Issue |
A&A
Volume 680, December 2023
|
|
---|---|---|
Article Number | A45 | |
Number of page(s) | 29 | |
Section | Astrophysical processes | |
DOI | https://doi.org/10.1051/0004-6361/202347298 | |
Published online | 12 December 2023 |
The short gamma-ray burst population in a quasi-universal jet scenario⋆
1
INAF – Osservatorio Astronomico di Brera, Via Emilio Bianchi 46, 23807 Merate (LC), Italy
e-mail: om.salafia@inaf.it
2
INFN – sezione di Milano-Bicocca, Piazza della Scienza 3, 20126 Milano (MI), Italy
3
Department of Astrophysics/IMAPP, Radboud University, 6525 AJ Nijmegen, The Netherlands
4
School of Physics and Astronomy, Monash University, Clayton, Victoria 3800, Australia
5
ARC Center of Excellence for Gravitational Wave Discovery – OzGrav, Hawthorn, Australia
Received:
27
June
2023
Accepted:
7
September
2023
We present a model of the short gamma-ray burst (SGRB) population under a ‘quasi-universal jet’ scenario in which jets can differ somewhat in their on-axis peak prompt emission luminosity, Lc, but share a universal angular luminosity profile, ℓ(θv) = L(θv)/Lc, as a function of the viewing angle, θv. The model was fitted, through a Bayesian hierarchical approach inspired by gravitational wave (GW) population analyses, to three observed SGRB samples simultaneously: the Fermi/GBM sample of SGRBs with spectral information available in the catalogue (367 events); a flux-complete sample of 16 Swift/BAT SGRBs that are also detected by the GBM and have a measured redshift; and a sample of SGRBs with a binary neutron star (BNS) merger counterpart, which only includes GRB 170817A at present. Particular care was put into modelling selection effects. The resulting model, which reproduces the observations, favours a narrow jet ‘core’ with half-opening angle θc = 2.1−1.4+2.4 deg (uncertainties hereon refer to 90% credible intervals from our fiducial ‘full sample’ analysis) whose peak luminosity, as seen on-axis, is distributed as a power law, p(Lc) ∝ Lc−A with A = 3.2−0.4+0.7, above a minimum isotropic-equivalent luminosity, Lc⋆ = 5−2+11 × 1051 erg s−1. For viewing angles larger than θc, the luminosity profile scales as a single power law, l ∝ θv−αL with αL = 4.7−1.4+1.2, with no evidence of a break, despite the model allowing for it. While the model implies an intrinsic ‘Yonetoku’ correlation between L and the peak photon energy, Ep, of the spectral energy distribution, its slope is somewhat shallower, Ep ∝ L0.4 ± 0.2, than the apparent one, and the normalisation is offset towards larger Ep due to selection effects. The implied local rate density of SGRBs (regardless of the viewing angle) is between about one hundred up to several thousand events per cubic gigaparsec per year, in line with the BNS merger rate density inferred from GW observations. Based on the model, we predict 0.2 to 1.3 joint GW+SGRB detections per year by the advanced GW detector network and Fermi/GBM during the O4 observing run.
Key words: relativistic processes / gamma-ray burst: general / methods: statistical
The source code and posterior samples behind this work are publicly available at https://github.com/omsharansalafia/grbpop.
© The Authors 2023
Open Access article, published by EDP Sciences, under the terms of the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
This article is published in open access under the Subscribe to Open model. Subscribe to A&A to support open access publication.
1. Introduction
Two main clusters of gamma-ray bursts (GRBs) have long been identified in the two-dimensional plane of duration versus hardness ratio1 of the large sample collected by the Burst Alert and Transient Source Experiment (BATSE) on board the Compton Gamma-Ray Observatory (Kouveliotou et al. 1993). The bimodality in the BATSE GRBs is also apparent when considering the durations only (more precisely, the time, T90, over which 5% to 95% of the background-subtracted counts are collected), with the histogram of the logarithms of the durations featuring two peaks separated by a valley at around T90 = 2 s. This has since become the customary separation between the ‘long’ (LGRB) and the ‘short’ (SGRB) events in the GRB class, although the actual position of the valley varies somewhat across different detectors (e.g. Bromberg et al. 2013). The evidence accumulated during the following decades painted a picture of two different progenitor systems: starting with GRB 980425 (Galama et al. 1998), a number of LGRBs have been firmly associated with type Ib/c core-collapse supernovae (e.g. Bloom et al. 2002; Malesani et al. 2004; Mirabal et al. 2006; Kann et al. 2011; Cano et al. 2014; D’Elia et al. 2018; Hu et al. 2021), securing the scenario of a massive star progenitor (Woosley 1993); the progenitors of SGRBs remained elusive for a longer time, though all hints consistently pointed (e.g. Nakar 2007; Fong & Berger 2013; Berger 2014; D’Avanzo 2015) to a compact binary merger progenitor (Eichler et al. 1989; Mochkovitch et al. 1993). Such a progenitor has been confirmed by the astounding association (Abbott et al. 2017a,b) of the SGRB 170817A (Goldstein et al. 2017) – detected by the Gamma-ray Burst Monitor (GBM; Meegan et al. 2009) on board the Fermi spacecraft and by the International Gamma-Ray Astrophysics Laboratory (INTEGRAL; Savchenko et al. 2017) – with the first-ever binary neutron star (BNS) merger detected by humankind, GW170817 (Abbott et al. 2019), whose gravitational wave (GW) signal was captured on 17 August 2017 by the Advanced Laser Interferometer Gravitational wave Observatory (aLIGO; Aasi et al. 2015) and localised thanks to Advanced Virgo (Acernese et al. 2014).
The properties of SGRBs – especially their shorter duration and harder spectrum with respect to LGRBs (Ghirlanda et al. 2009; Calderone et al. 2015) – make them very hard to detect with current facilities: of the more than 280 GRBs revealed by Fermi/GBM every year, only about 40 are SGRBs. The softer sensitivity band of the Burst Alert Telescope (BAT) on board the Neil GehrelsSwift Observatory (Gehrels et al. 2004), together with its smaller field of view, allows it to identify and localise only a handful of SGRBs per year. Moreover, even when localised by Swift, the fraction of SGRBs that end up with a secure redshift determination is relatively low, due to a combination of a fainter X-ray ‘afterglow’ (Costa et al. 1997) – whose detection is a requirement for a precise localisation – and a larger typical offset from the host galaxy (Fong et al. 2013, 2022; Fong & Berger 2013; D’Avanzo et al. 2014; Berger 2014) with respect to LGRBs, which renders the host galaxy identification ambiguous in cases where multiple galaxies stand at similar offsets from the position of the afterglow. As a result, the intrinsic properties of the SGRB population are much more uncertain than for LGRBs. Indeed, a variety of attempts at constraining the properties of the SGRB population throughout the years, in some cases with very different methodologies and reference samples, has yielded varied results (e.g. Schmidt 2001; Guetta & Piran 2005, 2006; Virgili et al. 2011; Yonetoku et al. 2014; D’Avanzo et al. 2014; Wanderman & Piran 2015; Shahmoradi & Nemiroff 2015; Ghirlanda et al. 2016; Zhang & Wang 2018; Paul 2018; Tan & Yu 2020). Some of these results are in clear tension with each other: in this work, we consider Wanderman & Piran (2015; hereafter W15) and Ghirlanda et al. (2016; hereafter G16), which reach very different conclusions, as our benchmarks.
A prominent challenge in unveiling the intrinsic properties of SGRBs is the uncertainty about which processes shape the luminosity function, that is, the probability distribution from which the luminosity of each event in the population is sampled. Lipunov et al. (2001) and Rossi et al. (2002) were the first to realise that the highly relativistic nature of GRB jets would make their angular structure an important factor in determining the luminosity function, in addition to the intrinsic spread in luminosities. Indeed, if the energy density and the typical Lorentz factor of a GRB jet are functions of the angular separation from the jet axis, then the apparent energetics are viewing-angle-dependent by virtue of relativistic aberration effects (e.g. Salafia et al. 2015). Since jets are isotropically oriented in space, this naturally produces a large spread in the apparent luminosities, with a well-defined dependence on the angular structure (Pescalli et al. 2015). In the presence of a narrow distribution of intrinsic jet luminosities, the luminosity function is then mainly shaped by the angular structure (e.g. Salafia et al. 2020; Tan & Yu 2020).
An angular profile in the jet properties arises naturally in essentially any physically viable jet formation scenario (see Salafia & Ghirlanda 2022 for a recent review). Moreover, a non-trivial jet energy density angular profile is required (e.g. Mooley et al. 2018; Lamb et al. 2018, 2019; Ghirlanda et al. 2019; Takahashi & Ioka 2020, 2021; Beniamini et al. 2022) to explain the observed properties of the non-thermal afterglow of the SGRB associated with GW170817. Hence, the question of whether the observed distribution of SGRB properties can be traced back, at least in part, to the differing viewing angles is of particular relevance, as it would provide a route to a unification of these sources and a way to disentangle the intrinsic diversity in their properties (and hence those of their progenitor) from the apparent diversity due to extrinsic factors, particularly the viewing angle.
In the future, joint GW-GRB observations will provide direct information on the structure of SGRB jets thanks to the measurements of the inclination of the merging binary’s angular momentum (which is most likely a proxy of the jet viewing angle) that can be inferred from GW analysis (Williams et al. 2018; Hayes et al. 2020; Farah et al. 2020; Biscoveanu et al. 2020). Still, such information must be combined self-consistently with that encoded in the SGRB population observed in gamma-rays only. In this work, we describe our investigation of the population properties of SGRBs within a ‘quasi-universal jet’ scenario. We assumed that all SGRB jets share the same angular profile of luminosity as a function of the viewing angle, while we allowed for a spread in the on-axis luminosities (as in e.g. Tan & Yu 2020; Hayes et al. 2023). This produces a particular parametrisation of the SGRB population properties, which we derive and describe in detail in Sects. 2.2 and 2.3. In order to constrain the parameters of this model, we considered the sample of SGRBs detected by Fermi and carefully modelled the underlying selection effects (Sect. 2.4). This allowed us to fit the model to the data through a hierarchical Bayesian approach (Sect. 2.5). In Sect. 3 we describe in detail the results of the fit, and in Sect. 4 we discuss several implications.
Throughout this work, we assume a flat Friedmann-Lemaître-Robertson-Walker cosmology with Planck Collaboration XIII (2016) parameters, that is, H0 = 67.74 km Mpc−1 s−1 and Ωm, 0 = 0.3075.
2. Methodology
2.1. Apparent versus intrinsic structure
The dominant form of energy in GRB jets and the processes that dissipate such energy, leading to the observed ‘prompt’ gamma-ray emission are still a matter of debate (see Kumar & Zhang 2015, for a recent review). Still, regardless of the particular dissipation and emission process, the observed emission is affected by relativistic aberration effects in a way that depends on the Lorentz factor profile and the viewing angle (e.g. Woods & Loeb 1999; Salafia et al. 2015). For example, an observer looking, from a viewing angle θv (angle between the line of sight and the jet axis), at an axisymmetric jet with a bulk Lorentz factor profile Γ(θ) (where θ is the angle from the jet axis) that radiates an energy per unit solid angle dE/dΩ(θ), would measure an isotropic-equivalent gamma-ray energy (Salafia et al. 2015),
where ϕ is the azimuthal angle of a spherical coordinate system whose z-axis coincides with the jet axis, is the Doppler factor, β(θ, ϕ) is the local jet bulk velocity vector – with a magnitude
– and eLoS is a unit vector pointing to the observer. Hence, when considering the emitted energy in gamma-rays, the ‘apparent structure’ Eiso(θv) depends on the intrinsic structure (Γ(θ),dE/dΩ(θ)) through a functional that is not invertible in general. The situation is even more nuanced when considering the luminosity, as the effective angular profile Liso(θv) depends also on the degree of overlap between different pulses (Salafia et al. 2016), and hence on the intrinsic variability.
For these reasons, in the absence of strong theoretical constraints on the intrinsic jet structure and on the dissipation and emission processes, the most straightforward approach – which we adopt in this work – is that of parameterising directly the apparent luminosity structure L(θv) (we drop the ‘iso’ suffix from here on for simplicity), which also reduces the number of parameters. An assessment of the intrinsic jet structures that are compatible with a given apparent luminosity profile can be then carried out a posteriori.
2.2. Statistical model of an SGRB population with a quasi-universal jet
Within our framework, we describe each SGRB by four physical quantities, namely its viewing angle θv, its peak isotropic-equivalent luminosity L, the photon energy Ep at the peak of the spectral energy distribution (SED; i.e. the νFν spectrum) and its redshift z. These collectively represent what we refer to as the source parameter vector, . For most events, the viewing angle is unknown and we therefore consider a reduced source parameter vector λsrc = (L, Ep, z). We assumed the diversity in these parameters to be the combined result of intrinsic heterogeneity in the physical properties of jets within the population and extrinsic diversity induced by the differing viewing angles under which these jets are observed.
In order to represent the intrinsic heterogeneity in SGRB jets, we opted for parameterising the joint probability density P(Lc, Ep, c) of their on-axis (‘core’) peak luminosity Lc and peak SED photon energy Ep, c as follows. We assumed Lc to be distributed as a power law with index −A, with a lower exponential cutoff below , namely
where indicates a Gamma function, and the integrated probability is normalised to unity. This probability density is defined in such a way that it peaks at
regardless of the value of A. This choice of parametrisation follows from the fact that, in the quasi-universal jet scenario, the high end of the luminosity function reflects the luminosity distribution of jets seen close to on-axis (see Sect. 2.3), and the high-luminosity end has been found to be well described by a power law in most previous studies of the SGRB population (e.g. Schmidt 2001; Guetta & Piran 2006; Virgili et al. 2011; Yonetoku et al. 2014; D’Avanzo et al. 2014; Wanderman & Piran 2015; Ghirlanda et al. 2016; Zhang & Wang 2018; Paul 2018; Tan & Yu 2020). The reason for introducing a low-end cut-off follows from the quasi-universality assumption (in which case low luminosities are due to off-axis viewing angles rather than to intrinsically weak jets) and can be physically linked to a minimum jet luminosity required for the jet to break out from the progenitor vestige (see Salafia & Ghirlanda 2022 for a recent review).
The probability distribution on Ep, c, conditional on Lc, was assumed log-normal and centred at a Lc-dependent value , where y sets the slope of the relation. Hence,
where σc sets the dispersion of Ep, c around . This assumption allows for a ‘Yonetoku’ correlation2 with slope y between the logarithms of the on-axis peak SED photon energy and the luminosity, which may be induced, for example, by the underlying emission process. The log-normal form of the scatter around the relation was chosen for its simplicity. The case with no correlation (hence with log-normally distributed values of Ep, c, un-correlated with Lc) is represented by y = 0 and it is therefore naturally included. The joint probability density of the core quantities is
where the population parameter vector λpop contains , A,
, y, σc and all other parameters that fully specify the SGRB population model.
The next, key assumption of the model is that all jets share a universal ‘structure’ that specifies the dependence of L and Ep on the viewing angle θv. In practice, we assumed the viewing-angle-dependent luminosity and SED peak photon energy to be expressed as
where ℓ and η are functions of the viewing angle and of some parameters included in the λpop vector. These functions, which we assumed to be redshift-independent, define the universal apparent structure of the jet.
For a population of isotropically oriented jets, whose viewing angle probability distribution is P(θv) = sin θv, we have
The induced joint luminosity and peak photon energy distribution, given the isotropic viewing angles, is then given by
In general, this must be solved numerically, but in the next section we analyse two cases where the intrinsic dispersion is negligible (i.e. σc → 0 and A → ∞) and ℓ and η take simple forms, so that an analytical integration is possible: this will help in demonstrating the main features of the P(L, Ep) distribution induced by such a quasi-universal structure scenario.
As a consequence of the assumption of redshift independence of the jet structure parameters, the probability distribution of the population source parameters is Ppop(L, Ep, z | λpop) = P(L, Ep | λpop)P(z | λpop), where the redshift probability distribution can be expressed as
Here dV/dz is the differential comoving volume and is the SGRB rate density at redshift z. We parametrise the latter as a smoothly broken power law, namely
where a, b, and zp are free parameters and R0 is the local rate density of SGRBs with any viewing angle. This parametrisation reflects the functional form of the fitting function to the cosmic star formation rate (CSFR) from Madau & Dickinson (2014). Allowing for variations in its parameters a, b, and zp covers a variety of possible evolutionary histories that rise, peak and decay with increasing (1 + z), in a way that is agnostic of the progenitor nature, but can still accommodate, for example, typical rate evolutions obtained by convolving the cosmic star formation history with a merger delay time distribution under the assumption of a compact binary merger progenitor (as done in e.g. D’Avanzo et al. 2014; Wanderman & Piran 2015).
2.3. Apparent jet structure models and the implied luminosity-peak energy distribution
2.3.1. Luminosity function
A simple and widely adopted parametric form for the jet structure is a Gaussian one, namely
In the absence of a dispersion in the core quantities, which formally corresponds to the limit σc → 0 and A → ∞, and in the case where ℓ and η are monotonic, Eq. (7) reduces to a change of variables from θv to either L or Ep applied to the viewing angle probability P(θv) = sin θv, that is (Pescalli et al. 2015; Salafia & Ghirlanda 2022),
where ℓ−1 is the inverse function of ℓ and η−1 is the inverse function of η. In what follows, we show results using the first of the above equalities, which highlights the dependence on L, but the results using the second equality are entirely analogous and can be obtained by exchanging L ↔ Ep, θc ↔ θc, Ep and Lc ↔ Ep, c. In the Gaussian apparent structure case this yields, for L < Lc and Ep < Ep, c,
where the last approximate equality is valid for θv ≪ π/2, which corresponds to (or
). For typical values θc ≪ 1 (or θc, Ep ≪ 1), the exponential factor is tiny and hence the approximation applies to essentially all relevant luminosities and Ep’s. The luminosity function induced by a Gaussian universal apparent structure is therefore uniform in log(L), and the same applies to the Ep distribution.
The effect of a non-zero dispersion in the core quantities is equivalent to that of convolving the zero-dispersion P(L, Ep) probability density distribution with the probability density distribution of the core luminosity and peak SED photon energy P(Lc, Ep, c). In that sense, the probability density distribution of the core quantities acts essentially as a smoothing kernel: in the Gaussian apparent structure case, it introduces a smooth transition to a power law fall-off above and
. In more physical terms, and for typical parameters relevant to the quasi-universal jet scenario, the high-end of the luminosity function is shaped by the distribution of core luminosities, while below
the luminosity function is set by the jet (apparent) structure. The left-hand panel in Fig. 1 shows the luminosity function ϕ(L) = dP/dln(L) = L∫P(L, Ep) dEp for an example Gaussian apparent structure case, demonstrating the effect of introducing a core luminosity dispersion with three different values of the slope A.
![]() |
Fig. 1. Example luminosity functions ϕ(L) = dP/dln(L) induced by a Gaussian apparent structure model (Eqs. (10), left-hand panel) and a power law apparent structure model (Eqs. (13)) with a luminosity profile slope α = 3 (right-hand panel), both with a core half-opening angle θc = 0.1 rad and |
We can get some further insight by adopting a power law apparent structure model with a ‘uniform core’ within θv ≤ θc, namely
again with no dispersion. The change of variables approach can be applied to θv > θc, where the structure is monotonic; for θv ≤ θc it is sufficient to note that all observers see L = Lc and Ep = Ep, c, and the probability of having a viewing angle in this range is 1 − cos θc. Hence, the L − Ep distribution is
for and
, where
and
, and the last approximate equality is valid for
and
. Again, the alternate form, which highlights the dependence on Ep, can be obtained by the substitutions L ↔ Ep, αL ↔ αEp and Lc ↔ Ep, c. For L ≥ Lc, Ep ≥ Ep, c, we have
While in this limiting zero-dispersion case the jet core clearly extends over a zero-measure region of the (L, Ep) plane, a non-zero dispersion spreads this over a more physically sound, finite region of the plane. Hence, a power law apparent structure induces a power law luminosity function with a slope L−2/αL in the logarithm of L (Pescalli et al. 2015). Similar conclusions apply to the induced Ep distribution. The slope parameter αL, together with the core half-opening angle θc, control the extent of the luminosity function as a consequence of the limited physical viewing angle range 0 ≤ θv ≤ π/2. The right-hand panel in Fig. 1 shows the luminosity function in an example power law case with αL = 3.
The above derivation also shows that in a quasi-universal structured jet scenario ϕ(L) can never attain a positive slope, except at the low-luminosity end, or in a narrow luminosity range close to the core for small dispersions (unless αL < 0, which however does not seem a likely physical possibility). This is a feature of quasi-universal jet models.
2.3.2. L − Ep correlation induced by the jet structure
Since both ℓ and η are functions of the viewing angle, the quasi-universal structured jet model inherently implies a Yonetoku correlation (Yonetoku et al. 2004) between L and Ep within the observed population (e.g. Salafia et al. 2015), in addition to any intrinsic correlation that may hold between the core quantities Lc and Ep, c. The shape of the induced correlation can be obtained by eliminating θv in the average apparent structure functions, to obtain η(ℓ), and is already apparent in the Dirac-delta functions in Eqs. (12) and (14). In the Gaussian average apparent structure case, it yields , so that the slope of the correlation is set by the ratio of the scale parameters θc and θc, Ep over which L and Ep decay with the viewing angle. Similarly, in the power law case the relation is (Ep/Ep, c) = (L/Lc)αEp/αL. In general, the induced correlation is a power law (or a collection of power law branches) whenever the functional forms of ℓ and η are the same, and its slope is set by the ratio of the decay rates of ℓ and η.
2.3.3. Average apparent structure model adopted in this work
In this study, in order to endow the average apparent structure model with a high degree of flexibility (in the absence of strong theoretical constraints on the expected shape), we adopted a double smoothly broken power law model with a nearly constant core within θv < θc and a break at a wider angle θw, that is,
where the ‘smoothness’ parameter is set to s = 4, which makes the transitions between the power law branches relatively sharp, to compensate the fact that the intrinsic dispersion of core quantities tends to smooth out the induced breaks in the luminosity function. The implied L − Ep correlation has two branches, with slopes αEp/αL and βEp/βL, the break being around and
. Figure 2 demonstrates the features of such a model in detail, including the induced L and Ep probability distributions, based on an example choice of parameters.
![]() |
Fig. 2. Example double smoothly broken power law jet apparent structure model and induced L and Ep probability distributions. Panels a.1 and a.2 show the apparent jet structure functions |
With this choice of average apparent jet structure functions, the model features a total of 14 free parameters. We list these parameters in Table 1, along with brief definitions and with information on the priors adopted on each of them in the analysis, discussed later in the text.
Population model parameters and adopted priors.
2.4. Sample definition and selection effect modelling
In order to compare a population model with an observed sample, the selection effects that shape the latter must be taken into account. Here we describe our sample choice and the procedure that we employed to model the underlying selection effects. We adopted a description of the SGRB photon spectrum as a cut-off power law (Ghirlanda et al. 2004), dṄ/dE(E,Ep,obs,α) ∝ Eαexp[−(2+α)E/Ep,obs], where Ṅ represents the rate of photons hitting the detector, E is the photon energy, and Ep, obs = Ep/(1 + z). We set the low-energy photon index to α = −0.4, which is the median of the values reported in the Fermi/GBM online catalogue for SGRBs. The peak photon flux in the E0 − E1 keV band is then defined as
where dL is the luminosity distance and
We stress here again that we extended the customary 1 − 104 keV pseudo-bolometric rest-frame band to the wider 0.1 − 107 keV to include possible cases with very high Ep. In practice, we pre-computed3 ℰ[50 − 300] for Fermi/GBM and ℰ[15 − 150] for Swift/BAT over a uniformly spaced two-dimensional grid in (log Ep, log z) and then used two-dimensional linear interpolation to recover it and obtain the photon flux from Eq. (17) (or equivalently to obtain L from p and Ep, obs, by inverting the equation).
We considered three reference SGRB samples: (i) SGRBs detected by Fermi/GBM, with available spectral information in the public catalogue; (ii) SGRBs detected by both Swift/BAT and Fermi/GBM, with a number of additional cuts to reach a high completeness in redshift; and (iii) SGRBs detected by Fermi/GBM with a GW counterpart, which currently includes only GRB 170817A/GW170817. In what follows, we describe in detail the selection cuts of each sample and our approach to the modelling of selection effects.
2.4.1. Observer-frame sample: Fermi/GBM SGRBs with spectral information
Our ‘observer-frame’ sample includes GRBs detected by the GBM on board Fermi from the start of the mission up to mid July 2018, after which no spectral information is present in the online catalogue (von Kienlin et al. 2020) at the time of writing. Among these, we selected 367 events that are nominally short, that is, their duration T90 < 2 s, as our initial raw sample. As a first approach, we considered working with the sub-sample of bursts whose peak photon flux is larger than a ‘completeness’ threshold p[50 − 300] > plim, GBM, where the GBM flux in the 50–300 keV band is measured with 64 ms binning to best approximate the actual peak photon flux. By visual inspection we found that above plim, GBM = 3.5 cm−2 s−1 the observed N(> p[50 − 300]) distribution looks like a single power law (see Fig. 3), and hence we selected this as our completeness threshold. This is a practical approach historically employed to construct flux-complete samples. The corresponding detection probability can be modelled simply as
![]() |
Fig. 3. Inverse cumulative distribution of Fermi/GBM SGRB peak photon fluxes. The solid red line shows the inverse cumulative number of SGRBs detected by GBM with spectral information available in the catalogue, as a function of the peak photon flux measured on a 64 ms timescale in the 50–300 keV band. The pink band shows the one-sigma-equivalent Poisson error. The dashed black line shows a power law |
where Θ is the Heaviside step function, that is, Θ(x) = 1 if x > 0 and Θ(x) = 0 otherwise.
This selection criterion significantly reduces the sample size: out of a total of 367 SGRBs in the raw sample, only 212 have p[50 − 300] > 3.5 cm−2 s−1. Most importantly, the discarded events possibly probe the luminosity function down to lower luminosities, which is where most of the useful information on the jet structure resides in a quasi-universal jet scenario. Last, but not least, the only SGRB with reliable viewing angle information, that is, GRB 170817A, is not included in this flux-limited sample. In order to access the flux-incomplete part of the Fermi/GBM SGRB sample, we carefully constructed a detection probability Pdet(L, Ep, z) by simulating the response of the GBM NaI detectors to SGRBs with a broad range of characteristics, as described in detail in Appendix C. In this study, we compare the results obtained by using either of the strategies in modelling the selection effects: hereafter, we refer to the reduced sample of Fermi/GBM SGRBs with p[50 − 300] > 3.5 cm−2 s−1 as the ‘flux-limited sample’, and to the analysis adopting the associated simplified selection effect model as the ‘flux-limited sample analysis’. Conversely, the analysis performed adopting the simulated Fermi/GBM detection efficiency is referred to as the ‘full sample analysis’.
As a further countermeasure against possible biases, we applied an additional quality cut to both the above samples: we removed events with best-fit Ep, obs > 10 MeV or Ep, obs < 50 keV, which fall outside the spectral range where the effective area of the GBM detectors is optimal: this removes two events whose uncertainty on Ep, obs is very large, reducing the flux-limited sample to 210 events. In the full sample analysis, we also removed another ten events with a low best-fit peak 64 ms photon flux p[50 − 300] < 1 cm−2 s−1, all of which have very large errors on both p[50 − 300] and Ep, obs. This reduces the ‘full’ sample to 355 events. To reflect these further quality cuts, we updated our model detection probabilities by multiplying them by Θ(p[50 − 300] − 1 cm−2 s−1)Θ(Ep, obs − 50 keV)Θ(10 MeV − Ep, obs).
2.4.2. Rest-frame sample: Flux-complete sample of Swift/BAT SGRBs also observed by Fermi/GBM
For some of the events in our Fermi/GBM observer-frame sample, redshift information is available and can be used to constrain the population parameters better than can be done using the observer-frame information (p[50 − 300], Ep, obs) only. In order to avoid biases, on the other hand, any additional selection effects at play in the sub-sample with known redshift must be accounted for in the inference, that is, in the Pdet model. The redshift determination is a complex process that involves multiple facilities and depends not only on the prompt emission properties, but also on those of the afterglow. Therefore, modelling the associated selection effects is prohibitive. On the other hand, Salvaterra et al. (2012) and D’Avanzo et al. (2014) showed that it is possible to construct a sample with a selection that is easier to model, but that leads to a high redshift completeness. The selection involves two cuts that do not bias the redshift distribution, namely (i) a cut on the foreground interstellar dust extinction AV and (ii) a cut on the Swift X-Ray Telescope (XRT) slew time; plus a cut on the BAT 64 ms peak flux in the 15–150 keV band, p[15 − 150] > 3.5 cm−2 s−1, to ensure flux completeness. The original SGRB sample constructed in this way, known as the S-BAT4 (D’Avanzo et al. 2014), included 16 events, 11 of which had a measured redshift. Thanks to a considerable effort spent by the community, and in particular by Fong et al. (2022) and Nugent et al. (2022), in identifying SGRB host galaxies and measuring their redshifts, it has been recently possible to construct an extended ‘S-BAT4ext’ sample (Ferro et al. 2023; D’Avanzo et al., in prep.) with a more than doubled size and an increased redshift completeness. For this work, we adopt the S-BAT4ext sub-sample of 18 events that have been jointly detected by Fermi/GBM with T90 < 2 s (for consistency with our observer-frame sample). Of these, 16 have either a spectroscopic (12 events) or photometric (4 events) redshift measurement, as listed in Table 2. For these SGRBs we performed an independent analysis of their peak spectra, based on publicly available Fermi/GBM data, as described in Appendix B, obtaining posterior samples of their bolometric flux F and observed peak photon energy Ep, obs, adopting a prior π(F, Ep, obs)∝F−1. For the 12 events with a spectroscopic redshift, these were converted into samples of P(L, Ep, z | di) by simply fixing the redshift at the best fit value. For the four events with photometric redshift, we obtained samples of P(L, Ep, z | di) by computing and Ep, l, m = (1 + zl)Ep, obs, m, where
are Ns photometric redshift posterior samples from the Broad-band Repository for Investigating Gamma-ray burst Host galaxies Traits (BRIGHT) catalogue (Nugent et al. 2022),
are the corresponding luminosity distances under our assumed cosmology and
are Ms posterior samples from the spectral analysis. In both cases, the effective prior on the source parameters is π(L, Ep, z)∝(1 + z)−1L−1. The two events with unknown redshift were not included in the analysis.
Rest-frame SGRB sample.
The careful selection adopted to construct this sample allowed us to model the underlying selection effects by multiplying the GBM detection efficiency, Pdet, GBM, by the simple BAT detection efficiency:
where plim, BAT = 3.5 cm−2 s−1 is numerically identical to plim, GBM by pure chance.
2.4.3. Viewing angle sample: GRB 170817A/GW170817
The third and last sample we considered is that of Fermi/GBM SGRBs that have a GW counterpart produced by the inspiral of a BNS merger, from which a measurement of θv can be obtained under the assumption that the jet is launched along the direction of the total angular momentum. At present, the sample clearly consists of the single event GRB 170817A with its counterpart GW170817. We use dG17 to indicate the available data regarding the prompt emission and the GW signal of the event, and with dHOST the available data of the host galaxy NGC4993 (Coulter et al. 2017; Hjorth et al. 2017; Cantiello et al. 2018). We took the host galaxy spectroscopic redshift zH = 0.009783 (Coulter et al. 2017; Hjorth et al. 2017) as the redshift of the source, neglecting the small uncertainty on the actual cosmological redshift (i.e. corrected for the galaxy proper motion), and hence P(z | dHOST) = δ(z − zH). We obtained the posterior P(L, Ep | dG17, zH), whose 50% and 90% containment contours are shown in the left-hand panel of Fig. 4, from our own analysis of the peak spectrum (Appendix B). For what concerns the viewing angle, in order to break the distance-inclination degeneracy inherent in the BNS inspiral GW analysis, we proceeded similarly as in Mandel (2018, but keeping the cosmological parameters fixed, differently from them), as follows: we approximated the host galaxy luminosity distance r uncertainty as
![]() |
Fig. 4. L − Ep contours for our sub-sample of Fermi/GBM SGRBs detected also by Swift/BAT with a 64 ms peak flux p[15 − 150] > 3.5 cm−2 s−1 (see the main text for the full sample selection criteria). The main panel shows the contours that contain 50% (thick lines) and 90% (thin lines) of the posterior probability on (L, Ep) for bursts with a spectroscopically measured redshift, plus GRB 170817A (which does not belong to the GBM+BAT sample); the right-hand inset shows the corresponding contours for four events with a photometric redshift measurement (from Fong et al. 2022); the left-hand inset shows the contours for the remaining two SGRBs with an unknown redshift, constructed assuming a uniform prior on z in the range (10−4, 4). These two events were not included in the analysis. |
with μr = 40.7 Mpc and σr = 2.36 Mpc (mean and square sum of statistical and systematic errors from the measurements performed by Cantiello et al. 2018). We then computed the posterior on the viewing angle as
where P(θv, r | dG17) is the joint posterior on θv and r from the GW analysis. In practice, we obtained samples of the posterior probability density in Eq. (22) by re-sampling the publicly available posterior samples from the low-spin prior analysis of GW170817 performed by Abbott et al. (2019) with a weight equal to the right-hand side of Eq. (21) evaluated at the luminosity distance of each sample. The viewing angle posterior probability density obtained from a kernel density estimate on the resulting samples is shown in Fig. 5.
![]() |
Fig. 5. GRB 170817A viewing angle posterior probability distribution, assuming the jet to be aligned with the GW170817 binary total angular momentum. The blue line shows the posterior probability constructed using the posterior samples from the low-spin-prior GW analysis (Abbott et al. 2019), while the red line shows the result of conditioning on the host galaxy distance (Cantiello et al. 2018), as explained in the main text. |
We modelled the selection effects acting on this sample as the product of Pdet, GBM(L, Ep, z) times a GW detection efficiency Pdet, GW(θv, z). Since the time-volume surveyed by aLIGO and Advanced Virgo so far is dominated by that of the third observing run, O3 (LIGO Scientific Collaboration 2021), we constructed the detection efficiency assuming, for the sake of simplicity, the GW network sensitivity of O3, neglecting periods with a lower sensitivity. To do so, we retrieved the dataset made publicly available by the LIGO Scientific Collaboration et al. (2023) that contains information on the response of online GW search pipelines to a large number of simulated signals injected into O3 data. We re-weighted the BNS merger injections in that dataset to reflect a population with a primary mass distribution between m1, min = 1.2 M⊙ and m1, max = 2.1 M⊙ (similar to the preferred distribution from the GWTC-3 population analysis, Abbott et al. 2023) and a flat secondary mass distribution p(m2 | m1)∝Θ(m1 − m2)Θ(m2 − m1, min). Assuming the inclination ι and the jet viewing angle θv to be related by
we binned the injected signals into a number of two-dimensional bins in (θv, z) space centred at (θv, i, zj). Calling wk and ρk the weight and network signal-to-noise ratio (S/N) associated with the k-th injected signal, we estimated Pdet, GW(θv, z) at the centre of each bin as
where ℐi, j represents the set of indices k of injections whose viewing angle and redshift fall into the bin centred at (θv, i, zj). The GW detection efficiency on the rest of the (θv, z) space was obtained by two-dimensional linear interpolation of the resulting values. The relatively high cut ρ ≥ 12 includes GW170817 (Abbott et al. 2017a) and ensures that the detection can be represented by a simple cut in S/N, in analogy with the flux completeness cuts discussed previously.
2.5. Inference on the population properties
Within a Bayesian hierarchical approach, the posterior probability on the population parameters λpop can be written as (Eqs. (7) and (8) in Mandel et al. 2019, hereafter M19)
where the index i runs over the N events in the sample, di (the i-th element of d) represents the corresponding data in the detectors (i.e. Fermi/GBM, plus either Swift/BAT or the aLIGO and Advanced Virgo detectors in our case), π(λpop) is the prior on the population parameters, and is a normalisation factor. We indicate with 𝒩i = 𝒩i(di, λpop) the numerator of the fraction after the product symbol in the above equation, that is,
and with 𝒟 = 𝒟(λpop) the denominator, namely
This equation contains the detection efficiency Pdet(λsrc), which must be chosen consistently with the selection effects acting on each of the considered samples.
2.5.1. Numerator 𝒩i for observer-frame sample events
For events in our observer-frame sample, which have unknown redshifts, we approximated the posterior probability on the source parameters P(L, Ep, z | di) = P(di | λsrc)π(L, Ep)π(z)/P(di) by neglecting the uncertainty on the measured peak photon flux pi = p[50 − 300],i and peak photon energy Ep, obs, i (which is justified by the large sample size and by the quality cuts discussed in the previous section), and by assuming the posterior to simply reflect the prior π(z) along the z axis. In other words, we assumed P(L, Ep, z | di)∝π(z) when keeping L and Ep fixed: this is equivalent to stating that no redshift information is available. This leads to
The appropriate prior π(L, Ep) was obtained by applying a coordinate transform to a uniform prior in both pi and Ep, obs for these events, that is,
where |X| represents the determinant of matrix X, and . The integral in the numerator of Eq. (25) is then
and we note that the P(di) term eventually cancels out in Eq. (25) (as in M19).
2.5.2. Numerator 𝒩i for rest-frame sample events
For events in the rest-frame sample, in order to carry out the integral over z, L and Ep in 𝒩i, we employed the same Monte Carlo approximation as M19, that is,
where represent a total of Ns samples of the luminosity, peak photon energy, and redshift posterior of the GRB. In cases with a photometric redshift, the (L, Ep) posterior was obtained by combining the results from our analysis of the spectrum at the peak of the GRB with the redshift posterior samples from Nugent et al. (2022), obtained from the BRIGHT online catalogue4. In cases with a spectroscopic redshift, we simply kept zi, j fixed at the best-fit redshift. The prior π(L, Ep) in Eq. (31) is proportional to (1 + z)−1L−1, as discussed in Sect. 2.4.2.
2.5.3. Numerator 𝒩i for viewing angle sample events
For the viewing angle sample, that is, for GRB 170817A, it was necessary to proceed differently for the full sample and flux-limited sample analyses: the peak photon flux of the GRB, p[50 − 300] = 2.02 cm−2 s−1, is below the completeness cut plim, GBM = 3.5 cm−2 s−1 adopted in the flux-limited sample analysis; hence, in this case, the simple treatment of GBM selection effects is not adequate. In the full sample analysis, this is not a problem since the GBM detection efficiency model from Appendix C accounts for the smooth decrease in the detection efficiency at low peak photon fluxes. In the full-sample case, the correct form of 𝒩i can be obtained by reformulating our population model by including θv among the source parameters, . The population probability then becomes
, and the integral over θv is deferred to 𝒩i, namely
It is straightforward to verify that whenever di does not contain information on the viewing angle this leads to exactly the same definition of 𝒩i as before.
For GRB 170817A, we have (see Sect. 2.4.3)
where the prior π(θv) = sin θv corresponds to that used in the GW analysis. This leads to
Approximating again the integral with a Monte Carlo sum, we obtained
where are a total of Ns samples from our analysis of the peak spectrum of GRB 170817A and
are Ng posterior samples of the probability in Eq. (22). Clearly, the term P(dG17, dHOST) eventually cancels out in Eq. (25) as before.
In the flux-limited sample analysis, the information on the viewing angle, luminosity and peak photon energy of GRB 170817A/GW170817 can still be used to condition the apparent structure parameters before the flux-limited sample analysis is performed, because the structure must be consistent with what has been observed in that event. The starting point is
Application of Bayes’ theorem gives
but we have P(L, Ep, θv | λpop) = P(L, Ep | θv, λpop) sin θv and π(θv) = sin θv, which therefore leads to P(λpop | L, Ep, θv) = P(L, Ep | θv, λpop)π(λpop)/π(L, Ep). Substitution of this into Eq. (34) leads to
This is similar to 𝒩G17, but the redshift information is not used. This can be again approximated by Monte Carlo integration over samples drawn from the posterior P(L, Ep, θv | dG17, zH), namely
This result can then be used as a prior in the analysis of the observer-frame and rest-frame samples. The final posterior on λpop then becomes
Hence, the posterior takes a similar form as in the full-sample analysis case, the difference being in a missing P(zH | λpop)/𝒟 factor.
2.5.4. Denominator 𝒟 for events in the three samples
The denominator 𝒟 in the above expressions represents the fraction of events in the population that pass the sample selection criteria (M19), that is, the ‘accessible’ events according to the selection effects model. For the observer-frame sample events, the denominator takes the form
For the rest-frame sample, it is
For the viewing-angle sample, it is a four-dimensional integral, namely
2.6. Choice of priors
Our Bayesian inference approach requires the definition of a prior probability density over the ‘hyper’ parameter space the λpop vector belongs to. We chose independent priors on all parameters, except for the pair (θc, θw) for which we enforce θw > θc, and hence
As shown in Table 1, we selected mildly informative priors on most parameters, typically adopting a uniform or uniform-in-log prior over a relatively wide range that encompasses what we consider reasonable values. There are two exceptions: for the core half-opening angle θc and the break angle θw we adopted a prior that is uniform in the subtended solid angle, π(θ)∝ sin θ, since their role is that of setting the probability for a jet to be observed within the core or within the ‘wing’ where the break occurs; for the low-end cutoff of the core luminosity , we set a relatively high lower limit
≥ 3 × 1051 erg s−1, consistently with our quasi-universal structured jet scenario, where relatively low luminosities are produced by off-axis jets and not by intrinsically under-luminous jets.
2.7. Stochastic sampling of the population posterior
In order to realise our inference in practice, we sampled the posterior probability density on λpop in Eq. (25) using the publicly available python package emcee (Foreman-Mackey et al. 2013), which constitutes an efficient and flexibile implementation of the Goodman & Weare (2010) affine-invariant ensemble sampler5. Numerical integrals were performed using the trapezoidal rule over a four-dimensional grid with a resolution of 1000 linearly spaced points in θv between 0 and π/2 and 60 logarithmically spaced points in each of the remaining axes, in the domain L ∈ (1044, 1056) erg s−1, Ep ∈ (10−1, 107) keV and z ∈ (10−3, 101). We ensured that this resolution was sufficient by comparing the value of the posterior at a number of points in the parameter space with those obtained with a higher resolution of 100 points over each axis, and found a negligible difference as long as A < 10 and σc ≳ 0.05 dex (our prior limits fall well within these requirements).
We set up the sampler to employ 14 × 2 = 28 walkers, and ran it for 104 iterations for each analysis, for a total of 2.8 × 105 posterior samples. The autocorrelation length of the resulting chains, averaged over all walkers, is around 650. A corner plot showing the density of the samples in the 14-dimensional parameter space is shown in Fig. A.1, while summary statistics for each parameter are reported in Table 3. The results presented in the next section are constructed using 1000 random posterior samples from these chains, after discarding the first half as burn-in6.
Constraints on population model parameters from our analysis.
3. Results
Generally speaking, the full sample and flux-limited sample analyses yielded quite similar results, as demonstrated by the detailed posterior probability density distributions (Fig. A.1) and the summary in Table 3. The main notable differences in the full sample analysis versus the flux-limited sample analysis are a preference for a slightly larger on-axis SED peak photon energy Ep, c; a steeper slope, a, of the rate density evolution before peak; and a slightly better constrained redshift zp of the peak of the rate density evolution. The posterior distributions from the two analyses show large overlaps for all parameters. In what follows, we present a thorough description of the results, further highlighting the differences in the results from the two analyses when relevant.
3.1. Apparent structure
First of all, we focus on the apparent jet structure. Figure 6 shows the constraint we obtained on the ‘average apparent jet structure’ (central panel) and
(bottom panel) from the two analyses, with the insets showing the posterior distribution on the core dispersion parameters A and σc. In this and the following figures, solid lines refer to the full-sample analysis and dashed lines to the flux-limited sample analysis. The small panel at the top shows the posterior probability density on the logarithm of θc (in grey) and of θw (in green). The figure shows that the preferred apparent structure features a narrow core of ∼2 − 3 deg, outside of which the luminosity falls off approximately as
and the SED peak photon energy as
. The posterior on the transition angle θw rails against the upper boundary of its physical range, and the slopes βL and βEp at larger viewing angles are not well constrained by the available data (see Table 3). This indicates that the data are consistent with an apparent structure described by a single power law outside the core: this may disfavour somewhat the presence of a distinct dissipation mechanism that dominates the gamma-ray emission at viewing angles in the range 20 ≲ θv/deg ≲ 30 (e.g. cocoon shock breakout as proposed by Gottlieb et al. 2018) in the majority of SGRBs, unless the apparent structure of the combined emission components follows the same power law. A caveat here is that there are most likely not more than a handful of events in our samples that are observed at such large viewing angles (see Sects. 4.2 and 4.4), and therefore this result must be taken with a grain of salt.
![]() |
Fig. 6. Average apparent jet structure. The two larger panels show the constraints on the SGRB average apparent structure obtained from our analysis: solid lines show the medians of the projected posterior distributions of the jet structure functions at fixed θv from the full sample analysis (red: |
The structure is by construction consistent with the GRB 170817A luminosity and Ep at the relevant viewing angle, as shown by the light green contours in the figure. These contours represent the 50% and 90% credible regions of (θv, L) and (θv, Ep) for GRB 170817A, constructed by combining the viewing angle information from the GW analysis conditioned on the host galaxy distance (Sect. 2.4.3) and our GRB 170817A spectral analysis at peak (Appendix B).
The uncertainty band on the typical luminosity at θv ≪ θc is particularly narrow because of the chosen prior on
(see the discussion in Sect. 4.5). The larger uncertainty at θv ∼ θc arises because of the combined uncertainties on
, θc and αL.
The jet structure functions can also be projected onto the (L, Ep) plane: Fig. 7 shows the median relation and the corresponding 90% credible region for the two analyses. In both cases, the bulk of the SGRBs with known redshift (shown by red crosses) are close to the upper-right end of the relation, which indicates that they are observed close to on-axis in the model (more precisely, close to the edge of the core; see also Sect. 4.2). Interestingly, the relation lies above most of the SGRBs in the known-redshift sub-sample, indicating that selection effects play a major role in how the plane is populated, according to the model. We expand on this in Sect. 4.3.
![]() |
Fig. 7. Average apparent jet structure in the L − Ep plane. The solid turquoise line connects the points |
The inset in Fig. 7 shows the posterior probability density on the y parameter that sets the correlation between the on-axis luminosity and peak SED photon energy. Despite our model allowing for such a correlation, the posterior is fully compatible with y = 0, that is, the absence of an intrinsic correlation between Lc and Ep, c.
3.2. Luminosity function
Our analysis does not directly constrain the local rate density, R0, of SGRBs, because in our inference framework it is effectively only a normalisation factor as long as its prior is uniform in the logarithm (M19; Fishbach et al. 2018). In order to derive the local rate density implied by our results, we required the observed rate of events with p > plim to be equal to7Robs(> plim, GBM) = N(> plim, GBM)/ηGBMTGBM ≈ 27.6 yr−1, where N(> plim, GBM) = 212 is the number of SGRBs with p > plim, GBM in our sample, ηGBM = 0.59 is a factor that corrects for the accessible field of view and the duty cycle of GBM (Burns et al. 2016), and TGBM = 13 yr is the Fermi mission duration at the time of the last SGRB in the observer-frame sample. Given λpop, the local rate density is then
where Pdet, GBM is the flux-limited form from Eq. (19) and is independent of R0 (see Eq. (9)). We applied the above expression to our population posterior samples
to obtain an equal number of
samples. In turn, this allowed us to construct samples of the posterior distribution of the luminosity function,
, where ϕ(L, λpop) = L∫P(L, Ep | λpop) dEp.
Figure 8 shows the luminosity function of SGRBs obtained in this way from the full-sample analysis (solid red) and the flux-limited sample analysis (dashed orange). The corresponding results from W15 and G16 (their fiducial model ‘a’) are shown in blue and grey, respectively. The shape of the result is somewhat in between the W15 and G16 results, but the normalisation of the high-luminosity end is in better agreement with W15 than G16. The high-luminosity end slope (1 − A ∼ 2) is in agreement with both benchmark results, while the flattening below the break is reminiscent of that found by G16, but translated to a luminosity that is lower by almost one order of magnitude. At luminosities below ∼ 3 × 1051 erg s−1, the slope of the luminosity function is ∼0.4, which is flatter than W15 (but not by as much as G16) and more similar to the older results from Guetta & Piran (2005, 2006). The low-luminosity end below ∼1049 erg s−1 remains highly uncertain.
![]() |
Fig. 8. Luminosity function. The median of the posterior probability density of R0ϕ(L) = dR0/dln(L) (i.e. the local rate density per unit logarithm of the peak luminosity) from the full sample analysis is shown with a solid red line, while the dashed orange line refers to the flux-limited sample analysis. The shaded regions encompass the symmetric 90% credible interval at each fixed L. We show for comparison the corresponding results from W15 (blue) and G16 (grey – their fiducial model ‘a’). The credible region of W15 is computed assuming uncorrelated errors on their parameters. |
Overall, the luminosity function we recover bears some similarity with that obtained by Tan & Yu (2020), whose study also relies on a quasi-universal jet assumption and on a similar form of the core luminosity dispersion. This may indicate that our results are at least in part driven by the chosen parametrisations, which follow from the quasi-universal jet assumption: this could partially explain the difference between our results and those of W15 and G16, and hence also some of the mutual differences between those two previous studies. Other likely sources of difference between these results are the different reference samples and, most prominently, the treatment of selection effects, whose accuracy is vital in order to avoid potentially strong biases in the results (e.g. M09).
3.3. Rate density
The total local rate density, R0, of SGRBs (including all viewing angles) is not well constrained by our analyses, because of the difficulty in determining the actual extent of the luminosity function with the available data. The full-sample analysis yields (median and symmetric 90% credible interval), while the flux-limited sample analysis gives
(see Fig. 9, left-hand panel). Both are compatible with the local BNS merger rate derived from GW observations (Abbott et al. 2023), R0, BNS = 10 − 1700 Gpc−3 yr−1, and also with other recent estimates of the total, collimation-corrected SGRB rate based on different methods (e.g. Rouco Escorial et al. 2022, see Mandel & Broekgaarden 2022 for a review and further references). The fact that the derived SGRB rate leans towards the high-end of the BNS merger rate uncertainty interval can be interpreted as an indication that the fraction of BNS mergers that yield a jet must be high, in agreement with the results of Salafia et al. (2022), Sarin et al. (2022), Beniamini et al. (2019), and Ghirlanda et al. (2019).
![]() |
Fig. 9. Local rate density. In the left-hand panel, the solid red (respectively dashed orange) line shows the posterior probability density on the total local rate density, R0, of SGRBs with a luminosity L > 1044 erg s−1 (which encompasses all viewing angles) from our analysis using the full (respectively flux-limited) sample. The grey shaded band shows the constraint on the local rate density of BNS mergers from GW observations, as derived in Abbott et al. (2023). In the right-hand panel, the solid red and dashed orange lines show the local rate density for events with a luminosity L > 1050 erg s−1 in our two analysis setups. For comparison, we show the corresponding local rate densities from W15 and G16 above the same minimum luminosity. |
In order to compare our local rate density with those presented in the literature, we also computed the rate density of events above a minimum luminosity Lmin = 1050 erg s−1. The right-hand panel in Fig. 9 shows the result from our two analyses, which yield (full sample) and
(flux-limited sample), compared with those obtained by integrating the W15 and G16 luminosity functions over the same luminosities. All results are in agreement with each other, placing the local rate density of luminous SGRBs around one event per Gpc3 yr, with roughly one order of magnitude of uncertainty.
Our model also constrains the evolution of the rate density with redshift, , using the parametrisation given in Eq. (9). Figure 10 shows the symmetric 90% credible interval of the posterior probability distribution of
, at each fixed z, considering only events with L ≥ 1050 erg s−1 (solid red: full-sample analysis; dashed orange: flux-limited sample analysis). Lowering the minimum luminosity would increase the uncertainty on the normalisation, but leave the shape identical, as a consequence of our assumption of no evolution of the jet structure parameters with redshift. For comparison, we show the corresponding results from W15 and G16, where the shaded regions only account for the uncertainty on the local rate density, and thus represent an underestimate of the actual uncertainty in these models. The shape of the constraint is in qualitative agreement with the result of G16, who find an evolution that is compatible with the expectations from compact binary merger progenitors but is in strong disagreement with the sharp cut-off in the rate density at z > 0.9 found by W15. We believe that this disagreement stems from a bias induced by the limited treatment of selection effects in that work, with a similar impact as that described in Bryant et al. (2021) for non-parametric methods.
![]() |
Fig. 10. Rate density evolution. The figure shows the redshift evolution of the rate density from both our full sample (solid red lines) and flux-limited sample (dashed orange lines) analysis, limited to events above a minimum luminosity Lmin = 1050 erg s−1. The corresponding evolutions from W15 and G16 are shown in blue and grey, respectively. The thick lines show the median of the posterior predictive distribution at each fixed redshift, while the shaded areas encompass the symmetric 90% credible interval. |
The low-redshift scaling of the SGRB rate, (1 + z)a with , is steeper than that of the CSFR at low redshift, CSFR ∝ (1 + z)2.7 (Madau & Dickinson 2014), while the constraint on the peak of the SGRB rate
indicates a preference for a rate that peaks at larger redshift than the CSFR (whose peak is at z ∼ 1.9). This is difficult to explain with either very short delay times between star formation and BNS mergers (which would suggest that the SGRB rate should trace the CSFR) or long delay times, which would shift the peak of the SGRB to lower redshift than the CSFR peak (though the uncertainty on the SGRB redshift peak could allow for the long delay time interpretation). This apparent discrepancy could point to a redshift-dependent evolution of the yield of merging BNSs per unit star formation or a redshift-dependent evolution of the fraction of BNS mergers yielding SGRBs. Such effects could plausibly be caused by metallicity-dependent variations in stellar and binary evolution, including in NS masses. We note, however, that three out of four SGRBs with a photometric redshift in our rest-frame sample, namely 170127B, 180727A and 191031D, have median redshift larger than 1.9. If we remove the four SGRBs with a photometric redshift from our sample, we obtain generally similar results (with larger error bars) except for the redshift evolution, whose preferred low-redshift slope and peak become more consistent with the Madau & Dickinson (2014) CSFR (but clearly with larger error bars: a ∼ 3 ± 1.5, zp ∼ 1.9 ± 1), which could reconcile the result with the expectations for BNS mergers with short delay times. Thus, the redshift distribution is sensitive to the reliability of these photometric redshifts.
3.4. Comparison with the three reference samples
Figure 11 compares the distributions of 64 ms peak photon flux p[50 − 300] and observed peak photon energy Ep, obs distributions predicted by our model (using the median of the λpop posterior distribution) with the observer-frame Fermi/GBM sample. For the flux-limited sample analysis, we limit the comparison to the sub-sample of events with p[50 − 300] > plim, GBM. The figure demonstrates an excellent agreement both in the joint p[50 − 300] − Ep, obs distribution and in the individual distributions. The fact that the shape of the low-end of the inverse cumulative log(Nobs)−log(p[50 − 300]) distribution is well reproduced (panel b in the figure) indirectly demonstrates the ability of our Fermi/GBM detection efficiency model to accurately reproduce the selection effects of the full sample.
![]() |
Fig. 11. Observer-frame constraints and best-fit model predictions. Panel a.0 shows the predicted distribution of Fermi/GBM SGRBs on the (p[50 − 300], Ep, obs) plane for our best-fit model, with progressively lighter contours containing 50%, 90%, 99% and 99% of the events. Red dots show the observed data reported in the Fermi/GBM catalogue that pass our additional quality cuts, while grey points show the events that are discarded. We additionally show the error bars with thin grey lines for those events with a relative error larger than 50% on either quantity, or both. Panels a.1 and a.2 show the predicted (solid blue line) and observed (solid red cumulative histogram, with the pink region showing the 90% confidence region that stems from statistical uncertainties on spectral fitting parameters) cumulative distributions of p[50 − 300] (panel a.1) and Ep, obs (panel a.2) for events that pass the quality cuts. Panel b shows the inverse cumulative distribution of p[50 − 300], highlighting the behaviour at the high-flux end, which follows the expected |
In Fig. 12 we also compare the distribution of L, Ep and z of our rest-frame sample (Sect. 2.4.2), in red, with those predicted by the population model with the parameter constraints from the full sample analysis (blue) and flux-limited sample analysis (grey). In panels a.1, a.2 and b we show the 90% credible bands that stem from the λpop uncertainty. In this case, we apply the detection efficiency model described in Appendix C to both the full sample and flux-limited sample analysis results in order to compute the predicted distributions. In panel a.0 we additionally show the only event in our viewing angle sample, GRB 170817A (orange cross), along with the 50%, 90%, 99% and 99.9% containment contours (shades of orange) of the distribution of SGRBs detected by Fermi/GBM and with a BNS merger counterpart detected by aLIGO and Advanced Virgo with the O3 sensitivity, as predicted by the population model with the best-fit parameters from the full sample analysis.
![]() |
Fig. 12. Rest-frame sample, viewing angle sample, and model predictions. This is similar to Fig. 11, but for rest-frame quantities L and Ep. Panel a.0 shows the predicted smallest regions in the (L, Ep) plane where 50%, 90%, 99% and 99.9% of the SGRBs that are detected by Fermi/GBM and by Swift/BAT with p[15 − 150] > plim, BAT are located, according to our model, using the parameters at the median of the λpop posterior (filled blue contours: full sample analysis; dashed grey contours: flux-limited sample analysis). Red crosses show the SGRBs with a measured redshift in our rest-frame sample. Panels a.1, a.2 and b show the predicted (blue solid lines: full sample; grey dashed lines: flux-limited sample) and observed (red cumulative histograms, with pink shaded areas showing the 90% confidence regions) cumulative distributions of luminosities (panel a.1), Ep (panel a.2), and redshifts (panel b). In panels a.1, a.2 and b, the 90% credible bands stemming from the uncertainty on λpop are shown with blue shading (full sample) and dotted lines (flux-limited sample). The additional orange contours in panel a.0 show the 50%, 90%, 99% and 99.9% containment regions for SGRBs with a BNS merger counterpart detected by Fermi/GBM and by the aLIGO and Advanced Virgo detectors with O3 sensitivity. The orange cross shows the position of GRB 170817A on the plane. |
4. Discussion
4.1. Jet intrinsic structures compatible with the derived apparent structure
4.1.1. Jet total luminosity and energy
Keeping in mind that L represents the luminosity at the peak of the light curve, the actual time-averaged gamma-ray luminosity can be written as ⟨L⟩=ξL, where we take the reference value ξ = 0.3 ξ−0.5, which is the median of the peak flux to average flux ratios in the GBM sample. The prompt emission energy conversion efficiency is ϵγ = 0.1 ϵγ, −1, where the reference value is based on the results of Beniamini et al. (2016). Therefore, the two factors compensate each other to some extent, and the jet core total (i.e. prior to dissipation that leads to gamma-ray emission) average isotropic-equivalent energy output rate is . If the jet duration is T = 1 T0 s, the core isotropic-equivalent jet energy is then, by definition,
. This is compatible with typical estimates of the core isotropic-equivalent energy of the GRB 170817A jet, which fall in the range 1051–3 × 1052erg s−1 (see e.g. Fig. S6 in Ghirlanda et al. 2019), and in line with cosmological SGRBs in general, which typically fall in a similar range (e.g. Rouco Escorial et al. 2022; Fong et al. 2015). The total jet energy is of the order of
. Also, this value is in line with those inferred from afterglow modelling of cosmological SGRBs (Rouco Escorial et al. 2022).
4.1.2. Angular structure
As discussed in Sect. 2.1, the relationship between the intrinsic jet structure and the apparent luminosity angular profile is not straightforward and dependent on the underlying dissipation and emission mechanism. Nevertheless, we can get some insight on the impact of our findings on the intrinsic structure of SGRB jets by adopting some simplifying assumptions: (i) the ratio ξ of the peak luminosity to the average luminosity does not depend on the viewing angle (i.e. the light curve shape is preserved when changing the viewing angle), and (ii) the observed duration of the emission is dominated by the central engine activity time, and hence it is also independent of the viewing angle. Under these assumptions, which likely hold only in a limited range of viewing angles close to the jet core, the peak luminosity scales with the viewing angle in the same way as the isotropic-equivalent energy (Eq. (1)). We further assumed a power law profile for the jet total isotropic-equivalent energy, Etot, iso, and a bulk Lorentz factor, Γ, that is, we set Etot, iso(θ)∝(θ/θc)−sE and Γ(θ)∝(θ/θc)−sΓ for θ > θc. It is likely that the gamma-ray efficiency ϵγ is also a function of the angle from the jet axis8, so that we also set ϵγ(θ)∝(θ/θc)−sϵ, with sϵ > 0. Therefore, the isotropic-equivalent energy radiated in gamma-rays at each angle goes as Erad, iso(θ)∝(θ/θc)−sE − sϵ.
In principle, using Eq. (1) one can derive the profile of the observed isotropic-equivalent energy Eγ, iso(θv) (and hence peak luminosity L, given our assumptions) as a function of the viewing angle given the slopes sE, sϵ, sΓ, the core bulk Lorentz factor Γc and the core half-opening angle θc. On the other hand, we can simplify the problem by noting that at relatively small viewing angles the emission is dominated by material along the line of sight, provided that Γc is relatively large, say Γc ≳ 100. In this regime, Eγ, iso(θv)∼Erad, iso(θ = θv) (Rossi et al. 2002); thus, αL ∼ sE + sϵ and hence sE ≲ αL. In the same regime, . The
here is the comoving peak SED photon energy and is likely positively correlated with Γ, and therefore sΓ ≲ αEp.
Using the upper end of the 90% credible intervals for αL and αEp from our full sample analysis, these arguments therefore lead to the upper limits sE ≲ 6 and sΓ ≲ 3. We stress again that, given the assumptions, these results only hold for viewing angles close to the jet core.
While these limits clearly conflict with a ‘top-hat’ jet structure (which would correspond to sE, sΓ → ∞), they are still in agreement with the rather steep kinetic energy profiles and shallower Lorentz factor profiles found in studies of the GRB 170817A afterglow (e.g. Hotokezaka et al. 2019; Ghirlanda et al. 2019; Mooley et al. 2022). The approximate θ−3 scaling of the jet isotropic-equivalent kinetic energy found in recent numerical simulations of SGRB jets (e.g. Gottlieb et al. 2020, 2021, 2022) is also compatible with these limits, even though it seems to conflict with the former findings based on the GRB 170817A afterglow.
4.2. Viewing angles of Fermi/GBM SGRBs with known redshifts
Through our population model it is possible to derive a viewing angle probability for any SGRB using only the information on its luminosity and spectral peak energy. Here we focus on SGRBs with a measured redshift in our rest-frame sample. The posterior probability on the viewing angle θv, i of the i-th SGRB in the sample (represented by data di in the data vector d) is
where the last equality follows from Monte Carlo approximation of the integrals, with being samples from the P(Li, Ep, i | di) posterior obtained from the spectral analysis of the SGRB, and
being samples from the population posterior.
Figure 13 shows the resulting population-informed viewing angle posterior probability densities for the SGRBs in our rest-frame sample. The constraints from the full sample and flux-limited sample analyses are in general agreement, with most jets likely viewed a few degrees from the jet axis. Focussing on the full sample analysis results, four SGRBs have population-informed viewing angles that are larger than 2 deg at 95% credibility: GRB 080905A, GRB 131004A, GRB 160821B, and, unsurprisingly, GRB 170817A. The median and 90% credible interval of our population-informed viewing angle estimate for GRB 160821B is , which is compatible with the estimate
by Troja et al. (2019) based on afterglow modelling. The population-informed estimate for GRB 170817A is
, in excellent agreement with afterglow-based estimates that include the information on the centroid proper motion from Very Long Baseline Interferometry imaging (e.g. Mooley et al. 2018, 2022; Hotokezaka et al. 2019; Ghirlanda et al. 2019; Fernández et al. 2022; Govreen-Segal & Nakar 2023), which find viewing angles in the range 15 ≲ θv/deg ≲ 25, and also with the estimate θv = 18 ± 8 deg by Mandel (2018, 68% credible interval) based on a similar method as that employed in Sect. 2.4.3, but which includes a marginalisation over the cosmological parameters.
![]() |
Fig. 13. Viewing angle posterior probability densities for Fermi/GBM SGRBs with known redshifts. For each SGRB with known redshift in our sample, we show the posterior probability density dP/dlnθv, i = θv, iP(θv, i | d) (see Eq. (43)) constructed using 300 population posterior samples from the full sample analysis (solid lines) or the flux-limited sample analysis (dashed lines). Tick marks indicate the 5th, 50th and 95th percentiles of each posterior probability. Individual SGRBs are assigned the same colours as in Fig. 4. |
4.3. The impact of selection effects on the L–Ep plane
Our result, shown in Fig. 7, that the bulk of the SGRB population is located upwards of the apparent SGRB Yonetoku correlation came with a surprise to us. In order to demonstrate the role played by selection effects in this context, we show with grey shades in Fig. 14 the detection efficiency that represents the selection effects acting on our rest-frame sample (Sect. 2.4.2), averaged over redshift, that is,
![]() |
Fig. 14. Visualisation of the impact of selection effects on the L–Ep plane. The grey filled contours in the figure show the Fermi/GBM SGRB detection efficiency averaged over redshift, assuming the redshift distribution to be described by our model with parameters corresponding to the median of the population posterior from the full sample analysis (values corresponding to different shades of grey are shown in the colour bar). Purple contours show the areas on the plane that contain 50%, 90%, 99%, 99.9%, 99.99%, and 99.999% of SGRBs in the Universe according to the model. The thick dashed purple line shows the |
where Pdet, GBM(L, Ep, z) is that from the full sample analysis (but the discussion remains unchanged when adopting that from the flux-limited sample analysis) and in this discussion we keep λpop fixed at the median of the full sample analysis posterior. The purple contours in the figure contain 50%, 90%, 99%, 99.9%, 99.99% and 99.999% of SGRBs in the Universe according to our model. For reference we also show the relation with a thick, purple dashed line. This ‘intrinsic’ distribution of SGRBs in the L–Ep plane is distorted by selection effects into the distribution represented by cyan contours, which contain 50%, 90%, 99%, and 99.9% of the SGRBs that pass the rest-frame sample selection criteria according to the model. The distribution represented by cyan contours can be understood as the product of the intrinsic (purple) distribution times the redshift averaged detection efficiency (grey filled contours). As in previous figures, red crosses mark the positions of SGRBs with measured z in our rest-frame sample, while the blue cross marks GRB 170817A. Focussing on SGRBs with L > 1050 erg s−1, these reside in a part of the plane where the ⟨Pdet⟩z contours are roughly parallel and equally spaced. In other words, the gradient
is roughly constant across the region occupied by the observed SGRBs in the plane, with a GBM detection efficiency that decreases by almost four orders of magnitude along the direction of this gradient. The variation in the density of the SGRBs in the rest-frame sample along the same direction, on the other hand, is clearly much less than four orders of magnitude. This suggests that the intrinsic density of SGRBs in this plane must increase steeply in the direction opposite to the ⟨Pdet⟩z gradient, in order for the variation in the intrinsic density of SGRBs to compensate the dramatic decrease in the detection efficiency. Hence, selection effects likely play a major role in shaping the observed L–Ep correlation in SGRBs. This conclusion and also, intriguingly, the slope and dispersion of the intrinsic L–Ep correlation we obtained, are in agreement with those found by Palmerio & Daigne (2021) for LGRBs (see Sect. 5.1 of that paper). This might be taken as an indication of a common universal luminosity and Ep angular profile in the two populations.
On a different note, it is worth stressing that the existence of a tail of SGRBs with low luminosities L ≲ 1050 erg s−1 but very high SED peak photon energies Ep ≳ 3 MeV, predicted by our population model and visible in the figure, must be taken with a grain of salt. Our parametrisation is constructed in such a way that the dispersion of Ep around the viewing-angle-dependent average is symmetric and identical at all viewing angles, so that the low-L, high-Ep tail is merely a result of the choice of parametrisation, being unobservable with the current instrumentation and therefore not observationally constrained.
4.4. Joint SGRB + GW detection predictions
We can produce predictions for the rate of coincident detections of SGRBs and GWs by Fermi/GBM and the ground-based GW detector network from the population model, leveraging the fact that it accounts for the jet luminosity as a function of the viewing angle. We focus here on a network consisting of the two aLIGO detectors (Hanford and Livingston) plus the Virgo detector (an ‘HLV’ network), and assume the projected sensitivities in the upcoming O4 observing run9. We assumed all SGRBs to be produced by BNS mergers with non-spinning components and the jets to be aligned with the orbital angular momentum. In practice, we constructed an HLV O4 GW detection efficiency Pdet, GW(z, θv) as a function of redshift and viewing angle as follows: we binned the simulated BNS mergers from Colombo et al. (2022; selecting only those that produce a jet according to their criteria, which are ∼50% in their population) in the (θv, z) plane and we computed the fraction fGW, i, j with a network S/N ρnet ≥ 12 in each bin (we assumed 100% duty cycle for all detectors for simplicity). We then estimated Pdet, GW(z, θv) by linearly interpolating the fGW, i, j’s on a grid with nodes corresponding to the centres of the bins. Samples of the cumulative joint detection rate were then computed based on 100 random population posterior samples as
and similarly
Figure 15 shows the median and symmetric 90% credible range over the population posterior samples of ṄGBM+GW(<z) (left-hand panels) and ṄGBM+GW(>θv) (right-hand panels) using posterior samples from the full sample analysis (top panels) and flux-limited sample analysis (bottom panels). The plots also show the same result for the Fermi/GBM detection only (i.e. setting pdet GW = 1) for comparison. All rates are normalised to the Fermi/GBM observed SGRB detection rate of 40 yr−1 and hence include the limited field of view and duty cycle of the instrument. The total observed SGRB + GW rates are for the full sample analysis and
for the flux-limited sample analysis (median and 90% credible range). These predictions are in agreement with, but slightly more optimistic than, the rate
estimated by Colombo et al. (2022), and are generally in line with other recent predictions from the literature (e.g. Mogushi et al. 2019; Howell et al. 2019; Saleem 2020; Yu et al. 2021; Patricelli et al. 2022), including those from the LIGO-Virgo Collaboration (Abbott et al. 2022).
![]() |
Fig. 15. Projected joint SGRB + GW rates for O4. The top (respectively bottom) panels refer to the results from the full (respectively flux-limited) sample analysis. Left-hand panels show the predicted rate Ṅobs(<z) of SGRBs within a redshift z detected by Fermi/GBM alone (blue: median; cyan: 90% credible range) or in coincidence with a network consisting of the two LIGO and the Virgo detectors with the projected O4 sensitivity (red: median; orange: 90% credible range). The right-hand panels show the corresponding rates Ṅobs(>θv) for events seen at a viewing angle larger than θv. |
4.5. Effect of decreasing the lower bound of the prior on
As stated in Sect. 2.6, we chose the lower bound of the prior on the typical on-axis luminosity to be consistent with our assumption that the high end of the luminosity function is shaped by on-axis events, while intermediate and low luminosities are due to off-axis events and depend on the apparent structure. In order to investigate the effect of relaxing that assumption, we ran the full sample analysis one additional time with a much looser bound
≥ 1048 erg s−1. As expected (given the fact that the marginalised posterior on
rails against the lower bound in the results described above), this results in a posterior that shows a mild preference for
∼ 1050 erg s−1, but with significant posterior support all the way down to 1048 erg s−1 and up to 1052 erg s−1, showing that the typical on-axis luminosity is poorly constrained by the available data within the proposed model. On the other hand, the lower
, the larger the local rate R0 needed to reproduce the GBM observed SGRB rate: if we additionally require R0 ≤ 1700 Gpc−3 yr−1, to reflect the upper limit on the BNS merger rate from the GW population analysis (Abbott et al. 2023), then the posterior becomes again consistent with that obtained with the original prior. We thus conclude that our prior, in addition to being a consequence of the assumed quasi-universal jet scenario, has a similar impact as the requirement that the SGRB local rate does not exceed the BNS merger rate. This suggests that future improved constraints on this rate through GW observations will positively impact our ability to constrain the SGRB population properties, including their typical apparent jet structure.
4.6. Difficulties in defining a ‘clean’ sample of GRBs from compact binary mergers
When selecting our sample, we focused on events with T90 shorter than 2 seconds for simplicity, and for the practical reason that the spectral parameters (photon index, observed Ep, obs, peak photon flux from the spectral analysis) in the Fermi/GBM online catalogue are given with 64 ms binning only for events with T90 < 2 s. On the other hand, this selection may result in a sample that includes events from both the main progenitor classes, that is, compact binary mergers and collapsars (e.g. Zhang et al. 2009; Bromberg et al. 2013), with the contamination from the latter class being difficult to quantity. Even if we did not test explicitly for the dependence of our results on this choice, the fact that the high-end of the luminosity function we obtain agrees relatively well with that obtained by W15 (who adopted a much more restrictive criterion in an attempt to select a sample of pure ‘non-collapsar’ GRBs) lends support to the conclusion that any potential contamination from collapsars in our sample does not impact the results significantly, given the present uncertainties.
Conversely, some GRBs with T90 > 2 s are now widely accepted as being the result of a compact binary merger rather than a collapsar (e.g. GRB211211A, Rastinejad et al. 2022; Mei et al. 2022; Gompertz et al. 2023). Hence, the definition of a clean sample of GRBs from compact binary mergers is not straightforward. We believe that the best approach to this kind of problem in the future will be that of modelling the entire GRB population as a mixture of the two classes, jointly fitting the two sub-populations in a hierarchical model, similarly to what is currently done for binary black hole merger GW population analyses (e.g. Bouffanais et al. 2019; Wong et al. 2021; Zevin et al. 2021).
As a final note, we caution that the duration of an SGRB must eventually increase with the viewing angle: at large enough viewing angles, the observed duration of a single pulse can exceed the duration of the central engine activity (e.g. Salafia et al. 2016). Therefore, our sample selection could contain a bias against events with a large viewing angle. In order to address this problem, the duration of the emission, and its dependence on the viewing angle, must be included in the model as well.
4.7. Peak luminosity versus average luminosity
Population studies of GRBs, whose luminosity varies erratically without a clear-cut minimum variability timescale during the prompt emission, must confront the issue of defining the relevant luminosity whose distribution is to be modelled. Given the stochastic nature of the light curves, the time-averaged luminosity is arguably the most relevant quantity from a physical point of view; on the other hand, the detectability (and hence the selection effects) depends more closely on the peak luminosity, so that from a practical point of view this is the most important quantity to be modelled, which has therefore become the standard in the field. To our knowledge, the relation between the two distributions, that is, the peak luminosity function ϕ(L) = dP/dlnL and the average luminosity function Φ(⟨L⟩) = dP/dln⟨L⟩, has never been investigated. Quite clearly, since L = ξ−1⟨L⟩ and the factor ξ−1 ≥ 1 can differ between two SGRBs with the same ⟨L⟩ due to the stochasticity in the light curve, then the L distribution is necessarily broader than the ⟨L⟩ one. In that sense, the ‘core luminosity dispersion’ in our model, parametrised as in Eq. (2), accounts at least partially for the dispersion due to these effects. At large viewing angles, on the other hand, the broadening of individual pulses is likely going to smooth out the light curve (Salafia et al. 2016), hence reducing this kind of scatter. Hence, another improvement over our approach could be that of including a viewing-angle-dependent scatter in L (and similarly in Ep), which may improve the recovery of the actual average luminosity and peak photon energy profiles ℓ and η.
5. Summary and conclusions
The observations of GW170817 and the GRB170817A afterglow provided clear support for the presence of a jet viewed off-axis and endowed with a non-trivial angular structure. The inferred properties of the jet’s core were found to be consistent with those typically derived from the afterglows of SGRBs. This prompted the question of whether jets underlying SGRBs could be very similar to each other on average, with a large part of the diversity due to the geometric choice of a viewing angle rather than intrinsic variations in jet structure and energetics. In this work we have shown that a good description of the observed SGRB population can be obtained within such a scenario. The implied typical jet properties are consistent with those inferred from the GRB 170817A afterglow and from the larger population of SGRBs with a known distance, adding stability to the foundations of a unification programme for SGRBs under the quasi-universal jet scenario, and more generally to our physical understanding of these phenomena.
The inferred jet structure features a ∼2 deg uniform core within which the observer sees a large typical SED peak photon energy (Ep ∼ 5 MeV) and luminosity (L ≳ 5 × 1051 erg s−1). Outside the core, the luminosity falls off with the viewing angle as a steep power law (slope αL ∼ 4.7), while Ep decreases as a relatively shallow power law (αEp ∼ 2). No evidence for a break in these power laws has been found with the present data and analysis approach. While we find no clear support for a correlation between the on-axis luminosity, Lc, and the on-axis peak SED photon energy, Ep, c, the combined viewing angle dependence of L and Ep induces a correlation for events viewed outside the core, . In the observed sample, we find that this correlation is distorted by selection effects.
The inferred local rate density of SGRBs (at all viewing angles) is compatible with that of BNS mergers as inferred from GW population studies, suggesting they are the dominant progenitors. The model shows a preference for a strong rate density evolution with redshift: the rate density steeply increases as at low redshifts, plateaus towards a maximum near z ∼ 2.2, and declines at higher redshifts, where it is poorly constrained. These results, on the other hand, may be driven by the rather large redshifts of three out of four SGRBs with a photometric redshift in our rest-frame sample: if photometric redshifts are excluded from the analysis, the redshift evolution becomes consistent with that of the CSFR, and hence with progenitor binaries that merge rapidly after formation. Based on the model and on the projected sensitivity of the aLIGO and Advanced Virgo network, we predict around 0.2 to 1.3 joint SGRB and GW detections per year during the O4 observing run.
Through the population model, it is possible to derive a population-informed viewing angle estimate for every SGRB whose intrinsic luminosity and peak photon energy are reasonably constrained. The estimates obtained for SGRBs with a known redshift in our sample indicate that most of them are viewed close to the edge of the core (either just within the core or slightly outside it), with a few exceptions with a somewhat larger viewing angle. The largest viewing angle is clearly that of GRB 170817A, for which we estimate , in excellent agreement with the estimates based on the afterglow and the superluminal motion seen in Very Long Baseline Interferometry observations.
A unification of SGRBs under a quasi-universal jet scenario would call for a relatively narrow progenitor parameter space, which can eventually help in pinpointing the long-debated jet-launching mechanism and the nature of the central engine. As demonstrated by the amount of information contained in the single GW170817 event, future multi-messenger observations of BNS mergers and their jets will be of the utmost importance to the success of this programme.
In X-ray and γ-ray astronomy, detectors typically identify ‘events’ (interactions between photons and the active part of the detector) in different energy channels. The hardness ratio is generally defined as the ratio of the event counts in a higher-energy (‘harder’) channel (or group of channels) to that in a lower-energy (‘softer’) channel.
The name follows from the apparent correlation between log(L) and log(Ep) in LGRBs originally found by Yonetoku et al. (2004).
While Fermi/GBM is sensitive over a larger energy band, and the results in the catalogue usually refer to the 10–1000 keV band, the 50–300 keV band is where most of the online GRB trigger algorithms look for excess (von Kienlin et al. 2020), so that the flux in that band is the most relevant for what concerns the modelling of the GBM detection – see Appendix C.
The python code used to produce the results and the main figures in this work is open source and publicly available at https://github.com/omsharansalafia/grbpop.
The population parameter posterior samples are stored at https://zenodo.org/record/8160783.
https://dcc.ligo.org/T2200043-v3/public. We conservatively did not consider KAGRA (Somiya 2012) due to its much lower expected sensitivity.
The data were obtained from https://heasarc.gsfc.nasa.gov/W3Browse/fermi/fermigtrig.html, where 527 trigdata files for bursts with T90 < 2 s were available. Some of the events had apparently incorrect background estimates, which were much larger than the actual counts recorded, resulting in negative background-subtracted counts at peak. Others had corrupted trigdata files. After removing these problematic events, we were left with 449 valid SGRBs.
Acknowledgments
The authors acknowledge Paolo D’Avanzo for support in building the S-BAT4ext sub-sample used in this work, and Ruben Salvaterra for insightful comments that helped deepening our understanding of some of the results. O.S.S. thanks Riccardo Buscicchio for many amusing and illuminating discussions. M.E.R. acknowledges support from the research programme Athena with project number 184.034.002, which is financed by the Dutch Research Council (NWO). G.G. acknowledges the Italian Ministry for University and Research for support through the ‘premiale’ grant 1.05.06.13, and the Agenzia Spaziale Italiana for support through the ASI-NuStar grant 1.05.04.95. I.M. is a recipient of the Australian Research Council Future Fellowship FT190100574. The color set used to identify SGRBs in the rest-frame sample is from https://colorbrewer2.org (Harrower & Brewer 2003).
References
- Aasi, J., Abbott, B. P., Abbott, R., Abbott, T., et al. 2015, CQG, 32, 074001 [Google Scholar]
- Abbott, B. P., Abbott, R., Abbott, T. D., et al. 2017a, ApJ, 848, L13 [CrossRef] [Google Scholar]
- Abbott, B. P., Abbott, R., Abbott, T. D., et al. 2017b, ApJ, 848, L12 [Google Scholar]
- Abbott, B. P., et al. (LIGO Scientific Collaboration& Virgo Collaboration) 2019, Phys. Rev. X, 9, 011001 [NASA ADS] [Google Scholar]
- Abbott, R., Abbott, T. D., Acernese, F., et al. 2022, ApJ, 928, 186 [NASA ADS] [CrossRef] [Google Scholar]
- Abbott, B. P., et al. (LIGO Scientific Collaboration, Virgo Collaboration,& KAGRA Collaboration) 2023, Phys. Rev. X, 13, 011048 [NASA ADS] [Google Scholar]
- Acernese, F., Agathos, M., Agatsuma, K., Aisa, D., et al. 2014, CQG, 32, 024001 [Google Scholar]
- Beniamini, P., Nava, L., & Piran, T. 2016, MNRAS, 461, 51 [NASA ADS] [CrossRef] [Google Scholar]
- Beniamini, P., Petropoulou, M., Barniol Duran, R., & Giannios, D. 2019, MNRAS, 483, 840 [NASA ADS] [CrossRef] [Google Scholar]
- Beniamini, P., Gill, R., & Granot, J. 2022, MNRAS, 515, 555 [NASA ADS] [CrossRef] [Google Scholar]
- Berger, E. 2014, ARA&A, 52, 43 [CrossRef] [Google Scholar]
- Biscoveanu, S., Thrane, E., & Vitale, S. 2020, ApJ, 893, 38 [NASA ADS] [CrossRef] [Google Scholar]
- Bloom, J. S., Kulkarni, S. R., Price, P. A., et al. 2002, ApJ, 572, L45 [NASA ADS] [CrossRef] [Google Scholar]
- Bouffanais, Y., Mapelli, M., Gerosa, D., et al. 2019, ApJ, 886, 25 [NASA ADS] [CrossRef] [Google Scholar]
- Bromberg, O., Nakar, E., Piran, T., & Sari, R. 2013, ApJ, 764, 179 [NASA ADS] [CrossRef] [Google Scholar]
- Bryant, C. M., Osborne, J. A., & Shahmoradi, A. 2021, MNRAS, 504, 4192 [NASA ADS] [CrossRef] [Google Scholar]
- Burns, E., Connaughton, V., Zhang, B.-B., et al. 2016, ApJ, 818, 110 [NASA ADS] [CrossRef] [Google Scholar]
- Calderone, G., Ghirlanda, G., Ghisellini, G., et al. 2015, MNRAS, 448, 403 [NASA ADS] [CrossRef] [Google Scholar]
- Cano, Z., de Ugarte Postigo, A., Pozanenko, A., et al. 2014, A&A, 568, A19 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Cantiello, M., Jensen, J. B., Blakeslee, J. P., et al. 2018, ApJ, 854, L31 [Google Scholar]
- Colombo, A., Salafia, O. S., Gabrielli, F., et al. 2022, ApJ, 937, 79 [NASA ADS] [CrossRef] [Google Scholar]
- Costa, E., Frontera, F., Heise, J., et al. 1997, Nature, 387, 783 [NASA ADS] [CrossRef] [Google Scholar]
- Coulter, D. A., Foley, R. J., Kilpatrick, C. D., et al. 2017, Science, 358, 1556 [NASA ADS] [CrossRef] [Google Scholar]
- D’Avanzo, P. 2015, J. High Energy Astrophys., 7, 73 [CrossRef] [Google Scholar]
- D’Avanzo, P., Salvaterra, R., Bernardini, M. G., et al. 2014, MNRAS, 442, 2342 [Google Scholar]
- D’Elia, V., Campana, S., D’Aì, A., et al. 2018, A&A, 619, A66 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- de Ugarte Postigo, A., Kann, D. A., Izzo, L., et al. 2020, GRB Coordinates Network, 29132, 1 [NASA ADS] [Google Scholar]
- Eichler, D., Livio, M., Piran, T., & Schramm, D. N. 1989, Nature, 340, 126 [NASA ADS] [CrossRef] [Google Scholar]
- Farah, A., Essick, R., Doctor, Z., Fishbach, M., & Holz, D. E. 2020, ApJ, 895, 108 [NASA ADS] [CrossRef] [Google Scholar]
- Fernández, J. J., Kobayashi, S., & Lamb, G. P. 2022, MNRAS, 509, 395 [Google Scholar]
- Ferro, M., Brivio, R., D’Avanzo, P., et al. 2023, A&A, 678, A142 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Fishbach, M., Holz, D. E., & Farr, W. M. 2018, ApJ, 863, L41 [NASA ADS] [CrossRef] [Google Scholar]
- Fong, W., & Berger, E. 2013, ApJ, 776, 18 [NASA ADS] [CrossRef] [Google Scholar]
- Fong, W., Berger, E., Chornock, R., et al. 2013, ApJ, 769, 56 [NASA ADS] [CrossRef] [Google Scholar]
- Fong, W., Berger, E., Margutti, R., & Zauderer, B. A. 2015, ApJ, 815, 102 [NASA ADS] [CrossRef] [Google Scholar]
- Fong, W.-F., Nugent, A. E., Dong, Y., et al. 2022, ApJ, 940, 56 [NASA ADS] [CrossRef] [Google Scholar]
- Foreman-Mackey, D., Hogg, D. W., Lang, D., & Goodman, J. 2013, PASP, 125, 306 [Google Scholar]
- Galama, T. J., Vreeswijk, P. M., van Paradijs, J., et al. 1998, Nature, 395, 670 [Google Scholar]
- Gehrels, N., Chincarini, G., Giommi, P., et al. 2004, ApJ, 611, 1005 [Google Scholar]
- Ghirlanda, G., Ghisellini, G., & Celotti, A. 2004, A&A, 422, L55 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Ghirlanda, G., Nava, L., Ghisellini, G., Celotti, A., & Firmani, C. 2009, A&A, 496, 585 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Ghirlanda, G., Salafia, O. S., Pescalli, A., et al. 2016, A&A, 594, A84 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Ghirlanda, G., Salafia, O. S., Paragi, Z., et al. 2019, Science, 363, 968 [NASA ADS] [CrossRef] [Google Scholar]
- Goldstein, A., Veres, P., Burns, E., et al. 2017, ApJ, 848, L14 [CrossRef] [Google Scholar]
- Gompertz, B. P., Ravasio, M. E., Nicholl, M., et al. 2023, Nat. Astron., 7, 67 [Google Scholar]
- Goodman, J., & Weare, J. 2010, Commun. Appl. Math. Comput. Sci., 5, 65 [Google Scholar]
- Gottlieb, O., Nakar, E., Piran, T., & Hotokezaka, K. 2018, MNRAS, 479, 588 [NASA ADS] [Google Scholar]
- Gottlieb, O., Bromberg, O., Singh, C. B., & Nakar, E. 2020, MNRAS, 498, 3320 [NASA ADS] [CrossRef] [Google Scholar]
- Gottlieb, O., Nakar, E., & Bromberg, O. 2021, MNRAS, 500, 3511 [Google Scholar]
- Gottlieb, O., Moseley, S., Ramirez-Aguilar, T., et al. 2022, ApJ, 933, L2 [NASA ADS] [CrossRef] [Google Scholar]
- Govreen-Segal, T., & Nakar, E. 2023, MNRAS, 524, 403 [NASA ADS] [CrossRef] [Google Scholar]
- Guetta, D., & Piran, T. 2005, A&A, 435, 421 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Guetta, D., & Piran, T. 2006, A&A, 453, 823 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Harrower, M., & Brewer, C. A. 2003, Cartographic J., 40, 27 [NASA ADS] [CrossRef] [Google Scholar]
- Hayes, F., Heng, I. S., Veitch, J., & Williams, D. 2020, ApJ, 891, 124 [NASA ADS] [CrossRef] [Google Scholar]
- Hayes, F., Heng, I. S., Lamb, G., et al. 2023, ApJ, 954, 92 [NASA ADS] [CrossRef] [Google Scholar]
- Hjorth, J., Levan, A. J., Tanvir, N. R., et al. 2017, ApJ, 848, L31 [NASA ADS] [CrossRef] [Google Scholar]
- Hotokezaka, K., Nakar, E., Gottlieb, O., et al. 2019, Nat. Astron., 3, 940 [NASA ADS] [CrossRef] [Google Scholar]
- Howell, E. J., Ackley, K., Rowlinson, A., & Coward, D. 2019, MNRAS, 485, 1435 [NASA ADS] [CrossRef] [Google Scholar]
- Hu, Y. D., Castro-Tirado, A. J., Kumar, A., et al. 2021, A&A, 646, A50 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Kann, D. A., Klose, S., Zhang, B., et al. 2011, ApJ, 734, 96 [NASA ADS] [CrossRef] [Google Scholar]
- Kouveliotou, C., Meegan, C. A., Fishman, G. J., et al. 1993, ApJ, 413, L101 [NASA ADS] [CrossRef] [Google Scholar]
- Kumar, P., & Zhang, B. 2015, Phys. Rep., 561, 1 [Google Scholar]
- Lamb, G. P., Mandel, I., & Resmi, L. 2018, MNRAS, 481, 2581 [CrossRef] [Google Scholar]
- Lamb, G. P., Lyman, J. D., Levan, A. J., et al. 2019, ApJ, 870, L15 [NASA ADS] [CrossRef] [Google Scholar]
- Levan, A. J., Fynbo, J. P. U., Hjorth, J., et al. 2009, GRB Coordinates Network, 9958, 1 [NASA ADS] [Google Scholar]
- LIGO Scientific Collaboration, Virgo Collaboration, KAGRA Collaboration, et al. 2021, ArXiv e-prints [arXiv:2111.03606] [Google Scholar]
- LIGO Scientific Collaboration, Virgo Collaboration,& KAGRA Collaboration 2023, https://doi.org/10.5281/zenodo.7890437 [Google Scholar]
- Lipunov, V. M., Postnov, K. A., & Prokhorov, M. E. 2001, Astron. Rep., 45, 236 [NASA ADS] [CrossRef] [Google Scholar]
- Madau, P., & Dickinson, M. 2014, ARA&A, 52, 415 [Google Scholar]
- Malesani, D., Tagliaferri, G., Chincarini, G., et al. 2004, ApJ, 609, L5 [Google Scholar]
- Mandel, I. 2018, ApJ, 853, L12 [NASA ADS] [CrossRef] [Google Scholar]
- Mandel, I., & Broekgaarden, F. S. 2022, Liv. Rev. Relativ., 25, 1 [NASA ADS] [CrossRef] [Google Scholar]
- Mandel, I., Farr, W. M., & Gair, J. R. 2019, MNRAS, 486, 1086 [Google Scholar]
- Meegan, C., Lichti, G., Bhat, P. N., et al. 2009, ApJ, 702, 791 [Google Scholar]
- Mei, A., Banerjee, B., Oganesyan, G., et al. 2022, Nature, 612, 236 [NASA ADS] [CrossRef] [Google Scholar]
- Mirabal, N., Halpern, J. P., An, D., Thorstensen, J. R., & Terndrup, D. M. 2006, ApJ, 643, L99 [NASA ADS] [CrossRef] [Google Scholar]
- Mochkovitch, R., Hernanz, M., Isern, J., & Martin, X. 1993, Nature, 361, 236 [NASA ADS] [CrossRef] [Google Scholar]
- Mogushi, K., Cavaglià, M., & Siellez, K. 2019, ApJ, 880, 55 [NASA ADS] [CrossRef] [Google Scholar]
- Mooley, K. P., Deller, A. T., Gottlieb, O., et al. 2018, Nature, 561, 355 [Google Scholar]
- Mooley, K. P., Anderson, J., & Lu, W. 2022, Nature, 610, 273 [NASA ADS] [CrossRef] [Google Scholar]
- Nakar, E. 2007, Phys. Rep., 442, 166 [NASA ADS] [CrossRef] [Google Scholar]
- Nugent, A. E., Fong, W.-F., Dong, Y., et al. 2022, ApJ, 940, 57 [NASA ADS] [CrossRef] [Google Scholar]
- Palmerio, J. T., & Daigne, F. 2021, A&A, 649, A166 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Patricelli, B., Bernardini, M. G., Mapelli, M., et al. 2022, MNRAS, 513, 4159 [NASA ADS] [CrossRef] [Google Scholar]
- Paul, D. 2018, MNRAS, 477, 4275 [NASA ADS] [CrossRef] [Google Scholar]
- Pescalli, A., Ghirlanda, G., Salafia, O. S., et al. 2015, MNRAS, 447, 1911 [NASA ADS] [CrossRef] [Google Scholar]
- Planck Collaboration XIII. 2016, A&A, 594, A13 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Rastinejad, J. C., Gompertz, B. P., Levan, A. J., et al. 2022, Nature, 612, 223 [NASA ADS] [CrossRef] [Google Scholar]
- Rau, A., McBreen, S., & Kruehler, T. 2009, GRB Coordinates Network, 9353, 1 [NASA ADS] [Google Scholar]
- Rossi, E., Lazzati, D., & Rees, M. J. 2002, MNRAS, 332, 945 [NASA ADS] [CrossRef] [Google Scholar]
- Rouco Escorial, A., Fong, W.-F., Berger, E., et al. 2022, ArXiv e-prints [arXiv:2210.05695] [Google Scholar]
- Salafia, O. S., & Ghirlanda, G. 2022, Galaxies, 10, 93 [NASA ADS] [CrossRef] [Google Scholar]
- Salafia, O. S., Ghisellini, G., Pescalli, A., Ghirlanda, G., & Nappo, F. 2015, MNRAS, 450, 3549 [NASA ADS] [CrossRef] [Google Scholar]
- Salafia, O. S., Ghisellini, G., Pescalli, A., Ghirlanda, G., & Nappo, F. 2016, MNRAS, 461, 3607 [NASA ADS] [CrossRef] [Google Scholar]
- Salafia, O. S., Barbieri, C., Ascenzi, S., & Toffano, M. 2020, A&A, 636, A105 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Salafia, O. S., Colombo, A., Gabrielli, F., & Mandel, I. 2022, A&A, 666, A174 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Saleem, M. 2020, MNRAS, 493, 1633 [NASA ADS] [CrossRef] [Google Scholar]
- Salvaterra, R., Campana, S., Vergani, S. D., et al. 2012, ApJ, 749, 68 [NASA ADS] [CrossRef] [Google Scholar]
- Sarin, N., Lasky, P. D., Vivanco, F. H., et al. 2022, Phys. Rev. D, 105, 083004 [NASA ADS] [CrossRef] [Google Scholar]
- Savchenko, V., Ferrigno, C., Kuulkers, E., et al. 2017, ApJ, 848, L15 [NASA ADS] [CrossRef] [Google Scholar]
- Schmidt, M. 2001, ApJ, 559, L79 [NASA ADS] [CrossRef] [Google Scholar]
- Shahmoradi, A., & Nemiroff, R. J. 2015, MNRAS, 451, 126 [NASA ADS] [CrossRef] [Google Scholar]
- Somiya, K. 2012, CQG, 29, 124007 [NASA ADS] [CrossRef] [Google Scholar]
- Takahashi, K., & Ioka, K. 2020, MNRAS, 497, 1217 [NASA ADS] [CrossRef] [Google Scholar]
- Takahashi, K., & Ioka, K. 2021, MNRAS, 501, 5746 [NASA ADS] [Google Scholar]
- Tan, W.-W., & Yu, Y.-W. 2020, ApJ, 902, 83 [NASA ADS] [CrossRef] [Google Scholar]
- Troja, E., Castro-Tirado, A. J., Becerra González, J., et al. 2019, MNRAS, 489, 2104 [Google Scholar]
- Virgili, F. J., Zhang, B., O’Brien, P., & Troja, E. 2011, ApJ, 727, 109 [NASA ADS] [CrossRef] [Google Scholar]
- von Kienlin, A., Meegan, C. A., Paciesas, W. S., et al. 2020, ApJ, 893, 46 [Google Scholar]
- Wanderman, D., & Piran, T. 2015, MNRAS, 448, 3026 [NASA ADS] [CrossRef] [Google Scholar]
- Williams, D., Clark, J. A., Williamson, A. R., & Heng, I. S. 2018, ApJ, 858, 79 [NASA ADS] [CrossRef] [Google Scholar]
- Wong, K. W. K., Breivik, K., Kremer, K., & Callister, T. 2021, Phys. Rev. D, 103, 083021 [Google Scholar]
- Woods, E., & Loeb, A. 1999, ApJ, 523, 187 [NASA ADS] [CrossRef] [Google Scholar]
- Woosley, S. E. 1993, ApJ, 405, 273 [Google Scholar]
- Yonetoku, D., Murakami, T., Nakamura, T., et al. 2004, ApJ, 609, 935 [Google Scholar]
- Yonetoku, D., Nakamura, T., Sawano, T., Takahashi, K., & Toyanago, A. 2014, ApJ, 789, 65 [NASA ADS] [CrossRef] [Google Scholar]
- Yu, J., Song, H., Ai, S., et al. 2021, ApJ, 916, 54 [NASA ADS] [CrossRef] [Google Scholar]
- Zevin, M., Bavera, S. S., Berry, C. P. L., et al. 2021, ApJ, 910, 152 [Google Scholar]
- Zhang, G. Q., & Wang, F. Y. 2018, ApJ, 852, 1 [Google Scholar]
- Zhang, B., Zhang, B.-B., Virgili, F. J., et al. 2009, ApJ, 703, 1696 [NASA ADS] [CrossRef] [Google Scholar]
Appendix A: Additional details on the results
A.1. Corner plot
Figure A.1 shows the full, 14-parameter corner plot of both the full sample analysis (magenta) and the flux-limited sample analysis (light blue).
![]() |
Fig. A.1. Corner plot of the posterior probability densities from the two analyses. The full sample analysis is shown in magenta, while the flux-limited sample analysis is shown in blue. The histograms on the diagonal represent the marginalised posterior probability densities constructed from the posterior samples, with the solid vertical lines marking the medians and the vertical dashed lines delimiting the symmetric 90% credible interval (i.e. the 5th and the 95th percentiles – note that these are not shown if they differ by less than one bin size from the nearest edge of the allowed range). The contours in the remaining panels show the one, two, three and four sigma credible areas from the two-dimensional marginalised joint posterior probability densities, with the dots showing the intersections of the medians of the corresponding one-dimensional marginalised posterior probability densities. |
Appendix B: Spectral analysis of Fermi/GBM SGRBs with known redshifts
For each Fermi/GBM SGRB with a known redshift in our sample, we analysed the spectrum at the peak flux of the light curve binned with a 64 ms timescale. As starting time of the interval selected for the spectral analysis, we used the one reported in the GBM Catalogue (referred to as ’Flux_64_Time’10).
The spectral data files and corresponding latest response matrix files (rsp2) were obtained from the online High Energy Astrophysics Science Archive Research Center (HEASARC) archive. We used the public software GTBURST to extract the spectral data. As part of the standard procedure, we selected the spectral data of the two most illuminated NaI detectors with a viewing angle smaller than 60° and the most illuminated BGO detector. In particular, we considered the energy channels in the range 10–900 keV for the NaI detectors, and 0.3–40 MeV for the BGO detector. We used intercalibration factors among the detectors, scaled to the most illuminated NaI and free to vary within 30%. To model the background, we manually selected time intervals before and after the burst and modelled them with a polynomial function whose best-fitting order is automatically found by GTBURST.
The spectral analysis has been performed with the public software XSPEC (v. 12.12.1). In the fitting procedure, we used the PG-Statistic, valid for Poisson data with a Gaussian background. Each peak flux spectrum was analysed with a cutoff power law model, typically used for the analysis of SGRBs spectra, and it consists of three parameters: the low-energy photon index α, the characteristic energy Ecut (from which the peak energy of the spectrum can be derived as Ep = Ecut(2 − α)) and the normalisation. To obtain the luminosity (and its uncertainties) directly from the fit, the cutoff power law model is multiplied by the CLUMIN function available in XSPEC11, which computes the luminosity of a specific model component, provided the redshift value of the source. The source-frame energy band over which luminosity is calculated is 1 keV-10 MeV. Since the CLUMIN model is used, the normalisation of the cutoff power law model has been kept fixed to 1 for all the spectra analysed. The parameters left free to vary in each fit are the following: the luminosity in the 1 keV-10 MeV energy range, the low-energy photon index α and the characteristic energy Ecut, alongside the two inter-calibration factors for the GBM detectors. Best-fit values and confidence ranges on these parameters have been derived within XSPEC, through the built-in Markov chain Monte Carlo algorithm (using the CHAIN command). The 50% and 90% contours for Lpeak, iso and Ep derived from the spectral analysis of each SGRB analysed in this work are reported in Fig. 4.
Appendix C: Fermi/GBM short gamma-ray burst detection efficiency
In order to construct a reliable estimate of the SGRB detection efficiency of Fermi/GBM, a detailed simulation of its response to an event of that class is needed. The onboard trigger algorithms of GBM, described for example in von Kienlin et al. 2020 (vK20 hereafter), monitor the counts (binned over a given timescale Δt) recorded in a subset of the 8 energy channels of the NaI detectors12. Each algorithm looks for an excess in the background-subtracted, binned counts CΔt(t)−BΔt(t) (where Bt is an estimate of the counts due to the background), over a certain multiple nσ of the standard deviation σbkg, Δt(t) of the counts recorded in a time interval immediately preceding t. Multiple algorithms operating on the same channels and with the same binning timescale are run in parallel, differing from each other only by small time offsets, which improves the triggering efficiency by minimising cases with sub-optimally placed bins. A burst trigger is initiated whenever an excess is recorded at a consistent time in at least two of the NaI detectors by any of the running algorithms.
Since a sufficient condition for a trigger is that the peak counts exceed the threshold, it is enough to simulate the peak counts generated by a putative source. On the other hand, since different algorithms operate with different binning timescales Δt, and since SGRBs are highly variable, it is necessary that the peak counts generated by the same source, but measured with a different Δt, reflect the actual peak count ratios induced by the temporal behaviour of the events of that class. The simulated counts must then be compared with realistic background counts, and this must be done in each of the NaI detectors and consistently for each of the triggering algorithms. In what follows, we describe our approach to these problems.
C.1. NaI background
In order to simulate a realistic background in each of the channels of the NaI detectors, we extracted the observed count rates from publicly available GBM daily data13, at a randomly sampled time, and multiplied them by Δt to get the counts. In particular, we considered ctime data relative to two selected days without triggers, 210530 and 220928, and extracted the count rates of the 12 NaI detectors in the eight energy channels at random times, excluding periods when the detectors were turned off. The resulting probability density distributions of count rates in the 50-300 keV band (channels 3-4, combined from all 12 detectors) is shown in Fig. C.1 (red histogram), along with that for the 25-50 keV band (channel 2), for comparison.
![]() |
Fig. C.1. Probability density of the NaI background count rates in the 25-50 and 50-300 keV bands, constructed by sampling GBM ctime day data from days 210530 and 220928 at random times, combined from all 12 detectors. |
C.2. Detector response to an SGRB
We assumed the photon spectrum (number of SGRB photons per unit time, per unit area, per unit photon energy as measured at the Fermi/GBM position) of every SGRB to be time-independent and well described by the same cut-off power law as used in the population modelling, namely
which corresponds to the COMP model commonly used in Fermi/GBM data analyses (vK20). We use f(E) to indicate the above spectrum normalised to unity in the 50 − 300 keV band, that is,
The expected counts in the j-th channel (j = 1, ..., 8) of the k-th NaI Fermi/GBM detector (k = 1, ..., 12), with a binning timescale Δt, produced by an SGRB with a peak photon flux p[50 − 300] (with 64 ms binning) and spectral parameters α and Ep, obs, located at a sky position (θ, ϕ) (polar and azimuthal angles with respect to the spacecraft z-axis), were computed as
where Emin, j (Emax, j) is the lower (upper) photon energy bound of channel j as given in Fig. 8 of Meegan et al. 2009 (M09 hereafter), Aeff is the NaI detector effective area at incidence angle Θk(θ, ϕ) measured with respect to the vector normal to the detector’s entrance window (see Appendix C.4), and Rpflx(Δt) is the ratio of the count rate peak flux measured with Δt binning to that measured with 64 ms binning. The incidence angle can be computed knowing the orientation of the k-th NaI detector with respect to the spacecraft z-axis (Table 1 of M09).
The ratio Rpflx(Δt) was constructed by extracting the count rate light curves in the 50-300 keV band (corresponding to channels 3 and 4, where the effective area is maximal) from the publicly available trigdata files14 of all Fermi/GBM SGRBs. For each burst, we extracted the background-subtracted count rate light curves in channels j = 3, 4, where k ∈ T means that we summed only over detectors for which an excess was found,
represents the count rate and
the estimated background count rate (as reported in the trigdata file). Figure C.2 shows the count rate light curves obtained in this way. We then computed
![]() |
Fig. C.2. Background-subtracted count rate light curves in the 50-300 keV band for SGRBs observed by Fermi/GBM. Each light curve is plotted with a colour that depends on its peak count rate, as shown in the colour bar on the right. More details are given in the text. |
where the time coordinate here is defined in such a way that t = 0 corresponds to the maximum of . The resulting ratios for all the SGRBs analysed are shown in Fig. C.3. For each simulated SGRB, we randomly picked one of the Rpflx(Δt) curves shown there.
![]() |
Fig. C.3. Ratio of SGRB count rate peak flux measured with a Δt binning over that measured with a 64 ms binning. This is constructed using Fermi/GBM trigdata files as explained in the text. Each thin red line corresponds to a single observed SGRB, while the light blue lines mark the 1st, 5th, 25th, 75th, 95th and 99th percentiles. |
For each considered detection algorithm, we computed the background-subtracted counts as , where j runs over the channels monitored by the algorithm,
represents a randomly picked matrix of NaI background counts as explained in Appendix C.1, and 𝒫(n) represents a random sample from a Poisson distribution with mean n. We then set
(which is a good approximation given the background count distributions shown in Fig. C.1) and considered a trigger in the k-th detector if (Ck, Δt − Bk, Δt) > nσσbkg, Δt, k. The simulated SGRB was marked as detected whenever at least two NaI detectors had at least one algorithm that triggered.
C.3. Detection algorithms
In our framework, each detection algorithm is characterised by a range of monitored channels, a binning timescale and a factor nσ that sets the threshold for triggering. When compared to the actual triggering algorithms running on board Fermi/GBM (vK20), we do not consider a time offset, as our definition of Rpflx(Δt) essentially corresponds to always taking the offset that maximises the peak count rates for each timescale. Hence, each of our algorithms effectively covers all the GBM algorithms with the same channels and binning timescales, but different offsets. Table C.1 reports the 6 algorithms that we implemented in our framework. These are limited to the 50-300 keV band because, by inspection of the trigdata files, we verified that essentially all the SGRBs detected by GBM triggered one of the algorithms operating in that band. We implemented all algorithms that operate on timescales Δt ≥ 64 ms, because trigdata files do not provide count rates with a finer resolution. We do not consider this as a limiting factor, since the variability of the SGRB light curves over such short timescales is hardly important to our purposes.
Trigger algorithms considered in our framework†.
C.4. NaI detector effective area
We decomposed the effective area of a NaI detector into its value for zero incidence angle Aeff(0, E), times a dimensionless function that captures its variation with the incidence angle, Aeff(Θ, E)≡Aeff(0, E)𝒜(Θ, E). We computed Aeff(0, E) by linearly interpolating the measurements reported Fig. 11 of M09 on a log-log plane, as shown in panel (a) of Fig. C.4. To model 𝒜(Θ, E) we considered the measurements reported in Fig. 12 of M09 (shown in panel (b) of Fig. C.4), relative to the three reference photon energies Eref, 1 = 32 keV, Eref, 2 = 279 keV and Eref, 3 = 662 keV, and normalised these data to the peak (which corresponds to Θ = 0 in each case). We assumed an ansatz analytical form,
![]() |
Fig. C.4. NaI detector effective area model. Panel (a) shows the NaI detector effective area measurements (red crosses), for zero incidence angle, reported in M09. The pink line shows our adopted interpolation. Panel (b) shows the measured effective area (M09) for different photon incidence angles at three reference photon energies (black circles: 32 keV; blue squares: 279 keV; red triangles: 662 keV). The solid lines show our best-fitting model. Panel (c) shows the best-fit values of parameters c1, c2 and c3 of our model (Eq. C.5) for the angular dependence of the effective area at the three reference energies (coloured circles). The dashed lines show our assumed behaviour of ci as a function of the photon energy. |
and fit it to the data at each reference photon energy by minimising the sum squares of the residuals, obtaining the values reported in Table C.2 (the best-fit models are shown by solid lines in panel (b) of Fig. C.4 – we note that c2 remains unconstrained at Eref, 1, because c3 ∼ 0). In order to extend the model to other photon energies, we then assumed a linear dependence of the parameters c1, c2 and c3 on the natural logarithm of the photon energy normalised to 32 keV, ϵ = ln(E/32 keV), namely ci = qi + miϵ, and obtained a good description of the data with q1 = 1.19, m1 = −0.308, q2 = 0.4, m2 = 0, q3 = 0.00938 and m3 = 0.173, as shown in panel (c) of Fig. C.4 (in the case of c3, since it is naturally positive definite, we took c3 = max(q3 + m3ϵ, 0)). We therefore used .
Best-fit parameter values for the GBM NaI incidence-angle-dependent effective area ansatz function.
C.5. Resulting detection efficiency
Within the framework described in the preceding sections, we computed the detection efficiency ηdet, 3D(p[50 − 300], Ep, obs, α) at a number of points on a (p[50 − 300], Ep, obs, α) three-dimensional grid. For each point of the grid, we simulated a large number of SGRBs with isotropic (θ, ϕ) positions in the Fermi sky, with randomly sampled NaI backgrounds, by repeatedly following the procedure outlined above. Finally, we estimated ηdet, 3D(p[50 − 300], Ep, obs, α) as the fraction of simulated SGRBs that yielded a detection over the total.
Panels (a) in Fig. C.6 show the resulting detection efficiency as a function of p[50 − 300] for a number of values of Ep, obs and α. Panels (b) additionally show the detection efficiency averaged over the observed distribution of Ep, obs (panel b.1), α (panel b.2) or both (panel b.3). The averaging is done by constructing kernel density estimates w(Ep, obs) and w(α) of the distributions of these quantities from the best-fit values reported in the GBM spectral catalogue, limiting to the events for which a valid value for the corresponding COMP model parameter was available. The distributions are shown in Fig. C.5. The averaged ηdet, 3D is then obtained as
![]() |
Fig. C.5. Distributions of the low-energy photon index, α, and peak SED photon energy, Ep, obs, within the observed Fermi/GBM sample of SGRBs successfully modelled with the COMP model at peak flux. Blue histograms show density estimates constructed by binning the best-fit values reported in the GBM catalogue. Red lines show the corresponding kernel density estimates. |
![]() |
Fig. C.6. Fermi/GBM detection efficiency for SGRBs. Panels (a) show the detection efficiency ηdet, 3D for a fixed value of Ep, obs (reported on top of each panel) and different photon indices α (values shown in the legend), as a function of the 64 ms peak photon flux p[50 − 300]. Panels (b) are obtained by averaging ηdet, 3D over the observed distribution of Ep, obs (panel b.1), that of α (panel b.2), or both (panel b.3). In panels b.1 and b.2, the solid lines are contours of constant averaged ηdet, 2D, with the values reported along each line. |
The figure demonstrates that the dependence of the detection efficiency on the low-energy photon index is essentially negligible, except for extreme cases where Ep, obs lies well below the 50-300 keV band. For these cases, on the other hand, the values p[50 − 300] reported on the x-axis would correspond to unrealistically large bolometric fluxes. The dependence on Ep, obs is relevant only for very low values of this quantity. For the purposes of our study, we define the detection probability, expressed as a function of the source intrinsic parameters λsrc = (L, Ep, z), as
where the peak photon flux p is computed as defined in Eq. 17, setting α = −0.4. In practice, this is obtained by linear interpolation of ηdet, 2D, Ep, obs over its grid. The result, for a number of fixed peak luminosities, is shown in Fig. C.7.
![]() |
Fig. C.7. Fermi/GBM triggering efficiency for SGRBs as a function of their isotropic-equivalent peak luminosity, L (annotated within each panel), SED peak photon energy, Ep, and redshift, z, assuming a low-energy photon index α = −0.4. |
All Tables
Best-fit parameter values for the GBM NaI incidence-angle-dependent effective area ansatz function.
All Figures
![]() |
Fig. 1. Example luminosity functions ϕ(L) = dP/dln(L) induced by a Gaussian apparent structure model (Eqs. (10), left-hand panel) and a power law apparent structure model (Eqs. (13)) with a luminosity profile slope α = 3 (right-hand panel), both with a core half-opening angle θc = 0.1 rad and |
In the text |
![]() |
Fig. 2. Example double smoothly broken power law jet apparent structure model and induced L and Ep probability distributions. Panels a.1 and a.2 show the apparent jet structure functions |
In the text |
![]() |
Fig. 3. Inverse cumulative distribution of Fermi/GBM SGRB peak photon fluxes. The solid red line shows the inverse cumulative number of SGRBs detected by GBM with spectral information available in the catalogue, as a function of the peak photon flux measured on a 64 ms timescale in the 50–300 keV band. The pink band shows the one-sigma-equivalent Poisson error. The dashed black line shows a power law |
In the text |
![]() |
Fig. 4. L − Ep contours for our sub-sample of Fermi/GBM SGRBs detected also by Swift/BAT with a 64 ms peak flux p[15 − 150] > 3.5 cm−2 s−1 (see the main text for the full sample selection criteria). The main panel shows the contours that contain 50% (thick lines) and 90% (thin lines) of the posterior probability on (L, Ep) for bursts with a spectroscopically measured redshift, plus GRB 170817A (which does not belong to the GBM+BAT sample); the right-hand inset shows the corresponding contours for four events with a photometric redshift measurement (from Fong et al. 2022); the left-hand inset shows the contours for the remaining two SGRBs with an unknown redshift, constructed assuming a uniform prior on z in the range (10−4, 4). These two events were not included in the analysis. |
In the text |
![]() |
Fig. 5. GRB 170817A viewing angle posterior probability distribution, assuming the jet to be aligned with the GW170817 binary total angular momentum. The blue line shows the posterior probability constructed using the posterior samples from the low-spin-prior GW analysis (Abbott et al. 2019), while the red line shows the result of conditioning on the host galaxy distance (Cantiello et al. 2018), as explained in the main text. |
In the text |
![]() |
Fig. 6. Average apparent jet structure. The two larger panels show the constraints on the SGRB average apparent structure obtained from our analysis: solid lines show the medians of the projected posterior distributions of the jet structure functions at fixed θv from the full sample analysis (red: |
In the text |
![]() |
Fig. 7. Average apparent jet structure in the L − Ep plane. The solid turquoise line connects the points |
In the text |
![]() |
Fig. 8. Luminosity function. The median of the posterior probability density of R0ϕ(L) = dR0/dln(L) (i.e. the local rate density per unit logarithm of the peak luminosity) from the full sample analysis is shown with a solid red line, while the dashed orange line refers to the flux-limited sample analysis. The shaded regions encompass the symmetric 90% credible interval at each fixed L. We show for comparison the corresponding results from W15 (blue) and G16 (grey – their fiducial model ‘a’). The credible region of W15 is computed assuming uncorrelated errors on their parameters. |
In the text |
![]() |
Fig. 9. Local rate density. In the left-hand panel, the solid red (respectively dashed orange) line shows the posterior probability density on the total local rate density, R0, of SGRBs with a luminosity L > 1044 erg s−1 (which encompasses all viewing angles) from our analysis using the full (respectively flux-limited) sample. The grey shaded band shows the constraint on the local rate density of BNS mergers from GW observations, as derived in Abbott et al. (2023). In the right-hand panel, the solid red and dashed orange lines show the local rate density for events with a luminosity L > 1050 erg s−1 in our two analysis setups. For comparison, we show the corresponding local rate densities from W15 and G16 above the same minimum luminosity. |
In the text |
![]() |
Fig. 10. Rate density evolution. The figure shows the redshift evolution of the rate density from both our full sample (solid red lines) and flux-limited sample (dashed orange lines) analysis, limited to events above a minimum luminosity Lmin = 1050 erg s−1. The corresponding evolutions from W15 and G16 are shown in blue and grey, respectively. The thick lines show the median of the posterior predictive distribution at each fixed redshift, while the shaded areas encompass the symmetric 90% credible interval. |
In the text |
![]() |
Fig. 11. Observer-frame constraints and best-fit model predictions. Panel a.0 shows the predicted distribution of Fermi/GBM SGRBs on the (p[50 − 300], Ep, obs) plane for our best-fit model, with progressively lighter contours containing 50%, 90%, 99% and 99% of the events. Red dots show the observed data reported in the Fermi/GBM catalogue that pass our additional quality cuts, while grey points show the events that are discarded. We additionally show the error bars with thin grey lines for those events with a relative error larger than 50% on either quantity, or both. Panels a.1 and a.2 show the predicted (solid blue line) and observed (solid red cumulative histogram, with the pink region showing the 90% confidence region that stems from statistical uncertainties on spectral fitting parameters) cumulative distributions of p[50 − 300] (panel a.1) and Ep, obs (panel a.2) for events that pass the quality cuts. Panel b shows the inverse cumulative distribution of p[50 − 300], highlighting the behaviour at the high-flux end, which follows the expected |
In the text |
![]() |
Fig. 12. Rest-frame sample, viewing angle sample, and model predictions. This is similar to Fig. 11, but for rest-frame quantities L and Ep. Panel a.0 shows the predicted smallest regions in the (L, Ep) plane where 50%, 90%, 99% and 99.9% of the SGRBs that are detected by Fermi/GBM and by Swift/BAT with p[15 − 150] > plim, BAT are located, according to our model, using the parameters at the median of the λpop posterior (filled blue contours: full sample analysis; dashed grey contours: flux-limited sample analysis). Red crosses show the SGRBs with a measured redshift in our rest-frame sample. Panels a.1, a.2 and b show the predicted (blue solid lines: full sample; grey dashed lines: flux-limited sample) and observed (red cumulative histograms, with pink shaded areas showing the 90% confidence regions) cumulative distributions of luminosities (panel a.1), Ep (panel a.2), and redshifts (panel b). In panels a.1, a.2 and b, the 90% credible bands stemming from the uncertainty on λpop are shown with blue shading (full sample) and dotted lines (flux-limited sample). The additional orange contours in panel a.0 show the 50%, 90%, 99% and 99.9% containment regions for SGRBs with a BNS merger counterpart detected by Fermi/GBM and by the aLIGO and Advanced Virgo detectors with O3 sensitivity. The orange cross shows the position of GRB 170817A on the plane. |
In the text |
![]() |
Fig. 13. Viewing angle posterior probability densities for Fermi/GBM SGRBs with known redshifts. For each SGRB with known redshift in our sample, we show the posterior probability density dP/dlnθv, i = θv, iP(θv, i | d) (see Eq. (43)) constructed using 300 population posterior samples from the full sample analysis (solid lines) or the flux-limited sample analysis (dashed lines). Tick marks indicate the 5th, 50th and 95th percentiles of each posterior probability. Individual SGRBs are assigned the same colours as in Fig. 4. |
In the text |
![]() |
Fig. 14. Visualisation of the impact of selection effects on the L–Ep plane. The grey filled contours in the figure show the Fermi/GBM SGRB detection efficiency averaged over redshift, assuming the redshift distribution to be described by our model with parameters corresponding to the median of the population posterior from the full sample analysis (values corresponding to different shades of grey are shown in the colour bar). Purple contours show the areas on the plane that contain 50%, 90%, 99%, 99.9%, 99.99%, and 99.999% of SGRBs in the Universe according to the model. The thick dashed purple line shows the |
In the text |
![]() |
Fig. 15. Projected joint SGRB + GW rates for O4. The top (respectively bottom) panels refer to the results from the full (respectively flux-limited) sample analysis. Left-hand panels show the predicted rate Ṅobs(<z) of SGRBs within a redshift z detected by Fermi/GBM alone (blue: median; cyan: 90% credible range) or in coincidence with a network consisting of the two LIGO and the Virgo detectors with the projected O4 sensitivity (red: median; orange: 90% credible range). The right-hand panels show the corresponding rates Ṅobs(>θv) for events seen at a viewing angle larger than θv. |
In the text |
![]() |
Fig. A.1. Corner plot of the posterior probability densities from the two analyses. The full sample analysis is shown in magenta, while the flux-limited sample analysis is shown in blue. The histograms on the diagonal represent the marginalised posterior probability densities constructed from the posterior samples, with the solid vertical lines marking the medians and the vertical dashed lines delimiting the symmetric 90% credible interval (i.e. the 5th and the 95th percentiles – note that these are not shown if they differ by less than one bin size from the nearest edge of the allowed range). The contours in the remaining panels show the one, two, three and four sigma credible areas from the two-dimensional marginalised joint posterior probability densities, with the dots showing the intersections of the medians of the corresponding one-dimensional marginalised posterior probability densities. |
In the text |
![]() |
Fig. C.1. Probability density of the NaI background count rates in the 25-50 and 50-300 keV bands, constructed by sampling GBM ctime day data from days 210530 and 220928 at random times, combined from all 12 detectors. |
In the text |
![]() |
Fig. C.2. Background-subtracted count rate light curves in the 50-300 keV band for SGRBs observed by Fermi/GBM. Each light curve is plotted with a colour that depends on its peak count rate, as shown in the colour bar on the right. More details are given in the text. |
In the text |
![]() |
Fig. C.3. Ratio of SGRB count rate peak flux measured with a Δt binning over that measured with a 64 ms binning. This is constructed using Fermi/GBM trigdata files as explained in the text. Each thin red line corresponds to a single observed SGRB, while the light blue lines mark the 1st, 5th, 25th, 75th, 95th and 99th percentiles. |
In the text |
![]() |
Fig. C.4. NaI detector effective area model. Panel (a) shows the NaI detector effective area measurements (red crosses), for zero incidence angle, reported in M09. The pink line shows our adopted interpolation. Panel (b) shows the measured effective area (M09) for different photon incidence angles at three reference photon energies (black circles: 32 keV; blue squares: 279 keV; red triangles: 662 keV). The solid lines show our best-fitting model. Panel (c) shows the best-fit values of parameters c1, c2 and c3 of our model (Eq. C.5) for the angular dependence of the effective area at the three reference energies (coloured circles). The dashed lines show our assumed behaviour of ci as a function of the photon energy. |
In the text |
![]() |
Fig. C.5. Distributions of the low-energy photon index, α, and peak SED photon energy, Ep, obs, within the observed Fermi/GBM sample of SGRBs successfully modelled with the COMP model at peak flux. Blue histograms show density estimates constructed by binning the best-fit values reported in the GBM catalogue. Red lines show the corresponding kernel density estimates. |
In the text |
![]() |
Fig. C.6. Fermi/GBM detection efficiency for SGRBs. Panels (a) show the detection efficiency ηdet, 3D for a fixed value of Ep, obs (reported on top of each panel) and different photon indices α (values shown in the legend), as a function of the 64 ms peak photon flux p[50 − 300]. Panels (b) are obtained by averaging ηdet, 3D over the observed distribution of Ep, obs (panel b.1), that of α (panel b.2), or both (panel b.3). In panels b.1 and b.2, the solid lines are contours of constant averaged ηdet, 2D, with the values reported along each line. |
In the text |
![]() |
Fig. C.7. Fermi/GBM triggering efficiency for SGRBs as a function of their isotropic-equivalent peak luminosity, L (annotated within each panel), SED peak photon energy, Ep, and redshift, z, assuming a low-energy photon index α = −0.4. |
In the text |
Current usage metrics show cumulative count of Article Views (full-text article views including HTML views, PDF and ePub downloads, according to the available data) and Abstracts Views on Vision4Press platform.
Data correspond to usage on the plateform after 2015. The current usage metrics is available 48-96 hours after online publication and is updated daily on week days.
Initial download of the metrics may take a while.