ArticlePDF Available

Selection Effects in Gamma-Ray Burst Correlations: Consequences on the Ratio between Gamma-Ray Burst and Star Formation Rates

Authors:

Abstract

Gamma Ray Bursts (GRBs) visible up to very high redshift have become attractive targets as potential new distance indicators. It is still not clear whether the relations proposed so far originate from an unknown GRB physics or result from selection effects. We investigate this issue in the case of the $L_X-T^*_a$ correlation (hereafter LT) between the X-ray luminosity $L_X (T_a)$ at the end of the plateau phase, $T_a$, and the rest frame time $T^{*}_a$. We devise a general method to build mock data sets starting from a GRB world model and taking into account selection effects on both time and luminosity. This method shows how not knowing the efficiency function could influence the evaluation of the intrinsic slope of any correlation and the GRB density rate. We investigate biases (small offsets in slope or normalization) that would occur in the LT relation as a result of truncations, possibly present in the intrinsic distributions of $L_X$ and $T^*_a$. We compare these results with the ones in Dainotti et al. (2013) showing that in both cases the intrinsic slope of the LT correlation is $\approx -1.0$. This method is general, therefore relevant to investigate if any other GRB correlation is generated by the biases themselves. Moreover, because the farthest GRBs and star-forming galaxies probe the reionization epoch, we evaluate the redshift-dependent ratio $\Psi(z)=(1+z)^{\alpha}$ of the GRB rate to star formation rate (SFR). We found a modest evolution $-0.2\leq \alpha \leq 0.5$ consistent with Swift GRB afterglow plateau in the redshift range $0.99<z<9.4$.
arXiv:1412.3969v3 [astro-ph.HE] 7 Jan 2015
Selection effects in Gamma Ray Bursts correlations: consequences
on the ratio between GRB and star formation rates
Dainotti, M. G. 1,2,3, Del Vecchio, R. 3, Shigehiro, N. 1, Capozziello, S. 4,5,6
ABSTRACT
Gamma Ray Bursts (GRBs) visible up to very high redshift have become at-
tractive targets as potential new distance indicators. It is still not clear whether
the relations proposed so far originate from an unknown GRB physics or result
from selection effects. We investigate this issue in the case of the LXT
acor-
relation (hereafter LT) between the X-ray luminosity LX(Ta) at the end of the
plateau phase, Ta, and the rest frame time T
a. We devise a general method to
build mock data sets starting from a GRB world model and taking into account
selection effects on both time and luminosity. This method shows how not know-
ing the efficiency function could influence the evaluation of the intrinsic slope
of any correlation and the GRB density rate. We investigate biases (small off-
sets in slope or normalization) that would occur in the LT relation as a result
of truncations, possibly present in the intrinsic distributions of LXand T
a. We
compare these results with the ones in Dainotti et al. (2013) showing that in both
cases the intrinsic slope of the LT correlation is ≈ −1.0. This method is general,
therefore relevant to investigate if any other GRB correlation is generated by
the biases themselves. Moreover, because the farthest GRBs and star-forming
galaxies probe the reionization epoch, we evaluate the redshift-dependent ratio
Ψ(z) = (1 + z)αof the GRB rate to star formation rate (SFR). We found a
1Astrophysical Big Bang Laboratory, Riken, 2-1 Hirosawa, Wako, Saitama 351-0198, Japan,
maria.dainotti@riken.jp.
2Physics Department, Stanford University, Via Pueblo Mall 382, Stanford, CA, USA, E-mail: mdain-
ott@stanford.edu
3Obserwatorium Astronomiczne, Uniwersytet Jagiello´nski, ul. Orla 171, 31-501 Krak´ow, E-mails: delvec-
chioroberta@hotmail.it, dainotti@oa.uj.edu.pl, mariagiovannadainotti@yahoo.it
4Dipartimento di Fisica, Universit´adi Napoli ”Federico II”, Compl. Univ. di Monte S. Angelo, Edicio
G, Via Cinthia, I-80126 Napoli, Italy, E-mail: capozziello@na.infn.it
5INFN Sez. di Napoli, Complesso Universitario di Monte S. Angelo, Via Cinthia, Edicio N, 80126 Napoli,
Italy
6Gran Sasso Science Institute (INFN), viale F. Crispi 7, I-67100 L’Aquila, Italy
– 2 –
modest evolution 0.2α0.5 consistent with Swift GRB afterglow plateau
in the redshift range 0.99 < z < 9.4.
Subject headings: stars: gamma-ray burst: general, statistics, methods: data
analysis.
1. INTRODUCTION
GRBs are the farthest sources, seen up to redshift z= 9.46 (Cucchiara et al. 2011), and
if emitting isotropically they are also the most powerful, (with Eiso 1054 erg s1), objects
in the Universe. Notwithstanding the variety of their peculiarities, some common features
may be identified by looking at their light curves. A crucial breakthrough in this area has
been the observation of GRBs by the Swift satellite, launched in 2004. With the instru-
ments on board, the Burst Alert Telescope (BAT, 15-150 keV), the X-Ray Telescope (XRT,
0.3-10 keV), and the Ultra-Violet/Optical Telescope (UVOT, 170-650 nm), Swift provides
a rapid follow-up of the afterglows in several wavelengths with better coverage than previ-
ous missions. Swift observations have revealed a more complex behavior of the light curves
afterglow (O’Brien et al. 2006; Sakamoto et al. 2007) that can be divided into two, three
and even more segments in the afterglows. The second segment, when it is flat, is called
plateau emission. A significant step forward in determining common features in the after-
glow light curves was made by fitting them with an analytical expression (Willingale et al.
2007), called hereafter W07. This provides the opportunity to look for universal features
that could provide a redshift independent measure of the distance from the GRB, as in
studies of correlations between GRB isotropic energy and peak photon energy of the νFν
spectrum, Eiso Epeak , (Lloyd and Petrosian 1999; Amati et al. 2009), the beamed total en-
ergy EγEpeak (Ghirlanda et al. 2004, 2006), the Luminosity-Variability, L-V (Norris et al.
2000; Fenimore and Ramirez-Ruiz 2000), L-Epeak (Yonetoku et al. 2004) and Luminosity- τ
lag (Schaefer 2003).
Dainotti et al. (2008, 2010), using the W07 phenomenological law for the light curves of
long GRBs, discovered a formal anti-correlation between the X-ray luminosity at the end
of the plateau LXand the rest frame plateau end-time, T
a=Tobs
a/(1 + z), where T
ais
in seconds and LXis in erg/s. The normalization and the slope parameters aand bare
constants obtained by the D’Agostini fitting method (D’Agostini 2005). Dainotti et al.
(2011a) attempted to use the LT correlation as a possible redshift estimator, but the paucity
of the data and the scatter prevents from a definite conclusion at least for a sample of
62 GRBs. In addition, a further step to better understand the role of the plateau emis-
sion has been made with the discovery of new significant correlations between LX, and
– 3 –
the mean luminosities of the prompt emission, < Lγ ,prompt >(Dainotti et al. 2011b). The
LT anticorrelation is also a useful test for theoretical models such as the accretion models,
(Cannizzo and Gehrels 2009; Cannizzo et al. 2011), the magnetar models (Dall’Osso et al.
2011; Bernardini et al. 2012a,b; Rowlinson et al. 2010, 2013, 2014), the prior emission model
(Yamazaki 2009), the unified GRB and AGN model (Nemmen et al. 2012) and the fireshell
model (Izzo et al. 2013). Moreover, Hasco¨et et al. (2014) and van Eerten (2014b) consider
both the LT and the LX-< Lγ,prompt >correlation to discriminate among several models
proposed for the origin of plateau. In Leventis et al. (2014) and in van Eerten (2014a) a
smooth energy injection through the reverse shock has been presented as a plausible expla-
nation for the origin of the LT correlation. Furthermore, also other authors were able to
reproduce and use the LT correlation to extend it in the optical band (Ghisellini et al. 2009),
to extrapolated it into correlations of the prompt emission (Sultana et al. 2012) and to use
the same methodology to build an analogous correlation in the prompt (Qi and Lu 2012).
Finally, it has been applied as a cosmological tool (Cardone et al. 2009, 2010; Dainotti et al.
2013a; Postnikov et al. 2014). Impacts of detector thresholds on cosmological standard can-
dles have also been considered (Shahmoradi and Nemiroff 2009; Petrosian and Lloyd 1998;
Petrosian et al. 1999; Petrosian 2002; Cabrera et al. 2007). However, because of large dis-
persion (Butler et al. 2010; Yu et al. 2009) and absence of good calibration none of these
correlations allow the use of GRBs as good standard candles as it has been done e.g. with
type Ia Supernovae. An important statistical technique to study selection effects for treating
data truncation in GRB correlations is the Efron and Petrosian (1992) method. Another
way to study the same problem in GRB correlations, derived modeling the high-energy
properties of GRBs, have been reported in Butler et al. (2010). In the latter paper it has
been shown that well-known examples of these correlations have common features indicative
of strong contamination by selection effects. We compare this procedure with the method
introduced by Efron & Petrosian (1992) and applied to the LT correlation (Dainotti et al.
2013b). The paper is organized as follows: section 2 introduces the relation between GRB
and SFR, section 2.1 is dedicated to the analysis of a GRB scaling relation, in particular we
consider the LT correlation as an example, but the procedure described can be adopted for
any other correlation. In section 3 we describe how to build the GRB samples, in section
4 we analyze the redshift evolution of the slope and normalization of the LT correlation.
In section 5 we study the selection effects related to simulated samples assuming different
normalization and slope values. Then, in section 6 we draw conclusions on the intrinsic slope
of the LT correlation and on the evaluation of the redshift-dependent ratio between GRB
and star formation rates.
– 4 –
2. The relation between GRB rate and the star formation rate
In order to understand the relation between GRBs and the star formation, it is often
assumed that the GRB rate (RGRB) is proportional to the SFR then the predicted distribu-
tion of the GRB redshift is compared to the observed distribution (Totani 1997; Mao and Mo
1998; Wijers et al. 1998; Porciani and Madau 2001; Natarajan et al. 2005; Jakobsson et al.
2006; Daigne and Mochkovitch 2007; Le and Dermer 2007; Coward 2007; Mao 2010). How-
ever, this relationship is not an easy task to handle, because some studies show that GRBs
do not seem to trace the star formation unbiasedly (Lloyd and Petrosian 1999). Namely, the
ratio between the RGRB and SFR, RGRB/SFR, increases with redshift (Kistler et al. 2008;
uksel and Kistler 2007) significantly. This means that GRBs are more frequent for a given
star formation rate density at earlier times. In fact, while observations consistently show
that the comoving rate density of star formation is nearly constant in the interval 1 z4
(Hopkins and Beacom 2006), the comoving rate density of GRBs appears evolving distinctly.
In our approach we explicitly take into consideration this issue when we fit the observed
GRB rate with the model. Selection effects involved in a GRB sample are of two kinds :
the GRB detection and localization; and the redshift determination through spectroscopy
and photometry of the GRB afterglow or the host galaxy. These problems have been ob-
ject of extensive study in literature (Bloom 2003; Fiore et al. 2007; Guetta and Della Valle
2007). Moreover, the Swift trigger, is very complex and the sensitivity of the detector is
very difficult to parameterize exactly (Band 2006), but in this case not dealing with prompt
peak energy we do not have to take into consideration the double truncation present in data
(Lloyd and Petrosian 1999). In the case of plateau it is easier, since an effective luminosity
threshold appears to be present in the data which can be approximated by a 0.310 keV
energy flux limit Flim 2×1012 erg cm2s1(Dainotti et al. 2013b). The luminosity thresh-
old is then Llim = 4πD2
L(z, M, H0)Flim , where DLis the luminosity distance to the burst.
Throughout the paper, we assume a flat universe with ΩM= 0.28, Ωλ= 0.72 and H0=70
km s1Mpc1. In our approach below several models are considered and then the one that
best matches the GRB rate with star formation rate has been chosen.
2.1. GRBS WORLD MODEL
We derive a model capable of reproducing the observed Swift GRB rate as a function
of redshift, luminosity and time of the plateau emission.
Rest frame time and luminosity at the end of the GRB plateau emission show strong cor-
relations as discovered by Dainotti et al. (2008) and later updated by Dainotti et al. (2010,
2011a,b, 2013a). Therefore, all these quantities must be considered in deriving reliable rates.
– 5 –
We characterize the GRB rate as a product of terms involving the redshift z of the bursts,
the isotropic equivalent luminosity release (0.3-10 keV) LXand the duration T
a.
Let us assume that a scaling relation exists so that the luminosity LX(Ta) for a GRB
with time scale T
aat redshift z is given by :
λ=α0+αττ+αζζ(1)
where we have introduced the compact notation
λ=log LX(T
a)
τ=log [T
a/(1 + z)]
ζ=log (1 + z).
(2)
and the term ζaccounts for redshift evolution. The luminosity is normalized by the
unit of 1 erg s1and the time by the unit of 1 s, so that non dimensional quantities are
considered. All the observables in this model are computed in the rest frame, because we
are testing the role played by selection effects in the rest frame, being the LT correlation
rest frame corrected. Independently, on the physical interpretation of this relation, (in fact,
there are several models that can reproduce it as we have mentioned in the introduction) we
can nevertheless expect GRBs to follow Equation 1 with a scatter σλ. Moreover, the zero
point α0may be known only up to a given uncertainty σα. Following the approach of Butler
et al. (2010) applied for prompt correlations, we assume that λcan be approximated by
Gaussian distribution with mean λ0, expressed in Equation 1, and the variance σint as the
intrinsic scatter of the correlation. We also write the probability that a GRB with given (τ,
ζ) values has a luminosity λas follows:
Pλ(λ, τ, ζ )exp[1
2[λ(α0+αττ+αζζ)
σint
]2] (3)
with σ2
int =σ2
λ+σ2
α, with σαthe uncertainty of the α0value and σλthe uncertainty on
the luminosity value.
The approximation of a Gaussian distribution both for the luminosity and time is motivated
by the goodness of the fit which gives a probability P= 0.46 and P= 0.61 respectively,
see Fig. 1 and 2. We note that the mean, (indicated with <>)< Ta>= 3.35 (s) with a
variance σTa= 0.77 (s) and < LX>= 48.04 (erg/s) with a variance σLX= 1.37 (erg/s) are
represented respectively in Fig. 1 and 2.
– 6 –
Fig. 1.— Probability Density Distribution of T
a, the rest frame end time of the plateau, for GRBs
observed from 2005 January until 2014 July analyzed following the Dainotti et al. (2013a) approach with a
superimposed best fit of the Gaussian distribution.
Fig. 2.— Probability Density Distribution of LX(Ta) at the end of the plateau for GRBs observed from
2005 January until 2014 July analyzed following the Dainotti et al. (2013a) approach with a superimposed
best fit of the Gaussian distribution.
– 7 –
In order to obtain the number of GRBs with a given luminosity λ, we need to integrate
over the distributions of τand ζ. We will assume, for simplicity, that τfollows a truncated
Gaussian law.
Pτ(τ)(exp[1
2(ττ0
στ)2]τL< τ < τU
0ττLor τ τU
(4)
where τLand τUindicate respectively the lower limit and upper limit of the observed τ
distribution and τ0is the mean value of this distribution. The limits of τare taken from an
updated sample of T
acomposed of 176 GRBs afterglows, with firm redshift determination,
from January 2005 till July 2014. The analysis follows the criteria adopted in Dainotti et al.
(2013a).
If we assume that the GRBs trace the cosmic star formation rate, we can model their
redshift distribution following Butler et al. (2010) as:
Pz(z)˙ρ(z)
1 + z
dV
dz (5)
where ˙ρ(z) is the comoving GRB rate density, V is the universal volume, and the factor
(1 + z) accounts for cosmic time dilatation and
dV
dz r2(z)
E(z)(6)
with r(z) the comoving distance and E(z) = H(z)/H0the Hubble parameter normalized
to its present day value.
Collecting the different terms, we can finally write the true, detector-independent event N
differential rate, for z, log T
aand log LX, as:
dN
dλdτdz Ψ(z)Pλ(λ, τ , ζ)Pτ(τ)Pz(z).(7)
We here note that we have introduced the term of the evolution in redshift, Ψ(z) =
(1 + z)α, following the approach of Lloyd & Petrosian (1999), Dermer (2007) and Robertson
& Ellis (2012). In Dermer (2007) assuming that the emission properties of GRBs do not
change with time, they find that the Swift data can only be fitted if the comoving rate
density of GRB sources exhibits positive evolution to z > 35. In our approach we
introduce evolution starting from z0.99.
– 8 –
So using the above expression for Pτ, we find that the number of GRBs with luminosity
in the range (λ,λ+) and redshift between zand z+dz is:
dN
dλdz Ψ(z)˙ρ(z)(dV /dz)
1 + zFτU− FτL
p8πσ2
τ
exp[1
2[λ(α0+αττ+αζζ)
pσ2
int
]2] (8)
where FτUand FτLare the error functions1of the lower and upper limit of the time
distribution. Note that Equation 8 is defined up to an overall normalization constant which
can be solved by imposing that the integral of dN/dλdz over (λ,z) gives the total number
of observed GRBs. Actually, this is not known since we do not observe all GRBs, but only
those passing a given set of selection criteria. However, we will be only interested in the
fraction of GRBs in a cell in the 2D (λ,z) space so that we do not need this quantity.
We are aware that we don’t map out the true LT relation given selection effects and the
observed LT relation. Doing this would require modeling the selection of the GRB sample
itself (using the gamma ray threshold) and also seeking to understand the tie between the
GRB flux and the afterglow LX. However, the relation between flux and LXhas been
already studied by Dainotti et al. (2013a) and reported briefly in the previous section. Here
we computed the new limit related to the updated sample, as it has been shown in the middle
panel of Fig 3.
3. SIMULATING THE GRB SAMPLES
The GRBs rate given by Equation 8 has been derived by implicitly assuming that all the
GRBs can be detected notwithstanding their observable properties. This is actually not the
case. As an example, we will consider hereafter the LT correlation although the formalism
and the method we will develop can be easily extended to whatever scaling law. For the
LT case, there are two possible selection effects. First, each detector has an efficiency which
is not the same for all the luminosities. Only GRBs with λ > λL, where λLis the lowest
detectable luminosity for a given instrument, can be detected while all the GRBs with λ
larger than a threshold luminosity λUwill be found.
Moreover, it is likely that the efficiency of the detector is not constant, but is rather a
1We remind that the usual definition of the error function is
erf (x) = 2
πZx
0
et2dt. (9)
– 9 –
function of the luminosity. We will therefore introduce an efficiency function Eλ(λ) whose
functional expression is not known in advance, but can only take values in the range (0,1).
A second selection effect is related to the time duration of the GRB. Indeed, in order to
be included in the sample used to calibrate the LT correlation, the GRB afterglow has to
be measured over a sufficiently long time scale to make possible to fit the data and extract
the relevant quantities. If τis too small, as it has been shown in Dainotti et al. (2013a)
the minimum rest frame time is 14 s, few points will be available for the fit, while, on the
contrary, large τvalues will give rise to afterglow light curves which could be well sampled
by the data. Again, we can parametrize these effects introducing a second efficiency function
Eτ(τ) so that the final observable rate is the following:
dNobs
dλdz dN
dλdz × Eλ(λ)Eτ(τ).(10)
We point out that our formulation, which takes into account of the efficiency functions
Eλ(λ) and Eλ(τ) in the final observed GRB rate is similar to the approach by Robertson &
Ellis (2012) in Equation 1, in which the additional factor K is presented. K is equivalent to
our Eλ(λ) and Eλ(τ).
It is worth noting that Equation 10 is actually still a simplified description. Indeed, it
is in principle possible that other selection effects take place involving observable quantities
not considered here, as for example βand the redshift. However, these parameters enter
in the determination of λso that one can (at least in a first order approximation) convert
selection cuts on them in a single efficiency function depending only on λ(for the dependence
of the flux on the redshift see left panel of Fig. 3). However, as we can see from Fig. 3 β
is constant with redshift, and there is no correlation between those two quantities, in fact
the Spearman correlation coefficient is ρ=0.062. Nevertheless, Equation 10 provides a
Fig. 3.— Flux at the end of the plateau phase, Flux(Ta), (left panel) and the spectral index, β, (right
panel) as a function of redshift. The limiting luminosity, log LXvs 1 + zshows (middle panel) two lines, one
for the limiting flux, FSwif t,lim = 1014 erg cm2s1and the other one is the most suitable for a plateau
duration of 104s, which is 2 ×1012 erg cm2s1.
– 10 –
reasonably accurate description of the observable GRB rate.
In order to evaluate Equation 10 there are different quantities to determine. First, we
need to set the scaling coefficients (α0,ατ,αζ) and the intrinsic scatter σint. Second, the
mean and variance of the τdistribution (τ0,στ) has to be given. Finally, an expression for
the cosmic SFR ˙ρ(z) has to be assigned. None of these quantities is actually available. In
principle, one could assume a SFR law and fit for the model parameters to a large enough
GRBs sample with measured (λ,τ,ζ) values. To this end, one should know the selection
function Eλ(λ)Eτ(τ) which is not the case. Studies of how light curves would appear to a
gamma-ray detector here on Earth have been performed (Kocevski and Petrosian 2013). In
this paper the prompt emission pulses are investigated and the conclusion is that even a
perfect detector that observes over a limited energy range would not faithfully measure the
expected time dilation effects on a GRB pulse as a function of redshift.
Fig. 4.— GRBs rate density using method of Li (2008) and the observed GRBs rate density obtained by
the linear efficiency functions (upper panel), and the polynomial efficiency function (lower panel) with the
redshifts distribution of our data sample.
Nevertheless, here we study detector threshold effects on afterglow properties. Our
– 11 –
aim is to investigate how the ignorance of the efficiency function bias the estimate of the
correlation coefficients. We can therefore rely on simulated samples based on a realistic
intrinsic rate. We proceed as schematically outlined below.
(i) We assume that the available data represent reasonably well the intrinsic τdistribution
so that we can infer (τ0,στ) from the data themselves. We set τL,U =τ0±5στthus
symmetrically cutting the Gaussian distribution at its extreme ends.
(ii) Based on the shape of the cosmic SFR (Hopkins and Beacom 2006), we assume a
broken power law for the comoving GRB rate density:
˙ρ(z)
(1 + z)g0zz0
(1 + z)g1z0zz1
(1 + z)g2zz1
(11)
where the relative normalizations are set so that ˙ρ(z) is continuous at z0= 0.97 and
z1and (z0, z1) = (0.97,4.50), (g0, g1, g2) = (3.4,0.3,8.0). Moreover, besides the
equation 11, we employed other shapes of the SFR (Li 2008; Robertson and Ellis 2012;
Kistler et al. 2013) to obtain the observed GRBs rate density. The one used by (Li
2008) is:
˙ρ(z) = a+b×Log(1 + z).(12)
The a and b parameters are :
(a, b) =
(1.70,3.30) z0.993
(0.727,0.0549) 0.993 z3.80
(2.35,4.46) z3.80
(13)
Robertson & Ellis (2012) defined the SFR as:
˙ρ(z) = a+b(z/c)f
1 + (z/c)d+g, (14)
where they have a= 0.009Myr1M pc3,b= 0.27Myr1Mpc3,c= 3.7, d= 7.4,
and g= 103Myr1Mpc3.
Instead, Kistler et al. (2013) defined the SFR as :
– 12 –
˙ρ(z) = ˙ρ0×[(1 + z)+ ( 1 + z
B)+ (1 + z
C)]1
ψ,(15)
with slopes a= 3.4, b=0.3, and c=2.5, breaks at z1= 1 and z2= 4 corresponding
to B= (1 + z1)1a
b5160 and C= (1 + z1)(ba)
c×(1 + z2)(1b)
c11.5, and ψ=10.
Finally, we compare the fitted functions obtained with these four methods with our data
distribution. The most realible fits for our parameters is the SFR used by Li (2008),
see Fig. 4 where the best fit among linear (upper panel) and polynomial (lower panel)
ǫ(λ) functions are considered. Moreover, we adopted the constraints for the redshift
dependent ratio between SFR and GRB rate adopted by Robertson & Ellis (2012).
In this paper a modest evolution (e.g.,Ψ(z)(1 + z)α) with 0.2α1.5, where
the peak probability occurs for α0.5 is consistent with the long GRB prompt data
(P0.9). These values can be explained if GRBs occur primarily in low-metallicity
galaxies which are proportionally more numerous at earlier times. We note that in our
approach we assumed no evolution at low redshift for z0.99 consistently with the
posterior probability in Robertson & Ellis (2012) in which no evolution is possible at
the 2-σlevel. However, because a constant Ψ(z) is also ruled out (Robertson and Ellis
2012), then we fit the normalization parameters and the evolution factors obtaining
Ψ(z)(1 + z)0.2for 0.993 z3.8 and Ψ(z)(1 + z)0.5for z3.8. These
values of the evolution are compatible with Robertson et al. (2012). Regarding the
observed GRB rate we obtained that the best efficiency functions are possible both for
two polynomial and two linear as we show in Fig. 4. Table 1 and 2 show the probability
that the density rate match the afterglow plateau GRB rate assuming those efficiency
functions
(iii) For given (α0,ατ,αζ,σint) values, we divide the 2D space (λ,z) in Mcells and, for
each cell, compute the fraction of GRBs in it as:
fsim(λi, zi) = Rλi+∆λ
λiλRzi+∆z
zizdz dN
dλdz
Rλmax
λmin Rzmax
zmin dz dN
dλdz
(16)
where we set
(λmin, λmax ) = (42.0,52.0),(zmin, zmax) = (0,10).(17)
We find more efficient to change variable from zto ζwhen dividing the 2D space in
10 ×10 square cells.
(iv) For each given cell, we generate Nij =fsim(λi, ζj)× Nsim GRBs (with Nsim the total
number of objects to simulate) by randomly sampling (λ,ζ) within the cell boundaries
and computing τby solving Eq. 1.
– 13 –
(v) To take into account of the selection effects, for each GRB, we generate two random
numbers (uτ,uλ) uniformly sampling the range (0, 1) and only retain the GRB if
uτ≤ Eτ(τ) and uλ≤ Eλ(λ). Note that, as a consequence of this cut, the final number
Nobs of observed GRBs is smaller than the input one Nsim.
(vi) Finally, for each one of the Nobs selected GRBs, we generate new (τobs,λobs ) values
extracting from Gaussian distributions centered on the simulated (τ,λ) values and
with a 1% variance. We also associate to each GRB an error set in such a way to be
similar to what is actually obtained for GRBs having comparable (τ,λ) values.
The above procedure allows us to build simulated GRBs sample taking into account
both the intrinsic properties of any scaling relation and the selection effects induced by the
instrumental setup. Moreover, we have referred to an actual GRBs sample in order to set
both the limits on (τ,ζ,λ) and the typical measurement errors. Therefore, we can rely on
these simulated samples to investigate the impact of selection effects on the recovered slope
and intrinsic scatter of the given correlation. To this end, the last ingredient we need is a
functional expression for the efficiency functions. Since these are largely unknown, we are
forced to make some arbitrary guess. Therefore, we consider two different cases. First, we
assume that there is no selection on τ, i.e. we set Eτ= 1. Two functional expressions are
then used for Eλ, namely a power law:
Eλ(λ) =
0λ < λL
(λλL
λUλL)EλλLλλU
1λ > λU
(18)
and a fourth order polynomial, i.e. :
Eλ(λ) =
0λ < λL
E1˜
λ+E2˜
λ2+E3˜
λ3+E4˜
λ4
E1+E2+E3+E4λLλλU
1λ > λU
(19)
with ˜
λ= (λλL)/(λUλL). We try different arbitrary choices for the parameters
entering both expressions of Eλin order to investigate to which extent the results depend
on the exact choice of the efficiency function, see Fig. 5. In a second step, we abandon the
assumption Eτ= 1, to assume for it the same functional expression used for Eλ, with the
same choices for the parameters, but different upper and lower limits depending on τUand
– 14 –
Fig. 5.— The first 5 panels represent examples of the efficiency function for the linear case versus lumi-
nosities of the GRBs, λ, in our data sample, while the last 5 panels the efficiency functions for the fourth
order polynomial. The linear functions as well the polynomial ones are computed according to Eq. 18, and
Eq. 19.
– 15 –
Fig. 6.— The first 5 panels represent examples of the efficiency function for the linear case versus the
times, τof the GRBs in our data sample, while the last 5 panels the efficiency functions for the fourth order
polynomial. The linear functions as well the polynomial ones are computed according to Eq. 18, and Eq.
19.
– 16 –
τL, see Fig. 6.
4. REDSHIFT EVOLUTION ON THE NORMALIZATION AND SLOPE
PARAMETERS
Fig. 7.— ατand normalization α0using a linear function α0=0.22x+ 52.31 (left panel) and ατ=
0.10x1.38 (right panel).
Fig. 8.— ατand normalization α0using a polynomial function α0= 55.87 8.13x+ 5.53x21.48x3+0.13x4
(left panel) and ατ=2.35 + 2.13x1.39x2+ 0.37x30.03x4(right panel).
As we have already mentioned in the previous paragraph the polynomial and the linear
model for the ǫ(λ) are unknown, then assumptions need to be made. We chose these forms,
because both normalization and slope of the LT correlation depend on the redshift either
with a polynomial or with a simple power law. Therefore, these choices for the selection
functions take into account of this redshift dependence. Namely, we consider a model redshift
– 17 –
Id λLλUEλPGRB,rate
PL1 44.34 50.86 1.25 104
PL2 43.64 49.87 2.99 0.003
PL3 43.77 50.74 1.65 0.53
PL4 44.77 49.59 2.04 0.001
PL5 44.14 50.83 0.23 0.54
Table 1: Efficiency function parameters for the power-law Eλand no cut on τi.e. Eτ= 1.
PGRB,rate is the goodness of fit between our data and the observed GRBs rate density,
thus how the data well fit the observed GRBs density rate. To compute the probability
we compute the χ2test that performs a statistical hypothesis test in which the sampling
distribution of the test statistic is a χ2distribution when the null hypothesis is true, in order
to determine whether there is a significant difference between the expected frequencies and
the observed frequencies.
Id λLλUE1E2E3E4PGRB,rate
PoL1 44.90 49.14 0.46 0.01 0.24 0.80 0.54
PoL2 41.10 50.23 0.60 0.95 0.05 0.53 104
PoL3 43.57 49.09 0.71 0.79 0.07 0.34 0.019
PoL4 44.37 49.52 0.51 0.03 0.46 0.78 0.15
PoL5 43.03 50.06 0.79 0.36 0.63 0.40 0.001
Table 2: Same as table 1 but for the polynomial functions. PGRB,r ate is the goodness of fit
between our data and the observed GRB rate density.
– 18 –
dependence because of the corresponding dependence of both luminosity and time. This has
been already shown in Dainotti et al. (2013a) and currently inthe middle panel of Fig. 3
for the updated data sample. To study the behavior of the redshift evolution we plot the
slope and the normalization values versus the redshift. These are obtained from the average
values for the data set divided into 5 bins, see figure 7 and into 12 bins, see figure 8. As we
can see from both figures 7 and 8 the normalization parameter α0decreases as the redshift
increases, while the slope parameter ατshows the opposite trend. Goodness of the fits is
given by the probability P= 0.79 for the data set divided in 5 bins and P= 0.87 for the
one divided into 12 bins for the linear case, while for the polynomial model P= 0.99 and
P= 0.94 for the data set divided in 5 and 12 bins respectively. These results show that both
polynomial and linear fit are possible.
5. IMPACT OF SELECTION EFFECTS
The simulated samples generated as described above are input to the same Bayesian
fitting procedure we use with real data. For each input (ατ,αζ,α0,σint) parameters, we
simulate 50 GRBs sample setting Nsim = 200, while the number of observed GRBs depend
on the efficiency function used. We fit these samples assuming no redshift evolution in Eq.
1, i.e. forcing αζ= 0 in the fit so that, for each simulated sample, the fitting procedure
returns both the best fit and the median and 68% confidence range of the parameters (ατ,
α0,σint). In order to investigate whether the selection effects impact the recovery of the
input scaling laws, we fit linear relations of the form:
xf=axinp +b(20)
where xinp is the input value and xfcan be either the best fit (denoted as xbf ) or
the median xf it value. When fitting the above linear relation, we use the χ2minimization
for xbf , while a weighted fit is performed for xf it with weights ωi= 12
iwhere σiis the
symmetrized 1σerror. Note that the label ihere runs over the simulations performed for
each given efficiency function.
5.1. No redshift evolution
Here, we consider input models with αζ= 0, i.e. no redshift evolution of the scaling law
(1). It is worth noting that such an assumption is actually well motivated since it has been
demonstrated in Dainotti et al. (2013a) that luminosity is almost not affected by redshift
– 19 –
Fig. 9.— Fitted vs input (ατ,α0,σint ) parameters obtained with the power law function. The first three
panels refer to the best fit values, while the other three show the median values with the 1σerror bars. Solid
red line is the best fit line while blue dashed is the no bias line when xinp =xf.
– 20 –
evolution, while time becomes to undergo redshift evolution for high redshift only. From our
point of view, however, this case allows us to directly quantify the impact of the efficiency
functions on the recovery of the scaling correlation parameters since any deviation will only
be due to the selection effects and not to any attempt of compensating the missed evolution
with z.
5.1.1. No selection on τ(Eτ= 1)
We start by considering the idealized case of no selection of τ, i.e., we force Eτ= 1,
and set the Eλparameters as listed in Tables 1 and 2 for the power-law and polynomial
expressions, respectively. As an example, figure 9 shows the results for the efficiency function,
while Table 3 summarizes the (a, b) coefficients of the linear fit between the input and
recovered quantities. The closer is ato 1, the less is the parameter biased, while b6= 0
should not be taken as evidence for bias. This result is in perfect agreement with the
intrinsic correlation slope, which is 1.07+0.09
0.14 (Dainotti et al. 2013b), when we consider as
the best choice for the selection functions the ones that returns values of acloser to 1. If
Equation 20 is fulfilled, we can estimate the relative bias as:
x
x=xinp xf
xinp
= 1 ab
xinp
,(21)
so that we can accept b6= 0 if xinp is much larger than b. This is indeed the case for
xinp =α0which takes typical values (50) much larger than the bones in Table 3.
From the proximity between solid and dashed lines, which represent respectively the best fit
line and the no bias line when xinp =xf, in the corresponding panels of Fig. 9, we see that,
for the power-law efficiency function (and no cut on τ), both the slope and the zero point
of the scaling relation are correctly recovered. The reason why is that the relative bias is
Id (a, b)τ
bf (a, b)τ
f it (a, b)α0
bf (a, b)α0
f it (a, b)σ
bf (a, b)σ
f it
x
x
PL1 (0.953,0.010) (0.959,0.013) (0.928,3.688) (1.000,-0.073) (0.593,0.354) (0.616,0.355) 0.004
PL2 (0.914,-0.008) (0.873,-0.024) (1.013,-0.836) (0.989,0.292) (0.689,0.299) (0.643,0.341) 0.002
PL3 (0.880,-0.052) (0.937,-0.013) (0.965,1.729) (1.008,-0.513) (0.683,0.340) (0.664,0.352) 0.003
PL4 (0.946,0.024) (0.964,0.024) (0.995,-0.076) (1.086,-4.905) (0.614,0.364) (0.585,0.380) 0.006
PL5 (0.916,-0.030) (0.962,0.004) (1.033,-2.067) (0.828,9.095) (0.716,0.333) (0.679,0.356) 0.005
Table 3: Slope aand zero point bof the fitted vs input parameters for both the best fit and median
values (labeled with subscripts bf and f it, respectively). The upperscript denotes the parameter fitted with
(τ,α0,σ) referring to (ατ,α0,σint ), respectively. x
xis the bias for each efficiency function considered.
– 21 –
negligible small notwithstanding the values of the parameters setting Eλ. This is particularly
true if one relies on the median values as estimate since they are typically consistent with
the no bias line within less than 2σ.
The above results have been obtained considering a power-law Eλso that it is worth inves-
tigating whether they critically depend on this assumption. We have therefore repeated the
analysis for the polynomial Eλmodels in Table 2 obtaining the results in Table 4. A compar-
ison with the values in Table 4 shows that the (a, b) coefficients are similar so that one could
preliminarily conclude that the shape of the efficiency function does not play a major role
in the determination of the bias. Actually, although the functional expressions are different,
both the power-law and the polynomial selection functions are qualitatively similar with Eλ
increasing with λover a comparable range. Although such a behaviour is likely common to
any reasonable Eλ, we can not exclude a priori that non monotonic selection functions do
actually exist. What would the results be in such a case is not clear so that we prefer to be
cautious and conclude that the bias is roughly the same whichever monotonic Eλ(λ) func-
tion is used, but not for all the possible Eλfunctions. For non-monotonic shape of selection
function, see Stern et al. (2001), in which an assumed detection efficiency function, defined
as the ratio of the number of detected test bursts to the number of test bursts applied to
the data versus the expected peak count rate, is given by:
E(ce) = 0.70 ×[1 exp[(ce
ce,0
)2]]ν,(22)
where ce,0= 0.097 counts s1cm2and ν= 2.34 are two constants. However, quoting
from Stern et al. (2001), the best possible efficiency quality has still not yet been achieved
because in fact the detection efficiency depends on the peak count rate rather then on the
time-integrated signal.
Id (a, b)τ
bf (a, b)τ
f it (a, b)α0
bf (a, b)α0
f it (a, b)σ
bf (a, b)σ
f it
x
x
PoL1 (0.950,0.020) (0.928,0.011) (1.099,-5.545) (1.178,-9.935) (0.647,0.350) (0.673,0.349) 0.004
PoL2 (1.095,0.128) (1.075,0.105) (1.030,-1.662) (1.023,-1.278) (0.681,0.337) (0.625,0.372) 0.0008
PoL3 (0.984,0.063) (0.936,0.015) (0.711,15.258) (0.988,0.376) (0.741,0.289) (0.685,0.336) 0.007
PoL4 (0.870,-0.025) (0.969,0.052) (0.730,14.103) (0.734,13.872) (0.630,0.351) (0.582,0.381) 0.009
PoL5 (1.004,0.069) (0.972,0.030) (0.963,11.772) (1.002,-0.374) (0.581,0.384) (0.549,0.402) 0.18
Table 4: Same as Table 3 but for the polynomial Eλmodel.
– 22 –
5.1.2. Selection cuts on both τand λ
We now consider the case where the total selection function may be factorized as
E(τ, λ) = Eτ(τ)Eλ(λ) with both Ex(x) functions being given by power-law or fourth or-
der polynomial expressions. We consider 10 different arbitrary choices for both cases. Note
that we have to increase Nsim to 300 in order to have Nobs = 80 100 as for the models
discussed in the previous subsection.
Table 5 gives the (a,b) coefficients for the different models considered. A comparison with
Table 3 shows that, on average, the bias on the parameters is roughly the same with the
median values giving smaller deviations and significant bias on σint only. A more detailed
analysis, however, shows that, while, in the Eτ= 1 case, biases larger than 5% are of the
order of 10%. Namely, from Table 3 and 4 we show that the relative biases, x
x, both in
the linear and the polynomial case, give very small values from 0.2% to 0.9%, with the only
exception of 1 polynomial function, in which the bias is 18%, thus giving P= 10% of having
larger bias than 5%. If we consider selection cuts on both τand λwe notice in Table 5 that
number of bias whose value is greater than 5% are 4, thus increasing their probability to
occur (P40%). This can be qualitatively explained noting that a cut on λonly removes
the points in the luminosity axis thus possibly shifting the best fit relation, but not changing
the slope. On the contrary, removing points also along the horizontal τaxis can change
the slope ατtoo and hence also affects (α0,σint) because of the correlation among these
parameters and ατ. Similar results are obtained when both Exfunctions are modeled with
fourth order polynomials so that we will not discuss this case here. We stress that when x
x
are larger than 6%, than the slope of the correlation is farther from 1.0 compared to cases
in which x
x0.06. In fact, in the first case the slopes values range from 0.86 to 0.91, see
the functions P LT a4 and P LT a6 in Table 5. These values are not compatible in 1 σwith
the claimed intrinsic slope of the correlation, 1.07+0.09
0.14. If we consider, instead, the lowest
x
x, then we obtain ranges of a= (0.94,0.99) thus showing full compatibility in 1 σwith the
intrinsic slope. In this way we have quantitatively confirmed the existence of the LXT
a
correlation with the same intrinsic slope as in Dainotti et al. (2013a) if appropriate selection
functions are chosen.
6. Conclusions
Here we built a general method to evaluate selection effects for GRB correlations not
knowing a priori the efficiency function of the detector used. We have tested this method
on the LT correlation. We chose a set of GRBs and assuming Gaussian distributions for
the variables involved, for luminosity and time, and also a particular shape for the GRBs
– 23 –
rate density. We simulated a mock sample of data in order to consider the selection ef-
fects of the detectors. As we can see in paragraph 3 assuming the correct observed GRBs
rate density shape was not an easy task. In fact, we explored different methods (Li 2008;
Robertson and Ellis 2012; Kistler et al. 2013; Hopkins and Beacom 2006) that use several
SFR shapes to understand which one best matches the afterglow plateau data distribution
including the selection functions, see Fig. 4. The most reliable fits for the GRB plateau
data is the SFR used by Li (2008), while the best efficiency functions for ǫ(λ) that match
the GRB density rate can be both two polynomial and two linear, see Fig. 4. Table 1 and
2 show the probability that the density rate fits the afterglow plateau GRB rate assuming
those efficiency functions. However, we assumed there could be selection effects both for
luminosity and time. In particular, the bias is roughly the same whichever monotonic effi-
ciency function for the luminosity detection Eλis taken. From Table 3 and 4 we show that
the relative biases, x
x, both in the linear and the polynomial case, give very small values
from 0.2% to 0.9%, with the only exception of 1 polynomial function, in which the bias is
18%, thus giving P= 10% of having larger bias than 5%. If we consider selection cuts on
both τand λwe notice in Table 5 that number of bias whose value is greater than 5% are
4, thus increasing the probability of having such biases (P40%). In addition, we studied
selection effects in the LT correlation assuming also a combination of the luminosity and time
detection efficiency functions. Different values for the parameters of the efficiency functions
in the detectors are taken into account as described in the paragraph 5. This gives distinct
fit values that inserted in Equation 20 allow to study the scattering of the correlation and its
selection effects. We have quantitatively confirmed the existence of the LXT
acorrelation
with the same intrinsic slope as in Dainotti et al. (2013a) if appropriate selection functions
are chosen. In particular, when x
xare larger than 6%, than the slope of the correlation is
farther from 1.0 compared to cases in which x
x0.06. The lowest x
xleads to ranges
of a= (0.94,0.99) thus showing full compatibility in 1 σwith the intrinsic slope. Finally,
the fact that the correlation is not generated by the biases themselves is a significant and
further step towards considering a set of GRBs as standard candles and their possible and
useful application as a cosmological tool.
7. Acknowledgements
This work made use of data supplied by the UK Swift Science Data Centre at the Uni-
versity of Leicester. We are particularly grateful to Cardone, V.F. for the initial contribution
to this work. M.G.D. and S.N. are grateful to the iTHES Group discussions at Riken. M.D
is grateful to the support from the JSPS Foundation, (No. 25.03786). N. S. is grateful to
JSPS (No.24.02022, No.25.03018, No.25610056, No.26287056) & MEXT(No.26105521), R.
– 24 –
D.V. is grateful to 2012/04/A/ST9/00083.
REFERENCES
L. Amati, F. Frontera, and C. Guidorzi. Extremely energetic Fermi gamma-ray bursts
obey spectral energy correlations. A&A, 508:173–180, December 2009. doi: 10.1051/
0004-6361/200912788.
D. L. Band. Postlaunch Analysis of Swift’s Gamma-Ray Burst Detection Sensitivity. ApJ,
644:378–384, June 2006. doi: 10.1086/503326.
M. G. Bernardini, R. Margutti, J. Mao, E. Zaninoni, and G. Chincarini. The X-ray light
curve of gamma-ray bursts: clues to the central engine. A&A, 539:A3, March 2012a.
doi: 10.1051/0004-6361/201117895.
M. G. Bernardini, R. Margutti, E. Zaninoni, and G. Chincarini. A universal scaling for short
and long gamma-ray bursts: EX,iso - E,iso - Epk.MNRAS, 425:1199–1204, September
2012b. doi: 10.1111/j.1365-2966.2012.21487.x.
J. S. Bloom. Is the Redshift Clustering of Long-Duration Gamma-Ray Bursts Significant?
AJ, 125:2865–2875, June 2003. doi: 10.1086/374945.
N. R. Butler, J. S. Bloom, and D. Poznanski. The Cosmic Rate, Luminosity Function, and
Intrinsic Correlations of Long Gamma-Ray Bursts. ApJ, 711:495–516, March 2010.
doi: 10.1088/0004-637X/711/1/495.
J. I. Cabrera, C. Firmani, V. Avila-Reese, G. Ghirlanda, G. Ghisellini, and L. Nava. Spectral
analysis of Swift long gamma-ray bursts with known redshift. MNRAS, 382:342–355,
November 2007. doi: 10.1111/j.1365-2966.2007.12374.x.
J. K. Cannizzo and N. Gehrels. A New Paradigm for Gamma-ray Bursts: Long-term Accre-
tion Rate Modulation by an External Accretion Disk. ApJ, 700:1047–1058, August
2009. doi: 10.1088/0004-637X/700/2/1047.
J. K. Cannizzo, E. Troja, and N. Gehrels. Fall-back Disks in Long and Short Gamma-Ray
Bursts. ApJ, 734:35, June 2011. doi: 10.1088/0004-637X/734/1/35.
V. F. Cardone, S. Capozziello, and M. G. Dainotti. An updated gamma-ray bursts Hubble
diagram. MNRAS, 400:775–790, December 2009. doi: 10.1111/j.1365-2966.2009.
15456.x.
– 25 –
V. F. Cardone, M. G. Dainotti, S. Capozziello, and R. Willingale. Constraining cosmological
parameters by gamma-ray burst X-ray afterglow light curves. MNRAS, 408:1181–
1186, October 2010. doi: 10.1111/j.1365-2966.2010.17197.x.
D. Coward. Open issues with the gamma-ray burst redshift distribution. New A Rev., 51:
539–546, September 2007. doi: 10.1016/j.newar.2007.03.003.
A. Cucchiara, A. J. Levan, D. B. Fox, N. R. Tanvir, T. N. Ukwatta, E. Berger, T. Kr¨uhler,
A. K¨upc¨u Yolda¸s, X. F. Wu, K. Toma, J. Greiner, F. E. Olivares, A. Rowlinson,
L. Amati, T. Sakamoto, K. Roth, A. Stephens, A. Fritz, J. P. U. Fynbo, J. Hjorth,
D. Malesani, P. Jakobsson, K. Wiersema, P. T. O’Brien, A. M. Soderberg, R. J.
Foley, A. S. Fruchter, J. Rhoads, R. E. Rutledge, B. P. Schmidt, M. A. Dopita,
P. Podsiadlowski, R. Willingale, C. Wolf, S. R. Kulkarni, and P. D’Avanzo. A
Photometric Redshift of z ˜ 9.4 for GRB 090429B. ApJ, 736:7, July 2011. doi:
10.1088/0004-637X/736/1/7.
G. D’Agostini. Fits, and especially linear fits, with errors on both axes, extra variance of
the data points and other complications. ArXiv Physics e-prints, November 2005.
F. Daigne and R. Mochkovitch. The low-luminosity tail of the GRB distribution: the case
of GRB 980425. A&A, 465:1–8, April 2007. doi: 10.1051/0004-6361:20066080.
M. G. Dainotti, V. F. Cardone, and S. Capozziello. A time-luminosity correlation for γ-
ray bursts in the X-rays. MNRAS, 391:L79–L83, November 2008. doi: 10.1111/j.
1745-3933.2008.00560.x.
M. G. Dainotti, R. Willingale, S. Capozziello, V. Fabrizio Cardone, and M. Ostrowski.
Discovery of a Tight Correlation for Gamma-ray Burst Afterglows with ”Canonical”
Light Curves. ApJ, 722:L215–L219, October 2010. doi: 10.1088/2041-8205/722/2/
L215.
M. G. Dainotti, V. Fabrizio Cardone, S. Capozziello, M. Ostrowski, and R. Willingale. Study
of Possible Systematics in the L*X-T*aCorrelation of Gamma-ray Bursts. ApJ, 730:
135, April 2011a. doi: 10.1088/0004-637X/730/2/135.
M. G. Dainotti, M. Ostrowski, and R. Willingale. Towards a standard gamma-ray burst: tight
correlations between the prompt and the afterglow plateau phase emission. MNRAS,
418:2202–2206, December 2011b. doi: 10.1111/j.1365-2966.2011.19433.x.
M. G. Dainotti, V. F. Cardone, E. Piedipalumbo, and S. Capozziello. Slope evolution of
GRB correlations and cosmology. MNRAS, 436:82–88, November 2013a. doi: 10.
1093/mnras/stt1516.
– 26 –
M. G. Dainotti, V. Petrosian, J. Singal, and M. Ostrowski. Determination of the Intrinsic
Luminosity Time Correlation in the X-Ray Afterglows of Gamma-Ray Bursts. ApJ,
774:157, September 2013b. doi: 10.1088/0004-637X/774/2/157.
S. Dall’Osso, G. Stratta, D. Guetta, S. Covino, G. De Cesare, and L. Stella. Gamma-ray
bursts afterglows with energy injection from a spinning down neutron star. A&A,
526:A121, February 2011. doi: 10.1051/0004-6361/201014168.
B. Efron and V. Petrosian. A simple test of independence for truncated data with applications
to redshift surveys. ApJ, 399:345–352, November 1992. doi: 10.1086/171931.
E. E. Fenimore and E. Ramirez-Ruiz. Redshifts For 220 BATSE Gamma-Ray Bursts De-
termined by Variability and the Cosmological Consequences. ArXiv Astrophysics
e-prints, April 2000.
F. Fiore, D. Guetta, S. Piranomonte, V. D’Elia, and L. A. Antonelli. Selection effects shaping
the gamma ray burst redshift distributions. A&A, 470:515–522, August 2007. doi:
10.1051/0004-6361:20077157.
G. Ghirlanda, G. Ghisellini, and D. Lazzati. The Collimation-corrected Gamma-Ray Burst
Energies Correlate with the Peak Energy of Their νFnu Spectrum. ApJ, 616:331–338,
November 2004. doi: 10.1086/424913.
G. Ghirlanda, G. Ghisellini, and C. Firmani. Gamma-ray bursts as standard candles to
constrain the cosmological parameters. New Journal of Physics, 8:123, July 2006.
doi: 10.1088/1367-2630/8/7/123.
G. Ghisellini, M. Nardini, G. Ghirlanda, and A. Celotti. A unifying view of gamma-ray
burst afterglows. MNRAS, 393:253–271, February 2009. doi: 10.1111/j.1365-2966.
2008.14214.x.
D. Guetta and M. Della Valle. On the Rates of Gamma-Ray Bursts and Type Ib/c Super-
novae. ApJ, 657:L73–L76, March 2007. doi: 10.1086/511417.
R. Hasco¨et, F. Daigne, and R. Mochkovitch. The prompt-early afterglow connection in
gamma-ray bursts: implications for the early afterglow physics. MNRAS, 442:20–27,
July 2014. doi: 10.1093/mnras/stu750.
A. M. Hopkins and J. F. Beacom. On the Normalization of the Cosmic Star Formation
History. ApJ, 651:142–154, November 2006. doi: 10.1086/506610.
– 27 –
L. Izzo, G. B. Pisani, M. Muccino, J. A. Rueda, Y. Wang, C. L. Bianco, A. V. Penacchioni,
and R. Ruffini. A common behavior in the late X-ray afterglow of energetic GRB-
SN systems. In A. J. Castro-Tirado, J. Gorosabel, and I. H. Park, editors, EAS
Publications Series, volume 61 of EAS Publications Series, pages 595–597, July 2013.
doi: 10.1051/eas/1361095.
P. Jakobsson, A. Levan, J. P. U. Fynbo, R. Priddey, J. Hjorth, N. Tanvir, D. Watson, B. L.
Jensen, J. Sollerman, P. Natarajan, J. Gorosabel, J. M. Castro Cer´on, K. Pedersen,
T. Pursimo, A. S. ´
Arnad´ottir, A. J. Castro-Tirado, C. J. Davis, H. J. Deeg, D. A.
Fiuza, S. Mikolaitis, and S. G. Sousa. A mean redshift of 2.8 for Swift gamma-ray
bursts. A&A, 447:897–903, March 2006. doi: 10.1051/0004-6361:20054287.
M. D. Kistler, H. Y¨uksel, J. F. Beacom, and K. Z. Stanek. An Unexpectedly Swift Rise in the
Gamma-Ray Burst Rate. ApJ, 673:L119–L122, February 2008. doi: 10.1086/527671.
M. D. Kistler, H. Yuksel, and A. M. Hopkins. The Cosmic Star Formation Rate from the
Faintest Galaxies in the Unobservable Universe. ArXiv e-prints, May 2013.
D. Kocevski and V. Petrosian. On the Lack of Time Dilation Signatures in Gamma-Ray
Burst Light Curves. ApJ, 765:116, March 2013. doi: 10.1088/0004-637X/765/2/116.
T. Le and C. D. Dermer. On the Redshift Distribution of Gamma-Ray Bursts in the Swift
Era. ApJ, 661:394–415, May 2007. doi: 10.1086/513460.
K. Leventis, R. A. M. J. Wijers, and A. J. van der Horst. The plateau phase of gamma-ray
burst afterglows in the thick-shell scenario. MNRAS, 437:2448–2460, January 2014.
doi: 10.1093/mnras/stt2055.
L.-X. Li. The X-ray transient 080109 in NGC 2770: an X-ray flash associated with a
normal core-collapse supernova. MNRAS, 388:603–610, August 2008. doi: 10.1111/j.
1365-2966.2008.13461.x.
N. M. Lloyd and V. Petrosian. Distribution of Spectral Characteristics and the Cosmological
Evolution of Gamma-Ray Bursts. ApJ, 511:550–561, February 1999. doi: 10.1086/
306719.
J. Mao. A Theoretical Investigation of Gamma-ray Burst Host Galaxies. ApJ, 717:140–146,
July 2010. doi: 10.1088/0004-637X/717/1/140.
S. Mao and H. J. Mo. The nature of the host galaxies for gamma-ray bursts. A&A, 339:
L1–L4, November 1998.
– 28 –
P. Natarajan, B. Albanna, J. Hjorth, E. Ramirez-Ruiz, N. Tanvir, and R. Wijers. The
redshift distribution of gamma-ray bursts revisited. MNRAS, 364:L8–L12, November
2005. doi: 10.1111/j.1745-3933.2005.00094.x.
R. S. Nemmen, M. Georganopoulos, S. Guiriec, E. T. Meyer, N. Gehrels, and R. M. Sam-
bruna. A Universal Scaling for the Energetics of Relativistic Jets from Black Hole
Systems. Science, 338:1445–, December 2012. doi: 10.1126/science.1227416.
J. P. Norris, G. F. Marani, and J. T. Bonnell. Connection between Energy-dependent Lags
and Peak Luminosity in Gamma-Ray Bursts. ApJ, 534:248–257, May 2000. doi:
10.1086/308725.
P. T. O’Brien, R. Willingale, J. Osborne, M. R. Goad, K. L. Page, S. Vaughan, E. Rol,
A. Beardmore, O. Godet, C. P. Hurkett, A. Wells, B. Zhang, S. Kobayashi, D. N. Bur-
rows, J. A. Nousek, J. A. Kennea, A. Falcone, D. Grupe, N. Gehrels, S. Barthelmy,
J. Cannizzo, J. Cummings, J. E. Hill, H. Krimm, G. Chincarini, G. Tagliaferri,
S. Campana, A. Moretti, P. Giommi, M. Perri, V. Mangano, and V. LaParola.
The Early X-Ray Emission from GRBs. ApJ, 647:1213–1237, August 2006. doi:
10.1086/505457.
V. Petrosian. New Statistical Methods for Analysis of Large Surveys: Distributions and
Correlations. In R. F. Green, E. Y. Khachikian, and D. B. Sanders, editors, IAU Col-
loq. 184: AGN Surveys, volume 284 of Astronomical Society of the Pacific Conference
Series, page 389, 2002.
V. Petrosian and N. M. Lloyd. The nu Fnu Peak Energy of Gamma Ray Bursts, and its
Correlation with Fluence and Peak Flux. In American Astronomical Society Meeting
Abstracts, volume 30 of Bulletin of the American Astronomical Society, page 1380,
December 1998.
V. Petrosian, N. Lloyd, and A. Lee. Cosmological Signatures in Temporal and Spectral
Characteristics of Gamma-Ray Bursts. In J. Poutanen and R. Svensson, editors,
Gamma-Ray Bursts: The First Three Minutes, volume 190 of Astronomical Society
of the Pacific Conference Series, page 235, 1999.
C. Porciani and P. Madau. On the Association of Gamma-Ray Bursts with Massive Stars:
Implications for Number Counts and Lensing Statistics. ApJ, 548:522–531, February
2001. doi: 10.1086/319027.
S. Postnikov, M. G. Dainotti, X. Hernandez, and S. Capozziello. Nonparametric Study of the
Evolution of the Cosmological Equation of State with SNeIa, BAO, and High-redshift
GRBs. ApJ, 783:126, March 2014. doi: 10.1088/0004-637X/783/2/126.
– 29 –
S. Qi and T. Lu. Toward Tight Gamma-Ray Burst Luminosity Relations. ApJ, 749:99, April
2012. doi: 10.1088/0004-637X/749/2/99.
B. E. Robertson and R. S. Ellis. Connecting the Gamma Ray Burst Rate and the Cosmic
Star Formation History: Implications for Reionization and Galaxy Evolution. ApJ,
744:95, January 2012. doi: 10.1088/0004-637X/744/2/95.
A. Rowlinson, P. T. O’Brien, N. R. Tanvir, B. Zhang, P. A. Evans, N. Lyons, A. J. Levan,
R. Willingale, K. L. Page, O. Onal, D. N. Burrows, A. P. Beardmore, T. N. Uk-
watta, E. Berger, J. Hjorth, A. S. Fruchter, R. L. Tunnicliffe, D. B. Fox, and
A. Cucchiara. The unusual X-ray emission of the short Swift GRB 090515: evi-
dence for the formation of a magnetar? MNRAS, 409:531–540, December 2010. doi:
10.1111/j.1365-2966.2010.17354.x.
A. Rowlinson, P. T. O’Brien, B. D. Metzger, N. R. Tanvir, and A. J. Levan. Signatures of
magnetar central engines in short GRB light curves. MNRAS, 430:1061–1087, April
2013. doi: 10.1093/mnras/sts683.
A. Rowlinson, B. P. Gompertz, M. Dainotti, P. T. O’Brien, R. A. M. J. Wijers, and A. J. van
der Horst. Constraining properties of GRB magnetar central engines using the ob-
served plateau luminosity and duration correlation. MNRAS, 443:1779–1787, Septem-
ber 2014. doi: 10.1093/mnras/stu1277.
T. Sakamoto, J. E. Hill, R. Yamazaki, L. Angelini, H. A. Krimm, G. Sato, S. Swindell,
K. Takami, and J. P. Osborne. Evidence of Exponential Decay Emission in the Swift
Gamma-Ray Bursts. ApJ, 669:1115–1129, November 2007. doi: 10.1086/521640.
B. E. Schaefer. Explaining the Gamma-Ray Burst Epeak Distribution. ApJ, 583:L71–L74,
February 2003. doi: 10.1086/368106.
A. Shahmoradi and R. Nemiroff. How Real Detector Thresholds Create False Standard
Candles. In C. Meegan, C. Kouveliotou, and N. Gehrels, editors, American Institute of
Physics Conference Series, volume 1133 of American Institute of Physics Conference
Series, pages 425–427, May 2009. doi: 10.1063/1.3155940.
B. E. Stern, Y. Tikhomirova, D. Kompaneets, R. Svensson, and J. Poutanen. An Off-
Line Scan of the BATSE Daily Records and a Large Uniform Sample of Gamma-Ray
Bursts. ApJ, 563:80–94, December 2001. doi: 10.1086/322295.
J. Sultana, D. Kazanas, and K. Fukumura. Luminosity Correlations for Gamma-Ray Bursts
and Implications for Their Prompt and Afterglow Emission Mechanisms. ApJ, 758:
32, October 2012. doi: 10.1088/0004-637X/758/1/32.
– 30 –
T. Totani. Cosmological Gamma-Ray Bursts and Evolution of Galaxies. ApJ, 486:L71–L74,
September 1997. doi: 10.1086/310853.
H. van Eerten. Self-similar relativistic blast waves with energy injection. MNRAS, 442:
3495–3510, August 2014a. doi: 10.1093/mnras/stu1025.
H. J. van Eerten. Gamma-ray burst afterglow plateau break time-luminosity correlations
favour thick shell models over thin shell models. MNRAS, 445:2414–2423, December
2014b. doi: 10.1093/mnras/stu1921.
R. A. M. J. Wijers, J. S. Bloom, J. S. Bagla, and P. Natarajan. Gamma-ray bursts from stellar
remnants - Probing the universe at high redshift. MNRAS, 294:L13–L17, February
1998. doi: 10.1046/j.1365-8711.1998.01328.x.
R. Willingale, P. T. O’Brien, J. P. Osborne, O. Godet, K. L. Page, M. R. Goad, D. N.
Burrows, B. Zhang, E. Rol, N. Gehrels, and G. Chincarini. Testing the Standard
Fireball Model of Gamma-Ray Bursts Using Late X-Ray Afterglows Measured by
Swift. ApJ, 662:1093–1110, June 2007. doi: 10.1086/517989.
R. Yamazaki. Prior Emission Model for X-ray Plateau Phase of Gamma-Ray Burst After-
glows. ApJ, 690:L118–L121, January 2009. doi: 10.1088/0004-637X/690/2/L118.
D. Yonetoku, T. Murakami, T. Nakamura, R. Yamazaki, A. K. Inoue, and K. Ioka. Gamma-
Ray Burst Formation Rate Inferred from the Spectral Peak Energy-Peak Luminosity
Relation. ApJ, 609:935–951, July 2004. doi: 10.1086/421285.
B. Yu, S. Qi, and T. Lu. Gamma-ray Burst Luminosity Relations: Two-Dimensional Versus
Three-Dimensional Correlations. ApJ, 705:L15–L19, November 2009. doi: 10.1088/
0004-637X/705/1/L15.
H. Y¨uksel and M. D. Kistler. Enhanced cosmological GRB rates and implications for cos-
mogenic neutrinos. Phys. Rev. D, 75(8):083004, April 2007. doi: 10.1103/PhysRevD.
75.083004.
This preprint was prepared with the AAS L
A
T
E
X macros v5.2.
– 31 –
Id (a, b)τ
bf (a, b)τ
f it (a, b)α0
bf (a, b)α0
f it (a, b)σ
bf (a, b)σ
f it
x
x
PLTa1 (0.842,-0.080) (0.920,-0.016) (1.099,-5.615) (1.064,-3.650) (0.615,0.363) (0.585,0.387) 0.04
PLTa2 (0.972,-0.016) (0.963,-0.027) (1.112,-5.980) (1.148,-7.746) (0.720,0.306) (0.674,0.335) 0.005
PLTa3 (0.960,-0.025) (0.974,-0.005) (0.985,0.791) (1.022,-1.228) (0.725,0.329) (0.690,0.345) 0.004
PLTa4 (0.877,-0.021) (0.933,-0.005) (1.005,-0.611) (0.959,2.196) (0.702,0.304) (0.644,0.345) 0.09
PLTa5 (0.860,-0.056) (0.901,-0.026) (0.911,4.592) (0.838,8.655) (0.687,0.338) (0.613,0.382) 0.06
PLTa6 (0.904,-0.010) (0.919,-0.003) (0.734,14.060) (0.397,32.119) (0.678,0.326) (0.680,0.330) 0.08
PLTa7 (0.915,-0.044) (0.998,0.035) (1.119,-6.577) (1.119,-6.474) (0.736,0.311) (0.722,0.327) 0.02
PLTa8 (0.967,-0.003) (0.961,-0.017) (1.027,-1.485) (1.178,-9.515) (0.731,0.287) (0.705,0.313) 0.03
PLTa9 (0.933,-0.005) (0.904,-0.031) (1.026,-1.530) (0.804,10.289) (0.641,0.357) (0.670,0.352) 0.06
PLTa10 (1.040,0.056) (0.997,0.028) (0.753,13.308) (0.822,9.345) (0.736,0.329) (0.713,0.344) 0.04
Table 5: Same as Table 3 but for the selection cuts on both (τ,λ) and power-law Exfunctions.
... Another great advantage of having more GRBs with redshift is the possibility of using GRBs as standardized candles with empirical relations between distance-dependent and intrinsic properties of GRBs. Among the earliest of these efforts, is the Dainotti relation (Dainotti et al. 2008(Dainotti et al. , 2015(Dainotti et al. , 2017, a roughly inversely proportional relationship between the rest-frame time at the end of the plateau phase (T a /(1 + z)) and its corresponding luminosity (L a ). Later Dainotti et al. (2013) showed that via the use of the Efron and Petrosian method (Efron & Petrosian 1992), this relation is intrinsic and not due to selection biases. ...
Article
Full-text available
Gamma-ray bursts (GRBs), due to their high luminosities, are detected up to a redshift of 10, and thus have the potential to be vital cosmological probes of early processes in the Universe. Fulfilling this potential requires a large sample of GRBs with known redshifts, but due to observational limitations, only 11% have known redshifts ( z ). There have been numerous attempts to estimate redshifts via correlation studies, most of which have led to inaccurate predictions. To overcome this, we estimated GRB redshift via an ensemble-supervised machine-learning (ML) model that uses X-ray afterglows of long-duration GRBs observed by the Neil Gehrels Swift Observatory. The estimated redshifts are strongly correlated (a Pearson coefficient of 0.93) and have an rms error, namely, the square root of the average squared error 〈Δ z ² 〉, of 0.46 with the observed redshifts showing the reliability of this method. The addition of GRB afterglow parameters improves the predictions considerably by 63% compared to previous results in peer-reviewed literature. Finally, we use our ML model to infer the redshifts of 154 GRBs, which increase the known redshifts of long GRBs with plateaus by 94%, a significant milestone for enhancing GRB population studies that require large samples with redshift.
... Our estimation illustrates the difficulties present in these predictions based on biased samples with uncertain selection effects, as has been discussed by Dainotti et al. ( 2015 ) and Dainotti, Petrosian & Bowden ( 2021 ). ...
Article
We collected the optical light curve data of 227 gamma-ray bursts (GRBs) observed with the TAROT, COATLI, and RATIR telescopes. These consist of 133 detections and 94 upper limits. We constructed average light curves in the observer and rest frames in both X-rays (from Swift/XRT) and in the optical. Our analysis focused on investigating the observational and intrinsic properties of GRBs. Specifically, we examined observational properties, such as the optical brightness function of the GRBs at T = 1000 seconds after the trigger, as well as the temporal slope of the afterglow. We also estimated the redshift distribution for the GRBs within our sample. Of the 227 GRBs analysed, we found that 116 had a measured redshift. Based on these data, we calculated a local rate of ρ0 = 0.2 Gpc−3 yr−1 for these events with z < 1. To explore the intrinsic properties of GRBs, we examined the average X-ray and optical light curves in the rest frame. We use the afterglowpy library to generate synthetic curves to constrain the parameters typical of the bright GRB jet, such as energy (〈E0〉 ∼ 1053.6 erg), opening angle (〈θcore〉 ∼ 0.2 rad), and density (〈n0〉 ∼ 10−2.1 cm−3). Furthermore, we analyse microphysical parameters, including the fraction of thermal energy in accelerated electrons (〈εe〉 ∼ 10−1.37) and in the magnetic field (〈εB〉 ∼ 10−2.26), and the power-law index of the population of non-thermal electrons (〈p〉 ∼ 2.2).
... Our estimation illustrates the difficulties present in these predictions based on biased samples with uncertain selection effects, as has been discussed by Dainotti et al. (2021Dainotti et al. ( , 2015. ...
Preprint
We collected the optical light curve data of 227 gamma-ray bursts (GRBs) observed with the TAROT, COATLI, and RATIR telescopes. These consist of 133 detections and 94 upper limits. We constructed average light curves in the observer and rest frames in both X-rays (from {\itshape Swift}/XRT) and in the optical. Our analysis focused on investigating the observational and intrinsic properties of GRBs. Specifically, we examined observational properties, such as the optical brightness function of the GRBs at $T=1000$ seconds after the trigger, as well as the temporal slope of the afterglow. We also estimated the redshift distribution for the GRBs within our sample. Of the 227 GRBs analysed, we found that 116 had a measured redshift. Based on these data, we calculated a local rate of $\rho_0=0.2$ Gpc$^{-3}$ yr$^{-1}$ for these events with $z<1$. To explore the intrinsic properties of GRBs, we examined the average X-ray and optical light curves in the rest frame. We use the {\scshape afterglowpy} library to generate synthetic curves to constrain the parameters typical of the bright GRB jet, such as energy (${\langle} {E_{0}}{\rangle}\sim 10^{53.6}$~erg), opening angle (${\langle}\theta_\mathrm{core}{\rangle}\sim 0.2$~rad), and density (${\langle}n_\mathrm{0}{\rangle}\sim10^{-2.1}$ cm$^{-3}$). Furthermore, we analyse microphysical parameters, including the fraction of thermal energy in accelerated electrons (${\langle}\epsilon_e{\rangle}\sim 10^{-1.37}$) and in the magnetic field (${\langle}\epsilon_B{\rangle}\sim10^{-2.26}$), and the power-law index of the population of non-thermal electrons (${\langle}p{\rangle}\sim 2.2$).
... The plateau features have attracted attention due to their use in building relevant correlations with the plateau parameters and their application as cosmological tools. Specifically, Dainotti et al. (2008Dainotti et al. ( , 2010bDainotti et al. ( , 2011; ; Dainotti et al. (2015Dainotti et al. ( , 2017 and Li et al. (2018b) explored the luminosity at the end of the plateau, L X,a vs. its rest-frame time T * X,a (known as the Dainotti relation or 2D L-T relation) 1 . The 2D relation has also been discovered in optical plateau emissions (Dainotti et al. 2020b. ...
Article
Full-text available
Gamma-ray bursts (GRBs), as they are observed at high redshift ( z = 9.4), are vital to cosmological studies and investigating Population III stars. To tackle these studies, we need correlations among relevant GRB variables with the requirement of small uncertainties on their variables. Thus, we must have good coverage of GRB light curves (LCs). However, gaps in the LC hinder the precise determination of GRB properties and are often unavoidable. Therefore, extensive categorization of GRB LCs remains a hurdle. We address LC gaps using a stochastic reconstruction, wherein we fit two preexisting models (the Willingale model; W07; and a broken power law; BPL) to the observed LC, then use the distribution of flux residuals from the original data to generate data to fill in the temporal gaps. We also demonstrate a model-independent LC reconstruction via Gaussian processes. At 10% noise, the uncertainty of the end time of the plateau, its correspondent flux, and the temporal decay index after the plateau decreases by 33.3%, 35.03%, and 43.32% on average for the W07, and by 33.3%, 30.78%, 43.9% for the BPL, respectively. The uncertainty of the slope of the plateau decreases by 14.76% in the BPL. After using the Gaussian process technique, we see similar trends of a decrease in uncertainty for all model parameters for both the W07 and BPL models. These improvements are essential for the application of GRBs as standard candles in cosmology, for the investigation of theoretical models, and for inferring the redshift of GRBs with future machine-learning analyses.
... The plateau features have attracted attention due to their use in building relevant correlations with the plateau parameters and their application as cosmological tools. Specifically, Dainotti et al. (2008Dainotti et al. ( , 2010bDainotti et al. ( , 2011; ; Dainotti et al. (2015Dainotti et al. ( , 2017 and Li et al. (2018b) explored the luminosity at the end of the plateau, L X,a vs. its rest-frame time T * X,a (known as the Dainotti relation or 2D L-T relation) 1 . The 2D relation has also been discovered in optical plateau emissions (Dainotti et al. 2020b. ...
Preprint
Full-text available
Gamma-Ray Bursts (GRBs), being observed at high redshift (z = 9.4), vital to cosmological studies and investigating Population III stars. To tackle these studies, we need correlations among relevant GRB variables with the requirement of small uncertainties on their variables. Thus, we must have good coverage of GRB light curves (LCs). However, gaps in the LC hinder the precise determination of GRB properties and are often unavoidable. Therefore, extensive categorization of GRB LCs remains a hurdle. We address LC gaps using a 'stochastic reconstruction,' wherein we fit two pre-existing models (Willingale 2007; W07 and Broken Power Law; BPL) to the observed LC, then use the distribution of flux residuals from the original data to generate data to fill in the temporal gaps. We also demonstrate a model-independent LC reconstruction via Gaussian Processes. At 10% noise, the uncertainty of the end time of the plateau, its correspondent flux, and the temporal decay index after the plateau decreases, on average, by 33.3% 35.03%, and 43.32%, respectively for the W07, and by 33.3%, 30.78%, 43.9% for the BPL. The slope of the plateau decreases by 14.76% in the BPL. After using the Gaussian Process technique, we see similar trends of a decrease in uncertainty for all model parameters for both the W07 and BPL models. These improvements are essential for the application of GRBs as standard candles in cosmology, for the investigation of theoretical models and for inferring the redshift of GRBs with future machine learning analysis.
... The measured quantities for the GRBs are the redshift z, the characteristic time scale T * X which marks the end of the plateau emission, the measured X-ray energy flux F X at T * X , the measured γ-ray energy flux F peak in the peak of the prompt emission over a 1 s interval, and the X-ray photon indices of the plateau phase α plateau and of the prompt emission α prompt . We make use of the 50 Platinum GRBs, spanning the redshift range 0.553 ≤ z ≤ 5.0, introduced in Dainotti et al. (2020), that we previously studied (Cao et al., 2022c), and a new LGRB95 sample consisting of 95 long GRBs, spanning 0.297 ≤ z ≤ 9.4 and also taken from Dainotti et al. (2020), as well as the combined LGRB145 data set of 145 GRBs, to test whether they are better described by the three-dimensional (3D) fundamental plane (Dainotti) correlation between the peak prompt luminosity, the luminosity at the end of the plateau emission, and its rest frame duration (Dainotti et al., 2016(Dainotti et al., , 2021aSrinivasaragavan et al., 2020) or by the two-dimensional (2D) Dainotti correlation between the luminosity at the end of the plateau emission and its rest frame duration (Dainotti et al., 2008(Dainotti et al., , 2011b(Dainotti et al., , 2013a(Dainotti et al., , 2015b, 2 and to constrain cosmological-model and GRB-correlation parameters. The Platinum sample is a compilation of the higher-quality (lower intrinsic dispersion) GRBs considered in Dainotti et al. (2020), and are tabulated in Table B.1. ...
Preprint
The current expansion of the Universe has been observed to be accelerating, and the widely accepted spatially-flat concordance model of general relativistic cosmology attributes this phenomenon to a constant dark energy, a cosmological constant, which is measured to comprise about 70% of the total energy budget of the current Universe. However, observational discrepancies and theoretical puzzles have raised questions about this model, suggesting that alternative cosmological models with non-zero spatial curvature and/or dark energy dynamics might provide better explanations. To explore these possibilities, we have conducted a series of studies using standardized, lower-redshift observations to constrain six different cosmological models with varying degrees of flatness and dark energy dynamics. Through comparing these observations with theoretical predictions, we aim to deepen our understanding of the evolution of the Universe and shed new light on its mysteries. Our data provide consistent cosmological constraints across all six models, with some suggesting the possibility of mild dark energy dynamics and slight spatial curvature. However, these joint constraints do not rule out the possibility of dark energy being a cosmological constant and the spatial hypersurfaces being flat. Overall, our findings contribute to the ongoing efforts to refine our understanding of the Universe and its properties, and suggest that multiple cosmological models remain viable.
Article
Gamma-ray bursts (GRBs) can be probes of the early Universe, but currently, only 26% of GRBs observed by the Neil Gehrels Swift Observatory have known redshifts ( z ) due to observational limitations. To address this, we estimated the GRB redshift (distance) via a supervised statistical learning model that uses optical afterglow observed by Swift and ground-based telescopes. The inferred redshifts are strongly correlated (a Pearson coefficient of 0.93) with the observed redshifts, thus proving the reliability of this method. The inferred and observed redshifts allow us to estimate the number of GRBs occurring at a given redshift (GRB rate) to be 8.47–9 yr ⁻¹ Gpc ⁻¹ for 1.9 < z < 2.3. Since GRBs come from the collapse of massive stars, we compared this rate with the star formation rate, highlighting a discrepancy of a factor of 3 at z < 1.
Article
Gamma Ray Bursts (GRB) are among the brightest objects in the Universe and hence can be observed up to a very high redshift. Properly calibrated empirical correlations between intensity and spectral correlations of GRBs can be used to estimate the cosmological parameters. However, the possibility of the evolution of GRBs with the redshift is a long-standing puzzle. In this work, we used 162 long-duration GRBs to determine whether GRBs below and above a certain redshift have different properties. The GRBs are split into two groups, and we fit the Amati relation for each group separately. Our findings demonstrate that estimations of the Amati parameters for the two groups are substantially dissimilar. We perform simulations to investigate whether the selection effects could cause the difference. Our analysis shows that the differences may be either intrinsic or due to systematic errors in the data, and the selection effects are not their true origin. However, in-depth analysis with a new data set comprised of 119 long GRBs shows that intrinsic scatter may partly be responsible for such effects.
Article
The recent ∼4 σ Hubble constant, H0, tension is observed between the value of H0 from the Cosmic Microwave Background (CMB) and Type Ia Supernovae (SNe Ia). It is a decade since this tension is excruciating the modern astrophysical community. To shed light on this problem is key to consider probes at intermediate redshifts between SNe Ia and CMB and reduce the uncertainty on H0. Toward these goals, we fill the redshift gap by employing Gamma-Ray Bursts (GRBs) and Quasars (QSOs), reaching z = 9.4 and z = 7.6, respectively, combined with Baryonic Acoustic Oscillations (BAO) and SNe Ia. To this end, we employ the ‘Dainotti GRB 3D relation’ among the rest-frame end time of the X-ray plateau emission, its corresponding luminosity, and the peak prompt luminosity, and the ‘Risaliti-Lusso’ QSO relation between ultraviolet and X-ray luminosities. We inquire the commonly adopted Gaussianity assumption on GRBs, QSOs, and BAO. With the joint sample, we fit the flat Λ Cold Dark Matter model with both the Gaussian and the newly discovered likelihoods. We also investigate the impact of the calibration assumed for Pantheon and Pantheon + SNe Ia on this analysis. Remarkably, we show that only GRBs fulfill the Gaussianity assumption. We achieve small uncertainties on the matter density parameter ΩM and H0. We find H0 values compatible within 2 σ with the one from the Tip of the Red Giant Branch. Finally, we show that the cosmological results are heavily biased against the arbitrary calibration choice for SNe Ia.
Article
Full-text available
An intrinsic correlation has been identified between the luminosity and duration of plateaus in the X-ray afterglows of Gamma-Ray Bursts (GRBs; Dainotti et al. 2008), suggesting a central engine origin. The magnetar central engine model predicts an observable plateau phase, with plateau durations and luminosities being determined by the magnetic fields and spin periods of the newly formed magnetar. This paper analytically shows that the magnetar central engine model can explain, within the 1$\sigma$ uncertainties, the correlation between plateau luminosity and duration. The observed scatter in the correlation most likely originates in the spread of initial spin periods of the newly formed magnetar and provides an estimate of the maximum spin period of ~35 ms (assuming a constant mass, efficiency and beaming across the GRB sample). Additionally, by combining the observed data and simulations, we show that the magnetar emission is most likely narrowly beamed and has $\lesssim$20% efficiency in conversion of rotational energy from the magnetar into the observed plateau luminosity. The beaming angles and efficiencies obtained by this method are fully consistent with both predicted and observed values. We find that Short GRBs and Short GRBs with Extended Emission lie on the same correlation but are statistically inconsistent with being drawn from the same distribution as Long GRBs, this is consistent with them having a wider beaming angle than Long GRBs.
Article
Full-text available
We study the dark energy equation of state as a function of redshift in a non-parametric way, without imposing any {\it a priori} $w(z)$ (ratio of pressure over energy density) functional form. As a check of the method, we test our scheme through the use of synthetic data sets produced from different input cosmological models which have the same relative errors and redshift distribution as the real data. Using the luminosity-time $L_{X}-T_{a}$ correlation for GRB X-ray afterglows (the Dainotti et al. correlation), we are able to utilize GRB sample from the {\it Swift} satellite as probes of the expansion history of the Universe out to $z \approx 10$. Within the assumption of a flat FLRW universe and combining SNeIa data with BAO constraints, the resulting maximum likelihood solutions are close to a constant $w=-1$. If one imposes the restriction of a constant $w$, we obtain $w=-0.99 \pm 0.06$ (consistent with a cosmological constant) with the present day Hubble constant as $H_{0}=70.0 \pm 0.6$ ${\rm km} \, {\rm s}^{-1} {\rm Mpc}^{-1}$ and density parameter as $\Omega_{\Lambda 0}=0.723 \pm 0.025$, while non-parametric $w(z)$ solutions give us a probability map which is centred at $H_{0}=70.04 \pm 1$ ${\rm km} \, {\rm s}^{-1} {\rm Mpc}^{-1}$ and $\Omega_{\Lambda 0}=0.724 \pm 0.03$. Our chosen GRB data sample with full correlation matrix allows us to estimate the amount, as well as quality (errors) of data, needed to constrain $w(z)$ in the redshift range extending an order of magnitude in beyond the farthest SNeIa measured.
Article
Full-text available
Gamma-ray bursts (GRBs) observed up to redshifts z > 9.4 can be used as possible probes to test cosmological models. Here we show how changes of the slope of the luminosity $L^*_{\rm X}$–break time $T^*_a$ correlation in GRB afterglows, hereafter the LT correlation, affect the determination of the cosmological parameters. With a simulated data set of 101 GRBs with a central value of the correlation slope that differs on the intrinsic one by a 5σ factor, we find an overestimated value of the matter density parameter, ΩM, compared to the value obtained with Type Ia supernovae, while the Hubble constant, H0, best-fitting value is still compatible in 1σ compared to other probes. We show that this compatibility of H0 is due to the large intrinsic scatter associated with the simulated sample. Instead, if we consider a subsample of high-luminosity GRBs (High L), we find that the evaluation of both H0 and ΩM is not more compatible in 1σ and ΩM is underestimated by 13 per cent. However, the High L sample choice reduces dramatically the intrinsic scatter of the correlation, thus possibly identifying this sample as the standard canonical ‘GRBs’ confirming previous results presented by Dainotti et al. Here, we consider the LT correlation as an example, but this reasoning can also be extended for all other GRB correlations. In the literature so far, GRB correlations are not corrected for redshift evolution and selection biases; therefore, we are not aware of their intrinsic slopes and consequently how far the use of the observed correlations can influence the derived ‘best’ cosmological settings. Therefore, we conclude that any approach that involves cosmology should take into consideration only intrinsic correlations and not the observed ones.
Article
Gamma-ray burst (GRB) afterglow plateaus can be explained in the thick shell model via some form of energy injection or later or slower ejecta, and in the thin shell model via two-jet models where the deceleration of the slower component marks the end of the plateau. A number of correlations between observables have been found to exist for GRB afterglows, linking ejecta energy to prompt and afterglow energy release and linking plateau phase optical and X-ray luminosity to the end times of the plateaus. It is shown here that the observed correlations rule out basic thin shell models but not the basic thick shell model. In the thick shell case, both forward shock and reverse shock dominated outflows are shown to be consistent with the correlations, using randomly generated samples of thick shell model afterglows.
Article
γ-ray bursts (GRBs) have recently attracted much attention as a possible way to extend the Hubble diagram to a very high redshift. However, the large scatter in their intrinsic properties prevents directly using them as a distance indicator so that the hunt is open for a relation involving an observable property to standardize GRBs in the same way as the Phillips law makes it possible to use Type Ia supernovae as standardizable candles. We use here the data on the X-ray decay curve and spectral index of a sample of GRBs observed with the Swift satellite. These data are used as input to a Bayesian statistical analysis looking for a correlation between the X-ray luminosity LX(Ta) and the time constant Ta of the afterglow curve. We find a linear relation between log [LX(Ta)] and log [Ta/(1 +z)] with an intrinsic scatter σint= 0.33 comparable to previously reported relations. Remarkably, both the slope and the intrinsic scatter are almost independent on the matter density ΩM and the constant equation of state w of the dark energy component thus suggesting that the circularity problem is alleviated for the LX–Ta relation.
Article
We present analytic calculations of synchrotron radiation from the forward and the reverse shock of gamma-ray burst blast waves, in the thick-shell scenario (i.e. when the reverse shock is relativistic). We show that this scenario can naturally account for the plateau phase, observed early in the afterglows of about half the bursts detected by Swift. We generalize our approach to include power-law luminosity of the central engine and show that when radiation from both regions (forward and reverse shock) is taken into account, a wide range of possibilities emerge, including chromatic and achromatic breaks, frequency-dependent spectral evolution during the injection break and widely varying decay indices in different bands. For both the forward and the reverse shock, we derive formulas for the spectral parameters and the observed flux in different power-law segments of the spectrum, as a function of observer time. We explore the Fb-tb relation (between the observed time of the end of the plateau phase and the flux at that point) in the framework of the presented model and show that model predictions favour the reverse shock as the dominant source of emission in both optical and X-rays. As case studies, we present simultaneous fits to X-ray and optical/IR afterglow data of GRB 080928 and GRB 090423. We identify the end of the plateau phase with the cessation of energy injection and infer the corresponding upper limits to central-engine activity, which are about 1 h for the former and 1.5 h for the latter. We conclude that smooth energy injection through the reverse shock is a plausible explanation for the plateau phase of gamma-ray burst afterglows. During that phase, radiation from the reverse shock is likely to be important, or even dominant, and should be taken into account when fitting model parameters to observations.
Article
A sufficiently powerful astrophysical source with power-law luminosity in time will give rise to a self-similar relativistic blast wave with a reverse shock travelling into the ejecta and a forward shock moving into the surrounding medium. Once energy injection ceases and the last energy is delivered to the shock front, the blast wave will transit into another self-similar stage depending only on the total amount of energy injected. I describe the effect of limited duration energy injection into environments with density depending on radius as a power law, emphasizing optical/X-ray Gamma-ray Burst afterglows as applications. The blast wave during injection is treated analytically, the transition following last energy injection with one-dimensional simulations. Flux equations for synchrotron emission from the forward and reverse shock regions are provided. The reverse shock emission can easily dominate, especially with different magnetizations for both regions. Reverse shock emission is shown to support both the reported X-ray and optical correlations between afterglow plateau duration and end time flux, independently of the luminosity power-law slope. The model is demonstrated by application to bursts 120521A and 090515, and can accommodate their steep post-plateau light-curve slopes.
Article
The early X-ray afterglow of gamma-ray bursts revealed by Swift carried many surprises. We focus in this paper on the plateau phase whose origin remains highly debated. We confront several newly discovered correlations between prompt and afterglow quantities (isotropic emitted energy in gamma-rays, luminosity and duration of the plateau) to several models proposed for the origin of plateaus in order to check if they can account for these observed correlations. We first show that the scenario of plateau formation by energy injection into the forward shock leads to an efficiency crisis for the prompt phase and therefore study two possible alternatives: the first one still takes place within the framework of the standard forward shock model but allows for a variation of the microphysics parameters to reduce the radiative efficiency at early times; in the second scenario the early afterglow results from a long-lived reverse shock. Its shape then depends on the distribution of energy as a function of Lorentz factor in the ejecta. In both cases, we first present simple analytical estimates of the plateau luminosity and duration and then compute detailed light curves. In the two considered scenarios we find that plateaus following the observed correlations can be obtained under the condition that specific additional ingredients are included. In the forward shock scenario, the preferred model supposes a wind external medium and a microphysics parameter epsilon_e that first varies as n^{-\nu} (n being the external density), with \nu~1 to get a flat plateau, before staying constant below a critical density n_0. To produce a plateau in the reverse shock scenario the ejecta must contain a tail of low Lorentz factor with a peak of energy deposition at \Gamma >~ 10.
Article
The comprehensive statistical analysis of Swift X-ray light curves, collecting data from six years of operation, revealed the existence of a universal scaling among the isotropic energy emitted in the rest-frame 1 - 104 keV energy band during the prompt emission (Eγ, iso ), the peak of the prompt emission energy spectrum (Epk) and the X-ray energy emitted in the 0.3-10 keV observed energy band (EX, iso). In this paper, we show that this three-parameter correlation is robust and does not depend on our definition of EX, iso. It is shared by long, short and low-energetic gamma-ray bursts (GRBs), and thus reflects the existence of some properties which are shared by the GRB class as a whole. A more intuitive way to express the three-parameter correlation is in the form of a two-parameter correlation between the GRB efficiency and Epk. We speculate that the physical origin of such a relation is connected with the outflow Lorentz factor.
Article
Gamma-ray bursts (GRBs) are tremendous explosions visible across most of the universe, certainly out to redshifts of z = 4.5 and likely out to z ~ 10. Recently, GRBs have been found to have a roughly constant explosive energy as well as to have two luminosity indicators (the spectral lag time and the variability) that can be used to derive the burst's luminosity distance from the gamma-ray light curve alone. There currently exists enough information to calibrate luminosity distances and independent redshifts for nine bursts. From these, a GRB Hubble diagram can be constructed, in which the observed shape of the curve provides a record of the expansion history of our universe. The current nine-burst diagram is sparse, yet formal limits can be placed on the mass density of a flat universe. This first GRB Hubble diagram provides a proof of concept for a new technique in cosmology at very high redshifts. With the launch of the Swift satellite in 2003, we should get ~120 bursts to produce a Hubble diagram impervious to all effects of dust extinction and out to redshifts impossible to reach by any other method.