<<

arXiv:1806.07471v1 [gr-qc] 19 Jun 2018 a unu il hoywt Classical with Quantum ) eateto hsc,NN,NreinUiest fSineand Science of University Norwegian NTNU, , of Department 1 3 2 b orsodn uhr mi drs:bo-sture.skagerstam@n address: Email author. Corresponding mi drs:[email protected] address: Email [email protected] address: Email ) oSueK Skagerstam K. Bo-Sture eateto pc,ErhadEvrnet hlesUniversit Chalmers Environment, and , of Department ia lcrccret h classical prese The the in current. derived electric were sical gauge Coulomb the in fields magnetic msinado bopinprocesses. micros absorption and and/or Einstein emission simpl of radiation a gravitational for quadrupole provides also framework quantum-mechanical -mome lin The classical retarded conserved a properly with and relativity causal general classical Co the a in of space- consideration emergence Minkowski these flat extend a around we deviations , quantum present the In emerged. ~ idpnetcua n eaddepcainvle fthe of values expectation retarded and causal -independent napeiu ok[]adi em fa xc quantum-mechan exact an of terms in and [1] work previous a In ore ierzdQuantum Linearized - Sources c ) od nvriyClee ..Bx21,N60 od,Norway Molde, N-6402 2110, Box P.O. College, University Molde E429 G¨oteborg, Sweden 96 SE-412 1 ,a rnhi,Norway Trondheim, ) , alEi Eriksson Karl-Erik ~ idpnetMxelseutoste naturally then equations Maxwell’s -independent Abstract tnu.no tmtno ste outlined. then is tensor ntum oi pnaeu spontaneous copic 2 aie lsia hoyof theory classical earized eodqatzdelectro- second-quantized ,b olna gravitational linear to s ) c facnevdclas- conserved a of nce praht classical to approach e , lm-iegue The gauge. ulomb-like e .Rekdal K. Per clframework, ical ehooy N-7491 Technology, fTechnology, of y 3 ,c ) 1. Introduction

In electro-dynamics it is natural to introduce gauge-dependent scalar and vector poten- tials. These potentials do not have to be local in space and time. It can then be a rather delicate issue to verify that gauge-independent obey the physical constraint of and that they also are properly retarded. Attention to this issue is often discussed in a classical framework. For instance one then shows in what manner various choices of gauge give rise to the same electro-magnetic field strengths (for recent discussions see, e.g., Refs. [2–10]). In a previous publication [1] we have shown that the time-evolution as dictated by quantum , for second-quantized electro-magnetic fields in the presence of a clas- sical conserved electric current, automatically solves these issues. The classical theory of Maxwell then naturally emerges in line with more general S-matrix arguments due to Wein- berg [11]. We also showed that various exact ~-independent radiative processes like the classical Vavilov-Cherenkovˇ radiation [12, 13] can be obtained in a straightforward man- ner [14–16]. In the present paper we extend the work of Ref. [1] to linear gravitational quantum deviations around a flat -time in a Coulomb-like gauge in the presence of a classical and conserved energy-momentum tensor. The time-evolution as dictated by , for a suitably defined second-quantized gravitational field, then auto- matically solves the issues of causality and retardation in a similar manner as in the case of electro-dynamics. It will, however, be argued that this is only true for certain propagating physical degrees of freedom and not for all degrees of freedom of the gravitational field as is sometimes claimed in the literature (see, e.g., Ref. [5], and Ref. [17] for various historical re- marks). The corresponding Coulomb-like degrees of freedom, like the ’s , are then neither retarded nor causal. In fact, they are shown to be instantaneous. The dynamical equations can, as in the case of electro-dynamics [1], be reduced to a system of decoupled harmonic oscillators with space-time dependent external . No pre-defined global causal order is assumed other than the deterministic time-evolution as prescribed by the Schr¨odinger equation. The classical theory of Einstein then naturally emerges in terms of expectation values of the second-quantized gravitational field for any initial again in line with more general S-matrix arguments due to Weinberg [11] and by Boulware and Deser [18]. The paper is organized as follows. In Section 2 we recall, for reasons of completeness, the classical version of Einstein’s theory of weak gravitational fields in vacuum in the presence of a space-time dependent conserved energy-momentum tensor, and a proper set of propagating degrees of freedom is constructed. The quantized degrees of freedom around a flat Minkowski

1 space-time in the presence of a conserved space-time dependent energy-momentum tensor is outlined in Section 3. In this section we also make some comments on the construction of a conserved energy-momentum tensor for these quantized gravitational degrees of freedom and the Weinberg-Witten theorem [19]. In Section 4 we consider the issues of causality and retardation in terms of quantum-mechanical averages of the second-quantized gravitational degrees of freedom and geodesic deviation. The emergence of the classical weak-field limit of Einstein’s general is then outlined. In Section 5 we also discuss, in a quantum-mechanical framework, Einstein’s classical gravitational quadrupole radiation as well as the microscopic spontaneous graviton emission process from a quantum source in terms of an excited hydrogen-like atom. In Section 6 finally, we present some conclusions and remarks. In an Appendix, we present some calculational techniques as made use of in obtaining decay rates for the emission of .

2. Weak Classical Gravitational Fields

We consider classical weak deviations from the flat Minkowski space-time in terms of the metric tensor, i.e., we write

gµν = ηµν + hµν , (2.1) where h 1 for all µ and ν. We will make use of the conventions that Greek indices run | µν|≪ from 1 to 4, with the space-time x-coordinate components xµ =(x1, x2, x3, x4) (x, x4 = ct) ≡ as well as ∂ = ∂/∂xµ, and the diagonal metric η has the signature (1, 1, 1, 1). The µ µν − Minkowski metric ηµν is used to raise or lower space-time Greek indices. When convenient, we will make use of the notation f(x) f(x, t) for space-time dependent fields. Apart from ≡ using the index 4 instead of 0 for time components, we follow the conventions of Ref. [20]. In general, the fundamental classical Einstein field equation takes the form 1 G R η R = κT , (2.2) µν ≡ µν − 2 µν − µν with κ =8πG/c4 . (2.3)

In the weak-field limit, the Riemann-Christoffel curvature tensor Rλµνκ is approximated by 1 ∂2h ∂2h ∂2h ∂2h R(1) = λν µν λκ + µκ . (2.4) λµνκ 2 ∂xκ∂xµ − ∂xκ∂xλ − ∂xµ∂xν ∂xν ∂xλ   λ λ In general, Rµν R µνλ as well as R R λ. The Gµν is therefore approxi- (1) ≡ ≡ mated by Gµν as defined by

2 (1) 1 α β αβ β α α β α β αβ Gµν = δµ δν + η ∂µ∂ν δν ∂µ∂ δµ ∂ν ∂ + ηµν ∂ ∂ η hαβ , (2.5) 2 − − − !    α 2 2 2 (1) µ (1) with  ∂ ∂α = ∂ /∂(ct) . Gµν is then such that ∂ Gµν = 0 and therefore we must ≡ ∇ − µ impose conservation of the energy-momentum tensor, i.e., ∂ Tµν = 0. The Riemann-Christoffel curvature tensor Eq.(2.4) has a local gauge invariance, i.e., it is invariant under infinitesimal coordinate transformations

x x′ = x ξ , (2.6) µ → µ µ − µ where we only consider ξµ and terms like ∂νξµ to first order. This is so since Eq.(2.6) and the tensor properties of gµν imply

h h′ = h + ∂ ξ + ∂ ξ . (2.7) µν → µν µν µ ν ν µ (1) It then follows that Gµν has the same invariance. For our purposes, we impose Lorentz- covariant equations of and regard (2.7) as a conventional local gauge-transformation of the gauge-field degrees of freedom hµν . It is now well-known that the components G(1) = κT of Einstein’s equation Eq.(2.2) µ4 − µ4 in the weak-field limit, required to have conservation of the energy-momentum tensor, i.e., µ ∂ Tµν = 0, contain at most time-derivatives of first order and can therefore be used as initial conditions for second-order (also see, e.g., Section 7 in Ref. [20] and for more recent discussions Refs. [21,22]). We therefore consider the components G(1) = κT µ4 − µ4 to be constraint equations similar to the constraint equation E = 2φ = ρ/ǫ in ∇ · −∇ 0 electro-dynamics in the Coulomb gauge (see, e.g., Ref. [1]). The components G(1) = κT µ4 − µ4 can therefore be used to eliminate degrees of freedom from the ten degrees of freedom of hµν to be clarified in more detail below. Instead of making use of an extension of the Helmholtz decomposition of vector fields (see, e.g., Ref. [1] for an elementary discussion and further references) to tensor fields, we find it convenient and more transparent to analyze the corresponding constraints by considering the Fourier transform of Einstein’s equation (2.2) in the weak-field limit. We therefore define

4 −ik·x Tµν (k)= d xe Tµν (x) , (2.8) Z and 4 −ik·x hµν (k)= d xe hµν (x) . (2.9) Z

3 Eqs.(2.2) and (2.5) therefore lead to

δαδβk2 + ηαβk k δβk kα δαk kβ + η kαkβ ηαβk2 h (k)=2κT (k) , (2.10) µ ν µ ν − ν µ − µ ν µν − αβ µν    2 2 2 (1) where k k k . The gauge-invariance of Gµν under the transformation (2.7) now allows ≡ − 4 us to introduce four Coulomb-like or radiation gauge conditions ∂ihiµ(x) = 0, i.e.,

kihiµ(k)=0 , (2.11) where Latin indices run from 1 to 3. With the gauge choice Eq.(2.11), we find from Eq.(2.10) the constraint equations 2 k hi4(k)+ kik4hll(k)=2κTi4(k) , (2.12) as well as k2h (k) k2h (k)= κT µ(k) . (2.13) 44 − ll µ The transversality gauge-condition (2.11), conservation of the energy-momentum tensor, i.e., µν 2 kµT (k) = 0, as well as Eq.(2.12), then imply that k k4hll(k)=2κkiTi4(k)=2κk4T44(k). We therefore obtain 2 k hll(k)=2κT44(k) . (2.14) By making use of Eq.(2.14), we can therefore write Eq.(2.12) in the following form where the transversality of hi4(k) is explicit:

2 k hi4(k)=2κPijTj4(k) . (2.15)

Here Pij is given by P δ kˆ kˆ = ǫ∗(k; λ)ǫ (k; λ) , (2.16) ij ≡ ij − i j i j λ X expressed, for later purposes, in terms of polarization vectors ǫ(k; λ), with λ = for ± complex-valued circular polarization and with λ = 1, 2 for real-valued linear polarization, obeying the transversality condition kˆ ǫ(k; λ) = 0, and where we have defined the unit · vector kˆ k/ k . The circular polarization degrees of freedom obey the rule ǫ( k; ) = ≡ | | − ± ǫ∗(k; ). In the case of linear polarization the unit vectors ǫ(k; λ) are such that ǫ( k; λ)= ± − ( 1)λ+1ǫ(k; λ) and − 1 ǫ(k; )= ǫ(k;1) iǫ(k;2) . (2.17) ± √2 ± Conservation of the energy-momentum tensor allows us, further more, to rewrite the con- straint equation (2.13) in the form

1 k2h (k)=2κ P T (k) T µ(k) . (2.18) 44 lm lm − 2 µ   4 2 2 In the Newtonian limit, Eq.(2.18) reduces to k h44(k) = κT44(k)=8πGρ(k)/c in terms of the density ρ(k), i.e., h (x) = 2Φ(x)/c2, as it should, where Φ(x) is the classical 44 − Newtonian gravitational potential such that 2Φ(x)=4πGρ(x). ∇ The components hi4(k) and h44(k) are therefore fixed by the constraint equations (2.15)

and (2.18). The spatial trace hll(k) is, in addition, determined by these constraints according to Eq.(2.14). When transformed back to space-time coordinates these constraints will only involve spatial derivatives evaluated at equal time. The constraint equation G(1) = κT µ4 − µ4 has therefore been solved for. Here we remark that the Coulomb-like degrees of freedom

hi4(x), h44(x), and hll(x) are not retarded and, in fact, they are instantaneous in time.

The dynamical Fourier transformed wave-equation for the components hij(k) that follows from Eq.(2.10) can conveniently now be written in the following form:

1 k2 h (k) P h (k) =2κP T (k) . (2.19) ij − 2 ij ll ij;lm lm  

Here we have defined the tensor projection Pij;lm by

1 P P P + P P P P , (2.20) ij;lm ≡ 2 il jm jl im − ij lm  

such that Pij;lmPlm;kn = Pij;kn. Both sides of Eq.(2.19) are now by construction transverse, symmetric, and traceless in the indices ij. The gauge-invariant combination 1 χ (k) h (k) P h (k) , (2.21) ij ≡ ij − 2 ij ll then contains two-degrees of freedom corresponding to the degrees of freedom of an on- shell -two massless field, i.e., a propagating gravitational field. In terms of a formal inverse Fourier transform δT (δ ∂ ∂ / 2) of the projection operator P , as defined by ij ≡ ij − i j ∇ ij T T Eq.(2.16), the transverse, traceless, and gauge-invariant spatial components hij (x, t) of the

tensor hµν (x, t) can now be written in the well known form (see, e.g., Ref. [23])

T T T T hij (x, t)= δij;klhkl(x, t) . (2.22)

Here we have defined the projection operator 1 δT T δT δT δT δT , (2.23) ij;kl ≡ ik jl − 2 ij kl which in Eq.(2.22), of course, can be expressed in terms of explicit non-local space- T T (see, e.g., Ref. [1]). It then follows that hij (x, t) = χij(x, t). Similarly, apart from a factor 2κ, the inverse Fourier transform of the right-hand side of Eq.(2.19) corresponds to T T T T the transverse and traceless part Tij (x, t) = δij;klTkl of the tensor Tij(x, t). At this point

5 one remarks that Eq.(2.19) does not mathematically prescribe the space-time causal and/or

retardation properties of χij(x, t) as, e.g., discussed by Rohrlich [5]. In Section 4 we will, however, verify that quantum mechanics leads to the correct causal and retarded behaviour

of suitable expectation values of a second-quantized version of the field χij(x, t).

3. The Second-Quantized

A second-quantized version of the components χij(x) in Eq.(2.21) takes a conventional form, i.e.,

~ 2 2 κc ik·x ∗ ∗ −ik·x χij(x)= ǫij(k; λ)akλ e + ǫij(k; λ)akλ e , (3.1) Vωk kλ s X   with λ = 1, 2 or λ = for gravitons with k x = k x ω t using k = ω /c = k , ± · · − k 4 k | | analogous to the second-quantized transverse electro-magnetic potential AT (x) for in the Coulomb gauge (see, e.g., Ref. [1]). The presence of the factor 2κ is similar to

the appearance of the factor 1/2ε0 in the second-quantized electro-magnetic field potential

AT (x) and also makes χij(x) dimensionless. An overall numerical factor is then determined from, e.g., a suitable canonical commutation relation (see, e.g., Refs. [11, 24, 25]) or, as will be verified below, from consistency with the classical Einstein field equation in terms of

expectation values. It is assumed that the free-field Hamiltonian H0 used in the definition of the interaction picture takes the form

~ ∗ H0 = ωk(akλakλ +1/2) . (3.2) kλ X The components of the two, in general complex-valued, polarization tensors ǫij(k; λ) in Eq.(3.1) are normalized in such a way that

∗ ′ ∗ ′ Tr ǫ (k; λ)ǫ(k; λ ) ǫ (k; λ)ǫ (k; λ )= δ ′ . (3.3) ≡ kl lk λλ k,l   X An explicit representation in terms of the real-valued, symmetric and transverse linear polarization unit vectors ǫ(k; λ), with λ =1, 2, in Section 2 is 1 ǫij(k;1) = ǫi(k;1)ǫj(k;1) ǫi(k;2)ǫj(k;2) , (3.4) √2 −  and 1 ǫij(k;2) = ǫi(k;1)ǫj(k;2)+ ǫi(k;2)ǫj(k;1) . (3.5) √2  With, e.g., k = (0, 0,k) then the matrix elements of ǫij(k; λ) are given by ǫ11(k;1) = ǫ (k;1) = 1/√2, ǫ (k;2) = ǫ (k;2) = 1/√2, and ǫ (k; λ) = 0 for i = 1, 2, 3 [11]. − 22 12 21 i3 6 By construction ǫ (k; λ) obey the rule ǫ ( k; λ)=( 1)λ+1ǫ (k; λ). In passing we notice ij ij − − ij that the tensor of a classical metric deviation χij(x) has been reported in a recent and remarkable precise test in terms of observed classical gravitational waves from the binary coalescence GW170814 [26]. In terms of the normalization Eq.(3.3), we

remark that the tensor projection operator Pij;lm as defined by Eq.(2.20) can be obtained as follows P = P (kˆ) ǫ∗ (k; λ)ǫ (k; λ) . (3.6) ij;lm ij;lm ≡ ij lm λ , X=1 2 If we express the quantum field χij(x) in terms of complex circular polarization we can make use of real-valued transverse photon linear polarization unit vectors ǫ(k; λ) and Eq.(2.17). We can then write 1 ǫij(k; ) ǫi(k; )ǫj(k; )= (ǫij(k;1) iǫij(k;2)) , (3.7) ± ≡ ± ± √2 ±

∗ ∗ such that ǫ (k; )= ǫ ( k; ). The corresponding creation operators ak± are given by ij ± ij − ±

∗ 1 ∗ ∗ ak± = (ak iak ) . (3.8) √2 1 ± 2 We then observe that, by construction,

∗ ∗ ∗ ǫij(k; λ)akλ = ǫij(k; λ)akλ , (3.9) λ , λ ± X=1 2 X=

which simply states that the quantum field χij(x) in Eq.(3.1) does not depend on the actual realization of the choice of polarization degrees of freedom. Under a rotation around the k-axis with an angle θ one now readily finds that ak± exp( i2θ)ak± due to the rotation → ± properties of the polarization tensors ǫij(k; λ), for λ = 1, 2. From this we conclude that ∗ single graviton states k ak± 0 carry intrinsic helicities 2~. | ±i ≡ | i ± In addition to the free field Hamiltonian H0 in Eq.(3.2), and analogous to the second- of the electro-magnetic field in the Coulomb gauge, we can now construct a momentum operator P and a helicity operator Σ according to

∗ P ~kak ak , (3.10) ≡ λ λ kλ X and ~ ˆ ∗ ∗ Σ 2 k ak+ak+ ak−ak− . (3.11) ≡ k − X  The complete set of commuting operators H0, P , and Σ are now such that diagonal one- particle states k carry all the appropriate quantum numbers of a spin-two particle, i.e., | ±i

7 a graviton. A complete set of physical Fock-states can then be generated in a conventional manner. In addition to the intrinsic spin angular momentum Σ, photon as well as graviton states can also carry conventional orbital angular momentum L, which for photons plays an important role in many current contexts (see, e.g., Ref. [27] and references cited therein), but will not be of concern in the present paper. For our purposes, the Eqs.(3.2),(3.10), and (3.11) are given by construction. Nevertheless, a gauge-invariant energy-momentum tensor for the gravitational field χij(x) can be found, as first discussed in various forms and in the context of classical gravitational waves by Einstein [28], and is defined by (also see, e.g., Refs. [23, 25, 29–33] for other considerations)

1 1 t (x) ∂ χ (x)∂ χ (x) η ∂αχ (x)∂ χ (x) . (3.12) µν ≡ 4κ µ ij ν ij − 2 µν ij α ij   At the classical level, one procedure to motivate Eq.(3.12) is, e.g., to consider the Einstein (2) tensor Gµν to second-order in the deviation hµν , denoted by Gµν . An effective energy- (2) momentum tensor for the gravitational degrees of freedom can then be defined by Gµν /κ − (see, e.g., Ref. [31], Section 7.6, and references cited therein), a procedure also used in various studies of low energy effective [34]. By a straightforward calculation in the

gauge Eq.(2.11), and by considering the source-free limit Tµν 0, we then obtain a gauge- (grav) → invariant energy-momentum tensor tµν for the free gravitational field χij(x), i.e.,

1 1 t(grav)(x) χ (x)∂ ∂ χ (x) η χ (x)∂α∂ χ (x) . (3.13) µν ≡ −2κ ij µ ν ij − 2 µν ij α ij  

With the normalization of second-quantized graviton field χij(x) with two polarization de- grees of freedom in Eq.(3.1), an additional factor 1/2 must then be included in Eq.(3.13). When performing space-integration over the quantization volume V one verifies that Eqs.(3.12) and (3.13) then leads to the same result. Conventional conservation of energy-momentum can now be expressed in terms of expectation values of the effective and conserved energy- momentum tensor T + t . µν h µν i The energy-momentum tensor tµν in Eq.(3.12) is, by construction, now such that the quantized gravitational field χij in Eq.(3.1) leads to the correct diagonal free-field Hamilto- nian, i.e., 3 H0 = d xt44(x, t) , (3.14) ZV analogous to the case of the free second-quantized transverse electro-magnetic field AT . A momentum operator P can also be obtain from Eq.(3.12) in terms of t = (t41, t42, t43) according to 1 P = d3x t(x, t) . (3.15) c ZV 8 The orbital angular momentum is then given by L = d3xx t(x, t)/c. The helicity operator × Σ can also be expressed in terms of the of quantumR field χij by taking the properties of χij under rotations into account in a standard manner, but will not be of importance in the present context. The total angular momentum is then given J = L + Σ analogous to the case of the second-quantized radiation field AT (x). We observe that in many classical expositions one neglect the second term in the energy- momentum Eq.(3.12) and performs an averaging procedure over several characteristic wave- lengths of gravitational radiation (see, e.g., Refs. [22, 23, 30–32]). We, however, follow a conventional second-quantized field theory procedure making use of Eq.(3.12) in order to verify, e.g., Eq.(3.14). Furthermore, and according to the Weinberg-Witten theorem [19], massless particles with helicity larger than one cannot carry a Lorentz-covariant energy- momentum tensor. As a theorem this appears to contradict the construction above. The theorem follows from a consideration of the rotational properties of matrix elements like k t k around the k-axis. Due to the properties of the polarization tensor ǫ (k; λ) h− ±| µν| ±i ij in Eq.(3.1) it, however, follows from an explicit calculation of such on-shell matrix elements, making use of the definition of tµν according to Eq.(3.12) and Eq.(3.1) for the quantum field

χij(x), that such matrix elements actually vanish identically and in this sense the Weinberg- Witten theorem is therefore avoided.

4. The Causality Issue

The well-known interaction picture Hamiltonian for a classical current j(x, t) interaction with the second-quantized and transverse electro-magnetic field AT (x, t) takes an analogous form for the gravitational field χij(x, t) (see, e.g., Ref. [11, 25]), i.e.,

1 H (t)= d3xT (x, t)χ (x, t) I −2 ij ij ZV ~ 2 κc −iωkt ∗ ∗ iωkt ∗ = ǫij(k; λ)akλ e Tij(k, t)+ ǫij(k; λ)akλ e Tij(k, t) , (4.1) − 2Vωk kλ s X   where we make use of the spatial Fourier transform

T (k, t) d3x eik·x T (x, t) . (4.2) ij ≡ ij ZV The total Hamiltonian of the system will, in general, also involve instantaneous act-at-a- distance c-number valued Newtonian-like terms (see, e.g., Ref. [25]) which, however, will not be of any concern in the present paper.

9 ∗ Due to the linear dependence of akλ and akλ in Eq.(4.1), it is now straightforward to exactly solve for the unitary (see, e.g., Refs. [35–38] for early discussions). Indeed, with ψ(t) exp(itH /~) ψ(t) , where we for convenience make the choice t =0 | iI ≡ 0 | i 0 for the initial time, the state ψ(t) is such that, | iI d ψ(t) i~ | iI = H (t) ψ(t) , (4.3) dt I | iI where for observables (t) exp(itH /~) exp( itH /~) , (4.4) OI ≡ 0 O − 0 and where, in our case, H is given by Eq.(3.14). The time-evolution for ψ(t) is then given 0 | iI by i i t ψ(t) = exp φ(t) exp dt′ H (t′) ψ(0) , (4.5) | iI ~ −~ I | i    Z0  for any initial pure state ψ(0) . The c-number phase φ(t), which will be of no importance | i in our considerations, can be computed according to

t t t′ 1 ′ ′ ′ 1 ′ ′′ ′′ ′ iφ(t)= ~ dt N(t ),HI(t ) = ~ dt dt HI (t ),HI (t ) , (4.6) 2 0 2 0 0 Z h i Z Z h i

with t ′ ′ N(t) dt HI(t ) , (4.7) ≡ 0 ′ ′ Z since [ N(t ),HI(t ) ] now is a c-number. Apart from the phase-factor φ(t), the time-evolution in the interaction picture can there- fore be expressed in terms of a conventional multi-mode displacement operator U(α) as (see, e.g., Refs. [39–41]) given by

t i ′ ′ ∗ ∗ U(α) exp dt H (t ) = exp αk (t)ak αk (t)ak , (4.8) ≡ −~ I λ λ − λ λ 0 kλ  Z  Y   where, in our case,

t ~ 2 ′ i κc ′ iωkt ∗ ∗ ′ αk (t) dt e ǫ (k; λ)T (k, t ) . (4.9) λ ≡ ~ 2Vω ij ij s k Z0 Since expectation values are independent of the picture used, i.e.,

(t) ψ(t) ψ(t) = ψ(t) (t) ψ(t) , (4.10) hO i ≡ Sh |O| iS I h |OI | iI and by upon acting with the displacement operator U(α) in Eq.(4.8) on, e.g., an initial vacuum state, we can now evaluate the exact expectation value of second-quantized field

10 χij(x, t) in a manner similar to the analysis for QED situation [1]. We then obtain the following ~-independent weak-field limit result

′ t sin ω (t t ) 1 ′ k χ (x, t) =2κc2 d3x′ P (kˆ)eik·(x−x ) dt′ − T (x′, t′) ij ij;lm   lm h i k V 0 ωk Z X Z t =2κc2 d3x′ dt′G (x x′, t t′)T TT(x′, t′) , (4.11) R − − ij Z Z0 3 3 in the large volume V limit, with k = V d k/(2π) , and making use of partial integra- x tions. Here GR( , t) is the retardedP Green’sR function 3 d k k·x sin(ωt) δ(t x /c) G (x, t) ei Θ(t)= −| | , (4.12) R ≡ (2π)3 ω 4πc2 x Z | | for t 0. Under time-reversal t t we remark that we can keep the expressions above ≥ → − with t t , i.e., we preserve time-reversal invariance provided that the source T (x, t) → | | ij is also time-reversed. In the derivation of Eq.(4.11) we have made use of Eq.(3.6) in the summation over the polarization degrees of freedom. A more explicitly form of Eq.(4.11) is the well-known causal and properly retarded expression

′ ′ 4G T TT(x , t x x /c) χ (x, t) = d3x′ ij −| − | . (4.13) h ij i c4 x x′ Z | − | The quantum mechanical expectation value χ (x, t) in Eq.(4.13) obeys the classical equa- h ij i tion of motion

 χ (x, t) = 2κT TT(x, t) . (4.14) h ij i − ij Eq.(4.14) is, of course, the Fourier transform of Einstein’s classical dynamical Eq.(2.19). If the initial state is changed from the vacuum state to an arbitrary pure state an additional contribution to Eq.(4.13) appears corresponding to a homogeneous solution of the wave equa- tion Eq.(2.19) for χ . Together with the constraint equations G(1) = κT , required to ij µ4 − µ4 have the energy-momentum tensor Tµν conserved, the complete set of the classical and lin- earized Einstein’s equations thus emerge from the quantum-mechanical framework outlined above analogous to the situation in QED [1].

The gauge-invariant Riemann-Christoffel tensor Rλµνκ(x, t) in the weak field limit ac- cording to Eq.(2.4) plays an important role in, e.g., the detection of gravitational waves (see, e.g., Refs. [20–23, 29–32, 42]). Its causal and retardation properties in the far-field limit of a sufficiently localize energy-momentum tensor Tµν(x, t) are therefore of importance. In the case of two test particles their spatial separation, i.e., the corresponding geodesic deviation,

11 is then determined by, e.g., the components Rl4m4(x, t) which we now express in terms of a space-time Fourier transform of Eq.(2.4), i.e., 1 R(1) (k)= k2h (k) k k h (k) k k h (k)+ k k h (k) . (4.15) l4m4 2 4 lm − l 4 m4 − m 4 l4 l m 44  In terms of hlm(k)= χlm(k)+ Plmhnn(k)/2 we obtain k2 κ k2 R(1) (k)= 4 χ (k)+ 4 P T (k) l4m4 2 lm k2 2 lm 44  1 k k P T (k)+ k P T (k) + k k P T (k) T µ(k) . (4.16) − 4 m lj j4 l mj j4 l m ij ij − 2 µ  where we have made use of the constraints Eqs.(2.14), (2.15), and(2.18). When inverting (1) 3 the Fourier transform Rl4m4(k) we make use of the fact that the integrals d xT4m(x, t) are time-independent for a sufficiently well localized energy-momentum tensorR Tµν (x, t). The (1) far-field limit of Rl4m4(x, t) is then obtained from Eq.(4.16) with the properly causal and retarded gauge-invariant result 1 ∂2 R(1) (x, t) = χ (x, t) . (4.17) h l4m4 i −2c2 ∂t2 h lm i For quantum fields in general, intrinsic quantum fluctuations are present [43, 44] and the related considerations for photons applies also for quantum states of gravitons. Since (1) the Riemann-Christoffel tensor Rl4m4(x, t) has now been promoted to an interaction-picture quantum-field, we remark that intrinsic quantum fluctuations will be associated with expec- tation values like in Eq.(4.17). As long as the equations of motion are linear averaging pro- cedures over a space-time resolution can be implemented as in the case of electro-magnetic interactions as discussed in, e.g., Ref. [1]. Since gravity couples to an effective conserved energy-momentum tensor non-linearities naturally emerge [11, 18, 34] and the correspond- ing averaging procedures require clarification which, however, goes beyond the scope of the present paper (in this context also see Ref. [45]) .

5. Gravitational Radiation Processes

The emission of soft photons and/or gravitons [24,25] can be investigated in terms of the formalism discussed above. The classical expression for gravitational quadrupole emission from a given, sufficiently well localized, source can, in addition, be obtained from the classical expression Eq.(4.13) in a standard manner (see, e.g., Refs. [20, 23, 29, 30, 32, 42]). We, however, notice the following. In view of diagonal form of the free-field Hamiltonian according to Eqs.(3.2) or (3.14), and the generated by the interaction Eq.(4.1),

12 it is now actually quite easy to find the time-dependence of the energy of the gravitons

emitted. We first recall that conservation of the energy-momentum tensor Tµν leads to the well-known identity

1 ∂2 d3xT (x, t)= d3xx x T (x, t) , (5.1) ij 2c2 ∂t2 i j 44 Z Z for a sufficiently well-localized source Tµν (x, t). By enforcing the dipole approximation

T (k, t) d3xeik·xT (x, t) d3xT (x, t) , (5.2) ij ≡ ij ≈ ij Z Z and after a summation over polarizations using Eq.(3.6), one finds, disregarding the divergent

vacuum contribution in H0 in Eq.(3.2), that

2 1 ~κc2 ~ kˆ E(t) I ψ(t) H0 ψ(t) I = ωk ~ Pij;lm( ) ≡ h | | i k s2Vωk ! X t t ′ ′′ dt′ dt′′eiωk(t −t )T (t′)T (t′′) , (5.3) × ij lm Zt0 Zt0 where used has been made of Eq.(4.9). Here we have defined

T (t) d3xT (x, t) , (5.4) ij ≡ ij Z which now can be expressed in terms of quadrupole moments according to Eq.(5.1). We now extend the time interval to ( , ), and define the Fourier transform D (ω) of the −∞ ∞ kl 3 2 2 d xxixkT44(x, t)/c in terms of the energy density T44(x, t) = c ρ(x, t) according to R ∞ dt iωt Dkl(ω) e Dkl(t) , (5.5) ≡ −∞ 2π Z where

D (t) d3xx x ρ(x, t) . (5.6) kl ≡ k l Z It then follows that 4πG ∞ E(t)= dωω6Q∗ (ω)Q (ω) . (5.7) 5c5 kl kl Z0 The angular average in Eq.(5.3) has than been obtained by making use of the result dΩ 2 P (kˆ)D (ω)D∗ (ω)= Q∗ (ω)Q (ω) , (5.8) 4π ij;lm ij lm 5 ij ij Z

13 expressed in terms of the gravitational quadrupole tensor Qkl(ω) as defined by 1 Q (ω) D (ω) δ D (ω) . (5.9) kl ≡ kl − 3 kl nn The total emitted by the classical source according to Eq.(5.7) agrees with well-known results (see, e.g., Ref. [20], Section 10). According to Eq.(5.3) we now also observe that

∞ t t 2 G ∂ ′ ′′ E(t)= dω dt′ dt′′Q¨ (t′)Q¨ (t′′) eiω(t −t ) , (5.10) 5πc5 ij ij ∂t′∂t′′ Z0 Zt0 Zt0 where the angular average has been obtained by making use of a similar expressions like

Eqs.(5.8) for Dkl(t) with 1 Q (t) D (t) δ D (t) . (5.11) kl ≡ kl − 3 kl nn If we now, e.g., imagine that the source is non-zero in the finite time-interval [0, t] par- tial integrations can be carried out in Eq.(5.10), and we then obtain the famous Einstein expression [28] quadrupole gravitational radiated power P (t) dE(t)/dt, i.e., ≡ ∞ t ...... ′ ...... G ′ ′ iω(t −t) G P (t)= 5 dω dt Qij(t)Qij(t )e = 5 Qij(t)Qij(t) . (5.12) 5πc −∞ 5c Z Zt0 If this power will be observed at a large distance R from the classical source, the time- parameter t in Eq.(5.12) should, of course, be replaced by the retarded time t R/c. − Furthermore, we can now also make use of the interaction Eq.(4.1) in order to evaluate the total decay rate Γ for the microscopic transition i f , with i = a 0 and | i→| i | i | ii⊗| i f = a kλ , of of a graviton with the energy ~ω from an | i | f i⊗| i k excited hydrogen atom in the initial atomic state a to the final atomic state a to first- | ii | f i order in time-dependent in a standard manner. We then make use of T (x, t)= m c2δ(3)(x x(t)) in Eq.(5.1), where x(t) is the of the at time 44 e − t in the interaction picture. The relevant matrix element for the spontaneous one-graviton transition from the hydrogen atomic state a = nlm = 3dm to the final atomic ground | ii | i | i state a = 1s is then given by | f i | i 2 meω 16πG~ f H (0) i = if ǫ∗ (k; λ) a x x a . (5.13) I 4c Vω lm f l m i h | | i r k h | | i Here ~ω E E = ~ω corresponds to energy conservation, and where, for the well if ≡ i − f k localized quantum atomic states, we make use of the dipole approximation k x 0. By · ≈ summing over all polarization degrees of freedom and directions of the emitted graviton,

14 making use of Eq.(3.6), the decay rate Γ, for large time-scales t, with finite ω ω t 1, | k − if | ≫ can then be obtained in a straightforward manner with the result (see Appendix A)

2π V 38 Gm2a4 ω5 Γ d3k δ(ω ω ) i H (0) f 2 = e B if , (5.14) ≡ ~2 (2π)3 k − if |h | I | i| 213 5~c5 λ Z X in the large volume V limit, independent of the energy-degeneracy m and where a = ~/αm c is the Bohr radius for a reduced mass m of the atomic hydrogen two- B e ≈ e body system. This result agrees with Ref. [46] and with the semi-classical result as reported in Ref. [47], but not with results as quoted in Refs. [20, 48]. If nkλ gravitons are present in the initial state, the rate Γ in Eq.(5.14) should be multiplied by the factor (1 + nkλ) corresponding to stimulated emission. Absorption graviton processes can be studied in a similar manner.

6. Final Remarks

A quantum-mechanical framework offers a platform to study causality and retardation issues in the classical theory of Maxwell [1] as well as in the weak-field limit of Einstein’s general theory of relativity. Second-quantization of physical degrees of freedom and current conservation for a classical source, leads to the well established classical theory for electro- magnetism in terms of expectation values of quantum fields. Inherent and unavoidable quantum fluctuations of these expectation values will always be present [43–45]. As we have now explicitly verified, the same reasoning applies for a second-quantized gravitational field around a flat metric in the presence of a sufficiently weak and conserved classical energy-momentum tensor. The weak-field limit of Einstein’s general theory of relativity than naturally emerges expressed also in terms of expectation values of quantum fields. The overwhelming experimental support for Maxwell’s classical theory does not neces- sarily imply the existence of photons as quantum states and doubts on the existence of such quantum states are sometimes put forward (see, e.g., Ref. [49]). However, it is clear that the quantum-mechanical derivation of the classical theory necessarily implies the existence of single particle quantum states corresponding to a photon. The recent experimental spec- tacular discovery of gravitational radiation [26, 50–54] is an additional verification of the correctness of Einstein’s general theory of relativity. Outstanding as such observations are, they do not necessarily prove the existence of gravitons as physical quantum states. In view of electro-magnetism and photons as quantum states, one would, however, encounter a fun- damental difficulty in understanding a possible physical non-existence of gravitons. This is so since the arguments we have presented concerning the emergence of the weak-field limit of Einstein’s classical theory of gravity from quantum mechanics, follow the same line of

15 reasoning as in the case of electro-magnetism [1]. Therefore, as for the existence of photons, they necessarily predict the existence of gravitons. Various arguments have, however, been put forward that, due to the weak coupling of gravitons to , it may, nevertheless, be impossible to detect single gravitons. In this context we observe that one may, in addition to the analysis of Dyson [55], also consider the possible detectability of quantum states of ∗ n gravitons in the form of, e.g., a Fock state n (ak ) 0 /√n! for a macroscopically large | i ≡ λ | i number n. Such quantum states have, of course, a trivial expectation value of the quantized gravitational field Eq.(3.1) but non-zero intrinsic quantum fluctuations. Related discussions on the existence of gravitons can be found in, e.g., Refs. [46, 56, 57]. We have also observed that various classical ~-independent expressions for weak pertur- bations around a flat space-time easily and naturally follow from the quantum-mechanical framework provided the classical source has a conserved energy-momentum tensor. In the weak-field limit such expressions are supposed to be exact results. We now observe that the power spectrum of gravitons emitted from a classical source, with a conserved energy- momentum tensor, is exactly obtained from first-order time-dependent perturbation theory as can be seen by inspection of Eq.(5.3). It may come as a surprise that a first-order quantum-mechanical calculation can give an exact ~-independent answer. An explanation of this, as it seems, remarkable fact can now be traced back to the factorization of the time- evolution operator in terms of a displacement operator for quantum states in the interaction picture according to Eq.(4.5) making use of Eq.(4.1). The phase φ(t) then contains the non- perturbative effects of all higher-order corrections to the first-order result. As a matter of fact, similar features are known to happen also in some other situations. As is well-known, the famous differential cross-section for Rutherford scattering can be obtained exactly by making use of the first-order Born approximation. All higher order corrections will then contribute with an overall phase for probability amplitudes which follows from the exact solution (see the excellent discussion in Ref. [58]). The classical Thomson cross-section for low-energy scattering on a charged particle is also exactly obtained from a Born approx- imation due to the existence of an exact low-energy theorem in quantum electro-dynamics (see, e.g., the discussion in Section 8 of Ref. [59]). As alluded to above, the weak coupling of gravitons to matter makes it perhaps of a more academic issue to discuss various features of quantum states for single graviton. Recently it has, however, been observed that well-aligned pairs of photons, with the same helicity quantum numbers, can emulate the quantum-mechanical properties of spin-two gravitons [60]. In such a sense one may even consider entangled pairs of such emulated gravitons and, e.g., carry out EPR-correlation experiments in terms of the CHSH Bell-like inequalities [61,62] in a similar manner as in the famous Aspect et al. [63] realization of entangled pairs

16 of photons. For photons the θ-angle dependence under rotations around the k-direction according to ak± exp( iθ)ak±, using the definition Eq.(3.8), will then be replaced by → ± θ 2θ for emulated gravitons. The corresponding quantum-mechanical CHSH Bell-like → correlations to be measured for entangled pairs of emulated gravitons can then easily be worked out simply by the replacement θ 2θ but will not be discussed in detail here [64]. → for gravitons can therefore, at least, be contemplated in terms of emulated gravitons. Finally, we speculate, that if the sources of the emulated gravitons [60] can be described by a conserved energy-momentum tensor, the theoretical arguments presented in this exposition should apply. The characteristic κ could then, of course, take a completely different value and thereby open up a possible window for experimental investigations of classical as well quantum-mechanical features of emulated gravitons in a completely different context.

17 Appendix

A. Decay Processes

The decay rate Γ is defined by

dΓ m2ω5 Γ dΩ = G e if Q, (A.1) ≡ dΩ ~c5 λ , Z X=1 2 where, for the one-graviton emission process of a hydrogen-like atom 3dm 0 1s | i⊗| i→| i ⊗ kλ , we have defined | i dΩ Q D∗ D ǫ (k; λ)ǫ∗ (k; λ) . (A.2) ≡ 4π ij kl ij kl λ , Z X=1 2

The matrix elements Dij are then obtained from the transition matrix elements Eq.(5.13) in the main text, i.e.,

D d3xψ∗ (x)x x ψ (x) . (A.3) ij ≡ 3dm i j 1s Z In the case with m = 0 the only non-zero components are then D /2 = D = D = 33 − 11 − 22 2 7 15 aB 3 /2 in terms of the Bohr radius aB as in Ref. [46]. It can be verified that Γ does not dependp on the quantum number m as must be the case. The polarization sum in Eq.(A.2) is now analogous to Eq.(5.8) in the main text where we now have defined the quadrupole

matrix elements Qkl by 1 Q D δ D , (A.4) kl ≡ kl − 3 kl nn in terms of Eq.(A.3). By combining Eqs.(A.1)-(A.4) the result Eq.(5.14) in the main text is obtained. Even though such a decay is most likely impossible to observe, this analysis shows that we have properly normalized the graviton quantum field χij in Eq.(3.1) since the same analysis for a purely classical source leads to the well-known Einstein as discussed in the main text.

18 ACKNOWLEDGMENT

B.-S. S. expresses his gratitude to the late J. D. Jackson for, in 2002, pointing out a sign error in an earlier publication (see Ref. [6]) as well as to the CERN laboratory, and in particular the head of the CERN Scientific Information Service J. Vigen, for providing hospitality during the final preparation of the present work. The research by B.-S. S. was supported in part by Molde University College and NTNU. K.-E. E gratefully acknowledges hospitality at Chalmers University of Technology. The research of P. K. R. was supported by Molde University College. The authors are also grateful for the hospitality provided for at Department of Space, Earth and Environment, Chalmers University of Technology, Sweden. REFERENCES

1. B.-S. K. Skagerstam, K. E. Eriksson, and P. K. Rekdal, “Causality in with Classical Sources - ”, arXiv:1801.09947v1 [quant-ph] 30 Jan 2018 (submitted for publication).

2. O. L. Brill and B. Goodman, “Causality in the Coulomb Gauge ”, Am. J. Phys. 35, 832-837 (1967).

3. C. W. Gardiner and P. D. Drummond, “Causality in the Coulomb gauge: A Direct Proof ”, Phys. Rev. A 38, 4897-4898 (1988).

4. J. D. Jackson, “Classical Electrodynamics ”, Third Edition, John Wiley and Sons (New York, 1998).

5. F. Rohrlich, “Causality, the Coulomb Field, and Newton’s Law of ”, Am. J. Phys. 70, 411-414 (2001).

6. J. D. Jackson, “From Lorentz to Coulomb and Other Explicit Gauge Transforma- tions ”, Am. J. Phys. 70, 917-928 (2002). As cited in this reference, a sign error in Eq.(9) in B.-S. Skagerstam, “A Note on the Poincar´eGauge ”, Am. J. Phys. 51, 1148-1149 (1983) was correctly observed by J. D. Jackson.

7. A. M. Stewart, “Vector Potential of the Coulomb Gauge ”, Eur. J. Phys. 24, 519-524 (2003).

8. V. Hnizdo, “Comment on ’Vector Potential of the Coulomb Gauge’ ”, Eur. J. Phys. 25, L21-L22 (2004).

9. J. A. Heras, “How Potentials in Different Gauges Yield the Same Retarded Electric and Magnetic Fields ”, Am. J. Phys. 75, 176-183 (2007).

19 10. R. Dickinson, J. Forshaw, and P. Millington, “Probabilities and Signaling in Quantum Field Theory ”, Phys. Rev. D 93, 065054-1-12(2016) and “Working Directly with Probabilities in Quantum Field Theory ”, in “8th International Workshop DICE2016: - Matter - Quantum Mechanics ”, IOP Conf. Series: Journal of Physics: Conf. Series 880 012041, 1-8 (2017).

11. S. Weinberg, “Photons and Gravitons in Perturbation Theory: Derivation of Maxwell’s and Einstein’s Equations ”, Phys. Rev. 138, B988-B1002 (1965).

12. P. A. Cherenkov,ˇ “The Visible Radiation of Pure Liquids Caused by γ-Rays ”, Dokl. Akad. Nauk SSSR 2, 451-457 (1934).

13. I. M. Frank and I. E. Tamm, “The Coherent Radiation of a Fast Electron in a Medium ”, Dokl. Akad. Nauk SSSR 14, 107-113 (1937).

14. V. L. Ginzburg, ” of Luminous Radiation From an Electron Traveling Uniformly in a Medium ”, J. Phys. USSR 2, 441-425 (1940). Recent discussions and reviews are given by V. L. Ginzburg in ”V. Radiation by Uniformly Moving Sources: Vavilov-Cherenkovˇ Effect, Doppler Effect in a Medium, Transition Radiation and Associated Phenomena) ”, Progress in Optics 32, pp.267-312 (1993), ”Radiation by Uniformly Moving Sources (Vavilov - Cherenkovˇ Effect, Transition Radiation, and Other Phenomena) ”, Physics - Uspekhi 39 No. 10, 973-982 (1996), ”Some Remarks on the Radiation of Charges and Multipoles Uniformly Moving in a Medium ”, Physics - Uspekhi 45 No. 3, 341-344 (2002), and Chap. 7 in “Applications of Electrodynamics in and Astrophysics ”, Gordon and Breach Sci. Pub. (New York, 1989).

15. E. G.Harris, “A Pedestrian Approach to Quantum Field Theory ”, Wiley-InterScience (N. Y., 1972).

16. D. Marcuse, “Principles of Quantum Electronics ”, Academic Press (N. Y., 1980).

17. T. Rothman, “The Secret History of Gravitational Waves ”, Am. Sci. 106 97-103 (2018).

18. D. G. Boulware and S. Deser, “Classical Derived from Quantum Gravity ”, Ann. Phys. (N.Y.) 89 193-240 (1975).

19. S. Weinberg and E. Witten, “Limits on Massless Particles ”, Phys. Lett. 96 B, 59-62 (1980). For a recent pedagogical account see, e.g., F. Loebbert,“The Weinberg-Witten Theorem on Massless Particles: An Essay ”, Ann. Phys. (Berlin) 17, 803-829 (2008).

20. S. Weinberg, “Gravitation and Cosmology: Principles and Applications of the General

20 Theory of Relativity ”, John Wiley & and Sons (New York, 1972) (a more recent account by S. Weinberg is “Cosmology ”, Oxford University Press (New York, 2008)).

21. E. Bertschinger, “Cosmological Dynamics ” in “Cosmology and Large Scale Struc- ture ”, pp. 273-347, “Proc. of Les Houches Summer School ”, Session LX, Eds. R. Schaeffer, J. Silk, M. Spiro, and J. Zinn-Justin, Elsevier (Amsterdam, 1996), and “Cosmological Dynamics ”, arXiv:astro-ph/9503125, 1995.

22. E.´ E.´ Flanagan and S. A. Hughes, “The Basics of Theory ”, New J. Phys. 7 204, 1-52 (2005).

23. C. W. Misner, K. S. Thorne, and J. A. Wheeler, “Gravitation ”, W. H. Freeman and Company (San Francisco, 1973).

24. S. Weinberg, “Infrared Photons and Gravitons ”, Phys. Rev. 140, B516-B524 (1965).

25. C. Alveg˚ard, K. E. Eriksson, and C. H¨ogfors, “Soft Gravitation Radiation ”, Physica Scripta 17, 95-102 (1978). (Due to unfortunate printing errors, some references are missing. In particular Ref.28 in this publication is our Ref. [24]. Also the reference to S. Weinberg after their Eq.(4.18) should be to our Ref. [11]).

26. B. P. Abbott et al. (LIGO Scientific Collaboration and Virgo Collaboration), “GW170814 : A Three-Detector Observation of Gravitational Waves from a Binary Black Hole Coalescence ”, Phys. Rev. Lett. 119, 141101-1-16 (2017).

27. “Optical Angular Momentum ”, Eds. L. Allen, S. M. Barnett, and M. J. Padgett, Institute of Physics (Bristol, England, 2003); S. Franke-Arnold, L. Allen, and M. J. Padgett, ‘Advances in Optical Angular Momentum ”, Laser & Photon. Rev. 2, No. 4, 299-313 (2008); A. M. Yao and M. J. Padgett, “Orbital Angular Momentum: Origins, Behavior and Applications ”, Advances in Optics and Photonics 3, 161-204 (2011); A. E. Willner et al., “Optical Communications Using Orbital Angular Momentum Beams ”, Advances in Optics and Photonics 7, 66-106 (2015).

28. A. Einstein, “N¨aherungsweise Integration der Feldgleichungen der Gravitation, N¨aherungsweise Integration der Feldgleichungen der Gravitation ”, Preuss. Akad. Wiss. Berlin, 688-696 (1916) and in “Uber¨ Gravitationswellen ”, and ibid., 154-167 (1918). A numerical factor of two in front of the quadrupole formula is, however, missing.

29. L. D. Landau and E. M. Lifshitz, ”The Classical Theory of Fields ”, Fourth Edition, Vol. 2 in ”Course on Theoretical Physics ”, Elsevier Science Inc. (Oxford, 1975).

30. H. C. Ohanian, “Gravitation and SpaceTime ”, W. W. Norton & Company (New

21 York, 1976).

31. S. Caroll, “SpaceTime and Geometry - An Introduction to General Relativity ”, Ad- dison Wesley (San Francisco, 2004).

32. M. Maggiore, “Gravitational Waves - Theory and Experiments ”, Vol. 1, Oxford University Press (New York, 2008).

33. S. A. Balbusa, “Simplified Derivation of the Gravitational Wave Stress Tensor from the Linearized ”, PNAS 42 1166211666 (2016).

34. J. F. Donoghue and B. R. Holstein, “Low Energy Theorems of Quantum Gravity From Effective Field Theory ”, J. Phys. G: Nucl. Part. Phys. 42 103102 (45pp) (2015).

35. R. P. Feynman, “An Operator Calculus Having Applications in Quantum Electrody- namics ”, Phys. Rev. 84, 108-128 (1951).

36. R. J. Glauber, ”Some Notes on Multiple- Processes ”, Phys. Rev. 84, 395-400 (1951); ”Coherent and Incoherent States of the Radiation Field ”, Phys. Rev. 131, 2766-2788 (1963), ”Optical and Photon Statistics ” in ”Quantum Optics and Electronics ”, Les Houches, pp. 65-185621, Eds. A. Blandin, C. DeWitt-Morette, and C. Cohen-Tannoudji, Gordon and Breach (New York, 1965), and in ”Quantum Theory of Optical Coherence ”, Wiley-VCH (Weinheim, 2007).

37. P. Carruthers and M. M. Nieto, “Coherent States and the Forced Harmonic Oscilla- tor ”, Am. J. Phys. 33, 537-544 (1965)(reprinted in Ref. [39]).

38. K. E. Eriksson, ”Summation Methods for Radiative Corrections ” in: M. L´evy (Ed.), ”Carg`ese Lectures in Physics ”, Vol. 2, pp. 245-274, Gordon and Breach (New York, 1967).

39. J. R. Klauder and B.-S. Skagerstam, “Coherent States - Applications in Physics and Mathematical Physics ”, World Scientific (Singapore, 1985 and Beijing 1988).

40. B.-S. Skagerstam, “Coherent States - Some Applications in Quantum Field Theory and ” in “Coherent States: Past, Present, and the Future ”, pp. 469- 506, Eds. D. H. Feng, J. R. Klauder and M. R. Strayer, World Scientific (Singapore, 1994); J. R. Klauder, ”The Current State of Coherent States ”, contribution to the 7th ICSSUR Conference, Boston, 2001 (arXiv:quant-ph/0110108).

41. V. V. Dodonov and V. I. Man’ko, “Theory of Nonclassical States of Light ” , Taylor & Francis (London, 2003).

42. J. D. E. Creighton and W. G. Anderson, “Gravitational-Wave Physics and Astronomy

22 - Introduction to Theory, Experiment and Data Analysis ”, Wiley Series in Cosmology (Wiley-VCH, Weinheim, 2011).

43. N. Bohr and L. Rosenfeld, “Field and Measurements in Quantum Electrody- namics ”, Phys. Rev. 78 794-798 (1950) and references therein.

44. A. S. Wightman, “La Th´eorie Quantique Locale et la Th´eorie Quantiques des Champs ”, Ann. Inst. Poincar´e 1, No.4, 403-420 (1964).

45. During the final preparation of the present paper we became aware about the work by Q. Wang, Z. Zhu, and W. G. Unruh, “How the Huge Energy of Quantum Vacuum Gravitates to Drive the Slow Accelerating Expansion of the ”, Phys. Rev. D 95, 103504-1–103504-34 (2017), where the fundamental role of unavoidable quantum fluctuations may play a fundamental role in cosmology.

46. S. Bough and T. Rothman, ”Graviton Emission and Absorption by Atomic Hydro- gen ”, arXiv:gr-qc/0605052 v1, 2006, and in ”Aspects of Graviton Detection: Graviton Emission and Absorption by Atomic Hydrogen ”, Class. Quantum Grav. 23, 5839 - 5852 (2006).

47. C. Kiefer, “Quantum Gravity ”, Third Edition, Oxford University Press (Oxford, 2012).

48. M. D. Scadron, “Advanced Quantum Theory ”, Springer-Verlag (New York, 1979).

49. W. E. Lamb, Jr., “Anti-Photon ”, Appl. Phys. B60, 77-84 (1995).

50. B. P. Abbott et al. (LIGO Scientific Collaboration and Virgo Collaboration), “Ob- servation of Gravitational Waves from a Binary Black Hole Merger ”, Phys. Rev. Lett. 116, 061102-1-12 (2016).

51. B. P. Abbott et al. (LIGO Scientific Collaboration and Virgo Collaboration), “As- trophysical Implications of the Binary Black Hole Merger GW150914 ”, Astrophys. J. Lett. 818, L22, 1-15 (2016).

52. B. P. Abbott et al. (LIGO Scientific Collaboration and Virgo Collaboration), “GW151226: Observation of Gravitational Waves from a 22-Solar-Mass Binary Black Hole Coalescencer ”, Phys. Rev. Lett. 116, 241103-1-7 (2016).

53. B. P. Abbott et al. (LIGO Scientific Collaboration and Virgo Collaboration), “GW170104: Observation of a 50-Solar-Mass Binary Black Hole Coalescence at Red- shift 0.2 ”, Phys. Rev. Lett. 118, 221101-1-17 (2017).

54. B. P. Abbott et al. (LIGO Scientific Collaboration and Virgo Collabora-

23 tion),“GW170817 : Observation of Gravitational Waves from a Binary Inspiral ”, Phys. Rev. Lett. 119, 161101-1-18 (2017).

55. F. J. Dyson, “The World on a ”, in New York Review of Books 51 No.8, May 13 Issue (2004). Also see “Is a Graviton Detectable? and references cited therein, in “Proceedings of the XVIIth International Congress on Mathematical Physics ”, pp. 670-682, Ed. A. Jensen, World Scientific (Singapore, 2013).

56. S. Bough and T. Rothman, ”Can Gravitons be Detected ”, arXiv:gr-qc/0601043 v3, 2006 and in Found. Phys. 36 1801-1825 (2006).

57. L. M Sokolowski and A. Staruszkiewicz, ”On the Issue of Gravitons ”, arXiv:gr-qc/0606111 v2, 2006, and in Class. Quantum Grav. 23, 59075917 (2006).

58. K. Gottfried, “Quantum Mechanics: Vol I: Fundamentals ”, W. A. Benjamin, Inc. (New York, 1966).

59. C. Itzykson, “Soft Quanta ”, Ann. Phys. 7, 59-83 (1972). Also discussed in Section 1-3-2 of C. Itzykson and J.-B. Zuber, “Relativistic Quantum Theory ”, McGraw-Hill (New York, 1980). All radiative corrections to Thomson scattering vanishes in the low-energy limit. This theorem goes back to W. Thirring, “Radiative Corrections in the Non-Relativistic Limit ”, Phil. Mag. 41 1193-1194 (1950). More details can be found in Section 19.3 of J. D. Bjorken and S. D. Drell, “Relativistic Quantum Fields ”, McGraw-Hill (New York, 1965).

60. I. Fernandez-Corbaton, M. Cirio, A. B¨use, L. Lamata, E. Solano, and G. Molina- Terriza, “Quantum Emulation of Gravitational Waves ”, Scientific Reports 5:11538, 1-10 (2015)

61. J. S. Bell, “On the Einstein-Podolsky-Rosen Paradox ”, Physics 1, 19-200 (1965).

62. J. F. Clauser, M. A. Horne, A. Shimony and R. A. Holt, “Proposed Experiment to Test Local Hidden-Variable ”, Phys. Rev. Lett. 23, 880-884, (1969).

63. A. Aspect, P. Grangier, and G. Roger, “Experimental Realization of Einstein- Podolsky-Rosen-Bohm Gedankenexperiment: A New Violation of Bell’s Inequalities”, Phys. Rev. Lett. 49, 91-94, (1982); A. Aspect, “Bell’s Inequality Test: More Ideal Than Ever ”, Nature 398, 189-190 (1999) and “Bell’s Theorem: The Naive View of an Experimentalist ” published in “Quantum Unspeakables From Bell to ”, Eds. R. A. Bertlmann and A. Zeilinger, Springer (Berlin, 2002). Also available at arXiv:quant-ph/0402001 and references cited therein.

64. B.-S. Skagerstam,“EPR-Correlations for Emulated Gravitons”, to appear.

24