Hadamard states for bosonic quantum field theory on globally hyperbolic spacetimes
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 in the sense of Duistermaat-Hörmander [Rad1996, DH1972]. Inspired by the construction of the corresponding advanced and retarded Green operator as done in [BGP2007], we construct the remaining two Green operators locally in terms of Hadamard series. Afterwards, we provide the global construction of , 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 associated to spacetime regions 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 -algebras and injective -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 and morphisms , where
-
•
is a globally hyperbolic Lorentzian manifold,
-
•
is a finite-dimensional, real vector bundle over with a non-degenerate inner product,
-
•
is a formally self-adjoint, Green hyperbolic operator,
-
•
time-orientation preserving, isometric embedding with open and causally compatible,
-
•
is a fiberwise isometric vector bundle isomorphism over such that are related via , where the restriction of to .
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 (see [Bär2015] for a thorough discussion of these operators). For , Theorem 3.5 of [BG2011] provides the exact sequence
(1.1) |
and hence, it leads to a covariant functor into the category of symplectic vector spaces with objects essentially given by the solution space of the field equation
(1.2) |
represents test sections in , more precisely smooth sections with compact support, and those with only spacelike compact support, meaning that it is contained in the causal future and past of some compact subset of . The -product of test sections is induced by the non-degenerate inner product on .
For a bosonic quantum field theory, we take the -representation of , i.e. a pair consisting of a -algebra and a map from into the unitary elements of such that is generated as a -algebra by and the Weyl relations hold:
(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, 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 on , where can be thought of as the expectation value of the observable in the state . The induced GNS-representation provides the familiar framework of a state space with observables as bounded operators and a cyclic vector (see section 2.3 of [BR2002] for details). Hence, the selection of a distinct vacuum is shifted to that of an algebraic state . Particularly adapted to free quantum fields are the so-called quasifree states generated by
for some scalar product on , by which is determined up to unitary equivalence. For these states, the unitary operators constitute a strongly continuous one-parameter group and thus, field operators are given by the self-adjoint generators due to Stone’s theorem. Furthermore, there is a dense domain such that for all , so polynomials of field operators are well-defined on . Hence, the Weyl relations (1.3) imply the familiar canonical commutator relations and for all , the -point-function of the state 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
(1.4) |
so it reproduces and hence . Indeed, we have for odd , and for even, it is given by some polynomial in the elements of . Thinking of 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 to , the scalar product corresponds to a bidistribution with
(1.5) |
With regard to (1.2) and (1.4), a quasifree state is therefore determined by and .
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 , 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
(1.6) |
where
(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 , involving the GNS-representation induced by . Also for non-quasifree states, one can formulate (1.6) as a constraint on the whole -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 satisfies the Hadamard condition if and only if it is of the form
(1.8) |
with the distinguished parametrices in the sense of Duistermaat-Hörmander (Theorem 6.5.3 in [DH1972]). With the primed diagonal, they are characterized by
(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 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 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 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 -dimensional vector space with non-degenerate inner product of index , we adopt the notations and conventions of [BGP2007], that is, for instance, the signature and the squared distance . The two connected components of the set of timelike vectors then determine a time-orientation, where we define the elements of to be future (past) directed. Correspondingly, we set , whose non-zero elements we call "lightlike" and "causal", respectively. Leaving out "" means the union of both components, i.e. and similarly and . Non-causal vectors are referred to as "spacelike".
For a -dimensional time-oriented Lorentzian manifold and , we write and for the corresponding chronological/lightlike/causal future/past of . These sets comprise all points that can be reached from 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 , we define and similarly . For the definitions of different types of subsets of 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 are always assumed to be spacelike.
For some real or complex finite-dimensional vector bundle over , the spaces of -, -, -sections in as well as distributional sections with values in some finite-dimensional space , including their (singular) support, convergence, order etc., are defined as in section 1.1 of [BGP2007]. For the volume density induced by the Lorentzian metric, another vector bundle over and some linear differential operator, the formally transposed operator of is given by
If is equipped with a non-degenerate inner product , which induces the -product and the isomorphism , we call formally self-adjoint if for all , that is, . Furthermore, is a wave operator if its principal symbol is given by on 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 , i.e. if , and bidistributions given by and called Schwartz kernel of . It is represented by a distributional section in the bundle over , whose fibers we identify via
(2.1) |
Definition 2.1.
Let be a linear differential operator. A linear and sequentially continuous operator is called
-
•
left parametrix for if is smoothing,
-
•
right parametrix for if is smoothing,
-
•
two-sided parametrix or just parametrix for if is left and right parametrix for ,
-
•
Green operator for if .
A bidistribution with for all is called
-
•
parametrix for at if ,
-
•
fundamental solution for at if .
Note that is a parametrix for if and only if its Schwartz kernel is a parametrix for at all . Especially, it is a Green operator for if and only if is a fundamental solution for and for all and .
2.1 Cauchy problems
A crucial feature of globally hyperbolic Lorentzian manifolds is a well-posed Cauchy problem for wave operators on smooth sections, i.e. for any Cauchy hypersurface with normal field , which is timelike, and Cauchy data , the Cauchy problem
(2.5) |
has a unique solution , which is supported in and depends continuously on the data (see section 3.2 of [BGP2007] and chapter 3 of [BF2009]). For a vector bundle over some further globally hyperbolic Lorentzian manifold , recall (2.1) for the definition of the vector bundle over .
Theorem 2.2.
Let be globally hyperbolic Lorentzian manifolds with Cauchy hypersurfaces and unit normal fields . Furthermore, let denote linear differential operators of second order acting on smooth sections in vector bundles over , which admit well-posed Cauchy problems and only lightlike characteristic directions. Then, for all and with , there is a unique section solving
(2.12) |
Let and denote the subset of elements satisfying . Then the map , which sends the Cauchy data to the unique solution of (2.12), is a linear and continuous operator .
Proof.
For all and , the Cauchy problem
(2.16) |
has a unique solution depending smoothly on the data by well-posedness of (2.5). Thus, it remains to determine from and to show that then is automatically fulfilled. For all , we define smooth sections and as solutions of
(2.17) |
By adapting the proof of Proposition A.1 of [FNW1981], we obtain smooth sections in over and , respectively, and, following the same lines, depends smoothly on . Hence, we found some solving (2.16), which yields the initial data of a solution of (2.12):
Note that and commute because they act on different factors of . Therefore, (2.16) and (2.17) imply that and satisfy the same Cauchy problem:
that is due to well-posedness. Clearly, this solution is unique since trivial Cauchy data in (2.12) lead to trivial data in (2.17) and therefore in (2.16), which implies for all and hence . The proof of stability follows the same lines as for (2.5), that is Theorem 3.2.12 of [BGP2007]: Obviously, the map
is linear, injective and continuous, and since (2.12) admits a solution for each data in , the closed subset is contained in . Due to continuity of differential operators, the subspace is also closed, so we obtain a continuous and bijective map between Fréchet spaces, whose inverse is continuous by the open mapping theorem. ∎
It seems that this procedure directly generalizes to operators of order , for which derivatives up to order have to be provided as data. Furthermore, symmetry of the data is inherited by the solution:
Corollary 2.3.
Additionally, we investigate the propagation of a family of singular solutions from a neighborhood of to all of 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 by checking Hörmander’s criterion.
Theorem 2.4.
Let be a globally hyperbolic Lorentzian manifold, a Cauchy hypersurface, a real or complex vector bundle over and a wave operator. Furthermore, let be relatively compact, and for all , let have spacelike compact support and only lightlike singular directions. Moreover, we assume for fixed . Then the Cauchy problem
has a unique solution , which has spacelike compact support and provides a smooth section for each .
Proof.
Let be a Cauchy time function on such that (Theorem 1.3.13 of [BGP2007]). Therefore, the normal directions of are timelike and do not match the singular directions of , so and are well-defined and compactly supported distributions on for all due to Hörmander’s criterion ((8.2.3) of [Hör1990]). Recall that any compactly supported distribution lies in some Sobolev space (see e.g. (31.6) of [Tre1967]), and hence, and for some . Thus, for all , Corollary 14 of [BTW2015] provides a unique solution
where this intersection is commonly referred to as the space of finite -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 implies continuity of the map of distributions for all .
For a differential operator, let denote the distribution . Since and act on different factors, is linearly and continuously mapped to , that is, commutes with (see the proof of Proposition A.1 in [FNW1981]). In particular, the map
is continuous due to smoothness of the first arrow. This holds for all differential operators , which provides smoothness of for fixed . ∎
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 on , 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 . It is not hard to check that these objects are invariant under the special orthochronous Lorentz group and their singular support is cointained in the light cone . This makes them directly comparable outside meaning that we need to understand -invariant distributions supported on .
For , we consider the submersions , where . Then the pullback maps distributions on to -invariant distributions on and moreover establishes a close connection to the well-known classification of distributions on supported in :
Theorem 2.5 (Théorème 2 of [Met1954]).
For any pair with , there is a -invariant distribution given by . Conversely, for any -invariant , we find a pair with such that .
For that, it is important to note that the proof of Theorem 2.5 particularly demonstrates surjectivity of .
Theorem 2.6 (Théorème 1 of [Met1954]).
Any -invariant with is of the form with for only finitely many .
There are two immediate consequences: First, every -invariant distribution supported in has to be of the form
(2.20) |
where only finitely many coefficients are non-zero. Recall that and the push-forward along a submersion is given by integration along the fibers. This leads to the second consequence, which is that every -invariant measure supported on is of the form (section IX.8 of [RS1975]), where
(2.21) |
The most known examples of -invariant supported on are certain Riesz distributions (not all of them) and in fact, there are no other: For all with , they are defined as continuous functions via
(2.24) |
One directly checks holomorphicity in and calculates , which provides an extension as distributions to all of via . Moreover, as shown in section 13.2 of [DK2010], we have and
(2.25) |
Therefore, due to (2.20), every -invariant distribution supported on is given by a linear combination of Riesz distributions:
(2.26) |
The properties of the Riesz distributions combined with the uniqueness of reveal as the advanced and retarded fundamental solution for on . In order to identify , we introduce the functions for , where stands for the direction from which the branch cut along is approached. A meromorphic extension to all of as distributions is given by
(2.27) |
which is in fact holomorphic on all of , and we calculate (see section III.3 of [GS1967] for proofs). This already ensures all crucial properties:
Proposition 2.7.
The distributions (2.27) are symmetric and -invariant. Moreover, for all natural numbers , we have
(2.28) |
and for , fundamental solutions for are given by
(2.29) |
Proof.
Symmetry and -invariance is obvious for and thus on all of , it follows from meromorphicity and the identity theorem. Holomorphicity on ensures , which leads to
Particularly at , it implies
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 -invariant distribution on , which solves . The spectral condition then leads to the general form ([RS1975, Ste2000]) and we choose . Here denotes the Fourier transformation with used in the exponential, i.e. ordinary Fourier transformation in space and inverse Fourier transformation in time direction , and the corresponding inverse operation. The maps
(2.30) |
are analytic in the complex forward/backward tube , they are -invariant and related via . Hence, for each , we obtain for some , and one directly calculates in the distributional sense, that is
(2.31) |
Due to the inverse Cauchy Schwarz inequality, maps to , and hence for all . Let denote the square root of with appropriately chosen sign such that again -invariance yields . In particular, is mapped to
(2.32) |
which lets us evaluate the integrals and thus derive a local expression for :
(2.33) |
It is a well-known procedure to extract the advanced and retarded fundamental solution and :
We extract the causal propagator from the antisymmetric part of , which turns out to be supported only in and is a solution of d’Alembert’s equation as well since . Hörmander’s criterion allows us to multiply with the step function with respect to the time coordinate and integration by parts reveals . Therefore, the aforesaid fundamental solutions are given by
(2.34) |
Due to their support properties, they are unique, so we directly conclude and .
Following Radzikowski’s results, the remaining two fundamental solutions are given by
(2.35) |
for which one directly calculates and outside of . It follows that the distributions and are symmetric and -invariant solutions of d’Alembert’s equation. Moreover, their support is contained in and they are homogeneous of degree , 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:
(2.36) |
3 Local Construction
3.1 Families of Riesz-like distributions
We proceed with the local construction of Hadamard bidistributions in a setting as in Definition 1.1 with a wave operator. Inspired by the local construction of the advanced and retarded parametrices for in [Gün1988, BGP2007], we introduce families of distributions similar to the Riesz distributions (2.24) but containing instead, so for , we introduce the distributions
(3.1) |
Very similar to , we have leading to a holomorphic extension to all of as distributions, for which then, moreover, analogous relations hold:
Proposition 3.1.
For all , we have
-
(1)
,
-
(2)
,
-
(3)
,
-
(4)
,
-
(5)
,
-
(6)
if , then are distributions of order at most .
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
(5): Follows from for all . ∎
With regard to (5), we omit the "" if . A crucial property of the Riesz distributions is , which, due to (4), fails to be true for in the even-dimensional case. However, it holds for odd because in that case, leads to . In even dimensions, we need a further family that is "more singular at even numbers", so for , we introduce
(3.2) |
Indeed, the zeros of and the poles of the prefactor in (3.2) compensate, and hence, exist as distributions for all , i.e. are meromorphic functions with simple poles at for fixed . From the definition follows that (1), (2), (3), (6) of Proposition 3.1 also hold for and in addition, we have and at all non-pole-integers
(3.5) |
Remark 3.2.
Following section 1.4 of [BGP2007], we now transfer the families locally to . For and geodesically starshaped with respect to , let
denote the squared Lorentz distance to and the distortion function (section 1.3 of [BGP2007]), which is related to the van-Vleck-Morette-determinant via . On , we define
(3.6) |
which provides holomorphic maps and, analogously, meromorphic maps with simple poles at .
Proposition 3.3.
Let and be geodesically starshaped with respect to . Then, for all , we have:
-
(1)
For , the maps are continuous on and given by
(3.7) -
(2)
,
-
(3)
,
-
(4)
.
-
(5)
For , (3.6) yield distributions of order at most . Moreover, there is an open neighborhood of and some such that
-
(6)
Let be an open neighborhood of such that is geodesically starshaped with respect to all . Furthermore, let and such that is compact for all . Then .
-
(7)
For all , the map is holomorphic on .
On the domain of holomorphicity of , the statements (1) - (7) remain true, when we replace and by and .
-
(8)
For an odd integer, we have .
-
(9)
.
-
(10)
For all , we have and .
Proof.
Corollary 3.4.
Proposition 3.3 leads to the following relations:
-
•
For odd, we have .
-
•
For moreover convex, i.e. starshaped w.r.t. all , and , we have
(3.8) and similarly for .
3.2 The Hadamard series
Let be a real vector bundle over , geodesically starshaped with respect to some and 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 at by taking the following ansatz of a formal Hadamard series:
(3.11) |
with coefficients yet to be determined. For , we identify with -valued test functions (see section 2.1 of [BGP2007]), so is (formally) understood as a distribution mapping to the complexified fiber . Similar to the procedure in chapter 2 of [BGP2007], we determine by formally demanding , which, by imposing the initial condition and by means of (A.1) and Corollary B.1, leads to the transport equations
(3.12) | ||||
(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 , which is therefore free to choose. Hence, even if (3.11) converges, the requirement determines only up to smooth solutions with arising from different choices of .
Proposition 3.6.
Proof.
The transport equations (3.12) and for half-integer also (3.15) coincide with (2.3) of [BGP2007]. Therefore, and for also are the Hadamard coefficients given by (3.16) and (3.17) due to Proposition 2.3.1 of [BGP2007]. For integer , we can apply the same proof for with replaced by everywhere, for which the same procedure then leads to (3.17). ∎
Corollary 3.7.
For convex, equipped with a non-degenerate inner product and formally self-adjoint, we have symmetry of for all and, in case of symmetric , of for half-integer in the sense of (A.3).
Remark 3.8.
Note that in the odd-dimensional case, leads to for all , whereas, as a consequence of the coupling with , for even dimensions, we have , in general.
More remarkably, the ’s fail to be symmetric in the even-dimensional case, i.e. for integer , even if we chose 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 always denote a convex domain, i.e. geodesically starshaped with respect to all . 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 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 at rather than fundamental solutions. To be more precise, for , some sequence and with , we define
(3.20) |
where
(3.25) |
Proposition 3.9.
For any relatively compact domain and any smooth choice of , there is a sequence such that (3.20) yield well-defined distributions for all , and
-
(i)
,
-
(ii)
with ,
-
(iii)
for all ,
-
(iv)
they are of order at most ,
-
(v)
there is a constant such that for all and .
The proofs of Lemma 2.4.1 - 2.4.4 of [BGP2007] only employ smoothness of the Hadamard coefficients and if , so replacing there by proves the Proposition in the odd-dimensional case. Similarly, we obtain convergence in of the -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 as Schwartz kernels, we extract the corresponding operators
(3.26) | ||||
(3.27) |
which are bounded due to compactness of .
Let be equipped with some non-degenerate inner product and 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 in the sense of Duistermaat-Hörmander.
Corollary 3.10.
Let and be the operators (3.26) arising from two different choices of and . Then is a smoothing operator on .
Proof.
Note that in terms of the operators (3.26), (3.27), Proposition 3.9 (iii) reads with smoothing. Hence, are left parametrices for and due to formal self-adjointness of , they also provide right parametrices:
Proposition 3.11.
For formally self-adjoint, the operators define two-sided parametrices for .
Proof.
We just have to show that yield right parametrices. From the symmetry properties of and ((3.8) and Theorem A.7) directly follows
(3.28) |
for all and . This works analogously for the logarithmic and -terms in (3.20). Furthermore, the series involving the coefficients are given by convergent power series , which yield smooth sections in . Altogether, we obtain the decomposition with representing the symmetric -part, i.e. , and the smooth -part. Hence, is symmetric up to smoothing in the sense
(3.29) |
from which we directly deduce the claim:
If is formally self-adjoint, then so is the operator
(3.30) |
due to symmetry of its Schwartz kernel by (3.28).
So far, we found two-sided parametrices 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 by directly comparing the corresponding Hadamard series. More precisely, we confirm that 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 represent a Feynman and an anti-Feynman parametrix.
Proposition 3.12.
Proof.
The first two claims are immediate conclusions from being two-sided parametrices and being symmetric, so we proceed with the Hadamard singularity structure.
Let such that , and for even , let either or . Then
(3.33) |
due to (3.2) and Proposition 3.3. denote the Riesz distributions (A.10), which are considered as bidistributions in the canonical way. Define
(3.34) |
such that the Hadamard series (3.31) takes the form
(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 used in [SV2001] are related with via (see Remark 2.3.2 of [BGP2007]). In addition, with the notation , we find as defined in (A.17), and hence,
For all , we choose in (3.25), so the series in (3.37) truncated at coincide with given in Appendix A.1 of [SV2001]. The remainder term is then of regularity and corresponds to in Definition 5.1 of [SV2001].
It remains to identify the singular terms (3.34) with given by (5.3) in [SV2001] up to some global factor, which is in the odd- and in the even-dimensional case. Moreover, note that for the squared Lorentzian distance in the definition of , the convention is used, whereas in Appendix A.1, it is .
Let . By definition of and as well as (2.35), we have with Wightman’s solution for . Then (2.33) leads to
in the distributional sense with . Since for odd , this coincides with . Furthermore, one directly calculates away from . For with , we obtain . Similar to (2.32), we set , which yields for all , and hence,
On the other hand, since are homogeneous distributions of degree , (3.34) provides
Of course, both expressions have to be understood in the distributional sense. Since is diffeomorphic to for all , their difference corresponds to a -invariant distributions on Minkowski space , which is supported on the light cone and homogeneous of degree . Therefore, by the classification (2.26), it has to vanish everywhere and thus, Theorem 5.8 of [SV2001] ensures that is of Hadamard form in the sense of (1.6).
It remains to show (3.32). According to (3.34), we define
(3.38) |
such that for (3.32), we obtain the expression (3.37) with replaced by and it suffices to show smoothness of the bidistributions (3.38). For , this follows directly from the definitions of as pullbacks of along a diffeomorphism and (2.35). On the other hand, one directly calculates that is given by the constant . ∎
Since are determined merely up to smooth sections, without loss of generality, we regard (3.32) as the equality
(3.39) |
3.4 Local fundamental solutions and Hadamard bisolutions
It remains to construct actual local bisolutions for from the parametrices following the lines of section 2.4 of [BGP2007]. By Proposition 3.9 (ii), we have , and hence, fundamental solutions are obtained by inverting the operators . Indeed, if , that is, for chosen "small enough", (3.27) provides isomorphisms for all with bounded inverses given by the Neumann series
(3.40) |
This means that all -norms of the series exist, which follows from compactness of and smoothness of . 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.
More precisely, is admissible if there is a choice of and such that (3.41) holds for the corresponding . Lemma 2.4.8 of [BGP2007] shows that for admissible, the corresponding operators are isomorphisms with bounded inverses (3.40).
Proposition 3.14.
For any admissible , the operators
fulfill , and hence, the distributions given by
(3.42) |
yield fundamental solutions for at . Furthermore, are smoothing operators.
Proof.
From now on, let be always equipped with some non-degenerate inner product, formally self-adjoint and admissible.
Proposition 3.15.
The operators are smoothing.
Proof.
Note that . Since are bounded and has a smooth Schwartz kernel, they extend to bounded maps
Hence, are bounded and therefore smoothing. ∎
It follows that yield anti-Feynman and Feynman parametrices for on . Moreover, their Schwartz kernels determine a real-valued bidistribution via
(3.43) |
which has the right singularity structure and is a solution for in the second argument, meaning and .
Proposition 3.16.
Proof.
Since is admissible, we obtain fundamental solutions at each , and furthermore, (3.43) provides for all . Moreover, as a causal subdomain of a globally hyperbolic Lorentzian manifold, is globally hyperbolic on its own right (Lemma A.5.8 of [BGP2007]). Hence, for a Cauchy hypersurface of with unit normal field , there is a unique smooth solution of
By continuous dependence on the Cauchy data, defines an -valued distribution for all . Furthermore, for all since it satisfies the trivial Cauchy problem.
It remains to check the wave front set, i.e. smoothness of . Since and yield parametrices for , the sections given by , and are smooth, and hence, is the solution of a Cauchy problem with smooth Cauchy data, which is smooth by Theorem 2.2.
∎
Altogether, any choice of parametrices in the sense of Proposition 3.9 leads to a bisolution with singularity structure given by in the sense of (1.9), so we constructed bisolutions with the Hadamard singularity structure on every :
Theorem 3.17.
Let be a globally hyperbolic Lorentzian manifold, an admissible domain, a real vector bundle with non-degenerate inner product and a formally self-adjoint wave operator. For the bisolution given by Proposition 3.16 and the advanced and retarded Green operator on , the bisolution is of Hadamard form.
Proof.
Due to smoothness of and , we obtain smoothness of , so Proposition 3.12 ensures the Hadamard property of . ∎
Remark 3.18.
For and , let denote the formal Hadamard series (3.11) with all terms removed and . Then Lemma 2.4.2 of [BGP2007] with replaced by and Lemma C.2 for the logarithmic part show that represents a -section over . This and the corresponding result for (Proposition 2.5.1 in [BGP2007]) ensure that 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 , and in this final section, the construction of global bisolutions , which locally coincide with those up to smooth bisolutions and thus inherit their singularity structure, is tackled. Assuming to be Riemannian and the validity of Theorem 6.6.2 of [DH1972] for sections in , we furthermore prove the existence of a smooth bisolution such that is symmetric and positive, i.e. a Hadamard two-point-function.
4.1 Global construction of symmetric bisolutions
For globally hyperbolic, we fix a Cauchy hypersurface and two locally finite covers , of it by admissible subsets of with if and only if . Without loss of generality, we assume that is a Cauchy hypersurface of for all . Then yields a causal normal neighborhood of in the sense of Lemma 2.2 of [KW1991]. By paracompactness of and the Hopf-Rinow-Theorem, we find an exhaustion of by finite subsets such that the relatively compact sets exhaust and every compact subset of is contained in some . Besides that, causality of implies and therefore,
(4.1) |
where stands for the Cauchy development of the respective set. It follows that every inextendible causal curve in meets exactly ones, so is a Cauchy hypersurface of , i.e. is globally hyperbolic. In addition, for all , we choose the corresponding local bisolutions obtained by Theorem 3.17 such that .
Proposition 4.1.
For each , there is a bisolution on satisfying .
Proof.
Let such that , and with . Then is a well-defined distribution with spacelike compact support on for all , since . With regard to Theorem 2.4, we define as the unique solution of
(4.5) |
which moreover depends smoothly on in the sense for fixed . Furthermore, global hyperbolicity of ensures by Theorem 2.2, since the difference solves the trivial Cauchy problem on .
Let and hence for all . It follows that , and if , which leads to . Consequently, it satisfies the trivial Cauchy problem, so we have , that is, represents a bisolution.
∎
This definition of is independent of the choice of in an appropriate sense: Let be another cut-off with and corresponding bisolution . Then is a bisolution with Cauchy data on given by . Recall that , so causality of yields for all , and hence, is contained in . Since satisfies the trivial Cauchy problem on , i.e. , it is a smooth bisolution by Theorem 2.2. Therefore, and differ merely by some smooth bisolution.
Next, we prove the existence of a compatible choice of bisolutions , meaning that they coincide on the overlaps . In this way, these compatible bisolutions assemble to a well-defined object on . 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 , let denote the sheaf given by the germs of the smooth sections in (see Example 5.2 in [War1983]). For the open cover of , the -simplices correspond to the non-empty ()-times intersections
with faces obtained by leaving out one in the intersection, respectively. An -cochain is a map that assigns to each non-empty a section of over , which we identify with the elements of . The space of -cochains is denoted by , where if , and the coboundary operator is defined by
It follows that for all and we set . These modules are trivial for all by some well-known construction (e.g. p. 202 in [War1983]), employing that admits a partition of unity subordinate to the locally finite cover :
Lemma 4.2.
For all , we have
Proof.
By choice of , the cover is locally finite. Let denote a partition of unity subordinate to and . Then, for each , the smooth section is supported in , and thus, via extension by zero, we consider it as an element of . In this way, we obtain homomorphisms via
which satisfy
Hence, implies , that is, . ∎
Lemma 4.3.
For all , there is a bisolution such that
Proof.
For , we consider the bisolution . For all , Proposition 3.9 provides parametrices on the relative compact domains such that for , Propositions 3.16 and 4.1 yield
(4.6) |
Such exist for all and exhausts , so we have smoothness on . Furthermore, as is causal and a neighborhood of a Cauchy hypersurface, fulfills a Cauchy problem with smooth Cauchy data and hence is smooth on all of by Theorem 2.2.
Therefore, recalling the identification of sections of with smooth sections in , the map represents a Čech-1-cochain, which moreover is a cocycle, since
for all . Thus, Lemma 4.2 ensures the existence of such that , and hence,
Recall that is a Cauchy hypersurface of for all and thus, each determines a bisolution via Theorem 2.2. On the other hand, due to causality of , we have a well-posed Cauchy problem on , and consequently, , since their Cauchy data coincide. This proves the claim:
For a partition of unity subordinate to , a well-defined bisolution on is given via
(4.7) |
Since a locally finite cover, for each , 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 , we directly read off from (4.7) that
(4.8) |
Hence, two different constructions of such a bisolution on differ only by a smooth bisolution.
Proposition 4.4.
There is a bisolution such that
(4.9) |
Proof.
Let be the bisolution on defined by (4.7) and recall that is an open neighborhood of . For all , we define as the unique solution of
This yields a smooth section, which leads to a bisolution since , and hence, solves the trivial Cauchy problem. Furthermore, we have , so (4.9) follows from (4.8) and Proposition 4.1. ∎
Corollary 4.5.
There is a smooth bisolution such that
Proof.
Theorem 4.6.
Let be a globally hyperbolic Lorentzian manifold, a real vector bundle with non-degenerate inner product over and a formally self-adjoint wave operator. Furthermore, let denote the advanced and retarded Green operator for and the symmetric bisolution given by Proposition 4.4 and Corollary 4.5. Then
(4.10) |
is a Hadamard bisolution, and a Feynman and an anti-Feynman Green operator for is determined by
(4.11) |
Proof.
For each , let be given as in (4.6). It follows that in the sense of (1.9) from Proposition 3.15, and moreover, we have
by Propositions 3.16 and 4.1 as well as (4.8). This holds for all and hence, is of Hadamard form in a causal normal neighborhood of due to Proposition 3.12. Therefore, 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 and thus stays unaffected when multiplying with smooth Hadamard coefficients. Hence, a Feynman and an anti-Feynman parametrix for are given by
which are even Green operators since are and is a bisolution. ∎
4.2 Positivity
It remains to show that can be chosen as a positive bisolution meaning that there is some smooth, symmetric bisolution such that for all . 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 be a smooth manifold, a real or complex vector bundle over with non-degenerate inner product and a properly supported pseudodifferential operator. For the definitions of being of real principal type in , pseudo-convexity of with respect to and the bicharacteristic relation of , 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 and , according to Theorem 6.5.3 of [DH1972], there are distinguished parametrices associated to the respective components of , where denotes the diagonal in . For a wave operator, they correspond to Feynman and anti-Feynman parametrices, respectively.
Definition 4.7.
Let be a smooth manifold, a real or complex vector bundle with non-degenerate inner product and a formally self-adjoint, properly supported pseudodifferential operator of real principal type in such that is pseudo-convex with respect to . Then is called of positive propagator type if there exists some such that the bidistribution satisfies
Note that 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 is of positive propagator type for the trivial line bundle . Note that the proof of this theorem employs positivity of for the directional derivatives on , and by applying certain operators, allowing one to keep track of the singularity structure of the corresponding parametrices, the general case is reduced to . 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 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 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 to be of positive propagator type, we need to show that can be actually chosen as a symmetric bisolution. It turns out that the existence of a pair , a well-posed Cauchy problem and
(4.12) |
is sufficient. For some Cauchy hypersurface of , the idea is to use as initial data on in order to determine a smooth bisolution via Theorem 2.2 and then following the lines of section 3.3 of [GW2015].
Let be the embedding map and the corresponding pullback to the initial data on , i.e.
(4.13) |
Clearly, is surjective and we have . Furthermore, for any differential operator with well-posed Cauchy problem, yields a bijection . The transposed map 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], corresponds to for any , and hence, is a map
denotes the distributions with wave front set contained in the closed cone , 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 if its wave front set does not contain the orthogonal directions mentioned above. Hence, for all closed cones with , (4.13) extends to a map
where contains the projections of onto . Let
(4.14) |
denote the inner product on with the induced volume density and the corresponding isomorphism . For Green-hyperbolic, the exact sequence (1.1) provides and thus a further bijection , which transfers to a Green operator on the space of initial data via
(4.15) |
It is not hard to deduce the Cauchy evolution operator mapping initial data to the solution of the corresponding homogeneous Cauchy problem. Moreover, [Dim1980] and, for the vector-valued case, [BS2019] provide the expression , from which uniqueness and surjectivity of lead to the particular expression .
Theorem 4.8.
Let be a globally hyperbolic Lorentzian manifold, a Riemannian vector bundle and 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 are given by (4.12) and that can be chosen as actual Green operators .
Then there is a real-valued and symmetric bisolution such that is smooth and
Proof.
The desired real-valued bisolution is given by
(4.16) |
and we show the claimed properties. With regard to Corollary 4.5 and without loss of generality, we assume to be symmetric, and furthermore, there is some such that by assumption on . Because this is also true for , we assume symmetry of as well, that is, for all .
Recall that Green operators map to , so for fixed , (4.16) provides a smooth section in . It follows that for each , we obtain a well-defined -valued distribution , which solves , and hence, exclusively contains lightlike directions by assumption on . By , the restriction of to yields a well-defined distribution on for any Cauchy hypersurface . This means that, due to Theorem 8.2.13 of [Hör1990], the operator associated to (4.16) can be applied to , , and for the result, we obtain
Since and , both contributions on the right hand side are empty. Hence, represents a map , so it follows that is smooth for fixed . With the adjoint operator , we eventually obtain a well-defined bidistribution via
(4.17) |
The bisection determines smooth and symmetric Cauchy data on and thus a smooth and symmetric bisolution by Theorem 2.2 (the first order analogon works completely similarly). Using the short-hand notation and , this yields for the corresponding bidistributions (4.17), and we show that positivity is preserved under the restriction to , i.e. for all . Theorem 8.2.3 of [Hör1990] provides a sequence such that in , and consequently, , so continuity of ensures
(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 , does not only descend to a well-defined bilinear form on by being a bisolution, but also to via
By following the lines of Proposition 3.9 of [GW2015] and employing , this allows us to trace back the claimed positivity property to (4.18). More precisely, for all , we have
which proves the theorem. ∎
In the case of formally self-adjoint wave operators, the existence of and is ensured by Theorem 4.6, so Theorem 4.8 leads to the final result:
Theorem 4.9.
Let be a globally hyperbolic Lorentzian manifold, a Riemannian vector bundle and a formally self-adjoint wave operator of positive propagator type. Then there exists a bidistribution such that
yields a Hadamard two-point-function, where denotes the advanced and retarded Green operator for . This means that has the Hadamard singularity structure (1.6) and satisfies
for all .
Moreover, a Feynman and an anti-Feynman Green operator are given by (4.11).
Note that, in general, 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 ’s in Lemma 4.3 and the 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 in a setting as in Definition 1.1 with 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 .
Let be a non-empty and convex domain, that is time-orientable, and let denote the -compatible connection on , meaning that
(A.1) |
It follows that for some uniquely determined endomorphism field and the connection-d’Alembert operator (see section 1.5 of [BGP2007]). Then the Hadamard coefficients for are defined as the unique solutions of the transport equations
(A.2) |
with and for all (Proposition 2.3.1 of [BGP2007]). Recall the identification (2.1) of as homomorphisms with fiberwise transposed operator . We are going to show symmetry in the sense
(A.3) |
One can see the Hadamard coefficients as a measure of the deviation of from , and indeed, "adding" does not change them: Let , over which we consider the same vector bundle , and . For the corresponding Hadamard coefficients , we obtain the initial condition and the transport equations
which are clearly solved by leading to
(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 . We directly conclude that, as a function of , the Levi-Civita connection on and, due to (A.1), the -compatible connection on 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 , the distortion function and the geodesic considered as a map
(A.5) |
Lemma A.1.
The -parallel transport along is analytic as a map
(A.6) |
Proof.
For fixed and , consider the parallel section in along for all , which therefore satisfies the ODE’s
(A.7) |
The columns of the corresponding fundamental matrix are given by linearly independent solutions of (A.7), so we have and the solution of (A.7) takes the form
From the definition of , we read off , and hence, the map is analytic. Moreover, implies
so is analytic for fixed and Osgood’s Lemma [Osg1898] proves the claim. ∎
Proposition A.2.
The map is analytic on for all .
Proof.
Analyticity of the zeroth Hadamard coefficient can be directly read off from
and we proceed via induction. By analyticity of , clearly is analytic if the prior coefficient is. Similarly, is analytic as a composition of analytic maps (recall that is positive). Therefore, the integrand of
is analytic in and uniformly continuous in on . 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 . ∎
Now the general case of smooth is tackled by analytic approximation of the coefficients, for which we quote
Proposition A.3 (Proposition 2.1 of [Mor1999]).
Let be a real, smooth and connected manifold with non-singular metric .
-
(a)
For any local chart of and any relatively compact domain with , there is a sequence of real and analytic (with respect to ) metrics with the same signature as , which are defined on some neighborhood of such that in , that is, all derivatives of converge uniformly on :
-
(b)
For any as in (a) and additionally any , there is an and a family of open neighborhoods of such that for any , and is a local base of the topology of . Moreover, for all , both and are common convex neighborhoods of for all metrics and .
Proposition A.4.
Let be relatively compact and a sequence of real and analytic metrics defined in a neighborhood of with the same signature as such that and are convex with respect to all and and in . For the corresponding Hadamard coefficients, we obtain for all and .
Proof.
The assumption directly provides , and with regard to the geodesic equation with converging right hand side, we similarly obtain as smooth maps on their domain of existence. Then the inverse function theorem provides as smooth maps on and, as a consequence, of the Lorentzian distance and the distortion function . Eventually, we have for the connecting geodesic (A.5).
It remains to investigate the parallel transport. For all , convergence of and leads to for the matrices defined in (A.7), and hence,
as smooth maps . Thus, we can directly conclude convergence of the zeroth Hadamard coefficient
(A.8) |
as smooth maps and, in particular, in .
We proceed inductively. Due to in , (A.8) implies in and consequently, . Therefore, the integrand in the expression of the first Hadamard coefficient
converges to the one in the expression of , and as a smooth function in , it is integrable on the compact interval . Hence, due to majorized convergence, the integral converges as well, so we have
which is . Recursively, this implies for all and . ∎
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 denote the Hadamard coefficients associated to . Employing formal self-adjointness of in the transport equations provides the relation
(A.9) |
According to section 1.4 of [BGP2007], we define
(A.10) |
which ensures the identification for . Due to Proposition 2.4.6 of [BGP2007], they comprise Hadamard series, which yield advanced and retarded parametrices for at each on any relatively compact domain . More precisely, for any integer and cut-off function with , there is a sequence such that
(A.13) |
represent well-defined distributions and . Furthermore, we have for fixed , so regarded as bidistributions and due to compactness of , they provide continuous operators
(A.14) |
Consequently, yield left parametrices for with , and for the advanced and retarded Green operator for , the differences 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, represent an advanced and a retarded parametrix for in the sense of Duistermaat-Hörmander.
Proposition A.5.
For all convex and relatively compact domains , the maps
(A.15) |
define smooth sections in over .
Proof.
By Lemma 3.4.4 of [BGP2007], the advanced and retarded Green operators for are given by , so formal self-adjointness of and uniqueness of lead to such that the operators are smoothing:
Lemma A.6.
Let and assume that for all quadruples as introduced in the beginning of the chapter with odd spacetime dimension and all lightlike related , we have
Then this equality holds for all and all .
Proof.
Consider the setting with odd spacetime dimension , let be causally related and choose such that . It follows that are lightlike related in and thus
by assumption. Therefore, Proposition A.4 provides .
Let be an analytic approximation of and the corresponding Hadamard coefficients. Write , which depends analytically on due to Proposition A.2 and vanishes on , so the identity theorem for analytic maps implies on all of . Furthermore, by Proposition A.4, we have and therefore for all , which proves the claim in the case of odd spacetime dimension.
For even-dimensional settings , this can be deduced from , which is odd-dimensional, and Proposition A.4.
Since this works for any relatively compact and convex domain , by uniqueness of the Hadamard coefficients, an appropriate exhaustion of by such subsets proves the claim on all of .
∎
Theorem A.7.
Let be a Lorentzian manifold of dimension , a real or complex vector bundle over with non-degenerate inner product, a formally self-adjoint wave operator and a convex domain. Then the Hadamard coefficients are symmetric in the sense
(A.16) |
Proof.
Let be odd. For all with , Lemma 1.4.2 (1) of [BGP2007] provides the recursion
(A.17) |
The proof of Lemma 2.4.2 in [BGP2007] shows that the sum defines a smooth section, which has to vanish for lightlike related since . Due to , this leads to
and it follows from Lemma A.6 that for , (A.16) is true also for even and on all of .
Now let again be odd, and for some , assume (A.16) to hold for all , i.e. the smooth section (A.15) is given by
Analogously, we obtain if , so again, applying Lemma A.6 completes the proof by induction. ∎
B Derivation of transport equations
Since Definition A.1 of the -compatible connection implies a product rule for , for odd , a straight forward calculation provides
For even , we need to relate the logarithmic and non-logarithmic parts, for which we first prove the following technical Lemma:
Lemma B.1.
Let be even and with . Then we have
and for
Proof.
This leads to the following calculation:
C Proof of Proposition 3.9 for even dimensional spacetimes
Note that in the proofs of all following Lemmas, denotes a generic constant, i.e. its particular value can change from one line to another.
Lemma C.1.
For all and , there is some such that for all , we have
Proof.
We start with calculating
Since for and due to , this yields
∎
Lemma C.2.
For any open and relatively compact domain and , there is a sequence such that for all the series
(C.18) |
converges in . In particular, for , this defines a continuous section over and a smooth section over .
Proof.
Since and even, is a smooth and a continuous section over , so every single summand of (C.18) is at least continuous, individually. Due to for all and by Lemma C.1, we have
Since for , we can choose such that
and (C.18) converges in . Now let and such that is of -regularity. Set , so by Lemma C.1 we have
(C.19) |
and Lemma 1.1.11 and 1.1.12 of [BGP2007] yield
Hence, for all , we demand
(C.20) |
so the summand can be estimated by and (C.18) converges in . Note that for each , we impose only finitely many conditions on , namely one for each , which are satisfied by some positive number. Hence, for each , there is a sufficiently small number such that (C.20) is fulfilled for all .
Since all summands are smooth on and the series converges in all -norms, it defines a smooth section on .
∎
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 can be chosen such that
for some .
Proof.
Let , so for all and . Due to Lemma 1.1.10 of [BGP2007], we can exchange with the sum, so the transport equations (3.12) imply
Recall that the transport equations (3.12) and (3.15) are derived from the requirement that applied to (3.11) is a telescoping series, that is,
Hence, the right hand side becomes
with
Then, for , the transport equations (3.12) for yield
(C.21) |
On the right hand side, every summand individually yields a smooth section, since both and as well as all derivatives of vanish in a neighborhood of , which contains the singular support of . Thus, vanishes on to arbitrary order and it remains to show convergence in all -norms, which again for the -part is provided by the proof of Lemma 2.4.3 of [BGP2007]. Therefore, we concentrate on
For fixed , let and . Then Lemma 1.1.12 of [BGP2007] implies for the summand of
so we additionally demand
Then, for all , the -norm of almost all summands (without the first ) of is bounded by and thus, we have convergence in for all , i.e. defines a smooth section in over .
The treatment for is completely identical, so we directly turn to the summand of :
Set , so we have . Then again Lemma 1.1.12 of [BGP2007] and Lemma C.19 yield
so we obtain
Hence, for all we demand
as well as for all that
Then the -norm of almost all summands of (without the first ) is bounded by , so the series converges in all -norms and is therefore smooth. Note that for each we again added only finitely many conditions. ∎
Finally, we show that the ’s can be chosen such that for all , the parametrices are distributions of degree at most .
Lemma C.4.
There is a sequence , for which we find some such that
Furthermore, for fixed , the map is smooth.
Proof.
We show the claim only for the logarithmic part, i.e. , since for the other two sums the proof of Lemma 2.4.4 of [BGP2007] applies identically. By Lemma C.2, we have and thus,
for all and , so the constant can be chosen via .
Since Proposition 3.3 (6) directly applies also to and is smooth on with compact, for every , the map
is smooth. Therefore, also is smooth for all and the remaining term is by Lemma C.2. This holds for all and hence, is smooth. ∎
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 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 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