Gravitational Wave Signatures of Gauged Baryon and Lepton Number
Abstract
We demonstrate that novel types of gravitational wave signatures arise in theories with new gauge symmetries broken at high energy scales. For concreteness, we focus on models with gauged baryon number and lepton number, in which neutrino masses are generated via the type I seesaw mechanism, leptogenesis occurs through the decay of a heavy right-handed neutrino, and one of the new baryonic fields is a good dark matter candidate. Depending on the scalar content of the theory, the gravitational wave spectrum consists of contributions from cosmic strings, domain walls, and first order phase transitions. We show that a characteristic double-peaked signal from domain walls or a sharp domain wall peak over a flat cosmic string background may be generated. Those new signatures are within the reach of future experiments, such as Cosmic Explorer, Einstein Telescope, DECIGO, Big Bang Observer, and LISA.
I Introduction
In the Standard Model Glashow (1961); Higgs (1964); Englert and Brout (1964); Weinberg (1967); Salam (1968); Fritzsch et al. (1973); Gross and Wilczek (1973); Politzer (1973), baryon number and lepton number are accidental global symmetries of the Lagrangian of an unknown origin. Although one expects both of those symmetries to be violated if grand unification is realized in Nature Georgi and Glashow (1974); Fritzsch and Minkowski (1975), so far no sign of related processes, such as proton decay, have been observed in experiments Abe et al. (2017), which excludes the minimal non-supersymmetric version of those theories. If grand unification does not happen, then global baryon and lepton number may be a low-energy manifestation of some more fundamental gauge symmetries unbroken at high energy scales. Indeed, this line of reasoning is supported by the self-consistency of quantum theories of gravity, in which only gauge symmetries can be properly accommodated, unless unnatural conditions are introduced Kallosh et al. (1995).
The first attempt to promote baryon and lepton number to the status of gauge symmetries dates back to the 1970s Pais (1973), and was followed by further theoretical efforts throughout the subsequent two decades Rajpoot (1988); Foot et al. (1989); Carone and Murayama (1995); Georgi and Glashow (1996). However, the first phenomenologically viable model of this type was constructed only recently in Fileviez Perez and Wise (2010), and later modified to avoid all current experimental bounds Duerr et al. (2013); Fileviez Perez et al. (2014). The idea of gauging baryon and lepton number was later successfully incorporated into a supersymmetric framework Arnold et al. (2013), theories unifying baryon number and color Fornal et al. (2015); Fornal and Tait (2016), and generalized to a non-Abelian gauged lepton number Fornal et al. (2017). Models with gauged baryon and lepton numbers have very attractive features: they explain the stability of the proton, have a natural realization of the seesaw mechanism for neutrino masses, contain an attractive baryonic dark matter candidate Duerr and Fileviez Perez (2015); Ohmer and Patel (2015); Fileviez Perez et al. (2019), and can accommodate high scale leptogenesis Fileviez Perez et al. (2021). Thus far, in all of the existing formulations of theories with gauged baryon and lepton number, each of the symmetries was broken by the vacuum expectation value of a single scalar. However, there is no reason to expect that the scalar sector is this minimal.
To this end, in this paper we investigate ways to probe the composition of high-scale symmetry breaking sectors, i.e., when at least one of the two sectors consists of more than one scalar breaking the symmetry. Although we focus on the class of theories with gauged baryon and lepton number, most of our analysis is general and can be applied to other theories with two broken gauge symmetries. Conventional particle physics experiments are not able to differentiate between the two scenario, or even probe them at all if the symmetry breaking scale is high. Nevertheless, as we demonstrate below, gravitational wave detectors have opened up a completely new set of opportunities to probe such models.
A renaissance period for gravitational wave physics was initiated by the first direct detection of a gravitational wave signal coming from a black hole merger by the Laser Interferometer Gravitational Wave Observatory (LIGO) within the LIGO/Virgo collaboration Abbott et al. (2016). By now, over one hundred of such events, involving also neutron stars, have been recorded. Those discoveries provide an ideal opportunity to test general relativity, but they are not directly related to particle physics. The gravitational waves which enable probing particle physics models, although not yet discovered, are expected to come in the form of a stochastic gravitational wave background produced in the early Universe by phenomena such as inflation Turner (1997), first order phase transitions Kosowsky et al. (1992), domain walls Hiramatsu et al. (2010) and cosmic strings Vachaspati and Vilenkin (1985a); Sakellariadou (1990). Although for such signals to be detectable at LIGO the underlying particle physics models require a large fine-tuning of parameters, future gravitational wave experiments, such as the Laser Interferometer Space Antenna (LISA) Amaro-Seoane et al. (2017), Cosmic Explorer Reitze et al. (2019), Einstein Telescope Punturo et al. (2010), DECIGO Kawamura et al. (2011), and Big Bang Observer Crowder and Cornish (2005), will be sensitive to more generic scenarios.
The most model-independent stochastic gravitational wave background comes from cosmic strings, which are topological defects formed via the Kibble mechanism Kibble (1976) upon a spontaneous breaking of a symmetry. They correspond to one-dimensional field configurations along the direction in which the symmetry remains unbroken. The dynamics of the produced cosmic string network provides a long-lasting source of gravitational radiation resulting in a mostly flat stochastic gravitational wave background, with its strength dependent only on the scale of the breaking. Cosmic string signatures have been considered in the context of grand unified theories Buchmuller et al. (2019); King et al. (2020), neutrino seesaw models Blanco-Pillado and Olum (2017); Ringeval and Suyama (2017); Cui et al. (2018, 2019); Guedes et al. (2018); Dror et al. (2020); Zhou and Bian (2020), new physics at the high scale Gouttenoire et al. (2020a), as well as baryon and lepton number violation Fornal and Shams Es Haghi (2020). For a review of gravitational waves signatures of cosmic strings see Gouttenoire et al. (2020b), and for the constraints from LIGO/Virgo data see Abbott et al. (2021).
The other topological defects which can be produced in the early Universe are domain walls, created when a symmetry is spontaneously broken. They are two-dimensional field configurations existing at the boundaries of regions corresponding to different vacua. In order for domain walls not to overclose the Universe, they need to annihilate away. This is possible when there exists a small energy density difference between the two vacua (the so-called potential bias). Domain wall annihilation leads to a stochastic gravitational wave background which is peaked at some frequency, but its strength and the peak frequency are, as in the case of cosmic strings, independent of the exact particle physics details of the model – the spectrum is determined by only two parameters: the scale of the symmetry breaking and the potential bias. Domain wall signatures have been considered in many theoreies beyond the Standard Model, including new electroweak scale physics Eto et al. (2018a, b); Chen et al. (2020); Battye et al. (2020), supersymmetry Kadota et al. (2015), axions Craig et al. (2021); Blasi et al. (2023), grand unification Dunsky et al. (2022), models with left-right symmetry Borah and Dasgupta (2022), baryon/lepton number violation Fornal et al. (2023), and models of leptogenesis Barman et al. (2022). The physics of domain walls and the expected gravitational wave spectrum are reviewed in Saikawa (2017); the bounds on domain walls from LIGO/Virgo data can be found in Jiang and Huang (2022).
The most model-dependent gravitational wave signatures arise from cosmological first order phase transitions. Those occur when the effective potential develops a new minimum with a lower energy density than the high-temperature one. If there exists a potential barrier between the two minima, the transition is first order and bubbles of true vacuum are being nucleated in various points in space. Such bubbles of true vacuum expand, eventually filling up the entire Universe. Gravitational waves are emitted from bubble collisions, turbulence, and sound shock waves in the primordial plasma generated by the violent expansion of the bubbles. The position of the gravitational wave peak is highly dependent on the temperature at which bubble nucleation occurs. First order phase transitions have been analyzed in a plethora of particle physics models, including, again, electroweak scale new physics Grojean and Servant (2007); Vaskonen (2017); Dorsch et al. (2017); Bernon et al. (2018); Chala et al. (2018); Angelescu and Huang (2019); Alves et al. (2019); Han et al. (2021); Benincasa et al. (2022), supersymmetry Craig et al. (2020); Fornal et al. (2021), axions Dev et al. (2019); Von Harling et al. (2020); Delle Rose et al. (2020), grand unification Croon et al. (2019); Huang et al. (2020); Okada et al. (2021), baryon/lepton number violation Hasegawa et al. (2019); Fornal and Shams Es Haghi (2020)), neutrino seesaw models Brdar et al. (2019); Okada and Seto (2018); Di Bari et al. (2021); Zhou et al. (2022), new flavor physics Greljo et al. (2020); Fornal (2021), dark gauge groups Schwaller (2015); Breitbach et al. (2019); Croon et al. (2018); Hall et al. (2020), models with conformal invariance Ellis et al. (2020a); Kawana (2022), and dark matter Baldes (2017); Azatov et al. (2021); Costa et al. (2022a, b); Fornal and Pierre (2022); Kierkla et al. (2023); Azatov et al. (2022). A comprehensive review of gravitational waves from first order phase transitions can be found in Caldwell et al. (2022); Athron et al. (2023), while the most recent constraints from LIGO/Virgo data were derived in Badger et al. (2023). For recent progress on supercooled phase transitions see Ellis et al. (2019); Lewicki and Vaskonen (2020a, b); Ellis et al. (2020a).
In this work, we examine how gravitational wave signals from domain walls, cosmic strings, and phase transitions interplay with each other, producing novel features in the expected spectrum. The two new gravitational wave signatures which have not been considered in the literature so far are: Two coexisting signals from domain wall annihilation, forming a characteristic sharp double-peak in the spectrum; Domain wall signal over a flat cosmic string contribution, leading to an unusually-shaped peak. Although we focus on a specific model, our results involving cosmic strings and domain walls, in particular signatures and , are general, since they do not depend on the details of the model.
II The model
The gravitational wave signatures we propose to search for, as will be discussed in Section VII, are anticipated in a large class of models with a two-step symmetry breaking pattern. In this paper, to provide a concrete realization of such scenarios, we focus on a model with gauged baryon and lepton number, based on the gauge group
(1) |
Below we describe the possible symmetry breaking patterns in the model and the new particles along with their masses.
Scalar sector
As mentioned in Section I, we consider extending the usual single-scalar symmetry breaking sector to possibly include two scalars per each symmetry breaking. Each of the fields breaking the symmetry comes in the representation
(2) |
while each of the scalars breaking is
(3) |
Within this framework, there are four possible cases:
-
breaks and breaks ,
-
, break and breaks ,
-
breaks and , break ,
-
, break and , break .
We assume that the mixed terms involving scalars breaking different gauge groups have negligible coefficients. This implies that for a given , if only one scalar breaks the symmetry, the scalar potential is
(4) |
whereas if two scalars participate in symmetry breaking,
(5) | |||||
The scalars develop the following vacuum expectation values,
(6) |
Whenever two scalars take part in the symmetry breaking, we define This way one can collectively describe the breaking scale as , and the breaking scale as , independent of whether the vacuum expectation value comes from a single scalar or two scalars.
If lepton number is broken at a higher scale than baryon number, the symmetry breaking pattern is:
followed by the usual electroweak symmetry breaking by the Standard Model Higgs. We note that the order of and breaking may be reversed.
Fermion sector
To provide a concrete quantitative example, we consider the model with gauged and proposed in Fileviez Perez et al. (2014), which involves the minimal fermionic particle content for a theory with such a gauge group. It is straightforward to check that all gauge anomalies are cancelled if the Standard Model quark fields , , and lepton fields , are augmented by
(7) |
where is the family index. Among the fields above, are the right-handed neutrinos, whereas is a Majorana dark matter candidate discussed in Section III.
Particle masses
The scalars and generate masses for the new fermions through the following Lagrangian terms,
(8) | |||||
which provide vector-like masses to the new fermions, as well as introduce a type I seesaw mechanism for the neutrinos. For example, assuming that the Yukawa couplings are and , the measured neutrino mass splittings are reproduced if . The mass matrices for the new fermions are provided in Fileviez Perez et al. (2014).
The spontaneous breaking of and leads to the appearance of vector gauge bosons and . Given the charges of the scalars breaking the two symmetries, the corresponding masses are
(9) |
where and are the and gauge couplings, respectively, whose values are free parameters.
III Dark matter and matter-antimatter asymmetry
Apart from providing a natural framework accommodating a type I seesaw mechanism generating small neutrino masses via breaking, the model also contains a phenomenologically viable dark matter candidate Fileviez Perez et al. (2014); Ohmer and Patel (2015) and can account for the matter-antimatter asymmetry through leptogenesis Fileviez Perez et al. (2021). We discuss the most relevant aspects of those highlights of the model below.
Dark matter
After breaking, there remains a residual discrete symmetry under which the new fermions transform as
(10) |
If is the lightest of the new fermions, there is no decay channel available for it, thus it becomes a good candidate for particle dark matter.
It was argued in Duerr and Fileviez Perez (2015); Fileviez Perez et al. (2019) that in models with gauged baryon and lepton number consistency with the dark matter relic abundance of Aghanim et al. (2020) imposes an upper bound on the breaking scale. In particular, if the dark matter annihilation happens via the resonant -channel process
(11) |
a dependence between the parameters , , and arises, and the perturbativity requirement leads to . This was the reason why in Fornal and Shams Es Haghi (2020), where the gravitational wave signal from a model with gauged baryon and lepton number was considered, a low scale of was imposed.
However, as was demonstrated in Ohmer and Patel (2015), in the model we are considering other dark matter annihilation channels remain unsuppressed, including the nonresonant -channel process
(12) |
whose cross section can be sufficiently large to explain the dark matter relic density. Therefore, the arguments in Duerr and Fileviez Perez (2015); Fileviez Perez et al. (2019) do not apply in our case, and the scale of breaking can be high. Alternatively, the aforementioned bound on the breaking scale can always be avoided by assuming nonthermal dark matter production.
Leptogenesis
There cannot exist any primordial baryon or lepton number asymmetry above the scales of and breaking. An excess of matter over antimatter can only arise once one of those two symmetries is broken. A natural setting to achieve this below the scale of breaking is offered by high-scale leptogenesis (see Davidson et al. (2008) and references therein), in which a lepton number asymmetry is generated through the out-of-equilibrium decays of the lightest right-handed neutrino,
(13) |
The asymmetry is introduced through the standard interference between the tree-level diagram for the process in Eq. (13) and the one-loop diagrams involving , , and the two heavier right-handed neutrinos , in the loop.
The generated lepton asymmetry is calculated by solving the Boltzmann equations for the evolution of the lightest right-handed neutrino abundance (where is the particle density and is the co-moving entropy density) and the asymmetry Buchmuller et al. (2005),
(14) |
where , the term accounts for decays and inverse decays, represents scatterings, describes the washout effects, and is the asymmetry parameter. In our case, the Boltzmann equations are slightly different than in the standard leptogenesis scenario, since the right-handed neutrinos have an extra interaction with the gauge boson. Those equations were solved in Fileviez Perez et al. (2021), and the amount of the generated lepton asymmetry was determined for various values of model parameters.
The produced lepton asymmetry is then partially converted into a baryon asymmetry through the electroweak sphalerons. Above the scale of breaking the sphaleron-induced interactions have the form
(15) |
and it was shown in Duerr and Fileviez Perez (2015) that if the breaking of occurs close to the electroweak scale, then the final baryon asymmetry predicted by the model is given by
(16) |
This can explain the observed baryon-to-photon ratio Workman et al. (2022)
(17) |
if the scale of breaking satisfies the relation
(18) |
The requirement for such a high symmetry breaking scale provides the desired setting accommodating the type I seesaw mechanism for the neutrinos.
IV Cosmic string spectrum
Spontaneous breaking of a gauge symmetry leads to the production of topological defects in the form of cosmic strings Kibble (1976), which correspond to one-dimensional field configurations along the direction of the unbroken symmetry. The network of produced cosmic strings is described collectively by the string tension , equal to the energy stored in a string per unit length, and depends solely on the scale at which the gauge symmetry is broken Vilenkin and Shellard (2000); Gouttenoire et al. (2020b),
(19) |
where is the Planck mass, the gravitational constant , and the winding number was taken to be one. The constraints from the cosmic microwave background measurements set an upper limit on the string tension of Ade et al. (2014), which corresponds to the following bound on the scale of symmetry breaking,
(20) |
Through its dynamics, the cosmic string network provides a long-lasting source of gravitational radiation and leads to a stochastic gravitational wave background roughly constant across a wide range of frequencies.
Dynamics of cosmic strings
There are two main processes governing the behavior of a cosmic string network: formation of string loops and stretching due to the expansion of the Universe. The first of those contributions, creation of string loops, happens when long strings intersect and intercommute. The newly created string loops oscillate and emit gravitational radiation, mainly from cusps and kinks propagating through the string loop, and from kink-kink collisions Olum and Blanco-Pillado (2000); Moore et al. (2002).
A competition between these two effects leads to the so-called scaling regime, in which there is a large number of string loops and a small number of Hubble-size strings Kibble (1985); Bennett and Bouchet (1988, 1989); Albrecht and Turok (1989); Allen and Shellard (1990). There is a continuous flow of energy from long strings to string loops, and then to gravitational radiation through their decays. This gravitational radiation makes up a fixed fraction of the energy density of the Universe Hindmarsh and Kibble (1995).
To describe this process quantitatively, we follow the steps outlined in Cui et al. (2019); Gouttenoire et al. (2020b). We consider a string loop created at time with initial length , where is a constant loop size parameter. The loop oscillates emitting gravitational waves with frequencies
(21) |
where is a positive integer. Rescaling this result by the scale factor , one obtains the currently observed frequency,
(22) |
where is the time of the emission and is the time today. The spectrum of the emitted gravitational waves from a single oscillating string loop is given by Blanco-Pillado et al. (2014); Blanco-Pillado and Olum (2017)
(23) |
where for the contribution from cusps , from kinks , and from kink-kink collisions , while the overall factor Vachaspati and Vilenkin (1985b). Due to the constant emission of gravitational radiation, the string loop shrinks and its length at the time of the gravitational wave emission is
(24) |
causing the loop to vanish after the time .
The only model-dependent quantity describing the cosmic string network is the loop distribution function for the created loops. Adopting the well-established model developed in Martins and Shellard (1996a, b, 2002), describing the string network just by the mean string velocity and the correlation length, leads in the scaling regime to the following formula,
(25) |
where the constant depends on the era in the evolution of the Universe (for radiation , whereas for the matter dominated era Cui et al. (2019)).
Gravitational wave spectrum
The stochastic gravitational wave background generated by the dynamics of the cosmic string network is Cui et al. (2019); Gouttenoire et al. (2020b)
(26) | |||||
where
(27) |
In Eqs. (26) and (27) the parameters are the following: is the fraction of the loops contributing to the gravitational wave signal, estimated to be Blanco-Pillado et al. (2014) since the majority of the energy is lost by long strings going into highly boosted smaller loops which provide only a subdominant contribution; is the critical density of the Universe; the loop size parameter provides an accurate estimate of the loop size distribution Blanco-Pillado et al. (2014); Blanco-Pillado and Olum (2017); is the time at which the cosmic string network was formed, related to the density of the Universe at that time via Gouttenoire et al. (2020b), is the instance when the loop was produced, and is the Heaviside step function. We also note that, as argued in Cui et al. (2019); Gouttenoire et al. (2020b), the largest contribution to the gravitational wave signal comes from the cusps.


The resulting stochastic gravitational wave background is presented in Fig. 1 for four values of the symmetry breaking scale, in the range of frequencies relevant for the upcoming gravitational wave experiments, whose sensitivities are denoted by the colored regions. If , the signal can be seen by all the detectors: LISA Amaro-Seoane et al. (2017), DECIGO Kawamura et al. (2011), Big Bang Observer Crowder and Cornish (2005), Einstein Telescope Punturo et al. (2010), and Cosmic Explorer Reitze et al. (2019). This is illustrated in more detail in Fig. 2, which shows the reach of each experiment in terms of the symmetry breaking scale leading to a cosmic string signal. The lower bound is detector-specific, whereas the upper bound reflects the cosmic microwave background constraint in Eq. (20).
The cosmic string network will be produced if either or is spontaneously broken by a single scalar. Therefore, among the cases enumerated in Section II, the gravitational wave signals discussed here are relevant in case for both baryon and lepton number breaking, in case only for lepton number breaking, and in case only for baryon number breaking.
V Domain wall spectrum
Another kind of topological defects, appearing when a symmetry is broken by two scalars, are domain walls. As in the case of breaking discussed in Fornal and Pierre (2022), the effective potential , which at high temperature has just one vacuum at , at lower temperature develops four vacua. They come in two pairs related via a gauge transformation, , and only two of them, say and , correspond to disconnected manifolds, and thus are physically distinct Ginzburg and Krawczyk (2005); Battye et al. (2011). When a transition takes place, patches of the Universe end up in either one of those vacua, leading to the creation of domain walls, i.e., two-dimensional field configurations on the boundaries of and .
If the two vacua have identical energy densities, domain walls remain stable and considerably affect the evolution of the Universe, introducing unacceptably large density fluctuations Saikawa (2017). Therefore, for a phenomenologically viable scenario, the symmetry between the two vacua needs to be softly broken. It cannot be strongly broken, since then patches of the Universe would transition preferentially to the lower energy density vacuum and domain walls would not form. In our case, the soft breaking of the symmetry removing the degeneracy between the vacua is provided by the terms involving , and in the Lagrangian in Eq. (5).
Dynamics of domain walls
The profile of the domain wall configuration is the solution to the equation Chen et al. (2020)
(28) |
where the -axis was chosen to be perpendicular to the domain wall, and the boundary conditions are,
(29) |
As mentioned earlier, there are only two parameters which describe the gravitational wave spectrum from domain walls. The first of them is the domain wall tension given by
(30) |
In the model we are considering, to a good approximation
(31) |
The other parameter is the potential bias , i.e., the energy density difference between the vacua, in our case equal to
(32) |
The created domain walls are unstable and undergo annihilation, emitting gravitational radiation, provided that Saikawa (2017)
(33) |
If in Eq. (32) the term involving is the dominant one, then Eq. (33) takes the form . For example, if the symmetry breaking scale is , this implies . An independent constraint arises from the necessity of domain wall annihilation happening before Big Bang nucleosynthesis, so the ratios of the produced elements are not altered, but the bound in Eq. (33) remains stronger.
Gravitational wave spectrum
Domain wall annihilation leads to a stochastic gravitational wave background given by Kadota et al. (2015); Saikawa (2017),
(34) | |||||
where for the area parameter we used and for the efficiency parameter we adopted the value Hiramatsu et al. (2014), denotes the step function, and the peak frequency is
(35) |
The slope of the signal falls to the left of the peak when moving toward lower frequencies, and falls like to the right of the peak when moving toward higher frequencies. The cosmic microwave background constraint on the strength of the signal at the peak is Clarke et al. (2020), which translates to the following condition on the parameters,
(36) |
stronger than the bound imposed by Eq. (33).


Several examples of gravitational wave spectra from domain wall annihilation, plotted using Eq. (34), are shown in Fig. 3 for representative values of the parameters and . The reach of the upcoming gravitational wave detectors is also shown, including LISA, Big Bang Observer, DECIGO, Einstein Telescope, and Cosmic Explorer. A more detailed look at their sensitivity is provided by Fig. 4, which shows the full parameter space that can be probed by those experiments. The lower bound on the domain wall parameter is a reflection of the cosmic microwave background constraint from Eq. (36). The parameter depends in general on all three fundamental Lagrangian parameters , , and through the relation in Eq. (32). Under the assumption that the term involving is dominant, the experimental sensitivity plot in the plane would be the same as in figure 4 of Fornal et al. (2023).
VI First order phase transition spectrum
Perhaps the most anticipated stochastic gravitational wave signal to be discovered is the one generated by a first order phase transition in the early Universe, predicted in a large class of theories beyond the Standard Model. Such a signature in the case of breaking has been considered in Fornal and Shams Es Haghi (2020), but in this work we adopt a different assumption for the Yukawa couplings and keep our analysis more general, so that it can be applied to both gauged and gauged . A first order phase transition can occur either when the symmetry breaking sector consists of a single scalar, or contains multiple scalars. Since the generalization is straightforward, we concentrate on the case with a single scalar.
Effective potential
The effective potential for the background field consists of the tree-level part, the one-loop Coleman-Weinberg zero temperature correction, and the finite temperature contribution. Upon imposing the condition that the minimum of the zero temperature potential and the mass of remain at their tree-level values (i.e., the cutoff regularization scheme), the effective potential is given by
(37) |
In the expression above the sums are over all particles charged under the including the Goldstone bosons , are the field-dependent masses, is the number of degrees of freedom for a given particle (, , ), is similar but includes only scalars and longitudinal components of vector bosons (, , ), and for the Goldstones one needs to replace . We will assume that all new Yukawa couplings are small, so that the only relevant field-dependent masses are
(38) |
In the limit , the thermal masses are
(39) |
where for gauged baryon number and , while for gauged lepton number and .

For a range of and values the effective potential develops a vacuum at (true vacuum) with a lower energy density than the high temperature vacuum at (false vacuum), separated by a potential bump, which are precisely the conditions needed for a first order phase transition to take place. The changing shape of the effective potential is shown in Fig. 5 for a particular choice of parameters, in the case of gauged baryon number.
Dynamics of the phase transition
A first order phase transition from the false vacuum to the true vacuum of a given patch of the Universe corresponds to the nucleation of a bubble which then starts expanding. This process is initiated at the nucleation temperature , which is determined from the condition that the bubble nucleation rate Linde (1983) becomes comparable with the Hubble expansion,
(40) |
where is the Euclidean action given by
(41) |
with being the solution of the bubble equation,
(42) |
subject to the boundary conditions
(43) |
Since , Eq. (40) becomes
(44) |
The phase transition parameters relevant for determining the gravitational wave signal are: the bubble wall velocity , the nucleation temperature , the phase transition strength , and its duration . In our analysis we assume , but other choices are also possible Espinosa et al. (2010); Caprini et al. (2016). The other three parameters, , , and , are determined from the behavior of the effective potential with changing temperature. As such, those parameters encode information about the details of the particle physics model considered.
The phase transition strength is calculated as the ratio of the energy density difference between the false and true vacuum, and that of radiation, both taken at nucleation temperature,
(45) |
Those two quantities are obtained from the relations
(46) |
where is the number of degrees of freedom active when the bubbles are nucleated. The parameter , related to the time scale of the phase transition, is determined via
(47) |
Gravitational wave spectrum
The dynamics of the nucleated bubbles generates gravitational waves through sound shock waves in the early Universe plasma, bubble collisions, and magnetohydrodynamic turbulence. The expected contribution of each of those processes to the stochastic gravitational wave background was determined through numerical simulations, and the corresponding empirical formulas were derived. The resulting gravitational wave signal is the sum of the three contributions,
(48) |
The expected signal from sound waves is Hindmarsh et al. (2014); Caprini et al. (2016)
(49) | |||||
where is the peak frequency, is the fraction of the latent heat transformed into the bulk motion of the plasma Espinosa et al. (2010), and is the suppression factor Ellis et al. (2020b); Guo et al. (2021),
(50) |
The signal from bubble wall collisions is Kosowsky et al. (1992); Huber and Konstandin (2008); Caprini et al. (2016) (see Lewicki and Vaskonen (2021) for recent updates)
(51) | |||||
where now is the peak frequency and is the fraction of the latent heat deposited into the bubble front Kamionkowski et al. (1994),
(52) |
The final contribution is provided by turbulence Caprini and Durrer (2006); Caprini et al. (2009),
(53) | |||||
where Caprini et al. (2016), while the peak frequency and the parameter are
(54) |

The resulting gravitational wave signals from the first order phase transition triggered by breaking are shown in Fig. 6 for several symmetry breaking scales: (light brown curve), (brown curve), (purple curve), and (black curve). The gauge coupling was assumed to be , and the quartic coupling was chosen to be . Spectra with peaks at larger frequencies correspond to higher symmetry breaking scales.
The effect of the suppression factor reducing the sound wave contribution in Eq. (49) is that the bubble collision and turbulence components become distinguishable in the spectrum. Although the main peak is still due to sound waves, the slope at lower frequencies is dominated by bubble collisions, whereas for higher frequencies the turbulence contribution visibly changes the slope of the curve. Without the suppression factor, the spectrum would be determined, to a good approximation, just by the sound wave component in the region relevant for future detectors.
Depending on the Lagrangian parameters, the signal may be detectable in upcoming gravitational wave experiments: LISA, DECIGO, Big Bang Observer, Einstein Telescope, and Cosmic Explorer. To study this more quantitatively, in Fig. 7 we show part of the parameter space for which the signal can be detected in each experiment when . Specifically, the upper boundary for each detector corresponds to a signal-to-noise ratio of five upon a single year of data taking, while the lower boundary arises when either is too large to satisfy Eq. (44) or the new vacuum has an energy density larger than that of the high temperature vacuum.

As mentioned earlier, our analysis of the first order phase transition signal from breaking differs from the one in Fornal and Shams Es Haghi (2020) in several aspects. The fermionic particle content in Eq. (II) is different, and we chose the corresponding Yukawa couplings to be small, which is a more minimal scenario than in Fornal and Shams Es Haghi (2020). Our analysis is also applicable to breaking, since we calculate the thermal masses in the general case – this result will be used in Section VII. Finally, when determining the gravitational wave signal we included the effect of bubble collisions, which was not considered in Fornal and Shams Es Haghi (2020), but which increases the reach of upcoming detectors due to the enhancement of the signal in the lower frequency region.
VII Gravitational wave signatures
In this section we demonstrate the diversity of gravitational wave signatures expected within the framework of the model, searchable in near-future experiments. The cases enumerated in Section II, corresponding to the possible structures of the scalar sectors, give rise to the coexistence in the spectrum of the following gravitational wave signals from first order phase transitions (PT), cosmic strings (CS), and domain walls (DW):
We discuss below explicit examples of how those signatures are realized in our model. Two of them, and have not been considered in the literature before, whereas , and have been already proposed. The case does not give rise to any new features, since the signal is dominated by the cosmic string contribution from the higher symmetry breaking due to the flatness of the cosmic string spectrum.
Domain walls + domain walls


A new gravitational wave signature arises when each of the two symmetries is broken by two scalars, leading to the production of domain walls at two different energy scales during the evolution of the Universe. The signal consists of two sharp domain wall peaks. The slope on the left side of each peak depends on the frequency like , whereas the slope on the right side of the peak falls like . There is a nontrivial structure created between the two peaks, which can be used to distinguish this type of signal from others. If the two symmetry breaking scales are high, this signature can be searched for in all the upcoming gravitational wave experiments we considered: LISA, DECIGO, Big Bang Observer, Einstein Telescope, and Cosmic Explorer. A realization of this scenario in our model is shown in Fig. 8, where the parameters for the breaking were chosen to be and , whereas for the breaking they are and .
Cosmic strings + domain walls
Another gravitational wave signature, not considered in the literature before, is realized when one of the symmetries is broken by one scalar, leading to cosmic string production, whereas the other is broken by two scalars, resulting in domain wall creation. If the two symmetry breaking scales are high, their contributions may overlap and produce a very unusual domain wall peak over the cosmic string background. An example is shown in Fig. 9, where the symmetry breaking scale for was chosen to be , whereas the parameters for breaking are and . For this particular selection of parameters, Big Bang Observer and DECIGO can probe the peak area, but for lower breaking scales this structure is accessible to LISA, whereas higher symmetry breaking scales would make it detectable by Einstein Telescope and Cosmic Explorer.

Phase transition + cosmic strings
If one of the symmetries is broken by a single scalar at the high scale, and the other symmetry is broken by either one or two scalars at the low scale, this can lead to a gravitational wave signature consisting of a phase transition bump over a cosmic string background. This signature was first proposed in Fornal and Shams Es Haghi (2020) in the context of a different gauged baryon and lepton number model with a well-motivated large hierarchy between symmetry breaking scales, and recently considered in another scenario Ferrer et al. (2023). In Fig. 10 we show an example of such a signal, where is broken at the scale , whereas the scale of breaking is . The other parameter values for this particular plot are and . As in the previous case, by changing the scale of breaking the bump can shift and become searchable not only by Big Bang Observer and DECIGO, but also by Cosmic Explorer, Einstein Telescope, or LISA.
Phase transition + phase transition
Independently of the scalar sector structure, the breaking of two gauge symmetries can always result in a gravitational wave signal with two first order phase transition peaks. Such a signature is generically expected in theories with a multistep symmetry breaking pattern, and has been proposed for various models of new physics Angelescu and Huang (2019); Greljo et al. (2020); Fornal (2021). In Fig. 11 a realization of this signature is shown in the case of our model, assuming that the symmetry is broken by one scalar at the scale (the other parameters are and ), and the symmetry is broken also by one scalar at the scale (with and ). We note that for the two contributions appropriate formulas for the thermal masses were adopted, according to Eq. (VI). Such a signal can be searched for in all the future gravitational wave detectors we considered.

Phase transition + domain wall
The final qualitatively different signature consists also of two peaks, but this time one coming from a first order phase transition and the second one arising from domain wall annihilation. Such a signal has very recently been proposed in Fornal et al. (2023). In our model it can be realized if there is a large hierarchy between the and symmetry breaking scales. Figure 12 shows a realization of this scenario when is broken by two scalars at the scale (with a potential bias ), whereas is broken by one scalar at the scale (with the other parameters being and ). As pointed out in Fornal et al. (2023), the two peaks may appear in a different order, which would happen for a breaking scale of and a breaking scale of . In both scenarios, the signature can be searched for in the upcoming gravitational wave detectors we focused on.
Although the signatures discussed above can be realized for any pattern of symmetry breaking, the phenomenologically more attractive scenarios involve broken at the high scale, so that the bound in Eq. (18) is satisfied and the theory can successfully accommodate leptogenesis, as discussed in Section III. Additionally, with no motivation for a low breaking scale in the model we are considering, the new signatures shown in Figs. 8 and 9 can be naturally realized, and are quite appealing given the reach of the upcoming gravitational wave experiments.

VIII Conclusions and Outlook
It is truly extraordinary that gravitational wave astronomy can join forces with elementary particle physics to search for answers to fundamental questions about the structure of the Universe and its earliest stages of evolution. Indeed, processes happening at energies too high to be probed by conventional particle physics detectors (such as high scale leptogenesis and seesaw mechanism) can leave a remarkable imprint through the primordial gravitational wave background emitted soon after the Big Bang. Detecting such a signal would bring us closer to discovering which, if any, of the proposed Standard Model extensions addressing the outstanding questions about dark matter, matter-antimatter asymmetry, or neutrino masses, is realized in Nature.
A stochastic gravitational wave background is expected to originate in the early Universe within the framework of many particle physics models through first order phase transitions, cosmic string dynamics and domain wall annihilation. In particular, explaining the matter-antimatter asymmetry puzzle requires a first order phase transition to happen, indicating the huge importance of stochastic gravitational wave searches. The literature referred to in Section I contains analyzes of such signatures in theories beyond the Standard Model, however, the majority of the works focus on one single component at a time, generally not looking at the possible interplay between the contributions from different sources.
In this paper we highlighted the importance of searches for novel gravitational wave signatures arising when multiple components are present in the spectrum and add up producing new features in the signal. Such unique signatures are expected in theories with more than one symmetry breaking, and result from the interplay between the contributions from first order phase transitions, cosmic strings, and/or domain walls. The new gravitational wave signals we propose to look for are: Double-sharp-peak structure from domain walls produced when two gauge symmetries are broken by multiple scalars; Domain wall peak over a cosmic string plateau when one symmetry is broken by a single scalar and the other symmetry is broken by multiple scalars.
Although we demonstrate how those signatures arise in a specific model with gauged baryon and lepton number, our results are applicable to a much wider class of theories with two gauge symmetries broken at different energy scales. Indeed, the new signals consist of the cosmic string and domain wall contributions, thus they are fairly model-independent, since the cosmic string component depends only on the symmetry breaking scale, whereas the domain wall contribution depends on the symmetry breaking scale and the potential bias. Our results can also be extended to models with non-Abelian gauge groups. As already suggested in Fornal et al. (2023), it would be interesting to investigate the case when one of the symmetries is broken by two scalar triplets, as this can result in the production of cosmic strings Hindmarsh et al. (2016), and could perhaps lead to new signals involving contributions from all three processes: first order phase transitions, cosmic string dynamics, and domain wall annihilation.
The gravitational wave signatures discussed here can be searched for in upcoming experiments, including LISA, Big Bang Observer, DECIGO, Cosmic Explorer, and Einstein Telescope, enabling those detectors to probe the structure of high-scale symmetry breaking sectors. This is especially relevant for theories of leptogenesis such as the model we considered, in which, contrary to Duerr et al. (2013); Fornal and Shams Es Haghi (2020), the scale of symmetry breaking is not bounded from above and can also be high, allowing for signals and to be generated.
Finally, it is worth mentioning that a spontaneous breaking of a single gauge symmetry can by itself lead to gravitational wave signatures combining signals from a phase transition and cosmic strings, or a phase transition and domain walls. Given the sensitivity of the experiments considered, the symmetry breaking scale would have to be for the combined signal to be discoverable. The cosmic string contribution would then be detectable by Big Bang Observer and DECIGO, the phase transition peak could be seen by Cosmic Explorer and Einstein Telescope, and the domain wall peak would be visible in LISA. Investigating this in more detail is an interesting follow-up project, and could be tied to gravitational wave experiments sensitive to lower frequencies, such as the pulsar timing arrays: NANOGrav Arzoumanian et al. (2018), PPTA Manchester et al. (2013), EPTA Ferdman et al. (2010), IPTA Hobbs et al. (2010), or SKA Weltman et al. (2020).
Acknowledgments
This research was supported by the National Science Foundation under Grant No. PHY-2213144.
References
- Glashow (1961) S. L. Glashow, Partial Symmetries of Weak Interactions, Nucl. Phys. 22, 579–588 (1961).
- Higgs (1964) P. W. Higgs, Broken Symmetries and the Masses of Gauge Bosons, Phys. Rev. Lett. 13, 508–509 (1964).
- Englert and Brout (1964) F. Englert and R. Brout, Broken Symmetry and the Mass of Gauge Vector Mesons, Phys. Rev. Lett. 13, 321–323 (1964).
- Weinberg (1967) S. Weinberg, A Model of Leptons, Phys. Rev. Lett. 19, 1264–1266 (1967).
- Salam (1968) A. Salam, Weak and Electromagnetic Interactions, 8th Nobel Symposium Lerum, Sweden, May 19-25, 1968, Conf. Proc. C 680519, 367–377 (1968).
- Fritzsch et al. (1973) H. Fritzsch, M. Gell-Mann, and H. Leutwyler, Advantages of the Color Octet Gluon Picture, Phys. Lett. B 47, 365–368 (1973).
- Gross and Wilczek (1973) D. J. Gross and F. Wilczek, Ultraviolet Behavior of Nonabelian Gauge Theories, Phys. Rev. Lett. 30, 1343–1346 (1973).
- Politzer (1973) H. D. Politzer, Reliable Perturbative Results for Strong Interactions? Phys. Rev. Lett. 30, 1346–1349 (1973).
- Georgi and Glashow (1974) H. Georgi and S. L. Glashow, Unity of All Elementary Particle Forces, Phys. Rev. Lett. 32, 438–441 (1974).
- Fritzsch and Minkowski (1975) H. Fritzsch and P. Minkowski, Unified Interactions of Leptons and Hadrons, Annals Phys. 93, 193–266 (1975).
- Abe et al. (2017) K. Abe et al. (Super-Kamiokande), Search for Proton Decay via and in 0.31 Megaton·Years Exposure of the Super-Kamiokande Water Cherenkov Detector, Phys. Rev. D 95, 012004 (2017), arXiv:1610.03597 [hep-ex] .
- Kallosh et al. (1995) R. Kallosh, A. D. Linde, D. A. Linde, and L. Susskind, Gravity and Global Symmetries, Phys. Rev. D 52, 912–935 (1995), arXiv:hep-th/9502069 .
- Pais (1973) A. Pais, Remark on Baryon Conservation, Phys. Rev. D 8, 1844–1846 (1973).
- Rajpoot (1988) S. Rajpoot, Gauge Symmetries of Electroweak Interactions, Int. J. Theor. Phys. 27, 689 (1988).
- Foot et al. (1989) R. Foot, G. C. Joshi, and H. Lew, Gauged Baryon and Lepton Numbers, Phys. Rev. D 40, 2487–2489 (1989).
- Carone and Murayama (1995) C. D. Carone and H. Murayama, Realistic Models with a Light U(1) Gauge Boson Coupled to Baryon Number, Phys. Rev. D 52, 484–493 (1995), arXiv:hep-ph/9501220 .
- Georgi and Glashow (1996) H. Georgi and S. L. Glashow, Decays of a Leptophobic Gauge Boson, Phys. Lett. B 387, 341–345 (1996), arXiv:hep-ph/9607202 .
- Fileviez Perez and Wise (2010) P. Fileviez Perez and M. B. Wise, Baryon and Lepton Number as Local Gauge Symmetries, Phys. Rev. D 82, 011901 (2010), [Erratum: Phys.Rev.D 82, 079901 (2010)], arXiv:1002.1754 [hep-ph] .
- Duerr et al. (2013) M. Duerr, P. Fileviez Perez, and M. B. Wise, Gauge Theory for Baryon and Lepton Numbers with Leptoquarks, Phys. Rev. Lett. 110, 231801 (2013), arXiv:1304.0576 [hep-ph] .
- Fileviez Perez et al. (2014) P. Fileviez Perez, S. Ohmer, and H. H. Patel, Minimal Theory for Lepto-Baryons, Phys. Lett. B 735, 283–287 (2014), arXiv:1403.8029 [hep-ph] .
- Arnold et al. (2013) J. M. Arnold, P. Fileviez Perez, B. Fornal, and S. Spinner, B and L at the Supersymmetry Scale, Dark Matter, and R-Parity Violation, Phys. Rev. D 88, 115009 (2013), arXiv:1310.7052 [hep-ph] .
- Fornal et al. (2015) B. Fornal, A. Rajaraman, and T. M. P. Tait, Baryon Number as the Fourth Color, Phys. Rev. D 92, 055022 (2015), arXiv:1506.06131 [hep-ph] .
- Fornal and Tait (2016) B. Fornal and T. M. P. Tait, Dark Matter from Unification of Color and Baryon Number, Phys. Rev. D 93, 075010 (2016), arXiv:1511.07380 [hep-ph] .
- Fornal et al. (2017) B. Fornal, Y. Shirman, T. M. P. Tait, and J. Rittenhouse West, Asymmetric Dark Matter and Baryogenesis from , Phys. Rev. D 96, 035001 (2017), arXiv:1703.00199 [hep-ph] .
- Duerr and Fileviez Perez (2015) M. Duerr and P. Fileviez Perez, Theory for Baryon Number and Dark Matter at the LHC, Phys. Rev. D 91, 095001 (2015), arXiv:1409.8165 [hep-ph] .
- Ohmer and Patel (2015) S. Ohmer and H. H. Patel, Leptobaryons as Majorana Dark Matter, Phys. Rev. D 92, 055020 (2015), arXiv:1506.00954 [hep-ph] .
- Fileviez Perez et al. (2019) P. Fileviez Perez, E. Golias, R.-H. Li, and C. Murgui, Leptophobic Dark Matter and the Baryon Number Violation Scale, Phys. Rev. D 99, 035009 (2019), arXiv:1810.06646 [hep-ph] .
- Fileviez Perez et al. (2021) P. Fileviez Perez, C. Murgui, and A. D. Plascencia, Baryogenesis via Leptogenesis: Spontaneous B and L Violation, Phys. Rev. D 104, 055007 (2021), arXiv:2103.13397 [hep-ph] .
- 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] .
- Turner (1997) M. S. Turner, Detectability of Inflation Produced Gravitational Waves, Phys. Rev. D 55, R435–R439 (1997), arXiv:astro-ph/9607066 .
- Kosowsky et al. (1992) A. Kosowsky, M. S. Turner, and R. Watkins, Gravitational Radiation from Colliding Vacuum Bubbles, Phys. Rev. D 45, 4514–4535 (1992).
- Hiramatsu et al. (2010) T. Hiramatsu, M. Kawasaki, and K. Saikawa, Gravitational Waves from Collapsing Domain Walls, JCAP 05, 032 (2010), arXiv:1002.1555 [astro-ph.CO] .
- Vachaspati and Vilenkin (1985a) T. Vachaspati and A. Vilenkin, Gravitational Radiation from Cosmic Strings, Phys. Rev. D 31, 3052 (1985a).
- Sakellariadou (1990) M. Sakellariadou, Gravitational Waves Emitted from Infinite Strings, Phys. Rev. D 42, 354–360 (1990), [Erratum: Phys. Rev. D 43, 4150 (1991)].
- Amaro-Seoane et al. (2017) P. Amaro-Seoane et al. (LISA), Laser Interferometer Space Antenna, (2017), arXiv:1702.00786 [astro-ph.IM] .
- Reitze et al. (2019) D. Reitze et al., Cosmic Explorer: The U.S. Contribution to Gravitational-Wave Astronomy beyond LIGO, Bull. Am. Astron. Soc. 51, 035 (2019), arXiv:1907.04833 [astro-ph.IM] .
- Punturo et al. (2010) M. Punturo et al., The Einstein Telescope: A Third-Generation Gravitational Wave Observatory, Class. Quant. Grav. 27, 194002 (2010).
- Kawamura et al. (2011) S. Kawamura et al., The Japanese Space Gravitational Wave Antenna: DECIGO, Class. Quant. Grav. 28, 094011 (2011).
- Crowder and Cornish (2005) J. Crowder and N. J. Cornish, Beyond LISA: Exploring Future Gravitational Wave Missions, Phys. Rev. D 72, 083005 (2005), arXiv:gr-qc/0506015 .
- Kibble (1976) T. W. B. Kibble, Topology of Cosmic Domains and Strings, J. Phys. A 9, 1387–1398 (1976).
- Buchmuller et al. (2019) W. Buchmuller, V. Domcke, H. Murayama, and K. Schmitz, Probing the Scale of Grand Unification with Gravitational Waves, (2019), arXiv:1912.03695 [hep-ph] .
- King et al. (2020) S. F. King, S. Pascoli, J. Turner, and Y.-L. Zhou, Gravitational Waves and Proton Decay: Complementary Windows into GUTs, (2020), arXiv:2005.13549 [hep-ph] .
- Blanco-Pillado and Olum (2017) J. J. Blanco-Pillado and K. D. Olum, Stochastic Gravitational Wave Background from Smoothed Cosmic String Loops, Phys. Rev. D 96, 104046 (2017), arXiv:1709.02693 [astro-ph.CO] .
- Ringeval and Suyama (2017) C. Ringeval and T. Suyama, Stochastic Gravitational Waves from Cosmic String Loops in Scaling, JCAP 12, 027 (2017), arXiv:1709.03845 [astro-ph.CO] .
- Cui et al. (2018) Y. Cui, M. Lewicki, D. E. Morrissey, and J. D. Wells, Cosmic Archaeology with Gravitational Waves from Cosmic Strings, Phys. Rev. D 97, 123505 (2018), arXiv:1711.03104 [hep-ph] .
- Cui et al. (2019) Y. Cui, M. Lewicki, D. E. Morrissey, and J. D. Wells, Probing the Pre-BBN Universe with Gravitational Waves from Cosmic Strings, JHEP 01, 081 (2019), arXiv:1808.08968 [hep-ph] .
- Guedes et al. (2018) G. S. F. Guedes, P. P. Avelino, and L. Sousa, Signature of Inflation in the Stochastic Gravitational Wave Background Generated by Cosmic String Networks, Phys. Rev. D 98, 123505 (2018), arXiv:1809.10802 [astro-ph.CO] .
- Dror et al. (2020) J. A. Dror, T. Hiramatsu, K. Kohri, H. Murayama, and G. White, Testing the Seesaw Mechanism and Leptogenesis with Gravitational Waves, Phys. Rev. Lett. 124, 041804 (2020), arXiv:1908.03227 [hep-ph] .
- Zhou and Bian (2020) R. Zhou and L. Bian, Gravitational Waves from Cosmic Strings and First-Order Phase Transition, (2020), arXiv:2006.13872 [hep-ph] .
- Gouttenoire et al. (2020a) Y. Gouttenoire, G. Servant, and P. Simakachorn, BSM with Cosmic Strings: Heavy, up to EeV Mass, Unstable Particles, JCAP 07, 016 (2020a), arXiv:1912.03245 [hep-ph] .
- Fornal and Shams Es Haghi (2020) B. Fornal and B. Shams Es Haghi, Baryon and Lepton Number Violation from Gravitational Waves, Phys. Rev. D 102, 115037 (2020), arXiv:2008.05111 [hep-ph] .
- Gouttenoire et al. (2020b) Y. Gouttenoire, G. Servant, and P. Simakachorn, Beyond the Standard Models with Cosmic Strings, JCAP 07, 032 (2020b), arXiv:1912.02569 [hep-ph] .
- Abbott et al. (2021) R. Abbott et al. (LIGO Scientific, Virgo, KAGRA), Constraints on Cosmic Strings Using Data from the Third Advanced LIGO–Virgo Observing Run, Phys. Rev. Lett. 126, 241102 (2021), arXiv:2101.12248 [gr-qc] .
- Eto et al. (2018a) M. Eto, M. Kurachi, and M. Nitta, Constraints on Two Higgs Doublet Models from Domain Walls, Phys. Lett. B 785, 447–453 (2018a), arXiv:1803.04662 [hep-ph] .
- Eto et al. (2018b) M. Eto, M. Kurachi, and M. Nitta, Non-Abelian Strings and Domain Walls in Two Higgs Doublet Models, JHEP 08, 195 (2018b), arXiv:1805.07015 [hep-ph] .
- Chen et al. (2020) N. Chen, T. Li, Z. Teng, and Y. Wu, Collapsing Domain Walls in the Two-Higgs-Doublet Model and Deep Insights from the EDM, JHEP 10, 081 (2020), arXiv:2006.06913 [hep-ph] .
- Battye et al. (2020) R. A. Battye, A. Pilaftsis, and D. G. Viatic, Domain Wall Constraints on Two-Higgs-Doublet Models with Symmetry, Phys. Rev. D 102, 123536 (2020), arXiv:2010.09840 [hep-ph] .
- Kadota et al. (2015) K. Kadota, M. Kawasaki, and K. Saikawa, Gravitational Waves from Domain Walls in the Next-to-Minimal Supersymmetric Standard Model, JCAP 10, 041 (2015), arXiv:1503.06998 [hep-ph] .
- Craig et al. (2021) N. Craig, I. Garcia Garcia, G. Koszegi, and A. McCune, P Not PQ, JHEP 09, 130 (2021), arXiv:2012.13416 [hep-ph] .
- Blasi et al. (2023) S. Blasi, A. Mariotti, A. Rase, A. Sevrin, and K. Turbang, Friction on ALP Domain Walls and Gravitational Waves, JCAP 04, 008 (2023), arXiv:2210.14246 [hep-ph] .
- Dunsky et al. (2022) D. I. Dunsky, A. Ghoshal, H. Murayama, Y. Sakakihara, and G. White, GUTs, Hybrid Topological Defects, and Gravitational Waves, Phys. Rev. D 106, 075030 (2022), arXiv:2111.08750 [hep-ph] .
- Borah and Dasgupta (2022) D. Borah and A. Dasgupta, Probing Left-Right Symmetry via Gravitational Waves from Domain Walls, Phys. Rev. D 106, 035016 (2022), arXiv:2205.12220 [hep-ph] .
- Fornal et al. (2023) B. Fornal, K. Garcia, and E. Pierre, Testing Unification and Dark Matter with Gravitational Waves, (2023), arXiv:2305.12566 [hep-ph] .
- Barman et al. (2022) B. Barman, D. Borah, A. Dasgupta, and A. Ghoshal, Probing High Scale Dirac Leptogenesis via Gravitational Waves from Domain Walls, Phys. Rev. D 106, 015007 (2022), arXiv:2205.03422 [hep-ph] .
- Saikawa (2017) K. Saikawa, A Review of Gravitational Waves from Cosmic Domain Walls, Universe 3, 40 (2017), arXiv:1703.02576 [hep-ph] .
- Jiang and Huang (2022) Y. Jiang and Q.-G. Huang, Implications for Cosmic Domain Walls from the First Three Observing Runs of LIGO-Virgo, Phys. Rev. D 106, 103036 (2022), arXiv:2208.00697 [astro-ph.CO] .
- Grojean and Servant (2007) C. Grojean and G. Servant, Gravitational Waves from Phase Transitions at the Electroweak Scale and Beyond, Phys. Rev. D 75, 043507 (2007), arXiv:hep-ph/0607107 .
- Vaskonen (2017) V. Vaskonen, Electroweak Baryogenesis and Gravitational Waves from a Real Scalar Singlet, Phys. Rev. D 95, 123515 (2017), arXiv:1611.02073 [hep-ph] .
- Dorsch et al. (2017) G. C. Dorsch, S. J. Huber, T. Konstandin, and J. M. No, A Second Higgs Doublet in the Early Universe: Baryogenesis and Gravitational Waves, JCAP 05, 052 (2017), arXiv:1611.05874 [hep-ph] .
- Bernon et al. (2018) J. Bernon, L. Bian, and Y. Jiang, A New Insight into the Phase Transition in the Early Universe with Two Higgs Doublets, JHEP 05, 151 (2018), arXiv:1712.08430 [hep-ph] .
- Chala et al. (2018) M. Chala, C. Krause, and G. Nardini, Signals of the Electroweak Phase Transition at Colliders and Gravitational Wave Observatories, JHEP 07, 062 (2018), arXiv:1802.02168 [hep-ph] .
- Angelescu and Huang (2019) A. Angelescu and P. Huang, Multistep Strongly First Order Phase Transitions from New Fermions at the TeV Scale, Phys. Rev. D 99, 055023 (2019), arXiv:1812.08293 [hep-ph] .
- Alves et al. (2019) A. Alves, T. Ghosh, H.-K. Guo, K. Sinha, and D. Vagie, Collider and Gravitational Wave Complementarity in Exploring the Singlet Extension of the Standard Model, JHEP 04, 052 (2019), arXiv:1812.09333 [hep-ph] .
- Han et al. (2021) X.-F. Han, L. Wang, and Y. Zhang, Dark Matter, Electroweak Phase Transition, and Gravitational Waves in the Type II Two-Higgs-Doublet Model with a Singlet Scalar Field, Phys. Rev. D 103, 035012 (2021), arXiv:2010.03730 [hep-ph] .
- Benincasa et al. (2022) N. Benincasa, L. Delle Rose, K. Kannike, and L. Marzola, Multistep Phase Transitions and Gravitational Waves in the Inert Doublet Model, (2022), arXiv:2205.06669 [hep-ph] .
- Craig et al. (2020) N. Craig, N. Levi, A. Mariotti, and D. Redigolo, Ripples in Spacetime from Broken Supersymmetry, JHEP 21, 184 (2020), arXiv:2011.13949 [hep-ph] .
- Fornal et al. (2021) B. Fornal, B. Shams Es Haghi, J.-H. Yu, and Y. Zhao, Gravitational Waves from Mini-Split SUSY, Phys. Rev. D 104, 115005 (2021), arXiv:2104.00747 [hep-ph] .
- Dev et al. (2019) P. S. B. Dev, F. Ferrer, Y. Zhang, and Y. Zhang, Gravitational Waves from First-Order Phase Transition in a Simple Axion-Like Particle Model, JCAP 11, 006 (2019), arXiv:1905.00891 [hep-ph] .
- Von Harling et al. (2020) B. Von Harling, A. Pomarol, O. Pujolas, and F. Rompineve, Peccei-Quinn Phase Transition at LIGO, JHEP 04, 195 (2020), arXiv:1912.07587 [hep-ph] .
- Delle Rose et al. (2020) L. Delle Rose, G. Panico, M. Redi, and A. Tesi, Gravitational Waves from Supercool Axions, JHEP 04, 025 (2020), arXiv:1912.06139 [hep-ph] .
- Croon et al. (2019) D. Croon, T. E. Gonzalo, and G. White, Gravitational Waves from a Pati-Salam Phase Transition, JHEP 02, 083 (2019), arXiv:1812.02747 [hep-ph] .
- Huang et al. (2020) W.-C. Huang, F. Sannino, and Z.-W. Wang, Gravitational Waves from Pati-Salam Dynamics, Phys. Rev. D 102, 095025 (2020), arXiv:2004.02332 [hep-ph] .
- Okada et al. (2021) N. Okada, O. Seto, and H. Uchida, Gravitational Waves from Breaking of an Extra in Grand Unification, PTEP 2021, 033B01 (2021), arXiv:2006.01406 [hep-ph] .
- Hasegawa et al. (2019) T. Hasegawa, N. Okada, and O. Seto, Gravitational Waves from the Minimal Gauged Model, Phys. Rev. D 99, 095039 (2019), arXiv:1904.03020 [hep-ph] .
- Brdar et al. (2019) V. Brdar, A. J. Helmboldt, and J. Kubo, Gravitational Waves from First-Order Phase Transitions: LIGO as a Window to Unexplored Seesaw Scales, JCAP 02, 021 (2019), arXiv:1810.12306 [hep-ph] .
- Okada and Seto (2018) N. Okada and O. Seto, Probing the Seesaw Scale with Gravitational Waves, Phys. Rev. D 98, 063532 (2018), arXiv:1807.00336 [hep-ph] .
- Di Bari et al. (2021) P. Di Bari, D. Marfatia, and Y.-L. Zhou, Gravitational Waves from First-Order Phase Transitions in Majoron Models of Neutrino Mass, JHEP 10, 193 (2021), arXiv:2106.00025 [hep-ph] .
- Zhou et al. (2022) R. Zhou, L. Bian, and Y. Du, Electroweak Phase Transition and Gravitational Waves in the Type-II Seesaw Model, JHEP 08, 205 (2022), arXiv:2203.01561 [hep-ph] .
- Greljo et al. (2020) A. Greljo, T. Opferkuch, and B. A. Stefanek, Gravitational Imprints of Flavor Hierarchies, Phys. Rev. Lett. 124, 171802 (2020), arXiv:1910.02014 [hep-ph] .
- Fornal (2021) B. Fornal, Gravitational Wave Signatures of Lepton Universality Violation, Phys. Rev. D 103, 015018 (2021), arXiv:2006.08802 [hep-ph] .
- Schwaller (2015) P. Schwaller, Gravitational Waves from a Dark Phase Transition, Phys. Rev. Lett. 115, 181101 (2015), arXiv:1504.07263 [hep-ph] .
- Breitbach et al. (2019) M. Breitbach, J. Kopp, E. Madge, T. Opferkuch, and P. Schwaller, Dark, Cold, and Noisy: Constraining Secluded Hidden Sectors with Gravitational Waves, JCAP 07, 007 (2019), arXiv:1811.11175 [hep-ph] .
- Croon et al. (2018) D. Croon, V. Sanz, and G. White, Model Discrimination in Gravitational Wave spectra from Dark Phase Transitions, JHEP 08, 203 (2018), arXiv:1806.02332 [hep-ph] .
- Hall et al. (2020) E. Hall, T. Konstandin, R. McGehee, H. Murayama, and G. Servant, Baryogenesis From a Dark First-Order Phase Transition, JHEP 04, 042 (2020), arXiv:1910.08068 [hep-ph] .
- Ellis et al. (2020a) J. Ellis, M. Lewicki, and V. Vaskonen, Updated Predictions for Gravitational Waves Produced in a Strongly Supercooled Phase Transition, JCAP 11, 020 (2020a), arXiv:2007.15586 [astro-ph.CO] .
- Kawana (2022) K. Kawana, Cosmology of a Supercooled Universe, Phys. Rev. D 105, 103515 (2022), arXiv:2201.00560 [hep-ph] .
- Baldes (2017) I. Baldes, Gravitational Waves from the Asymmetric-Dark-Matter Generating Phase Transition, JCAP 05, 028 (2017), arXiv:1702.02117 [hep-ph] .
- Azatov et al. (2021) A. Azatov, M. Vanvlasselaer, and W. Yin, Dark Matter Production from Relativistic Bubble Walls, JHEP 03, 288 (2021), arXiv:2101.05721 [hep-ph] .
- Costa et al. (2022a) F. Costa, S. Khan, and J. Kim, A Two-Component Dark Matter Model and its Associated Gravitational Waves, JHEP 06, 026 (2022a), arXiv:2202.13126 [hep-ph] .
- Costa et al. (2022b) F. Costa, S. Khan, and J. Kim, A Two-Component Vector WIMP – Fermion FIMP Dark Matter Model with an Extended Seesaw Mechanism, JHEP 12, 165 (2022b), arXiv:2209.13653 [hep-ph] .
- Fornal and Pierre (2022) B. Fornal and E. Pierre, Asymmetric Dark Matter from Gravitational Waves, Phys. Rev. D 106, 115040 (2022), arXiv:2209.04788 [hep-ph] .
- Kierkla et al. (2023) M. Kierkla, A. Karam, and B. Swiezewska, Conformal Model for Gravitational Waves and Dark Matter: A Status Update, JHEP 03, 007 (2023), arXiv:2210.07075 [astro-ph.CO] .
- Azatov et al. (2022) A. Azatov, G. Barni, S. Chakraborty, M. Vanvlasselaer, and W. Yin, Ultra-Relativistic Bubbles from the Simplest Higgs Portal and their Cosmological Consequences, JHEP 10, 017 (2022), arXiv:2207.02230 [hep-ph] .
- Caldwell et al. (2022) R. Caldwell et al., Detection of Early-Universe Gravitational-Wave Signatures and Fundamental Physics, Gen. Rel. Grav. 54, 156 (2022), arXiv:2203.07972 [gr-qc] .
- Athron et al. (2023) P. Athron, C. Balazs, A. Fowlie, L. Morris, and L. Wu, Cosmological Phase Transitions: From Perturbative Particle Physics to Gravitational Waves, (2023), arXiv:2305.02357 [hep-ph] .
- Badger et al. (2023) C. Badger, B. Fornal, K. Martinovic, A. Romero, K. Turbang, H. Guo, A. Mariotti, M. Sakellariadou, A. Sevrin, F.-W. Yang, and Y. Zhao, Probing Early Universe Supercooled Phase Transitions with Gravitational Wave Data, Phys. Rev. D 107, 023511 (2023), arXiv:2209.14707 [hep-ph] .
- 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 (2019), arXiv:1903.09642 [hep-ph] .
- Lewicki and Vaskonen (2020a) M. Lewicki and V. Vaskonen, On Bubble Collisions in Strongly Supercooled Phase Transitions, Phys. Dark Univ. 30, 100672 (2020a), arXiv:1912.00997 [astro-ph.CO] .
- Lewicki and Vaskonen (2020b) M. Lewicki and V. Vaskonen, Gravitational Wave Spectra from Strongly Supercooled Phase Transitions, Eur. Phys. J. C 80, 1003 (2020b), arXiv:2007.04967 [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] .
- Davidson et al. (2008) S. Davidson, E. Nardi, and Y. Nir, Leptogenesis, Phys. Rept. 466, 105–177 (2008), arXiv:0802.2962 [hep-ph] .
- Buchmuller et al. (2005) W. Buchmuller, P. Di Bari, and M. Plumacher, Leptogenesis for Pedestrians, Annals Phys. 315, 305–351 (2005), arXiv:hep-ph/0401240 .
- Workman et al. (2022) R. L. Workman et al. (Particle Data Group), Review of Particle Physics, PTEP 2022, 083C01 (2022).
- Vilenkin and Shellard (2000) A. Vilenkin and E. P. S. Shellard, Cosmic Strings and Other Topological Defects, Cambridge University Press (2000).
- Ade et al. (2014) P. A. R. Ade et al. (Planck), Planck 2013 Results. XXV. Searches for Cosmic Strings and Other Topological Defects, Astron. Astrophys. 571, A25 (2014), arXiv:1303.5085 [astro-ph.CO] .
- Olum and Blanco-Pillado (2000) K. D. Olum and J. J. Blanco-Pillado, Radiation from Cosmic String Standing Waves, Phys. Rev. Lett. 84, 4288–4291 (2000), arXiv:astro-ph/9910354 .
- Moore et al. (2002) J. N. Moore, E. P. S. Shellard, and C. J. A. P. Martins, On the Evolution of Abelian-Higgs String Networks, Phys. Rev. D 65, 023503 (2002), arXiv:hep-ph/0107171 .
- Kibble (1985) T. W. B. Kibble, Evolution of a System of Cosmic Strings, Nucl. Phys. B 252, 227 (1985), [Erratum: Nucl.Phys.B 261, 750 (1985)].
- Bennett and Bouchet (1988) D. P. Bennett and F. R. Bouchet, Evidence for a Scaling Solution in Cosmic-String Evolution, Phys. Rev. Lett. 60, 257–260 (1988).
- Bennett and Bouchet (1989) D. P. Bennett and F. R. Bouchet, Cosmic-String Evolution, Phys. Rev. Lett. 63, 2776–2779 (1989).
- Albrecht and Turok (1989) A. Albrecht and N. Turok, Evolution of Cosmic String Networks, Phys. Rev. D 40, 973–1001 (1989).
- Allen and Shellard (1990) B. Allen and E. P. S. Shellard, Cosmic-String Evolution: A Numerical Simulation, Phys. Rev. Lett. 64, 119–122 (1990).
- Hindmarsh and Kibble (1995) M. B. Hindmarsh and T. W. B. Kibble, Cosmic Strings, Rept. Prog. Phys. 58, 477–562 (1995), arXiv:hep-ph/9411342 .
- Blanco-Pillado et al. (2014) J. J. Blanco-Pillado, K. D. Olum, and B. Shlaer, The Number of Cosmic String Loops, Phys. Rev. D 89, 023512 (2014), arXiv:1309.6637 [astro-ph.CO] .
- Vachaspati and Vilenkin (1985b) T. Vachaspati and A. Vilenkin, Gravitational Radiation from Cosmic Strings, Phys. Rev. D 31, 3052–3058 (1985b).
- Martins and Shellard (1996a) C. J. A. P. Martins and E. P. S. Shellard, String Evolution with Friction, Phys. Rev. D 53, 575–579 (1996a), arXiv:hep-ph/9507335 .
- Martins and Shellard (1996b) C. J. A. P. Martins and E. P. S. Shellard, Quantitative String Evolution, Phys. Rev. D 54, 2535–2556 (1996b), arXiv:hep-ph/9602271 .
- Martins and Shellard (2002) C. J. A. P. Martins and E. P. S. Shellard, Extending the Velocity Dependent One Scale String Evolution Model, Phys. Rev. D 65, 043514 (2002), arXiv:hep-ph/0003298 .
- Ginzburg and Krawczyk (2005) I. F. Ginzburg and M. Krawczyk, Symmetries of Two Higgs Doublet Model and CP Violation, Phys. Rev. D 72, 115013 (2005), arXiv:hep-ph/0408011 .
- Battye et al. (2011) R. A. Battye, G. D. Brawn, and A. Pilaftsis, Vacuum Topology of the Two Higgs Doublet Model, JHEP 08, 020 (2011), arXiv:1106.3482 [hep-ph] .
- Hiramatsu et al. (2014) T. Hiramatsu, M. Kawasaki, and K. Saikawa, On the Estimation of Gravitational Wave Spectrum from Cosmic Domain Walls, JCAP 02, 031 (2014), arXiv:1309.5001 [astro-ph.CO] .
- Clarke et al. (2020) T. J. Clarke, E. J. Copeland, and A. Moss, Constraints on Primordial Gravitational Waves from the Cosmic Microwave Background, JCAP 10, 002 (2020), arXiv:2004.11396 [astro-ph.CO] .
- Linde (1983) A. D. Linde, Decay of the False Vacuum at Finite Temperature, Nuclear Physics B 216, 421 – 445 (1983).
- Espinosa et al. (2010) J. R. Espinosa, T. Konstandin, J. M. No, and G. Servant, Energy Budget of Cosmological First-Order Phase Transitions, JCAP 06, 028 (2010), arXiv:1004.4187 [hep-ph] .
- Caprini et al. (2016) C. Caprini et al., Science with the Space-Based Interferometer eLISA. II: Gravitational Waves from Cosmological Phase Transitions, JCAP 04, 001 (2016), arXiv:1512.06239 [astro-ph.CO] .
- Hindmarsh et al. (2014) M. Hindmarsh, S. J. Huber, K. Rummukainen, and D. J. Weir, Gravitational Waves from the Sound of a First Order Phase Transition, Phys. Rev. Lett. 112, 041301 (2014), arXiv:1304.2433 [hep-ph] .
- Ellis et al. (2020b) J. Ellis, M. Lewicki, and J. M. No, Gravitational Waves from First-Order Cosmological Phase Transitions: Lifetime of the Sound Wave Source, JCAP 07, 050 (2020b), arXiv:2003.07360 [hep-ph] .
- Guo et al. (2021) H.-K. Guo, K. Sinha, D. Vagie, and G. White, Phase Transitions in an Expanding Universe: Stochastic Gravitational Waves in Standard and Non-Standard Histories, JCAP 01, 001 (2021), arXiv:2007.08537 [hep-ph] .
- Huber and Konstandin (2008) S. J. Huber and T. Konstandin, Gravitational Wave Production by Collisions: More Bubbles, JCAP 09, 022 (2008), arXiv:0806.1828 [hep-ph] .
- Lewicki and Vaskonen (2021) M. Lewicki and V. Vaskonen, Gravitational Waves from Colliding Vacuum Bubbles in Gauge Theories, Eur. Phys. J. C 81, 437 (2021), arXiv:2012.07826 [astro-ph.CO] .
- Kamionkowski et al. (1994) M. Kamionkowski, A. Kosowsky, and M. S. Turner, Gravitational Radiation from First Order Phase Transitions, Phys. Rev. D 49, 2837–2851 (1994), arXiv:astro-ph/9310044 .
- Caprini and Durrer (2006) C. Caprini and R. Durrer, Gravitational Waves from Stochastic Relativistic Sources: Primordial Turbulence and Magnetic Fields, Phys. Rev. D 74, 063521 (2006), arXiv:astro-ph/0603476 .
- Caprini et al. (2009) C. Caprini, R. Durrer, and G. Servant, The Stochastic Gravitational Wave Background from Turbulence and Magnetic Fields Generated by a First-Order Phase Transition, JCAP 12, 024 (2009), arXiv:0909.0622 [astro-ph.CO] .
- Ferrer et al. (2023) F. Ferrer, A. Ghoshal, and M. Lewicki, Imprints of a Supercooled Universe in the Gravitational Wave Spectrum from a Cosmic String Network, (2023), arXiv:2304.02636 [astro-ph.CO] .
- Hindmarsh et al. (2016) M. Hindmarsh, K. Rummukainen, and D. J. Weir, New Solutions for Non-Abelian Cosmic Strings, Phys. Rev. Lett. 117, 251601 (2016), arXiv:1607.00764 [hep-th] .
- Arzoumanian et al. (2018) Z. Arzoumanian et al. (NANOGRAV), The NANOGrav 11-Year Data Set: Pulsar-Timing Constraints on the Stochastic Gravitational-Wave Background, Astrophys. J. 859, 47 (2018), arXiv:1801.02617 [astro-ph.HE] .
- Manchester et al. (2013) R. N. Manchester et al., The Parkes Pulsar Timing Array Project, Publ. Astron. Soc. Austral. 30, e017 (2013), arXiv:1210.6130 [astro-ph.IM] .
- Ferdman et al. (2010) R. D. Ferdman et al., The European Pulsar Timing Array: Current Efforts and a LEAP Toward the Future, Classical and Quantum Gravity 27, 084014 (2010), arXiv:1003.3405 [astro-ph.HE] .
- Hobbs et al. (2010) G. Hobbs et al., The International Pulsar Timing Array Project: Using Pulsars as a Gravitational Wave Detector, Classical and Quantum Gravity 27, 084013 (2010), arXiv:0911.5206 [astro-ph.SR] .
- Weltman et al. (2020) A. Weltman et al., Fundamental Physics with the Square Kilometre Array, Publ. Astron. Soc. Austral. 37, e002 (2020), arXiv:1810.02680 [astro-ph.CO] .