Rate of dark emission from positron in massive

Ermal Rrapaj,1, 2, ∗ Andre Sieverding,1 and Yong-Zhong Qian1 1School of Physics and Astronomy, University of Minnesota, Minneapolis, MN 55455, USA 2Department of Physics, University of California, Berkeley, CA 94720, USA We calculate the rate of production of dark from electron-positron pair annihilation in hot and dense characteristic of progenitors. Given the non-linear dependence of the emission rate on the mass and current astrophysical constraints on the dark photon parameter space, we focus on the mass range of 1–10 MeV. For the conditions under consideration both mixing with the in-medium photon and plasma effects on the electron dispersion relation are non-negligible and are explored in detail. We perform our calculations to the leading order in the fine-structure constant. Transverse and longitudinal photon modes are treated separately given their different dispersion relations. We consider the implications for the evolution of massive stars when dark photons decay either into of the or of the dark sector.

I. INTRODUCTION annihilation is highly favored as stars get sufficiently hot while still at densities far below those of a proto- Observations of galaxy rotation curves, the motion of . Under these conditions we expect pair annihilation ∗ galaxies in clusters, the ΛCDM model of the universe, to be the major emission channel of A . and empirical evidence from the bullet cluster are among among nuclei is a sub-leading source of dark photon the many indicators of the existence of (DM) emission and will be considered in future investigations. that interacts with ordinary matter through gravitational Here, we systematically include medium effects in cal- interactions [1–3]. Many proposed DM models also nat- culating dark photon emission from electron positron urally predict non-gravitational interactions. Here we fo- pair annihilation in massive stars and compare the rates cus on a class of these models, in which DM is part of with those of emission. In particular, we pro- a neutral hidden sector. This sector interacts with stan- vide a complete consistent treatment of plasma effects dard model (SM) particles through the exchange of light on photon and electron dispersion relations. Given that vector that couple to SM conserved currents [4– the dark photon is thermally produced, we focus on the 7]. Dark matter is charged under a local U(1) symmetry mass range 1–10 MeV, which is relevant for the tem- in which the mediator couples to the SM peratures inside massive stars, and is unconstrained by ∗ Q, and is described by the -one field Aµ, called the supernova considerations [11, 12]. We aim to provide a dark photon, which mixes kinetically with the standard self-contained treatment that can serve as a reference for photon Aµ [8]. numerical implementation in stellar evolution. In addi- The Lagrangian of this system is tion, we qualitatively discuss how the dark photon emis- sion might affect massive star evolution. 1 µν 1 ∗ ∗µν  ∗ µν 1 2 ∗ ∗µ = Aµν A A A A A + m ∗ A A , We start by analysing the mixing of the dark photon L −4 − 4 µν − 2 µν 2 A µ with the standard photon in section II A. In section II B which also includes the mass term for the gauge ∗ ∗ ∗ we focus on dark photon emission from electron-positron and where Aµν = ∂µAν ∂ν Aµ,A = ∂µA ∂ν A . − µν ν − µ pair annihilation. In section III A we compare dark pho- The emission and absorption of dark photons are due to ton emission with neutrino emission for conditions typical the oscillations between these two vector bosons. In the of massive stars. We also consider how the dark photon stellar medium, both transverse and longitudinal photon can affect the evolution of these stars. We intend to modes exist and both mix with the dark photon. Because employ the rates provided here in stellar evolution calcu- they have different dispersion relations, they behave dif- arXiv:1904.10567v1 [hep-ph] 23 Apr 2019 lations in subsequent works to quantify the effects of the ferently and we study both. Plasma effects in the kinetic dark photon. mixing have been considered previously both in the exist- ing constraints on the dark photons from SN 1987A and from our sun [9–12]. The purpose of this work is to re- visit the emission rates and mean free path for the dark II. DARK PHOTON PRODUCTION photon for density and temperature ranges relevant for the evolution of massive stars (& 8M ) and study possi- A. Mixing between dark photons and photons ble implications for the unconstrained parameter space. In the core of a supernova, -nucleon ∗ To study the mixing of the two gauge bosons, the dark bremsstrahlung is the dominant process of A pro- photon field is redefined as Bµ = A∗µ Aµ, which trans- duction. However, for pre-supernova massive stars, pair forms the kinetic mixing into a mass− mixing:

1 1 1 ∗ µν µν 2 µ 2 [email protected] = Aµν A Bµν B + m ∗ (Bµ A ) . L −4 − 4 2 A − 2

Following the same procedure as in [9], we perform a drastically only during core collapse, quickly reaching ∼ perturbative expansion in . While production of Bµ is of 14 MeV. Thus, mA∗ mT during most of the stellar order 2, the difference between Bµ and the original dark evolution. Nevertheless, we employ Eq. (4) in numerical photon A∗µ is of order 4, which for practical purposes calculations without any approximation. −9 is negligible (we assume  10 ). The decay width Γ While mT depends only on the properties of the and the imaginary part of the≤ polarization tensor Π [13] medium, i.e., electron density and temperature, the lon- are related by gitudinal plasma frequency ωL is an implicit function of the photon momentum k. For momentum values less Im[Π] = ωΓ = ω(Γabs Γems), (1) than − − − s where ω is the dark photon . By employing de- 3  1 1 + v   ω/T ∗ tailed balance, Γabs = e Γems, the emission rate can kmax = ωP 2 log 1 , v 2v∗ 1 v∗ − be expressed as ∗ − Im[Π] ωL is the solution of the following transcendental equa- Γems = . tion, − ω(eω/T 1) − 2 2 2 2v∗k v∗k + 3ωP ωL = . (5) 2 ωL+v∗k 3ωP log( ) 1 ωL−v∗k − A∗ A A∗ For k > kmax, ωL = k. The plasma frequency ωP and effective electron speed v∗ are given as: Z 2 2 2 4α ke 1 ke FIG. 1. Mixing between dark and standard photon at order ωP = dEe (1 2 )(fe− + fe+ ), O(2) π Ee − 3 Ee Z 2 2 4 2 4α ke 5 ke k ω1 = dEe ( 2 4 )(fe− + fe+ ), Figure 1 shows the leading order diagram of the mixing π Ee 3 Ee − Ee ω1 of the two types of photons, which gives v∗ = . ωP 2 4 2  mA∗ 4 ∗ ΠA =mA∗ + 2 2 + ( ). In the high density or high temperature limit, v∗ 1. ω k ΠA O ∼ − − Unlike mT , because ωL is momentum dependent, it is The dark photon emission rate is not easy to judge the regions of density and temperature where ωL is smaller or larger than the dark photon mass. 2 4  mA∗ ΓAems In our calculations, we always solve the transcendental Γ ∗ = . A ems 2 2 2 2 ω/T 2 2 (ω k Re[ΠA]) + ω (e 1) (ΓAems) equation in order to perform the integrations required for − − − (2) emissivities. For further details on the photon dispersion relation in medium we refer the reader to [14]. The real part of the photon polarization tensor is calcu- From Eqs. (2) and (3) the emission rates for the two lated at one loop level in perturbation theory [14] and is modes are different for transverse and longitudinal modes: 2 4 T T  mA∗ ΓAems Γ ∗ = , T 2 A ems 2 2 2 2 ω/T 2 T 2 Re[Π ] =m , (m ∗ m ) + ω (e 1) (Γ ems) A T A − T − A 2 2 2 (3) 2 4 4 L L ω k 2 mA∗ 2 L  mA∗ ω ΓAems Γ ∗ = . Re[ΠA] = −2 ωL = 2 ωL. A ems 4 2 2 2 6 ω/T 2 L 2 ω ω m ∗ (ω ω ) + ω (e 1) (Γ ems) A − L − A The thermal mass for the transverse mode mT and plasma frequency for the longitudinal mode ωL are cal- (6) culated based on the expressions found in [14]. As the dark photon oscillates into a virtual photon, the virtual It is interesting to note that for a small dark photon mass, photon has the same dispersion relation as the dark pho- the emission rate for the transverse mode is inversely 2 2 2 proportional to that of the photon, whereas for a large ton, ω k = m ∗ . The transverse thermal mass is − A dark photon mass, the two rates are proportional. Z ∞ 2 2 4α ke m = dk (f − + f + ), (4) T π e E e e 0 e B. Dark photon emission from e−e+ annihilation where fe± are the Fermi-Dirac distribution functions for and positrons. For most of the evolution of a We calculate the absorption rate and obtain the emis- massive star, the photon thermal mass is less than 1 MeV sion rate by multiplying the former with the detailed bal- and is mainly a function of density. This mass changes ance factor e−ω/T . As the virtual photon decays into 3 standard model particles, here we employ the dark pho- Neutrino emission through pair annihilation, photo- ton dispersion relation. We emphasize that A∗ in Fig. 2 neutrino process, decay, bremsstrahlung, and is the virtual photon on the dark photon mass shell (a recombination has been studied in detail and standard real photon cannot decay into e−e+ [14]). The decay rate results are available in the literature [14, 16–21]. Us- [13] is ing these standard neutrino cooling rates we compute the evolution for stars with initial masses of 12, 25, and Z Z 3 Z 3 1 1 d pe− d pe+ 30 M and with solar metallicity using the same stellar ΓA∗→e−e+ = dΩω 3 3 2ω 4π 2Ee− (2π) 2Ee+ (2π) evolution code as in [22]. For the conditions covered by 2 these evolutionary tracks, i.e., core temperature and den- ∗ − + (1 f − )(1 f + ) |MA →e e | − e − e sity, we compare the net energy loss rates due to dark 4 (4) (2π) δ (K Pe− Pe+ ). photons and . Before performing a quantita- × − − tive comparison we provide a qualitative assessment of As the transverse and longitudinal photon modes behave which stellar conditions favor the dark photon emission. differently in medium, we calculate the squared matrix el- In Fig. 3 we identify three regions where we expect dark ement for each mode. The detailed expressions are given photon emission to be suppressed due to low tempera- in Appendix A. tures (below the solid line), high density (to the right of the dashed line), and small dark photon mass (to the right of the dot-dashed line), respectively. For relatively e− hot and not too dense environments we can expect dark photon emission to compete with neutrino emission. A∗ e+ FIG. 2. (Dark) virtual photon decay into electron positron pair QA* Q

* T mA < 2me

Plasma effects on the electron and positron dispersion g relations are included to be consistent with the correc- o l T e tion to the photon dispersion relation described in sec- tion II A. In particular, the electron mass gets a correc- T me tion of order α1/2 [15], q m˜ =m /2 + m2/4 + m2 , e e e eff logYe α 2 2 2 meff = (π T + µe). 2 FIG. 3. Conditions regarding dark photon emission. Region below solid line: temperature is very low and there are very This effect was included in [9], and as that work focused few electron-positron pairs. Region to the right of the dashed on a smaller dark photon mass relevant for solar condi- line: density is very high and there are few positrons even tions, pair annihilation was not considered. In this work when temperature is high. Region to the right of the vertical we usem ˜ e in the relevant expressions [see Eq. (A2)]. This dot-dashed line: dark photon mass is lower than the mass of mass can be as high as 12 MeV in a supernova core [12]. the pair in the medium. Only for the region in the upper left corner is dark photon emission expected to potentially exceed neutrino emission. III. IMPLICATIONS FOR STELLAR EVOLUTION In Fig. 4, we show the ratio of the energy loss rates −9 Q˙ A∗ /Q˙ νν¯ for the coupling strength ε = 10 in the dark A. Comparison with neutrino emission photon mass range 1–10 MeV. The stellar evolutionary tracks calculated using neutrino cooling only are also shown in the same figure. There is a high sensitivity The rate of total energy emitted from dark photons is to the dark photon mass for the range 1–5 MeV. In the 5–10 MeV range, the rates are similar and become rel- Z d3k ˙ ∗ ∗ evant at higher temperatures for higher masses. In all QA = 3 ωΓA ems. (2π) the cases shown in Fig. 4, dark photon emission becomes relevant for relatively high temperatures and not very We use the polarization averaged rate, high densities, in agreement with the qualitative assess- ment shown in Fig. 3. Indeed, for smaller dark photon 1 T L  Γ ∗ = Γ ∗ + Γ ∗ . A ems 3 A ems A ems masses we see the expected suppression when electrons 4

electron-positron pair, it decays back into such a pair and 10.00 10.00 is reabsorbed by standard matter. In this scenario, from 9.75 9.75

) ) an astrophysical perspective, it could be understood as a

K 9.50 K 9.50 / /

T T short-lived that transports energy from the inner ( 9.25 ( 9.25 0 0 1 1 region of a star to the outer regions. g 9.00 g 9.00 o o l l However, if the dark photon is an actual portal to 8.75 8.75

8.50 8.50 the dark sector, its impact on stellar evolution is quite 1 3 5 7 9 11 1 3 5 7 9 11 3 3 radical. It could decay, for instance, into a pair of log10 (Ye /gcm ) log10 (Ye /gcm ) dark , and this decay may dominate over its (a) mA∗ = 1.21 MeV. (b) mA∗ = 2 MeV. re-absorption by standard model particles. This sce- nario has been explored in electron beam dumps for ex- 10.00 10.00 perimental constraints [23, 24]. For stellar evolution, 9.75 9.75

) ) dark photons would act as a cooling agent in this sce-

K 9.50 K 9.50 / / T T ( 9.25 ( 9.25 nario. If there is emission, there must also be absorp- 0 0 1 1

g 9.00 g 9.00 tion. From the rates calculated in the previous sections, o o l l 8.75 8.75 the mean free path for dark photon absorption is given by −1 8.50 8.50 λ = Γ ∗ − + /v . For typical stellar conditions, elec- 1 3 5 7 9 11 1 3 5 7 9 11 A →e e ω 3 3 log10 (Ye /gcm ) log10 (Ye /gcm ) trons and positrons are non-degenerate as densities are not too high. As the pre-supernova temperature reaches (c) m ∗ = 3 MeV. (d) m ∗ = 4 MeV. A A maximal values of 0.1 MeV, and dark photons are ther- ∼ 10.00 10.00 mally produced, we expect them to be non-relativistic p ∗ 9.75 9.75 with a typical momentum value 2mAT . Under these

) ) ∼

K 9.50 K 9.50 conditions, decay into a standard electron-positron pair / / T T ( 9.25 ( 9.25 can be approximated by decay in and the mean 0 0 1 1

g 9.00 g 9.00 free path is given by o o l l 8.75 8.75 s 8.50 8.50 2 2 2 1 3 5 7 9 11 1 3 5 7 9 11 −1 2 (mA∗ + 2me) 4me 3 3 (7) log10 (Ye /gcm ) log10 (Ye /gcm ) λA∗→e−e+ =α 1 2 . 3k − mA∗ (e) mA∗ = 5 MeV. (f) mA∗ = 10 MeV. From Eq. (7) and the stellar conditions we deduce a mean FIG. 4. Ratio of dark photon emissivity from pair annihi- free path of the order of 108–109 cm for a reduced cou- lation to neutrino emissivity as function of electron density pling strength  = 10−9. Were this the dominant decay and temperature assuming the reduced coupling strength is  = 10−9. Stellar evolutionary tracks for a 30 M (red dot- mode, dark photons would indeed act as a mechanism for redistributing energy from the central part of a star dashed), 25 M (green solid), and 12 M (blue dashed) are also shown. The contour lines represent ratios of 10 (solid), 1 to regions of lower density. This effect has been stud- (dashed) and 0.1 (dot-dashed). ied in core collapse supernovae to further constrain the parameter space [25]. However, if the dark photon could decay into a pair of in medium become too massive. Figure 4 shows that the dark fermions (XX¯), this channel might be much more dark photon emission would dominate over the neutrino efficient. As a consequence, the supernova constraint emission during the advanced stellar burning phases. For from energy redistribution would not apply in this case. the conditions reached in the stellar core the dark photon As an illustration, we can assume mX me and the cou- energy loss rate can exceed the neutrino losses by several pling of the dark photon to dark fermions∼ (dark electron- orders of magnitude. We therefore expect dark photons positron pair) to be the same as the standard QED cou- in the mass range of 1–10 MeV to affect the last stages pling. The ratio of the two mean free paths would be 2 −18 of stellar evolution. In particular, for small mass values λA∗→XX¯ /λA∗→e−e+  = 10 . Once produced, the of 1–3 MeV, the dark photon emission starts to overtake dark photon would essentially∼ always decay into the dark neutrino emission already towards the end of the carbon sector. In this scenario, it is effectively an additional cool- burning phase. ing mechanism. In future works we plan to explore the effects of this additional cooling on stellar evolution in detail. B. Dark photon decay

The way in which the dark photon can affect the stellar IV. CONCLUSIONS environment depends crucially on what happens after the particle is emitted. If the dark photon is not a portal We have calculated the dark photon emission rates for to the dark sector, but merely one more particle yet to both transverse and longitudinal modes by taking into be discovered, and if its mass is larger than that of the consideration all plasma effects at leading order in the 5

fine structure constant. We have compared the energy ACKNOWLEDGMENTS loss rates for dark photon emission from electron-positron annihilation with standard neutrino cooling in massive stars and found that the former can easily overtake the We thank Sam McDermott for discussion. We also latter for dark photon masses in the range 1–4 MeV. If acknowledge Alexander Heger for providing KEPLER, the dark photon is a portal to the dark sector, it becomes the hydrodynamics code for stellar evolution, and for an additional cooling mechanism. We intend to investi- helping set up the stellar evolution calculations. This gate the detailed effects of the combined cooling by dark work was supported in part by the NSF (PHY-1630782), photons and neutrinos on stellar evolution in follow-up the Heising-Simons Foundation (2017-228), and the DOE work. (DE-FG02-87ER40328).

Appendix A: Transverse and longitudinal dark photon production from electron–positron pair annihilation

To obtain the rates for these two modes we need to decompose the scattering matrix using projection operators

T qiqj T T T P = (δij ),P = P = P = 0, ij − q2 00 0µ µ0 2 2 i j L k ω k k Pµν = ( 2 , 2 2 ), mA∗ mA∗ k where P L is three-longitudinal and P T is three-transverse to the momentum ~k of the photon [26]. The squared scattering matrix element for each mode is

T 2 2 2 T T  µ ν  ∗ − + =e   (q, λ) (q, λ) (p/ +m ˜ e)γ (p/ m˜ e)γ |MA →e e | µ ν 2 3 − λ 2 2 T  µ ν ν µ 2 µν  =4e  P (p p + p p (m ˜ + p2 p3)g ) µν 2 3 2 3 − e · ~ ~ 2 2 (~pe− k)(~pe+ k)  =32πα m˜ + p − p + · · + (~p − ~p + ) , e e · e − k2 e · e L 2 2 2 X L L  µ ν  ∗ − + =e   (q, λ) (q, λ) (p/ +m ˜ e)γ (p/ m˜ e)γ |MA →e e | µ ν 2 3 − λ 2 2 L  µ ν ν µ 2 µν  =4e  P (p p + p p (m ˜ + p2 p3)g ) µν 2 3 2 3 − e · 2 2 ~ ~ 2 2 k ω (~pe− k)(~pe+ k) ω ~ ~ m˜ e + pe− pe+  =32πα 2 Ee− Ee+ + 2 · 2 · 2 [Ee− (~pe+ k) + Ee+ (~pe− k)] + · . mA∗ mA∗ k − mA∗ · · 2 By employing energy-momentum conservation, the above results can be simplified as

T 2 2 A∗→e−e+ 2 mA∗ 1 2 2 2  |M | =m ˜ + (ω + k ) ωE − + E − , 32πα2 e k2 4 − e e L 2 2 A∗→e−e+ mA∗ 1 2 2  |M | = m ∗ + ωE − E − . 32πα2 k2 − 4 A e − e The integration over the phase-space can be performed analytically to give

+ − Fn(k, mA∗ , µe,T ) = n(ω, µe,T,E − ) n(ω, µe,T,E − ), (A1) F e − F e where Z n    n(ω, µe, T, x) = dx x 1 f(x, µe,T ) 1 f(ω x, µe,T ) , F − × − − − q 2 2 ω = k + mA∗ , v s u  2 2 ± u 2 1 m˜ e Ee− =tm˜ e + ω 1 4 2 k . 4 − mA∗ ± 6

Equation (A1) can be expressed analytically for all values of n. The first two expressions are

x−µe ! T  1 + e t  (ω, µ , T, x) = log x , 0 e −ω/T ω+µ−x F 1 e 1 + e T − −  x−µe !  T 1 + e t x−µe ω−x+µe 1(ω, µe, T, x) = log + T Li2( e T ) T Li2( e T ) , −ω/T − ω−x+µe F 1 e 1 + e T − − − − where Lin(x) is the poly-logarithmic function of order n of x. Expressions for higher values of n can be obtained in a similar fashion. Finally, the emission rates can be written as

2 2 T −ω/T 2αmA∗ 2 X T Γ − + ∗ (ω, k) = e  c F (k, m ∗ , µ ,T ), e e →A ωk3 n n A e n=0 2 2 (A2) L −ω/T 2αmA∗ 2 X L Γ − + ∗ (ω, k) = e  c F (k, m ∗ , µ ,T ), e e →A ωk3 n n A e n=0 where

2 T k 2 1 2 2 T T c0 = 2 m˜ e + (ω + k ), c1 = ω, c2 = 1, m ∗ 4 A − (A3) L 1 2 L L c = m ∗ , c = ω, c = 1. 0 −4 A 1 2 − We have verified that these expressions agree with those in [11].

[1] K. A. Olive, in and cosmology: The [17] Y. Kohyama, N. Itoh, and H. Munakata, Astrophys. J. quest for physics beyond the standard model(s). (2003), 310, 815 (1986). pp. 797–851, astro-ph/0301505. [18] N. Itoh, T. Adachi, M. Nakagawa, Y. Kohyama, and [2] K. Freese, EAS Publ. Ser. 36, 113 (2009), 0812.4005. H. Munakata, Astrophys. J. 339, 354 (1989). [3] M. Roos, arXiv e-prints arXiv:1001.0316 (2010), [19] N. Itoh, H. Mutoh, A. Hikita, and Y. Kohyama, Astro- 1001.0316. phys. J. 395, 622 (1992). [4] B. Holdom, Physics Letters B 166, 196 (1986), ISSN [20] N. Itoh, H. Hayashi, A. Nishikawa, and Y. Kohyama, 0370-2693. Astrophys. J. 102, 411 (1996). [5] S. Rajpoot, Phys. Rev. D 40, 2421 (1989). [21] M. Haft, G. Raffelt, and A. Weiss, Astrophys. J. 425, [6] A. E. Nelson and N. Tetradis, Phys. Lett. B221, 80 222 (1994), astro-ph/9309014. (1989). [22] S. Woosley and A. Heger, Physics Reports 442, 269 [7] B. Batell, P. deNiverville, D. McKeen, M. Pospelov, and (2007), ISSN 0370-1573, the Hans Bethe Centennial Vol- A. Ritz, Phys. Rev. D90, 115014 (2014). ume 1906-2006. [8] J. Jaeckel and A. Ringwald, Ann. Rev. Nucl. Part. Sci. [23] E. Izaguirre, G. Krnjaic, P. Schuster, and N. Toro, Phys. 60, 405 (2010). Rev. D88, 114015 (2013), 1307.6554. [9] J. Redondo and G. Raffelt, JCAP 1308, 034 (2013), [24] B. Batell, R. Essig, and Z. Surujon, Phys. Rev. Lett. 113, 1305.2920. 171802 (2014), 1406.2698. [10] E. Rrapaj and S. Reddy, Phys. Rev. C94, 045805 (2016), [25] A. Sung, H. Tu, and M.-R. Wu (2019), 1903.07923. 1511.09136. [26] H. A. Weldon, Phys. Rev. D 26, 1394 (1982). [11] J. H. Chang, R. Essig, and S. D. McDermott, JHEP 01, 107 (2017), 1611.03864. [12] E. Hardy and R. Lasenby, JHEP 02, 033 (2017), 1611.05852. [13] H. A. Weldon, Phys. Rev. D 28, 2007 (1983). [14] E. Braaten and D. Segel, Phys. Rev. D48, 1478 (1993), hep-ph/9302213. [15] E. Braaten, Astrophys. J. 392, 70 (1992). [16] H. Munakata, Y. Kohyama, and N. Itoh, Astrophys. J. 296, 197 (1985).