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

Cosmological consequences of a scalar field with oscillating equation of state. III. Unifying inflation with dark energy and small tensor-to-scalar ratio

S. X. Tian [email protected] Department of Astronomy, Beijing Normal University, 100875 Beijing, China    Zong-Hong Zhu [email protected] Department of Astronomy, Beijing Normal University, 100875 Beijing, China School of Physics and Technology, Wuhan University, 430072 Wuhan, China
Abstract

We investigate the inflationary consequences of the oscillating dark energy model proposed by Tián [Phys. Rev. D 101, 063531 (2020)], which aims to solve the cosmological coincidence problem with multi-accelerating Universe (MAU). We point out that the inflationary dynamics belong to slow-roll inflation. The spectral index of scalar perturbations and the tensor-to-scalar ratio rr are shown to be consistent with current Planck measurements. Especially, this model predicts r107r\sim 10^{-7}, which is far below the observation limits. This result motivates us to explore the smallness of rr in the general MAU. We propose a quintessential generalization of the original model and prove r<0.01r<0.01 in general. The null detection to date of primordial gravitational waves provides a circumstantial evidence for the MAU. After the end of inflation, the scalar field rolls toward infinity instead of a local minimum, and meanwhile its equation of state is oscillating with an average value larger than 1/31/3. In this framework, we show that gravitational particle creation at the end of inflation is capable of reheating the Universe.

I Introduction

The present standard cosmological model is known as the Lambda Cold Dark Matter (Λ\LambdaCDM) model, in which the cosmological constant Λ\Lambda is generally referred to as the simplest explanation of the late-time cosmic acceleration [1, 2]. However, simpleness is not an indicator of naturalness, and the latter seems to be more important for a physical theory 111Simpleness represents mathematical complexity and naturalness represents the probability of choosing by nature.. The Λ\LambdaCDM model is unnatural due to an incredible coincidence: matter dominated the Λ\Lambda by many orders of magnitude through most of the cosmic history, but they are comparable at today [4]. Dynamical dark energy provides an alternative explanation to the cosmic acceleration, some of which can alleviate the coincidence problem with the tracker property [5, 6, 7, 8]. The core argument is that the tracker property allows the initial conditions of dark energy to vary within a wide range without affecting the late-time evolution. However, the problem is not solved as the relative dark energy density is still much less than 11 through most of the history in the tracker models (see Fig. 6 in [7] for an example).

Just releasing the initial conditions of dark energy cannot solve the coincidence problem. Intuitively, what we need may be a multi-accelerating Universe (MAU), in which the normal matters and dark energy alternately dominated each other across the whole cosmic history. The present day belongs to one of the transitions, and thus the coincidence problem disappears. This scenario can be realized by oscillating dark energy, which was first studied in [9] with a sinusoidal-modulated exponential potential V(ϕ)exp(λϕ)[1+Asin(Bϕ)]V(\phi)\propto\exp(-\lambda\phi)[1+A\sin(B\phi)]. Follow-up works mainly focus on exploring other forms of the oscillating dark energy, e.g., multiple scalar fields realization [10], everpresent Λ\Lambda [11], and parameterized equation of state (EOS) [12, 13, 14, 15, 16]. Note that the Lagrangian can be reconstructed from a given EOS [17, 18, 19, 20] and the result generally has a specified dependence on the Hubble constant H0H_{0}. This H0H_{0}-dependence is unattractive for a physical theory (see discussions in [1, 5, 6, 9]), and should not appear in the desired model. In [21], we proposed a new oscillating dark energy model with the potential

V(ϕ)=V0exp[λ1+λ22ϕα(λ1λ2)2sinϕα].V(\phi)=V_{0}\exp\left[-\frac{\lambda_{1}+\lambda_{2}}{2}\phi-\frac{\alpha(\lambda_{1}-\lambda_{2})}{2}\sin\frac{\phi}{\alpha}\right]. (1)

Meanwhile we showed that the parameter space of {λ1+λ2>4\{\lambda_{1}+\lambda_{2}>4, 0<λ2<0.390<\lambda_{2}<0.39, α=𝒪(1)\alpha=\mathcal{O}(1) and V0V_{0} is arbitrary}\} is able to explain the late-time cosmic acceleration and solve the coincidence problem. In this model, the theoretically preferred Universe evolves in an oscillating scaling manner in the radiation epoch and in a chaotic accelerating manner in the matter epoch [22]. Mathematically, the radiation-matter transition induces a process of period-doubling bifurcation to chaos.

Besides the late-time acceleration, inflation is the other widely accepted accelerating phase in the Universe [23]. These two accelerating phases may be driven by the same field. This idea emerged in the late 1980s [5, 6] and a concrete model, named quintessential inflation, is proposed in [24] after the observational confirmation of dark energy [25, 26]. The MAU can also provide such a unified picture as its name implies. On this topic, previous work, e.g., [12, 17, 20], has studied the cosmic background evolution based on the parameterized oscillating EOS or its reconstructed Lagrangian. However, in the era of precision cosmology [27], the performance of the MAU in inflation merits further discussion. In this paper, we study in detail the inflation model driven by the scalar field described by Eq. (1), and explore the general predictions of this kind of model.

This paper is organized as follows. In Sec. II, using the slow-roll approximation, we analyze the background evolution and the primordial inhomogeneities generated during inflation for our model. In particular, we prove an upper limit of the tensor-to-scalar ratio for a class of general MAU model. Section III analyzes the post-inflationary background evolution without considering the reheating process. Section IV shows that all matters of the familiar Universe can be from gravitational reheating in our model. Discussion is presented in Sec. V. Throughout this paper, we adopt the SI units and inherit the notations in [21], e.g., the slope λV/V\lambda\equiv-V^{\prime}/V, d/dϕ{}^{\prime}\equiv{\rm d}/{\rm d}\phi, ϕ\phi is dimensionless and [V]=length2[V]={\rm length}^{-2}.

II Slow-roll inflation

The early Universe is assumed to be dominated by the scalar field with the potential Eq. (1). As shown in Fig. 2(a), the potential changes periodically between flat (λλ2\lambda\approx\lambda_{2}) and steep (λλ1\lambda\approx\lambda_{1}). A natural idea is that the flat part can drive the slow-roll inflation [28, 29, 30] while the steep part can end it. To be consistent with the slow-roll conditions but without loss of generality, we assume ϕαπ\phi\approx\alpha\pi (λλ2\lambda\approx\lambda_{2}) at the beginning of inflation. With this setting, inflation would end around ϕ2απ\phi\approx 2\alpha\pi (λλ1\lambda\approx\lambda_{1}). Here we present detailed analysis of this inflation process.

II.1 Inflationary expansion

The first issue we concern is the total e-folding number during inflation. The action of this physical system reads S=κ1d4xg[R/2μϕμϕ/2V(ϕ)]S=\kappa^{-1}\int{\rm d}^{4}x\sqrt{-g}\left[R/2-\partial^{\mu}\phi\partial_{\mu}\phi/2-V(\phi)\right], where V(ϕ)V(\phi) is given by Eq. (1). We further assume the Universe is flat, and then the cosmic evolution equations corresponding to the above action can be written as

H2=ϕ˙2/6+c2V/3,\displaystyle H^{2}=\dot{\phi}^{2}/6+c^{2}V/3, (2a)
a¨/a=ϕ˙2/3+c2V/3,\displaystyle\ddot{a}/a=-\dot{\phi}^{2}/3+c^{2}V/3, (2b)
ϕ¨+3Hϕ˙+c2V=0.\displaystyle\ddot{\phi}+3H\dot{\phi}+c^{2}V^{\prime}=0. (2c)

Under the slow-roll conditions {ϕ˙2c2V\{\dot{\phi}^{2}\ll c^{2}V and |ϕ¨||c2V|}|\ddot{\phi}|\ll|c^{2}V^{\prime}|\}, Eq. (2) can be simplified to H2=c2V/3H^{2}=c^{2}V/3 and 3Hϕ˙=c2V3H\dot{\phi}=-c^{2}V^{\prime}. Dividing these two equations, we obtain

1adadϕ=1λ=2λ1+λ2+(λ1λ2)cosϕα.\frac{1}{a}\frac{{\rm d}a}{{\rm d}\phi}=\frac{1}{\lambda}=\frac{2}{\lambda_{1}+\lambda_{2}+(\lambda_{1}-\lambda_{2})\cos\frac{\phi}{\alpha}}. (3)

Integrating Eq. (3) gives the total e-folding number

Ntot=ϕiϕedϕλ=2αλ1λ2arctan(λ2λ1tanϕ2α)|ϕiϕe,N_{\rm tot}=\int_{\phi_{i}}^{\phi_{e}}\frac{{\rm d}\phi}{\lambda}=\frac{2\alpha}{\sqrt{\lambda_{1}\lambda_{2}}}\arctan(\sqrt{\frac{\lambda_{2}}{\lambda_{1}}}\left.\tan\frac{\phi}{2\alpha})\right|_{\phi_{i}}^{\phi_{e}}, (4)

where the subscripts ii and ee denote the beginning and end of inflation respectively, and the second equality holds for ϕiαπ\phi_{i}\geqslant\alpha\pi due to the singularity of the tangent function. Note that the above result is insensitive to the value of ϕe\phi_{e}. For example, if λ21\lambda_{2}\ll 1, the integral from 3απ/23\alpha\pi/2 to 2απ2\alpha\pi is approximately equal to 2α/λ12\alpha/\lambda_{1}, which is much smaller than the leading term. For the case of ϕi=απ\phi_{i}=\alpha\pi and λ21\lambda_{2}\ll 1, Eq. (4) gives

Ntot=απλ1λ2+𝒪(1),N_{\rm tot}=\frac{\alpha\pi}{\sqrt{\lambda_{1}\lambda_{2}}}+\mathcal{O}(1), (5)

where 𝒪(1)\mathcal{O}(1) represents the minor influence of the exact value of ϕe\phi_{e}. Changing ϕi\phi_{i} slightly does not affect the order of magnitude of the above result. Considering the typical values of λ1\lambda_{1} and α\alpha (see Sec. I), Eq. (5) requires λ2𝒪(104)\lambda_{2}\lesssim\mathcal{O}(10^{-4}) for successful inflation.

II.2 Linear perturbations

The second issue we concern is the primordial inhomogeneities generated during inflation. The source is quantum fluctuations [31, 32, 33, 34, 35, 36] and its evolution is governed by the Mukhanov-Sasaki equation [37, 38, 39]. The main result is characterized by spectral index and amplitude. The slow-roll approximation is suitable for dealing with this problem [40, 41, 42, 43]. In our conventions, the potential slow-roll parameters are defined as ϵV1/2(V/V)2\epsilon_{V}\equiv 1/2(V^{\prime}/V)^{2} and ηVV′′/V\eta_{V}\equiv V^{\prime\prime}/V. We are concentrate on the lowest order result, where the spectral index of scalar perturbations ns=1+2ηV6ϵVn_{\rm s}=1+2\eta_{V}-6\epsilon_{V} and the tensor-to-scalar ratio r=16ϵVr=16\epsilon_{V}. The use of potential slow-roll parameters facilitates the analysis based on the potential Eq. (1).

At the level of approximation we are considering, for Eq. (1), direct calculation gives

ns\displaystyle n_{\rm s} =1[(λ1λ2)24cos2ϕα+λ12λ222cosϕα\displaystyle=1-\left[\frac{(\lambda_{1}-\lambda_{2})^{2}}{4}\cos^{2}\frac{\phi_{\ast}}{\alpha}+\frac{\lambda_{1}^{2}-\lambda_{2}^{2}}{2}\cos\frac{\phi_{\ast}}{\alpha}\right.
λ1λ2αsinϕα+(λ1+λ2)24],\displaystyle\qquad\quad\left.-\frac{\lambda_{1}-\lambda_{2}}{\alpha}\sin\frac{\phi_{\ast}}{\alpha}+\frac{(\lambda_{1}+\lambda_{2})^{2}}{4}\right], (6a)
r\displaystyle r =2[λ1+λ2+(λ1λ2)cos(ϕ/α)]2,\displaystyle=2\left[\lambda_{1}+\lambda_{2}+(\lambda_{1}-\lambda_{2})\cos(\phi_{\ast}/\alpha)\right]^{2}, (6b)

where the subscript \ast denotes horizon crossing. Integrating Eq. (3) from ϕ\phi_{\ast} to ϕe\phi_{e} gives

N=2αλ1λ2arctan(λ2λ1tanϕ2α),N_{\ast}=-\frac{2\alpha}{\sqrt{\lambda_{1}\lambda_{2}}}\arctan(\sqrt{\frac{\lambda_{2}}{\lambda_{1}}}\tan\frac{\phi_{\ast}}{2\alpha}), (7)

in which we neglect the minor influence of the exact value of ϕe\phi_{e} and assume ϕ\phi_{\ast} is slightly larger than απ\alpha\pi. As we will see, this assumption is suitable for our analysis. Solving the above equation for ϕ\phi_{\ast}, we obtain

ϕ=2α[πarctan(λ1/λ2tanλ1λ2N2α)].\phi_{\ast}=2\alpha\left[\pi-\arctan(\sqrt{\lambda_{1}/\lambda_{2}}\tan\frac{\sqrt{\lambda_{1}\lambda_{2}}N_{\ast}}{2\alpha})\right]. (8)

The range of ϕ\phi_{\ast} given by Eq. (8) satisfies our requirement. In principle, for a given NN_{\ast}, we can use Eq. (8) to calculate ϕ\phi_{\ast}, and then use Eq. (6) to calculate nsn_{\rm s} and rr. Note that NN_{\ast} given by Eq. (7) is the e-folding number from horizon crossing to inflation ending. For most of the inflation models, e.g., chaotic inflation [30], we have 50N6050\lesssim N_{\ast}\lesssim 60 [27]. However, models that unifying inflation with dark energy may require a slightly larger NN_{\ast}. For example, the quintessential inflation generally requires N>60N_{\ast}>60 [44, 45]. In this section (especially in Fig. 1), we assume 50N7050\lesssim N_{\ast}\lesssim 70 for our model. The theoretical uncertainty of NN_{\ast} mainly comes from the modeling of the reheating process [46, 47, 48]. In Sec. IV, we present a specific reheating mechanism for our model. After that, a self-consistent calculation of NN_{\ast} is presented in the Appendix.

Before performing the calculations based on Eqs. (6) and (8), here we first use Taylor expansion to estimate the values of nsn_{\rm s} and rr. Substituting Eq. (8) into Eq. (6a), expanding the result around λ2=0\lambda_{2}=0 and N=+N_{\ast}=+\infty, and keeping the 𝒪(N1)\mathcal{O}(N_{\ast}^{-1}) terms, we obtain

ns=11N(4π2β23π4β4180π6β67560+),n_{\rm s}=1-\frac{1}{N_{\ast}}\left(4-\frac{\pi^{2}\beta^{2}}{3}-\frac{\pi^{4}\beta^{4}}{180}-\frac{\pi^{6}\beta^{6}}{7560}+\cdots\right), (9a)
where β=λ1λ2N/(απ)\beta=\sqrt{\lambda_{1}\lambda_{2}}N_{\ast}/(\alpha\pi). Note that, inspired by Eq. (5), we assumed β=𝒪(1)\beta=\mathcal{O}(1) in the third step to derive Eq. (9a). Planck 2018 observation gives ns0.965n_{\rm s}\approx 0.965 [27], which corresponds to the expression in above parentheses approximately equal to 22, i.e., β0.74\beta\approx 0.74. Here β<1\beta<1 (N<απ/λ1λ2N_{\ast}<\alpha\pi/\sqrt{\lambda_{1}\lambda_{2}}) indicates ϕ>απ\phi_{\ast}>\alpha\pi. Therefore, it is reasonable to assume β=𝒪(1)\beta=\mathcal{O}(1) and ϕ>απ\phi_{\ast}>\alpha\pi in the previous derivations. Considering the typical values of λ1\lambda_{1}, α\alpha and NN_{\ast}, β0.74\beta\approx 0.74 requires λ2=𝒪(104)\lambda_{2}=\mathcal{O}(10^{-4}). This result is consistent with the constraint given by Eq. (5) for successful inflation. For rr, similar calculation gives
r=𝒪(1)×128α4λ12N4,r=\mathcal{O}(1)\times\frac{128\alpha^{4}}{\lambda_{1}^{2}N_{\ast}^{4}}, (9b)

which indicates generally r107r\sim 10^{-7}. Planck 2018 [27] together with BICEP2/Keck Array BK15 [49] data requires r<0.06r<0.06 at 95%95\% CL. Therefore, for both nsn_{\rm s} and rr, the predictions of the inflation model described by Eq. (1) and current observations are in good agreement.

Fig. 1 plots Planck 2018+BK15 measurements and the results given by Eqs. (6) and (9). Comparing the (ns,rn_{\rm s},r) values of the three marked points in the blue and black solid lines with N=50N_{\ast}=50, we conclude that Eq. (9) is a good approximation of Eq. (6). The red and blue regions confirm the agreement between theory and observations. Note that Planck 2018+BK15 measurements rely on the Λ\LambdaCDM model. For self-consistency, we should replace the Λ\LambdaCDM model with the oscillating dark energy model in the parameter constraints. However, as shown in [15], the influences of oscillating dark energy on the late-time evolution of cosmic inhomogeneities are generally small. For a rough estimate, it seems reasonable to adopt the Λ\LambdaCDM-dependent results of the primordial inhomogeneities. The self-consistent constraints will be presented in a later paper.

Refer to caption
Figure 1: Predictions of the inflation model described by Eq. (1) in the plane (ns,r)(n_{\rm s},r) together with Planck 2018+BK15 measurements [27]. The red regions are the Planck 2018+BK15 68%68\% and 95%95\% CL contours. The blue region shows the result given by Eq. (6) with λ1=4.5\lambda_{1}=4.5, α=0.6\alpha=0.6, 50N7050\leqslant N_{\ast}\leqslant 70 and 0.3β0.950.3\leqslant\beta\leqslant 0.95, which corresponds to 2.8×105λ22.9×1042.8\times 10^{-5}\leqslant\lambda_{2}\leqslant 2.9\times 10^{-4} for N=50N_{\ast}=50. We vary λ2\lambda_{2} in the calculations and mark the corresponding β\beta in the plot. The black solid line shows the result given by Eq. (9) with the same parameters and β\beta-marks of the top blue line. For clarity, we multiply Eq. (9b) by 1010 to separate the black and blue plots. As Eq. (1) can be regarded as a modification of the exponential potential, we plot the predictions of the corresponding power-law inflation [47] for comparison (ns=1r/8n_{\rm s}=1-r/8, black dashed line).

The amplitude of the primordial inhomogeneities is related to the energy scale of inflation. In slow-roll approximation, the spectrum of scalar perturbations generated during inflation can be written as 𝒫(k)=As(k/k)ns1\mathcal{P}_{\mathcal{R}}(k)=A_{\rm s}(k/k_{\ast})^{n_{\rm s}-1}, where the amplitude

As=κH28π2cϵ(kcaeH)ns1,A_{\rm s}=\frac{\hbar\kappa H^{2}}{8\pi^{2}c\epsilon}\left(\frac{k_{\ast}c}{a_{e}H}\right)^{n_{\rm s}-1}, (10)

and kk is the comoving wavenumber, kk_{\ast} denotes the characteristic wavenumber that crossed the horizon at a=aa=a_{\ast}, the values of ϵ\epsilon and nsn_{\rm s} are evaluated at horizon crossing. In principle, HH should be evaluated at inflation ending as indicated by the exact solution of power-law inflation. However, the decrease in HH during inflation is small (especially in our model) and we can ignore its variation. In our conventions, we have

H2c2V03exp[λ1+λ22απ]c2V0103,H^{2}\approx\frac{c^{2}V_{0}}{3}\exp\left[-\frac{\lambda_{1}+\lambda_{2}}{2}\alpha\pi\right]\sim\frac{c^{2}V_{0}}{10^{3}}, (11)

where the last approximation adopts the typical values of the parameters. The relation between kk_{\ast} and NN_{\ast} is

eNaaeaHaeHe=kcaeHe.e^{-N_{\ast}}\equiv\frac{a_{\ast}}{a_{e}}\approx\frac{a_{\ast}H_{\ast}}{a_{e}H_{e}}=\frac{k_{\ast}c}{a_{e}H_{e}}. (12)

Substituting the above results and ns12/Nn_{\rm s}\approx 1-2/N_{\ast} into Eq. (10), we obtain

AsκH28π2cϵe22GH2c5ϵ.A_{\rm s}\approx\frac{\hbar\kappa H^{2}}{8\pi^{2}c\epsilon}e^{2}\approx 2\frac{\hbar GH^{2}}{c^{5}\epsilon}. (13)

Planck 2018 data gives As2×109A_{\rm s}\approx 2\times 10^{-9} [27]. At the time of horizon crossing, considering ϵ=r/16108\epsilon=r/16\sim 10^{-8} in our model, we obtain the following equivalent results

H\displaystyle H 3×109tP1,\displaystyle\sim 3\times 10^{-9}\,t_{\rm P}^{-1}, (14a)
V0\displaystyle V_{0} 1014lP2,\displaystyle\sim 10^{-14}\,l_{\rm P}^{-2}, (14b)
ρinf\displaystyle\rho_{\rm inf} 1018ρP,\displaystyle\sim 10^{-18}\,\rho_{\rm P}, (14c)

where tP=G/c5t_{\rm P}=\sqrt{\hbar G/c^{5}} is the Planck time, lP=G/c3l_{\rm P}=\sqrt{\hbar G/c^{3}} is the Planck length, ρinf\rho_{\rm inf} is the energy density of inflaton and ρP=c7/(G2)\rho_{\rm P}=c^{7}/(\hbar G^{2}) is the Planck energy density. The above result shows ρinfρP\rho_{\rm inf}\ll\rho_{\rm P} in our model. Actually, ρinfρP\rho_{\rm inf}\ll\rho_{\rm P} is a general requirement for all the slow-roll inflation models. In the early 1980s, people thought this is a serious problem of the inflation theory (see the discussions in the Abstract of [32, 34, 35]). However, this issue has gradually lost people’s attention in the follow-up development. The mainstream view of current cosmology community seems to admit that ρinfρP\rho_{\rm inf}\ll\rho_{\rm P} is not a problem. We disagree with that. In our model, a much smaller rr does exacerbate the problem, i.e., we need a much smaller ρinf\rho_{\rm inf}. However, the essence of this problem remains unchanged.

II.3 Smallness of rr in the general MAU

Equation (1) behaves well in the inflationary stage. Especially, it predicts an extremely small rr. It is natural to ask whether the small rr is a general consequence of the MAU. We can generalize Eq. (1) to

V(ϕ)=V0exp[λ1+λ22ϕα(λ1λ2)2f(ϕα)],V(\phi)=V_{0}\exp\left[-\frac{\lambda_{1}+\lambda_{2}}{2}\phi-\frac{\alpha(\lambda_{1}-\lambda_{2})}{2}f(\frac{\phi}{\alpha})\right], (15)

where f(x)f(x) is periodic with period approximately 2π2\pi and 1df/dx1-1\leqslant{\rm d}f/{\rm d}x\leqslant 1. The model proposed in [9] corresponds to fln[1+Asin(Bϕ)]f\propto\ln[1+A\sin(B\phi)]. Without loss of generality, we can assume 0λ2<λ10\leqslant\lambda_{2}<\lambda_{1} and α=𝒪(1)\alpha=\mathcal{O}(1), which is necessary to solve the coincidence problem [21]. The slope of this potential satisfies λ2λλ1\lambda_{2}\leqslant\lambda\leqslant\lambda_{1}. It is natural to assume that λ\lambda is increasing from horizon crossing to inflation ending. Then, integrating Eq. (3) gives

N=ϕϕedϕλ(ϕ)ϕϕedϕλ1λ,N_{\ast}=\int_{\phi_{\ast}}^{\phi_{e}}\frac{{\rm d}\phi}{\lambda(\phi)}\lesssim\int_{\phi_{\ast}}^{\phi_{e}}\frac{{\rm d}\phi}{\lambda_{\ast}}\approx\frac{1}{\lambda_{\ast}}, (16)

where λ=λ(ϕ)\lambda_{\ast}=\lambda(\phi_{\ast}) and the last approximation uses α=𝒪(1)\alpha=\mathcal{O}(1). Equation (16) gives λ1/N\lambda_{\ast}\lesssim 1/N_{\ast}. Then we obtain

r=16ϵ=8λ28/N2.r=16\epsilon=8\lambda_{\ast}^{2}\lesssim 8/N_{\ast}^{2}. (17)

For the typical value N=60N_{\ast}=60, the above equation gives r0.002r\lesssim 0.002. The upper limit may change slightly in a specific model due to the missing coefficient. Conservatively, we may conclude r<0.01r<0.01 for the general MAU model. For the traditional inflation models, some of them predict r>0.01r>0.01, e.g., chaotic inflation [30] and other models discussed in [50], while some predict r<0.01r<0.01, e.g., Starobinsky R2R^{2} inflation [51] and Higgs inflation [52]. Observational constraints on rr cannot distinguish our model from other traditional models that predict r<0.01r<0.01. However, our calculations provide a physical motivation for the small rr: if one want to use the MAU to solve the coincidence problem and unify inflation with dark energy in the framework of Eq. (15), then r<0.01r<0.01.

Currently, Planck 2018+BK15 data gives the strongest limit r<0.06r<0.06 at 95%95\% CL [27]. There are some ongoing and upcoming cosmic microwave background polarization experiments to detect primordial gravitational waves, such as BICEP/Keck [49], SPTpol [53], AliCPT [54], LiteBIRD [55], CMB-S4 [56] and so on. Part of these experiments may reach an upper limit of r<0.001r<0.001. Observational confirmation of r>0.01r>0.01 will make observations disfavor the MAU scenario. In contrast, if future observations limit rr to less than 0.010.01, then this can be regarded as a circumstantial evidence for the MAU.

III Post-inflationary evolution

Generally, the single scalar field slow-roll inflation ends once the potential becomes steep, and then the scalar field oscillates near the minimum of its potential [57]. However, things will be different in our model as there is no local minimum of the potential (note that 0<λ2λλ10<\lambda_{2}\leqslant\lambda\leqslant\lambda_{1}). A monotonic potential will drive the scalar field roll to infinity. Here we present a quantitative analysis of this process. In this section, reheating and its backreaction on the background dynamics [57] are not taken into account.

Here the slow-roll approximation is no longer applicable due to the non-negligible kinetic energy of the scalar field. Inspired by [22], what we need may be the dynamical system form of Eq. (2). Introducing the dimensionless variables x1ϕ˙/(6H)x_{1}\equiv\dot{\phi}/(\sqrt{6}H) and ν6(λ2V′′/V)\nu\equiv\sqrt{6}(\lambda^{2}-V^{\prime\prime}/V), Eq. (2) can be rewritten as

dx1dN\displaystyle\frac{{\rm d}x_{1}}{{\rm d}N} =(1x12)(62λ3x1),\displaystyle=(1-x_{1}^{2})(\frac{\sqrt{6}}{2}\lambda-3x_{1}), (18a)
dλdN\displaystyle\frac{{\rm d}\lambda}{{\rm d}N} =νx1,\displaystyle=\nu x_{1}, (18b)
dνdN\displaystyle\frac{{\rm d}\nu}{{\rm d}N} =3x1α2(λ1+λ22λ),\displaystyle=\frac{3x_{1}}{\alpha^{2}}(\lambda_{1}+\lambda_{2}-2\lambda), (18c)

where Nln(a/ai)N\equiv\ln(a/a_{i}). A constraint equation for λ\lambda and ν\nu is given by Eq. (9) in [21]. To characterize the evolution of this system, we use the EOS wϕ=2x121w_{\phi}=2x_{1}^{2}-1 [21] and the Hubble slow-roll parameters [42]

ϵH\displaystyle\epsilon_{H} H˙H2=ϕ˙22H2=3x12,\displaystyle\equiv-\frac{\dot{H}}{H^{2}}=\frac{\dot{\phi}^{2}}{2H^{2}}=3x_{1}^{2}, (19a)
ηH\displaystyle\eta_{H} ϕ¨Hϕ˙=36λ(1x12)2x1,\displaystyle\equiv-\frac{\ddot{\phi}}{H\dot{\phi}}=3-\frac{\sqrt{6}\lambda(1-x_{1}^{2})}{2x_{1}}, (19b)

where the above derivations following the definitions use Eq. (2). To lowest order, we have ϵH=ϵV\epsilon_{H}=\epsilon_{V}, ηH=ηVϵV\eta_{H}=\eta_{V}-\epsilon_{V} and ns=1+2ηH4ϵHn_{\rm s}=1+2\eta_{H}-4\epsilon_{H}. The use of Hubble slow-roll parameters facilitates the analysis based on the dynamical system Eq. (18).

To get a first glance of the system’s property, based on Eq. (18), Fig. 2 plot the evolution of wϕw_{\phi}, ϵH\epsilon_{H} and ηH\eta_{H} for a representative set of model parameters. This figure shows that inflation ends naturally in our model as we expected. Especially, inflation will not happen again when the scalar field passes through the second flat part (ϕ3απ\phi\approx 3\alpha\pi in our settings, see the first paragraph in Sec. II). The reason is the scalar field accumulates sufficient kinetic energy as it passes through the first steep part (ϕ2απ\phi\approx 2\alpha\pi) and then it will quickly pass through the following flat parts.

Refer to caption
Figure 2: Evolution of wϕw_{\phi}, ϵH\epsilon_{H} and ηH\eta_{H}. The model parameters are λ1=4.5\lambda_{1}=4.5, λ2=1.73×104\lambda_{2}=1.73\times 10^{-4} (β=0.74\beta=0.74), α=0.6\alpha=0.6 and N=50N_{\ast}=50. For the initial conditions, we assume ϕi=απ\phi_{i}=\alpha\pi, which corresponds to λ=λ2\lambda=\lambda_{2} and ν=0\nu=0 at the beginning of inflation. In addition, we assume the slow-roll initial condition for x1x_{1}, and then Eq. (19a) gives x1=ϵV(ϕi)/3=λ2/6x_{1}=\sqrt{\epsilon_{V}(\phi_{i})/3}=\lambda_{2}/\sqrt{6}. The second vertical dashed line corresponds to N=απ/λ1λ2=67.56N=\alpha\pi/\sqrt{\lambda_{1}\lambda_{2}}=67.56 [near inflation ending, see Eq. (5)]. Strictly speaking, inflation ends at N=69.90N=69.90, at which ϵH=1\epsilon_{H}=1 and a¨/a=0\ddot{a}/a=0. The first vertical dashed line corresponds to N=67.56NN=67.56-N_{\ast} (horizon crossing). In the subplot (d), the minor ticks from 10i-10^{-i} to 10i1-10^{-i-1} correspond to 10i×(1/2,1/3,,1/9)-10^{-i}\times(1/2,1/3,\cdots,1/9) respectively. The spectral index ns=1+2ηH4ϵHn_{\rm s}=1+2\eta_{H}-4\epsilon_{H} is given in the subplot (b). For the above settings, ns=0.961n_{\rm s}=0.961 at horizon crossing [meanwhile Eq. (9a) gives a consistent result ns=0.960n_{\rm s}=0.960]. The motion of the scalar field in its potential is illustrated in the subplot (a).

After the end of inflation, Fig. 2 shows the existence of oscillation. Here we investigate the oscillation frequency and the average EOS w¯ϕ\overline{w}_{\phi} during oscillation. Similar to [21, 22], the result of exponential potential may provide direct clues. For V(ϕ)exp(λϕ)V(\phi)\propto\exp(-\lambda\phi), where λ\lambda is a positive constant, the system’s evolution equation is given by Eq. (18a), and the stable critical point is x1=λ/6x_{1}=\lambda/\sqrt{6} if λ<6\lambda<\sqrt{6} and x1=1x_{1}=1 if λ6\lambda\geqslant\sqrt{6} [58]. For the dynamical system Eq. (18), considering wϕ=2x121w_{\phi}=2x_{1}^{2}-1 and the results obtained for the pure exponential potential, we may be able to find an approximate EOS

w¯ϕ,app={(λ1+λ2)2/121if λ1+λ2<26,1if λ1+λ226,\overline{w}_{\phi,{\rm app}}=\left\{\begin{array}[]{ll}(\lambda_{1}+\lambda_{2})^{2}/12-1&\textrm{if $\lambda_{1}+\lambda_{2}<2\sqrt{6}$},\\ 1&\textrm{if $\lambda_{1}+\lambda_{2}\geqslant 2\sqrt{6}$},\end{array}\right. (20)

during oscillation. As wϕ1w_{\phi}\leqslant 1 for all the quintessence models, we guess that the oscillation only occurs when w¯ϕ<1\overline{w}_{\phi}<1, i.e., approximately λ1+λ2<26\lambda_{1}+\lambda_{2}<2\sqrt{6}. Otherwise, if w¯ϕ=1\overline{w}_{\phi}=1, there will be no upper room for wϕw_{\phi} to be oscillating. To find an approximation of the oscillation frequency within this parameter space, we can use the Fourier series method to solve Eq. (18). Details can be found in the appendix of [22], and the only difference is that the background values should be x1=(λ1+λ2)/24x_{1}=(\lambda_{1}+\lambda_{2})/\sqrt{24}, λ=(λ1+λ2)/2\lambda=(\lambda_{1}+\lambda_{2})/2 and ν=0\nu=0. Note that this setting is consistent with the λ\lambda-ν\nu constraint given by Eq. (9) in [21] as the background means λ1=λ2\lambda_{1}=\lambda_{2}. The result of the oscillation frequency is

fos=λ1+λ24απ.f_{\rm os}=\frac{\lambda_{1}+\lambda_{2}}{4\alpha\pi}. (21)

Note that, in principle, we can only prove Eqs. (20) and (21) hold for λ1λ21\lambda_{1}-\lambda_{2}\ll 1, which is outside the model’s viable parameter space.

Fig. 3 plots the numerical results of w¯ϕ\overline{w}_{\phi} and the Fourier transform x~1(f)\tilde{x}_{1}(f), together with Eqs. (20) and (21). When λ1λ2\lambda_{1}\approx\lambda_{2} (λ1λ21\lambda_{1}-\lambda_{2}\ll 1), for both w¯ϕ\overline{w}_{\phi} and fosf_{\rm os}, the numerical and approximate results are in good agreement as we expected. The difference between the numerical results and the approximations grows as λ1\lambda_{1} increases. However, the difference miraculously becomes negligible when λ14\lambda_{1}\gtrsim 4. Especially, the boundary λ1+λ2=264.899\lambda_{1}+\lambda_{2}=2\sqrt{6}\approx 4.899 is verified. Therefore, for the viable parameter space (λ1+λ2>4\lambda_{1}+\lambda_{2}>4), Eqs. (20) and (21) are good approximations of the exact results. Equation (20) together with λ1+λ2>4\lambda_{1}+\lambda_{2}>4 give w¯ϕ>1/3\overline{w}_{\phi}>1/3, i.e., w¯ϕ\overline{w}_{\phi} is larger than the EOS of radiation when the Universe is dominated by the oscillating scalar field. If radiation is generated during this period, then its relative energy density will increase with cosmic expansion and the Universe will gradually enter the radiation era. Fig. 3 also shows that there is no period-doubling bifurcation and chaos in the three-dimensional dynamical system Eq. (18). This is different with the properties of the four-dimensional dynamical system discussed in [22].

Refer to caption
Refer to caption
Figure 3: Left: Fourier transform of x1(N)x_{1}(N). The result is normalized by the signal length. What we concern is the location of the peak rather than its amplitude. Right: w¯ϕ\overline{w}_{\phi} versus λ1\lambda_{1}. The numerical result is given by w¯ϕ,num=(N2N1)1N1N2wϕ(N~)dN~\overline{w}_{\phi,{\rm num}}=(N_{2}-N_{1})^{-1}\int_{N_{1}}^{N_{2}}w_{\phi}(\tilde{N}){\rm d}\tilde{N}, which characterizes the density decrease rate (see Eq. (1) in [59]). The model parameters are λ2=0.1\lambda_{2}=0.1, α=0.6\alpha=0.6 and λ1\lambda_{1} can be found in the figures. The initial conditions are x1=0.5x_{1}=0.5, λ=(λ1+λ2)/2\lambda=(\lambda_{1}+\lambda_{2})/2 and ν=ν+(λ)\nu=\nu_{+}(\lambda) (see Eq. (9) in [21]). We numerically solve Eq. (18) in N[0,105]N\in[0,10^{5}], perform Fourier transform and calculate w¯ϕ,num\overline{w}_{\phi,{\rm num}} in N[500,105]N\in[500,10^{5}]. The numerical results are plotted in red, while Eqs. (20) and (21) are plotted in blue. The gray dashed line corresponds to the boundary λ1+λ2=26\lambda_{1}+\lambda_{2}=2\sqrt{6}.

Terminology might be mentioned. In the literature, the decelerating post-inflationary period may be called “deflationary” [60] or “kination” [61]. In our model, the post-inflationary dynamics is dominated by the scalar field with an average EOS w¯ϕ>1/3\overline{w}_{\phi}>1/3. In this period, both the kinetic and potential energy of the scalar field are not negligible. Therefore, we call this period “deflation” instead of “kination”.

IV Gravitational reheating

At the end of the inflation, the Universe is very cold due to the fast cosmic expansion. A mechanism is needed to reheat the Universe [23]. In modern cosmology, the reheating is typically assumed to occur through the decay of inflaton, which is sourced by the nonminimal coupling of inflaton and other fields [62, 63, 64, 65, 66, 67]. A less commonly discussed case is gravitational reheating, i.e., particle creation in curved spacetime. This scenario was first investigated in [68] and applied to the quintessential inflation [24]. One important feature of the gravitational reheating is that it does not need to introduce any nonminimal coupling. Here we study whether the gravitational reheating can successfully create the hot Universe in our model. Especially, we concentrate on the radiation temperature TrehT_{\rm reh} at the end of deflation, i.e., the beginning of radiation era. The requirement is Treh1011KT_{\rm reh}\gtrsim 10^{11}\,{\rm K} (10MeV\sim 10\,{\rm MeV} in natural units, i.e., the temperature at which primordial nucleosynthesis starts [46]).

The Universe is assumed to undergo a transition from inflation (nearly de Sitter spacetime) to deflation (decelerating phase). For a massless scalar field, the created energy density at the end of inflation from the inflation-deflation transition is

ρr(ae)=Rρinf2/ρP,\rho_{\rm r}(a_{e})=R\rho_{\rm inf}^{2}/\rho_{\rm P}, (22)

where R0.01R\sim 0.01 for arbitrary power-law deflation [68, 69], and ρinf\rho_{\rm inf} is given by Eq. (14c) in our model. This creation can be regarded as an instantaneous process. Based on the calculations presented in Sec. III, it is natural to assume that Eq. (22) applies to our model. There may be many scalar fields in the Universe and each one contributes 0.010.01 to RR. Following [24], hereafter we assume 0.01R10.01\lesssim R\lesssim 1, which corresponds to Tr(ae)1023KT_{\rm r}(a_{e})\sim 10^{23}\,{\rm K}, where Tr(ae)T_{\rm r}(a_{e}) is the radiation temperature at the end of inflation (after reheating). During deflation, the radiation energy density ρr=ρr(ae)(ae/a)4\rho_{\rm r}=\rho_{r}(a_{e})\cdot(a_{e}/a)^{4} and the inflaton energy density ρϕρinf(ae/a)3(1+w¯ϕ)\rho_{\phi}\approx\rho_{\rm inf}\cdot(a_{e}/a)^{3(1+\overline{w}_{\phi})}, where 1/3<w¯ϕ11/3<\overline{w}_{\phi}\leqslant 1. Thus, the ratio of ρr\rho_{\rm r} and ρϕ\rho_{\phi} is

ρr/ρϕ1018R(a/ae)3w¯ϕ1.\rho_{\rm r}/\rho_{\phi}\approx 10^{-18}\cdot R\cdot(a/a_{e})^{3\overline{w}_{\phi}-1}. (23)

Radiation era thus begins at the scale factor

arehae(1018/R)1/(3w¯ϕ1).a_{\rm reh}\approx a_{e}\cdot(10^{18}/R)^{1/(3\overline{w}_{\phi}-1)}. (24)

Meanwhile the radiation temperature TrehT_{\rm reh} is

Treh\displaystyle T_{\rm reh} =[15c33ρr(areh)π2kB4]1/4\displaystyle=\left[\frac{15c^{3}\hbar^{3}\rho_{\rm r}(a_{\rm reh})}{\pi^{2}k_{\rm B}^{4}}\right]^{1/4}
={1.6×1014R0.75Kif w¯ϕ=1,4.1×1012R0.84Kif w¯ϕ=0.9,2.2×1010R0.96Kif w¯ϕ=0.8,\displaystyle=\left\{\begin{array}[]{ll}1.6\times 10^{14}R^{0.75}\,{\rm K}&\textrm{if $\overline{w}_{\phi}=1$,}\\ 4.1\times 10^{12}R^{0.84}\,{\rm K}&\textrm{if $\overline{w}_{\phi}=0.9$,}\\ 2.2\times 10^{10}R^{0.96}\,{\rm K}&\textrm{if $\overline{w}_{\phi}=0.8$,}\end{array}\right. (28)

where kBk_{\rm B} is the Boltzmann constant. Therefore, it is possible to realize Treh1011KT_{\rm reh}\gtrsim 10^{11}\,{\rm K} with suitable parameters in our model. For example, parameters with w¯ϕ>0.825\overline{w}_{\phi}>0.825 if R=1R=1 and w¯ϕ>0.903\overline{w}_{\phi}>0.903 if R=0.01R=0.01. Considering Eq. (20) and λ21\lambda_{2}\ll 1, we obtain λ1>4.68\lambda_{1}>4.68 and λ1>4.78\lambda_{1}>4.78 respectively.

V Discussion

Regarding the late-time acceleration as one of the accelerating phases in the MAU provides an attractive solution to the coincidence problem. Meanwhile, MAU provides a natural unification of inflation and dark energy. In this paper, we analyze the inflationary consequences of the MAU model described by Eq. (1). This model with λ2=𝒪(104)\lambda_{2}=\mathcal{O}(10^{-4}) gives sufficient e-folding number, suitable nsn_{\rm s} and extremely small rr. The post-inflationary expansion is decelerating with an average EOS w¯ϕ>1/3\overline{w}_{\phi}>1/3. In this framework, we show that gravitational particle creation at the end of inflation is sufficient to reheat the hot Universe.

Together with our previous works, here we summarize two observational probes for the general MAU scenario. One probe is the primordial gravitational waves: MAU generally predicts r<0.01r<0.01. The physical origin of this upper limit is that the coincidence problem requires the increment of ϕ\phi cannot be too large during one accelerating phase (including inflation) [21]. During inflation, in order to obtain sufficient e-folding number, the potential must be very flat, which results in small ϵV\epsilon_{V} and rr. The other probe is the Hubble expansion rate at cosmic dawn (denoted as HcdH_{\rm cd}) measured by the 21 cm signals [22, 70]. In the MAU, the scalar field is comparable with the normal matters across the whole cosmic history, and thus HcdH_{\rm cd} should be much larger than that in the Λ\LambdaCDM model. We expect future observations can prove or disprove these two predictions. But note that the MAU is not the unique model to obtain small rr or large HcdH_{\rm cd}. For example, Starobinsky R2R^{2} inflation also gives small rr [51, 27] and some interacting dark energy models may give large HcdH_{\rm cd} [71]. What we have done is to provide a physical motivation (i.e., the cosmological coincidence problem) for the small rr and large HcdH_{\rm cd}.

There are several unnatural features of our model. In principle, arbitrary V0V_{0} can be used to explain the late-time cosmic acceleration [21]. However, in order to obtain a suitable AsA_{\rm s}, we have to specify an extremely small value for V0V_{0} relative to the Planck scale values. This is unnatural (a fine-tuning problem), as in the conventional inflation models [32, 34, 35]. In addition, the dimensionless parameter λ2=𝒪(104)\lambda_{2}=\mathcal{O}(10^{-4}) indicates the other minor fine-tuning problem in our model. We believe that nature (cosmology) prefers models that only introduce Planck scale parameters and dimensionless parameters of order unity. Inspired by [24], those two fine-tuning problems may can be addressed in models with an extra inflation field χ\chi. In the new model, both V0V_{0} and λ2\lambda_{2} are not constants but slowly varying functions of χ\chi. This possibility will be explored in the future.

Acknowledgements

This work was supported by the National Natural Science Foundation of China under Grants Nos. 11633001, 11920101003 and 12021003, the Strategic Priority Research Program of the Chinese Academy of Sciences under Grant No. XDB23000000 and the Interdiscipline Research Funds of Beijing Normal University. S.X.T. was supported by the Initiative Postdocs Supporting Program under Grant No. BX20200065.

Appendix A Self-consistent calculation of NN_{\ast}

In Sec. II, we assumed 50N7050\leqslant N_{\ast}\leqslant 70 without considering the post-inflationary dynamics. However, specification of the reheating mechanism (see Sec. IV) enables us to self-consistently calculate NN_{\ast}. The purpose of this Appendix is to do so. As we show in the following, our model requires a slightly larger NN_{\ast} (N60N_{\ast}\gtrsim 60), which is similar to the result obtained for the quintessential inflation [44, 45]. This similarity mainly comes from the same reheating mechanism and similar deflationary dynamics in our model and the quintessential inflation.

During inflation, the mode kk_{\ast} crossed the horizon at kc=aHk_{\ast}c=a_{\ast}H_{\ast}. The standard way to calculate NN_{\ast} is to write down

kca0H0=aHa0H0=aaeaea0HH0,\displaystyle\frac{k_{\ast}c}{a_{0}H_{0}}=\frac{a_{\ast}H_{\ast}}{a_{0}H_{0}}=\frac{a_{\ast}}{a_{e}}\frac{a_{e}}{a_{0}}\frac{H_{\ast}}{H_{0}}, (29)

where the subscript 0 means today. Compared with Eq. (5.13) in [72], here we do not introduce the scale factor areha_{\rm reh}. The reason is, as we mentioned in Sec. IV, the reheating is instantaneous at the end of inflation. The created radiation energy density is given by Eq. (22). After reheating, the radiation energy density is proportional to a4a^{-4}, which gives

aea0=[ρr(a0)ρr(ae)]1/4=exp(52.414lnR).\frac{a_{e}}{a_{0}}=\left[\frac{\rho_{\rm r}(a_{0})}{\rho_{\rm r}(a_{e})}\right]^{1/4}=\exp\left(-52.4-\frac{1}{4}\ln R\right). (30)

Equation (14a) gives H/H0=e120.6H_{\ast}/H_{0}=e^{120.6}, where we adopt H0=70km/s/MpcH_{0}=70\,{\rm km}/{\rm s}/{\rm Mpc}. Substituting the above results into Eq. (29), we obtain

Nlnaea=68.214lnRlnkca0H0.N_{\ast}\equiv\ln\frac{a_{e}}{a_{\ast}}=68.2-\frac{1}{4}\ln R-\ln\frac{k_{\ast}c}{a_{0}H_{0}}. (31)

For the typical values 0.01R10.01\leqslant R\leqslant 1 and k/a0=0.002Mpc1k_{\ast}/a_{0}=0.002\,{\rm Mpc}^{-1} [27], we obtain 67.2N66.167.2\geqslant N_{\ast}\geqslant 66.1.

References