Next Article in Journal
Observation and Data Reduction of the Brown Dwarf 2MASSW J0746425 + 200032 by Five-Hundred-Meter Aperture Spherical Radio Telescope
Previous Article in Journal
Mission Target: Tetraquark Mesons of Flavour-Cryptoexotic Type
Previous Article in Special Issue
The Central Engine of GRB170817A and the Energy Budget Issue: Kerr Black Hole versus Neutron Star in a Multi-Messenger Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Concept of Assessment of LIV Tests with THESEUS Using the Gamma-Ray Bursts Detected by Fermi/GBM

by
Anastasia Tsvetkova
1,2,*,
Luciano Burderi
1,3,4,
Alessandro Riggio
1,3,5,
Andrea Sanna
1,3,4 and
Tiziana Di Salvo
6
1
Dipartimento di Fisica, Universitá degli Studi di Cagliari, SP Monserrato-Sestu, km 0.7, 09042 Monserrato, Italy
2
Ioffe Institute, Politekhnicheskaya 26, 194021 St. Petersburg, Russia
3
INFN, Sezione di Cagliari, Cittadella Universitaria, 09042 Monserrato, Italy
4
INAF—Osservatorio Astronomico di Cagliari, Via della Scienza 5, 09047 Selargius, Italy
5
INAF—Istituto di Astrofisica Spaziale e Fisica Cosmica di Palermo, Via U. La Malfa 153, 90146 Palermo, Italy
6
Dipartimento di Fisica e Chimica, Universitá degli Studi di Palermo, Via Archirafi 36, 90123 Palermo, Italy
*
Author to whom correspondence should be addressed.
Universe 2023, 9(8), 359; https://doi.org/10.3390/universe9080359
Submission received: 15 May 2023 / Revised: 13 July 2023 / Accepted: 21 July 2023 / Published: 31 July 2023
(This article belongs to the Special Issue GRBs Phenomenology, Models and Applications: A Beginner Guide)

Abstract

:
According to Einstein’s special relativity theory, the speed of light in a vacuum is constant for all observers. However, quantum gravity effects could introduce its dispersion depending on the energy of photons. The investigation of the spectral lags between the gamma-ray burst (GRB) light curves recorded in distinct energy ranges could shed light on this phenomenon: the lags could reflect the variation of the speed of light if it is linearlydependent on the photon energy and a function of the GRB redshift. We propose a methodology to start investigating the dispersion law of light propagation in a vacuum using GRB light curves. This technique is intended to be fully exploited using the GRB data collected with THESEUS.

1. Introduction

According to Einstein’s special relativity theory, a proper length is Lorentz-contracted by a factor of γ 1 = [ 1 ( v / c ) 2 ] 1 / 2 as observed from the reference frame moving at speed v relative to the rest frame. However, various spacetime theories, e.g., some string or loop quantum gravity theories (see, e.g., [1,2,3]), imply the existence of a minimum spatial length of the order of the Plank length l Pl = G / c 3 = 1.6 × 10 33 cm [4]. Lorentz invariance is a fundamental property of both the standard model of particle physics and general relativity. In general relativity, a locally inertial reference frame where the Lorentz symmetry is fulfilled can always be chosen. Some quantum gravity (QG) theories predict the Lorentz invariance violation (LIV) at the Planck energy scale ( E Pl = c 5 / G 1.22 × 10 19 GeV) as there exists a minimum spatial length l min = α l Pl (where α 1 is a dimensionless constant inherent to a particular spacetime theory), which, e.g., in the string theories, corresponds to the string length. Therefore, the Lorentz contraction is limited by this spatial scale (see [4] for a review).
There are several frameworks implying LIV, e.g., string theory [5,6], noncommutative spacetime [7,8], Brane worlds [9], Hořava–Lifshitz gravity [10]. LIV was considered in the gravitational context for the first time in [11], where the so-called standard model extension was developed. The Bumblebee1 models, which are the simplest cases of theories including the spontaneous breaking of Lorentz symmetry, are effective field QG theories describing a vector field with a non-zero vacuum expectation value and involving the vacuum condensate [5,6]. The spontaneous symmetry breaking preserves both the geometric constraints and conservation laws or quantities required by the general relativity theory or Riemannian geometry. Regarding gravity, LIV can happen if a vector field ruled by a potential exhibiting a minimum rolls to its vacuum expectation value, similar to the Higgs mechanism [11]. This “bumblebee” vector thus takes an explicit (four-dimensional) orientation, and preferred-frame effects may emerge [12]. The generalized uncertainty principle (GUP) states that, in quantum theory, if the quantities are incompatible, they are mutually dependent, and measuring one observable may yield some information about its incompatible partners. GUP is based on a momentum-dependent modification of the standard dispersion relation, which is supposed to produce LIV [13,14]. See, e.g., [15,16,17,18,19,20,21] for recent advances in the Bumblebee and GUP models.
One of the LIV effects, relevant to astrophysics, is the existence of a dispersion law for the photon speed c (see, e.g., [22]). However, LIV is not a mandatory property of all QG theories: in some of them, e.g., in the spacetime uncertainty principle [23] or in the quantum spacetime [24], LIV is not expected. Despite photon velocity dispersion, the Lorentz invariance is not violated and the dispersion law is a second-order effect relative to the ratio of the photon energy to the QG energy scale. To obtain an idea of the nature of the speed of light dispersion due to LIV, e.g., within the Liouville string approach, one can consider a vacuum as a non-trivial medium containing “foamy” quantum gravity fluctuations whose origin can be imagined as processes that involve the pair creation of virtual black holes. In this concept, one can verify that the massless particles of different energies can excite vacuum fluctuations differently as they propagate through the quantum gravity medium, producing a non-trivial dispersion relation of Lorentz “non-covariant” form, similarly to the thermal medium [25]. For more details regarding this concept, see, e.g., [26].
Since gamma-ray bursts (GRBs) are characterized by high-energy emission, large cosmological distances, and temporal variability at short (≲10 ms) timescales, they have been applied as powerful tools in LIV searches (see, e.g., [25,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49]) for more than two decades. In [25], it was first suggested to test LIV using a comparison of the arrival times of GRB photons detected in distinct energy ranges. In [29,30,50], the authors exploited the spectral lag between high-energy (∼31 GeV) and low-energy photons of GRB 090510 to derive lower limits on the linear and quadratic QG energy: E QG , 1 > ( 1 10 ) × E Pl = 1.22 × ( 10 19 10 20 ) GeV and E QG , 2 > ( 1 10 ) × E Pl = 1.3 × 10 11 GeV, respectively. In [51], E QG > 10 18 GeV was obtained based on the observation of GRB 080916C. For GRB 190114C [52], the linear and quadratic LIV were constrained using the time delays of TeV photons: E QG , 1 > 0.58 × 10 19 GeV ( E QG , 1 > 0.55 × 10 19 GeV) and E QG , 2 > 0.63 × 10 11 GeV ( E QG , 1 > 0.56 × 10 11 GeV) for the subluminal (superluminal) case [43]. In [53], the authors constrained the linear and quadratic LIV for the set of 32 Fermi/GBM GRBs with known redshifts and characterized by the positive-to-negative transition of the spectral lag: E QG , 1 = 1.5 × 10 14 GeV for the linear case and E QG , 2 = 8 × 10 5 GeV for the quadratic case.
The phenomenon of GRBs remains puzzling, although much progress has been made. Both the light curves and the spectra among GRBs vary significantly. It is generally believed that collisions between relativistic shells ejected from an active central engine produce pulses in GRB light curves [54]. The collision of the slower-moving shell with the second, faster shell ejected later produces a shock that dissipates internal energy and accelerates the particles that emit the GRB radiation. As of now, two “physical” classes of GRBs are distinguished (see, e.g., [55]): the merger-origin Type I GRBs [56,57,58,59], which are usually short, with a duration of less than 2 s [60,61], and spectrally hard, and the collapser-origin Type II GRBs [62,63,64,65], characterized by longer durations.
The spectral lag, which is known as the difference in arrival time between high-energy and low-energy photons, is a common phenomenon occurring during the GRB prompt emission phase [66,67,68,69] and in high-energy astrophysics in general (e.g., [67,70]). The authors of [71] were the first to analyze the spectral lags of GRBs. It was found that a soft lag, i.e., the hard photons arriving first, dominates in long GRBs [67,69,72,73], and some GRBs have significantly different spectral lags in early and late epochs [74]. It was shown that the lags are correlated with the GRB luminosity [67] and the jet break times in afterglow light curves [75]; the spectral lags of long GRBs are correlated with the pulse duration, while the spectral lags of short GRBs are not [76]. The GRB spectral lags can be explained within several physical models, such as the curvature effect of a relativistic jet and rapidly expanding spherical shell [77,78,79,80,81]. Regardless of its physical origin, a spectral lag is an important GRB parameter as it may help to distinguish between long and short GRBs: long bursts have large lags, while short bursts have relatively smaller or negligible lags [82,83,84].
To show the feasibility of the search for the dispersion of the speed of the light in the data collected by THESEUS [85], one can start performing similar research using the data of already commissioned missions. In this paper, we describe a methodology for the testing of the dispersion of the light speed, and we intend to apply it to the Fermi/GBM data and the set of GRBs with known redshifts. The paper is organized as follows. We start with a brief description of the instrumentation and data in Section 2. The methodology that involves constraining LIV using GRBs is described in Section 3. In Section 4, we discuss the THESEUS mission in the context of the LIV tests using GRBs. Section 5 concludes the paper.

2. Instrumentation and Data

Among several space-based detectors able to collect GRB data, the most prolific are Swift/BAT (BAT; [86]), Fermi/GBM (GBM; [87]), and Konus-Wind (KW; [88]). More details regarding the BAT, GBM, KW, and other GRB detectors’ design and performance can be found in [89]. However, since Swift/BAT collects GRB data in a narrow energy band of 15–350 keV, and Konus-Wind records GRB light curves in three fixed energy windows, the Fermi/GBM GRB time histories seem to be the most suitable for the testing of the speed of light variance.
Launched in June 2008, the Fermi Gamma-Ray Space Telescope [90] harbors two scientific instruments: the Gamma-Ray Burst Monitor (GBM) and the Large Area Telescope (LAT; [91]). The LAT covers the 30 MeV–300 GeV band, while the GBM, intended to detect and study GRBs, is sensitive within the 8 keV–30 MeV energy range, extending the spectral band over which bursts are observed downwards to the hard X-ray range. GBM comprises twelve NaI(Tl) detectors covering an energy range of 8 keV 1 MeV and two bismuth-germanate (BGO) scintillation detectors sensitive within the 150 keV to 30 MeV band that observe the whole sky not occluded by the Earth (>8 sr).
The primary scientific data produced by GBMs can be summarized as a time history and spectra, which are provided as temporally pre-binned (CTIME and CSPEC) or temporally unbinned time tagged events (TTE). These data types are produced as “snippets” for every trigger and are also provided continuously. The CTIME data are collected in 8 energy channels with a 256 ms time resolution, while the CSPEC data are recorded in 128 channels with an 8.192 s time resolution. TTEs for each detector are recorded with time precision down to 2 μ s, in 128 energy channels, matching the CSPEC ones, which gives an excellent opportunity to bin the data in time and energy in a suitable way. From 2008 through November 2012, TTE were available only during a 330 s interval: from 30 s before the burst trigger to 300 s after the burst trigger. Since November 2012, GBM flight software has produced a new data type, continuous TTE (CTTE), available at all times that the instrument is operating.
To date, the GBM has triggered almost 3500 GRBs, among which almost 3000 have T 90 > 2 s.

3. Methodology

We start the research by computing the redshifts and the rest-frame spectral lags for all GBM-triggered GRBs. Then, thanks to the central limit theorem, we can consider the distribution of the lags as normal and compute the mean and variance. The large number of “measurements” is expected to significantly increase the accuracy of the spectral lags and constraints on E QG . Another approach would be to constrain the QG energy (see Section 3.1) first and then estimate its mean values and variance.

3.1. Brief Introduction to the Basics of the Speed of Light Variance

The velocity dispersion law for photons can be expressed as a function of its observer-frame energy E obs in units of the QG energy scale E QG , at which the quantum nature of gravity becomes important:
v phot / c 1 = ξ E obs E QG n ,
where v phot is the group velocity of a photon wave-packet, E QG = ζ m Pl c 2 = ζ E Pl , m Pl = 2.176 × 10 5 g is the Planck mass, ζ α 1 1 expresses the significance of the QG effects, ξ ± 1 is a dimensionless constant inherent to a particular QG theory, and the index n denotes the order of the first relevant term of the small parameter E obs E QG . This expression takes into account that high-energy photons can travel faster (superluminal, ξ = + 1 ) or slower (subluminal, ξ = 1 ) than low-energy ones [92].
The difference in the arrival times of photons emitted at the same time in the same place is
Δ t QG = ξ D trav c Δ E obs E QG n ,
where D trav is the comoving distance traversed by a massless particle, emitted at redshift z and traveling down to redshift 0.

3.2. Observed Spectral Lag as a Function of Redshift

For a given observer-frame energy E obs , the total spectral lag τ total , obs ( E obs , z ) can be split into two terms:
τ total , obs ( E obs , z ) = τ int , obs ( E obs , z ) + τ QG , obs ( E obs , z ) ,
where τ int , obs is the observed intrinsic spectral lag, which corresponds to the intrinsic rest-frame lag
τ int , rf ( E obs ) = τ int , obs ( E rf ) / ( 1 + z )
induced by the GRB central engine emission mechanism and assumed to be independent of the photon source redshift z. τ QG , obs is the lag induced by the QG effects discussed above, and and E rf is the photon energy in the rest frame of its source. Following [93], τ QG , obs can be expressed as a function of the GRB rest-frame energy:
τ QG ( E rf , z ) = ξ 1 H 0 E rf ζ E pl n 1 + n 2 1 1 + z n × × 0 z ( 1 + z ) n d z Ω m ( 1 + z ) 3 + Ω Λ ,
where H 0 is the Hubble constant, Ω m is the matter density parameter, and Ω Λ is the dark energy density parameter, i.e., the parameters of the standard ΛCDM model.
Experimentally, the total observed spectral lag τ total , obs ( E rf , z ) can be computed by cross-correlating the GRB light curves recorded in the redshift-dependent energy windows corresponding to the fixed rest-frame energy windows as E obs = E rf / ( 1 + z ) , with the GRB light curve collected at the lowest possible energy channel, where the lag induced by QG is negligible, e.g., it is ∼ μ s in the 5–20 keV energy range, while, in the higher-energy bands, it is ∼ms.
The total observed spectral lag τ total , obs ( E rf , z ) should follow the relation obtained by inserting Equations (4) and (5) into Equation (3):
τ total , obs ( E rf , z ) = τ int , obs ( E rf ) + ξ 1 H 0 E rf ζ E pl n 1 + n 2 1 1 + z n 0 z ( 1 + z ) n d z Ω m ( 1 + z ) 3 + Ω Λ .
The transformation of spectral lags from the observer frame back to the rest frame2 can be performed by dividing by the redshift factor ( 1 + z ) :
τ total , rf ( E rf , z ) = τ total , obs ( E rf , z ) / ( 1 + z ) .
Thus, the GRB rest-frame spectral lag obeys the following relation:
τ total , rf ( E rf , z ) = τ int , obs ( E rf ) ( 1 + z ) + ξ 1 H 0 E rf ζ E pl n 1 + n 2 1 1 + z n + 1 0 z ( 1 + z ) n d z Ω m ( 1 + z ) 3 + Ω Λ .
Let us define a function
u ( z ) = 1 + n 2 1 1 + z n + 1 0 z ( 1 + z ) n d z Ω m ( 1 + z ) 3 + Ω Λ .
Then, the experimentally determined lag τ total , obs ( E rf , z ) will follow the relation
τ total , rf ( E rf , z ) = τ int , rf ( E rf ) + ξ 1 H 0 E rf ζ E pl n u ( z ) .
The dependence of τ total , rf ( E rf , z ) on u ( z ) is expected to be linear, with the intercept corresponding to the intrinsic lag and the slope proportional to the ratio of the rest-frame photon energy to the QG energy ζ E pl raised to the n-th power, which is the first significant term in the series expansion of the quantum gravity dispersion relation. An argument in support of the independence of τ int , rf from z is the absence of the prominent cosmological evolution of GRB energetics [89,94], which indicates that the GRB central engine does not evolve significantly with z. However, if τ int , rf is dependent on z, one can fit both τ int , rf ( z ) and τ QG ( z ) simultaneously to the data, as was done in [53].
In [53], it was found that the behavior of the lags follows some key statistical properties described in terms of log-normal and Gaussian distributions around average values. These empirical functions describing the spectral lags (as a function of energy) for each of the 32 Fermi/GBM GRBs in the sample of bursts with known redshifts that exhibit the lag transition phenomenon are shown in Figure 1 (as derived from Figure 1 of [53]). In the method proposed here, we make the reasonable assumption that the intrinsic lags (that dominate in magnitude over the small delays induced by QG effects; see the upper limits for the first- and second-order QG delays shown as blue and orange curves in Figure 1) do not correlate with the redshift as the distance of a given GRB from us does not affect the emission properties in, e.g., the fireball model. This means that, averaging over a large sample of GRBs at different redshifts, the intrinsic delays will cluster around the common value that defines the intrinsic average rest-frame lag.

3.3. The Technique of Averaging over the Sample

Relation (10) shows that, for a given GRB rest-frame energy E rf , the lag τ total , rf ( E rf ) depends only on the GRB redshift z through the function u ( z ) defined in Equation (9). Therefore, fixing E rf for an ensemble of N GRBs with known redshift, one can compute the function τ total , rf ( E rf ) from the observed data for all GRBs of the ensemble, i.e., it is possible to obtain a set of N experimentally computed values of τ total , rf ( E rf ) . This ensemble of N values can be fitted as a function of u ( z ) through Equation (10) to obtain the best fit values of the intrinsic lag in the GRB rest frame τ int , rf , and the coefficient of the QG-induced delay at the GRB rest-frame energy E rf
τ i n t , r f ( E rf ) BEST and ξ 1 H 0 E rf ζ E pl n BEST = ξ 1 H 0 α E rf E pl n BEST = ϕ ( α E rf ) BEST .
Once the values of ϕ ( α E rf ) BEST are obtained for all E rf , these values can be plotted as a function of s ( E rf ) = ( E rf / E pl ) n and subsequently linearly fitted through the equation
ϕ ( α E rf ) = α n H 0 s ( E rf ) = Δ QG s ( E rf ) ,
to obtain the best fit value of the strength of the QG effect Δ QG = α n / H 0 . We note that this technique allows us to combine the whole ensemble of N GRBs to obtain a unique measure of the strength of the QG effect, whose uncertainty σ Δ QG , in the absence of other systematic errors, only depends on the (Poissonian) statistics of the whole ensemble of N GRBs and therefore improves as the inverse square root of N:
σ Δ QG 1 N .
Consequently, the precision of the measurement of the QG effect’s strength can be improved by increasing the size of the analyzed sample.

3.4. GRB Intrinsic Spectral Lags vs. Quantum Gravity Effects

The GRB spectral lag can be caused by a mixture of two effects: the QG one and the one inherent to the fireball model. The latter is due to the curvature effect, i.e., the kinematic effect caused by the fact that the observer looks at an increasingly off-axis annulus area relative to the line-of-sight [77,78,79,95,96,97,98,99,100]. Softer low-energy radiation comes from the off-axis annulus area with smaller Doppler factors and is delayed for the observer with respect to on-axis emission due to the geometric curvature of the shell.
A competing hypothesis is that the traditional view based on the high-latitude emission “curvature effect” of a relativistic jet cannot explain spectral lags. Instead, spectral peaks should be swept across the observing energy range in a specific manner to account for the observed spectral lags. A simple physical model that implies synchrotron radiation from a rapidly expanding outflow can explain GRB spectral lags [81]. This model requires the following conditions to be fulfilled: (1) the emission radius has to be large (over several 10 14 cm from the central engine), in the optically thin region, well above the photosphere; (2) the γ -ray photon spectrum is curved (as observed); (3) the magnetic field strength in the emitting region decreases with the radius as the region expands in space, which is consistent with an expanding jet; and (4) the emission region itself undergoes rapid bulk acceleration as the prompt emission is produced. These requirements are consistent with a Poynting-flux-dominated jet abruptly dissipating magnetic energy at a large distance from the engine. The aforementioned theories successfully explain the positive spectral lags. Nevertheless, the rarely observed negative lags remain a more intriguing phenomenon that can be used to infer the different radiation mechanisms [101,102] or emission regions [103] of low- and high-energy photons.
For a given GRB, the intrinsic delay inherent to the GRB emission could mimic the genuine quantum gravity effect, making these two effects difficult to disentangle. However, currently, there is no evidence for a correlation between the GRB intrinsic delays and the distances to its sources. For example, in [94,104], where the largest sample of GRBs with known redshifts detected by a single instrument in a wide energy range is studied, the significance of the cosmological evolution of GRB energetics is ≲2 σ . Meanwhile, the delays induced by a photon dispersion law are proportional both to the light travel distance (a function of redshift) and to the differences in the energy of the photons. This dual dependence on energy and redshift could be the unique feature of a genuine QG effect. As suggested in [105], given an adequate collection area, GRBs, once their redshifts are known, are potentially excellent tools to search for the first-order dispersion law for photons.

3.5. Computation of Spectral Lags

To avoid distortions due to the fact that the shape of the light curve changes with the energy, we suggest fixing the energy channels in which the light curves are recorded to certain values in the rest frame. In this case, the corresponding observer-frame values of the channel boundaries will be E obs = E rest / ( 1 + z ) , i.e., redshift-dependent. The first step to test the LIV effects with the suggested technique would be apply it to the GBM data. Thus, we propose to use the following energy bands to record the light curves. Given the 9 keV lower boundary of the GBM spectral window, one has to select the rest-frame channels starting from, at least, 60 keV, to allow for bursts with redshifts up to z = 5 . This boundary can be shifted towards a higher value to allow high-redshift GRBs to contribute to the study. However, we should mention that the majority of GRBs have redshifts z < 5 (see Figure 2). Some examples of the pseudo-logarithmic channels that could be used are 60–100 keV, 100–160 keV, 160–250 keV, 250–400 keV, 400–600 keV, 600–900 keV, or 60–80 keV, 80–100 keV, 100–130 keV, 130–160 keV, 160–200 keV, 200–250 keV, 250–320 keV, 320–400 keV, 400–500 keV, 500–650 keV, 650–900 keV. We found these numbers of channels to be reasonable in terms of SNR based on the research of [53], carried out for 32 Fermi/GBM GRBs.
Since the GRB energetics are usually considered on the logarithmic scale, we suggest adopting the geometric mean of the lower and upper boundaries of an energy band, E phot = E min × E max , as a proxy for the average energy of photons in the given range.
Since the spectral lag distributions of the short and long GRBs significantly differ [76], and these two types of bursts belong to distinct classes of progenitors, they have different intrinsic spectral lags. Thus, we suggest studying them separately.

3.6. Obtaining GRB Redshifts

Since the suggested technique of testing LIV using GRBs strongly relies on prior knowledge of the burst redshift, it is necessary either to measure it directly from the observations in optics or estimate it using the prompt emission parameters. GRB redshift measurements based on the detection of emission lines or absorption features of GRB host galaxies imposed on the afterglow continuum, or performed photometrically, are widespread. However, there are other methods to obtain the redshift estimates, e.g., the “pseudo-redshift” (pseudo-z) technique based on the spectral properties of GRB prompt high-energy emission [111], using well-known correlations such as, for example, the Norris correlation (spectral lag vs. isotropic peak luminosity; [67]), the Amati correlation (rest-frame peak energy vs. isotropic energy release; [112]), the isotropic peak luminosity vs. temporal variability correlation [113,114], the Yonetoku (the rest-frame peak energy vs. the isotropic peak luminosity; [115]) correlation, etc., or the method of searching for a minimum on the intrinsic hydrogen column density versus the redshift plane (see, e.g., [116]).
Nowadays, the machine learning (ML) approach to redshift estimation is becoming popular in astrophysics (see, e.g., [117,118,119,120]). Supervised ML is a data mining method based on prior knowledge of a “training” data set, on which we can build models predicting the parameter under consideration, a “validation” set, which provides an unbiased evaluation of a model’s fit while tuning the model’s hyperparameters, and a “test” data set necessary for an unbiased evaluation of the final model fit.
Considering only spectroscopic and photometric redshifts, there were ≳420 GRBs with reliably measured redshifts by the middle of 2022 (for a list of GRBs with measured redshifts, see [94,104,121,122,123], the Gamma-Ray Burst Online Index3, Jochen Greiner’s GRB table 4, and the references therein). Using one of the aforementioned techniques, one can estimate the redshift of any burst based on its temporal or spectral parameters and energetics. For example, [115,124,125] used various correlations to obtain unknown GRB redshifts from GRB observables, while [126] used the random forest algorithm to estimate GRB redshifts.

4. Discussion

THESEUS is a mission aimed at increasing the discovery rate of the high-energy transient phenomena over the entirety of cosmic history and fully exploiting GRBs to explore the early Universe [85,127]. THESEUS is likely to become a cornerstone of multi-messenger and time-domain astrophysics thanks to its exceptional payload, providing wide and deep sky monitoring in a broad energy range (0.3 keV–20 MeV); focusing capabilities in the soft X-ray band, providing large grasp and a high angular resolution; and onboard near-IR capabilities for immediate transient identification and redshift determination.
The THESEUS payload is planned to include the following instrumentation: (1) the X-Gamma-Ray Imaging Spectrometer (XGIS, 2 keV–20 MeV): a set of two coded-mask cameras using monolithic X-gamma-ray detectors based on bars of silicon diodes coupled with a crystal scintillator, granting a ∼2 sr field of view (FoV) and source location accuracy of ∼10 in the 2–150 keV band, as well as a >4 sr FoV at energies > 150 keV, with a few μ s timing resolution; (2) a Soft X-Ray Imager (SXI, 0.3–5 keV): a set of two lobster-eye telescope units, covering a total FOV of ∼0.5 sr with source location accuracy 2 ; (3) an infrared telescope (IRT, 0.7–1.8 μ m): a 0.7 m class IR telescope with a 15 × 15 FOV, for a fast response, with both imaging (I, Z, Y, J, and H) and spectroscopic (resolving power, R∼400, through 2 × 2 grism) capabilities.
Thanks to the unique combination of a wide 0.3 keV–10 MeV energy range, remarkable sensitivity, and exceptionally high counting statistics, THESEUS heralds a new era in the multi-wavelength studies of GRBs, providing the community with a sample of GRBs with known redshifts of unprecedented size, which, in turn, will not only allow the use of GRBs as cosmological tools but also shed light on one of the most challenging aspects of QG theory, the systematic study of which is still beyond the current instrumental capabilities. The capability of THESEUS to detect and localize GRBs, as well as measure their redshifts, will essentially surpass those of the current missions.
The left panel of Figure 3 shows the expected detection rate of long GRBs by THESEUS compared with observed GRBs. The orange histogram depicts the cumulative distribution of the GRBs detected by the SXI and/or XGIS (the bursts with measured z are marked in purple), while the blue histogram presents the distribution of GRBs with known redshifts detected from 2005 to the end of 2020. The distribution of the GRBs detected by THESEUS was acquired based on the anticipated IRT capabilities and on the assumption of a ground follow-up rate of 50% for the GRBs at z < 5 . It is noticeable that THESEUS is expected to detect an order of magnitude more bursts than Swift does, especially in the high-redshift domain ( z > 6 ) [128]. It is expected that the redshifts for the majority of GRBs detected by THESEUS will be measured (onboard or on the ground). The cumulative distribution plotted in the right panel of Figure 3 represents the annual detection rate of short GRBs by XGIS, not corrected for mission observation efficiency. THESEUS is supposed to acquire a statistically significant sample of short GRBs, including high-redshift ( z 4–5) events. Considering that the distribution of the spectral lags of short GRBs and long GRBs differs significantly, advances in research on short GRBs are very important for such sophisticated studies of the QG effects as we discuss in this paper. Thus, the sample of GRBs with measured redshifts obtained by THESEUS is the most promising for the application of the described technique to study LIV.

5. Conclusions

Various QG theories predict LIV, which can manifest itself as the dispersion of the speed of light. The method that we propose to disentangle and constrain the QG delays from the intrinsic spectral lags in GRB light curves is based on the assumption of the constancy of the rest-frame intrinsic spectral lags and on the linear dependence of the GRB spectral lag on both the photon energy and function of the GRB redshift. The ability to collect a large sample of GRBs with known redshifts is crucial for this type of study, as the precision of the QG effect measurement can be improved by expanding the data set. Currently, redshifts are measured spectroscopically or photometrically for 500 GRBs. Thus, indirect estimates of the redshifts from the prompt emission observables are necessary to obtain a large GRB sample for LIV studies using the commissioned instruments. The sample of GRBs collected by Fermi/GBM could provide a promising opportunity to apply the aforementioned technique, thanks to its extensive trigger statistics (≳3500 GRBs up to date) and sophisticated data acquired with a high temporal and spectral resolution, which could allow precise measurements of the rest-frame spectral lags.
The THESEUS mission is likely to initiate a breakthrough in this field of fundamental physics as, thanks to the combination of its unique characteristics, the observatory will collect one order of magnitude more samples of GRBs with known redshifts than are currently available, which will not only allow the use of GRBs as cosmological tools but will also enable us to constrain the QG theories. Moreover, thanks to its capability to detect the GRB emission in the relatively soft energy band of 0.3–5 keV, THESEUS could provide a unique opportunity not only to constrain the empirical and physical GRB models but also to expand the data range, providing more accurate constraints on the QG energy from the lag-energy plane. Due to its high sensitivity, THESEUS will also allow advances in the study of short GRBs; in particular, XGIS will be able to detect short GRBs up to z∼4–5, which is important as the spectral lags of short GRBs essentially differ from the ones of long bursts. Thus, the THESEUS mission could make a significant contribution to the study of the QG effects using GRBs.

Author Contributions

Conceptualization, L.B., A.T., A.R., A.S. and T.d.S.; methodology, L.B., T.d.S., A.T., A.R. and A.S.; writing—original draft preparation, A.T.; writing—review and editing, A.T., L.B., T.d.S., A.S. and A.R.; supervision, L.B.; project administration, A.S.; funding acquisition, L.B. and A.S. All authors have read and agreed to the published version of the manuscript.

Funding

Some of the authors wish to thank ASI and INAF (agreements ASI-UNI-Ca 2016- 13-U.O and ASI- INAF 2018-10-hh.0), the Italian Ministry of Education, University and Research (MIUR), Italy (HERMES-TP project) and the EU (HERMES-SP Horizon 2020 Research and Innovation Project under grant agreement 821896) for the financial support within the HERMES project. L.B., A.S. and A.T. acknowledge funding from the Italian Ministry of University and Research (MUR), PRIN 2017 (prot. 20179ZF5KS), “The new frontier of multi-messenger astrophysics: follow-up of electromagnetic transient counterparts of gravitational wave sources” (PI: E. Capellaro). T.d.S., L.B. and A.S. also acknowledge the financial support of PRIN-INAF 2019 within the project “Probing the geometry of accretion: from theory to observations” (PI: Belloni).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
GRB Gamma-ray burst
GUP Generalized uncertainty principle
FoV Field of view
LIV Lorentz invariance violation
QG Quantum gravity

Notes

1
The name of the model was inspired by the insect, whose ability to fly has been questioned theoretically.
2
This value reflects only the mathematical transformation from the observer frame to the rest frame as, at the moment of emission of two photons, the QG lag between them equals zero.
3
4

References

  1. Rovelli, C.; Smolin, L. Knot theory and quantum gravity. Phys. Rev. Lett. 1988, 61, 1155–1158. [Google Scholar] [CrossRef] [PubMed]
  2. Rovelli, C.; Smolin, L. Loop space representation of quantum general relativity. Nuclear Phys. B 1990, 331, 80–152. [Google Scholar] [CrossRef]
  3. Rovelli, C. Loop Quantum Gravity. Living Rev. Relativ. 1998, 1, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Hossenfelder, S. Minimal Length Scale Scenarios for Quantum Gravity. Living Rev. Relativ. 2013, 16, 2. [Google Scholar] [CrossRef] [Green Version]
  5. Kostelecký, V.A.; Samuel, S. Spontaneous breaking of Lorentz symmetry in string theory. Phys. Rev. D 1989, 39, 683–685. [Google Scholar] [CrossRef] [Green Version]
  6. Kostelecký, V.A.; Samuel, S. Gravitational phenomenology in higher-dimensional theories and strings. Phys. Rev. D 1989, 40, 1886–1903. [Google Scholar] [CrossRef] [Green Version]
  7. Carroll, S.M.; Harvey, J.A.; Kostelecký, V.A.; Lane, C.D.; Okamoto, T. Noncommutative Field Theory and Lorentz Violation. Phys. Rev. Lett. 2001, 87, 141601. [Google Scholar] [CrossRef] [Green Version]
  8. Ferrari, A.F.; Gomes, M.; Nascimento, J.R.; Passos, E.; Petrov, A.Y.; da Silva, A.J. Lorentz violation in the linearized gravity. Phys. Lett. B 2007, 652, 174–180. [Google Scholar] [CrossRef] [Green Version]
  9. Santos, V.; Almeida, C.A.S. On gravity localization under Lorentz violation in warped scenario. Phys. Lett. B 2013, 718, 1114–1118. [Google Scholar] [CrossRef] [Green Version]
  10. Hořava, P. Quantum gravity at a Lifshitz point. Phys. Rev. D 2009, 79, 084008. [Google Scholar] [CrossRef] [Green Version]
  11. Kostelecký, V.A. Gravity, Lorentz violation, and the standard model. Phys. Rev. D 2004, 69, 105009. [Google Scholar] [CrossRef] [Green Version]
  12. Bertolami, O.; Páramos, J. Vacuum solutions of a gravity model with vector-induced spontaneous Lorentz symmetry breaking. Phys. Rev. D 2005, 72, 044001. [Google Scholar] [CrossRef] [Green Version]
  13. Tawfik, A.N.; Magdy, H.; Ali, A.F. Lorentz invariance violation and generalized uncertainty principle. Phys. Part. Nucl. Lett. 2016, 13, 59–68. [Google Scholar] [CrossRef] [Green Version]
  14. Lambiase, G.; Scardigli, F. Lorentz violation and generalized uncertainty principle. Phys. Rev. D 2018, 97, 075003. [Google Scholar] [CrossRef] [Green Version]
  15. Kanzi, S.; Sakallı, İ. GUP modified Hawking radiation in bumblebee gravity. Nuclear Phys. B 2019, 946, 114703. [Google Scholar] [CrossRef]
  16. Övgün, A.; Jusufi, K.; Sakallı, I. Exact traversable wormhole solution in bumblebee gravity. Phys. Rev. D 2019, 99, 024042. [Google Scholar] [CrossRef] [Green Version]
  17. Kanzi, S.; Sakallı, I. Greybody radiation and quasinormal modes of Kerr-like black hole in Bumblebee gravity model. Eur. Phys. J. C 2021, 81, 501. [Google Scholar] [CrossRef]
  18. Delhom, A.; Nascimento, J.R.; Olmo, G.J.; Petrov, A.Y.; Porfírio, P.J. Metric-affine bumblebee gravity: Classical aspects. Eur. Phys. J. C 2021, 81, 287. [Google Scholar] [CrossRef]
  19. Gogoi, D.J.; Dev Goswami, U. Quasinormal modes and Hawking radiation sparsity of GUP corrected black holes in bumblebee gravity with topological defects. J. Cosmol. Astropart. Phys. 2022, 2022, 029. [Google Scholar] [CrossRef]
  20. Gogoi, D.J.; Dev Goswami, U. Tideless traversable wormholes surrounded by cloud of strings in f(R) gravity. J. Cosmol. Astropart. Phys. 2023, 2023, 027. [Google Scholar] [CrossRef]
  21. Neves, J.C.S. Kasner cosmology in bumblebee gravity. Ann. Phys. 2023, 454, 169338. [Google Scholar] [CrossRef]
  22. Amelino-Camelia, G. Are We at the Dawn of Quantum-Gravity Phenomenology? In Toward Quantum Gravity; Kowalski-Glikman, J., Ed.; Springer: Berlin/Heidelberg, Germany, 2000; Volume 541, p. 1. [Google Scholar]
  23. Burderi, L.; Di Salvo, T.; Iaria, R. Quantum clock: A critical discussion on spacetime. Phys. Rev. D 2016, 93, 064017. [Google Scholar] [CrossRef] [Green Version]
  24. Sanchez, N.G. New Quantum Structure of Space-Time. Gravit. Cosmol. 2019, 25, 91–102. [Google Scholar] [CrossRef] [Green Version]
  25. Amelino-Camelia, G.; Ellis, J.; Mavromatos, N.E.; Nanopoulos, D.V.; Sarkar, S. Tests of quantum gravity from observations of γ-ray bursts. Nature 1998, 393, 763–765, Erratum in Nature 1998, 395, 525. [Google Scholar] [CrossRef] [Green Version]
  26. Amelino-Camelia, G.; Ellis, J.; Mavromatos, N.E.; Nanopoulos, D.V. Distance Measurement and Wave Dispersion in a Liouville-String Approach to Quantum Gravity. Int. J. Mod. Phys. A 1997, 12, 607–623. [Google Scholar] [CrossRef] [Green Version]
  27. Ellis, J.; Mavromatos, N.E.; Nanopoulos, D.V.; Sakharov, A.S. Quantum-gravity analysis of gamma-ray bursts using wavelets. Astron. Astrophys. 2003, 402, 409–424. [Google Scholar] [CrossRef]
  28. Ellis, J.; Mavromatos, N.E.; Nanopoulos, D.V.; Sakharov, A.S.; Sarkisyan, E.K.G. Robust limits on Lorentz violation from gamma-ray bursts. Astropart. Phys. 2006, 25, 402–411. [Google Scholar] [CrossRef] [Green Version]
  29. Abdo, A.A.; Ackermann, M.; Ajello, M.; Asano, K.; Atwood, W.B.; Axelsson, M.; Baldini, L.; Ballet, J.; Barbiellini, G.; Baring, M.G.; et al. A limit on the variation of the speed of light arising from quantum gravity effects. Nature 2009, 462, 331–334. [Google Scholar] [CrossRef] [Green Version]
  30. Vasileiou, V.; Granot, J.; Piran, T.; Amelino-Camelia, G. A Planck-scale limit on spacetime fuzziness and stochastic Lorentz invariance violation. Nat. Phys. 2015, 11, 344–346. [Google Scholar] [CrossRef]
  31. Zhang, S.; Ma, B.Q. Lorentz violation from gamma-ray bursts. Astropart. Phys. 2015, 61, 108–112. [Google Scholar] [CrossRef] [Green Version]
  32. Pan, Y.; Gong, Y.; Cao, S.; Gao, H.; Zhu, Z.H. Constraints on the Lorentz Invariance Violation with Gamma-Ray Bursts via a Markov Chain Monte Carlo Approach. Astrophys. J. 2015, 808, 78. [Google Scholar] [CrossRef] [Green Version]
  33. Xu, H.; Ma, B.Q. Light speed variation from gamma ray burst GRB 160509A. Phys. Lett. B 2016, 760, 602–604. [Google Scholar] [CrossRef] [Green Version]
  34. Xu, H.; Ma, B.Q. Light speed variation from gamma-ray bursts. Astropart. Phys. 2016, 82, 72–76. [Google Scholar] [CrossRef] [Green Version]
  35. Chang, Z.; Li, X.; Lin, H.N.; Sang, Y.; Wang, P.; Wang, S. Constraining Lorentz invariance violation from the continuous spectra of short gamma-ray bursts. Chin. Phys. C 2016, 40, 045102. [Google Scholar] [CrossRef] [Green Version]
  36. Wei, J.J.; Zhang, B.B.; Shao, L.; Wu, X.F.; Mészáros, P. A New Test of Lorentz Invariance Violation: The Spectral Lag Transition of GRB 160625B. Astrophys. J. Lett. 2017, 834, L13. [Google Scholar] [CrossRef] [Green Version]
  37. Ganguly, S.; Desai, S. Statistical significance of spectral lag transition in GRB 160625B. Astropart. Phys. 2017, 94, 17–21. [Google Scholar] [CrossRef] [Green Version]
  38. Liu, Y.; Ma, B.Q. Light speed variation from gamma ray bursts: Criteria for low energy photons. Eur. Phys. J. C 2018, 78, 825. [Google Scholar] [CrossRef] [Green Version]
  39. Zou, X.B.; Deng, H.K.; Yin, Z.Y.; Wei, H. Model-independent constraints on Lorentz invariance violation via the cosmographic approach. Phys. Lett. B 2018, 776, 284–294. [Google Scholar] [CrossRef]
  40. Ellis, J.; Konoplich, R.; Mavromatos, N.E.; Nguyen, L.; Sakharov, A.S.; Sarkisyan-Grinbaum, E.K. Robust constraint on Lorentz violation using Fermi-LAT gamma-ray burst data. Phys. Rev. D 2019, 99, 083009. [Google Scholar] [CrossRef] [Green Version]
  41. Wei, J.J. New constraints on Lorentz invariance violation with polarized gamma-ray bursts. Mon. Not. RAS 2019, 485, 2401–2406. [Google Scholar] [CrossRef] [Green Version]
  42. Pan, Y.; Qi, J.; Cao, S.; Liu, T.; Liu, Y.; Geng, S.; Lian, Y.; Zhu, Z.H. Model-independent Constraints on Lorentz Invariance Violation: Implication from Updated Gamma-Ray Burst Observations. Astrophys. J. 2020, 890, 169. [Google Scholar] [CrossRef] [Green Version]
  43. Acciari, V.A.; Ansoldi, S.; Antonelli, L.A.; Arbet Engels, A.; Baack, D.; Babić, A.; Banerjee, B.; Barres de Almeida, U.; Barrio, J.A.; Becerra González, J.; et al. Bounds on Lorentz Invariance Violation from MAGIC Observation of GRB 190114C. Phys. Rev. Lett. 2020, 125, 021301. [Google Scholar] [CrossRef] [PubMed]
  44. Du, S.S.; Lan, L.; Wei, J.J.; Zhou, Z.M.; Gao, H.; Jiang, L.Y.; Zhang, B.B.; Liu, Z.K.; Wu, X.F.; Liang, E.W.; et al. Lorentz Invariance Violation Limits from the Spectral-lag Transition of GRB 190114C. Astrophys. J. 2021, 906, 8. [Google Scholar] [CrossRef]
  45. Agrawal, R.; Singirikonda, H.; Desai, S. Search for Lorentz Invariance Violation from stacked Gamma-Ray Burst spectral lag data. J. Cosmol. Astropart. Phys. 2021, 2021, 029. [Google Scholar] [CrossRef]
  46. Wei, J.J.; Wu, X.F. Testing fundamental physics with astrophysical transients. Front. Phys. 2021, 16, 44300. [Google Scholar] [CrossRef]
  47. Bartlett, D.J.; Desmond, H.; Ferreira, P.G.; Jasche, J. Constraints on quantum gravity and the photon mass from gamma ray bursts. Phys. Rev. D 2021, 104, 103516. [Google Scholar] [CrossRef]
  48. Xiao, S.; Xiong, S.L.; Wang, Y.; Zhang, S.N.; Gao, H.; Zhang, Z.; Cai, C.; Yi, Q.B.; Zhao, Y.; Tuo, Y.L.; et al. A Robust Estimation of Lorentz Invariance Violation and Intrinsic Spectral Lag of Short Gamma-Ray Bursts. Astrophys. J. Lett. 2022, 924, L29. [Google Scholar] [CrossRef]
  49. Desai, S.; Agrawal, R.; Singirikonda, H. Search for Lorentz invariance violation using Bayesian model comparison applied to Xiao et al. GRB spectral lag catalog. Eur. Phys. J. C 2023, 83, 63. [Google Scholar] [CrossRef]
  50. Vasileiou, V.; Jacholkowska, A.; Piron, F.; Bolmont, J.; Couturier, C.; Granot, J.; Stecker, F.W.; Cohen-Tanugi, J.; Longo, F. Constraints on Lorentz invariance violation from Fermi-Large Area Telescope observations of gamma-ray bursts. Phys. Rev. D 2013, 87, 122001. [Google Scholar] [CrossRef] [Green Version]
  51. Abdo, A.A.; Ackermann, M.; Arimoto, M.; Asano, K.; Atwood, W.B.; Axelsson, M.; Baldini, L.; Ballet, J.; Band, D.L.; Barbiellini, G.; et al. Fermi Observations of High-Energy Gamma-Ray Emission from GRB 080916C. Science 2009, 323, 1688. [Google Scholar] [CrossRef] [PubMed]
  52. Acciari, V.A.; Ansoldi, S.; Antonelli, L.A.; Arbet Engels, A.; Baack, D.; Babić, A.; Banerjee, B.; Barres de Almeida, U.; Barrio, J.A.; et al.; MAGIC Collaboration Teraelectronvolt emission from the γ-ray burst GRB 190114C. Nature 2019, 575, 455–458. [Google Scholar] [CrossRef]
  53. Liu, Z.K.; Zhang, B.B.; Meng, Y.Z. Spectral Lag Transition of 32 Fermi Gamma-Ray Bursts and Their Application on Constraining Lorentz Invariance Violation. Astrophys. J. 2022, 935, 79. [Google Scholar] [CrossRef]
  54. Rees, M.J.; Meszaros, P. Unsteady Outflow Models for Cosmological Gamma-Ray Bursts. Astrophys. J. Lett. 1994, 430, L93. [Google Scholar] [CrossRef]
  55. Zhang, B.; Zhang, B.B.; Virgili, F.J.; Liang, E.W.; Kann, D.A.; Wu, X.F.; Proga, D.; Lv, H.J.; Toma, K.; Mészáros, P.; et al. Discerning the Physical Origins of Cosmological Gamma-ray Bursts Based on Multiple Observational Criteria: The Cases of z = 6.7 GRB 080913, z = 8.2 GRB 090423, and Some Short/Hard GRBs. Astrophys. J. 2009, 703, 1696–1724. [Google Scholar] [CrossRef] [Green Version]
  56. Blinnikov, S.I.; Novikov, I.D.; Perevodchikova, T.V.; Polnarev, A.G. Exploding Neutron Stars in Close Binaries. Sov. Astron. Lett. 1984, 10, 177–179. [Google Scholar]
  57. Paczynski, B. Gamma-ray bursters at cosmological distances. Astrophys. J. Lett. 1986, 308, L43–L46. [Google Scholar] [CrossRef]
  58. Eichler, D.; Livio, M.; Piran, T.; Schramm, D.N. Nucleosynthesis, neutrino bursts and γ-rays from coalescing neutron stars. Nature 1989, 340, 126–128. [Google Scholar] [CrossRef]
  59. Paczynski, B. Cosmological gamma-ray bursts. Acta Astron. 1991, 41, 257–267. [Google Scholar]
  60. Mazets, E.P.; Golenetskii, S.V.; Ilinskii, V.N.; Panov, V.N.; Aptekar, R.L.; Gurian, I.A.; Proskura, M.P.; Sokolov, I.A.; Sokolova, Z.I.; Kharitonova, T.V. Catalog of cosmic gamma-ray bursts from the KONUS experiment data. Astrophys. Space Sci. 1981, 80, 3–83. [Google Scholar] [CrossRef]
  61. Kouveliotou, C.; Meegan, C.A.; Fishman, G.J.; Bhat, N.P.; Briggs, M.S.; Koshut, T.M.; Paciesas, W.S.; Pendleton, G.N. Identification of Two Classes of Gamma-Ray Bursts. Astrophys. J. Lett. 1993, 413, L101. [Google Scholar] [CrossRef]
  62. Woosley, S.E. Gamma-Ray Bursts from Stellar Mass Accretion Disks around Black Holes. Astrophys. J. 1993, 405, 273. [Google Scholar] [CrossRef]
  63. Paczyński, B. Are Gamma-Ray Bursts in Star-Forming Regions? Astrophys. J. Lett. 1998, 494, L45–L48. [Google Scholar] [CrossRef] [Green Version]
  64. MacFadyen, A.I.; Woosley, S.E. Collapsars: Gamma-Ray Bursts and Explosions in “Failed Supernovae”. Astrophys. J. 1999, 524, 262–289. [Google Scholar] [CrossRef] [Green Version]
  65. Woosley, S.E.; Bloom, J.S. The Supernova Gamma-Ray Burst Connection. Annu. Rev. Astron Astrophys. 2006, 44, 507–556. [Google Scholar] [CrossRef] [Green Version]
  66. Norris, J.P.; Share, G.H.; Messina, D.C.; Dennis, B.R.; Desai, U.D.; Cline, T.L.; Matz, S.M.; Chupp, E.L. Spectral Evolution of Pulse Structures in Gamma-Ray Bursts. Astrophys. J. 1986, 301, 213. [Google Scholar] [CrossRef]
  67. Norris, J.P.; Marani, G.F.; Bonnell, J.T. Connection between Energy-dependent Lags and Peak Luminosity in Gamma-Ray Bursts. Astrophys. J. 2000, 534, 248–257. [Google Scholar] [CrossRef]
  68. Band, D.L. Gamma-Ray Burst Spectral Evolution through Cross-Correlations of Discriminator Light Curves. Astrophys. J. 1997, 486, 928–937. [Google Scholar] [CrossRef]
  69. Chen, L.; Lou, Y.Q.; Wu, M.; Qu, J.L.; Jia, S.M.; Yang, X.J. Distribution of Spectral Lags in Gamma-Ray Bursts. Astrophys. J. 2005, 619, 983–993. [Google Scholar] [CrossRef] [Green Version]
  70. Zhang, Y.H.; Treves, A.; Celotti, A.; Chiappetti, L.; Fossati, G.; Ghisellini, G.; Maraschi, L.; Pian, E.; Tagliaferri, G.; Tavecchio, F. Four Years of Monitoring Blazar PKS 2155-304 with BeppoSAX: Probing the Dynamics of the Jet. Astrophys. J. 2002, 572, 762–785. [Google Scholar] [CrossRef] [Green Version]
  71. Cheng, L.X.; Ma, Y.Q.; Cheng, K.S.; Lu, T.; Zhou, Y.Y. The time delay of gamma-ray bursts in the soft energy band. Astron. Astrophys. 1995, 300, 746. [Google Scholar]
  72. Wu, B.; Fenimore, E. Spectral Lags of Gamma-Ray Bursts From Ginga and BATSE. Astrophys. J. Lett. 2000, 535, L29–L32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Norris, J.P.; Bonnell, J.T.; Kazanas, D.; Scargle, J.D.; Hakkila, J.; Giblin, T.W. Long-Lag, Wide-Pulse Gamma-Ray Bursts. Astrophys. J. 2005, 627, 324–345. [Google Scholar] [CrossRef] [Green Version]
  74. Hakkila, J.; Giblin, T.W. Quiescent Burst Evidence for Two Distinct Gamma-Ray Burst Emission Components. Astrophys. J. 2004, 610, 361–367. [Google Scholar] [CrossRef]
  75. Salmonson, J.D.; Galama, T.J. Discovery of a Tight Correlation between Pulse Lag/Luminosity and Jet-Break Times: A Connection between Gamma-Ray Bursts and Afterglow Properties. Astrophys. J. 2002, 569, 682–688. [Google Scholar] [CrossRef]
  76. Yi, T.; Liang, E.; Qin, Y.; Lu, R. On the spectral lags of the short gamma-ray bursts. Mon. Not. RAS 2006, 367, 1751–1756. [Google Scholar] [CrossRef] [Green Version]
  77. Ioka, K.; Nakamura, T. Peak Luminosity-Spectral Lag Relation Caused by the Viewing Angle of the Collimated Gamma-Ray Bursts. Astrophys. J. Lett. 2001, 554, L163–L167. [Google Scholar] [CrossRef]
  78. Shen, R.F.; Song, L.M.; Li, Z. Spectral lags and the energy dependence of pulse width in gamma-ray bursts: Contributions from the relativistic curvature effect. Mon. Not. RAS 2005, 362, 59–65. [Google Scholar] [CrossRef] [Green Version]
  79. Lu, R.J.; Qin, Y.P.; Zhang, Z.B.; Yi, T.F. Spectral lags caused by the curvature effect of fireballs. Mon. Not. RAS 2006, 367, 275–289. [Google Scholar] [CrossRef] [Green Version]
  80. Shenoy, A.; Sonbas, E.; Dermer, C.; Maximon, L.C.; Dhuga, K.S.; Bhat, P.N.; Hakkila, J.; Parke, W.C.; Maclachlan, G.A.; Eskandarian, A.; et al. Probing Curvature Effects in the Fermi GRB 110920. Astrophys. J. 2013, 778, 3. [Google Scholar] [CrossRef] [Green Version]
  81. Uhm, Z.L.; Zhang, B. Toward an Understanding of GRB Prompt Emission Mechanism. I. The Origin of Spectral Lags. Astrophys. J. 2016, 825, 97. [Google Scholar] [CrossRef] [Green Version]
  82. Norris, J.P. Gamma-Ray Bursts: The Time Domain. Astrophys. Space Sci. 1995, 231, 95–102. [Google Scholar] [CrossRef]
  83. Ryde, F. Interpretations of gamma-ray burst spectroscopy. I. Analytical and numerical study of spectral lags. Astron. Astrophys. 2005, 429, 869–879. [Google Scholar] [CrossRef]
  84. Norris, J.P.; Bonnell, J.T. Short Gamma-Ray Bursts with Extended Emission. Astrophys. J. 2006, 643, 266–275. [Google Scholar] [CrossRef]
  85. Amati, L.; O’Brien, P.; Götz, D.; Bozzo, E.; Tenzer, C.; Frontera, F.; Ghirlanda, G.; Labanti, C.; Osborne, J.P.; Stratta, G.; et al. The THESEUS space mission concept: Science case, design and expected performances. Adv. Space Res. 2018, 62, 191–244. [Google Scholar] [CrossRef] [Green Version]
  86. Gehrels, N.; Chincarini, G.; Giommi, P.; Mason, K.O.; Nousek, J.A.; Wells, A.A.; White, N.E.; Barthelmy, S.D.; Burrows, D.N.; Cominsky, L.R.; et al. The Swift Gamma-Ray Burst Mission. Astrophys. J. 2004, 611, 1005–1020. [Google Scholar] [CrossRef] [Green Version]
  87. Meegan, C.; Lichti, G.; Bhat, P.N.; Bissaldi, E.; Briggs, M.S.; Connaughton, V.; Diehl, R.; Fishman, G.; Greiner, J.; Hoover, A.S.; et al. The Fermi Gamma-ray Burst Monitor. Astrophys. J. 2009, 702, 791–804. [Google Scholar] [CrossRef] [Green Version]
  88. Aptekar, R.L.; Frederiks, D.D.; Golenetskii, S.V.; Ilynskii, V.N.; Mazets, E.P.; Panov, V.N.; Sokolova, Z.J.; Terekhov, M.M.; Sheshin, L.O.; Cline, T.L.; et al. Konus-W Gamma-Ray Burst Experiment for the GGS Wind Spacecraft. Space Sci. Rev. 1995, 71, 265–272. [Google Scholar] [CrossRef] [Green Version]
  89. Tsvetkova, A.; Svinkin, D.; Karpov, S.; Frederiks, D. Key Space and Ground Facilities in GRB Science. Universe 2022, 8, 373. [Google Scholar] [CrossRef]
  90. Thompson, D.J.; Wilson-Hodge, C.A. Fermi Gamma-ray Space Telescope. arXiv 2022, arXiv:2210.12875. [Google Scholar]
  91. Atwood, W.B.; Abdo, A.A.; Ackermann, M.; Althouse, W.; Anderson, B.; Axelsson, M.; Baldini, L.; Ballet, J.; Band, D.L.; Barbiellini, G.; et al. The Large Area Telescope on the Fermi Gamma-Ray Space Telescope Mission. Astrophys. J. 2009, 697, 1071–1102. [Google Scholar] [CrossRef] [Green Version]
  92. Amelino-Camelia, G.; Smolin, L. Prospects for constraining quantum gravity dispersion with near term observations. Phys. Rev. D 2009, 80, 084017. [Google Scholar] [CrossRef] [Green Version]
  93. Jacob, U.; Piran, T. Lorentz-violation-induced arrival delays of cosmological particles. J. Cosmol. Astropart. Phys. 2008, 2008, 031. [Google Scholar] [CrossRef]
  94. Tsvetkova, A.; Frederiks, D.; Golenetskii, S.; Lysenko, A.; Oleynik, P.; Pal’shin, V.; Svinkin, D.; Ulanov, M.; Cline, T.; Hurley, K.; et al. The Konus-Wind Catalog of Gamma-Ray Bursts with Known Redshifts. I. Bursts Detected in the Triggered Mode. Astrophys. J. 2017, 850, 161. [Google Scholar] [CrossRef] [Green Version]
  95. Fenimore, E.E.; Madras, C.D.; Nayakshin, S. Expanding Relativistic Shells and Gamma-Ray Burst Temporal Structure. Astrophys. J. 1996, 473, 998. [Google Scholar] [CrossRef] [Green Version]
  96. Salmonson, J.D. On the Kinematic Origin of the Luminosity-Pulse Lag Relationship in Gamma-Ray Bursts. Astrophys. J. Lett. 2000, 544, L115–L117. [Google Scholar] [CrossRef] [Green Version]
  97. Kumar, P.; Panaitescu, A. Afterglow Emission from Naked Gamma-Ray Bursts. Astrophys. J. Lett. 2000, 541, L51–L54. [Google Scholar] [CrossRef] [Green Version]
  98. Qin, Y.P. Doppler effect of gamma-ray bursts in the fireball framework. Astron. Astrophys. 2002, 396, 705–713. [Google Scholar] [CrossRef]
  99. Qin, Y.P.; Zhang, Z.B.; Zhang, F.W.; Cui, X.H. Characteristics of Profiles of Gamma-Ray Burst Pulses Associated with the Doppler Effect of Fireballs. Astrophys. J. 2004, 617, 439–460. [Google Scholar] [CrossRef] [Green Version]
  100. Dermer, C.D. Curvature Effects in Gamma-Ray Burst Colliding Shells. Astrophys. J. 2004, 614, 284–292. [Google Scholar] [CrossRef] [Green Version]
  101. Li, Z. Prompt GeV Emission from Residual Collisions in Gamma-Ray Burst Outflows: Evidence from Fermi Observations of Grb 080916c. Astrophys. J. 2010, 709, 525–534. [Google Scholar] [CrossRef] [Green Version]
  102. Zhang, B.B.; Zhang, B.; Liang, E.W.; Fan, Y.Z.; Wu, X.F.; Pe’er, A.; Maxham, A.; Gao, H.; Dong, Y.M. A Comprehensive Analysis of Fermi Gamma-ray Burst Data. I. Spectral Components and the Possible Physical Origins of LAT/GBM GRBs. Astrophys. J. 2011, 730, 141. [Google Scholar] [CrossRef] [Green Version]
  103. Toma, K.; Wu, X.F.; Mészáros, P. An Up-Scattered Cocoon Emission Model of Gamma-Ray Burst High-Energy Lags. Astrophys. J. 2009, 707, 1404–1416. [Google Scholar] [CrossRef] [Green Version]
  104. Tsvetkova, A.; Frederiks, D.; Svinkin, D.; Aptekar, R.; Cline, T.L.; Golenetskii, S.; Hurley, K.; Lysenko, A.; Ridnaia, A.; Ulanov, M. The Konus-Wind Catalog of Gamma-Ray Bursts with Known Redshifts. II. Waiting-Mode Bursts Simultaneously Detected by Swift/BAT. Astrophys. J. 2021, 908, 83. [Google Scholar] [CrossRef]
  105. Burderi, L.; Di Salvo, T.; Riggio, A.; Gambino, A.F.; Sanna, A.; Fiore, F.; Amarilli, F.; Amati, L.; Ambrosino, F.; Amelino-Camelia, G.; et al. GrailQuest and HERMES: Hunting for gravitational wave electromagnetic counterparts and probing space-time quantum foam. In Proceedings of the Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Online, 14–18 December 2020; Volume 11444, p. 114444. [Google Scholar] [CrossRef]
  106. Hopkins, A.M. On the Evolution of Star-forming Galaxies. Astrophys. J. 2004, 615, 209–221. [Google Scholar] [CrossRef] [Green Version]
  107. Bouwens, R.J.; Illingworth, G.D.; Labbe, I.; Oesch, P.A.; Trenti, M.; Carollo, C.M.; van Dokkum, P.G.; Franx, M.; Stiavelli, M.; González, V.; et al. A candidate redshift z ≈ 10 galaxy and rapid changes in that population at an age of 500 Myr. Nature 2011, 469, 504–507. [Google Scholar] [CrossRef] [Green Version]
  108. Hanish, D.J.; Meurer, G.R.; Ferguson, H.C.; Zwaan, M.A.; Heckman, T.M.; Staveley-Smith, L.; Bland-Hawthorn, J.; Kilborn, V.A.; Koribalski, B.S.; Putman, M.E.; et al. The Survey for Ionization in Neutral Gas Galaxies. II. The Star Formation Rate Density of the Local Universe. Astrophys. J. 2006, 649, 150–162. [Google Scholar] [CrossRef] [Green Version]
  109. Thompson, R.I.; Eisenstein, D.; Fan, X.; Dickinson, M.; Illingworth, G.; Kennicutt, R.C.J. Star Formation History of the Hubble Ultra Deep Field: Comparison with the Hubble Deep Field-North. Astrophys. J. 2006, 647, 787–798. [Google Scholar] [CrossRef] [Green Version]
  110. Li, L.X. Star formation history up to z = 7.4: Implications for gamma-ray bursts and cosmic metallicity evolution. Mon. Not. RAS 2008, 388, 1487–1500. [Google Scholar] [CrossRef] [Green Version]
  111. Atteia, J.L. A simple empirical redshift indicator for gamma-ray bursts. Astron. Astrophys. 2003, 407, L1–L4. [Google Scholar] [CrossRef]
  112. Amati, L.; Frontera, F.; Tavani, M.; in’t Zand, J.J.M.; Antonelli, A.; Costa, E.; Feroci, M.; Guidorzi, C.; Heise, J.; Masetti, N.; et al. Intrinsic spectra and energetics of BeppoSAX Gamma-Ray Bursts with known redshifts. Astron. Astrophys. 2002, 390, 81–89. [Google Scholar] [CrossRef]
  113. Reichart, D.E.; Lamb, D.Q.; Fenimore, E.E.; Ramirez-Ruiz, E.; Cline, T.L.; Hurley, K. A Possible Cepheid-like Luminosity Estimator for the Long Gamma-Ray Bursts. Astrophys. J. 2001, 552, 57–71. [Google Scholar] [CrossRef] [Green Version]
  114. Fenimore, E.E.; Ramirez-Ruiz, E. Redshifts For 220 BATSE Gamma-Ray Bursts Determined by Variability and the Cosmological Consequences. arXiv 2000, arXiv:astro–ph/0004176. [Google Scholar]
  115. Yonetoku, D.; Murakami, T.; Nakamura, T.; Yamazaki, R.; Inoue, A.K.; Ioka, K. Gamma-Ray Burst Formation Rate Inferred from the Spectral Peak Energy-Peak Luminosity Relation. Astrophys. J. 2004, 609, 935–951. [Google Scholar] [CrossRef] [Green Version]
  116. Ghisellini, G.; Haardt, F.; Campana, S.; Lazzati, D.; Covino, S. Redshift Determination in the X-ray Band of Gamma-Ray Bursts. Astrophys. J. 1999, 517, 168–173. [Google Scholar] [CrossRef] [Green Version]
  117. D’Isanto, A.; Polsterer, K.L. Photometric redshift estimation via deep learning—Generalized and pre-classification-less, image based, fully probabilistic redshifts. Astron. Astrophys. 2018, 609, A111. [Google Scholar] [CrossRef]
  118. Dainotti, M.; Petrosian, V.; Bogdan, M.; Miasojedow, B.; Nagataki, S.; Hastie, T.; Nuyngen, Z.; Gilda, S.; Hernandez, X.; Krol, D. Gamma-ray Bursts as distance indicators through a machine learning approach. arXiv 2019, arXiv:1907.05074. [Google Scholar]
  119. Lee, J.; Shin, M.S. Estimation of Photometric Redshifts. I. Machine-learning Inference for Pan-STARRS1 Galaxies Using Neural Networks. Astron. J. 2021, 162, 297. [Google Scholar] [CrossRef]
  120. Momtaz, A.; Salimi, M.H.; Shakeri, S. Estimating the Photometric Redshifts of Galaxies and QSOs Using Regression Techniques in Machine Learning. arXiv 2022, arXiv:2201.04391. [Google Scholar]
  121. Gruber, D.; Greiner, J.; von Kienlin, A.; Rau, A.; Briggs, M.S.; Connaughton, V.; Goldstein, A.; van der Horst, A.J.; Nardini, M.; Bhat, P.N.; et al. Rest-frame properties of 32 gamma-ray bursts observed by the Fermi Gamma-ray Burst Monitor. Astron. Astrophys. 2011, 531, A20. [Google Scholar] [CrossRef] [Green Version]
  122. Atteia, J.L.; Heussaff, V.; Dezalay, J.P.; Klotz, A.; Turpin, D.; Tsvetkova, A.E.; Frederiks, D.D.; Zolnierowski, Y.; Daigne, F.; Mochkovitch, R. The Maximum Isotropic Energy of Gamma-ray Bursts. Astrophys. J. 2017, 837, 119. [Google Scholar] [CrossRef] [Green Version]
  123. Minaev, P.Y.; Pozanenko, A.S. The Ep,I-Eiso correlation: Type I gamma-ray bursts and the new classification method. Mon. Not. RAS 2021, 492, 1919–1936, Erratum in Mon. Not. RAS 2021, 504, 926–927. [Google Scholar] [CrossRef] [Green Version]
  124. Lloyd-Ronning, N.M.; Fryer, C.L.; Ramirez-Ruiz, E. Cosmological Aspects of Gamma-Ray Bursts: Luminosity Evolution and an Estimate of the Star Formation Rate at High Redshifts. Astrophys. J. 2002, 574, 554–565. [Google Scholar] [CrossRef] [Green Version]
  125. Kocevski, D.; Liang, E. Quantifying the Luminosity Evolution in Gamma-Ray Bursts. Astrophys. J. 2006, 642, 371–381. [Google Scholar] [CrossRef] [Green Version]
  126. Ukwatta, T.N.; Woźniak, P.R.; Gehrels, N. Machine-z: Rapid machine-learned redshift indicator for Swift gamma-ray bursts. Mon. Not. RAS 2016, 458, 3821–3829. [Google Scholar] [CrossRef] [Green Version]
  127. Amati, L.; O’Brien, P.T.; Götz, D.; Bozzo, E.; Santangelo, A.; Tanvir, N.; Frontera, F.; Mereghetti, S.; Osborne, J.P.; Blain, A.; et al. The THESEUS space mission: Science goals, requirements and mission concept. Exp. Astron. 2021, 52, 183–218. [Google Scholar] [CrossRef]
  128. Ghirlanda, G.; Salvaterra, R.; Toffano, M.; Ronchini, S.; Guidorzi, C.; Oganesyan, G.; Ascenzi, S.; Bernardini, M.G.; Camisasca, A.E.; Mereghetti, S.; et al. Gamma ray burst studies with THESEUS. Exp. Astron. 2021, 52, 277–308. [Google Scholar] [CrossRef]
Figure 1. The dependence of the spectral lag on the energy window for the GRB sample studied in [53]. The fits with a smoothly broken power law (SBPL) are shown by black solid lines. Blue and orange dotted lines denote the maximally allowed LIV-induced lags in linear and quadratic cases, thereby defining the lower limits on the QG energy. This figure is adapted from Figure 1 from [53] (see Section 3 of this work for details). ©AAS. Reproduced with permission.
Figure 1. The dependence of the spectral lag on the energy window for the GRB sample studied in [53]. The fits with a smoothly broken power law (SBPL) are shown by black solid lines. Blue and orange dotted lines denote the maximally allowed LIV-induced lags in linear and quadratic cases, thereby defining the lower limits on the QG energy. This figure is adapted from Figure 1 from [53] (see Section 3 of this work for details). ©AAS. Reproduced with permission.
Universe 09 00359 g001
Figure 2. The cosmological GRB formation rate (GRBFR) derived in [104], superposed onto the star formation rate (SFR) data from the literature. The gray points show the SFR data from [106,107,108,109]. The marked line denotes the SFR approximation from [110]. The GRBFR normalization is equal for all four data sets and the GRBFR points have been shifted arbitrarily to match the SFR at ( 1 z ) 3.5 . Figure 5b from [104]. ©AAS. Reproduced with permission.
Figure 2. The cosmological GRB formation rate (GRBFR) derived in [104], superposed onto the star formation rate (SFR) data from the literature. The gray points show the SFR data from [106,107,108,109]. The marked line denotes the SFR approximation from [110]. The GRBFR normalization is equal for all four data sets and the GRBFR points have been shifted arbitrarily to match the SFR at ( 1 z ) 3.5 . Figure 5b from [104]. ©AAS. Reproduced with permission.
Universe 09 00359 g002
Figure 3. Left plot: Observed GRBs with known redshifts measured in 2005–2020 (blue line and filled cyan area representing 1 σ uncertainty) superimposed on the anticipated frequency of detection of long GRBs by THESEUS (orange histogram). The purple hatched histogram shows the GRBs that are expected to have a redshift measured by either THESEUS or ground-based facilities’ telescopes. The model that fits the observed distribution used to make predictions for THESEUS is represented by the green curve. THESEUS is expected to detect one to two orders of magnitude more GRBs than Swift at any redshift, and most importantly in the high-redshift range ( z > 6 ). Right plot: Cumulative redshift distribution of short GRBs detectable with THESEUS/XGIS per year of mission. Theoretically, short GRBs can be detected at high redshifts z > 4 with a rate of ∼1 event per year. This figure is adopted from [128].
Figure 3. Left plot: Observed GRBs with known redshifts measured in 2005–2020 (blue line and filled cyan area representing 1 σ uncertainty) superimposed on the anticipated frequency of detection of long GRBs by THESEUS (orange histogram). The purple hatched histogram shows the GRBs that are expected to have a redshift measured by either THESEUS or ground-based facilities’ telescopes. The model that fits the observed distribution used to make predictions for THESEUS is represented by the green curve. THESEUS is expected to detect one to two orders of magnitude more GRBs than Swift at any redshift, and most importantly in the high-redshift range ( z > 6 ). Right plot: Cumulative redshift distribution of short GRBs detectable with THESEUS/XGIS per year of mission. Theoretically, short GRBs can be detected at high redshifts z > 4 with a rate of ∼1 event per year. This figure is adopted from [128].
Universe 09 00359 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tsvetkova, A.; Burderi, L.; Riggio, A.; Sanna, A.; Di Salvo, T. A Concept of Assessment of LIV Tests with THESEUS Using the Gamma-Ray Bursts Detected by Fermi/GBM. Universe 2023, 9, 359. https://doi.org/10.3390/universe9080359

AMA Style

Tsvetkova A, Burderi L, Riggio A, Sanna A, Di Salvo T. A Concept of Assessment of LIV Tests with THESEUS Using the Gamma-Ray Bursts Detected by Fermi/GBM. Universe. 2023; 9(8):359. https://doi.org/10.3390/universe9080359

Chicago/Turabian Style

Tsvetkova, Anastasia, Luciano Burderi, Alessandro Riggio, Andrea Sanna, and Tiziana Di Salvo. 2023. "A Concept of Assessment of LIV Tests with THESEUS Using the Gamma-Ray Bursts Detected by Fermi/GBM" Universe 9, no. 8: 359. https://doi.org/10.3390/universe9080359

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop