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

Gravitons and Temperature Fluctuation Correlations from Inflation

L. H. Ford ford@cosmos.phy.tufts.edu Institute of Cosmology, Department of Physics and Astronomy, Tufts University, Medford, Massachusetts 02155, USA    I-Tai Ho yidaihoap@gmail.com Department of Physics, National Taiwan Normal University, Taipei 116, Taiwan    Chun-Hsien Wu chunwu@scu.edu.tw Department of Physics, Soochow University, Taipei 111, Taiwan
Abstract

Inflationary tensor perturbations are treated as arising from a bath of gravitons produced by quantum particle creation at the end of inflation. We calculate the correlation function of the CMB temperature fluctuations produced by these gravitons in a model with an infrared cut off. The CMB photons are emitted from within a last scattering shell of finite thickness in redshift. We find the correlation function in terms of the separation of a pair of spacetime points of emission in both angle and redshift. In both variables, there is a significant amount of anti-correlation. The anti-correlation minimum has a relative magnitude compared to the central correlation maximum of about 20%20\% in angle and 50%50\% in redshift.

I Introduction

This work will re-examine inflationary tensor perturbations from a viewpoint which emphasizes their quantum origin and nature. Starobinsky S79 was apparently the first author to suggest that a period of deSitter expansion could produce an observable spectrum of gravity wave perturbations in the subsequent universe. This early work was soon followed by contributions from several authors, including Rubakov, Sazhin, and Veryaskin RSV82 , Fabbri and Pollack FP83 , Abbott and Wise AW84b , Abbott and Harari AH86 , Polnarev Polnanev , and Allen A88 .

A key result found in Refs. FP83 ; Polnanev is that a classical gravitational wave propagating in an expanding universe induces an anisotropic pattern of redshifts on photons also propagating in this universe. These are the tensor perturbations in the Cosmic Microwave Background (CMB) which may be observable. When these anisotropically redshifted photons scatter from electrons, a particular polarization pattern (the B-modes) is induced and may serve as an observational signal of primordial gravity waves. For a review, see Kamionkowski and Kovetz KK16 .

Inflation does not produce a single mode gravitational wave, but rather a spectrum of wave numbers and polarization states. This mixture is often described as a stochastic bath of gravity waves. However, a deeper approach involves quantum creation of gravitons by the expansion of the universe, a process first described by Parker Parker69 and applied to graviton creation in Refs. Grishchuk:1975 ; FordParker:1977 . For a review, see Ref. pp-review . Calculations of graviton creation at the end of inflation were given by Abbott and Harari AH86 and by Allen A88 . The quantum state of the created gravitons is predicted to be highly nonclassical, specifically a squeezed vacuum state GS89 .There has been an extensive discussion in the literature as to whether there is a quantum to classical transition and to whether the gravitons undergo decoherence PS96 ; GS19 ; Kanno22 ; Burgess23 ; Ning23 .

In the present paper, we explore the hypothesis that no quantum to classical transition occurs, and the gravitons remain in a nonclassical state when they interact with the CMB photons. The outline of the paper is the following: In Sect. II we review quantum graviton creation at the end of inflation, and discuss the resulting quantum state in Sect. III. The effects of the gravitons on the CMB photons is treated in Sect. IV. Section V contains our main results on the temperature correlation function and its physical implications. Section VI gives a summary.

II Graviton creation at the end of inflation

This section will review the basic principles of quantum particle creation in an expanding universe and application to inflationary cosmology.

II.1 Gravitons in a Spatially Flat Universe

Here we consider the spatially flat Friedman-Robertson-Walker spacetime with metric

ds2=dt2+a2(t)d𝐱2=a2(η)(dη2+d𝐱2),ds^{2}=-dt^{2}+a^{2}(t)d{\bf x}^{2}=a^{2}(\eta)\,\bigl{(}-d\eta^{2}+d{\bf x}^{2}\bigr{)}\,, (1)

where tt and η\eta are the comoving and conformal time coordinates, respectively. We are interested in treating gravitons propagating on this background spacetime. It was shown by LifshitzLifshitz that in the transverse, tracefree gauge, which eliminates all gauge freedom, each of the two independent polarization states is equivalent to a massless, minimally coupled scalar field which satisfies

φ=μμφ=0.\Box\varphi=\nabla_{\mu}\,\nabla^{\mu}\varphi=0\,. (2)

The positive norm plane wave solutions of this equation may be taken to be

f𝐤(𝐱,η)=16πPei𝐤𝐱a(η)(2π)3χk(η),f_{\bf k}({\bf x},\eta)={\frac{\sqrt{16\pi}\,\ell_{P}\,e^{i{\bf k\cdot x}}}{a(\eta)\sqrt{(2\pi)^{3}}}}\,\,\chi_{k}(\eta), (3)

where P\ell_{P} is the Planck length and we use the normalization for graviton modes, which is a factor of 16πP\sqrt{16\,\pi}\,\ell_{P} times that for the scalar field, Here χk(η)\chi_{k}(\eta) satisfies

d2χkdη2+[k216a2(η)R(η)]χk=0,{\frac{d^{2}\chi_{k}}{d\eta^{2}}}+\left[k^{2}-\frac{1}{6}\,a^{2}(\eta)\,R(\eta)\right]\chi_{k}=0\,, (4)

where R(η)R(\eta) is the scalar curvature.

First consider the spacetime of the inflationary model, deSitter space, where we may take

a(η)=1Hη,a(\eta)=-\frac{1}{H\,\eta}\,, (5)

where HH is the Hubble parameter and η0<η<ηR<0\eta_{0}<\eta<\eta_{R}<0. Here η0\eta_{0} denotes the time when inflation began and ηR\eta_{R} that when it ends in reheating. Here the scalar curvature is R=12H2R=12\,H^{2}. The solutions of Eq. (4) may be expressed in terms of Hankel functions as

χk(η)=η3/22k[c1(k)H3/2(1)(kη)+c2(k)H3/2(2)(kη)],\chi_{k}(\eta)=\frac{\eta^{3/2}}{\sqrt{2\,k}}\left[c_{1}(k)\,H^{(1)}_{3/2}(k\eta)+c_{2}(k)\,H^{(2)}_{3/2}(k\eta)\right]\,, (6)

where |c2(k)|2|c1(k)|2=1.|c_{2}(k)|^{2}-|c_{1}(k)|^{2}=1.

If we set c1=0c_{1}=0 and c2=1c_{2}=1, the resulting mode function

χkBD(η)=η3/22kH3/2(2)(kη)=12k(1ikη)eikη,\chi^{BD}_{k}(\eta)=\frac{\eta^{3/2}}{\sqrt{2\,k}}\,H^{(2)}_{3/2}(k\eta)=\frac{1}{\sqrt{2\,k}}\,\left(1-\frac{i}{k\,\eta}\right)\;{\rm e}^{-i\,k\eta}\,, (7)

defines the Bunch-Davies vacuum state of deSitter space BD78 . In the case of a massive scalar field, it is the unique deSitter invariant state, and is analogous to the Minkowski vacuum, the unique Lorentz invariant state in flat spacetime. However, the massless scalar field has a infrared divergent two-point function, φ(x)φ(x)\langle\varphi(x)\,\varphi(x^{\prime}), so the Bunch-Davies state is not well defined for either the massless scalar or the graviton fields. There is a class of states which are free of infrared divergences FordParker:1977b ; FV86 . If we require that c1(k)+c2(k)0c_{1}(k)+c_{2}(k)\rightarrow 0 as k0k\rightarrow 0, then the two-point function is infrared finite. This class may be said to be Bunch-Davies-like if we also have c1(k)0c_{1}(k)\approx 0 and |c2(k)|1|c_{2}(k)|\approx 1 for kkck\gtrsim k_{c} for some small cutoff scale kck_{c}. We can view this scale as set by some initial condition before the onset of inflation. If kck_{c} is sufficiently small that the associated modes have not entered the horizon since the end of inflation, then they seem to play no role in present cosmological observations. In this case, we may take Eq. (7) as the initial graviton mode before reheating.

II.2 The transition from Inflation to a Radiation Dominated Universe

The creation of gravitons at the end of inflation may be viewed as gravitational particle creation described by a Bogolubov transformation. Here we take the in-modes to be the Bunch-Davies-like modes given by Eqs. (3) and (7). In the subsequent radiation dominated phase, the scalar curvature vanishes, R=0R=0, so the positive unit norm mode functions become

F𝐤(𝐱,η)=16πPei(𝐤𝐱kη)2k(2π)3a(η).F_{\bf k}({\bf x},\eta)=\frac{\sqrt{16\pi}\,\ \ell_{P}\,e^{i({\bf k\cdot x}-k\,\eta)}}{\sqrt{2k\,(2\pi)^{3}}\,a(\eta)}\,. (8)

We are interested in very long wavelength gravitons modes which are far outside the horizon at the end of inflation, so kHk\ll H. These modes are not sensitive to the finite duration of reheating, so we assume a sharp transition in which the scale factor is given by Eq. (5) for η<H1\eta<-H^{-1} and by

a(η)=Hη+2a(\eta)=H\,\eta+2 (9)

for η>H1\eta>-H^{-1}. Note that we take the reheating time to be ηR=H1\eta_{R}=-H^{-1} and have set a(ηR)=1a(\eta_{R})=1.

The mode functions in the in and out regions are related by a Bogolubov transformation of the form

f𝐤(𝐱,η)=αkF𝐤(𝐱,η)+βkF𝐤(𝐱,η).f_{\bf k}({\bf x},\eta)=\alpha^{*}_{k}\,F_{\bf k}({\bf x},\eta)+\beta^{*}_{k}\,F^{*}_{\bf k}({\bf x},\eta)\,. (10)

where

|αk|2|βk|2=1,|\alpha_{k}|^{2}-|\beta_{k}|^{2}=1\,, (11)

The Bogolubov coefficients may be found by using Eq. (10) and its first time derivative at η=ηR\eta=\eta_{R}, using the fact that both a(η)a(\eta) and a(η)a^{\prime}(\eta) are continuous at this point. The results AH86 ; A88 are

|αk|2=1+H44k4,|\alpha_{k}|^{2}=1+\frac{H^{4}}{4\,k^{4}}\,, (12)

and

|βk|2=H44k4.|\beta_{k}|^{2}=\frac{H^{4}}{4\,k^{4}}\,. (13)

The mean number of gravitons created in a given mode is Nk=|βk|2\langle N_{k}\rangle=|\beta_{k}|^{2}. Here that number is very large. The energy spectrum of created gravitons is scale invariant in the sense that the graviton energy density in wave number interval d3kd^{3}k is proportional to k|βk|2d3kdk/kk\,|\beta_{k}|^{2}\,d^{3}k\propto dk/k, which independent of the choice of scale for kk. Note that in a spatially flat expanding universe, linear three-momentum is conserved, so the gravitons are created in pairs with equal and opposite wave vectors, 𝐤{\bf k} and 𝐤{-\bf k}

III The Quantum State of the Created Gravitons

III.1 Representation in the Out-Fock Space

The quantum particle creation process is being described in the Heisenberg picture, so the quantum state of the system, |ψ0|\psi_{0}\rangle, does not change in time. However, its description is very different in the deSitter space in-region from that in the radiation dominated out-region. In the in-region, it is one of several infrared finite choices of vacuum state. In the out-region, it is a highly excited state with very large numbers of particles in modes for which kHk\ll H.

The quantum state of the created particles in the out-Fock space may be shown to be

|ψ0=c0{n𝐤}𝐤γkn𝐤|{n𝐤},|\psi_{0}\rangle=c_{0}\sum_{\{n_{\bf k}\}}\prod_{\bf k}\gamma_{k}^{n_{\bf k}}\;|\{n_{\bf k}\}\rangle\,, (14)

where γk=βk/αk\gamma_{k}=\beta_{k}/\alpha_{k}. Here {n𝐤}\{n_{\bf k}\} denote a set of occupation numbers; for each mode 𝐤{\bf k} there is non-negative integer n𝐤n_{\bf k} which gives the number of particles in mode 𝐤{\bf k} and the number of particles in mode 𝐤{-\bf k}. In our enumeration, we may require kz0k_{z}\geq 0 to avoid over counting. The overall constant c0c_{0} satisfies

|c0|=(𝐤|αk|)1=𝐤1|γk|2,|c_{0}|=\left(\prod_{\bf k}|\alpha_{k}|\right)^{-1}=\prod_{\bf k}\sqrt{1-|\gamma_{k}|^{2}}\,, (15)

where the latter form follows from Eq. (11). This quantum state is a multi-mode squeezed vacuum state, containing all possible numbers of pairs of correlated particles. Strictly, Eq. (14) applies to the massless scalar field. For gravitons, we need to include the polarization among the mode labels. This will be explicitly addressed in Sect IV.1, where a sum over polarizations will be performed.

In the single mode case, only one type of pair is present. Here we have

|ψ01=1|γk|2n=0γkn|n,n,|\psi_{0}\rangle_{1}=\sqrt{1-|\gamma_{k}|^{2}}\,\sum_{n=0}^{\infty}\gamma_{k}^{n}\;|n,n\rangle\,, (16)

where |n,n|n,n\rangle denotes a state containing nn pairs, with one member of the pair in mode 𝐤{\bf k}, and the other in mode 𝐤{-\bf k}.

The expectation value of the field operator vanishes in this state,

ψ0|φ(x)|ψ0=0,\langle\psi_{0}|\varphi(x)|\psi_{0}\rangle=0, (17)

due to the fact that m,m|a𝐤|n,nm,m|n1,n=0\langle m,m|a_{\bf k}|n,n\rangle\propto\langle m,m|n-1,n\rangle=0 for all integers mm and nn. This holds for both the one-mode case of Eq. (16), and the multi-mode case of Eq. (14). As a result, the state |ψ01|\psi_{0}\rangle_{1} does not describe a classical wave.

III.2 The Observable Part of the State

It is well known in quantum theory that outcomes of observations can depend crucially upon the details of the question which is addressed, or the specific measurement which is undertaken. At least in the cosmological case, we can only observe one member of a pair of correlated gravitons. This can be seen if we adopt a set of wave packet mode functions, rather than plane waves with a precisely defined wave vector. Consider a set of wave packets with a finite but small bandwidth, Δkk=|𝐤|\Delta k\ll k=|{\bf k}|, where 𝐤{\bf k} is the peak wave vector of a packet. These packets will have a spatial extent of order Δx1/Δk\Delta x\approx 1/\Delta k.

In the wave packet basis, the correlated pairs of gravitons are created into a pair of packets peaked about values of 𝐤{\bf k} with equal magnitude, but opposite sign. This means that the two packets leave the region of their creation in opposite directions. A given observer will be able to observe at most one of these gravitons. This is illustrated in Fig, 1 for the case of gravitons which influence the redshifts of CMB photons.

Refer to caption
Figure 1: A spacetime diagram illustrating the history of graviton (gg) and photon (γ\gamma) wavepackets which may produce observable features in the CMB. Here ηR\eta_{R}, ηD\eta_{D}, and η0\eta_{0} are the conformal times associated with reheating, decoupling, and the present, respectively. At event 11, entangled gravitons are created, with one packet traveling to event 22, where it influences a photon packet, which reaches Earth at 33. The other entangled graviton mode goes in the opposite direction, and reaches the last scattering region at η=ηD\eta=\eta_{D} at event 44. There it influences photons which reach Earth at event 55, when η=η1\eta=\eta_{1}, about 10910^{9} years after event 33. At the same time that the photons from 22 are observed, there are photons arriving from the opposite direction which were emitted at event 77 and were influenced by gravitons created at 66.

Suppose we are making measurements of local field operators in a spacetime region {\cal R}, and using a complete set of wave packets with mode labels {j}\{j\}. It is convenient to divide the set {j}\{j\} into two subsets, {j1}\{j_{1}\} and {j2}\{j_{2}\}, where the j1j_{1} modes are nonzero in {\cal R} , and the j2j_{2} modes vanish there and do not contribute to local expectation values in {\cal R}. Now we can rewrite Eq. (14) as

|ψ0=c0{nj}jγknj|{nj1},{nj2},.|\psi_{0}\rangle=c_{0}\sum_{\{n_{j}\}}\prod_{j}\gamma_{k}^{n_{j}}\;|\{n_{j_{1}}\},\{n_{j_{2}}\},\rangle\,. (18)

Equation (16) is the special case where there is one j1j_{1} mode and one j2j_{2} mode.

If ii and \ell are two modes in set j1j_{1}, then the expectation values in state |ψ0|\psi_{0}\rangle of the associated creation and annihilation operators are

ai=ai=aia=aia=0,\langle a_{i}\rangle=\langle a_{i}^{\dagger}\rangle=\langle a_{i}a_{\ell}\rangle=\langle a^{\dagger}_{i}a^{\dagger}_{\ell}\rangle=0\,, (19)

but

aia=δi|βi|2.\langle a^{\dagger}_{i}a_{\ell}\rangle=\delta_{i\ell}\,|\beta_{i}|^{2}\,. (20)

This implies that expectation values of linear field operators vanish in this state, but those of quadratic operators can be nonzero. Furthermore, the effects of the state |ψ0|\psi_{0}\rangle in {\cal R} are equivalent to a density matrix.

Let aa and aa^{\dagger} be associated with any given mode in set j1j_{1}, and define a coordinate by

x=12(a+a)x=\frac{1}{\sqrt{2}}(a+a^{\dagger}) (21)

and a conjugate momentum by

p=i2(aa).p=\frac{i}{\sqrt{2}}(a-a^{\dagger})\,. (22)

Then x=p=0\langle x\rangle=\langle p\rangle=0  , but the ubcertainties are nonzero: Δx=x2=Δp\Delta x=\sqrt{x^{2}}=\Delta p. Furthermore the uncertainty relation becomes

ΔxΔp=|β|2+12.\Delta x\;\Delta p=|\beta|^{2}+\frac{1}{2}\,. (23)

In general, |ψ0|\psi_{0}\rangle is not a state of minimum uncertainty, unlike the Stoller squeezed states, and will have very large uncertainty when |βj|1|\beta_{j}|\gg 1. This implies large fluctuations, as evidenced the large variance of linear field operators.

III.3 An Electromagnetic Analog Model

There is a simple analog involving a squeezed state of photons which is illustrated in Fig. 2. A strong classical electromagnetic wave in a nonlinear material with a nonzero second or third order susceptibility creates a time dependent effective index of refraction for vacuum electromagnetic modes. This leads to mixing of positive and negative frequencies and quantum creation of photons, an effect analogous to the quantum creation of gravitons in inflation discussed in Sect. II. The photons are emitted in correlated pairs in opposite directions from the material in a quantum state of the form of |ψ0|\psi_{0}\rangle given in Eq. (18).

Refer to caption
Figure 2: The quantum creation of photons in a squeezed state in a crystal of nonlinear material is illustrated. A strong classical electromagnetic wave enters the crystal on the left. Correlated pairs of photons are emitted from the crystal in opposite directions. Experiments performed in regions {\cal R} are sensitive to only one member of each pair.

III.4 Scalar Correlation Functions

In this subsection, we focus on the case of a single massless, minimally coupled scalar field φ(x)\varphi(x), which is equivalent to one polarization state of the graviton field in the transverse tracefree gauge. Assume that we are in the radiation dominated era, when the mode functions may be given by Eq. (8) and the field operator may be expressed as

φ(x)=𝐤[a𝐤F𝐤(x)+a𝐤F𝐤(x)],\varphi(x)=\sum_{\bf k}\left[a_{\bf k}F_{\bf k}(x)+a^{\dagger}_{\bf k}F^{*}_{\bf k}(x)\right]\,, (24)

where the a𝐤a_{\bf k} satisfy the relations in Eqs. (19) and (20) in a state of the form of |ψ0|\psi_{0}\rangle. Next define a correlation operator by

κ(x,x)=φ(x)φ(x)\kappa(x,x^{\prime})=\varphi(x)\,\varphi(x^{\prime}) (25)

and a correlation function by

C(x,x)=κ(x,x).C(x,x^{\prime})=\langle\kappa(x,x^{\prime})\rangle\,. (26)

Recall that φ(x)=0\langle\varphi(x)\rangle=0, so C(x,x)C(x,x) is the variance of the fluctuations of φ(x)\varphi(x). If |βj|21|\beta_{j}|^{2}\gg 1, then

C(x,x)2j|βj|2Re,[Fj(x)Fj(x)],C(x,x^{\prime})\approx 2\,\sum_{j}|\beta_{j}|^{2}\,\rm Re,\left[F^{*}_{j}(x^{\prime})\,F_{j}(x)\right]\,, (27)

where we ignore a term of order |βj|0|\beta_{j}|^{0}.

Consider the case where a single mode jj gives the dominant contribution to C(x,x)C(x,x^{\prime}), so that C(x,x)Fj(x)C(x,x^{\prime})\propto F_{j}(x), and the correlation function has the form of a classical wave as a function of xx for fixed xx^{\prime}. This is similar to the case of a coherent state. However, unlike a highly excited coherent state, here the fluctuations of the correlation operator are very large:

κ2(x,x)=2κ(x,x)2+4|βj|4|Fj(x)Fj(x)|2+O(|βj|2),\langle\kappa^{2}(x,x^{\prime})\rangle=2\langle\kappa(x,x^{\prime})\rangle^{2}+4|\beta_{j}|^{4}\,|F_{j}(x^{\prime})\,F_{j}(x)|^{2}+O(|\beta_{j}|^{2})\,, (28)

so the fractional variance of KK is of order one

κ2(x,x)κ(x,x)2κ(x,x)2=O(|βj|0).\frac{\langle\kappa^{2}(x,x^{\prime})\rangle-\langle\kappa(x,x^{\prime})\rangle^{2}}{\langle\kappa(x,x^{\prime})\rangle^{2}}=O(|\beta_{j}|^{0})\,. (29)

The fluctuations are illustrated in Fig. 3. Because of the large fractional fluctuations, the correlation effects can only be observed in an ensemble average.

Refer to caption
Figure 3: The correlation fluctuations are illustrated. The mean value of the correlation operator κ\kappa, which is the correlation function, undergoes periodic oscillations. However, the uncertainty, Δκ\Delta\kappa, is of the same order as the mean value.

IV Effects of Gravity Waves and of Gravitons on CMB Photons

IV.1 Differential Redshifts due to a Classical Gravity Wave

IV.1.1 ++ Polarization

The metric may be written as

ds2=a2(η)[dη2+(1+h+)dx2+(1h+)dy2+dz2],ds^{2}=a^{2}(\eta)\,[-d\eta^{2}+(1+h_{+})\,dx^{2}+(1-h_{+})\,dy^{2}+dz^{2}]\,, (30)

where h+=h+(ηz)h_{+}=h_{+}(\eta-z). The equation for the four-velocity uμu^{\mu} of a timelike geodesic is

duμdτ=Γαβμuαuβ,\frac{du^{\mu}}{d\tau}=-\Gamma^{\mu}_{\alpha\beta}\,u^{\alpha}u^{\beta}\,, (31)

where τ\tau is the observer’s proper time, and dτ=dt=adηd\tau=dt=a\,d\eta. The connection coefficient which we need is, to first order in h+h_{+},

Γηηη=aa.\Gamma^{\eta}_{\eta\eta}=\frac{a^{\prime}}{a}\,. (32)

Hence the geodesic equation has the solution uμ=(a1,0,0,0)u^{\mu}=(a^{-1},0,0,0). This tells us that the comoving observers are at rest in the coordinates of Eq. (30). Let qμq^{\mu} be the 4-dimensional wave vector of a photon. The angular frequency of this photon in the comoving observer’s frame is

ω=qμuμ=a(η)qη.\omega=-q^{\mu}\,u_{\mu}=a(\eta)\,q^{\eta}\,. (33)

We may find the wave vector qμq^{\mu} either from geodesic equation, or the null vector condition, qμqμ=0q^{\mu}\,q_{\mu}=0. Here we use the latter. In the metric of Eq. (30) we find

qη=(1+h+)(qx)2+(1h+)(qy)2+(qz)2.q^{\eta}=\sqrt{(1+h_{+})\,(q^{x})^{2}+(1-h_{+})\,(q^{y})^{2}+(q^{z})^{2}}\,. (34)

In the isotropic case where h+=0h_{+}=0, we have qμ=q0μq^{\mu}=q_{0}^{\mu}, where

q0μ=(q0η,q0x,q0y,q0z)=ω0a2(η)(1,sinθcosϕ,sinθsinϕ,cosθ).q_{0}^{\mu}=(q_{0}^{\eta},q_{0}^{x},q_{0}^{y},q_{0}^{z})=\omega_{0}\,a^{-2}(\eta)\,(1,\sin\theta\cos\phi,\sin\theta\sin\phi,\cos\theta)\,. (35)

Here θ\theta and ϕ\phi are the direction angles for the photon in a frame where the gravity wave propagates in the +z+z-direction, and ω0\omega_{0} is the photon angular frequency when a=1a=1 The factor of a2(η)a^{-2}(\eta) arises because qμ=ω0dxμ/dλq^{\mu}=\omega_{0}\,dx^{\mu}/d\lambda and λ\lambda is an affine parameter. The latter may be taken to be given by dλ=a2dηd\lambda=a^{2}\,d\eta. (See, for example, Visser23 .)

To first order in h+h_{+}, we may place the above expressions for the zero order spatial components q0μq_{0}^{\mu} in the right hand side of Eq. (34), Taylor expand, and use Eq. (33) to find

ω+ω0a(η)(1+12h+sin2θcos2ϕ).\omega_{+}\approx\frac{\omega_{0}}{a(\eta)}\,\left(1+\frac{1}{2}h_{+}\,\sin^{2}\theta\,\cos 2\phi\right)\,. (36)

The pre-factor, ω0/a(η){\omega_{0}}/{a(\eta)}, is the redshifted frequency in an isotropic universe, and the term proportional to h+h_{+} gives the fractional anisotropic redshift due to the gravity wave. Specifically, we define the fractional redshift by

Δω+=ω+ω0/a(η)ω0/a(η)=12h+sin2θcos2ϕ.\Delta\omega_{+}=\frac{\omega_{+}-\omega_{0}/a(\eta)}{\omega_{0}/a(\eta)}=\frac{1}{2}h_{+}\,\sin^{2}\theta\,\cos 2\phi\,. (37)

Apart from a sign difference, this agrees with the result of Polnarev Polnanev which is quoted in Ref. KK16

IV.1.2 ×\times Polarization

Now the metric may be written as

ds2=a2(η)[dη2+dx2+2h×dxdy+dy2+dz2],ds^{2}=a^{2}(\eta)\,[-d\eta^{2}+dx^{2}+2h_{\times}dx\,dy+dy^{2}+dz^{2}]\,, (38)

where h×=h×(ηz)h_{\times}=h_{\times}(\eta-z). If we repeat the procedure in the previous subsection, the result is

ω×ω0a(η)(1+12h×sin2θsin2ϕ).\omega_{\times}\approx\frac{\omega_{0}}{a(\eta)}\,\left(1+\frac{1}{2}h_{\times}\,\sin^{2}\theta\,\sin 2\phi\right)\,. (39)

Note that if we let ϕϕ+π/4\phi\rightarrow\phi+\pi/4, then ω+ω×\omega_{+}\rightarrow\omega_{\times}, as a rotation by π/4\pi/4 interchanges the two polarizations. Now the fractional redshift becomes

Δω×=12h×sin2θsin2ϕ.\Delta\omega_{\times}=\frac{1}{2}h_{\times}\,\sin^{2}\theta\,\sin 2\phi\,. (40)

Note that if h+2=h×2=h2h_{+}^{2}=h_{\times}^{2}=h^{2}, then

(Δω+)2+(Δω×)2=14h2sin4θ.(\Delta\omega_{+})^{2}+(\Delta\omega_{\times})^{2}=\frac{1}{4}h^{2}\,\sin^{4}\theta\,. (41)

Thus the dependence upon the angle ϕ\phi disappears from a sum over polarization states.

IV.2 The Quantized Graviton Field

The graviton field operator may be expanded in terms of mode functions as

hμν(𝐱,η)=𝐤,λ[eμν(𝐤,λ)F𝐤(𝐱,η)a𝐤,λ+H.c.],h^{\mu\nu}({\bf x},\eta)=\sum_{{\bf k},\lambda}\left[{\rm e}^{\mu\nu}({\bf k},\lambda)\,F_{\bf k}({\bf x},\eta)\,a_{{\bf k},\lambda}+H.c.\right]\,, (42)

where eμν(𝐤,λ){\rm e}^{\mu\nu}({\bf k},\lambda) is the polarization tensor for mode 𝐤,λ{\bf k},\lambda. The expectation value of hμνh^{\mu\nu} in a coherent state for mode 𝐤,λ{{\bf k},\lambda} is a classical gravitational wave. Here we use a linear polarization basis, so λ=+,×\lambda=+,\times, and the polarization tensors are implicitly defined by Eqs. (30) and (38).

The graviton correlation function may be expressed as

hμν(𝐱,η)hρσ(𝐱,η)=P2π2a(η)a(η)d3kk|βk|2λeμνeρσcos[(𝐤(𝐱𝐱)k(ηη)].\langle h^{\mu\nu}({\bf x},\eta)\,h^{\rho\sigma}({\bf x^{\prime}},\eta^{\prime})=\frac{\ell_{P}^{2}}{\pi^{2}\,a(\eta)\,a(\eta^{\prime})}\;\int\frac{d^{3}k}{k}\,|\beta_{k}|^{2}\,\sum_{\lambda}{\rm e}^{\mu\nu}{\rm e}^{\rho\sigma}\;\cos[({\bf k}\cdot({\bf x}-{\bf x^{\prime}})-k\,(\eta-\eta^{\prime})]\,. (43)

Here we have used Eqs. (8) and (20), as well as the convention that 𝐤d3k\sum_{\bf k}\rightarrow\int d^{3}k in the continuum limit.

V The Temperature Correlation Function

The graviton correlation function, Eq. (43) may be used to compute a CMB temperature fluctuation correlation function using the results of Sect. IV.1. Equation (41) was derived for classical gravity waves, but also applies expectation values in a graviton bath if we replace h2h^{2} by an expectation value of a quadratic graviton operator. Specifically, we can convert Eq. (43) into a result for the fractional temperature fluctuation correlation function:

C(𝐫,η,η)=P24π2a(η)a(η)d3kksin4χ|βk|2cos[(𝐤𝐫k(ηη)],C({\bf r},\eta,\eta^{\prime})=\frac{\ell_{P}^{2}}{4\,\pi^{2}\,a(\eta)\,a(\eta^{\prime})}\;\int\frac{d^{3}k}{k}\,\sin^{4}\chi\,|\beta_{k}|^{2}\,\cos[({\bf k}\cdot{\bf r}-k\,(\eta-\eta^{\prime})]\,, (44)

where 𝐫=𝐱𝐱{\bf r}={\bf x}-{\bf x^{\prime}}, and we are interested in the correlation between photons emitted at spacetime point (η,𝐱)(\eta,{\bf x}) with those emitted at (η,𝐱)(\eta^{\prime},{\bf x^{\prime}}) . Here χ\chi denotes the angle between our line of sight and the graviton wave vector 𝐤{\bf k}. We have used Eq. (41) to make the replacement λeμνeρσ14sin4χ\sum_{\lambda}{\rm e}^{\mu\nu}{\rm e}^{\rho\sigma}\rightarrow\frac{1}{4}\,\sin^{4}\chi.

V.1 The infrared Cutoff k0k_{0}

If we use Eq. (13), with |βk|2k4|\beta_{k}|^{2}\propto k^{-4}, then the integral in Eq. (44) will diverge at the lower limit. This issue is most likely distinct from the infrared divergence discussed in Sect. II.1 which required a cutoff at k=kek=k_{e}. The latter was required because gravitons and massless minimally coupled scalar fields necessarily break deSitter symmetry. The cutoff at k=kek=k_{e} arises from a memory of pre-inflationary conditions which is was not erased by a finite amount of inflation. The infrared divergence carried by |βk|2|\beta_{k}|^{2} is most likely due to a breakdown of the particle interpretation of the out-Fock space. Recall that |βk|2|\beta_{k}|^{2} is interpreted as the mean number of gravitons in mode kk in the out-region. However, it is not clear that graviton number is meaningful for modes whose wavelength is far greater than the horizon size. We propose to modify Eq. (44) by restricting the integration to modes with k>k0k>k_{0}, where k0k_{0} is the wave number associated with the horizon size at the time when the gravitons interact with the CMB photons. Thus, Eq. (44) is replaced by

C(𝐫,η,η)=H4P216π2a(η)a(η)k0dkk3dΩksin4χcos[(𝐤𝐫k(ηη)].C({\bf r},\eta,\eta^{\prime})=\frac{H^{4}\,\ell_{P}^{2}}{16\,\pi^{2}\,a(\eta)\,a(\eta^{\prime})}\;\int^{\infty}_{k_{0}}\frac{dk}{k^{3}}\,\int d\Omega_{k}\,\sin^{4}\chi\,\cos[({\bf k}\cdot{\bf r}-k\,(\eta-\eta^{\prime})]\,. (45)

The angles involved in the dΩkd\Omega_{k} integration are illustrated in Fig. 4. First note that cosχ=𝐤^𝐪^=cosθcosϕ\cos\chi=\hat{\bf k}\cdot\hat{\bf q}=\cos\theta\,\cos\phi. Thus sin2χ=1(1c2)cos2ϕ\sin^{2}\chi=1-(1-c^{2})\,\cos^{2}\phi, with c=cosθc=\cos\theta and we have

02πsin4χdϕ=π4(3+2c2+3c4).\int_{0}^{2\pi}\sin^{4}\chi\,d\phi=\frac{\pi}{4}\,(3+2c^{2}+3c^{4}). (46)
Refer to caption
Figure 4: The various vectors appearing in Eq. (45) are illustrated in a particular coordinate system. The separation vector 𝐫{\mathbf{r}} between the two space points lies along the zz-axis. The photon wave vector 𝐪{\mathbf{q}} lies along the xx-axis, which is the direction from the point of emission of the photon to the observer. The graviton wave vector 𝐤{\mathbf{k}} lies along a direction given by the polar angle θ\theta and the azimuthal angle ϕ\phi.

Now we can write

C(𝐫,η,η)=H4P215πa(η)a(η)k02K(Δη,r),C({\bf r},\eta,\eta^{\prime})=\frac{H^{4}\,\ell_{P}^{2}}{15\,\pi\,a(\eta)\,a(\eta^{\prime})\,k_{0}^{2}}\;K(\Delta\eta,r)\,, (47)

where we define a reduced correlation function by

K(Δη,r)=156411𝑑c(3+2c2+3c4)1duu3cos[k0(crΔη)u],K(\Delta\eta,r)=\frac{15}{64}\,\int_{-1}^{1}dc\,(3+2c^{2}+3c^{4})\int_{1}^{\infty}\frac{du}{u^{3}}\,\cos[k_{0}(c\,r-\Delta\eta)\,u]\,, (48)

which is normalized by K(0,0)=1K(0,0)=1. The reduced correlation function may be expressed in terms of the functions CnC_{n} and SnS_{n} are defined by

Cn(α)=1duuncos(αu),C_{n}(\alpha)=\int_{1}^{\infty}\frac{du}{u^{n}}\,\cos(\alpha\,u)\,, (49)

and

Sn(α)=1duunsin(αu).S_{n}(\alpha)=\int_{1}^{\infty}\frac{du}{u^{n}}\,\sin(\alpha\,u)\,. (50)

These functions may be expressed in terms of trigonometric and cosine-integral functions and are given in the Appendix.

V.2 Results: Case 𝐫=0{\bf r}=0

In this subsection, we deal with the simpler case where 𝐫=0{\bf r}=0, but Δη=ηη\Delta\eta=\eta-\eta^{\prime} is nonzero. This is the case where the CMB photons arrive from the same direction, but were emitted at different times and hence different redshifts. In this case there is no angle dependence inside the cosine function in Eq. (48), and we find

K(k0Δη,0)=2C3(k0Δη),K(k_{0}\,\Delta\eta,0)=2\,C_{3}(k_{0}\,\Delta\eta)\,, (51)

which is plotted in Fig. 5. This function describes the temperature fluctuations correlations for photons arriving from the same direction, but different redshifts, with the pre-factor of 1/[a(η)a(η)]1/[a(\eta)\,a(\eta^{\prime})] in Eq. (45) removed.

Note that the region in which the inflationary gravitons interact with the CMB photons is actually a last scattering shell of finite thickness. This thickness is determined by the time required for the electrons and protons in the primordial plasma to combine into neutral hydrogen atoms. It has been estimated by Hadzhiyska and Spergel HS19 to be of order Δη19Mpc\Delta\eta\approx 19\,{\rm Mpc}, which is equivalent to Δz90\Delta z\approx 90 in redshift.

Refer to caption
Figure 5: The reduced correlation function, K(k0Δη,0)K(k_{0}\,\Delta\eta,0), for coincident spatial points is plotted. This gives the temperature correlations, without the effects of the scale factor changes, between regions at different redshifts, but the same spatial location. Note that regions separated by k0Δη2k_{0}\,\Delta\eta\approx 2 are strongly anti-correlated.
Refer to caption
Figure 6: This plot shows the temperature fluctuation correlations between regions at the same redshift, but different spatial locations and seen at different angles. Here regions separated by k0Δr3k_{0}\,\Delta r\approx 3 are somewhat anti-correlated.

V.3 Results: Case 𝐫0{\bf r}\not=0

When 𝐫0{\bf r}\not=0, the argument of the cosine function in Eq. (45) depends upon θ\theta. In this case, we have

02π𝑑ϕ11𝑑csin4χcos[k(crΔη)]=4πk5r5cos(kΔη)[kr(2k2r29)cos(kr)+(95k2r2+k4r4)sin(kr)].\int_{0}^{2\pi}d\phi\int_{-1}^{1}dc\,\sin^{4}\chi\,\cos[k(cr-\Delta\eta)]=\frac{4\pi}{k^{5}\,r^{5}}\,\cos(k\Delta\eta)\,[kr(2k^{2}r^{2}-9)\cos(kr)+(9-5k^{2}r^{2}+k^{4}r^{4})\sin(kr)]\,. (52)

We may now write Eq. (48) as

K(Δη,r)\displaystyle K(\Delta\eta,r) =\displaystyle= 158{2ρ2[C5(ρ+τ)+C5(ρτ)]9ρ4[C7(ρ+τ)+C7(ρτ)]\displaystyle\frac{15}{8}\;\biggl{\{}\frac{2}{\rho^{2}}\,[C_{5}(\rho+\tau)+C_{5}(\rho-\tau)]-\frac{9}{\rho^{4}}\,[C_{7}(\rho+\tau)+C_{7}(\rho-\tau)] (53)
+\displaystyle+ 9ρ5[S8(ρ+τ)+S8(ρτ)]5ρ3[S6(ρ+τ)+S6(ρτ)]+1ρ[S4(ρ+τ)+S4(ρτ)]}.\displaystyle\frac{9}{\rho^{5}}\,[S_{8}(\rho+\tau)+S_{8}(\rho-\tau)]-\frac{5}{\rho^{3}}\,[S_{6}(\rho+\tau)+S_{6}(\rho-\tau)]+\frac{1}{\rho}\,[S_{4}(\rho+\tau)+S_{4}(\rho-\tau)]\biggr{\}}\,.

with τ=k0Δη\tau=k_{0}\,\Delta\eta and ρ=k0r\rho=k_{0}\,r.

The case of spatial separation alone, K(0,r)K(0,r) is plotted in Fig. 6, and the general case is illustrated in Fig. 7.

Refer to caption
Figure 7: This plot illustrates the relative degrees of correlation and anti-correlation as a function of both spatial and temporal separation.

V.4 Physical Meaning of the Correlations

We have seen in the previous subsection that the temperature fluctuations produced by a graviton bath display distinctive correlations and anti-correlations, both in space and in time. The anti-correlations mean that a region whose temperature is above average, is more likely to be near a region which is cooler than average, and vice versa. The separation between the hotter and the cooler region can be either in angle or in redshift. The magnitude of the expected angular or redshift depends upon the unknown parameter k0k_{0}, and is given by k0Δr3k_{0}\,\Delta r\approx 3 for spatial or angular separations and by k0Δη2k_{0}\,\Delta\eta\approx 2 for temporal or redshift separations. Note that the ratio of these two expected separations, Δr/Δη1.5\Delta r/\Delta\eta\approx 1.5, is independent of k0k_{0}. As was emphasized in Sec. III.4 and illustrated in Fig. 3, correlation functions are averages of highly fluctuating quantities. This means that the temperature anti-correlations can only be observed by averaging over several pairs of regions.

VI Summary

We have presented a description of inflationary tensor perturbations in terms of gravitons created by quantum particle creation at the end of inflation. These gravitons are in a multi-mode squeezed vacuum state describing pairs of created gravitons. However, only one member of each pair is observable, and the resulting quantum state exhibits large fluctuations and is equivalent to a density matrix . Our description introduces two infrared cutoffs. The first is the parameter kck_{c} introduced in Sect. II.1, and depends upon initial conditions before the onset of inflation. This parameter is expected to be very small and its effects to be unobservable. The second cutoff is the parameter k0k_{0} introduced in Sect. V.1, and determined by the range of validity of the concept of graviton number in the out-region. It is an undetermined in our analysis, but is expected to be determined by the horizon size at the time of last scattering. We have followed the approach of previous authors, and calculated the CMB temperature anisotropies from solutions of the photon geodesic equations. This leads to the temperature fluctuation correlation function in terms of the graviton correlation function, Eq. (43). The cutoff k0k_{0} is used to make these expressions infrared finite.

An alternative approach may be to use the geodesic deviation equation to study the perturbations of the photon geodesics. This would replace the correlation function for the graviton field by that for the associated Riemann tensor. The letter contains additional derivatives which are expected to render it infrared finite. This approach will be the topic for future research.

Within our present approach, we are able to obtain an explicit expression, Eq. (53), as a function of the temporal separation Δη\Delta\eta and spatial separation rr of a pair of emission points. Recall that rr is equivalent to an angular separation on the sky, and Δη\Delta\eta is equivalent to a redshift separtion within the last scattering shell. The plots of this function in Figs. 5, 6, and 7, reveal significant anti-correlations in both angle and redshift. This means that a hot region will tend to be associated with a nearby cooler region, and vice versa. As was emphasized in Sect. III.4, these anti-correlations will appear only in averages over many pairs of emission points. However, if the anti-correlations can be observed, it will provide strong evidence for the existence of inflationary gravitons.

Acknowledgements.
We would like to thank J.T. Hsiang, B.L. Hu, and K.W. Ng for helpful discussions. LHF would like to thank the Department of Physics, Soochow University for hospitality while part of this work was performed. This work was supported in part by the National Science Foundation under Grant PHY-2114024.

Appendix A Integrals Appearing in the Correlation Functions

Here we give the explicit forms for some of the integrals which appear in the correlation function, Eq. (53). For nn odd, Cn(α)C_{n}(\alpha), defined in Eq. (49), is given by (See Formula 3.761.8 in Ref, GR .)

C2m+1(α)=α2m(2m)!{k=12m(2mk)!α2mk+1cos[α+(k1)π2]+(1)m+1ci(α)},C_{2m+1}(\alpha)=\frac{\alpha^{2m}}{(2m)!}\,\left\{\sum_{k=1}^{2m}\frac{(2m-k)!}{\alpha^{2m-k+1}}\,\cos[\alpha+(k-1)\frac{\pi}{2}]+(-1)^{m+1}\,{\rm ci}(\alpha)\right\}\,, (54)

where ci(α)=α𝑑tcos(t)/t{\rm ci}(\alpha)=-\int_{\alpha}^{\infty}dt\,\cos(t)/t is the cosine integral function. The special cases of interest to us are the following:

C3(α)=12[cos(α)αsin(α)+α2ci(α)],C_{3}(\alpha)=\frac{1}{2}\,\left[\cos(\alpha)-\alpha\,\sin(\alpha)+\alpha^{2}\,{\rm ci}(\alpha)\right]\,, (55)
C5(α)=14![3!cos(α)2αsin(α)α2cos(α)+α3sin(α)α4ci(α)],C_{5}(\alpha)=\frac{1}{4!}\,\left[3!\,\cos(\alpha)-2\alpha\,\sin(\alpha)-\alpha^{2}\,\cos(\alpha)+\alpha^{3}\,\sin(\alpha)-\alpha^{4}\,{\rm ci}(\alpha)\right]\,, (56)

and

C7(α)=16![5!cos(α)4!αsin(α)3!α2cos(α)+2α3sin(α)+α4cos(α)α5sin(α)+α6ci(α)].C_{7}(\alpha)=\frac{1}{6!}\,\left[5!\,\cos(\alpha)-4!\alpha\,\sin(\alpha)-3!\,\alpha^{2}\,\cos(\alpha)+2\,\alpha^{3}\,\sin(\alpha)+\alpha^{4}\,\cos(\alpha)-\alpha^{5}\,\sin(\alpha)+\alpha^{6}\,{\rm ci}(\alpha)\right]\,. (57)

For nn even, Sn(α)S_{n}(\alpha), defined in Eq. (50), is given by (See Formula 3.761.3 in Ref, GR .)

S2m1(α)=α2m1(2m1)!{k=12m1(2mk)!α2mksin[α+(k1)π2]+(1)mci(α)},S_{2m1}(\alpha)=\frac{\alpha^{2m-1}}{(2m-1)!}\,\left\{\sum_{k=1}^{2m-1}\frac{(2m-k)!}{\alpha^{2m-k}}\,\sin[\alpha+(k-1)\frac{\pi}{2}]+(-1)^{m}\,{\rm ci}(\alpha)\right\}\,, (58)

which leads to

S4(α)=16[2sin(α)+αcos(α)α2sin(α)+α3ci(α)],S_{4}(\alpha)=\frac{1}{6}\,\left[2\,\sin(\alpha)+\alpha\,\cos(\alpha)-\alpha^{2}\,\sin(\alpha)+\alpha^{3}\,{\rm ci}(\alpha)\right]\,, (59)
S6(α)=15![4!sin(α)+3!αcos(α)2α2sin(α)α3cos(α)+α4sin(α)α5ci(α)],S_{6}(\alpha)=\frac{1}{5!}\,\left[4!\,\sin(\alpha)+3!\,\alpha\,\cos(\alpha)-2\,\alpha^{2}\,\sin(\alpha)-\alpha^{3}\,\cos(\alpha)+\alpha^{4}\,\sin(\alpha)-\alpha^{5}\,{\rm ci}(\alpha)\right]\,, (60)

and

S8(α)=17![6!sin(α)+5!αcos(α)4!α2sin(α)3!α3cos(α)+2α4sin(α)+α5cos(α)α6sin(α)+α7ci(α)].S_{8}(\alpha)=\frac{1}{7!}\,\left[6!\,\sin(\alpha)+5!\,\alpha\,\cos(\alpha)-4!\,\alpha^{2}\,\sin(\alpha)-3!\,\alpha^{3}\,\cos(\alpha)+2\,\alpha^{4}\,\sin(\alpha)+\alpha^{5}\,\cos(\alpha)-\alpha^{6}\,\sin(\alpha)+\alpha^{7}\,{\rm ci}(\alpha)\right]\,. (61)

References

  • (1) A.A. Starobinsky, Spectrum of relict gravitational waves and the early universe, Pis’ma Zh. Teor. Fiz. 30, 719-723 (1979) [JETP Lett. 30, 682-685 (1979)].
  • (2) V.A. Rubakov, M.V. Sazhin, and A.V. Veryaskin, Graviton creation in the inflationary universe and the grand unification scale, Phys. Lett. 115B, 189-192 (1982).
  • (3) F. Fabbri and M.D. Pollack, The effect of primordially produced gravitons upon the anisotropy of the cosmological microwave background radiation, Phys. Lett. 125B, 445-448 (1983).
  • (4) L.F. Abbott and M.B. Wise, Constraints on generalized inflationary cosmologies, Nucl. Phys. B244, 641-548 (1984).
  • (5) L.F. Abbott and M.B. Wise, Anisotropy of the microwave background in the inflationary cosmology, Phys. Lett. 135B, 279-282 (1984).
  • (6) L.F. Abbott and Harari, Graviton production in inflationary cosmology, Nucl. Phys. B264, 487-492 (1986).
  • (7) A.G. Polnarev, Sov. Astron. 29, 607 (1986).
  • (8) B. Allen, Stochastic gravity-wave background in inflationary-universe models, Phys. Rev. D 37, 2078-2085 (1988).
  • (9) M. Kamionkowski and E.D. Kovetz, The Quest for B Modes from Inflationary Gravitational Waves, Annu. Rev, Aston. Astrophys. 54, 227-69 (2016).
  • (10) L. Parker, Quantized fields and particle creation in expanding Universes. I, Phys. Rev. 183, 1057 (1969).
  • (11) L. P. Grishchuk, Amplification of gravitational waves in an isotropic universe, Zh. Eksp. Teor. Fiz. 67, 825 (1974) [Sov. Phys. JETP 40, 409 (1975)].
  • (12) L. H. Ford and L. Parker, Quantized gravitational wave perturbations in Robertson-Walker universes, Phys. Rev. D 16, 1601 (1977).
  • (13) L.H. Ford, Cosmological Particle Production: A Review, Rep. Prog. Phys. 84, 116901 (2021) , arXiv:2112.02444.
  • (14) L. P. Grishchuk and Y.V. Sidorov, On the quantum state of relic gravitons, Class. Quantum Grav. 6, L161 (1989).
  • (15) D. Polarski and A. A. Starobinsky, Semiclassicality and decoherence of cosmological perturbations, Class. Quantum Grav. 13, 377 (1996)
  • (16) J-O Gong and M-S Seo, Quantum non-linear evolution of inflationary tensor perturbations, JHEP 2019, 21 (2019), arXiv:1903.12295
  • (17) S. Kanno, J. Soda and K. Ueda, Conversion of squeezed gravitons into photons during inflation, Phys. Rev. D 105, 083508 (2022), arXiv:2207.05734.
  • (18) C. P. Burgess, R. Holman, G. Kaplanek, J.Martin, and V. Vennin, Minimal decoherence from inflation, JCAP 07 (2023) 022 [ arXiv:2211.11046 ]
  • (19) S. Ning. C. M. Soub, and Y. Wangb, On the decoherence of primordial gravitons, JHEP06(2023)101
  • (20) E. M. Lifshitz, On the Gravitational Stability of the Expanding Universe, Zh. Eksp. Teor. Fiz. 16, 587 (1946) [J. Phys. USSR 10, 116 (1946)].
  • (21) T.S. Bunch and P.C.W. Davies, Quantum field theory in de Sitter space - Renormalization by point-splitting, Proc. Roy Soc. London 360, 117-134 (1978).
  • (22) L. H. Ford and L. Parker, Infrared divergences in a class of Robertson-Walker universes, Phys. Rev. D 16, 245 (1977).
  • (23) L.H. Ford and A. Vilenkin, Global Symmetry Breaking in Two-Dimensional Flat Spacetime and in de Sitter Spacetime, Phys. Rev. D 33, 2833-2839 (1986).
  • (24) M. Visser, Efficient computation of null affine parameters , Universe 9, 521 (2023) [arXiv:2211.07835].
  • (25) I. S. Gradshteyn and I. M. Ryzhik, Table of integrals, series, and products, 7th ed. Academic Press (2007).
  • (26) B. Hadzhiyska and D. Spergel, Phys. Rev. D 99, 043537 (2019) [arXiv:1808.04083].