MNRAS 000, 1–10 (2021) Preprint 4 January 2022 Compiled using MNRAS LATEX style file v3.0 An INTEGRAL/SPI View of Reticulum II: Particle Dark Matter and Primordial Black Holes Limits in the MeV Range Thomas Siegert,1,4? Celine Boehm,2 Francesca Calore,3 Roland Diehl,4 Martin G. H. Krause,5 Pasquale D. Serpico,3 and Aaron C. Vincent6 1Institut für Theoretische Physik und Astrophysik, Universität Würzburg, Campus Hubland Nord, Emil-Fischer-Str. 31, 97074 Würzburg, Germany 2Sydney Consortium for Particle Physics and Cosmology, School of Physics, The University of Sydney, NSW 2006, Australia 3Univ. Grenoble Alpes, Univ. Savoie Mont Blanc, CNRS, LAPTh, F-74940 Annecy, France 4Max-Planck-Institut für extraterrestrische Physik, Gießenbachstraße, 85741 Garching b. München, Germany 5Centre for Astrophysics Research, Department of Physics, Astronomy and Mathematics, University of Hertfordshire, College Lane, Hatfield, Hertfordshire AL10 9AB, UK 6Department of Physics, Enigneering Physics and Astronomy, Queen’s University, Kingston, ON, K7L 3N6, Canada Accepted XXX. Received 4 January 2022; in original form ZZZ ABSTRACT Reticulum II (Ret II) is a satellite galaxy of the Milky Way and presents a prime target to investigate the nature of dark matter (DM) because of its high mass-to-light ratio. We evaluate a dedicated INTEGRAL observation campaign data set to obtain γ-ray fluxes from Ret II and compare those with expectations from DM. Ret II is not detected in the γ-ray band 25–8000 keV, and we derive a flux limit of . 10−8 erg cm−2 s−1. The previously reported 511 keV line is not seen, and we find a flux limit of . 1.7 × 10−4 ph cm−2 s−1. We construct spectral models for primordial black hole (PBH) evaporation and annihilation/decay of particle DM, and subsequent annihilation of e+s produced in these processes. We exclude that the totality of DM in Ret II is made of a monochromatic distribution of PBHs of masses . 8 × 1015 g. Our limits on the velocity-averaged DM annihilation cross section into e+e− are 〈σv〉 . 5 × 10−28 (mDM/MeV)2.5 cm3 s−1. We conclude that analysing isolated targets in the MeV γ-ray band can set strong bounds on DM properties without multi-year data sets of the entire Milky Way, and encourage follow-up observations of Ret II and other dwarf galaxies. Key words: galaxies: individual: Reticulum II – gamma-rays: galaxies – dark matter – black hole physics 1 INTRODUCTION Dark matter (DM) is the most abundant component of the matter content of the Universe. Evidence for DM exists at different scales, from galaxy to cosmological background radiation, however the very nature of this elusive component remains a mystery (see Bertone & Hooper 2018, and references therein for a historical review).Owing to the vast parameter space for DMcandidates and possible interactions, different astroparticle observables can be used to tackle the nature of DM, from cosmic surveys to near field cosmology, from high-energy radiation to gravitational waves (see Alves Batista et al. 2021, for a state-of-the-art review). In particular, signals of DM in high-energy radiation from radio frequencies to γ-rays are expected from different targets. The choice of the ‘best’ target of interest is determined, on the one hand, by maximising the expected DM signal, and, on the other hand, by reducing the usually strong astrophysical fore- and background. Despite their abundance, dwarf galaxies are dim objects, some- times including only a few thousand stars, thus resulting among the most pristine, chemically unevolved, and possible DM dominated self-gravitating structures known (Simon 2019). As a result, they are a primary target for near-field cosmology and indirect searches ? E-mail: tho.siegert@gmail.com for DM, benefiting of a high (putative) exotic signal to astrophys- ical background ratio (e.g., Winter et al. 2016). High-energy pho- tons in the X- and γ-ray band from such objects may come from the annihilation of conventional DM candidates like weakly inter- acting massive particles (WIMPs), from decays of alternative DM candidates like sterile neutrinos, or from the Hawking evaporation process (Hawking 1971) of sufficiently light black holes, produced in the early universe. Interestingly, Siegert et al. (2016c) reported a 3.1σ signal of a 511 keV line from the direction of the dwarf galaxy Reticulum II (Ret II; distance d = 30 ± 2 kpc; Koposov et al. 2015). The line flux from this previousworkwas extraordinarily high, F511 = (1.7± 0.5) × 10−3 ph cm−2 s−1, which would make Ret II the most luminous 511 keV source in the sky. The signal has been inter- preted in terms of a recent binary neutron star (NS) merger (Fuller et al. 2018), possibly outshining the entire galaxy. Ret II has also been reported to show a 2.3–3.7σ excess at GeV energies (Geringer- Sameth et al. 2015) which might point to a dark matter (DM) origin of the signal. Follow-up analyses found no flux enhancement above the diffuse GeV emission (Albert et al. 2017). Intrigued by these findings, the INTEGRAL (Winkler et al. 2003) satellite performed a dedicated observation campaign to follow up the signal. In this work, we analyse the new 1Ms data together with archival data from the coded-mask γ-ray spectrometer onboard INTEGRAL, SPI (Ve- drenne et al. 2003). We use our findings to set limits on possible DM © 2021 The Authors 2 Thomas Siegert et al. models, including primordial black holes (PBHs) and annihilation or decay of (light) DM particles. This paper is structured as follows: In Sec. 2, we describe the new observations, and how SPI data are analysed in general. We build photon-emission models expected from the evaporation of PBHs and annihilating/decaying particle DM in Sec. 3. Here, we also include the possible annihilation of positrons (e+s) as a DM product into the galaxy emission spectra. Our results are presented in Sec. 4. We conclude in Sec. 5. 2 NEW OBSERVATION CAMPAIGN The previous analysis by Siegert et al. (2016c) focussed on the search of 511 keV signals from the 39 satellite galaxies of the Milky Way (MW) known at the time. In their data set, the entire sky was mod- elled and the diffuse 511 keV emission from the MW itself was included. Before the new observation campaign, Ret II had never been targeted for observations by SPI because 1) the galaxy was only discovered in 2015 (Koposov et al. 2015), and 2) other known X- or γ-ray sources are separated from Ret II by at least 18◦. However, thanks to SPI’s wide fully-coded field of view (FCFOV) of 16◦×16◦ (30◦ × 30◦ partially coded), Ret II data were collected also before its discovery. Outside the full coding region, the effective area of the instrument sharply drops which placed previous Ret II observations right onto exposure edges with a cumulative on-target observation time of 550 ks, split over ten years. While these effect were taken into account in the analysis, dedicated observations, pointed directly to Ret II, would be suitable to validate earlier findings. 2.1 Data Set Our new data set includes previous observations between 2003 and 2013 and adds ∼ 1Ms during the Ret II campaign in 2018. We selected all pointed observations (pointings) whose distance to Ret II is less than 16◦ so that our target is always in the FCFOV. This amounts to 603 pointings with an average observation time of 3300 s each. The mean dead time of a working detector is about 18.2%. Since the launch of the INTEGRAL mission in October 2002, four of the 19 SPI detectors have failed, reducing the effective area by ≈ 20%. The total dead-time- and efficiency-corrected exposure time then amounts to 1.38Ms. We use SPI’s bandpass between 25 and 8000 keV, and perform our analysis in 29 logarithmic energy bins, plus a single narrow bin for the 511 keV line from 508 to 514 keV. 2.2 General Analysis Method SPI data analysis relies on the comparison of expected detector il- luminations, that is, patterns in the 19-detector data space, with the measured count rate per pointing and energy. Fluxes are estimated by maximising the Poisson likelihood, L (D |M) = ∏ p m dp p exp(−mp) dp! , (1) where dp are the measured counts in pointing p, and mp is the model expectation. MeV γ-ray measurements are described by a two component model, one representing the instrumental background mBG, and one representing the celestial emissionmSKY. Instrumental background originates from cosmic-ray interactions with the satellite material and leads to continuum (C) and line (L) backgrounds (Diehl et al. 2018). Spatial emission models are convolved with the imaging response function of SPI, that is, the mask coding, leading to a 19- element vector of expected source counts for each pointing. The total model reads mp = mBGp + m SKY p = = ∑ t,t′  ∑ i∈{C,L } βi,tRBGp,i + NS∑ j=1 αj,t′RSKYp;lbM lb j (φ j )  , (2) where RBG p,i is the backgroundmodel (response) appropriate for point- ing p, scaled by the amplitudes βi,t for background components i ∈ {C, L} varying on timescales t, RSKY p;lb is the coded-mask response in pointing p, converting the j = 1 . . . NS sky models Mlbj (φ) from image space lb to the data space. The source models may have additional parameters φ j , for ex- ample defining the position of point sources, and are scaled by the amplitudes αj,t′ , possibly varying on another time scale t ′. Given SPI’s angular resolution of 2.7◦, and the expected maximum extent of Ret II’s DM halo of 1◦ (Evans et al. 2016), we model the galaxy as point source, MRet II(l, b) = δ(l − l0)δ(b − b0), (3) with the Galactic coordinates of Ret II (l0, b0) = (266.30,−49.74). All other sources that can be expected in our data set, EXO0748- 676, ESO33-2, SMCX-1, and LMCX-4 (c.f. Bouchet et al. 2011) are modelled as point sources as well. The fitted parameters in Eq. (2) are βi,t and αj,t′ . The fit is performed with spimodfit (Halloin 2009), which creates a spectrum by looping over the energy bins defined in our data set. Details about the procedure are found in Siegert et al. (2019). 2.3 Spectral Fitting The spectrum created in this way is subject to the instrument disper- sion, that is, the probability that a photon with initial energy Ei is measured at a final energy E f owing to scattering in the instrument. In order to correct for the intrinsic dispersion, expected spectral models dF dE are convolved with the energy-redistribution matrix H(Ei, E f ) that forward-folds the models into the appropriate data space. Thus, the spectral models read dF ( E f ;ψ ) dE f = ∫ dEiH(Ei, E f ) dF (Ei ;ψ)dEi , (4) with dFdE f being the folded model which is compared to the extracted flux values from Sec. 2.2, and dFdEi is the intrinsic source model that depends on a set of spectral parametersψ. We use theMulti-Mission- Maximum-Likelihood (3ML, Vianello et al. 2015) framework to perform spectral fits. Details about the choices of prior probabilities and parameter ranges are given in AppendixA. We note that, ideally, the steps explained in Secs. 2.2 and 2.3 should be performed in one single step to enhance the sensitivity of the parameter estimation, which is, however, not yet implemented in the current software release. 3 EXPECTATIONS IN THE MEV BAND No MeV emission has ever been detected from a (dwarf) satellite galaxy. Therefore, the expected signals have a large variety, ranging from γ-ray line emission due to radioactive decays and (subsequent) e+ emission. Population synthesis models estimate a 1.8MeV γ-ray MNRAS 000, 1–10 (2021) MeV Dark Matter Limits in Reticulum II 3 line flux from the decay of 26Al of the order of 10−6 ph cm−2 s−1 from the Large Magellanic Cloud (LMC), about one order of magni- tude below SPI’s sensitivity Diehl et al. (2018). Because the expected line flux is related to the star formation and supernova rate, any other satellite galaxy, and in particular Ret II, can be expected to show no significant excess at 1.8MeV. Cordier et al. (2004) performed a search for a 511 keV line owing to DM annihilation/decay from the Sgr dwarf galaxy, but found no excess. Siegert et al. (2016c) ex- tended this search to all satellite galaxies of the MW, which resulted in one 3σ signal from the direction of Ret II out of 39 tested galax- ies. Where this signal, if true, comes from is unknown, and might point to different origins in terms of NS mergers (Fuller et al. 2018), accreting X-ray binaries (Siegert et al. 2016a), or DM (Boehm et al. 2004). All these scenarios would show additional emission features in the MeV band besides a sole 511 keV line. In particular additional broad γ-ray lines from other nucleosynthesis products inNSmergers, continuum emission from a population of X-ray binaries, or radia- tive corrections in the final products from DM decay/annihilation or PBH decay. Therefore, we extract the spectrum of Ret II and nearby sources using the logarithmic energy binning defined in Sec. 2.1, plus a single bin for the 511 keV line. In this way, we can assess the previously found signal from Ret II in a model-independent fashion, and determine model dependent parameters using the additionally- expected components in Sec. 3.1. 3.1 Modelling the MeV Band in Ret II We focus on the γ-ray spectra expected from the evaporation of PBHs in the mass range 1014–1018 g (see Sec. 3.1.1) and annihi- lating/decaying particle DM in the mass range 0.05–300MeV (see Sec. 3.1.2). These black hole and particle mass ranges would result in soft γ-ray (MeV) emission with characteristic spectra that INTE- GRAL/SPI is sensitive to. The absolute fluxes as well as detailed spectral shapes of such DM signals are strongly model dependent and are based on unknown astro- and particle-physics input such as the size of the galaxy’s halo, particle cross sections, or the magnitude of secondary particles involved. Using literature values (see below), a DM signal from Ret II that imprints in the MeV photon band would bewithin reach in our renewed observation data set. Other signals, for example the γ-ray emission from old NS mergers remnants, are not part of our predictions because the expected emission of one 10 kyr old remnant in Ret II is about onemillion times lower (Korobkin et al. 2020) than the instrument sensitivity. While emission from GeV DM would also imprint in the MeV band, SPI is most sensitive to annihi- lation/decay to e+e− and subsequent emission processes considering the final state radiation (FSR) of the pairs as well as the annihilation of e+s during propagation and after thermalisation. Therefore, we restrict our models to the two channels e+e− and γγ, and present one µ+µ− example for comparison. We follow the formalism of Fortin et al. (2009) throughout the paper. If the DM microscopic emission process (e.g., annihilation cross- section 〈σv〉) does not depend on kinetic properties (notably veloc- ity distribution) the flux from DM haloes factorises into a ‘particle physics’ (owing the underlying DM model) and an ‘astrophysics’ component (which depends on the distribution of DM in astrophysi- cal systems) (Bergström et al. 1998) as dFγ ( Eγ; κ,M,ψ, n ) dEγ = κ 4piM dNγ ( Eγ;ψ ) dEγ︸ ︷︷ ︸ particle physics: PBHorDM ∫ los drdΩρnhalo(r,Ω)︸ ︷︷ ︸ astrophysics: D or J , (5) Model κ M n ψ PBH fPBH MBH 1→ D { fPs, E max kin } DM Annihilation 〈σv〉 2mDM 2→ J { fPs, E max kin } DM Decay 1/τ mDM 1→ D { fPs, E max kin } Table 1. Parameters for considered DM models in Eq. (5). where κ, M , and ψ describe the parameters of interest for the specific cases of PBH evaporation, self-conjugated DM annihilation or decay (see Tab. 1), and the last term is the D- (n = 1) or J-factor (n = 2) of the galaxy. They describe the extent and normalisation of the emission, calculated by an integration of the galaxy’s halo profile ρhalo(r,Ω) over the line of sight (los). The different scaling for decay (evaporation) vs. annihiliation originates from the fact that while for the former process only one particle (PBH) induces the γ-ray emission, for the latter two particles are needed for the annihilation to take place, and therefore the emissivity scales as ρ2halo(r,Ω). The photon spectrum dNγ (Eγ ;ψ)dEγ is related to the photon density of states as 1Nγ dNγ dEγ = 1κ dκ dEγ so that the number of emitted photons per process is Nγ = ∫ dEγ dNγ dEγ . Nγ is thus normalising the source spectrum, Eq. (5), in terms of the totally-emitted photons. We use the range of D- and J-factors of Ret II from the lit- erature (Bonnivard et al. 2015; Evans et al. 2016; Albert et al. 2017), in particular D ≈ (1–4) × 1018 GeV cm−2, and J ≈ (0.2– 3.7) × 1019 GeV2 cm−5. The conversion of the D- and J-factor into other often-used units are D = (0.9–3.5) × 10−2M pc−2 = (1.9– 7.3) × 10−6 g cm−2, and J = (0.6–8.4) × 10−3M2 pc−5 = (0.8– 11.8) × 10−29 g2 cm−5. We use these ranges as uniform priors in our spectral fits to include uncertainties in the DM halo as well as the distance to Ret II. Depending on the DM model, the parameters of interest change and are listed in Tab. 1. Here, fPBH is the fraction of DM made of PBHs (unitless), 〈σv〉 is the velocity-averaged DM annihilation cross section in units of cm3 s−1, τ = Γ−1 is the decay time of a DM particle in units of s, MBH is the mass of PBHs in units of g, and mDM is the DM particle mass in units of keV. The additional spectral model parameters ψ = { fPs, Emaxkin } appear when e+ annihilation is included (see below). 3.1.1 Primordial Black Hole Evaporation The spectrum of an evaporating black hole (BH) of mass MBH due to Hawking radiation (Hawking 1975; Page & Hawking 1976) is( dNi dEi ) BH = 1 2pi Γi(Ei,MBH) exp ( Ei TBH ) − (−1)2si . (6) In Eq. (6), Ei is the energy of particle i and si its spin. Γi(Ei,MBH) is the ‘greybody’ factor that alters the expected blackbody distribution of possible emitted particles, given the BH temperature TBH = ~c3 8piGkBMBH = 6 × 10−8 ( M MBH ) K = 1.06 ( 1016 g MBH ) MeV, (7) where ~ is reduced Planck’s constant,G is the gravitational constant, c is the speed of light, and kB is Boltzmann’s constant. We use BlackHawk (Arbey & Auffinger 2019) to calculate the spectra of all relevant particles in the energy range between 1 keV and 1GeV. BlackHawk allows to include secondary particle production due to hadronization, fragmentation, decay, and other processes as a result of BH evaporation, which we add to our spectra. MNRAS 000, 1–10 (2021) 4 Thomas Siegert et al. 101 102 103 104 E [keV] 10 7 10 6 10 5 10 4 10 3 10 2 10 1 100 E × F [k eV ph cm 2 s 1 ke V 1 ] PBH Evaporation (primary) PBH Evaporation (secondary) Positronium Decay Annihilation in Flight Annihilation in Flight (weighted) Internal Bremsstrahlung Internal Bremsstrahlung (weighted) Total Spectrum (MBH = 1016 g, D = 2.5 × 1018 GeV cm 2, fPs = 1.0, Ekin, maxe = 4.0 MeV) 102 104 Ekine [keV] 10 1 101 Ek in e × F e [1 0 3 ] 101 102 103 104 E [keV] 10 7 10 6 10 5 10 4 10 3 10 2 10 1 100 E × F [k eV ph cm 2 s 1 ke V 1 ] PBH Evaporation (primary) PBH Evaporation (secondary) Positronium Decay Annihilation in Flight Annihilation in Flight (weighted) Internal Bremsstrahlung Internal Bremsstrahlung (weighted) Total Spectrum (MBH = 3 × 1015 g, D = 2.5 × 1018 GeV cm 2, fPs = 0.1, Ekin, maxe = 10.0 MeV) 102 104 Ekine [keV] 10 1 101 Ek in e × F e [1 0 3 ] Figure 1. Expected γ-ray spectra from PBH evaporation and subsequent an- nihilation of positron-electron pairs, plus internal bremsstrahlung of emitted pairs. The two panels show different model parameters (see legends) for a D- factor of 2.5×1018 GeV cm−2, consistent with Reticulum II. The small insets show the distribution of electrons and positrons from evaporation, together with the maximum kinetic energy Emaxkin considered by the shaded area, and the mean kinetic energy of the electron/positron population (dashed line). The shaded area of the electron/positron distribution corresponds to the shaded areas in the γ-ray spectra. The production of e+s from BH evaporation has no kinematic threshold based on the mass of the BH. However, the number of e+s produced that may annihilate subsequently can lead to a 20–8000 keV flux stronger than the evaporation signal itself which is given by a BH mass of. 1.1× 1017 g. Given the particle spectrum of e+s from BlackHawk, the differential e+ flux in a galaxy with D-factor D is given by dFe dEe = 1 4piMBH dNe dEe D, (8) so that the total possible e+e−-annihilation flux FAnn in that galaxy is FAnn = ∫ Emaxkin 0 dEkin dFe dEe,kin . (9) Here, Emaxkin is the maximum kinetic energy of a e + that annihilates within the galaxy and does not escape into the intergalactic medium. For the MW, Emaxkin has been estimated to be around 3–7MeV (Bea- com & Yüksel 2006; Sizun et al. 2006). We caution, however, that these estimates may be inaccurate because the statistical methods to compare the expected spectrumwith the measurements of two differ- ent instruments are not rigorously correct. Therefore, we leave Emaxkin as a free parameter in our spectral fits. Because the transport and environmental conditions in Ret II are largely unknown, this proce- dure effectively takes into account these uncertainties. For reference, we discuss the two extreme cases of no and maximal e+ annihila- tion in our results, Sec. 4, together with the intermediate solution of marginalising over Emaxkin . The total annihilation flux is composed of three components, FAnn = F511 + FoPs + FIA, (10) where F511 is the flux in the 511 keV line due to direct annihilation with e−s and intermediate formation of Positronium (Ps), FoPs is the three-photon decay flux of ortho-Ps, and FIA is the total flux of e+s annihilating in flight (IA) before stopping/thermalising. The ortho-Ps and the line flux are related by quantum statistics as FoPs = 9 fPs 8 − 6 fPs F511 =: r32F511, (11) with fPs being the Ps fraction, that is, the number of e+s forming Ps before annihilating (Leventhal et al. 1978). The scaling factor r32 ranges between 0 ( fPs = 0) and 4.5 ( fPs = 1). In general fPs depends on the annihilation conditions, for example the gas in which e+s anni- hilate, which is a function of the temperature, density, and ionisation fraction, among others (e.g., Jean et al. 2006; Churazov et al. 2005, 2011; Siegert et al. 2016b). Since the true conditions in Ret II are unknown, we leave fPs as free parameters in our spectral fits. The ortho-Ps spectrum ( dNγ dEγ ) oPs has been calculated by Ore & Powell (1949), and we model the 511 keV line, ( dNγ dEγ ) 511 as a Gaussian at 511 keV with a width according to SPI’s energy resolution (Diehl et al. 2018; Attié et al. 2003). Beacom & Yüksel (2006) derived a relation between FIA and the 511 keV line flux, FIA = 1 1 − 34 fPs 1 − P P F511 =: rIAF511, (12) where P = P(Ekin,me) is the probability of a e+ with initial kinetic energy Ekin to annihilate before stopping/thermalising. Therefore, the total annihilation flux can be expressed as FAnn = (1 + r32 + rIA)F511. (13) The survival probability in Eq. (12) is given in general by P(Ekin,me) = exp ©­­«−nX ∫ Ekin me dE σ(E) dE(E,nX )dx ª®®¬ , (14) where σ(E) is the total annihilation cross section, and dE(E,nX )dx is the stopping power (energy loss rate, cooling function) of a e+ propa- gating in a mediumwith density nX . For simplicity, we only consider Coulomb losses which are expected to dominate up to 100MeV for MW-like conditions (Jean et al. 2009), dE(E,nX )dx Coulomb ∝ nX , so that nX cancels. The total IA spectrum for mono-energetic injection of e+s thus reads( dFγ dEγ )mono IA = F511 1 1 − 34 fPs 1 P ( dNγ dEγ ) IA , (15) with( dNγ dEγ )mono IA = nX 2me ∫ Ekin E1 dE P(Ekin, E) dσ(E,Eγ)dE dE(E,nX )dx . (16) where the integration limit E1 = max ( (Eγ−me )2+E2γ (Eγ−me )+Eγ me, Eγ ) is MNRAS 000, 1–10 (2021) MeV Dark Matter Limits in Reticulum II 5 bound to obey energy conservation (Svensson 1982; Beacom&Yük- sel 2006). We note that ∫ dEγ ( dNγ dEγ ) IA = 1 − P. Since in the case of PBHs, the injection energy is not mono-energetic, we calculate a weighted photon spectrum given the distribution of kinetic energies from BlackHawk as( dNγ dEγ ) IA = ∫ Emaxkin 0 dEkin dNe dEe ( dNγ (Ekin) dEγ )mono IA∫ Emaxkin 0 dEkin dNe dEe . (17) Radiative corrections in the production of e+e−-pairs can be de- scribed as internal bremsstrahlung (IB)which can be calculatedwith- out detailed knowledge of the previous process. This final state radia- tion (FSR) arises as a result of the production itself and is independent of the astrophysical environment. The photon source spectrum of IB per 511 keV line photon is given by Beacom et al. (2005),( dFγ dEγ )mono IB = 1 2 1 1 − 34 fPs ( dNγ dEγ )mono IB , (18) with( dNγ dEγ )mono IB = α 2pi [ M2 + (M − 2Eγ)2 M2Eγ ln ( M(M − 2Eγ) m2e )] , (19) where here, M = Ekin + me is the particles’ injection energy from PBH evaporation. Because the e+-injection spectrum is again not mono-energetic, we weight over expected energy distribution such that( dNγ dEγ ) IB = ∫ Emaxkin 0 dEkin dNe dEe ( dNγ (Ekin) dEγ )mono IB∫ Emaxkin 0 dEkin dNe dEe . (20) Finally, the total photon source spectrum of PBH evaporation in- cluding subsequent e+-annihilation and IB in a galaxy is the sum the components:( dFγ dEγ ) tot PBH = ( dFγ dEγ ) PBH + ( dFγ dEγ ) IB + ( dFγ dEγ ) Ann = = ( dFγ dEγ ) PBH + ( dFγ dEγ ) IB + ∑ i∈A ( dFγ dEγ ) i , (21) with A = {511, oPs, IA}. ( dFγ dEγ ) PBH is a function of MBH, D and fPBH, and ( dFγ dEγ ) Ann is a function of fPs and Emaxkin . We show two examples of the total emission spectrum from PBH evaporation in Ret II in Fig. 1. 3.1.2 Particle Dark Matter Annihilation and Decay We consider light (mDM . 300MeV) DM particles that either anni- hilate or decay directly into standard model particles, leading to an expected photon emission spectrum. Therefore we evaluate the cases (i) DM + DM→ e+ + e− + (γ)FSR, (ii) DM + DM→ γ + γ, (iii) DM→ e+ + e− + (γ)FSR, (iv) (a) DM→ γ + γ, and (b) DM→ Y + γ, whereY is any relativistic particle, for example a sterile neutrino that decays into γ + ν. The photon spectra of the FSR in cases 1. and 3. are identical to the IB spectrum,( dNγ dEγ ) FSR ≡ ( dNγ dEγ ) IB , (22) 101 102 103 104 E [keV] 10 7 10 6 10 5 10 4 10 3 10 2 10 1 100 E × F [k eV ph cm 2 s 1 ke V 1 ] DM Annihilation (FSR) Positronium Decay Annihilation in Flight Total Spectrum (MDM = 3.8 MeV, v = 3 × 10 27 cm3 s 1 J = 2.0 × 1019 GeV2 cm 5, fPs = 1.0, Ekin, maxe = 10.0 MeV) 102 104 Ekine [keV] 10 2 10 1 100 Ek in e × F e 101 102 103 104 E [keV] 10 7 10 6 10 5 10 4 10 3 10 2 10 1 100 E × F [k eV ph cm 2 s 1 ke V 1 ] DM Annihilation (FSR, e + e ) DM Annihilation (FSR, + ) DM Annihilation ( ) ×10 5 Positronium Decay Annihilation in Flight Total Spectrum (MDM = 10.3 MeV, v = 3 × 10 24 cm3 s 1 J = 2.0 × 1019 GeV2 cm 5, fPs = N/A, Ekin, maxe = 1 MeV) 102 104 Ekine [keV] 10 1 101 Ek in e × F e Figure 2. Expected γ-ray spectra from light dark matter annihilation and subsequent annihilation of positron-electron pairs, similar to Fig. 1. The two panels show different model parameters (see legends) for a J-factor of 2 × 1019 GeV2 cm−5. The lower panel also shows model expectations from final state radiation of the µ+µ−-channel and direct decay into two photons. No Ps decay or in flight annihilation would be expected in the latter case. however now,M = 2mDM for annihilation andM = 1mDM for decay. Another factor of 12 is required in Eq. (19) for Dirac DM annihila- tion/decay to account for distinct particle-antiparticle DMwith equal densities. The expected photon spectrum of direct annihilation/decay into two photons (cases 2. and 4.) is a delta-function,( dNγ dEγ ) γγ = 2δ ( Eγ − M2 ) . (23) Case 4. (b) differs from case 4. (a) only by an additional factor of 2. We model Eq. (23) as Gaussians with the energy resolution of SPI. Similar to the PBH case, the emitted e+s from DM particles in a galaxy’s halo might annihilate during (IA) and after (Ps) propagation in the galaxy, or escape into the intergalacticmedium if their injection energy is larger than a threshold Emaxkin . The only difference is now that the energy distribution of produced pairs is mono-energetic, so that either all or none of the pairs lead to additional e+-annihilation related spectra. Given the equivalence in Eq. (22), one can estimate the total 511 keV line flux from DM particle annihilation as FAnn511 = ( 1 − 3 4 fPs ) J〈σv〉 2pim2DM , (24) and from DM particle decay as FDec511 = ( 1 − 3 4 fPs ) D 2pimDMτ . (25) The remaining spectral components from ortho-Ps and annihilation MNRAS 000, 1–10 (2021) 6 Thomas Siegert et al. 45 60 75 90 105 120 135 90 75 60 45 30 15 LMC X-4 SMC X-1 ESO 33-2 EXO 0748-676 Reticulum II 25-68 keV 45 60 75 90 105 120 135 90 75 60 45 30 15 LMC X-4 SMC X-1 ESO 33-2 EXO 0748-676 Reticulum II 504-514 keV 45 60 75 90 105 120 135 90 75 60 45 30 15 LMC X-4 SMC X-1 ESO 33-2 EXO 0748-676 Reticulum II 940-2000 keV Figure 3. Count residuals projected back to the sky by applying the imaging response backwards. The radial and azimuthal coordinates are the Galactic latitude and longitude, respectively. Shown are the bands 25–68 keV, 508–514 keV, and 940–2000 keV as examples. Only LMC X-4 is significantly detected in the first energy band. in flight are the same as in Sec. 3.1.1, except that now the mono- energetic cases are considered. The total source photon spectrum is again the sum of the components, here FSR, 511 keV line, ortho-Ps, and in-flight annihilation. We show examples of the expected DM annihilation spectra in Fig. 2. 4 RESULTS We examine our model fits to measured data through a careful anal- ysis of the residuals in different data space dimensions. Herein, it proves valuable to apply a back-projection onto the sky of the count residuals. For cases of background model imperfections, we would discover irregular or large-scale patterns on the sky, whereas for im- perfections of the sky model (e.g., missing a point source) would show up as a single peak at the source’s location. In Fig. 3, we show the residual count map of an instrumental-background-only fit in three example energy bands. Only the high-mass X-ray binary LMCX-4 is detected with a significance of 21σ up to an energy of ∼ 200 keV. The other sources, and in particular Ret II, are not seen in any energy band. The count residuals show no strong patterns as expected from an otherwise empty field. In the 511 keV band, the residuals appear more structured, which is a result of the reduced count rate compared to other bands. For reference, the particularly bright spot between ESO33-2 and LMCX-4 at (l, b) = (−78◦,−37◦) has a significance of ∼ 2.5σ, and is thus considered a background fluctuation. Details about the 511 keV line from Ret II are given in Sec. 4.2. If not otherwise stated, upper bounds on fluxes are given as 99.85th percentile (3σ). 4.1 Total Spectrum We show the spectrum measured by SPI from the direction of Ret II, spatially modelled as a point source, Eq. (3), in Fig. 4. Because the source is not detected in any of the energy bins, we plot 2σ upper limits. For comparison, we show several models that are excluded, given these fluxes. Detailed exclusion plots are found in the following sections. From a spectral fit with a generic power-law, ∝ ( Eγ 1MeV )ν , we derive an upper bound on the total flux in the band 20–8000 keV of 10−8 erg cm−2 s−1. 4.2 511 keV Line Wefind a 511 keV line flux limit of 1.7×10−4 ph cm−2 s−1. This value is about one order of magnitude smaller than the 3σ signal reported 101 102 103 104 Energy [keV] 10 2 10 1 100 101 102 103 E2 × F E [k eV 2 cm 2 s 1 ke V 1 ] Powerlaw ( = 0.65) DM Annihilation (FSR; MDM = 4 MeV, v = 4 × 10 25 cm3 s 1) PBH (primaries, MBH = 2 × 1015 g) PBH (secondaries) NS merger (×106; @10 kyr) SPI ( < 2 ) Figure 4. SPI spectrum from the position of Reticulum II. Shown are 2σ upper limits on the flux as a function of photon energy. A selection of excluded models is shown (see Secs. 4.3 and 4.4). by Siegert et al. (2016c) of (1.7 ± 0.5) × 10−3 ph cm−2 s−1. Given the enhanced exposure time and the source being inside the FCFOV all the time thanks to the new dedicated observation campaign, the improvement in sensitivity is plausible. The previous detection of Ret II might be due to an intrinsic vari- ability in time, as could be expected from the outburst of a micro- quasar (Siegert et al. 2016a) or a NS merger (Fuller et al. 2018), for example. We therefore split the data set into different time bins according to chance observations of nearby targets as well as the new observation campaign. The resulting 511 keV light curve is shown in Fig. 5. Considering only observations in which Ret II is inside the FCFOV results in upper limits consistent with the previous measure- ment by Siegert et al. (2016c). We find that the significance of the 511 keV line rises up to ∼ 2σ until the end of the older data set. Including more data after 2013 and in particular the additional 1Ms in 2018 (IJD ∼ 6700), leads to a greatly reduced upper flux bound as well as a lower significance. Assuming the source to be variable in time, the upper limit on the 511 keV line is ∼ 6.1×10−4 ph cm−2 s−1 – barely consistent with the previous estimate. Another factor here is the analysed spatial region: While Siegert et al. (2016c) included the entire sky with the expected diffuse 511 keV emission (four compo- nents) plus 39 dwarf galaxies, in this work we narrowed the region of interest and number of components included (Ret II plus 4 nearby MNRAS 000, 1–10 (2021) MeV Dark Matter Limits in Reticulum II 7 2000 3000 4000 5000 6000 7000 IJD [MJD-51544] 10 4 10 3 F 5 11 [p hc m 2 s 1 ] new observation campaign previous data set Siegert et al. (2016c) this work (variable source) this work (constant source) SPI data (3 upper limits) Figure 5. Lightcurve of the 511 keV line from Ret II. No positive excess is found at any time. Therefore, shown are 3σ upper limits in the band 508– 514 keV. The inferred flux from Siegert et al. (2016c) is shown in comparison to the new INTEGRAL observation campaign. Additional serendipitous ob- servations between IJD 4800 and 6600 are also included. sources). This reduces the covariances between the fitted components which can result into a slightly changed significance. We conclude that the 3σ signal fromRet II by Siegert et al. (2016c) wasmost likely due to an instrumental background fluctuation, paired with short exposures outside the FCFOV. We can not, however, ex- clude that the previously reported signal was of astrophysical origin. 4.3 Primordial Black Holes We perform fits to the extracted spectrum, Fig. 4, with Eq. (21), and sub-cases thereof. Here we emphasise that we do not solve for the interesting parameters until a certain figure of merit is above a thresh- old, but determine the joint posterior distributions of all the free pa- rameters in Eq. (21). That means we include the uncertainties on D and estimate fPBH conditioned on all the other unknown parameters, and illustrate the limits on fPBH as a function ofMBH, as it is typically done.However herewe consider the 95th percentile in fPBH-direction of the marginalised posterior pi( fPBH,MBH |yi, σi) with yi and σi as the flux values and corresponding uncertainties of energy bin i in the spectrum. See AppendixA for details. In Fig. 6, we show our upper bounds on fPBH, and compare them to estimates from the recent literature, including the MW (Laha 2019; Laha et al. 2020), cosmic rays (Boudaud & Cirelli 2019), the cosmic microwave background (CMB, Clark et al. 2017), and the cosmic γ-ray background (CGB Iguaz et al. 2021). The primary component of PBH evaporation alone cannot con- strain fPBH in Ret II. Including the secondary particles, the upper bound on fPBH = 1 excludes monochromatic PBH distributions with masses of . 0.4 × 1016 g. Assuming that all e+s created by PBHs in Ret II annihilate strengthens the bound on fPBH, then excluding masses . 2.9 × 1016 g. Since this assumption is too optimistic, we included the limiting factor for e+ annihilation, that all e+s above the threshold Emaxkin escape from the galaxy and do not contribute to the annihilation spectrum. This describes our most realistic bound on fPBH = 1, which excludes masses . 0.8 × 1016 g. Laha (2019) performed a similar estimate for the flux of the 511 keV line in the MW, however assumed that the entire annihi- lation flux, from integrating over the e+-distribution of PBHs, ends up in the 511 keV line. Since the Ps fraction in the MW is within a range 0.97–1.00, the 511 keV line flux from PBHs is at least four times smaller, because most of the annihilation flux goes into the three-photon decay of ortho-Ps. In return this means that the bounds on fPBH from Laha (2019) should be shifted by about a factor of four 1015 1016 1017 MBH [g] 10 1 100 f P BH (9 5t h pe rc en til e) PBH (prim.) PBH (sec.) PBH (sec.) + Ann. PBH (sec.) + lim. Ann. + AiF + IB Laha+2020; MW; SPI Boudaud+2019; MW; Voyager 1 Iguaz+2021; CGB Clark+2017; CMB Figure 6. Fraction of DM that can be constituted of a monochromatic PBH distribution with mass MBH in Ret II (solid lines), compared to literature estimates (dashed). The circles mark the masses when the 95th percentile of the posterior pi( fPBH, MBH |yi, σi ) in each case is reached. Stars mark the literature limits of 100% PBH DM. The most realistic bound is found for the case of PBH evaporation, including secondary particles, limited Ps annihilation, annihilation in flight, and internal bremsstrahlung (purple). ‘vertically’,which decreases themass at fPBH = 1 from∼ 8.9×1016 g to ∼ 6.4 × 1016 g. We test this by making the same assumptions with our Ret II spectrum, and find fPBH = 1 is excluded for masses MBH . 4.0 × 1016 g. Our conservative and realistic bounds are about one order of mag- nitude in PBHmass below previous estimates. Nevertheless, our data set is considerably shorter than all the other data sets, often compris- ing more than a decade of observation time. Thanks to the much smaller region of interest in the case of Ret II, basically constituting a point source of SPI, our systematic uncertainties can be consid- ered smaller compared to studies of the MW. The entire Galactic (DM) profile is rather uncertain, and also the Galactic centre itself bears problems when DM models are applied, as is often the case for diffuse emission in the GeV band. While our bounds are not yet as constraining, we show that reasonable progress can be made with even a small data set. In particular, we show that when all uncertain- ties on the spectral parameters are included also more conservative and realistic bounds can result. 4.4 Particle Dark Matter Similar to the PBH case, we perform a fit to the extracted Ret II spec- trum and marginalise over the uncertain parameters to determine the posteriors pi(〈σv〉,mDM |yi, σi) and pi(τ,mDM |yi, σi) for DM annihi- lation and decay, respectively. In what follows, the excluded regions consider the 95th percentile of the posteriors in 〈σv〉- and τ-direction, respectively. 4.4.1 Annihilation In Fig. 7, we show the excluded region of the velocity-averaged DM annihilation cross section into e+e−, followed by FSR and subsequent e+ annihilation. Without additional e+ annihilation, the bounds from Ret II are barely comparable to literature values from other γ-ray experiments (Essig et al. 2013) or the CMB (Slatyer 2016). When limited e+ annihilation is included, we can improve the limits in the MNRAS 000, 1–10 (2021) 8 Thomas Siegert et al. 100 101 102 mDM [MeV] 10 31 10 29 10 27 10 25 10 23 10 21 10 19 v [c m 3 s 1 ] (9 5t h pe rc en til e) + (FSR, this work) e + e (FSR, this work) e + e (FSR + Ps, this work) e + e (FSR, Essig+2013; MW; SPI) e + e (FSR, Essig+2013; MW; COMPTEL) e + e (CMB, Slatyer+2016; Planck) Figure 7. DM annihilation cross section into e+e− with subsequent FSR and possible annihilation of the pairs in Ret II (case 1. in Sec. 3.1.2; gray lines), compared to literature limits. The shaded regions are excluded. 10 2 10 1 100 101 102 mDM [MeV] 10 32 10 30 10 28 10 26 10 24 10 22 10 20 v [c m 3 s 1 ] (9 5t h pe rc en til e) (this work) (Laha et al. 2020; SPI, MW) (Ng et al. 2019; NuSTAR, M31) (Slatyer et al. 2016; Planck, CMB) Figure 8. Similar to Fig. 7 but for DM + DM→ γ + γ. DM mass range 0.5–1.0MeV by up to two orders of magnitude. We find a general trend of the annihilation cross section at tree level of 〈σv〉 . 5×10−28 (mDM/MeV)2.5 cm3 s−1, based on the shape of the measured limits as a function of energy in Fig. 7. For completeness, we show the limits on DM annihilation into µ+µ−, being hardly constrained by Ret II data. The cross section limit for the case DM + DM→ γ + γ is shown in Fig. 8. Our limits on DM annihilation into two photons are compa- rable to those from NuSTAR observations of M31 (Ng et al. 2019), and about one order of magnitude worse than fromMWobservations with SPI (Laha et al. 2020). Because our data set extends to the full energy range of SPI, we can set however limits out tomDM ≤ 8MeV, varying between 10−28 and 10−25 cm3 s−1. However, limits from the CMB (Slatyer 2016) are still more stringent at there energies. 4.4.2 Decay The properties of decaying DM particles cannot be inferred directly from the previous case of annihilating DM. First, the kinematic thresholds for standard model particle production is shifted by a factor of two which leads to altered (shifted) spectral shapes in par- 100 101 102 mDM [MeV] 1020 1021 1022 1023 1024 1025 1026 1027 [s ] ( 95 th p er ce nt ile ) e + e (FSR, this work) e + e (FSR + Ps, this work) e + e (FSR, Essig+2013; MW; SPI) e + e (FSR, Essig+2013; MW; COMPTEL) e + e (CMB, Slatyer+2017; Planck) Figure 9. Similar to Fig. 7 but for DM→ e+ + e− + (γ)FSR. 10 2 10 1 100 101 102 mDM [MeV] 1017 1019 1021 1023 1025 1027 1029 [s ] ( 95 th p er ce nt ile ) (this work) (Laha+2020; MW; SPI) (Essig+2013; MW; HEAO) (Essig+2013; MW; SPI) (Essig+2013; MW; COMPTEL) (Ng+2019; M31; NuSTAR) (CMB, Slatyer+2017; Planck) Figure 10. Similar to Fig. 7 but forDM→ γ + γ. For the caseDM→ γ + Y, the limits are smaller by a factor of 2. ticular for FSR and IA. Second, the J-factor of Ret II is about twice as uncertain as its D-factor, which allows to populate different regions in the full posterior. We therefore perform spectral fits for the DM decay case separately to infer bounds on the DM lifetime τ. This also results in slightly worse bounds when considering decay compared to annihilation (cf. NuSTAR limits). In Fig. 9 we show the exclusion region of τ as a function of the DM mass. Similar to the DM annihilation case, FSR alone is barely constraining, but can be pushed by including e+ annihilation. We find that DM particles decaying into e+e− pairs must have a lifetime longer than 1–5× 1024 s between 1 and 2MeV. In the mass range 1– 2MeV, our bounds are as strong as the limits from the CMB Slatyer & Wu (2017). Above this mass range, previous limits from SPI and COMPTEL in the MW (Essig et al. 2013), and in particular those from the CMB are more constraining. Finally, our DM decay time limits considering the direct decay into two photons is not improving beyond previous constraints. This is expected from the pure narrow line sensitivity of SPI. MNRAS 000, 1–10 (2021) MeV Dark Matter Limits in Reticulum II 9 5 SUMMARY AND CONCLUSION Ret II is not seen in soft γ-rays between 0.025 and 8MeV, up to an upper flux limit of 10−8 erg cm−2 s−1 (5 × 10−10 erg cm−2 s−1 within 20–80 keV). The previously reported 511 keV line signal in Ret II has not been consolidated in this work, and we provide an upper limit on its flux of 1.7 × 10−4 ph cm−2 s−1. Given the distance to the galaxy, this constrains the e+-annihilation rate (including Ps formation and annihilation in flight) to 0.9–3.7×1043 e+ s−1, at most the annihilation rate of the MW1. We calculated parametrised spectral models for PBH evaporation and DM annihilation/decay with subsequent e+ annihilation and fit- ted these models to the total spectrum from Ret II. Our conservative limit on the monochromatic PBH mass distribution constituting the entirety of DM in Ret II is 0.8 × 1016 g. We improve the limits on the velocity-averaged DM annihilation cross section of decay time of DM particles into e+e− in the range 0.5–2MeV. For annihilation or decay into two photons, our limits are weaker than previous es- timates. We provide the currently only limits on the cross section 〈σv〉γγ in the range 3–8MeV based on γ-ray observations, ranging between 10−28 and 10−25 cm3 s−1. While our estimates are at most on the same level as previous constraints, we arrive at our conclusions with a very small data set (two weeks) compared to the multi-year or even decade-long obser- vations. With our detailed spectral modelling including the effects of possible e+ annihilation and the proper statistical treatment of known unknowns, we show that dedicated observations of selected targets provide valuable information about the DM conundrum. We encour- age follow-up observations of Ret II and other dwarf galaxies with INTEGRAL, and foresee a bright future for soft γ-ray instruments to come, such as COSI (Tomsick et al. 2019), to contribute significantly in the research for the nature of DM. With its more than one order of magnitude increased sensitivity, in the 0.5–5MeV band COSI can reach and surpass the strongest CMB limits within its nominal two- year mission from observations of dwarf galaxies only. Thanks to its better spatial resolution as well as its more uniform coverage of the MeV sky, disentangling fore- and background components of the Milky Way itself will result in a clearer separation of putative DM signals from the still weakly constrained MeV spectrum. SOFTWARE OSA/spimodfit (Halloin 2009), BlackHawk (Arbey & Auffinger 2019), numpy (Oliphant 2006), matplotlib (Hunter 2007), astropy (Collaboration et al. 2013), scipy (Virtanen et al. 2019), 3ML (Vianello et al. 2015),MultiNest (Feroz & Hobson 2008; Feroz et al. 2009, 2019). ACKNOWLEDGEMENTS Thomas Siegert is supported by the German Research Foundation (DFG-Forschungsstipendium SI 2502/3-1). We thank Ranjan Laha, Kenny Ng, Shunsaku Horiuchi, Marco Cirelli, Sam McDermott, Tracy Slatyer, and Joaquim Iguaz for providing limits on dark matter properties, and Oleg Korobkin for γ-ray spectra from neutron star 1 We note that in Siegert et al. (2016c), there is a typo, reducing the estimate for Ret II by a factor of ten – the actual annihilation rate in the previous work should be on the order of 1044 e+ s−1. mergers. We thank John Beacom and Hassan Yüksel for details on the in-flight annihilation spectrum. DATA AVAILABILITY The data underlying this article will be shared on reasonable request to the corresponding author. REFERENCES Albert A., et al., 2017, The Astrophysical Journal, 834, 110 Alves Batista R., et al., 2021, arXiv.org, p. arXiv:2110.10074 Arbey A., Auffinger J., 2019, The European Physical Journal C, 79, 693 Attié D., et al., 2003, Astronomy & Astrophysics, 411, L71 Bayes T., 1763, Proceedings of the Royal Society of London, Philosophical Transactions of the Royal Society, 53, 370 Beacom J. F., Yüksel H., 2006, Physical Review Letters, 97, 071102 Beacom J. F., Bell N. F., Bertone G., 2005, Physical Review Letters, 94, 171301 Bergström L., Ullio P., Buckley J. H., 1998, Astroparticle Physics, 9, 137 Bertone G., Hooper D., 2018, Reviews of Modern Physics, 90, 045002 Boehm C., Hooper D., Silk J., Cassé M., Paul J., 2004, Physical Review Letters, 92, 101301 Bonnivard V., et al., 2015, The Astrophysical Journal Letters, 808, L36 Bouchet L., Strong A.W., Porter T. A., Moskalenko I. V., Jourdain E., Roques J.-P., 2011, The Astrophysical Journal, 739, 29 Boudaud M., Cirelli M., 2019, Physical Review Letters, 122, 041104 Churazov E., Sunyaev R., Sazonov S., Revnivtsev M., Varshalovich D., 2005, Monthly Notices of the Royal Astronomical Society, 357, 1377 Churazov E., Sazonov S., Tsygankov S., Sunyaev R., Varshalovich D., 2011, Monthly Notices of the Royal Astronomical Society, 411, 1727 Clark S. J., Dutta B., Gao Y., Strigari L. E., Watson S., 2017, Physical Review D, 95, 083006 Collaboration A., et al., 2013, Astronomy & Astrophysics, 558, A33 Cordier B., et al., 2004, in Schoenfelder V., Lichti G., Winkler C., eds, 5th INTEGRAL Workshop on the INTEGRAL Universe. p. 581 Diehl R., et al., 2018, Astronomy & Astrophysics, 611, A12 Essig R., Kuflik E., McDermott S. D., Volansky T., Zurek K. M., 2013, Journal of High Energy Physics, 11, 193 Evans N. W., Sanders J. L., Geringer-Sameth A., 2016, arXiv.org, 93, 103512 Feroz F., Hobson M. P., 2008, Monthly Notices of the Royal Astronomical Society, 384, 449 Feroz F., Hobson M. P., Bridges M., 2009, Monthly Notices of the Royal Astronomical Society, 398, 1601 Feroz F., Hobson M. P., Cameron E., Pettitt A. N., 2019, The Open Journal of Astrophysics, 2, 10 Fortin J.-F., Shelton J., Thomas S., Zhao Y., 2009, arXiv.org, p. arXiv:0908.2258 Fuller G. M., Kusenko A., Radice D., Takhistov V., 2018, arXiv.org Geringer-Sameth A., Walker M. G., Koushiappas S. M., Koposov S. E., Belokurov V., Torrealba G., Evans N. W., 2015, Physical Review Letters, 115, 081101 Halloin H., 2009, |spimodfit| Explanatory Guide and Users Manual. Max Planck Institut für extraterrestrische Physik Max Planck Institut für ex- traterrestrische Physik, Giessenbachstraße 1, 85748 Garching, Germany, version 2.9 edn Hawking S., 1971, Monthly Notices of the Royal Astronomical Society, 152, 75 Hawking S. W., 1975, Communications In Mathematical Physics, 43, 199 Hunter J. D., 2007, Computing in Science & Engineering, 9, 90 Iguaz J., Serpico P. D., Siegert T., 2021, arXiv.org, p. arXiv:2104.03145 Jean P., Knoedlseder J., GillardW., Guessoum N., Ferrière K., Marcowith A., Lonjou V., Roques J. P., 2006, Astronomy & Astrophysics, 445, 579 Jean P., Gillard W., Marcowith A., Ferrière K., 2009, Astronomy & Astro- physics, 508, 1099 MNRAS 000, 1–10 (2021) 10 Thomas Siegert et al. Koposov S. E., Belokurov V., Torrealba G., Evans N. W., 2015, arXiv.org, 805, 130 Korobkin O., et al., 2020, The Astrophysical Journal, 889, 168 Laha R., 2019, Physical Review Letters, 123, 251101 Laha R., Muñoz J. B., Slatyer T. R., 2020, Physical Review D, 101, 123514 Leventhal M., MacCallum C. J., Stang P. D., 1978, Astrophysical Journal, 225, L11 Ng K. C. Y., Roach B. M., Perez K., Beacom J. F., Horiuchi S., Krivonos R., Wik D. R., 2019, Physical Review D, 99, 083005 Oliphant T. E., 2006, A guide to NumPy. Vol. 1, Trelgol Publishing USA Ore A., Powell J. L., 1949, Physical Review, 75, 1696 Page D. N., Hawking S. W., 1976, Astrophysical Journal, 206, 1 Siegert T., et al., 2016a, Nature, 531, 341 Siegert T., Diehl R., Khachatryan G., Krause M. G. H., Guglielmetti F., Greiner J., Strong A. W., Zhang X., 2016b, Astronomy & Astrophysics, 586, A84 Siegert T., Diehl R., Vincent A. C., Guglielmetti F., Krause M. G. H., Boehm C., 2016c, Astronomy & Astrophysics, 595, A25 Siegert T., Diehl R., Weinberger C., Pleintinger M. M. M., Greiner J., Zhang X., 2019, Astronomy & Astrophysics, 626, A73 Simon J. D., 2019, Annual Review of Astronomy and Astrophysics, 57, 375 Sizun P., Cassé M., Schanne S., 2006, Phys. Rev. D, 74, 063514 Slatyer T. R., 2016, Physical Review D, 93, 023527 Slatyer T. R., Wu C.-L., 2017, Physical Review D, 95, 023010 Svensson R., 1982, ] 10.1086/160081, 258, 321 Tomsick J. A., et al., 2019, arXiv.org, p. arXiv:1908.04334 Vedrenne G., et al., 2003, Astronomy & Astrophysics, 411, L63 Vianello G., et al., 2015, arXiv.org, p. arXiv:1507.08343 Virtanen P., et al., 2019, arXiv.org, p. arXiv:1907.10121 Winkler C., et al., 2003, Astronomy & Astrophysics, 411, L1 Winter M., Zaharijaš G., Bechtol K., Vandenbroucke J., 2016, The Astro- physical Journal Letters, 832, L6 APPENDIX A: DETAILS ON SPECTRAL FITS The error bars in the extracted spectrum in Sec. 4.1, Fig. 4 approxi- mately follow a normal distribution. Therefore, we use the likelihood Lnormal(D |M(ψ)) = 30∏ i=1 1√ 2piσi exp ( −1 2 [ yi − mi(ψ) σi ]2) , (A1) with yi as themeasured flux in energy bin i = 1 . . . 30 in the spectrum, σi the corresponding uncertainties, and mi(ψ) the forward-folded model, Eq. (4), that depends on a set of spectral parameters ψ. Since we want to include the astrophysical and particle physics uncertainties of our source, we construct the joint posterior of all N parameters ψ1 . . . ψN through Bayes’ theorem (Bayes 1763) as pi(ψ1 . . . ψN |yi, σi) ∝ Lnormal(yi, σi |ψ1 . . . ψN )pi(ψ1 . . . ψN ). (A2) Here, pi(ψ1 . . . ψN ) is the joint prior distribution of the individ- ual parameters. We use independent priors so that pi(ψ1 . . . ψN ) = pi(ψ1) · · · pi(ψN ), that is, each parameter obtains its individual prior probability distribution function. By integrating out the parameters ψ3 to ψN , for example, we can construct the marginalised joint pos- terior distribution of ψ1 and ψ2, pi(ψ1, ψ2 |yi, σi) = ∫ · · · ∫ dψ2 · · ·ψN pi(ψ1 . . . ψN |yi, σi). (A3) To obtain the upper bound on ψ1 as a function of ψ2, we integrate the marginalised posterior pi(ψ1, ψ2 |yi, σi) in ψ1-direction and equate to 103 104 105 106 mDM [keV c 2] 10 29 10 27 10 25 10 23 10 21 v ee [c m 3 s 1 ] v ub(mDM) MultiNest Samples Figure A1. Marginalised posterior distribution pi(〈σv〉,mDM |yi, σi ) (red dots), together with upper bounds of the annihilation cross section as a func- tion of the DM mass (black line). a certain percentile P, P = ∫ ψub1 (ψ2) ψmin1 dψ1pi(ψ1, ψ2 |yi, σi). (A4) Solving Eq. (A4) for ψub1 (ψ2) provides the upper bound on ψ1 as function of ψ2. A written example considering the case of annihilating DM into e+e− with subsequent FSR and e+ annihilation would calculate the upper bound on the velocity-averaged annihilation cross sec- tion 〈σv〉 as a function of the DM mass mDM. Thus first, the join posterior pi(〈σv〉,mDM, J, fPs, Emaxkin |yi, σi) is estimated, then marginalised over the J-factor, the Ps fraction and the kinetic energy threshold to avoid escape, pi(〈σv〉,mDM |yi, σi) =∫ Jmax Jmin dJ ∫ fPs,max fPs,min dfPs ∫ Emaxkin,max Emaxkin,min dEmaxkin pi(〈σv〉,mDM, J, fPs, Emaxkin |yi, σi), (A5) and finally integrated up to the bound 〈σv〉ub(mDM) by P = ∫ 〈σv〉ub(mDM) 0 d〈σv〉pi(〈σv〉,mDM |yi, σi). (A6) We use MultiNest (Feroz & Hobson 2008; Feroz et al. 2009, 2019) in 3ML to evaluate the joint posterior distributions, marginalisations, and upper bounds. We show the samples of this example in Fig. A1. We list our prior distributions and parameter ranges for all considered models in Tab. A1. This paper has been typeset from a TEX/LATEX file prepared by the author. MNRAS 000, 1–10 (2021) MeV Dark Matter Limits in Reticulum II 11 Model κ M D J fPs Emaxkin PBH (prim.) U (0, 1) logU (1014, 1018) U (1018, 4 × 1018) – – – PBH (sec.) . . . . . . . . . – – – PBH (sec.) + Ann. . . . . . . . . . – U (0, 1) – PBH (sec.) + lim. Ann. . . . . . . . . . – . . . logU (103, 105) 2DM→ e+e−γ logU (3 × 10−34, 3 × 10−20) logU (511, 106) – U (2 × 1018, 4 × 1019) U (0, 1) logU (103, 105) 2DM→ 2γ . . . logU (25, 8000) – . . . – – DM→ e+e−γ logU (1015, 1024) logU (1022, 106) U (1018, 4 × 1018) – U (0, 1) logU (103, 105) DM→ 2γ logU (1015, 1029) logU (50, 16000) . . . – – – Table A1. Prior distributions for considered DM models in Eq. (5). MNRAS 000, 1–10 (2021)