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

Can Gibbons-Hawking Radiation and Inflation Arise Due to Spacetime Quanta?

Naouel Boulkaboul Independent Researcher

I INTRODUCTION

Black hole radiation and inflation are ones of the most intriguing aspects in cosmology. The first one being theorized by Stephen Hawking [1], has been attracted a considerable interest [2][7]\cite[cite]{[\@@bibref{}{2}{}{}]}-\cite[cite]{[\@@bibref{}{7}{}{}]}. Along with Unruh radiation that is “felt” by an accelerated observer[8][11]\cite[cite]{[\@@bibref{}{8}{}{}]}-\cite[cite]{[\@@bibref{}{11}{}{}]}, Hawking radiation was originally derived based on Bogoliubov’s method [12], ever since many approaches [14][18]\cite[cite]{[\@@bibref{}{13}{}{}]}-\cite[cite]{[\@@bibref{}{17}{}{}]} including the complex path (or Hamilton-Jacobi) method [19]-[20] have been developed to derive such a radiation. Hawking radiation is described as a tunneling effect of particles across the black hole’s horizon [13]. In fact, such a thermal radiation is related to any geometrical background that possesses a horizon (i.e. the cosmological horizon of a black hole, de Sitter space and even the one of an accelerated observer).

The hypothesis of inflation, on the other hand, has been postulated by Alan Guth [21][22]\cite[cite]{[\@@bibref{}{20}{}{}]}-\cite[cite]{[\@@bibref{}{21}{}{}]}. His model relies on the assumption that the very early Universe has gone through a period of accelerated expansion that preceded the standard radiation-dominated era. Such a period of accelerated expansion offers a physical explanation of the cosmology’s biggest puzzles that the standard cosmological scenario cannot explain. Inflation drives any initially curved spacetime towards the spatial flatness observed today, hence answering the question: “Why would the universe be perfectly spatially flat?”, it brings together all causally disconnected regions and extends the causal horizon beyond the present Hubble length, in such a away to answer the question: “Why would the universe have the same temperature everywhere?” and it also brings a satisfactory solution to the magnetic monopoles problem, answering the question: “Why are there no leftover high-energy relics?” To date, many simple as well as complex inflationary models have been proposed [23][30]\cite[cite]{[\@@bibref{}{22}{}{}]}-\cite[cite]{[\@@bibref{}{29}{}{}]}. In their simplest picture, inflationary models are based on a scalar field (inflaton field) with a fine-tuning potential, i.e., a flat potential. The latter is constrained by slow roll conditions, i.e., in order to trigger inflation the scalar field’s potential energy must dominate over its kinetic and gradient energy that can prevent its starting. This fine-tuning requirement puts the viability of such inflationary models into question.

To contribute to the ongoing literature on the two topics, firstly we address the Gibbons-Hawking radiation [31][32]\cite[cite]{[\@@bibref{}{30}{}{}]}-\cite[cite]{[\@@bibref{}{31}{}{}]} in anti-de Sitter space by incorporating a quantized spacetime. We want to stress, however, that our contribution is to show that this radiation has nothing to do with the tunnelling particle but it is due to the space curved geometry, precisely its quantum nature. Secondly we propose, inspired by the Tsallis q-formalism [33][35]\cite[cite]{[\@@bibref{}{32}{}{}]}-\cite[cite]{[\@@bibref{}{34}{}{}]}, a new inflationary model, with a minimum of fine-tuning, that we will call the q-inlfation. The latter, which is devoid of singularities, is triggered implicitly by the spacetime quanta but invokes the contribution of the “cosmological constant” carried by the quanta, explicitly. We believe in the importance of the present work because it kills two birds (Gibbons-Hawking radiation and inflation) with one stone (spacetime quantization). Throughout the paper, we consider the metric signature ++---. Moreover, the units are chosen with c==1c=\hbar=1.

II THE ORIGIN OF GIBBONS-HAWKING RADIATION

In this section we derive the well-known Gibbons-Hawking temperature [31][32]\cite[cite]{[\@@bibref{}{30}{}{}]}-\cite[cite]{[\@@bibref{}{31}{}{}]} by means of a spacetime quantization framework firstly introduced by L.C. Céleri et al. [36]-[37]. Their study, in which they show that Unruh effect can be obtained without changing the reference frame, focuses on deriving Unruh radiation in the accelerated fields scenario. For our purpose however, let us consider the two-sheet hyperboloid (see Fig.1) governed by the parameterization

x0=l2+r2cosh(t/l),x_{0}=\sqrt{l^{2}+r^{2}}cosh(t/l), (1)
x1=l2+r2sinh(t/l),x_{1}=\sqrt{l^{2}+r^{2}}sinh(t/l), (2)
xi=rzi,2i4.x_{i}=rz_{i},\qquad 2\leq i\leq 4. (3)

where ziz_{i} gives the standard embedding of the 2-sphere in R3\textbf{R}^{3}. Embedding this in the five-dimensional Minkowski metric

dsM2=dx02μ=14dxμdxμ,ds^{2}_{M}=dx_{0}^{2}-\sum\limits_{\mu=1}^{4}dx_{\mu}dx^{\mu}, (4)

yields

ds2=(1+r2l2)dt211+r2l2dr2r2[dθ2+sin2θdϕ2],ds^{2}=-\big{(}1+\frac{r^{2}}{l^{2}}\big{)}dt^{2}-\frac{1}{1+\frac{r^{2}}{l^{2}}}dr^{2}-r^{2}[d\theta^{2}+sin^{2}\theta d\phi^{2}], (5)

One may notice the strange signature ()(----) of the metric, one common way to reproduce the metric with the appropriate Lorentzian signature, is making use of the Wick rotation titt\to it. For our convenience however we will use the real Wick rotation defined in [38] (and references therein), proven to be more adequate especially for curved spacetimes. Hence, we set the above line element as

ds2=η00dt2+ημνdXμdXν,μ,ν=1,2,3ds^{2}=\eta_{00}dt^{2}+\eta_{\mu\nu}dX^{\mu}dX^{\nu},\qquad\mu,\nu=1,2,3 (6)

which means “singling out” a proper time tt, where η00=(1+r2/l2)\eta_{00}=-\big{(}1+{r^{2}}/{l^{2}}\big{)}, X1=r,X2=θ,X3=ϕX^{1}=r,X^{2}=\theta,X^{3}=\phi and the corresponding components of the metric tensor ημν\eta_{\mu\nu} are η11=1/(1+r2/l2),η22=r2,η33=r2sin2θ\eta_{11}=-1/({1+{r^{2}}/{l^{2}}}),\eta_{22}=-r^{2},\eta_{33}=-r^{2}sin^{2}\theta.
Anti-de Sitter metric with Lorentzian signature is then recovered under mapping the 00-component of the metric as η00η00\eta_{00}\mapsto-\eta_{00} while keeping ημν\eta_{\mu\nu} unchanged, thus we get

dsAdS2=(1+r2l2)dt211+r2l2dr2r2[dθ2+sin2θdϕ2],ds_{AdS}^{2}=\big{(}1+\frac{r^{2}}{l^{2}}\big{)}dt^{2}-\frac{1}{1+\frac{r^{2}}{l^{2}}}dr^{2}-r^{2}[d\theta^{2}+sin^{2}\theta d\phi^{2}], (7)

Note that, unlike de Sitter metric, this metric doesn’t possess a cosmological horizon. In order to quantize spacetime we follow the formulation of accelerated quantum field theory proposed by L.C. Céleri et al. [37]. This can be achieved starting from the four-dimensional two-sheet hyperboloid’s equation

x02μ=14xμxμ=l2,x^{2}_{0}-\sum\limits_{\mu=1}^{4}x_{\mu}x^{\mu}=l^{2}, (8)

where x0.x4x_{0}....x_{4} are the Cartesian coordinates in Minkowski space in which Anti-de Sitter space is embedded, and ll is a nonzero constant with dimensions of length (the radius of the hyperboloid’s curvature).

Now we can introduce the time and position operators given by

x0=iω=iω,xμ=ikμ=ikμ,x_{0}=i\frac{\partial}{\partial\omega}=i\partial_{\omega},\qquad x_{\mu}=i\frac{\partial}{\partial{k^{\mu}}}=i\partial_{k^{\mu}}, (9)

Making use of Eq. 8, a wave equation similar to the Klein-Gordon equation can be obtained

[ω2kμ2+l2]φ(ω,kμ)=0,\bigg{[}\partial^{2}_{\omega}-\partial^{2}_{{{k^{\mu}}}}+l^{2}\bigg{]}{\varphi}({\omega,k^{\mu}})=0, (10)

with (ω,kμ)(\omega,k^{\mu}) being the five momentum. We can now construct the Lagrange density by going backward from the Euler-Lagrange equation 10

=12[(ωφ)2(kμφ)2l2φ2],\mathcal{L}=\frac{1}{2}\bigg{[}(\partial_{\omega}\varphi)^{2}-(\partial_{{{k^{\mu}}}}\varphi)^{2}-l^{2}\varphi^{2}\bigg{]}, (11)

The corresponding Hamiltonian is then given by

H~=12(kμφ)2+12Π2+12l2φ2,\tilde{H}=\frac{1}{2}(\partial_{k^{\mu}}\varphi)^{2}+\frac{1}{2}\Pi^{2}+\frac{1}{2}l^{2}\varphi^{2}, (12)

where Π(ω,kμ)\Pi(\omega,k^{\mu}) is the conjugate momentum for φ(ω,kμ)\varphi(\omega,k^{\mu}) defined as

Π(ω,kμ)=ωφ(ω,kμ),\Pi({\omega,k^{\mu}})=\partial_{\omega}\varphi({\omega,k^{\mu}}), (13)

In what follows we will use the notation x and k for our space and momentum vector, respectively. The equation 10 has a general solution of the form

φ(ω,k)=d4x[axux(ω,k)+axux(ω,k)],\varphi(\omega,\textbf{\emph{k}})=\int d^{4}{\emph{x}}\bigg{[}a_{\textbf{\emph{x}}}u_{\textbf{\emph{x}}}(\omega,\textbf{\emph{k}})+a^{\dagger}_{\textbf{\emph{x}}}u^{*}_{\textbf{\emph{x}}}(\omega,\textbf{\emph{k}})\bigg{]}, (14)

By letting the canonical equal-energy commutation relations be

[φ(ω,k),φ(ω,k)]=[Π(ω,k),Π(ω,k)]=0,[\varphi(\omega,{\textbf{\emph{k}}}),\varphi(\omega,{\textbf{\emph{k}}}^{\prime})]=[\Pi(\omega,{\textbf{\emph{k}}}),\Pi(\omega,{\textbf{\emph{k}}}^{\prime})]=0, (15)
[φ(ω,k),Π(ω,k)]=iδ(kk),[\varphi(\omega,{\textbf{\emph{k}}}),\Pi(\omega,{\textbf{\emph{k}}}^{\prime})]=i\delta({\textbf{\emph{k}}}-{\textbf{\emph{k}}}^{\prime}), (16)

the creation and annihilation operators satisfy

[ax,ax]=δ(xx),[ax,ax]=[ax,ax]=0,[a_{\textbf{\emph{x}}},a^{\dagger}_{\textbf{\emph{x}}^{\prime}}]=\delta({\textbf{\emph{x}}}-{\textbf{\emph{x}}}^{\prime}),\enspace[a_{\textbf{\emph{x}}},a_{\textbf{\emph{x}}^{\prime}}]=[a^{\dagger}_{\textbf{\emph{x}}},a^{\dagger}_{\textbf{\emph{x}}^{\prime}}]=0, (17)
Refer to caption
Figure 1: Parameterization of two-sheet hyperboloid in the coordinates 13\ref{static coord1}-\ref{static coord3}. The x3{x}_{3} and x4{x}_{4} axes are suppressed.

Thus, the scalar field operator φ(ω,k)\varphi(\omega,\textbf{\emph{k}}) in momentum space, can be represented in the form

φ(ω,k)=1(2π)2x0d4x[axei(k.xωx0)+axei(k.xωx0)],\varphi(\omega,\textbf{\emph{k}})=\frac{1}{(2\pi)^{2}\sqrt{x_{0}}}\int d^{4}{\emph{x}}\bigg{[}a_{\textbf{\emph{x}}}e^{i(\textbf{\emph{k}}.\textbf{\emph{x}}-\omega x_{0})}+a^{\dagger}_{\textbf{\emph{x}}}e^{-i(\textbf{\emph{k}}.\textbf{\emph{x}}-\omega x_{0})}\bigg{]}, (18)

ux=ei(k.xωx0)u_{\textbf{\emph{x}}}=e^{i(\textbf{\emph{k}}.\textbf{\emph{x}}-\omega x_{0})} and ux=ei(k.xωx0)u^{*}_{\textbf{\emph{x}}}=e^{-i(\textbf{\emph{k}}.\textbf{\emph{x}}-\omega x_{0})} are identified as positive and negative “frequency” solutions, respectively. That is, the “frequency” is defined as x0=x2+l2x_{0}=\sqrt{{\textbf{\emph{x}}}^{2}+l^{2}}. axa_{\textbf{\emph{x}}} and axa^{\dagger}_{\textbf{\emph{x}}} are, respectively, the annihilation and creation operators that annihilates and creates excitations at spacetime point (x0,x)(x_{0},{\textbf{\emph{x}}}). The associated vacuum states |0>|0\big{>} are defined by ax|0>=0a_{{\textbf{\emph{x}}}}|0\big{>}=0 and ax|0>=|x>a^{\dagger}_{{\textbf{\emph{x}}}}|0\big{>}=|{\textbf{\emph{x}}}\big{>}, respectively. The states ax|0>a^{\dagger}_{{\textbf{\emph{x}}}}|0\big{>} are interpreted as single particle states with time x0x_{0} and position x. One assumes, as seems reasonable since the wave equation is defined in momentum space, that the excitation “carries” a specific x0x_{0} and x. Moreover, it is not difficult to notice that the excitations are located on the hyperboloid’s upper sheet. This can be interpreted in the following way, “spacetime excitations” carrying x0>0x_{0}>0 constitute the hyperboloid’s upper sheet.

Establishing the theory in momentum space first, we shall now move on to coordinate space. To this end, we introduce a field χAdS\chi^{{AdS}} that we will call Anti-de Sitter field. The latter can be expanded in the basis {ux,ux}\{u_{\textbf{\emph{x}}},u^{*}_{\textbf{\emph{x}}}\} as

χAdS(ω,k)=d4x[ψxux(ω,k)+ψxux(ω,k)],\chi^{{AdS}}(\omega,{\textbf{\emph{k}}})=\int d^{4}x\big{[}\psi_{\textbf{\emph{x}}}u_{\textbf{\emph{x}}}(\omega,{\textbf{\emph{k}}})+\psi^{\dagger}_{\textbf{\emph{x}}}u^{*}_{\textbf{\emph{x}}}(\omega,{\textbf{\emph{k}}})\big{]}, (19)

where ψ(τ,x)=d4k[akϕk(τ,x)+(ak)ϕk(τ,x)]\psi(\tau,{\textbf{\emph{x}}})=\int d^{4}k\big{[}a_{\textbf{\emph{k}}}\phi_{\textbf{\emph{k}}}(\tau,{\textbf{\emph{x}}})+\big{(}a_{\textbf{\emph{k}}}\big{)}^{\dagger}\phi^{*}_{\textbf{\emph{k}}}(\tau,{\textbf{\emph{x}}})\big{]} is a Klein-Gordon (K-G) field operator that satisfies the canonical commutation relations [ψ(τ,x),ψ(τ,x’)]=δ(xx’)[\psi(\tau,{\textbf{\emph{x}}}),\psi^{\dagger}(\tau,{\textbf{\emph{x'}}})]=\delta({\textbf{\emph{x}}}-{\textbf{\emph{x'}}}) and [ψ(τ,x),Π(τ,x)]=iδ(xx)[\psi(\tau,{\textbf{\emph{x}}}),\Pi(\tau,{\textbf{\emph{x}}}^{\prime})]=i\delta({\textbf{\emph{x}}}-{\textbf{\emph{x}}}^{\prime}), with (τ,x\tau,{\textbf{\emph{x}}}) being the usual Minkowski coordinates. On one hand, the operators ψx\psi_{{\textbf{\emph{x}}}} and ψx\psi^{\dagger}_{\textbf{\emph{x}}} act as Fourier coefficients in the field operator χAdS\chi^{{AdS}}’s expansion; and on the other hand, they act as Klein-Gordon annihilation and creation operators that annihilates/creates K-G particles at (x0,x)(x_{0},{\textbf{\emph{x}}}). Note that the field ζAdS(x0,x)=ψxux(ω,k)+ψxux(ω,k)\zeta^{{AdS}}(x_{0},{\textbf{\emph{x}}})=\psi_{{\textbf{\emph{x}}}}u_{{\textbf{\emph{x}}}}(\omega,{\textbf{\emph{k}}})+\psi^{\dagger}_{\textbf{\emph{x}}}u^{*}_{\textbf{\emph{x}}}(\omega,{\textbf{\emph{k}}}) obeys the Klein-Gordon equation.

A more plausible way to write down the field ζAdS\zeta^{{AdS}} is to expand it in terms of the pair {ϕk,ϕk}τx0\{\phi_{\textbf{\emph{k}}},\phi^{*}_{\textbf{\emph{k}}}\}_{{}_{\tau\to x_{0}}}

ζAdS(x0,x)=d4k[akAdSϕk(x0,x)+(akAdS)ϕk(x0,x)],\zeta^{{AdS}}(x_{0},{\textbf{\emph{x}}})=\int d^{4}k\big{[}a^{{}^{AdS}}_{\textbf{\emph{k}}}\phi_{\textbf{\emph{k}}}(x_{0},{\textbf{\emph{x}}})+\big{(}a^{{}^{AdS}}_{\textbf{\emph{k}}}\big{)}^{\dagger}\phi^{*}_{\textbf{\emph{k}}}(x_{0},{\textbf{\emph{x}}})\big{]}, (20)

akAdSa_{\textbf{\emph{k}}}^{{}^{AdS}} and (akAdS)\big{(}a^{{}^{AdS}}_{\textbf{\emph{k}}}\big{)}^{\dagger} are annihilation and creation operators with respect to the modes ϕk\phi_{\textbf{\emph{k}}} and ϕk\phi^{*}_{\textbf{\emph{k}}}. The corresponding vacuum states are then defined as akAdS|0>AdS=0a_{\textbf{\emph{k}}}^{{}^{AdS}}|0\big{>}_{AdS}=0 and (akAdS)=|k>AdS(a_{\textbf{\emph{k}}}^{{}^{AdS}})^{\dagger}=|{\textbf{\emph{k}}}\big{>}_{AdS}, respectively. Note that the excitations associated to (akAdS)|0>AdS(a_{\textbf{\emph{k}}}^{{}^{AdS}})^{\dagger}|0\big{>}_{AdS} carry an energy ω\omega while those corresponding to ζAdS|0>AdS\zeta^{{AdS}}|0\big{>}_{AdS} are located at x0=x2+l2x_{0}=\sqrt{{\textbf{\emph{x}}}^{2}+l^{2}}. Thus, the field operator is associated with single particles with “length” ll and mass mm in case where the field is massive, i.e., ω=k2+m2\omega=\sqrt{{\textbf{\emph{k}}}^{2}+m^{2}}. Given that the field operators in momentum space are related to those in coordinate space by Fourier transform, one may intuitively assume that excitations carrying a specific ω\omega and k in coordinate space correspond to excitations that carry a specific x0x_{0} and x in momentum space. Thus, we may use the terminologies particles and spacetime quanta interchangeably.

For more details on the quantization procedure, the reader is referred to Refs.[36]-[37]. Since we are interested in Gibbons-Hawking temperature, the thermal effect of anti-de Sitter hyperbolic geometry was computed by means of the field correlation function, following mostly Ref. [37]. Thus, using Eq. 20 the two-time correlation function <ζAdS(x0)ζAdS(x0)>0AdS{{\big{<}\zeta^{{AdS}}(x_{0})\zeta^{{AdS}}(x^{\prime}_{0})\big{>}}_{0_{AdS}}} for a massless field in the vacuum state |0>AdS|0\big{>}_{AdS} reads

<ζAdS(x0)ζAdS(x0)>0AdS=1π1Δx02μ=14ΔxμΔxμ,{{\big{<}\zeta^{{AdS}}(x_{0})\zeta^{{AdS}}(x^{\prime}_{0})\big{>}}_{0_{AdS}}}=\frac{1}{\pi}\frac{1}{\Delta{x_{0}}^{2}-\sum\limits_{\mu=1}^{4}\Delta x_{\mu}\Delta x^{\mu}}, (21)

where Δx0=x0(t)x0(t)\Delta x_{0}=x^{\prime}_{0}(t^{\prime})-x_{0}(t) and Δxμ=xμ(t)xμ(t)\Delta x_{\mu}=x^{\prime}_{\mu}(t^{\prime})-x_{\mu}(t). From the relations Eqs. 13\ref{static coord1}-\ref{static coord3} it follows that

Δx02μ=14ΔxμΔxμ=4l2sinh2(tt2l),\Delta{x_{0}}^{2}-\sum\limits_{\mu=1}^{4}\Delta x_{\mu}\Delta x^{\mu}=-4l^{2}sinh^{2}\bigg{(}\frac{t^{\prime}-t}{2l}\bigg{)}, (22)

where we have made the approximation r<<lr<<l so that the metric 7 reduces to the conventional Friedmann-Lemaître-Robertson-Walker (FLRW) metric describing our universe in its static form. It turns out that the correlation function <ζAdS(x0)ζAdS(x0)>0AdS{{\big{<}\zeta^{{AdS}}(x_{0})\zeta^{{AdS}}(x^{\prime}_{0})\big{>}}_{0_{AdS}}} 21 takes the form

<ζAdS(x0)ζAdS(x0)>0AdS=14πl2csch2(tt2l),{{\big{<}\zeta^{{AdS}}(x_{0})\zeta^{{AdS}}(x^{\prime}_{0})\big{>}}_{0_{AdS}}}=-\frac{1}{4\pi l^{2}}csch^{2}\bigg{(}\frac{t^{\prime}-t}{2l}\bigg{)}, (23)

That is a thermal-field correlation function with temperature T=12πlT=\frac{1}{2\pi l} which is nothing more than Gibbons-Hawking temperature.

It is worth mentioning that the relation 23 gives a rigorous physical meaning: The vacuum’s thermal fluctuations arise due to the presence of anti-de Sitter field ζAdS\zeta^{AdS} whose quanta “carry” a “length” ll. In the spirit of our analysis, Gibbons-Hawking radiation is an intrinsic property of the hyperbolic geometry and not of the tunnellling particle. It might be known that the complex path method fails markedly to point out this feature, in the sense that the computations depend strongly on the tunnelling particle (particularly on its spin). Nonetheless, it has been proved for many backgrounds that the tunnelling of particles with different spins (s = 0, 1/2, 1 and s = 3/2) always yields the same temperature [39].

The results obtained till now, give rise to the following questions: Is the observed cosmic background radiation (CMB) nothing but the Gibbons-Hawking radiation? Is it the result of spacetime creation from nothing? Can inflation be generated by the creation of spacetime quanta? answering these questions is beyond the scope of this paper. They, the questions will be addressed later on and they will be explored further in an upcoming paper.

III q-INFLATIONARY MODEL

This section is dedicated to study the properties (i.e. power spectrum and spectral index) of large scale vacuum fluctuations in the q-inflationary scenario. The latter, which depends on the non extensive parameter qq measuring the deviation from the usual exponential expansion, is a strategy to (i) skirt scalar fields with fine-tuned potentials, (ii) skirt an exponential expansion that lasts forever and (iii) ensure a nearly but not perfectly scale-invariant spectrum, in agreement with observational data.

Dealing again with the metric

ds2=(1+r2l2)dt2[11+r2l2dr2+r2(dθ2+sin2θdϕ2)],ds^{2}=-\bigg{(}1+\frac{r^{2}}{l^{2}}\bigg{)}dt^{2}-\bigg{[}\frac{1}{1+\frac{r^{2}}{l^{2}}}dr^{2}+r^{2}(d\theta^{2}+sin^{2}\theta d\phi^{2})\bigg{]}, (24)

then making the following transformations [40]

r~=ai1r1+r2/l2et/l,t~=t+l2ln(1+r2l2),θ~=θ,ϕ~=ϕ,\tilde{r}=\frac{a_{i}^{-1}r}{\sqrt{1+r^{2}/l^{2}}}e^{-t/l},\qquad\tilde{t}=t+\frac{l}{2}ln\bigg{(}1+\frac{r^{2}}{l^{2}}\bigg{)},\qquad\tilde{\theta}=\theta,\qquad\tilde{\phi}=\phi, (25)

leads, under the Wick rotation mentioned in Section II, to the line element

ds2=dt~2ai2e2t~/l[dr~2+r~2dθ~2+r~2sin2θ~dϕ~2],ds^{2}=d\tilde{t}^{2}-a_{i}^{2}e^{2\tilde{t}/l}\big{[}d\tilde{r}^{2}+\tilde{r}^{2}d\tilde{\theta}^{2}+\tilde{r}^{2}sin^{2}\tilde{\theta}d\tilde{\phi}^{2}\big{]}, (26)

where aia_{i} is a constant denoting the initial value of the scale factor a(t~)=aiet~/la(\tilde{t})=a_{i}e^{\tilde{t}/l} at t~=0\tilde{t}=0. Finally, introducing the space coordinates x~1,x~2\tilde{x}^{1},\tilde{x}^{2} and x~3\tilde{x}^{3} which are related to r~,θ~\tilde{r},\tilde{\theta} and ϕ~\tilde{\phi} by the usual equations connecting Cartesian coordinates and polar coordinates in Euclidean space, Eq.26 may be written as

ds2=dt~2ai2e2t~/ldx~2,ds^{2}=d\tilde{t}^{2}-a_{i}^{2}e^{2\tilde{t}/l}d\tilde{\textbf{\emph{x}}}^{2}, (27)

Now let the scale factor aa being generalized to a(t~)=ai[1+(1q)Λ3t~]11qa(\tilde{t})=a_{i}\bigg{[}1+(1-q)\sqrt{\frac{\Lambda}{3}}{\tilde{t}}\bigg{]}^{\frac{1}{1-q}} (provided that q1q\neq 1), with Λ3=1l\sqrt{\frac{\Lambda}{3}}=\frac{1}{l}. From now on, we will call the mentioned scale factor, a q-scale factor that denotes a q-expansion of the space concerned. It should be noted that in the limiting case q1q\to 1, the usual exponential expansion is recovered. To restrict ourselves to an accelerating universe, values of q>2q>2 are excluded since they correspond to a decelerating universe.

Based on these assumptions, the line element 27 takes the following form

ds2=dt~2ai2[1+(1q)Λ3t~]21qdx~2,ds^{2}=d\tilde{t}^{2}-a_{i}^{2}\bigg{[}1+(1-q){\sqrt{\frac{\Lambda}{3}}{\tilde{t}}}\bigg{]}^{\frac{2}{1-q}}d\tilde{\textbf{\emph{x}}}^{2}, (28)

where the metric tensor

gαβ=[10000ai[1+(1q)Λ3t~]11q0000ai[1+(1q)Λ3t~]11q0000ai[1+(1q)Λ3t~]11q]g_{\alpha\beta}=\begin{bmatrix}1&0&0&0\\[3.00003pt] 0&-a_{i}\bigg{[}1+(1-q){\sqrt{\frac{\Lambda}{3}}{\tilde{t}}}\bigg{]}^{\frac{1}{1-q}}&0&0\\[3.00003pt] 0&0&-a_{i}\bigg{[}1+(1-q){\sqrt{\frac{\Lambda}{3}}{\tilde{t}}}\bigg{]}^{\frac{1}{1-q}}&0\\[3.00003pt] 0&0&0&-a_{i}\bigg{[}1+(1-q){\sqrt{\frac{\Lambda}{3}}{\tilde{t}}}\bigg{]}^{\frac{1}{1-q}}\end{bmatrix} (29)

is diagonal. Note the only Riemann tensor’s covariant derivatives that survive are Ri0i0;0=R0i0i;0=Ri00i;0=R0ii0;0=123giit~3R_{i0i0;0}=R_{0i0i;0}=-R_{i00i;0}=-R_{0ii0;0}=\frac{1}{2}\frac{\partial^{3}g_{ii}}{\partial{\tilde{t}}^{3}} (with i=1,2,3i=1,2,3), so that one can easily check that the Bianchi identities, i.e. Rαβγδ;λ+Rαβλγ;δ+Rαβδλ;γ=0R_{\alpha\beta\gamma\delta;\lambda}+R_{\alpha\beta\lambda\gamma;\delta}+R_{\alpha\beta\delta\lambda;\gamma}=0 still hold when incorporating the non-extensive parameter qq.

From Einstein’s field equations

Rαβ12gαβ=8πGTαβR_{\alpha\beta}-\frac{1}{2}g_{\alpha\beta}\mathcal{R}=8\pi GT_{\alpha\beta} (30)

with Tβα=diag[ρ¯,p¯,p¯,p¯]T^{\alpha}_{\beta}={\rm{diag}}[\bar{\rho},-\bar{p},-\bar{p},-\bar{p}], one can get the corresponding deformed matter sector defined by an energy density

ρ¯=38πGΛ3[1+(1q)Λ3t~]2\bar{\rho}=\frac{3}{8\pi G}{\frac{\Lambda}{3}}\bigg{[}1+(1-q)\sqrt{\frac{\Lambda}{3}}\tilde{t}\bigg{]}^{-2} (31)

and a pressure

p¯=(2q+1)18πGΛ3[1+(1q)Λ3t~]2=(2q+1)ρ¯3\bar{p}=-(2q+1)\frac{1}{8\pi G}{\frac{\Lambda}{3}}\bigg{[}1+(1-q)\sqrt{\frac{\Lambda}{3}}\tilde{t}\bigg{]}^{-2}=-(2q+1)\frac{\bar{\rho}}{3} (32)

where during inflation ordinary matter sector is negligible compared to the component triggering inflation. It is easily checked that, in the limiting case q1q\to 1, the standard relations ρ¯=Λ8πG\bar{\rho}=\frac{\Lambda}{8\pi G} and p¯=ρ¯\bar{p}=-\bar{\rho}, for a cosmological constant Λ\Lambda, are recovered. It is worth noting that the relations 31 and 32 fulfill the conservation equations for the energy-momentum tensor TαβT^{\alpha\beta} given by T;βαβ=0T^{\alpha\beta}_{;\beta}=0. That is

ρ¯˙+3H(ρ¯+p¯)=0,\dot{\bar{\rho}}+3H(\bar{\rho}+\bar{p})=0, (33)

where HH is the expansion rate defined as H=a˙aH=\frac{\dot{a}}{a} (with over-dot being the derivative over time t~\tilde{t}).
We consider now the dynamics of a scalar fluctuations δζAdS(η~,x~)\delta\zeta^{{}^{{AdS}}}(\tilde{\eta},\tilde{\textbf{\emph{x}}}) of our previously introduced field ζAdS\zeta^{{}^{{AdS}}}, in the new coordinate system (t~,x~)(\tilde{t},\tilde{\textbf{\emph{x}}}). Expanding the scalar fluctuations δζAdS\delta\zeta^{{}^{{AdS}}}, which will be our inflaton field, in Fourier modes yields

δζAdS(η~,x~)=𝑑k[akAdSeik.x~δvk(η~)+(akAdS)eik.x~δvk(η~)]\delta\zeta^{{}^{{AdS}}}(\tilde{\eta},\tilde{\textbf{\emph{x}}})=\int d\textbf{\emph{k}}[a^{AdS}_{\textbf{\emph{k}}}e^{i\textbf{\emph{k}}.\tilde{\textbf{\emph{x}}}}\delta v_{\textbf{\emph{k}}}(\tilde{\eta})+(a^{AdS}_{\textbf{\emph{k}}})^{\dagger}e^{-i\textbf{\emph{k}}.\tilde{\textbf{\emph{x}}}}\delta v^{*}_{\textbf{\emph{k}}}(\tilde{\eta})] (34)

where η~\tilde{\eta} is the conformal time, from here on we omit the superscript ``AdS"``AdS" in order to maintain simple notations. In an inflationary background, the mode function δvk\delta v_{\textbf{\emph{k}}} satisfies the following equation

δvk′′+2δvk+k2δvk+a2Vδvk=0,\delta v_{\textbf{\emph{k}}}^{{}^{\prime\prime}}+2\mathcal{H}\delta v_{\textbf{\emph{k}}}^{{}^{\prime}}+k^{2}\delta v_{\textbf{\emph{k}}}+a^{2}\frac{\partial{V}}{\partial{\delta v_{\textbf{\emph{k}}}}}=0, (35)

where the prime denotes derivative with respect to the conformal time η~\tilde{\eta}, =a/a\mathcal{H}=a^{\prime}/a and VV is the potential of the scalar field. Firstly, we consider the simple case of an exponential expansion (q1q\to 1), for which H=Λ3H=\sqrt{\frac{\Lambda}{3}} is time-independent and a(η~)=1Hη~a(\tilde{\eta})=-\frac{1}{H\tilde{\eta}}.
Equation. 35 can be recast into

δσk′′+(k2a′′a+mζ2a2)δσk=0,\delta\sigma_{\textbf{\emph{k}}}^{{}^{\prime\prime}}+\bigg{(}k^{2}-\frac{a^{\prime\prime}}{a}+m_{\zeta}^{2}a^{2}\bigg{)}\delta\sigma_{\textbf{\emph{k}}}=0, (36)
δσk′′+[k21η~2(ν214mζ2H2)]δσk=0,\delta\sigma_{\textbf{\emph{k}}}^{{}^{\prime\prime}}+\bigg{[}k^{2}-\frac{1}{\tilde{\eta}^{2}}\big{(}\nu^{2}-\frac{1}{4}-\frac{m^{2}_{\zeta}}{H^{2}}\big{)}\bigg{]}\delta\sigma_{\textbf{\emph{k}}}=0, (37)

with σ=a(η~)v\sigma=a(\tilde{\eta})v, mζm_{\zeta} being the mass of the scalar field and and ν2=94\nu^{2}=\frac{9}{4}. In the following, we will neglect the last term of Eq. 37 by taking into account the assumption mζ22H2<<1\frac{m_{\zeta}^{2}}{2H^{2}}<<1.
The generic solution to Eq. 36 is

δσk=η~[c1(k)Hν(1)(kη~)+c2Hν(2)(kη~)],\delta\sigma_{\textbf{\emph{k}}}=\sqrt{-\tilde{\eta}}\big{[}c_{1}(k)H^{(1)}_{\nu}(-k\tilde{\eta})+c_{2}H^{(2)}_{\nu}(-k\tilde{\eta})\big{]}, (38)

where Hν(1)H^{(1)}_{\nu} and Hν(2)H^{(2)}_{\nu} are the Hankel’s functions of the first and second kind, respectively. Imposing that in the ultraviolet regime k>>aHk>>aH, equation.36 admits a plane wave solution eikη~/2ke^{-ik\tilde{\eta}}/\sqrt{2k} that is expected in flat spacetime and having the following known form in hand

Hν(1)(kη~>>1)2πkη~ei(kη~π2νπ4),H^{(1)}_{\nu}(-k\tilde{\eta}>>1)\sim\sqrt{-\frac{2}{\pi k\tilde{\eta}}}e^{i(-k\tilde{\eta}-\frac{\pi}{2}\nu-\frac{\pi}{4})}, (39)

we set c2(k)=0c_{2}(k)=0 and c1(k)=π2ei(ν+12)π2c_{1}(k)=\frac{\sqrt{\pi}}{2}e^{i(\nu+\frac{1}{2})\frac{\pi}{2}}. The exact solution then reads

δσk=π2ei(ν+12)π2η~Hν(1)(kη~),\delta\sigma_{\textbf{\emph{k}}}=\frac{\sqrt{\pi}}{2}e^{i(\nu+\frac{1}{2})\frac{\pi}{2}}\sqrt{-\tilde{\eta}}H^{(1)}_{\nu}(-k\tilde{\eta}), (40)

The expansion quickly redshifts short-wavelength vacuum fluctuations until their wavelengths go beyond the horizon size H1H^{-1}, hence the quantum modes cease to evolve and “freeze out” as classical fluctuations. Here we are interested in fluctuations of wavelengths that are much larger than the Hubble horizon (i.e. k<<aH\emph{k}<<aH). Thus, on super-horizon scales we have Hν(1)(kη~<<1)2/πeiπ22ν32(Γ(ν)/Γ(3/2))(kη~)νH^{(1)}_{\nu}(-k\tilde{\eta}<<1)\sim\sqrt{2/\pi}e^{-i\frac{\pi}{2}}2^{\nu-\frac{3}{2}}(\Gamma(\nu)/\Gamma(3/2))(-k\tilde{\eta})^{-\nu}, which in turn yields

δσk=ei(ν12)π22ν32Γ(ν)Γ(3/2)12k(kη~)12ν,\delta\sigma_{\textbf{\emph{k}}}=e^{i(\nu-\frac{1}{2})\frac{\pi}{2}}2^{\nu-\frac{3}{2}}\frac{\Gamma(\nu)}{\Gamma(3/2)}\frac{1}{\sqrt{2k}}(-k\tilde{\eta})^{\frac{1}{2}-\nu}, (41)

Going back to the variable δvk\delta v_{\textbf{\emph{k}}}, the fluctuation on super-horizon scales is given by

δvkH2k3(kaH)32ν,\delta v_{\textbf{\emph{k}}}\simeq\frac{H}{\sqrt{2{\emph{k}}^{3}}}\bigg{(}\frac{\emph{k}}{aH}\bigg{)}^{{\frac{3}{2}-\nu}}, (42)

which, for ν=32\nu=\frac{3}{2}, reads

δvkH2k3,\delta v_{\textbf{\emph{k}}}\simeq\frac{H}{\sqrt{2\emph{k}^{3}}}, (43)

A useful quantity to describe the perturbations properties is the so-called power spectrum. It is derived from the average amplitude of the inflationary perturbations

<0|δζ(t~,x~)δζ(t~,x~)|0>=dkkk32π2|δvk|2,\big{<}0|{\delta\zeta}^{\dagger}(\tilde{t},\tilde{\textbf{\emph{x}}})\delta\zeta(\tilde{t},\tilde{\textbf{\emph{x}}})|0\big{>}=\int\frac{d\emph{k}}{k}\frac{\emph{k}^{3}}{2\pi^{2}}|\delta v_{\textbf{\emph{k}}}|^{2}, (44)

Thus, the power spectrum of vacuum fluctuations is defined as

𝒫k=k32π2|δvk|2=(H2π)2,\mathcal{P}_{\emph{k}}=\frac{\emph{k}^{3}}{2\pi^{2}}|\delta v_{\textbf{\emph{k}}}|^{2}=\bigg{(}\frac{H}{{2\pi}}\bigg{)}^{2}, (45)

The spectral index nsn_{s} is calculated throughout the logarithmic derivative of the power spectrum

ns1=dln𝒫kdlnk,n_{s}-1=\frac{dln\mathcal{P}_{\emph{k}}}{dln{\emph{k}}}, (46)

which yields ns=1n_{s}=1 for an exponential expansion where H=Λ3H=\sqrt{\frac{\Lambda}{3}}. This corresponds to a scale-invariant spectrum.
For q1q\neq 1 however, the expansion rate HH is given by

H=Λ3[1+(1q)Λ3t~]1,H=\sqrt{\frac{\Lambda}{3}}\bigg{[}1+(1-q)\sqrt{\frac{\Lambda}{3}}\tilde{t}\bigg{]}^{-1}, (47)

while the term a′′/aa^{\prime\prime}/a appearing in Eq. 36 reads

a′′a=2ϵη~2(1ϵ)21η~2(2+3ϵ),\frac{a^{\prime\prime}}{a}=\frac{2-\epsilon}{\tilde{\eta}^{2}(1-\epsilon)^{2}}\simeq\frac{1}{\tilde{\eta}^{2}}(2+3\epsilon), (48)

where we have used the fact that, for a Hubble rate that changes with time as H˙=(1q)H2\dot{H}=-(1-q)H^{2}, the scale factor, for small values of ϵ=1q\epsilon=1-q, takes the form a=1Hη~11ϵa=-\frac{1}{H\tilde{\eta}}\frac{1}{1-\epsilon}. Substituting Eq. 48 into Eq. 36, we get Eq. 37 with ν32+ϵ\nu\simeq\frac{3}{2}+\epsilon. So that, from Eq. 42, the solution takes the form

δvkH2k3(kaH)(1q),\delta v_{\emph{k}}\simeq\frac{H}{\sqrt{2{\emph{k}}^{3}}}\bigg{(}\frac{\emph{k}}{aH}\bigg{)}^{-(1-q)}, (49)

It turns out that the corresponding power spectrum, making use of Eq. 44, reads

𝒫k=(H2π)2(kaH)2(1q)=(12πΛ3[1+(1q)Λ3t~]1)2(kaH)2(1q),\mathcal{P}_{\emph{k}}=\bigg{(}\frac{H}{{2\pi}}\bigg{)}^{2}\bigg{(}\frac{\emph{k}}{aH}\bigg{)}^{-2(1-q)}=\bigg{(}\frac{1}{2\pi}\sqrt{\frac{\Lambda}{3}}\bigg{[}1+(1-q)\sqrt{\frac{\Lambda}{3}}\tilde{t}\bigg{]}^{-1}\bigg{)}^{2}\bigg{(}\frac{\emph{k}}{{aH}}\bigg{)}^{-2(1-q)}, (50)

It is easily checked, from Eq. 50, that the spectrum of the Bunch-Davies vacuum 45 is recovered in the limiting case q1q\to 1. On the other hand, the spectrum’s amplitude is slightly diminished regarding that of the Bunch-Davies vacuum for t~3Λ\tilde{t}\sim\sqrt{\frac{3}{\Lambda}}. However, the amplitude is drastically diminished for t~>>3Λ\tilde{t}>>\sqrt{\frac{3}{\Lambda}}, which means that the effect of the parameter qq is appreciable only at late times, i.e., end of inflation.

As a side note, inflation is due to the “cosmological constant” Λ/3=1/l\sqrt{{{\Lambda}}/{3}}={1}/{l}, which is nothing more than the inverse of the hyperboloid curvature’s radius. The physical meaning of this result, from the standpoint of spacetime quantization (i.e. x02=x2+l2x_{0}^{2}=\textbf{\emph{x}}^{2}+l^{2}), is that spacetime excitations (see Sec. II) “carrying” a “length” ll or a “curvature” 1/l21/l^{2} are implicitly triggering the q-exponential inflation. Furthermore, there would be no inflation regardless of its nature (exponential or linear..) if there was no excitation, i.e., x0=0x_{0}=0.

IV OBSERVATIONAL CONSTRAINTS ON THE q-PARAMETER

The value of the spectral index measured by the Planck collaboration including the results of WMAP and those based on the investigation of baryon acoustic oscillations (BAO) [41][43]\cite[cite]{[\@@bibref{}{40}{}{}]}-\cite[cite]{[\@@bibref{}{42}{}{}]}, is required to lie in the range [0.9550.98][0.955-0.98] at 95%95\% CL, which rules out the scale invariant spectrum (i.e. ns=1n_{s}=1) at more than 5σ\sigma confidence level. Based on this finding, we can safely discard inflationary models for which q=1q=1. Hence, we must pick out the range within which the value of qq must lie. For such a purpose, let us put ζ(t~,x~)=ζ0(t~)+δζ(t~,x~)\zeta(\tilde{t},\tilde{\textbf{\emph{x}}})=\zeta_{0}(\tilde{t})+\delta\zeta(\tilde{t},\tilde{\textbf{\emph{x}}}), where ζ0(t~)\zeta_{0}(\tilde{t}) represents the homogeneous background part of the field whereas δζ(t~,x~)\delta\zeta(\tilde{t},\tilde{\textbf{\emph{x}}}) denotes the spatially fluctuating part. During inflation, the universe was assumed to be in a Friedmann-Lemaître-Robertson-Walker state with small inhomogeneities. Roughly speaking, scalar perturbations about a homogeneous universe filled with an energy density can be described by perturbations in the energy density as well as perturbations in the metric

ds2=(1+2Φ)dt~2a2(t~)(12Φ)[dr~2+r~2(dθ~2+sin2θ~dϕ~2)],ds^{2}=(1+2\Phi)d\tilde{t}^{2}-a^{2}(\tilde{t})(1-2\Phi)[d\tilde{r}^{2}+\tilde{r}^{2}(d\tilde{\theta}^{2}+sin^{2}\tilde{\theta}d\tilde{\phi}^{2})], (51)

where we have considered only scalar perturbations Φ(t~,x~)\Phi(\tilde{t},\tilde{\textbf{\emph{x}}}). The latter plays a similar role as that of the Newtonian potential used to describe weak gravitational fields (A. Linde, 2005, p.176) 111See A. D. Linde, Particle physics and inflationary cosmology, Contemp. Concepts Phys.5(2005)1-362. (i.e. Schwarzschild metric). a(t~)a(\tilde{t}) is the q-scale factor described above.
A useful gauge-invariant quantity for characterizing scalar perturbations during inflation is the curvature perturbation \mathcal{R} defined as

=Φ+Hδζζ˙0,\mathcal{R}=\Phi+H\frac{\delta\zeta}{\dot{\zeta}_{0}}, (52)

Working in the spatially flat gauge (i.e. Φ=0\Phi=0), the corresponding dimensionless power spectrum reads

𝒫|Φ=0=H2ζ˙02(H2π)2(kaH)2(1q),\mathcal{P_{R}}\bigg{|}_{\Phi=0}=\frac{H^{2}}{\dot{\zeta}_{0}^{{}^{{}^{2}}}}\bigg{(}{\frac{H}{2\pi}}\bigg{)}^{2}\bigg{(}\frac{\emph{k}}{aH}\bigg{)}^{-2(1-q)}, (53)

Since the scalar field ζ0\zeta_{0} is our inflaton field, it must obey the equation of state given in Eq. 32. Consequently, the energy density and pressure due to ζ0\zeta_{0} are given by the following relations

ρ¯=ζ˙02/2+V(ζ0),\bar{\rho}={\dot{\zeta}_{0}}^{{}^{{}^{2}}}/2+V({\zeta}_{0}), (54)
p¯=ζ˙02/2V(ζ0),\bar{p}={\dot{\zeta}_{0}}^{{}^{{}^{2}}}/2-V({\zeta}_{0}), (55)

Thus, one can get, using the above relations along with Eq. 32 the form of ζ˙0{\dot{\zeta}_{0}}

ζ˙02=2(1q)3ρ¯{\dot{\zeta}_{0}}^{{}^{{}^{2}}}=\frac{2(1-q)}{3}\bar{\rho} (56)

where ρ¯=3mp28πH2\bar{\rho}=\frac{3m_{p}^{2}}{8\pi}H^{2} is the homogeneous background energy density, it should be noted that the fluctuations in TμνT_{\mu\nu}, have been considered to be negligible compared with the energy density ρ¯\bar{\rho}. Substituting the relation 56 into Eq.53, the power spectrum 𝒫\mathcal{P_{R}} becomes

𝒫|Φ=0=4π(1q)mp2(H2π)2(kaH)2(1q)=𝒜(kaH)2(1q),\mathcal{P_{R}}\bigg{|}_{\Phi=0}=\frac{4\pi}{(1-q)m_{p}^{2}}\bigg{(}{\frac{H}{2\pi}}\bigg{)}^{2}\bigg{(}\frac{\emph{k}}{aH}\bigg{)}^{-2(1-q)}=\mathcal{A_{R}}\bigg{(}\frac{\emph{k}}{aH}\bigg{)}^{-2(1-q)}, (57)

where mp=1Gm_{p}=\frac{1}{\sqrt{G}} is the Planck mass and 𝒜\mathcal{A_{R}} being the amplitude of the power spectrum. The logarithmic derivative of 𝒫\mathcal{P_{R}} is then expressed as

n1=dln𝒫(k)dlnk=2(1q),n_{\mathcal{R}}-1=\frac{dln\mathcal{P_{R}}(\emph{k})}{dln{\emph{k}}}=-2(1-q), (58)

Since the observational range of nn_{\mathcal{R}} is [0.9550.98][0.955-0.98], qq is restricted to be within the range 0.9775q0.990.9775\leq q\leq 0.99. Furthermore, the tensor-to-scalar ratio rr is defined as

r=𝒫𝒯𝒫=163(1q),r=\frac{\mathcal{P_{T}}}{\mathcal{P_{R}}}=\frac{16}{3}(1-q), (59)

where 𝒫𝒯=8.8π3mp2(H2π)2(kaH)2(1q)\mathcal{P_{T}}=8.\frac{8\pi}{3m_{p}^{2}}\bigg{(}{\frac{H}{2\pi}}\bigg{)}^{2}\bigg{(}\frac{\emph{k}}{aH}\bigg{)}^{-2(1-q)} is the well-known tensor perturbations spectrum. The above relation restricts the range of the tensor-to-scalar ratio to 0.053r0.120.053\leq r\leq 0.12, which is consistent with the limit r<0.12r<0.12 (at 95% CL) set by Planck including BAO data [44].

Another interesting feature of the inflationary cosmology is reheating. The latter describes production of the standard particles after the accelerated inflationary era where the universe has gone through supercooling. In the standard reheating process, the inflaton field decays perturbatively into a set of particles and it starts oscillating around the minimum of its potential during the decay mechanism [45]. This reheating mechanism does not work for our proposed field. Nonetheless, reheating can be achieved if one assumes that, at the end of inflation, the energy density ρ¯\bar{\rho} is converted instantaneously into radiation ρ¯ρrπ230N(T)T4\bar{\rho}\sim\rho_{r}\sim\frac{\pi^{2}}{30}N(T)T^{4}. It follows that the reheating temperature, the value of which can be computed at the end of inflation, takes the form

TRH|t~=t~end=[908π3N(TRH)mp2Hend2]14,T_{RH}\bigg{|}_{\tilde{t}=\tilde{t}_{end}}=\bigg{[}\frac{90}{8\pi^{3}N(T_{RH})}m_{p}^{2}H^{2}_{end}\bigg{]}^{\frac{1}{4}}, (60)

where N(TRH)N(T_{RH}) is the effective number of degrees of freedom at T=TRHT=T_{RH}, with N(TRH)102103N(T_{RH})\sim 10^{2}-10^{3}.
Inspecting Eq. 60, it is worth noting that the reheating temperature TRHT_{RH} and Gibbons-Hawking temperature TGH=1πΛ12T_{GH}=\frac{1}{\pi}\sqrt{\frac{\Lambda}{12}} are related via

TRH=[180π2mp4l4(1q)e2N(1q)N(TRH)𝒜]14TGH,T_{RH}=\bigg{[}\frac{180\pi^{2}m_{p}^{4}l^{4}(1-q)e^{-2N_{*}(1-q)}}{N(T_{RH})}\mathcal{A}_{\mathcal{R}}\bigg{]}^{\frac{1}{4}}T_{GH}, (61)

where we have used the fact that the Hubble scale at the end of inflation, HendH_{end} , and the Hubble scale HH_{*}, i.e. H2=𝒜(1q)mp2πH_{*}^{2}=\mathcal{A}_{\mathcal{R}}(1-q)m_{p}^{2}\pi, at the pivot scale kk_{*} which leaves the horizon NN_{*} e-folds before the end of inflation, are related via N=ttendH𝑑tN_{*}=\int_{t_{*}}^{t_{end}}Hdt, namely

eN(1q)=HHend,e^{N_{*}(1-q)}=\frac{H_{*}}{H_{end}}, (62)

It should be noted that one can compute NN_{*}, assuming an instantaneous reheating, i.e. ρRH=ρend\rho_{{}_{RH}}=\rho_{end}, from the matching equation

N=62.396+ln(Hmp)14lnρendmp4N_{*}=62.396+{\rm{ln}}\bigg{(}\frac{H_{*}}{m_{p}}\bigg{)}-\frac{1}{4}{\rm{ln}}\frac{\rho_{end}}{m^{4}_{p}} (63)

which is drawn from Eq. (20) of Ref. [46], with ρend=3mp28πHend2\rho_{end}=\frac{3m_{p}^{2}}{8\pi}H^{2}_{end}. Combining Eqs. 62 and 63, one can compute NN_{*} and hence derive TRHT_{RH} via Eq. 60. Note that CMB data constraint H2/(1q)H_{*}^{2}/(1-q) through the amplitude of the anisotropies 𝒜\mathcal{A}_{\mathcal{R}}, as well as qq from the spectral index. Consequently, one may expect CMB data to also provide some information on NN_{*} and TRHT_{RH}.

In what follows we will examine the observational constraints on the free parameters {q,l,t}\{q,l,t_{*}\} of the model, and the corresponding derived parameters {r,N,TRH}\{r,N_{*},T_{RH}\}.

V DATA ANALYSIS

In order to impose constraints on the q-inflationary model’s parameters, we have used a modified version of the Boltzmann CAMB code [47][49]\cite[cite]{[\@@bibref{}{46}{}{}]}-\cite[cite]{[\@@bibref{}{48}{}{}]} and the Monte Carlo Markov Chain (MCMC) analysis [50] provided by the publicly available CosmoMC package222http://cosmologist.info/cosmomc/. Therefore, the inflationary sector has been modified by plugging in the following power spectrum parameterizations

𝒫(k)=1(1q)πmp2l2[1+(1q)t~/l]2(kk)2(1q)\mathcal{P}_{\mathcal{R}}(k)=\frac{1}{(1-q)\pi m_{p}^{2}}{l^{-2}[1+(1-q)\tilde{t}_{*}/l]^{-2}}\bigg{(}\frac{k}{k_{*}}\bigg{)}^{-2(1-q)} (64)
𝒫t(k)=r.1(1q)πmp2l2[1+(1q)t~/l]2(kk)nt\mathcal{P}_{t}(k)=r.\frac{1}{(1-q)\pi m_{p}^{2}}{l^{-2}[1+(1-q)\tilde{t}_{*}/l]^{-2}}\bigg{(}\frac{k}{k_{*}}\bigg{)}^{n_{t}} (65)

where kk_{*} denotes an arbitrary pivot scale and nt=2(1q)n_{t}=-2(1-q) is the tilt of the tensor power spectrum. Thus, to constrain the model parameters {qq, ll, tt_{*}}, and derive their respective posterior probability distributions, we have used Planck 2018 data set (TT, TE, EE+lowTEB) [51] with low ll likelihood (0l290\leq l\leq 29) and high ll likelihood (30l250830\leq l\leq 2508), estimated using commander. It is worth noting that, instead of q,lq,l and t~\tilde{t}_{*}, we have constrained ns=2q1,ln[105l]n_{s}=2q-1,{\rm{ln}}[10^{-5}l] and ln[105t~]{\rm{ln}}[10^{-5}\tilde{t}_{*}] in such away to use the standard parameterization provided in the CAMB code, i.e. [l+(1q)t~]1/As[l+(1-q)\tilde{t}_{*}]\propto 1/\sqrt{A_{s}}. The free parameters were set as uniform priors, along with the other parameters of the standard cosmological model: baryon density (Ωbh2\Omega_{b}h^{2}), cold dark matter density (Ωch2\Omega_{c}h^{2}), Thomson scattering optical depth due to re-ionization (τ\tau), and angular size of horizon (θ\theta). The priors are summarized in Table.1. Due to their negligible effect on the CMB power spectrum, the effective number of neutrinos NνN_{\nu}, Helium mass fraction YpY_{p} and the width of re-ionization were kept fixed at their default values 3.046,0.243.046,0.24 and 0.50.5 respectively. We have also fixed the pivot scale to k=0.05Mpc1k_{*}=0.05\rm{Mpc}^{-1}.

Parameter Lower limit Upper limit
Ωbh2\Omega_{b}h^{2} 0.0050.005 0.10.1
Ωch2\Omega_{c}h^{2} 0.0010.001 0.990.99
θ\theta 0.50.5 1010
τ\tau 0.010.01 0.80.8
ns=2q1n_{s}=2q-1 0.80.8 1.21.2
ln[105t~]\rm{ln}[10^{-5}\emph{$\tilde{t}_{*}$}] 90-90 74-74
ln[105l]\rm{ln}[10^{-5}\emph{$l$}] 91.62-91.62 70-70
Table 1: Uniform priors used in MCMC parameters’ estimation.

It is noteworthy to mention that during the MCMC analysis we encountered a degeneracy problem in the parameter space {l,t~}\{l,\tilde{t}_{*}\}. This is illustrated in Fig. 2 where the posterior probability distribution of the parameters concerned can exhibit multiple peaks and/or subpeaks describing a multi-modal pattern. Due to this degeneracy, CosmoMC sampling of the q-inflation undergoes a longer time to find the right mixing between parameters and thereby reaching the desired convergence is considerably a slack task. To overcome the problem of degeneracy between the parameters ll and tt_{*} it can be useful to simply discard the parameter ll. Doing so is justified since for an extremely huge number of total e-folds the Hubble rate near to the end of inflation can be approximated by H[(1q)t~]1H_{*}\simeq[(1-q)\tilde{t}_{*}]^{-1}333At the end of inflation we can safely set (1q)t~end/l>>1(1-q)\tilde{t}_{end}/l>>1, so that Hend[(1q)t~end]1H_{end}\simeq[(1-q)\tilde{t}_{end}]^{-1}, with t~=e(1q)Nt~end\tilde{t}_{*}=e^{-(1-q)N_{*}}\tilde{t}_{end}. In the following we will only discuss the results obtained from the latter case.

One-dimensional and two-dimensional marginalised posterior distributions for the model’s parameters, are shown in Fig. 3, while values at 95% CL and best fit values are quoted in Table. 2. For comparative purposes, Table 2 also displays the values at 95% and the best fit for the standard normal inflation (ΛCDM+r\Lambda CDM+r). Furthermore, Figure 4 depicts the probability distribution functions and the marginalized confidence regions at 68%68\% and 95%95\% for the two models, where the crimson contours denote the constraints on the q-inflation while the cyan contours denote the constraints on ΛCDM+r\Lambda CDM+r model. It is worth noting that, since they don’t possess the same number of free parameters the two models are compared through the well-known Akaike information criterion (AIC) [52]

AIC=2lnmax+2N=χmin2+2N,{\rm{AIC}}=-2{\rm{ln}}\mathcal{L}_{max}+2N=\chi_{min}^{2}+2N, (66)

where max\mathcal{L}_{max} is the maximum likelihood and NN is the number of free parameters of the model. The difference in AIC values (ΔAIC<2\Delta{\rm{AIC}}<2) has an evidence support for the model under consideration.

Note that, unlike that of the standard normal inflation, the tensor-to-scalar ratio rr of the q-inflation is not considered as a free parameter but calculated as of function of nsn_{s} throughout the relation 59. Furthermore, a consistency relation between the tensor-to-scalar ratio rr and the tilt of the tensor power spectrum ntn_{t}, namely nt=3r/8n_{t}=-3r/8, has been used. The latter differs from the usual consistency relation by a factor of 33. Since they are functions of a single free parameter nsn_{s}, rr and ntn_{t} are strongly correlated and hence are more accurately constrained. The number of e-folds N=57.60N_{*}=57.60 lies within the range 50<N<6050<N_{*}<60 [53] expected for an instantaneous reheating stage. Moreover, the reheating temperature TRHT_{RH} is found to be about 3.9×1016GeV3.9\times 10^{16}GeV, and therefore inflation is halted at t~end=8.5×1037s\tilde{t}_{end}=8.5\times 10^{-37}s, from which it follows that l<<3.78×1030ml<<3.78\times 10^{-30}m.

Apart from the rest of the models parameters, we see that the best fit tensor-to-scalar ratio rr is 0.079 for the q-inflation and 0.022 for the standard normal inflation. Thus, the magnitude of rr can be of help in distinguishing the standard normal inflation from the q-inflation. Meanwhile, the value of the tensor-to-scalar ratio cannot yet be determined precisely from the present Planck data.

q-inflation ΛCDM+r\Lambda CDM+r
Parameter 95% limits Best fit 95% limits Best fit
nsn_{s} 0.97070.0065+0.00610.9707^{+0.0061}_{-0.0065} 0.97050.9705 0.96660.0090+0.00920.9666^{+0.0092}_{-0.0090} 0.96700.9670
ln[105t~]{\rm{ln}}[10^{-5}\tilde{t}_{*}] 75.880.31+0.32-75.88^{+0.32}_{-0.31} 75.90-75.90 - -
qq 0.98540.0033+0.00310.9854^{+0.0031}_{-0.0033} 0.98520.9852 - -
rr 0.0780.016+0.0170.078^{+0.017}_{-0.016} 0.0790.079 0.0290.029+0.0400.029^{+0.040}_{-0.029} 0.0220.022
ntn_{t} 0.02930.0065+0.0061-0.0293^{+0.0061}_{-0.0065} 0.0296-0.0296 0.00360.0050+0.0037-0.0036^{+0.0037}_{-0.0050} 0.0028-0.0028
ln[1010As]{\rm{ln}}[10^{10}A_{s}] 3.0870.056+0.0623.087^{+0.062}_{-0.056} 3.0833.083 3.0860.062+0.0623.086^{+0.062}_{-0.062} 3.0883.088
Ωbh2\Omega_{b}h^{2} 0.022310.00029+0.000300.02231^{+0.00030}_{-0.00029} 0.022250.02225 0.022240.00030+0.000310.02224^{+0.00031}_{-0.00030} 0.022260.02226
Ωch2\Omega_{c}h^{2} 0.11850.0024+0.00230.1185^{+0.0023}_{-0.0024} 0.11890.1189 0.11950.0028+0.00280.1195^{+0.0028}_{-0.0028} 0.11890.1189
100θMC100\theta_{MC} 1.040890.00061+0.000611.04089^{+0.00061}_{-0.00061} 1.040921.04092 1.040790.00062+0.000611.04079^{+0.00061}_{-0.00062} 1.04069
τ\tau 0.0780.028+0.0310.078^{+0.031}_{-0.028} 0.0750.075 0.0760.032+0.0320.076^{+0.032}_{-0.032} 0.0770.077
NN_{*} 57.590.14+0.1457.59^{+0.14}_{-0.14} 57.6057.60 - -
TRHT_{RH}444Here, TRHT_{RH} is expressed in m1m^{-1} ( 3.150.15+0.13)1031\left(\,3.15^{+0.13}_{-0.15}\,\right)\cdot 10^{31} 3.1510313.15\cdot 10^{31} - -
χmin2\chi_{min}^{2} 13004.94613004.946 13001.54813001.548
AIC{\rm{AIC}} 13016.94613016.946 13015.54813015.548
Table 2: Constraints on cosmological parameters for the q-inflationary model compared with ΛCDM+r\Lambda CDM+r model, obtained using Planck data 2018 TT+TE+EE+lowTEB.
Refer to caption
Figure 2: 2D joint posterior probability distributions and 1D marginal posterior probability distribution of the q-inflation parameters qq, ln(105l){\rm{ln}}(10^{-5}l) and ln(105t){\rm{ln}}(10^{-5}t_{*}).
Refer to caption
Figure 3: 2D joint posterior probability distributions and 1D marginal posterior probability distribution of the q-inflation parameters qq and ln(105t){\rm{ln}}(10^{-5}t_{*}) as well as related derived parameters (r,nt,Nr,n_{t},N_{*} and TRHT_{RH}).
Refer to caption
Figure 4: Probability distribution functions and marginalized confidence regions at 68%68\% and 95%95\% for the parameters Ωbh2,Ωch2,H0,ln(1010As),ns,r\Omega_{b}h^{2},\Omega_{c}h^{2},H_{0},{\rm{ln}}(10^{10}A_{s}),n_{s},r and ntn_{t} of the q-inflation (crimson contours) and ΛCDM+r\Lambda CDM+r model (cyan contours).

VI CLOSING REMARKS

Despite the fact that the obtained results may allow the q-inflation to be a plausible candidate for describing the early universe, the current model is just a toy model and much enhancement has yet to be done. In particular, the reheating mechanism must be further developed. As one may notice, the current model, similarly to the power-law inflation, has an exit problem, i.e. inflation does not end via slow roll violation, i.e. ϵ=1q=const\epsilon=1-q=const. Nevertheless, as has been stated by Lucchin et al. [54] for a power-law inflation, one can assume that at a particular time t~RH=t~end\tilde{t}_{{}_{RH}}=\tilde{t}_{end}, the model no longer holds in such a way that a rapid reheating process takes place. Note that, in the current study, we have neglected the term mζ2/2H2{m_{\zeta}^{2}}/{2H^{2}} appearing in Eq. 37, for the sake of simplicity. Nevertheless, despite its negligible effect one may take it into account. Doing so may lead to slight differences in the values of qq, rr, NN_{*} as well as TRHT_{RH}.

Within the context of the q-inflationary scenario, “although somehow hypothetical”, one may emphasize that “spacetime quanta” may be relevant to such an inflation, inasmuch as the cosmological constant, i.e., the hyperboloid curvature radius triggering the inflation is carried by the quanta. If spacetime is really quantized then the present work can be seen as a little step towards understanding (i) its quantum nature and (ii) why does time move forward. In this paper, we considered only quanta with “frequency” x0>0\emph{x}_{0}>0, covering the hyperboloid’s upper sheet but quanta that cover the lower sheet, i.e., x0<0\emph{x}_{0}<0 must also be taken into account, this is similar to K-G particle with negative energy ω<0\omega<0 that is considered as anti-particle having a positive energy. That is to say, if a parallel universe where time goes backward does exist then particles with negative energy may “live” there, but they may be seen as anti-particles in our universe.

VII ACKNOWLEDGMENTS

I would like to thank Antony Lewis for kindly providing the numerical codes CosmoMC and CAMB. I am also grateful to Sukannya Bhattacharya and Savvas Nesseris for valuable instructions about numerical issues regarding CosmoMC, and I am particularly indebted to an anonymous reviewer for insightful comments.

References

  • [1] S. W. Hawking, Nature (London) 248, 30 (1974); Commun. Math. Phys. 43, 199 (1975).
  • [2] S. Weinfurtner. Quantum simulation of black-hole radiation. Nature, 569, 634-635 (2019)
  • [3] J. R. M. de Nova, K. Golubkov, V. I. Kolobov, and J. Steinhauer. Observation of thermal Hawking radiation and its temperature in an analogue black hole. Nature, 569, 7758 (2019).
  • [4] J. Drori, Y. Rosenberg, D. Bermudez, Y. Silberberg, and U. Leonhardt, Phys. Rev. Lett. 122, 010404 (2019)
  • [5] P. Kraus and F. Wilczek, Nucl. Phys. B 433, 403 (1995).
  • [6] M.K. Parikh and F. Wilczek, Phys. Rev. Lett. 85, 5042 (2000).
  • [7] K. Srinivasan and T. Padmanabhan, Phys. Rev. D 60, 024007 (1999).
  • [8] J. Hu, L. Feng, Z. Zhang and C. Chin, Quantum simulation of Unruh radiation, Nat. Phys. (2019).
  • [9] F. Scardigli, M. Blasone, G. Luciano and R. Casadio, Modified Unruh effect from generalized uncertainty principle, Eur. Phys. J. C 78, 728 (2018).
  • [10] J.A. Rosabal, New perspective on the Unruh effect, Phys. Rev. D 98, 056015 (2018).
  • [11] W. G. Unruh, Phys. Rev. D 14, 870 (1976).
  • [12] G.W. Gibbons and S.W. Hawking, Phys. Rev. D 15, 2738 (1977).
  • [13] D. Chen, H. Wu, H. Yang and S. Yang, Int. J. Mod. Phys. A 29, 1430054 (2014).
  • [14] T. Damour and R. Ruffini, Phys. Rev. D 20, 239 (1976).
  • [15] S. Sannan, Gen. Relativ. Gravit. 20, 239 (1988).
  • [16] P. Kraus and F. Wilczek, Nucl. Phys. B 437, 231 (1995).
  • [17] M. K. Parikh and F. Wilczek, Phys. Rev. Lett. 85, 5042 (2000).
  • [18] K. Srinivasan and T. Padmanabhan, Phys. Rev. D 60, 024007 (1999).
  • [19] M. Angheben, M. Nadalini, L. Vanzo and S. Zerbini, JHEP 05, 014 (2005).
  • [20] R. Kerner and R.B. Mann, Phys. Rev. D 73, 104010 (2006).
  • [21] A. H. Guth, Phys. Rev. D, 23, 347 (1981).
  • [22] A. H. Guth and So-Young Pi, Phys. Rev. Lett., 49, 111 (1982).
  • [23] C. Armendariz-Picon, T. Damour, and V. F. Mukhanov, Phys. Lett., B 458, 209 (1999).
  • [24] A. D. Linde, Phys. Lett., B 108, 389 (1982).
  • [25] W. H. Kinney, Eternal Inflation and the Refined Swampland Conjecture, Phys. Rev. Lett. 122 (2019).
  • [26] A. de la Fuente, P. Saraswat, and R. Sundrum, Natural Inflation and Quantum Gravity, Phys. Rev. Lett. 114. 15, 151303 (2015).
  • [27] E.O. Kahya, B. Pourhasssan and S. Uraz, Phys. Rev. D 92,103511 (2015).
  • [28] Z. Yi and Y. Gong, On the constant-roll inflation, JCAP 1803, 052 (2018).
  • [28] H. Motohashi and A.A. Starobinsky, JCAP 11, 025 (2019).
  • [30] A. Mohammadi, K. Saaidi and T. Golanbari, Tachyon constant-roll inflation, Phys.Rev. D 97, 083006 (2018).
  • [31] G. W. Gibbons and S. W. Hawking, Cosmological Event Horizons, Thermodynamics, And Particle Creation, Phys. Rev. D 15, 2738 (1977).
  • [32] J. R. Gott, Nature 295, 304 (1982).
  • [33] S. Umarov, C. Tsallis and S. Steinberg, Milan J. Math. 76 (2008).
  • [34] K. Ourabah and M. Tribeche, Phys. Rev. E 89, 062130 (2014).
  • [35] A. Plastino and M. C. Rocca, Chin, Phys. C 42, 053102 (2018).
  • [36] L.C. Céleri and V.I. Kiosses, Canonical field quantization in momentum space and the corresponding Unruh effect, https://arxiv.org /abs /1712 .05206 (2017).
  • [37] L.C. Céleri and V. I Kiosses, Unruh effect as a result of quantization of spacetime, Phys. Lett. B 781, 611 (2018).
  • [38] A. Dasgupta, The real Wick rotations in quantum gravity, JHEP 0207, 062 (2002).
  • [39] H. Erbin and V. Lahoche, Universality of Tunneling Particles in Hawking Radiation, Phys. Rev. D 98, 104001 (2018).
  • [40] K. Yagdjian and A. Galstian. Fundamental solutions for the Klein-Gordon equation in de Sitter spacetime. Comm. Math. Phys., 285 (1) (2009).
  • [41] Akrami et al. (Planck Collaboration), Planck 2018 results.X. Constraints on inflation, arXiv: 1807.06211.
  • [42] P. A. R. Ade et al. (Planck Collaboration), Astron. Astrophys. 594, A20 (2016).
  • [43] P. A. R. Ade et al. (BICEP2 and Keck Array Collaborations), Phys. Rev. Lett. 116, 031302 (2016).
  • [44] P. A. R. Ade, et. al., Planck 2013 results. XXII. Constraints on inflation, arXiv:1303.5082.
  • [45] R. Allahverdi, R. Brandenberger, F.-Y. Cyr-Racine and A. Mazumdar, Ann. Rev. Nucl. Part. Sci., 60, 27 (2010).
  • [46] J. Martin and C. Ringeval, First CMB Constraints on the Inflationary Reheating Temperature, Phys. Rev. D 82, 023511 (2010).
  • [47] U. Seljak and M. Zaldarriaga, Astrophys. J. 469, 437 (1996).
  • [48] A. Lewis and A. Challinor, Phys. Rev. D 66 (2), 023531 (2002).
  • [49] A. Lewis, A. Challinor and A. Lasenby, Astrophys. J. 538, 473 (2000).
  • [50] A. Lewis, S. Bridle, Phys. Rev. D. 66 (10), 103511 (2002).
  • [51] Planck collaboration, N. Aghanim et al., Planck 2018 results. VI. Cosmological parameters, 1807.06209.
  • [52] H. Akaike, IEEE T. Automat. Contr. 19, 716 (1974).
  • [53] P. A. R. Ade et al. (Planck), Astron. Astrophys. 571, A22, 1303.5082 (2014).
  • [54] S. Lucchin, and S. Matarrese, Phys. Rev. D 32, 1316 (1985).