilde {H}}^{1}\left({\mathbb{R}}^{n}\right)}^{2}$):

which we may also write as

where Cg',a',b' contains LT)-norms of at most first-order derivatives of the coefficients g, a, b. Applying Gronwall's inequality then results in an energy estimate for the spatial derivative vector ∇u,

Equation (3.12)

where ${\tilde {\beta }}_{1}{:=}{C}_{\mathrm{min}}^{-1}{C}_{{g}^{\prime },{a}^{\prime },{b}^{\prime }}$. Employing the basic energy estimate (3.8) to estimate ${{\Vert}u\left(s,\cdot \right){\Vert}}_{{L}^{2}\left({\mathbb{R}}^{n}\right)}^{2}$ and ${{\Vert}{\partial }_{s}u\left(s,\cdot \right){\Vert}}_{{L}^{2}\left({\mathbb{R}}^{n}\right)}^{2}$ in terms of the restrictions tilde u0, tilde u1 and Pu, we arrive at

To prepare for the extension to the Colombeau solutions, we proceed by discussing higher-order energy estimates. Applying the operator ${\partial }_{{x}_{l}}$ to equation (3.9) and afterwards multiplying by g00 we essentially get a system of the form P(∇2u) = Q(2)2u) + R(2)2Pu), where Q2 is a PDO of order 3, containing time derivatives of at most order 1, and R is a purely spatial PDO of order 2. Here we have used the notation

Thus the terms produced by Q(2)2u) are either lower-order terms in ∇2u or terms with less than two spatial derivatives of u. In the first case, they can be dealt with just as the generic lower-order terms appearing in P(∇2u). In the second case, they can be interpreted as source terms on the right-hand side, just as the term R(2)2Pu). More generally, we obtain

Equation (3.13)

where Q(r) is a PDO of order r + 1, containing time derivatives of at most order 1, and R(r) is a purely spatial PDO of order r. Moreover, the coefficients of Q(r) and R(r) depend on spatial derivatives of the metric g of at most order r. It is important to note that all principal coefficients of Q(r) depend only on gμν, aμ, b and on derivatives gμν,ρ (0 ⩽ ρn). In particular, they can be viewed as lower-order terms of an operator Pr, which has the same principal symbol as P. The energy estimate following from (3.13) is of the form

Equation (3.14)

where βr depends only on gμν, aμ, b and on derivatives gμν,ρ.

Summing up, we obtain energy estimates for ∇ru and ∂tru for $r\in {\mathbb{N}}_{0}$. Equation (3.1) itself then immediately provides an estimate for ${\partial }_{t}^{2}u$ as well. Similarly, equation (3.9) allows us to estimate ${\partial }_{t}^{2}{\partial }_{{x}_{k}}u$ in terms of (mixed) derivatives of order ⩽ 1 in the time variable. More generally, equation (3.13) directly provides estimates for ${\partial }_{t}^{2}{\nabla }^{r}u$ ($r\in \mathbb{N}$) in terms of (mixed) derivatives of order ⩽ 1 in the time variable. To get estimates for mixed derivatives ${\partial }_{t}^{m}{\nabla }^{r}u$ of order m ⩾ 3 in the time variable, we may apply the operator ${\partial }_{t}^{l}$ to equation (3.13), yielding

Equation (3.15)

where ${\tilde {Q}}^{\left(r,l\right)}$ is a PDO of order r + l with time derivatives only up to order l − 1 and ${\tilde {R}}^{\left(r,l\right)}$ is a PDO of order r + l − 1 with time derivatives up to order l − 1. The first term in $P\left({\nabla }^{r}{\partial }_{t}^{l}u\right)$ is ${g}_{00}{\nabla }^{r}{\partial }_{t}^{l+2}u$. Thus, starting with l = 1, we immediately obtain energy estimates for all mixed derivatives of the form ${\nabla }^{r}{\partial }_{t}^{3}u$. In the next step, we can show energy estimates for ${\nabla }^{r}{\partial }_{t}^{4}u$, and proceed iteratively to get energy estimates for all mixed derivatives ${\nabla }^{r}{\partial }_{t}^{m}u$ as well. These estimates are direct consequences of (3.15), once estimates for the terms with time derivatives of lower order have been established (there is no need for a Gronwall argument). However, a perhaps more elegant viewpoint is that for any $r,l\in \mathbb{N}$, equation (3.15) also implies ${\tilde {H}}^{1}$-estimates for ${\nabla }^{r}{\partial }_{t}^{l}u$ in terms of initial data and source terms via the energy estimate (3.8) for the operator P. Again, it is important to note that the principal coefficients of ${\tilde {Q}}^{\left(r,l\right)}$ and ${\tilde {R}}^{\left(r,l\right)}$ only depend on gμν, aj, b and on first derivatives gμν,ρ. Higher-order derivatives of the coefficients will only enter in the lower order terms and thus do not show up in the exponent after applying Gronwall's inequality. In the following lemma we refer to Colombeau theoretic notions which are summarised in appendix A.

Lemma 3.2. We consider the initial value problem (3.4) with generalised functions as coefficients and data and assume that

  • (a)  
    The components of g as well as the lower-order coefficients aμ and b belong to ${\mathcal{G}}_{{L}^{\infty }}\left({\mathbb{R}}^{n+1}\right)$,
  • (b)  
    There exist constants τmin > 0 and λmin > 0 such that ${g}_{00}^{\varepsilon }\left(t,x\right){\geqslant}{\tau }_{\mathrm{min}}{\left(\mathrm{log}\left(1/\varepsilon \right)\right)}^{-1}$ and ${\sum }_{i,j}{h}_{ij}^{\varepsilon }\left(t,x\right)\overline{{\xi }_{i}}{\xi }_{j}{\geqslant}{\lambda }_{\mathrm{min}}{\left(\mathrm{log}\left(1/\varepsilon \right)\right)}^{-1}{\vert \xi \vert }^{2}$ for all $\left(t,x\right)\in {\mathbb{R}}^{n+1},\xi \in {\mathbb{C}}^{n}$ and for all ɛ < ɛ0,
  • (c)  
    All coefficients gμν, aμ, b as well as all derivatives gμν,ρ are of L-log-type.

Then, given initial data ${u}_{0},{u}_{1}\in {\mathcal{G}}_{{L}^{2}}\left({\mathbb{R}}^{n}\right)$ and right-hand side $f\in {\mathcal{G}}_{{L}^{2}}\left({{\Omega}}_{T}\right)$, the Cauchy problem (3.4) has a unique solution $u\in {\mathcal{G}}_{{L}^{2}}\left({{\Omega}}_{T}\right)$.

Proof. We fix a symmetric representative of g and representatives of all lower-order coefficients, initial data, and right-hand side. As noted above, we have smooth solutions ${u}_{\varepsilon }\in {C}^{\infty }\left(\left[0,T\right],{H}^{\infty }\left({\mathbb{R}}^{n}\right)\right)$ to the corresponding classical initial value problems for each ɛ < ɛ0. To show moderateness, we apply lemma 3.1 and obtain

where $\left({\mathrm{e}}^{{\beta }^{\varepsilon }}\right)$ is moderate thanks to the log-type condition on the coefficients and their first derivatives as well as the positivity condition (b). Thus the estimate shows that $\left({{\Vert}{u}_{\varepsilon }{\Vert}}_{{H}^{1}\left({{\Omega}}_{T}\right)}\right)$ and $\left({{\Vert}\nabla {u}_{\varepsilon }{\Vert}}_{{H}^{1}\left({{\Omega}}_{T}\right)}\right)$ are moderate. For r ⩾ 2, the higher-order estimates (3.14) for the spatial derivative vector ∇ruɛ then imply that $\left({{\Vert}{\nabla }^{r}{u}_{\varepsilon }{\Vert}}_{{H}^{1}\left({{\Omega}}_{T}\right)}\right)$ is also moderate (for any $r\in \mathbb{N}$), since all principal coefficients of the operators on the right-hand side of equation (3.13) only depend on $\left({g}_{\mu \nu }^{\varepsilon }\right),\left({a}_{\mu }^{\varepsilon }\right),\left({b}^{\varepsilon }\right)$ and on first derivatives $\left({g}_{\mu \nu ,\rho }^{\varepsilon }\right)$, all of which are of L-log-type.

An iterative application of the higher-order estimates for spatial derivatives (3.14),

starting with r ⩾ 2 then establishes moderateness of ${{\Vert}{\nabla }^{r}{u}_{\varepsilon }{\Vert}}_{{H}^{1}\left({{\Omega}}_{T}\right)}$ for all $r\in {\mathbb{N}}_{0}$, since $1/{C}_{\mathrm{min}}^{\varepsilon }$ is of logarithmic growth in ɛ (and thus moderate), ${C}_{\mathrm{max}}^{\varepsilon }$, ${C}_{{g}^{\prime }}^{\varepsilon }$, ${C}_{{g}_{\varepsilon }^{\left(r\right)},{a}_{\varepsilon }^{\left(r\right)},{b}_{\varepsilon }^{\left(r\right)}}$ are of moderate growth and ${\beta }_{r}^{\varepsilon }$ is of L-log-type (where ${\beta }_{r}^{\varepsilon }$ depends on ${C}_{\mathrm{min}}^{\varepsilon }$ and (first derivatives of) gɛ, aɛ, and bɛ). Finally, the corresponding energy estimates for mixed derivatives following from equation (3.15) yield moderateness of ${{\Vert}{\nabla }^{r}{\partial }_{t}^{l}{u}_{\varepsilon }{\Vert}}_{{H}^{1}\left({{\Omega}}_{T}\right)}$ as well. Note that only the principal coefficients will be exponentiated in the Gronwall inequality. Thus the important observation in all these higher-order estimates is that the principal coefficients on both sides of equation (3.15) only depend on ${g}_{\mu \nu }^{\varepsilon },{a}_{\mu }^{\varepsilon },{b}^{\varepsilon }$ and on derivatives ${g}_{\mu \nu ,\rho }^{\varepsilon }$, all of which are of L-log-type.

In total this implies that for all $r,l\in {\mathbb{N}}_{0}$ there exists $m\in {\mathbb{N}}_{0}$ such that ${{\Vert}{\nabla }^{r}{\partial }_{t}^{l}{u}_{\varepsilon }{\Vert}}_{{L}^{2}\left({{\Omega}}_{T}\right)}^{2}=O\left({\varepsilon }^{-m}\right)$ as ɛ → 0 and thus $\left[{\left({u}_{\varepsilon }\right)}_{\varepsilon { >}0}\right]\in {\mathcal{G}}_{{L}^{2}}\left({{\Omega}}_{T}\right)$. To show uniqueness of the generalised solution in ${\mathcal{G}}_{{L}^{2}}\left({{\Omega}}_{T}\right)$, we assume negligible initial data $\left({\tilde {{u}_{0}}}_{\varepsilon }\right),\left({\tilde {{u}_{1}}}_{\varepsilon }\right)\in {\mathcal{N}}_{{L}^{2}}\left({\mathbb{R}}^{n}\right)$ and right-hand side $\left({\tilde {f}}_{\varepsilon }\right)\in {\mathcal{N}}_{{L}^{2}}\left({{\Omega}}_{T}\right)$. Then the same energy estimates we used for moderateness yield negligibility of the solution $\left[{\left({\tilde {u}}_{\varepsilon }\right)}_{\varepsilon { >}0}\right]$. □

In the next lemma we want to identify conditions on the coefficients and data of a low-regularity Cauchy problem (3.4) such that the corresponding generalised Cauchy problem, obtained via regularisation, has a unique solution $u\in {\mathcal{G}}_{{L}^{2}}\left({{\Omega}}_{T}\right)$ and, moreover, this solution admits a distributional shadow.

Lemma 3.3. We consider the Cauchy problem (3.4) where all coefficients are supposed to be globally bounded and Lipschitz continuous on ${\mathbb{R}}^{n+1}$ and satisfy estimates according to (3.2) and (3.3). Then, given initial data $\left({u}_{0},{u}_{1}\right)\in {H}^{2}\left({\mathbb{R}}^{n}\right){\times}{H}^{1}\left({\mathbb{R}}^{n}\right)$ and right-hand side $f\in {L}^{2}\left(\left[0,T\right],{H}^{1}\left({\mathbb{R}}^{n}\right)\right)$, we consider the corresponding generalised Cauchy problem obtained via convolution regularisation of all coefficients and data. Then the unique generalised solution according to lemma 3.2 has a distributional shadow $\tilde {u}\in {C}^{0}\left(\left[0,T\right],{H}^{2}\left({\mathbb{R}}^{n}\right)\right)\cap {C}^{1}\left(\left[0,T\right],{H}^{1}\left({\mathbb{R}}^{n}\right)\right)\cap {H}^{2}\left({{\Omega}}_{T}\right)$ which is also the unique weak solution.

Proof. We note that standard convolution regularizations of the coefficients and data provide ɛ-parametrized families of smooth approximations which guarantee the assumptions in lemma 3.2. Thus, we obtain a unique generalised solution. We aim at showing that any representative (uɛ) of the solution is a Cauchy net. To this end we apply the variant of the energy estimate in lemma 3.1, implicit in the proof (see also (3.8)), with order of spatial Sobolev norms reduced by one (except for the initial data): applying the operator Pɛ to the difference ${u}_{\varepsilon }-{u}_{\tilde {\varepsilon }}$, we obtain

Equation (3.16)

where βɛ is bounded uniformly in ɛ thanks to the hypotheses on the non-smooth coefficients. We may write

Considering the regularity of the initial data and right-hand side, it is easy to see that (uɛ) is a Cauchy net, if for all η > 0 there exists ɛ0 such that ${{\Vert}\left({P}_{\varepsilon }-{P}_{\tilde {\varepsilon }}\right){u}_{\tilde {\varepsilon }}{\Vert}}_{{L}^{2}\left({{\Omega}}_{T}\right)}{< }\eta $ for all $\tilde {\varepsilon }{< }\varepsilon {\leqslant}{\varepsilon }_{0}$. We have

Since all coefficients are bounded and Lipschitz continuous, we obtain converge of their regularising families in the L-norm, hence it suffices to show that ${{\Vert}{u}_{\tilde {\varepsilon }}{\Vert}}_{{H}^{2}\left({{\Omega}}_{T}\right)}$ is bounded uniformly in ɛ. First, observe that the energy estimate (3.8) implies that ${{\Vert}{u}_{\tilde {\varepsilon }}{\Vert}}_{{H}^{1}\left({{\Omega}}_{T}\right)}=O\left(1\right)$ as ɛ → 0. The energy estimate for ∇u, inequality (3.12),

then implies that ${{\Vert}{u}_{\tilde {\varepsilon }}{\Vert}}_{{H}^{2}\left({{\Omega}}_{T}\right)}=O\left(1\right)$ as ɛ → 0 as well. Going back to (3.16), we thus have shown that (uɛ) is indeed a Cauchy net in the norm ${L}^{\infty }\left(\left[0,T\right],{H}^{1}\left({\mathbb{R}}^{n}\right)\right)$ and (∂tuɛ) is a Cauchy net in the norm ${L}^{\infty }\left(\left[0,T\right],{L}^{2}\left({\mathbb{R}}^{n}\right)\right)$. Hence the unique generalised solution $u\in {\mathcal{G}}_{{L}^{2}}\left({{\Omega}}_{T}\right)$ has a distributional limit $\tilde {u}\in C\left(\left[0,T\right],{H}^{1}\left({\mathbb{R}}^{n}\right)\right)\cap {C}^{1}\left(\left[0,T\right],{L}^{2}\left({\mathbb{R}}^{n}\right)\right)$. However, we can deduce even better regularity from the properties of the net (uɛ). For any t ∈ [0, T] and any $\varphi \in \mathcal{D}\left({\mathbb{R}}^{n}\right)$, we have

since ${{\Vert}{u}_{\varepsilon }\left(t,\cdot \right){\Vert}}_{{H}^{2}\left({\mathbb{R}}^{n}\right)}=O\left(1\right)$ as ɛ → 0 and we already have uniform convergence of (uɛ(t, ⋅)) in ${H}^{1}\left({\mathbb{R}}^{n}\right)$ by the Cauchy net estimate and hence as a distribution. It follows that ${\partial }_{{x}_{j}}{\partial }_{{x}_{k}}{u}_{0}\left(t,\cdot \right)\in {L}^{2}\left({\mathbb{R}}^{n}\right)$ and therefore ${u}_{0}\in {C}^{0}\left(\left[0,T\right],{H}^{2}\left({\mathbb{R}}^{n}\right)\right)$. Moreover we have

and thus $\tilde {u}\in {C}^{1}\left(\left[0,T\right],{H}^{1}\left({\mathbb{R}}^{n}\right)\right)$. Similarly we can show that ${\partial }_{{x}_{\mu }}{\partial }_{{x}_{k}}\tilde {u}\in {L}^{2}\left({{\Omega}}_{T}\right)$ for 0 ⩽ μn and 1 ⩽ kn. In addition, multiplying the equation Pɛuɛ = fɛ by ${\left({g}_{00}^{\varepsilon }\right)}^{-1}$, it is easy to see that ${{\Vert}{\partial }_{t}^{2}{u}_{\varepsilon }{\Vert}}_{{L}^{2}\left({{\Omega}}_{T}\right)}=O\left(1\right)$ as ɛ → 0, implying that ${\partial }_{t}^{2}\tilde {u}\in {L}^{2}\left({{\Omega}}_{T}\right)$ as well. Summing up, we have obtained a distributional shadow $\tilde {u}$ of the generalised solution with

In the last part of the proof we show that the distributional shadow $\tilde {u}$ of the generalised solution is the unique weak solution to the Cauchy problem. The proof follows the line of arguments in the proof of [14, corollary 4.6]. First note that both $\tilde {u}$ and ${\partial }_{t}\tilde {u}$ are continuous and thus, by construction of $\tilde {u}$, the initial conditions are satisfied. The Cauchy net estimate (3.16) implies that ${u}_{\varepsilon }\to \tilde {u}$ as ɛ → 0 in the norm H1T). Our aim is to prove that $P\;\tilde {u}=f$ in a suitable weak sense. Due to boundedness and Lipschitz continuity of the coefficients we immediately obtain L2-convergence of all first-order terms and H1-convergence of all zero-order terms:

We claim that

weakly* in L2T). Since there exists C > 0 such that ${{\Vert}{u}_{\varepsilon }{\Vert}}_{{H}^{2}\left({{\Omega}}_{T}\right)}{\leqslant}C$ for all ɛ < ɛ0, we know that the set $\left\{{u}_{1/n}\vert \;n\in \mathbb{N}\right\}$ is bounded in H2T). By the weak compactness theorem, this implies that there exists a weakly* convergent subsequence ${\left({u}_{1/{n}_{k}}\right)}_{k\in \mathbb{N}}$ with limit $\tilde {v}\in {H}^{2}\left({{\Omega}}_{T}\right)$ (cf [23, theorem 6.64]) and indeed $\tilde {v}=\tilde {u}$ since we already know from the Cauchy net estimate that ${u}_{1/{n}_{k}}\to \tilde {u}$ in H1T) as k. Therefore ${\partial }_{\mu }{\partial }_{\nu }{u}_{1/{n}_{k}}$ converges weakly* to ${\partial }_{\mu }{\partial }_{\nu }\tilde {u}$ in L2T) and we have for any φL2T):

as k. To see this, observe that the first term goes to zero because the coefficients of ${P}_{1/{n}_{k}}$ converge to those of $\tilde {P}$ in LT) as k and ${\left({u}_{1/{n}_{k}}\right)}_{k\in \mathbb{N}}$ is bounded in H2T); the second term vanishes as well in the limit k since ${u}_{1/{n}_{k}}$ converges to $\tilde {u}$ in H2T) as k. We provide the explicit calculation for the g00-term (the others can be treated similarly):

This shows that for any φL2T),

and thus $\tilde {u}$ is indeed a weak solution of the initial value problem.

To show uniqueness of the weak solution, we suppose that there exists another solution $\tilde {w}\in {H}^{2}\left({{\Omega}}_{T}\right)\cap C\left(\left[0,T\right],{H}^{1}\left({\mathbb{R}}^{n}\right)\right)\cap {C}^{1}\left(\left[0,T\right],{L}^{2}\left({\mathbb{R}}^{n}\right)\right)$ such that $P\;\tilde {w}=f$ and $\left(\tilde {w},{\partial }_{t}\tilde {w}\right){\vert }_{t=0}=\left({u}_{0},{u}_{1}\right)$. We may regularise this solution so that ${w}_{\varepsilon }\to \tilde {w}$ as ɛ → 0 in H2T) and (wɛ, ∂twɛ)|t=0 → (u0, u1) in ${H}^{1}\left({\mathbb{R}}^{n}\right){\times}{L}^{2}\left({\mathbb{R}}^{n}\right)$. The following estimate then shows that ${P}_{\varepsilon }{w}_{\varepsilon }\to P\;\tilde {w}=f$ in L2T) as ɛ → 0:

Here we have only used the H2T)-convergence of wɛ to $\tilde {w}$ as ɛ → 0. Denoting by ${\left({u}_{\varepsilon }\right)}_{\varepsilon }$ a representative of the generalised solution and applying the basic energy estimate (3.8) to the difference uɛwɛ then yields

Letting ɛ → 0 we obtain $\tilde {u}=\tilde {w}$ in $C\left(\left[0,T\right],{H}^{1}\left({\mathbb{R}}^{n}\right)\right)\cap {C}^{1}\left(\left[0,T\right],{L}^{2}\left({\mathbb{R}}^{n}\right)\right)$. □

Remark 3.4. The required conditions for the existence of a distributional shadow (and weak solution) are weaker than those that would be required in a similar result based on transforming the equation (3.1) into a first-order system as in [24], since the lower-order coefficients of this system would contain derivatives of the principal coefficients of equation (3.1) and therefore ${g}_{\mu \nu }\in {C}^{1,1}\left({\mathbb{R}}^{n+1}\right)$ would be necessary (instead of mere Lipschitz continuity). However, for a wave equation derived from the Laplace–Beltrami operator of a Lorentzian metric g,

where ${{\Gamma}}_{\mu \nu }^{\rho }=\frac{1}{2}{g}^{\rho \sigma }\left({g}_{\sigma \mu ,\nu }+{g}_{\sigma \nu ,\mu }-{g}_{\mu \nu ,\sigma }\right)$, the lower-order coefficients contain derivatives of the metric and thus the metric has to be C1,1 anyway in order to obtain a distributional shadow of the generalised solution.

Lemma 3.5. Consider the Cauchy problem (3.4) where all coefficients are globally bounded and Lipschitz continuous on ${\mathbb{R}}^{n+1}$ and satisfy (3.2) and (3.3). Then, given vanishing initial data and right-hand side $f\in {L}^{2}\left(\left[0,T\right]\right),{H}^{1}\left({\mathbb{R}}^{n}\right)\left.\right)$ with supp(f) ⊂ J+(K) for some compact set K, the unique distributional weak solution ū+ of the advanced problem given by lemma 3.3 satisfies the causal support condition

Proof. Let ${u}_{\varepsilon }^{+}$ be the unique solution of the corresponding advanced problem for the regularised metric gɛ where the right-hand side f is not necessarily smooth. Note that upon taking K slightly larger, we may assume that $\mathrm{supp}\left(f\right)\subset {J}_{\varepsilon }^{+}\left(K\right)$ for ɛ sufficiently small.

By lemmas 3.1 and 3.2 we can obtain ${u}_{\varepsilon }^{+}$ as a limit α → 0 of ${u}_{\varepsilon ,\alpha }^{+}$ where ${\left({u}_{\varepsilon ,\alpha }^{+}\right)}_{\alpha { >}0}$ is a Colombeau representative of the unique generalised solution with fixed smooth gɛ-coefficients and right-hand side being the class of a convolution regularisation ${\left({f}_{\alpha }\right)}_{\alpha { >}0}$. By the smooth theory we have for every ɛ and every α that

Equation (3.17)

At fixed ɛ and as α → 0 we have in terms of monotonically decreasing sets ⋂α>0supp(fα) = supp(f), i.e. supp(fα) ↘ supp(f). By closedness of the causal relation [3, lemma A.5.5] we obtain that

Equation (3.18)

Let ψ be a test function such that $\mathrm{supp}\left(\psi \right)\cap {J}_{\varepsilon }^{+}\left(\mathrm{supp}\left(f\right)\right)=\varnothing $. By (3.17) and (3.18) there exists some α0 > 0 such that $\mathrm{supp}\left(\psi \right)\cap {J}_{\varepsilon }^{+}\left(\mathrm{supp}\left({f}_{\alpha }\right)\right)=\varnothing $ for all α < α0. Therefore $\langle {u}_{\varepsilon ,\alpha }^{+},\psi \rangle =0$ for every α < α0; taking the limit α → 0 we obtain $\langle {u}_{\varepsilon }^{+},\psi \rangle =0$ and therefore

Now let ψ have support disjoint from J+(supp(f)). Then, by the causal properties of the Chruściel–Grant regularisation (cf proposition A.1), we have that supp(ψ) is also disjoint from ${J}_{\varepsilon }^{+}\left(\mathrm{supp}\left(f\right)\right)$ for every ɛ > 0.

Therefore, $\langle {u}_{\varepsilon }^{+},\psi \rangle =0$ for every such ψ and for every ɛ > 0, showing that $\mathrm{supp}\left({u}^{+}\right)\subseteq {J}_{\varepsilon }^{+}\left(\mathrm{supp}\left(f\right)\right)$ for every ɛ > 0, hence supp(ū+) ⊂ J+(supp(f)). □

Remark 3.6. More generally for non-zero initial data one has

The proof of this follows from the above result by recasting the smoothed version of the problem as an equivalent inhomogeneous problem with zero initial data and a right-hand side incorporating u0 and u1.

Theorem 3.7. Consider the Cauchy problem (3.4) where all coefficients are globally bounded and Lipschitz continuous on ${\mathbb{R}}^{n+1}$ and satisfy (3.2) and (3.3). Then, given initial data $\left({u}_{0},{u}_{1}\right)\in {H}^{2}\left({\mathbb{R}}^{n}\right){\times}{H}^{1}\left({\mathbb{R}}^{n}\right)$ and $f\in {L}^{2}\left(\left[0,T\right]\right),{H}^{1}\left({\mathbb{R}}^{n}\right)$, there exists a unique distributional weak solution $u\in C\left(\left[0,T\right],{H}^{2}\left({\mathbb{R}}^{n}\right)\right)\;\cap {C}^{1}\left(\left[0,T\right],{H}^{1}\left({\mathbb{R}}^{n}\right)\right)\cap {H}^{2}\left({{\Omega}}_{T}\right)$. Furthermore u satisfies the causal support condition

Proof. The existence and uniqueness of the weak solution follow from lemma 3.3 and the causal support condition follows from lemma 3.5 together with remark 3.6 applied to both the past and future. □

4. The Cauchy problem for C1,1 globally hyperbolic spacetimes

4.1. Causality results for C1,1 spacetimes

In this paper we will be considering solutions of the wave equation on orientable spacetimes (M, g) endowed with a C1,1 metric. Note that although the metric is only C1,1 we will always assume that the manifold has a smooth structure. The concept of global hyperbolicity (for smooth metrics) was introduced by Leray [25] as a condition to ensure the existence of unique solutions to hyperbolic equations and in particular the Cauchy problem for the wave equation is well-posed for smooth globally hyperbolic spacetimes [3, theorem 3.2.11 p 84ff]. For our situation it is therefore natural to consider globally hyperbolic spacetimes with C1,1 metrics. Global hyperbolicity is the strongest of the conditions on (M, g) in the causal hierarchy of spacetimes [26] and recently there has been considerable interest in looking at the causal properties of low-regularity spacetimes [8, 18]. It was shown explicitly by Chrusćiel [27] that essentially all of causality theory for smooth spacetimes goes through to the C2 case. However in the proofs of these results an important role is played by the Gauss lemma and the existence of totally normal (convex) neighbourhoods whose existence is not at all obvious in the C1,1 case. However it was shown in [8, 28, 29] that such neighbourhoods do exist and the exponential map gives a local lipeomorphism. This enables essentially the whole of the results of standard smooth causality theory to go through to the C1,1 case (see [8, 29] for details).

For spacetimes with a C2 metric there are four equivalent notions of global hyperbolicity (see for example [26, section 3.11 p 340ff.]). These are:

  • (a)  
    Compactness of the causal diamonds and causality4,
  • (b)  
    Compactness of the space of causal curves connecting two points and causality [25],
  • (c)  
    Existence of a Cauchy hypersurface,
  • (d)  
    The metric splitting of the spacetime.

For C1,1 spacetimes we will adopt the first definition. However for non-totally imprisoned [31] C0 spacetimes (and hence in particular for globally hyperbolic C1,1 spacetimes) these four definitions remain equivalent [17]. See also [18, theorem 2.45] for a more general notion formulated in terms of closed cone structures.

In our constructions below we will make use of time functions and temporal functions. A time function is a function that is strictly increasing along every causal curve while a temporal function has the additional property that its gradient is everywhere past-directed and timelike. It is shown by Minguzzi [18 theorem 2.30] (see also [32]) that for a stably causal closed cone structure (and hence in particular for a C1,1 globally hyperbolic spacetime) there exists a smooth temporal function $t:M\to \mathbb{R}$. Furthermore every globally hyperbolic closed cone structure is the domain of dependence of a stable Cauchy surface (see definition below) Σ, so that M = D(Σ) and that M is the topological product $\mathbb{R}{\times}{\Sigma}$ where the first projection is t and the level surfaces Στ = {xM : t(x) = τ} are diffeomorphic to Σ [18, theorem 2.42]. So that although the metric is only C1,1 the topological splitting remains smooth.

In the case of a smooth metric Bernal and Sánchez [33] show that given a smooth spacelike Cauchy hypersurface Σ there exists a smooth temporal function t such that Σ = t−1(0). However in the case of a non-smooth metric the temporal function they construct will not be smooth. To generalise the results of [33] to the non-smooth case we need the concept of a stable Cauchy hypersurface introduced by Minguzzi in [18]. These are Cauchy hypersurfaces which are also Cauchy hypersurfaces for some metric g' ≻ g with strictly wider lightcones than g.

Bernard and Suhr [34, corollary 2.4] show that a smooth spacelike Cauchy hypersurface is a stable Cauchy hypersurface and that furthermore one can construct a smooth temporal function such that Σ = t−1(0) [34, theorem 1]. A full discussion of this issue is given in the paper by Minguzzi [35]. The approach in [35] is complementary to that in [34] and consists of using topological arguments to show that the causal cones can be widened while preserving the Cauchy property of the hypersurface. One may then use the methods of Bernal and Sánchez [33] to construct a smooth time function with Σ = t−1(0) which as shown in [18] is a smooth temporal function for the original spacetime (M, g). See [35, theorem 2.22] for details. Indeed given two smooth spacelike Cauchy hypersurfaces Σ0 and Σ1 with Σ1J+0)\Σ0 one can find a smooth temporal function t that interpolates between them so that Σ0t−1(0) and Σ1t−1(1) [35, theorem 2.23].

4.2. Existence and uniqueness

In this section we extend the results of section 3 to a globally hyperbolic C1,1 spacetime (M, g). The main result is theorem 4.7 which establishes the existence and uniqueness of ${H}_{\mathrm{loc}}^{2}\left(M\right)$ solutions to the wave equation for a globally hyperbolic C1,1 spacetime. We start by obtaining an energy inequality which we use to establish uniqueness and the causal support properties of solutions to the wave equation.

Lemma 4.1. (energy inequality). [4, lemma 7.4.4]. Let (M, g) be an (n + 1)-dimensional C1,1 globally hyperbolic spacetime with Σ a smooth spacelike n-dimensional Cauchy surface and t a smooth temporal function such that Σ = Σ0 := t−1(0). Let UM be an open set with compact closure and let U+ := UJ+(Σ) be such that ∂UŪ+ is achronal. Then if u is a (weak) ${H}_{\mathrm{loc}}^{2}\left(M\right)$ solution of □gu = f where $f\in {L}_{\mathrm{loc}}^{2}\left(M\right)$ then

Equation (4.1)

where ${U}_{\tau }=\left\{q\in U:0{\leqslant}t\left(q\right){\leqslant}\tau \right\}$.

Proof. To establish the energy inequality we follow Hawking and Ellis by applying the divergence theorem to an enhanced energy–momentum tensor [4, lemma 7.4.4]. Let

be the energy momentum tensor of the scalar field. Then ${\left(1\right){T}}_{\alpha \beta }$ has vanishing divergence and satisfies the dominant energy condition. We now follow [4] and modify this by adding on the term

to obtain

which still satisfies the dominant energy condition. Let ξα = ∇αt, then we obtain the required inequality by applying the divergence theorem to Sαβξα over the region ${U}_{\tau }=\left\{q\in U:0{\leqslant}t\left(q\right){\leqslant}\tau \right\}$. In order to do this we require that div(Sαβξα) should be integrable with respect to the volume form νg, and this is guaranteed by the compactness of Ū and the fact that our solution is in ${H}_{\mathrm{loc}}^{2}\left(M\right)$. In fact it is enough that the weak solutions have two derivatives in ${L}_{\mathrm{loc}}^{2}\left(M,g\right)$ if the metric and the timelike vector field are in the space C0,1 (see [36]). The boundary of Uτ consists of three parts; the level surface ΣτŪ+, the level surface Σ0Ū+ and the remainder which we denote $\mathcal{H}$. Because of the dominant energy condition and the fact that ∂UŪ+ is achronal, the contribution to the surface integral from $\mathcal{H}$ is positive. We therefore obtain the following inequality:

Equation (4.2)

where νg is the volume form on Uτ given by g, and μτ is the volume form induced by g on Στ.

We now define an energy type integral

Then on Ū this is equivalent [37] to the restricted Sobolev norm (B.3)

Note that since the solution u is in ${H}_{\mathrm{loc}}^{2}\left(M\right)$ we have well-defined traces in ${\tilde {H}}^{1}\left({\Sigma}\right)$. In terms of the energy norm we may write (4.2) in the form

Repeated application of the Cauchy–Schwarz inequality and □gu = f then gives [37]

Equation (4.3)

which on applying Gronwall's inequality gives

In terms of the Sobolev type norms this gives

We now use the energy inequality (4.1) to prove uniqueness of the solution as well as the causal support properties of the solution to the Cauchy problem.

Proposition 4.2. (uniqueness).Let (M, g) be a (connected, oriented, time oriented,) globally hyperbolic (n + 1)-dimensional Lorentzian manifold with C1,1 metric and Σ a smooth n-dimensional spacelike Cauchy surface. Let u be a (weak) ${H}_{\mathrm{loc}}^{2}\left(M\right)$ solution of

with $f\in {H}_{\mathrm{loc}}^{1}\left(M\right)$, which satisfies the initial conditions

with (u0, u1) ∈ H2(Σ) × H1(Σ). Then u is unique.

Proof. Let qM and without loss of generality suppose that qI+(Σ). Then since our spacetime is globally hyperbolic we may find a p such that qI(p) ∩ I+(Σ) := U where by C1,1 causality theory U has compact closure [8]. Now suppose there exist two solutions u and tilde u to the above initial value problem. Then applying lemma 4.1 to hat u := utilde u over the region U+ gives

Hence ${{\Vert}\hat{u}{\Vert}}_{{\tilde {H}}^{1}\left({{\Sigma}}_{\tau }\cap {U}^{+}\right)}=0$ so that hat u and ∇hat u vanish in U+. Since q is arbitrary the solution is unique in D+(Σ). A similar result applies to D(Σ), so we have uniqueness in the whole of M = D(Σ). □

Proposition 4.3. (causal support). Let (M, g) be a connected, oriented, time oriented, globally hyperbolic (n + 1)-dimensional Lorentzian manifold with C1,1 metric, Σ0 a smooth spacelike n-dimensional Cauchy surface and t a smooth temporal function such that t−1(0) = Σ0. Let $u\in {C}^{0}\left(\mathbb{R},{H}^{2}\left({{\Sigma}}_{t}\right)\right)\cap {C}^{1}\left(\mathbb{R},{H}^{1}\left({{\Sigma}}_{t}\right)\right)\cap {H}_{\mathrm{loc}}^{2}\left(M\right)$ be a (weak) solution to the initial value problem

Then $\mathrm{supp}\left(u\right)\subset J\left(\mathrm{supp}\left({u}_{0}\right)\cup \mathrm{supp}\left({u}_{1}\right)\cup \mathrm{supp}\left(f\right)\right)$.

Proof. We prove the result for J+. A similar proof holds for J. Let $V={J}^{+}\left(\mathrm{supp}\left({u}_{0}\right)\cup \mathrm{supp}\left({u}_{1}\right)\cup \mathrm{supp}\left(f\right)\right)$ and suppose qM\V. Then we may find a point pD+0) such that qI(p) ∩ I+0) and u and ∇u vanish on J(p) ∩ Σ0 and f vanishes on J(p) ∩ J+0). Now let U = I(p) ∩ I+0) and apply (4.3) on this region to obtain

But by the choice of p the right-hand side vanishes, so u must also vanish in the region U with τ in the range 0 ⩽ τt(p). Hence u vanishes on a neighbourhood of q. Since q was arbitrary, u vanishes on M\V which proves the result. □

To establish existence on M, we need the following two lemmas from Ringström [5]. In both cases the proof given in [5] for the smooth case goes through to that of a C1,1 metric unchanged.

Lemma 4.4. (Ringström [5, lemma 12.5]). Let (M, g) be an (n + 1)-dimensional Lorentzian manifold with C1,1 metric and Σ a smooth spacelike n-dimensional submanifold. If pS there is a chart (U, x) with pU and x = (x0, x1, ..., xn) such that qU ∩ Σ if and only if qU and x0(q) = 0. Furthermore we may choose x so that $\frac{\partial }{\partial {x}^{0}}$ is the future directed unit normal to Σ for q ∈ Σ ∩ U.

If we fix epsilon > 0 and let ${g}_{\mu \nu }{:=}g\left(\frac{\partial }{\partial {x}^{\mu }},\frac{\partial }{\partial {x}^{\nu }}\right)$, then we can assume U to be such that |g0i| ⩽ epsilon, i = 1, ..., n on U. If we let a = g00(p) and b > 0 be such that gij(p) regarded as a positive definite matrix is bounded below by b (i.e. gij(p)ξiξj > b|ξ|2) then we may assume that g00(q) < a/2 and gij(q) regarded as a positive definite matrix is bounded below by b/2 for qU.

Lemma 4.5. (Ringström [5, lemma 12.16]). Let (M, g) be a (connected, oriented, time oriented,) globally hyperbolic (n + 1)-dimensional Lorentzian manifold with C1,1 metric and Σ a smooth spacelike n-dimensional Cauchy surface. Let t be a (smooth) temporal function (as given by [34, 35]) with ${t}^{-1}\left(0\right)={{\Sigma}}_{{t}_{0}}$. If $p\in {{\Sigma}}_{{t}_{0}}$ there is an epsilon > 0 and open neighbourhoods U, W of p such that

  • (a)  
    The closure of W is compact and contained in U;
  • (b)  
    If qW and τ ∈ [t0epsilon, t0 + epsilon], then J+τ) ∩ J(q) is compact and contained in U;
  • (c)  
    There is a chart (U, ϕ) with ϕ = (x0, ..., xn) and x0 = t, such that there exist a, b > 0 with g00(q) < − a and gij(q)ξiξjijξiξj for qU;
  • (d)  
    For any compact KU there is a C1,1 matrix valued function h on ${\mathbb{R}}^{n+1}$ such that hμν = gμ,νϕ−1 on ϕ−1(K) and such that there are positive constants a1, b1 c1 with h00 ⩽ −a1, hijξiξjb1δijξiξj and |hμν| ⩽ c1 on all of ${\mathbb{R}}^{n+1}$.

Note that although the proof is identical to that in [5] it relies on the C1,1 causality results of [8] and [34, theorem 1]. Note also that in point (d) above the matrix valued function is only C1,1 rather than smooth as it is in [5, lemma 12.16].

We are now in a position to establish existence.

Proposition 4.6. (existence for compactly supported source and initial data). Let (M, g) be a time oriented (n + 1)-dimensional Lorentzian manifold with C1,1 metric and Σ a smooth spacelike n-dimensional hypersurface. Let t be a smooth temporal function with t−1(0) = Σ and let n be the future directed timelike unit normal to Σ. Given initial data $\left({u}_{0},{u}_{1}\right)\in {H}_{\mathrm{comp}}^{2}\left({\Sigma}\right){\times}{H}_{\mathrm{comp}}^{1}\left({\Sigma}\right)$ and source $f\in {H}_{\mathrm{comp}}^{1}\left(M\right)$, then there exists a (weak) solution $u\in {C}^{0}\left(\mathbb{R},{H}^{2}\left({{\Sigma}}_{t}\right)\right)\cap {C}^{1}\left(\mathbb{R},{H}^{1}\left({{\Sigma}}_{t}\right)\right)\cap {H}_{\mathrm{loc}}^{2}\left(M\right)$ to the initial value problem

Proof. We again closely follow Ringström [5, theorem 12.17]. Let K1 ⊂ Σ be a compact set such that supp(u0) ∪ supp(u1) ⊂ K1 and K2M a compact set such that supp(f) ⊂ K2. Let t1 > 0 and define ${R}_{{t}_{1}}$ to be the set of q such that 0 ⩽ t(q) ⩽ t1. Then ${R}_{{t}_{1}}$ is closed and ${K}_{3}={K}_{2}\cap {R}_{{t}_{1}}$ is compact. The union of I+(p) for pI(Σ) is an open cover of K1K3, so there is a finite number of points p1, ..., p such that the I+(pi) are a finite subcover of K1K3. Note that the set $F={\bigcup }_{i=1}^{\ell }{J}^{+}\left({p}_{i}\right)\cap {J}^{-}\left({{\Sigma}}_{{t}_{1}}\right)$ is compact and that if there is a solution in ${R}_{{t}_{1}}$ it has to be zero in ${R}_{{t}_{1}}{\backslash}F$ by proposition 4.3. We now show that there is a solution in the compact set ${R}_{{t}_{1}}\cap F$.

Let Fτ := F ∩ Στ. Let 0 ⩽ τ < t1 and assume we have a solution in the function space specified in the proposition up to time τ, i.e. on Rτ or for every Rs with 0 ⩽ s < τ. We now extend the solution to the future of τ. For every pFτ there are neighbourhoods Up, Wp and epsilonp with the properties of lemma 4.5. By compactness there is a finite number of points ${\tilde {p}}_{i},\dots ,\;{\tilde {p}}_{N}$ such that the ${W}_{{\tilde {p}}_{i}}$ cover Fτ. Let $0{< }{\epsilon}{\leqslant}\mathrm{min}\left\{{{\epsilon}}_{{\tilde {p}}_{1}},\dots ,\;{{\epsilon}}_{{\tilde {p}}_{N}}\right\}$ be such that

Equation (4.4)

for all s ∈ [τepsilon, τ + epsilon]. Now let s1 ∈ [τepsilon, τ] be such that there is a solution, in the function space specified in the proposition up to and including s1 and let pFs for any s ∈ [s1, τ + epsilon]. Then ${K}_{p}{:=}{J}^{-}\left(p\right)\cap {J}^{+}\left({{\Sigma}}_{{s}_{1}}\right)$ is compact and contained in one of the charts, say $\left({U}_{{\tilde {p}}_{k}},\phi \right)$. Let $\chi \in {C}_{0}^{\infty }\left({U}_{{\tilde {p}}_{k}}\right)$ be such that χ(q) = 1 for all qKp. Then we use our solution up to time s1 to define new initial data on ${{\Sigma}}_{{s}_{1}}$ given by ${\tilde{u}}_{0}{:=}\left(\chi u\right){\vert }_{{{\Sigma}}_{{s}_{1}}}\in {H}^{2}\left({{\Sigma}}_{{s}_{1}}\right)$ and ${\tilde{u}}_{1}{:=}\left(\chi {\nabla }_{n}u\right){\vert }_{{S}_{{s}_{1}}}\in {H}^{1}\left({{\Sigma}}_{{s}_{1}}\right)$ and source $\tilde {f}{:=}\chi f$. These all have their support within ${U}_{{\tilde {p}}_{k}}$ so we may use the chart $\left({U}_{{\tilde {p}}_{k}},\phi \right)$ to regard these as data and source on the whole of ${\mathbb{R}}^{n}$ and ${\mathbb{R}}^{n+1}$ respectively. We may also extend the Lorentz metric gμνϕ−1 to a Lorentz matrix-valued function hμν on the whole of ${\mathbb{R}}^{n+1}$ which coincides with gμνϕ−1 on Kp. We may therefore regard the tildered version as an initial value problem on ${\mathbb{R}}^{n+1}$. The third condition of lemma 4.5 and the fact that the solution is in ${C}^{0}\left(\mathbb{R},{H}^{2}\left({{\Sigma}}_{t}\right)\right)\cap {C}^{1}\left(\mathbb{R},{H}^{1}\left({{\Sigma}}_{t}\right)\right)\cap {H}_{\mathrm{loc}}^{2}\left(M\right)$ ensures that we may apply lemma 3.3 to obtain a solution on ${\mathbb{R}}^{n+1}$ which on $\phi \left({U}_{{\tilde {p}}_{k}}\right)$ may be transferred back to give a solution on Kp. In the region ${V}_{p}{:=}{I}^{-}\left(p\right)\cap {J}^{+}\left({{\Sigma}}_{{s}_{1}}\right)$ we define u to be this solution. In the region VpVq then uniqueness ensures that the two potential solutions coincide. We now define O1 to be the union of the Vp for pFs, s ∈ [s1, τ + epsilon] then the above construction defines a unique solution in O1. Note that the interior of O1 contains Fs for all s ∈ (s1, τ + epsilon). Now define O2 to be the set of points for which s1t(q) < τ + epsilon and for which qF. We want to define the solution to be zero in this set, however we need to check that there is no contradiction for points in both O1 and O2. If qO2O1 with t(q) > s1 then both u and ∇u vanish at ${J}^{-}\left(q\right)\cap {S}_{{s}_{1}}$ and f vanishes in ${J}^{-}\left(q\right)\cap {J}^{+}\left({{\Sigma}}_{{s}_{1}}\right)$. Furthermore there is an r such that qVrO1. So by uniqueness the solution defined on O1 has to vanish for a sufficiently small neighbourhood at q.

In summary we have shown that if there exists a solution for all s < τ or up to time τ in the required function space, we get a solution in the same space on the larger region Rτ+epsilon for some epsilon > 0. Let $\mathcal{A}$ be the set of s ∈ [0, ) such that there is a solution up to time s. Taking τ = 0 in the above we have a solution on Repsilon so $\mathcal{A}$ is not empty. We have also shown that if $\tau \in \mathcal{A}$ then a solution exists for [0, τ + epsilon), so that for any τ > 0 we may find an open interval containing τ in which a solution exists. Thus $\mathcal{A}$ is open in the relative topology of [0, ). Finally we note that by definition $\mathcal{A}$ is also closed in [0, ) because it contains its limit points. Then this set is open, closed and non-empty, so it must be the whole of [0, ) and we have a solution for all future times. By an analogous argument interchanging past and future we also have a solution for all past times and hence on the whole of M. □

Theorem 4.7. (global existence and uniqueness). Let (M, g) be a connected, oriented, time oriented (n + 1)-dimensional Lorentzian globally hyperbolic manifold with C1,1 metric and Σ a smooth spacelike n-dimensional Cauchy hypersurface. Let t be a smooth temporal function with t−1(0) = Σ and let n be the future directed timelike unit normal to Σ.

Given initial data (u0, u1) ∈ H2(Σ) × H1(Σ) and source $f\in {C}^{0}\left(\mathbb{R},{H}^{1}\left({{\Sigma}}_{t}\right)\right)\cap {H}_{\mathrm{loc}}^{1}\left(M\right)$ then there exists a unique (weak) solution $u\in {C}^{0}\left(\mathbb{R},{H}^{2}\left({{\Sigma}}_{t}\right)\right)\cap {C}^{1}\mathbb{R},{H}^{1}\left({{\Sigma}}_{t}\right)\cap {H}_{\mathrm{loc}}^{2}\left(M\right)$ to the initial value problem

Moreover $\mathrm{supp}\left(u\right)\subset J\left(\mathrm{supp}\left({u}_{0}\right)\cup \mathrm{supp}\left({u}_{1}\right)\cup \mathrm{supp}\left(f\right)\right)$.

Proof. Let p be any point to the future of Σ. Then Kp = J(p) ∩ J+(Σ) is a compact set. Let $\chi \in {C}_{0}^{\infty }\left(M\right)$ be such that χ(q) = 1 for all qKp. Now define f' = χf, u0' = χu0 and u1' = χu1. Then by proposition 4.6 there is a unique solution u' to the primed initial value problem. Now set u = u' in Vp := I(p) ∩ J+(Σ). Similarly given some other point $\tilde {p}$ to the future of Σ there is a unique solution tilde u' to the initial value problem in ${V}_{\tilde {p}}{:=}{I}^{-}\left(\tilde {p}\right)\cap {J}^{+}\left({\Sigma}\right)$ and we may set u = tilde u' in ${V}_{\tilde {p}}$. If we now consider some further point $q\in {V}_{p}\cap {V}_{\tilde {p}}$ then by uniqueness the two potential solutions, given by u' and tilde u' respectively, agree in ${V}_{p}\cap {V}_{\tilde {p}}$, so there is no contradiction in setting u = u' in Vp and u = tilde u' in ${V}_{\tilde {p}}$ since u' = tilde u' in the intersection. Thus we have a solution on the whole of D+(Σ). A similar argument gives us a solution on D(Σ) and hence on the whole of M. The solution is unique by proposition (4.2) and satisfies the causal support condition by proposition (4.3).

We also want to show that the initial value problem is well-posed in the sense of corollary 4.11 below. For solutions in ${H}_{\mathrm{loc}}^{1}\left(M\right)$ this follows immediately from the energy estimate (4.1). But well-posedness in ${H}_{\mathrm{loc}}^{2}\left(M\right)$ requires a higher order estimate which we now establish.

Lemma 4.8. For each compact subset KM there exists a δ > 0 with the following property: if (u0, u1) ∈ H2(Σ) × H1(Σ) with supp(uj) ⊂ K ∩ Σ (j = 1, 2) and fH1(M) with supp(f) ⊂ K, then the (weak) solution $u\in {C}^{0}\left(\mathbb{R},{H}^{2}\left({{\Sigma}}_{t}\right)\right)\cap {C}^{1}\left(\mathbb{R},{H}^{1}\left({{\Sigma}}_{t}\right)\right)\cap {H}_{\mathrm{loc}}^{2}\left(M\right)$ of □gu = f with initial data (u0, u1) satisfies the energy inequality

Equation (4.5)

Proof. We use a similar approach to that in the existence proof given in [3, theorem 3.2.11].

For every pK there are neighbourhoods Up, Wp and epsilonp > 0 with the properties of lemma 4.5. By compactness there is a finite number of points p1, ..., pN such that the corresponding ${W}_{{p}_{j}}$ cover K. Now let ${\left\{{\chi }_{j}\right\}}_{j=1}^{N}$ be a partition of unity of K subordinate to the ${W}_{{p}_{j}}$.

We now define

so that

We also define

Let uj be the (weak) solution of the initial value problem

We will employ (an implicit choice of a temporal function in terms of) the diffeomorphism $M\cong \mathbb{R}{\times}{\Sigma}$ and slightly abuse notation from now on by considering all functions u, uj etc to be defined already on products I × Σ, where I is some open interval, thus suppressing the transfers of functions via restrictions of the underlying global diffeomorphism.

By point two of lemma 4.5 there exists an ${{\epsilon}}_{{p}_{j}}{ >}0$ such that

Equation (4.6)

Let $\delta =\mathrm{min}\left\{{{\epsilon}}_{{p}_{1}},\dots ,\;{{\epsilon}}_{{p}_{N}}\right\}$. Given the solutions uj of the local problem we may extend them by zero on all of (−2δ, 2δ) × Σ and sum them to give our unique solution

Since by (4.6) each of the uj lie entirely within some chart $\left({U}_{{p}_{j}},{\phi }_{{p}_{j}}\right)$ we may regard the initial value problem as one on $I{\times}{\mathbb{R}}^{n}$ where I is an interval chosen sufficiently large such that the images of (−2δ, 2δ) × Σ under all the ${\phi }_{{p}_{j}}$ are contained in $I{\times}{\mathbb{R}}^{n}$.

Then the third condition of lemma 4.5 enables us to transfer the basic energy estimate according to lemma 3.1 from $I{\times}{\mathbb{R}}^{n}$ to ones for uj on UjM to give

Now u = juj so that

where we may replace ${{\Vert}f{\Vert}}_{{L}^{2}\left(\left[0,\delta \right],{H}^{1}\left({\Sigma}\right)\right)}$ by the larger value ${{\Vert}f{\Vert}}_{{H}^{1}\left(M\right)}$, since $f\in {H}_{\mathrm{comp}}^{1}\left(M\right)$. □

We remark that in the above proof (and formulation of the result) we have replaced the norm ${{\Vert}f{\Vert}}_{{L}^{2}\left(\left[0,\delta \right],{H}^{1}\left({\Sigma}\right)\right)}$ by ${{\Vert}f{\Vert}}_{{H}^{1}\left(M\right)}$, which is valid for $f\in {H}_{\mathrm{comp}}^{1}\left(M\right)$, to avoid the need to specify a particular choice of a temporal function.

Proposition 4.9. (global higher energy estimates). Let (M, g) be a time oriented (n + 1)-dimensional Lorentzian manifold with C1,1 metric and Σ a smooth spacelike n-dimensional hypersurface. Let t be a smooth temporal function with Σ = t−1(0) and let n be the future directed timelike unit normal to Σ.

Given initial data $\left({u}_{0},{u}_{1}\right)\in {H}_{\mathrm{comp}}^{2}\left({\Sigma}\right){\times}{H}_{\mathrm{comp}}^{1}\left({\Sigma}\right)$ and source $f\in {H}_{\mathrm{comp}}^{1}\left(M\right)$, then the (weak) solution $u\in {C}^{0}\left(\mathbb{R},{H}^{2}\left({{\Sigma}}_{t}\right)\right)\cap {C}^{1}\left(\mathbb{R},{H}^{1}\left({{\Sigma}}_{t}\right)\right)\cap {H}_{\mathrm{loc}}^{2}\left(M\right)$ satisfies

for any interval [0, T].

Proof. We first use lemma 4.8 to obtain an estimate for the data ${\hat{u}}_{0}{:=}u{\vert }_{{{\Sigma}}_{\delta }}$ and ${\hat{u}}_{1}{:=}{\nabla }_{n}u{\vert }_{{{\Sigma}}_{\delta }}$ induced by u on Σδ. It follows from (4.5) and the fact that $u\in {C}^{0}\left(\mathbb{R},{H}^{2}\left({{\Sigma}}_{t}\right)\right)\cap {C}^{1}\left(\mathbb{R},{H}^{1}\left({{\Sigma}}_{t}\right)\right)\cap {H}_{\mathrm{loc}}^{2}\left(M\right)$ that

Now applying lemma 4.8 to the initial surface Σδ we obtain a $\hat{\delta }{ >}\delta $ such that

Combining the two energy inequalities on [0, δ] and $\left[\delta ,\hat{\delta }\right]$ we have

This shows that we may extend the energy inequality from [0, δ] to the larger time interval $\left[0,\hat{\delta }\right]$ by repeatedly applying lemma 4.8.

Similarly, for every τ ∈ [0, T] we may find a δ(τ) > 0 such that lemma 4.8 applies to the time interval (τδ(τ), τ + δ(τ)) with initial data given on Στ. By compactness, finitely many intervals (τkδ(τk), τk + δ(τk)) (k = 1, ..., m) cover [0, T] and the energy inequalities on these may be combined. □

We now use the above proposition to obtain a spacetime energy inequality.

Proposition 4.10. (higher order spacetime energy estimates). Let (M, g) be a time oriented (n + 1)-dimensional Lorentzian manifold with C1,1 metric and Σ a smooth spacelike n-dimensional hypersurface. Given initial data $\left({u}_{0},{u}_{1}\right)\in {H}_{\mathrm{comp}}^{2}\left({\Sigma}\right){\times}{H}_{\mathrm{comp}}^{1}\left({\Sigma}\right)$ and source $f\in {H}_{\mathrm{comp}}^{1}\left(M\right)$, then the (weak) solution u satisfies the energy-inequality

Equation (4.7)

for any compact KM.

Proof. Without loss of generality we may assume that K ⊂ [0, T] × Σ. Due to the regularity of $u\in {C}^{0}\left(\mathbb{R},{H}^{2}\left({{\Sigma}}_{t}\right)\right)\cap {C}^{1}\left(\mathbb{R},{H}^{1}\left({{\Sigma}}_{t}\right)\right)$ we have control of the second order spatial derivatives, the second order mixed derivatives and the lower order terms

Equation (4.8)

Equation (4.9)

In order to obtain the required estimate we also need to control ∂ttu in the L2(K) norm. Using □gu = f we have

From (4.8) and (4.9), the regularities of f and g, we obtain an L2 estimate for ∂ttu,

Combining the above with proposition 4.9 completes the proof. □

Equation (4.7) implies the following result.

Corollary 4.11. The solution to the Cauchy problem described in theorem 4.7 is well-posed in the sense that the solution map

is continuous in the topologies coming from the respective Sobolev spaces (see appendix B).

5. Green operators for C1,1 spacetimes

In this section we will define Green operators for □g on globally hyperbolic manifolds M with C1,1 metrics. We will show existence and uniqueness of Green operators via the existence of solutions to the wave equation with appropriate regularity and causal support. We define below the notion of generalised hyperbolicity which will give us the required conditions in this situation.

A spacetime (M, g) is said to satisfy the condition of generalised hyperbolicity if the inhomogeneous wave equation for zero Cauchy data is well-posed and causal. The precise choice of function spaces in the definition of well-posedness depends upon the regularity of the metric. In our case we give the following definition:

Definition 5.1. (generalised hyperbolicity). A C1,1 spacetime (M, g) satisfies the condition of generalised hyperbolicity if the following conditions hold

  • (a)  
    Existence: for every $f\in {H}_{\mathrm{comp}}^{1}\left(M\right)$ there exists a unique future solution ${u}^{+}\in {H}_{\mathrm{loc}}^{2}\left(M,g\right)$ such that
    which satisfies the causal support condition supp(u+) ⊂ J+(supp(f)). Note that this condition implies that on a Cauchy surface to the past of supp(f) one must have zero initial data. However the particular choice of such a Cauchy surface makes no difference to the solution (see proof of theorem 5.2 below for more details).
  • (b)  
    Uniqueness of causal solution: for every $f\in {H}_{\mathrm{comp}}^{1}\left(M\right)$ there exists a unique past solution ${u}^{-}\in {H}_{\mathrm{loc}}^{2}\left(M,g\right)$ such that
    which satisfies supp(u) ⊂ J(supp(f)) where we can choose any Cauchy surface to the future of supp(f) and solve the Cauchy problem going back in time.
  • (c)  
    Well-posedness: we require that the maps fu+ and fu are continuous maps from ${H}_{\mathrm{comp}}^{1}\left(M\right)\to {H}_{\mathrm{loc}}^{2}\left(M\right)$ in the sense that ${{\Vert}{u}^{{\pm}}{\Vert}}_{{H}^{2}\left(K\right)}$C±${{\Vert}f{\Vert}}_{{H}^{1}\left(M\right)}$ holds for any compact subset KM with suitable constants C± > 0.

Theorem 5.2. Let (M, g) be a globally hyperbolic time oriented (n + 1)-dimensional Lorentzian manifold with C1,1 metric and Σ a smooth spacelike n-dimensional Cauchy surface. Then (M, g) satisfies the condition of generalised hyperbolicity.

Proof. Theorem 4.7 shows that a globally hyperbolic C1,1 spacetime satisfies the condition of generalised hyperbolicity to the future by considering the forward initial value problem

where $f\in {H}_{\mathrm{comp}}^{1}\left(M\right)$ and Σ+ is a smooth spacelike Cauchy hypersurface such that J+(supp(f)) ∩ Σ+ = .

Note: if we were to choose some other smooth spacelike Cauchy hypersurface ${\tilde {{\Sigma}}}_{+}$, which also satisfies ${J}^{+}\left(\mathrm{supp}\left(f\right)\right)\cap {\tilde {{\Sigma}}}_{+}=\varnothing $, then the corresponding solution is the same, since the divergence theorem arguments used in lemma 4.1 apply and yield that the solution must vanish in the region between Σ+ and ${\tilde {{\Sigma}}}_{+}$.

Similarly, theorem 4.7 shows it satisfies the condition of generalised hyperbolicity to the past by considering the backwards initial value problem

where $f\in {H}_{\mathrm{comp}}^{1}\left(M\right)$ and Σ is a smooth spacelike Cauchy hypersurface such that J(supp(f)) ∩ Σ = . Again the solution is independent of the choice of Cauchy surface as long as it satisfies the causal support condition. □

5.1. Green operators

The definition of the Green operators in the non-smooth setting will require us to choose suitable spaces of functions as domain and range (see theorem 5.9). We therefore define the following spaces:

Equation (5.1)

Remark 5.3. Note that none of the spaces defined above depend upon the choice of background metric used in the definition of the Sobolev spaces.

Definition 5.4. A linear map

satisfying the properties

  • (a)  
    ${\square }_{g}{G}^{+}={\text{id}}_{{H}_{\mathrm{comp}}^{1}\left(M\right)}$,
  • (b)  
    ${G}^{+}{\square }_{g}{\vert }_{{V}_{0}}={\text{id}}_{{V}_{0}}$,
  • (c)  
    supp(G+(f)) ⊂ J+(supp(f)) for all $f\in {H}_{\mathrm{comp}}^{1}\left(M\right)$,

is called an advanced Green operator for □g. A retarded Green operator G is defined similarly.

Remark 5.5. 

  • (a)  
    Clearly, the regularity condition in the definition of the space V0 was chosen to guarantee that □gf, given fV0 belongs to the domain ${H}_{\mathrm{comp}}^{1}\left(M\right)$ of the Green operators.
  • (b)  
    In several proofs below we will show the identities in properties (a) and (b) to hold weakly, i.e., when evaluated on test functions in $\mathcal{D}\left(M\right)$. However it can be shown using the results of Hörmander [38, theorem 1.25] that if two such Sobolev functions have the same effect on test functions then they are actually equal as Sobolev functions.
  • (c)  
    The function spaces ${H}_{\mathrm{loc}}^{2}\left(M\right)$ and ${H}_{\mathrm{comp}}^{1}\left(M\right)$ used as target space and domain for the Green operators are in perfect accordance with the theory of so-called regular fundamental solutions for hyperbolic operators with constant coefficients5 (cf [39, section 12.5]) as we sketch briefly in the following: let M be (n + 1)-dimensional Minkowski space so that □ has the symbol p(τ, ξ) = τ2 − |ξ|2, which is hyperbolic with respect to the directional vectors (±1, 0) and produces the temperate weight function $\tilde {p}\left(\tau ,\xi \right){:=}\sqrt{{\sum }_{\vert \alpha \vert {\geqslant}0}{\vert {\partial }^{\alpha }p\left(\tau ,\xi \right)\vert }^{2}}=\sqrt{{\left({\tau }^{2}-{\xi }^{2}\right)}^{2}+4\left({\tau }^{2}+{\xi }^{2}+2\right)}{\geqslant}\sqrt{1+{\tau }^{2}+{\xi }^{2}}=:{w}_{1}\left(\tau ,\xi \right)$. The unique fundamental solution E± with support in the half space where ±t ⩾ 0 belongs to ${B}_{\infty ,\tilde {p}}^{\mathrm{loc}}$, i.e., for every test function ϕ the Fourier transform $\mathcal{F}\left(\phi {E}_{{\pm}}\right)$ times $\tilde {p}$ is measurable and bounded. The advanced and retarded Green operators are then given by convolution G±f = E±*f for every $f\in {H}_{\mathrm{comp}}^{1}\left({\mathbb{R}}^{n+1}\right)={\mathcal{E}}^{\prime }\left({\mathbb{R}}^{n+1}\right)\cap {B}_{2,{w}_{1}}$, where ${B}_{2,{w}_{1}}=\left\{u\in {\mathcal{S}}^{\prime }\vert {w}_{1}\cdot \mathcal{F}u\in {L}^{2}\right\}={H}^{1}\left({\mathbb{R}}^{n+1}\right)$. Finally, we may apply [39, theorem 12.5.3 or theorem 10.1.24] to obtain ${G}^{{\pm}}f\in {B}_{2,\tilde {p}\cdot {w}_{1}}^{\mathrm{loc}}\subseteq {B}_{2,{w}_{1}^{2}}^{\mathrm{loc}}={H}_{\mathrm{loc}}^{2}\left({\mathbb{R}}^{n+1}\right)$, since $\tilde {p}\left(\tau ,\xi \right){w}_{1}\left(\tau ,\xi \right){\geqslant}{w}_{1}^{2}\left(\tau ,\xi \right)=1+{\tau }^{2}+{\xi }^{2}$.

We next show that the advanced and retarded Green operators (if they exist) are adjoints of one another. To do this we use the following lemma.

Lemma 5.6. Given $\chi ,\varphi \in {H}_{\mathrm{loc}}^{2}\left(M,g\right)$ and supp(χ) ∩ supp(φ) compact. Then, we have

Equation (5.2)

The proof of the lemma follows from using integration by parts twice and the support properties given in the hypothesis. Note that the specified regularity of the metric g and of the functions is needed in order to use the L2 inner product. We may now prove the following theorem.

Theorem 5.7. Given Green operators satisfying conditions (a) and (c) of definition 5.4 and $\chi ,\varphi \in {H}_{\mathrm{comp}}^{1}\left(M\right)$ we have that

Proof. First, notice that if $\chi ,\varphi \in {H}_{\mathrm{comp}}^{1}\left(M\right)$ we have that ${G}^{+}\left(\chi \right),{G}^{-}\left(\varphi \right)\in {H}_{\mathrm{loc}}^{2}\left(M,g\right)$. Moreover, G+(χ) ∩ G(φ) ⊂ J+(supp(χ)) ∩ J(supp(φ)) is compact by the global hyperbolicity condition.

Hence,

We are now in a position to prove the main result about existence of Green operators.

Theorem 5.8. Let (M, g) be a spacetime that satisfies the definition of generalised hyperbolicity (definition 5.1). Then there exist unique continuous advanced and retarded Green operators for □g on M.

Proof. We will only discuss the advanced Green operator, the existence and the properties of the retarded Green operator follow from an analogous argument with the roles of future and past interchanged.

Existence: we define the linear map

which sends a source function f to the (unique) advanced weak solution u+. That such a u+ exists and is unique is a consequence of generalised hyperbolicity. Property (a) in definition 5.4 is immediate. In addition, the energy estimate (4.7) shows that G+ is a continuous operator. It remains to prove properties (b) and (c) in definition 5.4.

Property (b): let fV0 and $v\in \mathcal{D}\left(M\right)$, then

where we have used theorem 5.7 and property (a) for the retarded Green operator G. Thus the weak form of the required identity holds, which implies G+g(f) = f for every fV0 (see remark 5.5)(b).

Property (c) follows because supp(G+(f))= supp(u+) = supp(u+) ⊂ J+(supp(f)) by proposition 4.3.

Uniqueness: let ${\tilde {G}}^{+}$ be another linear operator satisfying definition 5.4. Given $f\in {H}_{\mathrm{comp}}^{1}\left(M\right)$ we have that $v{:=}{\tilde {G}}^{+}\left(f\right)$ satisfies □gv = f and supp(v) ⊂ J+(supp(f)). Since $f\in {H}_{\mathrm{comp}}^{1}\left(M\right)$, supp(v) ⊂ J+(supp(f)) and M is globally hyperbolic, there is a smooth timelike Cauchy surface Σ to the past of the support of f where the Cauchy data vanishes, i.e., v = 0 and ∇nv = 0. Hence, v is a solution to the zero initial data forward Cauchy problem on Σ. By uniqueness we must have v = u+ so we can conclude that ${\tilde {G}}^{+}\left(f\right)={G}^{+}\left(f\right)$ for all $f\in {H}_{\mathrm{comp}}^{1}\left(M\right)$. □

We now show that the low-regularity Green operators satisfy an exact sequence result similar to that in the smooth case [3, theorem 3.4.7].

Theorem 5.9. Let M be a connected time-oriented globally hyperbolic Lorentzian manifold that satisfies definition 5.1. Define the causal propagator as

Then the image of G is contained in Vsc and the following complex is exact:

Proof of theorem 5.9. First we show that the sequence is a complex: we have from the definitions ${G}^{+}{\square }_{g}{\vert }_{{V}_{0}}={G}^{-}{\square }_{g}{\vert }_{{V}_{0}}={\text{id}}_{{V}_{0}}$ and ${\square }_{g}{G}^{+}={\square }_{g}{G}^{-}={\text{id}}_{{H}_{\mathrm{comp}}^{1}\left(M\right)}$, therefore Ggϕ = 0 for all ϕV0 and □g = 0 for all ψU0.

  • Exactness at V0, i.e., injectivity of □g: let ϕV0 be such that □gϕ = 0. By compactness of the support there is a smooth spacelike Cauchy hypersurface Σ such that ϕ = 0 and ∇nϕ = 0 on Σ. Therefore, ϕ is a solution to the Cauchy problem with vanishing initial data and source. Uniqueness of the solution implies ϕ = 0.
  • Exactness at U0: let ϕU0 be such that ϕ ∈ ker(G), i.e., G+(ϕ) = G(ϕ). Define ψ := G+(ϕ) = G(ϕ), hence $\psi \in {H}_{\mathrm{loc}}^{2}\left(M\right)$ and □gψ = ϕ. Moreover, ψ is compactly supported in M because supp(ψ) ⊂ supp(G+(ϕ)) ∩ supp(G(ϕ)) ⊂ J+(supp(ϕ)) ∩ J(supp(ϕ)) and the latter is compact due to global hyperbolicity. Thus there exists $\psi \in {H}_{\mathrm{loc}}^{2}\left(M\right)$ such that □gψ = ϕU0 and supp(ψ) is compact. Hence, ψV0 and ϕ ∈ Im(G).
  • Exactness at Vsc: let ϕ ∈ ker(□g) and ϕVsc. Without loss of generality we may assume that supp(ϕ) ⊂ I+(K) ∪ I(K) for some compact set6 K of M. Using a partition of unity {χ, χ+} subordinate to {I(K), I+(K)} we let ϕ1 = χϕ and ϕ2 = χ+ϕ, thus ϕ = ϕ1 + ϕ2. Then supp(ϕ1) ⊂ J(K) and supp(ϕ2) ⊂ J+(K), hence ϕ1, ϕ2Vsc.

Define ψ := −□gϕ1 = □gϕ2. Then supp(ψ) is compact because supp(ψ) ⊂ J(K) ∩ J+(K). Moreover, $\psi \in {H}_{\mathrm{loc}}^{1}\left(M\right)$ since ϕVsc. Combining these two observations, we conclude that $\psi \in {H}_{\mathrm{comp}}^{1}\left(M\right)$ and therefore G+(ψ) is defined.

For arbitrary $\chi \in \mathcal{D}\left(M\right)$ we have

which shows G+(ψ) = ϕ2, where we have made use of the fact that the supports of Gχ and ϕ2 intersect in a compact set due to global hyperbolicity. Similarly, G(ψ) = −ϕ1 and therefore G(ψ) = G+(ψ) − G(ψ) = ϕ2 + ϕ1 = ϕ. In summary, we may conclude that there exists ψU0 satisfying G(ψ) = ϕ.

5.2. Restrictions

We briefly discuss the restriction of Green operators to causally compatible subsets Ω ⊂ M, that is, sets such that

We have the following theorem (cf [3, proposition 3.5.1]).

Theorem 5.10. Let M be a time oriented connected globally hyperbolic manifold with a C1,1 Lorentzian metric, G+ be the advanced Green operator for □g and Ω ⊂ M be a causally compatible open subset. Then we may define an advanced Green operator for the restriction of □g to Ω by

where φext denotes the extension of ϕ by zero. Similar results hold for G.

Remark 5.11. We denote the restriction of □g to Ω by ${\tilde {\square }}_{g}$. Notice that for all $u\in {H}_{\mathrm{loc}}^{2}\left(M\right)$ we have ${\tilde {\square }}_{g}\left(u{\vert }_{{\Omega}}\right)={\square }_{g}{\vert }_{{\Omega}}\left(u{\vert }_{{\Omega}}\right)=\left({\square }_{g}u\right){\vert }_{{\Omega}}$ and for all uH2(Ω) with supp(u) ⊆ Ω we have ${\left({\tilde {\square }}_{g}u\right)}_{\text{ext}}={\square }_{g}\left({u}_{\text{ext}}\right)$.

Proof of theorem 5.10. Property (a): let $f\in {H}_{\mathrm{comp}}^{1}\left(M\right)$ with supp(f) ⊆ Ω, then

Property (b): let fV0 with supp(f) ⊂ Ω, then

Property (c): for $f\in {H}_{\mathrm{comp}}^{1}\left(M\right)$ with supp(f) ⊆ Ω we have

6. Quantisation functors

In this section we discuss suitable categories and functors as in the smooth case that will allow us to construct the algebra of observables of the quantum theory.

6.1. The functor SYMPL and the categories GENHYP and SYMPLVECT

This subsection defines a category based on the analytic results in the previous sections and a functor assigning to each object a symplectic space.

Definition 6.1. Let GENHYP denote the category whose objects are 3-tuples (M, G+, G) where M is a time oriented connected generalised globally hyperbolic manifold as in definition 5.1 and G+, G are the unique Green operators of □g. Let $X=\left({M}_{1},{G}_{1}^{+},{G}_{1}^{-}\right)$ and $Y=\left({M}_{2},{G}_{2}^{+},{G}_{2}^{-}\right)$ be two objects in GENHYP, then Mor(X, Y) consists of all smooth maps ι : M1M2 which are time-orientation preserving isometric embeddings such that ι(M1) ⊂ M2 is a causally compatible open subset.

Remark 6.2. Theorem 5.2 shows that time oriented globally hyperbolic spacetimes with C1,1 metrics are objects in this category.

Before considering quantisation we prove the following result on compatibility of Green operators.

Theorem 6.3. Let M1 and M2 be as in definition 6.1, then the following diagram commutes:

Proof. Theorem 5.10 shows that ${\tilde {G}}^{{\pm}}\left(\phi \right){:=}{G}_{2}^{{\pm}}\left({\phi }_{\text{ext}}\right){\vert }_{{M}_{1}}$ is a Green operator. By uniqueness, this operator has to be equal to ${G}_{1}^{{\pm}}$ and the result follows. □

Remark 6.4. In the smooth setting [3] the category LORFUND is defined as the category with objects being 5-tuples (M, F, G+, G, P), where M is a Lorentzian manifold, F is a real vector bundle over M with non-degenerate inner product, P is a formally self-adjoint normally hyperbolic operator acting on sections in F and G+, G are the advanced and retarded Green operators for P. The morphisms consist of maps ι such that ι : M1M2 is a time-orientation preserving isometric embedding such that ι(M1) ⊂ M2 is a causally compatible open subset [3]. Moreover, given the condition of global hyperbolicity one can form the category GLOBHYP where objects are 3-tuples (M, F, P), where M is a Lorentzian manifold, F is a real vector bundle over M with non-degenerate inner product, P is a formally self-adjoint normally hyperbolic operator acting on sections in F. The morphisms are then given by maps ι such that ι : M1M2 is a time-orientation preserving isometric embedding such that ι(M1) ⊂ M2 is a causally compatible open subset. The existence and uniqueness of Green operators allow us to form a functor from GLOBHYP to LORFUND [3].

We now use the Green operators in order to construct a symplectic vector space. Let (M, G+, G) be an object of GENHYP and define

by

where G = G+G is the causal propagator (see theorem 5.9). Then $\tilde {\omega }$ is bilinear and skew-symmetric by theorem 5.7. However, $\tilde {\omega }$ is degenerate because ker(G) is nontrivial. Moreover, using theorem 5.9 we have that

Therefore on the quotient space V(M) = U0/ker(G) = U0/□gV0 the degenerate form $\tilde {\omega }$ induces a symplectic form which we denote by ω.

Remark 6.5. It follows from corollary 4.11 that G is continuous so that ker G is a closed subspace and hence V(M) is a normed space (and in particular, Hausdorff). See the discussion section for more details on this point.

Finally, we need a functor SYMPL : GENHYP → SYMPLVECT, where SYMPLVECT is the category whose objects are symplectic vector spaces with morphisms given by symplectic maps, i.e., linear maps A such that ω1(f, g) = ω2(Af, Ag). The following theorem shows the existence of such a functor.

Theorem 6.6. Let $X=\left({M}_{1},{G}_{1}^{+},{G}_{1}^{-}\right)$ and $Y=\left({M}_{2},{G}_{2}^{+},{G}_{2}^{-}\right)$ be two objects in GENHYP and f ∈ Mor(X, Y) be a morphism. Then $\text{ext}:{H}_{\mathrm{comp}}^{1}\left({M}_{1}\right)\to {H}_{\mathrm{comp}}^{1}\left({M}_{2}\right)$ maps the null space ker(G1) into the null space ker(G2) and hence induces a continuous symplectic linear map V(M1) → V(M2).

Proof. Let ϕ ∈ ker(G1) then $\phi ={\square }_{{g}_{1}}\psi $ for some ψV0(M1) where we have used theorem 5.9.

From the fact that ${G}_{2}\left({\phi }_{\text{ext}}\right)={G}_{2}\left({\left({\square }_{{g}_{1}}\psi \right)}_{\text{ext}}\right)={G}_{2}{\square }_{{g}_{2}}{\psi }_{\text{ext}}=0$ we see that ext( ker(G1)) ⊂ ker(G2). Hence, ext induces a linear map from V(M1) → V(M2). Moreover, for $\phi ,\psi \in {H}_{\mathrm{comp}}^{1}\left({M}_{1}\right)$ we have on taking representatives

Therefore, ext induces a symplectic map from V(M1) to V(M2). This induced map is also continuous (with the respective quotient topologies), because it is the composition π2E of two continuous maps, namely $E:V\left({M}_{1}\right)={H}_{\mathrm{comp}}^{1}\left({M}_{1}\right)/\mathrm{ker}\left({G}_{1}\right)\to {H}_{\mathrm{comp}}^{1}\left({M}_{2}\right)$, the factor map of ext, and π2, the quotient map ${H}_{\mathrm{comp}}^{1}\left({M}_{2}\right)\to {H}_{\mathrm{comp}}^{1}\left({M}_{2}\right)/\mathrm{ker}\left({G}_{2}\right)=V\left({M}_{2}\right)$. □

6.2. The functor CCR and the categories C*-ALG and QUASILOCALALG

In this section we closely follow [3] and define the algebraic structures that will be required to represent the observables of the quantum theory. The definitions are algebraic in nature and do not require any further analytical considerations with respect to the regularity of solutions to the Cauchy problem. Nevertheless, the C1,1 causality theory is required and will be mentioned below when it is used. Another modification with respect to the smooth case is that when considering the symplectic space (V, ω) in the smooth theory one has $V\left(M\right)=\mathcal{D}\left(M\right)/\mathrm{ker}\;G$ where G = G+G is a map $G:\mathcal{D}\left(M\right)\to {C}_{\mathrm{s}\mathrm{c}}^{\infty }\left(M\right)$. Employing the short-hand notation [f] for the class f + ker(G) in U0/ker(G), the symplectic form is given by $\omega \left(\left[f\right],\left[h\right]\right)={\left(f,Gh\right)}_{{L}^{2}\left(M,g\right)}$ whereas in this section we will have $V\left(M\right)={H}_{\mathrm{comp}}^{1}\left(M\right)/\mathrm{ker}\;G$ where now $G:{H}_{\mathrm{comp}}^{1}\left(M\right)\to {V}_{\text{sc}}$ and the symplectic form is given by $\omega \left(\left[f\right],\left[h\right]\right)={\left(f,Gh\right)}_{{L}^{2}\left(M,g\right)}$.

We now introduce the definition of a Weyl system and a CCR-representation of (V, ω).

Definition 6.7. A Weyl system of the symplectic vector space (V, ω) consists of a C*-algebra $\mathcal{A}$ with unit and a map $W:V\to \mathcal{A}$ such that for all φ, ψV,

  • (a)  
    W(0) = 1,
  • (b)  
    W(−φ) = W(φ)*,
  • (c)  
    W(φ) ⋅ W(ψ) = e−iω(φ,ψ)/2W(φ + ψ).

A Weyl system $\left(\mathcal{A},W\right)$ of a symplectic vector space (V, ω) is called a CCR-representation of (V, ω) if $\mathcal{A}$ is generated as a C*-algebra by the elements W(φ), φV. In this case we call $\mathcal{A}$ a CCR-algebra of (V, ω) and write it as CCR(V, ω).

It is always possible to construct a CCR-representation (CCR(V, ω), W) for any symplectic vector space (V, ω) (see [3, example 4.2.2]). Moreover, the construction is categorical in the sense that if (V1, ω1) and (V2, ω2) are two symplectic vector spaces and S : V1V2 is a symplectic linear map, then there exists a unique injective ∗-morphism CCR(S) : CCR(V1, ω1) → CCR(V2, ω2).

The proof can be found in corollary 4.2.11 in [3].

From uniqueness of the map CCR(S) it is possible to define a functor

where C*-ALG is the category whose objects are C*-algebras and whose morphisms are injective unit preserving ∗-morphisms.

A set I is called a directed set with orthogonality relation, if it carries a partial order ⩽ and a symmetric relation ⊥ between its elements such that:

  • (a)  
    For all α, βI there exists a γI with αγ and βγ,
  • (b)  
    For every αI there is a βI with αβ,
  • (c)  
    If αβ and βγ, then αγ,
  • (d)  
    If αβ and αγ, then there exists a δI such that βδ, γδ and αδ.

Sets of this type allow to define the objects and morphisms of the category QUASILOCALALG.

Definition 6.8. The objects of the category QUASILOCALALG are bosonic quasi-local C*-algebras which are pairs $\left(\mathcal{U},{\left\{{\mathcal{U}}_{\alpha }\right\}}_{\alpha \in I}\right)$ of a C*-algebra $\mathcal{U}$ and a family ${\left\{{\mathcal{U}}_{\alpha }\right\}}_{\alpha \in I}$ of C*-subalgebras, where I is a directed set with orthogonality relation such that the following holds:

  • (a)  
    ${\mathcal{U}}_{\alpha }\subset {\mathcal{U}}_{\beta }$ whenever αβ,
  • (b)  
    $\mathcal{U}=\overline{{\bigcup }_{\alpha }{\mathcal{U}}_{\alpha }}$,
  • (c)  
    The algebras ${\mathcal{U}}_{\alpha }$ have a common unit 1,
  • (d)  
    If αβ, then the commutators of elements from ${\mathcal{U}}_{\alpha }$ with those of ${\mathcal{U}}_{\beta }$ are trivial.

A morphism between two quasi-local C*-algebras $\left(\mathcal{U},{\left\{{\mathcal{U}}_{\alpha }\right\}}_{\alpha \in I}\right)$ and $\left(\mathcal{V},{\left\{{\mathcal{V}}_{\beta }\right\}}_{\beta \in J}\right)$ is defined as a pair (φ, Φ) where ${\Phi}:\mathcal{U}\to \mathcal{V}$ is a unit-preserving C*-morphism and φ : IJ is a map such that

  • (a)  
    φ is monic, i.e., if α1α2 in I then φ(α1) ⩽ φ(α2) in J,
  • (b)  
    φ preserves orthogonality, i.e., if α1α2 in I, then φ(α1) ⊥ φ(α2),
  • (c)  
    ${\Phi}\left(\mathcal{U}\right)\subset {\mathcal{V}}_{\varphi \left(\alpha \right)}$ for all αI.

In the remainder of this section we discuss a functor from GENHYP to QUASILOCALALG. Let (M, G+, G) be an object in GENHYP and

The relation OO' means that O and O' are causally independent, i.e., there is no causal curve connecting a point in $\overline{O}$ to a point in $\overline{{O}^{\prime }}$.

Remark 6.9. The proof that the set I is a directed set with orthogonality relation requires results from causality theory in a low regularity setting [8, 17, 19] to obtain lemma A.5.11 in [3] and the existence of smooth time functions [34, 35] to obtain proposition A.5.13 in [3]. Properties (a) and (b) follow upon taking α = M, β = . Property (c) follows from the observation that OO' implies J(O) ⊂ J(O'), and Property (d) is implied by lemma 4.4.8 in [3] with the appropriate modifications of lemma A.5.11 and proposition A.5.13 therein.

For any non-empty set OI take the restriction of the operator □g to the region O. Due to causal compatibility of OM the restriction of Green operators G+, G to the region O yields Green operators ${G}_{O}^{+},{G}_{O}^{-}$. Therefore, we get an object $\left(O,{G}_{O}^{+},{G}_{O}^{-}\right)$ for each OI. For O1O2 the inclusion induces a morphism ${\iota }_{{O}_{2},{O}_{1}}$ in the category GENHYP. This morphism is given by the embedding O1O2. Let ${\alpha }_{{O}_{2},{O}_{1}}$ denote the morphism $\text{CCR}{\circ}\text{SYMPL}\left({\iota }_{{O}_{2},{O}_{1}}\right)$ in C*-ALG and recall that ${\alpha }_{{O}_{2},{O}_{1}}$ is an injective unit preserving ∗-morphism.

We set for OI,

and for OI, O, M,

for O = M define

which is the algebra of CCR(VM, ωM) generated by all the ${\mathcal{U}}_{O}$; for O = , set ${\mathcal{U}}_{\varnothing }=\mathbb{C}$.

Now we assign to any morphism in GENHYP a morphism between quasi-local algebras in QUASILOCALALG: consider a morphism $\iota :\left(M,{G}^{+},{G}^{-}\right)\to \left(N,{\tilde {G}}^{+},{\tilde {G}}^{-}\right)$ in GENHYP. Let I1, I2 denote the index sets associated to M, N respectively. We define a map φ : I1I2 by MN and O1ι(O1) if O1M. Since ι is an embedding such that ι(M) ⊂ N is causally compatible, the map φ is monotonic and preserves causal independence. Therefore, (φ, Φ) with Φ = CCR◦SYMPL(ι) is the required morphism. To be precise we have the following result.

Theorem 6.10. The assignment $\left(M,{G}^{+},{G}^{-}\right)\to \left({\mathcal{U}}_{M},{\left\{{\mathcal{U}}_{O}\right\}}_{O\in I}\right)$ and ι → (φ, Φ) yields a functor QUANT from GENHYP to QUASILOCALALG.

Proof. A detailed proof can be found in [3, lemma 4.4.10, theorem 4.4.11 and lemma 4.4.13]. □

Remark 6.11. In the low regularity setting the proof above requires one to consider elements $\phi \in {H}_{\mathrm{comp}}^{1}\left(O\right)$ rather than $\phi \in \mathcal{D}\left(O\right)$ in lemma 4.4.10 and the low regularity quotient ${H}_{\mathrm{comp}}^{1}\left(M\right)/\mathrm{ker}\left(G\right)$ instead of $\mathcal{D}\left(M\right)/\mathrm{ker}\left(G\right)$ with C1,1 causality theory in lemma 4.4.13 in [3].

6.3. The Haag–Kastler axioms

In this subsection we show that the functor QUANT given by theorem 6.10 satisfies the Haag–Kastler axioms.

Theorem 6.12. The functor QUANT : GENHYP → QUASILOCALALG satisfies the Haag–Kastler axioms, i.e., for every object (M, G+, G) in GENHYP the corresponding quasi-local C*-algebra $\left({\mathcal{U}}_{M},{\left\{{\mathcal{U}}_{\mathcal{O}}\right\}}_{\mathcal{O}\in I}\right)$ satisfies:

  • (a)  
    If O1O2 then ${\mathcal{U}}_{{\mathcal{O}}_{1}}\subset {\mathcal{U}}_{{O}_{2}}$ for all O1, O2I.
  • (b)  
    ${\mathcal{U}}_{M}=\overline{{\bigcup }_{O\in I,O\ne M,\varnothing }{\mathcal{U}}_{\mathcal{O}}}$.
  • (c)  
    ${\mathcal{U}}_{M}$ is simple.
  • (d)  
    The ${\mathcal{U}}_{O}$'s have a common unit 1.
  • (e)  
    For all O1, O2I with $J\left(\overline{{O}_{1}}\right)\cap \overline{{O}_{2}}=\varnothing $ the subalgebras ${\mathcal{U}}_{{O}_{1}},{\mathcal{U}}_{{O}_{2}}$ commute.
  • (f)  
    (Time-slice axiom) let O1O2 be nonempty elements of I admitting a common smooth spacelike Cauchy hypersurface, then ${\mathcal{U}}_{{O}_{1}}={\mathcal{U}}_{{O}_{2}}$.
  • (g)  
    Let O1, O2I and let the Cauchy development D(O2) be relatively compact in M. If O1D(O2), then ${\mathcal{U}}_{{O}_{1}}\subset {\mathcal{U}}_{{O}_{2}}$.

The proof will be based on the following two lemmas.

Lemma 6.13. Let O be a causally compatible globally hyperbolic open subset of a C1,1 globally hyperbolic manifold M. Assume there exists a smooth spacelike Cauchy hypersurface Σ of O which is also a Cauchy hypersurface of M, let h be a smooth Cauchy time function on O and KM be compact. Assume that there exists $t\in \mathbb{R}$ with KI+(h−1(t)). Then there is a smooth function ρ : M → [0, 1] such that

  • (a)  
    ρ = 1 on a neighbourhood of K,
  • (b)  
    supp(ρ) ∩ J(K) ⊂ M is compact, and
  • (c)  
    $\overline{\left\{x\in M\vert 0{< }\rho \left(x\right){< }1\right\}}\cap {J}^{-}\left(K\right)$ is compact and contained in O.

The proof of lemma 6.13 in the C1,1 setting can be carried out following that of [3] with suitable modifications using results of low regularity causality theory [8, 17, 19].

We reproduce the main argument of the proof as done in [3].

Proof. By assumption there are real numbers t, t+ in the range of h such that $K\subset {I}^{{\pm}}\left({S}_{{t}_{{\pm}}}\right)$ where St = h−1(t). Since St is a Cauchy surface of O and since O and M admit a common Cauchy hypersurface, from C1,1 causality theory it follows that S± are also Cauchy hypersurfaces of M. Since J±(S±) are disjoint closed subsets of M there exists a smooth function ρ : M → [0, 1] such that $\rho {\vert }_{{J}^{+}\left({S}_{{t}_{+}}\right)}=1$ and $\rho {\vert }_{{J}^{-}\left({S}_{{t}_{-}}\right)}=0$.

The function ρ satisfies the first property stated in the lemma because $K\subset {I}^{+}\left({S}_{{t}_{+}}\right)$.

Since $\rho {\vert }_{{J}^{-}\left({S}_{{t}_{-}}\right)}=0$, we have $\mathrm{supp}\left(\rho \right)\subset {J}^{+}\left({S}_{{t}_{-}}\right)$. Moreover, in the C1,1 setting the causal relationship is still closed and implies that ${J}^{+}\left({S}_{{t}_{-}}\right)\cap {J}^{-}\left(K\right)$ is compact and therefore supp(ρ) ∩ J(K) ⊂ M is compact. This shows that the second property is satisfied.

The last property follows from two observations. The first one is that the closed set $\overline{\left\{x\in M\vert 0{< }\rho \left(x\right){< }1\right\}}\cap {J}^{-}\left(K\right)$ is contained in the compact set supp(ρ) ∩ J(K) ⊂ M. The second observation is that ${J}^{+}\left({S}_{{t}_{-}}\right)\cap {J}^{-}\left({S}_{{t}_{+}}\right)\subset O$ which implies $\overline{\left\{x\in M\vert 0{< }\rho \left(x\right){< }1\right\}}\cap {J}^{-}\left(K\right)$ is contained in O. The statement that ${J}^{+}\left({S}_{{t}_{-}}\right)\cap {J}^{-}\left({S}_{{t}_{+}}\right)\subset O$ follows from the characterisation of Cauchy hypersurfaces as surfaces that are met exactly once by every inextendible causal curve. This characterisation also holds in the C1,1 setting. □

Lemma 6.14. Let (M, G+, G) be an object of GENHYP and O be a causally compatible globally hyperbolic open subset of M. Assume that there exists a Cauchy hypersurface Σ which is also a Cauchy hypersurface of M. Let φU0, then there exist χV0 and ψU0 such that supp(ψ) ⊂ O and φ = ψ + □gχ.

Proof. Let h be a Cauchy time function on O. Fix tt+ in the range of h. Then the subsets Σ = h−1(t), Σ+ = h−1(t+) are Cauchy hypersurfaces of M. Hence every inextendible timelike curve in M meets Σ, Σ+. Since tt+, the set {I+), I+)} is a finite open cover of M. Let {f+, f} be a smooth partition of unity subordinated to this cover. In particular, supp(f±) ⊂ I±). Set K± := supp(f±φ) = supp(φ) ∩ supp(f±). Then K± is a compact subset of M satisfying K±I±). Applying lemma 6.13 we obtain two smooth functions ρ± : M → [0, 1] satisfying

  • (a)  
    ρ± = 1 on a neighbourhood of K±,
  • (b)  
    supp(ρ±) ∩ J(K±) ⊂ M is compact, and
  • (c)  
    $\overline{\left\{x\in M\vert 0{< }{\rho }_{{\pm}}\left(x\right){< }1\right\}}\cap {J}^{\mp }\left({K}_{{\pm}}\right)$ is compact and contained in O.

Set χ± := ρ±G(f±φ), χ := χ+χ and ψ := φ − □gχ. Since supp(G(f±φ)) ⊂ J(K), the support of χ± is contained in supp(ρ±) ∩ J(K±) which is compact by the second property of ρ±. Since ρ± and f± are smooth by construction, we have ${\chi }_{{\pm}}\in {H}_{\mathrm{comp}}^{2}\left(M\right)$. Moreover,

which implies ${\square }_{g}{\chi }_{{\pm}}\in {H}_{\mathrm{loc}}^{1}\left(M\right)$. Notice that □gρ± is not smooth but C0,1.

Now ψ is the difference of ${H}_{\mathrm{loc}}^{1}\left(M\right)$ functions so it remains to show that supp(ψ) is compact and contained in O. By the first property of ρ±, one has χ± := G(f±φ) in a neighbourhood of K±. Moreover, f±φ = 0 on {ρ± = 0}. Hence, □gχ± = f±φ on {ρ± = 0} ∪ {ρ± = 1}. Therefore, f±φ − □gχ± vanishes outside $\overline{\left\{x\in M\vert 0{< }{\rho }_{{\pm}}\left(x\right){< }1\right\}}$, i.e., $\mathrm{supp}\left({f}_{{\pm}}-{\square }_{g}{\chi }_{{\pm}}\right)\subset \overline{\left\{0{< }{\rho }_{{\pm}}\left(x\right){< }1\right\}}$. By the definitions of χ±, f± one also has supp(f±φ − □gχ±) ⊂ J(K±), hence $\mathrm{supp}\left({f}_{{\pm}}\varphi -{\square }_{g}{\chi }_{{\pm}}\right)\subset {J}^{\mp }\left({K}_{{\pm}}\right)\cap \overline{\left\{0{< }{\rho }_{{\pm}}\left(x\right){< }1\right\}}$ which is compact and contained in O by the third property of ρ±. Therefore, ψU0 with supp(ψ) ⊂ O. Moreover, φψ = □gχU0 which gives χV0. □

Proof of theorem 6.12. The first, fourth and fifth axiom follow from the definition of the quasi-local C*-algebra, the definition of the set I and [3, lemma 4.4.10]. The second axiom follows from [3, lemma 4.4.13 ] and the third axiom follows from [3, remark 4.5.3]. Remark 6.11 mentions the necessary modifications of those lemmas in the C1,1 setting.

It therefore remains to prove the time-slice axiom. Let O1O2 be nonempty causally compatible globally hyperbolic subsets of M admitting a common smooth spacelike Cauchy hypersurface Σ. Let [ϕ] ∈ V(O2). Then lemma 6.14 applied to M := O2 and O = O1 yields χV0, ψU0 such that ϕ = ψext + □gχ with supp(ψ) ⊂ O1. Since, ${\square }_{g}\chi \in \mathrm{ker}\left({G}_{{O}_{2}}\right)$ we have [ϕ] = [ψext], that is, [ϕ] is the image of the symplectic linear map V(O1) → V(O2) induced by the inclusion ι : O1O2. Therefore, the map is surjective, and hence an isomorphism of symplectic topological vector spaces. This isomorphism functorially induces an isomorphism of C*-algebras, hence ${\mathcal{U}}_{{O}_{1}}={\mathcal{U}}_{{O}_{2}}$. This proves the time-slice axiom. Finally, the seventh axiom can be deduced from the first and the sixth axiom [3, theorem 4.5.1]. □

7. Discussion

In this paper we have constructed Green operators for spacetime metrics of regularity C1,1. The function spaces for the domain and range of the Green operators play a fundamental role in low regularity spacetimes and our choices for these spaces were motivated by the following two requirements: global well-posedness of the Cauchy problem and employing Sobolev spaces, such as ${H}_{\mathrm{loc}}^{k}\left(M\right)$ and ${H}_{\mathrm{comp}}^{k}\left(M\right)$ ($k\in {\mathbb{N}}_{0}$), that do not depend on a Riemannian background metric. We have shown that the quotient space V(M) = U0/□gV0 can be used to construct quasi-local C*-algebras that satisfy the Haag–Kastler axioms, so that in a quantum theoretic setting the self-adjoint elements in these C*-algebras can be associated with the observables of the theory.

7.1. Topological issues

Let us describe the quotient vector space U0/□gV0 in some more detail for the globally hyperbolic case, where we have ker(G) = Im(□g) as a consequence of the spectral sequence given in theorem 5.9, thus U0/□gV0 = U0/ker(G) in this case. Recall that G is a linear map U0Vsc and let G0 denote the associated map from the quotient U0/ker(G) to Im(G) ⊆ Vsc, defined by G0(ϕ + ker(G)) := for every ϕU0. Therefore, G0 is linear and bijective by construction and we arrive at the following chain of (algebraic) isomorphisms of vector spaces

Equation (7.1)

Recall that the analogue of (7.1) in the smooth globally hyperbolic case, as discussed in [3], is

showing also that the quotient is isomorphic to the space of solutions to the homogeneous wave equation.

The question arises whether the isomorphism in the middle part of (7.1), obtained via the factored map G0, is topological, where the quotient U0/ker(G) is equipped with the finest topology such that the canonical surjection π : U0U0/ker(G), ϕϕ + ker(G) is continuous. Note that by continuity of G we have that ker(G) is closed in the normed space U0, hence U0/ker(G) is a normed space (in particular, Hausdorff). Furthermore, G0 is continuous by construction and the continuity of G, thus it remains to be checked whether the inverse of G0 is continuous, or, equivalently, whether G0 is an open map.

Remark 7.1. We note that by [40, chapter III, proposition 1.2], the factored map G0 is a topological isomorphism if and only if G is open as a map from U0 to Im(G) (with the relative topology on the latter). In case of Fréchet spaces such a property for G could be deduced conveniently via an open mapping principle or from a closed image criterion, but observe that neither ${U}_{0}={H}_{\mathrm{comp}}^{1}\left(M\right)$ nor Im(G) is complete (with respect to the metric inherited from the Banach space H1(M) and the Fréchet space ${H}_{\mathrm{loc}}^{2}\left(M\right)$, respectively).

We choose a finer topology σ on Vsc to make ${\square }_{g}:\left({V}_{\text{sc}},\sigma \right)\to {H}_{\mathrm{loc}}^{1}\left(M\right)$ continuous by adding the semi-norms ${p}_{\chi }\left(\phi \right){:=}{{\Vert}\chi \cdot {\square }_{g}\phi {\Vert}}_{{H}^{1}}$ ($\chi \in \mathcal{D}\left(M\right)$) to those on Vsc inherited from ${H}_{\mathrm{loc}}^{2}\left(M\right)$. Note that this has no effect on the subspace Im(G) ⊆ Vsc, since Im(G) ⊆ ker(□g) in the complex of maps in theorem 5.9 (even equality holds due to global hyperbolicity). In fact, σ is precisely the coarsest topology that is finer than the ${H}_{\mathrm{loc}}^{2}\left(M\right)$-topology on Vsc, which we denote by τ2, and renders □g continuous as a map ${V}_{\text{sc}}\to {H}_{\mathrm{loc}}^{1}\left(M\right)$, i.e., σ is the supremum (in the lattice of topologies on Vsc) of τ2 and the initial (projective) topology τ1 with respect to □g. Therefore, we have continuity of G : U0 → (Vsc, σ), since G is continuous U0 → (Vsc, τ2) by corollary 4.11 and also continuous U0 → (Vsc, τ1) due to the obvious continuity of □gG = 0 from U0 into ${H}_{\mathrm{loc}}^{1}\left(M\right)$.

Lemma 7.2. The inverse of G0 : U0/ker(G) → Im(G), ϕ + ker(G) ↦ , is continuous.

Proof. We will show that ${G}_{0}^{-1}$ can be written as the composition ${G}_{0}^{-1}=\pi {\circ}P{\circ}Z$ of three continuous linear maps. The map π : U0U0/ker(G) is the canonical surjection, which is continuous by construction. It remains to construct suitable continuous maps P and Z with PZ : Im(G) → U0 and such that G0πPZ = idIm(G) and $\pi {\circ}P{\circ}Z{\circ}{G}_{0}={\text{id}}_{{U}_{0}/\mathrm{ker}\left(G\right)}$.

Let ${V}_{\text{sc}}^{{\pm}}{:=}\left\{\phi \in {V}_{\text{sc}}\vert \mathrm{supp}\left(\phi \right)\subseteq {J}^{{\pm}}\left(K\right)\;\text{for}\;\text{some}\;\text{compact}\;\text{subset}\;K\subseteq M\right\}$ and define the subspace

which we equip with the trace of the product topology stemming from σ.

Construction of P: we consider P : WU0, given by P(ϕ, ϕ+) := (□gϕ+ − □gϕ)/2. Note that a priori, P(ϕ, ϕ+) is only in ${H}_{\mathrm{loc}}^{1}\left(M\right)$ and we have to show that P(ϕ, ϕ+) has compact support, thus belongs to ${U}_{0}={H}_{\mathrm{comp}}^{1}\left(M\right)$. To prove this, observe that ϕ = ϕ + ϕ+ ∈ ker(□g) implies □gϕ = −□gϕ+, hence = □gϕ+ = −□gϕ. Let K and K+ be compact subsets of M with supp(ϕ±) ⊆ J±(K±), then we have supp() ⊆ supp(ϕ) ∩ supp(ϕ+) ⊆ J(K) ∩ J+(K+), where J(K) ∩ J+(K+) is compact by global hyperbolicity ([3], lemma A.5.7]). The continuity of P is clear by construction of the topology σ.

Construction of Z: as a preparation we will first construct two continuous maps ${S}^{{\pm}}:{V}_{\text{sc}}\to {V}_{\text{sc}}^{{\pm}}$, such that ϕ = Sϕ + S+ϕ holds for every ϕVsc and, moreover,

Equation (7.2)

Choose a smooth spacelike Cauchy surface Σ ⊆ M and let $t:M\to \mathbb{R}$ be a smooth temporal function such that Σ = t−1(0) (compare the earlier discussion on causality in C1,1-spacetimes). We obtain an open covering of M by the two sets O := {xM|t(x) < 1} and O+ := {xM|t(x) > −1} and choose a subordinate partition of unity χ, χ+C(M), i.e., supp(χ±) ⊆ O± and χ + χ+ = 1. We define S±ϕ := χ±ϕ, then the relation ϕ = Sϕ + S+ϕ holds by construction and the continuity of S± is clear from continuity of multiplication by fixed smooth functions with respect to (localised) Sobolev norms. It remains to show that ${S}^{{\pm}}\in {V}_{\text{sc}}^{{\pm}}$ for every ϕVsc and equation (7.2) is true.

Let ϕVsc and KM be compact such that supp(ϕ) ⊆ J(K) ∪ J+(K). Then supp(χ+ϕ) ⊆ O+ ∩ (J(K) ∪ J+(K)) ⊆ (O+J(K)) ∪ J+(K). Note that O+J(K) is relatively compact by [3, corollary A.5.4], since O+J+) holds with Σ := t−1(−1) (note that the time function is strictly increasing along causal curves). Therefore, with some compact set K+ containing K as well as O+J(K) we obtain supp(χ+ϕ) ⊆ J+(K+), thus ${S}^{+}\phi \in {V}_{\text{sc}}^{+}$. The reasoning for ${S}^{-}\phi \in {V}_{\text{sc}}^{-}$ is analogous.

For the proof of (7.2) we start by noting that ϕ ∈ Im(G) = ker(□g) implies 0 = □gϕ = □gSϕ + □gS+ϕ, so that the part with S in (7.2) follows once the equation for S+ is shown. Recall that we have ${\square }_{g}{S}^{+}\phi =-{\square }_{g}{S}^{-}\phi \in {H}_{\mathrm{comp}}^{1}\left(M\right)$ from the reasoning in the construction of P above. Moreover, for every test function ψ on M we have that supp(Gψ) ∩ supp(S±ϕ) ⊆ J(supp(ψ)) ∩ J±(K) for some compact set K, hence global hyperbolicity guarantees that the supports of Gψ and S+ϕ as well as those of G+ψ and Sϕ always have compact intersection. To summarise, we may apply theorem 5.7 and lemma 5.6 to obtain the following chain of weak equalities

which implies G±gS+ϕ = ±S±ϕ and therefore GgS+ϕ = G+gS+ϕGgS+ϕ = S+ϕ − (−Sϕ) = S+ϕ + Sϕ = ϕ. Thus, equation (7.2) is proved and concludes the preparatory construction of S±.

Finally, we turn to the definition of the map Z. Observe that ϕ ∈ Im(G) = ker(□g) ⊆ Vsc implies (Sϕ, S+ϕ) ∈ W, which allows to set := (Sϕ, S+ϕ) for every ϕ ∈ Im(G) and obtain a continuous linear map Z : Im(G) → W.

We complete the proof by showing that πPZ is the inverse of G0.

  • The relation G0πPZ = idIm(G) holds, since for every ϕ ∈ Im(G) we have
    where we may apply (7.2) to rewrite the last term as $\frac{1}{2}\left(\phi +\phi \right)=\phi $.
  • We finally show that the equation $\pi {\circ}P{\circ}Z{\circ}{G}_{0}={\text{id}}_{{U}_{0}/\mathrm{ker}\left(G\right)}$ is true. Let fU0, then
    and in the last term we may replace ${f}_{1}{:=}\frac{1}{2}\left({\square }_{g}{S}^{+}Gf-{\square }_{g}{S}^{-}Gf\right)$ by f, since thanks to (7.2) the difference is in the kernel: $G\left({f}_{1}-f\right)=\frac{1}{2}\left(Gf+Gf\right)-Gf=Gf-Gf=0$. □

Proposition 7.3. For a globally hyperbolic C1,1 spacetime (M, g), we obtain a topological isomorphism U0/ker(G) ≅ Im(G) according to (7.1), where Vsc carries the topology σ.

Remark 7.4. We are not using an inductive limit construction for the topology on Vsc as, e.g., in [41], because we preferred to stay with questions of convergence and continuity in the simpler realm of local Sobolev norms. Moreover, in the above context, we would otherwise not have a topological isomorphism of Im(G) with U0/ker(G), since we decided coherently that U0 should inherit the norm topology from H1(M), thus rendering ${U}_{0}={H}_{\mathrm{comp}}^{1}\left(M\right)$ normed, but incomplete. However, the basic constructions of quantisation for the associated symplectic (quotient) vector spaces do not require completeness.

7.2. An equivalent symplectic structure

An analogous construction of the CCR representation can be achieved using a symplectic structure on the vector space of solutions to the homogeneous problem parametrised by their initial data [1]. In that context, one defines a symplectic structure Ξ on ker(□g) given by Ξ(ϕ, ψ) = Σ(u1v0v1u0)μh where (u0, u1), (v0, v1) are compactly supported smooth initial data induced by the smooth solutions ϕ, ψ respectively on the Cauchy hypersurface Σ. Moreover, the Weyl system generated by the symplectic space (ker(□g), Ξ) is isomorphic to the Weyl system generated by (U0/ker(G), ω) [1, 42]. In the C1,1 setting this isomorphism remains true with suitable modifications.

To be precise, using theorem 5.9 we know that ker(□g) = Im(G). Moreover, for any smooth spacelike Cauchy hypersurface Σ, if ϕ = G(f) then $\phi {\vert }_{{\Sigma}}\in {H}_{\mathrm{comp}}^{2}\left({\Sigma}\right)$ and ${\nabla }_{n}\phi {\vert }_{{\Sigma}}\in {H}_{\mathrm{comp}}^{1}\left({\Sigma}\right)$. This follows from the observation that ϕVsc and is the difference of two solutions to the Cauchy problem with zero initial data, which by theorem 4.7 belong to the space ${C}^{0}\left(\mathbb{R},{H}^{2}\left({{\Sigma}}_{t}\right)\right)\cap {C}^{1}\left(\mathbb{R},{H}^{1}\left({{\Sigma}}_{t}\right)\right)$. Therefore, given any smooth spacelike Cauchy hypersurface Σ, we define for ϕ, ψ ∈ ker □g with u0 := ϕ|Σ, u1 := ∇nϕ|Σ, v0 := ψ|Σ, v1 := ∇nψ|Σ, hence $\left({u}_{0},{u}_{1}\right),\left({v}_{0},{v}_{1}\right)\in {H}_{\mathrm{comp}}^{2}\left({\Sigma}\right){\times}{H}_{\mathrm{comp}}^{1}\left({\Sigma}\right)$, the skew-symmetric bilinear form

It follows from linearity, the uniqueness of solutions to the Cauchy problem, and direct computations that Ξ is symplectic where to show non-degeneracy one tests with elements of the form (0, u1) and (v0, 0), i.e., with u0 = 0 and v1 = 0 and employs uniqueness in the Cauchy problem (cf [42]).

We show that ΞΣ does not depend on Σ: this follows from the divergence theorem in a region bounded by two Cauchy hypersurfaces Σ1, Σ2 and the conservation of the current ${j}^{\mu }\left(\phi ,\psi \right)={g}^{\mu \nu }\left(\phi {\nabla }_{\nu }\psi -\psi {\nabla }_{\nu }\phi \right)$. Explicitly we have for any ϕ, ψ ∈ ker(□g)

and

Therefore,

so we will drop the Σ from the notation of Ξ. Notice that the ${H}_{\mathrm{loc}}^{2}$ regularity is required in order to make sense of the divergence of the current.

Finally, we show that the linear bijective factor map G0 of G, as defined before (7.1), provides a symplectic map from (U0/ker(G), ω) to (ker(□g), Ξ).

Proposition 7.5. Let the symplectic vector spaces (ker(□g), Ξ), (U0/ker(G), ω) and the factor map G0 be defined as above. Then we have for every f, f' ∈ U0 with ϕ = G0([f']) = Gf', ψ = G0([f]) = Gf ∈ ker(□g),

Proof. Without loss of generality we may consider $M\cong \mathbb{R}{\times}{\Sigma}$ and suppose that supp(f) ⊂ (t1, t2) × Σ for some real t1 < t2. Then we have for every ϕ ∈ ker(□g) upon integrating by parts twice,

Using the fact that □gϕ = 0 and that ${{\Sigma}}_{{t}_{1}}$ is disjoint7 from supp(G+f) we obtain

Similarly, from the causal properties again we have that ${{\Sigma}}_{{t}_{2}}$ and supp(G(f)) are disjoint. Therefore ${G}^{+}f{\vert }_{{{\Sigma}}_{{t}_{2}}}=Gf{\vert }_{{{\Sigma}}_{{t}_{2}}}$ and ${\nabla }_{n}{G}^{+}f{\vert }_{{{\Sigma}}_{{t}_{2}}}={\nabla }_{n}Gf{\vert }_{{{\Sigma}}_{{t}_{2}}}$ which gives

Recalling that ψ = G(f) and t1 < t < t2 in supp(f) we obtain

Here, we use also the assumption G(f') = ϕ to proceed with

We have established a symplectomorphism between the spaces (ker(□g), Ξ) and (U0/ker(G), ω). This implies that the functor CCR will give isomorphic C*-algebras in the quantisation. Therefore, the result shows that one can use either the elements of U0/ker(G) or those of ker(□g) to construct the algebra of quantum observables.

7.3. The physical quantum states

Finally, in order to construct a full quantum field theory in a low regularity spacetime, a suitable choice of quantum states must be made. In the algebraic quantisation method, a quantum state Λ is a normalised positive linear functional on the quasi-local C*-algebra. In particular, given a real scalar product $\mu :\mathrm{ker}\left({\square }_{g}\right){\times}\mathrm{ker}\left({\square }_{g}\right)\to \mathbb{R}$ satisfying ${\vert {\Xi}\left(\phi ,\psi \right)\vert }^{2}{\leqslant}\frac{1}{4}\mu \left(\phi ,\psi \right)\mu \left(\phi ,\psi \right)$ for all ϕ, ψ ∈ ker(□g), we define a quasi-free state by ${{\Lambda}}_{\mu }\left(W\left(\phi \right)\right)={\mathrm{e}}^{\frac{1}{2}\mu \left(\phi ,\phi \right)}$. As a consequence of the GNS construction, a quasi-free state on a C*-algebra possesses a natural Fock space structure [1]. Moreover, if additional symmetries exist such as the existence of a timelike Killing vector field one can define ground states and thermal equilibrium states at finite temperature as quasi-free states that can be represented as elements of a suitable Fock space.

In the absence of symmetries a common candidate for the physical quantum states in the smooth case are the quasi-free states that satisfy the microlocal spectrum condition. This condition allows via a renormalization procedure to define the energy momentum tensor of the state [1].

To specify the microlocal spectrum condition, we need to define appropriate subsets of T*(M × M)\0, i.e., the cotangent bundle with the zero section removed, and the two-point function of the state Λμ, which is a distribution on M × M. Let

where $\left({x}_{1},\eta \right)\sim \left({x}_{2},\tilde {\eta }\right)$ means that η, $\tilde {\eta }$ are cotangent to the null geodesic γ at x1, x2 respectively, and parallel transports of each other along γ. The value of the two-point function of a state Λμ acting on the elements of the algebra defined by ϕ and ψ is

Using the isomorphism between ker(□g) and V(M) the two-point function can be seen to induce a bidistribution on spacetime, i.e., ${{\Lambda}}_{2}\in {\mathcal{D}}^{\prime }\left(M{\times}M\right)$.

Definition 7.6. A quasi-free state ΛH on the algebra of observables satisfies the microlocal spectrum condition if its two-point function Λ2H is a distribution ${\mathcal{D}}^{\prime }\left(M{\times}M\right)$ and satisfies the following wavefront set condition

where $W{F}^{\prime }\left({{\Lambda}}_{2H}\right){:=}\left\{\left({x}_{1},\eta ;{x}_{2},-\tilde {\eta }\right)\in {T}^{{\ast}}\left(M{\times}M\right);\left({x}_{1},\eta ;{x}_{2},\tilde {\eta }\right)\in WF\left({\omega }_{2H}\right)\right\}$.

The states that satisfy the microlocal spectrum condition8 are called Hadamard states and their class includes ground states and KMS states ([13, 44, 45]). An alternative method of constructing a Fock space makes use of the SJ vacuum states. These are quasi-free states that in general do not satisfy this condition [46].

In the low regularity setting we require a generalisation of Hadamard states. A larger class of states, called adiabatic states of order N and characterised in terms of their Sobolev wavefront set, has been obtained by Junker and Schrohe [47]. These states are natural candidates to replace the Hadamard states in spacetimes with limited regularity. In particular, quantum ground states have been constructed in static spacetimes using semigroup techniques [48] and they can be described as adiabatic states [49]. We briefly recall the definition of this class of states and of the Sobolev wavefront set.

Definition 7.7. A quasi-free state ΛN on the algebra of observables is called an adiabatic state of order $N\in \mathbb{R}$ if its two-point function Λ2N is a bidistribution that satisfies for every $s{\leqslant}N+\frac{3}{2}$ the Hs-wavefront set condition

where WFs denotes the refinement of the notion of the wavefront set in terms of Sobolev regularity ([43]), i.e., (x, ξ) ∉ WFs(u) if and only if u = u1 + u2 with u1Hs and (x, ξ) ∉ WF(u2).

Acknowledgments

All authors acknowledge financial support through FWF (Austrian Science Fund) project no. P28770, JAV acknowledges financial support from STFC in the UK through Grant No. ST/R00045X/1 and YSS would like to thank the Max Planck Institute of Mathematics in Bonn and Leibniz University Hannover for their financial support during the writing of this paper.

Appendix A.: Regularisation methods and generalised functions

In this section we gather a minimum of notions required from the theory of Colombeau generalised functions and regularisation methods for Lorentzian metrics. For a comprehensive introduction to the theory of Colombeau algebras we refer to [15, 50], the details about the approximation results for Lorentzian metrics can be found in [19, 26, 51].

Let E be a locally convex topological vector space whose topology is given by the family of semi-norms ${\left\{{p}_{j}\right\}}_{j\in J}$. The elements of

and

are called E-moderate and E-negligible, respectively. Defining operations component-wise turns ${\mathcal{N}}_{E}$ into a vector subspace of ${\mathcal{M}}_{E}$. We define the generalised functions based on E as the quotient ${\mathcal{G}}_{E}{:=}{\mathcal{M}}_{E}/{\mathcal{N}}_{E}$. If E is a differential algebra, then ${\mathcal{N}}_{E}$ is an ideal in ${\mathcal{M}}_{E}$ and ${\mathcal{G}}_{E}$ is a differential algebra as well, called the Colombeau algebra based on E.

Let Ω be an open subset of ${\mathbb{R}}^{n}$. By choosing E = C(Ω) with the topology of uniform convergence of all derivatives one obtains the standard Colombeau algebra ${\mathcal{G}}_{{C}^{\infty }\left({\Omega}\right)}=\mathcal{G}\left({\Omega}\right)$; here we will mainly use $E={H}^{\infty }\left({\Omega}\right)=\left\{h\in {C}^{\infty }\left(\overline{{\Omega}}\right):{\partial }^{\alpha }h\in {L}^{2}\left({\Omega}\right)\,\forall \alpha \in {\mathbb{N}}_{0}^{n}\right\}$ with the family of semi-norms

or $E={W}^{\infty ,\infty }\left({\Omega}\right)=\left\{h\in {C}^{\infty }\left(\overline{{\Omega}}\right):{\partial }^{\alpha }h\in {L}^{\infty }\left({\Omega}\right)\,\,\forall \alpha \in {\mathbb{N}}^{n}\right\}$ with the family of semi-norms

We employ the notation

Colombeau algebras contain the distributions as a linear subspace, though not every element of a Colombeau algebra is a regularisation of a distribution. Their elements are equivalence classes of nets of smooth functions, $\mathcal{G}\left({\Omega}\right)\ni u=\left[{\left({u}_{\varepsilon }\right)}_{\varepsilon }\right]$. We say that a Colombeau function u is associated with a distribution $w\in {\mathcal{D}}^{\prime }\left({\Omega}\right)$ if some (and hence every) representative ${\left({u}_{\varepsilon }\right)}_{\varepsilon }$ converges to w in ${\mathcal{D}}^{\prime }\left({\Omega}\right)$. The distribution w represents the macroscopic behaviour of u and is called the distributional shadow of u.

A generalised function $u\in \mathcal{G}\left({\Omega}\right)$ is said to be of L-log-type if

Logarithmic growth conditions on the coefficients of a differential equation are typical in statements on existence and uniqueness of generalised solutions. These results are usually derived from a detailed analysis of regularisation techniques and Colombeau solutions often lead to very weak solutions in the sense of [16].

A related methodology of regularisation is used in the approximation results of [51] which show how to approximate a globally hyperbolic C1,1 metric by a smooth family of globally hyperbolic metrics while controlling the causal structure. We recall from [26, sections 3.8.2], [19, section 1.2] that for two Lorentzian metrics g1, g2, we say that g2 has strictly wider light cones than g1, denoted by

Equation (A.1)

Thus any g1-causal vector is g2-timelike. The key result is [19, proposition 1.2], which we give here in the strengthened version of [51, proposition 2.3].

Proposition A.1. Let (M, g) be a C0-spacetime and let h be some smooth background Riemannian metric on M. Then for any ɛ > 0, there exist smooth Lorentzian metrics ǧɛ and ĝɛ on M such that ǧɛgĝɛ and dh(ǧɛ, g) + dh(ĝɛ, g) < ɛ, where

Equation (A.2)

Moreover, ĝɛ(p) and ǧɛ(p) depend smoothly on $\left(\varepsilon ,p\right)\in {\mathbb{R}}^{+}{\times}M$, and if gC1,1 then letting gɛ be either ǧɛ or ĝɛ, we additionally have:

  • (a)  
    For any compact subset KM there exists a sequence ɛj ↘ 0 such that ${\hat{g}}_{{\varepsilon }_{j+1}}\prec {\hat{g}}_{{\varepsilon }_{j}}$ on K (resp. ${\check{g} }_{{\varepsilon }_{j}}\prec {\check{g} }_{{\varepsilon }_{j+1}}$ on K) for all $j\in {\mathbb{N}}_{0}$.
  • (b)  
    If g' is a continuous Lorentzian metric with gg' (resp. g' ≺ g) then ĝɛ (resp. ǧɛ) can be chosen such that gĝɛg' (resp. g' ≺ǧɛg) for all ɛ.
  • (c)  
    There exist sequences of smooth Lorentzian metrics ǧjgĝj ($j\in \mathbb{N}$) such that dh(ǧj, g) + dh(ĝj, g) < 1/j and ǧjǧj+1 as well as ĝj+1ĝj for all $j\in \mathbb{N}$.
  • (d)  
    If g is C1,1 and globally hyperbolic then the ĝɛ (and ǧɛ) can be chosen to be globally hyperbolic.
  • (e)  
    If g is C1,1 then the regularisations can in addition be chosen such that they converge to g in the C1-topology and such that their second derivatives are bounded, uniformly in ɛ on compact sets.

Remark A.2. In our application the main point we will need compared to [19, section 1.2] is property (d) which guarantees that for globally hyperbolic metrics there exist approximations with strictly narrower (wider) lightcones that are themselves globally hyperbolic. Extending methods of [52], it was shown in [53] that global hyperbolicity is stable in the interval topology. Consequently, if g is a smooth, globally hyperbolic Lorentzian metric, then there exists some smooth globally hyperbolic metric g' ≺ g (resp. g' ≻ g). Constructing ĝɛ resp. ĝj as in (b) then automatically gives globally hyperbolic metrics (cf [53], section II]).

Appendix B.: Function spaces

The (real) Hilbert space L2(M, g) is used in the section on Green operators to formulate adjointness properties and is defined as follows: recall that for any Lorentzian manifold (M, g) we have a unique positive density μg on M [54, proposition 2.1.15], which has the local coordinate expression $\sqrt{\vert \mathrm{det}\left({g}_{ij}\right)\vert }\;\vert \mathrm{d}{x}^{0}\wedge \cdots \wedge \;\mathrm{d}{x}^{n}\vert $; in case of a Lipschitz continuous metric g the density μg is continuous and induces a positive Borel measure on M, which we employ to define the corresponding L2 space and denote it by L2(M, g). If M is orientable, then we have a global volume form νg on M [55, chapter 7] from which the density μg can be obtained.

We consider the (real) Sobolev spaces Hm(M) for a nonnegative integer m to be defined with respect to some chosen smooth Riemannian background metric on M as described in [56], i.e., by completion of the space of (real) smooth functions whose covariant derivatives up to order m are square integrable with respect to the positive Borel measure on M associated with the Riemannian metric (cf [57, section III.3]; it can be written in terms of a global Riemannian volume form, if M is orientable). Recall ([56], theorems 2.7 and 2.8]) that the space $\mathcal{D}\left(M\right)$ (of smooth compactly supported test functions) is dense in H1(M), if M is complete with respect to the chosen Riemannian metric, and also in Hm(M) for m ⩾ 2, if, in addition, Riemannian curvature bounds hold as well.

On a compact manifold M, the definition of Hm(M) is independent of the chosen Riemannian background metric ([56], proposition 2.3]) and, similarly, one concludes that Sobolev norms induced by two different Riemannian metrics on functions with support contained in a fixed compact subset are equivalent. This observation guarantees that the following two spaces are independent of the chosen background metric, namely the compactly supported Sobolev functions ${H}_{\mathrm{comp}}^{m}\left(M\right){:=}\left\{f\in {H}^{m}\left(M\right)\vert \mathrm{supp}\left(f\right)\;\text{is}\;\text{compact}\;\text{in}\;M\right\}$ and the local space ${H}_{\mathrm{loc}}^{m}\left(M\right){:=}\left\{f:M\to \mathbb{R}\;\text{measurable}\vert \forall \varphi \in \mathcal{D}\left(M\right):\varphi f\in {H}^{m}\left(M\right)\right\}$ (in fact, $\varphi f\in {H}_{\mathrm{comp}}^{m}\left(M\right)$ in the latter case).

In the context of the function space topologies for the current paper, we simply consider ${H}_{\mathrm{comp}}^{m}\left(M\right)$ as a subspace of the Banach space Hm(M), hence it is normed and not complete. One could equip ${H}_{\mathrm{comp}}^{m}\left(M\right)$ with a complete (non-metrizable) locally convex vector space topology, e.g., as in [41] or [58, part II, chapter 31], via a strict inductive limit construction which turns it into a so-called (LF)-space, but we prefer to formulate our results more directly in terms of the inherited Sobolev norm.

For ${H}_{\mathrm{loc}}^{m}\left(M\right)$ we have the family of semi-norms $f{\mapsto}{{\Vert}\varphi f{\Vert}}_{{H}^{m}\left(M\right)}$, parametrised by $\varphi \in \mathcal{D}\left(M\right)$, which provides us with a Fréchet space topology on ${H}_{\mathrm{loc}}^{m}\left(M\right)$ (cf [58, part II, chapter 31] or [41]). We clearly have ${H}^{m}\left(M\right)\subseteq {H}_{\mathrm{loc}}^{m}\left(M\right)$ (with continuous embedding).

If K is a compact subset of M and $f\in {H}_{\mathrm{loc}}^{m}\left(M\right)$ we occasionally abuse the notation and write ${{\Vert}f{\Vert}}_{{H}^{m}\left(K\right)}$ to mean the value obtained when the integrals defining ${{\Vert}\varphi f{\Vert}}_{{H}^{m}\left(M\right)}$ are only evaluated on K and $\varphi \in \mathcal{D}\left(M\right)$ is a cut-off such that φ = 1 on K. (No cut-off is required if fHm(M).)

In case of M = (0, T) × Σ we may choose the Riemannian background metric in the form dt ⊗ dt + γ, where γ is a Riemannian metric on Σ. We will then often consider a function vL2((0, T) × Σ) as a map tv(t) from the interval into the Hilbert space L2(Σ) in the sense that v(t)(x) = v(t, x) holds pointwise for continuous v. Thanks to Fubini's theorem, we may then write

Equation (B.1)

For general constructions with measurable functions valued in Banach spaces we refer to [59, 60]; in particular we will make use of the isomorphism L2((0, T) × Σ) ≅ L2((0, T), L2(Σ)) [59, theorem 8.28]. If v is differentiable and interpreted as a function tv(t), we will occasionally denote the partial derivative ∂tv by $\dot {v}$ and write ∂t for the corresponding vector field on (0, T) × Σ. The space Ck([0, T]; Hm(Σ)) (k a nonnegative integer) consists of all k times continuously (strongly) differentiable functions (if k = 0, simply continuous functions) v : [0, T] → Hm(Σ) with finite norm

Equation (B.2)

We have Ck([0, T]; Hm(Σ)) ⊆ L2((0, T), Hm(Σ)) (with continuous embedding).

In place of a bounded time interval we will occasionally consider the basic spacetime to be $\mathbb{R}{\times}{\Sigma}$ and deal with function spaces of Bochner measurable maps from $\mathbb{R}$ to some of the Sobolev-type Hilbert spaces (cf [59, chapter 8]), in particular, ${L}^{2}\left(\mathbb{R},{H}^{1}\left({\Sigma}\right)\right)$. We will then use the notation ${L}_{\mathrm{loc}}^{2}\left(\mathbb{R},{H}^{1}\left({\Sigma}\right)\right)$ for the set of all Bochner-measurable functions $v:\mathbb{R}\to {H}^{1}\left({\Sigma}\right)$ such that for every compact subinterval $I\subset \mathbb{R}$ the restriction v|I belongs to L2(I, H1(Σ)).

In looking at energy estimates on $\mathbb{R}{\times}{\Sigma}$ we will also need versions of the Sobolev norms where the derivatives are taken in both the space and time directions but the integration and volume form are confined to the t = τ level hypersurfaces Sτ := {τ} × Σ. These norms will be denoted by

Equation (B.3)

where $\tilde {\nabla }$ is the covariant derivative with respect to the spatial background metric γ and μτ is the Riemannian measure on Sτ which is just that given by the spatial metric γ.

Finally let us adapt the basic function space structures to the situation of a general globally hyperbolic C1,1 spacetime (M, g) with Cauchy hypersurface Σ, where we suppose that—according to the discussion in the subsection on C1,1 causality theory—we have chosen a smooth temporal function $t:M\to \mathbb{R}$ such Σ = t−1(0) and a corresponding diffeomorphism ${\Phi}:M\to \mathbb{R}{\times}{\Sigma}$. For $\tau \in \mathbb{R}$ denote the corresponding level surface by Στ := t−1(τ) = Φ−1({τ} × Σ), hence Σ0 = Σ, and consider again a background Riemannian metric of the form h = dt ⊗ dt + γ on the product manifold $\mathbb{R}{\times}{\Sigma}$, which in turn provides us with the convenient background metric Φ*h on M. In the sequel, all Sobolev spaces on submanifolds of M or $\mathbb{R}{\times}{\Sigma}$ will be considered to be defined via Riemannian metrics induced by Φ*h or h, respectively. Let Φτ denote the induced diffeomorphism Στ → Σ, i.e., Φ(x) := (τ, Φτ(x)) for every x ∈ Στ.

We will commit another abuse of notation and a somewhat naive simplification in defining now the spaces Ck(I, Hmt)) for the case of a compact interval I = [0, T] or for $I=\mathbb{R}$ without using the full theory of more sophisticated constructions in terms of sections, e.g., as in [41]. Let Bm(I) denote the set of all maps u : I → ⋃τIHmτ) such that u(τ) ∈ Hmτ) for every τI. Then we have that uBm(I) implies (equivalently) $u\left(\tau \right){\circ}{{\Phi}}_{\tau }^{-1}\in {H}^{m}\left({\Sigma}\right)$ for every τI. We define Ck(I, Hmt)) to be the subset of those elements uBm(I) such that the map $\tau {\mapsto}u\left(\tau \right){\circ}{{\Phi}}_{\tau }^{-1}$ belongs to Ck(I, Hm(Σ)).

For elements uCk(I, Hmt)) we can then also define the norms over spatial domains, but involving derivatives in space and time directions, such as ${{\Vert}u{\Vert}}_{{\tilde {H}}^{m}\left({{\Sigma}}_{\tau }\right)}$ via the corresponding ${{\Vert}.{\Vert}}_{{\tilde {H}}^{m}\left({S}_{\tau }\right)}$-norm evaluated for the associated map $\tau {\mapsto}u\left(\tau \right){\circ}{{\Phi}}_{\tau }^{-1}$ in Ck(I, Hm(Σ)).

Note that the definition of the spaces Ck(I, Hmt)) depends on the splitting $M\cong \mathbb{R}{\times}{\Sigma}$ and on the choice of temporal function. However, the reasoning in the main text tries to use the temporal function only in intermediate calculations and afterwards gives formulations of results essentially in 'pure' spacetime terms without recourse to the splitting.

Footnotes

  • As shown by Bernal and Sánchez the requirement of strong causality in the classical definition of global hyperbolicity can be weakened to only require causality [30].

  • Of course, in case of the wave operator we even have explicit representations for the advanced and retarded fundamental solutions E+ and E, e.g., in [38, sections 6.2 and 7.4], but these are not required here.

  • We may take I+(K) rather than J+(K) by replacing an initial choice of compact set $\tilde {K}$ with a slightly larger K for which ${J}^{+}\left(\tilde {K}\right)\subset {I}^{+}\left(K\right)$. Specifically we can take K = Ū where U is any open set containing $\tilde {K}$ with compact closure. Similar remarks apply to I(K).

  • Because supp(G+f) ⊆ J+(supp(f)) ⊆ (t1 + ɛ, ) × Σ for some ɛ > 0.

  • The singular structure of these states is equal to the singular structure of the Feynman parametrix minus the retarded parametrix [13]. The Feynman and retarded parametrix are examples of the distinguished parametrices of the wave equation [43]. In the smooth setting one can construct them using geodesically convex domains. Therefore, for this type of constructions in the non-smooth setting the regularity of the exponential map is crucial [8].

Please wait… references are loading.
10.1088/1361-6382/ab839a