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

Hadamard states for bosonic quantum field theory on globally hyperbolic spacetimes

Max Lewandowski
Institute of Mathematics, Potsdam University
Karl-Liebknecht-Str. 24-25, 14476 Potsdam, Germany
E-mail: [email protected]
Abstract

According to Radzikowski’s celebrated results, bisolutions of a wave operator on a globally hyperbolic spacetime are of Hadamard form iff they are given by a linear combination of distinguished parametrices i2(G~aFG~F+G~AG~R)\frac{i}{2}\big{(}\widetilde{G}_{aF}-\widetilde{G}_{F}+\widetilde{G}_{A}-\widetilde{G}_{R}\big{)} in the sense of Duistermaat-Hörmander [Rad1996, DH1972]. Inspired by the construction of the corresponding advanced and retarded Green operator GA,GRG_{A},G_{R} as done in [BGP2007], we construct the remaining two Green operators GF,GaFG_{F},G_{aF} locally in terms of Hadamard series. Afterwards, we provide the global construction of i2(G~aFG~F)\frac{i}{2}\big{(}\widetilde{G}_{aF}-\widetilde{G}_{F}\big{)}, which relies on new techniques like a well-posed Cauchy problem for bisolutions and a patching argument using Čech cohomology. This leads to global bisolutions of Hadamard form, each of which can be chosen to be a Hadamard two-point-function, i.e. the smooth part can be adapted such that, additionally, the symmetry and the positivity condition are exactly satisfied.

1 Introduction

1.1 Quantum field theory on curved space times

Quantum field theory on curved spacetimes is a semiclassical theory, which investigates the coupling of a quantum field with classical gravitation. It already predicts remarkable effects such as particle creation by the curved spacetime itself, a phenomenon most prominently represented by Hawking’s evaporation of black holes [Haw1975] and the Unruh effect [Ful1973, Dav1975, Unr1976]. Due to their rather small, generally trivial, isometry group, curved spacetimes lack an invariant concept of energy, so a distinct vacuum and consequently the notion of particles turn out to be non-sensible in general curved spacetimes [Dav1984, Wal1994]. This situation is best addressed by the algebraic approach to quantum field theory [HK1964, Dim1980], where first of all observables are introduced as elements of rather abstract algebras 𝒜(O)𝒜(M)\mathcal{A}(O)\subset\mathcal{A}(M) associated to spacetime regions OMO\subset M in a local and covariant manner, and states will join the game only later as certain functionals on these algebras.
The categorical framework of locally covariant quantum field theory [BFV2003] has established as a suitable generalization of the principles of AQFT to curved spacetimes. Instead of fixing a spacetime and its symmetries from the beginning, a whole category of spacetimes is considered with arrows given by certain isometric embeddings. The actual QFT is then represented by a covariant functor to the category of CC^{*}-algebras and injective CC^{*}-homomorphisms fulfilling adapted Haag-Kastler-axioms. For the category of spacetimes, we will adopt the setting of [BG2011]:

Definition 1.1.

The category GlobHypGreen consists of objects (M,E,P)(M,E,P) and morphisms (f,F)(f,F), where

  • MM is a globally hyperbolic Lorentzian manifold,

  • EE is a finite-dimensional, real vector bundle over MM with a non-degenerate inner product,

  • P:C(M,E)C(M,E)P\colon C^{\infty}(M,E)\rightarrow C^{\infty}(M,E) is a formally self-adjoint, Green hyperbolic operator,

  • f:M1M2f\colon M_{1}\rightarrow M_{2} time-orientation preserving, isometric embedding with f(M1)M2f(M_{1})\subset M_{2} open and causally compatible,

  • FF is a fiberwise isometric vector bundle isomorphism over ff such that P1,P2P_{1},P_{2} are related via P1res=resP2P_{1}\circ\mathrm{res}=\mathrm{res}\circ P_{2}, where res(φ):=F1φf\mathrm{res}(\varphi):=F^{-1}\circ\varphi\circ f the restriction of φC(M2,E2)\varphi\in C^{\infty}(M_{2},E_{2}) to M1M_{1}.

Because of their good causal and analytic properties (no causal loops, foliation by Cauchy hypersurfaces, well-posed Cauchy problem), globally hyperbolic Lorentzian manifolds have proven to be a reasonable model for curved spacetimes. As further data, we consider real vector bundles, which excludes for instance charged fields, and the large class of Green hyperbolic operators implying the existence of an advanced and retarded Green operator GA,GRG_{A},G_{R} (see [Bär2015] for a thorough discussion of these operators). For G:=GAGRG:=G_{A}-G_{R}, Theorem 3.5 of [BG2011] provides the exact sequence

{0}𝒟(M,E)P𝒟(M,E)GCsc(M,E)PCsc(M,E),\displaystyle\{0\}\longrightarrow\mathscr{D}(M,E)\stackrel{{\scriptstyle P}}{{\longrightarrow}}\mathscr{D}(M,E)\stackrel{{\scriptstyle G}}{{\longrightarrow}}C^{\infty}_{sc}(M,E)\stackrel{{\scriptstyle P}}{{\longrightarrow}}C^{\infty}_{sc}(M,E), (1.1)

and hence, it leads to a covariant functor into the category of symplectic vector spaces with objects (V,σ)(V,\sigma) essentially given by the solution space of the field equation

V:=𝒟(M,E)/kerGkerP|Csc,σ([φ],[ψ]):=(Gφ,ψ)M.\displaystyle V:=\mathscr{D}(M,E)/\ker G\cong\ker P\big{|}_{C^{\infty}_{sc}},\qquad\sigma\big{(}[\varphi],[\psi]\big{)}:=(G\varphi,\psi)_{M}. (1.2)

𝒟(M,E)\mathscr{D}(M,E) represents test sections in EE, more precisely smooth sections with compact support, and Csc(M,E)C^{\infty}_{sc}(M,E) those with only spacelike compact support, meaning that it is contained in the causal future and past of some compact subset of MM. The L2L^{2}-product (,)M(\cdot,\cdot)_{M} of test sections is induced by the non-degenerate inner product on EE.
For a bosonic quantum field theory, we take the CCR\mathrm{CCR}-representation of (V,σ)(V,\sigma), i.e. a pair (w,A)(w,A) consisting of a CC^{*}-algebra AA and a map ww from VV into the unitary elements of AA such that AA is generated as a CC^{*}-algebra by {w(x)}xV\{w(x)\}_{x\in V} and the Weyl relations hold:

w(x)w(y)=ei2σ(x,y)w(x+y),x,yV.\displaystyle w(x)w(y)=e^{-\frac{i}{2}\sigma(x,y)}w(x+y),\qquad x,y\in V. (1.3)

This construction goes back to [Man1968] (see also section 4.2 of [BGP2007] and 5.2.2.2 of [BR2002]) and it is unique in an appropriate sense. Therefore, altogether, (M,E,P)CCR(𝒟(M,E)/kerG)(M,E,P)\mapsto\mathrm{CCR}\big{(}\mathscr{D}(M,E)/\ker G\big{)} provides the desired functor and Haag-Kastler’s axioms are satisfied (Theorem 3.10 of [BG2011])

1.2 States, quasifree states and Hadamard states

We introduce states in the theory as normed and positive functionals τ\tau on CCR(V)\mathrm{CCR}(V), where τ(a)\tau(a) can be thought of as the expectation value of the observable aa in the state τ\tau. The induced GNS-representation (πτ,τ,Ωτ)(\pi_{\tau},\mathscr{H}_{\tau},\Omega_{\tau}) provides the familiar framework of a state space τ\mathscr{H}_{\tau} with observables as bounded operators πτ(a)\pi_{\tau}(a) and a cyclic vector Ωτ\Omega_{\tau} (see section 2.3 of [BR2002] for details). Hence, the selection of a distinct vacuum is shifted to that of an algebraic state τ\tau. Particularly adapted to free quantum fields are the so-called quasifree states generated by

τ(w(x))=e12η(x,x),xV,\displaystyle\tau\big{(}w(x)\big{)}=e^{-\frac{1}{2}\eta(x,x)},\qquad x\in V,

for some scalar product η\eta on VV, by which (πτ,τ,Ωτ)(\pi_{\tau},\mathscr{H}_{\tau},\Omega_{\tau}) is determined up to unitary equivalence. For these states, the unitary operators {πτ(w(tx))}t\big{\{}\pi_{\tau}\big{(}w(tx)\big{)}\big{\}}_{t\in\mathbb{R}} constitute a strongly continuous one-parameter group and thus, field operators Φτ(x)\Phi_{\tau}(x) are given by the self-adjoint generators due to Stone’s theorem. Furthermore, there is a dense domain DττD_{\tau}\subset\mathscr{H}_{\tau} such that ranΦτ(x)DτdomΦτ(x)\mathrm{ran}\,\Phi_{\tau}(x)\subset D_{\tau}\subset\mathrm{dom}\,\Phi_{\tau}(x) for all xx, so polynomials of field operators are well-defined on DτD_{\tau}. Hence, the Weyl relations (1.3) imply the familiar canonical commutator relations [Φτ(x),Φτ(y)]=iσ(x,y)idτ\big{[}\Phi_{\tau}(x),\Phi_{\tau}(y)\big{]}=i\sigma(x,y)\,\text{id}_{\mathscr{H}_{\tau}} and for all nn\in\mathbb{N}, the nn-point-function of the state τn(x1,,xn):=Φτ(x1)Φτ(xn)Ωτ,Ωττ\tau_{n}(x_{1},\ldots,x_{n}):=\big{\langle}\Phi_{\tau}(x_{1})\ldots\Phi_{\tau}(x_{n})\Omega_{\tau},\Omega_{\tau}\big{\rangle}_{\mathscr{H}_{\tau}} represents a well-defined distribution (see section 4.2 of [BG2011] for precise definitions and proofs). In particular, the two-point-function is of the form

τ2(x,y)=η(x,y)+i2σ(x,y),x,yV,\displaystyle\tau_{2}(x,y)=\eta(x,y)+\frac{i}{2}\sigma(x,y),\qquad x,y\in V, (1.4)

so it reproduces η\eta and hence τ\tau. Indeed, we have τn=0\tau_{n}=0 for odd nn, and for nn even, it is given by some polynomial in the elements of {τ2(xi,xj)}i,j=1,,n\{\tau_{2}(x_{i},x_{j})\}_{i,j=1,\ldots,n}. Thinking of τn\tau_{n} as the propagation of the state of the field, the focus on quasifree states corresponds to the perception that this propagation is essentially given by independent one-particle-propagations, which legitimates them as the natural objects to look at when dealing with free quantum fields (see chapter 17 of [DG2013] for an overview of quasifree states). Going back from (V,σ)(V,\sigma) to (M,E,P)(M,E,P), the scalar product η\eta corresponds to a bidistribution S:𝒟(M,E)×𝒟(M,E)S\colon\mathscr{D}(M,E)\times\mathscr{D}(M,E)\rightarrow\mathbb{R} with

S[Pψ1,ψ2]=0=S[ψ1,Pψ2],S[ψ1,ψ2]=S[ψ2,ψ1],S[ψ,ψ]0.\displaystyle S[P\psi_{1},\psi_{2}]=0=S[\psi_{1},P\psi_{2}],\qquad S[\psi_{1},\psi_{2}]=S[\psi_{2},\psi_{1}],\qquad S[\psi,\psi]\geq 0. (1.5)

With regard to (1.2) and (1.4), a quasifree state is therefore determined by SS and GG.
Despite all physically motivated restrictions on the field so far, there is still a huge variety of possible states, so we need to look for constraints also on this level. A reasonable demand would be renormalizability of τ2\tau_{2}, most prominently represented by the expectation value of the energy momentum tensor, since products of distributions are in general ill-defined. In flat quantum field theory, this would be carried out by subtracting the vacuum expectation value setting us back to the problem of non-existence of a distinct vacuum. On the other hand, this procedure merely requires regularity of differences of expectation values, and indeed, J. Hadamard’s theory of second order hyperbolic equations [Had1923] led to a family of bisolutions with fixed singular part in the sense that the difference of any two such bisolutions is smooth (see [Wal1994] and, more recently, [Hac2016] for details, as well as [DF2008] for the concrete renormalization). Accordingly, a state is called a Hadamard state if its two-point-function has the Hadamard singularity structure, which, by now, has been shown to be invariant under Cauchy evolution [FSW1978]. Futhermore, any globally hyperbolic spacetime admits a large class of pure Hadamard states [FNW1981, SV2001].
However, the first mathematically precise definition of the Hadamard singularity structure has been specified only in [KW1991], in which the authors also show that for a wide class of spacetimes the Hadamard property singles out an invariant quasifree state. Moreover, in any spatially compact spacetime ("closed universes"), all Hadamard states, more specifically their GNS representations, comprise one unitary equivalence class, which, for general spacetimes, suggests a certain "local equivalence" of all possible notions of a vacuum state [Ver1994]. In addition, Hadamard states yield finite fluctuations for all Wick polynomials [BF2000], which makes them relevant also for the perturbative construction of interacting fields (see also [HW2015, Rej2016, Düt2019] and references therein). Consequently, Hadamard states are by now considered a reasonable counterpart of Minkowski finite energy states and the Hadamard condition an appropriate generalization of the energy condition for Minkowski quantum field theory. Note that the replacement of a distinct vacuum state by a whole class of states somehow reflects the essence of general relativity: Just like there is no preferred coordinate system, the concept of vacuum and particles as absolute quantities has to be re-evaluated and eventually downgraded to one choice among many.
It was Radzikowski who showed that for the massive scalar field the global Hadamard condition is equivalent to a certain requirement on the wave front set of the two-point-function [Rad1992, Rad1996], namely

WF(τ2)={(p,ξ;q,ζ)(TM×TM)\{0}|(p,ξ)(q,ζ),ξ is future-directed},\displaystyle\text{WF}(\tau_{2})=\big{\{}(p,-\xi;q,\zeta)\in\big{(}T^{*}M\times T^{*}M\big{)}\backslash\{0\}\leavevmode\nobreak\ \big{|}\leavevmode\nobreak\ (p,\xi)\sim(q,\zeta),\leavevmode\nobreak\ \xi\text{ is future-directed}\big{\}}, (1.6)

where

(p,ξ)(q,ζ) lightlike geodesic c:IMandt,tI:c(t)=p,c(t)=q,c˙(t)=ξ,c˙(t)=ζ.\displaystyle\begin{split}(p,\xi)\sim(q,\zeta)\qquad\Longleftrightarrow\qquad\begin{array}[]{cl}&\exists\text{ lightlike geodesic }c\colon I\rightarrow M\enspace\text{and}\enspace t,t^{\prime}\in I\colon\\[5.69054pt] &c(t)=p,\enspace c(t^{\prime})=q,\enspace\dot{c}(t)=\xi^{\sharp},\enspace\dot{c}(t^{\prime})=\zeta^{\sharp}.\end{array}\end{split} (1.7)

Note that, unlike the criterion given in [KW1991], this is a local condition, which Sahlmann and Verch generalized to sections in general vector bundles [SV2001]. It includes Hadamard states of the Dirac field in the sense of [Köh1995, Kra2000, Hol2001], which have been used, for instance, for a mathematical rigorous description of the chiral anomaly [BS2016]. In addition, [SVW2002] proposed an even more elegant characterization of the Hadamard property in terms of Hilbert space valued distributions φΦτ[φ]Ωττ\varphi\mapsto\Phi_{\tau}[\varphi]\Omega_{\tau}\in\mathscr{H}_{\tau}, involving the GNS-representation induced by τ\tau. Also for non-quasifree states, one can formulate (1.6) as a constraint on the whole nn-point-function, which is compatible with the special case of quasifree states [San2010]. Moreover, in analytic spacetimes, this generalized Hadamard condition can be sharpened to a condition on the analytic wave front set, thereby implying the Reeh-Schlieder-property [SVW2002]. Likewise, for non-globally hyperbolic spacetimes, there is a formulation of the Hadamard condition via restriction to globally hyperbolic subregions. Hadamard states have therefore been studied in connection with the Casimir effect and on anti-de Sitter spacetime (see [DNP2014, DFM2018] and references therein). By using the weaker concept of Sobolev wave front sets, a definition of adiabatic states on globally hyperbolic spacetimes similar to (1.6) is given in [JS2002], thus implying that Hadamard states are adiabatic.
However, most importantly for the purpose of this work, the Hadamard condition in the form (1.6) allows us to employ the techniques of microlocal analysis provided by Duistermaat and Hörmander [DH1972]. Soon after Radzikowski’s work, Junker derived pure Hadamard states for the massive scalar field on spatially compact globally hyperbolic spacetimes, using a factorization of the Klein-Gordon operator by pseudo-differential operators [Jun1996, Jun2002]. Gérard, Wrochna et al. generalized this construction to a large class of spacetimes [GOW2017] and even gauge fields [GW2015]. Furthermore, they proved the existence of (not necessarily pure) Hadamard states [GW2014] in a much more concrete manner than [FNW1981]. See [Gér2019] for a recent review of these techniques.
On the other hand, there have been further proposals for physically reasonable states like the Sorkin-Johnston-states [AAS2012], which in general lack the Hadamard property [FV2012]. Nevertheless, a modification of their construction produces Hadamard states [BF2014]. For a contemporary synopsis concerning preferred vacuum states on general spacetimes, the nature of the Hadamard property and this construction in particular, see also [Few2018]. Apart from these rather general prescriptions, Hadamard states have been constructed explicitly for a large variety of spacetimes with special (asymptotic) symmetries, and furthermore, well-established states have been tested for the Hadamard property (see the introduction sections of [GW2014, GOW2017] and the references therein as well as section 8.4 of [FV2015] and 2.4 of [Hac2016] for an overview).

1.3 This work

In his seminal work [Rad1996], Radzikowski already realized that a bidistribution H~\widetilde{H} satisfies the Hadamard condition if and only if it is of the form

H~=i2(G~aFG~F+G~AG~R)\displaystyle\widetilde{H}=\frac{i}{2}\big{(}\widetilde{G}_{aF}-\widetilde{G}_{F}+\widetilde{G}_{A}-\widetilde{G}_{R}\big{)} (1.8)

with G~aF,G~F,G~A,G~R\widetilde{G}_{aF},\widetilde{G}_{F},\widetilde{G}_{A},\widetilde{G}_{R} the distinguished parametrices in the sense of Duistermaat-Hörmander (Theorem 6.5.3 in [DH1972]). With Δ:={(p,ξ;p,ξ)}\Delta^{\prime}:=\big{\{}(p,\xi;p,-\xi)\big{\}} the primed diagonal, they are characterized by

WF(G~A)=Δ{(p,ξ)(q,ζ),qJ+(p)},WF(G~R)=Δ{(p,ξ)(q,ζ),qJ(p)},WF(G~F)=Δ{(p,ξ)(q,ζ),t>t},WF(G~aF)=Δ{(p,ξ)(q,ζ),t<t}.\displaystyle\begin{split}\text{WF}(\widetilde{G}_{A})=\Delta^{\prime}\cup\big{\{}(p,\xi)\sim(q,-\zeta),\leavevmode\nobreak\ q\in J_{+}(p)\big{\}},&\quad\text{WF}(\widetilde{G}_{R})=\Delta^{\prime}\cup\big{\{}(p,\xi)\sim(q,-\zeta),\leavevmode\nobreak\ q\in J_{-}(p)\big{\}},\\[5.69054pt] \text{WF}(\widetilde{G}_{F})=\Delta^{\prime}\cup\big{\{}(p,\xi)\sim(q,\zeta),\leavevmode\nobreak\ t>t^{\prime}\big{\}},&\quad\text{WF}(\widetilde{G}_{aF})=\Delta^{\prime}\cup\big{\{}(p,\xi)\sim(q,\zeta),\leavevmode\nobreak\ t<t^{\prime}\big{\}}.\end{split} (1.9)

However, he remarked that it is not clear how to prove that one may choose the smooth part such that (1.5) is exactly satisfied [Rad1992], an issue closely related to the question, whether G~aF,G~F\widetilde{G}_{aF},\widetilde{G}_{F} can be chosen as actual Green operators, which has been already addressed in section 6.6 of [DH1972]. Clearly, microlocal analysis proved to be a powerful tool for the investigation of singularities and indispensable for the results listed in the former paragraph. Nevertheless, for these remaining questions, the non-singular part of H~\widetilde{H} is primarily concerned, so different techniques are required.
This work is dedicated to provide such techniques and resolves the question for wave operators on globally hyperbolic spacetimes. It therefore gives a further and more constructive existence proof of, not necessarily pure, Hadamard states than [FNW1981]. By avoiding any kind of deformation argument, it covers situations involving, for example, analytic spacetimes or constraint equations as in general relativity, where such an argument is usually not applicable.
The starting point in chapter 3 is the local construction of G~aF,G~F\widetilde{G}_{aF},\widetilde{G}_{F} in terms of Hadamard series very much inspired by [BGP2007]. In chapter 4, these local objects are globalized and finally patched together by introducing a well-posed Cauchy problem for bisolutions and techniques from Čech cohomology theory. In the final chapter 5, we check existence of a "positive and symmetric choice", that is the smooth part of any Hadamard bisolution can be chosen such that it obtains the properties of a two-point-function. The necessary preparation is given in chapter 2 as well as a proof for the symmetry of the Hadamard coefficients in the vector valued case alternative to [Kam2019] in the Appendix.

2 Preliminaries

In this section, we provide basic notations and prove certain theorems needed in the later constructions. For any dd-dimensional vector space with non-degenerate inner product ,\left\langle\!\left\langle\cdot,\cdot\right\rangle\!\right\rangle of index 11, we adopt the notations and conventions of [BGP2007], that is, for instance, the signature (,+,,+)(-,+,\ldots,+) and the squared distance γ(x):=x,x\gamma(x):=-\left\langle\!\left\langle x,x\right\rangle\!\right\rangle. The two connected components I±I_{\pm} of the set of timelike vectors I:={γ(x)>0}I:=\{\gamma(x)>0\} then determine a time-orientation, where we define the elements of I+(I)I_{+}(I_{-}) to be future (past) directed. Correspondingly, we set C±:=I±,J±:=I±¯C_{\pm}:=\partial I_{\pm},\leavevmode\nobreak\ J_{\pm}:=\overline{I_{\pm}}, whose non-zero elements we call "lightlike" and "causal", respectively. Leaving out "±\pm" means the union of both components, i.e. I:=I+II:=I_{+}\cup I_{-} and similarly CC and JJ. Non-causal vectors are referred to as "spacelike".
For (M,g)(M,g) a dd-dimensional time-oriented Lorentzian manifold and pMp\in M, we write I±M(p),C±M(p)I_{\pm}^{M}(p),C_{\pm}^{M}(p) and J±M(p)J^{M}_{\pm}(p) for the corresponding chronological/lightlike/causal future/past of pp. These sets comprise all points that can be reached from pp via some timelike/lightlike/causal future/past directed differentiable curve, that is a curve with tangent vectors of the respective type at each point. For subsets AMA\subset M, we define I±M(A):=pAI±M(p)I_{\pm}^{M}(A):=\bigcup_{p\in A}I^{M}_{\pm}(p) and similarly J±M(A)J^{M}_{\pm}(A). For the definitions of different types of subsets of MM like future/past compact, geodesically starshaped, convex, causally compatible, causal, Cauchy hypersurface etc., we refer to section 1.3 of [BGP2007]. In this work, Cauchy hypersurfaces of MM are always assumed to be spacelike.
For EE some real or complex finite-dimensional vector bundle over MM, the spaces of CkC^{k}-, CC^{\infty}-, 𝒟\mathscr{D}-sections in EE as well as distributional sections 𝒟(M,E,W)\mathscr{D}(M,E,W)^{\prime} with values in some finite-dimensional space WW, including their (singular) support, convergence, order etc., are defined as in section 1.1 of [BGP2007]. For dV\,\text{d}V the volume density induced by the Lorentzian metric, FF another vector bundle over MM and P:C(M,E)C(M,F)P\colon C^{\infty}(M,E)\rightarrow C^{\infty}(M,F) some linear differential operator, the formally transposed operator Pt:C(M,F)C(M,E)P^{t}\colon C^{\infty}(M,F^{*})\rightarrow C^{\infty}(M,E^{*}) of PP is given by

(Ptφ)[ψ]:=φ[Pψ]=Mφ(Pψ)dV,ψ𝒟(M,E),φ𝒟(M,F).(P^{t}\varphi)[\psi]:=\varphi[P\psi]=\int_{M}\varphi\big{(}P\psi\big{)}\,\text{d}V,\qquad\psi\in\mathscr{D}(M,E),\leavevmode\nobreak\ \varphi\in\mathscr{D}(M,F^{*}).

If EE is equipped with a non-degenerate inner product ,\left\langle\cdot,\cdot\right\rangle, which induces the L2L^{2}-product (,)M(\cdot,\cdot)_{M} and the isomorphism Θ:EE\Theta\colon E\rightarrow E^{*}, we call PP formally self-adjoint if (Pψ1,ψ2)M=(ψ1,Pψ2)M(P\psi_{1},\psi_{2})_{M}=(\psi_{1},P\psi_{2})_{M} for all ψ1,ψ2\psi_{1},\psi_{2}, that is, P=Θ1PtΘP=\Theta^{-1}P^{t}\Theta. Furthermore, PP is a wave operator if its principal symbol is given by ξg(ξ,ξ)idE,\xi\mapsto g(\xi^{\sharp},\xi^{\sharp})\cdot\text{id}_{E}, on TMT^{*}M implying that wave operators are of second order (see section 1.5 of [BGP2007] for details).
L. Schwartz’ celebrated kernel theorem establishes a one-to-one-correspondence between sequentially continuous operators 𝒦:𝒟(M,E)𝒟(M,E)\mathcal{K}\colon\mathscr{D}(M,E^{*})\rightarrow\mathscr{D}(M,E^{*})^{\prime}, i.e. 𝒦φj𝒦φ\mathcal{K}\varphi_{j}\rightarrow\mathcal{K}\varphi if φjφ\varphi_{j}\rightarrow\varphi, and bidistributions K:𝒟(M,E)×𝒟(M,E)K\colon\mathscr{D}(M,E)\times\mathscr{D}(M,E^{*})\rightarrow\mathbb{R} given by K[ψ,φ]=(𝒦φ)[ψ]K[\psi,\varphi]=(\mathcal{K}\varphi)[\psi] and called Schwartz kernel of 𝒦\mathcal{K}. It is represented by a distributional section in the bundle EEE^{*}\boxtimes E over M×MM\times M, whose fibers we identify via

(EE)(p,q)=EpEqHom(Eq,Ep),(p,q)M×M.\displaystyle(E^{*}\boxtimes E)_{(p,q)}=E^{*}_{p}\otimes E_{q}\cong\textup{Hom}(E^{*}_{q},E^{*}_{p}),\qquad(p,q)\in M\times M. (2.1)
Definition 2.1.

Let P:C(M,E)C(M,E)P\colon C^{\infty}(M,E)\rightarrow C^{\infty}(M,E) be a linear differential operator. A linear and sequentially continuous operator 𝒬:𝒟(M,E)C(M,E)\mathcal{Q}\colon\mathscr{D}(M,E^{*})\rightarrow C^{\infty}(M,E^{*}) is called

  • left parametrix for PtP^{t} if 𝒬Pt|𝒟id\mathcal{Q}P^{t}\big{|}_{\mathscr{D}}-\text{id} is smoothing,

  • right parametrix for PtP^{t} if Pt𝒬idP^{t}\mathcal{Q}-\text{id} is smoothing,

  • two-sided parametrix or just parametrix for PtP^{t} if 𝒬\mathcal{Q} is left and right parametrix for PtP^{t},

  • Green operator for PtP^{t} if 𝒬Pt|𝒟=Pt𝒬=id\mathcal{Q}P^{t}\big{|}_{\mathscr{D}}=P^{t}\mathcal{Q}=\text{id}.

A bidistribution Q:𝒟(M,E)×𝒟(M,E)Q\colon\mathscr{D}(M,E)\times\mathscr{D}(M,E^{*})\rightarrow\mathbb{R} with pQ(p)[φ]C(M,E)p\mapsto Q(p)[\varphi]\in C^{\infty}(M,E^{*}) for all φ\varphi is called

  • parametrix for PP at pMp\in M if P(2)Q(p)δpC(M,E)P_{(2)}Q(p)-\delta_{p}\in C^{\infty}(M,E),

  • fundamental solution for PP at pMp\in M if P(2)Q(p)=δpP_{(2)}Q(p)=\delta_{p}.

Note that 𝒬\mathcal{Q} is a parametrix for PtP^{t} if and only if its Schwartz kernel is a parametrix for PtP^{t} at all pMp\in M. Especially, it is a Green operator for PtP^{t} if and only if Q(p)Q(p) is a fundamental solution for PP and P(1)t(Q()[φ])=φP^{t}_{(1)}\big{(}Q(\cdot)[\varphi]\big{)}=\varphi for all pp and φ\varphi.

2.1 Cauchy problems

A crucial feature of globally hyperbolic Lorentzian manifolds MM is a well-posed Cauchy problem for wave operators on smooth sections, i.e. for any Cauchy hypersurface Σ\Sigma with normal field ν\nu, which is timelike, and Cauchy data fC(M,E),u0,u1C(Σ,E)f\in C^{\infty}(M,E),u_{0},u_{1}\in C^{\infty}(\Sigma,E), the Cauchy problem

{Pu=f,u|Σ=u0,νu|Σ=u1,\displaystyle\left\{\begin{array}[]{cl}Pu&=f,\\[5.69054pt] u\big{|}_{\Sigma}&=u_{0},\\[5.69054pt] \nabla_{\nu}u\big{|}_{\Sigma}&=u_{1},\end{array}\right. (2.5)

has a unique solution uC(M,E)u\in C^{\infty}(M,E), which is supported in J(suppu0suppu1suppf)J\big{(}\text{supp}\,u_{0}\cup\text{supp}\,u_{1}\cup\text{supp}\,f\big{)} and depends continuously on the data (see section 3.2 of [BGP2007] and chapter 3 of [BF2009]). For FF a vector bundle over some further globally hyperbolic Lorentzian manifold NN, recall (2.1) for the definition of the vector bundle EFE\boxtimes F over M×NM\times N.

Theorem 2.2.

Let M,NM,N be globally hyperbolic Lorentzian manifolds with Cauchy hypersurfaces Σ,Ξ\Sigma,\Xi and unit normal fields μ,ν\mu,\nu. Furthermore, let P,QP,Q denote linear differential operators of second order acting on smooth sections in vector bundles E,FE,F over M,NM,N, which admit well-posed Cauchy problems and only lightlike characteristic directions. Then, for all uiC(Σ×Ξ,EF),i=1,,4,u_{i}\in C^{\infty}(\Sigma\times\Xi,E\boxtimes F),\leavevmode\nobreak\ i=1,...,4, and f,gC(M×N,EF)f,g\in C^{\infty}(M\times N,E\boxtimes F) with Qf=PgQf=Pg, there is a unique section uC(M×N,EF)u\in C^{\infty}(M\times N,E\boxtimes F) solving

{Pu=f,Qu=g,u|Σ×Ξ=u1,μu|Σ×Ξ=u2,νu|Σ×Ξ=u3,νμu|Σ×Ξ=u4.\displaystyle\left\{\begin{array}[]{cl}Pu&=f,\\[5.69054pt] Qu&=g,\\[5.69054pt] u\big{|}_{\Sigma\times\Xi}&=u_{1},\\[5.69054pt] \nabla_{\mu}u\big{|}_{\Sigma\times\Xi}&=u_{2},\\[5.69054pt] \nabla_{\nu}u\big{|}_{\Sigma\times\Xi}&=u_{3},\\[5.69054pt] \nabla_{\nu}\nabla_{\mu}u\big{|}_{\Sigma\times\Xi}&=u_{4}.\end{array}\right. (2.12)

Let Z:=(2C(M×N,EF))(4C(Σ×Ξ,EF))Z:=\big{(}\oplus^{2}C^{\infty}(M\times N,E\boxtimes F)\big{)}\oplus\big{(}\oplus^{4}C^{\infty}(\Sigma\times\Xi,E\boxtimes F)\big{)} and XX denote the subset of elements (f,g,u1,u2,u3,u4)(f,g,u_{1},u_{2},u_{3},u_{4}) satisfying Qf=PgQf=Pg. Then the map (f,g,u1,u2,u3,u4)u(f,g,u_{1},u_{2},u_{3},u_{4})\longmapsto u, which sends the Cauchy data to the unique solution uu of (2.12), is a linear and continuous operator XC(M×N,EF)X\longrightarrow C^{\infty}(M\times N,E\boxtimes F).

Proof.

For all qNq\in N and h0,h1C(Σ×N,EF)h_{0},h_{1}\in C^{\infty}(\Sigma\times N,E\boxtimes F), the Cauchy problem

{Puq=f(,q),uq|Σ=h0(,q),μuq|Σ=h1(,q),\displaystyle\left\{\begin{array}[]{cl}Pu_{q}&=f(\cdot,q),\\[5.69054pt] u_{q}\big{|}_{\Sigma}&=h_{0}(\cdot,q),\\[5.69054pt] \nabla_{\mu}u_{q}\big{|}_{\Sigma}&=h_{1}(\cdot,q),\end{array}\right. (2.16)

has a unique solution uqC(M,EFq)u_{q}\in C^{\infty}(M,E\otimes F_{q}) depending smoothly on the data by well-posedness of (2.5). Thus, it remains to determine h0,h1h_{0},h_{1} from u1,u2,u3,u4,gu_{1},u_{2},u_{3},u_{4},g and to show that then Qu=gQu=g is automatically fulfilled. For all σΣ,ξΞ\sigma\in\Sigma,\xi\in\Xi, we define smooth sections h0(σ,),h1(σ,)C(N,EσF)h_{0}(\sigma,\cdot),h_{1}(\sigma,\cdot)\in C^{\infty}(N,E_{\sigma}\otimes F) and h0(,ξ),h2(,ξ)C(M,EFξ)h_{0}(\cdot,\xi),h_{2}(\cdot,\xi)\in C^{\infty}(M,E\otimes F_{\xi}) as solutions of

{Ph0(,ξ)=f(,ξ),h0(,ξ)|Σ=u1(,ξ),μh0(,ξ)|Σ=u2(,ξ),{Ph2(,ξ)=(νf)(,ξ),h2(,ξ)|Σ=u3(,ξ),μh2(,ξ)|Σ=u4(,ξ),{Qh0(σ,)=g(σ,),h0(σ,)|Ξ=u1(σ,),νh0(σ,)|Ξ=u3(σ,),{Qh1(σ,)=(μg)(σ,),h1(σ,)|Ξ=u2(σ,),νh1(σ,)|Ξ=u4(σ,).\displaystyle\begin{split}\left\{\begin{array}[]{cl}Ph_{0}(\cdot,\xi)&=f(\cdot,\xi),\\[5.69054pt] h_{0}(\cdot,\xi)\big{|}_{\Sigma}&=u_{1}(\cdot,\xi),\\[5.69054pt] \nabla_{\mu}h_{0}(\cdot,\xi)\big{|}_{\Sigma}&=u_{2}(\cdot,\xi),\end{array}\right.\qquad&\qquad\left\{\begin{array}[]{cl}Ph_{2}(\cdot,\xi)&=\big{(}\nabla_{\nu}f\big{)}(\cdot,\xi),\\[5.69054pt] h_{2}(\cdot,\xi)\big{|}_{\Sigma}&=u_{3}(\cdot,\xi),\\[5.69054pt] \nabla_{\mu}h_{2}(\cdot,\xi)\big{|}_{\Sigma}&=u_{4}(\cdot,\xi),\end{array}\right.\\[5.69054pt] \left\{\begin{array}[]{cl}Qh_{0}(\sigma,\cdot)&=g(\sigma,\cdot),\\[5.69054pt] h_{0}(\sigma,\cdot)\big{|}_{\Xi}&=u_{1}(\sigma,\cdot),\\[5.69054pt] \nabla_{\nu}h_{0}(\sigma,\cdot)\big{|}_{\Xi}&=u_{3}(\sigma,\cdot),\end{array}\right.\qquad&\qquad\left\{\begin{array}[]{cl}Qh_{1}(\sigma,\cdot)&=\big{(}\nabla_{\mu}g\big{)}(\sigma,\cdot),\\[5.69054pt] h_{1}(\sigma,\cdot)\big{|}_{\Xi}&=u_{2}(\sigma,\cdot),\\[5.69054pt] \nabla_{\nu}h_{1}(\sigma,\cdot)\big{|}_{\Xi}&=u_{4}(\sigma,\cdot).\end{array}\right.\end{split} (2.17)

By adapting the proof of Proposition A.1 of [FNW1981], we obtain smooth sections h0,h1,h2h_{0},h_{1},h_{2} in EFE\boxtimes F over (M×Ξ)(Σ×N),Σ×N(M\times\Xi)\cup(\Sigma\times N),\Sigma\times N and M×ΞM\times\Xi, respectively, and, following the same lines, u(,q):=uqu(\cdot,q):=u_{q} depends smoothly on qq. Hence, we found some uC(M×N,EF)u\in C^{\infty}(M\times N,E\boxtimes F) solving (2.16), which yields the initial data of a solution of (2.12):

u|Σ×Ξ=h0|Σ×Ξ=u1,\displaystyle u\big{|}_{\Sigma\times\Xi}=h_{0}\big{|}_{\Sigma\times\Xi}=u_{1},\qquad μu|Σ×Ξ=h1|Σ×Ξ=u2,\displaystyle\nabla_{\mu}u\big{|}_{\Sigma\times\Xi}=h_{1}\big{|}_{\Sigma\times\Xi}=u_{2},
νu|Σ×Ξ=νh0|Σ×Ξ=u3,\displaystyle\nabla_{\nu}u\big{|}_{\Sigma\times\Xi}=\nabla_{\nu}h_{0}\big{|}_{\Sigma\times\Xi}=u_{3},\qquad νμu|Σ×Ξ=νh1|Σ×Ξ=u4.\displaystyle\nabla_{\nu}\nabla_{\mu}u\big{|}_{\Sigma\times\Xi}=\nabla_{\nu}h_{1}\big{|}_{\Sigma\times\Xi}=u_{4}.

Note that PP and QQ commute because they act on different factors of M×NM\times N. Therefore, (2.16) and (2.17) imply that QuQu and gg satisfy the same Cauchy problem:

{PQu=QPu=Qf=Pg,Qu|Σ×N=Qh0=g|Σ×N,μQu|Σ×N=Qμu|Σ×N=Qh1=μg|Σ×N,\left\{\begin{array}[]{cl}&PQu=QPu=Qf=Pg,\\[5.69054pt] &Qu\big{|}_{\Sigma\times N}=Qh_{0}=g\big{|}_{\Sigma\times N},\\[5.69054pt] &\nabla_{\mu}Qu\big{|}_{\Sigma\times N}=Q\nabla_{\mu}u\big{|}_{\Sigma\times N}=Qh_{1}=\nabla_{\mu}g\big{|}_{\Sigma\times N},\end{array}\right.

that is Qu=gQu=g due to well-posedness. Clearly, this solution uu is unique since trivial Cauchy data in (2.12) lead to trivial data in (2.17) and therefore in (2.16), which implies uq=0u_{q}=0 for all qq and hence u=0u=0. The proof of stability follows the same lines as for (2.5), that is Theorem 3.2.12 of [BGP2007]: Obviously, the map

Φ:C(M×N,EF)\displaystyle\Phi\colon\qquad C^{\infty}(M\times N,E\boxtimes F) Z\displaystyle\longrightarrow Z
u\displaystyle u (Pu,Qu,u|Σ×Ξ,μu|Σ×Ξ,νu|Σ×Ξ,νμu|Σ×Ξ)\displaystyle\longmapsto\big{(}Pu,Qu,u\big{|}_{\Sigma\times\Xi},\nabla_{\mu}u\big{|}_{\Sigma\times\Xi},\nabla_{\nu}u\big{|}_{\Sigma\times\Xi},\nabla_{\nu}\nabla_{\mu}u\big{|}_{\Sigma\times\Xi}\big{)}

is linear, injective and continuous, and since (2.12) admits a solution uu for each data in XX, the closed subset XZX\subset Z is contained in ranΦ\mathrm{ran}\,\Phi. Due to continuity of differential operators, the subspace Φ1(X)C(M×N,EF)\Phi^{-1}(X)\subset C^{\infty}(M\times N,E\boxtimes F) is also closed, so we obtain a continuous and bijective map Φ:Φ1(X)X\Phi\colon\Phi^{-1}(X)\rightarrow X between Fréchet spaces, whose inverse is continuous by the open mapping theorem. ∎

It seems that this procedure directly generalizes to operators of order kk, for which derivatives up to order k2k^{2} have to be provided as data. Furthermore, symmetry of the data is inherited by the solution:

Corollary 2.3.

With regard to the assumptions of Theorem 2.2, let (N,Ξ,ν)=(M,Σ,μ),F=E(N,\Xi,\nu)=(M,\Sigma,\mu),\leavevmode\nobreak\ F=E^{*} and EE be equipped with a non-degenerate inner product. Let Q=PtQ=P^{t}, and for all (p,q)M×M(p,q)\in M\times M, assume

f(p,q)=Θp1g(q,p)tΘq,u2(σ1,σ2)=Θσ11u3(σ2,σ1)tΘσ2,u1(σ1,σ2)=Θσ11u1(σ2,σ1)tΘσ2,u4(σ1,σ2)=Θσ11u4(σ2,σ1)tΘσ2\displaystyle\begin{split}f(p,q)=\Theta_{p}^{-1}g(q,p)^{t}\Theta_{q},\qquad u_{2}(\sigma_{1},\sigma_{2})=\Theta^{-1}_{\sigma_{1}}u_{3}(\sigma_{2},\sigma_{1})^{t}\Theta_{\sigma_{2}},\\[5.69054pt] u_{1}(\sigma_{1},\sigma_{2})=\Theta^{-1}_{\sigma_{1}}u_{1}(\sigma_{2},\sigma_{1})^{t}\Theta_{\sigma_{2}},\qquad u_{4}(\sigma_{1},\sigma_{2})=\Theta^{-1}_{\sigma_{1}}u_{4}(\sigma_{2},\sigma_{1})^{t}\Theta_{\sigma_{2}}\end{split} (2.18)

with fiberwise transposition :tHom(Eq,Ep)Hom(Ep,Eq)\leavevmode\nobreak\ {}^{t}\colon\textup{Hom}(E_{q},E_{p})\rightarrow\textup{Hom}(E_{p}^{*},E_{q}^{*}). Then the solution of (2.12) satisfies

u(p,q)=Θp1u(q,p)tΘq,p,qM.\displaystyle u(p,q)=\Theta_{p}^{-1}u(q,p)^{t}\Theta_{q},\qquad p,q\in M. (2.19)

Additionally, we investigate the propagation of a family of singular solutions from a neighborhood of Σ\Sigma to all of MM by applying the well-posed Cauchy problem for singular sections treated in [BTW2015]. For that, we just have to ensure the existence of the restriction to Σ\Sigma by checking Hörmander’s criterion.

Theorem 2.4.

Let MM be a globally hyperbolic Lorentzian manifold, ΣM\Sigma\subset M a Cauchy hypersurface, π:EM\pi\colon E\rightarrow M a real or complex vector bundle over MM and P:C(M,E)C(M,E)P\colon C^{\infty}(M,E)\rightarrow C^{\infty}(M,E) a wave operator. Furthermore, let OMO\subset M be relatively compact, and for all pOp\in O, let v(p)𝒟(M,E,Ep)v(p)\in\mathscr{D}(M,E,E_{p}^{*})^{\prime} have spacelike compact support and only lightlike singular directions. Moreover, we assume pv(p)[φ]C(M,E)p\mapsto v(p)[\varphi]\in C^{\infty}(M,E^{*}) for fixed φ𝒟(M,E)\varphi\in\mathscr{D}(M,E^{*}). Then the Cauchy problem

{Pu(p)=0,u(p)|Σ=v(p)|Σ,νu(p)|Σ=νv(p)|Σ,\displaystyle\left\{\begin{array}[]{cl}Pu(p)&=0,\\[5.69054pt] u(p)\big{|}_{\Sigma}&=v(p)\big{|}_{\Sigma},\\[5.69054pt] \nabla_{\nu}u(p)\big{|}_{\Sigma}&=\nabla_{\nu}v(p)\big{|}_{\Sigma},\end{array}\right.

has a unique solution u(p)𝒟(M,E,Ep)u(p)\in\mathscr{D}(M,E,E^{*}_{p})^{\prime}, which has spacelike compact support and provides a smooth section pu(p)[φ]p\mapsto u(p)[\varphi] for each φ𝒟(M,E)\varphi\in\mathscr{D}(M,E^{*}).

Proof.

Let t:Mt\colon M\rightarrow\mathbb{R} be a Cauchy time function on MM such that Σ=t1(0)\Sigma=t^{-1}(0) (Theorem 1.3.13 of [BGP2007]). Therefore, the normal directions of Σ\Sigma are timelike and do not match the singular directions of vv, so v(p)|Σv(p)\big{|}_{\Sigma} and νv(p)|Σ\nabla_{\nu}v(p)\big{|}_{\Sigma} are well-defined and compactly supported distributions on Σ\Sigma for all pp due to Hörmander’s criterion ((8.2.3) of [Hör1990]). Recall that any compactly supported distribution lies in some Sobolev space HckH_{c}^{k} (see e.g. (31.6) of [Tre1967]), and hence, v(p)|ΣHck(Σ,EpE)v(p)\big{|}_{\Sigma}\in H_{c}^{k}(\Sigma,E_{p}^{*}\otimes E) and νv(p)|ΣHck1(Σ,EpE)\nabla_{\nu}v(p)\big{|}_{\Sigma}\in H^{k-1}_{c}(\Sigma,E_{p}^{*}\otimes E) for some kk\in\mathbb{R}. Thus, for all pp, Corollary 14 of [BTW2015] provides a unique solution

u(p)Csc0(t(M),Hk(Σ);EpE)Csc1(t(M),Hk1(Σ);EpE),u(p)\in C_{sc}^{0}\big{(}t(M),H^{k}(\Sigma_{\cdot});E_{p}^{*}\otimes E\big{)}\cap C_{sc}^{1}\big{(}t(M),H^{k-1}(\Sigma_{\cdot});E_{p}^{*}\otimes E\big{)},

where this intersection is commonly referred to as the space of finite kk-energy sections (see section 1.7 of [BTW2015] for details about them). Moreover, the mapping of initial data to the solution is a linear homeomorphism, i.e. continuity of the restriction v(p)(v(p)|Σ,νv(p)|Σ)v(p)\mapsto\big{(}v(p)\big{|}_{\Sigma},\nabla_{\nu}v(p)\big{|}_{\Sigma}\big{)} implies continuity of the map of distributions T:v(p)u(p)T\colon v(p)\mapsto u(p) for all pp.
For DD a differential operator, let (D(1)v)(p)\big{(}D_{(1)}v\big{)}(p) denote the distribution φ(D(v()[φ]))(p)\varphi\mapsto\big{(}D(v(\cdot)[\varphi])\big{)}(p). Since PP and DD act on different factors, (D(1)v)(p)\big{(}D_{(1)}v\big{)}(p) is linearly and continuously mapped to (D(1)u)(p)\big{(}D_{(1)}u\big{)}(p), that is, TT commutes with D(1)D_{(1)} (see the proof of Proposition A.1 in [FNW1981]). In particular, the map

p(D(1)v)(p)[φ](D(1)u)(p)[φ]=(D(u()[φ]))(p),φ𝒟(M,E),p\longmapsto\big{(}D_{(1)}v\big{)}(p)[\varphi]\longmapsto\big{(}D_{(1)}u\big{)}(p)[\varphi]=\big{(}D(u(\cdot)[\varphi])\big{)}(p),\qquad\varphi\in\mathscr{D}(M,E^{*}),

is continuous due to smoothness of the first arrow. This holds for all differential operators DD, which provides smoothness of pu(p)[φ]p\mapsto u(p)[\varphi] for fixed φ\varphi. ∎

2.2 The prototype

The Hadamard condition in the form (1.6) is a local condition and, moreover, the singularity structure of a bisolution is related to the corresponding differential operator essentially via its principal symbol. On these grounds, we start with the prototype setting P=P=\Box on M=Minkd,d3M=\mathbb{R}^{d}_{\mathrm{Mink}},\leavevmode\nobreak\ d\geq 3, since, from the viewpoint of the singularity structure of the solutions, this already incorporates the characteristic properties of the solutions for the general setting of wave operators on curved spacetimes. In order to obtain a decomposition like (1.8), we study fundamental solutions for \Box. It is not hard to check that these objects are invariant under the special orthochronous Lorentz group +\mathcal{L}^{\uparrow}_{+} and their singular support is cointained in the light cone CC. This makes them directly comparable outside CC meaning that we need to understand +\mathcal{L}^{\uparrow}_{+}-invariant distributions supported on CC.
For ̊d:=d\{0}\mathring{\mathbb{R}}^{d}:=\mathbb{R}^{d}\backslash\{0\}, we consider the submersions γ±:=γ|J±c\gamma_{\pm}:=\gamma\big{|}_{J_{\pm}^{c}}, where J±c:=̊d\J±J_{\pm}^{c}:=\mathring{\mathbb{R}}^{d}\backslash J_{\pm}. Then the pullback γ±\gamma^{*}_{\pm} maps distributions on \mathbb{R} to +\mathcal{L}^{\uparrow}_{+}-invariant distributions on J±cJ_{\pm}^{c} and moreover establishes a close connection to the well-known classification of distributions on \mathbb{R} supported in {0}\{0\}:

Theorem 2.5 (Théorème 2 of [Met1954]).

For any pair T±𝒟()T_{\pm}\in\mathscr{D}(\mathbb{R})^{\prime} with T+|<0=T|<0T_{+}|_{\mathbb{R}_{<0}}=T_{-}|_{\mathbb{R}_{<0}}, there is a +\mathcal{L}^{\uparrow}_{+}-invariant distribution T𝒟(̊d)T\in\mathscr{D}(\mathring{\mathbb{R}}^{d})^{\prime} given by T|Jc:=γ±T±T|_{J_{\mp}^{c}}:=\gamma_{\pm}^{*}T_{\pm}. Conversely, for any +\mathcal{L}^{\uparrow}_{+}-invariant T𝒟(̊d)T\in\mathscr{D}(\mathring{\mathbb{R}}^{d})^{\prime}, we find a pair T±𝒟()T_{\pm}\in\mathscr{D}(\mathbb{R})^{\prime} with T+|<0=T|<0T_{+}|_{\mathbb{R}_{<0}}=T_{-}|_{\mathbb{R}_{<0}} such that T|Jc=γ±T±T|_{J_{\mp}^{c}}=\gamma_{\pm}^{*}T_{\pm}.

For that, it is important to note that the proof of Theorem 2.5 particularly demonstrates surjectivity of (γ±)(\gamma_{\pm})_{*}.

Theorem 2.6 (Théorème 1 of [Met1954]).

Any +\mathcal{L}^{\uparrow}_{+}-invariant T𝒟(d)T\in\mathscr{D}(\mathbb{R}^{d})^{\prime} with supp(T){0}\text{supp}\,(T)\subset\{0\} is of the form T=k=0bkkδ0T=\sum_{k=0}^{\infty}b_{k}\cdot\Box^{k}\delta_{0} with bk0b_{k}\neq 0 for only finitely many kk.

There are two immediate consequences: First, every +\mathcal{L}^{\uparrow}_{+}-invariant distribution TT supported in CC has to be of the form

T=k=0(ak+γ+δ0(k)+akγδ0(k)+bkkδ0),\displaystyle T=\sum_{k=0}^{\infty}\left(a^{+}_{k}\cdot\gamma_{+}^{*}\delta_{0}^{(k)}+a^{-}_{k}\cdot\gamma_{-}^{*}\delta_{0}^{(k)}+b_{k}\cdot\Box^{k}\delta_{0}\right), (2.20)

where only finitely many coefficients ak±,bka_{k}^{\pm},b_{k} are non-zero. Recall that γ±δ0(k)[φ]=((γ±)φ)(k)(0)\gamma_{\pm}^{*}\delta_{0}^{(k)}[\varphi]=\big{(}(\gamma_{\pm})_{*}\varphi\big{)}^{(k)}(0) and the push-forward along a submersion is given by integration along the fibers. This leads to the second consequence, which is that every +\mathcal{L}^{\uparrow}_{+}-invariant measure supported on C±C_{\pm} is of the form adΩ±0+bδ0a\,\text{d}\Omega^{0}_{\pm}+b\delta_{0} (section IX.8 of [RS1975]), where

dΩ±0[φ]=d1φ(±x^,x^)2x^dx^,φ𝒟(d).\displaystyle\,\text{d}\Omega^{0}_{\pm}[\varphi]=\int_{\mathbb{R}^{d-1}}\frac{\varphi\big{(}\pm\|\hat{x}\|,\hat{x}\big{)}}{2\|\hat{x}\|}\leavevmode\nobreak\ \,\text{d}\hat{x},\qquad\varphi\in\mathscr{D}(\mathbb{R}^{d}). (2.21)

The most known examples of +\mathcal{L}^{\uparrow}_{+}-invariant supported on C±C_{\pm} are certain Riesz distributions (not all of them) and in fact, there are no other: For all α\alpha\in\mathbb{C} with Re(α)>d\text{Re}\left(\alpha\right)>d, they are defined as continuous functions via

R±α(x)={2C(α,d)γ(x)αd2,xJ±,0,otherwise,,C(α,d)=2απ2d2Γ(α2)Γ(αd2+1).\displaystyle R_{\pm}^{\alpha}(x)=\left\{\begin{array}[]{cl}2C(\alpha,d)\cdot\gamma(x)^{\frac{\alpha-d}{2}},&x\in J_{\pm},\\[5.69054pt] 0,&\text{otherwise},\end{array}\right.,\qquad C(\alpha,d)=\frac{2^{-\alpha}\pi^{\frac{2-d}{2}}}{\Gamma\left(\frac{\alpha}{2}\right)\Gamma\left(\frac{\alpha-d}{2}+1\right)}. (2.24)

One directly checks holomorphicity in α\alpha and calculates R±α+2=R±α\Box R^{\alpha+2}_{\pm}=R^{\alpha}_{\pm}, which provides an extension as distributions to all of \mathbb{C} via R±α:=kR±α+2k,k>d2Re(α)R^{\alpha}_{\pm}:=\Box^{k}R^{\alpha+2k}_{\pm},\leavevmode\nobreak\ k>\frac{d}{2}-\text{Re}\left(\alpha\right). Moreover, as shown in section 13.2 of [DK2010], we have R±0=δ0R^{0}_{\pm}=\delta_{0} and

γ±δ0(k)=Γ(d2(k+1))22k+3dπd22R±d2(k+1)|Jc,k0.\displaystyle\gamma_{\pm}^{*}\delta_{0}^{(k)}=\frac{\Gamma\left(\frac{d}{2}-(k+1)\right)}{2^{2k+3-d}\pi^{\frac{d-2}{2}}}\cdot R^{d-2(k+1)}_{\pm}\Big{|}_{J_{\mp}^{c}},\qquad k\in\mathbb{N}_{0}. (2.25)

Therefore, due to (2.20), every +\mathcal{L}^{\uparrow}_{+}-invariant distribution TT supported on CC is given by a linear combination of Riesz distributions:

T=k=0(λk+R+d2(k+1)+λkRd2(k+1)+βkR±2k).\displaystyle T=\sum_{k=0}^{\infty}\left(\lambda_{k}^{+}\cdot R_{+}^{d-2(k+1)}+\lambda_{k}^{-}\cdot R_{-}^{d-2(k+1)}+\beta_{k}\cdot R_{\pm}^{-2k}\right). (2.26)

The properties of the Riesz distributions combined with the uniqueness of GA,GRG_{A},G_{R} reveal R±2R_{\pm}^{2} as the advanced and retarded fundamental solution for \Box on Minkd\mathbb{R}^{d}_{\mathrm{Mink}}. In order to identify GF,GaFG_{F},G_{aF}, we introduce the functions (γ±i0)α(\gamma\pm i0)^{\alpha} for Re(α)>0\text{Re}\left(\alpha\right)>0, where ±i0\pm i0 stands for the direction from which the branch cut along <0\mathbb{R}_{<0} is approached. A meromorphic extension to all of \mathbb{C} as distributions is given by

(γ±i0)α=k(γ±i0)α+k4kj=1k(α+j)(α+j+d22),\displaystyle(\gamma\pm i0)^{\alpha}=\frac{\Box^{k}(\gamma\pm i0)^{\alpha+k}}{4^{k}\prod_{j=1}^{k}(\alpha+j)\big{(}\alpha+j+\frac{d-2}{2}\big{)}}, (2.27)

which is in fact holomorphic on all of {Re(α)>d2}\big{\{}\text{Re}\left(\alpha\right)>-\frac{d}{2}\big{\}}, and we calculate Resα=d2(γ±i0)α=(i)d1πd2Γ(d2)δ0\mathop{\mathrm{Res}}\limits_{\alpha=-\frac{d}{2}}(\gamma\pm i0)^{\alpha}=(\mp i)^{d-1}\frac{\pi^{\frac{d}{2}}}{\Gamma\left(\frac{d}{2}\right)}\cdot\delta_{0} (see section III.3 of [GS1967] for proofs). This already ensures all crucial properties:

Proposition 2.7.

The distributions (2.27) are symmetric and +\mathcal{L}^{\uparrow}_{+}-invariant. Moreover, for all natural numbers m<d2m<\frac{d}{2}, we have

(γ±i0)m=(1)m1Γ(d2m)4mΓ(m)Γ(d2)mlog(γ±i0),\displaystyle(\gamma\pm i0)^{-m}=\frac{(-1)^{m-1}\leavevmode\nobreak\ \Gamma\left(\frac{d}{2}-m\right)}{4^{m}\leavevmode\nobreak\ \Gamma(m)\Gamma\left(\frac{d}{2}\right)}\cdot\Box^{m}\log(\gamma\pm i0), (2.28)

and for d>2d>2, fundamental solutions for \Box are given by

S±:=(±i)d+1Γ(d22)4πd2(γ±i0)2d2.\displaystyle S_{\pm}:=(\pm i)^{d+1}\frac{\Gamma\left(\frac{d-2}{2}\right)}{4\pi^{\frac{d}{2}}}\cdot\left(\gamma\pm i0\right)^{\frac{2-d}{2}}. (2.29)
Proof.

Symmetry and +\mathcal{L}^{\uparrow}_{+}-invariance is obvious for Re(α)>0\text{Re}\left(\alpha\right)>0 and thus on all of \mathbb{C}, it follows from meromorphicity and the identity theorem. Holomorphicity on {Re(α)>d2}\big{\{}\text{Re}\left(\alpha\right)>-\frac{d}{2}\big{\}} ensures (γ±i0)m=limαm(γ±i0)α(\gamma\pm i0)^{-m}=\mathop{\lim}\limits_{\alpha\rightarrow-m}(\gamma\pm i0)^{\alpha}, which leads to

(γ±i0)m=mddα|α=m(γ±i0)α+m4m(d2m)d22(1)m1j=1m1(mj)=(1)m1Γ(d2m)4mΓ(m)Γ(d2)mlog(γ±i0).\displaystyle(\gamma\pm i0)^{-m}=\frac{\Box^{m}\left.\frac{\,\text{d}}{\,\text{d}\alpha}\right|_{\alpha=-m}(\gamma\pm i0)^{\alpha+m}}{4^{m}\left(\frac{d}{2}-m\right)\ldots\frac{d-2}{2}\leavevmode\nobreak\ (-1)^{m-1}\prod^{m-1}_{j=1}(m-j)}=\frac{(-1)^{m-1}\leavevmode\nobreak\ \Gamma\left(\frac{d}{2}-m\right)}{4^{m}\leavevmode\nobreak\ \Gamma(m)\Gamma\left(\frac{d}{2}\right)}\cdot\Box^{m}\log(\gamma\pm i0).

Particularly at α=2d2\alpha=\frac{2-d}{2}, it implies

(γ±i0)2d2=limαd2(γ±i0)α+1=42d2limαd2(α+d2)(γ±i0)α=Resα=d2(γ±i0)α=(i)d+14πd2Γ(d22)δ0.\displaystyle\Box\left(\gamma\pm i0\right)^{\frac{2-d}{2}}=\mathop{\lim}\limits_{\alpha\rightarrow-\frac{d}{2}}\Box\left(\gamma\pm i0\right)^{\alpha+1}=4\cdot\frac{2-d}{2}\underbrace{\mathop{\lim}\limits_{\alpha\rightarrow-\frac{d}{2}}\left(\alpha+\frac{d}{2}\right)\left(\gamma\pm i0\right)^{\alpha}}_{=\mathop{\mathrm{Res}}\limits_{\alpha=-\frac{d}{2}}(\gamma\pm i0)^{\alpha}}=(\mp i)^{d+1}\frac{4\pi^{\frac{d}{2}}}{\Gamma\left(\frac{d-2}{2}\right)}\cdot\delta_{0}.\qed

Let us now come to the two-point-function in the prototype case, which is characterized by Wightman’s axioms and d’Alembert’s equation. In particular, the Hadamard condition corresponds to the spectral condition given by some constraint on the Fourier transform. Due to the symmetry conditions implied by these axioms, the two-point-function is completely determined by some +\mathcal{L}^{\uparrow}_{+}-invariant distribution WW on ̊d\mathring{\mathbb{R}}^{d}, which solves W=0\Box W=0. The spectral condition then leads to the general form W=adΩˇ0+,a,W=a\,\text{d}\widecheck{\Omega}_{0}^{+},\leavevmode\nobreak\ a\in\mathbb{R}, ([RS1975, Ste2000]) and we choose a=(2π)2d2a=(2\pi)^{\frac{2-d}{2}}. Here f^\widehat{f} denotes the Fourier transformation with ,\left\langle\!\left\langle\cdot,\cdot\right\rangle\!\right\rangle used in the exponential, i.e. ordinary Fourier transformation in space and inverse Fourier transformation in time direction e0e_{0}, and fˇ\widecheck{f} the corresponding inverse operation. The maps

Δ±:𝒯±,z±i(2π)d1Ceiz,pdΩ0(p)\displaystyle\Delta^{\pm}:\qquad\mathcal{T}_{\pm}\longrightarrow\mathbb{C},\qquad z\longmapsto\pm\frac{i}{(2\pi)^{d-1}}\int_{C_{\mp}}e^{i\left\langle\!\left\langle z,p\right\rangle\!\right\rangle}\,\text{d}\Omega^{\mp}_{0}(p) (2.30)

are analytic in the complex forward/backward tube 𝒯±:={zd|Im(z)I±}\mathcal{T}_{\pm}:=\left\{z\in\mathbb{C}^{d}\,\big{|}\,\text{Im}\left(z\right)\in I_{\pm}\right\}, they are +\mathcal{L}^{\uparrow}_{+}-invariant and related via Δ+(z)=Δ(z),zT+\Delta^{+}(z)=-\Delta^{-}(-z),\leavevmode\nobreak\ z\in T_{+}. Hence, for each z𝒯±z\in\mathcal{T}_{\pm}, we obtain Δ±(z)=Δ±(x±iεe0)=:Δε(x)\Delta^{\pm}(z)=\Delta^{\pm}(x\pm i\varepsilon e_{0})=:\Delta_{\varepsilon}(x) for some xd,ε>0x\in\mathbb{R}^{d},\leavevmode\nobreak\ \varepsilon>0, and one directly calculates W=ilimε0ΔεW=i\mathop{\lim}\limits_{\varepsilon\rightarrow 0}\Delta_{\varepsilon}^{-} in the distributional sense, that is

W(x)=1(2π)d1limε0C+eixiεe0,pdΩ0+(p).\displaystyle W(x)=\frac{1}{(2\pi)^{d-1}}\mathop{\lim}\limits_{\varepsilon\rightarrow 0}\int_{C_{+}}e^{i\left\langle\!\left\langle x-i\varepsilon e_{0},p\right\rangle\!\right\rangle}\,\text{d}\Omega^{+}_{0}(p). (2.31)

Due to the inverse Cauchy Schwarz inequality, γ\gamma maps 𝒯±\mathcal{T}_{\pm} to \0\mathbb{C}\backslash\mathbb{R}_{\geq 0}, and hence γ(z)\\sqrt{\gamma(z)}\in\mathbb{C}\backslash\mathbb{R} for all z𝒯±z\in\mathcal{T}_{\pm}. Let σ:𝒯±±:={±Im(z)>0}\sigma\colon\mathcal{T}_{\pm}\rightarrow\mathbb{H}_{\pm}:=\{\pm\text{Im}\left(z\right)>0\} denote the square root of γ\gamma with appropriately chosen sign such that again +\mathcal{L}^{\uparrow}_{+}-invariance yields Δ±(z)=Δ±(σ(z)e0),z𝒯±\Delta^{\pm}(z)=\Delta^{\pm}\big{(}\sigma(z)e_{0}\big{)},\leavevmode\nobreak\ z\in\mathcal{T}_{\pm}. In particular, z=x±iεe0z=x\pm i\varepsilon e_{0} is mapped to

σ(x±iεe0)=sgn(x0)γε±(x),γε±(x)\displaystyle\sigma(x\pm i\varepsilon e_{0})=\text{sgn}(x_{0})\sqrt{\gamma_{\varepsilon}^{\pm}(x)},\qquad\qquad\gamma_{\varepsilon}^{\pm}(x) :=γ(x±iεe0)=γ(x)ε2±2iεx0,\displaystyle:=\gamma(x\pm i\varepsilon e_{0})=\gamma(x)-\varepsilon^{2}\pm 2i\varepsilon x_{0}, (2.32)

which lets us evaluate the integrals Δε±\Delta^{\pm}_{\varepsilon} and thus derive a local expression for WW:

Δε±=±iΓ(d22)4πd2(γε±)2d2W=Γ(d22)4πd2limε0(γε)2d2.\displaystyle\Delta^{\pm}_{\varepsilon}=\pm\frac{i\Gamma\left(\frac{d-2}{2}\right)}{4\pi^{\frac{d}{2}}}\cdot\big{(}-\gamma^{\pm}_{\varepsilon}\big{)}^{\frac{2-d}{2}}\qquad\Longrightarrow\qquad W=\frac{\Gamma\left(\frac{d-2}{2}\right)}{4\pi^{\frac{d}{2}}}\cdot\mathop{\lim}\limits_{\varepsilon\rightarrow 0}\big{(}-\gamma^{-}_{\varepsilon}\big{)}^{\frac{2-d}{2}}. (2.33)

It is a well-known procedure to extract the advanced and retarded fundamental solution ΔA\Delta^{A} and ΔR\Delta^{R}:
We extract the causal propagator ΔC:=2iWa\Delta^{C}:=-2iW^{a} from the antisymmetric part of WW, which turns out to be supported only in J={γ(x)0}J=\big{\{}\gamma(x)\geq 0\big{\}} and is a solution of d’Alembert’s equation as well since Wa=Δ++ΔW^{a}=\Delta^{+}+\Delta^{-}. Hörmander’s criterion allows us to multiply ΔC\Delta^{C} with the step function H0H_{0} with respect to the time coordinate and integration by parts reveals (H0ΔC)=δ0\Box\big{(}H_{0}\cdot\Delta^{C}\big{)}=-\delta_{0}. Therefore, the aforesaid fundamental solutions are given by

ΔA:=(1H0)ΔC,ΔR:=H0ΔC.\displaystyle\Delta^{A}:=\left(1-H_{0}\right)\cdot\Delta^{C},\qquad\Delta^{R}:=-H_{0}\cdot\Delta^{C}. (2.34)

Due to their support properties, they are unique, so we directly conclude ΔA=R2\Delta^{A}=R^{2}_{-} and ΔR=R+2\Delta^{R}=R^{2}_{+}.
Following Radzikowski’s results, the remaining two fundamental solutions are given by

ΔF=iW+ΔA,ΔaF=iW+ΔR,\displaystyle\Delta^{F}=iW+\Delta^{A},\qquad\qquad\Delta^{aF}=-iW+\Delta^{R}, (2.35)

for which one directly calculates ΔF=S\Delta^{F}=S_{-} and ΔaF=S+\Delta^{aF}=S_{+} outside of CC. It follows that the distributions ΔFS\Delta^{F}-S_{-} and ΔaFS+\Delta^{aF}-S_{+} are symmetric and +\mathcal{L}^{\uparrow}_{+}-invariant solutions of d’Alembert’s equation. Moreover, their support is contained in CC and they are homogeneous of degree 2d2-d, so they have to vanish also on the light cone according to our prior characterization (2.26). Combining both equations (2.35) leads to the following form of the Wightman distribution, that is the Hadamard two-point-function in the prototype case:

W=i2(S+S+R2R+2).\displaystyle W=\frac{i}{2}\big{(}S_{+}-S_{-}+R^{2}_{-}-R^{2}_{+}\big{)}. (2.36)

3 Local Construction

3.1 Families of Riesz-like distributions

We proceed with the local construction of Hadamard bidistributions in a setting (M,E,P)(M,E,P) as in Definition 1.1 with PP a wave operator. Inspired by the local construction of the advanced and retarded parametrices for PP in [Gün1988, BGP2007], we introduce families of distributions similar to the Riesz distributions (2.24) but containing S±S_{\pm} instead, so for Re(α)>0\text{Re}\left(\alpha\right)>0, we introduce the distributions

L±α:=C(α,d)(γ±i0)αd2,C(α,d):=2απ2d2Γ(α2)Γ(αd2+1).\displaystyle L^{\alpha}_{\pm}:=C(\alpha,d)\cdot(\gamma\pm i0)^{\frac{\alpha-d}{2}},\qquad\qquad C(\alpha,d):=\frac{2^{-\alpha}\pi^{\frac{2-d}{2}}}{\Gamma\left(\frac{\alpha}{2}\right)\Gamma\left(\frac{\alpha-d}{2}+1\right)}. (3.1)

Very similar to R±αR^{\alpha}_{\pm}, we have L±α+2=L±α\Box L^{\alpha+2}_{\pm}=L^{\alpha}_{\pm} leading to a holomorphic extension to all of \mathbb{C} as distributions, for which then, moreover, analogous relations hold:

Proposition 3.1.

For all α\alpha\in\mathbb{C}, we have

  • (1)

    γL±α=α(αd+2)L±α+2\gamma\cdot L^{\alpha}_{\pm}=\alpha(\alpha-d+2)L^{\alpha+2}_{\pm},

  • (2)

    gradγL±α=2αgradL±α+2\text{grad}\gamma\cdot L^{\alpha}_{\pm}=2\alpha\leavevmode\nobreak\ \text{grad}L^{\alpha+2}_{\pm},

  • (3)

    L±α+2=L±α\Box L^{\alpha+2}_{\pm}=L_{\pm}^{\alpha},

  • (4)

    L±d2n=0,nL^{d-2n}_{\pm}=0,\enspace n\in\mathbb{N},

  • (5)

    L+d+2n=Ld+2n,n0L^{d+2n}_{+}=L^{d+2n}_{-},\enspace n\in\mathbb{N}_{0},

  • (6)

    if Re(α)>0\text{Re}\left(\alpha\right)>0, then L±αL^{\alpha}_{\pm} are distributions of order at most κd:=2d2\kappa_{d}:=2\cdot\left\lceil\frac{d}{2}\right\rceil.

Proof.

The proofs for (1) - (3) and (6) are similar to Proposition 1.2.4 of [BGP2007]. (4): Due to (3), integration by parts yields

L±d2n[φ]=L±d[nφ]=C(d,d)d(nφ)(x)dx=0,φ𝒟(d),n.L^{d-2n}_{\pm}[\varphi]=L^{d}_{\pm}\big{[}\Box^{n}\varphi\big{]}=C(d,d)\int_{\mathbb{R}^{d}}\big{(}\Box^{n}\varphi\big{)}(x)\,\text{d}x=0,\qquad\varphi\in\mathscr{D}(\mathbb{R}^{d}),\leavevmode\nobreak\ n\in\mathbb{N}.

(5): Follows from (γ±i0)n=γn(\gamma\pm i0)^{n}=\gamma^{n} for all n0n\in\mathbb{N}_{0}. ∎

With regard to (5), we omit the "±\pm" if α=d+2n\alpha=d+2n. A crucial property of the Riesz distributions is R±0=δ0R_{\pm}^{0}=\delta_{0}, which, due to (4), fails to be true for L±αL_{\pm}^{\alpha} in the even-dimensional case. However, it holds for odd dd because in that case, Γ(d22)Γ(4d2)=πsin(d22π)=(1)d+12π\Gamma\left(\frac{d-2}{2}\right)\cdot\Gamma\left(\frac{4-d}{2}\right)=\frac{\pi}{\sin\left(\frac{d-2}{2}\pi\right)}=(-1)^{\frac{d+1}{2}}\pi leads to L±2=S±L^{2}_{\pm}=S_{\pm}. In even dimensions, we need a further family that is "more singular at even numbers", so for α\{,d2,d,d+2,}\alpha\in\mathbb{C}\backslash\{\ldots,d-2,d,d+2,\ldots\}, we introduce

L~±α:=(±i)dα1sin(dα2π)L±α=C~(α,d)(γ±i0)αd2,C~(α,d):=(±i)dα1Γ(dα2)2απd2Γ(α2).\displaystyle\widetilde{L}^{\alpha}_{\pm}:=\frac{(\pm i)^{d-\alpha-1}}{\sin\left(\frac{d-\alpha}{2}\pi\right)}\cdot L^{\alpha}_{\pm}=\widetilde{C}(\alpha,d)\cdot(\gamma\pm i0)^{\frac{\alpha-d}{2}},\qquad\widetilde{C}(\alpha,d):=\frac{(\pm i)^{d-\alpha-1}\Gamma\left(\frac{d-\alpha}{2}\right)}{2^{\alpha}\leavevmode\nobreak\ \pi^{\frac{d}{2}}\leavevmode\nobreak\ \Gamma\left(\frac{\alpha}{2}\right)}. (3.2)

Indeed, the zeros of L±αL^{\alpha}_{\pm} and the poles of the prefactor in (3.2) compensate, and hence, L~±α\widetilde{L}^{\alpha}_{\pm} exist as distributions for all α=d2n,n\alpha=d-2n,\leavevmode\nobreak\ n\in\mathbb{N}, i.e. αL~±α[φ]\alpha\mapsto\widetilde{L}^{\alpha}_{\pm}[\varphi] are meromorphic functions with simple poles at α=d,d+2,\alpha=d,d+2,\ldots for fixed φ𝒟(d)\varphi\in\mathscr{D}(\mathbb{R}^{d}). From the definition follows that (1), (2), (3), (6) of Proposition 3.1 also hold for L~±α\widetilde{L}^{\alpha}_{\pm} and in addition, we have L~±2=S±\widetilde{L}^{2}_{\pm}=S_{\pm} and at all non-pole-integers

L~±α={L±α,dα odd,±iπLdnlog(γ±i0),α=d2n,n.\displaystyle\widetilde{L}^{\alpha}_{\pm}=\left\{\begin{array}[]{cl}L^{\alpha}_{\pm},&d-\alpha\text{ odd},\\[5.69054pt] \pm\frac{i}{\pi}L^{d}\Box^{n}\log(\gamma\pm i0),&\alpha=d-2n,\leavevmode\nobreak\ n\in\mathbb{N}.\end{array}\right. (3.5)
Remark 3.2.

Also for d=1d=1 and d=2d=2, (3.1) - (3.5) provide the respective fundamental solutions:

L2(x)=|x|2,L~±2=±i2πlog(γ±i0),L^{2}(x)=\frac{|x|}{2},\qquad\qquad\widetilde{L}^{2}_{\pm}=\pm\frac{i}{2\pi}\log(\gamma\pm i0),

such that L~±0=δ0\widetilde{L}^{0}_{\pm}=\delta_{0} also in these cases.

Following section 1.4 of [BGP2007], we now transfer the families {L±α}α,{L~±α}α\{d,d+2,}\{L_{\pm}^{\alpha}\}_{\alpha\in\mathbb{C}},\{\widetilde{L}_{\pm}^{\alpha}\}_{\alpha\in\mathbb{C}\backslash\{d,d+2,\ldots\}} locally to MM. For pMp\in M and ΩM\Omega\subset M geodesically starshaped with respect to pp, let

Γp(q):=γ(expp1(q)),μp(q):=|det(dexpp)|expp1(q)|,qΩ,\Gamma_{p}(q):=\gamma\big{(}\exp_{p}^{-1}(q)\big{)},\qquad\mu_{p}(q):=\big{|}\det\big{(}\,\text{d}\exp_{p}\big{)}\big{|}_{\exp_{p}^{-1}(q)}\big{|},\qquad q\in\Omega,

denote the squared Lorentz distance to pp and the distortion function (section 1.3 of [BGP2007]), which is related to the van-Vleck-Morette-determinant via Δ=1μ\Delta=\frac{1}{\mu}. On Ω\Omega, we define

L±Ω(α,p)[φ]:=L±α[(μpφ)expp],φ𝒟(Ω),\displaystyle L_{\pm}^{\Omega}(\alpha,p)[\varphi]:=L_{\pm}^{\alpha}\left[(\mu_{p}\varphi)\circ\exp_{p}\right],\qquad\varphi\in\mathscr{D}(\Omega), (3.6)

which provides holomorphic maps αL±Ω(α,p)[φ]\alpha\mapsto L_{\pm}^{\Omega}(\alpha,p)[\varphi] and, analogously, meromorphic maps αL~±Ω(α,p)[φ]\alpha\mapsto\widetilde{L}_{\pm}^{\Omega}(\alpha,p)[\varphi] with simple poles at α=d,d+2,\alpha=d,d+2,\ldots.

Proposition 3.3.

Let pMp\in M and ΩM\Omega\subset M be geodesically starshaped with respect to pp. Then, for all α\alpha\in\mathbb{C}, we have:

  • (1)

    For Re(α)>d\text{Re}\left(\alpha\right)>d, the maps pL±Ω(α,p)p\mapsto L^{\Omega}_{\pm}(\alpha,p) are continuous on Ω\Omega and given by

    L±Ω(α,p)=C(α,d)(Γp±i0)αd2.\displaystyle L^{\Omega}_{\pm}(\alpha,p)=C(\alpha,d)\left(\Gamma_{p}\pm i0\right)^{\frac{\alpha-d}{2}}. (3.7)
  • (2)

    ΓpL±Ω(α,p)=α(αd+2)L±Ω(α+2,p)\Gamma_{p}\cdot L^{\Omega}_{\pm}(\alpha,p)=\alpha(\alpha-d+2)\cdot L^{\Omega}_{\pm}(\alpha+2,p),

  • (3)

    gradΓpL±Ω(α,p)=2αgradL±Ω(α+2,p)\text{grad}\,\Gamma_{p}\cdot L^{\Omega}_{\pm}(\alpha,p)=2\alpha\leavevmode\nobreak\ \text{grad}L^{\Omega}_{\pm}(\alpha+2,p),

  • (4)

    L±Ω(α+2,p)=(Γp2d2α+1)L±Ω(α,p),α0\Box L^{\Omega}_{\pm}(\alpha+2,p)=\left(\frac{\Box\Gamma_{p}-2d}{2\alpha}+1\right)\cdot L^{\Omega}_{\pm}(\alpha,p),\quad\alpha\neq 0.

  • (5)

    For Re(α)>0\text{Re}\left(\alpha\right)>0, (3.6) yield distributions of order at most κd\kappa_{d}. Moreover, there is an open neighborhood UU of pp and some C>0C>0 such that

    |L±Ω(α,q)[φ]|CφCκd(Ω),qU,φ𝒟(Ω).\left|L^{\Omega}_{\pm}(\alpha,q)[\varphi]\right|\leq C\cdot\|\varphi\|_{C^{\kappa_{d}}(\Omega)},\qquad q\in U,\leavevmode\nobreak\ \varphi\in\mathscr{D}(\Omega).
  • (6)

    Let UΩU\subset\Omega be an open neighborhood of pp such that Ω\Omega is geodesically starshaped with respect to all qUq\in U. Furthermore, let Re(α)>0\text{Re}\left(\alpha\right)>0 and VCκd+k(U×Ω)V\in C^{\kappa_{d}+k}(U\times\Omega) such that suppV(q,)Ω\text{supp}\,V(q,\cdot)\subset\Omega is compact for all qUq\in U. Then qL±Ω(α,q)[V(q,)]Ck(U)q\mapsto L^{\Omega}_{\pm}(\alpha,q)[V(q,\cdot)]\in C^{k}(U).

  • (7)

    For all φCck(Ω)\varphi\in C^{k}_{c}(\Omega), the map αL±Ω(α,p)[φ]\alpha\mapsto L^{\Omega}_{\pm}(\alpha,p)[\varphi] is holomorphic on {Re(α)>d2k2}\left\{\text{Re}\left(\alpha\right)>d-2\lfloor\frac{k}{2}\rfloor\right\}.

On the domain of holomorphicity of L~±Ω(α,p)\widetilde{L}^{\Omega}_{\pm}(\alpha,p), the statements (1) - (7) remain true, when we replace L±Ω(α,p)L^{\Omega}_{\pm}(\alpha,p) and C(α,d)C(\alpha,d) by L~±Ω(α,p)\widetilde{L}^{\Omega}_{\pm}(\alpha,p) and C~(α,d)\widetilde{C}(\alpha,d).

  • (8)

    For dαd-\alpha an odd integer, we have L±Ω(α,p)=L~±Ω(α,p)L^{\Omega}_{\pm}(\alpha,p)=\widetilde{L}^{\Omega}_{\pm}(\alpha,p).

  • (9)

    L~±Ω(0,p)=δp\widetilde{L}_{\pm}^{\Omega}(0,p)=\delta_{p}.

  • (10)

    For all nn\in\mathbb{N}, we have L±Ω(d2n,p)=0L^{\Omega}_{\pm}(d-2n,p)=0 and L+Ω(d+2n2,p)=LΩ(d+2n2,p)L^{\Omega}_{+}(d+2n-2,p)=L^{\Omega}_{-}(d+2n-2,p).

Proof.

The proofs of (1) - (7) are similar to those of Proposition 1.4.2 of [BGP2007] and by employing the respective properties of L~±α\widetilde{L}^{\alpha}_{\pm}, they also apply here. Moreover, (8) follows directly from (3.5), (10) from 3.1 and (9) from L~±0=δ0\widetilde{L}^{0}_{\pm}=\delta_{0} and μp(p)=det(idTpM)=1\mu_{p}(p)=\det(\text{id}_{T_{p}M})=1. ∎

Corollary 3.4.

Proposition 3.3 leads to the following relations:

  • For dd odd, we have L±Ω(0,p)=δpL_{\pm}^{\Omega}(0,p)=\delta_{p}.

  • For Ω\Omega moreover convex, i.e. starshaped w.r.t. all pΩp\in\Omega, α\alpha\in\mathbb{C} and u𝒟(Ω×Ω)u\in\mathscr{D}(\Omega\times\Omega), we have

    ΩL±Ω(α,p)[u(p,)]dV(p)=ΩL±Ω(α,q)[u(,q)]dV(q)\displaystyle\int_{\Omega}L_{\pm}^{\Omega}(\alpha,p)[u(p,\cdot)]\,\text{d}V(p)=\int_{\Omega}L^{\Omega}_{\pm}(\alpha,q)[u(\cdot,q)]\,\text{d}V(q) (3.8)

    and similarly for L~±Ω(α)\widetilde{L}^{\Omega}_{\pm}(\alpha).

3.2 The Hadamard series

Let EE be a real vector bundle over MM, ΩM\Omega\subset M geodesically starshaped with respect to some pΩp\in\Omega and P:C(M,E)C(M,E)P\colon C^{\infty}(M,E)\rightarrow C^{\infty}(M,E) a wave operator. Adopting the approach pursued in section 5.2 of [Gar1964], we start the deduction of local expressions for Feynman and anti-Feynman parametrices for PP at pp by taking the following ansatz of a formal Hadamard series:

±(p):={k=0UpkL±Ω(2k+2,p)+k=d22WpkLΩ(2k+2,p),d odd,k=0d42UpkL~±Ω(2k+2,p)±iπk=d22(Upklog(Γp±i0)+Wpk)LΩ(2k+2,p),d even,\displaystyle\mathscr{L}_{\pm}(p):=\left\{\begin{array}[]{cl}\mathop{\sum}\limits_{k=0}^{\infty}U_{p}^{k}L^{\Omega}_{\pm}(2k+2,p)+\mathop{\sum}\limits_{k=\frac{d-2}{2}}^{\infty}W_{p}^{k}L^{\Omega}(2k+2,p),&d\text{ odd},\\[14.22636pt] \mathop{\sum}\limits_{k=0}^{\frac{d-4}{2}}U_{p}^{k}\widetilde{L}^{\Omega}_{\pm}(2k+2,p)\pm\frac{i}{\pi}\mathop{\sum}\limits_{k=\frac{d-2}{2}}^{\infty}\big{(}U_{p}^{k}\log(\Gamma_{p}\pm i0)+W_{p}^{k}\big{)}L^{\Omega}(2k+2,p),&d\text{ even},\end{array}\right. (3.11)

with coefficients Upk,WpkC(Ω,EpE)U_{p}^{k},W_{p}^{k}\in C^{\infty}\left(\Omega,E_{p}^{*}\otimes E\right) yet to be determined. For φ𝒟(Ω,E)\varphi\in\mathscr{D}(\Omega,E^{*}), we identify Upkφ,WpkφU_{p}^{k}\varphi,W_{p}^{k}\varphi with EpE_{p}^{*}-valued test functions (see section 2.1 of [BGP2007]), so ±(p)\mathscr{L}_{\pm}(p) is (formally) understood as a distribution mapping 𝒟(Ω,E)\mathscr{D}(\Omega,E^{*}) to the complexified fiber EpE_{p}^{*}\otimes_{\mathbb{R}}\mathbb{C}. Similar to the procedure in chapter 2 of [BGP2007], we determine Upk,WpkU_{p}^{k},W_{p}^{k} by formally demanding P±(p)=δpP\mathscr{L}_{\pm}(p)=\delta_{p}, which, by imposing the initial condition Up0(p)=idEpU_{p}^{0}(p)=\text{id}_{E_{p}^{*}} and by means of (A.1) and Corollary B.1, leads to the transport equations

2kPUpk1\displaystyle 2k\leavevmode\nobreak\ PU_{p}^{k-1} =gradΓpUpk(12Γpd+2k)Upk,k0,\displaystyle=\nabla_{\text{grad}\Gamma_{p}}U_{p}^{k}-\left(\frac{1}{2}\Box\Gamma_{p}-d+2k\right)U_{p}^{k},\qquad k\in\mathbb{N}_{0}, (3.12)
2kPWpk1\displaystyle 2k\leavevmode\nobreak\ PW_{p}^{k-1} ={gradΓpWpk(12Γpd+2k)Wpk,k+12,kd2,gradΓpWpk(12Γp+2kd)Wpk+2kPUpk1kd222Upk,k,kd2.\displaystyle=\left\{\begin{array}[]{cl}\nabla_{\text{grad}\Gamma_{p}}W_{p}^{k}-\left(\frac{1}{2}\Box\Gamma_{p}-d+2k\right)W_{p}^{k},&k+\frac{1}{2}\in\mathbb{N},\leavevmode\nobreak\ k\geq\frac{d}{2},\\[5.69054pt] \nabla_{\text{grad}\Gamma_{p}}W_{p}^{k}-\left(\frac{1}{2}\Box\Gamma_{p}+2k-d\right)W_{p}^{k}+\frac{2k\leavevmode\nobreak\ PU_{p}^{k-1}}{k-\frac{d-2}{2}}-2U_{p}^{k},&k\in\mathbb{N},\leavevmode\nobreak\ k\geq\frac{d}{2}.\end{array}\right. (3.15)

For the full calculation, we refer to section A.2 in the appendix of this work.

Remark 3.5.

Note that there is no constraint on Wpd22W_{p}^{\frac{d-2}{2}}, which is therefore free to choose. Hence, even if (3.11) converges, the requirement P±(p)=δpP\mathscr{L}_{\pm}(p)=\delta_{p} determines ±(p)\mathscr{L}_{\pm}(p) only up to smooth solutions k=0ck,d(WpkW~pk)Γk\mathop{\sum}\limits_{k=0}^{\infty}c_{k,d}(W_{p}^{k}-\widetilde{W}_{p}^{k})\leavevmode\nobreak\ \Gamma^{k} with Wpk,W~pkW_{p}^{k},\widetilde{W}_{p}^{k} arising from different choices of Wpd22W_{p}^{\frac{d-2}{2}}.

Proposition 3.6.

Let OΩO\subset\Omega be a non-empty domain such that Ω\Omega is geodesically starshaped with respect to all pOp\in O. For any Wd22C(O×Ω,EE)W_{\frac{d-2}{2}}\in C^{\infty}(O\times\Omega,E^{*}\boxtimes E), there are unique and smooth solutions of (3.12) and (3.15) given by

U0(p,q)\displaystyle U_{0}(p,q) =Πqpμ(p,q),\displaystyle=\frac{\Pi_{q}^{p}}{\sqrt{\mu(p,q)}},
Uk(p,q)\displaystyle U_{k}(p,q) =kU0(p,q)01tk1U0(p,ϕpq(t))1(P(2)Uk1)(p,ϕpq(t))dt,k1,\displaystyle=-kU_{0}(p,q)\int_{0}^{1}t^{k-1}U_{0}\big{(}p,\phi_{pq}(t)\big{)}^{-1}\big{(}P_{(2)}U_{k-1}\big{)}\big{(}p,\phi_{pq}(t)\big{)}\,\text{d}t,\qquad k\geq 1, (3.16)
Wk(p,q)\displaystyle W_{k}(p,q) =kU0(p,q)01tk1U0(p,ϕpq(t))1W^k1(p,ϕpq(t))dt,kd2.\displaystyle=-kU_{0}(p,q)\int_{0}^{1}t^{k-1}U_{0}\big{(}p,\phi_{pq}(t)\big{)}^{-1}\widehat{W}_{k-1}\big{(}p,\phi_{pq}(t)\big{)}\,\text{d}t,\qquad k\geq\frac{d}{2}. (3.17)

Πqp:EpEq\Pi^{p}_{q}\colon E_{p}\rightarrow E_{q} denotes the \nabla-parallel transport, ϕpq:[0,1]Ω\phi_{pq}\colon[0,1]\rightarrow\Omega the unique geodesic connecting p,qp,q (A.5) and

W^k1:={P(2)Wk1,k+12,kd2,P(2)(Wk1Uk1kd22)+Ukk,k,kd2.\displaystyle\widehat{W}_{k-1}:=\left\{\begin{array}[]{cl}P_{(2)}W_{k-1},&k+\frac{1}{2}\in\mathbb{N},\leavevmode\nobreak\ k\geq\frac{d}{2},\\[5.69054pt] P_{(2)}\left(W_{k-1}-\frac{U_{k-1}}{k-\frac{d-2}{2}}\right)+\frac{U_{k}}{k},&k\in\mathbb{N},\leavevmode\nobreak\ k\geq\frac{d}{2}.\end{array}\right.
Proof.

The transport equations (3.12) and for half-integer kk also (3.15) coincide with (2.3) of [BGP2007]. Therefore, UkU_{k} and for k+12k+\frac{1}{2}\in\mathbb{N} also WkW_{k} are the Hadamard coefficients given by (3.16) and (3.17) due to Proposition 2.3.1 of [BGP2007]. For integer kk, we can apply the same proof for WkW_{k} with PWpk1PW_{p}^{k-1} replaced by W^pk1\widehat{W}_{p}^{k-1} everywhere, for which the same procedure then leads to (3.17). ∎

Corollary 3.7.

For Ω\Omega convex, EE equipped with a non-degenerate inner product and PP formally self-adjoint, we have symmetry of UkU_{k} for all k0k\in\mathbb{N}_{0} and, in case of symmetric Wd22W_{\frac{d-2}{2}}, of WkW_{k} for half-integer kd2k\geq\frac{d}{2} in the sense of (A.3).

Remark 3.8.

Note that in the odd-dimensional case, Wpd22=0W_{p}^{\frac{d-2}{2}}=0 leads to Wpk=0W_{p}^{k}=0 for all kk, whereas, as a consequence of the coupling with UpkU_{p}^{k}, for even dimensions, we have Wpk0W_{p}^{k}\neq 0, in general.
More remarkably, the WkW_{k}’s fail to be symmetric in the even-dimensional case, i.e. for integer kk, even if we chose Wd22W_{\frac{d-2}{2}} to be symmetric. This phenomenon is closely related to the conformal trace anomaly, which indeed does not occur in odd-dimensional spacetimes [Wal1978, Wal1994, DF2008].

3.3 Local parametrices and Hadamard bidistributions

From now on, let ΩM\Omega\subset M always denote a convex domain, i.e. geodesically starshaped with respect to all pΩp\in\Omega. We referred to (3.11) as formal since in general, the series do not converge, and we merely employed it in order to extract the transport equations (3.12) - (3.15). Nevertheless, it leads to left parametrices by some well-known procedure [Fri1975, Gün1988, BGP2007], which smoothly cuts off ±(p)\mathscr{L}_{\pm}(p) away from its singular support and leads to convergent series on relatively compact domains. Due to the derivatives arising from the cut-off, this results in left parametrices for PP at pp rather than fundamental solutions. To be more precise, for N>d2N>\frac{d}{2}, some sequence {εk}kN(0,1]\{\varepsilon_{k}\}_{k\geq N}\subset(0,1] and σ𝒟([1,1],[0,1])\sigma\in\mathscr{D}\big{(}[-1,1],[0,1]\big{)} with σ|[12,12]1\sigma\big{|}_{\left[-\frac{1}{2},\frac{1}{2}\right]}\equiv 1, we define

~±(p):={k=0U~pkL±Ω(2k+2,p)+k=d22W~pkLΩ(2k+2,p),d odd,k=0d42UpkL~±Ω(2k+2,p)±iπk=d22(U~pklog(Γp±i0)+W~pk)LΩ(2k+2,p),d even,\displaystyle\widetilde{\mathscr{L}}_{\pm}(p):=\left\{\begin{array}[]{cl}\mathop{\sum}\limits_{k=0}^{\infty}\widetilde{U}_{p}^{k}L^{\Omega}_{\pm}(2k+2,p)+\mathop{\sum}\limits_{k=\frac{d-2}{2}}^{\infty}\widetilde{W}_{p}^{k}L^{\Omega}(2k+2,p),&d\text{ odd},\\[14.22636pt] \mathop{\sum}\limits_{k=0}^{\frac{d-4}{2}}U_{p}^{k}\widetilde{L}^{\Omega}_{\pm}(2k+2,p)\pm\frac{i}{\pi}\mathop{\sum}\limits_{k=\frac{d-2}{2}}^{\infty}\big{(}\widetilde{U}_{p}^{k}\log(\Gamma_{p}\pm i0)+\widetilde{W}_{p}^{k}\big{)}L^{\Omega}(2k+2,p),&d\text{ even},\end{array}\right. (3.20)

where

U~k:={Uk,k<N,(σΓεk)Uk,kN,W~k:={Wk,k<N,(σΓεk)Wk,kN.\displaystyle\widetilde{U}_{k}:=\left\{\begin{array}[]{cl}U_{k},&k<N,\\[5.69054pt] \left(\sigma\circ\frac{\Gamma}{\varepsilon_{k}}\right)\cdot U_{k},&k\geq N,\end{array}\right.\qquad\widetilde{W}_{k}:=\left\{\begin{array}[]{cl}W_{k},&k<N,\\[5.69054pt] \left(\sigma\circ\frac{\Gamma}{\varepsilon_{k}}\right)\cdot W_{k},&k\geq N.\end{array}\right. (3.25)
Proposition 3.9.

For any relatively compact domain OΩO\subset\Omega and any smooth choice of Wd22W_{\frac{d-2}{2}}, there is a sequence {εk}kN(0,1]\{\varepsilon_{k}\}_{k\geq N}\subset(0,1] such that (3.20) yield well-defined distributions for all pO¯p\in\overline{O}, and

  • (i)

    singsupp(~±(p))C(p)\text{sing}\,\text{supp}\,\left(\widetilde{\mathscr{L}}_{\pm}(p)\right)\subset C(p),

  • (ii)

    P~±(p)=δp+K±(p,)P\widetilde{\mathscr{L}}_{\pm}(p)=\delta_{p}+K_{\pm}(p,\cdot) with K±C(O¯×O¯,EE)K_{\pm}\in C^{\infty}(\overline{O}\times\overline{O},E^{*}\boxtimes E),

  • (iii)

    p~±(p)[φ]C(O¯,E)p\mapsto\widetilde{\mathscr{L}}_{\pm}(p)[\varphi]\in C^{\infty}(\overline{O},E^{*}) for all φ𝒟(O,E)\varphi\in\mathscr{D}(O,E^{*}),

  • (iv)

    they are of order at most κd\kappa_{d},

  • (v)

    there is a constant C>0C>0 such that |~±(p)[φ]|CφCκd(O,E)\big{|}\widetilde{\mathscr{L}}_{\pm}(p)[\varphi]\big{|}\leq C\|\varphi\|_{C^{\kappa_{d}}(O,E^{*})} for all pO¯p\in\overline{O} and φ𝒟(O,E)\varphi\in\mathscr{D}(O,E^{*}).

The proofs of Lemma 2.4.1 - 2.4.4 of [BGP2007] only employ smoothness of the Hadamard coefficients and σ(Γ(p,q)εk)=0\sigma\left(\frac{\Gamma(p,q)}{\varepsilon_{k}}\right)=0 if |Γ(p,q)|εk|\Gamma(p,q)|\geq\varepsilon_{k}, so replacing there R±ΩR^{\Omega}_{\pm} by L±ΩL^{\Omega}_{\pm} proves the Proposition in the odd-dimensional case. Similarly, we obtain convergence in CC^{\infty} of the WkW_{k}-part in even dimensions. However, for the logarithmic terms, we have to adapt the corresponding estimates, which is of purely technical nature and therefore removed to the Appendix. Considering ~±,K±\widetilde{\mathscr{L}}_{\pm},K_{\pm} as Schwartz kernels, we extract the corresponding operators

~±:\displaystyle\widetilde{\mathcal{L}}_{\pm}\colon\qquad 𝒟(O,E)C(O¯,E),φ(p~±(p)[φ]),\displaystyle\mathscr{D}(O,E^{*})\longrightarrow C^{\infty}(\overline{O},E^{*}),\qquad\varphi\longmapsto\big{(}p\mapsto\widetilde{\mathscr{L}}_{\pm}(p)[\varphi]\big{)}, (3.26)
𝒦±:\displaystyle\mathcal{K}_{\pm}\colon\qquad C0(O,E)C(O¯,E),u(pO¯K±(p,q)u(q)dV(q)),\displaystyle C^{0}(O,E^{*})\longrightarrow C^{\infty}(\overline{O},E^{*}),\qquad u\longmapsto\left(p\mapsto\int_{\overline{O}}K_{\pm}(p,q)u(q)\,\text{d}V(q)\right), (3.27)

which are bounded due to compactness of O¯\overline{O}.

Let EE be equipped with some non-degenerate inner product and PP be formally self-adjoint, which implies symmetry of the Hadamard coefficients (Theorem A.7). The aim of the rest of the section is to show that then, for all choices involved in Proposition 3.9, the corresponding operators (3.26) represent anti-Feynman and Feynman parametrices for PtP^{t} in the sense of Duistermaat-Hörmander.

Corollary 3.10.

Let ~±\widetilde{\mathcal{L}}_{\pm} and ~±\widetilde{\mathcal{L}}^{\prime}_{\pm} be the operators (3.26) arising from two different choices of N,Wd22,ON,W_{\frac{d-2}{2}},O and {εk}k\{\varepsilon_{k}\}_{k\in\mathbb{N}}. Then ~±~±\widetilde{\mathcal{L}}_{\pm}-\widetilde{\mathcal{L}}^{\prime}_{\pm} is a smoothing operator on OOO\cap O^{\prime}.

Proof.

The Schwartz kernels of these differences are given by the bidistributions

(p,q)(~±(p)~±(p))(q),(p,q)\longmapsto\big{(}\widetilde{\mathscr{L}}_{\pm}(p)-\widetilde{\mathscr{L}}^{\prime}_{\pm}(p)\big{)}(q),

which are smooth by Lemma 2.4.3 of [BGP2007] and Lemma C.2 since supp(σkσk)Γ1(0)=\text{supp}\,(\sigma_{k}-\sigma_{k}^{\prime})\cap\Gamma^{-1}(0)=\emptyset for all kk. ∎

Note that in terms of the operators (3.26), (3.27), Proposition 3.9 (iii) reads ~±Pt=id+𝒦±\widetilde{\mathcal{L}}_{\pm}P^{t}=\text{id}+\mathcal{K}_{\pm} with 𝒦±\mathcal{K}_{\pm} smoothing. Hence, ~±\widetilde{\mathcal{L}}_{\pm} are left parametrices for PtP^{t} and due to formal self-adjointness of PP, they also provide right parametrices:

Proposition 3.11.

For PP formally self-adjoint, the operators ~±\widetilde{\mathcal{L}}_{\pm} define two-sided parametrices for PtP^{t}.

Proof.

We just have to show that ~±\widetilde{\mathcal{L}}_{\pm} yield right parametrices. From the symmetry properties of L±Ω(α)L^{\Omega}_{\pm}(\alpha) and U~k\widetilde{U}_{k} ((3.8) and Theorem A.7) directly follows

OL±Ω(2k+2,p)[(U~k(p,)φ)(ψ(p))]dV(p)=OL±Ω(2k+2,q)[U~k(q,)Θψ(Θq1φ(q))]dV(q)\displaystyle\int_{O}L^{\Omega}_{\pm}(2k+2,p)\big{[}\big{(}\widetilde{U}_{k}(p,\cdot)\varphi\big{)}\big{(}\psi(p)\big{)}\big{]}\,\text{d}V(p)=\int_{O}L^{\Omega}_{\pm}(2k+2,q)\big{[}\widetilde{U}_{k}(q,\cdot)\Theta\psi\big{(}\Theta^{-1}_{q}\varphi(q)\big{)}\big{]}\,\text{d}V(q) (3.28)

for all φ𝒟(O,E),ψ𝒟(O,E)\varphi\in\mathscr{D}(O,E^{*}),\psi\in\mathscr{D}(O,E) and k0k\in\mathbb{N}_{0}. This works analogously for the logarithmic and L~±Ω\widetilde{L}^{\Omega}_{\pm}-terms in (3.20). Furthermore, the series involving the coefficients W~k\widetilde{W}_{k} are given by convergent power series j=0ajΓj\sum_{j=0}^{\infty}a_{j}\Gamma^{j}, which yield smooth sections in EEE^{*}\boxtimes E. Altogether, we obtain the decomposition ~±=𝒰±+𝒲\widetilde{\mathcal{L}}_{\pm}=\mathcal{U}_{\pm}+\mathcal{W} with 𝒰±\mathcal{U}_{\pm} representing the symmetric U~k\widetilde{U}_{k}-part, i.e. 𝒰±t=Θ1𝒰±Θ\mathcal{U}_{\pm}^{t}=\Theta^{-1}\mathcal{U}_{\pm}\Theta, and 𝒲\mathcal{W} the smooth W~k\widetilde{W}_{k}-part. Hence, ~±\widetilde{\mathcal{L}}_{\pm} is symmetric up to smoothing in the sense

~±=Θ~±tΘ1+𝒲Θ𝒲tΘ1smoothing,\displaystyle\widetilde{\mathcal{L}}_{\pm}=\Theta\widetilde{\mathcal{L}}_{\pm}^{t}\Theta^{-1}+\underbrace{\mathcal{W}-\Theta\mathcal{W}^{t}\Theta^{-1}}_{\text{smoothing}}, (3.29)

from which we directly deduce the claim:

Pt~±\displaystyle P^{t}\widetilde{\mathcal{L}}_{\pm} =Pt(Θ~±tΘ1+𝒲Θ𝒲tΘ1)=ΘP~±tΘ1+Pt𝒲tPtΘ𝒲tΘ1\displaystyle=P^{t}\big{(}\Theta\widetilde{\mathcal{L}}_{\pm}^{t}\Theta^{-1}+\mathcal{W}-\Theta\mathcal{W}^{t}\Theta^{-1}\big{)}=\Theta P\widetilde{\mathcal{L}}_{\pm}^{t}\Theta^{-1}+P^{t}\mathcal{W}^{t}-P^{t}\Theta\mathcal{W}^{t}\Theta^{-1}
=Θ(~±Pt)tΘ1+ΘP(Θ1𝒲𝒲tΘ1)=id+Θ𝒦±tΘ1+ΘP(Θ1𝒲𝒲tΘ1)smoothing.\displaystyle=\Theta\big{(}\widetilde{\mathcal{L}}_{\pm}P^{t}\big{)}^{t}\Theta^{-1}+\Theta P\big{(}\Theta^{-1}\mathcal{W}-\mathcal{W}^{t}\Theta^{-1}\big{)}=\text{id}+\underbrace{\Theta\mathcal{K}_{\pm}^{t}\Theta^{-1}+\Theta P\big{(}\Theta^{-1}\mathcal{W}-\mathcal{W}^{t}\Theta^{-1}\big{)}}_{\text{smoothing}}.\qed

If PP is formally self-adjoint, then so is the operator

~:=i2(~+~)\displaystyle\widetilde{\mathcal{L}}:=\frac{i}{2}(\widetilde{\mathcal{L}}_{+}-\widetilde{\mathcal{L}}_{-}) (3.30)

due to symmetry of its Schwartz kernel ~=i2(~+~)\widetilde{\mathscr{L}}=\frac{i}{2}\big{(}\widetilde{\mathscr{L}}_{+}-\widetilde{\mathscr{L}}_{-}\big{)} by (3.28).
So far, we found two-sided parametrices G~±,~±\widetilde{G}_{\pm},\widetilde{\mathcal{L}}_{\pm} given by Hadamard series (A.13), (3.20), and [SV2001] actually proved equivalence of the Hadamard condition (1.6) and that the bidistribution is given by a certain Hadamard series. This latter condition therefore allows us to express (1.8) in terms of G~±,~±\widetilde{G}_{\pm},\widetilde{\mathcal{L}}_{\pm} by directly comparing the corresponding Hadamard series. More precisely, we confirm that i2(~+~+G~+G~)\frac{i}{2}\big{(}\widetilde{\mathscr{L}}_{+}-\widetilde{\mathscr{L}}_{-}+\widetilde{G}_{+}-\widetilde{G}_{-}\big{)} is a Hadamard bidistribution, which, up to smooth errors, moreover is a bisolution with the right antisymmetric part. By examining a further linear combination of parametrices, it will follow that ~±\widetilde{\mathscr{L}}_{\pm} represent a Feynman and an anti-Feynman parametrix.

Proposition 3.12.

Let OΩO\subset\Omega be relatively compact, PP formally self-adjoint and ~±,~±\widetilde{\mathscr{R}}_{\pm},\widetilde{\mathscr{L}}_{\pm} the bidistributions given by (A.13) and (3.20). Then, for

H~:=i2(~+~+~~+),\displaystyle\widetilde{H}:=\frac{i}{2}\big{(}\widetilde{\mathscr{L}}_{+}-\widetilde{\mathscr{L}}_{-}+\widetilde{\mathscr{R}}_{-}-\widetilde{\mathscr{R}}_{+}\big{)}, (3.31)

the sections P(1)tH~,P(2)H~P^{t}_{(1)}\widetilde{H},P_{(2)}\widetilde{H} are smooth, the antisymmetric part of H~\widetilde{H} is given by i2(~~+)\frac{i}{2}\big{(}\widetilde{\mathscr{R}}_{-}-\widetilde{\mathscr{R}}_{+}\big{)} and H~\widetilde{H} has the Hadamard singularity structure (1.6). Furthermore,

~++~~~+C(O×O,EE).\displaystyle\widetilde{\mathscr{L}}_{+}+\widetilde{\mathscr{L}}_{-}-\widetilde{\mathscr{R}}_{-}-\widetilde{\mathscr{R}}_{+}\in C^{\infty}(O\times O,E^{*}\boxtimes E). (3.32)
Proof.

The first two claims are immediate conclusions from ~±,~±\widetilde{\mathscr{R}}_{\pm},\widetilde{\mathcal{L}}_{\pm} being two-sided parametrices and ~\widetilde{\mathscr{L}} being symmetric, so we proceed with the Hadamard singularity structure.
Let k,j0k,j\in\mathbb{N}_{0} such that kjk\geq j, and for even dd, let either j,kd22j,k\leq\frac{d-2}{2} or j,k>d22j,k>\frac{d-2}{2}. Then

R±Ω(2k+2)R±Ω(2j+2)=L±Ω(2k+2)L±Ω(2j+2)=L~±Ω(2k+2)L~±Ω(2j+2)=C(2k+2,d)C(2j+2,d)=:Kk,j,d0Γkj\displaystyle\frac{R^{\Omega}_{\pm}(2k+2)}{R^{\Omega}_{\pm}(2j+2)}=\frac{L^{\Omega}_{\pm}(2k+2)}{L^{\Omega}_{\pm}(2j+2)}=\frac{\widetilde{L}^{\Omega}_{\pm}(2k+2)}{\widetilde{L}^{\Omega}_{\pm}(2j+2)}=\underbrace{\frac{C(2k+2,d)}{C(2j+2,d)}}_{=:K_{k,j,d}\neq 0}\leavevmode\nobreak\ \Gamma^{k-j} (3.33)

due to (3.2) and Proposition 3.3. R±Ω(α)R^{\Omega}_{\pm}(\alpha) denote the Riesz distributions (A.10), which are considered as bidistributions in the canonical way. Define

HΩ(2):=i2(L~+Ω(2)L~Ω(2)+RΩ(2)R+Ω(2)),HΩ(d):=LΩ(d)2π(log(Γ+i0)+log(Γi0))+i2(RΩ(d)R+Ω(d))\displaystyle\begin{split}H^{\Omega}(2)&:=\frac{i}{2}\big{(}\widetilde{L}^{\Omega}_{+}(2)-\widetilde{L}^{\Omega}_{-}(2)+R^{\Omega}_{-}(2)-R^{\Omega}_{+}(2)\big{)},\\[5.69054pt] H^{\Omega}(d)&:=-\frac{L^{\Omega}(d)}{2\pi}\big{(}\log(\Gamma+i0)+\log(\Gamma-i0)\big{)}+\frac{i}{2}\big{(}R^{\Omega}_{-}(d)-R^{\Omega}_{+}(d)\big{)}\end{split} (3.34)

such that the Hadamard series (3.31) takes the form

H~={HΩ(2)k=0Kk,0,dU~kΓk,d odd,HΩ(2)k=0d42Kk,0,dU~kΓk+HΩ(d)k=d22Kk,d22,dU~kΓkd22,d even.\displaystyle\widetilde{H}=\left\{\begin{array}[]{cl}H^{\Omega}(2)\mathop{\sum}\limits_{k=0}^{\infty}K_{k,0,d}\cdot\widetilde{U}_{k}\Gamma^{k},&d\text{ odd},\\[5.69054pt] H^{\Omega}(2)\mathop{\sum}\limits_{k=0}^{\frac{d-4}{2}}K_{k,0,d}\cdot\widetilde{U}_{k}\Gamma^{k}+H^{\Omega}(d)\mathop{\sum}\limits_{k=\frac{d-2}{2}}^{\infty}K_{k,\frac{d-2}{2},d}\cdot\widetilde{U}_{k}\Gamma^{k-\frac{d-2}{2}},&d\text{ even}.\end{array}\right. (3.37)

We show that this is of Hadamard form in the sense of Definition 5.1 in [SV2001], where mostly the notations and conventions of [Gün1988] are adopted. In particular, the Hadamard coefficients U(k)U_{(k)} used in [SV2001] are related with UkU_{k} via 2kk!U(k)=Uk2^{k}k!\cdot U_{(k)}=U_{k} (see Remark 2.3.2 of [BGP2007]). In addition, with the notation (α,k):=2kΓ(α2+k)Γ(α2)(\alpha,k):=2^{k}\cdot\frac{\Gamma\left(\frac{\alpha}{2}+k\right)}{\Gamma\left(\frac{\alpha}{2}\right)}, we find (2j+2,kj)(2j+4d,kj)=Kk,j,d1(2j+2,k-j)\cdot(2j+4-d,k-j)=K_{k,j,d}^{-1} as defined in (A.17), and hence,

Kk,0,dUk\displaystyle K_{k,0,d}\cdot U_{k} =2kk!U(k)2kk!(4d,k)=U(k)(4d,k),\displaystyle=\frac{2^{k}\cdot k!\cdot U_{(k)}}{2^{k}\cdot k!\cdot(4-d,k)}=\frac{U_{(k)}}{(4-d,k)},
Kk,d22,dUk\displaystyle K_{k,\frac{d-2}{2},d}\cdot U_{k} =2kk!Γ(d2)U(k)4kd22k!Γ(k+2d2)=(2,d22)2k+d2(kd22)!U(k).\displaystyle=\frac{2^{k}\cdot k!\cdot\Gamma\left(\frac{d}{2}\right)\cdot U_{(k)}}{4^{k-\frac{d-2}{2}}\cdot k!\cdot\Gamma\left(k+2-\frac{d}{2}\right)}=\frac{\left(2,\frac{d-2}{2}\right)}{2^{k+d-2}\cdot\left(k-\frac{d-2}{2}\right)!}\cdot U_{(k)}.

For all nn\in\mathbb{N}, we choose Nn+d2N\geq n+\left\lceil\frac{d}{2}\right\rceil in (3.25), so the series in (3.37) truncated at k=n+d2k=n+\left\lceil\frac{d}{2}\right\rceil coincide with U,V(n),T(n)U,V^{(n)},T^{(n)} given in Appendix A.1 of [SV2001]. The remainder term is then of regularity CnC^{n} and corresponds to H(n)H^{(n)} in Definition 5.1 of [SV2001].
It remains to identify the singular terms (3.34) with G(1),G(2)G^{(1)},G^{(2)} given by (5.3) in [SV2001] up to some global factor, which is 2-2 in the odd- and 2(1)d22\cdot(-1)^{\frac{d}{2}} in the even-dimensional case. Moreover, note that for the squared Lorentzian distance in the definition of G(1),G(2)G^{(1)},G^{(2)}, the convention s=Γs=-\Gamma is used, whereas in Appendix A.1, it is s=Γs=\Gamma.

Let p,qOp,q\in O. By definition of L~±Ω(α,p)\widetilde{L}^{\Omega}_{\pm}(\alpha,p) and R±Ω(α,p)R^{\Omega}_{\pm}(\alpha,p) as well as (2.35), we have HΩ(2,p)=(expp)WH^{\Omega}(2,p)=\big{(}\exp_{p}\big{)}_{*}W with WW Wightman’s solution for (Minkd,)(\mathbb{R}^{d}_{\mathrm{Mink}},\Box). Then (2.33) leads to

HΩ(2,p)(q)=Γ(d22)4πd2limε0(Γ(p,q)+2iεq0+ε2)2d2H^{\Omega}(2,p)(q)=\frac{\Gamma\left(\frac{d-2}{2}\right)}{4\pi^{\frac{d}{2}}}\mathop{\lim}\limits_{\varepsilon\rightarrow 0}\big{(}-\Gamma(p,q)+2i\varepsilon\cdot q^{0}+\varepsilon^{2}\big{)}^{\frac{2-d}{2}}

in the distributional sense with q0=(expp1(q))0q^{0}=\big{(}\exp_{p}^{-1}(q)\big{)}^{0}. Since Γ(d22)Γ(1d22)=(1)d+12π\Gamma\left(\frac{d-2}{2}\right)\Gamma\left(1-\frac{d-2}{2}\right)=(-1)^{\frac{d+1}{2}}\pi for odd dd, this coincides with G(1)G^{(1)}. Furthermore, one directly calculates HΩ(d)G(2)=0H^{\Omega}(d)-G^{(2)}=0 away from Γ1(0)\Gamma^{-1}(0). For ϕpq(t):=expp(texpp1(q))\phi_{pq}(t):=\exp_{p}\big{(}t\,\exp_{p}^{-1}(q)\big{)} with t[0,1]t\in[0,1], we obtain Γ(p,ϕpq(t))=t2Γ(p,q)\Gamma\big{(}p,\phi_{pq}(t)\big{)}=t^{2}\cdot\Gamma(p,q). Similar to (2.32), we set Γε±(p,):=γε±expp1\Gamma_{\varepsilon}^{\pm}(p,\cdot):=\gamma_{\varepsilon}^{\pm}\circ\exp_{p}^{-1}, which yields Γε±(p,ϕpq(t))=t2Γεt±(p,q)\Gamma_{\varepsilon}^{\pm}\big{(}p,\phi_{pq}(t)\big{)}=t^{2}\cdot\Gamma_{\frac{\varepsilon}{t}}^{\pm}\big{(}p,q\big{)} for all t(0,1]t\in(0,1], and hence,

G(2)(p,ϕpq(t))=LΩ(d)πlimε0log(Γε±(p,ϕpq(t)))=2LΩ(d)πlogt+G(2)(p,q).G^{(2)}\big{(}p,\phi_{pq}(t))=-\frac{L^{\Omega}(d)}{\pi}\mathop{\lim}\limits_{\varepsilon\rightarrow 0}\log\big{(}-\Gamma_{\varepsilon}^{\pm}\big{(}p,\phi_{pq}(t)\big{)}\big{)}=-\frac{2L^{\Omega}(d)}{\pi}\cdot\log t+G^{(2)}(p,q).

On the other hand, since R±Ω(d),LΩ(d)R_{\pm}^{\Omega}(d),L^{\Omega}(d) are homogeneous distributions of degree 0, (3.34) provides

HΩ(d,p)(ϕpq(t))=LΩ(d)2π(2logt+2logt)+HΩ(d,p)(q)=2LΩ(d)πlogt+HΩ(d,p)(q).H^{\Omega}(d,p)\big{(}\phi_{pq}(t)\big{)}=-\frac{L^{\Omega}(d)}{2\pi}\big{(}2\log t+2\log t\big{)}+H^{\Omega}(d,p)(q)=-\frac{2L^{\Omega}(d)}{\pi}\cdot\log t+H^{\Omega}(d,p)(q).

Of course, both expressions have to be understood in the distributional sense. Since Ω\Omega is diffeomorphic to expp1(Ω)TpM\exp_{p}^{-1}(\Omega)\subset T_{p}M for all pΩp\in\Omega, their difference corresponds to a +\mathcal{L}^{\uparrow}_{+}-invariant distributions on Minkowski space d\mathbb{R}^{d}, which is supported on the light cone and homogeneous of degree 0. Therefore, by the classification (2.26), it has to vanish everywhere and thus, Theorem 5.8 of [SV2001] ensures that H~\widetilde{H} is of Hadamard form in the sense of (1.6).
It remains to show (3.32). According to (3.34), we define

AΩ(2):=i2(L~+Ω(2)+L~Ω(2)RΩ(2)R+Ω(2)),AΩ(d):=LΩ(d)2π(log(Γ+i0)log(Γi0))i2(RΩ(d)+R+Ω(d))\displaystyle\begin{split}A^{\Omega}(2)&:=\frac{i}{2}\big{(}\widetilde{L}^{\Omega}_{+}(2)+\widetilde{L}^{\Omega}_{-}(2)-R^{\Omega}_{-}(2)-R^{\Omega}_{+}(2)\big{)},\\[5.69054pt] A^{\Omega}(d)&:=-\frac{L^{\Omega}(d)}{2\pi}\big{(}\log(\Gamma+i0)-\log(\Gamma-i0)\big{)}-\frac{i}{2}\big{(}R^{\Omega}_{-}(d)+R^{\Omega}_{+}(d)\big{)}\end{split} (3.38)

such that for (3.32), we obtain the expression (3.37) with HΩ(2),HΩ(d)H^{\Omega}(2),H^{\Omega}(d) replaced by AΩ(2),AΩ(d)A^{\Omega}(2),A^{\Omega}(d) and it suffices to show smoothness of the bidistributions (3.38). For AΩ(2)A^{\Omega}(2), this follows directly from the definitions of L~±Ω(2),R±Ω(2)\widetilde{L}^{\Omega}_{\pm}(2),R^{\Omega}_{\pm}(2) as pullbacks of S±,R±2S_{\pm},R^{2}_{\pm} along a diffeomorphism and (2.35). On the other hand, one directly calculates that AΩ(d)A^{\Omega}(d) is given by the constant iC(d,d)-iC(d,d). ∎

Since ~±,~±\widetilde{\mathscr{L}}_{\pm},\widetilde{\mathscr{R}}_{\pm} are determined merely up to smooth sections, without loss of generality, we regard (3.32) as the equality

~++~=~~+.\displaystyle\widetilde{\mathscr{L}}_{+}+\widetilde{\mathscr{L}}_{-}=\widetilde{\mathscr{R}}_{-}-\widetilde{\mathscr{R}}_{+}. (3.39)

Now we can deduce from Theorem 5.1 of [Rad1996] and section 6.6 of [DH1972] that the operators ~±\widetilde{\mathcal{L}}_{\pm} represent anti-Feynman and Feynman parametrices for PtP^{t} in the sense of (1.9).

3.4 Local fundamental solutions and Hadamard bisolutions

It remains to construct actual local bisolutions SOS^{O} for PP from the parametrices ~±\widetilde{\mathcal{L}}_{\pm} following the lines of section 2.4 of [BGP2007]. By Proposition 3.9 (ii), we have ~±Pt|𝒟(O,E)=id+𝒦±\widetilde{\mathscr{L}}_{\pm}P^{t}\big{|}_{\mathscr{D}(O,E^{*})}=\text{id}+\mathcal{K}_{\pm}, and hence, fundamental solutions are obtained by inverting the operators id+𝒦±\text{id}+\mathcal{K}_{\pm}. Indeed, if vol(O¯)K±C0(O¯×O¯)<1\text{vol}(\overline{O})\cdot\|K_{\pm}\|_{C^{0}(\overline{O}\times\overline{O})}<1, that is, for OO chosen "small enough", (3.27) provides isomorphisms id+𝒦±:Cl(O¯,E)Cl(O¯,E)\text{id}+\mathcal{K}_{\pm}\colon C^{l}(\overline{O},E^{*})\longrightarrow C^{l}(\overline{O},E^{*}) for all l0l\in\mathbb{N}_{0} with bounded inverses given by the Neumann series

(id+𝒦±)1=j=0(𝒦±)j.\displaystyle(\text{id}+\mathcal{K}_{\pm})^{-1}=\sum_{j=0}^{\infty}(-\mathcal{K}_{\pm})^{j}. (3.40)

This means that all ClC^{l}-norms of the series exist, which follows from compactness of O¯\overline{O} and smoothness of K±K_{\pm}. The full proof coincides with the one of Lemma 2.4.8 of [BGP2007]. In the following, we restrict to such small domains:

Definition 3.13.

We call a relatively compact and causal subdomain OO of Ω\Omega admissible if Proposition 3.9 provides parametrices ~±\widetilde{\mathcal{L}}_{\pm} via (3.26) such that the smooth Schwartz kernel K±K_{\pm} of ~±Ptid\widetilde{\mathcal{L}}_{\pm}P^{t}-\text{id} fulfills

vol(O¯)K±C0(O¯×O¯)<1.\displaystyle\text{vol}(\overline{O})\cdot\|K_{\pm}\|_{C^{0}(\overline{O}\times\overline{O})}<1. (3.41)

More precisely, OO is admissible if there is a choice of {εk}k\{\varepsilon_{k}\}_{k} and Wd22W_{\frac{d-2}{2}} such that (3.41) holds for the corresponding K±K_{\pm}. Lemma 2.4.8 of [BGP2007] shows that for OO admissible, the corresponding operators ~±Pt=id+𝒦±\widetilde{\mathcal{L}}_{\pm}P^{t}=\text{id}+\mathcal{K}_{\pm} are isomorphisms with bounded inverses (3.40).

Proposition 3.14.

For any admissible OO, the operators

𝒮~±O:=(id+𝒦±)1~±:𝒟(O,E)C(O¯,E)\displaystyle\widetilde{\mathcal{S}}^{O}_{\pm}:=(\text{id}+\mathcal{K}_{\pm})^{-1}\widetilde{\mathcal{L}}_{\pm}\colon\qquad\mathscr{D}(O,E^{*})\rightarrow C^{\infty}(\overline{O},E^{*})

fulfill 𝒮~±OPt|𝒟(O,E)=id\widetilde{\mathcal{S}}^{O}_{\pm}P^{t}\big{|}_{\mathscr{D}(O,E^{*})}=\text{id}, and hence, the distributions S~±O(p),pO,\widetilde{S}_{\pm}^{O}(p),\leavevmode\nobreak\ p\in O, given by

S~±O(p)[φ]=((id+𝒦±)1~±φ)(p),φ𝒟(O,E),\displaystyle\widetilde{S}_{\pm}^{O}(p)[\varphi]=\big{(}(\text{id}+\mathcal{K}_{\pm})^{-1}\widetilde{\mathcal{L}}_{\pm}\varphi\big{)}(p),\qquad\varphi\in\mathscr{D}(O,E^{*}), (3.42)

yield fundamental solutions for PP at pp. Furthermore, 𝒬±:=(id+𝒦±)1id\mathcal{Q}_{\pm}:=(\text{id}+\mathcal{K}_{\pm})^{-1}-\text{id} are smoothing operators.

Proof.

The first claim follows from ~±Pt=id+𝒦±\widetilde{\mathcal{L}}_{\pm}P^{t}=\text{id}+\mathcal{K}_{\pm}. Moreover, Proposition 3.9 and Lemma 2.4.10 of [BGP2007], with ~±()[φ]\widetilde{\mathscr{R}}_{\pm}(\cdot)[\varphi] and F±Ω()[φ]F^{\Omega}_{\pm}(\cdot)[\varphi] replaced by ~±φ\widetilde{\mathcal{L}}_{\pm}\varphi and 𝒮~±Oφ\widetilde{\mathcal{S}}_{\pm}^{O}\varphi, show that (3.42) yield fundamental solutions. Finally, (3.40) directly yields 𝒬±=(id+𝒦±)1𝒦±\mathcal{Q}_{\pm}=(\text{id}+\mathcal{K}_{\pm})^{-1}\circ\mathcal{K}_{\pm}, which is smoothing since 𝒦±\mathcal{K}_{\pm} is, and (id+𝒦±)1(\text{id}+\mathcal{K}_{\pm})^{-1} is a continuous map C(M,E)C(M,E)C^{\infty}(M,E^{*})\rightarrow C^{\infty}(M,E^{*}). ∎

From now on, let EE be always equipped with some non-degenerate inner product, PP formally self-adjoint and OO admissible.

Proposition 3.15.

The operators 𝒮~±O~±\widetilde{\mathcal{S}}^{O}_{\pm}-\widetilde{\mathcal{L}}_{\pm} are smoothing.

Proof.

Note that 𝒮~±O~±=𝒬±~±\widetilde{\mathcal{S}}^{O}_{\pm}-\widetilde{\mathcal{L}}_{\pm}=\mathcal{Q}_{\pm}\widetilde{\mathcal{L}}_{\pm}. Since 𝒬±,~±\mathcal{Q}_{\pm},\widetilde{\mathcal{L}}_{\pm} are bounded and 𝒬±\mathcal{Q}_{\pm} has a smooth Schwartz kernel, they extend to bounded maps

𝒬±:𝒟(O,E)C(O¯,E),~±:(O,E)𝒟(O,E).\mathcal{Q}_{\pm}\colon\quad\mathscr{D}(O,E^{*})^{\prime}\rightarrow C^{\infty}(\overline{O},E^{*}),\qquad\qquad\widetilde{\mathcal{L}}_{\pm}\colon\quad\mathcal{E}(O,E^{*})^{\prime}\rightarrow\mathscr{D}(O,E^{*})^{\prime}.

Hence, 𝒬±~±:(O,E)C(O¯,E)\mathcal{Q}_{\pm}\widetilde{\mathcal{L}}_{\pm}\colon\mathcal{E}(O,E^{*})^{\prime}\rightarrow C^{\infty}(\overline{O},E^{*}) are bounded and therefore smoothing. ∎

It follows that 𝒮~±O\widetilde{\mathcal{S}}_{\pm}^{O} yield anti-Feynman and Feynman parametrices for PtP^{t} on OO. Moreover, their Schwartz kernels determine a real-valued bidistribution via

S~O[ψ,φ]:=i4(S~+O[ψ,φ]S~O[ψ,φ]S~+O[ψ,φ]¯+S~O[ψ,φ]¯),ψ𝒟(O,E),φ𝒟(O,E),\displaystyle\widetilde{S}^{O}[\psi,\varphi]:=\frac{i}{4}\left(\widetilde{S}_{+}^{O}[\psi,\varphi]-\widetilde{S}_{-}^{O}[\psi,\varphi]-\overline{\widetilde{S}_{+}^{O}[\psi,\varphi]}+\overline{\widetilde{S}_{-}^{O}[\psi,\varphi]}\right),\qquad\psi\in\mathscr{D}(O,E),\leavevmode\nobreak\ \varphi\in\mathscr{D}(O,E^{*}), (3.43)

which has the right singularity structure and is a solution for PP in the second argument, meaning WF(S~O)=WF(~)\text{WF}\big{(}\widetilde{S}^{O}\leavevmode\nobreak\ \big{)}=\text{WF}\big{(}\widetilde{\mathscr{L}}\leavevmode\nobreak\ \big{)} and P(2)S~O=0P_{(2)}\widetilde{S}^{O}=0.

Proposition 3.16.

Let MM be a globally hyperbolic Lorentzian manifold, π:EM\pi\colon E\rightarrow M a real vector bundle with non-degenerate inner product and P:C(M,E)C(M,E)P\colon C^{\infty}(M,E)\rightarrow C^{\infty}(M,E) a formally self-adjoint wave operator. Furthermore, let OMO\subset M be admissible and ~\widetilde{\mathscr{L}} denote the bidistribution given by Proposition 3.9 and (3.30). Then there is a bisolution SO:𝒟(O,E)×𝒟(O,E)S^{O}\colon\mathscr{D}(O,E)\times\mathscr{D}(O,E^{*})\rightarrow\mathbb{R} for PP with WF(SO)=WF(~)\text{WF}\big{(}S^{O}\big{)}=\text{WF}\big{(}\widetilde{\mathscr{L}}\leavevmode\nobreak\ \big{)}.

Proof.

Since OO is admissible, we obtain fundamental solutions S~±O(p)\widetilde{S}_{\pm}^{O}(p) at each pO¯p\in\overline{O}, and furthermore, (3.43) provides pS~O(p)[φ]C(O¯,E)p\mapsto\widetilde{S}^{O}(p)[\varphi]\in C^{\infty}(\overline{O},E^{*}) for all φ𝒟(O,E)\varphi\in\mathscr{D}(O,E^{*}). Moreover, as a causal subdomain of a globally hyperbolic Lorentzian manifold, OO is globally hyperbolic on its own right (Lemma A.5.8 of [BGP2007]). Hence, for Σ\Sigma a Cauchy hypersurface of OO with unit normal field ν\nu, there is a unique smooth solution of

{Pt(SO()[φ])=0,SO()[φ]|Σ=S~O()[φ]|Σ,ν(SO()[φ])|Σ=νS~O()[φ]|Σ.\displaystyle\left\{\begin{array}[]{cl}P^{t}\big{(}S^{O}(\cdot)[\varphi]\big{)}&=0,\\[5.69054pt] S^{O}(\cdot)[\varphi]\big{|}_{\Sigma}&=\widetilde{S}^{O}(\cdot)[\varphi]\big{|}_{\Sigma},\\[5.69054pt] \nabla_{\nu}\big{(}S^{O}(\cdot)[\varphi]\big{)}\big{|}_{\Sigma}&=\nabla_{\nu}\widetilde{S}^{O}(\cdot)[\varphi]\big{|}_{\Sigma}.\end{array}\right.

By continuous dependence on the Cauchy data, SO(p)S^{O}(p) defines an EpE_{p}^{*}-valued distribution for all pOp\in O. Furthermore, SO()[Ptφ]=0S^{O}(\cdot)[P^{t}\varphi]=0 for all φ\varphi since it satisfies the trivial Cauchy problem.
It remains to check the wave front set, i.e. smoothness of DO:=~SOD^{O}:=\widetilde{\mathscr{L}}-S^{O}. Since SO,S~OS^{O},\widetilde{S}^{O} and ~\widetilde{\mathscr{L}} yield parametrices for PP, the sections given by P(2)DOP_{(2)}D^{O}, P(1)tDOP^{t}_{(1)}D^{O} and ~S~O\widetilde{\mathscr{L}}-\widetilde{S}^{O} are smooth, and hence, DOD^{O} is the solution of a Cauchy problem with smooth Cauchy data, which is smooth by Theorem 2.2. ∎

Altogether, any choice of parametrices ~±\widetilde{\mathscr{L}}_{\pm} in the sense of Proposition 3.9 leads to a bisolution SOS^{O} with singularity structure given by i2(G~aFG~F)\frac{i}{2}(\widetilde{G}_{aF}-\widetilde{G}_{F}) in the sense of (1.9), so we constructed bisolutions with the Hadamard singularity structure on every O×OO\times O:

Theorem 3.17.

Let MM be a globally hyperbolic Lorentzian manifold, OMO\subset M an admissible domain, π:EM\pi\colon E\rightarrow M a real vector bundle with non-degenerate inner product and P:C(M,E)C(M,E)P\colon C^{\infty}(M,E)\rightarrow C^{\infty}(M,E) a formally self-adjoint wave operator. For SOS^{O} the bisolution given by Proposition 3.16 and G±G_{\pm} the advanced and retarded Green operator on OO, the bisolution HO:=SO+i2(G+G)H^{O}:=S^{O}+\frac{i}{2}\big{(}G_{+}-G_{-}\big{)} is of Hadamard form.

Proof.

Due to smoothness of SO~S^{O}-\widetilde{\mathscr{L}} and G±G~±G_{\pm}-\widetilde{G}_{\pm}, we obtain smoothness of HOH~H^{O}-\widetilde{H}, so Proposition 3.12 ensures the Hadamard property of HOH^{O}. ∎

Remark 3.18.

For l0l\in\mathbb{N}_{0} and pΩp\in\Omega, let ±N+l(p)\mathscr{L}_{\pm}^{N+l}(p) denote the formal Hadamard series (3.11) with all terms kN+lk\geq N+l removed and N+l:=i2(+N+lN+l)\mathscr{L}^{N+l}:=\frac{i}{2}\big{(}\mathscr{L}^{N+l}_{+}-\mathscr{L}^{N+l}_{-}\big{)}. Then Lemma 2.4.2 of [BGP2007] with R±ΩR^{\Omega}_{\pm} replaced by L±ΩL^{\Omega}_{\pm} and Lemma C.2 for the logarithmic part show that (p,q)(SO(p)N+l(p))(q)(p,q)\mapsto\big{(}S^{O}(p)-\mathscr{L}^{N+l}(p)\big{)}(q) represents a ClC^{l}-section over O×OO\times O. This and the corresponding result for G~±\widetilde{G}_{\pm} (Proposition 2.5.1 in [BGP2007]) ensure that HOH^{O} is given by a Hadamard series up to terms of arbitrarily high regularity.

4 Global Hadamard two-point-functions

So far, we constructed Hadamard bisolutions on products of certain small patches O×OO\times O, and in this final section, the construction of global bisolutions SS, which locally coincide with those SOS^{O} up to smooth bisolutions and thus inherit their singularity structure, is tackled. Assuming EE to be Riemannian and the validity of Theorem 6.6.2 of [DH1972] for sections in EE, we furthermore prove the existence of a smooth bisolution uu such that S+uS+u is symmetric and positive, i.e. a Hadamard two-point-function.

4.1 Global construction of symmetric bisolutions

For MM globally hyperbolic, we fix a Cauchy hypersurface ΣM\Sigma\subset M and two locally finite covers 𝒪:={Oi}iI\mathcal{O}:=\{O_{i}\}_{i\in I}, 𝒪:={Oi}iI\mathcal{O}^{\prime}:=\big{\{}O_{i}^{\prime}\big{\}}_{i\in I} of it by admissible subsets of MM with O¯iOj\overline{O}_{i}\subset O_{j}^{\prime} if and only if i=ji=j. Without loss of generality, we assume that OiΣO_{i}\cap\Sigma is a Cauchy hypersurface of OiO_{i} for all ii. Then N:=iIOiN:=\bigcup_{i\in I}O_{i} yields a causal normal neighborhood of Σ\Sigma in the sense of Lemma 2.2 of [KW1991]. By paracompactness of MM and the Hopf-Rinow-Theorem, we find an exhaustion {Am}m\{A_{m}\}_{m\in\mathbb{N}} of II by finite subsets such that the relatively compact sets Nm:=iAmOiN_{m}:=\bigcup_{i\in A_{m}}O_{i} exhaust NN and every compact subset of NN is contained in some NmN_{m}. Besides that, causality of OO implies OD(O)=D(OΣ)O\subset D(O)=D(O\cap\Sigma) and therefore,

NmiAmD(OiΣ)D(iAmOiΣ)=D(NmΣ),\displaystyle N_{m}\subset\bigcup_{i\in A_{m}}D(O_{i}\cap\Sigma)\subset D\left(\bigcup_{i\in A_{m}}O_{i}\cap\Sigma\right)=D(N_{m}\cap\Sigma), (4.1)

where DD stands for the Cauchy development of the respective set. It follows that every inextendible causal curve in NmN_{m} meets NmΣN_{m}\cap\Sigma exactly ones, so NmΣN_{m}\cap\Sigma is a Cauchy hypersurface of NmN_{m}, i.e. NmN_{m} is globally hyperbolic. In addition, for all iIi\in I, we choose the corresponding local bisolutions SOi,SOiS^{O_{i}^{\prime}},S^{O_{i}} obtained by Theorem 3.17 such that SOi|Oi×Oi=SOiS^{O_{i}^{\prime}}\big{|}_{O_{i}\times O_{i}}=S^{O_{i}}.

Proposition 4.1.

For each O𝒪O\in\mathcal{O}, there is a bisolution S^O\widehat{S}^{O} on O×MO\times M satisfying S^O|O×O=SO\widehat{S}^{O}\big{|}_{O\times O}=S^{O}.

Proof.

Let O𝒪O^{\prime}\in\mathcal{O}^{\prime} such that O¯O\overline{O}\subset O^{\prime}, and χ𝒟(O)\chi\in\mathscr{D}(O^{\prime}) with χ|O=1\chi\big{|}_{O}=1. Then χSO(p)\chi S^{O^{\prime}}(p) is a well-defined distribution with spacelike compact support on MM for all pOp\in O, since χSO(p)[φ]=SO(p)[χφ],\chi S^{O^{\prime}}(p)[\varphi]=S^{O^{\prime}}(p)[\chi\varphi], φ𝒟(M,E)\varphi\in\mathscr{D}(M,E^{*}). With regard to Theorem 2.4, we define S^O(p)𝒟(M,E,Ep)\widehat{S}^{O}(p)\in\mathscr{D}(M,E,E^{*}_{p})^{\prime} as the unique solution of

{PS^O(p)=0,S^O(p)|Σ=χSO(p)|Σ,νS^O(p)|Σ=ν(χSO(p))|Σ,\displaystyle\left\{\begin{array}[]{cl}P\widehat{S}^{O}(p)&=0,\\[5.69054pt] \widehat{S}^{O}(p)\big{|}_{\Sigma}&=\chi S^{O^{\prime}}(p)\big{|}_{\Sigma},\\[5.69054pt] \nabla_{\nu}\widehat{S}^{O}(p)\big{|}_{\Sigma}&=\nabla_{\nu}\big{(}\chi S^{O^{\prime}}(p)\big{)}\big{|}_{\Sigma},\end{array}\right. (4.5)

which moreover depends smoothly on pp in the sense pS^O(p)[φ]C(O,E)p\mapsto\widehat{S}^{O}(p)[\varphi]\in C^{\infty}(O,E^{*}) for fixed φ𝒟(M,E)\varphi\in\mathscr{D}(M,E^{*}). Furthermore, global hyperbolicity of OO ensures S^O|O×O=SO\widehat{S}^{O}\big{|}_{O\times O}=S^{O} by Theorem 2.2, since the difference solves the trivial Cauchy problem on O×OO\times O.
Let T(p)[φ]:=Pt(S^O()[φ])(p)T(p)[\varphi]:=P^{t}\big{(}\widehat{S}^{O}(\cdot)[\varphi]\big{)}(p) and hence T(p)𝒟(M,E,Ep)T(p)\in\mathscr{D}(M,E,E^{*}_{p})^{\prime} for all pOp\in O. It follows that PT(p)=0PT(p)=0, and T(p)[φ]=0=T(p)[νφ]T(p)[\varphi]=0=T(p)[\nabla_{\nu}\varphi] if suppφO\text{supp}\,\varphi\subset O, which leads to T(p)|Σ=νT(p)|Σ=0T(p)\big{|}_{\Sigma}=\nabla_{\nu}T(p)\big{|}_{\Sigma}=0. Consequently, it satisfies the trivial Cauchy problem, so we have T(p)=0T(p)=0, that is, S^O\widehat{S}^{O} represents a bisolution. ∎

This definition of S^O\widehat{S}^{O} is independent of the choice of χ\chi in an appropriate sense: Let χ~𝒟(O)\widetilde{\chi}\in\mathscr{D}(O^{\prime}) be another cut-off with χ~|O=1\widetilde{\chi}\big{|}_{O}=1 and corresponding bisolution S~O\widetilde{S}^{O}. Then D:=S^OS~OD:=\widehat{S}^{O}-\widetilde{S}^{O} is a bisolution with Cauchy data on (OΣ)×Σ(O\cap\Sigma)\times\Sigma given by (χχ~)SO(\chi-\widetilde{\chi})S^{O^{\prime}}. Recall that singsuppSOΓ1(0)(O×O)\text{sing}\,\text{supp}\,S^{O^{\prime}}\subset\Gamma^{-1}(0)\cap(O^{\prime}\times O^{\prime}), so causality of OO yields singsuppSO(p)|Σ(CM(p)OΣ)O\text{sing}\,\text{supp}\,S^{O^{\prime}}(p)\big{|}_{\Sigma}\subset\big{(}C^{M}(p)\cap O^{\prime}\cap\Sigma\big{)}\subset O for all pOp\in O, and hence, singsuppχSO|OΣ×Σ\text{sing}\,\text{supp}\,\chi S^{O^{\prime}}\big{|}_{O\cap\Sigma\times\Sigma} is contained in O×OO\times O. Since DD satisfies the trivial Cauchy problem on O×OO\times O, i.e. D|O×O=0D\big{|}_{O\times O}=0, it is a smooth bisolution by Theorem 2.2. Therefore, S^O\widehat{S}^{O} and S~O\widetilde{S}^{O} differ merely by some smooth bisolution.

ppOOOO^{\prime}Σ\SigmaD(p)=0D(p)=0CM(p)C^{M}(p)singsuppS^O(p)|Σ\text{sing}\,\text{supp}\,\widehat{S}^{O^{\prime}}(p)\big{|}_{\Sigma}

Next, we prove the existence of a compatible choice of bisolutions {S^Oi}iI\{\widehat{S}^{O_{i}}\}_{i\in I}, meaning that they coincide on the overlaps OiOj,i,jIO_{i}\cap O_{j},\leavevmode\nobreak\ i,j\in I. In this way, these compatible bisolutions assemble to a well-defined object on N×MN\times M. The tools for such a procedure are provided by Čech cohomology theory, for which we give a brief and purposive overview. For an introduction to this subject with details and proofs, we refer to section 5.33 of [War1983].
On N×MN\times M, let 𝒞\mathscr{C}^{\infty} denote the sheaf given by the germs of the smooth sections in EEE^{*}\boxtimes E (see Example 5.2 in [War1983]). For the open cover 𝒪M:={Oi×M}iI\mathcal{O}^{M}:=\{O_{i}\times M\}_{i\in I} of N×MN\times M, the nn-simplices correspond to the non-empty (n+1n+1)-times intersections

Oi0inM:=(Oi0Oin)×M,i0,,inI,O^{M}_{i_{0}\ldots i_{n}}:=\big{(}O_{i_{0}}\cap\ldots\cap O_{i_{n}}\big{)}\times M,\qquad i_{0},\ldots,i_{n}\in I,

with n+1n+1 faces {Oi0i^kinM}k=0,,n\big{\{}O^{M}_{i_{0}\ldots\hat{i}_{k}\ldots i_{n}}\big{\}}_{k=0,\ldots,n} obtained by leaving out one OiO_{i} in the intersection, respectively. An nn-cochain is a map that assigns to each non-empty Oi0inMO^{M}_{i_{0}\ldots i_{n}} a section of 𝒞\mathscr{C}^{\infty} over Oi0inMO^{M}_{i_{0}\ldots i_{n}}, which we identify with the elements of C(Oi0inM,EE)C^{\infty}\big{(}O^{M}_{i_{0}\ldots i_{n}},E^{*}\boxtimes E\big{)}. The space of nn-cochains is denoted by 𝒞n(𝒪M,𝒞)\mathcal{C}^{n}(\mathcal{O}^{M},\mathscr{C}^{\infty}), where 𝒞n:={0}\mathcal{C}^{n}:=\{0\} if n<0n<0, and the coboundary operator is defined by

n:𝒞n(𝒪M,𝒞)𝒞n+1(𝒪M,𝒞),(nfn)(Oi0in+1M):=k=0n+1(1)kfn(Oi0i^kin+1M)|Oi0in+1M.\partial_{n}\colon\quad\mathcal{C}^{n}(\mathcal{O}^{M},\mathscr{C}^{\infty})\longrightarrow\mathcal{C}^{n+1}(\mathcal{O}^{M},\mathscr{C}^{\infty}),\quad(\partial_{n}f_{n})\big{(}O^{M}_{i_{0}\ldots i_{n+1}}\big{)}:=\sum_{k=0}^{n+1}(-1)^{k}\cdot f_{n}\big{(}O^{M}_{i_{0}\ldots\hat{i}_{k}\ldots i_{n+1}}\big{)}\big{|}_{O^{M}_{i_{0}\ldots i_{n+1}}}.

It follows that n+1n=0\partial_{n+1}\circ\partial_{n}=0 for all n0n\in\mathbb{N}_{0} and we set Hn(𝒪M,𝒞):=kernrann1H^{n}(\mathcal{O}^{M},\mathscr{C}^{\infty}):=\frac{\ker\partial_{n}}{\mathrm{ran}\,\partial_{n-1}}. These modules are trivial for all nn\in\mathbb{N} by some well-known construction (e.g. p. 202 in [War1983]), employing that 𝒞\mathscr{C}^{\infty} admits a partition of unity subordinate to the locally finite cover 𝒪M\mathcal{O}^{M}:

Lemma 4.2.

For all nn\in\mathbb{N}, we have

Hn(𝒪M,𝒞)={0}.H^{n}(\mathcal{O}^{M},\mathscr{C}^{\infty})=\{0\}.
Proof.

By choice of 𝒪\mathcal{O}, the cover 𝒪M\mathcal{O}^{M} is locally finite. Let {χi}iI\{\chi_{i}\}_{i\in I} denote a partition of unity subordinate to 𝒪M\mathcal{O}^{M} and fn𝒞n(𝒪M,𝒞)f_{n}\in\mathcal{C}^{n}(\mathcal{O}^{M},\mathscr{C}^{\infty}). Then, for each iIi\in I, the smooth section χifn(OiMOi0in1M)\chi_{i}f_{n}\big{(}O^{M}_{i}\cap O^{M}_{i_{0}\ldots i_{n-1}}\big{)} is supported in OiMOi0in1MO^{M}_{i}\cap O^{M}_{i_{0}\ldots i_{n-1}}, and thus, via extension by zero, we consider it as an element of C(Oi0in1M,EE)C^{\infty}\big{(}O^{M}_{i_{0}\ldots i_{n-1}},E^{*}\boxtimes E\big{)}. In this way, we obtain homomorphisms hn:𝒞n(𝒪M,𝒞)𝒞n1(𝒪M,𝒞)h_{n}\colon\mathcal{C}^{n}(\mathcal{O}^{M},\mathscr{C}^{\infty})\rightarrow\mathcal{C}^{n-1}(\mathcal{O}^{M},\mathscr{C}^{\infty}) via

hn(fn)(Oi0in1M):=iIχifn(OiMOi0in1M)C(Oi0in1M,EE),h_{n}(f_{n})\big{(}O^{M}_{i_{0}\ldots i_{n-1}}\big{)}:=\sum_{i\in I}\chi_{i}f_{n}\big{(}O^{M}_{i}\cap O^{M}_{i_{0}\ldots i_{n-1}}\big{)}\in C^{\infty}\big{(}O^{M}_{i_{0}\ldots i_{n-1}},E^{*}\boxtimes E\big{)},

which satisfy

(hn+1(nfn))(Oi0inM)\displaystyle\big{(}h_{n+1}(\partial_{n}f_{n})\big{)}\big{(}O^{M}_{i_{0}\ldots i_{n}}\big{)} =iIχinf(OiMOi0inM)\displaystyle=\sum_{i\in I}\chi_{i}\cdot\partial_{n}f\big{(}O^{M}_{i}\cap O^{M}_{i_{0}\ldots i_{n}}\big{)}
=iIχifn(Oi0inM)+iIk=0n(1)k+1χifn(OiMOi0i^kinM)|Oi0inM\displaystyle=\sum_{i\in I}\chi_{i}f_{n}\big{(}O^{M}_{i_{0}\ldots i_{n}}\big{)}+\sum_{i\in I}\sum_{k=0}^{n}(-1)^{k+1}\cdot\chi_{i}f_{n}\big{(}O^{M}_{i}\cap O^{M}_{i_{0}\ldots\hat{i}_{k}\ldots i_{n}}\big{)}\big{|}_{O^{M}_{i_{0}\ldots i_{n}}}
=fn(Oi0inM)(n1hn(fn))(Oi0inM).\displaystyle=f_{n}\big{(}O^{M}_{i_{0}\ldots i_{n}}\big{)}-\big{(}\partial_{n-1}h_{n}(f_{n})\big{)}\big{(}O^{M}_{i_{0}\ldots i_{n}}\big{)}.

Hence, fnkernf_{n}\in\ker\partial_{n} implies fn=n1hn(fn)f_{n}=\partial_{n-1}h_{n}(f_{n}), that is, fnrann1f_{n}\in\mathrm{ran}\,\partial_{n-1}. ∎

Lemma 4.3.

For all iIi\in I, there is a bisolution hiC(Oi×M,EE)h_{i}\in C^{\infty}(O_{i}\times M,E^{*}\boxtimes E) such that

(S^Oi+hi)|OijM=(S^Oj+hj)|OijM,i,jI.\displaystyle\big{(}\widehat{S}^{O_{i}}+h_{i}\big{)}\big{|}_{O^{M}_{ij}}=\big{(}\widehat{S}^{O_{j}}+h_{j}\big{)}\big{|}_{O^{M}_{ij}},\qquad i,j\in I.
Proof.

For i,jIi,j\in I, we consider the bisolution hij:=S^Oi|OijMS^Oj|OijMh_{ij}:=\widehat{S}^{O_{i}}\big{|}_{O^{M}_{ij}}-\widehat{S}^{O_{j}}\big{|}_{O^{M}_{ij}}. For all mm\in\mathbb{N}, Proposition 3.9 provides parametrices ~±m\widetilde{\mathscr{L}}_{\pm}^{m} on the relative compact domains NmN_{m} such that for ~m:=i2(~+m~m)\widetilde{\mathscr{L}}^{m}:=\frac{i}{2}\big{(}\widetilde{\mathscr{L}}_{+}^{m}-\widetilde{\mathscr{L}}_{-}^{m}\big{)}, Propositions 3.16 and 4.1 yield

hij|Oij×Nm=S^Oi~mC(S^Oj~mC)C(Oij×Nm,EE).\displaystyle h_{ij}\big{|}_{O_{ij}\times N_{m}}=\underbrace{\widehat{S}^{O_{i}}-\widetilde{\mathscr{L}}^{m}}_{\in C^{\infty}}-\big{(}\underbrace{\widehat{S}^{O_{j}}-\widetilde{\mathscr{L}}^{m}}_{\in C^{\infty}}\big{)}\in C^{\infty}\big{(}O_{ij}\times N_{m},E^{*}\boxtimes E\big{)}. (4.6)

Such ~m\widetilde{\mathscr{L}}^{m} exist for all mm and {Nm}m\{N_{m}\}_{m\in\mathbb{N}} exhausts NN, so we have smoothness on Oij×NO_{ij}\times N. Furthermore, as OijO_{ij} is causal and NN a neighborhood of a Cauchy hypersurface, hijh_{ij} fulfills a Cauchy problem with smooth Cauchy data and hence is smooth on all of OijMO^{M}_{ij} by Theorem 2.2.
Therefore, recalling the identification of sections of 𝒞\mathscr{C}^{\infty} with smooth sections in EEE^{*}\boxtimes E, the map f1:OijMhijf_{1}\colon O^{M}_{ij}\mapsto h_{ij} represents a Čech-1-cochain, which moreover is a cocycle, since

(1f1)(OijkM)\displaystyle(\partial_{1}f_{1})(O^{M}_{ijk}) =hjk|OijkMhik|OijkM+hij|OijkM\displaystyle=h_{jk}\big{|}_{O_{ijk}^{M}}-h_{ik}\big{|}_{O_{ijk}^{M}}+h_{ij}\big{|}_{O_{ijk}^{M}}
=S^Oj|OijkMS^Ok|OijkMS^Oi|OijkM+S^Ok|OijkM+S^Oi|OijkMS^Oj|OijkM\displaystyle=\widehat{S}^{O_{j}}\big{|}_{O_{ijk}^{M}}-\widehat{S}^{O_{k}}\big{|}_{O_{ijk}^{M}}-\widehat{S}^{O_{i}}\big{|}_{O_{ijk}^{M}}+\widehat{S}^{O_{k}}\big{|}_{O_{ijk}^{M}}+\widehat{S}^{O_{i}}\big{|}_{O_{ijk}^{M}}-\widehat{S}^{O_{j}}\big{|}_{O_{ijk}^{M}}
=0\displaystyle=0

for all i,j,kIi,j,k\in I. Thus, Lemma 4.2 ensures the existence of f0:OiMh~iC(OiM,EE)f_{0}\colon O^{M}_{i}\mapsto\widetilde{h}_{i}\in C^{\infty}(O^{M}_{i},E^{*}\boxtimes E) such that 0f0=f1\partial_{0}f_{0}=f_{1}, and hence,

hij=f1(OijM)=0f0(OijM)=f0(OjM)|OijMf0(OiM)|OijM=h~j|OijMh~i|OijM,i,jI.\displaystyle h_{ij}=f_{1}(O^{M}_{ij})=\partial_{0}f_{0}(O^{M}_{ij})=f_{0}(O^{M}_{j})\big{|}_{O^{M}_{ij}}-f_{0}(O^{M}_{i})\big{|}_{O^{M}_{ij}}=\widetilde{h}_{j}\big{|}_{O^{M}_{ij}}-\widetilde{h}_{i}\big{|}_{O^{M}_{ij}},\qquad i,j\in I.

Recall that OiΣO_{i}\cap\Sigma is a Cauchy hypersurface of OiO_{i} for all iIi\in I and thus, each h~i\widetilde{h}_{i} determines a bisolution hiC(OiM,EE)h_{i}\in C^{\infty}\big{(}O^{M}_{i},E^{*}\boxtimes E\big{)} via Theorem 2.2. On the other hand, due to causality of OijO_{ij}, we have a well-posed Cauchy problem on OijMO^{M}_{ij}, and consequently, hj|OijMhi|OijM=hijh_{j}\big{|}_{O^{M}_{ij}}-h_{i}\big{|}_{O^{M}_{ij}}=h_{ij}, since their Cauchy data coincide. This proves the claim:

(S^Oi+hi)|OijM=(S^Oi+hjhij)|OijM=(S^Oj+hj)|OijM.\displaystyle\big{(}\widehat{S}^{O_{i}}+h_{i}\big{)}\big{|}_{O^{M}_{ij}}=\big{(}\widehat{S}^{O_{i}}+h_{j}-h_{ij}\big{)}\big{|}_{O^{M}_{ij}}=\big{(}\widehat{S}^{O_{j}}+h_{j}\big{)}\big{|}_{O^{M}_{ij}}.\qed

For a partition of unity {χi}iI\{\chi_{i}\}_{i\in I} subordinate to 𝒪M\mathcal{O}^{M}, a well-defined bisolution on N×MN\times M is given via

S^N[ψ,φ]:=iI(S^Oi+hi)[χiψ,φ],ψ𝒟(N,E),φ𝒟(M,E).\displaystyle\widehat{S}^{N}[\psi,\varphi]:=\sum_{i\in I}\big{(}\widehat{S}^{O_{i}}+h_{i}\big{)}[\chi_{i}\psi,\varphi],\qquad\psi\in\mathscr{D}(N,E),\leavevmode\nobreak\ \varphi\in\mathscr{D}(M,E^{*}). (4.7)

Since 𝒪M\mathcal{O}^{M} a locally finite cover, for each ψ\psi, only finitely many summands are non-zero. Moreover, due to Lemma (4.3), this definition does not depend on the choice of the partition, and for all ii, we directly read off from (4.7) that

S^N|Oi×MS^OiC(Oi×M,EE).\displaystyle\widehat{S}^{N}\big{|}_{O_{i}\times M}-\widehat{S}^{O_{i}}\in C^{\infty}(O_{i}\times M,E^{*}\boxtimes E). (4.8)

Hence, two different constructions of such a bisolution on N×MN\times M differ only by a smooth bisolution.

Proposition 4.4.

There is a bisolution S:𝒟(M,E)×𝒟(M,E)S\colon\mathscr{D}(M,E)\times\mathscr{D}(M,E^{*})\rightarrow\mathbb{R} such that

S|Oi×OiSOiC(Oi×Oi,EE),iI.\displaystyle S\big{|}_{O_{i}\times O_{i}}-S^{O_{i}}\in C^{\infty}\big{(}O_{i}\times O_{i},E^{*}\boxtimes E\big{)},\qquad i\in I. (4.9)
Proof.

Let S^N\widehat{S}^{N} be the bisolution on N×MN\times M defined by (4.7) and recall that NN is an open neighborhood of Σ\Sigma. For all φ𝒟(M,E)\varphi\in\mathscr{D}(M,E^{*}), we define S()[φ]S(\cdot)[\varphi] as the unique solution of

{Pt(S()[φ])=0,S()[φ]|Σ=S^N()[φ]|Σ,ν(S()[φ])|Σ=νS^N()[φ]|Σ.\left\{\begin{array}[]{cl}P^{t}\big{(}S(\cdot)[\varphi]\big{)}&=0,\\[5.69054pt] S(\cdot)[\varphi]\big{|}_{\Sigma}&=\widehat{S}^{N}(\cdot)[\varphi]\big{|}_{\Sigma},\\[5.69054pt] \nabla_{\nu}\big{(}S(\cdot)[\varphi]\big{)}\big{|}_{\Sigma}&=\nabla_{\nu}\widehat{S}^{N}(\cdot)[\varphi]\big{|}_{\Sigma}.\end{array}\right.

This yields a smooth section, which leads to a bisolution since S^N()[Ptφ]=0\widehat{S}^{N}(\cdot)[P^{t}\varphi]=0, and hence, S()[Ptφ]S(\cdot)[P^{t}\varphi] solves the trivial Cauchy problem. Furthermore, we have S|Oi×Oi=S^N|Oi×OiS\big{|}_{O_{i}\times O_{i}}=\widehat{S}^{N}\big{|}_{O_{i}\times O_{i}}, so (4.9) follows from (4.8) and Proposition 4.1. ∎

Corollary 4.5.

There is a smooth bisolution uC(M×M,EE)u\in C^{\infty}(M\times M,E^{*}\boxtimes E) such that

(Su)[ψ1,Θψ2]=(Su)[ψ2,Θψ1],ψ1,ψ2𝒟(M,E).(S-u)[\psi_{1},\Theta\psi_{2}]=(S-u)[\psi_{2},\Theta\psi_{1}],\qquad\psi_{1},\psi_{2}\in\mathscr{D}(M,E).
Proof.

For (ιS)[ψ1,Θψ2]:=S[ψ2,Θψ1](\iota S)[\psi_{1},\Theta\psi_{2}]:=S[\psi_{2},\Theta\psi_{1}], let u:=12(SιS)u:=\frac{1}{2}(S-\iota S). It follows that Su=ι(Su)S-u=\iota(S-u) and we show that uu is smooth. For all mm\in\mathbb{N}, let ~m\widetilde{\mathscr{L}}^{m} be given as in (4.6), i.e. ~m=ι~m\widetilde{\mathscr{L}}^{m}=\iota\widetilde{\mathscr{L}}^{m} due to symmetry and S~N|Nm×Nm~m\widetilde{S}^{N}\big{|}_{N_{m}\times N_{m}}-\widetilde{\mathscr{L}}^{m} smooth by Proposition 3.15. Therefore, uu is smooth on Nm×NmN_{m}\times N_{m} for all mm:

2u|Nm×Nm=S^NmιS^Nm+~m~m=S^Nm~mι(S^Nm~m)2u\big{|}_{N_{m}\times N_{m}}=\widehat{S}^{N_{m}}-\iota\widehat{S}^{N_{m}}+\widetilde{\mathscr{L}}^{m}-\widetilde{\mathscr{L}}^{m}=\widehat{S}^{N_{m}}-\widetilde{\mathscr{L}}^{m}-\iota\big{(}\widehat{S}^{N_{m}}-\widetilde{\mathscr{L}}^{m}\big{)}

and thus on N×NN\times N. Since uu is a bisolution and NN a neighborhood of Σ\Sigma, Theorem 2.2 ensures smoothness on all of M×MM\times M. ∎

Theorem 4.6.

Let MM be a globally hyperbolic Lorentzian manifold, π:EM\pi\colon E\rightarrow M a real vector bundle with non-degenerate inner product over MM and P:C(M,E)C(M,E)P\colon C^{\infty}(M,E)\rightarrow C^{\infty}(M,E) a formally self-adjoint wave operator. Furthermore, let G±G_{\pm} denote the advanced and retarded Green operator for PtP^{t} and SS the symmetric bisolution given by Proposition 4.4 and Corollary 4.5. Then

H:=S+i2(G+G)\displaystyle H:=S+\frac{i}{2}(G_{+}-G_{-}) (4.10)

is a Hadamard bisolution, and a Feynman and an anti-Feynman Green operator for PtP^{t} is determined by

GF=iS+12(G++G),GaF=iS+12(G++G).\displaystyle G_{F}=iS+\frac{1}{2}(G_{+}+G_{-}),\qquad\qquad G_{aF}=-iS+\frac{1}{2}(G_{+}+G_{-}). (4.11)
Proof.

For each mm\in\mathbb{N}, let ~m\widetilde{\mathscr{L}}^{m} be given as in (4.6). It follows that WF(~m)=WF(G~aFG~F)\text{WF}\big{(}\widetilde{\mathscr{L}}^{m}\big{)}=\text{WF}(\widetilde{G}_{aF}-\widetilde{G}_{F}) in the sense of (1.9) from Proposition 3.15, and moreover, we have

S|Nm×Nm~mC(Nm×Nm,EE)S\big{|}_{N_{m}\times N_{m}}-\widetilde{\mathscr{L}}^{m}\in C^{\infty}(N_{m}\times N_{m},E^{*}\boxtimes E)

by Propositions 3.16 and 4.1 as well as (4.8). This holds for all mm and hence, HH is of Hadamard form in a causal normal neighborhood NN of Σ\Sigma due to Proposition 3.12. Therefore, HH is globally Hadamard by Theorem 5.8 of [SV2001] or, to be more precise, by (i) of the subsequent Remark.
For (4.11) note that the proof in the scalar case, given by Theorem 5.1 of [Rad1996] and section 6.6 of [DH1972], exclusively employs the singularity structure of the parametrices. In our case, this is still determined by scalar distributions L±Ω,R±ΩL^{\Omega}_{\pm},R^{\Omega}_{\pm} and thus stays unaffected when multiplying with smooth Hadamard coefficients. Hence, a Feynman and an anti-Feynman parametrix for PtP^{t} are given by

±iH+G±=±iS+12(G++G),\pm iH+G_{\pm}=\pm iS+\frac{1}{2}(G_{+}+G_{-}),

which are even Green operators since G±G_{\pm} are and SS is a bisolution. ∎

4.2 Positivity

It remains to show that SS can be chosen as a positive bisolution meaning that there is some smooth, symmetric bisolution uu such that (S+u)[φ,Θφ]0(S+u)[\varphi,\Theta\varphi]\geq 0 for all φ𝒟(M,E)\varphi\in\mathscr{D}(M,E). For this, we need a bundle-valued version of Theorem 6.6.2 of [DH1972], so we restrict to a certain class of operators, for which this result holds.
Let MM be a smooth manifold, π:EM\pi\colon E\rightarrow M a real or complex vector bundle over MM with non-degenerate inner product and P:C(M,E)C(M,E)P\colon C^{\infty}(M,E)\rightarrow C^{\infty}(M,E) a properly supported pseudodifferential operator. For the definitions of PP being of real principal type in MM, pseudo-convexity of MM with respect to PP and the bicharacteristic relation CPC_{P} of PP, we adopt Definition 3.1 of [Den1982] as well as Definition 6.3.2 and (6.5.2) of [DH1972], respectively. Assuming those properties for MM and PP, according to Theorem 6.5.3 of [DH1972], there are distinguished parametrices Q~CP\Δ,Q~\widetilde{Q}_{C_{P}\backslash\Delta},\widetilde{Q}_{\emptyset} associated to the respective components of CP\ΔC_{P}\backslash\Delta, where Δ\Delta denotes the diagonal in CharP×CharP\mathrm{Char}\,P\times\mathrm{Char}\,P. For PP a wave operator, they correspond to Feynman and anti-Feynman parametrices, respectively.

Definition 4.7.

Let MM be a smooth manifold, π:EM\pi\colon E\rightarrow M a real or complex vector bundle with non-degenerate inner product and P:C(M,E)C(M,E)P\colon C^{\infty}(M,E)\rightarrow C^{\infty}(M,E) a formally self-adjoint, properly supported pseudodifferential operator of real principal type in MM such that MM is pseudo-convex with respect to PP. Then PP is called of positive propagator type if there exists some fC(M×M,EE)f\in C^{\infty}(M\times M,E^{*}\boxtimes E) such that the bidistribution T:=i2(Q~CP\ΔQ~)+fT:=\frac{i}{2}\big{(}\widetilde{Q}_{C_{P}\backslash\Delta}-\widetilde{Q}_{\emptyset}\big{)}+f satisfies

T[ψ,Θψ]0,ψ𝒟(M,E).T[\psi,\Theta\psi]\geq 0,\qquad\psi\in\mathscr{D}(M,E).

Note that ff is not demanded to be unique and in general, a positive propagator type operator will have many such sections. Theorem 6.6.2 of [DH1972] states that every such PP is of positive propagator type for EE the trivial line bundle M×M\times\mathbb{R}. Note that the proof of this theorem employs positivity of i2(G~aFG~F)\frac{i}{2}(\widetilde{G}_{aF}-\widetilde{G}_{F}) for the directional derivatives Dn:=ixn,n=0,,d1,D_{n}:=-i\frac{\partial}{\partial x_{n}},\leavevmode\nobreak\ n=0,\ldots,d-1, on C(d)C^{\infty}(\mathbb{R}^{d}), and by applying certain operators, allowing one to keep track of the singularity structure of the corresponding parametrices, the general case is reduced to DnD_{n}. Eventually, positivity holds up to smooth functions since there is no way to control this smooth part in terms of the singularity structure. However, in the setting of Definition 1.1 with EE assumed to be Riemannian, we can choose the same ansatz and basically the same procedure. This strongly suggests the assumption that wave operators acting on smooth sections in some general Riemannian vector bundle over a globally hyperbolic Lorentzian manifold are of positive propagator type. On the other hand, by Proposition 5.6 of [SV2001], the Hadamard bisolutions fail to be positive if the inner product on EE is not positive definite. Hence, anticipating the result of this section, wave operators acting on sections in a non-Riemannian vector bundle over a globally hyperbolic Lorentzian manifold are not of positive propagator type.
Assuming PP to be of positive propagator type, we need to show that ff can be actually chosen as a symmetric bisolution. It turns out that the existence of a pair GF,GaFG_{F},G_{aF}, a well-posed Cauchy problem and

Char(P)={(p,ξ)TM\{0}|gp(ξ,ξ)=0},CP={(p,ξ;q,η)(TM×TM)\{0}|(p,ξ)(q,ζ)}.\displaystyle\begin{split}\mathrm{Char}\,(P)&=\big{\{}(p,\xi)\in T^{*}M\backslash\{0\}\leavevmode\nobreak\ \big{|}\leavevmode\nobreak\ g_{p}(\xi^{\sharp},\xi^{\sharp})=0\},\\[5.69054pt] C_{P}&=\big{\{}(p,\xi;q,\eta)\in\big{(}T^{*}M\times T^{*}M\big{)}\backslash\{0\}\leavevmode\nobreak\ \big{|}\leavevmode\nobreak\ (p,\xi)\sim(q,\zeta)\big{\}}.\end{split} (4.12)

is sufficient. For Σ\Sigma some Cauchy hypersurface of MM, the idea is to use ff as initial data on Σ×Σ\Sigma\times\Sigma in order to determine a smooth bisolution uu via Theorem 2.2 and then following the lines of section 3.3 of [GW2015].
Let ι:ΣM\iota\colon\Sigma\hookrightarrow M be the embedding map and ρ:=(ι,ιν)\rho:=\big{(}\iota^{*},\iota^{*}\circ\nabla_{\nu}\big{)} the corresponding pullback to the initial data on Σ\Sigma, i.e.

ρ:C(M,E)C(Σ,EE),u(u|Σ,νu|Σ).\displaystyle\rho\colon\qquad C^{\infty}(M,E)\rightarrow C^{\infty}(\Sigma,E\oplus E),\qquad u\longmapsto\big{(}u\big{|}_{\Sigma},\nabla_{\nu}u\big{|}_{\Sigma}\big{)}. (4.13)

Clearly, ρ\rho is surjective and we have ρ(Csc(M,E))=𝒟(M,EE)\rho\big{(}C^{\infty}_{sc}(M,E)\big{)}=\mathscr{D}(M,E\oplus E). Furthermore, for any differential operator PP with well-posed Cauchy problem, ρ\rho yields a bijection kerPC(Σ,EE)\ker P\rightarrow C^{\infty}(\Sigma,E\oplus E). The transposed map ρt\rho^{t} is related to the pushforward along the embedding, which creates singular directions orthogonal to the embedded (spacelike) hypersurface. More precisely, according to Proposition 10.21 of [DK2010], ιφ\iota_{*}\varphi corresponds to φδΣ\varphi\delta_{\Sigma} for any φC(Σ,E)\varphi\in C^{\infty}(\Sigma,E), and hence, ρt\rho^{t} is a map

ρt:C(Σ,EE)𝒟NΣ(M,E).\displaystyle\rho^{t}\colon\qquad C^{\infty}(\Sigma,E^{*}\oplus E^{*})\longrightarrow\mathscr{D}_{N^{*}\Sigma}(M,E^{*})^{\prime}.

𝒟Γ\mathscr{D}_{\Gamma}^{\prime} denotes the distributions with wave front set contained in the closed cone ΓTM\{0}\Gamma\subset T^{*}M\backslash\{0\}, and we refer to section 8.2 of [Hör1990] for precise definitions and properties of these spaces. Due to Hörmander’s criterion ((8.2.3) of [Hör1990]), we can pull back a distribution along ι\iota if its wave front set does not contain the orthogonal directions mentioned above. Hence, for all closed cones ΓTM\{0}\Gamma\subset T^{*}M\backslash\{0\} with ΓNΣ=\Gamma\cap N^{*}\Sigma=\emptyset, (4.13) extends to a map

ρ:𝒟Γ(M,E)𝒟ιΓ(Σ,EE),u(χu[ρtχ]),\displaystyle\rho\colon\qquad\mathscr{D}_{\Gamma}(M,E^{*})^{\prime}\longrightarrow\mathscr{D}_{\iota^{*}\Gamma}(\Sigma,E^{*}\oplus E^{*})^{\prime},\qquad u\longmapsto(\chi\mapsto u[\rho^{t}\chi]),

where ιΓ:={(σ,dι|σt(ξ))|(ι(σ),ξ)Γ}TΣ\{0}\iota^{*}\Gamma:=\big{\{}\big{(}\sigma,\,\text{d}\iota|_{\sigma}^{t}(\xi)\big{)}\leavevmode\nobreak\ \big{|}\leavevmode\nobreak\ \big{(}\iota(\sigma),\xi\big{)}\in\Gamma\big{\}}\subset T^{*}\Sigma\backslash\{0\} contains the projections of ξΓ\xi\in\Gamma onto TΣT^{*}\Sigma. Let

(χ,ζ)Σ:=Σ(χ0,ζ0+χ1,ζ1)dVΣ,χ,ζ𝒟(Σ,EE),\displaystyle(\chi,\zeta)_{\Sigma}:=\int_{\Sigma}\big{(}\left\langle\chi_{0},\zeta_{0}\right\rangle+\left\langle\chi_{1},\zeta_{1}\right\rangle\big{)}\,\text{d}V_{\Sigma},\qquad\chi,\zeta\in\mathscr{D}(\Sigma,E\oplus E), (4.14)

denote the inner product on 𝒟(Σ,EE)\mathscr{D}(\Sigma,E\oplus E) with dVΣ\,\text{d}V_{\Sigma} the induced volume density and Θ~:=(Θ,Θ)\widetilde{\Theta}:=(\Theta,\Theta) the corresponding isomorphism EEEEE\oplus E\rightarrow E^{*}\oplus E^{*}. For PP Green-hyperbolic, the exact sequence (1.1) provides ranG=kerP|Csc\mathrm{ran}\,G=\ker P\big{|}_{C^{\infty}_{sc}} and thus a further bijection ρG:𝒟(M,E)/kerG𝒟(Σ,EE)\rho G\colon\mathscr{D}(M,E)/\ker G\rightarrow\mathscr{D}(\Sigma,E\oplus E), which transfers GG to a Green operator GΣG_{\Sigma} on the space of initial data 𝒟(Σ,EE)\mathscr{D}(\Sigma,E\oplus E) via

(ρGψ1,GΣρGψ2)Σ:=(ψ1,Gψ2)M,ψ1,ψ2𝒟(M,E).\displaystyle(\rho G\psi_{1},G_{\Sigma}\rho G\psi_{2})_{\Sigma}:=(\psi_{1},G\psi_{2})_{M},\qquad\psi_{1},\psi_{2}\in\mathscr{D}(M,E). (4.15)

It is not hard to deduce the Cauchy evolution operator UΣ:=GρGΣU_{\Sigma}:=-G\rho^{*}G_{\Sigma} mapping initial data (u0,u1)C(Σ,EE)(u_{0},u_{1})\in C^{\infty}(\Sigma,E\oplus E) to the solution uC(M,E)u\in C^{\infty}(M,E) of the corresponding homogeneous Cauchy problem. Moreover, [Dim1980] and, for the vector-valued case, [BS2019] provide the expression u=G(ρ1u0ρ0u1)u=G^{*}\big{(}\rho^{*}_{1}u_{0}-\rho^{*}_{0}u_{1}\big{)}, from which uniqueness and surjectivity of ρG:𝒟(M,E)𝒟(Σ,EE)\rho G\colon\mathscr{D}(M,E)\rightarrow\mathscr{D}(\Sigma,E\oplus E) lead to the particular expression GΣ(u0,u1)=(u1,u0)G_{\Sigma}(u_{0},u_{1})=(-u_{1},u_{0}).

Theorem 4.8.

Let MM be a globally hyperbolic Lorentzian manifold, π:EM\pi\colon E\rightarrow M a Riemannian vector bundle and P:C(M,E)C(M,E)P\colon C^{\infty}(M,E)\rightarrow C^{\infty}(M,E) a linear first- or second-order differential operator, which is of positive propagator type and admits a well-posed Cauchy problem. Assume that the characteristic set and the bicharacteristic relation of PP are given by (4.12) and that Q~CP\Δ,Q~\widetilde{Q}_{C_{P}\backslash\Delta},\widetilde{Q}_{\emptyset} can be chosen as actual Green operators QCP\Δ,QQ_{C_{P}\backslash\Delta},Q_{\emptyset}.
Then there is a real-valued and symmetric bisolution SS such that Si2(QCP\ΔQ)S-\frac{i}{2}\big{(}Q_{C_{P}\backslash\Delta}-Q_{\emptyset}\big{)} is smooth and

S[ψ,Θψ]0,ψ𝒟(M,E).\displaystyle S[\psi,\Theta\psi]\geq 0,\qquad\psi\in\mathscr{D}(M,E).
Proof.

The desired real-valued bisolution is given by

S[ψ,φ]:=i4(QCP\ΔQ)[ψ,φ]+i4(QCP\ΔQ)[ψ,φ]¯,ψ,𝒟(M,E),φ𝒟(M,E),\displaystyle S[\psi,\varphi]:=\frac{i}{4}\big{(}Q_{C_{P}\backslash\Delta}-Q_{\emptyset}\big{)}[\psi,\varphi]+\frac{i}{4}\overline{\big{(}Q_{C_{P}\backslash\Delta}-Q_{\emptyset}\big{)}[\psi,\varphi]},\qquad\psi,\in\mathscr{D}(M,E),\leavevmode\nobreak\ \varphi\in\mathscr{D}(M,E^{*}), (4.16)

and we show the claimed properties. With regard to Corollary 4.5 and without loss of generality, we assume SS to be symmetric, and furthermore, there is some fC(M×M,EE)f\in C^{\infty}(M\times M,E^{*}\boxtimes E) such that (S+f)[ψ,Θψ]0,ψ𝒟(M,E),(S+f)[\psi,\Theta\psi]\geq 0,\leavevmode\nobreak\ \psi\in\mathscr{D}(M,E), by assumption on PP. Because this is also true for f~[ψ1,Θψ2]:=12(f[ψ1,Θψ2]+f[ψ2,Θψ1])\widetilde{f}[\psi_{1},\Theta\psi_{2}]:=\frac{1}{2}\big{(}f[\psi_{1},\Theta\psi_{2}]+f[\psi_{2},\Theta\psi_{1}]\big{)}, we assume symmetry of ff as well, that is, f[ψ1,Θψ2]=f[ψ2,Θψ1]f[\psi_{1},\Theta\psi_{2}]=f[\psi_{2},\Theta\psi_{1}] for all ψ1,ψ2𝒟(M,E)\psi_{1},\psi_{2}\in\mathscr{D}(M,E).
Recall that Green operators map 𝒟(M,E)\mathscr{D}(M,E^{*}) to C(M,E)C^{\infty}(M,E^{*}), so for fixed φC(M,E)\varphi\in C^{\infty}(M,E^{*}), (4.16) provides a smooth section pS(p)[φ]p\mapsto S(p)[\varphi] in EE^{*}. It follows that for each pMp\in M, we obtain a well-defined EpE_{p}^{*}-valued distribution S(p)S(p), which solves PS(p)=0PS(p)=0, and hence, WF(S(p))\text{WF}\big{(}S(p)\big{)} exclusively contains lightlike directions by assumption on PP. By WF(S(p))NΣ=\text{WF}\big{(}S(p)\big{)}\cap N^{*}\Sigma=\emptyset, the restriction of S(p)S(p) to Σ\Sigma yields a well-defined distribution ρ(S(p))\rho\big{(}S(p)\big{)} on 𝒟(Σ,EE)\mathscr{D}(\Sigma,E^{*}\oplus E^{*}) for any Cauchy hypersurface Σ\Sigma. This means that, due to Theorem 8.2.13 of [Hör1990], the operator 𝒮\mathcal{S} associated to (4.16) can be applied to ρtχ𝒟NΣ(M,E)\rho^{t}\chi\in\mathscr{D}_{N^{*}\Sigma}(M,E^{*})^{\prime}, χ𝒟(Σ,EE)\chi\in\mathscr{D}(\Sigma,E^{*}\oplus E^{*}), and for the result, we obtain

WF(𝒮ρtχ){(p,ξ)|(p,ξ;q,0)WF(S)}{(p,ξ)|(q,ζ)WF(ρtχ):(p,ξ;q,ζ)WF(S)}.\text{WF}\big{(}\mathcal{S}\rho^{t}\chi\big{)}\subset\big{\{}(p,\xi)\leavevmode\nobreak\ \big{|}\leavevmode\nobreak\ (p,\xi;q,0)\in\text{WF}(S)\big{\}}\cup\big{\{}(p,\xi)\leavevmode\nobreak\ \big{|}\leavevmode\nobreak\ \exists(q,\zeta)\in\text{WF}(\rho^{t}\chi):\enspace(p,\xi;q,-\zeta)\in\text{WF}(S)\big{\}}.

Since WF(S)CP={(p,ξ)(q,ζ)}\text{WF}(S)\subset C_{P}=\big{\{}(p,\xi)\sim(q,\zeta)\big{\}} and WF(ρtχ)NΣ\text{WF}(\rho^{t}\chi)\subset N^{*}\Sigma, both contributions on the right hand side are empty. Hence, 𝒮ρt\mathcal{S}\rho^{t} represents a map 𝒟(Σ,EE)C(M,E)\mathscr{D}(\Sigma,E^{*}\oplus E^{*})\rightarrow C^{\infty}(M,E^{*}), so it follows that pρ(S(p))[χ]=(𝒮ρtχ)(p)p\mapsto\rho\big{(}S(p)\big{)}[\chi]=\big{(}\mathcal{S}\rho^{t}\chi\big{)}(p) is smooth for fixed χ\chi. With the adjoint operator ρ=Θ1ρtΘ~\rho^{*}=\Theta^{-1}\rho^{t}\widetilde{\Theta}, we eventually obtain a well-defined bidistribution SΣ:𝒟(Σ,EE)×𝒟(Σ,EE)S^{\Sigma}\colon\mathscr{D}(\Sigma,E\oplus E)\times\mathscr{D}(\Sigma,E^{*}\oplus E^{*})\rightarrow\mathbb{R} via

SΣ[λ,χ]:=S[ρλ,ρtχ]=ΣΘ~ρΘ1(S()[ρtχ])(σ)(λ(σ))dVΣ(σ).\displaystyle S^{\Sigma}[\lambda,\chi]:=S\big{[}\rho^{*}\lambda,\rho^{t}\chi\big{]}=\int_{\Sigma}\widetilde{\Theta}\rho\Theta^{-1}\Big{(}S(\cdot)\big{[}\rho^{t}\chi\big{]}\Big{)}(\sigma)\leavevmode\nobreak\ \big{(}\lambda(\sigma)\big{)}\,\text{d}V_{\Sigma}(\sigma). (4.17)

The bisection ff determines smooth and symmetric Cauchy data on Σ×Σ\Sigma\times\Sigma and thus a smooth and symmetric bisolution uu by Theorem 2.2 (the first order analogon works completely similarly). Using the short-hand notation Sf:=S+fS_{f}:=S+f and Su:=S+uS_{u}:=S+u, this yields SuΣ=SfΣS^{\Sigma}_{u}=S^{\Sigma}_{f} for the corresponding bidistributions (4.17), and we show that positivity is preserved under the restriction to Σ\Sigma, i.e. SfΣ[λ,Θ~λ]0S^{\Sigma}_{f}[\lambda,\widetilde{\Theta}\lambda]\geq 0 for all λ𝒟(Σ,EE)\lambda\in\mathscr{D}(\Sigma,E\oplus E). Theorem 8.2.3 of [Hör1990] provides a sequence (ψn)n𝒟(M,E)(\psi_{n})_{n\in\mathbb{N}}\subset\mathscr{D}(M,E) such that ψnρλ\psi_{n}\rightarrow\rho^{*}\lambda in 𝒟NΣ(M,E)\mathscr{D}_{N^{*}\Sigma}(M,E)^{\prime}, and consequently, ΘψnΘρλ=ρtΘ~λ\Theta\psi_{n}\rightarrow\Theta\rho^{*}\lambda=\rho^{t}\widetilde{\Theta}\lambda, so continuity of SfS_{f} ensures

SfΣ[λ,Θ~λ]=Sf[ρλ,ρtΘ~λ]=limnSf[ψn,Θψn]00.\displaystyle S^{\Sigma}_{f}[\lambda,\widetilde{\Theta}\lambda]=S_{f}\big{[}\rho^{*}\lambda,\rho^{t}\widetilde{\Theta}\lambda\big{]}=\mathop{\lim}\limits_{n\rightarrow\infty}\underbrace{S_{f}\big{[}\psi_{n},\Theta\psi_{n}\big{]}}_{\geq 0}\geq 0. (4.18)

The proof of Theorem 3.3.1 and Proposition 3.4.2 of [BGP2007] show that well-posedness of the Cauchy problem implies the existence of a unique advanced and retarded Green operator and hence exactness of the sequence (1.1). Thus, due to kerP=ranG\ker P=\mathrm{ran}\,G, SuS_{u} does not only descend to a well-defined bilinear form on 𝒟(M,E)/kerP\mathscr{D}(M,E)/\ker P by being a bisolution, but also to ranG\mathrm{ran}\,G via

Su[Gψ1,ΘGψ2]:=Su[ψ1,Θψ2],ψ1,ψ2𝒟(M,E).S^{\prime}_{u}[G\psi_{1},\Theta G\psi_{2}]:=S_{u}[\psi_{1},\Theta\psi_{2}],\qquad\psi_{1},\psi_{2}\in\mathscr{D}(M,E).

By following the lines of Proposition 3.9 of [GW2015] and employing G=GρGΣρGG=-G\rho^{*}G_{\Sigma}\rho G, this allows us to trace back the claimed positivity property to (4.18). More precisely, for all ψ1,ψ2𝒟(M,E)\psi_{1},\psi_{2}\in\mathscr{D}(M,E), we have

Su[ψ1,Θψ2]\displaystyle S_{u}[\psi_{1},\Theta\psi_{2}] =Su[Gψ1,ΘGψ2]=Su[GρGΣρGψ1,ΘGρGΣρGψ2]\displaystyle=S^{\prime}_{u}[G\psi_{1},\Theta G\psi_{2}]=S^{\prime}_{u}[G\rho^{*}G_{\Sigma}\rho G\psi_{1},\Theta G\rho^{*}G_{\Sigma}\rho G\psi_{2}]
=Su[ρGΣρGψ1,ΘρGΣρGψ2]=SuΣ[GΣρGψ1,Θ~GΣρGψ2]=SfΣ[GΣρGψ1,Θ~GΣρGψ2],\displaystyle=S_{u}[\rho^{*}G_{\Sigma}\rho G\psi_{1},\Theta\rho^{*}G_{\Sigma}\rho G\psi_{2}]=S^{\Sigma}_{u}[G_{\Sigma}\rho G\psi_{1},\widetilde{\Theta}G_{\Sigma}\rho G\psi_{2}]=S^{\Sigma}_{f}[G_{\Sigma}\rho G\psi_{1},\widetilde{\Theta}G_{\Sigma}\rho G\psi_{2}],

which proves the theorem. ∎

In the case of formally self-adjoint wave operators, the existence of GFG_{F} and GaFG_{aF} is ensured by Theorem 4.6, so Theorem 4.8 leads to the final result:

Theorem 4.9.

Let MM be a globally hyperbolic Lorentzian manifold, π:EM\pi\colon E\rightarrow M a Riemannian vector bundle and P:C(M,E)C(M,E)P\colon C^{\infty}(M,E)\rightarrow C^{\infty}(M,E) a formally self-adjoint wave operator of positive propagator type. Then there exists a bidistribution S:𝒟(M,E)×𝒟(M,E)S\colon\mathscr{D}(M,E)\times\mathscr{D}(M,E^{*})\rightarrow\mathbb{R} such that

H:=S+i2(G+G)\displaystyle H:=S+\frac{i}{2}(G_{+}-G_{-})

yields a Hadamard two-point-function, where G±G_{\pm} denotes the advanced and retarded Green operator for PtP^{t}. This means that WF(H)\text{WF}(H) has the Hadamard singularity structure (1.6) and satisfies

H[Pψ,φ]=0=H[ψ,Ptφ],H[ψ,φ]H[Θ1φ,Θψ]=i2(G+G)[ψ,φ],H[ψ,Θψ]0\displaystyle H[P\psi,\varphi]=0=H[\psi,P^{t}\varphi],\qquad H[\psi,\varphi]-H[\Theta^{-1}\varphi,\Theta\psi]=\frac{i}{2}(G_{+}-G_{-})[\psi,\varphi],\qquad H[\psi,\Theta\psi]\geq 0

for all ψ𝒟(M,E),φ𝒟(M,E)\psi\in\mathscr{D}(M,E),\leavevmode\nobreak\ \varphi\in\mathscr{D}(M,E^{*}).
Moreover, a Feynman and an anti-Feynman Green operator GF,GaFG_{F},G_{aF} are given by (4.11).

Note that, in general, SS is far from being unique, i.e. there may be many bidistributions with the required properties. Clearly, this is related to the non-uniqueness of the many choices of smooth sections during the construction, and in most cases, it is not at all obvious, how to find these sections practically. This particularly concerns the choice of the hih_{i}’s in Lemma 4.3 and the ff for operators of positive propagator type.
However, the overall reasoning provides a comparatively constructive alternative to the existence proofs, which are already present in the literature ([BF2014], [FNW1981], [GOW2017]). It starts most naturally with the Hadamard condition, so the form of the bidistributions is, up to smooth terms, determined right from the start. It therefore might provide a promising starting point for a possible classification of these states up to unitary equivalence of their respective GNS-representations. This and the identification of pure states in particular would require to investigate the choices of the said smooth sections.
Furthermore, the methods used here provide an alternative procedure to the classic deformation arguments since they rely on the ability to make modifications to the metric confined to certain spacetime regions. There are situations, where this is not applicable, for instance, in the case of linearized gravity, where the background spacetime must solve the Einstein equation, or similarly for linearizations of Yang-Mills theories. They also occur if one is restricted to analytic metrics.

Appendix

A Symmetry of the Hadamard coefficients

In this section, we prove the symmetry of the Hadamard coefficients UkU_{k} in a setting (M,E,P)(M,E,P) as in Definition 1.1 with PP a wave operator. This represents an alternative path to the more general case treated in [Kam2019], which is more geared to the specific situation of wave operators, and generalizes the well-known approach of Moretti [Mor1999, Mor2000] it to sections in EE.
Let ΩM\Omega\subset M be a non-empty and convex domain, that is time-orientable, and let \nabla denote the PP-compatible connection on EE, meaning that

gradfs=12(fPsP(fs)+fs),sC(M,E),fC(M).\displaystyle\nabla_{\text{grad}f}s=\frac{1}{2}\big{(}f\cdot Ps-P(f\cdot s)+\Box f\cdot s\big{)},\qquad s\in C^{\infty}(M,E),\leavevmode\nobreak\ f\in C^{\infty}(M). (A.1)

It follows that P=+BP=\Box^{\nabla}+B for some uniquely determined endomorphism field BB and =(tridE)TME\Box^{\nabla}=\left(\mathrm{tr}\otimes\text{id}_{E}\right)\circ\nabla^{T^{*}M\otimes E}\circ\nabla the connection-d’Alembert operator (see section 1.5 of [BGP2007]). Then the Hadamard coefficients UkC(Ω×Ω,EE),k0,U_{k}\in C^{\infty}(\Omega\times\Omega,E^{*}\boxtimes E),\leavevmode\nobreak\ k\in\mathbb{N}_{0}, for PP are defined as the unique solutions of the transport equations

gradΓpUpk(12Γpd+2k)Upk=2kPUpk1,k,\displaystyle\nabla_{\text{grad}\Gamma_{p}}U_{p}^{k}-\left(\frac{1}{2}\Box\Gamma_{p}-d+2k\right)U_{p}^{k}=2k\cdot PU_{p}^{k-1},\qquad k\in\mathbb{N}, (A.2)

with U0(p,p)=idEpU_{0}(p,p)=\text{id}_{E_{p}^{*}} and Upk:=Uk(p,)U^{k}_{p}:=U_{k}(p,\cdot) for all pΩp\in\Omega (Proposition 2.3.1 of [BGP2007]). Recall the identification (2.1) of Uk(p,q)U_{k}(p,q) as homomorphisms EqEpE_{q}^{*}\rightarrow E_{p}^{*} with fiberwise transposed operator Uk(p,q)tHom(Ep,Eq)U_{k}(p,q)^{t}\in\textup{Hom}(E_{p},E_{q}). We are going to show symmetry in the sense

Uk(p,q)=ΘpUk(q,p)tΘq1,p,qΩ,k0.\displaystyle U_{k}(p,q)=\Theta_{p}U_{k}(q,p)^{t}\Theta_{q}^{-1},\qquad p,q\in\Omega,\leavevmode\nobreak\ k\in\mathbb{N}_{0}. (A.3)

One can see the Hadamard coefficients as a measure of the deviation of (M,g,E,P)(M,g,E,P) from (Minkd,)(\mathbb{R}^{d}_{\mathrm{Mink}},\Box), and indeed, "adding" (,ds2,{0},2s2)\big{(}\mathbb{R},\,\text{d}s^{2},\{0\},\frac{\partial^{2}}{\partial s^{2}}\big{)} does not change them: Let (M^,g^):=(M×,g+ds2)(\widehat{M},\widehat{g}):=(M\times\mathbb{R},g+\,\text{d}s^{2}\big{)}, over which we consider the same vector bundle EE, and P^:=P2s2\widehat{P}:=P-\frac{\partial^{2}}{\partial s^{2}}. For the corresponding Hadamard coefficients U^k\widehat{U}_{k}, we obtain the initial condition U^0((p,s),(p,s))=idEp\widehat{U}_{0}\big{(}(p,s),(p,s)\big{)}=\text{id}_{E_{p}^{*}} and the transport equations

0=^gradg^Γ^(p,s)U^(p,s)k(12g^Γ^(p,s)(d+1)+2k)U^(p,s)k2kP^U^(p,s)k1\displaystyle 0=\widehat{\nabla}_{\text{grad}_{\widehat{g}}\widehat{\Gamma}_{(p,s)}}\widehat{U}_{(p,s)}^{k}-\left(\frac{1}{2}\Box_{\widehat{g}}\widehat{\Gamma}_{(p,s)}-(d+1)+2k\right)\widehat{U}_{(p,s)}^{k}-2k\widehat{P}\widehat{U}_{(p,s)}^{k-1}
=gradgΓpU^(p,s)k+2(ss)^sU^(p,s)k(12(gΓp+2)(d+1)+2k)U^(p,s)k2kPU^(p,s)k1+2k2s2U^(p,s)k1\displaystyle=\nabla_{\text{grad}_{g}\Gamma_{p}}\widehat{U}_{(p,s)}^{k}+2(s^{\prime}-s)\widehat{\nabla}_{\frac{\partial}{\partial s}}\widehat{U}_{(p,s)}^{k}-\left(\frac{1}{2}\big{(}\Box_{g}\Gamma_{p}+2\big{)}-(d+1)+2k\right)\widehat{U}_{(p,s)}^{k}-2kP\widehat{U}_{(p,s)}^{k-1}+2k\frac{\partial^{2}}{\partial s^{2}}\widehat{U}_{(p,s)}^{k-1}
=gradgΓpU^(p,s)k(12gΓpd+2k)U^(p,s)k2kPU^(p,s)k1+2(ss)^sU^(p,s)k+2k2s2U^(p,s)k1,\displaystyle=\nabla_{\text{grad}_{g}\Gamma_{p}}\widehat{U}_{(p,s)}^{k}-\left(\frac{1}{2}\Box_{g}\Gamma_{p}-d+2k\right)\widehat{U}_{(p,s)}^{k}-2kP\widehat{U}_{(p,s)}^{k-1}+2(s^{\prime}-s)\widehat{\nabla}_{\frac{\partial}{\partial s}}\widehat{U}_{(p,s)}^{k}+2k\frac{\partial^{2}}{\partial s^{2}}\widehat{U}_{(p,s)}^{k-1},

which are clearly solved by UpkU_{p}^{k} leading to

U^k((p,s),(q,s))=Uk(p,q),p,qΩ,s,s,k0.\displaystyle\widehat{U}_{k}\big{(}(p,s),(q,s^{\prime})\big{)}=U_{k}(p,q),\qquad p,q\in\Omega,\leavevmode\nobreak\ s,s^{\prime}\in\mathbb{R},\leavevmode\nobreak\ k\in\mathbb{N}_{0}. (A.4)

We adopt Moretti’s approach insofar as we start by considering wave operators with analytical coefficients, so in particular the metric as the principal symbol is assumed to be analytic, and deduce analyticity of (p,q)Uk(p,q)(p,q)\mapsto U_{k}(p,q). We directly conclude that, as a function of gg, the Levi-Civita connection on TMTM and, due to (A.1), the PP-compatible connection \nabla on EE are analytic as well, so the corresponding Christoffel symbols are. Applying basic ODE-theory and the analytic inverse function theorem (Theorem 1.4.3 of [KP1992]) ensure analyticity of the Lorentzian distance (p,q)Γ(p,q)=gp(expp1(q))(p,q)\mapsto\Gamma(p,q)=g_{p}\big{(}\exp_{p}^{-1}(q)\big{)}, the distortion function (p,q)μ(p,q)(p,q)\mapsto\mu(p,q) and the geodesic considered as a map

ϕ:[0,1]×Ω×ΩΩ,(t,p,q)expp(texpp1(q))=:ϕpq(t).\displaystyle\phi:\qquad[0,1]\times\Omega\times\Omega\longrightarrow\Omega,\qquad(t,p,q)\longmapsto\exp_{p}\big{(}t\,\exp_{p}^{-1}(q)\big{)}=:\phi_{pq}(t). (A.5)
Lemma A.1.

The \nabla-parallel transport along ϕpq\phi_{pq} is analytic as a map

[0,1]×Ω×ΩEE,(t,p,q)Πϕpq(t)p.\displaystyle[0,1]\times\Omega\times\Omega\longrightarrow E^{*}\boxtimes E,\qquad(t,p,q)\longmapsto\Pi_{\phi_{pq}(t)}^{p}. (A.6)
Proof.

For fixed pΩp\in\Omega and eEpe\in E_{p}, consider the parallel section sp(t,q):=Πϕpq(t)pes_{p}(t,q):=\Pi^{p}_{\phi_{pq}(t)}e in EE along ϕpq\phi_{pq} for all qΩq\in\Omega, which therefore satisfies the ODE’s

s˙pβ(t,q)=Γiαβ(ϕpq(t))ϕ˙pqi(t)=:Ap(t,q)αβspα(t,q),sp(0,q)=e.\displaystyle\dot{s}_{p}^{\beta}(t,q)=\underbrace{-\Gamma^{\beta}_{i\alpha}\big{(}\phi_{pq}(t)\big{)}\,\dot{\phi}^{i}_{pq}(t)}_{=:A_{p}(t,q)^{\beta}_{\alpha}}\,s_{p}^{\alpha}(t,q),\qquad s_{p}(0,q)=e. (A.7)

The columns of the corresponding fundamental matrix Φp(t,q)\Phi_{p}(t,q) are given by rk(E)\mathrm{rk}(E) linearly independent solutions of (A.7), so we have Φ˙p(t,q)=Ap(t,q)Φp(t,q)\dot{\Phi}_{p}(t,q)=A_{p}(t,q)\Phi_{p}(t,q) and the solution of (A.7) takes the form

sp(t,q)=Φp(t,q)Φp(0,q)1sp(0,q).s_{p}(t,q)=\Phi_{p}(t,q)\Phi_{p}(0,q)^{-1}s_{p}(0,q).

From the definition of sps_{p}, we read off Πϕpq(t)p=Φp(t,q)Φp(0,q)1\Pi^{p}_{\phi_{pq}(t)}=\Phi_{p}(t,q)\Phi_{p}(0,q)^{-1}, and hence, the map (t,q)Πϕpq(t)p(t,q)\longmapsto\Pi^{p}_{\phi_{pq}(t)} is analytic. Moreover, Πpr=(Πrp)1\Pi^{r}_{p}=\big{(}\Pi_{r}^{p}\big{)}^{-1} implies

Φp(1,r)Φp(0,r)1=(Φr(1,p)Φr(0,p)1)1=Φr(0,p)Φr(1,p)1,\Phi_{p}(1,r)\Phi_{p}(0,r)^{-1}=\left(\Phi_{r}(1,p)\Phi_{r}(0,p)^{-1}\right)^{-1}=\Phi_{r}(0,p)\Phi_{r}(1,p)^{-1},

so pΠrpp\mapsto\Pi^{p}_{r} is analytic for fixed rΩr\in\Omega and Osgood’s Lemma [Osg1898] proves the claim. ∎

Proposition A.2.

The map (p,q)Uk(p,q)(p,q)\mapsto U_{k}(p,q) is analytic on Ω×Ω\Omega\times\Omega for all k0k\in\mathbb{N}_{0}.

Proof.

Analyticity of the zeroth Hadamard coefficient can be directly read off from

(p,q)U0(p,q)=Πqpμ(p,q),(p,q)\longmapsto U_{0}(p,q)=\frac{\Pi^{p}_{q}}{\sqrt{\mu(p,q)}},

and we proceed via induction. By analyticity of PP, clearly (p,q)P(2)Uk1(p,Φpq(t))(p,q)\mapsto P_{(2)}U_{k-1}\big{(}p,\Phi_{pq}(t)\big{)} is analytic if the prior coefficient (p,q)Uk1(p,q)(p,q)\mapsto U_{k-1}(p,q) is. Similarly, (t,p,q)U0(p,ϕpq(t))1=μ(p,ϕpq(t))Πpϕpq(t)(t,p,q)\mapsto U_{0}\big{(}p,\phi_{pq}(t)\big{)}^{-1}=\sqrt{\mu\big{(}p,\phi_{pq}(t)\big{)}}\cdot\Pi_{p}^{\phi_{pq}(t)} is analytic as a composition of analytic maps (recall that μ\mu is positive). Therefore, the integrand of

Uk(p,q)=kU0(p,q)01tk1U0(p,ϕpq(t))1P(2)Uk1(p,ϕpq(t))dtU_{k}(p,q)=-kU_{0}(p,q)\int_{0}^{1}t^{k-1}\cdot U_{0}\big{(}p,\phi_{pq}(t)\big{)}^{-1}P_{(2)}U_{k-1}\big{(}p,\phi_{pq}(t)\big{)}\,\text{d}t

is analytic in (p,q)(p,q) and uniformly continuous in tt on [0,1][0,1]. Hence, taking the power series expression of the integrand, the sum and the integral can be swapped, which results in a uniformly converging power series for UkU_{k}. ∎

Now the general case of smooth PP is tackled by analytic approximation of the coefficients, for which we quote

Proposition A.3 (Proposition 2.1 of [Mor1999]).

Let MM be a real, smooth and connected manifold with non-singular metric gg.

  • (a)

    For any local chart (x,V)(x,V) of MM and any relatively compact domain OO with O¯V\overline{O}\subset V, there is a sequence {gn}n\{g^{n}\}_{n\in\mathbb{N}} of real and analytic (with respect to xx) metrics with the same signature as gg, which are defined on some neighborhood of O¯\overline{O} such that gngg^{n}\rightarrow g in CC^{\infty}, that is, all derivatives of gng^{n} converge uniformly on O¯\overline{O}:

    i,j=1,,D,α0D:maxvx(O¯)|(Dα(gnx1)ij)(v)(Dα(gx1)ij)(v)|0.\forall i,j=1,\ldots,D,\quad\alpha\in\mathbb{N}_{0}^{D}:\qquad\max_{v\in x(\overline{O})}\Big{|}\big{(}D^{\alpha}(g^{n}\circ x^{-1})_{ij}\big{)}(v)-\big{(}D^{\alpha}(g\circ x^{-1})_{ij}\big{)}(v)\Big{|}\longrightarrow 0.
  • (b)

    For any (x,V),O,{gn}n(x,V),O,\{g^{n}\}_{n\in\mathbb{N}} as in (a) and additionally any zOz\in O, there is an n0n_{0}\in\mathbb{N} and a family {Nzi}i\{N_{z}^{i}\}_{i\in\mathbb{R}} of open neighborhoods of zz such that NziN¯zjON_{z}^{i}\subset\overline{N}_{z}^{j}\subset O for any j>ij>i, and {Nzi}i\{N_{z}^{i}\}_{i\in\mathbb{R}} is a local base of the topology of MM. Moreover, for all ii\in\mathbb{R}, both NziN_{z}^{i} and N¯zi\overline{N}_{z}^{i} are common convex neighborhoods of zz for all metrics {gn}n>n0\{g^{n}\}_{n>n_{0}} and gg.

Proposition A.4.

Let OΩO\subset\Omega be relatively compact and {gn}n\{g^{n}\}_{n\in\mathbb{N}} a sequence of real and analytic metrics defined in a neighborhood of O¯\overline{O} with the same signature as gg such that OO and O¯\overline{O} are convex with respect to all gn,n,g^{n},\leavevmode\nobreak\ n\in\mathbb{N}, and gg and gngg^{n}\rightarrow g in CC^{\infty}. For {Ukn}n\{U_{k}^{n}\}_{n\in\mathbb{N}} the corresponding Hadamard coefficients, we obtain Ukn(p,q)Uk(p,q)U_{k}^{n}(p,q)\rightarrow U_{k}(p,q) for all k0k\in\mathbb{N}_{0} and p,qOp,q\in O.

Proof.

The assumption directly provides Γijk,nΓijk\Gamma^{k,n}_{ij}\rightarrow\Gamma^{k}_{ij}, and with regard to the geodesic equation with converging right hand side, we similarly obtain expnexp\exp^{n}\rightarrow\exp as smooth maps (t,ξ,p)expp(tξ)(t,\xi,p)\mapsto\exp_{p}(t\xi) on their domain of existence. Then the inverse function theorem provides (expn)1exp1\big{(}\exp^{n}\big{)}^{-1}\rightarrow\exp^{-1} as smooth maps on O¯×O¯\overline{O}\times\overline{O} and, as a consequence, of the Lorentzian distance ΓnΓ\Gamma^{n}\rightarrow\Gamma and the distortion function μnμ\mu^{n}\rightarrow\mu. Eventually, we have ϕnϕ\phi^{n}\rightarrow\phi for the connecting geodesic (A.5).
It remains to investigate the parallel transport. For all pOp\in O, convergence of Γijk,n\Gamma^{k,n}_{ij} and ϕn\phi^{n} leads to ApnApA_{p}^{n}\rightarrow A_{p} for the matrices defined in (A.7), and hence,

Πϕpqn(t)p,n=Φpn(t,q)Φpn(0,q)1Φp(t,q)Φp(0,q)1=Πϕpq(t)p\Pi^{p,n}_{\phi^{n}_{pq}(t)}=\Phi^{n}_{p}(t,q)\Phi^{n}_{p}(0,q)^{-1}\longrightarrow\Phi_{p}(t,q)\Phi_{p}(0,q)^{-1}=\Pi^{p}_{\phi_{pq}(t)}

as smooth maps [0,1]×O¯EpE[0,1]\times\overline{O}\rightarrow E_{p}^{*}\otimes E. Thus, we can directly conclude convergence of the zeroth Hadamard coefficient

U0n(p,)=Πp,nμpnΠpμp=U0(p,)\displaystyle U_{0}^{n}(p,\cdot)=\frac{\Pi^{p,n}_{\cdot}}{\sqrt{\mu^{n}_{p}}}\longrightarrow\frac{\Pi^{p}_{\cdot}}{\sqrt{\mu_{p}}}=U_{0}(p,\cdot) (A.8)

as smooth maps O¯EpE\overline{O}\rightarrow E^{*}_{p}\otimes E and, in particular, U0n(p,q)U0(p,q)U_{0}^{n}(p,q)\rightarrow U_{0}(p,q) in Hom(Eq,Ep)\textup{Hom}(E_{q}^{*},E_{p}^{*}).
We proceed inductively. Due to ϕpqnϕpq\phi_{pq}^{n}\rightarrow\phi_{pq} in C([0,1],O¯)C^{\infty}\big{(}[0,1],\overline{O}\big{)}, (A.8) implies U0n(p,)ϕpqnU0(p,)ϕpqU_{0}^{n}(p,\cdot)\circ\phi^{n}_{pq}\rightarrow U_{0}(p,\cdot)\circ\phi_{pq} in C([0,1],EpE)C^{\infty}\big{(}[0,1],E^{*}_{p}\otimes E\big{)} and consequently, PU0n(p,)ϕpqnPU0(p,)ϕpqPU_{0}^{n}(p,\cdot)\circ\phi^{n}_{pq}\rightarrow PU_{0}(p,\cdot)\circ\phi_{pq}. Therefore, the integrand in the expression of the first Hadamard coefficient

U1n(p,q)=kU0n(p,q)01U0n(p,ϕpqn(t))1P(2)U0n(p,ϕpqn(t))dtU_{1}^{n}(p,q)=-kU^{n}_{0}(p,q)\int_{0}^{1}U_{0}^{n}\big{(}p,\phi_{pq}^{n}(t)\big{)}^{-1}P_{(2)}U_{0}^{n}\big{(}p,\phi_{pq}^{n}(t)\big{)}\,\text{d}t

converges to the one in the expression of U1(p,q)U_{1}(p,q), and as a smooth function in tt, it is integrable on the compact interval [0,1][0,1]. Hence, due to majorized convergence, the integral converges as well, so we have

U0n(p,q)01U0n(p,ϕpqn(t))1P(2)U0n(p,ϕpqn(t))dtU0(p,q)01U0(p,ϕpq(t))1P(2)U0(p,ϕpq(t))dt,-U^{n}_{0}(p,q)\int_{0}^{1}U_{0}^{n}\big{(}p,\phi_{pq}^{n}(t)\big{)}^{-1}P_{(2)}U_{0}^{n}\big{(}p,\phi_{pq}^{n}(t)\big{)}\,\text{d}t\longrightarrow-U_{0}(p,q)\int_{0}^{1}U_{0}\big{(}p,\phi_{pq}(t)\big{)}^{-1}P_{(2)}U_{0}\big{(}p,\phi_{pq}(t)\big{)}\,\text{d}t,

which is U1(p,q)U_{1}(p,q). Recursively, this implies Ukn(p,q)Uk(p,q)U_{k}^{n}(p,q)\rightarrow U_{k}(p,q) for all kk\in\mathbb{N} and p,qOp,q\in O. ∎

It remains to show symmetry (A.3) in the analytic case, for which we leave Moretti’s path and present a novel approach using the construction in [BGP2007] instead.
Let VkC(Ω×Ω,EE),k0,V_{k}\in C^{\infty}(\Omega\times\Omega,E\boxtimes E^{*}),\leavevmode\nobreak\ k\in\mathbb{N}_{0}, denote the Hadamard coefficients associated to (M,g,E,Pt)(M,g,E^{*},P^{t}). Employing formal self-adjointness of PP in the transport equations provides the relation

Vk(p,q)=Θp1Uk(p,q)Θq,p,qΩ.\displaystyle V_{k}(p,q)=\Theta_{p}^{-1}U_{k}(p,q)\Theta_{q},\qquad p,q\in\Omega. (A.9)

According to section 1.4 of [BGP2007], we define

R±Ω(α,p)[φ]:=R±α[(μpφ)expp],φ𝒟(Ω),\displaystyle R^{\Omega}_{\pm}(\alpha,p)[\varphi]:=R^{\alpha}_{\pm}\big{[}(\mu_{p}\varphi)\circ\exp_{p}\big{]},\qquad\varphi\in\mathscr{D}(\Omega), (A.10)

which ensures the identification R±Ω(α,p)|J±Ω(p)=C(α,d)Γpαd2R^{\Omega}_{\pm}(\alpha,p)\big{|}_{J_{\pm}^{\Omega}(p)}=C(\alpha,d)\cdot\Gamma_{p}^{\frac{\alpha-d}{2}} for Re(α)>d\text{Re}\left(\alpha\right)>d. Due to Proposition 2.4.6 of [BGP2007], they comprise Hadamard series, which yield advanced and retarded parametrices ~±(p)\widetilde{\mathscr{R}}_{\pm}(p) for PP at each pO¯p\in\overline{O} on any relatively compact domain OΩO\subset\Omega. More precisely, for any integer N>d2N>\frac{d}{2} and cut-off function σ𝒟((1,1),[0,1])\sigma\in\mathscr{D}\big{(}(-1,1),[0,1]\big{)} with σ|[12,12]=1\sigma\big{|}_{\left[-\frac{1}{2},\frac{1}{2}\right]}=1, there is a sequence {εk}kN(0,1]\{\varepsilon_{k}\}_{k\geq N}\subset(0,1] such that

~±(p)=k=0U~k(p,)R±Ω(2k+2,p),U~k:={Uk,k<N,(σΓεk)Uk,kN,\displaystyle\widetilde{\mathscr{R}}_{\pm}(p)=\sum_{k=0}^{\infty}\widetilde{U}_{k}(p,\cdot)\leavevmode\nobreak\ R^{\Omega}_{\pm}(2k+2,p),\qquad\widetilde{U}_{k}:=\left\{\begin{array}[]{cl}U_{k},&k<N,\\[5.69054pt] \left(\sigma\circ\frac{\Gamma}{\varepsilon_{k}}\right)\cdot U_{k},&k\geq N,\end{array}\right. (A.13)

represent well-defined distributions and (p,q)(P~±(p)δp)(q)C(Ω×Ω,EE)(p,q)\mapsto\big{(}P\widetilde{\mathscr{R}}_{\pm}(p)-\delta_{p}\big{)}(q)\in C^{\infty}(\Omega\times\Omega,E^{*}\boxtimes E). Furthermore, we have p~±(p)[φ]C(O¯,E)p\mapsto\widetilde{\mathscr{R}}_{\pm}(p)[\varphi]\in C^{\infty}(\overline{O},E^{*}) for fixed φ𝒟(O,E)\varphi\in\mathscr{D}(O,E^{*}), so regarded as bidistributions and due to compactness of O¯\overline{O}, they provide continuous operators

G~±:𝒟(O,E)C(O¯,E),φ(p~(p)[φ]).\displaystyle\widetilde{G}_{\pm}\colon\qquad\mathscr{D}(O,E^{*})\rightarrow C^{\infty}(\overline{O},E^{*}),\qquad\varphi\longmapsto\big{(}p\mapsto\widetilde{\mathscr{R}}_{\mp}(p)[\varphi]\big{)}. (A.14)

Consequently, G~±\widetilde{G}_{\pm} yield left parametrices for PtP^{t} with suppG~±φJ±O¯(suppφ)\text{supp}\,\widetilde{G}_{\pm}\varphi\subset J_{\pm}^{\overline{O}}(\text{supp}\,\varphi), and for G±G_{\pm} the advanced and retarded Green operator for PtP^{t}, the differences G±G~±G_{\pm}-\widetilde{G}_{\pm} are smoothing (their integral kernels correspond to the last equation in the proof of Proposition 2.5.1 of [BGP2007], which is actually a smooth section). Therefore, G~±\widetilde{G}_{\pm} represent an advanced and a retarded parametrix for PtP^{t} in the sense of Duistermaat-Hörmander.

Proposition A.5.

For all convex and relatively compact domains OΩO\subset\Omega, the maps

(p,q)k=0((U~k(p,)V~k(,p)t)R±Ω(2k+2,p))(q)\displaystyle(p,q)\longmapsto\sum_{k=0}^{\infty}\left(\big{(}\widetilde{U}_{k}(p,\cdot)-\widetilde{V}_{k}(\cdot,p)^{t}\big{)}R^{\Omega}_{\pm}(2k+2,p)\right)(q) (A.15)

define smooth sections in EEE^{*}\boxtimes E over O×OO\times O.

Proof.

By Lemma 3.4.4 of [BGP2007], the advanced and retarded Green operators for PP are given by GtG^{t}_{\mp}, so formal self-adjointness of PP and uniqueness of G±G_{\pm} lead to G±=ΘGtΘ1G_{\pm}=\Theta G_{\mp}^{t}\Theta^{-1} such that the operators G~±ΘG~tΘ1\widetilde{G}_{\pm}-\Theta\widetilde{G}^{t}_{\mp}\Theta^{-1} are smoothing:

G~±ΘG~tΘ1=G~±G±smoothingΘ(G~Gsmoothing)tΘ1.\widetilde{G}_{\pm}-\Theta\widetilde{G}^{t}_{\mp}\Theta^{-1}=\underbrace{\widetilde{G}_{\pm}-G_{\pm}}_{\text{smoothing}}-\leavevmode\nobreak\ \Theta\big{(}\underbrace{\widetilde{G}_{\mp}-G_{\mp}}_{\text{smoothing}}\big{)}^{t}\Theta^{-1}.

Recalling G~±t[φ,ψ]=G~±[ψ,φ]\widetilde{G}^{t}_{\pm}[\varphi,\psi]=\widetilde{G}_{\pm}[\psi,\varphi], we just have to show that the Schwartz kernel of G~±t\widetilde{G}_{\pm}^{t} is given by the distribution

k=0Θp1V~k(,p)tΘR±Ω(2j+2,p),pO,\sum_{k=0}^{\infty}\Theta^{-1}_{p}\widetilde{V}_{k}(\cdot,p)^{t}\Theta\leavevmode\nobreak\ R^{\Omega}_{\pm}(2j+2,p),\qquad p\in O,

which can be directly deduced from Lemma 1.4.3 of [BGP2007] and (A.9). ∎

Lemma A.6.

Let k0k\in\mathbb{N}_{0} and assume that for all quadruples (M,g,E,P)(M,g,E,P) as introduced in the beginning of the chapter with odd spacetime dimension and all lightlike related p,qOp,q\in O, we have

Uk(p,q)=Vk(q,p)t.U_{k}(p,q)=V_{k}(q,p)^{t}.

Then this equality holds for all p,qΩp,q\in\Omega and all (M,g,E,P)(M,g,E,P).

Proof.

Consider the setting (M,g,E,P)(M,g,E,P) with odd spacetime dimension dd, let p,qp,q be causally related and choose a,a2a,a^{\prime}\in\mathbb{R}^{2} such that aa2=Γ(p,q)\|a-a^{\prime}\|^{2}=\Gamma(p,q). It follows that (p,a),(q,a)(p,a),(q,a^{\prime}) are lightlike related in (M×2,g+gEucl)\big{(}M\times\mathbb{R}^{2},g+g_{\mathrm{Eucl}}\big{)} and thus

U^k((p,a),(q,a))=V^k((q,a),(p,a))t\widehat{U}_{k}\big{(}(p,a),(q,a^{\prime})\big{)}=\widehat{V}_{k}\big{(}(q,a^{\prime}),(p,a)\big{)}^{t}

by assumption. Therefore, Proposition A.4 provides Uk(p,q)=Vk(q,p)tU_{k}(p,q)=V_{k}(q,p)^{t}.
Let {gn}n\{g^{n}\}_{n\in\mathbb{N}} be an analytic approximation of gg and Ukn,VknU_{k}^{n},V_{k}^{n} the corresponding Hadamard coefficients. Write Dkn(p,q):=Ukn(p,q)Vkn(q,p)tD^{n}_{k}(p,q):=U^{n}_{k}(p,q)-V^{n}_{k}(q,p)^{t}, which depends analytically on p,qp,q due to Proposition A.2 and vanishes on Γ1(0)\Gamma^{-1}\big{(}\mathbb{R}_{\geq 0}\big{)}, so the identity theorem for analytic maps implies Dkn=0D_{k}^{n}=0 on all of O×OO\times O. Furthermore, by Proposition A.4, we have Dkn(p,q)Dk(p,q)D^{n}_{k}(p,q)\rightarrow D_{k}(p,q) and therefore Dk(p,q)=0D_{k}(p,q)=0 for all p,qp,q, which proves the claim in the case of odd spacetime dimension.
For even-dimensional settings (M,g,E,P)(M,g,E,P), this can be deduced from (M×,g+ds2,E,P2s2)\big{(}M\times\mathbb{R},g+\,\text{d}s^{2},E,P-\frac{\partial^{2}}{\partial s^{2}}\big{)}, which is odd-dimensional, and Proposition A.4.
Since this works for any relatively compact and convex domain OΩO\subset\Omega, by uniqueness of the Hadamard coefficients, an appropriate exhaustion of Ω\Omega by such subsets proves the claim on all of Ω×Ω\Omega\times\Omega. ∎

Theorem A.7.

Let MM be a Lorentzian manifold of dimension dd, π:EM\pi\colon E\rightarrow M a real or complex vector bundle over MM with non-degenerate inner product, P:C(M,E)C(M,E)P\colon C^{\infty}(M,E)\rightarrow C^{\infty}(M,E) a formally self-adjoint wave operator and ΩM\Omega\subset M a convex domain. Then the Hadamard coefficients UkC(Ω×Ω,EE)U_{k}\in C^{\infty}(\Omega\times\Omega,E^{*}\boxtimes E) are symmetric in the sense

Uk(p,q)=ΘpUk(q,p)tΘq1,p,qΩ,k0.\displaystyle U_{k}(p,q)=\Theta_{p}U_{k}(q,p)^{t}\Theta_{q}^{-1},\qquad p,q\in\Omega,\leavevmode\nobreak\ k\in\mathbb{N}_{0}. (A.16)
Proof.

Let dd be odd. For all k,j0k,j\in\mathbb{N}_{0} with jkj\leq k, Lemma 1.4.2 (1) of [BGP2007] provides the recursion

R±Ω(2k+2,p)=Kk,j,dΓ(p,)kjR±Ω(2j+2,p),\displaystyle R^{\Omega}_{\pm}(2k+2,p)=K_{k,j,d}\Gamma(p,\cdot)^{k-j}\cdot R^{\Omega}_{\pm}(2j+2,p), (A.17)

with Kk,j,d\{0}K_{k,j,d}\in\mathbb{R}\backslash\{0\} as defined in (3.33), so the smooth sections (A.15) can be rewritten into

R±Ω(2,p)k=0Kk,0,d(U~k(p,)V~k(,p)t)Γ(p,)k.R^{\Omega}_{\pm}(2,p)\sum_{k=0}^{\infty}K_{k,0,d}\big{(}\widetilde{U}_{k}(p,\cdot)-\widetilde{V}_{k}(\cdot,p)^{t}\big{)}\Gamma(p,\cdot)^{k}.

The proof of Lemma 2.4.2 in [BGP2007] shows that the sum defines a smooth section, which has to vanish for lightlike related p,qp,q since singsuppR±Ω(2,p)=C±Ω(p)\text{sing}\,\text{supp}\,R^{\Omega}_{\pm}(2,p)=C_{\pm}^{\Omega}(p). Due to σ(Γ(p,q)εk)=σ(0)=1\sigma\left(\frac{\Gamma(p,q)}{\varepsilon_{k}}\right)=\sigma(0)=1, this leads to

0=k=0(U~k(p,q)V~k(q,p)t)Γ(p,q)k=U0(p,q)V0(q,p)t,0=\sum_{k=0}^{\infty}\big{(}\widetilde{U}_{k}(p,q)-\widetilde{V}_{k}(q,p)^{t}\big{)}\Gamma(p,q)^{k}=U_{0}(p,q)-V_{0}(q,p)^{t},

and it follows from Lemma A.6 that for k=0k=0, (A.16) is true also for even dd and on all of Ω×Ω\Omega\times\Omega.
Now let dd again be odd, and for some k0k_{0}\in\mathbb{N}, assume (A.16) to hold for all k=0,,k01k=0,\ldots,k_{0}-1, i.e. the smooth section (A.15) is given by

k=k0(U~k(p,)V~k(,p)t)R±Ω(2k+2,p)=R±Ω(2k0+2,p)k=k0Kk,k0,d(U~k(p,)V~k(,p)t)Γ(p,)kk0.\sum_{k=k_{0}}^{\infty}\big{(}\widetilde{U}_{k}(p,\cdot)-\widetilde{V}_{k}(\cdot,p)^{t}\big{)}R^{\Omega}_{\pm}(2k+2,p)=R^{\Omega}_{\pm}(2k_{0}+2,p)\sum_{k=k_{0}}^{\infty}K_{k,k_{0},d}\big{(}\widetilde{U}_{k}(p,\cdot)-\widetilde{V}_{k}(\cdot,p)^{t}\big{)}\Gamma(p,\cdot)^{k-k_{0}}.

Analogously, we obtain 0=Uk0(p,q)Vk0(q,p)t0=U_{k_{0}}(p,q)-V_{k_{0}}(q,p)^{t} if Γ(p,q)=0\Gamma(p,q)=0, so again, applying Lemma A.6 completes the proof by induction. ∎

B Derivation of transport equations

Since Definition A.1 of the PP-compatible connection \nabla implies a product rule for P(fs)P(f\cdot s), for odd dd, a straight forward calculation provides

L±Ω(0,p)\displaystyle L^{\Omega}_{\pm}(0,p) =δp=!P±(p)=k=0P(UpkL±Ω(2k+2,p))+k=d22P(WpkLΩ(2k+2,p))\displaystyle=\delta_{p}\stackrel{{\scriptstyle!}}{{=}}P\mathscr{L}_{\pm}(p)=\sum_{k=0}^{\infty}P\left(U_{p}^{k}L^{\Omega}_{\pm}(2k+2,p)\right)+\sum_{k=\frac{d-2}{2}}^{\infty}P\left(W_{p}^{k}L^{\Omega}(2k+2,p)\right)
=Up0L±Ω(2,p)2gradL±Ω(2,p)Up0+k=1(PUpk1L±Ω(2k,p)2gradL±Ω(2k+2,p)Upk+UpkL±Ω(2k+2,p))\displaystyle=U_{p}^{0}\leavevmode\nobreak\ \raisebox{0.0pt}{\text{{$\Box$}}}L_{\pm}^{\Omega}(2,p)-2\nabla_{\text{grad}L_{\pm}^{\Omega}(2,p)}U_{p}^{0}+\sum_{k=1}^{\infty}\left(PU_{p}^{k-1}L^{\Omega}_{\pm}(2k,p)-2\nabla_{\text{grad}L_{\pm}^{\Omega}(2k+2,p)}U_{p}^{k}+U_{p}^{k}\leavevmode\nobreak\ \raisebox{0.0pt}{\text{{$\Box$}}}L^{\Omega}_{\pm}(2k+2,p)\right)
+k=d2(PWpk1LΩ(2k,p)2gradLΩ(2k+2,p)Wpk+WpkLΩ(2k+2,p))\displaystyle\hskip 28.45274pt+\sum_{k=\frac{d}{2}}^{\infty}\left(PW_{p}^{k-1}L^{\Omega}(2k,p)-2\nabla_{\text{grad}L^{\Omega}(2k+2,p)}W_{p}^{k}+W_{p}^{k}\leavevmode\nobreak\ \raisebox{0.0pt}{\text{{$\Box$}}}L^{\Omega}(2k+2,p)\right)
=Up0L±Ω(2,p)2gradL±Ω(2,p)Up0+k=112k(2kPUpk1gradΓpUpk+(12Γpd+2k)Upk)L±Ω(2k,p)\displaystyle=U_{p}^{0}\leavevmode\nobreak\ \raisebox{0.0pt}{\text{{$\Box$}}}L_{\pm}^{\Omega}(2,p)-2\nabla_{\text{grad}L_{\pm}^{\Omega}(2,p)}U_{p}^{0}+\sum_{k=1}^{\infty}\frac{1}{2k}\left(2k\leavevmode\nobreak\ PU_{p}^{k-1}-\nabla_{\text{grad}\Gamma_{p}}U_{p}^{k}+\left(\frac{1}{2}\raisebox{0.0pt}{\text{{$\Box$}}}\Gamma_{p}-d+2k\right)U_{p}^{k}\right)L^{\Omega}_{\pm}(2k,p)
+k=d212k(2kPWpk1gradΓpWpk+(12Γpd+2k)Wpk)LΩ(2k,p).\displaystyle\hskip 28.45274pt+\sum_{k=\frac{d}{2}}^{\infty}\frac{1}{2k}\left(2k\leavevmode\nobreak\ PW_{p}^{k-1}-\nabla_{\text{grad}\Gamma_{p}}W_{p}^{k}+\left(\frac{1}{2}\raisebox{0.0pt}{\text{{$\Box$}}}\Gamma_{p}-d+2k\right)W_{p}^{k}\right)L^{\Omega}(2k,p).

For even dd, we need to relate the logarithmic and non-logarithmic parts, for which we first prove the following technical Lemma:

Lemma B.1.

Let dd be even and kk\in\mathbb{N} with kd2k\geq\frac{d}{2}. Then we have

grad(LΩ(2k+2,p)log(Γp±i0))=gradΓp4kLΩ(2k,p)log(Γp±i0)+gradΓp2k(2k+2d)LΩ(2k,p),\displaystyle\begin{split}&\text{grad}\left(L^{\Omega}(2k+2,p)\cdot\log(\Gamma_{p}\pm i0)\right)\\[5.69054pt] &\hskip 56.9055pt=\frac{\text{grad}\Gamma_{p}}{4k}\cdot L^{\Omega}(2k,p)\log(\Gamma_{p}\pm i0)+\frac{\text{grad}\Gamma_{p}}{2k(2k+2-d)}\cdot L^{\Omega}(2k,p),\end{split}
(LΩ(2k+2,p)log(Γp±i0))=Γp2d+2k2kLΩ(2k,p)log(Γp±i0)+Γp22d+4k+22k(kd22)LΩ(2k,p),\displaystyle\begin{split}&\raisebox{0.0pt}{\text{{$\Box$}}}\left(L^{\Omega}(2k+2,p)\cdot\log(\Gamma_{p}\pm i0)\right)\\[5.69054pt] &\hskip 56.9055pt=\frac{\frac{\Box\Gamma_{p}}{2}-d+2k}{2k}\cdot L^{\Omega}(2k,p)\cdot\log(\Gamma_{p}\pm i0)+\frac{\frac{\Box\Gamma_{p}}{2}-2d+4k+2}{2k\left(k-\frac{d-2}{2}\right)}\cdot L^{\Omega}(2k,p),\end{split}

and for k=d22k=\frac{d-2}{2}

grad(LΩ(d,p)log(Γp±i0))\displaystyle\text{grad}\left(L^{\Omega}(d,p)\cdot\log(\Gamma_{p}\pm i0)\right) =iπgradΓp2(d2)L~±Ω(d2,p)\displaystyle=\mp\frac{i\pi\cdot\text{grad}\Gamma_{p}}{2(d-2)}\cdot\widetilde{L}^{\Omega}_{\pm}(d-2,p)
(LΩ(d,p)log(Γp±i0))\displaystyle\raisebox{0.0pt}{\text{{$\Box$}}}\left(L^{\Omega}(d,p)\cdot\log(\Gamma_{p}\pm i0)\right) =iπ(Γp4)4(d2)L~±Ω(d2,p).\displaystyle=\mp\frac{i\pi\left(\raisebox{0.0pt}{\text{{$\Box$}}}\Gamma_{p}-4\right)}{4(d-2)}\cdot\widetilde{L}^{\Omega}_{\pm}(d-2,p).
Proof.

Proposition 3.3 (3) provides

grad(LΩ(2k+2,p)log(Γp±i0))\displaystyle\text{grad}\left(L^{\Omega}(2k+2,p)\cdot\log(\Gamma_{p}\pm i0)\right)
=log(Γp±i0)gradLΩ(2k+2,p)+LΩ(2k+2,p)gradlog(Γp±i0)\displaystyle\hskip 56.9055pt=\log(\Gamma_{p}\pm i0)\cdot\text{grad}L^{\Omega}(2k+2,p)+L^{\Omega}(2k+2,p)\cdot\text{grad}\log(\Gamma_{p}\pm i0)
=gradΓp4kLΩ(2k,p)log(Γp±i0)+ΓpLΩ(2k,p)2k(2k+2d)gradΓpΓp\displaystyle\hskip 56.9055pt=\frac{\text{grad}\Gamma_{p}}{4k}\cdot L^{\Omega}(2k,p)\log(\Gamma_{p}\pm i0)+\frac{\Gamma_{p}\cdot L^{\Omega}(2k,p)}{2k(2k+2-d)}\cdot\frac{\text{grad}\Gamma_{p}}{\Gamma_{p}}
=gradΓp4kLΩ(2k,p)log(Γp±i0)+gradΓp2k(2k+2d)LΩ(2k,p),\displaystyle\hskip 56.9055pt=\frac{\text{grad}\Gamma_{p}}{4k}\cdot L^{\Omega}(2k,p)\log(\Gamma_{p}\pm i0)+\frac{\text{grad}\Gamma_{p}}{2k(2k+2-d)}\cdot L^{\Omega}(2k,p),

and hence, gradΓp,gradΓp=4Γp\left\langle\text{grad}\Gamma_{p},\text{grad}\Gamma_{p}\right\rangle=-4\Gamma_{p} implies

(LΩ(2k+2,p)log(Γp±i0))\displaystyle\raisebox{0.0pt}{\text{{$\Box$}}}\left(L^{\Omega}(2k+2,p)\cdot\log(\Gamma_{p}\pm i0)\right)
=\mydiv(gradΓp4kLΩ(2k,p)log(Γp±i0)+gradΓp2k(2k+2d)LΩ(2k,p))\displaystyle\hskip 28.45274pt=-\mydiv\left(\frac{\text{grad}\Gamma_{p}}{4k}\cdot L^{\Omega}(2k,p)\log(\Gamma_{p}\pm i0)+\frac{\text{grad}\Gamma_{p}}{2k(2k+2-d)}\cdot L^{\Omega}(2k,p)\right)
=14k(ΓpLΩ(2k,p)log(Γp±i0)LΩ(2k2,p)log(Γp±i0)4(k1)gradΓp,gradΓp\displaystyle\hskip 28.45274pt=\frac{1}{4k}\left(\raisebox{0.0pt}{\text{{$\Box$}}}\Gamma_{p}\cdot L^{\Omega}(2k,p)\log(\Gamma_{p}\pm i0)-\frac{L^{\Omega}(2k-2,p)\log(\Gamma_{p}\pm i0)}{4(k-1)}\left\langle\text{grad}\Gamma_{p},\text{grad}\Gamma_{p}\right\rangle\right.
LΩ(2k,p)gradΓp,gradΓpΓp)+ΓpLΩ(2k,p)LΩ(2k2,p)4(k1)gradΓp,gradΓp2k(2k+2d)\displaystyle\hskip 56.9055pt\left.-L^{\Omega}(2k,p)\cdot\frac{\left\langle\text{grad}\Gamma_{p},\text{grad}\Gamma_{p}\right\rangle}{\Gamma_{p}}\right)+\frac{\raisebox{0.0pt}{\text{{$\Box$}}}\Gamma_{p}\cdot L^{\Omega}(2k,p)-\frac{L^{\Omega}(2k-2,p)}{4(k-1)}\cdot\left\langle\text{grad}\Gamma_{p},\text{grad}\Gamma_{p}\right\rangle}{2k(2k+2-d)}
=14k(Γp+2(2kd))LΩ(2k,p)log(Γp±i0)+(1k+Γp+2(2kd)2k(2k+2d))LΩ(2k,p)\displaystyle\hskip 28.45274pt=\frac{1}{4k}\big{(}\raisebox{0.0pt}{\text{{$\Box$}}}\Gamma_{p}+2(2k-d)\big{)}L^{\Omega}(2k,p)\log(\Gamma_{p}\pm i0)+\left(\frac{1}{k}+\frac{\raisebox{0.0pt}{\text{{$\Box$}}}\Gamma_{p}+2(2k-d)}{2k(2k+2-d)}\right)L^{\Omega}(2k,p)
=Γp2d+2k2kLΩ(2k,p)log(Γp±i0)+Γp22n+4k+22k(kd22)LΩ(2k,p).\displaystyle\hskip 28.45274pt=\frac{\frac{\Box\Gamma_{p}}{2}-d+2k}{2k}\cdot L^{\Omega}(2k,p)\log(\Gamma_{p}\pm i0)+\frac{\frac{\Box\Gamma_{p}}{2}-2n+4k+2}{2k\left(k-\frac{d-2}{2}\right)}\cdot L^{\Omega}(2k,p).

By Proposition 3.3 (1) LΩ(d,p)=C(d,d)=iπC~(d2,d)2(d2)L^{\Omega}(d,p)=C(d,d)=\mp\frac{i\pi\widetilde{C}(d-2,d)}{2(d-2)} is constant on Ω\Omega, so again using (3) yields

grad(LΩ(d,p)log(Γp±i0))\displaystyle\text{grad}\left(L^{\Omega}(d,p)\cdot\log(\Gamma_{p}\pm i0)\right) =iπC~(d2,d)2(d2)gradΓpΓp=iπgradΓp2(d2)L~±Ω(d2,p),\displaystyle=\mp\frac{i\pi\widetilde{C}(d-2,d)}{2(d-2)}\cdot\frac{\text{grad}\Gamma_{p}}{\Gamma_{p}}=\mp\frac{i\pi\cdot\text{grad}\Gamma_{p}}{2(d-2)}\cdot\widetilde{L}^{\Omega}_{\pm}(d-2,p),
(LΩ(d,p)log(Γp±i0))\displaystyle\raisebox{0.0pt}{\text{{$\Box$}}}\left(L^{\Omega}(d,p)\cdot\log(\Gamma_{p}\pm i0)\right) =±iπ2(d2)\mydiv(gradΓpL~±Ω(d2,p))\displaystyle=\pm\frac{i\pi}{2(d-2)}\mydiv\left(\text{grad}\Gamma_{p}\cdot\widetilde{L}^{\Omega}_{\pm}(d-2,p)\right)
=iπ2(d2)(ΓpL~±Ω(d2,p)gradΓp,gradΓp2(d4)L~±Ω(d4,p))\displaystyle=\mp\frac{i\pi}{2(d-2)}\left(\raisebox{0.0pt}{\text{{$\Box$}}}\Gamma_{p}\cdot\widetilde{L}^{\Omega}_{\pm}(d-2,p)-\frac{\left\langle\text{grad}\Gamma_{p},\text{grad}\Gamma_{p}\right\rangle}{2(d-4)}\cdot\widetilde{L}^{\Omega}_{\pm}(d-4,p)\right)
=iπ2(d2)(Γp+42(d4)(d4)(d4d+2))L~±Ω(d2,p)\displaystyle=\mp\frac{i\pi}{2(d-2)}\left(\raisebox{0.0pt}{\text{{$\Box$}}}\Gamma_{p}+\frac{4}{2(d-4)}\cdot(d-4)(d-4-d+2)\right)\cdot\widetilde{L}^{\Omega}_{\pm}(d-2,p)
=iπ(Γp4)2(d2)L~±Ω(d2,p).\displaystyle=\mp\frac{i\pi\left(\raisebox{0.0pt}{\text{{$\Box$}}}\Gamma_{p}-4\right)}{2(d-2)}\cdot\widetilde{L}^{\Omega}_{\pm}(d-2,p).\qed

This leads to the following calculation:

L~±Ω(0,p)=δp=!P±(p)\displaystyle\widetilde{L}^{\Omega}_{\pm}(0,p)=\delta_{p}\stackrel{{\scriptstyle!}}{{=}}P\mathscr{L}_{\pm}(p)
=k=0d42P(UpkL~±Ω(2k+2,p))±iπk=d22P(UpkLΩ(2k+2,p)log(Γp±i0)+WpkLΩ(2k+2,p))\displaystyle\hskip 8.5359pt=\sum_{k=0}^{\frac{d-4}{2}}P\left(U_{p}^{k}\widetilde{L}^{\Omega}_{\pm}(2k+2,p)\right)\pm\frac{i}{\pi}\sum_{k=\frac{d-2}{2}}^{\infty}P\left(U_{p}^{k}L^{\Omega}(2k+2,p)\log(\Gamma_{p}\pm i0)+W_{p}^{k}L^{\Omega}(2k+2,p)\right)
=Up0L~±Ω(2,p)2gradL~±Ω(2,p)Up0±iπ(Wpd22LΩ(d,p)=02gradLΩ(d,p)Wpd22=0)\displaystyle\hskip 8.5359pt=U_{p}^{0}\leavevmode\nobreak\ \raisebox{0.0pt}{\text{{$\Box$}}}\widetilde{L}^{\Omega}_{\pm}(2,p)-2\nabla_{\text{grad}\widetilde{L}^{\Omega}_{\pm}(2,p)}U_{p}^{0}\pm\frac{i}{\pi}\bigg{(}W_{p}^{\frac{d-2}{2}}\overbrace{\raisebox{0.0pt}{\text{{$\Box$}}}L^{\Omega}(d,p)}^{=0}-\overbrace{2\nabla_{\text{grad}L^{\Omega}(d,p)}W_{p}^{\frac{d-2}{2}}}^{=0}\bigg{)}
+k=1d42(UpkL~±Ω(2k+2,p)2gradL~±Ω(2k+2,p)Upk+L~±Ω(2k,p)PUpk1)+L~±Ω(d2,p)PUpd42\displaystyle\hskip 17.07182pt+\sum_{k=1}^{\frac{d-4}{2}}\left(U_{p}^{k}\leavevmode\nobreak\ \raisebox{0.0pt}{\text{{$\Box$}}}\widetilde{L}^{\Omega}_{\pm}(2k+2,p)-2\nabla_{\text{grad}\widetilde{L}^{\Omega}_{\pm}(2k+2,p)}U_{p}^{k}+\widetilde{L}^{\Omega}_{\pm}(2k,p)PU_{p}^{k-1}\right)+\widetilde{L}^{\Omega}_{\pm}(d-2,p)PU_{p}^{\frac{d-4}{2}}
±iπ{Upd22(LΩ(d,p)log(Γp±i0))2grad(LΩ(d,p)log(Γp±i0))Upd22\displaystyle\hskip 17.07182pt\pm\frac{i}{\pi}\bigg{\{}U_{p}^{\frac{d-2}{2}}\raisebox{0.0pt}{\text{{$\Box$}}}\!\left(L^{\Omega}(d,p)\log(\Gamma_{p}\pm i0)\right)-2\nabla_{\text{grad}\left(L^{\Omega}(d,p)\log(\Gamma_{p}\pm i0)\right)}U_{p}^{\frac{d-2}{2}}
+k=d2[LΩ(2k,p)log(Γp±i0)PUpk12grad(LΩ(2k+2,p)log(Γp±i0))Upk+Upk(LΩ(2k+2,p)log(Γp±i0))]\displaystyle\hskip 17.07182pt+\sum_{k=\frac{d}{2}}^{\infty}\left[L^{\Omega}(2k,p)\log(\Gamma_{p}\pm i0)PU_{p}^{k-1}-2\nabla_{\text{grad}\left(L^{\Omega}(2k+2,p)\log(\Gamma_{p}\pm i0)\right)}U_{p}^{k}+U_{p}^{k}\leavevmode\nobreak\ \raisebox{0.0pt}{\text{{$\Box$}}}\!\left(L^{\Omega}(2k+2,p)\log(\Gamma_{p}\pm i0)\right)\right]
+k=d2(WpkLΩ(2k+2,p)2gradLΩ(2k+2,p)Wpk+LΩ(2k,p)PWpk1)}\displaystyle\hskip 17.07182pt+\sum_{k=\frac{d}{2}}^{\infty}\left(W_{p}^{k}\leavevmode\nobreak\ \raisebox{0.0pt}{\text{{$\Box$}}}L^{\Omega}(2k+2,p)-2\nabla_{\text{grad}L^{\Omega}(2k+2,p)}W_{p}^{k}+L^{\Omega}(2k,p)PW_{p}^{k-1}\right)\bigg{\}}
=Up0L~±Ω(2,p)2gradL~±Ω(2,p)Up0+k=1d42[(12Γpd+2k)UpkgradΓpUpk+2kPUpk1]L~±Ω(2k,p)2k\displaystyle\hskip 8.5359pt=U_{p}^{0}\leavevmode\nobreak\ \raisebox{0.0pt}{\text{{$\Box$}}}\widetilde{L}^{\Omega}_{\pm}(2,p)-2\nabla_{\text{grad}\widetilde{L}^{\Omega}_{\pm}(2,p)}U_{p}^{0}+\sum_{k=1}^{\frac{d-4}{2}}\left[\left(\frac{1}{2}\raisebox{0.0pt}{\text{{$\Box$}}}\Gamma_{p}-d+2k\right)U_{p}^{k}-\nabla_{\text{grad}\Gamma_{p}}U_{p}^{k}+2kPU_{p}^{k-1}\right]\frac{\widetilde{L}^{\Omega}_{\pm}(2k,p)}{2k}
+((d2)PUpd42+(12Γp2)Upd22gradΓpUpd22)L~±Ω(d2,p)d2\displaystyle\hskip 17.07182pt+\left((d-2)PU_{p}^{\frac{d-4}{2}}+\left(\frac{1}{2}\raisebox{0.0pt}{\text{{$\Box$}}}\Gamma_{p}-2\right)U_{p}^{\frac{d-2}{2}}-\nabla_{\text{grad}\Gamma_{p}}U_{p}^{\frac{d-2}{2}}\right)\frac{\widetilde{L}^{\Omega}_{\pm}(d-2,p)}{d-2}
±iπk=d2[(12Γpd+2k)UpkgradΓpUpk+2kPUpk1]LΩ(2k,p)log(Γp±i0)2k\displaystyle\hskip 17.07182pt\pm\frac{i}{\pi}\sum_{k=\frac{d}{2}}^{\infty}\left[\left(\frac{1}{2}\raisebox{0.0pt}{\text{{$\Box$}}}\Gamma_{p}-d+2k\right)U_{p}^{k}-\nabla_{\text{grad}\Gamma_{p}}U_{p}^{k}+2kPU_{p}^{k-1}\right]\frac{L^{\Omega}(2k,p)\log(\Gamma_{p}\pm i0)}{2k}
±iπk=d2[(12Γp+2+4k2d)UpkgradΓpUpk]LΩ(2k,p)2k(kd22)\displaystyle\hskip 17.07182pt\pm\frac{i}{\pi}\sum_{k=\frac{d}{2}}^{\infty}\left[\left(\frac{1}{2}\raisebox{0.0pt}{\text{{$\Box$}}}\Gamma_{p}+2+4k-2d\right)U_{p}^{k}-\nabla_{\text{grad}\Gamma_{p}}U_{p}^{k}\right]\frac{L^{\Omega}(2k,p)}{2k\left(k-\frac{d-2}{2}\right)}
±iπk=d2[(12Γp+2kd)WpkgradΓpWpk+2kPWpk1]LΩ(2k,p)2k,\displaystyle\hskip 17.07182pt\pm\frac{i}{\pi}\sum_{k=\frac{d}{2}}^{\infty}\left[\left(\frac{1}{2}\raisebox{0.0pt}{\text{{$\Box$}}}\Gamma_{p}+2k-d\right)W_{p}^{k}-\nabla_{\text{grad}\Gamma_{p}}W_{p}^{k}+2kPW_{p}^{k-1}\right]\frac{L^{\Omega}(2k,p)}{2k},

from which one reads off the transport equations (3.12) and (3.15).

C Proof of Proposition 3.9 for even dimensional spacetimes

Note that in the proofs of all following Lemmas, cc denotes a generic constant, i.e. its particular value can change from one line to another.

Lemma C.1.

For all ll\in\mathbb{N} and βl+1\beta\geq l+1, there is some cl,βc_{l,\beta} such that for all 0<ε10<\varepsilon\leq 1, we have

dldtl(σ(tε)tβlogt)C0()ε(log1ε+π+1)cl,βσCl().\displaystyle\left\|\frac{\,\text{d}^{l}}{\,\text{d}t^{l}}\left(\sigma\left(\frac{t}{\varepsilon}\right)\cdot t^{\beta}\cdot\log t\right)\right\|_{C^{0}(\mathbb{R})}\leq\varepsilon\left(\log\frac{1}{\varepsilon}+\pi+1\right)\cdot c_{l,\beta}\cdot\|\sigma\|_{C^{l}(\mathbb{R})}.
Proof.

We start with calculating

djdtj(tβlogt)C0()\displaystyle\left\|\frac{\,\text{d}^{j}}{\,\text{d}t^{j}}\left(t^{\beta}\cdot\log t\right)\right\|_{C^{0}(\mathbb{R})}
=β(β1)(βj+1)tβjlogt+tβji=1j(ji)(1)i1(i1)!β(β1)(βj+i+1)C0()\displaystyle\hskip 14.22636pt=\left\|\beta(\beta-1)\ldots(\beta-j+1)t^{\beta-j}\log t+t^{\beta-j}\sum_{i=1}^{j}\binom{j}{i}(-1)^{i-1}(i-1)!\beta(\beta-1)\ldots(\beta-j+i+1)\right\|_{C^{0}(\mathbb{R})}
cj,β|t|βj(|logt|+1)\displaystyle\hskip 14.22636pt\leq c_{j,\beta}\cdot|t|^{\beta-j}\left(|\log t|+1\right)
cj,β|t|βj(|log|t||+π+1).\displaystyle\hskip 14.22636pt\leq c_{j,\beta}\cdot|t|^{\beta-j}\left(\big{|}\log|t|\big{|}+\pi+1\right).

Since σ(lj)(tε)=0\sigma^{(l-j)}\left(\frac{t}{\varepsilon}\right)=0 for |t|ε|t|\geq\varepsilon and εβlε\varepsilon^{\beta-l}\leq\varepsilon due to βl+1\beta\geq l+1, this yields

dldtl(σ(tε)tβlogt)C0()\displaystyle\left\|\frac{\,\text{d}^{l}}{\,\text{d}t^{l}}\left(\sigma\left(\frac{t}{\varepsilon}\right)\cdot t^{\beta}\cdot\log t\right)\right\|_{C^{0}(\mathbb{R})} j=0l(lj)cj,βσ(lj)(tε)εlj|t|βj(|log|t||+π+1)C0()\displaystyle\leq\sum_{j=0}^{l}\binom{l}{j}\cdot c_{j,\beta}\left\|\frac{\sigma^{(l-j)}\left(\frac{t}{\varepsilon}\right)}{\varepsilon^{l-j}}\cdot|t|^{\beta-j}\left(\big{|}\log|t|\big{|}+\pi+1\right)\right\|_{C^{0}(\mathbb{R})}
j=0l(lj)cj,βεβl(log1ε+π+1)σ(lj)C0()\displaystyle\leq\sum_{j=0}^{l}\binom{l}{j}\cdot c_{j,\beta}\cdot\varepsilon^{\beta-l}\left(\log\frac{1}{\varepsilon}+\pi+1\right)\cdot\left\|\sigma^{(l-j)}\right\|_{C^{0}(\mathbb{R})}
cl,βε(log1ε+π+1)σCl().\displaystyle\leq c_{l,\beta}\cdot\varepsilon\left(\log\frac{1}{\varepsilon}+\pi+1\right)\cdot\left\|\sigma\right\|_{C^{l}(\mathbb{R})}.

Lemma C.2.

For any open and relatively compact domain OΩO\subset\Omega and l0l\in\mathbb{N}_{0}, there is a sequence {εk}kN(0,1]\{\varepsilon_{k}\}_{k\geq N}\subset(0,1] such that for all l0l\geq 0 the series

(p,q)k=N+lU~k(p,q)log(Γ(p,q)±i0)L±Ω(2k+2,p)(q)\displaystyle(p,q)\longmapsto\sum_{k=N+l}^{\infty}\widetilde{U}_{k}(p,q)\log\big{(}\Gamma(p,q)\pm i0\big{)}\leavevmode\nobreak\ L^{\Omega}_{\pm}(2k+2,p)(q) (C.18)

converges in Cl(O¯×O¯,EE)C^{l}\left(\overline{O}\times\overline{O},E^{*}\boxtimes E\right). In particular, for l=0l=0, this defines a continuous section over O¯×O¯\overline{O}\times\overline{O} and a smooth section over (O¯×O¯)\Γ1(0)\big{(}\overline{O}\times\overline{O}\big{)}\backslash\Gamma^{-1}(0).

Proof.

Since kNd2k\geq N\geq\frac{d}{2} and dd even, Γ(p,q)kd22\Gamma(p,q)^{k-\frac{d-2}{2}} is a smooth and Γ(p,q)kd22log(Γ(p,q)±i0)\Gamma(p,q)^{k-\frac{d-2}{2}}\cdot\log\left(\Gamma(p,q)\pm i0\right) a continuous section over O¯×O¯\overline{O}\times\overline{O}, so every single summand of (C.18) is at least continuous, individually. Due to supp(σΓεk){Γ(p,q)<εk}\text{supp}\,\left(\sigma\circ\frac{\Gamma}{\varepsilon_{k}}\right)\subset\{\Gamma(p,q)<\varepsilon_{k}\} for all kNk\geq N and 0σ10\leq\sigma\leq 1 by Lemma C.1, we have

U~klog(Γ±i0)C(2k+2,d)Γkd22C0(O¯×O¯)ck,dUkC0(O¯×O¯)εk(log1εk+π+1).\displaystyle\left\|\widetilde{U}_{k}\log\big{(}\Gamma\pm i0\big{)}\cdot C(2k+2,d)\Gamma^{k-\frac{d-2}{2}}\right\|_{C^{0}\left(\overline{O}\times\overline{O}\right)}\leq c_{k,d}\left\|U_{k}\right\|_{C^{0}\left(\overline{O}\times\overline{O}\right)}\cdot\varepsilon_{k}\left(\log\frac{1}{\varepsilon_{k}}+\pi+1\right).

Since εklog1εk0\varepsilon_{k}\log\frac{1}{\varepsilon_{k}}\rightarrow 0 for εk0\varepsilon_{k}\rightarrow 0, we can choose εk\varepsilon_{k} such that

ck,dUkC0εk(log1εk+π+1)<2k,\displaystyle c_{k,d}\left\|U_{k}\right\|_{C^{0}}\cdot\varepsilon_{k}\left(\log\frac{1}{\varepsilon_{k}}+\pi+1\right)<2^{-k},

and (C.18) converges in C0C^{0}. Now let l>0l>0 and kN+lk\geq N+l such that Γ(p,q)kd22log(Γ(p,q)±i0)\Gamma(p,q)^{k-\frac{d-2}{2}}\log\left(\Gamma(p,q)\pm i0\right) is of ClC^{l}-regularity. Set ρk(t):=σ(tεk)tkd22\rho_{k}(t):=\sigma\left(\frac{t}{\varepsilon_{k}}\right)t^{k-\frac{d-2}{2}}, so by Lemma C.1 we have

ρkCl()cl,k,dεkσCl(),ρklogCl()cl,k,dεk(log1εk+π+1)σCl(),\displaystyle\|\rho_{k}\|_{C^{l}(\mathbb{R})}\leq c_{l,k,d}\cdot\varepsilon_{k}\|\sigma\|_{C^{l}(\mathbb{R})},\qquad\|\rho_{k}\cdot\log\|_{C^{l}(\mathbb{R})}\leq c_{l,k,d}\cdot\varepsilon_{k}\left(\log\frac{1}{\varepsilon_{k}}+\pi+1\right)\|\sigma\|_{C^{l}(\mathbb{R})}, (C.19)

and Lemma 1.1.11 and 1.1.12 of [BGP2007] yield

U~klog(Γ±i0)C(2k+2,d)Γkd22Cl(O¯×O¯)\displaystyle\left\|\widetilde{U}_{k}\log\big{(}\Gamma\pm i0\big{)}\cdot C(2k+2,d)\Gamma^{k-\frac{d-2}{2}}\right\|_{C^{l}\left(\overline{O}\times\overline{O}\right)} cl,k,dUk((ρklog)Γ)Cl(O¯×O¯)\displaystyle\leq c_{l,k,d}\|U_{k}\cdot\big{(}(\rho_{k}\cdot\log)\circ\Gamma\big{)}\|_{C^{l}\left(\overline{O}\times\overline{O}\right)}
1.1.11cl,k,dUkCl(O¯×O¯)(ρklog)ΓCl(O¯×O¯)\displaystyle\stackrel{{\scriptstyle\text{1.1.11}}}{{\leq}}c_{l,k,d}\|U_{k}\|_{C^{l}\left(\overline{O}\times\overline{O}\right)}\cdot\|(\rho_{k}\cdot\log)\circ\Gamma\|_{C^{l}\left(\overline{O}\times\overline{O}\right)}
1.1.12cl,k,dUkCl(O¯×O¯)maxj=0,,lΓCl(O¯×O¯)jρklogCl()\displaystyle\stackrel{{\scriptstyle\text{1.1.12}}}{{\leq}}c_{l,k,d}\|U_{k}\|_{C^{l}\left(\overline{O}\times\overline{O}\right)}\cdot\max_{j=0,\ldots,l}\|\Gamma\|^{j}_{C^{l}\left(\overline{O}\times\overline{O}\right)}\cdot\|\rho_{k}\cdot\log\|_{C^{l}(\mathbb{R})}
(C.19)cl,k,dσCl()maxj=0,,lΓCl(O¯×O¯)jUkCl(O¯×O¯)εklog(1εk+π+1).\displaystyle\stackrel{{\scriptstyle(\ref{EstRhok})}}{{\leq}}c_{l,k,d}\|\sigma\|_{C^{l}(\mathbb{R})}\cdot\max_{j=0,\ldots,l}\|\Gamma\|^{j}_{C^{l}\left(\overline{O}\times\overline{O}\right)}\cdot\|U_{k}\|_{C^{l}\left(\overline{O}\times\overline{O}\right)}\cdot\varepsilon_{k}\log\left(\frac{1}{\varepsilon_{k}}+\pi+1\right).

Hence, for all kNlk\geq N-l, we demand

cl,k,dUkCl(O¯×O¯)εklog(1εk+π+1)2k,\displaystyle c_{l,k,d}\|U_{k}\|_{C^{l}\left(\overline{O}\times\overline{O}\right)}\cdot\varepsilon_{k}\log\left(\frac{1}{\varepsilon_{k}}+\pi+1\right)\leq 2^{-k}, (C.20)

so the k.k. summand can be estimated by σCl()2kmaxj=0,,lΓCl(O¯×O¯)j\frac{\|\sigma\|_{C^{l}\left(\mathbb{R}\right)}}{2^{k}}\cdot\max_{j=0,\ldots,l}\|\Gamma\|_{C^{l}\left(\overline{O}\times\overline{O}\right)}^{j} and (C.18) converges in Cl(O¯×O¯,EE)C^{l}\left(\overline{O}\times\overline{O},E^{*}\boxtimes E\right). Note that for each kk, we impose only finitely many conditions on εk\varepsilon_{k}, namely one for each lkNl\leq k-N, which are satisfied by some positive number. Hence, for each kk, there is a sufficiently small number εk>0\varepsilon_{k}>0 such that (C.20) is fulfilled for all lkNl\leq k-N.
Since all summands are smooth on (O¯×O¯)\Γ1(0)\left(\overline{O}\times\overline{O}\right)\backslash\Gamma^{-1}(0) and the series converges in all ClC^{l}-norms, it defines a smooth section on (O¯×O¯)\Γ1(0)\left(\overline{O}\times\overline{O}\right)\backslash\Gamma^{-1}(0). ∎

Thus, we showed that (3.20) yield well-defined distributions with singular support on the light cone, i.e. property (i). Furthermore, (iii) follows from Proposition 3.3 (6). We proceed with (ii):

Lemma C.3.

The sequence (εk)kN\big{(}\varepsilon_{k}\big{)}_{k\geq N} can be chosen such that

P(2)~±(p)=δp+K±(p,)P_{(2)}\widetilde{\mathscr{L}}_{\pm}(p)=\delta_{p}+K_{\pm}(p,\cdot)

for some K±C(O¯×O¯,EE)K_{\pm}\in C^{\infty}\left(\overline{O}\times\overline{O},E^{*}\boxtimes E\right).

Proof.

Let σk:=σΓεkC(Ω×Ω)\sigma_{k}:=\sigma\circ\frac{\Gamma}{\varepsilon_{k}}\in C^{\infty}(\Omega\times\Omega), so U~k=σkUk\widetilde{U}_{k}=\sigma_{k}\cdot U_{k} for all kNk\geq N and suppσk{Γ(p,q)εk}\text{supp}\,\sigma_{k}\subset\big{\{}\Gamma(p,q)\leq\varepsilon_{k}\big{\}}. Due to Lemma 1.1.10 of [BGP2007], we can exchange PP with the sum, so the transport equations (3.12) imply

k=NP(2)σk(Uklog(Γ±i0)+Wk)L±Ω(2k+2)\displaystyle\sum_{k=N}^{\infty}P_{(2)}\leavevmode\nobreak\ \sigma_{k}\big{(}U_{k}\log(\Gamma\pm i0)+W_{k}\big{)}L^{\Omega}_{\pm}(2k+2)
=k=N((2)σk2grad(2)σk+σkP(2))(Uklog(Γ±i0)+Wk)L±Ω(2k+2)\displaystyle\hskip 85.35826pt=\sum_{k=N}^{\infty}\left(\raisebox{0.0pt}{\text{{$\Box$}}}_{(2)}\sigma_{k}-2\nabla_{\text{grad}_{\!(2)}\sigma_{k}}+\sigma_{k}P_{(2)}\right)\big{(}U_{k}\log(\Gamma\pm i0)+W_{k}\big{)}L^{\Omega}_{\pm}(2k+2)
=:Σ1+Σ2+k=NσkP(2)(Uklog(Γ±i0)+Wk)L±Ω(2k+2).\displaystyle\hskip 85.35826pt=:\Sigma_{1}+\Sigma_{2}+\sum_{k=N}^{\infty}\sigma_{k}P_{(2)}\big{(}U_{k}\log(\Gamma\pm i0)+W_{k}\big{)}L^{\Omega}_{\pm}(2k+2).

Recall that the transport equations (3.12) and (3.15) are derived from the requirement that P(2)P_{(2)} applied to (3.11) is a telescoping series, that is,

P(2)(Uklog(Γ±i0)+Wk)L±Ω(2k+2)=(log(Γ±i0)P(2)Uk+P(2)Wk)L±Ω(2k+2)(log(Γ±i0)P(2)Uk1+P(2)Wk1)L±Ω(2k).\displaystyle\begin{split}&P_{(2)}\big{(}U_{k}\log(\Gamma\pm i0)+W_{k}\big{)}L^{\Omega}_{\pm}(2k+2)\\[5.69054pt] &\hskip 14.22636pt=\big{(}\log(\Gamma\pm i0)\cdot P_{(2)}U_{k}+P_{(2)}W_{k}\big{)}L^{\Omega}_{\pm}(2k+2)-\big{(}\log(\Gamma\pm i0)\cdot P_{(2)}U_{k-1}+P_{(2)}W_{k-1}\big{)}L^{\Omega}_{\pm}(2k).\end{split}

Hence, the right hand side becomes

Σ1+Σ2σN(log(Γ±i0)P(2)UN1+P(2)WN1)L±Ω(2N)+Σ3\displaystyle\Sigma_{1}+\Sigma_{2}-\sigma_{N}\big{(}\log(\Gamma\pm i0)\cdot P_{(2)}U_{N-1}+P_{(2)}W_{N-1}\big{)}L^{\Omega}_{\pm}(2N)+\Sigma_{3}

with

Σ3:=k=N(σkσk+1)(log(Γ±i0)P(2)Uk+P(2)Wk)L±Ω(2k+2).\Sigma_{3}:=\sum_{k=N}^{\infty}(\sigma_{k}-\sigma_{k+1})\big{(}\log(\Gamma\pm i0)\cdot P_{(2)}U_{k}+P_{(2)}W_{k}\big{)}L^{\Omega}_{\pm}(2k+2).

Then, for K±(p,):=P(2)~±(p)δpK_{\pm}(p,\cdot):=P_{(2)}\widetilde{\mathscr{L}}_{\pm}(p)-\delta_{p}, the transport equations (3.12) for UkU_{k} yield

K±=(1σN)(log(Γ±i0)P(2)UN1+P(2)WN1)L±Ω(2N)+Σ1+Σ2+Σ3.\displaystyle K_{\pm}=(1-\sigma_{N})\big{(}\log(\Gamma\pm i0)\cdot P_{(2)}U_{N-1}+P_{(2)}W_{N-1}\big{)}L^{\Omega}_{\pm}(2N)+\Sigma_{1}+\Sigma_{2}+\Sigma_{3}. (C.21)

On the right hand side, every summand individually yields a smooth section, since both 1σN1-\sigma_{N} and σkσk+1\sigma_{k}-\sigma_{k+1} as well as all derivatives of σk\sigma_{k} vanish in a neighborhood of Γ1(0)\Gamma^{-1}(0), which contains the singular support of ~±\widetilde{\mathscr{L}}_{\pm}. Thus, K±K_{\pm} vanishes on Γ1(0)\Gamma^{-1}(0) to arbitrary order and it remains to show convergence in all ClC^{l}-norms, which again for the WW-part is provided by the proof of Lemma 2.4.3 of [BGP2007]. Therefore, we concentrate on

k=N((2)σk2grad(2)σk+σkP(2))Uklog(Γ±i0)L±Ω(2k+2)=:Σ1+Σ2+Σ3.\sum_{k=N}^{\infty}\left(\raisebox{0.0pt}{\text{{$\Box$}}}_{(2)}\sigma_{k}-2\nabla_{\text{grad}_{\!(2)}\sigma_{k}}+\sigma_{k}P_{(2)}\right)U_{k}\log(\Gamma\pm i0)\leavevmode\nobreak\ L^{\Omega}_{\pm}(2k+2)=:\Sigma_{1}^{\prime}+\Sigma_{2}^{\prime}+\Sigma_{3}^{\prime}.

For fixed l0l\in\mathbb{N}_{0}, let k2(l+1)+Nk\geq 2(l+1)+N and Sk:={εk2|Γ(p,q)|εk}S_{k}:=\left\{\frac{\varepsilon_{k}}{2}\leq\left|\Gamma(p,q)\right|\leq\varepsilon_{k}\right\}. Then Lemma 1.1.12 of [BGP2007] implies for the k.k. summand of Σ2\Sigma_{2}^{\prime}

grad(2)σkUklog(Γ±i0)L±Ω(2k+2)Cl(O¯×O¯)=grad(2)σkUklog(Γ±i0)L±Ω(2k+2)Cl(O¯×O¯Sk)\displaystyle\left\|\nabla_{\text{grad}_{\!(2)}\sigma_{k}}U_{k}\log\big{(}\Gamma\pm i0\big{)}\leavevmode\nobreak\ L^{\Omega}_{\pm}(2k+2)\right\|_{C^{l}\left(\overline{O}\times\overline{O}\right)}=\left\|\nabla_{\text{grad}_{\!(2)}\sigma_{k}}U_{k}\log\big{(}\Gamma\pm i0\big{)}\leavevmode\nobreak\ L^{\Omega}_{\pm}(2k+2)\right\|_{C^{l}\left(\overline{O}\times\overline{O}\cap S_{k}\right)}
cl,dUkCl+1(O¯×O¯Sk)σkCl+1(O¯×O¯Sk)Γkd22log(Γ±i0)Cl+1(O¯×O¯Sk)\displaystyle\hskip 14.22636pt\leq c_{l,d}\|U_{k}\|_{C^{l+1}\left(\overline{O}\times\overline{O}\cap S_{k}\right)}\cdot\|\sigma_{k}\|_{C^{l+1}\left(\overline{O}\times\overline{O}\cap S_{k}\right)}\cdot\left\|\Gamma^{k-\frac{d-2}{2}}\log\big{(}\Gamma\pm i0\big{)}\right\|_{C^{l+1}\left(\overline{O}\times\overline{O}\cap S_{k}\right)}
cl,dUkCl+1(O¯×O¯Sk)σCl+1()εkl+1ttkd22logtCl+1([εk2,εk])maxj=0,,l+1ΓCl+1(O¯×O¯Sk)2j\displaystyle\hskip 14.22636pt\leq c_{l,d}\|U_{k}\|_{C^{l+1}\left(\overline{O}\times\overline{O}\cap S_{k}\right)}\cdot\frac{\|\sigma\|_{C^{l+1}(\mathbb{R})}}{\varepsilon_{k}^{l+1}}\cdot\left\|t\mapsto t^{k-\frac{d-2}{2}}\cdot\log t\right\|_{C^{l+1}\left(\left[\frac{\varepsilon_{k}}{2},\varepsilon_{k}\right]\right)}\cdot\max_{j=0,\cdots,l+1}\|\Gamma\|^{2j}_{C^{l+1}\left(\overline{O}\times\overline{O}\cap S_{k}\right)}
cl,k,dUkCl+1(O¯×O¯Sk)σCl+1()εkl+1maxj=0,,l+1ΓCl+1(O¯×O¯Sk)2jmaxt([εk2,εk])|t|kd22(l+1)(|logt|+1)\displaystyle\hskip 14.22636pt\leq c_{l,k,d}\|U_{k}\|_{C^{l+1}\left(\overline{O}\times\overline{O}\cap S_{k}\right)}\frac{\|\sigma\|_{C^{l+1}(\mathbb{R})}}{\varepsilon_{k}^{l+1}}\max_{j=0,\cdots,l+1}\|\Gamma\|^{2j}_{C^{l+1}\left(\overline{O}\times\overline{O}\cap S_{k}\right)}\max_{t\in\left(\left[\frac{\varepsilon_{k}}{2},\varepsilon_{k}\right]\right)}|t|^{k-\frac{d-2}{2}-(l+1)}\left(|\log t|+1\right)
cl,k,dUkCl+1(O¯×O¯)εk(log1εk+π+1)σCl+1()maxj=0,,l+1ΓCl+1(O¯×O¯)2j,\displaystyle\hskip 14.22636pt\leq c_{l,k,d}\|U_{k}\|_{C^{l+1}\left(\overline{O}\times\overline{O}\right)}\cdot\varepsilon_{k}\left(\log\frac{1}{\varepsilon_{k}}+\pi+1\right)\|\sigma\|_{C^{l+1}(\mathbb{R})}\cdot\max_{j=0,\cdots,l+1}\|\Gamma\|^{2j}_{C^{l+1}\left(\overline{O}\times\overline{O}\right)},

so we additionally demand

cl,k,dUkCl+1(O¯×O¯)εk(log1εk+π+1)2k.c_{l,k,d}\cdot\left\|U_{k}\right\|_{C^{l+1}\left(\overline{O}\times\overline{O}\right)}\cdot\varepsilon_{k}\left(\log\frac{1}{\varepsilon_{k}}+\pi+1\right)\leq 2^{-k}.

Then, for all ll, the ClC^{l}-norm of almost all summands (without the first 2(l+1)+N2(l+1)+N) of Σ2\Sigma_{2}^{\prime} is bounded by 2kσCl+1()maxj=0,,l+1ΓCl+1(O¯×O¯)2j2^{-k}\cdot\|\sigma\|_{C^{l+1}(\mathbb{R})}\cdot\max_{j=0,\cdots,l+1}\|\Gamma\|^{2j}_{C^{l+1}\left(\overline{O}\times\overline{O}\right)} and thus, we have convergence in ClC^{l} for all ll, i.e. Σ2\Sigma_{2}^{\prime} defines a smooth section in EEE^{*}\boxtimes E over O¯×O¯\overline{O}\times\overline{O} .
The treatment for Σ1\Sigma_{1}^{\prime} is completely identical, so we directly turn to the k.k. summand of Σ3\Sigma_{3}^{\prime}:

(σkσk+1)L±Ω(2k+2,)log(Γ±i0)P(2)UkCl(O¯×O¯)\displaystyle\left\|(\sigma_{k}-\sigma_{k+1})L^{\Omega}_{\pm}(2k+2,\cdot)\log\left(\Gamma\pm i0\right)\cdot P_{(2)}U_{k}\right\|_{C^{l}\left(\overline{O}\times\overline{O}\right)}
cl,k,d(σkΓl+1log(Γ±i0)Cl+σk+1Γl+1log(Γ±i0)Cl)ΓkNlClP(2)UkCl.\displaystyle\hskip 28.45274pt\leq c_{l,k,d}\left(\left\|\sigma_{k}\leavevmode\nobreak\ \Gamma^{l+1}\log\left(\Gamma\pm i0\right)\right\|_{C^{l}}+\left\|\sigma_{k+1}\leavevmode\nobreak\ \Gamma^{l+1}\log\left(\Gamma\pm i0\right)\right\|_{C^{l}}\right)\left\|\Gamma^{k-N-l}\right\|_{C^{l}}\cdot\left\|P_{(2)}U_{k}\right\|_{C^{l}}.

Set ρkl(t):=σ(tεk)tl+1logt\rho_{kl}(t):=\sigma\left(\frac{t}{\varepsilon_{k}}\right)\cdot t^{l+1}\log t, so we have σkΓl+1log(Γ±i0)=ρklΓ\sigma_{k}\leavevmode\nobreak\ \Gamma^{l+1}\log\left(\Gamma\pm i0\right)=\rho_{kl}\circ\Gamma. Then again Lemma 1.1.12 of [BGP2007] and Lemma C.19 yield

σkΓl+1log(Γ±i0)Cl(O¯×O¯)\displaystyle\left\|\sigma_{k}\leavevmode\nobreak\ \Gamma^{l+1}\log\left(\Gamma\pm i0\right)\right\|_{C^{l}\left(\overline{O}\times\overline{O}\right)} clρklCl()maxj=0,,lΓCl(O¯×O¯)\displaystyle\leq c_{l}\cdot\|\rho_{kl}\|_{C^{l}(\mathbb{R})}\cdot\max_{j=0,\ldots,l}\|\Gamma\|_{C^{l}\left(\overline{O}\times\overline{O}\right)}
ck,lεk(log1εk+π+1)σCl()maxj=0,,lΓCl(O¯×O¯),\displaystyle\leq c_{k,l}\cdot\varepsilon_{k}\left(\log\frac{1}{\varepsilon_{k}}+\pi+1\right)\|\sigma\|_{C^{l}(\mathbb{R})}\cdot\max_{j=0,\ldots,l}\|\Gamma\|_{C^{l}\left(\overline{O}\times\overline{O}\right)},

so we obtain

Σ3Cl(O¯×O¯)cl,k,d(εk(log1εk+π+1)+εk+1(log1εk+1+π+1))σCl()\displaystyle\|\Sigma_{3}^{\prime}\|_{C^{l}\left(\overline{O}\times\overline{O}\right)}\leq c_{l,k,d}\left(\varepsilon_{k}\left(\log\frac{1}{\varepsilon_{k}}+\pi+1\right)+\varepsilon_{k+1}\left(\log\frac{1}{\varepsilon_{k+1}}+\pi+1\right)\right)\|\sigma\|_{C^{l}(\mathbb{R})}
maxj=0,,lΓCl(O¯×O¯)ΓkNlCl(O¯×O¯)P(2)UkCl(O¯×O¯).\displaystyle\hskip 170.71652pt\cdot\max_{j=0,\ldots,l}\|\Gamma\|_{C^{l}\left(\overline{O}\times\overline{O}\right)}\cdot\left\|\Gamma^{k-N-l}\right\|_{C^{l}\left(\overline{O}\times\overline{O}\right)}\cdot\left\|P_{(2)}U_{k}\right\|_{C^{l}\left(\overline{O}\times\overline{O}\right)}.

Hence, for all kN+lk\geq N+l we demand

cl,k,dεk(log1εk+π+1)P(2)UkCl(O¯×O¯)ΓkNlCl(O¯×O¯)2k1c_{l,k,d}\cdot\varepsilon_{k}\left(\log\frac{1}{\varepsilon_{k}}+\pi+1\right)\cdot\left\|P_{(2)}U_{k}\right\|_{C^{l}\left(\overline{O}\times\overline{O}\right)}\cdot\left\|\Gamma^{k-N-l}\right\|_{C^{l}\left(\overline{O}\times\overline{O}\right)}\leq 2^{-k-1}

as well as for all kN+l+1k\geq N+l+1 that

cl,k1,dεk(log1εk+π+1)P(2)Uk1Cl(O¯×O¯)ΓkNl1Cl(O¯×O¯)2k2.c_{l,k-1,d}\cdot\varepsilon_{k}\left(\log\frac{1}{\varepsilon_{k}}+\pi+1\right)\cdot\left\|P_{(2)}U_{k-1}\right\|_{C^{l}\left(\overline{O}\times\overline{O}\right)}\cdot\left\|\Gamma^{k-N-l-1}\right\|_{C^{l}\left(\overline{O}\times\overline{O}\right)}\leq 2^{-k-2}.

Then the ClC^{l}-norm of almost all summands of Σ3\Sigma_{3}^{\prime} (without the first l+Nl+N) is bounded by 2kσCl()maxj=0,,lΓCl(O¯×O¯)2^{-k}\cdot\|\sigma\|_{C^{l}(\mathbb{R})}\cdot\max_{j=0,\ldots,l}\|\Gamma\|_{C^{l}\left(\overline{O}\times\overline{O}\right)}, so the series converges in all ClC^{l}-norms and is therefore smooth. Note that for each kk we again added only finitely many conditions. ∎

Finally, we show that the εk\varepsilon_{k}’s can be chosen such that for all pO¯p\in\overline{O}, the parametrices ~±(p)\widetilde{\mathscr{L}}_{\pm}(p) are distributions of degree at most κd\kappa_{d}.

Lemma C.4.

There is a sequence (εk)kN(0,1](\varepsilon_{k})_{k\geq N}\subset(0,1], for which we find some C>0C>0 such that

|~±(p)[φ]|CφCκd(Ω),pO¯,φ𝒟(Ω,E).\displaystyle\left|\widetilde{\mathscr{L}}_{\pm}(p)[\varphi]\right|\leq C\cdot\|\varphi\|_{C^{\kappa_{d}}(\Omega)},\qquad p\in\overline{O},\leavevmode\nobreak\ \varphi\in\mathscr{D}\left(\Omega,E^{*}\right).

Furthermore, for fixed φ𝒟(Ω,E)\varphi\in\mathscr{D}\left(\Omega,E^{*}\right), the map p~±(p)[φ]p\mapsto\widetilde{\mathscr{L}}_{\pm}(p)[\varphi] is smooth.

Proof.

We show the claim only for the logarithmic part, i.e. f:=k=d22U~klog(Γ±i0)L±Ω(2k+2)f:=\sum^{\infty}_{k=\frac{d-2}{2}}\widetilde{U}_{k}\log(\Gamma\pm i0)L^{\Omega}_{\pm}(2k+2), since for the other two sums the proof of Lemma 2.4.4 of [BGP2007] applies identically. By Lemma C.2, we have fC0(O¯×O¯,EE)f\in C^{0}(\overline{O}\times\overline{O},E^{*}\boxtimes E) and thus,

|f(p)[φ]|fC0(O¯×O¯)vol(O¯)φC0(O¯×O¯)fCκd(O¯×O¯)vol(O¯)φCκd(O¯×O¯)|f(p)[\varphi]|\leq\|f\|_{C^{0}(\overline{O}\times\overline{O})}\cdot\text{vol}(\overline{O})\cdot\|\varphi\|_{C^{0}(\overline{O}\times\overline{O})}\leq\|f\|_{C^{\kappa_{d}}(\overline{O}\times\overline{O})}\cdot\text{vol}(\overline{O})\cdot\|\varphi\|_{C^{\kappa_{d}}(\overline{O}\times\overline{O})}

for all pOp\in O and φ𝒟(O,E)\varphi\in\mathscr{D}(O,E^{*}), so the constant can be chosen via C:=fCκd(O¯×O¯)vol(O¯)C:=\|f\|_{C^{\kappa_{d}}(\overline{O}\times\overline{O})}\cdot\text{vol}(\overline{O}).
Since Proposition 3.3 (6) directly applies also to log(Γp±i0)L±Ω(2k+2,p)\log(\Gamma_{p}\pm i0)L^{\Omega}_{\pm}(2k+2,p) and UkφU_{k}\varphi is smooth on O×OO\times O with supp(Upk)φ\text{supp}\,(U_{p}^{k})\varphi compact, for every kd22k\geq\frac{d-2}{2}, the map

pU~pklog(Γp±i0)L±Ω(2k+2,p)[φ],φ𝒟(O,E),p\longmapsto\widetilde{U}_{p}^{k}\log(\Gamma_{p}\pm i0)L^{\Omega}_{\pm}(2k+2,p)[\varphi],\qquad\varphi\in\mathscr{D}(O,E^{*}),

is smooth. Therefore, also k=d22l1U~pklog(Γp±i0)L±Ω(2k+2,p)[φ]\sum_{k=\frac{d-2}{2}}^{l-1}\widetilde{U}_{p}^{k}\log(\Gamma_{p}\pm i0)L^{\Omega}_{\pm}(2k+2,p)[\varphi] is smooth for all l>d2l>\frac{d}{2} and the remaining term k=lU~pklog(Γp±i0)L±Ω(2k+2,p)[φ]\sum_{k=l}^{\infty}\widetilde{U}_{p}^{k}\log(\Gamma_{p}\pm i0)L^{\Omega}_{\pm}(2k+2,p)[\varphi] is ClC^{l} by Lemma C.2. This holds for all ll and hence, pf(p)[φ]p\mapsto f(p)[\varphi] is smooth. ∎

Lemma C.4 shows properties (iv) and (v) of ~±\widetilde{\mathscr{L}}_{\pm}, so Proposition 3.9 is proved also for even dd.

References

  • [AAS2012] Afshordi, N., Aslanbeigi, S., Sorkin, R. D. J. - A distinguished vacuum state for a quantum field in a curved spacetime: formalism, features, and cosmology, J. High Energ. Phys.8, 137 (2012)
  • [Bär2015] Bär, C. - Green-hyperbolic operators on globally hyperbolic spacetimes, Comm. Math. Phys. Verlag: Springer Seiten: 1585-1615 Band: 333, no. 3 (2015)
  • [BF2009] Bär, C., Fredenhagen, K. (Editors) - Quantum Field Theory on Curved Spacetimes, Lecture Notes in Physics, Vol. 786, Springer Berlin Heidelberg 2009.
  • [BG2011] Bär, C., Ginoux, N. - Classical and Quantum Fields on Lorentzian Manifolds, Global Differential Geometry, Springer Proceedings in Math. 17, Springer-Verlag (2011), 359-400
  • [BGP2007] Bär, C., Ginoux, N., Pfäffle, F. - Wave Equations on Lorentzian Manifolds and Quantization, European Mathematical Society (2007)
  • [BS2016] Bär, C., Strohmaier, A. - A Rigorous Geometric Derivation of the Chiral Anomaly in Curved Backgrounds, Commun. Math., Phys. 347, 703 – 721 (2016)
  • [BS2019] Bär, C., Strohmaier, A. - An index theorem for Lorentzian manifolds with compact spacelike Cauchy boundary, American Journal of Mathematics 141(5) (2019), 1421–1455
  • [BTW2015] Bär, C., Tagne Wafo, R. - Initial value problems for wave equations on manifolds, Math. Phys. Anal. Geom. Verlag: Springer Seiten: Art.: 7 Band: 18, no. 1
  • [BD1984] Birrell, N. D., Davies, P. C. W. - Quantum fields in curved space, Second edition, Cambridge University Press, Cambridge, 1984
  • [BR2002] Bratteli, O., Robinson, D. W. - Operator Algebras and Quantum Statistical Mechanics I, Springer, Berlin Heidelberg (2002)
  • [BF2014] Brum, M., Fredenhagen, K. - ‘Vacuum-like’ Hadamard states for quantum fields on curved spacetimes, Class. Quantum Grav. 31 025024 (2014)
  • [BF2000] Brunetti, R., Fredenhagen, K. - Microlocal Analysis and Interacting Quantum Field Theories: Renormalization on Physical Backgrounds, Commun. Math. Phys. 208, 623 – 661 (2000)
  • [BFV2003] Brunetti, R., Fredenhagen, K., Verch, R. - The generally covariant locality principle – A new paradigm for local quantum field theory, Commun. Math. Phys. 237, 31 – 68 (2003)
  • [DFM2018] Dappiaggi, C., Ferreira, H. R. C., Marta, A. - Ground states of a Klein-Gordon field with Robin boundary conditions in global anti-de Sitter spacetime, Phys. Ref. D 98 025005 (2018)
  • [DNP2014] Dappiaggi, C., Nosari G., Pinamonti, N. - The Casimir effect from the point of view of algebraic quantum field theory, Math. Phys. Anal. Geom. 19, 12 (2016)
  • [Dav1975] Davies, P. C. R. - Scalar production in Schwarzschild and Rindler metrics, J. Phys. A: Math. Gen. 8 609–616 (1975)
  • [Dav1984] Davies, P. C. R. - Particles do not Exist, Quantum Theory of Gravity: Essays in Honor of the Sixtieth Birthday of B. S. deWitt, edited by S. M. Christensen (Adam Hiler, Bristol, UK, 1984)
  • [DF2008] Décanini, Y., Folacci, A. - Hadamard renormalization of the stress-energy tensor for a quantized scalar field in a general spacetime of arbitrary dimension, Phys. Ref. D 78 044025 (2008)
  • [Den1982] Dencker, N. - On the propagation of polarization sets for systems of real principal type, J. Funct. Anal. 46, 351 -– 372 (1982)
  • [DG2013] Dereziński, J., Gérard, C. - Mathematical Quantization and Quantum Fields, Cambridge Monographs in Mathematical Physics, Cambridge University Press, 2013
  • [Dim1980] Dimock, J. - Algebras of Local Observables on a Manifold, Comm. Math. Phys. 77 (1980), 219–228
  • [DH1972] Duistermaat, J. J., Hörmander, L. - Fourier Integral Operators II, Acta Mathematica 128, 183-269 (1972)
  • [DK2010] Duistermaat, J. J., Kolk, J. A. C. - Distributions - Theory and Applications, Springer Science+Business Media, LLC 2010
  • [Düt2019] Dütsch, M. - From Classical Field Theory to Perturbative Quantum Field Theory, Progress in Mathematical Physics, Springer International Publishing (2019)
  • [Few2018] Fewster, C. J. - The art of the state, Int. J. Mod. Phys. vol. D27 no.11 (2018)
  • [FV2012] Fewster, C. J., Verch, R. - On a recent construction of ‘vacuum-like’ quantum field states in curved spacetime, Class. Quantum Grav. 31 205017 (2012)
  • [FV2015] Fewster, C. J., Verch, R. - Algebraic quantum field theory in curved spacetimes, Advances in Algebraic Quantum Field Theory, Springer (2015)
  • [Fri1975] Friedlander, B. F. - The wave equation on a curved spacetime, Cambridge University Press, Cambridge-New York-Melbourne, 2010
  • [Ful1973] Fulling, S. A. - Nonuniqueness of Canonical Field Quantization in Riemannian Space-Time, Phys. Ref. D 7 2850–2862 (1973)
  • [FNW1981] Fulling, S. A., Narcowich, F., Wald, R. M. - Singularity Structure of the Two-Point-Function in Quantum Field Theory in Curved Spacetime. II, Ann. Phys. 136, 243 (1981)
  • [FSW1978] Fulling, S. A., Sweeney, M., Wald, R. M. - Singularity Structure of the Two-Point-Function in Quantum Field Theory in Curved Spacetime, Commun. Math. Phys 63, 257 (1978)
  • [Gar1964] Garabedian, P. R. - Partial Differential Equations. New York: Wiley, 1964
  • [GS1967] Gelfand, I. M., Schilow, G. E. - Verallgemeinerte Funktionen (Distributionen) I, VEB Deutscher Verlag der Wissenschaften Berlin (1967), Zweite Auflage
  • [Gér2019] Gérard, C. - Microlocal analysis of quantum fields on curved spacetimes, Lecture Notes, arXiv:1901.10175v1 [math.AP]
  • [GW2014] Gérard, C., Wrochna, M. - Construction of Hadamard states by pseudo-differential calculus, Comm. Math. Phys. 325 (2) (2014), 713–755
  • [GW2015] Gérard, C., Wrochna, M. - Hadamard states for the Linearized Yang-Mills Equation on Curved Spacetime, Commun. Math. Phys. 337, 253-320 (2015)
  • [GOW2017] Gérard, C., Oulghazi, O., Wrochna, M. - Hadamard states for the Klein-Gordon equation on Lorentzian manifolds of bounded geometry, Commun. Math. Phys. 352, 519-583 (2017)
  • [Gün1988] Günther, P. - Huygens’ principle and hyperbolic equations, Academic Press, Boston, 1988
  • [HK1964] Haag, R., Kastler, D. - An algebraic approach to quantum field theory, Journal of Mathematical Physics 5: 848–861 (1964)
  • [Hac2016] Hack, T.-P. - Cosmological Applications of Algebraic Quantum Field Theory in Curved Spacetimes, Springer (2016)
  • [Had1923] Hadamard, J. - Lectures on Cauchy’s Problem in Linear Partial Differential Equations, Yale University Press, New Haven (1923)
  • [Haw1975] Hawking, S. W. - Particle Creation by Black Holes, Commun. Math. Phys. 43, 199–220 (1975)
  • [Hol2001] Hollands, S. - The Hadamard Condition for Dirac Fields and Adiabatic States on Robertson-Walker Spacetimes, Commun. Math. Phys., 216, 635 – 661 (2001)
  • [HW2015] Hollands, S., Wald, R. M. - Quantum fields on curved spacetime, Phys. Rep., 574, 1 – 35 (2015)
  • [Hör1990] Hörmander, L. - The Analysis of Linear Partial Differential Operators I: Distribution Theory and Fourier Analysis, 2nd Edition, Springer-Verlag Berlin Heidelberg, 1990
  • [Jun1996] Junker, W. - Hadamard States, adiabatic vacua and the construction of physical states for scalar quantum fields on curved spacetime, Rev. Math. Phys. 8, (1996), 1091 – 1159
  • [Jun2002] Junker, W. - Erratum to "Hadamard States, adiabatic vacua and the construction of physical states for scalar quantum fields on curved spacetime", Rev. Math. Phys. 207 (2002), 511–517
  • [JS2002] Junker, W., Schrohe E. - Adiabatic Vacuum States on General Space-time Manifolds: Definition, Construction, and Physical Properties, Ann. Henri Poincaré, 3, 1113 – 1181 (2002)
  • [Kam2019] Kamiński, W. - Elementary proof of symmetry of the off-diagonal Seeley-deWitt (and related Hadamard) coefficients, arXiv:1904.03708v1 [math-ph] (2019)
  • [KW1991] Kay, B. S., Wald, R. M. - Theorems on the Uniqueness and Thermal Properties of Stationary, Nonsingular, Quasifree States on Spacetimes with a Bifurcate Killing Horizon,, Phys. Rep. 207, 49 (1991)
  • [Köh1995] Köhler, M. - The stress-energy tensor of a locally supersymmetric quantum field on a curved space-time, PhD-thesis, Hamburg 1995
  • [KP1992] Krantz, S. G., Parks, H. R. - A Primer of Real Analytic Functions, Birkhäuser Verlag (1992)
  • [Kra2000] Kratzert, K. - Singularity structure of the two point function of the free Dirac field on a globally hyperbolic spacetime, Ann. Phys. (Leipzig), 9, 475 – 498 (2000)
  • [Man1968] Manuceau, J. - C*-algèbre des Relations de Commutation, Ann. Inst. Henri Poincaré 2, 139 (1968)
  • [Met1954] Methée, P.-D. - Sur les distributions invariantes dans le groupe des rotations de Lorentz, Comment. Math. Helv. 28 (1954), 225 – 269
  • [Mor1999] Moretti, V. - Proof of the Symmetry of the Off-Diagonal Heat-Kernel and Hadamard’s Expansion Coefficients in General CC^{\infty} Riemannian Manifolds", Commun. Math. Phys. 208, 283 – 308 (1999)
  • [Mor2000] Moretti, V. - Proof of the Symmetry of the Off-Diagonal Hadamard/Seeley-deWitt’s Coefficients in CC^{\infty} Lorentzian Manifolds by a "Local Wick Rotation", Commun. Math. Phys. 212, 165 – 189 (2000)
  • [Osg1898] Osgood, W. F. - Note über analytische Functionen mehrerer Veränderlichen, Math. Ann. 52, p. 462 – 464 (1898)
  • [Rad1992] Radzikowski, M. J. - The Hadamard Condition and Kay’s Conjecture in (Axiomatic) Quantum Field Theory on Curved Spacetime, Ph. D. thesis, Princeton University (1992)
  • [Rad1996] Radzikowski, M. J. - Micro-local approach to the Hadamard condition in quantum field theory on curved space.time, Commun. Math. Phys. 179, 529 (1996)
  • [RS1975] Reed, S., Simon, B. - Methods of Modern Mathematical Physics vol. II: Fourier Analysis, Self-Adjointness, Academic Press, San Diego, 1975
  • [Rej2016] Rejzner, K. - Perturbative Algebraic Quantum Field Theory: an introduction for mathematicians, Mathematical Physics Studies, Springer (2016)
  • [San2010] Sanders, K. - Equivalence of the (generalised) Hadamard and microlocal spectrum condition for (generalised) free fields in curved spacetime, Commun. Math. Phys. 295, 485 – 501 (2010)
  • [SV2001] Sahlmann, H., Verch, R. - Microlocal spectrum condition and Hadamard form for vector-valued quantum fields in curved spacetime, Rev. Math. Phys. 13, 1203–1246 (2001)
  • [Ste2000] Steinmann, O. - Perturbative Quantum Electrodynamics and Axiomatic Field Theory, Berlin; New York. Springer, (c) 2000.
  • [SVW2002] Strohmaier, A., Verch, R., Wollenberg, M. - Microlocal analysis of quantum fields on curved spacetimes: Analytic wavefront sets and Reeh-Schlieder theorems, J.Math.Phys. 43 (2002) 5514-5530
  • [Tre1967] Treves, F. - Topological Vector Spaces, Distributions and Kernels, Academic Press Inc., 1967
  • [Unr1976] Unruh, W. G. - Notes on black-hole evaporation, Phys. Rev. D14, 870 (1976)
  • [Ver1994] Verch, R. - Local Definiteness, Primarity and Quasiequivalence of Quasifree Hadamard Quantum States in Curved Spacetime, Commun. Math. Phys. 160, 507–536 (1994)
  • [Wal1978] Wald, R. M. - Trace anomaly of a conformally invariant quantum field in curved spacetime, Phys. Rev. D17, 1477–1484 (1978)
  • [Wal1994] Wald, R. M. - Quantum Field Theory in Curved Spacetime and Black Hole Thermodynamics, University of Chicago Press (Chicago, 1994)
  • [War1983] Warner, F. W., Foundations of differentiable manifolds and Lie groups, Springer 2nd printing 1983