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

Supernova constraint on self-interacting dark sector particles

Allan Sung [email protected] Institute of Physics, Academia Sinica, Taipei, 11529, Taiwan Department of Physics, Princeton University, Princeton, New Jersey 08544, USA    Gang Guo [email protected] Institute of Physics, Academia Sinica, Taipei, 11529, Taiwan Key Laboratory of Dark Matter and Space Astronomy, Purple Mountain Observatory, Chinese Academy of Sciences, Nanjing 210033, China    Meng-Ru Wu [email protected] Institute of Physics, Academia Sinica, Taipei, 11529, Taiwan Institute of Astronomy and Astrophysics, Academia Sinica, Taipei, 10617, Taiwan
Abstract

We examine the constraints on sub-GeV dark sector particles set by the proto-neutron star cooling associated with the core-collapse supernova event SN1987a. Considering explicitly a dark photon portal dark sector model, we compute the relevant interaction rates of dark photon (AA^{\prime}) and dark fermion (χ\chi) with the Standard Model particles as well as their self-interaction inside the dark sector. We find that even with a small dark sector fine structure constant αD1\alpha_{D}\ll 1, dark sector self-interactions can easily lead to their own self-trapping. This effect strongly limits the energy luminosity carried away by dark sector particles from the supernova core and thus drastically affects the parameter space that can be constrained by SN1987a. We consider specifically two mass ratios mA=3mχm_{A^{\prime}}=3m_{\chi} and 3mA=mχ3m_{A^{\prime}}=m_{\chi} which represent scenarios where the decay of AA^{\prime} to χχ¯\chi\bar{\chi} is allowed or not. For mA=3mχm_{A^{\prime}}=3m_{\chi}, we show that this effect can completely evade the supernova bounds on widely-examined dark photon parameter space for a dark sector with αD107\alpha_{D}\gtrsim 10^{-7}. In particular, for the mass range mχ20m_{\chi}\lesssim 20 MeV, supernova bounds can only be applied to weakly self-interacting dark sector with αD1015\alpha_{D}\lesssim 10^{-15}. For 3mA=mχ3m_{A^{\prime}}=m_{\chi}, bounds in regions where αD107\alpha_{D}\gtrsim 10^{-7} for mχ20m_{\chi}\lesssim 20 MeV can be evaded similarly. Our findings thus imply that the existing supernova bounds on light dark particles can be generally eluded by a similar self-trapping mechanism. This also implies that nonstandard strongly self-interacting neutrino is not consistent with the SN1987a observation. Same effects can also take place for other known stellar bounds on dark sector particles.

I Introduction

The detection of 20 electron antineutrinos emitted from the core-collapse supernova (SN) explosion event, SN1987a Hirata et al. (1987); Bionta et al. (1987); Alekseev et al. (1988), not only broadly confirmed the prevalent SN theory, but also led to several important consequences to fundamental physics, including, e.g., bounds on the neutrino decay lifetime, the absolute masses of neutrinos, and nonstandard neutrino interactions Sato and Suzuki (1987); Spergel et al. (1987); Bahcall et al. (1987); Burrows and Lattimer (1987); Schramm and Truran (1990); Loredo and Lamb (2002); Frieman et al. (1988); Kolb and Turner (1989); Berezhiani and Smirnov (1989); Farzan (2003). In particular, important constraints on a variety of particles beyond the Standard Model (SM) including the axion, sterile neutrino, dark photon, etc., whose masses are sub-GeV, were derived Raffelt and Seckel (1988); Turner (1988); Mayle et al. (1988); Brinkmann and Turner (1988); Janka et al. (1996); Keil et al. (1997); Fischer et al. (2016); Carenza et al. (2019); Bollig et al. (2020); Lucente et al. (2020); Raffelt and Zhou (2011); Argüelles et al. (2019); Suliga et al. (2019); Syvolap et al. (2019); Mastrototaro et al. (2020); Suliga et al. (2020); Dent et al. (2012); Rrapaj and Reddy (2016); Mahoney et al. (2017); Chang et al. (2017); Hardy and Lasenby (2017); Chang et al. (2018); Guha et al. (2017, 2019); Ishizuka and Yoshimura (1990); Arndt and Fox (2003); Keung et al. (2014); Tu and Ng (2017); Dev et al. (2020); Croon et al. (2021); Camalich et al. (2020); Farzan (2003); Hanhart et al. (2001a, b); Hannestad et al. (2007); Freitas and Wyler (2007), which complement ongoing experimental searches for those particles. These constraints were based on the requirement that the exotic particles should not carry away an amount of energy from the cooling proto-neutron star (PNS) more than the inferred total energy carried by neutrinos, Eν3×1053E_{\nu}\simeq 3\times 10^{53} erg (see, however, a caution from Ref. Bar et al. (2020)). In addition to the SN cooling (more precisely, the PNS cooling) constraint, recent works also proposed new constraints on light dark photon or dark photon portal light dark matter (DM) based on other SN-related observables, such as the measured SN explosion energy Sung et al. (2019), the γ\gamma-rays Kazanas et al. (2014); DeRocco et al. (2019a), or the produced (semi-)relativistic dark matter flux arriving at the terrestrial detectors DeRocco et al. (2019b).

One important aspect in deriving the SN constraint on light dark sector (DS) particles is that their interaction with SM particles cannot be too strong for them being trapped inside the PNS. We note that previous studies always ignored the self-interactions between dark sector particles when deriving the SN bounds. However, if the abundance of dark sector particles inside the SN core can be as large as SM particles, and if the self-interaction cross section can be as large as the neutrino-nucleon scattering cross section O(1041)\sim O(10^{-41}) cm2 for neutrinos of 𝒪(10)\sim\mathcal{O}(10) MeV, dark sector particles can trap themselves inside the PNS. Consequently, SN bounds on self-interacting dark sector particles can be largely evaded (see also a very recent study discussing SN bound on axion-like particle portal light DM Darmé et al. (2020)). Note that self-interacting dark matter has been considered to be a viable option to resolve a number of tensions in the scale of galaxies or galaxy clusters, e.g., the core-cusp, too-big-to-fail, and the missing satellites problems; see, e.g., Refs. Spergel and Steinhardt (2000); Vogelsberger et al. (2012); Rocha et al. (2013); Tulin and Yu (2018). Extensive efforts investigating consequences of self-interacting or annihilating dark matter on various cosmological and astrophysical signatures have been pursued in recent years, e.g., Refs. Tulin et al. (2013); Bernal et al. (2016); Kawasaki et al. (2015); Bringmann et al. (2017); Elor et al. (2016); Ackermann et al. (2015); Abdallah et al. (2016); Feng et al. (2016); Leane et al. (2017); Kouvaris et al. (2016); Essig et al. (2019); Chang et al. (2019); Depta et al. (2019); Bernal et al. (2020); Foot and Vagnozzi (2015).

In this work, we aim to address the issue of SN constraints on self-interacting dark sector particles by taking into account their self-trapping effect in a systematic way for the first time. We use a widely examined dark photon portal dark sector model explicitly to compute all the relevant interaction cross sections and the decay rates. In principle, to determine precisely the dark sector particle fluxes emerging from the PNS requires solving full Boltzmann transport equations in a way similar to the neutrino transport problem in SNe (see e.g., a recent review Mezzacappa et al. (2020) and references therein).111Reference DeRocco et al. (2019b) adopted a Monte-Carlo based particle transport scheme to compute the light dark matter flux emitted from the PNS, without considering their potential self-interactions. Such approach demands intensive computational power to fully incorporate the scattering kernels and particle annihilation. Instead of directly pursuing full numerical simulations, we adopt an approximated approach to estimate the energy fluxes carried by dark photons and dark fermions evaluated in the nondiffuse regime and diffuse regime separately, and formulate a physically motivated criterion to switch from one regime to another. This approach allows us to estimate the effect of dark sector self-trapping on SN bounds for a wide range of parameter space, which turns out to be very important even for small couplings in the dark sector.

The rest of the paper is organized as follows. In Sec. II, we describe the underlying dark photon portal dark sector model, the considered SN model, and list the relevant interactions and decay processes that we included in this work. In Sec. III, we compute the energy luminosity of dark sector particles leaving the PNS in the nondiffuse and the diffuse regime, respectively, and formulate the criterion to switch from one regime to another. We apply this method to derive SN bounds on self-interacting dark sector particles in Sec. IV. Our conclusion and discussions of potential caveats as well as other implications are given in Sec. V. All detailed derivations of the cross sections, the decay rates, and the diffusion luminosity of dark sector particles are given in the Appendices. We adopt natural units with =c=1\hbar=c=1 throughout the paper unless explicitly specified.

II Models

II.1 Dark sector model

Mass Interaction Type Particle Coupling
mA<2mχm_{A^{\prime}}<2m_{\chi} AnpnpA^{\prime}np\rightarrow np Abs. SM ϵ2\epsilon^{2}
AeeγA^{\prime}e^{-}\rightarrow e^{-}\gamma Abs. SM ϵ2\epsilon^{2}
Aee+A^{\prime}\rightarrow e^{-}e^{+} Abs. SM ϵ2\epsilon^{2}
AAχχ¯A^{\prime}A^{\prime}\rightarrow\chi\bar{\chi} Abs. DS αD2\alpha_{D}^{2}
AχχAA^{\prime}\chi\rightarrow\chi A^{\prime} Sca. DS αD2\alpha_{D}^{2}
mA>2mχm_{A^{\prime}}>2m_{\chi} AnpnpA^{\prime}np\rightarrow np Abs. SM ϵ2\epsilon^{2}
AeeγA^{\prime}e^{-}\rightarrow e^{-}\gamma Abs. SM ϵ2\epsilon^{2}
Aee+A^{\prime}\rightarrow e^{-}e^{+} Abs. SM ϵ2\epsilon^{2}
Aχχ¯A^{\prime}\rightarrow\chi\bar{\chi} Abs. DS αD\alpha_{D}
Table 1: Relevant processes of dark photon interactions considered in this work. “Abs.” refers to a process which absorbs dark photon(s) or decay of dark photon. “Sca.” refers to a scattering process of a dark photon with a Standard Model (SM) particle or a dark sector (DS) particle. Only leading-order processes are included here. Note that for mA>2mχm_{A^{\prime}}>2m_{\chi}, Aχχ¯A^{\prime}\rightarrow\chi\bar{\chi} also accounts for contribution from AχAχA^{\prime}\chi\rightarrow A^{\prime}\chi [see Eq. (37) and text below for details].
Mass Interaction Type Particle Coupling
mA<2mχm_{A^{\prime}}<2m_{\chi} χχ¯npnp\chi\bar{\chi}np\rightarrow np Abs. SM ϵ2αD\epsilon^{2}\alpha_{D}
χχ¯ee+\chi\bar{\chi}\rightarrow e^{-}e^{+} Abs. SM ϵ2αD\epsilon^{2}\alpha_{D}
χχ¯γ\chi\bar{\chi}\rightarrow\gamma^{\ast} Abs. SM ϵ2αD\epsilon^{2}\alpha_{D}
χpχp\chi p\rightarrow\chi p Sca. SM ϵ2αD\epsilon^{2}\alpha_{D}
χeχe\chi e^{-}\rightarrow\chi e^{-} Sca. SM ϵ2αD\epsilon^{2}\alpha_{D}
χχ¯AA\chi\bar{\chi}\rightarrow A^{\prime}A^{\prime} Abs. DS αD2\alpha_{D}^{2}
χχχχ\chi\chi\rightarrow\chi\chi Sca. DS αD2\alpha_{D}^{2}
χχ¯χχ¯\chi\bar{\chi}\rightarrow\chi\bar{\chi} Sca. DS αD2\alpha_{D}^{2}
χAAχ\chi A^{\prime}\rightarrow A^{\prime}\chi Sca. DS αD2\alpha_{D}^{2}
mA>2mχm_{A^{\prime}}>2m_{\chi} χpχp\chi p\rightarrow\chi p Sca. SM ϵ2αD\epsilon^{2}\alpha_{D}
χeχe\chi e^{-}\rightarrow\chi e^{-} Sca. SM ϵ2αD\epsilon^{2}\alpha_{D}
χχ¯A\chi\bar{\chi}\rightarrow A^{\prime} Abs. DS αD\alpha_{D}
Table 2: Relevant processes of dark fermion interactions. Notations are the same as in Table 1. Only leading-order processes are included here. Note that for mA>2mχm_{A^{\prime}}>2m_{\chi}, χχ¯A\chi\bar{\chi}\rightarrow A^{\prime} also accounts for DS processes of AχAχA^{\prime}\chi\rightarrow A^{\prime}\chi and χχ¯χχ¯\chi\bar{\chi}\rightarrow\chi\bar{\chi}, as well as SM processes involving a pair of dark fermions [see Eq. (37) and text below for details].

We consider a dark photon portal DS model wherein the Dirac dark fermion χ\chi couples to dark photon AA^{\prime} and the dark photon kinetically mixes with the SM photon Holdom (1986); Okun (1982); Pospelov et al. (2008). The corresponding Lagrangian of the dark sector is given by

14FμνFμνϵ2FμνFμν+12mA2AμAμ+χ¯(i∂̸mχ)χ+gDχ¯χ,\begin{split}\mathcal{L}&\supset-\frac{1}{4}F^{\prime}_{\mu\nu}F^{\prime\mu\nu}-\frac{\epsilon}{2}F^{\prime}_{\mu\nu}F^{\mu\nu}+\frac{1}{2}m_{A^{\prime}}^{2}A^{\prime}_{\mu}A^{\prime\mu}\\ &+\bar{\chi}\left(i\not{\partial}-m_{\chi}\right)\chi+g_{D}\bar{\chi}\not{A^{\prime}}\chi,\end{split} (1)

where ϵ\epsilon is the mixing parameter, mAm_{A^{\prime}} is the dark photon mass, mχm_{\chi} is the mass of the dark fermion, and gDg_{D} is the DS coupling constant. We define the DS fine structure constant αDgD2/4π\alpha_{D}\equiv g_{D}^{2}/4\pi analogous to the electromagnetic fine structure constant αee2/4π\alpha_{e}\equiv e^{2}/4\pi.

Through the mixing of dark photon with the SM photon, dark photons and dark fermions can be produced via processes analogous to the SM electromagnetic ones. We follow Refs. Chang et al. (2017, 2018) to consider the following production channels for light dark photons and dark fermions inside the hot and dense interior of a PNS with a temperature 30\simeq 30 MeV and a core mass density 1014\gtrsim 10^{14} g cm-3. For the dark photon, we include the nucleon-nucleon bremsstrahlung npnpAnp\rightarrow npA^{\prime}, Compton-like interaction γeeA\gamma e^{-}\rightarrow e^{-}A^{\prime}, and electron-positron annihilation ee+Ae^{-}e^{+}\rightarrow A^{\prime}. For the dark fermion, we consider three pair-production channels including nucleon-nucleon bremsstrahlung npnpχχ¯np\rightarrow np\chi\bar{\chi}, electron-positron annihilation ee+χχ¯e^{-}e^{+}\rightarrow\chi\bar{\chi}, and plasmon decay γχχ¯\gamma^{\ast}\rightarrow\chi\bar{\chi}.

When the PNS interior is optically thin to dark photons or dark fermions, the rates of the above production channels directly determine the energy luminosity carried away by the dark particles. However, when the interactions between dark particles and the SM medium, as well as the self-interactions in the DS, are strong enough, dark photons and fermions can be trapped in the PNS. These interactions include the inverse processes of the above production channels, the dark photon (fermion) pair-annihilation AAχχ¯A^{\prime}A^{\prime}\leftrightarrow\chi\bar{\chi}, the DS Compton scattering AχχAA^{\prime}\chi\leftrightarrow\chi A^{\prime}, the scattering of dark fermions χχχχ\chi\chi\rightarrow\chi\chi and χχ¯χχ¯\chi\bar{\chi}\rightarrow\chi\bar{\chi}, as well as the dark fermion scattering with SM charged fermions χpχp\chi p\rightarrow\chi p and χeχe\chi e^{-}\rightarrow\chi e^{-}. These interactions determine the mean free path of the DS particles and thus the energy loss rate through the neutrinosphere in the diffuse regime. Moreover, when mA>2mχm_{A^{\prime}}>2m_{\chi}, the (inverse) decay of dark photon Aχχ¯A^{\prime}\leftrightarrow\chi\bar{\chi} needs to be considered. In this scenario, the inverse decay process dominates all the dark fermion pair-absorption processes and the DS self-interactions due to the resonances.222χχ¯χχ¯\chi\bar{\chi}\rightarrow\chi\bar{\chi} has a dark photon resonance and AχAχA^{\prime}\chi\rightarrow A^{\prime}\chi has a dark fermion resonance.; see Eq. (37) and text below in the Appendix VI for details. Other DS self-interactions without resonances are suppressed by an extra factor of αD\alpha_{D} compared to the (inverse) decay rate and thus can be neglected for mA>2mχm_{A^{\prime}}>2m_{\chi}. We list all the included interactions in this work in Tables 1 and 2 for dark photon and dark fermion, respectively.

Throughout this work, we will use perturbative calculation for the interaction rates. Since the temperature in the PNS is 30\sim 30 MeV, we only consider DS particles of masses below 1\simeq 1 GeV.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 1: Supernova profiles (density, temperature, electron chemical potential, and plasma frequency) used in this work from the 1818 MM_{\odot} progenitor model in Ref. Fischer et al. (2010), extracted at the time 1 s post the core bounce. The vertical dash line indicates where the neutrinosphere is located.

II.2 PNS cooling constraint and SN model

Without the presence of dark particles, long-term SN simulations predicted that the PNS cools by emitting neutrinos of all flavors in 10\sim 10 s. The total energy carried away by neutrinos is 3×1053\simeq 3\times 10^{53} erg, which is fixed by the available gravitational energy released to form a compact neutron star. The potential emission of exotic particles that may be produced inside the PNS will thus reduce the duration of the neutrino emission and can be constrained by the observed neutrino events from SN1987a. Based on comparisons with simulations including the emission of axions from the PNS, a well-known criterion (Raffelt’s criterion) was formulated to constrain the maximal energy luminosity carried by any exotic particles Raffelt (1990),

LDLν3×1052ergs1,L_{D}\leq L_{\nu}\equiv 3\times 10^{52}~{}{\rm erg~{}s}^{-1}, (2)

where LDL_{D} denotes the energy luminosity of dark emission, evaluated during the early PNS cooling phase, and LνL_{\nu} is approximately the time-averaged neutrino energy luminosity in SM.

In this work, we use the PNS density and temperature profile at 1 s post the SN core bounce obtained in Ref. Fischer et al. (2010) to compute the emission of dark photons and dark fermions. This PNS profile was widely used recently for similar purposes (see e.g., Refs. Chang et al. (2017); Sung et al. (2019)). Note that the choice of a particular SN model may introduce uncertainties of a factor of a few for the derived bounds Chang et al. (2017); Mahoney et al. (2017). Figure 1 shows the radial evolution of the density ρ\rho, temperature TT, the electron chemical potential μe\mu_{e}, and the plasma frequency ωp\omega_{p} [see Eq. (45)]. The position of the spectral-averaged neutrino decoupling sphere, i.e., neutrinosphere, is indicated by the vertical dash line at Rν22R_{\nu}\simeq 22 km. The density profile shows a monotonically decreasing behavior as a function of radius, while the temperature profile exhibits a peak at r11r\simeq 11 km, due to the inefficient compression heating at the densest core region. The electrons are highly degenerate inside the PNS. The plasma frequency ωp14\omega_{p}\simeq 14 MeV at the PNS center and decreases at larger radii. The plasma effect effectively alters the mixing between the dark photon and the SM photon differently for the transverse and the longitudinal polarizations An et al. (2013); Hardy and Lasenby (2017); Chang et al. (2017). We have included this effect throughout this work and give the details in Appendix VII.

III Luminosity of dark sector particles

In this section, we first describe how we compute the energy luminosity of DS particles leaving the PNS for the scenario where the DS self-trapping can be ignored (Sec. III.1) and for cases where they can be considered as diffusive due to self-trapping (Sec. III.2). We formulate the criteria that determine if a DS particle species is in diffuse regime or not in Sec. III.3. In Sec. III.4, we then apply our formalism to the adopted PNS profile to compute the total luminosity carried away by DS particles from the PNS interior.

In the rest of the paper, we denote the 4-momentum of AA^{\prime} by k=(ω,k)k=(\omega,\vec{k}), that of χ\chi by p=(E,p)p=(E,\vec{p}) and that of χ¯\bar{\chi} by p=(E,p)p^{\prime}=(E^{\prime},\vec{p}^{\prime}), unless noted otherwise.

III.1 Nondiffuse regime

In the nondiffuse regime, we consider the bulk emission rates of DS particles inside the neutrinosphere and the attenuation due to absorption and decay, following Refs. Chang et al. (2017); Sung et al. (2019). The luminosity of the dark photon is given by

LA=L,T0Rν4πr2𝑑rd3k(2π)3×gL,TωΓA,prodL,T(ω,r)eτL,T(ω,r),\begin{split}L_{A^{\prime}}&=\sum_{L,T}\int_{0}^{R_{\nu}}4\pi r^{2}dr\int\frac{d^{3}k}{(2\pi)^{3}}\\ &\quad\times g_{L,T}\omega\Gamma^{L,T}_{A^{\prime},\rm prod}(\omega,r)e^{-\tau_{L,T}(\omega,r)},\end{split} (3)

where gL=1g_{L}=1, gT=2g_{T}=2, and ΓA,prodL,T\Gamma^{L,T}_{A^{\prime},\rm prod} and τL,T\tau_{L,T} are the production rate and optical depth respectively. The exponential factor accounts for the absorption of dark photons by the medium and their decay. We separate the longitudinal (L) and transverse (T) modes because the medium effect modifies their dispersion relations and leads to different effective mixings of the two modes with the SM photon (see Appendix VII). The production rate ΓA,prodL,T\Gamma^{L,T}_{A^{\prime},\rm prod} is determined by the interactions involving the SM particles listed in Table 1. By detailed balance, the rates of each production process and its inverse process are related by ΓA,prodL,T=eω/TΓA,absL,T\Gamma^{L,T}_{A^{\prime},\rm prod}=e^{-\omega/T}\Gamma^{L,T}_{A^{\prime},\rm abs}, where ΓA,absL,T\Gamma^{L,T}_{A^{\prime},\rm abs} is the total absorption rate of the inverse process. Therefore,

ΓA,prodL,T=eω/T(ΓAnpnpL,T+ΓAeeγL,T+ΓAee+L,T).\Gamma^{L,T}_{A^{\prime},\rm prod}=e^{-\omega/T}(\Gamma^{L,T}_{A^{\prime}np\rightarrow np}+\Gamma^{L,T}_{A^{\prime}e^{-}\rightarrow e^{-}\gamma}+\Gamma^{L,T}_{A^{\prime}\rightarrow e^{-}e^{+}}). (4)

For the optical depth, we include the absorption and decay of AA^{\prime} by the same processes with SM particle as above and ignore those involving DS particles, to avoid double counting the total DS luminosity (see also later discussions in this subsection).333The pair-annihilation rate of AAχχ¯A^{\prime}A^{\prime}\rightarrow\chi\bar{\chi} is also ignored since the dark photon abundance inside the PNS is relatively small compared to that of SM particles in the nondiffuse regime. This gives

τL,T(ω,r)=f(r)rRνdr~vΓA,absL,T(ω,r~),\tau_{L,T}(\omega,r)=f(r)\int_{r}^{R_{\nu}}\frac{d\tilde{r}}{v}\Gamma_{A^{\prime},\rm abs}^{L,T}(\omega,\tilde{r}), (5)

where v=|k|/ωv=|\vec{k}|/\omega is the dark photon velocity, and f(r)f(r) is a geometric factor used in Chang et al. (2017) that effectively takes into account different path lengths of dark photons emitting locations to the neutrinosphere. The explicit forms of these absorption rates are given in Appendix VIII.

Next, we compute the luminosity of the dark fermion (χ\chi) in the nondiffuse regime as

Lχ=0Rν4πr2𝑑rd3p(2π)3gχEΓχ,prod(E,r),L_{\chi}=\int_{0}^{R_{\nu}}4\pi r^{2}dr\int\frac{d^{3}p}{(2\pi)^{3}}g_{\chi}E\,\Gamma_{\chi,\rm prod}(E,r), (6)

where gχ=2g_{\chi}=2 is the physical degrees of freedom of χ\chi, and Γχ,prod\Gamma_{\chi,\rm prod} is the production rate of χ\chi by the SM medium (see Table 2). Note that since χ\chi and χ¯\bar{\chi} are symmetric in our model, the total luminosity in the dark fermion pair is Lχ+Lχ¯=2LχL_{\chi}+L_{\bar{\chi}}=2L_{\chi}. Here we do not include the attenuation due to pair absorption. This is because in the nondiffuse regime, the dark fermion abundance below the neutrinosphere is very low, which suppresses the pair absorption of dark fermions.

If the dark fermions were in equilibrium with the medium, detailed balance could relate the production rates to the absorption rates by Γχ,prodeq=eE/TΓχ,abseq\Gamma^{\rm eq}_{\chi,\rm prod}=e^{-E/T}\Gamma^{\rm eq}_{\chi,\rm abs}. We take the equilibrium production rates Γχ,prodeq\Gamma^{\rm eq}_{\chi,\rm prod} as an approximation for the production rates Γχ,prod\Gamma_{\chi,\rm prod} used in Eq. (6). That is,

Γχ,prodeE/T(Γχχ¯npnp+Γχχ¯ee++Γχχ¯γ).\Gamma_{\chi,\rm prod}\simeq e^{-E/T}(\Gamma_{\chi\bar{\chi}np\rightarrow np}+\Gamma_{\chi\bar{\chi}\rightarrow e^{-}e^{+}}+\Gamma_{\chi\bar{\chi}\rightarrow\gamma^{\ast}}). (7)

This approximation in principle underestimates a bit the production rates of χ\chi and χ¯\bar{\chi} due to the assumed equilibrium occupation number, which effectively results in Pauli blocking. However, since the χ\chi and χ¯\bar{\chi} are always produced in pairs, the effective Pauli-blocking suppression of the rates is small due to their zero chemical potentials.

We note that when mA2mχm_{A^{\prime}}\geq 2m_{\chi}, the dark fermion pair production is in fact dominated by the decay of on-shell dark photons. Since in Eq. (3) we do not include the decay of AA^{\prime} to χ\chi and χ¯\bar{\chi} when AA^{\prime} are in nondiffuse regime, it leads to double counting of the total dark luminosity if we also include the dark fermion production. Thus, we do not consider the contribution of Eq. (6) when dark photons are in nondiffuse regime and when mA2mχm_{A^{\prime}}\geq 2m_{\chi}. We also note here that if dark photons are in the diffuse limit but dark fermions are not, then effectively the non-diffuse dark fermion luminosity can be determined by the decay rate of the trapped dark photons that are in thermal equilibrium with the SM medium (see Appendix VI).444Note that we consider separately the diffuse condition for the longitudinal and the transverse dark photons. Thus, we further multiply the nondiffuse dark fermion luminosity by a factor of 1/31/3 (2/32/3) if only the longitudinal (transverse) mode of the dark photons are trapped (see Appendix IX).

Before we discuss the detailed numerical results of the DS luminosity in the nondiffuse regime, let us provide an analytic estimation for the relevant region of the parameter space. First, for most cases, the production rates of longitudinal dark photons and dark fermions are suppressed by a factor of mA2/ω2m_{A^{\prime}}^{2}/\omega^{2} (see Appendix VIII) and by the coupling constant αD\alpha_{D}, respectively, compared to the production rate of the transverse dark photons. We can thus approximate the nondiffuse luminosity of DS particles by considering transverse dark photons only. Second, we consider for simplicity a homogeneous PNS with radius Rc10 kmR_{c}\simeq 10\mbox{ km}, temperature T30 MeVT\simeq 30\mbox{ MeV}, density ρ3×1014 g/cm3\rho\simeq 3\times 10^{14}\mbox{ g}/\mbox{cm}^{3} and electron fraction Ye0.3Y_{e}\simeq 0.3. We also assume that dark photons are relativistic. Taking the nucleon-nucleon bremsstrahlung AnpnpA^{\prime}np\rightarrow np without the plasma effects (see Appendix VIII.1), the luminosity of DS particles, LDL_{D}, is approximately

LDLν×(ϵ4×1010)2×exp(mA29 MeV)L_{D}\simeq L_{\nu}\times\left(\frac{\epsilon}{4\times 10^{-10}}\right)^{2}\times\exp\left(-\frac{m_{A^{\prime}}}{29\mbox{ MeV}}\right) (8)

for mA1m_{A^{\prime}}\lesssim 1 GeV.

III.2 Diffuse regime

In the diffuse regime, we assume that DS particles are in good thermal contact with the SM medium. Due to the temperature gradient, the DS phase space distributions are slightly anisotropic, which induces an outward energy flux. We use the radiative transfer equation for the DS particle energy flux through the neutrinosphere in Appendix X. The energy flux of a particle species ii is approximately given by

Li=2giRν2Tν33πdTdr|Rν1λi1(Rν)×mi/Tνξ3ξ2(miTν)2eξ(eξ±1)2dξ,\begin{split}L_{i}&=-\frac{2g_{i}R_{\nu}^{2}T_{\nu}^{3}}{3\pi}\left.\frac{dT}{dr}\right|_{R_{\nu}}\frac{1}{\langle\lambda^{-1}_{i}(R_{\nu})\rangle}\\ &\quad\times\int_{m_{i}/T_{\nu}}^{\infty}\xi^{3}\sqrt{\xi^{2}-\left(\frac{m_{i}}{T_{\nu}}\right)^{2}}\frac{e^{\xi}}{\left(e^{\xi}\pm 1\right)^{2}}d\xi,\end{split} (9)

where the upper (lower) sign is for fermions (bosons), gig_{i} is the physical degrees of freedom of particle ii, TνT_{\nu} is the temperature at the neutrinosphere, mim_{i} is the mass of particle ii, and λi1(r)\langle\lambda^{-1}_{i}(r)\rangle is the thermally averaged inverse mean free path (IMFP)555The IMFP used here is rescaled by the relative abundances of the DS particles. See Sec. III.3 for the definition. of particle ii at radius rr defined by

λi1(r)d3pfi(E,T(r))λi1(E,r)d3pfi(E,T(r)),\langle\lambda^{-1}_{i}(r)\rangle\equiv\frac{\int d^{3}p\,f_{i}(E,T(r))\lambda^{-1}_{i}(E,r)}{\int d^{3}p\,f_{i}(E,T(r))}, (10)

where fi(E,T(r))f_{i}(E,T(r)) is the distribution function of particle ii at radius rr. We distinguish between the absorptive and scattering IMFP, λi,abs1\lambda_{i,abs}^{-1} and λi,sca1\lambda_{i,sca}^{-1}, and define the total IMFP as

λi1(E,r)λi,abs1(E,r)(1±eE/T(r))+λi,sca1(E,r).\begin{split}\lambda_{i}^{-1}(E,r)&\equiv\lambda_{i,abs}^{-1}(E,r)(1\pm e^{-E/T(r)})\\ &\quad+\lambda_{i,sca}^{-1}(E,r).\end{split} (11)

As in the nondiffuse regime, we compute the energy fluxes of the longitudinal and transverse dark photon separately, and assume that the energy fluxes of χ\chi and χ¯\bar{\chi} are equal. Hence, the total energy loss rate in the DS particles is Ltot=LA,L+LA,T+2LχL_{tot}=L_{A^{\prime},L}+L_{A^{\prime},T}+2L_{\chi}. The IMFP calculations can be found in Appendix VIII.

We now estimate the relevant parameter space region in the limit where the DS self-interaction is the dominant opacity source. When mA>2mχm_{A^{\prime}}>2m_{\chi}, the dominant DS self-interaction is the (inverse) decay Aχχ¯A^{\prime}\leftrightarrow\chi\bar{\chi}. Given Tν3.9 MeVT_{\nu}\simeq 3.9\mbox{ MeV} and the temperature gradient |dT/dr|Rν6.2×104 MeV/m|dT/dr|_{R_{\nu}}\simeq 6.2\times 10^{-4}\mbox{ MeV}/\mbox{m}, if we fix mA/mχ=3m_{A^{\prime}}/m_{\chi}=3, the luminosities of the DS particles can be fitted by

LALν×(6.3×1012αD)(MeVmA)2×exp[49.0+(mA4.3MeV)2]\begin{split}L_{A^{\prime}}&\simeq L_{\nu}\times\left(\frac{6.3\times 10^{-12}}{\alpha_{D}}\right)\left(\frac{\rm MeV}{m_{A^{\prime}}}\right)^{2}\\ &\quad\times\exp\left[-\sqrt{49.0+\left(\frac{m_{A^{\prime}}}{4.3~{}\rm MeV}\right)^{2}}\right]\end{split} (12)

for mA25m_{A^{\prime}}\lesssim 25 MeV, and

Lχ+Lχ¯Lν×(2.1×1013αD)(MeVmχ)2×exp[(mχ1.9MeV)49.0+(mχ4.3MeV)2]\begin{split}L_{\chi}+L_{\bar{\chi}}&\simeq L_{\nu}\times\left(\frac{2.1\times 10^{-13}}{\alpha_{D}}\right)\left(\frac{\rm MeV}{m_{\chi}}\right)^{2}\\ &\quad\times\exp\left[\left(\frac{m_{\chi}}{1.9~{}\rm MeV}\right)-\sqrt{49.0+\left(\frac{m_{\chi}}{4.3~{}\rm MeV}\right)^{2}}\right]\end{split} (13)

for mχ30m_{\chi}\lesssim 30 MeV. If we take mA=9m_{A^{\prime}}=9 MeV and mχ=3m_{\chi}=3 MeV, then the total DS luminosity is

LDLν×(1.5×1016αD).L_{D}\simeq L_{\nu}\times\left(\frac{1.5\times 10^{-16}}{\alpha_{D}}\right). (14)

The corresponding χχ¯A\chi\bar{\chi}\rightarrow A^{\prime} cross section is666The estimated value here is the thermally averaged cross section evaluated at the center of PNS σχχ¯Aλχχ¯A1(r=0)/nχ(r=0)\sigma_{\chi\bar{\chi}\rightarrow A^{\prime}}\equiv\langle\lambda^{-1}_{\chi\bar{\chi}\rightarrow A^{\prime}}(r=0)\rangle/n_{\chi}(r=0).

σχχ¯A4.7×1041 cm2(αD1.5×1016).\sigma_{\chi\bar{\chi}\rightarrow A^{\prime}}\simeq 4.7\times 10^{-41}\mbox{ cm}^{2}\left(\frac{\alpha_{D}}{1.5\times 10^{-16}}\right). (15)

When mA<2mχm_{A^{\prime}}<2m_{\chi}, the diffuse luminosities of dark photons and dark fermions depend on the values of ϵ\epsilon and αD\alpha_{D}. For regions where ϵ\epsilon is small enough such that the dominating opacity source for dark photons is DS self-interaction AAχχ¯A^{\prime}A^{\prime}\rightarrow\chi\bar{\chi}, one may estimate the DS luminosity by considering the dark photon luminosity only. If we fix mA/mχ=1/3m_{A^{\prime}}/m_{\chi}=1/3, then the dark photon luminosity can be fitted by

LALν×(2.0×107αD)2×exp[(mA0.74MeV)49.0+(mA4.3MeV)2]\begin{split}L_{A^{\prime}}&\simeq L_{\nu}\times\left(\frac{2.0\times 10^{-7}}{\alpha_{D}}\right)^{2}\\ &\quad\times\exp\left[\left(\frac{m_{A^{\prime}}}{0.74~{}\rm MeV}\right)-\sqrt{49.0+\left(\frac{m_{A^{\prime}}}{4.3~{}\rm MeV}\right)^{2}}\right]\end{split} (16)

for mA25m_{A^{\prime}}\lesssim 25 MeV. Adopting mA=3 MeVm_{A^{\prime}}=3\mbox{ MeV} and mχ=9 MeVm_{\chi}=9\mbox{ MeV}, this gives

LDLν×(4.5×108αD)2.L_{D}\simeq L_{\nu}\times\left(\frac{4.5\times 10^{-8}}{\alpha_{D}}\right)^{2}. (17)

The corresponding χχχχ\chi\chi\rightarrow\chi\chi cross section is777The cross section is evaluated in the same way as Eq. (15).

σχχχχ9.7×1037 cm2(αD4.5×108)2,\sigma_{\chi\chi\rightarrow\chi\chi}\simeq 9.7\times 10^{-37}\mbox{ cm}^{2}\left(\frac{\alpha_{D}}{4.5\times 10^{-8}}\right)^{2}, (18)

and the corresponding AAχχ¯A^{\prime}A^{\prime}\rightarrow\chi\bar{\chi} cross section is

σAAχχ¯3.1×1039 cm2(αD4.5×108)2.\sigma_{A^{\prime}A^{\prime}\rightarrow\chi\bar{\chi}}\simeq 3.1\times 10^{-39}\mbox{ cm}^{2}\left(\frac{\alpha_{D}}{4.5\times 10^{-8}}\right)^{2}. (19)

These analytical formulas Eqs. (14)–(19) already illustrate that DS particles can be self-trapped by interaction cross sections of 𝒪(1040)\gtrsim\mathcal{O}(10^{-40}) cm2 such that the corresponding diffuse DS luminosity LDLνL_{D}\lesssim L_{\nu}. Only very weakly interacting DS particle with αD\alpha_{D} much smaller than those nominal values in Eqs. (14)–(19) can result in LD>LνL_{D}>L_{\nu}.

III.3 Diffuse criteria

Whether DS particles are in the nondiffuse or in the diffuse regime sensitively depends on their abundances in the PNS. For example, if the dark fermion abundance is significant, the contribution of their pair-absorption processes (e.g., χχ¯npnp\chi\bar{\chi}np\rightarrow np, χχ¯ee+\chi\bar{\chi}\rightarrow e^{-}e^{+}, and χχ¯γ\chi\bar{\chi}\rightarrow\gamma^{\ast}) to the dark fermion optical depth cannot be neglected. Furthermore, the DS abundances lead to significant DS self-interactions (e.g. χχχχ\chi\chi\rightarrow\chi\chi, χχ¯AA\chi\bar{\chi}\rightarrow A^{\prime}A^{\prime}, χAAχ\chi A^{\prime}\rightarrow A^{\prime}\chi,…) that can trap themselves in the PNS. These DS self-interactions delay the escape times for both the dark photon and the dark fermion, which enhances their abundances and optical depths. Therefore, the abundances of DS particles is critical in determining whether self-trapping is important. Below, we formulate the diffuse criteria in terms of the DS abundances and their IMFP at the neutrinosphere.

To obtain the exact DS abundances, detailed transport incorporating their production, scattering, and absorption needs be solved. Here we roughly estimate the abundance of each particle species by their production rate and a relevant timescale. The abundance of a DS particle species ii is approximated as

Ni=0Rν4πr2𝑑rd3pi(2π)3giΓi,prodΔti,\begin{split}N_{i}&=\int_{0}^{R_{\nu}}4\pi r^{2}dr\int\frac{d^{3}p_{i}}{(2\pi)^{3}}g_{i}\Gamma_{i,\mathrm{prod}}\Delta t_{i},\\ \end{split} (20)

where the quantity Δti\Delta t_{i} is subject to the following considerations. First, the longest possible time for DS particles to accumulate is the cooling time of the PNS. We set this upper bound of Δti\Delta t_{i} as tcool=1t_{\rm cool}=1 s. Second, the shortest possible timescale, i.e., the lower bound of Δti\Delta t_{i}, is estimated by the free-escaping timescale tfreeRν/vit_{\rm free}\equiv R_{\nu}/v_{i}. Another characteristic timescale here in the limit where a DS particle is trapped is the diffusion timescale ti,diffRν2/Dit_{i,\rm diff}\simeq R_{\nu}^{2}/D_{i} where Divi/λi1D_{i}\simeq v_{i}/\lambda^{-1}_{i} is the diffusion coefficient (see below for details). For tfree<ti,diff<tcoolt_{\rm free}<t_{i,\rm diff}<t_{\rm cool}, we then take ti,difft_{i,\rm diff} as Δti\Delta t_{i}. Combining these criteria, the relevant timescale Δti\Delta t_{i} used in Eq. (20) is given by

Δti={tfree,ifti,difftfree,ti,diff,iftfree<ti,diff<tcool,tcool,iftcoolti,diff.\Delta t_{i}=\begin{cases}t_{\rm free},\,{\rm if}~{}t_{i,\rm diff}\leq t_{\rm free},\\ t_{i,\rm diff},\,{\rm if}~{}t_{\rm free}<t_{i,\rm diff}<t_{\rm cool},\\ t_{\rm cool},\,{\rm if}~{}t_{\rm cool}\leq t_{i,\rm diff}.\end{cases} (21)

Since the PNS is not homogeneous, the diffusion of DS particles cannot be described by a constant DiD_{i}. As shown in Tables 1 and 2, the interactions relevant to the IMFP of DS particles can be categorized as SM-type or DS-type by particles involved in the interactions. We choose particular locations in the PNS where the IMFPs of each type are maximal to roughly estimate their contribution to the diffusion timescale.888Overestimating the IMFP makes Eq. (24) easier to satisfy. Because the luminosity in the diffuse regime is generally lower than that in the nondiffuse regime for the same parameters, our estimation leads to a conservative bound. For the SM-type interactions, we estimate the IMFP at the center (rSM=0kmr_{\rm SM}=0~{}\mathrm{km}) of the PNS where the density of SM particles is the largest. As for the DS-type interactions, we estimate the IMFP at rDS=11kmr_{\rm DS}=11\mathrm{km} where the temperature is the highest. This location has the largest DS particle density, if DS particles are in thermal equilibrium with the SM medium. Knowing the IMFPs of both types at these respective locations, we can then compute the diffusion timescale

ti,diffRν2vi[λ~i,SM1(rSM)+λ~i,DS1(rDS)]t_{i,\rm diff}\equiv\frac{R^{2}_{\nu}}{v_{i}}\left[\langle\tilde{\lambda}^{-1}_{i,SM}(r_{SM})\rangle+\langle\tilde{\lambda}^{-1}_{i,DS}(r_{DS})\rangle\right] (22)

where λ~i,SM1\tilde{\lambda}^{-1}_{i,SM} and λ~i,DS1\tilde{\lambda}^{-1}_{i,DS} are SM-type and DS-type contribution to the IMFP of particle ii. When computing the IMFP contribution from DS particles, λ~i,DS1\tilde{\lambda}^{-1}_{i,DS} in Eq. (22), we take the assumption that DS particles are in full thermal equilibrium with the SM medium. Combining Eqs. (20)–(22) then allows us to compute the DS abundance NiN_{i} for particle ii.

After deriving NiN_{i}, we then define a scaling factor ηiNi/Nieq\eta_{i}\equiv N_{i}/N_{i}^{\rm eq} for each dark particle species ii for Ni<NieqN_{i}<N_{i}^{\rm eq} (ηi=1\eta_{i}=1 if NiNeqN_{i}\geq N_{\rm eq}), where

Nieq0Rν4πr2𝑑rd3pi(2π)3gifi(Ei,T(r))\begin{split}N_{i}^{\rm eq}\equiv\int_{0}^{R_{\nu}}4\pi r^{2}dr\int\frac{d^{3}p_{i}}{(2\pi)^{3}}g_{i}f_{i}(E_{i},T(r))\end{split} (23)

is the corresponding total equilibrium abundance of ii. These scaling factors are then used to better estimate the IMFP, λi,DS1\lambda^{-1}_{i,DS} due to the DS particles self-interaction, used in Eq. (10) for computing the DS diffuse luminosities. We also use ηi\eta_{i} to calculate the thermally averaged total IMFP of ii at RνR_{\nu}, λi1(Rν)\langle{\lambda}^{-1}_{i}(R_{\nu})\rangle.

We are now finally ready to write down our diffuse criteria. We say that a DS particle species ii can be treated as in diffuse limit if the following conditions are satisfied:

Ni>Nieq,andλi1(Rν)>Rν1.N_{i}>N_{i}^{\rm eq}{,~{}\rm and~{}}\langle{\lambda}^{-1}_{i}(R_{\nu})\rangle>R_{\nu}^{-1}. (24)

The first condition ensures that a DS particle species ii is only considered to be diffusive when their production is efficient enough such that their amount can exceed the equilibrium number within the relevant timescale Δti\Delta t_{i}. The second condition requires that the IMFP of ii at RνR_{\nu}, where both the density and temperature are the lowest in the PNS, is large enough to trap particle ii inside the PNS (see Fig. 1).999Note that in computing the second criterion, we exclude the contribution from the dark photon decay via Aee+A^{\prime}\rightarrow e^{-}e^{+} This is because including this decay process leads to an artificially enhancement of the IMFP by several orders of magnitudes for dark photon heavier than 30\simeq 30 MeV, as the electron chemical potential μe15\mu_{e}\simeq 15 MeV at RνR_{\nu}. However, this process should be strongly Pauli-blocked for mA200m_{A^{\prime}}\lesssim 200 MeV in most region inside the PNS (see Fig. 1) where dark photons are most abundant in the diffuse limit. When Eq. (24) is satisfied, we use Eq. (9) to compute the DS diffuse luminosity for ii. Otherwise, we take the nondiffuse luminosities, Eqs. (3) and (6) for dark photons and dark fermions, respectively. Note that in the diffuse regime and when Δti=ti,diff\Delta t_{i}=t_{i,\rm diff}, the first condition in Eq. (24) is approximately equivalent to having the characteristic thermalization IMFP Keil et al. (2003); Shapiro and Teukolsky (1983) of DS particles with SM medium, λi,SM,abs1×λi,total1\sqrt{\lambda^{-1}_{i,SM,\rm abs}\times\lambda^{-1}_{i,\rm total}} averaged over the phase space and spatial volume inside RνR_{\nu}, being larger than Rν1R_{\nu}^{-1}. This means that the energy exchange between the DS particles and the SM medium, enhanced by the DS self-interactions, is efficient enough to keep themselves in thermal contact with SM medium.

We provide in Appendix IX a detailed work flow to additionally describe how we use the results in these sections to compute the dark sector luminosity for interested readers.

III.4 Numerical calculations

Refer to caption
Refer to caption
Figure 2: Luminosity of different dark sector particles as a function of αD\alpha_{D} for a fixed value of ϵ=108\epsilon=10^{-8}. In the left (right) panel, mA=9 MeVm_{A^{\prime}}=9\mbox{ MeV} and mχ=3 MeVm_{\chi}=3\mbox{ MeV} (mA=3 MeVm_{A^{\prime}}=3\mbox{ MeV} and mχ=9 MeVm_{\chi}=9\mbox{ MeV}). The horizontal dotted lines label the neutrino luminosity Lν=3×1052 erg/sL_{\nu}=3\times 10^{52}\mbox{ erg/s} used to set the supernova bound [Eq. (2)]. The vertical dashed lines indicate the transition from the nondiffuse (smaller αD)\alpha_{D}) to the diffuse (larger αD)\alpha_{D}) regime (see text for details). Note that χ\chi is in the nondiffuse limit for the range of αD\alpha_{D} shown in the right panel.
Refer to caption
Refer to caption
Figure 3: Luminosity of different dark sector particles as a function of ϵ\epsilon for fixed values of αD=1017\alpha_{D}=10^{-17} (left), 10910^{-9} (right). In the left (right) panel, mA=9 MeVm_{A^{\prime}}=9\mbox{ MeV} and mχ=3 MeVm_{\chi}=3\mbox{ MeV} (mA=3 MeVm_{A^{\prime}}=3\mbox{ MeV} and mχ=9 MeVm_{\chi}=9\mbox{ MeV}). The horizontal dotted lines label the neutrino luminosity Lν=3×1052 erg/sL_{\nu}=3\times 10^{52}\mbox{ erg/s} used to set the supernova bound [Eq. (2)]. The vertical dashed lines indicate the transition from the nondiffuse (smaller ϵ\epsilon) to the diffuse (larger ϵ\epsilon) regime (see text for details). Note that χ\chi is in the nondiffuse limit for the range of ϵ\epsilon shown in the left panel.

We now apply formulas derived in previous sections to compute the dark photon and dark fermion luminosities and examine how they depend on the coupling constants ϵ\epsilon and αD\alpha_{D}. Figure 2 shows the luminosities of different DS particles as functions of αD\alpha_{D} with ϵ=108\epsilon=10^{-8} for two different choices of DS masses. In the left panel, the masses are chosen such that the decay process Aχχ¯A^{\prime}\rightarrow\chi\bar{\chi} is allowed. In this scenario, dark photons of different polarization modes are trapped diffusively when αD5×1018\alpha_{D}\gtrsim 5\times 10^{-18}, while the dark fermions are diffusive when αD1017\alpha_{D}\gtrsim 10^{-17}. The nondiffuse luminosity of the dark photon for small αD\alpha_{D} mainly depends on the interaction with the SM particles and thus is independent of αD\alpha_{D}.101010The difference of 102.5\sim 10^{2.5} between the luminosities of the longitudinal and transverse dark photon is due to the plasma effects.. The dark fermion luminosity is proportional to αD\alpha_{D} in the diffuse limit, as analyzed in Appendix VI However, the dark fermion luminosity for αD5×1018\alpha_{D}\lesssim 5\times 10^{-18} is set to zero to avoid double counting because they are predominantly produced through the decay of on-shell dark photons (see discussions in Sec. III.1). In the diffuse regime, the luminosities of the DS particles are proportional to αD1\alpha_{D}^{-1} due to the self-trapping interactions effectively dominated by the (inverse) decay process Aχχ¯A^{\prime}\leftrightarrow\chi\bar{\chi}.

For the right panel in Fig. 2, we choose DS masses such that the decay process Aχχ¯A^{\prime}\rightarrow\chi\bar{\chi} is not allowed. In this case, dark photons of different polarization modes are in diffuse regime when αD8×109\alpha_{D}\gtrsim 8\times 10^{-9}, while dark fermions are always in the nondiffuse regime for the range of αD\alpha_{D} shown in the plot. The main reason leading to several orders of magnitude larger difference in αD\alpha_{D} here than the previous case is due to the extra αD\alpha_{D} dependence in the IMFP of DS self-interactions [see, e.g., Table 1 or Eqs. (14) and (17)]. Similar to the previous scenario, the nondiffuse luminosity of the dark photon is independent of αD\alpha_{D}. However, the luminosities in the diffuse regime scale as αD2\alpha_{D}^{-2}. Once again, this is because the dominant interaction is responsible for trapping the dark photon being AAχχ¯A^{\prime}A^{\prime}\rightarrow\chi\bar{\chi}, whose interaction rate is proportional to αD2\alpha_{D}^{2}. The dark fermion luminosity is significantly smaller than the dark photon luminosity for the range of αD\alpha_{D} in the plot, because the dark fermion production rate is proportional to ϵ2αD\epsilon^{2}\alpha_{D}, suppressed by an extra factor of αD\alpha_{D} when compared with the dark photon production rate.

In Fig. 3, we show the luminosities of the DS particles as functions of ϵ\epsilon with two choices of αD\alpha_{D} and same DS masses as those in Fig. 2. For the case with αD=1017\alpha_{D}=10^{-17} in the left panel where the decay process Aχχ¯A^{\prime}\rightarrow\chi\bar{\chi} is again allowed, the longitudinal (transverse) dark photons are trapped when ϵ3×109\epsilon\gtrsim 3\times 10^{-9} (ϵ2×109\epsilon\gtrsim 2\times 10^{-9}). For dark fermions, they just free stream independent of the value of ϵ\epsilon, because of the small value of αD\alpha_{D} (cf., Fig. 2). The nondiffuse luminosities of the dark photons of both modes are proportional to ϵ2\epsilon^{2} as determined by their production rates. The dark fermion luminosity for ϵ2×109\epsilon\lesssim 2\times 10^{-9} is set to zero for the same reason discussed above. The diffuse luminosities of dark photons are affected by both the interactions with the SM particles and the DS self-interactions. For ϵ107\epsilon\gtrsim 10^{-7}, dark photon–SM interactions dominate over DS self-interactions, so the dark photon luminosities are proportional to ϵ2\epsilon^{-2} as determined by the mean free path of the dark photon–SM interactions. Interestingly, the dark fermion luminosity becomes independent of ϵ\epsilon and can thus remain larger than LνL_{\nu} for large ϵ\epsilon. This is because the branching ratio of the dark photon absorption processes approaches to 11, such that dark fermions can be produced via the decay of trapped dark photons (see Appendix VI).

For the right panel in Fig. 3, we choose αD=109\alpha_{D}=10^{-9} for the case where Aχχ¯A^{\prime}\rightarrow\chi\bar{\chi} is not allowed. There, longitudinal (transverse) dark photons are diffusively trapped when ϵ3×107\epsilon\gtrsim 3\times 10^{-7} (ϵ2×107\epsilon\gtrsim 2\times 10^{-7}), while dark fermions are in the diffuse limit when ϵ2×105\epsilon\gtrsim 2\times 10^{-5}. The dark photon luminosities are proportional to ϵ2\epsilon^{2} for ϵ109\epsilon\lesssim 10^{-9}, where the optical depths are negligible. The diffuse luminosities of dark photons scale as ϵ2\epsilon^{-2} because the absorption processes by the SM medium are dominant. The dark fermion luminosity is much smaller than that of the dark photons until ϵ105\epsilon\gtrsim 10^{-5}, because the dark fermion production rate is suppressed by an additional factor of αD\alpha_{D} when compared with the dark photon production rate. The diffuse luminosity of the dark fermion is proportional to ϵ2\epsilon^{-2} for ϵ104\epsilon\gtrsim 10^{-4}, where χeχe\chi e^{-}\rightarrow\chi e^{-} dominates the IMFP at RνR_{\nu}.

Refer to caption
Refer to caption
Figure 4: Dark sector luminosity LDL_{D} contours in the αD\alpha_{D}-ϵ\epsilon plane for mA=9 MeVm_{A^{\prime}}=9\mbox{ MeV}, mχ=3 MeVm_{\chi}=3\mbox{ MeV} (left) and mA=3 MeVm_{A^{\prime}}=3\mbox{ MeV}, mχ=9 MeVm_{\chi}=9\mbox{ MeV} (right). Solid black (brown) curves indicate where LD=LνL_{D}=L_{\nu} (LD=0.1LνL_{D}=0.1L_{\nu}). Regions inside the LD=LνL_{D}=L_{\nu} are excluded by the supernova bound. Thin dash lines show where the transition from the nondiffuse to diffuse regime occur. See text for discussions on regions labeled (I), (II), and (A), (B).

IV Cooling bounds on self-interacting dark sector particles

In this section, we examine the excluded parameter space of the DS. Since the model has four free parameters, mAm_{A^{\prime}}, mχm_{\chi}, ϵ\epsilon, and αD\alpha_{D}, we choose to project the exclusion contours on various combinations of these parameters. As discussed in the previous sections, whether the decay of Aχχ¯A^{\prime}\rightarrow\chi\bar{\chi} is allowed affects the DS luminosity significantly. Here we choose two benchmark mass ratios mA/mχ=3m_{A^{\prime}}/m_{\chi}=3 and 1/31/3 to investigate the SN bounds for these two scenarios.

We first show the contours of the DS luminosity LD=LνL_{D}=L_{\nu} and LD=0.1LνL_{D}=0.1L_{\nu} on the αD\alpha_{D}ϵ\epsilon plane with fixed mAm_{A^{\prime}} and mχm_{\chi} in Fig. 4. The chosen masses are mA=9m_{A^{\prime}}=9 MeV and mχ=3m_{\chi}=3 MeV (mA=3m_{A^{\prime}}=3 MeV and mχ=9m_{\chi}=9 MeV) for the left (right) panel in the figure, which allows (forbids) Aχχ¯A^{\prime}\rightarrow\chi\bar{\chi}. Regions inside the LD=LνL_{D}=L_{\nu} contours indicate that the cooling bound [Eq. (2)] is violated. In the left panel where Aχχ¯A^{\prime}\rightarrow\chi\bar{\chi} is allowed, the excluded region exhibits an L shape, covering the parameter ranges of (I) αD1016\alpha_{D}\lesssim 10^{-16} for 1010ϵ10610^{-10}\lesssim\epsilon\lesssim 10^{-6}, as well as (II) 1019αD101610^{-19}\lesssim\alpha_{D}\lesssim 10^{-16} for ϵ106\epsilon\gtrsim 10^{-6}. For the majority of the parameter space in (I) (αD<1016\alpha_{D}<10^{-16} and 1010<ϵ<10610^{-10}<\epsilon<10^{-6}), the DS luminosity is mainly contributed by the nondiffuse dark photons (see also Figs. 2 and 3). In (II) (1019αD101610^{-19}\lesssim\alpha_{D}\lesssim 10^{-16} for ϵ106\epsilon\gtrsim 10^{-6}), LDL_{D} is mainly contributed by the nondiffuse dark fermions (see also Fig. 4). Close to the upper edge of the LD=LνL_{D}=L_{\nu} contour in both regions, self-trapping of DS particle takes effect such that LDαD1L_{D}\propto\alpha_{D}^{-1} decreases with increasing αD\alpha_{D} (see Fig. 2), leading to the horizontal edge at αD1016\alpha_{D}\sim 10^{-16}. The lower bound of αD1019\alpha_{D}\sim 10^{-19} for region (II) is due to the inefficient production of dark fermions. For the right panel in Fig. 4 where Aχχ¯A^{\prime}\rightarrow\chi\bar{\chi} is not allowed, two regions similarly exist. Region (A) with 109ϵ10610^{-9}\lesssim\epsilon\lesssim 10^{-6} and αD<108\alpha_{D}<10^{-8} receives dominant contribution from nondiffuse dark photons as region (I) in the left panel (see also Figs. 2 and 3). Similarly, the self-trapping of DS particles defines the upper edge of LD=LνL_{D}=L_{\nu} at αD108\alpha_{D}\sim 10^{-8}. For the narrow diagonal-shape region (B) at ϵ105\epsilon\gtrsim 10^{-5}, LDL_{D} is dominated by dark fermions. This shape is related to the fact that both the production and pair-absorption rates of dark fermions from the SM medium are proportional to ϵ2αD\epsilon^{2}\alpha_{D}, as discussed also in Sec. III.4.

Refer to caption
Refer to caption
Figure 5: The LD=LνL_{D}=L_{\nu} contour on the αD\alpha_{D}-mχm_{\chi} plane for a fixed value of ϵ=108\epsilon=10^{-8} and mass ratios mA/mχ=3m_{A^{\prime}}/m_{\chi}=3 (left), 1/31/3 (right). Regions below the contours are excluded by the SN bound. Above the contours, the self-trapping of dark sector particles limit their luminosities and cannot be excluded by the supernova bound.

For both scenarios shown in Fig. 4, there are specific values of αD\alpha_{D} above which the cooling criterion gives no constraint due to DS self-trapping. We now investigate the dependence of these critical values of αD\alpha_{D} on DS particle masses. Figure 5 shows the excluded regions on the αD\alpha_{D}-mχm_{\chi} plane with ϵ=108\epsilon=10^{-8} for fixed mass ratios mA=3mχm_{A^{\prime}}=3m_{\chi} (left panel) and 3mA=mχ3m_{A^{\prime}}=m_{\chi} (right panel). Note that regions below the contours are excluded. In the left panel where the decay Aχχ¯A^{\prime}\rightarrow\chi\bar{\chi} is allowed, the SN bound only weakly depends on αD\alpha_{D} for mχ10m_{\chi}\lesssim 10 MeV. On the other hand, αD\alpha_{D} increase sharply with mχm_{\chi} for mχ10m_{\chi}\gtrsim 10 MeV. The main reason is the DS particle abundances are insensitive to the mass for mχ30m_{\chi}\ll 30 MeV which is the typical temperature inside the PNS. Larger mχm_{\chi} (and mA)m_{A^{\prime}}) leads to smaller DS abundances, which in turn gives rise to a smaller IMFP of DS self-trapping for a given αD\alpha_{D}. Thus, the suppression of LDL_{D} due to DS self-trapping occurs at a larger αD\alpha_{D} for larger mχm_{\chi} (and mA)m_{A^{\prime}}). For mχ70m_{\chi}\gtrsim 70 MeV (mA210m_{A^{\prime}}\gtrsim 210 MeV), no SN bounds can be placed due to the inefficient production of DS particles. In the right panel where 3mA=mχ3m_{A^{\prime}}=m_{\chi}, the bound is completely determined by the dark photon only because the dark fermion production is further suppressed by an extra factor of αD\alpha_{D}. Here, the critical αD\alpha_{D} also increases with mχm_{\chi} sharply for mχ10m_{\chi}\gtrsim 10 MeV as in the left panel. Once again, this is resulting from smaller dark photon abundance for larger mχm_{\chi} (thus mAm_{A}^{\prime}) inside the PNS. It thus requires a larger αD\alpha_{D} for AAχχ¯A^{\prime}A^{\prime}\rightarrow\chi\bar{\chi} to self-trap the dark photons. Note that the maximal mχ800m_{\chi}\simeq 800 MeV below which SN bound exists corresponds to a maximal mA250m_{A}^{\prime}\simeq 250 MeV, consistent with the maximal mAm_{A}^{\prime} in the left panel.

Refer to caption
Refer to caption
Figure 6: The excluded regions (enclosed by the solid curves) on the ϵ\epsilon-mAm_{A^{\prime}} plane for different αD\alpha_{D} values, with fixed mass ratios mA/mχ=3m_{A^{\prime}}/m_{\chi}=3 (left), 1/31/3 (right). Note that we include bounds from considering the degree of freedom of dark photon only (dotted black curves) for comparison. The thin gray curve in the right panel shows regions excluded by the terrestrial experiments extracted from Andreas et al. (2012). These curves show that even with tiny values of αD\alpha_{D}, supernova bounds on dark photon parameter space can be largely affected.

The above examples clearly demonstrated how efficiently a small αD\alpha_{D} within the self-interacting DS can affect the SN bound. In Fig. 6, we further show the excluded regions in the ϵ\epsilon-mAm_{A^{\prime}} plane with two different values of αD\alpha_{D} for each choice of the DS mass ratio. Also shown are the bounds derived by considering the dark photon degree of freedom only, as well as the existing experimental constraints on dark photon (extracted from Ref. Andreas et al. (2012)). We refrain from showing other astrophysical or cosmological bounds on dark photon of similar masses; see e.g., Refs. Jaeckel et al. (2008); Mirizzi et al. (2009); Redondo and Raffelt (2013); An et al. (2013); Fradette et al. (2014); Sung et al. (2019); DeRocco et al. (2019a); An et al. (2020); Li et al. (2020); Sieverding et al. (2021). These plots show once again that even for small values of αD=1012\alpha_{D}=10^{-12} (left panel) and 10510^{-5} (right panel), the SN exclusion regions shrink significantly due to the self-trapping effects when compared to results derived by considering only dark photons without self-trapping. However, for very small values of αD\alpha_{D}, e.g., 101710^{-17} and 10910^{-9} shown in the left and right panels, respectively, the cooling bounds can be extended to larger ϵ\epsilon. This is due to the contribution of dark fermions through the decay of trapped dark photons, as discussed in earlier sections. Note that for αD=1012\alpha_{D}=10^{-12} in the left panel, the excluded region slightly extends to larger values of mAm_{A^{\prime}} for ϵ107\epsilon\gtrsim 10^{-7}. This is because the dark fermion luminosity depends on both the decay rate of Aχχ¯A^{\prime}\rightarrow\chi\bar{\chi} and the blackbody energy density of dark photons (see Appendix VI). The larger decay rate can compensate the smaller dark photon energy density with increasing mAm_{A^{\prime}}, provided that the branching ratio of the dark photon absorption processes is close to 11.

V Conclusion

In this work, we examined the SN bounds on self-interacting dark sector particles. Adopting a dark photon portal dark sector, we derived the relevant interaction cross sections and (inverse) decay rates for reactions listed in Tables 1 and 2. We then used these to compute the energy luminosities of dark sector particles in the nondiffuse and diffuse regimes separately, and formulated a simple criterion to connect these two regimes. The self-interaction of dark sector particles can efficiently trap themselves inside the proto-neutron star and thus suppress their energy luminosities.

Comparing the dark sector luminosity with the neutrino luminosity inferred from the SN1987a event, we derived SN bounds for two assumed dark photon to dark fermion mass ratios, mA/mχ=3m_{A^{\prime}}/m_{\chi}=3 and 1/31/3, which represent scenarios where Aχχ¯A^{\prime}\rightarrow\chi\bar{\chi} is allowed or not. For the former (later) case with mA/mχ=3m_{A^{\prime}}/m_{\chi}=3 (mA/mχ=1/3m_{A^{\prime}}/m_{\chi}=1/3), SN bounds only apply to weakly interacting dark sectors whose dark fine structure constant αD1015\alpha_{D}\lesssim 10^{-15} (107\lesssim 10^{-7}), for mχ𝒪(20)m_{\chi}\lesssim\mathcal{O}(20) MeV (see Figs. 4 and 5). The dominating dark sector cross sections for these αD\alpha_{D} values correspond to 1040\simeq 10^{-40} cm2. In particular, there is no SN bound for the former case for αD107\alpha_{D}\gtrsim 10^{-7} (Fig. 5). Our results differ from the previous analysis Chang et al. (2018) considering similar models. This is because Ref. Chang et al. (2018) assumed that the DS particles decouple from the SM medium at the surface where the χ\chi-p scattering becomes inefficient, and ignored the DS self-interactions which can trap themselves and help thermalize the DS particles with the SM medium.

Although the exact excluded regions in the DS parameter space should also depend on the chosen value of mA/mχm_{A^{\prime}}/m_{\chi}, which are unexplored in this work, our results demonstrated that when applying the supernova bounds to dark sector particles, their self-interactions, which can evade the bounds, must be taken into considerations. Our results here also imply that other stellar bounds, e.g., from the horizontal branch stars, tips of red giants, or white dwarfs, on dark sector particles may also be sensitive to the structure of the dark sector. Similarly, our results also indicate that for nonstandard strongly self-interacting neutrinos proposed to resolve the Hubble tension Kreisch et al. (2020), the needed strong self-interaction, 106\sim 10^{6} or 10910^{9} times stronger than the SM weak interaction, will likely lead to self-trapping of neutrinos and results in inconsistency with the SN1987a observation [cf. Eqs. (15) and (19)]. Moreover, we would like to point out that although the self-interaction of dark sector can completely evade the SN bound, new constraints may be further derived by considering their potential late-time heating to the remnant NS via decays or annihilations in a longer timescale. Furthermore, such self-trapping effect might provide an efficient mechanism to produce DM-admixed NS, which might have implication for the GW detection of binary NS merger events Ellis et al. (2018); Nelson et al. (2019); Bauswein et al. (2020). All these aspects are beyond the scope of this paper and can be further investigated in future work. We also note that our results indicate that self-interacting DM that can help solve the small-scale issues in galaxies cannot be excluded by the SN cooling bound, as the required χ\chi-χ\chi self-interaction cross section σχχ/mχ𝒪(1)\sigma_{\chi\chi}/m_{\chi}\sim\mathcal{O}(1) cm2 g-1 largely exceeds values that can be constrained by SN cooling [cf. Eq. (18)].

Finally, we comment on potential caveats in this work. First, our criterion of switching from the nondiffuse to diffuse regime is rather abrupt and can sometimes create non-negligible discontinuity in dark sector luminosities as shown in e.g., Figs. 2 and 3. In reality, the transition should be smooth and this can possibly introduce errors of a factor of a few in all our derived bounds. For example, while evaluating the luminosities in the nondiffuse regime, we only included absorption and decay processes in the opacity. In principle, scattering can somewhat reduce the energy luminosity of dark sector particles leaving the PNS, before the condition for diffusion is fully satisfied. Also, we used a sharp neutrinosphere as a boundary to estimate the luminosities in the diffuse regime. This may lead to some errors when the dark sector particles are not fully in the diffuse limit. Second, when we evaluated the diffusion timescale used to determine the diffuse criterion, we selected two specific locations where the IMFPs are largest for simplicity. This approximation may overestimate the diffusion time a bit. All these sources of uncertainties can only be addressed by performing a full numerical calculation of multidimensional Boltzmann transport and can be pursued in future. However, the main conclusion derived in this work – self-interactions inside the dark sector can crucially affect stellar bounds – should remain relatively solid and needs to be considered in all relevant studies.

Acknowledgements.
We thank Jae Hyeok Chang, Yen-Hsun Lin, Huitzu Tu, and Tse-Chun Wang for useful discussions. We also thank Luc Darme, Sam McDermott, Sunny Vagnozzi, and an anonymous referee for providing feedback during the reviewing process, which help improve this work. This work was supported in part by the Ministry of Science and Technology, Taiwan under Grant No. 108-2112-M-001-010, No. 109-2112-M-001-004, the Academia Sinica under Project No. AS-CDA-109-M11, and the Physics Division of National Center for Theoretical Sciences.

VI Narrow Width Approximation

Consider a process χχ¯Af1f2\chi\bar{\chi}\rightarrow A^{\prime\ast}\rightarrow f_{1}f_{2}... when mA>2mχm_{A^{\prime}}>2m_{\chi} so that the intermediate AA^{\prime} can be on shell. Let k=(ω,k)k=(\omega,\vec{k}) be the momentum of the intermediate AA^{\prime} and p=(E,p)p=(E,\vec{p}), p=(E,p)p^{\prime}=(E^{\prime},\vec{p}^{\prime}) be the momentum of the initial χ\chi and χ¯\bar{\chi} respectively. The spin-averaged amplitude squared is given by

||2¯=gD2k22{|ϵLJ|2(k2mA2)2+(ωΓL)2[1(ω2E|k|)2]+λ|ϵTλJ|2(k2mA2)2+(ωΓT)2[12+2mχ2k2+12(ω2E|k|)2]},\begin{split}\overline{|\mathcal{M}|^{2}}=g_{D}^{2}\cdot\frac{k^{2}}{2}&\left\{\frac{|\epsilon_{L}\cdot J|^{2}}{(k^{2}-m_{A^{\prime}}^{2})^{2}+(\omega\Gamma_{L})^{2}}\left[1-\left(\frac{\omega-2E}{|\vec{k}|}\right)^{2}\right]\right.\\ &\left.+\frac{\sum_{\lambda}|\epsilon_{T\lambda}\cdot J|^{2}}{(k^{2}-m_{A^{\prime}}^{2})^{2}+(\omega\Gamma_{T})^{2}}\left[\frac{1}{2}+\frac{2m_{\chi}^{2}}{k^{2}}+\frac{1}{2}\left(\frac{\omega-2E}{|\vec{k}|}\right)^{2}\right]\right\},\\ \end{split} (25)

where JμJ^{\mu} is the final state current that couples to the dark photon, ϵL\epsilon_{L}, ϵTλ\epsilon_{T\lambda} are the polarization vectors of the longitudinal and transverse dark photon with helicity index λ=1,2\lambda=1,2, and the thermal width of the dark photon ΓL,T\Gamma_{L,T} is given by Weldon (1983)

ΓL,T=(1eω/T)(ΓA,absL,T+ΓAχχ¯).\Gamma_{L,T}=(1-e^{-\omega/T})(\Gamma^{L,T}_{A^{\prime},\rm abs}+\Gamma_{A^{\prime}\rightarrow\chi\bar{\chi}}). (26)

In the limit ΓL,TmA\Gamma_{L,T}\ll m_{A^{\prime}}, we can approximate the Breit-Wigner distribution by a δ\delta function

1(k2mA2)2+(ωΓL,T)2πωΓL,Tδ(k2mA2).\frac{1}{(k^{2}-m_{A^{\prime}}^{2})^{2}+(\omega\Gamma_{L,T})^{2}}\rightarrow\frac{\pi}{\omega\Gamma_{L,T}}\delta(k^{2}-m_{A^{\prime}}^{2}). (27)

With the above approximation, the absorption rate of χ\chi becomes

Γχχ¯Af12Ed3p(2π)3gχ¯fχ¯2E𝑑Πf(2π)4δ4(p+ppf)||2¯116πE|p|ωω+𝑑ω(1+fA)gχ¯fχ¯[Br(ALf)||2¯χχ¯AL+Br(ATf)||2¯χχ¯AT],\begin{split}\Gamma_{\chi\bar{\chi}\rightarrow A^{\prime\ast}\rightarrow f}&\equiv\frac{1}{2E}\int\frac{d^{3}p^{\prime}}{(2\pi)^{3}}\frac{g_{\bar{\chi}}f_{\bar{\chi}}}{2E^{\prime}}d\Pi_{f}(2\pi)^{4}\delta^{4}(p+p^{\prime}-p_{f})\overline{|\mathcal{M}|^{2}}\\ &\simeq\frac{1}{16\pi E|\vec{p}|}\int_{\omega_{-}}^{\omega_{+}}d\omega(1+f_{A^{\prime}})g_{\bar{\chi}}f_{\bar{\chi}}\left[\mbox{Br}(A^{\prime}_{L}\rightarrow f)\overline{|\mathcal{M}|^{2}}_{\chi\bar{\chi}\rightarrow A^{\prime}_{L}}+\mbox{Br}(A^{\prime}_{T}\rightarrow f)\overline{|\mathcal{M}|^{2}}_{\chi\bar{\chi}\rightarrow A^{\prime}_{T}}\right],\end{split} (28)

where

ω±=mA22mχ2(E±|p|14mχ2mA2),\displaystyle\omega_{\pm}=\frac{m_{A^{\prime}}^{2}}{2m_{\chi}^{2}}\left(E\pm|\vec{p}|\sqrt{1-\frac{4m_{\chi}^{2}}{m_{A^{\prime}}^{2}}}\right), (29)
Br(AL,Tf)=ΓAfΓA,absL,T+ΓAχχ¯,\displaystyle\mbox{Br}(A^{\prime}_{L,T}\rightarrow f)=\frac{\Gamma_{A^{\prime}\rightarrow f}}{\Gamma^{L,T}_{A^{\prime},\rm abs}+\Gamma_{A^{\prime}\rightarrow\chi\bar{\chi}}}, (30)
ΓALf=12ω𝑑Πf(2π)4δ4(kpf)|ϵLJ|2,\displaystyle\Gamma_{A^{\prime}_{L}\rightarrow f}=\frac{1}{2\omega}\int d\Pi_{f}(2\pi)^{4}\delta^{4}(k-p_{f})|\epsilon_{L}\cdot J|^{2}, (31)
ΓATf=12ω𝑑Πf(2π)4δ4(kpf)12λ|ϵTλJ|2,\displaystyle\Gamma_{A^{\prime}_{T}\rightarrow f}=\frac{1}{2\omega}\int d\Pi_{f}(2\pi)^{4}\delta^{4}(k-p_{f})\frac{1}{2}\sum_{\lambda}|\epsilon_{T\lambda}\cdot J|^{2}, (32)
||2¯χχ¯AL=12gD2mA2[1(ω2E|k|)2],\displaystyle\overline{|\mathcal{M}|^{2}}_{\chi\bar{\chi}\rightarrow A^{\prime}_{L}}=\frac{1}{2}g_{D}^{2}m_{A^{\prime}}^{2}\left[1-\left(\frac{\omega-2E}{|\vec{k}|}\right)^{2}\right], (33)
||2¯χχ¯AT=gD2mA2[12+2mχ2mA2+12(ω2E|k|)2].\displaystyle\overline{|\mathcal{M}|^{2}}_{\chi\bar{\chi}\rightarrow A^{\prime}_{T}}=g_{D}^{2}m_{A^{\prime}}^{2}\left[\frac{1}{2}+\frac{2m_{\chi}^{2}}{m_{A^{\prime}}^{2}}+\frac{1}{2}\left(\frac{\omega-2E}{|\vec{k}|}\right)^{2}\right]. (34)

Note that the inverse decay rates into dark photons of different polarizations are

Γχχ¯AL,T=116πE|p|ωω+𝑑ω(1+fA)gχ¯fχ¯||2¯χχ¯AL,T.\Gamma_{\chi\bar{\chi}\rightarrow A^{\prime}_{L,T}}=\frac{1}{16\pi E|\vec{p}|}\int_{\omega_{-}}^{\omega_{+}}d\omega(1+f_{A^{\prime}})g_{\bar{\chi}}f_{\bar{\chi}}\overline{|\mathcal{M}|^{2}}_{\chi\bar{\chi}\rightarrow A^{\prime}_{L,T}}. (35)

Thus, we can relate the differential absorption rate and inverse decay rates by

dΓχχ¯AfdωBr(ALf)dΓχχ¯ALdω+Br(ATf)dΓχχ¯ATdω.\frac{d\Gamma_{\chi\bar{\chi}\rightarrow A^{\prime\ast}\rightarrow f}}{d\omega}\simeq\mbox{Br}(A^{\prime}_{L}\rightarrow f)\frac{d\Gamma_{\chi\bar{\chi}\rightarrow A^{\prime}_{L}}}{d\omega}+\mbox{Br}(A^{\prime}_{T}\rightarrow f)\frac{d\Gamma_{\chi\bar{\chi}\rightarrow A^{\prime}_{T}}}{d\omega}. (36)

With the above equation, the total pair-absorption rate of χ\chi is given by

Γχ,pair=fΓχχ¯Af=ωω+𝑑ω[dΓχχ¯ALdωfBr(ALf)+dΓχχ¯ATdωfBr(ATf)]=ωω+𝑑ω[dΓχχ¯ALdω+dΓχχ¯ATdω]=Γχχ¯A.\begin{split}\Gamma_{\chi,\rm pair}&=\sum_{f}\Gamma_{\chi\bar{\chi}\rightarrow A^{\prime\ast}\rightarrow f}\\ &=\int_{\omega_{-}}^{\omega_{+}}d\omega\,\left[\frac{d\Gamma_{\chi\bar{\chi}\rightarrow A^{\prime}_{L}}}{d\omega}\sum_{f}\mbox{Br}(A^{\prime}_{L}\rightarrow f)+\frac{d\Gamma_{\chi\bar{\chi}\rightarrow A^{\prime}_{T}}}{d\omega}\sum_{f}\mbox{Br}(A^{\prime}_{T}\rightarrow f)\right]\\ &=\int_{\omega_{-}}^{\omega_{+}}d\omega\,\left[\frac{d\Gamma_{\chi\bar{\chi}\rightarrow A^{\prime}_{L}}}{d\omega}+\frac{d\Gamma_{\chi\bar{\chi}\rightarrow A^{\prime}_{T}}}{d\omega}\right]\\ &=\Gamma_{\chi\bar{\chi}\rightarrow A^{\prime}}.\end{split} (37)

The total pair-absorption rate is equal to that of the inverse decay process χχ¯A\chi\bar{\chi}\rightarrow A^{\prime}. Therefore, the inverse decay process χχ¯A\chi\bar{\chi}\rightarrow A^{\prime} accounts for all the pair-absorption processes with dark photon resonances when mA>2mχm_{A^{\prime}}>2m_{\chi}. These processes include χχ¯npnp\chi\bar{\chi}np\rightarrow np, χχ¯ee+\chi\bar{\chi}\rightarrow e^{-}e^{+}, χχ¯γ\chi\bar{\chi}\rightarrow\gamma^{\ast}, and χχ¯χχ¯\chi\bar{\chi}\rightarrow\chi\bar{\chi}. Note that the dark Compton scattering χAAχ\chi A^{\prime}\rightarrow A^{\prime}\chi also admits an on-shell dark fermion χ¯\bar{\chi}. With the approximation analogous to Eq. (27), one can show that the scattering rate of χ\chi via χAAχ\chi A^{\prime}\rightarrow A^{\prime}\chi is equivalent to the inverse decay rate of χχ¯A\chi\bar{\chi}\rightarrow A^{\prime}, and the scattering rate of AA^{\prime} via the same process is equivalent to the decay rate of Aχχ¯A^{\prime}\rightarrow\chi\bar{\chi}. Thus, we do not include the dark Compton process in our IMFP calculations (see Tables 1 and 2).

With Eq. (36), we can derive the luminosity of χ\chi and χ¯\bar{\chi} in the nondiffuse regime

Lχ+Lχ¯𝑑V𝑑ωduAdωΓAχχ¯[13Br(ALf)+23Br(ATf)],L_{\chi}+L_{\bar{\chi}}\simeq\int dV\int d\omega\frac{du_{A^{\prime}}}{d\omega}\Gamma_{A^{\prime}\rightarrow\chi\bar{\chi}}\left[\frac{1}{3}\mbox{Br}(A^{\prime}_{L}\rightarrow f)+\frac{2}{3}\mbox{Br}(A^{\prime}_{T}\rightarrow f)\right], (38)

where uAu_{A^{\prime}} is the blackbody energy density of the dark photon. We shall make several comments on the above equation.

  1. 1.

    It is assumed that dark photons are in thermal equilibrium with the SM medium while the dark fermion streams freely. (This corresponds to large ϵ\epsilon and small αD\alpha_{D}.) In this scenario, Br(AL,TSM)1\mbox{Br}(A^{\prime}_{L,T}\rightarrow SM)\approx 1. So 1/31/3 of the dark fermion is produced by the longitudinal dark photon, while 2/32/3 of the dark fermion is produced by the transverse dark photon. However, if the dark photon escapes the PNS freely, this calculation is invalid, and the dark photon usually cannot decay into the dark fermions before escaping the PNS. Thus, the dark fermion luminosity would be negligible. Therefore, we compute the luminosity of χ\chi according to Eq. (97) for different scenarios.

  2. 2.

    In the limit of large ϵ\epsilon and small αD\alpha_{D}, the dark sector luminosity is mostly contributed by the dark fermion since the dark photon is trapped. The luminosity only depends on αD\alpha_{D} and does not depend on ϵ\epsilon in this limit. It is possible that the luminosity can exceed LνL_{\nu} for some values of αD\alpha_{D} even if ϵ\epsilon is large, as shown in the left panels of Figs. 3, 4, and 6.

  3. 3.

    The luminosity in this regime also depends on mAm_{A^{\prime}} through the blackbody energy density uAu_{A^{\prime}}. Therefore, for a larger value of αD\alpha_{D}, the constraint can extends to larger mAm_{A^{\prime}} values, as shown in the left panel of Fig. 6.

VII Plasma Effects

In the PNS, the hot (T30T\sim 30 MeV) and dense (ρ1014 g/cm3\rho\sim 10^{14}\mbox{ g}/\mbox{cm}^{3}) plasma consisting of electrons and nucleons can modify the dispersion relations of the SM photons and give rise to a longitudinally polarized propagation mode called plasmon. This plasma effect changes the SM photon propagator, so we must take it into account when calculating the interaction rates between SM and DS particles through photon–dark photon mixing. The effective mixing ϵk|L,T\epsilon_{k|L,T} and the plasma factor βk|L,T\beta_{k|L,T} for the SM current–dark photon interactions are defined by

βk|L,T2ϵk|L,T2ϵ2(k2)2(k2ReΠL,T)2+(ImΠL,T)2,\beta_{k|L,T}^{2}\equiv\frac{\epsilon_{k|L,T}^{2}}{\epsilon^{2}}\equiv\frac{(k^{2})^{2}}{(k^{2}-\mbox{Re}\Pi_{L,T})^{2}+(\mbox{Im}\Pi_{L,T})^{2}}, (39)

where k=(ω,k)k=(\omega,\vec{k}) is the 4-momentum of the dark photon with k2ω2|k|2k^{2}\equiv\omega^{2}-|\vec{k}|^{2}, and ΠL,T\Pi_{L,T} is the longitudinal (L) and transverse (T) polarization functions and is related to the SM photon polarization tensor Πμν=e2Aμ,Aν\Pi^{\mu\nu}=e^{2}\langle A^{\mu},A^{\nu}\rangle by

ΠL\displaystyle\Pi_{L} =k2|k|2Π00,\displaystyle=\frac{k^{2}}{|\vec{k}|^{2}}\Pi^{00}, (40)
ΠT\displaystyle\Pi_{T} =12(δijkikj|k|2)Πij.\displaystyle=\frac{1}{2}\left(\delta^{ij}-\frac{k^{i}k^{j}}{|\vec{k}|^{2}}\right)\Pi^{ij}. (41)

We would like to point out that our definition for ΠL\Pi_{L} aligns with An et al. (2013) but differs from Braaten and Segel (1993). For the on-shell dark photon, we replace the subscript “kk” with “mm” for the effective mixing, i.e., ϵm|L,T2=(ϵk|L,T2)|k2=mA2\epsilon_{m|L,T}^{2}=(\epsilon_{k|L,T}^{2})|_{k^{2}=m_{A^{\prime}}^{2}}, and likewise for the plasma factor. On the other hand, the DS current–SM photon counterpart of Eq. (39) is

ϵ¯k|L,T2ϵ2=(k2)2(k2mA2)2+(ωΓL,T)2β¯k|L,T2,\frac{\bar{\epsilon}_{k|L,T}^{2}}{\epsilon^{2}}=\frac{(k^{2})^{2}}{(k^{2}-m_{A^{\prime}}^{2})^{2}+(\omega\Gamma_{L,T})^{2}}\equiv\bar{\beta}_{k|L,T}^{2}, (42)

where ΓL,T=(1eω/T)ΓA,absL,T\Gamma_{L,T}=(1-e^{-\omega/T})\Gamma^{L,T}_{A^{\prime},\rm abs} is the thermal absorptive width of the dark photon Weldon (1983). Similarly, for the on-shell transverse photon or plasmon, we replace the subscript kk with mm for the effective mixing and the plasma factor. Note that the dispersion relation of the SM photon k2=ReΠL,Tk^{2}=\mathrm{Re}\Pi_{L,T} is a transcendental equation that can only be solved numerically. The derivation of Eqs. (39) and (42) in diagonalized mass basis can be found in Appendix B of Lin (2019).

In the medium consisting of relativistic and degenerate electron, the real parts of the scalar polarization functions are Braaten and Segel (1993)

ReΠL\displaystyle\mbox{Re}\Pi_{L} =3ωp2(1v2)v2[12vln(1+v1v)1],\displaystyle=\frac{3\omega_{p}^{2}(1-v^{2})}{v^{2}}\left[\frac{1}{2v}\ln{\left(\frac{1+v}{1-v}\right)}-1\right], (43)
ReΠT\displaystyle\mbox{Re}\Pi_{T} =3ωp22v2[11v22vln(1+v1v)],\displaystyle=\frac{3\omega_{p}^{2}}{2v^{2}}\left[1-\frac{1-v^{2}}{2v}\ln{\left(\frac{1+v}{1-v}\right)}\right], (44)

where v|k|/ωv\equiv|\vec{k}|/\omega and the plasma frequency in this limit is

ωp2=4αe3π(μe2+13π2T2).\omega_{p}^{2}=\frac{4\alpha_{e}}{3\pi}\left(\mu_{e}^{2}+\frac{1}{3}\pi^{2}T^{2}\right). (45)

With detailed balance ΓprodL,T=eω/TΓabsL,T\Gamma^{L,T}_{\rm prod}=e^{-\omega/T}\Gamma^{L,T}_{\rm abs}, the imaginary part of the polarization functions is

ImΠL,T=ω(ΓprodL,TΓabsL,T)=ωΓabsL,T(1eω/T),\mbox{Im}\Pi_{L,T}=-\omega(\Gamma^{L,T}_{\rm prod}-\Gamma^{L,T}_{\rm abs})=-\omega\Gamma^{L,T}_{\rm abs}(1-e^{-\omega/T}), (46)

where ΓprodL,T\Gamma^{L,T}_{\rm prod} (ΓabsL,T\Gamma^{L,T}_{\rm abs}) are the production (absorption) rates of the SM photon. When calculating the plasmon decay, we should include the renormalization factor for the polarization vectors of the external photons,

ϵ~L,Tμ=ZL,TϵL,Tμ,\tilde{\epsilon}_{L,T}^{\mu}=\sqrt{Z_{L,T}}\epsilon_{L,T}^{\mu}, (47)

where the renormalization factor is given by

ZL,T1=1ReΠL,Tω2|pole,Z_{L,T}^{-1}=1-\left.\frac{\partial\mbox{Re}\Pi_{L,T}}{\partial\omega^{2}}\right|_{\rm pole}, (48)

which can be calculated from Eqs. (43) and (44),

ZL1\displaystyle Z_{L}^{-1} =3ωp22ω2v2[11v22vln(1+v1v)]pole,\displaystyle=\frac{3\omega_{p}^{2}}{2\omega^{2}v^{2}}\left[1-\frac{1-v^{2}}{2v}\ln{\left(\frac{1+v}{1-v}\right)}\right]_{\rm pole}, (49)
ZT1\displaystyle Z_{T}^{-1} =13ωp22ω2v2[323v24vln(1+v1v)]pole.\displaystyle=1-\frac{3\omega_{p}^{2}}{2\omega^{2}v^{2}}\left[\frac{3}{2}-\frac{3-v^{2}}{4v}\ln{\left(\frac{1+v}{1-v}\right)}\right]_{\rm pole}. (50)

VIII Interaction Rates and Inverse Mean Free Path

The interaction rate of particle 11 via a process 1++n1++m1+...+n\rightarrow 1^{\prime}+...+m^{\prime} is given by

Γ=12E1i=1nd3pi(2π)3gifi2Eij=1md3pj(2π)31±fj2Ej||2¯(2π)4δ4(i=1npij=1mpj),\Gamma=\frac{1}{2E_{1}}\int\prod_{i=1}^{n}\frac{d^{3}p_{i}}{(2\pi)^{3}}\frac{g_{i}f_{i}}{2E_{i}}\prod_{j=1^{\prime}}^{m^{\prime}}\frac{d^{3}p_{j}}{(2\pi)^{3}}\frac{1\pm f_{j}}{2E_{j}}\overline{|\mathcal{M}|^{2}}(2\pi)^{4}\delta^{4}(\sum_{i=1}^{n}p_{i}-\sum_{j=1^{\prime}}^{m^{\prime}}p_{j}), (51)

where gig_{i} is the degrees of freedom of the initial state particles and fif_{i} and fjf_{j} are the distribution functions of the initial and final state particles. The upper (lower) sign in front of fjf_{j} is for bosonic (fermionic) final states, which takes into account the bosonic enhancement and Pauli-blocking effects. The above expression already assumes that all the particles are in thermal equilibrium. The IMFP of the process is related to the interaction rate by

λ~1=Γv1=E1|p1|Γ.\tilde{\lambda}^{-1}=\frac{\Gamma}{v_{1}}=\frac{E_{1}}{|\vec{p}_{1}|}\Gamma. (52)

In the following, we denote the 4-momentum of AA^{\prime} by k=(ω,k)k=(\omega,\vec{k}), that of χ\chi by p=(E,p)p=(E,\vec{p}), and that of χ¯\bar{\chi} by p=(E,p)p^{\prime}=(E^{\prime},\vec{p}^{\prime}), unless noted otherwise.

VIII.1 AnpnpA^{\prime}np\rightarrow np

The absorption rate via inverse nucleon bremsstrahlung AnpnpA^{\prime}np\rightarrow np is given by Chang et al. (2017)

ΓAnpnpL,T=323παeϵm|L,T2nnnpω3(πTmN)3/2σnp(2)(T)hL,T,\Gamma^{L,T}_{A^{\prime}np\rightarrow np}=\frac{32}{3\pi}\frac{\alpha_{e}\epsilon^{2}_{m|L,T}n_{n}n_{p}}{\omega^{3}}\left(\frac{\pi T}{m_{N}}\right)^{3/2}\left\langle\sigma_{np}^{(2)}(T)\right\rangle h_{L,T}, (53)

where nn,pn_{n,p} is the number density of neutrons and protons, respectively, mNm_{N} is the nucleon mass, σnp(2)(T)\left\langle\sigma_{np}^{(2)}(T)\right\rangle is the thermally averaged nn-pp cross section defined as Rrapaj and Reddy (2016)

σnp(2)(T)=120𝑑xx2ex11dcosθcm(1cosθcm)dσnpdcosθcm(Tcm=xT,cosθcm),\left\langle\sigma^{(2)}_{np}(T)\right\rangle=\frac{1}{2}\int_{0}^{\infty}dxx^{2}e^{-x}\int_{-1}^{1}d\cos{\theta_{cm}}(1-\cos{\theta_{cm}})\frac{d\sigma_{np}}{d\cos{\theta_{cm}}}\left(T_{cm}=xT,\cos{\theta_{cm}}\right), (54)

and hL,Th_{L,T} is defined as

hL\displaystyle h_{L} =mA2ω2,\displaystyle=\frac{m_{A^{\prime}}^{2}}{\omega^{2}}, (55)
hT\displaystyle h_{T} =1.\displaystyle=1. (56)

In the derivation of Eq. (53), the soft radiation approximation is used to connect the bremsstrahlung rate with the experimental nn-pp scattering cross section. Additionally, the nucleons are assumed to follow Maxwell-Boltzmann distribution, and the Pauli-blocking effect is ignored.

VIII.2 AeeγA^{\prime}e^{-}\rightarrow e^{-}\gamma

The absorption rate via the Compton-like process AeeγA^{\prime}e^{-}\rightarrow e^{-}\gamma is approximately Chang et al. (2017)

ΓAeeγL,T=8παe2ϵm|L,T2ne3EF2ωpωhL,T,\Gamma^{L,T}_{A^{\prime}e^{-}\rightarrow e^{-}\gamma}=\frac{8\pi\alpha_{e}^{2}\epsilon^{2}_{m|L,T}n_{e}}{3E_{F}^{2}}\sqrt{\frac{\omega_{p}}{\omega}}h_{L,T}, (57)

where nen_{e} is the electron number density, EF=(3π2ne)2/3+me2E_{F}=\sqrt{(3\pi^{2}n_{e})^{2/3}+m_{e}^{2}} is the electron Fermi energy, and hL,Th_{L,T} is defined in Eqs. (55) and (56). The approximated formula is valid for ω200\omega\lesssim 200 MeV and rRνr\leq R_{\nu} Chang et al. (2017).

VIII.3 Aee+A^{\prime}\rightarrow e^{-}e^{+}

The decay rates of dark photon via Aee+A^{\prime}\rightarrow e^{-}e^{+}, given that mA>2mem_{A^{\prime}}>2m_{e}, are given by

ΓAee+L\displaystyle\Gamma_{A^{\prime}\rightarrow e^{-}e^{+}}^{L} =ϵm|L2αemA24|k|ξ0ξ0𝑑ξ[1+exp(μeω/2Tω2Tξ)]1(1ω2|k|2ξ2),\displaystyle=\frac{\epsilon_{m|L}^{2}\alpha_{e}m_{A^{\prime}}^{2}}{4|\vec{k}|}\int_{-\xi_{0}}^{\xi_{0}}d\xi\left[1+\exp{\left(\frac{\mu_{e}-\omega/2}{T}-\frac{\omega}{2T}\xi\right)}\right]^{-1}\left(1-\frac{\omega^{2}}{|\vec{k}|^{2}}\xi^{2}\right), (58)
ΓAee+T\displaystyle\Gamma_{A^{\prime}\rightarrow e^{-}e^{+}}^{T} =ϵm|T2αemA24|k|ξ0ξ0𝑑ξ[1+exp(μeω/2Tω2Tξ)]1(12+2me2mA2+ω22|k|2ξ2),\displaystyle=\frac{\epsilon_{m|T}^{2}\alpha_{e}m_{A^{\prime}}^{2}}{4|\vec{k}|}\int_{-\xi_{0}}^{\xi_{0}}d\xi\left[1+\exp{\left(\frac{\mu_{e}-\omega/2}{T}-\frac{\omega}{2T}\xi\right)}\right]^{-1}\left(\frac{1}{2}+\frac{2m_{e}^{2}}{m_{A^{\prime}}^{2}}+\frac{\omega^{2}}{2|\vec{k}|^{2}}\xi^{2}\right), (59)

where the integral limit is

ξ0=|k|ω14me2mA2.\xi_{0}=\frac{|\vec{k}|}{\omega}\sqrt{1-\frac{4m_{e}^{2}}{m_{A^{\prime}}^{2}}}. (60)

VIII.4 χχ¯npnp\chi\bar{\chi}np\rightarrow np

The pair-annihilation rate via inverse nucleon bremsstrahlung χχ¯npnp\chi\bar{\chi}np\rightarrow np is

Γχχ¯npnp=16ϵ2αDαennnp3π2E(πTmN)3/2σnp(2)(T)0d|k|11dcosθ|k|2(χχ¯npnpL+2χχ¯npnpT)Eω2k2(1+eE/T),\Gamma_{\chi\bar{\chi}np\rightarrow np}=\frac{16\epsilon^{2}\alpha_{D}\alpha_{e}n_{n}n_{p}}{3\pi^{2}E}\left(\frac{\pi T}{m_{N}}\right)^{3/2}\left\langle\sigma_{np}^{(2)}(T)\right\rangle\int_{0}^{\infty}d|\vec{k}|\int_{-1}^{1}d\cos{\theta}\,\frac{|\vec{k}|^{2}(\mathcal{F}^{L}_{\chi\bar{\chi}np\rightarrow np}+2\mathcal{F}^{T}_{\chi\bar{\chi}np\rightarrow np})}{E^{\prime}\omega^{2}k^{2}(1+e^{E^{\prime}/T})}, (61)

where the momentum transfer is k=p+pk=p+p^{\prime} and cosθ=kp/|k||p|\cos{\theta}=\vec{k}\cdot\vec{p}/|\vec{k}||\vec{p}|. Note that E,pE,p (E,pE^{\prime},p^{\prime}) are for χ\chi (χ¯\bar{\chi}). Thus E=E2+|k|22|k||p|cosθE^{\prime}=\sqrt{E^{2}+|\vec{k}|^{2}-2|\vec{k}||\vec{p}|\cos{\theta}} and ω=E+E\omega=E+E^{\prime}. The two terms χχ¯npnpL\mathcal{F}^{L}_{\chi\bar{\chi}np\rightarrow np}, χχ¯npnpT\mathcal{F}^{T}_{\chi\bar{\chi}np\rightarrow np} in the integrand are

χχ¯npnpL\displaystyle\mathcal{F}^{L}_{\chi\bar{\chi}np\rightarrow np} =βk|L2β¯k|L2k2ω2[1(ω2E|k|)2],\displaystyle=\beta_{k|L}^{2}\bar{\beta}_{k|L}^{2}\cdot\frac{k^{2}}{\omega^{2}}\left[1-\left(\frac{\omega-2E}{|\vec{k}|}\right)^{2}\right], (62)
χχ¯npnpT\displaystyle\mathcal{F}^{T}_{\chi\bar{\chi}np\rightarrow np} =βk|T2β¯k|T2[12+2mχ2k2+12(ω2E|k|)2].\displaystyle=\beta_{k|T}^{2}\bar{\beta}_{k|T}^{2}\cdot\left[\frac{1}{2}+\frac{2m_{\chi}^{2}}{k^{2}}+\frac{1}{2}\left(\frac{\omega-2E}{|\vec{k}|}\right)^{2}\right]. (63)

Similar to the calculation of AnpnpA^{\prime}np\rightarrow np, we use soft radiation approximation for the dark fermion pair absorption and ignore the Pauli-blocking effect of the nucleons.

When mA>2mχm_{A^{\prime}}>2m_{\chi}, we use the narrow width approximation (NWA) detailed in Appendix VI to approximate the integral. The resulting formula is

Γχχ¯npnpNWA=8ϵ2αDαemA2nnnp3πE|p|(πTmN)3/2σnp(2)(T)ωω+𝑑ωχχ¯npnpL+2χχ¯npnpTω3[1+e(ωE)/T],\Gamma^{\rm NWA}_{\chi\bar{\chi}np\rightarrow np}=\frac{8\epsilon^{2}\alpha_{D}\alpha_{e}m_{A^{\prime}}^{2}n_{n}n_{p}}{3\pi E|\vec{p}|}\left(\frac{\pi T}{m_{N}}\right)^{3/2}\left\langle\sigma_{np}^{(2)}(T)\right\rangle\int_{\omega_{-}}^{\omega_{+}}d\omega\,\frac{\mathcal{F}^{L\ast}_{\chi\bar{\chi}np\rightarrow np}+2\mathcal{F}^{T\ast}_{\chi\bar{\chi}np\rightarrow np}}{\omega^{3}[1+e^{(\omega-E)/T}]}, (64)

where the two terms χχ¯npnpL\mathcal{F}^{L\ast}_{\chi\bar{\chi}np\rightarrow np}, χχ¯npnpT\mathcal{F}^{T\ast}_{\chi\bar{\chi}np\rightarrow np} in the integrand are

χχ¯npnpL\displaystyle\mathcal{F}^{L\ast}_{\chi\bar{\chi}np\rightarrow np} =βm|L2ΓLmA2ω2[1(ω2E|k|)2],\displaystyle=\frac{\beta_{m|L}^{2}}{\Gamma_{L}}\cdot\frac{m_{A^{\prime}}^{2}}{\omega^{2}}\left[1-\left(\frac{\omega-2E}{|\vec{k}|}\right)^{2}\right], (65)
χχ¯npnpT\displaystyle\mathcal{F}^{T\ast}_{\chi\bar{\chi}np\rightarrow np} =βm|T2ΓT[12+2mχ2mA2+12(ω2E|k|)2],\displaystyle=\frac{\beta_{m|T}^{2}}{\Gamma_{T}}\cdot\left[\frac{1}{2}+\frac{2m_{\chi}^{2}}{m_{A^{\prime}}^{2}}+\frac{1}{2}\left(\frac{\omega-2E}{|\vec{k}|}\right)^{2}\right], (66)

and we now have |k|=ω2mA2|\vec{k}|=\sqrt{\omega^{2}-m_{A^{\prime}}^{2}}. The dark photon widths must include the contribution from the dark photon decay via Aχχ¯A^{\prime}\rightarrow\chi\bar{\chi}. Thus we have ΓL,T=(1eω/T)(ΓA,absL,T+ΓAχχ¯)\Gamma_{L,T}=(1-e^{-\omega/T})(\Gamma^{L,T}_{A^{\prime},\rm abs}+\Gamma_{A^{\prime}\rightarrow\chi\bar{\chi}}). Momentum conservation demands that ω\omega is bounded by

ω±=mA22mχ2(E±|p|14mχ2mA2).\omega_{\pm}=\frac{m_{A^{\prime}}^{2}}{2m_{\chi}^{2}}\left(E\pm|\vec{p}|\sqrt{1-\frac{4m_{\chi}^{2}}{m_{A^{\prime}}^{2}}}\right). (67)

VIII.5 χχ¯ee+\chi\bar{\chi}\rightarrow e^{-}e^{+}

The pair-annihilation rate via χ(p)+χ¯(p)e(p3)+e+(p4)\chi(p)+\bar{\chi}(p^{\prime})\rightarrow e^{-}(p_{3})+e^{+}(p_{4}) is

Γχχ¯ee+=ϵ2αDαe4πE0d|k|11dcosθE3E3+𝑑E3|k|Ef(1f3)(1f4)(χχ¯ee+L+2χχ¯ee+T),\Gamma_{\chi\bar{\chi}\rightarrow e^{-}e^{+}}=\frac{\epsilon^{2}\alpha_{D}\alpha_{e}}{4\pi E}\int_{0}^{\infty}d|\vec{k}|\int_{-1}^{1}d\cos{\theta}\,\int_{E_{3-}}^{E_{3+}}dE_{3}\,\frac{|\vec{k}|}{E^{\prime}}f^{\prime}(1-f_{3})(1-f_{4})\,(\mathcal{F}^{L}_{\chi\bar{\chi}\rightarrow e^{-}e^{+}}+2\mathcal{F}^{T}_{\chi\bar{\chi}\rightarrow e^{-}e^{+}}), (68)

where we have the same definitions for cosθ\cos{\theta}, EE^{\prime}, and ω\omega as in Eq. (61). Moreover, ff^{\prime}, f3f_{3}, and f4f_{4} are the distribution functions of χ¯\bar{\chi}, ee^{-}, and e+e^{+}, respectively. That is,

f\displaystyle f^{\prime} =1eE/T+1,\displaystyle=\frac{1}{e^{E^{\prime}/T}+1}, (69)
f3\displaystyle f_{3} =1e(E3μe)/T+1,\displaystyle=\frac{1}{e^{(E_{3}-\mu_{e})/T}+1}, (70)
f4\displaystyle f_{4} =1e(E4+μe)/T+1,\displaystyle=\frac{1}{e^{(E_{4}+\mu_{e})/T}+1}, (71)

where the positron energy is expressed as E4=ωE3E_{4}=\omega-E_{3} by energy conservation. With the electron and positron distribution function included, we take the Pauli-blocking effect into account, which is significant near the center of the PNS. The two terms χχ¯ee+L\mathcal{F}^{L}_{\chi\bar{\chi}\rightarrow e^{-}e^{+}}, χχ¯ee+T\mathcal{F}^{T}_{\chi\bar{\chi}\rightarrow e^{-}e^{+}} in the integrand are

χχ¯ee+L\displaystyle\mathcal{F}^{L}_{\chi\bar{\chi}\rightarrow e^{-}e^{+}} =βk|L2β¯k|L2[1(ω2E|k|)2][1(ω2E3|k|)2],\displaystyle=\beta_{k|L}^{2}\bar{\beta}_{k|L}^{2}\left[1-\left(\frac{\omega-2E}{|\vec{k}|}\right)^{2}\right]\left[1-\left(\frac{\omega-2E_{3}}{|\vec{k}|}\right)^{2}\right], (72)
χχ¯ee+T\displaystyle\mathcal{F}^{T}_{\chi\bar{\chi}\rightarrow e^{-}e^{+}} =βk|T2β¯k|T2[12+2mχ2k2+12(ω2E|k|)2][12+2me2k2+12(ω2E3|k|)2].\displaystyle=\beta_{k|T}^{2}\bar{\beta}_{k|T}^{2}\left[\frac{1}{2}+\frac{2m_{\chi}^{2}}{k^{2}}+\frac{1}{2}\left(\frac{\omega-2E}{|\vec{k}|}\right)^{2}\right]\left[\frac{1}{2}+\frac{2m_{e}^{2}}{k^{2}}+\frac{1}{2}\left(\frac{\omega-2E_{3}}{|\vec{k}|}\right)^{2}\right]. (73)

Momentum conservation demands that E3E_{3} is bounded by

E3±=12[ω±|k|14me2k2].E_{3\pm}=\frac{1}{2}\left[\omega\pm|\vec{k}|\sqrt{1-\frac{4m_{e}^{2}}{k^{2}}}\right]. (74)

When mA>2mχm_{A^{\prime}}>2m_{\chi}, we use NWA (see Appendix VI) to approximate the annihilation rate. The resulting formula is

Γχχ¯ee+NWA=ϵ2αeαDmA48E|p|ωω+𝑑ωE3E3+𝑑E3f2(1f3)(1f4)ω|k|(χχ¯ee+L+2χχ¯ee+T),\Gamma^{\rm NWA}_{\chi\bar{\chi}\rightarrow e^{-}e^{+}}=\frac{\epsilon^{2}\alpha_{e}\alpha_{D}m_{A^{\prime}}^{4}}{8E|\vec{p}|}\int_{\omega_{-}}^{\omega_{+}}d\omega\int_{E_{3-}}^{E_{3+}}dE_{3}\frac{f_{2}(1-f_{3})(1-f_{4})}{\omega|\vec{k}|}(\mathcal{F}^{L\ast}_{\chi\bar{\chi}\rightarrow e^{-}e^{+}}+2\mathcal{F}^{T\ast}_{\chi\bar{\chi}\rightarrow e^{-}e^{+}}), (75)

where we now have |k|=ω2mA2|\vec{k}|=\sqrt{\omega^{2}-m_{A^{\prime}}^{2}}. The energies of χ¯\bar{\chi} and e+e^{+} on which their distribution functions depend are expressed as E=ωEE^{\prime}=\omega-E and E4=ωE3E_{4}=\omega-E_{3}. The two terms χχ¯ee+L\mathcal{F}^{L\ast}_{\chi\bar{\chi}\rightarrow e^{-}e^{+}}, χχ¯ee+T\mathcal{F}^{T\ast}_{\chi\bar{\chi}\rightarrow e^{-}e^{+}} in the integrand now are

χχ¯ee+L\displaystyle\mathcal{F}^{L\ast}_{\chi\bar{\chi}\rightarrow e^{-}e^{+}} =βm|L2ΓL[1(ω2E|k|)2][1(ω2E3|k|)2],\displaystyle=\frac{\beta_{m|L}^{2}}{\Gamma_{L}}\left[1-\left(\frac{\omega-2E}{|\vec{k}|}\right)^{2}\right]\left[1-\left(\frac{\omega-2E_{3}}{|\vec{k}|}\right)^{2}\right], (76)
χχ¯ee+T\displaystyle\mathcal{F}^{T\ast}_{\chi\bar{\chi}\rightarrow e^{-}e^{+}} =βm|T2ΓT[12+2mχ2mA2+12(ω2E|k|)2][12+2me2mA2+12(ω2E3|k|)2],\displaystyle=\frac{\beta_{m|T}^{2}}{\Gamma_{T}}\left[\frac{1}{2}+\frac{2m_{\chi}^{2}}{m_{A^{\prime}}^{2}}+\frac{1}{2}\left(\frac{\omega-2E}{|\vec{k}|}\right)^{2}\right]\left[\frac{1}{2}+\frac{2m_{e}^{2}}{m_{A^{\prime}}^{2}}+\frac{1}{2}\left(\frac{\omega-2E_{3}}{|\vec{k}|}\right)^{2}\right], (77)

where ΓL,T\Gamma_{L,T} is the same as in Eqs. (65) and (66). The limit of E3E_{3} now becomes

E3±=12[ω±|k|14me2mA2],E_{3\pm}=\frac{1}{2}\left[\omega\pm|\vec{k}|\sqrt{1-\frac{4m_{e}^{2}}{m_{A^{\prime}}^{2}}}\right], (78)

while the limit of ω\omega is the same as Eq. (67).

VIII.6 χχ¯γ\chi\bar{\chi}\rightarrow\gamma^{\ast}

Let ωL,T=ωL,T(|k|)\omega_{L,T}=\omega_{L,T}(|\vec{k}|) be the energy of the longitudinal and transverse SM photon as functions of |k||\vec{k}|, and mL,Tm_{L,T} be the effective mass satisfying mL,T2=ωL,T2|k|2=ReΠL,T(ωL,T,|k|)m_{L,T}^{2}=\omega_{L,T}^{2}-|\vec{k}|^{2}=\mathrm{Re}\Pi_{L,T}(\omega_{L,T},|\vec{k}|). Then the pair-annihilation rate can be expressed as

Γχχ¯γ=ϵ2αD4E|p|L,T0|k|d|k|ωL,TgL,Tχχ¯γL,TΘ(1cos2θL,T)[1+e(ωL,TE)/T](1eωL,T/T),\Gamma_{\chi\bar{\chi}\rightarrow\gamma^{\ast}}=\frac{\epsilon^{2}\alpha_{D}}{4E|\vec{p}|}\sum_{L,T}\int_{0}^{\infty}\frac{|\vec{k}|d|\vec{k}|}{\omega_{L,T}}\frac{g_{L,T}\mathcal{F}^{L,T}_{\chi\bar{\chi}\rightarrow\gamma^{\ast}}\Theta(1-\cos^{2}{\theta_{L,T}})}{[1+e^{(\omega_{L,T}-E)/T}](1-e^{-\omega_{L,T}/T})}, (79)

where gL=1g_{L}=1, gT=2g_{T}=2, and χχ¯γL\mathcal{F}^{L}_{\chi\bar{\chi}\rightarrow\gamma^{\ast}}, χχ¯γT\mathcal{F}^{T}_{\chi\bar{\chi}\rightarrow\gamma^{\ast}} are defined as

χχ¯γL\displaystyle\mathcal{F}^{L}_{\chi\bar{\chi}\rightarrow\gamma^{\ast}} =β¯m|L2ZLmL2[1(ωL2E|k|)2],\displaystyle=\bar{\beta}^{2}_{m|L}Z_{L}m_{L}^{2}\left[1-\left(\frac{\omega_{L}-2E}{|\vec{k}|}\right)^{2}\right], (80)
χχ¯γT\displaystyle\mathcal{F}^{T}_{\chi\bar{\chi}\rightarrow\gamma^{\ast}} =β¯m|T2ZTmT2[12+2mχ2mT2+12(ωT2E|k|)2],\displaystyle=\bar{\beta}^{2}_{m|T}Z_{T}m_{T}^{2}\left[\frac{1}{2}+\frac{2m_{\chi}^{2}}{m_{T}^{2}}+\frac{1}{2}\left(\frac{\omega_{T}-2E}{|\vec{k}|}\right)^{2}\right], (81)

where ZL,TZ_{L,T} is the renormalization factor for the longitudinal plasmon and transverse photon detailed in Appendix VII. The bosonic enhancement for the plasmon is included in Eq. (79). The step function in Eq. (79) is to ensure that the momenta are kinematically allowed. The definition of cosθL,T\cos{\theta_{L,T}} is

cosθL,T=2ωL,TEmL,T22|k||p|.\cos{\theta_{L,T}}=\frac{2\omega_{L,T}E-m_{L,T}^{2}}{2|\vec{k}||\vec{p}|}. (82)

We use NWA to approximate the annihilation rate when mA>2mχm_{A^{\prime}}>2m_{\chi}. The resulting expression is

Γχχ¯γ=πϵ2αDmA64E|p|L,TgL,Tχχ¯γL,TΘ(ω+ωL,T)Θ(ωL,Tω)[1+e(ωLE)/T](1eωL/T),\Gamma_{\chi\bar{\chi}\rightarrow\gamma^{\ast}}=\frac{\pi\epsilon^{2}\alpha_{D}m_{A^{\prime}}^{6}}{4E|\vec{p}|}\sum_{L,T}\frac{g_{L,T}\mathcal{F}^{L,T\ast}_{\chi\bar{\chi}\rightarrow\gamma^{\ast}}\Theta(\omega_{+}-\omega^{\ast}_{L,T})\Theta(\omega^{\ast}_{L,T}-\omega_{-})}{[1+e^{(\omega^{\ast}_{L}-E)/T}](1-e^{-\omega^{\ast}_{L}/T})}, (83)

where ωL,T\omega^{\ast}_{L,T} is the solution of the equation ReΠL,T(ω,ω2mA2)=mA2\mbox{Re}\Pi_{L,T}(\omega,\sqrt{\omega^{2}-m_{A^{\prime}}^{2}})=m_{A^{\prime}}^{2}. χχ¯γL,T\mathcal{F}^{L,T\ast}_{\chi\bar{\chi}\rightarrow\gamma^{\ast}} is now defined as

χχ¯γL\displaystyle\mathcal{F}^{L\ast}_{\chi\bar{\chi}\rightarrow\gamma^{\ast}} =1ωLΓL[1(ωL2E)2ωL2mA2]|dReΠL(ω,ω2mA2)dω|ωL|1,\displaystyle=\frac{1}{\omega^{\ast}_{L}\Gamma_{L}}\left[1-\frac{(\omega^{\ast}_{L}-2E)^{2}}{\omega^{\ast 2}_{L}-m_{A^{\prime}}^{2}}\right]\left|\left.\frac{d\mbox{Re}\Pi_{L}(\omega,\sqrt{\omega^{2}-m_{A^{\prime}}^{2}})}{d\omega}\right|_{\omega^{\ast}_{L}}\right|^{-1}, (84)
χχ¯γT\displaystyle\mathcal{F}^{T\ast}_{\chi\bar{\chi}\rightarrow\gamma^{\ast}} =1ωTΓT[12+2mχ2mA2+12(ωT2E)2ωT2mA2]|dReΠT(ω,ω2mA2)dω|ωT|1.\displaystyle=\frac{1}{\omega^{\ast}_{T}\Gamma_{T}}\left[\frac{1}{2}+\frac{2m_{\chi}^{2}}{m_{A^{\prime}}^{2}}+\frac{1}{2}\frac{(\omega^{\ast}_{T}-2E)^{2}}{\omega^{\ast 2}_{T}-m_{A^{\prime}}^{2}}\right]\left|\left.\frac{d\mbox{Re}\Pi_{T}(\omega,\sqrt{\omega^{2}-m_{A^{\prime}}^{2}})}{d\omega}\right|_{\omega^{\ast}_{T}}\right|^{-1}. (85)

ω±\omega_{\pm} is the same as Eq. (67) and ΓL,T\Gamma_{L,T} is the same as in Eqs. (65) and (66).

VIII.7 χeχe\chi e^{-}\rightarrow\chi e^{-} and χpχp\chi p\rightarrow\chi p

The scattering rate of χ(p)+e(p2)χ(p)+e(p4)\chi(p)+e^{-}(p_{2})\rightarrow\chi(p^{\prime})+e^{-}(p_{4}) is given by

Γχeχe=ϵ2αDαe4πEd|k|dcosθE2+𝑑E2|k|Ef2(1f)(1f4)(χeχeL+2χeχeT),\Gamma_{\chi e^{-}\rightarrow\chi e^{-}}=\frac{\epsilon^{2}\alpha_{D}\alpha_{e}}{4\pi E}\int d|\vec{k}|\,d\cos{\theta}\,\int_{E_{2+}}^{\infty}dE_{2}\,\frac{|\vec{k}|}{E^{\prime}}f_{2}(1-f^{\prime})(1-f_{4})\,(\mathcal{F}^{L}_{\chi e^{-}\rightarrow\chi e^{-}}+2\mathcal{F}^{T}_{\chi e^{-}\rightarrow\chi e^{-}}), (86)

where k=ppk=p-p^{\prime} is the momentum transfer, cosθ\cos{\theta} and EE^{\prime} are the same in Eq. (61) while ω=EE\omega=E-E^{\prime} instead. f2f_{2}, ff^{\prime}, and f4f_{4} are the distribution functions of the initial state ee^{-}, final state χ\chi, and final state ee^{-}, respectively. The final state electron energy is expressed as E4=ω+E2E_{4}=\omega+E_{2} by energy conservation. The two terms χeχeL\mathcal{F}^{L}_{\chi e^{-}\rightarrow\chi e^{-}}, χeχeT\mathcal{F}^{T}_{\chi e^{-}\rightarrow\chi e^{-}} are the same as Eqs. (72) and (73) except that E3-E_{3} is replaced with +E2+E_{2}, and E2E_{2} is bounded below by

E2+=12(|k|14me2k2ω).E_{2+}=\frac{1}{2}\left(|\vec{k}|\sqrt{1-\frac{4m_{e}^{2}}{k^{2}}}-\omega\right). (87)

Note that the momentum transfer is spacelike (k2<0k^{2}<0) for the χ\chi-ee^{-} scattering. In this scenario, we take the dark photon width to be zero, while Im ΠL,T\mbox{Im }\Pi_{L,T} is determined by the imaginary parts of Eqs. (43) and (44) with ωω+iε\omega\rightarrow\omega+i\varepsilon for an infinitesimal ε>0\varepsilon>0 Kuznetsov and Mikheev (2013).

The scattering rate via χpχp\chi p\rightarrow\chi p is the same as χeχe\chi e^{-}\rightarrow\chi e^{-} if the differences in masses and distribution functions of the proton and the electron are taken into account. However, we can approximate the χ\chi-pp scattering as an elastic scattering due to the fact that mpm_{p} is much larger than mχm_{\chi}. We also ignore the Pauli-blocking effect of proton. Thus the scattering rate of χ\chi-pp scattering can be approximated by

Γχpχp=4πϵ2αeαDnp|p|E(1+eE/T)11dcosθβk|L22E2|p|2(1cosθ)(k2mA2)2,\Gamma_{\chi p\rightarrow\chi p}=\frac{4\pi\epsilon^{2}\alpha_{e}\alpha_{D}n_{p}|\vec{p}|}{E(1+e^{-E/T})}\int_{-1}^{1}d\cos{\theta}\beta_{k|L}^{2}\cdot\frac{2E^{2}-|\vec{p}|^{2}(1-\cos{\theta})}{(k^{2}-m_{A^{\prime}}^{2})^{2}}, (88)

where the momentum transfer is k2|k|22|p|2(1cosθ)k^{2}\simeq-|\vec{k}|^{2}\simeq-2|\vec{p}|^{2}(1-\cos{\theta}). Note that in this static limit (ω0\omega\rightarrow 0), the contribution from the transverse mode, corresponding to the classical magnetic field, is suppressed. And the plasma factor βk|L2\beta_{k|L}^{2} of the longitudinal mode, corresponding to the static electric field, accounts for the screening effect in medium Kapusta and Gale (2006).

VIII.8 AχχAA^{\prime}\chi\rightarrow\chi A^{\prime} and AAχχ¯A^{\prime}A^{\prime}\leftrightarrow\chi\bar{\chi}

The spin-averaged matrix element squared of A(p1)+χ(p2)χ(p3)+A(p4)A^{\prime}(p_{1})+\chi(p_{2})\rightarrow\chi(p_{3})+A^{\prime}(p_{4}) is

||2¯AχχA=64π2αD23{(mA2+2mχ2)2[1(smχ2)2+1(tmχ2)2]+8(mχ4mA4)(smχ2)(tmχ2)+ 4(mχ2+mA2)(1smχ2+1tmχ2)(smχ2tmχ2+tmχ2smχ2)},\begin{split}\overline{|\mathcal{M}|^{2}}_{A^{\prime}\chi\rightarrow\chi A^{\prime}}&=\frac{64\pi^{2}\alpha_{D}^{2}}{3}\left\{(m_{A^{\prime}}^{2}+2m_{\chi}^{2})^{2}\left[\frac{1}{(s-m_{\chi}^{2})^{2}}+\frac{1}{(t-m_{\chi}^{2})^{2}}\right]+\frac{8(m_{\chi}^{4}-m_{A^{\prime}}^{4})}{(s-m_{\chi}^{2})(t-m_{\chi}^{2})}\right.\\ &\left.+\,4(m_{\chi}^{2}+m_{A^{\prime}}^{2})\left(\frac{1}{s-m_{\chi}^{2}}+\frac{1}{t-m_{\chi}^{2}}\right)-\left(\frac{s-m_{\chi}^{2}}{t-m_{\chi}^{2}}+\frac{t-m_{\chi}^{2}}{s-m_{\chi}^{2}}\right)\right\},\end{split} (89)

where s=(p1+p2)2s=(p_{1}+p_{2})^{2} and t=(p1p3)2t=(p_{1}-p_{3})^{2} are the Mandelstam variables. By crossing symmetry, the matrix element of A(p1)+A(p2)χ(p3)+χ¯(p4)A^{\prime}(p_{1})+A^{\prime}(p_{2})\rightarrow\chi(p_{3})+\bar{\chi}(p_{4}) is the same as Eq. (89) except that ss is replaced by u=(p1p4)2u=(p_{1}-p_{4})^{2} and the matrix element is multiplied by a factor of 2/3-2/3. (The minus sign is due to the crossing of one fermion state.) For χ(p1)+χ¯(p2)A(p3)+A(p4)\chi(p_{1})+\bar{\chi}(p_{2})\rightarrow A^{\prime}(p_{3})+A^{\prime}(p_{4}), we can again reuse Eq. (89), replace ss by uu, and multiply by a factor of 3/2-3/2.

VIII.9 χχ¯χχ¯\chi\bar{\chi}\rightarrow\chi\bar{\chi} and χχχχ\chi\chi\rightarrow\chi\chi

The spin-averaged matrix element squared of χ(p1)+χ¯(p2)χ(p3)+χ¯(p4)\chi(p_{1})+\bar{\chi}(p_{2})\rightarrow\chi(p_{3})+\bar{\chi}(p_{4}) is

||2¯χχ¯χχ¯=64π2αD2{(52mA44mχ2mA2+4mχ4)[1(tmA2)2+1(smA2)2]+4(mA4mχ4)(tmA2)(smA2)+(3mA24mχ2)[smA2(tmA2)2+tmA2(smA2)2]+6mA2(1tmA2+1smA2)+(smA2tmA2+1+tmA2smA2)2}.\begin{split}&\overline{|\mathcal{M}|^{2}}_{\chi\bar{\chi}\rightarrow\chi\bar{\chi}}=64\pi^{2}\alpha_{D}^{2}\left\{\left(\frac{5}{2}m_{A^{\prime}}^{4}-4m_{\chi}^{2}m_{A^{\prime}}^{2}+4m_{\chi}^{4}\right)\left[\frac{1}{(t-m_{A^{\prime}}^{2})^{2}}+\frac{1}{(s-m_{A^{\prime}}^{2})^{2}}\right]+\frac{4(m_{A^{\prime}}^{4}-m_{\chi}^{4})}{(t-m_{A^{\prime}}^{2})(s-m_{A^{\prime}}^{2})}\right.\\ &\left.+(3m_{A^{\prime}}^{2}-4m_{\chi}^{2})\left[\frac{s-m_{A^{\prime}}^{2}}{(t-m_{A^{\prime}}^{2})^{2}}+\frac{t-m_{A^{\prime}}^{2}}{(s-m_{A^{\prime}}^{2})^{2}}\right]+6m_{A^{\prime}}^{2}\left(\frac{1}{t-m_{A^{\prime}}^{2}}+\frac{1}{s-m_{A^{\prime}}^{2}}\right)+\left(\frac{s-m_{A^{\prime}}^{2}}{t-m_{A^{\prime}}^{2}}+1+\frac{t-m_{A^{\prime}}^{2}}{s-m_{A^{\prime}}^{2}}\right)^{2}\right\}.\end{split} (90)

Due to crossing symmetry, we can reuse Eq. (LABEL:eq:xx'xx') for the scattering process χ(p1)+χ(p2)χ(p3)+χ(p4)\chi(p_{1})+\chi(p_{2})\rightarrow\chi(p_{3})+\chi(p_{4}) with ss replaced by uu.

VIII.10 Aχχ¯A^{\prime}\leftrightarrow\chi\bar{\chi}

For the decay process Aχχ¯A^{\prime}\rightarrow\chi\bar{\chi}, we take into account the Pauli-blocking effect of the dark fermions. The decay rate is given by

ΓAχχ¯=αD(mA2+2mχ2)3ω|k|2T1eω/Tln{cosh[(1+ξ0)ω/4T]cosh[(1ξ0)ω/4T]},\Gamma_{A^{\prime}\rightarrow\chi\bar{\chi}}=\frac{\alpha_{D}(m_{A^{\prime}}^{2}+2m_{\chi}^{2})}{3\omega|\vec{k}|}\frac{2T}{1-e^{-\omega/T}}\ln{\left\{\frac{\cosh{\left[(1+\xi_{0})\omega/4T\right]}}{\cosh{\left[(1-\xi_{0})\omega/4T\right]}}\right\}}, (91)

where ξ0\xi_{0} is similar to Eq. (60),

ξ0=|k|ω14mχ2mA2.\xi_{0}=\frac{|\vec{k}|}{\omega}\sqrt{1-\frac{4m_{\chi}^{2}}{m_{A^{\prime}}^{2}}}. (92)

Note that in the limit T0T\rightarrow 0, Eq. (91) reduces to the decay rate in vacuum given by Eq. (2.6) in Chang et al. (2018).

The inverse decay rate of χχ¯A\chi\bar{\chi}\rightarrow A^{\prime}, taking into account the bosonic enhancement of the dark photon, is given by

Γχχ¯A=αD(mA2+2mχ2)2E|p|T1+eE/Tln{sinh(ω+/2T)cosh[(ωE)/2T]sinh(ω/2T)cosh[(ω+E)/2T]},\Gamma_{\chi\bar{\chi}\rightarrow A^{\prime}}=\frac{\alpha_{D}(m_{A^{\prime}}^{2}+2m_{\chi}^{2})}{2E|\vec{p}|}\frac{T}{1+e^{-E/T}}\ln{\left\{\frac{\sinh{(\omega_{+}/2T)}\cosh{[(\omega_{-}-E)/2T]}}{\sinh{(\omega_{-}/2T)}\cosh{[(\omega_{+}-E)/2T]}}\right\}}, (93)

where ω±\omega_{\pm} is the same as Eq. (67).

IX Detailed Work Flow for Computing DS Luminosity

  1. 1.

    We take the following steps to determine if a DS particle species is in the diffuse regime or not:

    1. (a)

      Compute λ~i1(Ei,r)\tilde{\lambda}^{-1}_{i}(E_{i},r), the IMFP of each species ii assuming all DS particle species are in thermal equilibrium with the SM medium at temperature T(r)T(r).

    2. (b)

      Compute Δti\Delta t_{i}, the escape timescale for each particle species ii with Eqs. (21) and (22).

    3. (c)

      Compute NiN_{i}, the estimated abundance of each particle species with Eq. (20), and obtain the relative abundance ηiNi/Nieq\eta_{i}\equiv N_{i}/N_{i}^{eq}, where NieqN_{i}^{\rm eq} is the abundance of DS particle species ii assuming they are in thermal equilibrium with the SM medium. If ηi>1\eta_{i}>1, we simply set ηi=1\eta_{i}=1.

    4. (d)

      Compute λi1\lambda^{-1}_{i}, the IMFP of each species rescaled by the relative abundance as

      λi1=possiblefinalstates(λ~ifinalstates1+jηjλ~i+jfinalstates1),\lambda^{-1}_{i}=\sum_{\rm possible\,final\,states}(\tilde{\lambda}^{-1}_{i\rightarrow\rm final\,states}+\sum_{j}\eta_{j}\tilde{\lambda}^{-1}_{i+j\rightarrow\rm final\,states}), (94)

      where jj runs over all possible DS particle species.

    5. (e)

      Check the diffuse criteria for the dark photons: if ηAL,T=1\eta_{A^{\prime}_{L,T}}=1 and λAL,T1(Rν)λ~AL,Tee+1(Rν)>Rν1\langle\lambda^{-1}_{A^{\prime}_{L,T}}(R_{\nu})\rangle-\langle\tilde{\lambda}^{-1}_{A^{\prime}_{L,T}\rightarrow e^{-}e^{+}}(R_{\nu})\rangle>R_{\nu}^{-1}, then AL,TA^{\prime}_{L,T} is in the diffuse limit. Otherwise, AL,TA^{\prime}_{L,T} is treated as nondiffuse particles.

    6. (f)

      Check the diffuse criteria for the dark fermion: if ηχ=1\eta_{\chi}=1 and λχ1(Rν)>Rν1\langle\lambda^{-1}_{\chi}(R_{\nu})\rangle>R_{\nu}^{-1}, then χ\chi is in the diffuse regime. Otherwise, χ\chi is treated as nondiffuse particles.

  2. 2.

    Depending on whether the DS particles are in the diffuse limit or not, we compute the luminosity of each particle species as follows:

    1. (a)

      Regardless of the DS masses, the dark photon luminosity is given by

      LAL,T={LAL,T,diff.,for diffuse AL,T [use Eq. (9)],LAL,T,nondiff.,for nondiffuse AL,T [use Eq. (3)].L_{A^{\prime}_{L,T}}=\begin{cases}L_{A^{\prime}_{L,T},\rm diff.},\,\mbox{for diffuse }A^{\prime}_{L,T}\mbox{ [use Eq.~{}\eqref{eq:flux}]},\\ L_{A^{\prime}_{L,T},\rm nondiff.},\,\mbox{for nondiffuse }A^{\prime}_{L,T}\mbox{ [use Eq.~{}\eqref{eq:LA'emis}]}.\end{cases} (95)
    2. (b)

      When mA<2mχm_{A^{\prime}}<2m_{\chi}, the dark fermion luminosity is given by

      Lχ={Lχ,diff.,for diffuse χ [use Eq. (9)],Lχ,nondiff.,for nondiffuse χ [use Eq. (6)].L_{\chi}=\begin{cases}L_{\chi,\rm diff.},\,\mbox{for diffuse }\chi\mbox{ [use Eq.~{}\eqref{eq:flux}]},\\ L_{\chi,\rm nondiff.},\,\mbox{for nondiffuse }\chi\mbox{ [use Eq.~{}\eqref{eq:Lchiemis}]}.\end{cases} (96)
    3. (c)

      When mA>2mχm_{A^{\prime}}>2m_{\chi}, the dark fermion luminosity is given by

      Lχ={Lχ,diff.,if χ is in the diffuse limit,Lχ,nondiff.,if only AL,AT are in the diffuse limit,23Lχ,nondiff.,if only AT is in the diffuse limit,13Lχ,nondiff.,if only AL is in the diffuse limit,0,if no DS particle species is in the diffuse limit.L_{\chi}=\begin{cases}L_{\chi,\rm diff.},\,\mbox{if }\chi\mbox{ is in the diffuse limit},\\ L_{\chi,\rm nondiff.},\,\mbox{if only }A^{\prime}_{L},A^{\prime}_{T}\mbox{ are in the diffuse limit},\\ \frac{2}{3}L_{\chi,\rm nondiff.},\,\mbox{if only }A^{\prime}_{T}\mbox{ is in the diffuse limit},\\ \frac{1}{3}L_{\chi,\rm nondiff.},\,\mbox{if only }A^{\prime}_{L}\mbox{ is in the diffuse limit},\\ 0,\,\mbox{if no DS particle species is in the diffuse limit}.\end{cases} (97)
    4. (d)

      The total DS luminosity is LX=LAL+LAT+2LχL_{\rm X}=L_{A^{\prime}_{L}}+L_{A^{\prime}_{T}}+2L_{\chi}.

X Radiative Transfer

We now derive the energy flux carried by the DS particles through a surface of radius rr near the neutrinosphere, assuming they are in the diffuse limit. The derivation follows from Appendix I of Prialnik (2000). The equation of radiative transfer is given by

1ρIrcosθ+κIj=0,\frac{1}{\rho}\frac{\partial I}{\partial r}\cos{\theta}+\kappa I-j=0, (98)

where ρ\rho is the matter density, II is the intensity of radiation per unit solid angle per unit frequency, θ\theta is the angle between the direction of radiation and the radial direction, κ\kappa is the opacity, and jj is the total radiation power emitted per unit mass per unit frequency. We distinguish between the opacity contributions from scattering processes and absorption processes κ=κs+κa\kappa=\kappa_{s}+\kappa_{a}, and between the radiation by scattering and radiation emitted by matter j=js+jemj=j_{s}+j_{em}. In equilibrium and isotropic environment,

Iiso,eq\displaystyle I^{\rm iso,eq} B=g(2π)3E2peE/T±1,\displaystyle\equiv B=\frac{g}{(2\pi)^{3}}\frac{E^{2}p}{e^{E/T}\pm 1}, (99)
jemiso,eq\displaystyle j_{em}^{\rm iso,eq} =κaB,\displaystyle=\kappa_{a}B, (100)

where gg is the degree of freedom, and the upper (lower) sign is for fermions (bosons). However, in an anisotropic environment, the relation between jemj_{em} and II is given by

jem=κa(1±eE/T)BκaeE/TI.j_{em}=\kappa_{a}\left(1\pm e^{-E/T}\right)B\mp\kappa_{a}e^{-E/T}I. (101)

In the second term, the minus sign for fermion is due to Pauli blocking, while the plus sign for boson is due to stimulated emission. Substitute the above equation back into Eq. (98),

1ρIrcosθ+κsIjs+κa(IB)=0,\frac{1}{\rho}\frac{\partial I}{\partial r}\cos{\theta}+\kappa_{s}I-j_{s}+\kappa_{a}^{\ast}(I-B)=0, (102)

where κaκa(1±eE/T)\kappa_{a}^{\ast}\equiv\kappa_{a}(1\pm e^{-E/T}). We assume that the radiation intensity II is very close to Iiso,eqI^{\rm iso,eq}. Therefore, we can expand II in terms of Legendre polynomials Pn(cosθ)P_{n}(\cos{\theta}) and substitute it back into Eq. (102). Keeping the terms up to n=1n=1, we obtain

IB1ρ(κa+κs)Brcosθ.I\simeq B-\frac{1}{\rho(\kappa_{a}^{\ast}+\kappa_{s})}\frac{\partial B}{\partial r}\cos{\theta}. (103)

It follows that the energy flux through the spherical surface of radius rr is

L(r)=4πr2IcosθdEdΩ=163π2r2dTdr1ρ(κa+κs)BT𝑑E=2gr23πT3dTdrm/Tξ3ξ2(mT)2λ1(E=ξT,r)eξ(eξ±1)2𝑑ξ,\begin{split}L(r)&=4\pi r^{2}\int I\cos{\theta}dEd\Omega\\ &=-\frac{16}{3}\pi^{2}r^{2}\frac{dT}{dr}\int\frac{1}{\rho(\kappa_{a}^{\ast}+\kappa_{s})}\frac{\partial B}{\partial T}dE\\ &=-\frac{2gr^{2}}{3\pi}\frac{T^{3}dT}{dr}\int_{m/T}^{\infty}\frac{\xi^{3}\sqrt{\xi^{2}-\left(\frac{m}{T}\right)^{2}}}{\lambda^{-1}(E=\xi T,r)}\frac{e^{\xi}}{\left(e^{\xi}\pm 1\right)^{2}}d\xi,\end{split} (104)

where mm is the mass of the particle carrying the radiation, and λ1\lambda^{-1} is the effective IMFP defined as

λ1=ρ(κa+κs).\lambda^{-1}=\rho(\kappa_{a}^{\ast}+\kappa_{s}). (105)

We can identify the absorptive IMFP λabs1\lambda_{\rm abs}^{-1} with ρκa\rho\kappa_{a} and the scattering IMFP λsca1\lambda_{\rm sca}^{-1} with ρκs\rho\kappa_{s}. Thus, we can express the effective IMFP as

λ1(E,r)λabs1(E,r)(1±eE/T)+λsca1(E,r).\lambda^{-1}(E,r)\equiv\lambda_{\rm abs}^{-1}(E,r)(1\pm e^{-E/T})+\lambda_{\rm sca}^{-1}(E,r). (106)

We note that since the integral in Eq. (104) is dominated by the energies with small λ1(E,r)\lambda^{-1}(E,r), it is possible that Eq. (104) could overestimate the energy flux if λ1(E,r)Rν\lambda^{-1}(E,r)\lesssim R_{\nu} for some energies EE. We found that it occurs near the switching of diffuse and nondiffuse regimes, and thus it leads to orders of magnitude jumps of energy luminosity. To avoid this caveat, we approximate the IMFP λ1(E,r)\lambda^{-1}(E,r) by the thermally averaged IMFP λ1(r)\langle\lambda^{-1}(r)\rangle defined as

λ1(r)=d3pf(E,T(r))λ1(E,r)d3pf(E,T(r)).\langle\lambda^{-1}(r)\rangle=\frac{\int d^{3}p\,f(E,T(r))\lambda^{-1}(E,r)}{\int d^{3}p\,f(E,T(r))}. (107)

Thus, the energy flux is approximately given by

L(r)2gr23πT3dTdr1λ1(r)m/Tξ3ξ2(mT)2eξ(eξ±1)2𝑑ξ.L(r)\simeq-\frac{2gr^{2}}{3\pi}\frac{T^{3}dT}{dr}\frac{1}{\langle\lambda^{-1}(r)\rangle}\int_{m/T}^{\infty}\xi^{3}\sqrt{\xi^{2}-\left(\frac{m}{T}\right)^{2}}\frac{e^{\xi}}{\left(e^{\xi}\pm 1\right)^{2}}d\xi. (108)

References