Enhancement of small-scale induced gravitational waves from the soliton/oscillon domination
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 . A higher value of , 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 at the confidence level [53]. Future CMB missions are expected to tighten the constraints on 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 .
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 . 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
(1) |
where the is the scale factor, 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 , the scale factor can be expressed as
(2) |
In the early Universe, effective self-resonance resulted in the amplification of perturbations of a real or complex scalar field . 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 . Solitons form rapidly from fragmented scalar field lumps, allowing their formation to be approximated as instantaneous, occurring at time . 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 and the total energy density of the Universe as , and we introduce the fractional energy denisity of solitons at as . If the solitons are formed during the eMD era, then . Conversely, if the solitons are formed during the eRD era, we have . In this paper, we primarily focus on the two cases where and .
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, , also decreases with time, and solitons rapidly decay once the charge reaches the critical value , a parameter that depends on the specific soliton model. For solitons formed during the eRD era, their energy density is negligible at . As the Universe expands, the energy density of radiation scales as while the energy density of matter scales as , 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 and the decay time as . The Universe is dominated by matter at , and we find the relationship .
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 . In this scenario, the Universe to be dominated by matter at , then solitons form at and dominate the Universe at , where , and finally decay at .
Similarly, we can find the relationship of the comoving wavenumbers with corresponding to different time nodes: the wavenumber at the time of soliton formation, ; the wavenumber when the Universe begins to be dominated by matter, ; the wavenumber when solitons start to dominate the Universe, ; and the wavenumber at the time of soliton decay, . Additionally, we define the characteristic wavenumber of solitons as , 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 . The lifetime of solitons can be described in terms of , where is defined as . If solitons were formed during eRD era, the hierarchy of the relevant scales reads
(3) |
and the relationship between them can be expressed through the parameters of the soliton model as follows
(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
(5) |
where we assume that since the formation of solitons is a sub-horizon process. The relationship between them at this time is
(6) |
We also obtain the evolution of the scale factor and conformal Hubble parameter as conformal time in different cases. For the solitons formed during the eRD era, the results can be written as
(7) |
(8) |
For the solitons formed during the eMD era,
(9) |
(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 , the power spectrum of soliton density perturbations can be expressed as [55]:
(11) |
where represents the density perturbations of . The smoothed spectrum of density fluctuations is valid up to . For smaller scales, the complex effects of granularity become important and we set as the cutoff scale.
The Fourier expansion of soliton density perturbations is
(12) |
By plugging eq. (12) into eq. (11), the power spectrum of can be expressed as
(13) |
If solitons form during the eRD era, i.e., , we define the density contrast of solitons, . In this case, isocurvature perturbations are given by
(14) |
where we have used . Thus, isocurvature fluctuations can be wiritten as
(15) |
and the dimensionless isocurvature power spectrum can then be expressed as
(16) |
During the eRD era, such isocurvature perturbations will convert into adiabatic perturbations. At the beginning of the sRD era, , we can write the Bardeen potential, , as a summation of the two components,
(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 are related to as
(18) |
As predicted by most well-motivated inflationary models, the power spectrum of is approximately scale-invariant,
(19) |
From Planck-2018 data, we apply the scalar spectrum amplitude , the pivot scale , and the scalar spectral index [56].
Then, we investigate the evolution of isocurvature perturbations in the eRD era and eMD era. At the linear order, scalar perturbations satisfies
(20) |
where the prime denotes the deviative with respect to conformal time and 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, does not evolve with time for the subhorizon modes. For , the mode enters the horizon during the eMD era and becomes constant. For , 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 by solving eq. (20)
(21) |
If solitons form during the eMD era, i.e., , we can obtain curvature perturbations directly from energy density perturbations of solitons using the relationship
(22) |
Considering the fact that is a constant during the eMD era, we obtain that
(23) |
where and are the Fourier forms of and . In this way, we can get the power spectrum of scalar perturbations induced by solitons as
(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 -dependent suppression factor, which depends on the specific decay behavior of the soliton, in general, can be expressed as [57]:
(25) |
where refers to the instantaneous transition value of scalar perturbations [58], and describes the decay behavior of the soliton which is dependent on the soliton model, and generally satisfies 111 In certain soliton models, such suppression does not occur, represented effectively by setting and ..
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
(26) |
where 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
(27) |
where the overline denotes the oscillation average and we have defined dimensionless parameter . is the kernel of induced GWs, which takes into account the time evolution of scalar perturbations,
(28) |
The Green’s function can be obtained by solving the equation
(29) |
In the following, we define , where is the transfer function and is the primordial value. Since the velocities of solitons are negligible, we can calculate in terms of the transfer function
(30) |
The GW energy spectrum is defined as the fraction of the GW energy density per logarithmic wavelength
(31) |
The gravitational wave (GW) source term peaks around 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 . Utilizing entropy conservation, the GW energy spectrum today, , can be expressed in terms of in the sRD era as [59, 60]
(32) |
where is the energy density fraction of radiation evaluated today [53]. In this paper, we use to denote the effective relativistic degrees of freedom , and .
III.2 Scalar induced GWs in scenarios of soliton formation
In the eMD Universe, we can divide the contribution of into three components, , and , 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 and , the dominant contribution to GWs comes from , and the GW energy spectrum can be expressed as
(33) |
As mentioned above, the scalar power spectrum can be written as
(34) |
If solitons form during the eRD era, the inflationary scalar perturbation modes, , that reenter the horizon during the eRD era, are significantly suppressed so that we can set the cutoff scale of at . If solitons form during the eMD era, primarily referring to oscillons in this paper, with 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 is also valid in this case. Taking into account that eq. (21) and eq. (24) are valid only for , we can obtain
(35) |
(36) |
Since , whether solitons were produced in the eMD era or the eRD era, the conmoving Hubble parameter in the sRD era can be approximately expressed as . We introduce a dimensionless variable , 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
(37) |
For simplicity, we define
(38) |
The general solution for the transfer function in the sRD era, preceded by an eMD era can be written as
(39) |
where and are the first and second spherical Bessel functions, and are constants which depend on . Demanding the continuity of the transfer function and its time dericative at , we can determine and as
(40) |
(41) |
In principle, using eq. (39) and eq. (30), we can directly obtain the GW energy density . 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 on the scales of and , and use this approximate form to calculate induced GWs.
In the small-scale approximation, and , the primary contributions to the kernel term come from the sum of trigonometric terms. These include the terms such as , , cosine integral terms like , and sine integral terms , where , , and represent the coefficients that depend on , and . Additionally, we define [6]. Furthermore, we introduce the cosine integral and sine integral functions, defined as:
(42) |
(43) |
In this case, we are interested only in the resonant part of induced GWs. Following the calculation of the kernel term of , the terms containing contribute the main part [37]
(44) |
For the large-scale approximation, considering the cutoff conditions of and , the integrals over and are primarily dominated by the region where is large and is small. Under these approximations, we find that
(45) |
Therefore, the energy spectrum of induced GWs can be expressed as
(46) |
These four terms represent the contribution of and on small and large scales, respectively. Using eq. (44) and eq. (45), we can derive the following expression
(47) |
(48) |
(49) |
(50) |
Here, the lower and upper limits of the integral are described by and , which come from the cutoff condition of and . The upper limit is a function of
(51) |
The result of eq. (46) also contains contributions from the cross terms between and . In most cases either or dominates the total , is mainly contributed from the dominant term and the cross term is negligible.
IV results and the constraints to soliton model








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 . 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 [61]. Thus, our results represent a conservative estimate of .
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 . 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 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, , 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 , 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 . Under these conditions, the induced GWs could potentially be detected by future GW observatories, as illustrated in Fig. 2 where we assume .
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 , and , 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, . A higher 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 . The limit given by the current and future observations on is [62]
(52) |
This gives the upper bound on GWs density at C.L.
(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:
(54) |
Fig. 4 illustrates the constraints imposed on the parameter space of the soliton model from . 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 .
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 .
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, . 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 .
In this work, our calculations assume small density perturbations remain in the linear regime, . 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
- Abbott et al. [2016] B. P. Abbott et al. (LIGO Scientific, Virgo), Observation of Gravitational Waves from a Binary Black Hole Merger, Phys. Rev. Lett. 116, 061102 (2016), arXiv:1602.03837 [gr-qc] .
- Cai et al. [2017] R.-G. Cai, Z. Cao, Z.-K. Guo, S.-J. Wang, and T. Yang, The Gravitational-Wave Physics, Natl. Sci. Rev. 4, 687 (2017), arXiv:1703.00187 [gr-qc] .
- Baumann et al. [2007] D. Baumann, P. J. Steinhardt, K. Takahashi, and K. Ichiki, Gravitational Wave Spectrum Induced by Primordial Scalar Perturbations, Phys. Rev. D 76, 084019 (2007), arXiv:hep-th/0703290 .
- Saito and Yokoyama [2009] R. Saito and J. Yokoyama, Gravitational wave background as a probe of the primordial black hole abundance, Phys. Rev. Lett. 102, 161101 (2009), [Erratum: Phys.Rev.Lett. 107, 069901 (2011)], arXiv:0812.4339 [astro-ph] .
- Ananda et al. [2007] K. N. Ananda, C. Clarkson, and D. Wands, The Cosmological gravitational wave background from primordial density perturbations, Phys. Rev. D 75, 123518 (2007), arXiv:gr-qc/0612013 .
- Kohri and Terada [2018] K. Kohri and T. Terada, Semianalytic calculation of gravitational wave spectrum nonlinearly induced from primordial curvature perturbations, Phys. Rev. D 97, 123532 (2018), arXiv:1804.08577 [gr-qc] .
- Cai et al. [2019a] R.-g. Cai, S. Pi, and M. Sasaki, Gravitational Waves Induced by non-Gaussian Scalar Perturbations, Phys. Rev. Lett. 122, 201101 (2019a), arXiv:1810.11000 [astro-ph.CO] .
- Cai et al. [2019b] R.-G. Cai, S. Pi, S.-J. Wang, and X.-Y. Yang, Resonant multiple peaks in the induced gravitational waves, JCAP 05, 013, arXiv:1901.10152 [astro-ph.CO] .
- Cai et al. [2020] R.-G. Cai, S. Pi, and M. Sasaki, Universal infrared scaling of gravitational wave background spectra, Phys. Rev. D 102, 083528 (2020), arXiv:1909.13728 [astro-ph.CO] .
- Espinosa et al. [2018a] J. R. Espinosa, D. Racco, and A. Riotto, A Cosmological Signature of the SM Higgs Instability: Gravitational Waves, JCAP 09, 012, arXiv:1804.07732 [hep-ph] .
- Domènech [2021] G. Domènech, Scalar Induced Gravitational Waves Review, Universe 7, 398 (2021), arXiv:2109.01398 [gr-qc] .
- Martin et al. [2020a] J. Martin, T. Papanikolaou, and V. Vennin, Primordial black holes from the preheating instability in single-field inflation, JCAP 01, 024, arXiv:1907.04236 [astro-ph.CO] .
- Martin et al. [2020b] J. Martin, T. Papanikolaou, L. Pinol, and V. Vennin, Metric preheating and radiative decay in single-field inflation, JCAP 05, 003, arXiv:2002.01820 [astro-ph.CO] .
- Choudhury et al. [2024] S. Choudhury, A. Karde, S. Panda, and M. Sami, Primordial non-Gaussianity from ultra slow-roll Galileon inflation, JCAP 01, 012, arXiv:2306.12334 [astro-ph.CO] .
- Khlopov et al. [1998] M. Y. Khlopov, R. V. Konoplich, S. G. Rubin, and A. S. Sakharov, Formation of black holes in first order phase transitions, (1998), arXiv:hep-ph/9807343 .
- Lozanov and Amin [2019] K. D. Lozanov and M. A. Amin, Gravitational perturbations from oscillons and transients after inflation, Phys. Rev. D 99, 123504 (2019), arXiv:1902.06736 [astro-ph.CO] .
- Liu et al. [2022] J. Liu, L. Bian, R.-G. Cai, Z.-K. Guo, and S.-J. Wang, Primordial black hole production during first-order phase transitions, Phys. Rev. D 105, L021303 (2022), arXiv:2106.05637 [astro-ph.CO] .
- Liu et al. [2023] J. Liu, L. Bian, R.-G. Cai, Z.-K. Guo, and S.-J. Wang, Constraining First-Order Phase Transitions with Curvature Perturbations, Phys. Rev. Lett. 130, 051001 (2023), arXiv:2208.14086 [astro-ph.CO] .
- Ellis et al. [2019] J. Ellis, M. Lewicki, J. M. No, and V. Vaskonen, Gravitational wave energy budget in strongly supercooled phase transitions, JCAP 06, 024, arXiv:1903.09642 [hep-ph] .
- Baker et al. [2021] M. J. Baker, M. Breitbach, J. Kopp, and L. Mittnacht, Primordial Black Holes from First-Order Cosmological Phase Transitions, (2021), arXiv:2105.07481 [astro-ph.CO] .
- Liu [2023] J. Liu, Distinguishing the nanohertz gravitational-wave sources by the observations of compact dark matter subhalos, Phys. Rev. D 108, 123544 (2023).
- Bhaumik et al. [2022] N. Bhaumik, A. Ghoshal, and M. Lewicki, Doubly peaked induced stochastic gravitational wave background: testing baryogenesis from primordial black holes, JHEP 07, 130, arXiv:2205.06260 [astro-ph.CO] .
- Sui et al. [2024] X.-B. Sui, J. Liu, X.-Y. Yang, and R.-G. Cai, Detecting the dark sector through scalar-induced gravitational waves, (2024), arXiv:2407.04220 [astro-ph.CO] .
- Cai et al. [2024] R.-G. Cai, Y.-S. Hao, and S.-J. Wang, Primordial black holes and curvature perturbations from false-vacuum islands, (2024), arXiv:2404.06506 [astro-ph.CO] .
- Flores et al. [2024] M. M. Flores, A. Kusenko, and M. Sasaki, Revisiting formation of primordial black holes in a supercooled first-order phase transition, (2024), arXiv:2402.13341 [hep-ph] .
- Abbott et al. [2021] R. Abbott et al. (KAGRA, Virgo, LIGO Scientific), Upper limits on the isotropic gravitational-wave background from Advanced LIGO and Advanced Virgo’s third observing run, Phys. Rev. D 104, 022004 (2021), arXiv:2101.12130 [gr-qc] .
- Ruan et al. [2020] W.-H. Ruan, Z.-K. Guo, R.-G. Cai, and Y.-Z. Zhang, Taiji program: Gravitational-wave sources, Int. J. Mod. Phys. A 35, 2050075 (2020), arXiv:1807.09495 [gr-qc] .
- Wu et al. [2021] Y.-L. Wu et al. (Taiji Scientific), China’s first step towards probing the expanding universe and the nature of gravity using a space borne gravitational wave antenna, Commun. Phys. 4, 34 (2021).
- Luo et al. [2016] J. Luo et al. (TianQin), TianQin: a space-borne gravitational wave detector, Class. Quant. Grav. 33, 035010 (2016), arXiv:1512.02076 [astro-ph.IM] .
- Barausse et al. [2020] E. Barausse et al., Prospects for Fundamental Physics with LISA, Gen. Rel. Grav. 52, 81 (2020), arXiv:2001.09793 [gr-qc] .
- Inomata et al. [2019a] K. Inomata, K. Kohri, T. Nakama, and T. Terada, Gravitational Waves Induced by Scalar Perturbations during a Gradual Transition from an Early Matter Era to the Radiation Era, JCAP 10, 071, [Erratum: JCAP 08, E01 (2023)], arXiv:1904.12878 [astro-ph.CO] .
- Lozanov and Takhistov [2023] K. D. Lozanov and V. Takhistov, Enhanced Gravitational Waves from Inflaton Oscillons, Phys. Rev. Lett. 130, 181002 (2023), arXiv:2204.07152 [astro-ph.CO] .
- White et al. [2021] G. White, L. Pearce, D. Vagie, and A. Kusenko, Detectable Gravitational Wave Signals from Affleck-Dine Baryogenesis, Phys. Rev. Lett. 127, 181601 (2021), arXiv:2105.11655 [hep-ph] .
- Assadullahi and Wands [2009] H. Assadullahi and D. Wands, Gravitational waves from an early matter era, Phys. Rev. D 79, 083511 (2009), arXiv:0901.0989 [astro-ph.CO] .
- Alabidi et al. [2013] L. Alabidi, K. Kohri, M. Sasaki, and Y. Sendouda, Observable induced gravitational waves from an early matter phase, JCAP 05, 033, arXiv:1303.4519 [astro-ph.CO] .
- Pearce et al. [2023] M. Pearce, L. Pearce, G. White, and C. Balazs, Gravitational Wave Signals From Early Matter Domination: Interpolating Between Fast and Slow Transitions, (2023), arXiv:2311.12340 [astro-ph.CO] .
- Inomata et al. [2019b] K. Inomata, K. Kohri, T. Nakama, and T. Terada, Enhancement of Gravitational Waves Induced by Scalar Perturbations due to a Sudden Transition from an Early Matter Era to the Radiation Era, Phys. Rev. D 100, 043532 (2019b), [Erratum: Phys.Rev.D 108, 049901 (2023)], arXiv:1904.12879 [astro-ph.CO] .
- Mukhanov [2005] V. Mukhanov, Physical Foundations of Cosmology (Cambridge University Press, Oxford, 2005).
- Rosen [1968] G. Rosen, Particlelike Solutions to Nonlinear Complex Scalar Field Theories with Positive-Definite Energy Densities, J. Math. Phys. 9, 996 (1968).
- Friedberg et al. [1976] R. Friedberg, T. D. Lee, and A. Sirlin, A Class of Scalar-Field Soliton Solutions in Three Space Dimensions, Phys. Rev. D 13, 2739 (1976).
- Coleman [1985] S. R. Coleman, Q-balls, Nucl. Phys. B 262, 263 (1985), [Addendum: Nucl.Phys.B 269, 744 (1986)].
- Kusenko [1997] A. Kusenko, Small Q balls, Phys. Lett. B 404, 285 (1997), arXiv:hep-th/9704073 .
- Amin [2010] M. A. Amin, Inflaton fragmentation: Emergence of pseudo-stable inflaton lumps (oscillons) after inflation, (2010), arXiv:1006.3075 [astro-ph.CO] .
- Kasuya et al. [2003] S. Kasuya, M. Kawasaki, and F. Takahashi, I-balls, Phys. Lett. B 559, 99 (2003), arXiv:hep-ph/0209358 .
- Copeland et al. [1995] E. J. Copeland, M. Gleiser, and H. R. Muller, Oscillons: Resonant configurations during bubble collapse, Phys. Rev. D 52, 1920 (1995), arXiv:hep-ph/9503217 .
- Broadhead and McDonald [2005] M. Broadhead and J. McDonald, Simulations of the end of supersymmetric hybrid inflation and non-topological soliton formation, Phys. Rev. D 72, 043519 (2005), arXiv:hep-ph/0503081 .
- Amin et al. [2012] M. A. Amin, R. Easther, H. Finkel, R. Flauger, and M. P. Hertzberg, Oscillons After Inflation, Phys. Rev. Lett. 108, 241302 (2012), arXiv:1106.3335 [astro-ph.CO] .
- del Corral [2024] D. del Corral, Self-resonance during preheating: The case of -attractor models, Annals Phys. 470, 169824 (2024), arXiv:2406.04017 [hep-th] .
- Lozanov and Amin [2018] K. D. Lozanov and M. A. Amin, Self-resonance after inflation: oscillons, transients and radiation domination, Phys. Rev. D 97, 023533 (2018), arXiv:1710.06851 [astro-ph.CO] .
- Aich et al. [2020] M. Aich, Y.-Z. Ma, W.-M. Dai, and J.-Q. Xia, How much primordial tensor mode is allowed?, Phys. Rev. D 101, 063536 (2020).
- Hou et al. [2013] Z. Hou, R. Keisler, L. Knox, M. Millea, and C. Reichardt, How massless neutrinos affect the cosmic microwave background damping tail, Phys. Rev. D 87, 083008 (2013).
- Wallisch [2018] B. Wallisch, Cosmological Probes of Light Relics, Ph.D. thesis, Cambridge U. (2018), arXiv:1810.02800 [astro-ph.CO] .
- Aghanim et al. [2020] N. Aghanim et al. (Planck), Planck 2018 results. VI. Cosmological parameters, Astron. Astrophys. 641, A6 (2020), [Erratum: Astron.Astrophys. 652, C4 (2021)], arXiv:1807.06209 [astro-ph.CO] .
- Di Valentino et al. [2018] E. Di Valentino et al. (CORE), Exploring cosmic origins with CORE: Cosmological parameters, JCAP 04, 017, arXiv:1612.00021 [astro-ph.CO] .
- Papanikolaou et al. [2021] T. Papanikolaou, V. Vennin, and D. Langlois, Gravitational waves from a universe filled with primordial black holes, JCAP 03, 053, arXiv:2010.11573 [astro-ph.CO] .
- Akrami et al. [2020] Y. Akrami et al. (Planck), Planck 2018 results. X. Constraints on inflation, Astron. Astrophys. 641, A10 (2020), arXiv:1807.06211 [astro-ph.CO] .
- Inomata et al. [2020a] K. Inomata, K. Kohri, T. Nakama, and T. Terada, Gravitational waves induced by scalar perturbations during a gradual transition from an early matter era to the radiation era, J. Phys. Conf. Ser. 1468, 012001 (2020a).
- Inomata et al. [2020b] K. Inomata, M. Kawasaki, K. Mukaida, T. Terada, and T. T. Yanagida, Gravitational Wave Production right after a Primordial Black Hole Evaporation, Phys. Rev. D 101, 123533 (2020b), arXiv:2003.10455 [astro-ph.CO] .
- Inomata et al. [2020c] K. Inomata, K. Kohri, T. Nakama, and T. Terada, Enhancement of gravitational waves induced by scalar perturbations due to a sudden transition from an early matter era to the radiation era, J. Phys. Conf. Ser. 1468, 012002 (2020c).
- Espinosa et al. [2018b] J. Espinosa, D. Racco, and A. Riotto, A cosmological signature of the sm higgs instability: gravitational waves, Journal of Cosmology and Astroparticle Physics 2018 (09), 012.
- Fernandez et al. [2024] N. Fernandez, J. W. Foster, B. Lillard, and J. Shelton, Stochastic Gravitational Waves from Early Structure Formation, Phys. Rev. Lett. 133, 111002 (2024), arXiv:2312.12499 [astro-ph.CO] .
- Cang et al. [2023] J. Cang, Y.-Z. Ma, and Y. Gao, Implications for Primordial Black Holes from Cosmological Constraints on Scalar-induced Gravitational Waves, Astrophys. J. 949, 64 (2023), arXiv:2210.03476 [astro-ph.CO] .