1 Introduction

The quantisation of the scalar field forms part of the basis for the subject of Algebraic Quantum Field Theory. While the main mathematical framework for the smooth setting was initiated more than 20 years ago, see e.g. [23, 28, 32, 39, 50], ongoing research continues to develop new techniques, particularly in connection with microlocal analysis [29, 37, 38], the importance of Hadamard states [20, 27, 40, 55, 58], locality and covariance [12, 26, 44], perturbation theory [11, 14, 34], Dirac fields [15, 24, 31, 33] and gauge theory [7, 13].

Moreover, it is now possible to approach certain mathematical questions related to quantum fields propagating in spacetimes of finite regularity. This is motivated by the deep foundational work on causality theory [8, 18, 42, 46] and advances in our understanding of nonlinear hyperbolic equations [17, 19, 41], which were needed as a first step towards a full understanding of Einstein’s equations as a well-posed Cauchy problem, which requires solutions that go beyond the smooth ones. Additionally, there are several astrophysical models of phenomena such as neutron stars, self-gravitating fluids and gravitational collapse that are not smooth [2, 16, 47].

The quantisation proceeds in two steps. First, one constructs an algebra of observables, then one represents this algebra on a Hilbert space of physical states.

A common candidate for such physical quantum states, \(\omega \), are quasifree states that satisfy the microlocal spectrum condition.

To state it, it is useful to introduce the sets

$$\begin{aligned}&C=\big \{(\tilde{x},\tilde{\xi },\tilde{y},\tilde{\eta })\in T^{*}(M\times M)\backslash 0; g^{ab}(\tilde{x})\tilde{\xi }_{a}{\tilde{\xi }}_{b}{=}g^{ab}(\tilde{y})\tilde{\eta }_{a}\tilde{\eta }_{b}=0, (\tilde{x},\tilde{\xi })\sim (\tilde{y},\tilde{\eta })\big \} \nonumber \\&C^{+}{=}\left\{ (\tilde{x},\tilde{\xi },\tilde{y},\tilde{\eta }){\in } C; \tilde{\xi }^{0}\ge 0,\tilde{\eta }^{0}\ge 0\right\} , \end{aligned}$$
(1.1)

where \((\tilde{x},\tilde{\xi })\sim (\tilde{y},\tilde{\eta })\) means that there is a null geodesic \(\gamma \) joining \(\tilde{x}\) and \(\tilde{y}\) such that \(\tilde{\xi },\tilde{\eta }\) are cotangent to the null geodesic \(\gamma \) at \(\tilde{x}\) resp. \(\tilde{y}\) and parallel transports of each other.

Using the above sets, one can define the microlocal spectrum condition as follows:

Definition 1.1

A quasifree state \(\omega _{H}\) on the algebra of observables satisfies the microlocal spectrum condition if its two-point function \(\omega ^{(2)}_{H}\) is a distribution in \({\mathcal {D}}'(M\times M)\) and satisfies the following wavefront set condition

$$\begin{aligned} WF'(\omega ^{(2)}_{H})=C^{+}, \end{aligned}$$

where \(WF'(\omega ^{(2)}_{H}):= \{(\tilde{x}, \tilde{\xi }; \tilde{y}, -\tilde{\eta }) \in T^{*}(M\times M); (\tilde{x}, \tilde{\xi }; \tilde{y}, \tilde{\eta }) \in WF(\omega ^{(2)}_{H})\}.\)

These states, called Hadamard states, have been constructed in the smooth setting. They encompass both ground and KMS states [29, 37]. Moreover, they are particularly well suited for point-splitting renormalisation, a technique used for calculating key physical quantities like the renormalised energy-momentum tensor [63, 64].

A central goal now is the construction of suitable quantum states in non-smooth scenarios following the techniques in [29, 38], which requires a thorough knowledge of the wavefront set of the causal propagator. This is the question we address in this article. To be precise, we characterise the wavefront set of the causal propagator of the Klein–Gordon operator in non-smooth globally hyperbolic spacetimes. The causal propagator is constructed using the inverses associated with the Cauchy problem, which makes it a classical propagator. It is worth noting that there exist other bisolutions such as the two-point functions described above, which are non-classical (see [22] for further details on this convention).

The microlocal analysis of the propagators of the wave equation and its parametrices in low-regularity spacetimes introduces several technical challenges due to the lack of a complete theory of Fourier integral operators with non-smooth symbols and amplitudes. However, progress has been made using the paradifferential calculus introduced by Bony [10] (see also [6, 45, 61]). In addition, Szeftel has constructed a parametrix which requires only control over the \(L^2\) curvature of the metric in order to prove the \(L^2\)-curvature conjecture related to Einstein’s field equations [41, 59]. Moreover, Tataru [60] has constructed parametrices of the wave equations in low regularity for metrics with \( C^{1,1}\) coefficients as a preliminary step to show suitable Strichartz estimates and analyse nonlinear PDE’s using phase space transforms. In addition, his results allowed even lower regularity at the expense of showing weaker results. Finally, we mention Smith’s construction of parametrices for the \(C^{1,1}\) case using wave packets [56] (see [65] for a parametrix construction using Gaussians). The contribution of our paper is establishing the microlocal singular structure of the causal propagator when the regularity of the spacetime is finite. The main theorems we prove are:

Theorem

(Theorem 5.1). Let (Mg) be a \(C^{\tau }\) globally hyperbolic spacetime with \(\tau >2\) and \(K_G\) the causal propagator of the Klein–Gordon operator P. Then,

$$\begin{aligned} WF'^{-2+\tau -{\epsilon }}(K_G)\subset C \end{aligned}$$

for every \({\epsilon }>0\), C as in Eq.(1.1),

and

Theorem

(Theorem 5.2). For a \(C^{\tau +2}\) globally hyperbolic spacetime with \(\tau >2\),

$$\begin{aligned} C\subset WF'^{-\frac{1}{2}}(K_G)\subset WF'^{\tau -\epsilon }(K_G)\subset C, \end{aligned}$$

and hence equality, holds for \(0<\epsilon <\tau +\frac{1}{2}\).

In the ultrastatic case, sharper results are available. For completeness, we state these in the Appendix, see Lemmas 6.5 and 6.7, Theorems 6.9 and 6.11.

1.1 The Smooth Setting

Consider a pair (Mg), where M is a smooth manifold and g is a smooth Lorentzian metric. The Klein–Gordon operator P on (Mg) is given by

$$\begin{aligned} P:=g^{\mu \nu }\nabla _{\mu }\nabla _{\nu }\phi +m^{2}\phi =(\square _g+m^2)\phi \end{aligned}$$
(1.2)

where \(g^{\mu \nu }\) is the inverse metric tensor, \(\nabla _{\mu }\) is the covariant derivative and m is a positive real number.

The starting point is the notion of advanced and retarded Green operators in this situation.

Definition 1.2

Let M be a time-oriented connected Lorentzian manifold and let P be the Klein–Gordon operator. An advanced Green operator \(G^+\) is a linear map \(G^+:{{\mathcal {D}}}(M) \rightarrow C^\infty (M)\) such that

  1. 1.

    \(P \circ G^+ =\textrm{id}_{{{\mathcal {D}}}(M)}\)

  2. 2.

    \(G^+ \circ P|_{{{\mathcal {D}}}(M)} =\textrm{id}_{{{\mathcal {D}}}(M)}\)

  3. 3.

    \({{{\,\textrm{supp}\,}}}(G^+\phi ) \subset J^+({{{\,\textrm{supp}\,}}}(\phi ))\) for all \(\phi \in {{\mathcal {D}}}(M)\).

A retarded Green operator \(G^-\) satisfies (1) and (2), but (3) is replaced by the condition \({{{\,\textrm{supp}\,}}}(G^-\phi ) \subset J^-({{{\,\textrm{supp}\,}}}(\phi ))\) for all \(\phi \in {{\mathcal {D}}}(M)\).

In [5, Corollary 3.4.3], it is shown 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

$$\begin{aligned} G:=G^+-G^- \end{aligned}$$

which maps \({{{\mathcal {D}}}(M)}\) to \(C^\infty _{\textrm{sc}}(M)\), the space of spatially compact maps, i.e. the smooth maps \(\phi \) such that there exists a compact subset \(K \subset M\) with \({{\,\textrm{supp}\,}}(\phi ) \subset J(K)\). If M is globally hyperbolic, then one has the following exact sequence [5, Theorem 3.4.7]:

figure a

Since G is a continuous linear operator, the Schwartz Kernel Theorem implies that there exists one and only one distribution \({\displaystyle K_G\in {{\mathcal {D}}}'(M\times M)}\) such that

$$\begin{aligned} K_G(u\otimes v)=\langle G(v),u\rangle , \quad u,v\in {\mathcal {D}}(M). \end{aligned}$$
(1.3)

It follows from Duistermaat and Hörmander’s characterisation using Fourier integral operators that the kernel \(K_G\) satisfies

$$\begin{aligned} WF'(K_G)=C. \end{aligned}$$
(1.4)

More explicitly, they showed that \(K_G\in I^{-\frac{3}{2}}(M\times M, C')\), where \(I^{\mu }(X,\Lambda )\) denotes the space of Lagrangian distributions of order \(\mu \) over the manifold X associated to the Lagrangian submanifold \(\Lambda \). In this case \(\Lambda =C'=\{(\tilde{x}, \tilde{\xi }; \tilde{y}, -\tilde{\eta }); (\tilde{x}, \tilde{\xi }; \tilde{y}, \tilde{\eta }) \in C\}\), see [25, Theorem 6.5.3]. Using [25, Theorem 5.4.1, Theorem 6.5.3], one obtains that in four dimensions, \(K_G\) belongs to the Sobolev space \(H_{loc}^{-\frac{1}{2}-\epsilon }(M\times M)\) for any \(\epsilon >0\). For details on the Sobolev spaces mentioned, see Sect. 6.1 and [36, Appendix B].

2 The Non-Smooth Setting

Next we will consider the case, where g is a non-smooth metric. We will specify the precise regularity in each section.

The definition of the Green operators in the non-smooth setting will require us to choose suitable spaces of functions based on Sobolev spaces as domain and range. We let

$$\begin{aligned} V_0&= \{\phi \in H_{{comp}}^{2}(M); P\phi \in H^{1}_{{comp}}(M) \}\nonumber \\ V_{sc}&= \{\phi \in H_{{loc}}^{2}(M); P\phi \in H^{1}_{{loc}}(M) \nonumber \\&\text { and } \text {supp}(\phi ) \subset J(K), \text { where }K \text { is a compact subset of } M\}. \end{aligned}$$
(2.1)

Definition 2.1

An advanced Green operator for the Klein–Gordon operator P is a linear map

$$\begin{aligned} G^{+}:H_{\text {comp}}^{1}(M)\rightarrow H_{{loc}}^{2}(M) \end{aligned}$$

satisfying the properties

  1. 1.

    \(PG^{+}=\text {id}_{H_{{comp}}^{1}(M)}\),

  2. 2.

    \( G^{+}P|_{V_0}=\text {id}_{V_0}\),

  3. 3.

    \({{\,\textrm{supp}\,}}(G^{+}(f))\subset J^{+}({{\,\textrm{supp}\,}}(f))\) for all \(f\in H_{{comp}}^{1}(M)\),

A retarded Green operator \(G^{-}\) is defined correspondingly.

It is shown in [36, Theorem 5.8] that these operators exist and are unique on Lorentzian manifolds that satisfy the condition of generalised hyperbolicity. This condition is satisfied in particular for \(C^{1,1}\) globally hyperbolic spacetimes. Moreover, one obtains a short exact sequence for the low-regularity causal propagator, \(G:=G^+-G^-\), similar to that in the smooth case

figure b

3 Pseudodifferential Operators with Non-Smooth Symbols

3.1 Symbol Classes

Let \(\{\psi _j; j=0,1,\ldots \}\) be a Littlewood-Paley partition of unity on \({\mathbb {R}}^n\), i.e. a partition of unity \(1=\sum _{j=0}^\infty \psi _j\), where \(\psi _0\equiv 1\) for \(|\xi |\le 1\) and \(\psi _0\equiv 0\) for \(|\xi |\ge 2\) and \(\psi _j(\xi ) = \psi _0(2^j\xi )-\psi _0(2^{1-j}\xi )\). The support of \(\psi _j\), \(j\ge 1\), then lies in an annulus around the origin of interior radius \(2^j\) and exterior radius \(2^{1+j}\).

Definition 3.1

(a) For \(\tau \in (0,\infty )\), the Hölder space \(C^\tau ({\mathbb {R}}^n)\) is the set of all functions f with

$$\begin{aligned} \Vert f\Vert _{C^\tau }:=\displaystyle \sum _{|\alpha |\le [\tau ]}\Vert \partial ^\alpha _{x}f\Vert _{L^\infty ({\mathbb {R}}^n)}+\displaystyle \sum _{|\alpha |= [\tau ]}\sup _{x\ne y}\frac{|\partial ^\alpha _{x}f(x)-\partial ^\alpha _{x}f(y)|}{|x-y|^{\tau -[\tau ]}}<\infty .\quad \end{aligned}$$
(3.1)

(b) For \(\tau \in {\mathbb {R}}\), the Zygmund space \(C^\tau _{*}({\mathbb {R}}^n)\) consists of all functions f with

$$\begin{aligned} \Vert f\Vert _{C^\tau _*}=\sup _j 2^{j\tau }\Vert \psi _j (D) f \Vert _{L^\infty }<\infty . \end{aligned}$$
(3.2)

Here, \(\psi _j(D)\) is the Fourier multiplier with symbol \(\psi _j\), i.e. \(\psi _j(D)u = {\mathcal {F}}^{-1}\psi _j{\mathcal {F}} u\), where \((\mathcal Fu)(\xi ) = (2\pi )^{-n/2} \int e^{-ix\xi } u(x)\, d^nx\) is the Fourier transform.

We have the following relations: \(C^\tau =C^\tau _{*}\) if \(\tau \notin {\mathbb {N}}\), and \(C^\tau \subset C^\tau _{*}\) if \(\tau \in {\mathbb {N}}\).

We next introduce symbol classes of finite Hölder or Zygmund regularity, following Taylor [61]. We use the notation \(\langle \xi \rangle :=(1+|\xi |^2)^{\frac{1}{2}}\), \(\xi \in \mathbb R^n\).

Definition 3.2

(a) Let \(0\le \delta <1\). A symbol \(p(x,\xi )\) belongs to \(C^\tau _* S^{m}_{1,\delta }:=C^\tau _*S^m_{1,\delta }({\mathbb {R}}^n\times {\mathbb {R}}^n)\) if

$$\begin{aligned} \Vert D^{\alpha }_{\xi }p(\cdot ,\xi )\Vert _{C_*^{\tau }}\le C_\alpha \langle \xi \rangle ^{m-|\alpha |+\tau \delta } \text { and } |D^{\alpha }_{\xi }p(x,\xi )|\le C_\alpha \langle \xi \rangle ^{m-|\alpha |}. \end{aligned}$$

(b) We obtain the symbol class \(C^\tau S^{m}_{1,\delta }:=C^\tau S^m_{1,\delta }({\mathbb {R}}^n\times {\mathbb {R}}^n)\) for \(\tau >0\) by requiring that

$$\begin{aligned} \Vert D^{\alpha }_{\xi }p(\cdot ,\xi )\Vert _{C^{s}} \le C_\alpha \langle \xi \rangle ^{m-|\alpha |+s\delta }, \quad 0\le s\le \tau . \end{aligned}$$

(c) A symbol \(p(x,\xi )\) is in \(C^{\tau }S_{cl}^{m}\) provided \(p(x,\xi )\in C^{\tau }S^{m}_{1,0}\) and \(p(x,\xi )\) has a classical expansion

$$\begin{aligned} p(x,\xi )\sim \sum _{j\ge 0}p_{m-j}(x,\xi ) \end{aligned}$$

in terms \(p_{m-j}\) homogeneous of degree \(m-j\) in \(\xi \) for \(|\xi |\ge 1\), in the sense that the difference between \(p(x,\xi )\) and the sum over \(0\le j< N\) belongs to \(C^{\tau }S^{m-N}_{1,0}\).

The pseudodifferential operator \(p(x,D_x)\) with the symbol \(p(x,\xi )\in C^\tau S^m_{1,\delta } \) is given by

$$\begin{aligned} \left( p(x,D_x)u\right) (x)=(2\pi )^{-n/2} \int _{{\mathbb {R}}^n}e^{ix\cdot \xi }p(x,\xi )({{{\mathcal {F}}}}{u})(\xi )d^n\xi , \quad u\in {\mathcal {S}}({\mathbb {R}}^n).\quad \end{aligned}$$
(3.3)

It extends to continuous maps

$$\begin{aligned} p(x,D_x): H^{s+m}({\mathbb {R}}^n)\rightarrow H^{s}({\mathbb {R}}^n), \quad -\tau (1-\delta )<s<\tau . \end{aligned}$$
(3.4)

While it is possible to extend the theory of pseudodifferential operators with non-smooth symbols to manifolds (see [1]), due to the local nature of our results it is a key point of this article that we can work entirely on \({\mathbb {R}}^n\).

3.2 Symbol Smoothing

Given \(p(x,\xi )\in C^{\tau }S_{1,\gamma }^{m}\) and \(\delta \in (\gamma ,1)\) let

$$\begin{aligned} p^{\#}(x,\xi )=\displaystyle \sum _{j=0}^{\infty }J_{\epsilon _{j}}p(x,\xi )\psi _j(\xi ). \end{aligned}$$
(3.5)

Here, \(J_\epsilon \) is the smoothing operator given by \((J_\epsilon f)(x)=(\phi (\epsilon D)f)(x)\) with \(\phi \in C^{\infty }_0({\mathbb {R}}^n)\), \(\phi (\xi )=1\) for \(|\xi |\le 1\), and we take \(\epsilon _j=2^{-j(\delta -\gamma )}\).

Letting \(p^{b}(x,\xi )=p(x,\xi )-p^{\#}(x,\xi )\), we obtain the decomposition

$$\begin{aligned} p(x,\xi )=p^{\#}(x,\xi )+p^{b}(x,\xi ), \end{aligned}$$
(3.6)

where \(p^{\#}(x,\xi )\in S^{m}_{1,\delta }\) and \(p^{b}(x,\xi )\in C^{\tau }S^{m-\tau (\delta -\gamma )}_{1,\delta }\).

The symbol estimates for \(p^\#\) are a consequence of the estimate

$$\begin{aligned} \Vert \partial _x^\beta J_\epsilon f\Vert _{L^\infty }\le {\left\{ \begin{array}{ll} C\Vert f\Vert _{C^\tau } \quad |\beta |\le \tau \\ C\epsilon ^{-(|\beta |-\tau )}\Vert f\Vert _{C^\tau } \quad |\beta |> \tau , \end{array}\right. } \end{aligned}$$

and that \(\epsilon _j=2^{-j(\delta -\gamma )}\). For details, see Proposition 1.3 E and Equation (1.3.21) in [61].

3.3 Microlocal Sobolev Regularity

Let \(p\in C^\tau S^{m}_{\rho ,\delta }\), \(\tau >0\), with \(\delta <\rho \). Suppose that there is a conic neighbourhood \(\Gamma \) of \((x_0,\xi _0)\) and constants \(c,C>0\) such that \(|p(x,\xi )|\ge c|\xi |^m\) for \((x,\xi ) \in \Gamma \), \(|\xi |\ge C\). Then, \((x_0,\xi _0)\) is called non-characteristic. If p has a homogeneous principal symbol \(p_m\), the condition is equivalent to \(p_m(x_0,\xi _0)\ne 0\). The complement of the set of non-characteristic points is the set of characteristic points denoted by \({{{\,\textrm{Char}\,}}}(p)\).

A distribution u is microlocally in \(H^{s}\) at \((x_0,\xi _0)\in T^*M\backslash 0\) if there exists a conic neighbourhood \({\Gamma _0}\) of \(\xi _0\) and a smooth function \(\varphi \in C_{0}^{\infty }(M)\) with \(\varphi (x_0)\ne 0\) such that

$$\begin{aligned} \int _{{\Gamma _0}}\langle \xi \rangle ^{2s}|{{{\mathcal {F}}}}(\varphi u)(\xi )|^{2}d^{n}\xi <\infty . \end{aligned}$$

Otherwise we say that \((x_0,\xi _0)\) lies in the \(H^s\)-wavefront set \(WF^s(u)\).

If u is microlocally in \(H^{s}\) in an open conic subset \({\Gamma }\subset T^*M\backslash 0\), we write \(u\in H^s_{mcl}({\Gamma })\).

3.4 Propagation of Singularities for Bisolutions of the Klein–Gordon Operator

A globally hyperbolic spacetime is of the form \({\mathbb {R}}\times \Sigma \), where \(\Sigma \) is not assumed to be compact, and we will write local coordinates in the form

$$\begin{aligned} \tilde{x}= (t,x), \tilde{y}=(s,y) \end{aligned}$$
(3.7)

and the associated covariables as

$$\begin{aligned} \tilde{\xi }= (\xi ^0,\xi ), \tilde{\eta }= (\eta ^{0},\eta ). \end{aligned}$$
(3.8)

On the product \(({\mathbb {R}}\times \Sigma )\times ( {\mathbb {R}}\times \Sigma )\), we use \((\mathbf{{x}},{\varvec{\xi }})\) with

$$\begin{aligned} \mathbf{{x}}= (\tilde{x},\tilde{y}), {\varvec{\xi }}=(\tilde{\xi },\tilde{\eta }). \end{aligned}$$
(3.9)

In the sequel, we shall apply the Klein–Gordon operator also to functions and distributions on \(M\times M\). Using the coordinates in Eqs. (3.7), (3.8) and (3.9), we distinguish the cases, where P acts on the first set of variables (tx) or on the second set (sy), and write \(P_{(t,x)}\) and \(P_{(s,y)}\), respectively. Explicitly,

$$\begin{aligned} P_{(t,x)}(\mathbf{{x}},D_{\mathbf{{x}}})= & {} P_{(t,x)}(\tilde{x},D_{\tilde{x}},\tilde{y},D_{\tilde{y}})=(\square _{g(\tilde{x})}+m^2)\otimes I\\ P_{(s,y)}(\mathbf{{x}},D_{\mathbf{{x}}})= & {} P_{(s,y)}(\tilde{x},D_{\tilde{x}},\tilde{y},D_{\tilde{y}})=I\otimes (\square _{g(\tilde{y})}+m^2) \end{aligned}$$

In particular,

$$\begin{aligned} {{\,\textrm{Char}\,}}(P_{(t,x)})= & {} {{\,\textrm{Char}\,}}(P)\times T^{*}M\cup \{(\mathbf{{x}},{\varvec{\xi }})\in T^{*}(M\times M)\backslash 0, \tilde{\xi }=0\}\nonumber \\ {{\,\textrm{Char}\,}}(P_{(s,y)})= & {} T^{*}M\times {{\,\textrm{Char}\,}}(P)\cup \{(\mathbf{{x}},{\varvec{\xi }})\in T^{*}(M\times M)\backslash 0, \tilde{\eta }=0\}.\nonumber \\ \end{aligned}$$
(3.10)

Theorem 3.3

Let the metric g be of class \(C^\tau \), \(\tau >1\), \(0\le \sigma <\tau -1\) and \(v\in H_{loc}^{2+\sigma -\tau +\epsilon }(M\times M)\) for some \({\epsilon }>0\) with \(P_{(t,x)}(\mathbf{{x}},D_\mathbf{{x}}){v}=0\). Then,

$$\begin{aligned} WF^{\sigma +2}({v})\subset {{\,\textrm{Char}\,}}(P_{(t,x)}). \end{aligned}$$

Proof

Being interested in the wavefront set of v near a point \(\mathbf{{x}}\), we multiply v by a function \(\varphi \in {\mathcal {D}}(M\times M)\) with \(\varphi \equiv 1\) near \(\mathbf{{x}}\) and consider \(\varphi v\). So we can assume that v has support in a small neighbourhood of \(\mathbf{{x}}\) contained in a single coordinate patch and consider v as an element of \(H^{2+\sigma -\tau +\epsilon }({\mathbb {R}}^4\times {\mathbb {R}}^4)\). In order to distinguish points \((\mathbf{{x}},{\varvec{\xi }})=(\tilde{x},\tilde{\xi },\tilde{y},\tilde{\eta })\) from their representation in local coordinates, we will write the latter in the form \((\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})=(\underline{\tilde{x}}, \underline{\tilde{\xi }}, \underline{\tilde{y}},\underline{\tilde{\eta }})\). In this local setting, \(P_{(t,x)}(\underline{\mathbf{{x}}},D_{\underline{\mathbf{{x}}}})\) is given by the symbol

$$\begin{aligned} P_{(t,x)}(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})=P_{(t,x)}(\underline{\tilde{x}}, \underline{\tilde{\xi }},\underline{\tilde{y}},\underline{\tilde{\eta }})=\underbrace{g^{\mu \nu } (\underline{\tilde{x}})\underline{\xi _\mu }\underline{\xi _\nu }}_{p_2(\underline{\mathbf{{x}}}, \underline{{\varvec{\xi }}})}+\underbrace{ig^{\mu \nu }(\underline{x}) \Gamma ^\rho _{\mu \nu }(\underline{x})\underline{\xi }_\rho }_{p_1(\underline{\mathbf{{x}}}, \underline{{\varvec{\xi }}})}+\underbrace{m^2}_{p_0(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})}.\nonumber \\ \end{aligned}$$
(3.11)

The symbol smoothing (Eq. (3.6)) on \(p_2,p_1\) gives a decomposition

$$\begin{aligned} p_2(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})&=p_2^{\#}(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})+p_2^{b}(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})\\ p_1(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})&=p_1^{\#}(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})+p_1^{b}(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})\nonumber \\ P_{(t,x)}(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})&=(p_2^{\#}(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})+p_1^{\#}(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}}))+p_2^{b}(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})+p_1^{b}(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})+p_0(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})\nonumber \\&=q^{\#}(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})+p_2^{b}(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})+p_1^{b}(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}}), \end{aligned}$$

where

$$\begin{aligned}{} & {} q^{\#}(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})=(p_2^{\#}(\underline{\mathbf{{x}}}, \underline{{\varvec{\xi }}})+p_1^{\#}(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})+p_0 (\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}}))\in S^{2}_{1,\delta }({\mathbb {R}}^8\times {\mathbb {R}}^8), \end{aligned}$$
(3.12)
$$\begin{aligned}{} & {} p_2^{b}(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})\in C^{\tau }S^{2-\tau \delta }_{1,\delta }({\mathbb {R}}^8\times {\mathbb {R}}^8) \quad p_1^{b}(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})\in C^{\tau -1}S^{1-(\tau -1)\delta }_{1,\delta }({\mathbb {R}}^8\times {\mathbb {R}}^8).\nonumber \\ \end{aligned}$$
(3.13)

Taking \(0\le \delta < 1\) so close to 1 that \(2-\tau \delta <2-\tau +\epsilon \) we have \({v}\in H^{2+\sigma -\tau \delta }({\mathbb {R}}^4\times {\mathbb {R}}^4)\) (notice this implies \({v}\in H^{1+\sigma -(\tau -1)\delta }({\mathbb {R}}^4\times {\mathbb {R}}^4)\)), and we have

$$\begin{aligned} q^{\#}(\underline{\mathbf{{x}}},D_{\underline{\mathbf{{x}}}})v=-(p_2^{b} (\underline{\mathbf{{x}}},D_{\underline{\mathbf{{x}}}})+p_1^{b}(\underline{\mathbf{{x}}}, D_{\underline{\mathbf{{x}}}}))v=f, \end{aligned}$$
(3.14)

where \(f\in H^{\sigma }({\mathbb {R}}^4\times {\mathbb {R}}^4)\), since \(p_2^b(\underline{\mathbf{{x}}},D_{\underline{\mathbf{{x}}}}) v\in H^{\sigma }({\mathbb {R}}^4\times {\mathbb {R}}^4)\) and \(p_1^b(\underline{\mathbf{{x}}},D_{\underline{\mathbf{{x}}}}) v\in H^{\sigma +1-\delta }({\mathbb {R}}^4\times {\mathbb {R}}^4)\).

Now if \((\underline{\tilde{x}_0},\underline{\tilde{\xi }_0},\underline{\tilde{y}_0}, \underline{\tilde{\eta }_0})\notin {{\,\textrm{Char}\,}}(P_{(t,x)})\), there are \(C,c>0\) such that

$$\begin{aligned} |P_{(t,x)}(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})|\ge c |\underline{{\varvec{\xi }}}|^2 \text { for } |\underline{{\varvec{\xi }}}|\ge C \end{aligned}$$

in a conical neighbourhood \(\Gamma \) that contains \((\underline{\tilde{x}_0},\underline{\tilde{\xi }_0},\underline{\tilde{y}_0},\underline{\tilde{\eta }_0})\).

Since \(p_2^{b}(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})\in C^{\tau }S^{2-\tau \delta }_{1,\delta }\) and \(p_1^{b}(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})\in C^{\tau -1}S^{1-(\tau -1)\delta }_{1,\delta }\), there exists a \(\tilde{C}>0\) such that

$$\begin{aligned} |q^{\#}(\underline{\mathbf{{x}}},\underline{{\varvec{\xi }}})|&\ge C (1+|\underline{{\varvec{\xi }}}|^2)-(1+|\underline{{\varvec{\xi }}}|^2)^{\frac{2-\tau \delta }{2}}-(1+|\underline{{\varvec{\xi }}}|^2)^{\frac{1-(\tau -1)\delta }{2}}\\&\ge \tilde{C} (1+|\underline{{\varvec{\xi }}}|^2) \text { for large } |\underline{{\varvec{\xi }}}|. \end{aligned}$$

Therefore, \((\underline{\tilde{x}_0},\underline{\tilde{\xi }_0},\underline{\tilde{y}_0},\underline{\tilde{\eta }_0})\notin {{\,\textrm{Char}\,}}(q^{\#})\).

Since \(q^{\#}\in S^{2}_{1,\delta }\) and \((\underline{\tilde{x}_0},\underline{\tilde{\xi }_0},\underline{\tilde{y}_0},\underline{\tilde{\eta }_0})\notin {{\,\textrm{Char}\,}}(q^{\#}) \), there is a microlocal parametrix with symbol \(\tilde{q}\in S^{-2}_{1,\delta }({\mathbb {R}}^8\times {\mathbb {R}}^8)\) such that

$$\begin{aligned} v+r(\underline{\mathbf{{x}}},D_{\underline{\mathbf{{x}}}})v&=\tilde{q} (\underline{\mathbf{{x}}},D_{\underline{\mathbf{{x}}}})q^{\#}(\underline{\mathbf{{x}}}, D_{\underline{\mathbf{{x}}}})v =\tilde{q}(\underline{\mathbf{{x}}},D_{\underline{\mathbf{{x}}}})f, \end{aligned}$$

where \((\underline{\tilde{x}_0},\underline{\tilde{\xi }_0},\underline{\tilde{y}_0},\underline{\tilde{\eta }_0})\notin WF(r(\underline{\mathbf{{x}}},D_{\underline{\mathbf{{x}}}})v)\) and \(\tilde{q}(\underline{\mathbf{{x}}},D_{\underline{\mathbf{{x}}}})f\in H^{\sigma +2}({\mathbb {R}}^4\times {\mathbb {R}}^4)\) which shows that \((\underline{\tilde{x}_0},\underline{\tilde{\xi }_0},\underline{\tilde{y}_0},\underline{\tilde{\eta }_0})\notin WF^{\sigma +2}({\tilde{q}}(\underline{\mathbf{{x}}},D_{\underline{\mathbf{{x}}}})f)\). Since

$$\begin{aligned} WF^{\sigma +2}(v)\subset WF^{\sigma +2}({\tilde{q}}(\underline{\mathbf{{x}}},D_{\underline{\mathbf{{x}}}})f) \cup WF(r(\underline{\mathbf{{x}}},D_{\underline{\mathbf{{x}}}})v), \end{aligned}$$
(3.15)

we see that \((\underline{\tilde{x}_0},\underline{\tilde{\xi }_0},\underline{\tilde{y}_0},\underline{\tilde{\eta }_0)}\notin WF^{\sigma +2}(v)\).

By definition of the wavefront set, this means that \((\mathbf{{x}}_0,{\varvec{\xi }}_0)\) is not in the wavefront set of v, considered as a distribution on \(M\times M\). \(\square \)

Remark 3.4

In the proof presented above, we showed that the microlocal results are local estimates, which can be done within a chart in the cotangent bundle \(T^*{\mathbb {R}}^8\). To streamline the discussion and avoid frequently alternating between the notation of the chart and the manifold, we will forego this distinction in Section 5. However, it is important to bear in mind that the proofs in that section are analogous to the one detailed above, involving localization within a chart.

Remark 3.5

Applying the symbol smoothing directly to \(P_{(t,x)}\in C^{\tau -1}S^2_{1,0}\) would leave us with \(P^b_{(t,x)}\in C^{\tau -1} S^{2-(\tau -1)\delta }_{1,\delta }\). The advantage of the decomposition in Theorem 3.3 with \(p_1^b\in C^{\tau -1}S^{1-(\tau -1)\delta }_{1,\delta }\) and \(p^b_2\in C^\tau S^{2-\tau \delta }_{1,\delta }\) is that the associated operators map a given \(u\in H^{2+s-\tau \delta }\) to \(H^s\) and \(H^{s+1-\delta }\), respectively, for \(-(1-\delta )(\tau -1)<s<\tau -1\), so that the sum is in \(H^s\) instead of \(H^{s-\delta }\).

The theorem, below, will be crucial for our main result. Proofs can be found in [61, Proposition 6.1.D] or [62, Proposition 11.4]. In [62, p.215], Taylor points out that Zygmund regularity \(C_*^2\) for the metric suffices.

Theorem 3.6

Let \(u\in {\mathcal {D}}'(M\times M)\) solve \(P_{(t,x)}u=f\). Let \(\gamma \) be an integral curve of the Hamiltonian vector field \(H_{p_{2}}\) with \(p_2\) as in Eq. (3.11). If for some \(s\in {\mathbb {R}}\), we have \(f\in H_{mcl}^s({\Gamma })\) and \(P_{(t,x)}^b u\in H_{mcl}^s({ \Gamma )}\), where \(\gamma \subset {\Gamma }\) with \({\Gamma }\) a conical neighbourhood and \(u\in H_{mcl}^{s+1}(\gamma (0))\), then \(u\in H_{mcl}^{s+1}({\gamma })\).

Remark 3.7

If \(u\in H_{comp}^{2+s-\tau \delta }\), then \(P_{(t,x)}^b u\in H^s\), see Remark (3.5). Moreover, using the divergence structure of the operator one can show that, if \(u\in H_{comp}^{1+s-\tau \delta }, f\in H^{s-1}, u\in H_{mcl}^{s}({\gamma }(0))\), then \(u\in H_{mcl}^{s}({\gamma })\) for \(-2(1-\delta )<s\le 2\); see [62, p.210] for details.

Remark 3.8

Notice that the \(s\in {\mathbb {R}}\) is constrained by the microlocal regularity of \(P^b_{(t,x)}u\) and not only that of f. In fact, one can use the stronger hypothesis that \(u\in H^{s-\tau \delta }_{comp} (U)\) for a suitable domain U, regularity \(\tau \) and \(\delta \in (0,1)\) in order to guarantee that \(P^b_{(t,x)}u\in H^s(U)\subset H^s_{mcl}(\Gamma )\)

4 Support and Global Regularity of \(K_G\)

The following two lemmas contain the main results of this section. The first lemma shows that only causally connected points belong to the support of \(K_G\). The second lemma establishes that \(K_G\in H_{loc}^{-1-\epsilon }(M\times M)\).

Lemma 4.1

Let \((\tilde{x},\tilde{y})\in M\times M\) be such that \({\tilde{x}}\) and \({\tilde{y}}\) are not causally related, i.e. \(\tilde{x}\notin J(\tilde{y})\). Then, \((\tilde{x},\tilde{y})\notin {{\,\textrm{supp}\,}}(K_G)\).

Proof

Since the support of \(K_G\) is the complement of the largest open set where \(K_G\) vanishes, it is enough to show that there are open neighbourhoods V of \({\tilde{x}}\) and U of \({\tilde{y}}\) such that \(K_G\) vanishes in \(W=V\times U\).

We construct the sets V and U as follows: For globally hyperbolic spacetimes, there exist a time function and a foliation by Cauchy surfaces, i.e. \(M={\mathbb {R}}\times \Sigma \), see [8, Theorem 1.1], [53, Theorem 5.9]. Let \(\tilde{x}\in \{t\}\times \Sigma \) and \(\tilde{y}\in \{s\}\times \Sigma \). Without loss of generality, we assume \(t\le s\). Since M is globally hyperbolic, \(J(\tilde{y})\cap (\{t\}\times \Sigma )\) is compact and by hypothesis does not contain \(\tilde{x}\). Therefore, there exists a neighbourhood \(\tilde{V}\) of \(\tilde{x}\) in \(\{t\}\times \Sigma \) such that \(\overline{\tilde{V}}\cap (J(\tilde{y})\cap (\{t\}\times \Sigma ))=\emptyset \). By symmetry, \(\tilde{y}\notin J(\overline{{\tilde{V}})}\cap (\{s\}\times \Sigma )=\overline{J({\tilde{V}})} \cap (\{s\}\times \Sigma )\), and we thus also find a neighbourhood \({\tilde{U}}\) of \(\tilde{y}\) in \(\{s\}\times \Sigma \) such that \({\tilde{U}}\cap J({\tilde{V}}) \cap (\{s\}\times \Sigma ) = \emptyset \).

Now we consider the total domain of dependence of both sets, i.e. \(D(\tilde{U})\) and \(D(\tilde{V})\).Footnote 1 Notice that \(J(D(\tilde{V}))\cap D(\tilde{U})=\emptyset \) and \(J(D(\tilde{U}))\cap D(\tilde{V})=\emptyset \). Otherwise, we could construct a causal curve between \(\tilde{U}\) and \(\tilde{V}\). We define \(V:=\text {Int}D(\tilde{V})\) and \(U:=\text {Int}D(\tilde{U})\), see Fig. 1.

Fig. 1
figure 1

\(U\cap J(V)=\emptyset \) and \(V\cap J(U)=\emptyset \)

Now we show that \(K_G\) vanishes in \(W=V\times U\): Choose smooth functions \(\psi \) and \(\phi \) with \({{\,\textrm{supp}\,}}(\psi )\subset V\) and \({{\,\textrm{supp}\,}}(\phi )\subset U\). Then,

$$\begin{aligned} K_G(\psi \otimes \phi )&=\langle G(\psi ),\phi \rangle =\int _MG(\psi )\phi \sqrt{g}dx\\&=\int _{J({{\,\textrm{supp}\,}}(\psi ))\cap {{\,\textrm{supp}\,}}(\phi )}G(\psi )\phi \sqrt{g}dx\\&=\int _{J(V)\cap U}G(\psi )\phi \sqrt{g}dx=0. \end{aligned}$$

\(\square \)

Remark 4.2

Notice that a totally analogous proof shows that \((\tilde{x},\tilde{y})\notin {{\,\textrm{supp}\,}}(K_{G^\pm })\) if \(\tilde{x}\notin J^\pm (\tilde{y})\).

Regarding the global regularity of the causal propagator for \(C^{1,1}\) globally hyperbolic spacetimes, we find a slightly weaker result compared to the smooth case. Nevertheless, in the ultrastatic setting we show that the same regularity as in the smooth setting holds (Lemma 6.11).

Lemma 4.3

Let (Mg) be a \(C^{1,1}\)-globally hyperbolic spacetime. Then \(K_G\in H_{loc}^{-1-\epsilon }(M\times M)\) for every \(\epsilon >0\).

Proof

We have to show that, given \(\psi _1,\psi _2\in {{{\mathcal {D}}}}(M)\), the Schwartz kernel of the product \(\psi _2G\psi _1\) is in \(H^{-1-\varepsilon }(M)\) for every \(\varepsilon >0\). Since the proof is local, we may assume (using possibly disconnected coordinate charts) that \(\psi _1\) and \(\psi _2\) have their support in the same coordinate neighbourhood for M. We will therefore work in \({\mathbb {R}}^4\), using the notation \(\psi _1, \psi _2\) and G also for the representations in local coordinates. In order to distinguish the standard variables and covariables on \({\mathbb {R}}^4\) from those chosen for M we shall denote them by \(\underline{x}\), \(\underline{\xi }\), etc. Moreover, we choose \(\psi _3,\psi _4\in {{{\mathcal {D}}}}({\mathbb {R}}^4)\) supported in the same coordinate chart, satisfying \(\psi _3\psi _2=\psi _2\) and \(\psi _4\psi _3=\psi _3\). Finally, we denote by \(\Lambda ^s\), \(s\in {\mathbb {R}}\), the pseudodifferential operator of order s with symbol \((1+|\underline{\xi }|^2)^{s/2}\) on \({\mathbb {R}}^4\).

We have

$$\begin{aligned} \psi _2 G\psi _1=\Lambda ^{1+\varepsilon }\Lambda ^{-1-\varepsilon }\psi _3\psi _2G\psi _1 =\Lambda ^{1+\varepsilon } (\psi _4+(1-\psi _4))\Lambda ^{-1-\varepsilon }\psi _3\psi _2 G\psi _1.\nonumber \\ \end{aligned}$$
(4.1)

The operator \(\psi _4\Lambda ^{-1-\varepsilon }\psi _3\psi _2 G\psi _1\) maps \(H^1({\mathbb {R}}^4)\) to \(H^{3+\varepsilon }_{comp}({\mathbb {R}}^4)\) and therefore is a Hilbert–Schmidt operator. Hence, it has an integral kernel in \(L^2({\mathbb {R}}^4\times {\mathbb {R}}^4)\). The operator \((1-\psi _4)\Lambda ^{-1-\varepsilon }\psi _3\) is obviously smoothing, since \(1-\psi _4\) and \(\psi _3\) have disjoint support. Hence, it maps \(H^2({\mathbb {R}}^4)\) to \(H^\infty ({\mathbb {R}}^4)= \bigcap _sH^s({\mathbb {R}}^4)\). But more is true: In the identity

$$\begin{aligned} \underline{x}_j(1-\psi _4)\Lambda ^{-1-\varepsilon } \psi _3 = (1-\psi _4) \Lambda ^{-1-\varepsilon } \underline{x}_j\psi _3 + (1-\psi _4)[\underline{x}_j,\Lambda ^{-1-\varepsilon }]\psi _3, \end{aligned}$$

both operators on the right hand side map \(H^2({\mathbb {R}}^4)\) to \(H^\infty ({\mathbb {R}}^4)\) (recall that \([\underline{x}_j,\Lambda ^{-1-\varepsilon }]\) has the symbol \(D_{\underline{\xi }_j}(1+|\underline{\xi }|^2)^{-1-\varepsilon }\)). Iterating this identity, we find that \((1+|\underline{x}|^{2N})(1-\psi _4)\Lambda ^{-1-\varepsilon } \psi _3\in {\mathcal {B}}(H^2({\mathbb {R}}^4),H^\infty ({\mathbb {R}}^4))\) for every \(N\in {\mathbb {N}}\). Hence, \((1-\psi _4)\Lambda ^{-1-\varepsilon } \psi _3\) maps \(H^2({\mathbb {R}}^4)\) to \({\mathcal {S}} ({\mathbb {R}}^4)\).

Therefore, it also has an integral kernel in \(L^2({\mathbb {R}}^4\times {\mathbb {R}}^4)\). Denote for the moment the \(L^2\)-integral kernel of \(\Lambda ^{-1-\varepsilon } \psi _3\psi _2G\psi _1\) by \(k_A=k_A(\underline{x},\underline{y})\). Then, the kernel \(k=k(\underline{x},\underline{y})\) of \(\psi _2G\psi _1\) is given by

$$\begin{aligned} \Lambda ^{1+\varepsilon }_{(\underline{x})} k_A(\underline{x},\underline{y}). \end{aligned}$$

Here, the notation \(\Lambda ^{1+\varepsilon }_{(x)} \) indicates that we view \(\Lambda ^{1+\varepsilon } \) as an operator on \({\mathbb {R}}^4\times {\mathbb {R}}^4\) that acts only with respect to the first copy of \({\mathbb {R}}^4\). In this sense, it is a pseudodifferential operator with symbol in the Hörmander class \(S^{1+\varepsilon }_{0,0}\) and thus maps \(L^2({\mathbb {R}}^4\times {\mathbb {R}}^4)\) to \(H^{-1-\varepsilon }({\mathbb {R}}^4\times {\mathbb {R}}^4)\). This shows the assertion. \(\square \)

Remark 4.4

Notice that since only the mapping properties of G were used we have also that \(K_{G^+},K_{G^-}\in H^{-1-\epsilon }_{loc}(M\times M)\).

5 Proof of the Main Theorems

A globally hyperbolic spacetime is given by a family of Riemannian metrics \(\{h_t\}_{t\in {\mathbb {R}}}\) on \(\Sigma \) and a function \(\beta (x,t)>0\) such that the spacetime metric (Mg), where \(M={\mathbb {R}}\times \Sigma \), is given by

$$\begin{aligned} ds^2=\beta ^2(t,x)dt^2-h_t, \end{aligned}$$
(5.1)

see [9, Theorem 1.1]. We will assume that the regularity of the spacetime metric g is \(C^\tau \).

In this section, we will prove the following results:

Theorem 5.1

Let (Mg) be a \(C^{\tau }\) globally hyperbolic spacetime with \(\tau >2\) and \(K_G\) the causal propagator of the Klein–Gordon operator P. Then,

$$\begin{aligned} WF'^{-2+\tau -{\epsilon }}(K_G)\subset C \end{aligned}$$

for every \({\epsilon }>0\), C as in Eq. (1.1).

Theorem 5.2

For a \(C^{\tau +2}\) globally hyperbolic spacetime with \(\tau >2\),

$$\begin{aligned} C\subset WF'^{-\frac{1}{2}}(K_G)\subset WF'^{\tau -\epsilon }(K_G)\subset C \end{aligned}$$

holds for \(0<\epsilon <\tau +\frac{1}{2}\).

Remark 5.3

In the non-smooth case, we cannot expect \(G(f)\in C^\infty (M)\) even if \(f\in {{{\mathcal {D}}}}(M)\) as a consequence of the fact that G(f) solves the homogeneous Cauchy problem. We know from [38, Proposition B.8] that for \(f\in {{\mathcal {D}}}(M)\),

$$\begin{aligned} WF^s(G(f))\subset \{(\tilde{x},\tilde{\xi })\in T^*M;(\tilde{x},\tilde{\xi },\tilde{y},0)\in WF^s(K_G) \text { for some } y\in M\}. \end{aligned}$$

Therefore, \(WF'^s(K_G)\) might contain points that are not in C.

Remark 5.4

Since \(K_G\) is antisymmetric, we have that for \(\rho (\tilde{x},\tilde{y})=(\tilde{y},\tilde{x})\), \(\rho ^*K_G=-K_G\). This implies that if \((\tilde{x},\tilde{\xi },\tilde{y},0)\in WF^s(K_G) \text { for some } y\in M\), then \((\tilde{y},0,\tilde{x},\tilde{\xi })\in WF^s(K_G) \text { for some } y\in M\).

5.1 Proof of Theorem 5.1

Let \(u\in H_{comp}^{1+s-\tau \delta }(M\times M)\) satisfy \(P_{(t,x)}(\mathbf{{x}},D_\mathbf{{x}})u=0\).

Then also

$$\begin{aligned} \partial _\nu \left( \sqrt{|g|}g^{\mu \nu }\partial _\mu u\right) =0. \end{aligned}$$

Using the decomposition \(\sqrt{|g|}g^{\mu \nu }\partial _\mu =(\sqrt{|g|}g^{\mu \nu }\partial _\mu )^\#+(\sqrt{|g|}g^{\mu \nu }\partial _\mu )^b\), we obtain

$$\begin{aligned} P_{(t,x)}(\mathbf{{x}},D_\mathbf{{x}})u=\frac{1}{\sqrt{|g|}}\partial _\nu \left( (\sqrt{|g|}g^{\mu \nu }\partial _\mu )^\#u+(\sqrt{|g|}g^{\mu \nu }\partial _\mu )^bu\right) . \end{aligned}$$
(5.2)

We state the behaviour outside the characteristic in this setting.

Lemma 5.5

For \(\tau >2\) and any \(\tilde{\epsilon }>0\),

$$\begin{aligned} WF^{-1-\tilde{\epsilon }+\tau }(K_G)\subset {{\,\textrm{Char}\,}}(P_{(t, x)})\cap {{\,\textrm{Char}\,}}(P_{(s, y)}). \end{aligned}$$
(5.3)

Proof

As the statement is microlocal, we can work in local coordinates in \(T^*({\mathbb {R}}^4\times {\mathbb {R}}^4)\) and consider \(\varphi K_G\) for \(\varphi \in {{{\mathcal {D}}}}({\mathbb {R}}^4\times {\mathbb {R}}^4)\) with \(\varphi =1\) near \(\mathbf{{x}}_0\).

Let \((\mathbf{{x}}_0,{\varvec{\xi }}_0)=(\tilde{x}_0,\tilde{\xi }_0,\tilde{y}_0,\tilde{\eta }_0)\not \in {{\,\textrm{Char}\,}}(P_{(t,x)})\)Footnote 2 Then, \(0<\sqrt{|g(\tilde{x})|}\) and \(|g^{\mu \nu }(\tilde{x})\sqrt{|g(\tilde{x})|}\xi _\mu \xi _\nu |\ge C|{\varvec{\xi }}|^2\) for suitable \(C>0\) in a conic neighbourhood of \((\mathbf{{x}}_0,{\varvec{\xi }}_0)\).

In particular, \((\mathbf{{x}}_0,{\varvec{\xi }}_0)\not \in {{\,\textrm{Char}\,}}(\partial _{\nu }(\sqrt{|g|}g^{\mu \nu }\partial _\mu )^\#)\), so there exists a microlocal parametrix \(\tilde{q}\in S^{-2}_{1,\delta }\) such that

$$\begin{aligned} \tilde{q}\partial _{\nu }(\sqrt{|g|}g^{\mu \nu }\partial _\mu )^\#=I+ r, \end{aligned}$$
(5.4)

where \(r(\mathbf{{x}},D_{\mathbf{{x}}})\) is microlocally smoothing near \((\mathbf{{x}}_0,{\varvec{\xi }}_0)\).

Since \(P_{(t,x)}(\mathbf{{x}}, D_\mathbf{{x}})K_G=0\), we have near \(\mathbf{{x}}_0\)

$$\begin{aligned} 0&=\partial _\nu (\sqrt{|g|}g^{\mu \nu }\partial _\mu )K_G \end{aligned}$$
(5.5)
$$\begin{aligned}&=\partial _\nu (\sqrt{|g|}g^{\mu \nu }\partial _\mu )^\#\varphi K_G+\partial _\nu (\sqrt{|g|}g^{\mu \nu }\partial _\mu )^b\varphi K_G, \end{aligned}$$
(5.6)

Since \((\sqrt{|g|}g^{\mu \nu }\xi _\mu )^b\in C^\tau S^{1-\tau \delta }_{1,\delta }\) for every \(0\le \delta <1\), we obtain a bounded map

$$\begin{aligned} \partial _\nu (\sqrt{|g|}g^{\mu \nu }\partial _\mu )^b:H^{s+1-\tau \delta }({\mathbb {R}}^4\times {\mathbb {R}}^4)\rightarrow H^{s-1}({\mathbb {R}}^4\times {\mathbb {R}}^4), \end{aligned}$$
(5.7)

\(-(1-\delta )\tau<s<\tau \delta \).

Since \(K_G\in H^{-1-\epsilon }_{loc}(M\times M)\) for every \(\epsilon >0\) by Lemma 4.3, we can choose \(\delta \) such that \(s=-2+\tau \delta -\epsilon >0\) so that by Eq. (5.5), we have locally

$$\begin{aligned} \partial _\nu (\sqrt{|g|}g^{\mu \nu }\partial _\mu )^\#\varphi K_G=-\partial _\nu (\sqrt{|g|}g^{\mu \nu }\partial _\mu )^b\varphi K_G\in H^{-3+\tau \delta -\epsilon }({\mathbb {R}}^4\times {\mathbb {R}}^4).\nonumber \\ \end{aligned}$$
(5.8)

Applying the microlocal parametrix \(\tilde{q}\), we obtain

$$\begin{aligned} \tilde{q}\partial _\nu (\sqrt{|g|}g^{\mu \nu }\partial _\mu )^\#\varphi K_G\in H^{-1+\tau \delta -\epsilon }({\mathbb {R}}^4\times {\mathbb {R}}^4). \end{aligned}$$
(5.9)

By Eq. (5.4), Eq. (5.9) equals

$$\begin{aligned} (I+r(\mathbf{{x}},D_{\mathbf{{x}}}))\varphi K_G. \end{aligned}$$
(5.10)

Hence, \(K_G\in H^{-1+\tau \delta -\epsilon }(M\times M)\) microlocally near \((\mathbf{{x}}_0,{\varvec{\xi }}_0)\), so that \((\mathbf{{x}}_0,{\varvec{\xi }}_0)\not \in WF^{-1+\tau \delta -\epsilon }(K_G)\) for any \(\epsilon >0, 0\le \delta <1\). Choosing \(\delta \) appropriately, we find that for every \(\tilde{\epsilon }>0\)

$$\begin{aligned} WF^{-1-\tilde{\epsilon }+\tau }(K_G)\subset {{\,\textrm{Char}\,}}(P_{(t,x)}). \end{aligned}$$
(5.11)

Arguing analogously for \(P_{(s,y)}\), we can see that

$$\begin{aligned} WF^{-1+\tau -\tilde{\epsilon }}(K_G)\subset&{{\,\textrm{Char}\,}}(P_{(t, x)})\cap {{\,\textrm{Char}\,}}(P_{(s, y)}). \end{aligned}$$
(5.12)

\(\square \)

Notice that

$$\begin{aligned} {{\,\textrm{Char}\,}}(P_{(t, x)})\cap {{\,\textrm{Char}\,}}(P_{(s, y)})=({{\,\textrm{Char}\,}}(P)\times {{\,\textrm{Char}\,}}(P))\cup {\mathcal {A}}\cup {\mathcal {B}}, \end{aligned}$$

where \({\mathcal {A}}:=\{(\tilde{x},0,\tilde{y},\tilde{\eta })\in T^*(M\times M):(\tilde{y},\tilde{\eta })\in {{\,\textrm{Char}\,}}(P)\}\) and \({\mathcal {B}}:=\{(\tilde{x},\tilde{\xi },\tilde{y},0) \in T^*(M\times M):(\tilde{x},\tilde{\xi })\in {{\,\textrm{Char}\,}}(P)\}\).

We will show now that the sets \({\mathcal {A}}\) and \({\mathcal {B}}\) do not belong to \(WF^{-2+\tau -\tilde{\epsilon }}(K_G)\). Nevertheless, for higher wavefront sets, that may not be the case, see Remarks 5.3 and 5.4.

In order to show the result, we will need the following lemma.

Lemma 5.6

\((\tilde{x},\tilde{\xi },\tilde{x},\mu \tilde{\xi })\notin WF^{-2+\tau -\tilde{\epsilon }}(K_{G^\pm })\) for \(\mu \ne -1\).

Proof

Consider a point \((\tilde{y},\tilde{\eta })\ne (\tilde{x},\tilde{\xi })\) on the null bicharacteristic \(\gamma (\tilde{x},\tilde{\xi })\), with \(\tilde{y}\in J^{-}(\tilde{x})\). Since \(PG^+=I\), it holds

$$\begin{aligned} K_I=K_{PG^+}=P_{(t,x)}K_{G^+} \end{aligned}$$
(5.13)

with wavefront set the conormal to the diagonal. As \(\mu \ne -1\), \((\tilde{x},\tilde{\xi },\tilde{x},\mu \tilde{\xi })\) is not part of it, and neither are the points of the set \(\gamma (\tilde{x},\tilde{\xi })\times \{(\tilde{x},\mu \tilde{\xi })\}\). Hence, there exists an open conic neighbourhood W of the set of all \((\tilde{z},\tilde{\zeta },\tilde{x},\mu \tilde{\xi })\in T^*(M\times M),\) where \((\tilde{z},\tilde{\zeta })\) lies on \(\gamma (\tilde{x},\tilde{\xi })\) between \((\tilde{x},\tilde{\xi })\) and \((\tilde{y},\tilde{\eta })\), that does not intersect \(WF(K_I)\). We can assume that the base point projection \(\Pi (W)\) is relatively compact. We choose \(\varphi \in {{\mathcal {D}}}(M\times M)\) with \(\varphi =1\) on \(\Pi W\). Then,

$$\begin{aligned} \emptyset =WF(K_I)\cap W = WF(P_{(t,x)}K_{G^+})\cap W. \end{aligned}$$
(5.14)

Moreover, \(P_{(t,x)}^{\#}(\varphi K_{G^+})=P_{(t,x)}(\varphi K_{G^+})-P_{(t,x)}^b(\varphi K_{G^+})\).

According to Remark 4.4, \(K_{G^+}\in H^{-1-\epsilon }_{loc}(M\times M)\) for every \(\epsilon >0\), therefore \(P_{(t,x)}^b(\varphi K_{G^+})\in H^{-3-\epsilon +\tau }\). We now apply Theorem 3.6 with \(u=\varphi K_{G^+}, s=-3-\tilde{\epsilon }+\tau \), \(\Gamma =W\), \(f=P_{(t,x)}K_{G^+}\in H^\infty _{mcl}(W)\), \(P_{(t,x)}^b(\varphi K_{G^+})\in H^{s}\). We have \(\varphi K_{G^+}\in H^\infty _{mcl}\) near \((\tilde{y},\tilde{\eta },\tilde{x},\mu \tilde{\xi })\), since \((\tilde{y},\tilde{x})\) is not in the support of \(K_{G^+}\). Hence, Theorem 3.6 implies that \(K_{G^+}\in H^{-2-\epsilon +\tau }_{mcl}\) also in a conic neighbourhood of \((\tilde{x},\tilde{\xi },\tilde{x},\mu \tilde{\xi })\), as this point lies on the integral curve of the Hamiltonian vector field for the principal symbol of \(P_{(t,x)}\). Hence, \((\tilde{x},\tilde{\xi },\tilde{x},\mu \tilde{\xi })\notin WF^{-2+\tau -\epsilon }(K_{G^+})\). In an analogous way, we see that \((\tilde{x},\tilde{\xi },\tilde{x},\mu \tilde{\xi })\notin WF^{-2+\tau -\epsilon } (K_{G^-})\) by considering a point \((\tilde{y},\tilde{\eta })\) on \(\gamma (\tilde{x},\tilde{\xi })\) with \(\tilde{y}\in J^+(\tilde{x})\). \(\square \)

Remark 5.7

Notice that the fact that the wavefront set of \(K_I\) is the conormal to the diagonal does not allow one to repeat the same argument in the case \((\tilde{x},\tilde{\xi },\tilde{x},-\tilde{\xi })\in WF^s(K_{G^+})\).

Remark 5.8

A similar argument holds for the case \((\tilde{x},\lambda \tilde{\xi },\tilde{x},\tilde{\xi })\notin WF^{-2+\tau -\tilde{\epsilon }} (K_{G^\pm })\) by using \(P_{(s,y)}\).

Lemma 5.9

For \(\tau >2\) and any \(\tilde{\epsilon }>0\),

$$\begin{aligned} WF^{-2+\tau -\tilde{\epsilon }}(K_G)\subset {{\,\textrm{Char}\,}}(P)\times {{\,\textrm{Char}\,}}(P). \end{aligned}$$
(5.15)

Proof

Using Lemma 5.5, we just need to show that there are no points from the sets \({\mathcal {A}}\) or \({\mathcal {B}}\). Let \((\tilde{x},\tilde{\xi },\tilde{y},0)\in {{{\mathcal {B}}}}\cap WF^{-2+\tau -\tilde{\epsilon }}(K_G)\) then by Theorem 3.6, we have that \((\gamma (\tilde{x},\tilde{\xi }),\tilde{y},0)\in WF^{-2+\tau -\tilde{\epsilon }}(K_G)\). Now \(\tilde{y}=(s_1,y_1)\) for some \(s_1\in {\mathbb {R}},y_1\in \Sigma \). By global hyperbolicity, \(\gamma (\tilde{x},\tilde{\xi })\) intersects \(\{s_1\}\times \Sigma \) in exactly one point with the covector \(\chi \ne 0\). Since causally separated points are not in \({{\,\textrm{supp}\,}}(K_G)\), the point of intersection has to be \((s_1,y_1)\). Hence, \((s_1,y_1,\chi ,s_1,y_1,0)\in WF^{-2+\tau -\tilde{\epsilon }}(K_G)\subset (WF^{{-}2{+}\tau -\tilde{\epsilon }}(K_{G^{+}})\cup WF^{-2{+}\tau -\tilde{\epsilon }} (K_{G^{-}}))\). This is a contradiction to Lemma 5.6. A similar argument holds for points in \({\mathcal {A}} \).\(\square \)

Remark 5.10

The existence of symmetries allows one to show that the Sobolev wavefront set in Lemma 5.5 is already disjoint from the sets \({{\mathcal {A}}}\) and \({{\mathcal {B}}}\). For example, if M is stationary, \(K_G\) is of the form \(K_G(t-s,x,y)\). Therefore, one has the additional equation \((\partial _t+\partial _s)K_G=0\), that implies \(WF^l(K_G)\subset {{\,\textrm{Char}\,}}(\partial _t+\partial _s)\) for \(l\in {\mathbb {R}}\). Moreover, \({{\,\textrm{Char}\,}}(\partial _t+\partial _s)\cap {{\mathcal {A}}}=\emptyset \) and \({{\,\textrm{Char}\,}}(\partial _t+\partial _s)\cap {{\mathcal {B}}}=\emptyset \). A similar argument holds in the case of a sufficiently spatially symmetric spacetime, e.g. cosmological space of the form \(ds^2=a(t)(-dt^2+dx^2+dy^2+dz^2)\). In this case, \(K_G\) is of the form \(K_G(t,s,x_1-x_2,y_1-y_2,z_1-z_2)\) due to the spatial invariance.

Now we establish that points above the diagonal are of a specific form.

Lemma 5.11

If \((\tilde{x},\tilde{\xi },\tilde{x},\tilde{\eta })\in WF^{-2+\tau -\tilde{\epsilon }}(K_G)\) for \(\tau >2\), and some \(\tilde{\epsilon }>0\), then \(\tilde{\eta }=-\tilde{\xi }\).

Proof

Suppose \({\tilde{\eta }}\) and \({\tilde{\xi }}\) are linearly independent, i.e. \(\tilde{\eta }\ne \mu \tilde{\xi }\) for \(\mu \in {\mathbb {R}}\). By Lemma (5.9) \(({\tilde{x}}, {\tilde{\xi }}, {\tilde{x}},{\tilde{\eta }})\in {{\,\textrm{Char}\,}}(P)\times {{\,\textrm{Char}\,}}(P)\). Now we choose a Cauchy hypersurface \(\Sigma _{ t_0}=\{t_0\}\times \Sigma \) such that the null geodesic with initial data \((\tilde{x},\tilde{\xi })\) and the null geodesic with initial data \((\tilde{x},\tilde{\eta })\) intersect it. These points of intersections are unique by global hyperbolicity. Moreover, using the condition \(\tilde{\eta }\ne \mu \tilde{\xi }\), we can choose \(\Sigma _ {t_0}\) such that these points are distinct. We denote these points by \((t_0,x_0), (t_0, y_0)\). Furthermore, they are not causally related. Now \(K_G\in H_{loc}^{-1-\epsilon }(M\times m)\) so \(\partial _{\nu }(\sqrt{|g|}g^{\mu \nu }\partial _\mu )^bK_G\in H^{-3-\epsilon +\tau \delta }({\mathbb {R}}^4\times {\mathbb {R}}^4)\) and therefore if \((\tilde{x},\tilde{\xi },\tilde{y},-\tilde{\eta })\in WF^{-2+\tau -{\epsilon }}(K_G)\) then \((\gamma (\tilde{x},\tilde{\xi }),\gamma (\tilde{y},-\tilde{\eta }))\in { WF^{-2-\epsilon +\tau }}(K_G)\) where \(\gamma (\tilde{x},\tilde{\xi })\) is the null bicharacteristic with initial data \((\tilde{x},\tilde{\xi })\) and \(\gamma (\tilde{x},\tilde{\eta })\) is the null bicharacteristic with initial data \((\tilde{x},\tilde{\eta })\).

In particular \((t_0,x_0, t_0, y_0)\in \Pi ({WF^{-\frac{1}{2}-\epsilon +\tau }}(K_G))\), where \(\Pi \) is the projection from \(T^*(M\times M)\) to \(M\times M\). However, this is a contradiction to Proposition 4.1, since \((t_0,x_0, t_0, y_0)\notin {{\,\textrm{supp}\,}}(K_G)\). Therefore, \(\tilde{\eta }=\mu \tilde{\xi }\).

Now as a consequence of the fact that \(K_G=K_{G^+}+K_{G^-}\) and \(WF^s(K_G)\subset WF^s(K_{G^+})\cup WF^s(K_{G^-})\) for all s, Lemma 5.6 implies that \(\mu =-1\). \(\square \)

Proof of Theorem 5.1

Let \((\tilde{x},\tilde{\xi },\tilde{y},-\tilde{\eta })\in WF^{-2+\tau -{\epsilon }}(K_{G})\). The propagation of singularities result (Theorem 3.6) implies that \((\gamma (\tilde{x},\tilde{\xi }),\gamma (\tilde{y},-\tilde{\eta }))\in { WF^{-2-\epsilon +\tau }} (K_G)\), where \(\gamma (\tilde{x},\tilde{\xi })\) is the null bicharacteristic with initial data \((\tilde{x},\tilde{\xi })\) and \(\gamma (\tilde{y},-\tilde{\eta })\) is the null bicharacteristic with initial data \((\tilde{y},-\tilde{\eta })\).

Now we choose a Cauchy surface \(\Sigma _{t_1}=\{t_1\}\times \Sigma \) and suppose that \((t_1,x_1,\tilde{\xi }_1,t_1,x_2,{\tilde{\xi }}_2)\in (\gamma (\tilde{x},\tilde{\xi }),\gamma (\tilde{y},-\tilde{\eta }))\cap (\Sigma _{t_1}^2)\). By Lemmas 4.1 and 5.9, \((t_1, x_1,\tilde{\xi }_1),(t_1, x_2,\tilde{\xi }_2)\in {{\,\textrm{Char}\,}}(P)\), \(x_1=x_2\), and \({\tilde{\xi }}_2 = -{\tilde{\xi }}_1\).

Next we define a curve \(\tilde{\gamma }:(-\infty ,\infty )\rightarrow M\) as follows. First, we shift the parametrization \(\lambda \) in the definition of the null bicharacteristics so that

$$\begin{aligned} \gamma (\tilde{x},\tilde{\xi })(t_1)= (t_1,x_1,\tilde{\xi }_1), \quad \gamma (\tilde{y},-\tilde{\eta })(t_1)= (t_1,x_1,-\tilde{\xi }_1). \end{aligned}$$

Then, we denote by \(\Pi :T^*M\rightarrow M\) the canonical projection and define two curves in M by

$$\begin{aligned} \gamma _1(\lambda ):=\Pi (\gamma (\tilde{x},\tilde{\xi })(\lambda )),\quad \gamma _2(\lambda ):=\Pi (\gamma (\tilde{y},-\tilde{\eta })(\lambda )). \end{aligned}$$

Notice that we have \({\gamma _1}(t_1)=(t_1,x_1), \dot{\gamma _1}(t_1)=g^{-1}(\tilde{\xi }_1,\cdot )\) and \({\gamma _2}(t_1)=(t_1,x_1),\dot{\gamma _2}({t_1})=g^{-1}(-\tilde{\xi }_1,\cdot )\). Moreover, we can assume that \(\tilde{x}= \gamma _1(a)\) and \(\tilde{y}=\gamma _2(b)\) for suitable \(a, b\in {\mathbb {R}}\) with \(a<t_1<b\).

Finally, let

$$\begin{aligned} \tilde{\gamma }(\lambda )={\left\{ \begin{array}{ll} \gamma _1(\lambda )&{} \lambda \in (-\infty , t_1]\\ -\gamma _2(\lambda )&{} \lambda \in (t_1,\infty ) \end{array}\right. } \end{aligned}$$
(5.16)

where \(-\gamma _2\) denotes the curve with opposite orientation.

Then, \({\tilde{\gamma }}(a) = \tilde{x}\), \({\tilde{\gamma }}(b) = \tilde{y}\); moreover, \(g(\cdot , {\dot{{\tilde{\gamma }}}})|_{T_{\tilde{x}}M}=\tilde{\xi }\), \(g(\cdot , {\dot{{\tilde{\gamma }}}})|_{T_{\tilde{y}}M}=\tilde{\eta }\), and therefore, \(\tilde{\gamma }\) is a null geodesic between \(\tilde{x}\) and \(\tilde{y}\) with cotangent vectors \(\tilde{\xi }\) at \(\tilde{x}\) and \(\tilde{\eta }\) at \(\tilde{y}\), i.e. \((\tilde{x},\tilde{\xi },\tilde{y},-\tilde{\eta })\in C':=\{(\tilde{x}, \tilde{\xi },\tilde{y}, -\tilde{\eta });(\tilde{x}, \xi ; \tilde{y}, \tilde{\eta }) \in C\}\), see Fig. 2.

This shows

$$\begin{aligned} {WF^{-2-\epsilon +\tau }}(K_G)\subset C' \end{aligned}$$
(5.17)

or, equivalently \({WF'^{-2-\epsilon +\tau }}(K_G)\subset C\). \(\square \)

Fig. 2
figure 2

\(\gamma _1\) is a null geodesic that satisfies \(\gamma (a)=\tilde{x},\dot{\gamma }_1(a)=g^{-1}(\xi ,\dot{)}\) and \(\gamma _2\) is a null geodesic that satisfies \(\gamma (b)=\tilde{y},\dot{\gamma }_2(b)=g^{-1}(-\eta ,\cdot )\)

5.2 Proof of Theorem 5.2

Now we show that C is contained in \(WF'^{-\frac{1}{2}}(K_G)\).

Lemma 5.12

Let P be the Klein–Gordon operator with \(g\in C^{\tau +2},\tau >2\). Then, \(C\subset WF'^{-\frac{1}{2}}(K_G)\)

Proof

Using Proposition C.1 of [28], see also [48], there exists an interpolating spacetime of regularity \(C^\tau \), \((\bar{M}, \bar{g})\), which satisfies the following conditions: There exist times \(t_1\) and \(t_2\) such that for \(t < t_1\), \((\bar{M}, \bar{g})\) is isometric to a neighbourhood of a Cauchy surface \(\tilde{\Sigma }\) of a smooth, globally hyperbolic spacetime \((M_s,g_s)\). Furthermore, for \(t > t_2\), \((\bar{M}, \bar{g})\) is isometric to a neighbourhood of a Cauchy surface \(\Sigma \) of the non-smooth spacetime \(({M}, {g})\).

Now if \(K_{\bar{G}}\) is the causal propagator associated to \((\bar{M}, \bar{g})\), its restriction to \(t<t_1\), denoted \(K_{\bar{G}}|_{t<t_1}\), corresponds to the smooth causal propagator [5, Proposition 3.5.1] and therefore

$$\begin{aligned} WF'(K_{\bar{G}}|_{t<t_1}) =\bar{C}\cap T^*( \{(t,x)\in \bar{M}; t<t_1 \}\times \{(t,x)\in \bar{M}; t<t_1\}), \end{aligned}$$

where \(\bar{C}\) denotes the canonical relationship associated to \(\bar{g}\).

Let \((\tilde{x},\tilde{\xi },\tilde{x},-\tilde{\xi })\in \bar{C}'\) in the non-smooth region, i.e. \(\tilde{x}=(t_3,x)\) with \(t_3>t_2\).

By global hyperbolicity, the base point projections of the null bicharacteristics \(\gamma (\tilde{x},\tilde{\xi })\) and \(\gamma (\tilde{x},-\tilde{\xi })\) intersect the hypersurface \(t=t_0<t_1\) at one unique point denoted w. Moreover, as a consequence of being in \(\bar{C}'\), we have \((w,\chi ,w,-\chi )=(\gamma (\tilde{x},\tilde{\xi })\times \gamma (\tilde{x},-\tilde{\xi }))\cap (\tilde{\Sigma }_{t_0}\times \tilde{\Sigma }_{t_0})\).

Since we are in the smooth part, smooth theory implies, in particular, that \((w,\chi ,w,-\chi )\in WF^{s}(K_{\bar{G}}|_{t<t_1})\) for \(-\frac{1}{2}\le s\) by combining [25, Theorem 6.5.3] and [38, Proposition B.10]. Now, an application of Theorem 3.6 gives \((\tilde{x},\tilde{\xi },\tilde{x},-\tilde{\xi })\in WF^{s}(K_{\bar{G}})\) for \(-\frac{1}{2}\le s\).

Furthermore, by [36, Theorem 5.10, Theorem 5.8], the restriction of \(K_{\bar{G}}\) to \(t>t_2\), denoted \(K_{\bar{G}}|_{t>t_2}\), in a neighbourhood of \(\Sigma _{t_3}\) is the same as the restriction of the non-smooth causal propagator, \(K_G\), associated to (Mg). Hence, \((\tilde{x},\tilde{\xi },\tilde{x},-\tilde{\xi })\in WF^{s}(K_{G})\).

Another application of Theorem 3.6 using the null bicharacteristics from (Mg) gives \(C'\subset WF^{-\frac{1}{2}}(K_G) \), i.e. \(C\subset WF'^{-\frac{1}{2}}(K_G) \). \(\square \)

Proof of Theorem 5.2

The combination of Lemma 5.12 and Theorem 5.1 gives the result. \(\square \)