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

Enhancement of small-scale induced gravitational waves from the soliton/oscillon domination

Xiao-Bin Sui [email protected] School of Fundamental Physics and Mathematical Sciences, Hangzhou Institute for Advanced Study, University of Chinese Academy of Sciences (HIAS-UCAS), Hangzhou 310024, China CAS Key Laboratory of Theoretical Physics, Institute of Theoretical Physics, Chinese Academy of Sciences, Beijing 100190, China University of Chinese Academy of Sciences, Beijing 100049, China    Jing Liu [email protected] International Centre for Theoretical Physics Asia-Pacific, University of Chinese Academy of Sciences, Beijing 100190, China Taiji Laboratory for Gravitational Wave Universe (Beijing/Hangzhou), University of Chinese Academy of Sciences, Beijing 100049, China    Rong-Gen Cai [email protected] Institute of Fundamental Physics and Quantum Technology, Ningbo University, Ningbo, 315211, China CAS Key Laboratory of Theoretical Physics, Institute of Theoretical Physics, Chinese Academy of Sciences, Beijing 100190, China School of Fundamental Physics and Mathematical Sciences, Hangzhou Institute for Advanced Study, University of Chinese Academy of Sciences (HIAS-UCAS), Hangzhou 310024, China
Abstract

We investigate density fluctuations and scalar-induced gravitational waves (GWs) arising from the production of long-lived solitons and oscillons, which can dominate the early Universe and drive reheating prior to the standard radiation-dominated era. Curvature perturbations are generated not only by the Poisson distribution of these solitons/oscillons but are also amplified during the early matter-dominated phase when the solitons/oscillons dominate. Scalar-induced GWs with characteristic energy spectra emerge from these amplified curvature perturbations, particularly at the sudden transition from matter domination to radiation domination. We analyze this scenario and its GW predictions in detail, focusing on the impact of the initial energy fraction, the lifetime, and the average separation of these solitons/oscillons. This provides a wealth of potential observational targets for various GW detectors. Furthermore, for inflation and preheating models that produce oscillons with large separations, we derive new constraints on these models based on the upper bounds on the effective number of relativistic degrees of freedom, as anticipated from future cosmic microwave background and large-scale structure observations.

I Introduction

Gravitational waves (GWs) have been attracting more and more attention on account of their potential to provide valuable information about both merger events of compact binaries and the early Universe  [1, 2]. Stochastic gravitational waves backgrounds (SGWBs) induced by curvature perturbations are one of the most important cosmological GW sources tied to the evolution of the early Universe, which are expected to generate both from vacuum quantum fluctuations during inflation [3, 4, 5, 6, 7, 8, 9, 10, 6, 11, 12, 13, 14] and the violent physical processes after inflation [15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25]. Such SGWBs can be explored through multiband GW projects such as LIGO/VIRGO/KAGRA [26], Taiji [27, 28], TianQin [29] and LISA [30].

In this work, we aim to utilize scalar-induced GWs induced by scalar perturbations to explore the early Universe and underlying new physics. Despite the successful agreement between the standard model of cosmology and observations, the transition from the inflationary epoch to the standard radiation-dominated (sRD) era remains poorly understood. Numerous studies suggest the existence of an early matter-dominated (eMD) era following inflation, prior to the sRD era [31, 32, 33, 34, 35, 36, 37]. In the eMD era, scalar perturbations remain constant even on sub-horizon scales [38]. If the transition from the eMD era to the sRD era is rapid, a sudden change takes place in the equation of state parameter for the Universe, leading to a significant enhancement in the production of scalar-induced GWs [31, 34, 35, 36, 37].

The nonlinear interaction of the matter field can result in intriguing physical phenomena, including the formation of solitons and oscillons. [39, 40, 41, 42, 43, 44, 45, 46, 43, 47, 48, 49]. The longevity of these compact objects can stem from various reasons. Some configurations, such as Q-balls, are stable due to the conservation of topological or nontopological charges [39, 40, 41, 42], while others are sustained by a dynamic equilibrium between attractive forces and dissipative forces, such as oscillons [43, 44, 45, 46, 43, 47]. The mass distribution of stable solitons/oscillons is sharply peaked, and they decay rapidly when they reach a critical value. Therefore, we can treat the transition from the eMD era to the sRD era as an instantaneous process. For simplicity, we will refer to them collectively as solitons in the following unless otherwise specified.

The separation of solitons increases along with the expansion of the Universe while the physical size remains fixed, which implies that the solitons can be treated as point-like particles for most of the eMD era. Since the formation of solitons is independent of each other, the statistics of solitons follow a Poisson distribution, resulting in additional curvature perturbations with a peaked power spectrum upon the nearly scale-invariant one from quantum fluctuations during inflation. Additionally, the solitons evolve akin to dust gas and can eventually dominate the Universe if their lifetimes significantly exceed the Hubble time. These induced curvature perturbations grow at the subhorizon scales during the matter-dominated era. The rapid transition from the eMD era to the sRD era also amplifies the GW production, leaving distinct imprints on the GW energy spectrum. Furthermore, we perform a detailed analysis of the soliton models and the influence on the GW predictions. The key parameters include the soliton lifetime, the initial energy fraction of solitons, and the mean separation of solitons. These free factors of models can affect the duration of the eMD era, the amplitude and the characteristic scale of Poisson-induced perturbations. The prediction of the GW energy spectra lies within the sensitivity curves of various GW detectors, where the peak frequency mainly depends on the energy scale of the Universe at the matter-radiation transition. Additionally, scalar-induced GWs that behave like a dark radiation component, thus contributing to the relativistic degrees of freedom ΔNeff\Delta N_{\text{eff}}. A higher value of ΔNeff\Delta N_{\text{eff}}, would delay the radiation-matter equality and change the size of the sound horizon [50], leaving distinct signatures in the cosmic microwave background (CMB) anisotropies [51], baryon acoustic oscillations, and Big Bang nucleosynthesis [52]. The latest Planck CMB data constrain ΔNeff<0.3\Delta N_{\text{eff}}<0.3 at the 95%95\% confidence level [53]. Future CMB missions are expected to tighten the constraints on ΔNeff\Delta N_{\text{eff}} by an order of magnitude [54]. We also constrain the parameter spaces of the soliton models using the upper bounds on scalar-induced GWs inferred from the cosmological constraints on ΔNeff\Delta N_{\text{eff}}.

This paper is organized as follows: In Sec. II, we explore the evolution of scalar perturbations in the context of the soliton formation. We calculate the background evolution in the cases where the solitons form during the eMD era and the eRD era, respectively. We then analyze the evolution of the scalar power spectrum in different stages. In Sec. III, we discuss the detailed calculations involved in deriving scalar-induced GWs. Section IV presents our predictions of GW energy spectra and the constraints on the soliton model based on the observations of ΔNeff\Delta N_{\text{eff}}. We conclude in Sec. V.

II power spectrum of the gravitational potential

II.1 background dynamics with a eMD era dominated by solitons

At the linear order, the perturbed metric in the Newtonian gauge reads

ds2=a2(η)((1+2Φ)dη2+(δij2Φδij+hij)dxidxj),ds^{2}=a^{2}(\eta)\left(-\left(1+2\Phi\right)d\eta^{2}+\left(\delta_{ij}-2\Phi\delta_{ij}+h_{ij}\right)dx^{i}dx^{j}\right)\,, (1)

where the a(η)a(\eta) is the scale factor, η\eta is the conformal time. We neglect vector perturbations, first-order GWs and anisotropic stress. According to the Friedmann equations and the continuity equation, in the Universe dominated by a perfect fluid with a general equation of state parameter ω\omega, the scale factor can be expressed as

a(η)=(c0η+c1)23ω+1.a(\eta)=(c_{0}\eta+c_{1})^{\frac{2}{3\omega+1}}\,. (2)

In the early Universe, effective self-resonance resulted in the amplification of perturbations of a real or complex scalar field ϕ\phi. As a consequence, the homogeneous scalar condensate fragments into lumps, leading to the formation of solitons, the long-lived, spatially localized structures. The solitons are distributed in space by Poisson statistics, with a mean physical separation dd. Solitons form rapidly from fragmented scalar field lumps, allowing their formation to be approximated as instantaneous, occurring at time η=ηf\eta=\eta_{\text{f}}. Before the formation of solitons, the Universe could have been dominated by either matter or radiation. To address this, we consider both scenarios separately. For convenience, we define the energy density of solitons as ρs\rho_{\text{s}} and the total energy density of the Universe as ρtot\rho_{\text{tot}}, and we introduce the fractional energy denisity of solitons at η=ηf\eta=\eta_{\text{f}} as Ωs,f=ρs,f/ρtot,f\Omega_{\text{s,f}}=\langle\rho_{\text{s,f}}\rangle/\rho_{\text{tot,f}}. If the solitons are formed during the eMD era, then Ωs,f=1\Omega_{\text{s,f}}=1. Conversely, if the solitons are formed during the eRD era, we have Ωs,f1\Omega_{\text{s,f}}\leq 1. In this paper, we primarily focus on the two cases where Ωs,f=1\Omega_{\text{s,f}}=1 and Ωs,f1\Omega_{\text{s,f}}\ll 1.

After formation, solitons gradually lose energy by emitting relativistic radiation at a slow rate and behave as non-relativistic matter. The conserved charge of solitons, QQ, also decreases with time, and solitons rapidly decay once the charge reaches the critical value QcrQ_{\text{cr}}, a parameter that depends on the specific soliton model. For solitons formed during the eRD era, their energy density is negligible at ηf\eta_{\text{f}}. As the Universe expands, the energy density of radiation scales as ρra4\rho_{\text{r}}\propto a^{-4} while the energy density of matter scales as ρma3\rho_{\text{m}}\propto a^{-3}, respectively. Therefore, for solitons with sufficiently long lifetimes, they could eventually dominate the Universe before decaying. Since the decay process of solitons is also rapid, we can also assume instantaneous decay. Therefore, we define the time when the solitons dominate the Universe as η=ηd\eta=\eta_{d} and the decay time as η=ηr\eta=\eta_{r}. The Universe is dominated by matter at η=ηm\eta=\eta_{\text{m}}, and we find the relationship ηr>ηm=ηd>ηf\eta_{\text{r}}>\eta_{\text{m}}=\eta_{\text{d}}>\eta_{\text{f}}.

Oscillons can also form in the eMD era. The post-inflationary Universe can host a long range of strongly nonlinear processes due to parametric resonance, which is referred to as preheating. The resonance fragments the inflaton condensate and the Universe is strongly inhomogeneous on sub-horizon scales. The energy density is transferred from the oscillating scalar field to the oscillons. Before oscillon formation, the inflaton performs rapid, damped oscillations about a quadratic minimum of the potential after inflation with the frequency much higher than the expansion rate of the Universe. During the oscillation period, the energy density already scales as a3a^{-3}. In this scenario, the Universe to be dominated by matter at η=ηm\eta=\eta_{\text{m}}, then solitons form at η=ηf\eta=\eta_{\text{f}} and dominate the Universe at ηd\eta_{\text{d}}, where ηf=ηd>ηm\eta_{\text{f}}=\eta_{\text{d}}>\eta_{\text{m}}, and finally decay at η=ηr\eta=\eta_{\text{r}}.

Similarly, we can find the relationship of the comoving wavenumbers with kaHk\sim aH corresponding to different time nodes: the wavenumber at the time of soliton formation, kfk_{\text{f}}; the wavenumber when the Universe begins to be dominated by matter, kmk_{\text{m}}; the wavenumber when solitons start to dominate the Universe, kdk_{\text{d}}; and the wavenumber at the time of soliton decay, krk_{\text{r}}. Additionally, we define the characteristic wavenumber of solitons as kUVadk_{\text{UV}}\equiv\frac{a}{d}, beyond which the fluid description of solitons becomes invalid; the average number of solitons within a Hubble event horizon at the time of their formation as n3n^{3}. The lifetime of solitons can be described in terms of NN, where NN is defined as NkfkrN\equiv\frac{k_{\text{f}}}{k_{\text{r}}}. If solitons were formed during eRD era, the hierarchy of the relevant scales reads

kUV>kf>km=kd>kr,k_{\text{UV}}>k_{\text{f}}>k_{\text{m}}=k_{\text{d}}>k_{\text{r}}\,, (3)

and the relationship between them can be expressed through the parameters of the soliton model as follows

kUVkf=n,kmkf=2Ωs,f,kfkr=N,\frac{k_{\text{UV}}}{k_{\text{f}}}=n\,,\quad\frac{k_{\text{m}}}{k_{\text{f}}}=2\Omega_{\text{s,f}}\,,\quad\frac{k_{\text{f}}}{k_{\text{r}}}=N\,, (4)

where the middle equation implies that the matter domination begins when the energy density of radiation and solitons are equal.

Similarly, if solitons form during the eMD era, we have

kUV>km>kf=kd>kr,k_{\text{UV}}>k_{\text{m}}>k_{\text{f}}=k_{\text{d}}>k_{\text{r}}\,, (5)

where we assume that kUV>kmk_{\text{UV}}>k_{\text{m}} since the formation of solitons is a sub-horizon process. The relationship between them at this time is

kUVkf=n,kfkr=N,kmkf=p.\frac{k_{\text{UV}}}{k_{\text{f}}}=n\,,\quad\frac{k_{\text{f}}}{k_{\text{r}}}=N\,,\quad\frac{k_{\text{m}}}{k_{\text{f}}}=p\,. (6)

We also obtain the evolution of the scale factor a(η)a(\eta) and conformal Hubble parameter \mathcal{H} as conformal time η\eta in different cases. For the solitons formed during the eRD era, the results can be written as

aeRD(η)=afηηf,aeMD(η)=af(η+ηm)24ηfηm,asRD(η)=af(ηr+ηm)(2ηηr+ηm)4ηfηm,\begin{split}&a_{\text{eRD}}(\eta)=a_{\text{f}}\frac{\eta}{\eta_{\text{f}}}\,,\\ &a_{\text{eMD}}(\eta)=a_{\text{f}}\frac{\left(\eta+\eta_{\text{m}}\right)^{2}}{4\eta_{\text{f}}\eta_{\text{m}}}\,,\\ &a_{\text{sRD}}(\eta)=a_{\text{f}}\frac{\left(\eta_{\text{r}}+\eta_{\text{m}}\right)\left(2\eta-\eta_{\text{r}}+\eta_{\text{m}}\right)}{4\eta_{\text{f}}\eta_{\text{m}}}\,,\end{split} (7)
eRD(η)=1η,eMD(η)=2η+ηm,sRD(η)=1ηηrηm2.\begin{split}&\mathcal{H}_{\text{eRD}}(\eta)=\frac{1}{\eta}\,,\\ &\mathcal{H}_{\text{eMD}}(\eta)=\frac{2}{\eta+\eta_{\text{m}}}\,,\\ &\mathcal{H}_{\text{sRD}}(\eta)=\frac{1}{\eta-\frac{\eta_{\text{r}}-\eta_{\text{m}}}{2}}\,.\end{split} (8)

For the solitons formed during the eMD era,

aeMD(η)=afη2ηf,asRD(η)=afηr(2ηηr)ηf2,\begin{split}&a_{\text{eMD}}(\eta)=a_{\text{f}}\frac{\eta^{2}}{\eta_{\text{f}}}\,,\\ &a_{\text{sRD}}(\eta)=a_{\text{f}}\frac{\eta_{\text{r}}\left(2\eta-\eta_{\text{r}}\right)}{\eta^{2}_{\text{f}}}\,,\end{split} (9)
eMD=2η,sRD=22ηηr.\begin{split}&\mathcal{H}_{\text{eMD}}=\frac{2}{\eta}\,,\\ &\mathcal{H}_{\text{sRD}}=\frac{2}{2\eta-\eta_{\text{r}}}\,.\end{split} (10)

II.2 power spectrum of the gravitational potential

Since solitons are rapidly diluted by cosmic expansion after their formation, while their physical size remains unchanged, they become well-separated and can be treated as non-interacting, point-like entities. Therefore, considering that the solitons are randomly distributed in space according to Poisson statistics with a mean physical separation of dd, the power spectrum of soliton density perturbations can be expressed as [55]:

δρs(x)δρs(x)=4π3(da)3ρs2δ(x+x),\langle\delta\rho_{\text{s}}(\textbf{x})\delta\rho_{\text{s}}(\textbf{x}^{\prime})\rangle=\frac{4\pi}{3}\left(\frac{d}{a}\right)^{3}\rho_{\text{s}}^{2}\delta(\textbf{x}+\textbf{x}^{\prime})\,, (11)

where δρs\delta\rho_{\text{s}} represents the density perturbations of ρs\rho_{\text{s}}. The smoothed spectrum of density fluctuations is valid up to kUVk_{\text{UV}}. For smaller scales, the complex effects of granularity become important and we set kUVk_{\text{UV}} as the cutoff scale.

The Fourier expansion of soliton density perturbations is

δρs(𝐱)ρs(𝐱)=d3𝐤(3π)3/2δ𝐤(t)ei𝐤𝐱.\frac{\delta\rho_{\text{s}}(\mathbf{x})}{\rho_{\text{s}}(\mathbf{x})}=\int\frac{d^{3}\mathbf{k}}{(3\pi)^{3/2}}\delta_{\mathbf{k}}(t)e^{i\mathbf{k}\mathbf{x}}\,. (12)

By plugging eq. (12) into eq. (11), the power spectrum of δ𝐤\delta_{\mathbf{k}} can be expressed as

𝒫δ(k)=23π(kkUV)3.\mathcal{P}_{\delta}(k)=\frac{2}{3\pi}\left(\frac{k}{k_{\text{UV}}}\right)^{3}\,. (13)

If solitons form during the eRD era, i.e., Ωs,f1\Omega_{\text{s,f}}\ll 1, we define the density contrast of solitons, δs=δρsρs\delta_{s}=\frac{\delta\rho_{\text{s}}}{\rho_{\text{s}}}. In this case, isocurvature perturbations are given by

S=δρsρs34δρrρrδρsρs,S=\frac{\delta\rho_{\text{s}}}{\rho_{\text{s}}}-\frac{3}{4}\frac{\delta\rho_{\text{r}}}{\rho_{\text{r}}}\approx\frac{\delta\rho_{\text{s}}}{\rho_{\text{s}}}\,, (14)

where we have used Ωs1\Omega_{s}\ll 1. Thus, isocurvature fluctuations can be wiritten as

Sfδρs,fρs,f,S_{f}\approx\frac{\delta\rho_{\text{s,f}}}{\rho_{\text{s,f}}}\,, (15)

and the dimensionless isocurvature power spectrum can then be expressed as

𝒫S(k)𝒫δ(k)=23π(kkUV)3.\mathcal{P}_{S}(k)\approx\mathcal{P}_{\delta}(k)=\frac{2}{3\pi}\left(\frac{k}{k_{\text{UV}}}\right)^{3}\,. (16)

During the eRD era, such isocurvature perturbations will convert into adiabatic perturbations. At the beginning of the sRD era, η=ηr\eta=\eta_{\text{r}}, we can write the Bardeen potential, Φk\Phi_{k}, as a summation of the two components,

Φk(ηr)=Φk,inf(ηr)+Φk,s(ηr),\Phi_{k}(\eta_{\text{r}})=\Phi_{k,\text{inf}}(\eta_{\text{r}})+\Phi_{k,\text{s}}(\eta_{\text{r}})\,, (17)

where the first component comes from quantum fluctuations during inflation, while the second component is due to the isocurvature perturbations introduced by solitons. At large scale, curvature perturbations ζ\zeta are related to Φ\Phi as

ΦkΦk=δ(k+k)2π2k3(3+3ω5+3ω)2𝒫ζ(k).\langle\Phi_{\textbf{k}}\Phi_{\textbf{k}^{\prime}}\rangle=\delta(\textbf{k}+\textbf{k}^{\prime})\frac{2\pi^{2}}{k^{3}}\left(\frac{3+3\omega}{5+3\omega}\right)^{2}\mathcal{P}_{\zeta}(k)\,. (18)

As predicted by most well-motivated inflationary models, the power spectrum of ζk,inf\zeta_{k,\text{inf}} is approximately scale-invariant,

𝒫ζ,inf=As(kk)ns1.\mathcal{P}_{\zeta,\text{inf}}=A_{s}\left(\frac{k}{k_{*}}\right)^{n_{\text{s}}-1}\,. (19)

From Planck-2018 data, we apply the scalar spectrum amplitude As=2.1×109A_{s}=2.1\times 10^{-9}, the pivot scale k=0.05k_{*}=0.05 Mpc1\mathrm{Mpc}^{-1}, and the scalar spectral index ns=0.965n_{\text{s}}=0.965 [56].

Then, we investigate the evolution of isocurvature perturbations in the eRD era and eMD era. At the linear order, scalar perturbations Φ\Phi satisfies

Φ′′+3(1+cs2)Φ+(2(1+3cs2)+2)Φcs2ΔΦ=a2ρs2Mpl2cs2S,\Phi^{\prime\prime}+3\mathcal{H}(1+c_{s}^{2})\Phi^{\prime}+\left(\mathcal{H}^{2}\left(1+3c_{s}^{2}\right)+2\mathcal{H}^{\prime}\right)\Phi-c_{s}^{2}\Delta\Phi=\frac{a^{2}\rho_{\text{s}}}{2M_{\text{pl}^{2}}}c_{s}^{2}S\,, (20)

where the prime denotes the deviative with respect to conformal time and cs2c_{s}^{2} is the sound speed of the Universe. Isocurvature perturbations act as a source term in the evolution of scalar perturbations and transform into adiabatic perturbations when the solitons dominate the Universe. In the eMD era, Φk\Phi_{k} does not evolve with time for the subhorizon modes. For k<ksk<k_{\text{s}}, the mode enters the horizon during the eMD era and becomes constant. For k>ksk>k_{\text{s}}, the mode enter the horizon during the eRD era and decreases until the eMD era begins. In addition, isocurvature perturbations will convert into adiabatic perturbations and we can obtain the power spectrum Φk,s\Phi_{k,\mathrm{s}} by solving eq. (20)

𝒫Φ,s(k,ηr)=23π(kkUV)3(5+49k2d2)2.\mathcal{P}_{\Phi,\text{s}}(k,\eta_{\text{r}})=\frac{2}{3\pi}\left(\frac{k}{k_{\text{UV}}}\right)^{3}\left(5+\frac{4}{9}\frac{k^{2}}{\mathcal{H}_{\text{d}}^{2}}\right)^{-2}\,. (21)

If solitons form during the eMD era, i.e., Ωs,f=1\Omega_{\text{s,f}}=1, we can obtain curvature perturbations directly from energy density perturbations of solitons using the relationship

ΔΦ3(Φ+Φ)=4πGa2δρs.\Delta\Phi-3\mathcal{H}\left(\Phi^{\prime}+\mathcal{H}\Phi\right)=4\pi Ga^{2}\delta\rho_{\text{s}}\,. (22)

Considering the fact that Φ\Phi is a constant during the eMD era, we obtain that

23k2d2Φk2Φk=δk.-\frac{2}{3}\frac{k^{2}}{\mathcal{H}^{2}_{\text{d}}}\Phi_{k}-2\Phi_{k}=\delta_{k}\,. (23)

where Φk\Phi_{k} and δk\delta_{k} are the Fourier forms of Φ\Phi and δρsρs\frac{\delta\rho_{\text{s}}}{\rho_{\text{s}}}. In this way, we can get the power spectrum of scalar perturbations induced by solitons as

𝒫Φ,s(k,ηr)=23π(kkUV)3(2+23k2d2)2.\mathcal{P}_{\Phi,\text{s}}(k,\eta_{\text{r}})=\frac{2}{3\pi}\left(\frac{k}{k_{\text{UV}}}\right)^{3}\left(2+\frac{2}{3}\frac{k^{2}}{\mathcal{H}_{\text{d}}^{2}}\right)^{-2}\,. (24)

Note that if the solitons decay is not a perfectly instantaneous process, the finite duration will affect the modes whose period is larger than the rate of transition. This gives rise to an additional kk-dependent suppression factor, which depends on the specific decay behavior of the soliton, in general, can be expressed as [57]:

ΦsRDΦsRDinstantS1/2(kkr)m,\frac{\Phi_{\text{sRD}}}{\Phi^{\text{instant}}_{\text{sRD}}}\approx S^{1/2}\left(\frac{k}{k_{\text{r}}}\right)^{m}\,, (25)

where ΦsRDinstant\Phi^{\text{instant}}_{\text{sRD}} refers to the instantaneous transition value of scalar perturbations [58], S𝒪(1)S\sim\mathcal{O}(1) and mm describes the decay behavior of the soliton which is dependent on the soliton model, and generally satisfies 1<m<0-1<m<0111 In certain soliton models, such suppression does not occur, represented effectively by setting S=1S=1 and m=0m=0..

III Scalar induced GWs

In this section, we focus on the energy spectrum of GWs induced by curvature perturbations we have obtained in the last section.

III.1 GWs at second order

Scalar and tensor perturbations of the metric coupled with each other at second order expansion of the Einstein equation. The equation of motion for tensor perturbations in the Fourier space can be written as

hk′′+2hk+k2hk=4Sk,h_{\textbf{k}}^{\prime\prime}+2\mathcal{H}h_{\textbf{k}}^{\prime}+k^{2}h_{\textbf{k}}=4S_{\textbf{k}}\,, (26)

where SkS_{\textbf{k}} represents the source term arising from first-order scalar perturbations. Using the Green’s function method, the power spectrum of tensor perturbations can be expressed as

𝒫h(η,k)=40𝑑v|1v|1+v𝑑u(4u2(1+v2u2)24vu)2I2¯(v,u,x)𝒫ζ(vk)𝒫ζ(uk),\mathcal{P}_{h}(\eta,k)=4\int_{0}^{\infty}dv\int_{|1-v|}^{1+v}du\left(\frac{4u^{2}-\left(1+v^{2}-u^{2}\right)^{2}}{4vu}\right)^{2}\overline{I^{2}}(v,u,x)\mathcal{P}_{\zeta}(vk)\mathcal{P}_{\zeta}(uk)\,, (27)

where the overline denotes the oscillation average and we have defined dimensionless parameter x=kηx=k\eta. I(v,u,x)I(v,u,x) is the kernel of induced GWs, which takes into account the time evolution of scalar perturbations,

I(v,u,x,xi)xixa(x~)a(x)kGk(x,x~)f(v,u,x~,xi)𝑑x~.I(v,u,x,x_{i})\equiv\int_{x_{i}}^{x}\frac{a(\tilde{x})}{a(x)}kG_{k}(x,\tilde{x})f(v,u,\tilde{x},x_{i})d\tilde{x}\,. (28)

The Green’s function Gk(x,x~)G_{\textbf{k}}(x,\tilde{x}) can be obtained by solving the equation

d2Gkdx2(x,x~)+(11a(x)d2a(x)dx2)Gk(x,x~)=δ(xx~).\frac{d^{2}G_{\textbf{k}}}{dx^{2}}(x,\tilde{x})+\left(1-\frac{1}{a(x)}\frac{d^{2}a(x)}{dx^{2}}\right)G_{\textbf{k}}(x,\tilde{x})=\delta(x-\tilde{x})\,. (29)

In the following, we define Φk(η)T(kη)ϕk\Phi_{\textbf{k}}(\eta)\equiv T(k\eta)\phi_{\textbf{k}}, where T(kη)T(k\eta) is the transfer function and ϕk\phi_{\textbf{k}} is the primordial value. Since the velocities of solitons are negligible, we can calculate f(u,v,x~,xi)f(u,v,\tilde{x},x_{i}) in terms of the transfer function T(kη)T(k\eta)

f(u,v,x~,xi)=325(1+ω)(2(5+3ω)T(vx~)T(ux~)+41(T(ux~)T(vx~)+T(vx~)T(ux~))+42T(vx~)T(ux~)).f(u,v,\tilde{x},x_{i})=\frac{3}{25(1+\omega)}\left(2(5+3\omega)T(v\tilde{x})T(u\tilde{x})+4\mathcal{H}^{-1}\left(T^{\prime}(u\tilde{x})T(v\tilde{x})+T^{\prime}(v\tilde{x})T(u\tilde{x})\right)+4\mathcal{H}^{-2}T^{\prime}(v\tilde{x})T^{\prime}(u\tilde{x})\right)\,. (30)

The GW energy spectrum is defined as the fraction of the GW energy density per logarithmic wavelength

ΩGW(η,k)=124(ka(η)H(η))2𝒫h(η,k)¯.\Omega_{\text{GW}}(\eta,k)=\frac{1}{24}\left(\frac{k}{a(\eta)H(\eta)}\right)^{2}\overline{\mathcal{P}_{h}(\eta,k)}\,. (31)

The gravitational wave (GW) source term peaks around ηr\eta_{\mathrm{r}} and rapidly diminishes thereafter. Once generated, GWs hardly interact between GW and the background plasma is negligible, and the GW energy density scales with the expansion of the Universe as a4a^{-4}. Utilizing entropy conservation, the GW energy spectrum today, ΩGW,0\Omega_{\text{GW},0}, can be expressed in terms of ΩGW\Omega_{\text{GW}} in the sRD era as [59, 60]

ΩGW,0h2(k)=0.39(gc106.75)1/3Ωr,0h2ΩGW(ηc,k),\Omega_{\text{GW},0}h^{2}(k)=0.39\left(\frac{g_{c}}{106.75}\right)^{-1/3}\Omega_{\text{r},0}h^{2}\Omega_{\text{GW}}(\eta_{c},k)\,, (32)

where Ωr,0h24.18×105\Omega_{\text{r},0}h^{2}\sim 4.18\times 10^{-5} is the energy density fraction of radiation evaluated today [53]. In this paper, we use gg to denote the effective relativistic degrees of freedom , and gcg(η=ηc)g_{c}\equiv g(\eta=\eta_{c}).

III.2 Scalar induced GWs in scenarios of soliton formation

In the eMD Universe, we can divide the contribution of I(v,u,x)I(v,u,x) into three components, IeRDI_{\text{eRD}}, IeMDI_{\text{eMD}} and IsRDI_{\text{sRD}}, which correspond to the GW production in different eras: the eRD era at first, followed by the eMD era and finally the sRD era. However, we only focus on GWs generated during the sRD era, because for both 𝒫Φinf\mathcal{P}_{\Phi_{\text{inf}}} and 𝒫Φs\mathcal{P}_{\Phi_{\text{s}}}, the dominant contribution to GWs comes from IsRDI_{\text{sRD}}, and the GW energy spectrum can be expressed as

ΩGW(η,k)=160𝑑v|1v|1+v𝑑u(4v2(1+v2u2)4uv)2sRD2¯(v,u,x,xr)𝒫ζ(kv)𝒫ζ(ku).\Omega_{\text{GW}}(\eta,k)=\frac{1}{6}\int_{0}^{\infty}dv\int_{|1-v|}^{1+v}du\left(\frac{4v^{2}-\left(1+v^{2}-u^{2}\right)}{4uv}\right)^{2}\overline{\mathcal{I}_{\text{sRD}}^{2}}(v,u,x,x_{\text{r}})\mathcal{P}_{\zeta}(kv)\mathcal{P}_{\zeta}(ku)\,. (33)

As mentioned above, the scalar power spectrum 𝒫Φ\mathcal{P}_{\Phi} can be written as

𝒫Φ(k,ηr)=𝒫Φ,inf(k,ηr)+𝒫Φ,s(k,ηr),\mathcal{P}_{\Phi}(k,\eta_{\text{r}})=\mathcal{P}_{\Phi,\text{inf}}(k,\eta_{\text{r}})+\mathcal{P}_{\Phi,\text{s}}(k,\eta_{\text{r}})\,, (34)

If solitons form during the eRD era, the inflationary scalar perturbation modes, Φk,inf\Phi_{k,\mathrm{inf}}, that reenter the horizon during the eRD era, are significantly suppressed so that we can set the cutoff scale of Φinf\Phi_{\text{inf}} at k=kmk=k_{\text{m}}. If solitons form during the eMD era, primarily referring to oscillons in this paper, Φk,inf\Phi_{k,\mathrm{inf}} with k>kmk>k_{\text{m}} correspond to perturbations that remained inside the horizon during inflation, which provide the initial conditions for the self-resonance of inflaton perturbations. Thus, the cutoff condition k<kmk<k_{\text{m}} is also valid in this case. Taking into account that eq. (21) and eq. (24) are valid only for k<kUVk<k_{\text{UV}}, we can obtain

𝒫Φ,inf(k,ηr)=(3+3ω5+3ω)2As(kk)ns1Θ(kmk),\mathcal{P}_{\Phi,\text{inf}}(k,\eta_{\text{r}})=\left(\frac{3+3\omega}{5+3\omega}\right)^{2}A_{s}\left(\frac{k}{k_{*}}\right)^{n_{\text{s}}-1}\Theta(k_{\text{m}}-k)\,, (35)
𝒫Φ,s(k,ηr)={23π(kkUV)3(5+49k2d2)2,if solitons form during the eRD era;23π(kkUV)3(2+23k2d2)2,if solitons form during the eMD era,\mathcal{P}_{\Phi,\text{s}}(k,\eta_{\text{r}})=\begin{cases}\frac{2}{3\pi}\left(\frac{k}{k_{\text{UV}}}\right)^{3}\left(5+\frac{4}{9}\frac{k^{2}}{\mathcal{H}_{\text{d}}^{2}}\right)^{-2},&\text{if solitons form during the eRD era};\\ \frac{2}{3\pi}\left(\frac{k}{k_{\text{UV}}}\right)^{3}\left(2+\frac{2}{3}\frac{k^{2}}{\mathcal{H}_{\text{d}}^{2}}\right)^{-2},&\text{if solitons form during the eMD era},\end{cases} (36)

Since ηrηm\eta_{\text{r}}\gg\eta_{\text{m}}, whether solitons were produced in the eMD era or the eRD era, the conmoving Hubble parameter \mathcal{H} in the sRD era can be approximately expressed as sRD22ηηr\mathcal{H}_{\text{sRD}}\approx\frac{2}{2\eta-\eta_{\text{r}}}. We introduce a dimensionless variable xr=kηrx_{\text{r}}=k\eta_{\text{r}}, which corresponds to the beginning of the sRD era. Based on the background evolution given by eq. (7) and eq. (9), the GW kernel can be obtained by solving the Green’s equation eq. (29) during the sRD era

IsRD(v,u,x,xr)=xrx𝑑x~x~xr/2xxr/2f(v,u,x~,xr)kGk(x~,x).I_{\text{sRD}}(v,u,x,x_{\text{r}})=\int_{x_{\text{r}}}^{x}d\tilde{x}\frac{\tilde{x}-x_{\text{r}}/2}{x-x_{\text{r}}/2}f(v,u,\tilde{x},x_{\text{r}})kG_{k}(\tilde{x},x)\,. (37)

For simplicity, we define

sRD(v,u,x,xr)=IsRD(v,u,x,xr)(xxr2).\mathcal{I}_{\text{sRD}}(v,u,x,x_{\text{r}})=I_{\text{sRD}}(v,u,x,x_{\text{r}})\left(x-\frac{x_{\text{r}}}{2}\right)\,. (38)

The general solution for the transfer function in the sRD era, preceded by an eMD era can be written as

T(x,xr)=33xxr/2(A(xr)j1(xxr/23)+B(xr)y1(xxr/23)),T(x,x_{\text{r}})=\frac{3\sqrt{3}}{x-x_{\text{r}}/2}\left(A(x_{\text{r}})j_{1}\left(\frac{x-x_{\text{r}}/2}{\sqrt{3}}\right)+B(x_{\text{r}})y_{1}\left(\frac{x-x_{\text{r}}/2}{\sqrt{3}}\right)\right)\,, (39)

where j1j_{1} and y1y_{1} are the first and second spherical Bessel functions, and A,BA,B are constants which depend on xrx_{r}. Demanding the continuity of the transfer function and its time dericative at η=ηr\eta=\eta_{\text{r}}, we can determine AA and BB as

A(xr)=xr23sin(xr23)136(xr236)cos(xr23),A(x_{\text{r}})=\frac{x_{\text{r}}}{2\sqrt{3}}\sin\left(\frac{x_{\text{r}}}{2\sqrt{3}}\right)-\frac{1}{36}\left(x_{\text{r}}^{2}-36\right)\cos\left(\frac{x_{\text{r}}}{2\sqrt{3}}\right)\,, (40)
B(xr)=136(xr236)sin(xr23)xr23cos(xr23).B(x_{\text{r}})=-\frac{1}{36}(x_{\text{r}}^{2}-36)\sin\left(\frac{x_{\text{r}}}{2\sqrt{3}}\right)-\frac{x_{\text{r}}}{2\sqrt{3}}\cos\left(\frac{x_{\text{r}}}{2\sqrt{3}}\right)\,. (41)

In principle, using eq. (39) and eq. (30), we can directly obtain the GW energy density ΩGW\Omega_{\text{GW}}. However, our aim is to establish a relationship between GWs and the parameters of the soliton model, which will further allow us to impose constraints on the soliton model parameters. Therefore, we seek to derive approximate analytical formulas for induced GWs. We then obtain the approximate form of sRD\mathcal{I}_{\text{sRD}} on the scales of kkrk\gg k_{\text{r}} and kkrk\ll k_{\text{r}}, and use this approximate form to calculate induced GWs.

In the small-scale approximation, kkrk\gg k_{r} and xxrx\gg x_{\text{r}}, the primary contributions to the kernel term sRD\mathcal{I}_{sRD} come from the sum of trigonometric terms. These include the terms such as A(u,v,xr)xr4sin(D±±xr)A(u,v,x_{\text{r}})x_{\text{r}}^{4}\sin\left(D_{\pm\pm}x_{\text{r}}\right), B(u,v,xr)xr4cos(D±±xr)B(u,v,x_{\text{r}})x_{\text{r}}^{4}\cos\left(D_{\pm\pm}x_{\text{r}}\right), cosine integral terms like C(u,v,xr)xr4Ci(D±±xr)C(u,v,x_{\text{r}})x_{\text{r}}^{4}\mathrm{Ci}\left(D_{\pm\pm}x_{\text{r}}\right), and sine integral terms D(u,v,xr)xr4Si(D±±)D(u,v,x_{\text{r}})x_{\text{r}}^{4}\mathrm{Si}\left(D_{\pm\pm}\right), where A(u,v,xr)A(u,v,x_{\text{r}}), B(u,v,xr)B(u,v,x_{\text{r}}), C(u,v,xr)C(u,v,x_{\text{r}}) and D(u,v,xr)D(u,v,x_{\text{r}}) represent the coefficients that depend on uu, vv and xrx_{\text{r}}. Additionally, we define D±±=u±v±333D_{\pm\pm}=\frac{u\pm v\pm\sqrt{3}}{3\sqrt{3}} [6]. Furthermore, we introduce the cosine integral and sine integral functions, defined as:

Ci(x)=x𝑑x~cosx~x~,\mathrm{Ci}(x)=-\int_{x}^{\infty}d\tilde{x}\frac{\cos\tilde{x}}{\tilde{x}}\,, (42)
Si(x)=0x𝑑x~sinx~x~.\mathrm{Si}(x)=\int_{0}^{x}d\tilde{x}\frac{\sin\tilde{x}}{\tilde{x}}\,. (43)

In this case, we are interested only in the resonant part of induced GWs. Following the calculation of the kernel term of sRD\mathcal{I}_{\text{sRD}}, the terms containing Ci\mathrm{Ci} contribute the main part [37]

eRD2¯|resxr85971968(u2+v23)4u2v2Ci((3u+3v3)xr6)2.\overline{\mathcal{I}_{\text{eRD}}^{2}}|_{\text{res}}\approx\frac{x_{\text{r}}^{8}}{5971968}\frac{\left(u^{2}+v^{2}-3\right)^{4}}{u^{2}v^{2}}\mathrm{Ci}\left(\left(\sqrt{3}u+\sqrt{3}v-3\right)\frac{x_{\text{r}}}{6}\right)^{2}\,. (44)

For the large-scale approximation, considering the cutoff conditions of 𝒫Φ,inf\mathcal{P}_{\Phi,\text{inf}} and 𝒫Φ,s\mathcal{P}_{\Phi,\text{s}}, the integrals over uu and vv are primarily dominated by the region where u+v1u+v-1 is large and uvu-v is small. Under these approximations, we find that

sRD2¯|lar(u+v1)4xr89102222(4Ci(xr2)2+(π2Si(xr2))2).\overline{\mathcal{I}^{2}_{\text{sRD}}}|_{\text{lar}}\approx\frac{(u+v-1)^{4}x_{\text{r}^{8}}}{9102222}\left(4\mathrm{Ci}\left(\frac{x_{\text{r}}}{2}\right)^{2}+\left(\pi-2\mathrm{Si}\left(\frac{x_{\text{r}}}{2}\right)\right)^{2}\right)\,. (45)

Therefore, the energy spectrum of induced GWs can be expressed as

ΩGW=ΩGW,inf,res+ΩGW,inf,lar+ΩGW,s,res+ΩGW,inf,lar.\Omega_{\text{GW}}=\Omega_{\text{GW,inf,res}}+\Omega_{\text{GW,inf,lar}}+\Omega_{\text{GW,s,res}}+\Omega_{\text{GW,inf,lar}}\,. (46)

These four terms represent the contribution of 𝒫Φ,inf\mathcal{P}_{\Phi,\text{inf}} and 𝒫Φ,s\mathcal{P}_{\Phi,\text{s}} on small and large scales, respectively. Using eq. (44) and eq. (45), we can derive the following expression

ΩGW,inf,lar=1.14×109(4Ci(xr2)2+(π2Si(xr2))2)S2As2xr8(kkr)4m(kk)2ns2𝑑v𝑑u(4v2(1+v2u2)2)2(u+v1)4(uv)2m+ns3,\begin{split}\Omega_{\text{GW,inf,lar}}=&1.14\times 10^{-9}\left(4\mathrm{Ci}\left(\frac{x_{\text{r}}}{2}\right)^{2}+\left(\pi-2\mathrm{Si}\left(\frac{x_{\text{r}}}{2}\right)\right)^{2}\right)S^{2}A_{s}^{2}x_{\text{r}}^{8}\left(\frac{k}{k_{\text{r}}}\right)^{4m}\left(\frac{k}{k_{*}}\right)^{2n_{s}-2}\\ &\int dv\int du\left(4v^{2}-(1+v^{2}-u^{2})^{2}\right)^{2}(u+v-1)^{4}(uv)^{2m+n_{s}-3}\,,\end{split} (47)
ΩGW,s,lar=2.6×1010(4Ci(xr2)2+(π2Si(xr2))2)S2xr8(kkr)4m(kkUV)6𝑑v𝑑u(4v2(1+v2u2)2)2(u+v1)4(uv)2m+11(a+bk2v2d2)2(a+bk2v2d2)2,\begin{split}\Omega_{\text{GW,s,lar}}=&2.6\times 10^{-10}\left(4\mathrm{Ci}\left(\frac{x_{\text{r}}}{2}\right)^{2}+\left(\pi-2\mathrm{Si}\left(\frac{x_{\text{r}}}{2}\right)\right)^{2}\right)S^{2}x_{\text{r}}^{8}\left(\frac{k}{k_{\text{r}}}\right)^{4m}\left(\frac{k}{k_{\text{UV}}}\right)^{6}\\ &\int dv\int du\left(4v^{2}-(1+v^{2}-u^{2})^{2}\right)^{2}(u+v-1)^{4}(uv)^{2m+1}\frac{1}{\left(a+b\frac{k^{2}v^{2}}{\mathcal{H}^{2}_{\text{d}}}\right)^{2}\left(a+b\frac{k^{2}v^{2}}{\mathcal{H}^{2}_{\text{d}}}\right)^{2}}\,,\end{split} (48)
ΩGW,inf,res=1.4×106(14)2m+nsS2As2xr7(kkr)4m(kk)2ns2smsm(1s2)2(3s2)2m+ns1,\Omega_{\text{GW,inf,res}}=1.4\times 10^{-6}\left(\frac{1}{4}\right)^{2m+n_{s}}S^{2}A_{s}^{2}x_{\text{r}}^{7}\left(\frac{k}{k_{\text{r}}}\right)^{4m}\left(\frac{k}{k_{*}}\right)^{2n_{s}-2}\int_{s_{\text{m}}}^{s_{\text{m}}}\left(1-s^{2}\right)^{2}\left(3-s^{2}\right)^{2m+n_{s}-1}\,, (49)
ΩGW,s,res=3.2×107S2b2(14)2m2xr7(kkr)4m(kkUV)6(dk)8sUVsUV𝑑s(1s2)2(3s2)2m1.\Omega_{\text{GW,s,res}}=3.2\times 10^{-7}\frac{S^{2}}{b^{2}}\left(\frac{1}{4}\right)^{2m-2}x_{\text{r}}^{7}\left(\frac{k}{k_{\text{r}}}\right)^{4m}\left(\frac{k}{k_{\text{UV}}}\right)^{6}\left(\frac{\mathcal{H}_{d}}{k}\right)^{8}\int_{-s_{\text{UV}}}^{s_{\text{UV}}}ds\left(1-s^{2}\right)^{2}\left(3-s^{2}\right)^{2m-1}\,. (50)

Here, the lower and upper limits of the integral are described by sms_{\text{m}} and sUVs_{\text{UV}}, which come from the cutoff condition of Θ(kmk)\Theta(k_{\text{m}}-k) and Θ(kUVk)\Theta(k_{\text{UV}}-k). The upper limit sUVs_{\text{UV}} is a function of kk

smax={1,kkmax21+3;2kmaxk3,21+3<kkmax23;0,kkmax>23.s_{\text{max}}=\begin{cases}1,&\frac{k}{k_{\text{max}}}\leq\frac{2}{1+\sqrt{3}}\,;\\ 2\frac{k_{\text{max}}}{k}-\sqrt{3},&\frac{2}{1+\sqrt{3}}<\frac{k}{k_{\text{max}}}\leq\frac{2}{\sqrt{3}}\,;\\ 0,&\frac{k}{k_{\text{max}}}>\frac{2}{\sqrt{3}}\,.\end{cases} (51)

The result of eq. (46) also contains contributions from the cross terms between 𝒫Φ,inf\mathcal{P}_{\Phi,\text{inf}} and 𝒫Φ,s\mathcal{P}_{\Phi,\text{s}}. In most cases either 𝒫Φ,inf\mathcal{P}_{\Phi,\text{inf}} or 𝒫Φ,s\mathcal{P}_{\Phi,\text{s}} dominates the total 𝒫Φ\mathcal{P}_{\Phi}, ΩGW\Omega_{\mathrm{GW}} is mainly contributed from the dominant term and the cross term is negligible.

IV results and the constraints to soliton model

Refer to caption
Refer to caption
Figure 1: This figure illustrates the temporal evolution of soliton density perturbations during the eMD era. The region above the black dashed line represents the scale at which these perturbations reach nonlinearity, providing a conservative estimate for our calculations. The left panel shows the scenario where solitons form during the eRD era, while the right panel illustrates their formation during the eMD era.
Refer to caption
Refer to caption
Figure 2: This figure shows the power spectra of scalar-induced GWs under various parameter choices for the soliton model and we set kf=0.001Hzk_{\text{f}}=0.001\mathrm{Hz}. The left panel represents the scenario where solitons form during the eRD era, while the right panel illustrates soliton formation during the eMD era. The dashed line indicates the scale at which soliton-induced density perturbations reach nonlinearity, making the present results conservative estimates for this portion of the induced GWs. For simplicity, we set S=1S=1 and m=12m=-\frac{1}{2}.
Refer to caption
Refer to caption
Figure 3: This figure illustrates differences in the GW energy spectra induced by solitons forming during the eMD era and the eRD era, where we keep other parameters the same and set kf=0.001Hzk_{\text{f}}=0.001\mathrm{Hz}. The pink solid line represents soliton formation during the eRD era, while the purple solid line corresponds to soliton formation during the eMD era. In the left panel, both scenarios assume an identical soliton formation number density and an equal duration of the soliton-dominated eMD era. In the right panel, both scenarios assume the same soliton formation number density and the same soliton lifetime. For simplicity, we set S=1S=1 and m=12m=-\frac{1}{2}.
Refer to caption
Refer to caption
Figure 4: This figure illustrates the constraints imposed by induced GWs on the soliton model. The orange curve denotes the constraint applied by PBB, while the yellow curve represents the constraint set by S4\mathrm{S4} on the soliton model. The black curve marks the threshold at which primordial density perturbations become nonlinear, beyond which our constraints no longer hold. In the left panel, we present the scenario where solitons form during the eRD: the solid line corresponds to the case of Ωs,f=0.1\Omega_{\text{s,f}}=0.1, and the dashed line to Ωs,f=0.05\Omega_{\text{s,f}}=0.05. The right panel shows the scenario where solitons form during the eMD era. For simplicity, we set S=1S=1 and m=12m=-\frac{1}{2}.

Before presenting our results, we first outline the relevant scales to which our discussion applies. During the eMD era, the near-zero pressure of the Universe leads to a growth in small-scale density perturbations over time. If this period of matter dominance extends too long, density perturbations will inevitably become nonlinear, and our conclusions will no longer be applicable. This limitation has been discussed in prior research, where the nonlinear scale of primordial density perturbations is denoted by knl𝒫ζ1/4k_{\text{nl}}\sim\mathcal{P}_{\zeta}^{-1/4}\mathcal{H}. However, for density perturbations induced by solitons, small-scale perturbations are significantly larger than primordial density perturbations, necessitating a careful assessment of the scales over which our conclusions remain valid. As shown in Fig. 1, soliton-induced density perturbations at small scales can reach nonlinearity relatively quickly during the eMD era compared to primordial density perturbations. At nonlinear scales, solitons interact to form small-scale structures, and on these scales, the gravitational wave power spectrum follows ΩGWk3/2\Omega_{\text{GW}}\sim k^{3/2}[61]. Thus, our results represent a conservative estimate of ΩGW\Omega_{\mathrm{GW}}.

A main finding of this paper is that even with a brief eMD era, a soliton-dominated early Universe can still produce sufficiently strong induced GWs. In Fig. 2, we present scalar-induced GWs generated with different soliton model parameters. Distinguished from previous studies, we do not require a long duration of the eMD era to amplify primordial scalar perturbations as efficient GW sources. The duration of the eMD era satisfies the condition kmηr𝒪(100)\frac{k_{\text{m}}}{\eta_{\text{r}}}\sim\mathcal{O}(100). Since the number of solitons within a Hubble horizon at the formation time spans in a large range, depending solely on the soliton model, Poisson-induced density perturbations can increase to the 𝒪(1)\mathcal{O}(1) level and produce strong GWs within the short duration of the eMD era.

We also consider the case that solitons may form during the eMD era. This scenario introduces several key differences: First, if the soliton forms during the eRD era, its density perturbation is expected to initially behave as isocurvature perturbations. As the soliton’s energy density increases over time, the isocurvature perturbations gradually transition into the adiabatic perturbations. However, if the soliton forms during the eMD era, the physical situation differs. In this case, an oscillating scalar field dominates the Universe, and solitons emerge through resonant fragmentation of the field. As a result, density perturbations of solitons behave as adiabatic perturbations from the moment of its formation. Since solitons capture most of the energy of scalar field, they immediately dominate the Universe upon formation. Second, if solitons form during the eMD era, they will directly govern the Universe’s dynamics. In this case, the lifetime of the solitons determines the duration of the eMD era. Thus, even with a very short soliton lifetime, N1N\sim 1, strong GWs can still be induced, as shown in Fig. 2. Finally, as shown in eq. (7) and eq. (9), the expansion rates of the Universe during the matter-dominated and radiation-dominated eras are different. This leads to the conclusion that, in the two scenarios where solitons form during the eMD era and the eRD era, respectively, the induced GWs will be stronger in the former scenario, even when both cases share the same soliton model parameters as shown in Fig. 3.

The peak of the induced gravitational wave power spectrum is located at kkUVk\sim k_{\text{UV}}, which in turn depends on the energy scale at the time of soliton formation and the number density at that time. Observational constraints on the early Universe require that the presence of the eMD era does not adversely affect primordial nucleosynthesis. Therefore, in the scenario we consider, solitons are assumed to form at an energy scale T>102GeVT>10^{2}\mathrm{GeV}. Under these conditions, the induced GWs could potentially be detected by future GW observatories, as illustrated in Fig. 2 where we assume kf=0.001Hzk_{\text{f}}=0.001\mathrm{Hz}.

In this section, we also discuss the constraints on soliton models based on the results we obtained. We discuss a general soliton model and parameterize the model in terms of NN, Ωs,f\Omega_{\text{s,f}} and nn, which we introduced in section II. From the cosmological point of view, GWs constitute of dark radiation and can be parameterised by an correction of the effective degree of freedom, ΔNeff=NeffNeffSM\Delta N_{\text{eff}}=N_{\text{eff}}-N^{\text{SM}}_{\text{eff}}. A higher NeffN_{\text{eff}} can delay radiation-to-matter equality and change the size of the sound horizon, which can leave features on CMB anisotropies, baryon acoustic oscillations and Big-Bang nucleosynthesis. In this way, we can get the GW upper bounds inferred from cosmological constraints on NeffN_{\text{eff}}. The limit given by the current and future observations on ΔNeff\Delta N_{\text{eff}} is [62]

ΔNeff={0.175,Planck+BAO+BBN(PBB),0.027,CMB Stage IV(S4).\Delta N_{\text{eff}}=\begin{cases}0.175,&Planck+\text{BAO}+\text{BBN}\,\text{(PBB)}\,,\\ 0.027,&\text{CMB Stage IV}\,\text{(S4)}\,.\end{cases} (52)

This gives the upper bound on GWs density at 95%95\% C.L.

ΩGW,bou={2.11×106,PBB,3.25×107,S4.\Omega_{\text{GW,bou}}=\begin{cases}2.11\times 10^{-6},&\text{PBB}\,,\\ 3.25\times 10^{-7},&\text{S4}\,.\end{cases} (53)

Since we are primarily interested in the peak value of the GWs, we focus only on the resonant component of the GWs. Utilizing eq. (53) we can derive the constraints on the parameter space of the soliton model as follows:

{n4m1N7+4mΩs,f8<1.515×103S2ΩGW,bou,If solitons form during the eRD era.,n4m1N7+4m<7.634×103S2ΩGW,bou,If solitons form during the eRD era..\begin{cases}n^{4m-1}N^{7+4m}\Omega^{8}_{\text{s,f}}<1.515\times 10^{3}\,S^{-2}\Omega_{\text{GW,bou}},&\text{If solitons form during the eRD era.}\,,\\ \,\,\,\,\,\,\,\,\,n^{4m-1}N^{7+4m}<7.634\times 10^{3}\,S^{-2}\Omega_{\text{GW,bou}},&\text{If solitons form during the eRD era.}\,.\end{cases} (54)

Fig. 4 illustrates the constraints imposed on the parameter space of the soliton model from ΔNeff\Delta N_{\text{eff}}. The models with large separation and long lifetime of solitions can be ruled out for the overproduction of GWs. The constraints on the soliton models become weaker for solitons formed in the eRD era because of the smaller peak value of Ωs,f\Omega_{\text{s,f}}.

Note that if the soliton lifetime is too long, the density perturbations during the eMD era will grow over time and eventually become nonlinear. As a result, we restrict our discussion to the case where kηr<450\frac{k}{\eta_{\text{r}}}<450.

V Conclusions

In this paper, we explore the scenario of an eMD era in the Universe driven by solitons and the generation of scalar-induced GWs. Distinguished from previous studies, we both consider soliton formation during both the eRD and eMD eras. In the scenario where solitons form during the eRD era, the Universe is initially radiation-dominated. Solitons then gradually form and start to dominate the Universe, marking the beginning of eMD. Solitons may also form during the eMD era, in which case the Universe is initially dominated by an oscillating scalar field. Since the solitons contain the majority of the energy of scalar field, they can be considered to dominate the Universe as soon as they form. Since the solitons formation is a random process, their spatial distribution generally follows a Poisson distribution, resulting in an induced power spectrum of density perturbations with an ultraviolet cutoff at a scale related to the average soliton separation, kUVk_{\text{UV}}. For solitons form during the eRD era, the Poisson distribution induces isocurvature perturbations that transit into curvature perturbations once solitons dominate the Universe. For solitons form in the eMD era, density perturbations of solitons in this case are adiabatic from the formation time. Density perturbations that originate both from inflation and the Poisson distribution increase during the soliton-dominated era and the corresponding GW sources reach the maximum at around the decay time of solitons, predicting characteristic GW energy spectra that are expected to be observed by multiband GW observers. We also impose new restrictions on the soliton models that overproduce GWs from the upper bound of NeffN_{\mathrm{eff}}.

In this work, our calculations assume small density perturbations remain in the linear regime, δρρ1\frac{\delta\rho}{\rho}\lesssim 1. However, energy density gradually increases over time during the eMD era and may readily reach nonlinear threshold for long-lived solitons. The solitons might also experience colliding with each other in the non-linear regime. In this paper, we present a conservative analysis without considering these nonlinear interactions. These effects would require a more detailed analysis, which we leave for future work.

Acknowledgements.
This work is supported in part by the National Key Research and Development Program of China Grants No. 2020YFC2201501 and No. 2021YFC2203002, in part by the National Natural Science Foundation of China Grants No. 12105060, No. 12147103, No. 12235019, No. 12075297 and No. 12147103, in part by the Science Research Grants from the China Manned Space Project with NO.

References