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

Gravitational Wave Signatures of Gauged Baryon and Lepton Number

Jessica Bosch Department of Chemistry and Physics, Barry University, Miami Shores, Florida 33161, USA    Zoraida Delgado Department of Chemistry and Physics, Barry University, Miami Shores, Florida 33161, USA    Bartosz Fornal Department of Chemistry and Physics, Barry University, Miami Shores, Florida 33161, USA    Alejandra Leon Department of Chemistry and Physics, Barry University, Miami Shores, Florida 33161, USA
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 U(1)\rm U(1) 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 U(1){\rm U}(1) 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 U(1){\rm U}(1) 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 U(1){\rm U}(1) 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 U(1){\rm U}(1) 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 Z2Z_{2} 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:(1)(1) Two coexisting signals from domain wall annihilation, forming a characteristic sharp double-peak in the spectrum; (2)(2) 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 (1)(1) and (2)(2), 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

SU(3)c×SU(2)L×U(1)Y×U(1)B×U(1)L.\displaystyle{\rm SU}(3)_{c}\times{\rm SU}(2)_{L}\times{\rm U}(1)_{Y}\times{\rm U}(1)_{B}\times{\rm U}(1)_{L}\ . (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 U(1)L{\rm U}(1)_{L} symmetry comes in the representation

ΦLi=(1,1,0,0,2),\displaystyle\Phi_{Li}=(1,1,0,0,2)\ , (2)

while each of the scalars breaking U(1)B{\rm U}(1)_{B} is

ΦBi=(1,1,0,3,3).\displaystyle\Phi_{Bi}=(1,1,0,3,3)\ . (3)

Within this framework, there are four possible cases:

  • (a)(a)

    ΦB\Phi_{B} breaks U(1)B{\rm U(1)}_{B} and ΦL\Phi_{L} breaks U(1)L{\rm U(1)}_{L},

  • (b)(b)

    ΦB1\Phi_{B1}, ΦB2\Phi_{B2} break U(1)B{\rm U(1)}_{B} and ΦL\Phi_{L} breaks U(1)L{\rm U(1)}_{L},

  • (c)(c)

    ΦB\Phi_{B} breaks U(1)B{\rm U(1)}_{B} and ΦL1\Phi_{L1}, ΦL2\Phi_{L2} break U(1)L{\rm U(1)}_{L},

  • (d)(d)

    ΦB1\Phi_{B1}, ΦB2\Phi_{B2} break U(1)B{\rm U(1)}_{B} and ΦL1\Phi_{L1}, ΦL2\Phi_{L2} break U(1)L{\rm U(1)}_{L}.

We assume that the mixed terms involving scalars breaking different U(1){\rm U}(1) gauge groups have negligible coefficients. This implies that for a given U(1){\rm U}(1), if only one scalar breaks the symmetry, the scalar potential is

V(Φ)\displaystyle V(\Phi) =\displaystyle= m2|Φ|2+λ|Φ|4,\displaystyle-m^{2}|\Phi|^{2}+\lambda|\Phi|^{4}\ , (4)

whereas if two scalars participate in symmetry breaking,

V(Φ1,Φ2)\displaystyle V(\Phi_{1},\Phi_{2}) =\displaystyle= m12|Φ1|2+λ1|Φ1|4m22|Φ2|2+λ2|Φ2|4\displaystyle-\,m_{1}^{2}|\Phi_{1}|^{2}+\lambda_{1}|\Phi_{1}|^{4}-m_{2}^{2}|\Phi_{2}|^{2}+\lambda_{2}|\Phi_{2}|^{4} (5)
+\displaystyle+ [(λ4|Φ1|2+λ5|Φ2|2+λ6Φ1Φ2)Φ1Φ2+h.c.]\displaystyle\left[(\lambda_{4}|\Phi_{1}|^{2}\!+\!\lambda_{5}|\Phi_{2}|^{2}\!+\!\lambda_{6}\Phi_{1}^{*}\Phi_{2})\Phi_{1}^{*}\Phi_{2}+{\rm h.c.}\right]
\displaystyle- (m122Φ1Φ2+h.c.)+λ3|Φ1|2|Φ2|2.\displaystyle(m_{12}^{2}\Phi_{1}^{*}\Phi_{2}+{\rm h.c.})+\lambda_{3}|\Phi_{1}|^{2}|\Phi_{2}|^{2}\ .

The scalars develop the following vacuum expectation values,

Φi=vi2.\displaystyle\langle\Phi_{i}\rangle=\frac{v_{i}}{\sqrt{2}}\ . (6)

Whenever two scalars take part in the symmetry breaking, we define vv12+v22.v\equiv\sqrt{v_{1}^{2}+v_{2}^{2}}\ . This way one can collectively describe the U(1)B{\rm U}(1)_{B} breaking scale as v=vBv=v_{B}, and the U(1)L{\rm U}(1)_{L} breaking scale as v=vLv=v_{L}, 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:

SU(3)c×SU(2)L×U(1)Y×U(1)B×U(1)LΦLi 0SU(3)c×SU(2)L×U(1)Y×U(1)BΦBi 0SU(3)c×SU(2)L×U(1)Y,\begin{array}[]{c}{\rm SU}(3)_{c}\times{\rm SU}(2)_{L}\times{\rm U}(1)_{Y}\times{\rm U}(1)_{B}\times{\rm U}(1)_{L}\\[3.0pt] \hskip 31.86707pt\bigg{\downarrow}\hskip 8.53581pt{\scriptstyle\langle\Phi_{Li}\rangle\ \neq\ 0}\\[12.0pt] {\rm SU}(3)_{c}\times{\rm SU}(2)_{L}\times{\rm U}(1)_{Y}\times{\rm U}(1)_{B}\\[3.0pt] \hskip 31.86707pt\bigg{\downarrow}\hskip 8.53581pt{\scriptstyle\langle\Phi_{Bi}\rangle\ \neq\ 0}\\[12.0pt] {\rm SU}(3)_{c}\times{\rm SU}(2)_{L}\times{\rm U}(1)_{Y}\ ,\\[7.0pt] \end{array}

followed by the usual electroweak symmetry breaking by the Standard Model Higgs. We note that the order of U(1)B{\rm U}(1)_{B} and U(1)L{\rm U}(1)_{L} breaking may be reversed.

Fermion sector

To provide a concrete quantitative example, we consider the model with gauged U(1)B{\rm U}(1)_{B} and U(1)L{\rm U}(1)_{L} 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 QLjQ_{L}^{j}, uRju_{R}^{j}, dRjd_{R}^{j} and lepton fields lLjl_{L}^{j}, eRje_{R}^{j} are augmented by

νRj\displaystyle\nu_{R}^{j} =\displaystyle= (1,1,0,0,1),\displaystyle(1,1,0,0,1)\ ,
ΨL\displaystyle\Psi_{L} =\displaystyle= (ψL+ψL0)=(1,2,12,32,32),\displaystyle\begin{pmatrix}\psi_{L}^{+}\\[2.0pt] \psi_{L}^{0}\end{pmatrix}=\left(1,2,\tfrac{1}{2},\tfrac{3}{2},\tfrac{3}{2}\right)\ ,
ΨR\displaystyle\Psi_{R} =\displaystyle= (ψR0ψR)=(1,2,12,32,32),\displaystyle\begin{pmatrix}\psi_{R}^{0}\\ \psi_{R}^{-}\end{pmatrix}=\left(1,2,\tfrac{1}{2},-\tfrac{3}{2},-\tfrac{3}{2}\right)\ ,
ΣL\displaystyle\Sigma_{L} =\displaystyle= 12(σ02σ+2σσ0)=(1,3,0,32,32),\displaystyle\tfrac{1}{2}\begin{pmatrix}\sigma^{0}&\sqrt{2}\,\sigma^{+}\\[2.0pt] \sqrt{2}\,\sigma^{-}&-\sigma^{0}\end{pmatrix}=\left(1,3,0,-\tfrac{3}{2},-\tfrac{3}{2}\right)\ ,
χL\displaystyle\chi_{L} =\displaystyle= (1,1,0,32,32),\displaystyle\left(1,1,0,-\tfrac{3}{2},-\tfrac{3}{2}\right)\ , (7)

where jj is the family index. Among the fields above, νRj\nu_{R}^{j} are the right-handed neutrinos, whereas χL\chi_{L} is a Majorana dark matter candidate discussed in Section III.

Particle masses

The scalars ΦBi\Phi_{Bi} and ΦLi\Phi_{Li} generate masses for the new fermions through the following Lagrangian terms,

\displaystyle-\mathcal{L} \displaystyle\supset i(YΨiΨ¯RΨLΦBi+YΣiTr(ΣL2)ΦBi+YχiχLχLΦBi)\displaystyle\sum_{i}\!\left(Y_{\Psi}^{i}\overline{\Psi}_{R}\Psi_{L}\Phi_{Bi}^{*}\!+\!Y_{\Sigma}^{i}{\rm Tr}(\Sigma_{L}^{2})\Phi_{Bi}\!+\!Y_{\chi}^{i}\chi_{L}\chi_{L}\Phi_{Bi}\right) (8)
+\displaystyle+ yνjl¯LjHνRj+iYνijνRjνRjΦLi+h.c.,\displaystyle y_{\nu}^{j}\,\bar{l}_{L}^{j}H\nu_{R}^{j}+\sum_{i}Y_{\nu}^{ij}\nu_{R}^{j}\nu_{R}^{j}\Phi_{Li}+{\rm h.c.}\ ,\ \ \ \

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 yνj1y_{\nu}^{j}\sim 1 and Yνj102Y_{\nu}^{j}\sim 10^{-2}, the measured neutrino mass splittings are reproduced if vL105PeVv_{L}\sim 10^{5}\ {\rm PeV}. The mass matrices for the new fermions are provided in Fileviez Perez et al. (2014).

The spontaneous breaking of U(1)L{\rm U}(1)_{L} and U(1)B{\rm U}(1)_{B} leads to the appearance of vector gauge bosons ZLZ_{L} and ZBZ_{B}. Given the charges of the scalars breaking the two symmetries, the corresponding masses are

mZL=2gLvL,mZB=3gBvB,\displaystyle m_{Z_{L}}=2g_{L}v_{L}\ ,\ \ \ \ \ \ \ \ m_{Z_{B}}=3g_{B}v_{B}\ , (9)

where gLg_{L} and gBg_{B} are the U(1)L{\rm U}(1)_{L} and U(1)B{\rm U}(1)_{B} 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 U(1)L{\rm U}(1)_{L} breaking, the model also contains a phenomenologically viable dark matter candidate χL\chi_{L} 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 U(1)B{\rm U}(1)_{B} breaking, there remains a residual discrete Z2{Z}_{2} symmetry under which the new fermions transform as

ΨL\displaystyle\Psi_{L} \displaystyle\to ΨL,Ψ¯RΨ¯R,\displaystyle-\Psi_{L}\ ,\ \ \ \ \overline{\Psi}_{R}\to-\overline{\Psi}_{R}\ ,
Σ\displaystyle\Sigma \displaystyle\to Σ,χLχL.\displaystyle-\Sigma\ ,\ \ \ \ \ \ \ \chi_{L}\to-\chi_{L}\ . (10)

If χL\chi_{L} 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 h2ΩDM0.12h^{2}\Omega_{\rm DM}\approx 0.12 Aghanim et al. (2020) imposes an upper bound on the U(1)B{\rm U(1)}_{B} breaking scale. In particular, if the dark matter annihilation happens via the resonant ss-channel process

χLχLZBq¯q,\displaystyle\chi_{L}\,\chi_{L}\to Z_{B}^{*}\to\bar{q}\,q\ , (11)

a dependence between the parameters vBv_{B}, YχY_{\chi}, and gBg_{B} arises, and the perturbativity requirement leads to gBvB20TeVg_{B}v_{B}\lesssim 20\ {\rm TeV}. 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 U(1)B{\rm U}(1)_{B} 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 tt-channel process

χLχLΦBΦB,\displaystyle\chi_{L}\,\chi_{L}\to\Phi_{B}\,\Phi_{B}\ , (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 U(1)B{\rm U}(1)_{B} breaking can be high. Alternatively, the aforementioned bound on the U(1)B{\rm U}(1)_{B} 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 U(1)B{\rm U}(1)_{B} and U(1)L{\rm U}(1)_{L} 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 U(1)L{\rm U}(1)_{L} 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,

N1HlL.\displaystyle N_{1}\to H\,l_{L}\ . (13)

The CPCP asymmetry is introduced through the standard interference between the tree-level diagram for the process in Eq. (13) and the one-loop diagrams involving HH, lLl_{L}, and the two heavier right-handed neutrinos N2N_{2}, N3N_{3} in the loop.

The generated lepton asymmetry is calculated by solving the Boltzmann equations for the evolution of the lightest right-handed neutrino abundance YN1=nN1/sY_{N_{1}}=n_{N_{1}}/s (where nN1n_{N_{1}} is the N1N_{1} particle density and ss is the co-moving entropy density) and the BLB\!-\!L asymmetry YBLY_{B-L} Buchmuller et al. (2005),

dYN1dz\displaystyle\frac{dY_{N_{1}}}{dz} =\displaystyle= (D+S)(YN1YN1eq),\displaystyle-(D+S)(Y_{N_{1}}-Y_{N_{1}}^{\rm eq})\ ,
dYBLdz\displaystyle\frac{dY_{B-L}}{dz} =\displaystyle= ϵ1D(YN1YN1eq)WYBL,\displaystyle-\epsilon_{1}D(Y_{N_{1}}-Y_{N_{1}}^{\rm eq})-WY_{B-L}\ , (14)

where z=mN1/Tz=m_{N_{1}}/T, the term DD accounts for decays and inverse decays, SS represents ΔL=1\Delta L=1 scatterings, WW describes the washout effects, and ϵ1\epsilon_{1} is the CPCP 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 ZLZ_{L} gauge boson. Those equations were solved in Fileviez Perez et al. (2021), and the amount of the generated lepton asymmetry ΔL\Delta L 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 U(1)B{\rm U}(1)_{B} breaking the sphaleron-induced interactions have the form

(QQQL)3Ψ¯RΨLΣL4,\displaystyle(QQQL)^{3}\overline{\Psi}_{R}\Psi_{L}\Sigma_{L}^{4}\ , (15)

and it was shown in Duerr and Fileviez Perez (2015) that if the breaking of U(1)B{\rm U}(1)_{B} occurs close to the electroweak scale, then the final baryon asymmetry predicted by the model is given by

|ΔB|=3299|ΔL|.\displaystyle|\Delta B|=\frac{32}{99}|\Delta L|\ . (16)

This can explain the observed baryon-to-photon ratio Workman et al. (2022)

η6×1010\displaystyle\eta\approx 6\times 10^{-10} (17)

if the scale of U(1)L{\rm U}(1)_{L} breaking satisfies the relation

vL4000PeV.\displaystyle v_{L}\gtrsim 4000\ {\rm PeV}\ . (18)

The requirement for such a high U(1)L{\rm U}(1)_{L} symmetry breaking scale provides the desired setting accommodating the type I seesaw mechanism for the neutrinos.

IV Cosmic string spectrum

Spontaneous breaking of a U(1){\rm U}(1) 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 μ\mu, equal to the energy stored in a string per unit length, and depends solely on the scale at which the U(1){\rm U}(1) gauge symmetry is broken Vilenkin and Shellard (2000); Gouttenoire et al. (2020b),

Gμ=2π(vMPl)2,\displaystyle G\mu=2\pi\left(\frac{v}{M_{Pl}}\right)^{2}, (19)

where MPl=1.22×1013PeVM_{Pl}=1.22\times 10^{13}\ {\rm PeV} is the Planck mass, the gravitational constant G=6.7×1039GeV2G=6.7\times 10^{-39}\ \rm GeV^{-2}, 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 Gμ107G\mu\lesssim 10^{-7} Ade et al. (2014), which corresponds to the following bound on the scale of symmetry breaking,

v1.5×109PeV.\displaystyle v\lesssim 1.5\times 10^{9}\ {\rm PeV}\ . (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 t0t_{0} with initial length l(t0)=αt0l(t_{0})=\alpha\,t_{0}, where α\alpha is a constant loop size parameter. The loop oscillates emitting gravitational waves with frequencies

f~=2kl,\displaystyle\tilde{f}=\frac{2k}{l}\ , (21)

where kk is a positive integer. Rescaling this result by the scale factor a(t)a(t), one obtains the currently observed frequency,

f=a(te)a(T)f~,\displaystyle f=\frac{a(t_{e})}{a(T)}\,\tilde{f}\ , (22)

where tet_{e} is the time of the emission and TT 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)

P(k,n)=ΓGμ2kn(p=11pn)1,\displaystyle P_{(k,n)}=\frac{\Gamma G\mu^{2}}{k^{n}}\bigg{(}\sum_{p=1}^{\infty}\frac{1}{p^{n}}\!\bigg{)}^{-1}, (23)

where for the contribution from cusps n=4/3n=4/3, from kinks n=5/3n=5/3, and from kink-kink collisions n=2n=2, while the overall factor Γ50\Gamma\simeq 50 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

l(te)=αt0ΓGμ(tet0),\displaystyle l(t_{e})=\alpha\,t_{0}-\Gamma G\mu\,(t_{e}-t_{0})\ , (24)

causing the loop to vanish after the time αt0/(ΓGμ){\alpha\,t_{0}}/{(\Gamma G\mu)}.

The only model-dependent quantity describing the cosmic string network is the loop distribution function F(l,t0)F(l,t_{0}) 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,

F(l,t0)=2Ceffαt04δ(lαt0),\displaystyle F(l,t_{0})=\frac{\sqrt{2}\,{C}_{\rm eff}}{\alpha\,t_{0}^{4}}\,\delta(l-\alpha\,t_{0})\ , (25)

where the constant Ceff{C}_{\rm eff} depends on the era in the evolution of the Universe (for radiation Ceff=5.4C_{\rm eff}=5.4, whereas for the matter dominated era Ceff=0.39C_{\rm eff}=0.39 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)

h2ΩCS(f)\displaystyle h^{2}\Omega_{\rm CS}(f) =\displaystyle= 2h2αρcα2fk,nkP(k,n)tFT𝑑teCeff(t0k)t0k 4\displaystyle\frac{2h^{2}\mathcal{F}_{\alpha}}{\rho_{c}\alpha^{2}f}\,\sum_{k,n}\,{k\,P_{(k,n)}}\int_{t_{F}}^{T}\!dt_{e}\ \frac{C_{\text{eff}}(t_{0k})}{t_{0k}^{\,4}}\ \ \ \ (26)
×\displaystyle\times (a(te)a(T))5(a(t0k)a(te))3θ(t0ktF),\displaystyle\left(\frac{a(t_{e})}{a(T)}\right)^{\!5}\left(\frac{a(t_{0k})}{a(t_{e})}\right)^{\!3}\theta(t_{0k}-t_{F})\ ,

where

t0k=1α(2kfa(te)a(T)+ΓGμte).\displaystyle t_{0k}=\frac{1}{\alpha}\left(\frac{2k}{f}\frac{a(t_{e})}{a(T)}+\Gamma G\mu\,t_{e}\right)\ . (27)

In Eqs. (26) and (27) the parameters are the following: α\mathcal{F}_{\alpha} is the fraction of the loops contributing to the gravitational wave signal, estimated to be α0.1\mathcal{F}_{\alpha}\approx 0.1 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; ρc\rho_{c} is the critical density of the Universe; the loop size parameter α=0.1\alpha=0.1 provides an accurate estimate of the loop size distribution Blanco-Pillado et al. (2014); Blanco-Pillado and Olum (2017); tFt_{F} is the time at which the cosmic string network was formed, related to the density of the Universe at that time via ρ(tF)=μ\sqrt{\rho(t_{F})}=\mu Gouttenoire et al. (2020b), t0kt_{0k} is the instance when the loop was produced, and θ(x)\theta(x) 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.

Refer to caption
Figure 1: Stochastic gravitational wave background from cosmic strings for four different symmetry breaking scales. Shaded regions correspond to the sensitivity of future gravitational wave detectors: LISA (green), DECIGO (blue), Big Bang Observer (purple), Einstein Telescope (red), and Cosmic Explorer (gray).
Refer to caption
Figure 2: Reach of future detectors in probing the scale of a U(1){\rm U}(1) symmetry breaking leading to the production of cosmic strings. The colors for each experiment correspond to those adopted in Fig. 1.

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 v104PeVv\gtrsim 10^{4}\ {\rm PeV}, 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 U(1)B{\rm U}(1)_{B} or U(1)L{\rm U}(1)_{L} 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 (a)(a) for both baryon and lepton number breaking, in case (b)(b) only for lepton number breaking, and in case (c)(c) only for baryon number breaking.

V Domain wall spectrum

Another kind of topological defects, appearing when a U(1){\rm U}(1) symmetry is broken by two scalars, are domain walls. As in the case of SU(2){\rm SU}(2) breaking discussed in Fornal and Pierre (2022), the effective potential V(ϕ1,ϕ2,T)V(\phi_{1},\phi_{2},T), which at high temperature has just one vacuum at (ϕ1,ϕ2)=(0,0)(\phi_{1},\phi_{2})=(0,0), at lower temperature develops four vacua. They come in two pairs related via a gauge transformation, ΦieiθΦi\Phi_{i}\to e^{i\theta}\Phi_{i}, and only two of them, say ϕvac1\vec{\phi}_{\rm vac1} and ϕvac2\vec{\phi}_{\rm vac2}, 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 ϕvac1\vec{\phi}_{\rm vac1} and ϕvac2\vec{\phi}_{\rm vac2}.

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 Z2{Z}_{2} 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 Z2{Z}_{2} symmetry removing the degeneracy between the vacua is provided by the terms involving m122m_{12}^{2}, λ4\lambda_{4} and λ5\lambda_{5} in the Lagrangian in Eq. (5).

Dynamics of domain walls

The profile of the domain wall configuration ϕdw(z)\vec{\phi}_{dw}(z) is the solution to the equation Chen et al. (2020)

d2ϕdw(z)dz2ϕVeff[ϕdw(z)]=0,\displaystyle\frac{d^{2}\vec{\phi}_{dw}(z)}{dz^{2}}-\vec{\nabla}_{\!\phi}V_{\rm eff}\big{[}\vec{\phi}_{dw}(z)\big{]}=0\ , (28)

where the zz-axis was chosen to be perpendicular to the domain wall, and the boundary conditions are,

ϕdw()=ϕvac1,ϕdw()=ϕvac2.\displaystyle\vec{\phi}_{dw}(-\infty)=\vec{\phi}_{\rm vac1}\ ,\ \ \ \ \ \vec{\phi}_{dw}(\infty)=\vec{\phi}_{\rm vac2}. (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 σ\sigma given by

σ=𝑑z[12(dϕdw(z)dz)2+Veff[ϕdw(z)]].\displaystyle\sigma=\int_{-\infty}^{\infty}dz\!\left[\frac{1}{2}\bigg{(}\frac{d\vec{\phi}_{dw}(z)}{dz}\bigg{)}^{2}+V_{\rm eff}\big{[}\vec{\phi}_{dw}(z)\big{]}\right]. (30)

In the model we are considering, to a good approximation

σv3.\displaystyle\sigma\,\sim\,v^{3}\ . (31)

The other parameter is the potential bias Δρ\Delta\rho, i.e., the energy density difference between the vacua, in our case equal to

Δρ=(m122+12λ4v12+12λ5v22)v1v2.\displaystyle\Delta\rho=\big{(}m_{12}^{2}+\tfrac{1}{2}\lambda_{4}v_{1}^{2}+\tfrac{1}{2}\lambda_{5}v_{2}^{2}\big{)}\,v_{1}v_{2}\ . (32)

The created domain walls are unstable and undergo annihilation, emitting gravitational radiation, provided that Saikawa (2017)

Δρ(σMPl)2.\displaystyle\Delta\rho\gtrsim\left(\frac{\sigma}{M_{Pl}}\right)^{2}\ . (33)

If in Eq. (32) the term involving m122m_{12}^{2} is the dominant one, then Eq. (33) takes the form m12v2/MPlm_{12}\gtrsim{v^{2}}/{M_{Pl}}. For example, if the symmetry breaking scale is v103PeVv\sim 10^{3}\ \rm PeV, this implies m12100MeVm_{12}\gtrsim 100\ {\rm MeV}. 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),

h2ΩDW(f)\displaystyle h^{2}\Omega_{\rm DW}(f) \displaystyle\approx 7.1×1033(σPeV3)4(TeV4Δρ)2(100g)13\displaystyle 7.1\times 10^{-33}\left(\frac{\sigma}{\rm PeV^{3}}\right)^{4}\left(\frac{\rm TeV^{4}}{\Delta\rho}\right)^{2}\left(\frac{100}{g_{*}}\right)^{\frac{1}{3}} (34)
×\displaystyle\times [(ffd)3θ(fdf)+(fdf)θ(ffd)],\displaystyle\!\!\bigg{[}\!\left(\frac{f}{f_{d}}\right)^{3}\!\theta(f_{d}-f)+\left(\frac{f_{d}}{f}\right)\theta(f-f_{d})\bigg{]},\ \ \ \ \ \ \

where for the area parameter we used 𝒜=0.8\mathcal{A}=0.8 and for the efficiency parameter we adopted the value ϵ~gw=0.7\tilde{\epsilon}_{\rm gw}=0.7 Hiramatsu et al. (2014),θ\theta denotes the step function, and the peak frequency fdf_{d} is

fd(0.14Hz)PeV3σΔρTeV4.\displaystyle f_{d}\approx(0.14\ {\rm Hz})\,\sqrt{\frac{\rm PeV^{3}}{\sigma}\frac{\Delta\rho}{\rm TeV^{4}}}\ . (35)

The slope of the signal falls f3\sim f^{3} to the left of the peak when moving toward lower frequencies, and falls like f1f^{-1} 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 h2Ω(f)<2.9×107h^{2}\Omega(f)<2.9\times 10^{-7} Clarke et al. (2020), which translates to the following condition on the parameters,

σΔρ2.5×1012PeV,\displaystyle\frac{\sigma}{\sqrt{\Delta\rho}}\lesssim 2.5\times 10^{12}\ \rm PeV\ , (36)

stronger than the bound imposed by Eq. (33).

Refer to caption
Figure 3: Stochastic gravitational wave background from domain walls for various symmetry breaking scales. Shaded regions correspond to the sensitivity of future gravitational wave detectors, as in Fig. 1.
Refer to caption
Figure 4: Regions of parameter space (v,Δρ)(v,\Delta\rho) for which the signal-to-noise ratio of the gravitational wave signal generated by domain wall annihilation is greater than five upon one year of data taking by various experiments. The choice of colors matches that in Fig. 3, including the color of the dots which correspond to the four curves.

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 vv and Δρ\Delta\rho. 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 Δρ\Delta\rho is a reflection of the cosmic microwave background constraint from Eq. (36). The parameter Δρ\Delta\rho depends in general on all three fundamental Lagrangian parameters m12m_{12}, λ4\lambda_{4}, and λ5\lambda_{5} through the relation in Eq. (32). Under the assumption that the term involving m12m_{12} is dominant, the experimental sensitivity plot in the plane (v,m12)(v,m_{12}) 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 U(1)B{\rm U}(1)_{B} 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 U(1)B{\rm U}(1)_{B} and gauged U(1)L{\rm U}(1)_{L}. 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 ϕ\phi 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 ϕ\phi remain at their tree-level values (i.e., the cutoff regularization scheme), the effective potential is given by

Veff(ϕ,T)=12λv2ϕ2+14λϕ4\displaystyle V_{\rm eff}(\phi,T)=-\frac{1}{2}\lambda v^{2}\phi^{2}+\frac{1}{4}\lambda\phi^{4}
+inimi2(ϕ)64π2{mi2(ϕ)[log(mi2(ϕ)mi2(v))32]+2mi2(v)}\displaystyle+\sum_{i}\frac{n_{i}m_{i}^{2}(\phi)}{64\pi^{2}}\bigg{\{}\!m_{i}^{2}(\phi)\!\left[\log\!\left(\frac{m_{i}^{2}(\phi)}{m_{i}^{2}(v)}\right)\!-\!\frac{3}{2}\right]\!+\!2m_{i}^{2}(v)\!\bigg{\}}
+T42π2ini0𝑑xx2log(1emi2(ϕ)/T2+x2)\displaystyle+\ \frac{T^{4}}{2\pi^{2}}\sum_{i}n_{i}\int_{0}^{\infty}dx\,x^{2}\log\left(1\mp e^{-\sqrt{{m_{i}^{2}(\phi)}/{T^{2}}+x^{2}}}\right)
+T12πjnj{mj3(ϕ)[mj2(ϕ)+Πj(T)]32}.\displaystyle+\ \frac{T}{12\pi}\sum_{j}n^{\prime}_{j}\left\{m_{j}^{3}(\phi)-\left[m_{j}^{2}(\phi)+\Pi_{j}(T)\right]^{\frac{3}{2}}\right\}\ . (37)

In the expression above the sums are over all particles charged under the U(1){\rm U}(1) including the Goldstone bosons χGB\chi_{\rm GB}, mi(ϕ)m_{i}(\phi) are the field-dependent masses, nin_{i} is the number of degrees of freedom for a given particle (nZ=3n_{Z^{\prime}}=3, nϕ=1n_{\phi}=1, nχGB=1n_{\chi_{\rm GB}}=1), njn_{j}^{\prime} is similar but includes only scalars and longitudinal components of vector bosons (nZ=1n^{\prime}_{Z^{\prime}}=1, nϕ=1n^{\prime}_{\phi}=1, nχGB=1n^{\prime}_{\chi_{\rm GB}}=1), and for the Goldstones one needs to replace mχGB(v)mϕ(v)m_{\chi_{\rm GB}}(v)\to m_{\phi}(v). We will assume that all new Yukawa couplings are small, so that the only relevant field-dependent masses are

mZ(ϕ)=xgϕ,mϕ(ϕ)=λ(3ϕ2v2),\displaystyle m_{Z^{\prime}}(\phi)=xg\phi\ ,\ \ \ \ m_{\phi}(\phi)=\sqrt{\lambda(3\phi^{2}-v^{2})}\ ,
mχGB(ϕ)=λ(ϕ2v2).\displaystyle m_{\chi_{\rm GB}}(\phi)=\sqrt{\lambda(\phi^{2}-v^{2})}\ . (38)

In the limit λg\lambda\ll g, the thermal masses are

ΠZ(T)\displaystyle\Pi_{Z^{\prime}}(T) =\displaystyle= 13(x2+92)g2T2,\displaystyle\frac{1}{3}\left(x^{2}+\frac{9}{2}\right)g^{2}T^{2}\ ,
Πϕ(T)\displaystyle\Pi_{\phi}(T) =\displaystyle= ΠχGB(T)=14x2g2T2,\displaystyle\Pi_{\chi_{\rm GB}}(T)=\frac{1}{4}x^{2}g^{2}T^{2}\ , (39)

where for gauged baryon number g=gBg=g_{B} and x=3x=3, while for gauged lepton number g=gLg=g_{L} and x=2x=2.

Refer to caption
Figure 5: The effective potential of the model, Veff(ϕB,T)V_{\rm eff}(\phi_{B},T), plotted for vB=100TeVv_{B}=100\ \rm TeV, gB=0.25g_{B}=0.25, λB=0.006\lambda_{B}=0.006, and several temperatures.

For a range of λ\lambda and gg values the effective potential develops a vacuum at ϕ0\phi\neq 0 (true vacuum) with a lower energy density than the high temperature vacuum at ϕ=0\phi=0 (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 TT_{*}, which is determined from the condition that the bubble nucleation rate Linde (1983) becomes comparable with the Hubble expansion,

(S(T)2πT)3/2T4eS(T)/TH(T)4,\displaystyle\bigg{(}\frac{S(T_{*})}{2\pi T_{*}}\bigg{)}^{3/2}T_{*}^{4}\,e^{-{S(T_{*})}/{T_{*}}}\approx H(T_{*})^{4}\ , (40)

where S(T)S(T) is the Euclidean action given by

S(T)=d3r[12(dϕbdr)2+Veff(ϕb,T)],\displaystyle S(T)=\int d^{3}r\left[\frac{1}{2}\left(\frac{d\phi_{b}}{dr}\right)^{2}+V_{\rm eff}(\phi_{b},T)\right], (41)

with ϕb\phi_{b} being the solution of the bubble equation,

d2ϕdr2+2rdϕdrdVeff(ϕ,T)dϕ=0,\displaystyle\frac{d^{2}\phi}{dr^{2}}+\frac{2}{r}\frac{d\phi}{dr}-\frac{dV_{\rm eff}(\phi,T)}{d\phi}=0\ , (42)

subject to the boundary conditions

dϕdr|r=0=0,ϕ()=ϕfalse.\displaystyle\frac{d\phi}{dr}\bigg{|}_{r=0}=0\ ,\ \ \ \ \ \phi(\infty)=\phi_{\rm false}\ . (43)

Since H(T)(T2/MPl)4π3g/45H(T)\approx(T^{2}/M_{Pl})\sqrt{4\pi^{3}g_{*}/45}, Eq. (40) becomes

S(T)T4log(MPlT)log[(4π3g45)2(2πTS(T))32].\displaystyle\frac{S(T_{*})}{T_{*}}\approx 4\log\!\left(\frac{M_{Pl}}{T_{*}}\right)\!-\!\log\left[\left(\frac{4\pi^{3}g_{*}}{45}\right)^{\!\!2}\!\left(\frac{2\pi\,T_{*}}{S(T_{*})}\right)^{\!\!\frac{3}{2}}\right]\!.\ \ (44)

The phase transition parameters relevant for determining the gravitational wave signal are: the bubble wall velocity vwv_{w}, the nucleation temperature TT_{*}, the phase transition strength α\alpha, and its duration 1/β~1/\tilde{\beta}. In our analysis we assume vw=cv_{w}=c, but other choices are also possible Espinosa et al. (2010); Caprini et al. (2016). The other three parameters, TT_{*}, α\alpha, and β~\tilde{\beta}, 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,

α=ρvac(T)ρrad(T).\displaystyle\alpha=\frac{\rho_{\rm vac}(T_{*})}{\rho_{\rm rad}(T_{*})}\ . (45)

Those two quantities are obtained from the relations

ρvac(T)\displaystyle\rho_{\rm vac}(T) =\displaystyle= Veff(ϕfalse,T)Veff(ϕtrue,T)\displaystyle V_{\rm eff}(\phi_{\rm false},T)-V_{\rm eff}(\phi_{\rm true},T)
\displaystyle- TT[Veff(ϕfalse,T)Veff(ϕtrue,T)],\displaystyle T\frac{\partial}{\partial T}{\left[V_{\rm eff}(\phi_{\rm false},T)-V_{\rm eff}(\phi_{\rm true},T)\right]}\ ,
ρrad(T)\displaystyle\rho_{\rm rad}(T) =\displaystyle= π230gT4,\displaystyle\frac{\pi^{2}}{30}g_{*}T^{4}\ , (46)

where gg_{*} is the number of degrees of freedom active when the bubbles are nucleated. The parameter β~\tilde{\beta}, related to the time scale of the phase transition, is determined via

β~=TddT(S(T)T)|T=T.\displaystyle\tilde{\beta}=T_{*}\frac{d}{dT}\bigg{(}\frac{S(T)}{T}\bigg{)}\bigg{|}_{T=T_{*}}\ . (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,

h2ΩPT(f)=h2Ωs(f)+h2Ωc(f)+h2Ωt(f).\displaystyle h^{2}\Omega_{\rm PT}(f)=h^{2}\Omega_{s}(f)+h^{2}\Omega_{c}(f)+h^{2}\Omega_{t}(f)\ . (48)

The expected signal from sound waves is Hindmarsh et al. (2014); Caprini et al. (2016)

h2Ωs(f)\displaystyle h^{2}\Omega_{s}(f)\, \displaystyle\approx 1.9×105β~(κsαα+1)2(100g)13Υ\displaystyle\,\frac{1.9\times 10^{-5}}{\tilde{\beta}}\left(\frac{\kappa_{s}\,\alpha}{\alpha+1}\right)^{2}\left(\frac{100}{g_{*}}\right)^{\frac{1}{3}}\Upsilon (49)
×\displaystyle\times (f/fs)3[1+0.75(f/fs)2]7/2,\displaystyle\frac{(f/f_{s})^{3}}{\big{[}1+0.75(f/f_{s})^{2}\big{]}^{7/2}}\ ,

where fsf_{s} is the peak frequency, κs\kappa_{s} is the fraction of the latent heat transformed into the bulk motion of the plasma Espinosa et al. (2010), and Υ\Upsilon is the suppression factor Ellis et al. (2020b); Guo et al. (2021),

fs\displaystyle f_{s} =\displaystyle= (0.19Hz)(T1PeV)(g100)16β~,\displaystyle(0.19\ {\rm Hz})\left(\frac{T_{*}}{1\ {\rm PeV}}\right)\left(\frac{g_{*}}{100}\right)^{\frac{1}{6}}\tilde{\beta}\ ,
κs\displaystyle\kappa_{s} =\displaystyle= α0.73+0.083α+α,\displaystyle\frac{\alpha}{0.73+0.083\sqrt{\alpha}+\alpha}\ ,
Υ\displaystyle\Upsilon\, =\displaystyle= 11[1+8π13β~(α+13κsα)12]12.\displaystyle 1-\frac{1}{\Big{[}1+\frac{8\pi^{\frac{1}{3}}}{\tilde{\beta}}\big{(}\frac{{\alpha+1}}{3\kappa_{s}\alpha}\big{)}^{\frac{1}{2}}\Big{]}^{\frac{1}{2}}}\ . (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)

h2Ωc(f)\displaystyle h^{2}\Omega_{c}(f)\, \displaystyle\approx 4.9×106β~2(κcαα+1)2(100g)13\displaystyle\,\frac{4.9\times 10^{-6}}{\tilde{\beta}^{2}}\left(\frac{\kappa_{c}\,\alpha}{\alpha+1}\right)^{2}\left(\frac{100}{g_{*}}\right)^{\frac{1}{3}} (51)
×\displaystyle\times (f/fc)2.81+2.8(f/fc)3.8,\displaystyle\frac{(f/f_{c})^{2.8}}{1+2.8(f/f_{c})^{3.8}}\ ,\ \ \ \ \ \

where now fcf_{c} is the peak frequency and κc\kappa_{c} is the fraction of the latent heat deposited into the bubble front Kamionkowski et al. (1994),

fc\displaystyle f_{c} =\displaystyle= (0.037Hz)(T1PeV)(g100)16β~,\displaystyle(0.037\ {\rm Hz})\left(\frac{T_{*}}{1\ {\rm PeV}}\right)\left(\frac{g_{*}}{100}\right)^{\frac{1}{6}}\tilde{\beta}\ ,
κc\displaystyle\kappa_{c} =\displaystyle= 42732α+0.72α1+0.72α.\displaystyle\frac{\frac{4}{27}\sqrt{\frac{3}{2}\alpha}+0.72\,\alpha}{1+0.72\,\alpha}\ . (52)

The final contribution is provided by turbulence Caprini and Durrer (2006); Caprini et al. (2009),

h2Ωt(f)\displaystyle h^{2}\Omega_{t}(f)\, \displaystyle\approx 3.4×104β~(ϵκsαα+1)32(100g)13\displaystyle\,\frac{3.4\times 10^{-4}}{\tilde{\beta}}\left(\frac{\epsilon\,\kappa_{s}\,\alpha}{\alpha+1}\right)^{\frac{3}{2}}\left(\frac{100}{g_{*}}\right)^{\frac{1}{3}} (53)
×\displaystyle\times (f/ft)3(1+8πf/f)(1+f/ft)11/3,\displaystyle\frac{({f}/{f_{t}})^{3}}{\big{(}1+{8\pi f}/{f_{*}}\big{)}\big{(}1+{f}/{f_{t}}\big{)}^{{11}/{3}}}\ ,

where ϵ=0.05\epsilon=0.05 Caprini et al. (2016), while the peak frequency ftf_{t} and the parameter ff_{*} are

ft\displaystyle f_{t} =\displaystyle= (0.27Hz)(g100)16(T1PeV)β~,\displaystyle(0.27\ {\rm Hz})\left(\frac{g_{*}}{100}\right)^{\frac{1}{6}}\left(\frac{T_{*}}{1\ {\rm PeV}}\right)\tilde{\beta}\ ,
f\displaystyle f_{*} =\displaystyle= (0.17Hz)(g100)16(T1PeV).\displaystyle(0.17\ {\rm Hz})\left(\frac{g_{*}}{100}\right)^{\frac{1}{6}}\left(\frac{T_{*}}{1\ {\rm PeV}}\right)\ . (54)
Refer to caption
Figure 6: Stochastic gravitational wave background from first order phase transitions triggered by U(1)B{\rm U}(1)_{B} breaking assuming the model parameters gB=0.25g_{B}=0.25 and λB=0.006\lambda_{B}=0.006, for several choices of the symmetry breaking scale. The shaded regions correspond to the sensitivity of future gravitational wave detectors, as in Fig. 1.

The resulting gravitational wave signals from the first order phase transition triggered by U(1)B{\rm U}(1)_{B} breaking are shown in Fig. 6 for several symmetry breaking scales: 10TeV10\ {\rm TeV} (light brown curve), 100TeV100\ {\rm TeV} (brown curve), 1PeV1\ {\rm PeV} (purple curve), and 10PeV10\ {\rm PeV} (black curve). The gauge coupling was assumed to be gB=0.25g_{B}=0.25, and the quartic coupling was chosen to be λB=0.006\lambda_{B}=0.006. Spectra with peaks at larger frequencies correspond to higher symmetry breaking scales.

The effect of the suppression factor Υ\Upsilon 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 (gB,λB)(g_{B},\lambda_{B}) parameter space for which the signal can be detected in each experiment when vB=1PeVv_{B}=1\ {\rm PeV}. 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 S(T)/TS(T)/T is too large to satisfy Eq. (44) or the new vacuum has an energy density larger than that of the high temperature vacuum.

Refer to caption
Figure 7: Regions of parameter space (gB,λB)(g_{B},\lambda_{B}) assuming vB=1PeVv_{B}=1\ {\rm PeV} where the gravitational wave signal from a first order phase transition has a signal-to-noise ratio greater than five upon one year of data taking in various experiments. The choice of colors matches that in Fig. 6, and the dot corresponds to the purple curve.

As mentioned earlier, our analysis of the first order phase transition signal from U(1)B{\rm U}(1)_{B} 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 Y=0.6Y=0.6 in Fornal and Shams Es Haghi (2020). Our analysis is also applicable to U(1)L{\rm U}(1)_{L} 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):

  • (a)(a)

    (PT+PT),(PT+CS),(CS+CS);{\rm(PT+PT),(PT+CS),(CS+CS);}

  • (b,c)(b,c)

    (PT+PT),(PT+CS),(PT+DW),(CS+DW);{\rm(PT+PT),(PT+CS),(PT+DW),(CS+DW);}

  • (d)(d)

    (PT+PT),(PT+DW),(DW+DW).{\rm(PT+PT),(PT+DW),(DW+DW).}

We discuss below explicit examples of how those signatures are realized in our model. Two of them, (DW+DW){\rm(DW+DW)} and (CS+DW){\rm(CS+DW)} have not been considered in the literature before, whereas (PT+CS){\rm(PT+CS)}, (PT+PT){\rm(PT+PT)} and (PT+DW){\rm(PT+DW)} have been already proposed. The case (CS+CS){\rm(CS+CS)} 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

Refer to caption
Figure 8: First novel gravitational wave signature of the model consisting of a double domain wall peak, realized when each of the U(1){\rm U}(1) symmetries is broken by two scalars – case (d)(d).
Refer to caption
Figure 9: Second novel gravitational wave signature consisting of a domain wall peak over a cosmic string background, realized when one U(1){\rm U}(1) symmetry is broken by one scalar and the other U(1){\rm U(1)} is broken by two scalars – cases (b)(b) and (c)(c).

A new gravitational wave signature arises when each of the two U(1){\rm U}(1) 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 f3\sim f^{3}, whereas the slope on the right side of the peak falls like 1/f\sim 1/f. 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 U(1)B{\rm U(1)}_{B} breaking were chosen to be vB=103PeVv_{B}=10^{3}\ {\rm PeV} and Δρ=105PeV4\Delta\rho=10^{-5}\ {\rm PeV^{4}}, whereas for the U(1)L{\rm U(1)}_{L} breaking they are vL=5×104PeVv_{L}=5\times 10^{4}\ {\rm PeV} and Δρ=1.6×105PeV4\Delta\rho=1.6\times 10^{5}\ {\rm PeV^{4}}.

Cosmic strings + domain walls

Another gravitational wave signature, not considered in the literature before, is realized when one of the U(1){\rm U}(1) symmetries is broken by one scalar, leading to cosmic string production, whereas the other U(1){\rm U}(1) 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 U(1)L{\rm U}(1)_{L} was chosen to be vL=106PeVv_{L}=10^{6}\ {\rm PeV}, whereas the parameters for U(1)B{\rm U}(1)_{B} breaking are vB=8×103PeVv_{B}=8\times 10^{3}\ {\rm PeV} and Δρ=2PeV4\Delta\rho=2\ {\rm PeV^{4}}. For this particular selection of parameters, Big Bang Observer and DECIGO can probe the peak area, but for lower U(1)B{\rm U}(1)_{B} breaking scales this structure is accessible to LISA, whereas higher symmetry breaking scales would make it detectable by Einstein Telescope and Cosmic Explorer.

Refer to caption
Figure 10: Gravitational wave signature with a first order phase transition peak over a cosmic string background, first proposed in Fornal and Shams Es Haghi (2020), realized when one U(1){\rm U}(1) is broken by only one scalar and the other U(1){\rm U}(1) is broken either by one or two scalars – cases (a)(a), (b)(b), (c)(c).

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 U(1)L{\rm U}(1)_{L} is broken at the scale vL=106PeVv_{L}=10^{6}\ {\rm PeV}, whereas the scale of U(1)B{\rm U}(1)_{B} breaking is vB=200TeVv_{B}=200\ \rm TeV. The other parameter values for this particular plot are gB=0.25g_{B}=0.25 and λB=0.006\lambda_{B}=0.006. As in the previous case, by changing the scale of U(1)B{\rm U}(1)_{B} 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 U(1){\rm U}(1) 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 U(1)B{\rm U}(1)_{B} symmetry is broken by one scalar at the scale vB=20TeVv_{B}=20\ {\rm TeV} (the other parameters are gB=0.25g_{B}=0.25 and λB=0.006\lambda_{B}=0.006), and the U(1)L{\rm U}(1)_{L} symmetry is broken also by one scalar at the scale vL=5PeVv_{L}=5\ {\rm PeV} (with gL=0.20g_{L}=0.20 and λL=0.0025\lambda_{L}=0.0025). 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.

Refer to caption
Figure 11: Gravitational wave signature consisting of two first order phase transition peaks, similar to the ones proposed in Angelescu and Huang (2019); Greljo et al. (2020); Fornal (2021), arising when the two U(1){\rm U}(1) symmetries are broken by any number of scalars – realized in cases (a)(a), (b)(b), (c)(c), (d)(d).

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 U(1)B{\rm U}(1)_{B} and U(1)L{\rm U}(1)_{L} symmetry breaking scales. Figure 12 shows a realization of this scenario when U(1)L{\rm U}(1)_{L} is broken by two scalars at the scale vL=3×104PeVv_{L}=3\times 10^{4}\ {\rm PeV} (with a potential bias Δρ=6.3×103PeV4\Delta\rho=6.3\times 10^{3}\ {\rm PeV^{4}}), whereas U(1)B{\rm U}(1)_{B} is broken by one scalar at the scale vB=20TeVv_{B}=20\ {\rm TeV} (with the other parameters being gB=0.25g_{B}=0.25 and λB=0.006\lambda_{B}=0.006). As pointed out in Fornal et al. (2023), the two peaks may appear in a different order, which would happen for a U(1)L{\rm U}(1)_{L} breaking scale of vL103PeVv_{L}\sim 10^{3}\ {\rm PeV} and a U(1)B{\rm U}(1)_{B} breaking scale of vB10PeVv_{B}\sim 10\ {\rm PeV}. 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 U(1)L{\rm U}(1)_{L} 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 U(1)B{\rm U}(1)_{B} 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.

Refer to caption
Figure 12: Gravitational wave signature containing a phase transition peak and a domain wall peak, proposed in Fornal et al. (2023), realized when at least one U(1){\rm U}(1) symmetry is broken by two scalars – cases (b)(b), (c)(c), (d)(d).

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: (1)(1) Double-sharp-peak structure from domain walls produced when two gauge symmetries are broken by multiple scalars; (2)(2) 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 U(1){\rm U}(1) 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 SU(2){\rm SU}(2) 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 U(1)B{\rm U(1)_{B}} symmetry breaking is not bounded from above and can also be high, allowing for signals (1)(1) and (2)(2) 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 1001000PeV\sim 100-1000\ {\rm PeV} 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