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

aainstitutetext: School of Physics, Huazhong University of Science and Technology
Wuhan, 430074, China
bbinstitutetext: School of Physics and Astronomy, University of Minnesota
Minneapolis, MN 55455, USA

Gravitational waves from bubble collisions in FLRW spacetime

Haowen Zhong,111Corresponding author a    Biping Gong a,1    Taotao Qiu [email protected] [email protected] [email protected]
Abstract

Stochastic gravitational wave background (SGWB) is a promising tool to probe the very early universe where the standard model of particle physics and cosmology are connected closely. As a possible component of SGWB, gravitational waves (GW) from bubble collisions during the first order cosmological phase transitions deserve comprehensive analyses. In 2017, Ryusuke Jinno and Masahiro Takimoto proposed an elegant analysis approach to derive the analytical expressions of energy spectra of GW from bubble collisions in Minkowski spacetime avoiding large-scale numerical simulations for the first timeJinno_2017 . However, they neglect the expansion of the universe and regard the duration of phase transitions as infinity in their derivation which could deviate their estimations from true values. For these two reasons, we give a new expression of GW spectra by adopting their method, switching spacetime background to FLRW spacetime, and considering a finite duration of phase transitions. By denoting σ\sigma as the fraction of the speed of phase transitions to the expansion speed of the universe, we find when σ\sigma is around 𝒪(10)\mathcal{O}(10), the maxima of estimated GW energy spectra drop by around 1 order of magnitude than the results given by their previous work. Even when σ=100\sigma=100, the maximum of GW energy spectrum is only 65%65\% of their previous estimation. Such a significant decrease may bring about new challenges for the detectability of GW from bubble collisions. Luckily, by comparing new spectra with PLI (power-law integrated) sensitivity curves of GW detectors, we find that the detection prospect for GW from bubble collisions is still promising for DECIGO, BBO, LISA, and TianQin in the foreseeable future.

Keywords:
Stochastic gravitational waves background, bubble collision, finite duration, FLRW spacetime

1 Introduction

Since the first gravitational wave signal has been detected by LIGO on September 14,th2015{}^{\text{th}},2015GW150914 , the study of GW is becoming more and more popular and important. As a brand new detection tool, GW plays a vital role to study astrophysics and cosmology in this golden era of multi-messenger astronomy. Although the GW events we have detected are all generated from the inspiral-merger-ringdown processes of binary systems, there exists abundant different GW sources especially in the very early universe which inspire our curiosity to a great extent, see Ref.Caprini_2018 and Ref.Maggiore_2018 for a comprehensive understanding of the whole picture. The superposition of tremendous amount of GW signals from the very early universe constitutes a stochastic gravitational wave background, where the GW from the first order cosmological phase transitions is a possible component and has been attracting us for a long time because of its relationship to the physics beyond standard model and also the formation of Primordial black holes(PBH), e.g. Espinosa_2008 ; Ashoorioon_2009 ; Amjad_2015 ; Amjad_2021 ; Das_2010 ; Huang_2016 ; Jinno_2017_2 ; Maxim_1998 ; Maxim_1999 ; Maxim_2008 .

Noticing that GW is a promising tool to probe the very early universe, which in turn is an ideal place to study particle physics, studying GW is of great significance from many different perspectives. As we know, the highest energy scale of LHC (Large Hardron Collider) LHC is around 10 TeV which is much less than the energy scale in the very early universe. So there has been no direct means to study particle physics processes happened in that period of time in earth-based laboratories in the foreseeable future. Besides, according to the thermal history of the universe, the oldest photons we can receive are CMB (Cosmic Micro Background) photons which were free from the Thomson scattering and propagated in the universe without restriction when the redshift was around 1100. However, the very early universe we want to study corresponds to the time when the redshift was much larger than 1100. From this point of view, we cannot use electromagnetic waves to probe the universe at that era, because the universe hasn’t been transparent for photons at that time. Nevertheless, GW provides a special probe to it from a definitely different dimension. After the generation of GW, it has nearly no interaction with the contents of the universeMaggiore_2008 , so GW contains a huge amount of precious information from the very early universe compelling us to explore the cosmology and particle physics by studying the properties of GW generated at that epoch.

The first order cosmological phase transitions proceed by the nucleation, expansion and collision of bubbles, at the same time, amount of GW signals are generated. Although nowadays people pay more attention to GW from sound wavesMark_2014 ; Mark_2015 ; Mark_2017 and regard it as the main source of GW from the first order phase transitions, there still are cases like runaway transitions, where the energy density stored in the scalar field can play a really important role i.e. the study of GW from bubble collisions cannot be neglected. The study on GW from bubble collisions needs to be traced back to 1990s, Kosowsky et al. did numerical simulations of bubble collisions in Minkowski spacetime with thin wall approximation and envelope approximation and gave energy spectra of GW Kosowsky_1992_1 ; Kosowsky_1992_2 ; Kosowsky_1993 ; Kosowsky_1994 . The subsequent works by others using numerical simulations can refer to Ref.Huber_2008 ; Weir_2018 . However, considering that large-scale numerical simulations spend lots of computational resources and time, people have been trying to understand the physics of bubble collisions from an analytical view as well. In 2008, Caprini et al. proposed an alternative analytical method to study the GW from bubble collisions relying on a different treatment of the nature of stochasticityCaprini_2008 . In 2017, Caprini’s ansatz for correlator function has been adopted by Jinno and TakimotoJinno_2017 who developed a brand new analysis method to study GW spectra from bubble collisions in Minkowski spacetime by studying the past cones of the spacetime points xx and yy which show up in two-point correlator of energy-momentum tensor Tij(x)Tij(y)\langle T_{ij}(x)T_{ij}(y)\rangle. Their method not only gives a perspicuous physical picture, but also saves computational sources by avoiding large-scale numerical simulations and averts accompanying numerical errors. However, their work is based on Minkowski spacetime rather than FLRW spacetime which is definitely a better choice to take the expansion of the universe into account. Besides, they neglect the effect brought about by the finite duration of the phase transitions which could make the energy density of GW they estimated larger than the real value.

In this paper, we discuss GW from bubble collisions during the first order cosmological phase transitions in FLRW spacetime during RD(Radiation Dominated) era via the analysis method proposed by Jinno and Takimoto in Ref.Jinno_2017 and also take the effect of finite duration of phase transitions into account. As the analyses in Minkowski spacetime, the core quantity for GW spectrum calculation is still the transverse and traceless part of two-point correlator of energy-momentum tensor Tij(x)Tij(y)\langle T_{ij}(x)T_{ij}(y)\rangle. With the help of thin wall approximation and envelope approximation, we successfully obtain the analytical expressions of GW spectra described by two triple integrations which can be integrated numerically. By comparing the GW spectra with PLI (power-law integrated) sensitivity curves of several space-borne GW detectors and pulsar timing arrays, we can estimate if GW from bubble collisions in FLRW spacetime during RD era can be detected or not, which has important practical significance for future detection. Although our derivation is limited to the phase transitions happened in RD era, the analytical expressions of GW spectra we obtain is model-independent, and the derivation in other eras can be extended easily by specifying the relationship between Hubble parameters, conformal time and scale factor.

The organization of this paper is as follows. In Sec.2, a brief retrospection and summary of Jinno and Takimoto’s work is given. In Sec.3, we introduce the basic ingredients that the subsequent analyses need, including our assumptions and derivations of GW spectra. In Sec.4, we obtain the analytical expressions of ΔF(s)\Delta^{\text{F}(s)} and ΔF(d)\Delta^{\text{F}(d)} which can both be written as a triple integration, where ΔF(s)\Delta^{\text{F}(s)} and ΔF(d)\Delta^{\text{F}(d)} are defined in Sec.2. At the end of this section, we also prove our integration expression of ΔF\Delta^{\text{F}} can return to Jinno and Takimoto’s expression analytically when σ\sigma\to\infty. In Sec.5, we introduce our method to determine the “effective duration” of phase transitions and then show the numerical estimations of ΔF(s)\Delta^{\text{F}(s)} and ΔF(d)\Delta^{\text{F}(d)} and a comparison between our results and the results obtained in Minkowski spacetime. To testify the detectability of GW, we compare GW spectra with PLI sensitivity curves of several GW detectors in the end of Sec.5. A thorough conclusion of this work and some extra discussion are given in Sec.6.

2 A Brief Retrospection of Jinno and Takimotos’ Work

In this section, we give a brief retrospection and summary of their work. Because our analyses and calculations are all based on their model, we summarize the assumptions and approximations adopted by them at here and directly list the result they’ve obtained. The calculation details won’t be shown in this section, so for those readers who are interested in those details can read Ref.Jinno_2017 to get a more in-depth understanding.

Their calculations are based on Minkowski spacetime, including tensor perturbation the metric can be written as :

ds2=dt2+(δij+2hij)dxidxj\text{d}s^{2}=-\text{d}t^{2}+(\delta_{ij}+2h_{ij})\text{d}x^{i}\text{d}x^{j} (1)

where hijh_{ij} are all transverse and traceless. The speed of light cc has been set to 1, we’ll adopt this convention as well in our own derivation. As for the convention of indices, the Greek indices run over 0,1,2,30,1,2,3 and the Latin indices run over 1,2,31,2,3 throughout our paper.

2.1 Assumptions and approximations

They adopt thin wall approximation and envelope approximation in their work. The so-called thin wall approximation means that we assume all energy of the bubble is located in the bubble wall with an infinitesimal width B\ell_{B}. With the help of this approximation, they write the energy momentum tensor TBT^{B} of the uncollided wall of a single bubble nucleated at xN(tN,𝒙N)x_{N}\equiv(t_{N},\bm{x}_{N}) as:

Tij(x)=ρ(x)(xxN)^i(xxN)^jT_{ij}(x)=\rho(x)\widehat{(x-x_{N})}_{i}\widehat{(x-x_{N})}_{j} (2)

where

ρ(x)={4π3rB(t)3κρ04πrB(t)2BrB(t)<|𝒙𝒙N|<rB(t)0Other regions\rho(x)=\left\{\begin{aligned} &\frac{\frac{4\pi}{3}r_{B}(t)^{3}\kappa\rho_{0}}{4\pi r_{B}(t)^{2}\ell_{B}}\quad r_{B}(t)<\absolutevalue{\bm{x}-\bm{x}_{N}}<r_{B}^{\prime}(t)\\ &\hskip 28.45274pt0\hskip 65.44142pt\text{Other regions}\end{aligned}\right. (3)

and

rB(t)=v(ttN)rB(t)=rB(t)+Br_{B}(t)=v(t-t_{N})\quad r_{B}^{\prime}(t)=r_{B}(t)+\ell_{B} (4)

At here, x(t,𝒙)x\equiv(t,\bm{x}) denotes the spacetime point, ^\hat{\bullet} indicates the unit vector in the same direction of \vec{\bullet}, vv is the speed of the bubble wall, ρ0\rho_{0} represents the energy density released by the phase transitions, and κ\kappa denotes the fraction of the energy localized in the bubble wall to the released energy during the phase transition.

Envelope approximation means that as long as two or more bubbles collide, the colliding parts of bubble walls and corresponding energy-momentum tensor vanish at once. With the guarantee of envelope approximation, every spatial point(𝒙\bm{x}) can be passed by bubble only once. Since once a spatial point enter the inner part of any single bubble, it will transform from the false vacuum state to the true vacuum state and the inverse process is forbidden. See FIG.1 for a straightforward understanding of this important approximation.

Refer to caption
Figure 1: 2D explanation of envelope approximation, the black lines represent bubble walls. Only uncollided parts of bubbles are regarded as GW sources, and the energy-momentum tensor of the collided parts of bubbles vanish at once.

Besides these two approximations, they denote the bubble nucleation rate per unit time per volume as:

Γ(t)=Γeβ(tt)\Gamma(t)=\Gamma_{\star}\text{e}^{\beta(t-t_{\star})} (5)

where tt_{\star} denotes a time point typically around the transition time, Γ\Gamma_{\star} is the transition rate at tt_{\star}, and β\beta is a constant.

In order to obtain the GW spectrum, they define equal-time correlator of GW and unequal-time correlator of energy momentum tensor given by Eq.(6), which we’ll also take use of in our own derivation.

h˙ij(t,𝒌)h˙ij(t,𝒒)=(2π)3δ(3)(𝒌𝒒)Ph˙(t,k)\displaystyle\langle\dot{h}_{ij}(t,\bm{k})\dot{h}_{ij}^{*}(t,\bm{q})\rangle=(2\pi)^{3}\delta^{(3)}(\bm{k}-\bm{q})P_{\dot{h}}(t,k) (6)
Πij(tx,𝒌)Πij(ty,𝒒)=(2π)3δ(3)(𝒌𝒒)Π(tx,ty,k)\displaystyle\langle\Pi_{ij}(t_{x},\bm{k})\Pi_{ij}^{*}(t_{y},\bm{q})\rangle=(2\pi)^{3}\delta^{(3)}(\bm{k}-\bm{q})\Pi(t_{x},t_{y},k)

Where Πij\Pi_{ij} is the transverse-traceless part of the energy-momentum tensor TijT_{ij}.

2.2 GW spectrum

According to the assumptions and approximations given by the previous subsection, they find the core quantity we need to obtain for describing the GW spectrum is the transverse and traceless part of Tij(x)Tij(y)\langle T_{ij}(x)T_{ij}(y)\rangle i.e. Πij(x)Πij(y)\langle\Pi_{ij}(x)\Pi_{ij}(y)\rangle. After figuring it out, we can get Π(tx,ty,k)\Pi(t_{x},t_{y},k) with the help of Eq.(6) and Fourier transformation. As long as we have Π(tx,ty,k)\Pi(t_{x},t_{y},k), we can calculate the dimensionless GW energy density per logarithmic frequency ΩGW\Omega_{\text{GW}} by:

ΩGW:=1ρtotdρGWdlnk\displaystyle\Omega_{\text{GW}}:=\frac{1}{\rho_{\text{tot}}}\frac{\text{d}\rho_{\text{GW}}}{\text{d}\ln k} =κ2(Hβ)2(α1+α)2Δ(k/β,v)κ2(α1+α)2ΔM(k/β,σ,v)\displaystyle=\kappa^{2}\Bigg{(}\frac{H_{\star}}{\beta}\Bigg{)}^{2}\Bigg{(}\frac{\alpha}{1+\alpha}\Bigg{)}^{2}\Delta(k/\beta,v)\equiv\kappa^{2}\Bigg{(}\frac{\alpha}{1+\alpha}\Bigg{)}^{2}\Delta^{\text{M}}(k/\beta,\sigma,v) (7)
σ:=βHΔM(k/β,σ,v)Δ(k/β,v)σ2\sigma:=\frac{\beta}{H_{\star}}\quad\quad\Delta^{\text{M}}(k/\beta,\sigma,v)\equiv\frac{\Delta(k/\beta,v)}{\sigma^{2}} (8)

The superscript M indicates Minkowski spacetime. The definition of Δ(k/β,v)\Delta(k/\beta,v) is given by

Δ(k/β,v)=34π2β2k3κ2ρ02titfdtxtitfdtycos[k(txty)]Π(tx,ty,k)\Delta(k/\beta,v)=\frac{3}{4\pi^{2}}\frac{\beta^{2}k^{3}}{\kappa^{2}\rho_{0}^{2}}\int_{t_{i}}^{t_{f}}\text{d}t_{x}\int_{t_{i}}^{t_{f}}\text{d}t_{y}\cos[k(t_{x}-t_{y})]\Pi(t_{x},t_{y},k) (9)

Here ρtot=ρ0+ρrad\rho_{\text{tot}}=\rho_{0}+\rho_{\text{rad}}, α:=ρ0/ρrad\alpha:=\rho_{0}/\rho_{\text{rad}} represents the fraction of the released energy density from the phase transitions to the energy density of radiation, HH_{\star} denotes the typical value of Hubble parameter around the transition time, tit_{i} and tft_{f} denote the start time and end time of phase transitions. σ\sigma is defined as the ratio of β\beta and HH_{\star}, which can characterize the speed of phase transitions to the expansion speed of the universe. For the convenience of latter comparison, we define a new variable ΔM\Delta^{\text{M}} at here, we’ll see that ΔM\Delta^{\text{M}} is the specific variable that we’ll use to compare with our results in Sec.5 rather than Δ\Delta itself. To prevent possible confusion, we’ll explicitly write superscript M to distinguish Δ\Delta and ΔM\Delta^{\text{M}}.

According to analyses, Δ\Delta can be divided into two parts as Δ=Δ(s)+Δ(d)\Delta=\Delta^{(s)}+\Delta^{(d)}. Δ(s)\Delta^{(s)} and Δ(d)\Delta^{(d)} correspond to the contribution to Δ\Delta from the single bubble case and double bubble case respectively. The so-called single bubble case and double bubble case need to be understood by considering the relative positions of bubble walls and two spatial points 𝒙\bm{x} and 𝒚\bm{y}. It is easy to realize that only if Tij(x)0T_{ij}(x)\neq 0 and Tij(y)0T_{ij}(y)\neq 0 can this situation contributes to Tij(x)Tij(y)\langle T_{ij}(x)T_{ij}(y)\rangle, which means that the spatial points 𝒙\bm{x} and 𝒚\bm{y} must be in the bubble wall(s) at the time points txt_{x} and tyt_{y}, respectively. So we can find that there are two possible situations. The first one is that 𝒙\bm{x} and 𝒚\bm{y} are passed by the same bubble at txt_{x} and tyt_{y} respectively, and the second one is that 𝒙\bm{x} and 𝒚\bm{y} are passed by two different bubbles at txt_{x} and tyt_{y} respectively. FIG.2 displays these two possible situations, the left panel and right panel correspond to the single bubble case and double bubble case, respectively. Because txt_{x} can be equal or not equal to tyt_{y}, for each case there are two more possibilities. The upper left panel shows the situation where tx=tyt_{x}=t_{y}. The spatial points 𝒙\bm{x} and 𝒚\bm{y} are passed by the same bubble at t=tx=tyt=t_{x}=t_{y}. The lower left panel shows the situation where the bubble firstly passes the spatial point 𝒙\bm{x} at t=txt=t_{x} and lately passes the spatial point 𝒚\bm{y} at t=tyt=t_{y}. The right panel can be understood similarly.

Refer to caption
Figure 2: Sketch of single bubble case and double bubble case. xx and yy are two spacetime points we are studying, bubble(s) is/are denoted by \mathcal{B}, the superscripts (s)(s) and (d)(d) distinguish the single bubble case and double bubble case. The subscripts txt_{x} and tyt_{y} denote t=txt=t_{x} or t=tyt=t_{y}, subscript 11 and 22 denote the first bubble and the second bubble in double bubble case. The left panel corresponds to the single bubble case, and the right panel characterizes the double bubble case. Note that we don’t require the spatial points 𝒙\bm{x} and 𝒚\bm{y} to stay in the bubble walls at the same time, so they can be in the bubble walls at different time as the lower panel of this figure displays.

In order to calculate Δ(s)\Delta^{(s)} and Δ(d)\Delta^{(d)}, we need to study the past cones of xx and yy to decide the permitted spacetime points for bubbles to nucleate. For each possible nucleation point, we need to figure out two quantities. The first one is the probability of such a situation, and the second one is the contribution to Tij(x)Tij(y)\langle T_{ij}(x)T_{ij}(y)\rangle from this situation. By summing up all of the possible situations, we can get the final result of Tij(x)Tij(y)\langle T_{ij}(x)T_{ij}(y)\rangle. In a word, calculation process can be described by TT(x,y)=situationP(situation)TT(x,y)|situation\displaystyle{\langle\text{TT}\rangle(x,y)=\sum_{\text{situation}}\text{P}(\text{situation})\text{TT}(x,y)\Big{|}_{\text{situation}}}. The detailed calculation idea can be found in their Appendix A.

As a result of their work, Jinno and Takimoto find that the GW spectrum can be described as two double integrations which can be calculated numerically. The shape of GW spectrum they’ve got is consistent with the result given by numerical simulations. They also give the fitting formula of Δ\Delta as:

Δ=Δpeakcl(ffpeak)3+(1clch)(ffpeak)1+ch(1fpeak)\Delta=\frac{\Delta_{\text{peak}}}{c_{l}\Big{(}\frac{f}{f_{\text{peak}}}\Big{)}^{-3}+(1-c_{l}-c_{h})\Big{(}\frac{f}{f_{\text{peak}}}\Big{)}^{-1}+c_{h}\Big{(}\frac{1}{f_{\text{peak}}}\Big{)}} (10)

with Δpeak=0.043,fpeak/β=1.24/2π\Delta_{\text{peak}}=0.043,f_{\text{peak}}/\beta=1.24/2\pi and (cl,ch)=(0.064,0.48)(c_{l},c_{h})=(0.064,0.48). We’ll use this formula to calculate Δ\Delta and compare it with our results in Sec.5.

3 Basic setup

Because our discussion is based on FLRW spacetime, let’s recall the FLRW metric at the very beginning:

ds2=a2(η)(dη2+dx2+dy2+dz2)\text{d}s^{2}=a^{2}(\eta)(-\text{d}\eta^{2}+\text{d}x^{2}+\text{d}y^{2}+\text{d}z^{2}) (11)

where x,y,zx,y,z are comoving coordinates, a(η)a(\eta) is scale factor and the argument η\eta is conformal time whose definition is given by:

dη:=dta(t)η:=tdta(t)\text{d}\eta:=\frac{\text{d}t}{a(t)}\Longrightarrow\eta:=\int^{t}\frac{\text{d}t}{a(t)} (12)

with tt being called cosmic time. In the latter parts of this section, we firstly specify the approximations and assumptions we use in this paper, and then we give the GW spectrum at the end of phase transitions and nowadays which we can detect directly.

3.1 Assumptions and approximations

In this work, we adopt thin wall approximation and envelope approximation as Jinno and Takimoto’s choice in their work. The concrete definitions of these two approximations have been discussed in Sec.2. According to Jinno and Takimoto’s setup, we can write the energy-momentum tensor of an uncollided bubble nucleated at xN(ηN,𝒙N)x_{N}\equiv(\eta_{N},\bm{x}_{N}) as:

Tij(x)=a2(η)(ρ(x)(xxN)^i(xxN)^j)T_{ij}(x)=a^{2}(\eta)\Bigg{(}\rho(x)\widehat{(x-x_{N})}_{i}\widehat{(x-x_{N})}_{j}\Bigg{)} (13)

where

ρ(x)={4π3rB(η)3κρ04πrB(η)2a(η)BrB(η)<Δ𝒙<rB(η)0Other regions\rho(x)=\left\{\begin{aligned} &\frac{\frac{4\pi}{3}r_{B}(\eta)^{3}\kappa\rho_{0}}{4\pi r_{B}(\eta)^{2}a(\eta)\ell_{B}}\quad r_{B}(\eta)<\Delta\bm{x}<r_{B}^{\prime}(\eta)\\ &\hskip 28.45274pt0\hskip 65.44142pt\text{Other regions}\end{aligned}\right. (14)

with Δ𝒙a(η)|𝒙𝒙N|\Delta\bm{x}\equiv a(\eta)\absolutevalue{\bm{x}-\bm{x}_{N}}. To remind, the expressions of TijT_{ij} given by Eq.(13) is a modeling rather than a rigorous expression by derivation. The physical meanings of ρ0\rho_{0} and κ\kappa are both the same as which has been given in the previous section. In our work, we only consider the situation where the speed of bubble walls equals to the speed of the light, i.e. v=c=1v=c=1. So we have:

rB(η)=ηηNrB(η)=rB(η)+Br_{B}(\eta)=\eta-\eta_{N}\quad r_{B}^{\prime}(\eta)=r_{B}(\eta)+\ell_{B} (15)

Note that B\ell_{B} is the comoving width of the bubble wall. For the future convenience, we define Ψij(η,𝒙)\Psi_{ij}(\eta,\bm{x}) and Ψij(η,𝒌)\Psi_{ij}(\eta,\bm{k}) as:

Ψij(η,𝒙):=TijTT(η,𝒙)a2(η)Ψij(η,𝒌):=TijTT(η,𝒌)a2(η)\Psi_{ij}(\eta,\bm{x}):=\frac{T^{\text{TT}}_{ij}(\eta,\bm{x})}{a^{2}(\eta)}\Longrightarrow\Psi_{ij}(\eta,\bm{k}):=\frac{T^{\text{TT}}_{ij}(\eta,\bm{k})}{a^{2}(\eta)} (16)

where superscript TT is the abbreviation of Transverse-Traceless. Ψij(η,𝒌)\Psi_{ij}(\eta,\bm{k}) and TijTT(η,𝒌)T_{ij}^{\text{TT}}(\eta,\bm{k}) are the Fourier modes of Ψij(η,𝒙)\Psi_{ij}(\eta,\bm{x}) and TijTT(η,𝒙)T_{ij}^{\text{TT}}(\eta,\bm{x}) with wave vector 𝒌\bm{k}, respectively. In this work, our conventions of Fourier transformation are given by d3xei𝒌𝒙\displaystyle{\int\text{d}^{3}xe^{i\bm{k}\cdot\bm{x}}} and d3k/(2π)3ei𝒌𝒙\displaystyle{\int\text{d}^{3}k/(2\pi)^{3}e^{-i\bm{k}\cdot\bm{x}}}. We can extract the TT parts of Tij(η,𝒌)T_{ij}(\eta,\bm{k}) by contracting it with Lambda tensor whose definition is given by Eq.(22).

The transition ratio per unit conformal time per comoving volume we use in our paper is given byCaprini_2008 :

Γ(η)=Γeβ~(ηη)\Gamma(\eta)=\Gamma_{\star}e^{\widetilde{\beta}(\eta-\eta_{\star})} (17)

where η\eta_{\star} denotes the conformal time corresponding to the start of phase transitions, and Γ\Gamma_{\star} is the value of Γ(η)\Gamma(\eta) at this specific time. Note that the definition of β~\widetilde{\beta} is given by β~:=a(η)β|η=η=\widetilde{\beta}:=a(\eta)\beta\Big{|}_{\eta=\eta_{\star}}=const. where β\beta is more common to be used in the literature e.g. Ref.Kosowsky_1993 . At here, it is necessary to explain that Eq.(17) only holds when ηη\eta\sim\eta_{\star}, so in our work, we consider a finite length of phase transitions to make sure this relationship can be satisfied. Actually, later we’ll see the choice of FLRW spacetime itself gives us a relatively natural way to determine the duration of phase transitions which will directly influence the final GW spectrum. In most literature, people estimate the duration of the phase transitions by β1\beta^{-1}, and in Jinno and Takimoto’s work they set the phase transitions start at ti=t_{i}=-\infty and end at tf=+t_{f}=+\infty which itself means that they regard the duration of phase transitions as infinite long in their actual derivation and numerical calculation. Although they give an argument to this point that this treatment doesn’t change their final results, in our view, a finite duration might be a better choice and could make the results more reliable. The exact method to determine the length of phase transitions will be discussed in the latter section.

Without loss of generality, we will set β~1\widetilde{\beta}\equiv 1 in the rest of this paper which is just a choice of the unit like setting the speed of light c1c\equiv 1. Besides, it’s also worth mentioning that although β~\widetilde{\beta} is a constant given a certain value of β\beta and aa i.e. given a specific physical scenario, we are not allowed to compare any dimensional physical quantity by comparing the number of them in different scenarios. Because the physical meaning of “1” in each particular scenario is different. However, by the virtue of dimensionless property of the quantity h02ΩGWh_{0}^{2}\Omega_{\text{GW}}, we can compare it safely in different physical scenarios.

3.2 GW power spectrum derivation

In this part we begin to derive the analytical expressions of GW spectrum using the model proposed by Ref.Jinno_2017 .

Let’s consider the equation of motion of the metric perturbation satisfy at first. As we have stated already, our spacetime background is FLRW spacetime. Including tensor perturbations, we have the spacetime metric as:

ds2=a2(η)[dη2+(δij+hij)dxidxj]\text{d}s^{2}=a^{2}(\eta)[-\text{d}\eta^{2}+(\delta_{ij}+h_{ij})\text{d}x^{i}\text{d}x^{j}] (18)

Here hijh_{ij} are all traceless and transverse. By calculating the Christoffel symbol, Riemman tensor and Ricci tensor, we have the linearized Einstein’s equation to the first order in hijh_{ij}, over the FLRW spacetime as Caprini_2018 :

h¨ij(𝒙,t)+3Hh˙ij(𝒙,t)2a2hij(𝒙,t)=16πGΨij(𝒙,t)\ddot{h}_{ij}(\bm{x},t)+3H\dot{h}_{ij}(\bm{x},t)-\frac{\nabla^{2}}{a^{2}}h_{ij}(\bm{x},t)=16\pi G\Psi_{ij}(\bm{x},t) (19)

Where HH is Hubble parameter at tt, 2ii\nabla^{2}\equiv\partial^{i}\partial_{i} indicates the Laplacian associated with comoving coordinates xix^{i} and ˙\dot{\bullet} denotes the derivative of \bullet respective to cosmic time tt. Define Hij(η,𝒙)=a(η)hij(η,𝒙)H_{ij}(\eta,\bm{x})=a(\eta)h_{ij}(\eta,\bm{x}) and Eq.(19) in Fourier space becomes

Hij′′(𝒌,η)+(k2a′′a)Hij(𝒌,η)=16πGa3Ψij(𝒌,η)H_{ij}^{\prime\prime}(\bm{k},\eta)+\Bigg{(}k^{2}-\frac{a^{\prime\prime}}{a}\Bigg{)}H_{ij}(\bm{k},\eta)=16\pi Ga^{3}\Psi_{ij}(\bm{k},\eta) (20)

where \bullet^{\prime} denotes the derivative of \bullet with respect to conformal time η\eta. For the convenience of subsequent derivation, we define ζkη\zeta\equiv k\eta at here. Since we assume that the phase transitions happen in RD era, we can use the relationship of scale factor and conformal time a(η)=χηa(\eta)=\chi\eta to simplify the equation of motion and obtain Eq.(21) where χ\chi is a constant.

d2Hij(𝒌,ζ)dζ2+Hij(𝒌,ζ)=16πGχ3k5ζ3Ψij(𝒌,ζ)\frac{\text{d}^{2}H_{ij}(\bm{k},\zeta)}{\text{d}\zeta^{2}}+H_{ij}(\bm{k},\zeta)=\frac{16\pi G\chi^{3}}{k^{5}}\zeta^{3}\Psi_{ij}(\bm{k},\zeta) (21)

As we have stated before, Ψij(η,𝒙)\Psi_{ij}(\eta,\bm{x}) is related to the TT part of energy-momentum tensor TijT_{ij}, and we can use Lambda tensor to extract the TT part of TijT_{ij}. The definition of Lambda tensor is given by:

{Λij,kl(𝒌^):=Pik(𝒌^)Pjl(𝒌^)12Pij(𝒌^)Pkl(𝒌^)Pij(𝒌^):=δij𝒌^i𝒌^j\left\{\begin{aligned} &\Lambda_{ij,kl}(\widehat{\bm{k}}):=P_{ik}(\widehat{\bm{k}})P_{jl}(\widehat{\bm{k}})-\frac{1}{2}P_{ij}(\widehat{\bm{k}})P_{kl}(\widehat{\bm{k}})\\ &P_{ij}(\widehat{\bm{k}}):=\delta_{ij}-\widehat{\bm{k}}_{i}\widehat{\bm{k}}_{j}\end{aligned}\right. (22)

It’s easy to check that the Lambda tensor has a basic property which is of great use:

Λij,kl(𝒌^)Λij,mn(𝒌^)=Λkl,mn(𝒌^)=Λmn,kl(𝒌^)\Lambda_{ij,kl}(\widehat{\bm{k}})\Lambda_{ij,mn}(\widehat{\bm{k}})=\Lambda_{kl,mn}(\widehat{\bm{k}})=\Lambda_{mn,kl}(\widehat{\bm{k}}) (23)

So we can rewrite Ψij\Psi_{ij} in terms of TijT_{ij} with the help of Lambda tensor:

Ψij(η,𝒌)=Λij,kl(𝒌)Tkl(η,𝒌)a2(η)\Psi_{ij}(\eta,\bm{k})=\frac{\Lambda_{ij,kl}(\bm{k})T_{kl}(\eta,\bm{k})}{a^{2}(\eta)} (24)

Assuming that the phase transitions begin at ζi=kηi\zeta_{i}=k\eta_{i} and end at ζf=kηf\zeta_{f}=k\eta_{f}. Ψij=0\Psi_{ij}=0 before the phase transitions. Let’s consider a time point ζi<ζ<ζf\zeta_{i}<\zeta<\zeta_{f} during the phase transitions, we can get the solution of HijH_{ij} by Green’s function method:

Hij(𝒌,ζ<ζf)=16πGχ3k5ζiζdψψ3sin(ζψ)Ψij(𝒌,ψ)H_{ij}(\bm{k},\zeta<\zeta_{f})=\frac{16\pi G\chi^{3}}{k^{5}}\int_{\zeta_{i}}^{\zeta}\text{d}\psi\ \psi^{3}\sin(\zeta-\psi)\Psi_{ij}(\bm{k},\psi) (25)

When phase transitions finish, Ψij=0\Psi_{ij}=0 (no bubbles anymore, all spacetime points stay in the true vacuum state), so we have

Hij(𝒌,ζ>ζf)=Aij(𝒌)cosζ+Bij(𝒌)sinζH_{ij}(\bm{k},\zeta>\zeta_{f})=A_{ij}(\bm{k})\cos\zeta+B_{ij}(\bm{k})\sin\zeta (26)

We require HijH_{ij} to satisfy the connection requirements at ζ=ζf\zeta=\zeta_{f}, i.e. Hij|ζfH_{ij}|_{\zeta_{f}} and Hij|ζfH^{\prime}_{ij}|_{\zeta_{f}} given by Eq.(25) and Eq.(26) should be equal to each other. Using these requirements, we can arrive at

{Aij(𝒌)=16πGχ3k5ζiζfdψψ3sin(ψ)Ψij(𝒌,ψ)Bij(𝒌)=16πGχ3k5ζiζfdψψ3cos(ψ)Ψij(𝒌,ψ)\left\{\begin{aligned} &A_{ij}(\bm{k})=\frac{16\pi G\chi^{3}}{k^{5}}\int_{\zeta_{i}}^{\zeta_{f}}\text{d}\psi\ \psi^{3}\sin(-\psi)\Psi_{ij}(\bm{k},\psi)\\ &B_{ij}(\bm{k})=\frac{16\pi G\chi^{3}}{k^{5}}\int_{\zeta_{i}}^{\zeta_{f}}\text{d}\psi\ \psi^{3}\cos(\psi)\Psi_{ij}(\bm{k},\psi)\end{aligned}\right. (27)

Insert Eq.(27) into Eq.(26), we can get the final expression of HijH_{ij}. Note that this solution only holds during the RD era, so our results cannot be extended to the phase transitions that happen in MD (Matter Dominated) or Λ\LambdaD (Λ\Lambda Dominated) eras.

Now let’s discuss the power spectrum of GW. First of all, we define the equal-time two-point correlator of GW by

hij(η,𝒌)(hij)(η,𝒒):=(2π)3δ(3)(𝒌𝒒)Ph(η,k)\langle h^{\prime}_{ij}(\eta,\bm{k})(h^{\prime}_{ij})^{*}(\eta,\bm{q})\rangle:=(2\pi)^{3}\delta^{(3)}(\bm{k}-\bm{q})P_{h^{\prime}}(\eta,k) (28)

Where \langle...\rangle denotes ensemble average and we will use spatial and time average to substitute ensemble average in the actual calculation. (2π)3(2\pi)^{3} is a coefficient putting at here out of convenience for future derivation. δ(3)(𝒌𝒒)\delta^{(3)}(\bm{k}-\bm{q}) shows up out of the homogeneity of the system. Interested readers could refer to Chapter 7 of Ref.Maggiore_2008 to know more details and more properties of stochastic gravitational wave background.

We define the unequal-time correlator of energy-momentum tensor by222For simplicity, we call TijT_{ij} and Ψij\Psi_{ij} both as energy-momentum tensor:

Ψij(ηx,𝒌)Ψij(ηy,𝒒):=(2π)3δ(3)(𝒌𝒒)Ψ(ηx,ηy,k)\langle\Psi_{ij}(\eta_{x},\bm{k})\Psi_{ij}^{*}(\eta_{y},\bm{q})\rangle:=(2\pi)^{3}\delta^{(3)}(\bm{k}-\bm{q})\Psi(\eta_{x},\eta_{y},k) (29)

we can expand the left hand side of Eq.(29) as:

Ψij(ηx,𝒌)Ψij(ηy,𝒒)=d3Xd3rei(𝒌𝒒)𝑿ei(𝒌+𝒒)𝒓2Ψij(ηx,𝒙)Ψij(ηy,𝒚)\langle\Psi_{ij}(\eta_{x},\bm{k})\Psi_{ij}^{*}(\eta_{y},\bm{q})\rangle=\int\text{d}^{3}X\text{d}^{3}re^{i(\bm{k}-\bm{q})\cdot\bm{X}}e^{i(\bm{k}+\bm{q})\cdot\frac{\bm{r}}{2}}\langle\Psi_{ij}(\eta_{x},\bm{x})\Psi_{ij}(\eta_{y},\bm{y})\rangle (30)

Where we have defined 𝑿:=(𝒙+𝒚)/2\bm{X}:=(\bm{x}+\bm{y})/2 and 𝒓:=𝒙𝒚\bm{r}:=\bm{x}-\bm{y} to simplify our derivation. Note that the correlator itself doesn’t rely on 𝒙\bm{x} or 𝒚\bm{y}, on the contrary, it only depends on (𝒙𝒚)𝒓(\bm{x}-\bm{y})\equiv\bm{r}. So we have

LHS=(2π)3δ(3)(𝒌𝒒)d3rei𝒌𝒓Ψij(ηx,𝒙)Ψij(ηy,𝒚)\text{LHS}=(2\pi)^{3}\delta^{(3)}(\bm{k}-\bm{q})\int\text{d}^{3}re^{i\bm{k}\cdot\bm{r}}\langle\Psi_{ij}(\eta_{x},\bm{x})\Psi_{ij}(\eta_{y},\bm{y})\rangle (31)

which tells us the factor Ψ(ηx,ηy,k)\Psi(\eta_{x},\eta_{y},k) appeared in Eq.(29) can be written as:

Ψ(ηx,ηy,k)=d3rei𝒌𝒓Ψij(ηx,𝒙)Ψij(ηy,𝒚)\Psi(\eta_{x},\eta_{y},k)=\int\text{d}^{3}re^{i\bm{k}\cdot\bm{r}}\langle\Psi_{ij}(\eta_{x},\bm{x})\Psi_{ij}(\eta_{y},\bm{y})\rangle (32)

We know that the energy density of GW is given byMTW :

ρGW(t/η)=h˙ij(t,𝒙)h˙ij(t,𝒙)32πG=hij(η,𝒙)hij(η,𝒙)32πGa2(η)\rho_{\text{GW}}(t/\eta)=\frac{\langle\dot{h}_{ij}(t,\bm{x})\dot{h}_{ij}(t,\bm{x})\rangle}{32\pi G}=\frac{\langle h_{ij}^{\prime}(\eta,\bm{x})h_{ij}^{\prime}(\eta,\bm{x})\rangle}{32\pi Ga^{2}(\eta)} (33)

Note that the t/ηt/\eta above means tt or η\eta rather than tt by η\eta, and here hijh_{ij} are all transverse and traceless. Now insert Eq.(26) and Eq.(27) into Eq.(28) and make an inverse Fourier transformation. Substitute the numerator of Eq.(33) by the result we’ve got, finally we can arrive at:

ρGW(η)=2Gχ6πa4(η)dkk6ζiζfζiζfdψdϕψ3ϕ3cos(ψϕ)Ψ(ψ,ϕ,k)\rho_{\text{GW}}(\eta)=\frac{2G\chi^{6}}{\pi a^{4}(\eta)}\int\frac{\text{d}k}{k^{6}}\int_{\zeta_{i}}^{\zeta_{f}}\int_{\zeta_{i}}^{\zeta_{f}}\text{d}\psi\text{d}\phi\ \psi^{3}\phi^{3}\cos(\psi-\phi)\Psi(\psi,\phi,k) (34)

Now we define the dimensionless energy density per logarithmic comoving wave number of GW by:

ΩGW(η,k):=1ρtotdρGWdlogk\Omega_{\text{GW}}(\eta,k):=\frac{1}{\rho_{\text{tot}}}\frac{\text{d}\rho_{\text{GW}}}{\text{d}\log k} (35)

The definition of ρtot\rho_{\text{tot}} is unchanged comparing with the definition given in Sec.2. Since we consider the spatial curvature equals to zero, the total energy of the universe specifically equals to the energy needs to close the universe i.e. ρcr=ρtot\rho_{\text{cr}}=\rho_{\text{tot}}. We assume that the phase transitions happen during RD era, so we can neglect the contribution of energy density by matter ρmat\rho_{\text{mat}} and cosmological constant ρΛ\rho_{\Lambda} at here.

Substitute ρGW\rho_{\text{GW}} by Eq.(34), we can rewrite ΩGW\Omega_{\text{GW}} as:

ΩGW(η,k)=κ2(α1+α)2ΔF(k/β~;σ;η)\Omega_{\text{GW}}(\eta,k)=\kappa^{2}\Bigg{(}\frac{\alpha}{1+\alpha}\Bigg{)}^{2}\Delta^{\text{F}}(k/\widetilde{\beta};\sigma;\eta) (36)

Here, we use superscript F to indicate FLRW spacetime. The dimensionless quantity σ:=β~=βH\sigma:=\frac{\widetilde{\beta}}{\mathcal{H}}=\frac{\beta}{H} denotes the fraction of speed of the phase transitions to the expansion speed of the universe as the same with the definition given in the previous section. Define τ\tau as the effective duration of phase transitions, since τ\tau itself is dependent on σ\sigma, different values of σ\sigma will influence the upper limit of integral of ηx\eta_{x} and ηy\eta_{y} in Eq.(38), so Δ\Delta is also a function of σ\sigma.

At this stage, we must remind readers again that our definition of ΔF\Delta^{\text{F}} is a little bit different with the definition of Δ\Delta given by the first line of Eq.(7). Our ΔF\Delta^{\text{F}} is a function of k/β~,σk/\widetilde{\beta},\sigma and η\eta, while the Δ\Delta defined by Jinno and Takimoto is a function of k/βk/\beta and the speed of bubble wall vv. The difference appears because of the different treatments of the duration of the phase transitions. In their previous work, they regard the phase transitions start at ti=t_{i}=-\infty and end at tf=+t_{f}=+\infty in their calculation, which indeed simplify their analyses and decouple σ\sigma from Δ\Delta, but also could bring about a practical problem in latter numerical integration. In the actual integration, we cannot use \infty as the upper limit, on the contrary, we have to artificially specify a finite value which is large enough to confirm the convergence of the integral. However, if we don’t know the specific relationship between τ\tau and σ\sigma, then we may specify a too large number as upper limit of integration which can definitely lead to a longer time for computation. On the other hand, if we pick a too small value as the upper limit of integration, the result we get will be inaccurate and lack of value for reference. Besides the reasons above, there is another important motivation to consider a finite duration of phase transitions. Guo et al. consider a finite life time of the sources in their work to study the GW spectrum from sound waves during the first order cosmological phase transitions in an expanding universe.Guo_2020 As a result, they find an additional suppression factor Υ(y)\Upsilon(y) should be included in the new expression of GW spectrum which could decrease GW spectrum significantly when the life time of sources is quite short. This point can be easily checked according to the FIG.15 of Ref.Guo_2020 . Be inspired by their work, we think it is necessary to take the finite duration of phase transitions into account. So in our work, τ\tau itself is a function of σ\sigma which induces that σ\sigma cannot be decoupled from ΔF\Delta^{\text{F}} anymore. As a result, the value of ΔF\Delta^{\text{F}} in our work cannot compare with Δ\Delta in Ref.Jinno_2017 directly. In fact ΔF\Delta^{\text{F}} and ΔM\Delta^{\text{M}} are two quantities we will make a comparison later, where ΔF\Delta^{\text{F}} represents our result and ΔM\Delta^{\text{M}} represents Jinno’s result.

Using Friedmann EquationWeinberg :

2=8πG3ρtota2(η)\mathcal{H}^{2}=\frac{8\pi G}{3}\rho_{\text{tot}}a^{2}(\eta) (37)

ΔF(k/β~;σ;η)\Delta^{\text{F}}(k/\widetilde{\beta};\sigma;\eta) can be rewritten as:

ΔF(k/β~;σ;η)=3k38(η)4π2κ2ρ02ηiηfηiηfdηxdηy×(ηx3ηy3cos[k(ηxηy)]Ψ(ηx,ηy,k))\Delta^{\text{F}}(k/\widetilde{\beta};\sigma;\eta)=\frac{3k^{3}\mathcal{H}^{8}(\eta)}{4\pi^{2}\kappa^{2}\rho_{0}^{2}}\int_{\eta_{i}}^{\eta_{f}}\int_{\eta_{i}}^{\eta_{f}}\text{d}\eta_{x}\text{d}\eta_{y}\times\Bigg{(}\eta_{x}^{3}\eta_{y}^{3}\cos[k(\eta_{x}-\eta_{y})]\Psi(\eta_{x},\eta_{y},k)\Bigg{)} (38)

Now, by comparing our Eq.(38) with Eq.(9), we can easily find the difference between these two equations out of two different choices of spacetime background. We’ll see the difference more clearly later from Eq.(57) and Eq.(64). At here, we adopt the Eq.(22) of Ref.Kosowsky_1994 to describe the functional relationship between κ\kappa and α\alpha for latter numerical estimation as a benchmark:

κ=11+0.715α[0.715α+4273α2]\kappa=\frac{1}{1+0.715\alpha}\Bigg{[}0.715\alpha+\frac{4}{27}\sqrt{\frac{3\alpha}{2}}\Bigg{]} (39)

Note that this equation only holds for the case of Jouguet detonation, readers can adopt the expressions given by Ref.Espinosa_2010 to do more estimations when α\alpha and vv take different values. In Ref.Espinosa_2010 , Espinosa et al. studied all of the bubble expansion regimes without specifying any particle physics model and gave the fitting formulas of κ(v,α)\kappa(v,\alpha) in their Appendix A. Note that since our derivation is only valid for phase transitions happen during the RD era, so we artificially restrict α[0,1]\alpha\in[0,1]. The work considering other eras and other ranges of α\alpha needs to be done in the future.

Now we can clearly discover that the most important quantity we need to focus on is ΔF(k/β~;σ;η)\Delta^{\text{F}}(k/\widetilde{\beta};\sigma;\eta). The coefficients κ\kappa and α\alpha should be uniquely specified by choosing a specific model. In our work, we don’t choose any specific model and regard α\alpha as a free variable, κ\kappa is a function of α\alpha described by Eq.(39). Through this approach, we can obtain a general understanding of the behavior of GW spectrum without being restricted by any specific model.

Note that the ΩGW\Omega_{\text{GW}} given by Eq.(35) denotes the dimensionless energy density of GW per logarithmic comoving wave number when GW has just been generated after the phase transitions rather than nowadays. Considering the effect of cosmological redshift yields:

h02ΩGW=1.60×105κ2(g106.75)1/3(α1+α)2ΔF(k/β~;σ;η)h_{0}^{2}\Omega_{\text{GW}}=1.60\times 10^{-5}\kappa^{2}\Bigg{(}\frac{g_{\star}}{106.75}\Bigg{)}^{-1/3}\Bigg{(}\frac{\alpha}{1+\alpha}\Bigg{)}^{2}\Delta^{\text{F}}(k/\widetilde{\beta};\sigma;\eta) (40)
f\displaystyle f =k2πa0=12π(kβ~)(β~)(aa0)H\displaystyle=\frac{k}{2\pi a_{0}}=\frac{1}{2\pi}\Bigg{(}\frac{k}{\widetilde{\beta}}\Bigg{)}\Bigg{(}\frac{\widetilde{\beta}}{\mathcal{H}}\Bigg{)}\Bigg{(}\frac{a_{\star}}{a_{0}}\Bigg{)}H_{\star} (41)
=2.63×106Hz(kβ~)(β~)(T100GeV)(g106.75)1/6\displaystyle=2.63\times 10^{-6}\text{Hz}\Bigg{(}\frac{k}{\widetilde{\beta}}\Bigg{)}\Bigg{(}\frac{\widetilde{\beta}}{\mathcal{H}}\Bigg{)}\Bigg{(}\frac{T_{\star}}{100\text{GeV}}\Bigg{)}\Bigg{(}\frac{g_{\star}}{106.75}\Bigg{)}^{1/6}

where kk is the magnitude of comoving wave vector 𝒌\bm{k}, TT_{\star} is the temperature after the phase transitions and gg_{\star} is the relativistic degrees of freedom in the universe corresponding to the temperature TT_{\star}.

3.3 Notations

For the simplicity of comparison, we adopt the notation system defined in Ref.Jinno_2017 . The only difference are listed below. We define two new time variables Ξ\Xi and ξ\xi rather than the choice of TT and tdt_{\text{d}}:

Ξ:=ηx+ηy2ξ:=ηxηy\Xi:=\frac{\eta_{x}+\eta_{y}}{2}\quad\xi:=\eta_{x}-\eta_{y} (42)

Besides these two variables, we only need to replace tt with η\eta in our derivation. The specific physical meaning of these quantities can refer to the FIG.3 and FIG.5 of Ref.Jinno_2017 . Out of the special property of FLRW metric, we find that light still move along the 45 degree line in the ηx\eta-x plane of the spacetime diagram, which indicates that we can still use the physical picture given by Jinno and Takimoto directly and make our analyses easier to follow.

4 Analysis and derivation

In this section, we begin to derive the analytical expressions of ΔF(s)\Delta^{\text{F}(s)} and ΔF(d)\Delta^{\text{F}(d)} by making use of the method proposed by Jinno and Takimoto.

We define the phase transitions start when the probability for a bubble to nucleate in a Hubble volume during a Hubble time reaches 𝒪(1)\mathcal{O}(1). Our definition of the end of phase transitions is a little bit tricky which will be shown later. Firstly, let’s consider the information that can be derived from our definition of the moment when the phase transitions start. According to the definition, we have:

Γ41\frac{\Gamma_{\star}}{\mathcal{H}_{\star}^{4}}\simeq 1 (43)

Without loss of generality, we can take the approximate equal sign as equal sign at here. So we have

Γ=4=1σ4\Gamma_{\star}=\mathcal{H}_{\star}^{4}=\frac{1}{\sigma^{4}} (44)

where we have set β~=1\widetilde{\beta}=1. Besides this, we know that the Hubble parameter related with scale factor during RD era by H(a)=χ/a2H(a)=\chi/a^{2}. So we can find that

β~=aβ=aσH=σχa=ση1\widetilde{\beta}=a_{\star}\beta=a_{\star}\sigma H_{\star}=\frac{\sigma\chi}{a_{\star}}=\frac{\sigma}{\eta_{\star}}\equiv 1 (45)

Now we have found that the conformal time corresponding to the start of the phase transitions is η=σ\eta_{\star}=\sigma. Insert η=σ,β~=1\eta_{\star}=\sigma,\widetilde{\beta}=1 and Γ=1/σ4\Gamma_{\star}=1/\sigma^{4} into Eq.(17), we have

Γ(η)=1σ4eησ\Gamma(\eta)=\frac{1}{\sigma^{4}}e^{\eta-\sigma} (46)

At this stage, we could see the advantage to adopt FLRW spacetime is twofold. On the one hand, we take a more realistic spacetime background which can reflect the expansion of the universe and then directly influence the final GW spectrum. On the other hand, we have a relatively natural way to connect η\eta and \mathcal{H} which could help us determine the start point of phase transitions and therefore we can try to figure out the duration of phase transitions rather than take ηi=\eta_{i}=-\infty and ηf=+\eta_{f}=+\infty or manually specify a cutoff which might bring about errors or lead to a longer integration time.

To prepare for the latter calculation, we need to study the false vacuum probability at first.

4.1 False vacuum probability

According to Eq.(42) and Eq.(43) of Ref.Jinno_2017 , we have the probability of two spacetime points staying in the false vacuum state as:

P(x,y)=eI(x,y)I(x,y)=Vxyd4zΓ(z)P(x,y)=e^{-I(x,y)}\quad I(x,y)=\int_{V_{xy}}\text{d}^{4}z\Gamma(z) (47)

Where the integral of I(x,y)I(x,y) can be expanded as two terms:

{Ix(y)=σηxyπ3rx3Γ(η)(2+cx)(1cx)2dη+ηxyηx4π3rx3Γ(η)dηIy(x)=σηxyπ3ry3Γ(η)(2cy)(1+cy)2dη+ηxyηy4π3ry3Γ(η)dη\left\{\begin{aligned} &I_{x}^{(y)}=\int_{\sigma}^{\eta_{xy}}\frac{\pi}{3}r_{x}^{3}\Gamma(\eta)(2+c_{x})(1-c_{x})^{2}\text{d}\eta+\int_{\eta_{xy}}^{\eta_{x}}\frac{4\pi}{3}r_{x}^{3}\Gamma(\eta)\text{d}\eta\\ &I_{y}^{(x)}=\int_{\sigma}^{\eta_{xy}}\frac{\pi}{3}r_{y}^{3}\Gamma(\eta)(2-c_{y})(1+c_{y})^{2}\text{d}\eta+\int_{\eta_{xy}}^{\eta_{y}}\frac{4\pi}{3}r_{y}^{3}\Gamma(\eta)\text{d}\eta\end{aligned}\right. (48)

After integration, we have I(x,y)=Ix(y)+Iy(x)I(x,y)=I_{x}^{(y)}+I_{y}^{(x)}:

I(x,y)=\displaystyle I(x,y)= π12rσ4{[12ξ2(Ξ22(Ξ+1)σ+2Ξ+σ2+2)3r2(ξ2+4(Ξ22(Ξ+1)σ\displaystyle\frac{\pi}{12r\sigma^{4}}\Bigg{\{}\Bigg{[}-12\xi^{2}\Bigg{(}\Xi^{2}-2(\Xi+1)\sigma+2\Xi+\sigma^{2}+2\Bigg{)}-3r^{2}\Big{(}\xi^{2}+4(\Xi^{2}-2(\Xi+1)\sigma (49)
+2Ξ+σ2+2))+r44r(3ξ2(Ξσ+1)+24(Ξσ+1)+4(Ξσ)2(Ξσ\displaystyle+2\Xi+\sigma^{2}+2)\Big{)}+r^{4}-4r\Big{(}3\xi^{2}(\Xi-\sigma+1)+24(\Xi-\sigma+1)+4(\Xi-\sigma)^{2}(\Xi-\sigma
+3))]+24eΞr2σ(ξ2r(r+4))+48reΞσcosh(ξ2)}\displaystyle+3)\Big{)}\Bigg{]}+24e^{\Xi-\frac{r}{2}-\sigma}\Big{(}\xi^{2}-r(r+4)\Big{)}+48re^{\Xi-\sigma}\cosh(\frac{\xi}{2})\Bigg{\}}

Before we go ahead, let’s explain the reason why we need to discuss the probability of spacetime points staying in the false vacuum. Remember that under the so-called envelope approximation, any spatial point can be passed by a bubble for only once. Assuming that a bubble wall passes spatial point 𝒙\bm{x} at η=ηx\eta=\eta_{x}, we must confirm that there are no bubbles nucleated inside the past cones of xx, otherwise the spatial point 𝒙\bm{x} must have already transformed from the false vacuum state to the true vacuum state before η=ηx\eta=\eta_{x} and cannot be passed by another bubble by the second time. For the sake of the reason above, we have to figure out the probability of spacetime points staying in the false vacuum at the very first step of our calculation.

4.2 Contributions from single bubble case and double bubble case

So far, we have every ingredient prepared to analyze the energy spectrum of GW. Just as the definition given by Jinno and Takimoto, ΔF\Delta^{\text{F}} can be decomposed as ΔF(s)+ΔF(d)\Delta^{\text{F}(s)}+\Delta^{\text{F}(d)}, where ΔF(s)\Delta^{\text{F}(s)} indicates the contribution to ΔF\Delta^{\text{F}} from single bubble case and ΔF(d)\Delta^{\text{F}(d)} denotes the contribution from double bubble case. The specific meaning of these two cases have been discussed in Sec.2. Here we only simply introduce the calculation procedure taking the single bubble case for an example and directly give our results, the detailed calculation ideas can refer to the Appendix A of Ref.Jinno_2017 . Remember that our ultimate target is to get the GW spectrum which can be calculated by Eq.(35) and Eq.(38). In order to use these two equations we must figure out Ψ(ηx,ηy,k)\Psi(\eta_{x},\eta_{y},k) defined by Eq.(32) at first, which requires us to find all of the possible spacetime points for bubbles to nucleate which can lead to non-vanishing energy-momentum tensor at two spacetime points xx and yy to make Ψij(x)Ψij(y)\langle\Psi_{ij}(x)\Psi_{ij}(y)\rangle nonzero. When we consider the single bubble case, we’ll easily find that only if the bubble nucleates at some special points which are belonging to the intersection part of the past cones of xx and yy (in Ref.Jinno_2017 , they call it δVxy\delta V_{xy}), may this situation contribute to ΔF(s)\Delta^{\text{F}(s)}. By taking the possibility for bubbles to nucleate in these spacetime points into consideration and calculating the contributions given by every single configuration, we can write the two-point correlator of energy-momentum tensor as below:

Ψij(ηx,𝒙)Ψij(ηy,𝒚)(s)=2π9r5σ4κ2ρ02P(ηx,ηy,r)×[12F0+14(1crk2)F1+116(1crk2)F2]\langle\Psi_{ij}(\eta_{x},\bm{x})\Psi_{ij}(\eta_{y},\bm{y})\rangle^{(s)}=\frac{2\pi}{9r^{5}\sigma^{4}}\kappa^{2}\rho_{0}^{2}P(\eta_{x},\eta_{y},r)\times\Bigg{[}\frac{1}{2}F_{0}+\frac{1}{4}(1-c_{rk}^{2})F_{1}+\frac{1}{16}(1-c_{rk}^{2})F_{2}\Bigg{]} (50)

where F0,F1F_{0},F_{1} and F2F_{2} are given by:

F0=\displaystyle F_{0}= (r2ξ2)2er2σ16{er2+σ[16(Ξσ)2(Ξ22(Ξ+2)σ+4Ξ+σ2+12)+384(Ξσ+1)\displaystyle-\frac{(r^{2}-\xi^{2})^{2}e^{-\frac{r}{2}-\sigma}}{16}\Bigg{\{}e^{\frac{r}{2}+\sigma}\Bigg{[}16(\Xi-\sigma)^{2}\left(\Xi^{2}-2(\Xi+2)\sigma+4\Xi+\sigma^{2}+12\right)+384(\Xi-\sigma+1)- (51)
8r2(Ξ22(Ξ+1)σ+2Ξ+σ2+2)+r4]32eΞ(r(r+6)+12)}\displaystyle 8r^{2}(\Xi^{2}-2(\Xi+1)\sigma+2\Xi+\sigma^{2}+2)+r^{4}\Bigg{]}-32e^{\Xi}(r(r+6)+12)\Bigg{\}}
F1=\displaystyle F_{1}= (r2ξ2)er2σ8{er2+σ[(80ξ2(Ξσ)[Ξ2(43σ)+Ξ3+Ξ(σ(3σ8)+12)σ((σ4)σ\displaystyle-\frac{(r^{2}-\xi^{2})e^{-\frac{r}{2}-\sigma}}{8}\Bigg{\{}e^{\frac{r}{2}+\sigma}\Bigg{[}(80\xi^{2}(\Xi-\sigma)\Big{[}\Xi^{2}(4-3\sigma)+\Xi^{3}+\Xi(\sigma(3\sigma-8)+12)-\sigma((\sigma-4)\sigma (52)
+12)+24]+1920ξ2+r4(ξ28(Ξ22(Ξ+1)σ+2Ξ+σ2+2))+8r2(3ξ2(Ξ22(Ξ\displaystyle+12)+24\Big{]}+1920\xi^{2}+r^{4}(\xi^{2}-8(\Xi^{2}-2(\Xi+1)\sigma+2\Xi+\sigma^{2}+2))+8r^{2}\Big{(}-3\xi^{2}(\Xi^{2}-2(\Xi
+1)σ+2Ξ+σ2+2)2(Ξσ)2(Ξ22(Ξ+2)σ+4Ξ+σ2+12)48(Ξσ+1))+3r6]\displaystyle+1)\sigma+2\Xi+\sigma^{2}+2)-2(\Xi-\sigma)^{2}(\Xi^{2}-2(\Xi+2)\sigma+4\Xi+\sigma^{2}+12)-48(\Xi-\sigma+1)\Big{)}+3r^{6}\Bigg{]}
+16eΞ(r2(r×(r(r+4)+12)+24)ξ2(r(r(r+12)+60)+120))}\displaystyle+16e^{\Xi}(r^{2}(r\times(r(r+4)+12)+24)-\xi^{2}(r(r(r+12)+60)+120))\Bigg{\}}
F2=\displaystyle F_{2}= 116{560ξ4(Ξ4+4Ξ3+12Ξ24(Ξ+1)σ3+6(Ξ(Ξ+2)+2)σ24(Ξ(Ξ(Ξ+3)+6)+6)σ\displaystyle-\frac{1}{16}\Bigg{\{}560\xi^{4}(\Xi^{4}+4\Xi^{3}+12\Xi^{2}-4(\Xi+1)\sigma^{3}+6(\Xi(\Xi+2)+2)\sigma^{2}-4(\Xi(\Xi(\Xi+3)+6)+6)\sigma
+24Ξ+σ4+24)+2r6(ξ2+4(Ξ22(Ξ+1)σ+2Ξ+σ2+2))+3r4(16ξ2(Ξ22(Ξ+1)σ\displaystyle+24\Xi+\sigma^{4}+24)+2r^{6}(\xi^{2}+4(\Xi^{2}-2(\Xi+1)\sigma+2\Xi+\sigma^{2}+2))+3r^{4}(16\xi^{2}(\Xi^{2}-2(\Xi+1)\sigma
+2Ξ+σ2+2)+ξ4+16(Ξσ)2(Ξ22(Ξ+2)σ+4Ξ+σ2+12)+384(Ξσ+1))\displaystyle+2\Xi+\sigma^{2}+2)+\xi^{4}+16(\Xi-\sigma)^{2}(\Xi^{2}-2(\Xi+2)\sigma+4\Xi+\sigma^{2}+12)+384(\Xi-\sigma+1))
120ξ2r2(ξ2Ξ22(Ξ+1)σ+2Ξ+σ2+2)+4(Ξσ)2(Ξ22(Ξ+2)σ+4Ξ+σ2+12)\displaystyle-120\xi^{2}r^{2}(\xi^{2}\Xi^{2}-2(\Xi+1)\sigma+2\Xi+\sigma^{2}+2)+4(\Xi-\sigma)^{2}(\Xi^{2}-2(\Xi+2)\sigma+4\Xi+\sigma^{2}+12)
+96(Ξσ+1))+(16ξ2(r(r(r(r+12)+84)+360)+720)r28(r(r(r(r+4)+20)+72)\displaystyle+96(\Xi-\sigma+1))+(16\xi^{2}(r(r(r(r+12)+84)+360)+720)r^{2}-8(r(r(r(r+4)+20)+72)
+144)r48ξ4(r(r(r(r+20)+180)+840)+1680))eΞr2σ+3r8}\displaystyle+144)r^{4}-8\xi^{4}(r(r(r(r+20)+180)+840)+1680))e^{\Xi-\frac{r}{2}-\sigma}+3r^{8}\Bigg{\}}

So Ψ(s)(ηx,ηy,k)\Psi^{(s)}(\eta_{x},\eta_{y},k) can be rewritten as following:

Ψ(s)(ηx,ηy,k)=d3rei𝒌𝒓2π9r5σ4κ2ρ02×P(ηx,ηy,r)[12F0+14(1crk2)F1+116(1crk2)2F2]\Psi^{(s)}(\eta_{x},\eta_{y},k)=\int\text{d}^{3}re^{i\bm{k}\cdot\bm{r}}\frac{2\pi}{9r^{5}\sigma^{4}}\kappa^{2}\rho_{0}^{2}\times P(\eta_{x},\eta_{y},r)\Bigg{[}\frac{1}{2}F_{0}+\frac{1}{4}(1-c_{rk}^{2})F_{1}+\frac{1}{16}(1-c_{rk}^{2})^{2}F_{2}\Bigg{]} (53)

Here crkc_{rk} is the shorthand of cos(𝒓^,𝒌^)\cos(\langle\hat{\bm{r}},\hat{\bm{k}}\rangle), where 𝒂,𝒃\langle\bm{a},\bm{b}\rangle denotes the angle between 𝒂\bm{a} and 𝒃\bm{b} rather than ensemble average.

The integration of ϕ\phi is trivial, since all of the terms show up in the integrand are independent with ϕ\phi. However, the integration of rr and θ\theta need to be paid more attention. Let’s consider the integration of rr first. The determination of lower and upper limit of the integration of rr needs to be careful. Note that the separation of xx and yy must be space-like, one can easily find that if the separation is time-like, there must exist a spatial point (𝒙\bm{x} or 𝒚\bm{y}) which has transformed from the false vacuum state to the true vacuum state before ηx\eta_{x} or ηy\eta_{y}. According to this requirement, we have rmin=|ηxηy|=|ξ|r_{\min}=\absolutevalue{\eta_{x}-\eta_{y}}=\absolutevalue{\xi}. Without loss of generality, we can assume ηx>ηy\eta_{x}>\eta_{y}, so we are permitted to throw away the absolute sign, and denote rmin=ξr_{\min}=\xi directly. The process of deciding the upper limit of rr is a little bit complicated. We can see FIG.3 for a straightforward understanding.

Refer to caption
Figure 3: Determination of the upper limit of rr in the integral Eq.(53). (ηx,x)(\eta_{x},\vec{x}) and (ηy,y)(\eta_{y},\vec{y}) denote two spacetime points, Sx,SyS_{x},S_{y} are their past cones, Σσ\Sigma_{\sigma} and Σσ+τ\Sigma_{\sigma+\tau} characterize two hyperplanes where η=σ\eta=\sigma and η=σ+τ\eta=\sigma+\tau respectively. Case 1 corresponds to the situation where ηxy>σ\eta_{xy}>\sigma, under such a circumstance, the bubble can nucleate after the beginning of the phase transitions and influence two spacetime points x,yx,y. Case 2 corresponds to the situation where ηxy<σ\eta_{xy}<\sigma, so the bubble must nucleate before the start of the phase transitions which is forbidden.

According to FIG.3, we can find that only if ηxy>σ\eta_{xy}>\sigma can this configuration contributes to the integration, so we have:

ηxy:=ηx+ηyr2σr2(Ξσ)\eta_{xy}:=\frac{\eta_{x}+\eta_{y}-r}{2}\geqslant\sigma\Longrightarrow r\leqslant 2(\Xi-\sigma) (54)

With the upper and low limits for the integration of rr being decided, we can perform integration and rewrite Ψ(s)(ηx,ηy,k)\Psi^{(s)}(\eta_{x},\eta_{y},k) as following:

Ψ(s)(ηx,ηy,k)=ξ2(Ξσ)dr4π2κ2ρ029r3σ4eI(x,y)×[j0(kr)F0+j1(kr)krF1+j2(kr)(kr)2F2]\Psi^{(s)}(\eta_{x},\eta_{y},k)=\int_{\xi}^{2(\Xi-\sigma)}\text{d}r\ \frac{4\pi^{2}\kappa^{2}\rho_{0}^{2}}{9r^{3}\sigma^{4}}e^{-I(x,y)}\times\Bigg{[}j_{0}(kr)F_{0}+\frac{j_{1}(kr)}{kr}F_{1}+\frac{j_{2}(kr)}{(kr)^{2}}F_{2}\Bigg{]} (55)

where j0(kr),j1(kr),j2(kr)j_{0}(kr),j_{1}(kr),j_{2}(kr) are spherical Bessel functions. Here we have used integration formula to perform the integration of θ\theta:

1+1dceicx=2j0(x)1+1dceicx(1c2)=4j1(x)x1+1dceicx(1c2)2=16j2(x)x2\int_{-1}^{+1}\text{d}c\ e^{icx}=2j_{0}(x)\quad\int_{-1}^{+1}\text{d}c\ e^{icx}(1-c^{2})=\frac{4j_{1}(x)}{x}\quad\int_{-1}^{+1}\text{d}c\ e^{icx}(1-c^{2})^{2}=\frac{16j_{2}(x)}{x^{2}} (56)

Recall Eq.(38) and insert Ψ(s)(ηx,ηy,k)\Psi^{(s)}(\eta_{x},\eta_{y},k) into it, we have

ΔF(s)(k/β~;σ;ηf)=2k33σ4(σ+τ)8σσ+τdΞ\displaystyle\Delta^{\text{F}(s)}(k/\widetilde{\beta};\sigma;\eta_{f})=\frac{2k^{3}}{3\sigma^{4}(\sigma+\tau)^{8}}\int_{\sigma}^{\sigma+\tau}\text{d}\Xi 0τdξξ2(Ξσ)dr×\displaystyle\int_{0}^{\tau}\text{d}\xi\int_{\xi}^{2(\Xi-\sigma)}\text{d}r\times (57)
[eI(x,y)r3(Ξ2ξ24)3cos(kξ)𝒥(s)(k,r,Ξ,ξ,σ)]\displaystyle\Bigg{[}\frac{e^{-I(x,y)}}{r^{3}}\Bigg{(}\Xi^{2}-\frac{\xi^{2}}{4}\Bigg{)}^{3}\cos(k\xi)\mathcal{J}^{(s)}(k,r,\Xi,\xi,\sigma)\Bigg{]}

where 𝒥(s)(k,r,Ξ,ξ,σ)\mathcal{J}^{(s)}(k,r,\Xi,\xi,\sigma) is defined by:

𝒥(s)(k,r,Ξ,ξ,σ)[j0(kr)F0+j1(kr)krF1+j2(kr)(kr)2F2]\mathcal{J}^{(s)}(k,r,\Xi,\xi,\sigma)\equiv\Bigg{[}j_{0}(kr)F_{0}+\frac{j_{1}(kr)}{kr}F_{1}+\frac{j_{2}(kr)}{(kr)^{2}}F_{2}\Bigg{]} (58)

Now we have got the energy spectrum contributed by the single bubble case successfully. We can see that the final expression of ΔF(s)\Delta^{\text{F}(s)} can be described by a triple integration, while in Jinno and Takimoto’s work ΔM(s)\Delta^{\text{M}(s)} is given by a double integration. At here, we can compare Eq.(57) with Eq.(54) of Ref.Jinno_2017 to see the difference brought by the different choices of spacetime background. Neglecting the difference of constant coefficients like 2/32/3 and 1/121/12, we can find the significant difference between two denominators. Factors (σ+τ)8(\sigma+\tau)^{8} and (Ξ2ξ2/4)3(\Xi^{2}-\xi^{2}/4)^{3} appear because we choose FLRW spacetime and hijh_{ij} thereby satisfy a different equation of motion which could make an impact to the final result. The factor τ\tau in (σ+τ)8(\sigma+\tau)^{8} especially shows the dilution effect led by the expansion of the universe. Although the collision of bubbles generally emits a bunch of energy, the lapse of time also keeps diluting it.

Next, we’re going to directly show the contribution to the total ΔF\Delta^{\text{F}} from double bubble case i.e. ΔF(d)\Delta^{\text{F}(d)}. Readers can turn to the Part D of Sec.III of Ref.Jinno_2017 for the calculation ideas and details .

As the same with the single bubble case, in order to obtain ΔF(d)\Delta^{\text{F}(d)}, we need to find the expression of Ψ(d)\Psi^{(d)}:

Ψ(d)(ηx,ηy,k)=16πξ2(Ξσ)r2drP(ηx,ηy,r)×x(d)(ηx,ηy,r)y(d)(ηx,ηy,r)j2(kr)(kr)2\Psi^{(d)}(\eta_{x},\eta_{y},k)=16\pi\int_{\xi}^{2(\Xi-\sigma)}r^{2}\text{d}rP(\eta_{x},\eta_{y},r)\times\mathcal{B}_{x}^{(d)}(\eta_{x},\eta_{y},r)\mathcal{B}_{y}^{(d)}(\eta_{x},\eta_{y},r)\frac{j_{2}(kr)}{(kr)^{2}} (59)

where x(d)(ηx,ηy,r)\mathcal{B}^{(d)}_{x}(\eta_{x},\eta_{y},r) and y(d)(ηx,ηy,r)\mathcal{B}^{(d)}_{y}(\eta_{x},\eta_{y},r) are given by:

x(d)(ηx,ηy,r)=\displaystyle\mathcal{B}_{x}^{(d)}(\eta_{x},\eta_{y},r)= πκρ0(r2ξ2)er2σ24r3σ4(er2+σ(8ξ×(Ξσ)(Ξ2+Ξ(32σ)+(σ3)σ\displaystyle-\frac{\pi\kappa\rho_{0}(r^{2}-\xi^{2})e^{-\frac{r}{2}-\sigma}}{24r^{3}\sigma^{4}}(e^{\frac{r}{2}+\sigma}(-8\xi\times(\Xi-\sigma)(\Xi^{2}+\Xi(3-2\sigma)+(\sigma-3)\sigma (60)
+6)48ξ+2r2(ξ(Ξσ+1)2(Ξ22(Ξ+1)σ+2Ξ+σ2+2))+r4)\displaystyle+6)-48\xi+2r^{2}(\xi(\Xi-\sigma+1)-2(\Xi^{2}-2(\Xi+1)\sigma+2\Xi+\sigma^{2}+2))+r^{4})
+4eΞ(12ξ+r(6ξ+r(ξ+r+2))))\displaystyle+4e^{\Xi}(12\xi+r(6\xi+r(\xi+r+2))))
y(d)(ηx,ηy,r)=\displaystyle\mathcal{B}_{y}^{(d)}(\eta_{x},\eta_{y},r)= πκρ0(r2ξ2)er2σ24r3σ4(er2+σ(8ξ×(Ξσ)(Ξ2+Ξ(32σ)+(σ3)σ+6)\displaystyle-\frac{\pi\kappa\rho_{0}(r^{2}-\xi^{2})e^{-\frac{r}{2}-\sigma}}{24r^{3}\sigma^{4}}(e^{\frac{r}{2}+\sigma}(8\xi\times(\Xi-\sigma)(\Xi^{2}+\Xi(3-2\sigma)+(\sigma-3)\sigma+6) (61)
+48ξ2r2(σ(ξ+4Ξ+4)+ξΞ+ξ+2Ξ2+4Ξ+2σ2+4)+r4)+4eΞ\displaystyle+48\xi-2r^{2}(-\sigma(\xi+4\Xi+4)+\xi\Xi+\xi+2\Xi^{2}+4\Xi+2\sigma^{2}+4)+r^{4})+4e^{\Xi}
×(r2(r+2)ξ(r(r+6)+12)))\displaystyle\times(r^{2}(r+2)-\xi(r(r+6)+12)))

Rearrange the equations above, we’ll find

Ψ(ηx,ηy,k)(d)=ξ2(Ξσ)dr16π3κ2ρ02r4σ8P(ηx,ηy,r)𝒥(d)(k,r,Ξ,ξ,σ)\Psi(\eta_{x},\eta_{y},k)^{(d)}=\int_{\xi}^{2(\Xi-\sigma)}\text{d}r\ \frac{16\pi^{3}\kappa^{2}\rho_{0}^{2}}{r^{4}\sigma^{8}}P(\eta_{x},\eta_{y},r)\mathcal{J}^{(d)}(k,r,\Xi,\xi,\sigma) (62)

where we have defined three new quantities whose definitions are given below:

x/y(d)=πκρ0r3σ4~x/y(d)𝒥(d)(k,r,Ξ,ξ,σ)=~x(d)(ηx,ηy,r)~y(d)(ηx,ηy,r)j2(kr)(kr)2\mathcal{B}^{(d)}_{x/y}=\frac{\pi\kappa\rho_{0}}{r^{3}\sigma^{4}}\widetilde{\mathcal{B}}^{(d)}_{x/y}\quad\mathcal{J}^{(d)}(k,r,\Xi,\xi,\sigma)=\widetilde{\mathcal{B}}_{x}^{(d)}(\eta_{x},\eta_{y},r)\widetilde{\mathcal{B}}_{y}^{(d)}(\eta_{x},\eta_{y},r)\frac{j_{2}(kr)}{(kr)^{2}} (63)

Insert Eq.(62-63) into Eq.(38) and then we have the final form of Δ(d)(k/β~;σ;ηf)\Delta^{(d)}(k/\widetilde{\beta};\sigma;\eta_{f}) given by Eq.(64):

ΔF(d)(k/β~;σ;ηf)=24πk3σ8(σ+τ)8σσ+τdΞ\displaystyle\Delta^{\text{F}(d)}(k/\widetilde{\beta};\sigma;\eta_{f})=\frac{24\pi k^{3}}{\sigma^{8}(\sigma+\tau)^{8}}\int_{\sigma}^{\sigma+\tau}\text{d}\Xi 0τdξξ2(Ξσ)dr×\displaystyle\int_{0}^{\tau}\text{d}\xi\int_{\xi}^{2(\Xi-\sigma)}\text{d}r\times (64)
[(Ξ2ξ24)3cos(kξ)eI(x,y)r4𝒥(d)(k,r,Ξ,ξ,σ)]\displaystyle\Bigg{[}\Bigg{(}\Xi^{2}-\frac{\xi^{2}}{4}\Bigg{)}^{3}\cos(k\xi)\frac{e^{-I(x,y)}}{r^{4}}\mathcal{J}^{(d)}(k,r,\Xi,\xi,\sigma)\Bigg{]}

Adding ΔF(s)\Delta^{\text{F}(s)} and ΔF(d)\Delta^{\text{F}(d)} up, we can get the analytical expression of ΔF\Delta^{\text{F}}, which is specifically the quantity we want to obtain. At this step, we have no idea about how to go forward by analysis anymore. The analytical expressions of two triple integrations given by Eq.(53) and Eq.(62) are too complicated that we cannot integrate them directly. However, we can integrate them numerically to see the behavior of ΔF\Delta^{\text{F}} versus different values of σ\sigma. We’ll show the result in the next section and make comparison between our results with the results in Minkowski spacetime.

So far we have successfully obtained the analytical expressions of ΔF\Delta^{\text{F}}, but how can we know the validity of our derivation? One of the good approaches is to consider a limiting case: Does our results converge to the results given in Minkowski spacetime when σ\sigma\to\infty ? When σ\sigma\to\infty, we have τ/σ0\tau/\sigma\to 0 which means that the impact brought by the expansion of the universe disappears, as a result we don’t need to consider it anymore. So in principle, our results are supposed to converge to the expressions given in Minkowski spacetime in the limiting case of large σ\sigma. We can prove it in two steps. Firstly, let’s recall Eq.(38) and compare it with Eq.(9). We can find that if we ignore the difference between Ψ(ηx,ηy,k)\Psi(\eta_{x},\eta_{y},k) and Π(tx,ty,k)\Pi(t_{x},t_{y},k), then the only difference between two integrations is that ΔF\Delta^{\text{F}} has an additional factor 8ηx3ηy3\mathcal{H}^{8}\eta_{x}^{3}\eta_{y}^{3}. By derivation, we have:

σ3σ3(σ+τ)88(ηf)ηx3ηy3(σ+τ)3(σ+τ)3(σ+τ)8\frac{\sigma^{3}\sigma^{3}}{(\sigma+\tau)^{8}}\leqslant\mathcal{H}^{8}(\eta_{f})\eta_{x}^{3}\eta_{y}^{3}\leqslant\frac{(\sigma+\tau)^{3}(\sigma+\tau)^{3}}{(\sigma+\tau)^{8}} (65)

When σ\sigma\to\infty, the phase transitions finish instantly , as a result τ/σ0\tau/\sigma\to 0, so we can rewrite Eq.(65) as:

8(ηf)ηx3ηy31σ2(for large σ)\mathcal{H}^{8}(\eta_{f})\eta_{x}^{3}\eta_{y}^{3}\simeq\frac{1}{\sigma^{2}}\quad\text{(for large $\sigma$)} (66)

At this stage, we have successfully extracted the factor 1/σ21/\sigma^{2} out of the expression of ΔF\Delta^{\text{F}}. Recall that in Eq.(7) we specially define ΔM\Delta^{\text{M}} by dividing Δ\Delta with σ2\sigma^{2} to make the definition of ΔM\Delta^{\text{M}} and ΔF\Delta^{\text{F}} consistent with each other. Now the only integrand left is cos[k(ηxηy)]Ψ(ηx,ηy,k)\cos[k(\eta_{x}-\eta_{y})]\Psi(\eta_{x},\eta_{y},k) which looks really similar to cos[k(txty)]Π(tx,ty,k)\cos[k(t_{x}-t_{y})]\Pi(t_{x},t_{y},k) and the next step we need to do is to prove Ψ(ηx,ηy,k)Π(tx,ty,k)\Psi(\eta_{x},\eta_{y},k)\to\Pi(t_{x},t_{y},k) when σ\sigma\to\infty. However, don’t forget a very important point, in our derivation the duration of phase transitions τ\tau is finite while in Jinno and Takimoto’s previous paper, they regard it as infinity to simplify their derivation. When σ\sigma\to\infty, we will also have τ\tau\to\infty although these two infinity are not the same order. So we can naturally set the upper limit of integral ηf=σ+τ\eta_{f}=\sigma+\tau as \infty. The lower limit also can be set as -\infty because when η<σ\eta<\sigma, there doesn’t exist the collision of bubbles, so they won’t contribute to the integral. So we can set ηi=\eta_{i}=-\infty, ηf=+\eta_{f}=+\infty and ignore the difference of tt and η\eta, actually we’ll obtain the totally same expressions of I(x,y),F0,F1I(x,y),F_{0},F_{1} and F2F_{2} as in Ref.Jinno_2017 , thereby leading to the same expressions of Ψ(ηx,ηy,k)\Psi(\eta_{x},\eta_{y},k) and Π(tx,ty,k)\Pi(t_{x},t_{y},k). Since tt and η\eta are both integration variables, ++cos[k(ηxηy)]Ψ(ηx,ηy,k)dηxdηy\displaystyle{\int_{-\infty}^{+\infty}\int_{-\infty}^{+\infty}\cos[k(\eta_{x}-\eta_{y})]\Psi(\eta_{x},\eta_{y},k)\text{d}\eta_{x}\text{d}\eta_{y}} certainly equals to ++cos[k(txty)]Π(tx,ty,k)dtxdty\displaystyle{\int_{-\infty}^{+\infty}\int_{-\infty}^{+\infty}\cos[k(t_{x}-t_{y})]\Pi(t_{x},t_{y},k)\text{d}t_{x}\text{d}t_{y}}. Now we can definitely confirm that our results indeed converge to the results given in Minkowski spacetime by taking large σ\sigma. Later, we’ll directly give the numerical simulation to show this point.

From the instructions above, actually we can clearly see that the difference of the final results of GW spectrum are contributed by two sources. The first one is the different choices of spacetime metric: our derivation takes the expansion of the universe into consideration, the factor 8(ηf)ηx3ηy3\mathcal{H}^{8}(\eta_{f})\eta_{x}^{3}\eta_{y}^{3} in integrand shows up for this reason. When σ𝒪(1)\sigma\sim\mathcal{O}(1) this term can play an important role and dilute the energy density of GW. The second source is the different treatments of phase transitions duration. Intuitively, the approximation of infinite duration works worse when σ\sigma is really small, so to consider a more realistic length of phase transitions is important.

5 Numerical Calculation

In the previous section, we have obtained the final form of ΔF(s)\Delta^{\text{F}(s)} and ΔF(d)\Delta^{\text{F}(d)} given by Eq.(57) and Eq.(64). However, we have not discussed the way to get the “duration” of phase transitions. Now, let’s observe the Eq.(57) and Eq.(64), there is a special factor(σ+τ)8(\sigma+\tau)^{8} in the denominator of both expressions of ΔF(s)\Delta^{\text{F}(s)} and ΔF(d)\Delta^{\text{F}(d)}. This factor coming from 8(ηfσ+τ)\mathcal{H}^{8}(\eta_{f}\equiv\sigma+\tau) directly shows the influence of the expansion of the universe. Although the collision of bubbles emits energy, the expansion of the universe dilute it all the time. In another perspective, we can rewrite 8(σ+τ)\mathcal{H}^{8}(\sigma+\tau) as 1/[1(σ+τ)]81/[\mathcal{H}^{-1}(\sigma+\tau)]^{8} where 1(σ+τ)\mathcal{H}^{-1}(\sigma+\tau) can be understood as the Hubble horizon at time σ+τ\sigma+\tau. As time elapses, the Hubble horizon keeps growing while the released energy from bubble collisions won’t increase all the time, as a result, the energy density of GW from bubble collisions in this process will increase first and decrease later. If we only pay attention to the triple integrations in the Eq.(57) and Eq.(64), we can find that two integrals increase first and converge to specific values at last with increasing τ\tau, nevertheless the prefactor decreases monotonously in this process. So we can expect the value of ΔF=ΔF(s)+ΔF(d)\Delta^{\text{F}}=\Delta^{\text{F}(s)}+\Delta^{\text{F}(d)} will increase first and decrease after it reaching the maximum. Before the maximum point, the new released energy is more powerful than the dilution effect of the expansion of the universe, so we can regard that the phase transition has not finished. After the maximum point, the new released energy is not enough to compensate for the dilution effect of the expansion of the universe, so we can regard that the phase transition has finished. Now, let’s call τ\tau as “effective duration” of phase transitions. For a given σ\sigma, we need to study the behavior of ΔF\Delta^{\text{F}} first to figure out the exact value of effective duration τ\tau and then to calculate the corresponding energy density. In a word, we define τ\tau by the following equation:

τ:=argmax𝜏(ΔF(s)+ΔF(d))\tau:=\underset{\tau}{\text{argmax}}\Bigg{(}\Delta^{\text{F}(s)}+\Delta^{\text{F}(d)}\Bigg{)} (67)

After numerically finding the exact value of τ\tau, we can integrate ΔF(s)\Delta^{\text{F}(s)} and ΔF(d)\Delta^{\text{F}(d)} numerically to study the concrete dependence of ΔF\Delta^{\text{F}} on σ\sigma. At the same time, we are supposed to compare our results with the results given by Ref.Jinno_2017 to estimate the influence brought by the differences between two background spacetime and finite duration.

Refer to caption
Figure 4: Plot of GW spectrum. The solid curves indicate our results of ΔF\Delta^{\text{F}} in FLRW spacetime, while the dashed curves indicates Jinno and Takimoto’s results of ΔM\Delta^{\text{M}} in Minkowski spacetime. When σ=2𝒪(1)\sigma=2\sim\mathcal{O}(1), two curves deviate a lot. With increasing σ\sigma, two curves become more and more close. When σ200\sigma\gtrsim 200, two curves almost coincide with each other as our expectation.

Firstly, a comparison between our results in FLRW spacetime with a finite duration and results in Minkowski spacetime with an infinite duration is given by FIG.4. Here, we have shown 7 pairs of curves where solid curves indicate our results of ΔF\Delta^{\text{F}} estimated by Eq.(57) and Eq.(64) in FLRW spacetime while dashed curves indicate Jinno and Takimoto’s results of ΔM\Delta^{\text{M}} estimated by Eq.(10) in Minkowski spacetime. It’s apparent that when σ\sigma is relative small i.e. we cannot neglect the expansion of the universe, the dashed curves and solid curves deviate a lot. When σ\sigma generally increase, we can also find the deviation between the dashed curves and solid curves decreases as our expectation. When σ200\sigma\gtrsim 200, two curves almost coincide with each other.

Refer to caption
Figure 5: The step plot of the fraction of the maximum value of ΔF\Delta^{\text{F}} to the maximum value of ΔM\Delta^{\text{M}} versus σ\sigma. When σ𝒪(10)\sigma\lesssim\mathcal{O}(10), the corresponding GW spectrum is significantly influenced by the expansion of the universe. Even when σ100\sigma\sim 100, the GW spectrum is still be depressed by 50%\% compared to the results estimated in Minkowski spacetime. When σ\sigma increases, this ratio increase as well and will finally converge to 100%100\% when σ\sigma\to\infty, in which regime the expansion of the universe can be totally neglected.

According to our numerical estimation, we obtain the behavior of the quantity max(ΔF)\max(\Delta^{\text{F}}) over max(ΔM)\max(\Delta^{\text{M}}) versus σ\sigma shown by FIG.5. Noticing that with the guarantee of Eq.(40), ΔF\Delta^{\text{F}} is proportional to h02ΩGWh_{0}^{2}\Omega_{\text{GW}}, so the significant decrease in ΔF\Delta^{\text{F}} can directly lead to weaker signals for detection. Observing FIG.5, we can clearly find the trend of this curve is the same as our previous expectation. When σ\sigma is small i.e. the expansion speed of the universe is comparable to the speed of phase transitions, the corresponding GW spectrum is significantly influenced. When σ=1\sigma=1 a.k.a. β~=\widetilde{\beta}=\mathcal{H}, the ratio is only equal to 0.2%\%. Even when σ=100\sigma=100, this ratio is still as small as 64.5%64.5\% which shows the big impact brought by the expansion of the universe and also the consideration of finite duration of phase transitions. When σ\sigma keeps increasing, this ratio will also increase monotonously and finally converge to 100%\% when σ\sigma\to\infty. As our analytical derivation in the end of the last section, when σ\sigma\to\infty, our results should return to the results estimated in Minkowski spacetime and our numerical calculation successfully shows this point.

Since GW spectrum is indeed depressed compared to Jinno and Takimoto’s estimation in Minkowski spacetime when we consider a more realistic spacetime background and also take the finite duration of phase transitions into account, we are supposed to review the detectability of GW from bubble collisions in FLRW spacetime again by comparing GW spectra with PLI sensitivity curves of GW detectors. Out of the frequency band of GW we are considering about, we neglect the terrestrial-based detectors like LIGO and Virgo whose sensitive band is around hundred hertz. We choose five space-borne GW detector proposals at here including TianQinTianQin , LISALISA , DECIGODECIGO , B-DECIGODECIGO and BBOBBO and three Pulsar Timing Arrays including EPTAEPTA , NanoGravNG and SKASKA . The results are shown by FIG.6.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 6: Plot of GW spectra and PLI sensitivity curvesPLI of EPTAEPTA , NanoGravNG , SKASKA , TianQinTianQin , DECIGODECIGO , LISALISA , B-DECIGODECIGO , DECIGODECIGO and BBOBBO . At here, we didn’t choose any specific model of phase transitions, so σ,T,α\sigma,T,\alpha are three free variables which can change without restrictions. The solid lines are calculated by numerically integrating and dotted lines are given by fitting.

According to FIG.6, we can find that although the maximum values of GW energy density do decrease a lot than in Minkowski spacetime, there still are many possible parameter combinations whose corresponding GW can be detected by TianQin, LISA, DECIGO and BBO which is really an impressing result. Even when σ=100\sigma=100, the energy density of GW from bubble collision is still large enough to be detected in the promising future. Again, we remind readers that for the little understanding of the precise physical processes happen in the era of phase transitions, we don’t adopt any specific physical model to describe it. Under such a circumstance, σ\sigma, α\alpha and TT_{\star} are all free parameters. However, we still need to confirm that the phase transitions happen during the RD era, since our derivation is based on this elementary assumption.

Then we can go forward a small step to calculate the possible parameter combinations whose corresponding GW can be detected by the above detectors. Observing the shape of the sensitivity curves shown in FIG.6, we can find that as long as the curves of GW spectra fall above the sensitivity curves of BBO, this GW signal may be detected. Based on such a consideration, we can get the FIG.7.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 7: Plotting of the parameter combinations in σα\sigma-\alpha plane with different values of TT_{\star}. The upper left Lyons blue area denotes the pairs of (σ,α,T)(\sigma,\alpha,T_{\star}) whose corresponding GW can not be detected, while the lower right salmon pink area denotes the pairs of (σ,α,T)(\sigma,\alpha,T_{\star}) whose corresponding GW can be detected.

Observing FIG.7, we can see that as the temperature after the phase transitions TT_{\star} becomes higher and higher, the area of parameter combinations whose corresponding GW can be detected (salmon pink parts in FIG.7) is becoming bigger and bigger, which is definitely what we could predict before the calculation. Recall Eq.(57), Eq.(64) and combine our numerical estimations, we can see in two situations where the values of ΔF\Delta^{\text{F}} decrease monotonously. Firstly, ΔF\Delta^{\text{F}} drops with the increase of σ\sigma, when σ2\sigma\gtrsim 2. Secondly, ΔF\Delta^{\text{F}} drops with the decrease of σ\sigma, when σ2\sigma\lesssim 2. We can try to understand this phenomenon by considering about two limiting cases. Note that σ\sigma denotes the speed of the phase transitions with respect to the speed of the expansion of the universe. If σ\sigma\to\infty i.e. the phase transitions finish instantly, we can assume that every single spacetime point transform from the false vacuum state into the true vacuum state at once where the spherical symmetry of the system is conserved and there are almost no uncollided bubbles left, so there isn’t any detectable GW signal. On the other hand, if σ0\sigma\to 0 which means the phase transitions proceed very slowly, we still cannot expect a detectable GW signal for the low probability of bubble collisions. So we can expect that when σ2\sigma\simeq 2, the needed α\alpha is the smallest. For other cases, we all need a larger α\alpha to make sure the resulting GW could be detected. When σ>2\sigma>2, we can clearly see that the larger σ\sigma is, the larger α\alpha is needed to have a detectable signal.

Now, let’s discuss the difference between the four sub-figures of FIG.7. The only difference of the parameter for these four figures is the value of TT_{\star}, we can discover that with the increase of TT_{\star}, the corresponding salmon pink area becomes bigger. Actually, we can recall Eq.(41) to obtain a straightforward understanding. When TT_{\star} is larger than 100100 GeV, the frequency of GW will increase and move toward the sensitive band of space-borne GW detectors. Although the values of h02ΩGWh_{0}^{2}\Omega_{\text{GW}} will be lower with higher TT_{\star}, the increase of the GW frequency can totally compensate for this negative impact.

To remind, at here we only adopt a special case of κ(α)\kappa(\alpha) for calculation whose results should be regarded as a benchmark rather than holding in all different cases. Nevertheless, the method we’ve taken is very straightforward and could be realized easily.

6 Conclusion and Discussion

In this work, we explored the energy spectra of GW from bubble collisions in FLRW spacetime during RD era with a finite duration of phase transitions. We adopted the analysis method proposed by Ryusuke Jinno and Masahiro Takimoto in 2017 and extended it to a conformal flat spacetime background.

In the very first beginning, we gave a brief retrospection and summary of Jinno and Takimoto’s work. For the convenience of readers, we also listed some useful equations and results of their work. Then by adopting their elegant approach, we derived the analytical expression of ΔF(s)\Delta^{\text{F}(s)} and ΔF(d)\Delta^{\text{F}(d)} as Eq.(57) and Eq.(64), respectively. Later, we discussed the behavior of ΔF\Delta^{\text{F}} qualitatively and claimed that it would increase to the maximum value first and decrease later when τ\tau increases. By the virtue of this property, we defined the so-called “effective duration” of phase transitions by finding the specific τ\tau which could maximize ΔF\Delta^{\text{F}}. By numerical integration, we found that the maximum values of Δ\Delta depressed to only 10%10\% of Jinno and Takimoto’s estimation in Minkowski spacetime when σ𝒪(10)\sigma\sim\mathcal{O}(10). Even when σ\sigma is large as 100, the ratio max(ΔF)/max(ΔM)\max{(\Delta^{\text{F}})}/\max{(\Delta^{\text{M}})} is still as small as 64.5%. The significant difference shows up out of two different reasons. Firstly, we chose a more realistic spacetime metric—FLRW metric—to describe an expanding universe. When the speed of phase transitions and the speed of the expansion of the universe was comparable, we could not neglect the “dilution” effect brought by the expansion of the universe. Secondly, we didn’t assume the phase transitions start from ηi=\eta_{i}=-\infty and end at ηf=\eta_{f}=\infty, on the contrary, we defined an effective duration of phase transitions to realize numerical estimation.

The big decrease in the energy density of GW might bring about some new challenges for future detection. So we thought it was necessary to review the detectablity of GW generated from bubble collisions in FLRW spacetime. We regarded σ\sigma, α\alpha and TT_{\star} as three free variables and compared the GW spectra corresponding to different parameter combinations with the PLI sensitivity curves of several gravitational waves detectors. After comparison, we found that albeit the overall decrease of GW spectra, there were still many possible parameter combinations whose corresponding GW spectra curves fell above the sensitivity curves of GW detectors. Considering the shape of the sensitivity curves of GW detectors, we could find that as long as the GW spectrum curves fell above the sensitivity curve of BBO, the corresponding GW might be detected by us in the future. Based on such a criterion, we got FIG.7 and found that with the increase of TT_{\star}, the possibility for us to detect GW increased as well. Note that not every point (σ,α,T)(\sigma,\alpha,T_{\star}) in FIG.7 is permitted by the theories of the first order cosmological phase transitions, provided we have chosen a specific model to describe the phase transitions, and then α,σ\alpha,\sigma and TT_{\star} are all specified uniquely. However, since we have little knowledge about the real physics processes happened around transition time, we don’t think it is necessary to choose a model. On the other hand, our result is very general. Given a specific model, we can calculate the energy spectrum and compare it with PLI sensitivity curves of GW detectors to testify the detectability of GW.

Besides these, there are many other works waiting to be done. Firstly, our derivation is limited to the phase transitions happen in RD era and the calculation in other eras are also demanded. Secondly, we only considered the situation where the bubble wall moved with the speed of light. Actually, in many models, the speed of the bubble wall doesn’t need to be cc, so it is a problem should be focused on the future research. Finally, we still used envelope approximation in our work, while there are several paper which have abandoned this approximation, see e.g.Jinno_2019 . In a word, more detailed and elaborate discussions of GW from bubble collisions in FLRW spacetime still are needed to be done.

7 Acknowledgements

We thank Ryusuke Jinno in particular for his idea, reading our manuscript and fruitful discussion. This work is supported by the National SKA Program of China (2020SKA0120300), the National Key Research and Development Program of China (No. 2020YFC2201400),and also the National Natural Science Foundation of China under Grants No. 11653002, No. 11875141 and the National Key R&D Program of China (2021YFC2203100).

References

  • [1] B. P. Abbott, R. Abbott, T. D. Abbott, et al. Observation of gravitational waves from a binary black hole merger. Phys. Rev. Lett., 116:061102, Feb 2016.
  • [2] Pau Amaro-Seoane, Heather Audley, Stanislav Babak, et al. Laser Interferometer Space Antenna, 2017.
  • [3] Z. Arzoumanian, P. T. Baker, A. Brazier, et al. The NANOGrav 11 year data set: Pulsar-timing constraints on the stochastic gravitational-wave background. The Astrophysical Journal, 859(1):47, may 2018.
  • [4] A Ashoorioon and T Konstandin. Strong electroweak phase transitions without collider traces. Journal of High Energy Physics, 2009(07):086–086, jul 2009.
  • [5] Amjad Ashoorioon. Exit from inflation with a first-order phase transition and a gravitational wave blast. Physics Letters B, 747:446–453, 2015.
  • [6] Amjad Ashoorioon, Abasalt Rostami, and Javad T. Firouzjaee. Examining the end of inflation with primordial black hole mass distribution and gravitational waves. Phys. Rev. D, 103:123512, Jun 2021.
  • [7] Moritz Breitbach, Joachim Kopp, Eric Madge, Toby Opferkuch, and Pedro Schwaller. Dark, cold, and noisy: constraining secluded hidden sectors with gravitational waves. Journal of Cosmology and Astroparticle Physics, 2019(07):007–007, jul 2019.
  • [8] Chiara Caprini, Ruth Durrer, and Géraldine Servant. Gravitational wave generation from bubble collisions in first-order phase transitions: An analytic approach. Phys. Rev. D, 77:124015, Jun 2008.
  • [9] Chiara Caprini and Daniel G Figueroa. Cosmological backgrounds of gravitational waves. Classical and Quantum Gravity, 35(16):163001, jul 2018.
  • [10] Kip S. Thorne Charles W. Misner and John Archibald Wheeler. Gravitation. W. H. Freeman and Company, United States of America, 1973.
  • [11] Vincent Corbin and Neil J Cornish. Detecting the cosmic gravitational wave background with the Big Bang Observer. Classical and Quantum Gravity, 23(2435), 2006.
  • [12] Daniel Cutting, Mark Hindmarsh, and David J. Weir. Gravitational waves from vacuum first-order phase transitions: From the envelope to the lattice. Phys. Rev. D, 97:123513, Jun 2018.
  • [13] Subinoy Das, Patrick J. Fox, Abhishek Kumar, et al. The dark side of the electroweak phase transition. Journal of High Energy Physics, 2010(11):108, 2010.
  • [14] S. Dittmaier, C. Mariotti, G. Passarino, et al. Handbook of LHC Higgs Cross Sections: 1. Inclusive Observables, 2011.
  • [15] J. R. Espinosa, T. Konstandin, J. M. No, and M. Quirós. Some cosmological implications of hidden sectors. Phys. Rev. D, 78:123528, Dec 2008.
  • [16] José R Espinosa, Thomas Konstandin, José M No, et al. Energy budget of cosmological first-order phase transitions. Journal of Cosmology and Astroparticle Physics, 2010(06):028–028, jun 2010.
  • [17] Huai-Ke Guo, Kuver Sinha, Daniel Vagie, and Graham White. Phase Transitions in an Expanding Universe: Stochastic Gravitational Waves in Standard and Non-Standard Histories. JCAP, 01:001, 2021.
  • [18] Mark Hindmarsh, Stephan J. Huber, Kari Rummukainen, et al. Gravitational waves from the sound of a first order phase transition. Phys. Rev. Lett., 112:041301, Jan 2014.
  • [19] Mark Hindmarsh, Stephan J. Huber, Kari Rummukainen, et al. Numerical simulations of acoustically generated gravitational waves at a first order phase transition. Phys. Rev. D, 92:123009, Dec 2015.
  • [20] Mark Hindmarsh, Stephan J. Huber, Kari Rummukainen, et al. Shape of the acoustic gravitational wave power spectrum from a first order phase transition. Phys. Rev. D, 96:103520, Nov 2017.
  • [21] Fa Peng Huang, Youping Wan, DongGang Wang, et al. Hearing the echoes of electroweak baryogenesis with gravitational wave detectors. Phys. Rev. D, 94:041702, Aug 2016.
  • [22] Stephan J Huber and Thomas Konstandin. Gravitational wave production by collisions: more bubbles. Journal of Cosmology and Astroparticle Physics, 2008(09):022, sep 2008.
  • [23] G. H. Janssen, G. Hobbs, M. McLaughlin, et al. Gravitational wave astronomy with the SKA, 2014.
  • [24] Ryusuke Jinno, Thomas Konstandin, and Henrique Rubira. A hybrid simulation of gravitational wave production in first-order phase transitions. Journal of Cosmology and Astroparticle Physics, 2021(04):014, apr 2021.
  • [25] Ryusuke Jinno, Kazunori Nakayama, and Masahiro Takimoto. Gravitational waves from the first order phase transition of the higgs field at high energy scales. Phys. Rev. D, 93:045024, Feb 2016.
  • [26] Ryusuke Jinno and Masahiro Takimoto. Gravitational waves from bubble collisions: An analytic derivation. Phys. Rev. D, 95:024009, Jan 2017.
  • [27] Ryusuke Jinno and Masahiro Takimoto. Probing a classically conformal BLB-L model with gravitational waves. Phys. Rev. D, 95:015020, Jan 2017.
  • [28] Ryusuke Jinno and Masahiro Takimoto. Gravitational waves from bubble dynamics: beyond the envelope. Journal of Cosmology and Astroparticle Physics, 2019(01):060–060, jan 2019.
  • [29] Marc Kamionkowski, Arthur Kosowsky, and Michael S. Turner. Gravitational radiation from first-order phase transitions. Phys. Rev. D, 49:2837–2851, Mar 1994.
  • [30] M. Yu. Khlopov, R. V. Konoplich, S. G. Rubin, and A. S. Sakharov. First order phase transitions as a source of black holes in the early universe. Grav. Cosmol., 2:S1, 1999.
  • [31] Maxim Yu. Khlopov. Primordial Black Holes. Res. Astron. Astrophys., 10:495–528, 2010.
  • [32] R. V. Konoplich, S. G. Rubin, A. S. Sakharov, and M. Yu. Khlopov. Formation of black holes in first-order phase transitions as a cosmological test of symmetry-breaking mechanisms. Phys. Atom. Nucl., 62:1593–1600, 1999.
  • [33] Arthur Kosowsky and Michael S. Turner. Gravitational radiation from colliding vacuum bubbles: Envelope approximation to many-bubble collisions. Phys. Rev. D, 47:4372–4391, May 1993.
  • [34] Arthur Kosowsky, Michael S. Turner, and Richard Watkins. Gravitational radiation from colliding vacuum bubbles. Phys. Rev. D, 45:4514–4535, Jun 1992.
  • [35] Arthur Kosowsky, Michael S. Turner, and Richard Watkins. Gravitational waves from first-order cosmological phase transitions. Phys. Rev. Lett., 69:2026–2029, Oct 1992.
  • [36] L. Lentati, S. R. Taylor, C. M. F. Mingarelli, et al. European Pulsar Timing Array limits on an isotropic stochastic gravitational-wave background. Monthly Notices of the Royal Astronomical Society, 453(3):2576–2598, 08 2015.
  • [37] Jun Luo, Li Sheng Chen, Hui Zong Duan, et al. TianQin: a space-borne gravitational wave detector. Classical and Quantum Gravity, 33(3):035010, jan 2016.
  • [38] Michele Maggiore. Gravitational Waves Volume 1: Theory and Experiments. Oxford University Press, 2008.
  • [39] Michele Maggiore. Gravitational Waves Volume 2: Astrophysics and Cosmology. Oxford University Press, 2018.
  • [40] Shuichi Sato, Seiji Kawamura, Masaki Ando, et al. The status of DECIGO. Journal of Physics: Conference Series, 840:012010, may 2017.
  • [41] Eric Thrane and Joseph D. Romano. Sensitivity curves for searches for gravitational-wave backgrounds. Phys. Rev. D, 88:124032, Dec 2013.
  • [42] Steven Weinberg. Cosmology. Oxford University Press, 2008.