33email: [email protected] 44institutetext: Xue-Feng Wu 55institutetext: Purple Mountain Observatory, Chinese Academy of Sciences, Nanjing 210023, China 66institutetext: School of Astronomy and Space Sciences, University of Science and Technology of China, Hefei 230026, China
66email: [email protected]
Tests of Lorentz Invariance
Abstract
Lorentz invariance is a fundamental symmetry of both Einstein’s theory of general relativity and quantum field theory. However, deviations from Lorentz invariance at energies approaching the Planck scale are predicted in many quantum gravity theories seeking to unify the force of gravity with the other three fundamental forces of matter. Even though any violations of Lorentz invariance are expected to be very small at observable energies, they can increase with energy and accumulate to detectable levels over large distances. Astrophysical observations involving high-energy emissions and long baselines can therefore offer exceptionally sensitive tests of Lorentz invariance. With the extreme features of astrophysical phenomena, it is possible to effectively search for signatures of Lorentz invariance violations (LIV) in the photon sector, such as vacuum dispersion, vacuum birefringence, photon decay, and photon splitting. All of these potential signatures have been studied carefully using different methods over the past few decades. This chapter attempts to review the status of our current knowledge and understanding of LIV, with particular emphasis on a summary of various astrophysical tests that have been used to set lower limits on the LIV energy scales.
Keywords:
Lorentz invariance, Astroparticle physics, Quantum gravity, Relativity and gravitation1 Introduction
Einstein’s theory of general relativity gives a classical description of gravity, and the Standard Model of particle physics is a well-tested quantum theoretical model of particles and all fundamental forces except gravity. They provide an excellent description of nature together at experimentally attainable energies. However, at the Planck scale ( GeV), a more fundamental quantum theory of gravity is required, which would be a giant leap towards unification of all fundamental forces. A correct and consistent unification scheme has long been the Holy Grail of modern physics. Several quantum gravity (QG) theories that attempt to unify general relativity and the standard Model at the Planck scale predict that Lorentz invariance may be broken at this energy scale, perhaps due to the underlying quantized nature of spacetime 1989PhRvD..39..683K ; 1991NuPhB.359..545K ; 1995PhRvD..51.3923K ; 2005LRR…..8….5M ; 2005hep.ph….6054B ; 2013LRR….16….5A ; 2014RPPh…77f2901T ; 2021FrPhy..1644300W . Therefore, the dedicated experimental searches for violations of Lorentz invariance may help to clear the path to a unification theory. Such experimental tests obtained from a wide range of systems have been compiled in Ref. 2011RvMP…83…11K .
Although any deviations from Lorentz invariance are supposed to be very tiny at attainable energies , they can increase with energy and over large distances due to cumulative process in particle propagation. Astrophysical observations involving high energies and long baselines can therefore provide the most sensitive tests of Lorentz invariance. In the photon sector, the potential signatures of Lorentz invariance violations (LIV) include vacuum dispersion, vacuum birefringence, photon decay, photon splitting, and so on 2008ApJ…689L…1K ; 2001APh….16…97S .
Vacuum dispersion would produce an energy-dependent velocity of light, which would lead to arrival-time differences between promptly emitted photons with different energies traveling over astrophysical distances. Astrophysical time-of-flight measurements can therefore be used to test Lorentz invariance (e.g., Refs. 1998Natur.393..763A ; 2006APh….25..402E ; 2008JCAP…01..031J ; 2008ApJ…689L…1K ; 2009PhRvD..80a5020K ; 2009Sci…323.1688A ; 2009Natur.462..331A ; 2012APh….36…47C ; 2012PhRvL.108w1103N ; 2013PhRvD..87l2001V ; 2013APh….43…50E ; 2015PhRvD..92d5016K ; 2015APh….61..108Z ; 2017ApJ…834L..13W ; 2017ApJ…842..115W ; 2017ApJ…851..127W ; 2019PhRvD..99h3009E ; PhysRevLett.125.021301 ; 1999PhRvL..83.2108B ; Kaaret1999 ). Similarly, the effects of vacuum birefringence would accumulate over cosmological distances resulting in a measurable rotation of the polarization plane of linearly polarized photons as a function of energy. Thus, Lorentz invariance can also be tested with astrophysical polarization measurements (e.g., Refs. 1990PhRvD..41.1231C ; 1998PhRvD..58k6002C ; 2001PhRvD..64h3007G ; 2001PhRvL..87y1304K ; 2006PhRvL..97n0401K ; 2007PhRvL..99a1601K ; 2013PhRvL.110t1601K ; 2003Natur.426Q.139M ; 2004PhRvL..93b1101J ; 2007MNRAS.376.1857F ; 2009JCAP…08..021G ; 2011PhRvD..83l1301L ; 2011APh….35…95S ; 2012PhRvL.109x1104T ; 2013MNRAS.431.3550G ; 2014MNRAS.444.2776G ; 2016MNRAS.463..375L ; 2017PhRvD..95h3013K ; 2019PhRvD..99c5045F ; 2019MNRAS.485.2401W ). Since time-of-flight measurements are less sensitive than polarization measurements by a factor , where is the energy of the light, time-of-flight measurements typically place less stringent limits on LIV 2009PhRvD..80a5020K . However, numerous predicted signals of LIV have no vacuum birefringence, so constraints from time-of-flight measurements are indispensable in a broad-based search for nonbirefringent Lorentz-violating effects.
Besides the measurements of time-of-flight and polarization, astrophysical spectrum measurements can also be used to test Lorentz invariance (e.g., Refs. 2001APh….16…97S ; 2017PhRvD..95f3001M ; 2019JCAP…04..054A ; 2019EPJC…79.1011S ; 2020PhRvL.124m1101A ; 2021arXiv210507927C ; 2021arXiv210612350T ). This is because superluminal LIV allows photons to decay at high energies. Photons decay into electron-positron pairs proceed over very short distances (centimeters or less) once above the energy threshold of the decay process, which would result in a hard cutoff in the gamma-ray spectra of astrophysical sources 2017PhRvD..95f3001M . Another superluminal LIV decay process, photon splitting into multiple photons, also predicts a cutoff at the highest energy part of the photon spectra of astrophysical sources 2019JCAP…04..054A ; 2019EPJC…79.1011S . Experimental searches for LIV-induced spectral cutoff have therefore been carried out in the observed spectra of several astrophysical sources 2017PhRvD..95f3001M ; 2019JCAP…04..054A ; 2019EPJC…79.1011S ; 2020PhRvL.124m1101A ; 2021arXiv210507927C ; 2021arXiv210612350T .
In this chapter, we summarize the current status on astrophysical tests of Lorentz invariance and attempt to chart the future of the subject. We start by reviewing the recent achievements in sensitivity of vacuum dispersion time-of-flight measurements. After that, we overview the progresses on LIV tests through the effects of vacuum birefringence, photon decay, and photon splitting. Finally, we present a summary and future prospect.
2 Vacuum dispersion
2.1 Modified photon dispersion relation
Compared to the standard energy-momentum relationship in Einstein’s special relativity, the introduction of LIV can induce modifications to the photon dispersion relation in vacuum, which can be described using a Taylor series 1998Natur.393..763A :
(1) |
where and are the energy and momentum of the photon, respectively, is the Lorentz-invariant speed of light, and is a dimensionless expansion coefficient. Also, is the hypothetical energy scale at which QG effects become significant. At small energies , the sum is dominated by the lowest-order term of the series. Considering the sensitivity of current detectors, only the first two leading terms ( or ) are of interest for independently experimental searches. They are often referred to as linear and quadratic energy-dependent modifications, respectively. The energy-dependent photon group velocity can therefore be derived as
(2) |
where the sign allows for the “superluminal” () or “subluminal” () scenarios.
Owing to the energy dependence of , two photons with different energies radiated simultaneously from the same source would travel at different speeds, thus arriving at the observer at different times. For example, in the subluminal case (where high-energy photons are slower), the photon with a higher energy (denoted by ) would arrive with respect to the photon with a lower energy () by a time delay 2008JCAP…01..031J
(3) |
where and represent the observed arrival times of the high- and low-energy photons, respectively, and is the redshift of the source. Here , , and are cosmological parameters of the standard CDM model.
In principle, Lorentz invariance can be tested using the observed time delays of photons with different energies arising in astronomical sources. However, it is difficult to exclude the possibility that the energy-dependent delays are of pure astrophysical origins. In order to prove the existence of LIV, one needs to systematically collect numerous delay events and demonstrate that they all match the prediction of LIV (Eq. (3)).
On the contrary, it is relatively simple to place limits on the non-existence of LIV at a certain energy scale. A smaller-than-expected observed delay of high-energy photons with respect to low-energy photons can be used to discard a given model. More specifically, a conservative lower limit on the QG energy scale can be placed under the assumption that the observed time delay is mainly contributed by the LIV effects.
2.2 Present constraints from time-of-flight measurements
It is obvious from Eq. (3) that a potential LIV-induced time delay is predicted to be an increasing function of the photon energy and the source distance. Additionally, a short timescale of the signal variability provides a natural reference time to measure the energy-dependent delay more easily. The greatest sensitivities on can therefore be expected from those violent astrophysical phenomena with high-energy emissions, large distances, and short time delays. The first attempt to constrain LIV by exploiting the rapid variations of gamma-ray emissions from astrophysical sources at cosmological distances was presented in Ref. 1998Natur.393..763A . As of now, transient or variable sources such as gamma-ray bursts (GRBs), active galactic nuclei (AGNs), and pulsars form a group of excellent candidates for searching for the LIV-induced vacuum dispersion.
Gamma-ray bursts
GRBs are among the most distant gamma-ray transients in the Universe and their prompt emission light curves vary on subsecond timescales. As such, GRBs are ideal probes that one can use to perform LIV tests. There have been lots of LIV studies by applying various analysis techniques on the time-of-flight measurements of GRBs.
Some of the pre-Fermi constraints are those by Ref. Ellis2003 using BATSE GRBs; by Ref. 2004ApJ…611L..77B using RHESSI observations of GRB 021206; by Ref. 2008ApJ…676..532B using HETE-2 GRBs; by Ref. 2008GReGr..40.1731L using INTEGRAL GRBs; by Ref. 2006JCAP…05..017R using Konus-Wind and Swift observations of GRB 051221A; by Refs. 2006APh….25..402E ; 2008APh….29..158E using BATSE, HETE-2, and Swift GRBs; and by Ref. 2017ApJ…851..127W using Swift GRBs. Much more stringent constraints on LIV, however, have been obtained using Fermi observations, entirely thanks to the unprecedented sensitivity for detecting GRB prompt MeV/GeV emission by the Large Area Telescope (LAT) onboard the Fermi satellite. These tighter constraints include those by the Fermi Gamma-Ray Burst Monitor (GBM) and LAT Collaborations using GRBs 080916C 2009Sci…323.1688A and 090510 2009Natur.462..331A ; by Ref. 2009PhRvD..80k6005X using GRB 090510; and by Refs. 2010APh….33..312S ; 2012APh….36…47C ; 2012PhRvL.108w1103N ; 2013PhRvD..87l2001V ; 2019PhRvD..99h3009E using multiple Fermi GRBs. Remarkably, Ref. 2009Natur.462..331A studied the time-of-flight effect by analyzing the arrival time lag between the highest energy (31 GeV) photon and low-energy photons from GRB 090510. The burst was localized at . Fermi/LAT detected this 31 GeV photon 0.829 s after the GBM trigger. Assuming the 31 GeV photon was emitted at the GBM trigger time, the time delay between the highest energy photon and low-energy (trigger) photons is then ms. This gives the most conservative constraint on the linear LIV energy scale . If the 31 GeV photon is assumed to be associated with the contemporaneous MeV spike, one therefore has ms, which gives the most radical constraint . These results obviously disfavor the linear LIV models requiring , and are in good agreement with Lorentz invariance. Subsequently, Ref. 2013PhRvD..87l2001V adopted three different techniques to constrain the degree of dispersion observed in the data of four Fermi/LAT GRBs. For the subluminal case, their constraints from GRB 090510, namely and for linear and quadratic leading-order LIV-induced vacuum dispersion, further improved previous results by a factor of . On 2019 January 14, the Major Atmospheric Gamma Imaging Cherenkov (MAGIC) telescopes discovered a gamma-ray signal above 0.2 TeV from GRB 190114C, recording the most energetic photons ever detected from a GRB 2019Natur.575..455M . The MAGIC Collaboration used this unique observation to test the dependence of the speed of light in vacuum on its energy PhysRevLett.125.021301 . They obtained competitive lower limits on the quadratic LIV energy scale, namely () for the subluminal (superluminal) case.
Generally, Lorentz invariance has been tested with high accuracy using the spectral lags111Spectral lag is defined as the arrival time delay between correlated photons with different energies (or between light curves in different energy bands). of GRBs. A key challenge in such time-of-flight tests, however, is to distinguish an intrinsic time lag at the source from a delay induced by LIV. Possible source-intrinsic effects caused by the unknown emission mechanism could cancel-out or enhance the LIV-induced time delay, which would impact the reliability of the resulting constraints on LIV.
Any effects of LIV would increase with the redshift of the source (see Eq. (3)), whereas source-intrinsic effects might be independent of the redshift. Therefore, Refs. Ellis2003 ; 2006APh….25..402E ; 2008APh….29..158E proposed working on a statistical sample of GRBs with different redshifts to disentangle the intrinsic time lag problem. For each GRB, Ref. 2006APh….25..402E extracted the spectral lag in the light curves recorded in the selected observer-frame energy bands 25–55 and 115–320 keV. To account for the unknown intrinsic time lag, they fitted the observed arrival time delays of 35 GRBs with the inclusion of a term specified in the GRB source frame. That is, the observed spectral lags are fitted by two components, , reflecting the contributions from both the LIV and source-intrinsic effects 2006APh….25..402E . Such parametrization allows one to derive a simple linear fitting function:
(4) |
where
(5) |
is a non-linear redshift function which is related to the specific cosmological model, the slope is connected to the energy scale of LIV, and the intercept stands for the unknown intrinsic time lag inherited from the sources. Also, is the dimensionless expansion rate at , where the standard flat CDM model with parameters and is adopted. Note that the linear LIV correction () for the subluminal case () was considered in this work. Ref. 2008APh….29..158E compiled the rescaled time lags extracted from 35 light curve pairs as functions of the variable and performed a linear regression analysis of the available data. The linear fit corresponds to . With the corresponding upper limit on the slope parameter , as well as the energy difference between two observer-frame energy bands , the 95% confidence-level lower limit on the linear LIV energy scale is GeV 2008APh….29..158E . Going beyond the CDM cosmology, Refs. 2009CQGra..26l5007B ; 2015ApJ…808…78P extended this analysis to other different cosmological models, finding the result is only weakly sensitive to the background cosmology. Some cosmology-independent techniques were furthermore applied to this kind of analysis 2018PhLB..776..284Z ; 2020ApJ…890..169P .
It is worth pointing out that Ref. 2006APh….25..402E extracted the spectral lag of each GRB in the light curves between two fixed observer-frame energy bands. However, due to the fact that different GRBs have different redshift measurements, these two observer-frame energy bands actually correspond to a different pair of energy bands in the GRB source frame 2012MNRAS.419..614U , thus potentially introducing an artificial energy dependence to the extracted spectral lag and/or systematic uncertainties to the search for LIV-induced lags.222For those cosmological sources with various redshifts, their observer-frame quantities can be quite different than the corresponding source-frame ones. However, note that if we focus on the observer-frame quantities of individual cosmological sources, there is no such a problem. This problem can be solved by choosing two appropriate energy bands fixed in the GRB source frame and estimating the observed time lag for two projected observer-frame energy bands by the relation , Ref. 2012MNRAS.419..614U showed that the correlation between observer-frame lags and source-frame lags for the same GRB sample has a large scatter, supporting that the source-frame lag can not be directly represented by the observer-frame lag. Therefore, when analyzing the LIV effects in a large sample of cosmological sources with different redshifts, it would be good to use the spectral lags extracted in the source frame. Ref. 2015MNRAS.446.1129B investigated the source-frame spectral lags of 56 GRBs observed by Swift/Burst Alert Telescope (BAT). This sample has redshifts ranging from 0.35 (GRB 061021) to 5.47 (GRB 060927). For each GRB, they extracted light curves for two observer-frame energy bands corresponding to the fixed source-frame energy bands 100–150 and 200–250 keV. These two specific source-frame energy bands were chosen so that after transforming to the observer frame (based on the redshift of each GRB, i.e., and keV) they still lie in the detectable energy range of Swift/BAT. Finally, for each light-curve pairs, Ref. 2015MNRAS.446.1129B used the improved cross-correlation function analysis technology to estimate the spectral lag. Note that the energy difference between the mid-points of the two source-frame energy bands is fixed at 100 keV, whereas in the observer frame (as expected), the energy difference varies depending on the redshift of each GRB. For example, the energy difference is 16 keV in GRB 060927 and 74 keV in GRB 061021. This is in contrast to the spectral lag extractions performed in the observer frame, where the energy difference is treated as a constant 2012MNRAS.419..614U . Ref. 2017ApJ…851..127W first took advantage of the source-frame spectral lags of 56 GRBs presented in Ref. 2015MNRAS.446.1129B to study the LIV effects. For the subluminal case, the linear LIV-induced time delay between two observer-frame energy bands with the energy difference is given by
(6) |
where is the source-frame energy difference. Similar to the treatment of Ref. 2006APh….25..402E , one can formulate the intrinsic time delay problem in terms of linear regression:
(7) |
where is the extracted spectral lag for two fixed source-frame energy bands 100–150 and 200–250 keV,
(8) |
is a function of the redshift, and is the slope in . In order to probe the LIV effects, Ref. 2017ApJ…851..127W performed a linear fit to the versus data. With the optimized and the source-frame energy difference , the 95% confidence-level lower limit on the linear LIV energy scale is GeV 2017ApJ…851..127W . This is a step forward in the investigation of LIV effects, since all previous studies used arbitrary observer-frame energy bands.
Another point to note is that in the treatment of Ref. 2006APh….25..402E an unknown constant was assumed to be the intrinsic time lag in the linear fitting function (see Eq. 4). That is, their work based on the assumption that all GRBs have the same intrinsic time delay. However, since the durations of GRBs span about five or six orders of magnitude, it is unlikely that high-energy photons emitted from different GRBs (or from the same GRB) have the same intrinsic time lag relative to the emission time of low-energy photons 2016ChPhC..40d5102C . As an improvement, Ref. 2012APh….36…47C estimated the intrinsic time delay between high- and low-energy photons emitted from each GRB by using the magnetic jet model. However, the magnetic jet model relies on some given theoretical parameters, and this introduces uncertainties on the LIV results.
In 2017, Ref. 2017ApJ…834L..13W used multiple spectral lags from GRB 160625B to constrain LIV in the photon sector. GRB 160625B was detected by the Fermi satellite on 2016 June 25, consisting of three different isolated sub-bursts with unusually high photon statistics. Since the second sub-burst of GRB 160625B is especially bright, it is easy to extract its high-quality light curves in different energy bands (see Figure 1). With multi-photon energy bands, Ref. 2017ApJ…834L..13W calculated the spectral lags in the light curves recorded in the lowest-energy band (10–12 keV) relative to any other light curves with higher-energy bands, finding that the observed spectral lag increases at MeV and then gradually decreases in the energy range MeV. The lag behavior is very peculiar, being an obvious transition from positive to negative spectral lags discovered within a burst (see Figure 2). Conventionally, a positive spectral lag means an earlier arrival time for photons with higher energies, whereas a negative spectral lag corresponds to a delayed arrival of the higher-energy photons. Ref. 2017ApJ…834L..13W proposed that the spectral-lag transition of GRB 160625B provides a great opportunity to distinguish LIV-induced propagation effects from any source-intrinsic time delay in the emission of photons at different energy bands. Since the time delay caused by LIV may probably be accompanied by an intrinsic energy-dependent time delay arise from the unknown emission mechanism of the source, the observed spectral lag should be comprised of two contributions,
(9) |
As the observed spectral lags of most GRBs have a positive energy dependence (i.e., high-energy photons arrive earlier than low-energy ones; see 2017ApJ…844..126S ; 2018ApJ…865..153L ), Ref. 2017ApJ…834L..13W suggested that the observer-frame correlation between the intrinsic time lag and the energy can be expressed approximately by a power-law with positive dependence,
(10) |
where and , and keV is the median value of the reference lowest-energy band (10–12 keV). Also, when the subluminal case () is considered, the LIV effects predict a negative spectral lag (i.e., high-energy photons travel slower than low-energy ones). As the Lorentz-violating term takes the lead at higher energies, the positive correlation of the lag with energy would gradually become an anti-correlation. The contributions from both the intrinsic energy-dependent time lag and the LIV-induced time delay can therefore result in the observed lag behavior with an apparent transition from positive to negative lags 2017ApJ…834L..13W . By directly fitting the spectral lag behavior of GRB 160625B, Ref. 2017ApJ…834L..13W obtained both comparatively robust limits on the QG energy scales and a reasonable formulation of the intrinsic energy-dependent time lag. The best-fit theoretical curves for the linear and quadratic LIV models are shown in Figure 2. The confidence-level lower limits on the linear and quadratic LIV energy scales are and , respectively. The multi-photon spectral-lag data of GRB 160625B has also been used to constrain the Lorentz-violating coefficients of the Standard Model Extension 2017ApJ…842..115W . Recently, Ref. 2021ApJ…906….8D showed decisive evidence that GRB 190114C is an additional burst having a well-defined transition from positive to negative lags, providing another opportunity not only to disentangle the intrinsic time lag problem but also to place robust constraints on LIV. Although the spectral-lag transitions of GRB 160625B and GRB 190114C do not currently have the best sensitivity to LIV constraints, the presented method, when applied to future bright short GRBs with similar lag features, may derive more stringent constraints on LIV.
Active galactic nuclei
Because of their rapid flux variability (ranging from minutes to years), very-high-energy (VHE, ) emissions, and large distances, flaring AGNs have also been considered as promising sources for LIV studies. It should be underlined that searching for LIV-induced vacuum dispersions using both GRBs and VHE flares of AGNs is of great fundamental interest. GRBs can be detected up to very high redshifts (), but with very limited statistics of photons above a few tens of GeV. Conversely, AGN flares can be observed by ground-based detectors with large statistics of photons up to a few tens of TeV. Due to the absorption of VHE gamma-rays by extragalactic background light, TeV measurements are limited to sources with relatively low redshifts. Thus, GRBs and AGNs are complementary to each other in testing LIV, and they enable us to test different energy and redshift ranges.
There have been some competitive constraints on LIV-induced vacuum dispersions using VHE gamma-ray observations of bright AGN flares, including the Whipple analysis of the flare of Mrk 421 1999PhRvL..83.2108B , the H.E.S.S. analyses of the flares of PKS 2155-304 2008PhRvL.101q0402A ; 2011APh….34..738H and PG 1553+113 2015ApJ…802…65A , and the MAGIC and H.E.S.S. analyses of the flares of Mrk 501 2008PhLB..668..253M ; 2009APh….31..226M ; 2019ApJ…870…93A . For the linear LIV effect considering a subluminal case, the limit obtained from the PKS 2155-304 flare data by H.E.S.S. is the best time-of-flight limit obtained with an AGN, yielded 2011APh….34..738H . For the quadratic LIV effect, the limits obtained from H.E.S.S.’s observation of the Mrk 501 flare are the most stringent constraints ever obtained with an AGN, namely () for the subluminal (superluminal) case 2019ApJ…870…93A .
Pulsars
Gamma-ray pulsars are a third class of astrophysical sources that are often used for LIV time-of-flight tests. There are several reasons why they are a tempting target for these types of studies. First, due to the stable and cyclical nature of pulsar emission, sensitivity to LIV can be systematically planned and improved by observing longer. Second, as the timing of the pulsar is widely studied throughout the electromagnetic spectrum, energy-dependent time delays caused by propagation effects can be more easily distinguished from source-intrinsic effects. Moreover, although being observed several orders of magnitude closer than GRBs or AGNs, the detection of VHE emission from pulsars will compensate for a short distance when it comes to testing the quadratic LIV term.
First limits on LIV using gamma-ray emission from the galactic Crab pulsar were obtained from the observation of CGRO/EGRET at energies above 2 GeV Kaaret1999 , and then improved by VERITAS using VHE gamma-rays reaching up to 120 GeV 2011ICRC….7..256O ; 2013ICRC…33.2768Z . Remarkably, Ref. 2017ApJS..232….9M used the Crab pulsar emission observed up to TeV energies by MAGIC to put new constraints on the energy scale . The 95% confidence-level lower limits for the subluminal (superluminal) case are () and () for linear and quadratic LIV effects, respectively 2017ApJS..232….9M .
Source(s) | Instrument | Technique | [GeV] | [GeV] | Refs. |
GRB | |||||
9 GRBsa | BATSE+OSSE | Wavelets | () | () | Ellis2003 |
GRB 021206b | RHESSI | Peak times at different energies | () | 2004ApJ…611L..77B | |
15 GRBs | HETE-2 | Wavelets | () | 2008ApJ…676..532B | |
11 GRBs | INTEGRAL | Likelihood | () | 2008GReGr..40.1731L | |
GRB 051221A | Konus-Wind+Swift | Peak times at different energies | () | () | 2006JCAP…05..017R |
35 GRBsc | BATSE+HETE-2 | Observer-frame spectral lags | () | 2006APh….25..402E ; 2008APh….29..158E | |
+Swift | |||||
56 GRBs | Swift | Source-frame spectral lags | () | 2017ApJ…851..127W | |
GRB 080916C | Fermi GBM+LAT | Associating a 13.2 GeV photon with | () | 2009Sci…323.1688A | |
the trigger time | |||||
GRB 090510 | Fermi GBM+LAT | Associating a 31 GeV photon with | () | 2009Natur.462..331A | |
the start of the first GBM pulse | |||||
GRB 090510 | Fermi/LAT | PairView+Likelihood | () | () | 2013PhRvD..87l2001V |
+Sharpness-Maximization Method | () | () | |||
GRB 190114C | MAGIC | Likelihood | () | () | PhysRevLett.125.021301 |
() | () | ||||
GRB 160625B | Fermi/GBM | Spectral-lag transition | () | () | 2017ApJ…834L..13W |
GRB 190114C | Fermi/GBM | Spectral-lag transition | () | () | 2021ApJ…906….8D |
AGN | |||||
Mrk 421 | Whipple | Binning | () | 1999PhRvL..83.2108B | |
PKS 2155-304 | H.E.S.S. | Modified cross correlation function | () | () | 2008PhRvL.101q0402A |
PKS 2155-304 | H.E.S.S. | Likelihood | () | () | 2011APh….34..738H |
PG 1553+113 | H.E.S.S. | Likelihood | () | () | 2015ApJ…802…65A |
() | () | ||||
Mrk 501 | MAGIC | Energy cost function | () | () | 2008PhLB..668..253M |
Mrk 501 | MAGIC | Likelihood | () | () | 2009APh….31..226M |
Mrk 501 | H.E.S.S. | Likelihood | () | () | 2019ApJ…870…93A |
() | () | ||||
Pulsar | |||||
Crab | CGRO/EGRET | Pulse arrival times | () | Kaaret1999 | |
in different energy bands | |||||
Crab | VERITAS | Likelihood | () | () | 2011ICRC….7..256O |
Crab | VERITAS | Dispersion Cancellation | () | 2013ICRC…33.2768Z | |
() | |||||
Crab | MAGIC | Likelihood | () | () | 2017ApJS..232….9M |
() | () |
aLimit obtained taking no account of the factor in the integral of Eq. (5).
bThe spectral and temporal properties of GRB 021206 were used to estimate the pseudo redshift.
cThe Limits of Ref. 2006APh….25..402E were corrected in Ref. 2008APh….29..158E
by taking account of the factor in the integral of Eq. (5).
The most important results obtained to date with vacuum dispersion time-of-flight measurements of various astrophysical events are listed in Table 1. The most stringent limits so far on have been obtained with Fermi/LAT’s observation of GRB 090510. The limits set are () for a linear, and () for a quadratic LIV, for the subluminal (superluminal) case 2013PhRvD..87l2001V . Obviously, these time-of-flight studies using gamma-ray emissions in the GeV–TeV range provide us at present with the best opportunity to search for Planck-scale modifications of the photon dispersion relation. Nevertheless, while they offer meaningful constraints for the linear () modification, they are still much weaker for deviations that appear at the quadratic () order.
3 Vacuum birefringence
3.1 General formulae
The Charge-Parity-Time (CPT) theorem states that the physical laws are invariant under charge conjugation, parity transformation, and time reversal. In some QG theories that invoke LIV, the CPT theorem no longer holds. In the absence of Lorentz symmetry, the CPT invariance, if needed, should be imposed as an additional assumption of the theory. In the photon sector, these theories invoke a Lorentz- and CPT-violating dispersion relation of the form 2003PhRvL..90u1601M
(11) |
where the sign corresponds to the helicity, i.e., two different circular polarization states, and is a dimensionless parameter that needs to be bound. Note that the parameter exactly vanishes in LIV but CPT invariant theories. Therefore, in this sense, such tests might be less broad and general than those tests based on vacuum dispersion.
The linear polarization can be decomposed into right- or left-handed circular polarization states. If , then the dispersion relation (11) indicates that photons with opposite circular polarizations have slightly different phase velocities and therefore propagate with different speeds, leading to an energy-dependent rotation of the polarization vector of a linearly polarized light. This effect is known as vacuum birefringence. The rotation angle during the propagation from the source at redshift to the observer is expressed as 2011PhRvD..83l1301L ; 2012PhRvL.109x1104T
(12) |
where is the observed photon energy, and is a function related to the cosmological model, which reads (in the standard flat CDM model)
(13) |
Generally, it is hard to know the intrinsic polarization angles for photons emitted with different energies from a given astrophysical source. If one possessed this information, evidence for vacuum birefringence (i.e., an energy-dependent rotation of the polarization plane of linearly polarized photons) could be directly obtained by measuring differences between the known intrinsic polarization angles and the actual observed polarization angles at different energies. Even in the absence of such knowledge, however, the birefringent effect can still be tested by polarized sources at arbitrary redshifts. This is because a large degree of birefringence would add opposite oriented polarization vectors, effectively washing out most, if not all, of the net polarization of the signal. Therefore, the polarization detections with high significance levels can put upper bounds on the energy-dependent birefringent effect.
3.2 Present constraints from polarization measurements
Observations of linear polarization from distant astrophysical sources can be used to place upper bounds on the birefringent parameter . The vacuum birefringence constraints stem from the fact that if the rotation angles of photons with different energies (Eq. 12) differ by more than over an energy range (), then the net polarization of the signal would be significantly depleted and could not be as high as the observed level. Thus, the detection of highly polarized photons implies that the differential rotation angle should be smaller than , i.e.,
(14) |
Previously, Ref. 2001PhRvD..64h3007G used the observed linear polarization of ultraviolet light from a galaxy 3C 256 at a distance of around 300 Mpc to set an upper bound of . Ref. 2008PhRvD..78j3003M derived a tighter constraint of using the hard X-ray polarization observation from the Crab Nebula. Much stronger limits of were obtained by Refs. 2003Natur.426Q.139M ; 2004PhRvL..93b1101J using a report of polarized gamma-rays observed in the prompt emission from GRB 021206 2003Natur.423..415C . However, this claimed polarization detection has been refuted by subsequent re-analyses of the same data 2004MNRAS.350.1288R ; 2004ApJ…613.1088W . Using the reported detection of polarized soft gamma-ray emission from GRB 041219A 2007ApJS..169…75K ; 2007A&A…466..895M ; 2009ApJ…695L.208G , Refs. 2011PhRvD..83l1301L ; 2011APh….35…95S obtained a stringent upper limit on of and , respectively. But again, the previous reports of the gamma-ray polarimetry for GRB 041219A have been disputed (see 2012PhRvL.109x1104T for more explanations), and the arguments for the resulting constraints given by Refs. 2011PhRvD..83l1301L ; 2011APh….35…95S are still open to questions.
Contrary to those controversial reports, the detection of gamma-ray linear polarization by the Gamma-ray burst Polarimeter (GAP) onboard the Interplanetary Kite-craft Accelerated by Radiation Of the Sun (IKAROS) is fairly credible and thus can be used to set a reliable limit on the birefringent parameter 2012PhRvL.109x1104T . IKAROS/GAP clearly detected linear polarizations in the prompt gamma-ray emission of three GRBs, with a polarization degree of for GRB 100826A 2011ApJ…743L..30Y , for GRB 110301A 2012ApJ…758L…1Y , and for GRB 110721A 2012ApJ…758L…1Y . The detection significance is , , and , respectively. With the assumption that , Ref. 2012PhRvL.109x1104T applied the polarization data of these three GRBs to derive a severe upper bound on in the order of . Note that since the redshifts of these GRBs were not measured, Ref. 2012PhRvL.109x1104T used a luminosity indicator for GRBs to estimate the possible redshifts. Utilizing the real redshift determination () together with the polarimetric data of GRB 061122, Ref. 2013MNRAS.431.3550G derived a more stringent limit () on the possibility of LIV based on the vacuum birefringent effect. The current deepest limit of was obtained by Ref. 2014MNRAS.444.2776G using the most distant polarized burst GRB 140206A at redshift .
All the polarization constraints presented above were based on the assumption that the observed polarization degree would be severely suppressed for a given energy range if the differential rotation angle is too large, regardless of the intrinsic polarization fraction at the corresponding rest-frame energy range. However, Ref. 2016MNRAS.463..375L gave a detailed calculation on the evolution of GRB polarization induced by the vacuum birefringent effect, and showed that, even if is as large as , more than 60% of the initial polarization degree can be conserved. This is in conflict with the general perception that could not be larger than when high polarization is observed. Thus, Ref. 2016MNRAS.463..375L suggested that it is inappropriate to simply use as the upper limit of to constrain the vacuum birefringent effect. Applying their formulae for calculating the polarization evolution to two true GRB events, Ref. 2016MNRAS.463..375L obtained the most conservative limits on the birefringent parameter from the polarimetric data of GRB 061122 and GRB 110721A as and , respectively. Following the analysis method proposed in Ref. 2016MNRAS.463..375L and using the latest detections of prompt emission polarization in GRBs, Ref. 2019MNRAS.485.2401W improved existing bounds on a deviation from Lorentz invariance through the vacuum birefringent effect by factors ranging from two to ten.
Instead of requiring the more complicated and indirect assumption that , some authors simply assume that all photons in the observed energy band are emitted with the same (unknown) intrinsic polarization angle 2007MNRAS.376.1857F ; 2020EPJP..135..527W ; 2021Galax…9…44Z . If the rotation angle of the linear polarization plane arising from the birefringent effect is considered here, the observed linear polarization angle at a certain with an intrinsic polarization angle should be
(15) |
Since is assumed to be an unknown constant, we expect to observe the birefringent effect as an energy-dependent linear polarization vector. Such an observation could give a robust limit on the birefringent parameter . Ref. 2007MNRAS.376.1857F searched for the energy-dependent change of the polarization angle in the spectropolarimetric observations of the optical afterglows of GRB 020813 and GRB 021004. By fitting the multiband polarimetric data of these two GRBs with Eq. (15), Ref. 2007MNRAS.376.1857F obtained constraints on both and (see Figure 3). At the confidence level, the joint constraint on from two GRBs is (see also 2020EPJP..135..527W ). Ref. 2021Galax…9…44Z applied the same analysis method to multiband optical polarimetry of five blazars and obtained constraints with similar accuracy. It is clear from Eq. (12) that the larger the distance of the polarized source, and the higher the energy band of the polarization observation, the greater the sensitivity to small values of . As expected, less stringent constraints on were obtained from the optical polarization data 2007MNRAS.376.1857F ; 2020EPJP..135..527W ; 2021Galax…9…44Z .

Source | Distance | Polarimeter | Energy Ranga | Refs. | ||
Galaxy 3C 256 | Mpc | Spectropolarimeter | Ultraviolet | 2001PhRvD..64h3007G | ||
Crab Nebula | kpc | INTEGRAL/SPI | 100–1000 keV | 2008PhRvD..78j3003M | ||
GRB | ||||||
GRB 021206 | light yearsc | RHESSI | 150–2000 keV | b | 2003Natur.426Q.139M | |
GRB 021206 | Gpcd | RHESSI | 150–2000 keV | b | 2004PhRvL..93b1101J | |
GRB 041219A | e | INTEGRAL/IBIS | 200–800 keV | b | 2011PhRvD..83l1301L | |
GRB 041219A | f | INTEGRAL/SPI | 100–350 keV | b | 2011APh….35…95S | |
GRB 100826A | f | IKAROS/GAP | 70–300 keV | 2012PhRvL.109x1104T | ||
GRB 110301A | f | IKAROS/GAP | 70–300 keV | 2012PhRvL.109x1104T | ||
GRB 110721A | f | IKAROS/GAP | 70–300 keV | 2012PhRvL.109x1104T | ||
GRB 061122 | INTEGRAL/IBIS | 250–800 keV | 2013MNRAS.431.3550G | |||
GRB 140206A | INTEGRAL/IBIS | 200–400 keV | 2014MNRAS.444.2776G | |||
GRB 061122 | INTEGRAL/IBIS | 250–800 keV | 2016MNRAS.463..375L | |||
GRB 110721A | f | IKAROS/GAP | 70–300 keV | 2016MNRAS.463..375L | ||
GRB 061122 | INTEGRAL/IBIS | 250–800 keV | 2019MNRAS.485.2401W | |||
GRB 100826A | f | IKAROS/GAP | 70–300 keV | 2019MNRAS.485.2401W | ||
GRB 110301A | f | IKAROS/GAP | 70–300 keV | 2019MNRAS.485.2401W | ||
GRB 110721A | IKAROS/GAP | 70–300 keV | 2019MNRAS.485.2401W | |||
GRB 140206A | INTEGRAL/IBIS | 200–400 keV | 2019MNRAS.485.2401W | |||
GRB 160106A | f | AstroSat/CZTI | 100–300 keV | 2019MNRAS.485.2401W | ||
GRB 160131A | AstroSat/CZTI | 100–300 keV | 2019MNRAS.485.2401W | |||
GRB 160325A | f | AstroSat/CZTI | 100–300 keV | 2019MNRAS.485.2401W | ||
GRB 160509A | AstroSat/CZTI | 100–300 keV | 2019MNRAS.485.2401W | |||
GRB 160802A | f | AstroSat/CZTI | 100–300 keV | 2019MNRAS.485.2401W | ||
GRB 160821A | f | AstroSat/CZTI | 100–300 keV | 2019MNRAS.485.2401W | ||
GRB 160910A | f | AstroSat/CZTI | 100–300 keV | 2019MNRAS.485.2401W | ||
GRB 020813 | LRISp | 3500–8800 Å | 1.8%–2.4% | (–1.4) | 2007MNRAS.376.1857F | |
GRB 021004 | VLT | 3500–8600 Å | (–1.4) | 2007MNRAS.376.1857F | ||
GRB 020813 | LRISp | 3500–8800 Å | 1.8%–2.4% | (–1.1) | 2020EPJP..135..527W | |
GRB 021004 | VLT | 3500–8600 Å | (–1.1) | 2020EPJP..135..527W | ||
Blazar | ||||||
3C 66A | Photopolarimeter | 14.1%–33.7% | 2021Galax…9…44Z | |||
S5 0716+714 | Photopolarimeter | 15.1%–20.1% | 2021Galax…9…44Z | |||
OJ 287 | Photopolarimeter | 5.2%–22.8% | 2021Galax…9…44Z | |||
MK 421 | Photopolarimeter | 1.7%–3.5% | 2021Galax…9…44Z | |||
PKS 2155-304 | Photopolarimeter | 3%–7% | 2021Galax…9…44Z |
aThe energy range in which polarization is observed.
bThe claimed polarization detections have been refuted.
cThe distance of GRB 021206 was assumed to be light years.
dThe distance of GRB 021206 was assumed to be 0.5 Gpc.
eThe lower limit to the photometric redshift of GRB 041219A was adopted.
fThe redshifts of these GRBs were estimated by the empirical luminosity relation.
Table 2 presents a summary of astrophysical polarization constraints on a possible LIV through the vacuum birefringent effect. The current best limits on the birefringent parameter, , have been obtained by the detections of gamma-ray linear polarization of GRBs 2013MNRAS.431.3550G ; 2014MNRAS.444.2776G ; 2016MNRAS.463..375L ; 2019MNRAS.485.2401W .
4 Photon decay and photon splitting
There are various forms of modified dispersion relation for different particles and underlying QG theories. One of the Lorentz-violating dispersion relations for photons takes the generalized form 1998Natur.393..763A
(16) |
where is the leading order of the modification from the underlying QG theory, is the th order LIV parameter, and the sign refers to the superluminal () and subluminal () cases. For , limits on can be interpreted in terms of the QG energy scale,
(17) |
4.1 Photon decay
In the superluminal LIV case, photons can decay into electron-positron pairs, . The resulting decay rates are rapid and effective at energies where the process is allowed. Once above the energy threshold, this photon decay process would lead to a hard cutoff in the gamma-ray spectrum without any high-energy photons being observed 2017PhRvD..95f3001M . The threshold for any order is given by
(18) |
where represents the electron mass 2017PhRvD..95f3001M . One can see from Eqs. (17) and (18) that the upper limits on (lower limits on ) become tighter with the increase in the observed photon energy by a factor of ( for lower limits on ).
4.2 Photon splitting
A second superluminal LIV decay process is photon splitting to multiple photons, . The dominant splitting process with quadratic modification () is the photon decay into three photons () 2019JCAP…04..054A . The decay rate of photon splitting (in units of energy) is 2019JCAP…04..054A ; 2019EPJC…79.1011S
(21) |
which is much smaller than the photon decay rate . Despite the lack of an energy threshold, this process is kinematically allowed whenever (i.e., superluminal LIV). It becomes significant when photons travel through astrophysical distances and also predicts a hard cutoff at high photon energies in astrophysical spectra.
If we find the evidence of a cutoff in the spectrum of an energetic astrophysical source at distance , we can equate the mean free path of a photon to the source distance. That is, we take , with translated to units of . The corresponding LIV limit is therefore given by
(22) |
which is a function of the highest photon energy. Once again, this photon splitting in flight from the source provides a direct way to constrain the quadratic LIV energy scale that mainly relies on the highest energy photons observed.
4.3 Present constraints from spectral cutoff
If there is no sign of clear cutoff at the highest energy end of the photon spectra of astrophysical sources, then stringent lower limits on the LIV energy scale can be derived. There have been some resulting constraints on LIV-induced spectral cutoff using sub-PeV measurements of gamma-ray spectra from astrophysical sources, including the HEGRA observations of the Crab Nebula spectrum 2017PhRvD..95f3001M ; 2019JCAP…04..054A , the Tibet observations of the Crab Nebula spectrum 2019EPJC…79.1011S , the HAWC observations of the spectra of the Crab and three other sources 2020PhRvL.124m1101A , and the LHAASO observations of the spectra of the Crab and four other sources 2021arXiv210507927C ; 2021arXiv210612350T .
For instance, the LHAASO Collaboration studied the superluminal LIV effect using the observations of unprecedentedly high-energy gamma rays, with a rigorous statistical technique and a careful assessment of the systematic uncertainties 2021arXiv210612350T . Recently, 12 ultra-high-energy gamma-ray Galactic sources above 100 TeV have been detected by LHAASO 2021Natur.594…33C . The highest energy photon-like event from the source LHAASO J2032+4102 is about 1.4 PeV. The second highest energy photon-like event from LHAASO J0534+2202 (Crab Nebula) is about 0.88 PeV. Since higher energy photons may provide better limits on the LIV energy scale, the LHAASO Collaboration adopted the two sources J2032+4102 and J0534+2202 for their purpose 2021arXiv210612350T . Two general spectral forms are assumed for both sources, i.e., a log-parabolic form and a power-law form. Both the photon decay () and photon splitting () processes predict that the energy spectrum of a source has a quasi-sharp cutoff at the highest energy part. If LIV does exist, the expected spectrum of the source can then be expressed as 2021arXiv210612350T
(23) |
where is the flux normalization, and are the spectral indices, TeV is a reference energy, stands for the Heaviside step function, and is the cutoff energy. The above expression is for the log-parabolic spectrum. When , the log-parabolic spectrum reduces to the power-law spectrum. The spectral parameters can be optimized by maximizing the likelihood function, i.e.,
(24) |
where represents the -th energy bin, is the number of observed events from the source, is the expected signal from the chosen energy spectrum, and is the estimated background. For each , Ref. 2021arXiv210612350T obtained the best-fit spectral parameters (, , and ) and the corresponding likelihood value. The statistical significance of the existence of such a cutoff was calculated using a test statistic (TS) variable, which is the log-likelihood ratio of the fit with a cutoff and the fit with no cutoff,
(25) |
where the null hypothesis (without the LIV effect) is the case with and the alternative hypothesis (with the LIV effect) corresponds to a finite . Because the spectral fits do not indicate a significant preference for , a lower limit on was proceed to set, below which there is weak or no evidence for the photon decay. This lower limit on also serves as an upper limit on the observed photon energy . Then Eqs. (19), (20), and (22) directly lead to lower limits on and .
Experiment | Source | Distance | Refs. | ||||
[kpc] | [TeV] | [GeV] | [GeV] | [GeV] | |||
HEGRA | Crab Nebula | 2 | 56 | — | 2017PhRvD..95f3001M | ||
Crab Nebula | 2 | 75 | — | — | 2019JCAP…04..054A | ||
Tibet | Crab Nebula | 2 | 140 | — | — | 2019EPJC…79.1011S | |
HAWC | J1825–134 | 1.55 | 244 | 2020PhRvL.124m1101A | |||
J1907+063 | 2.37 | 218 | |||||
Crab Nebula | 2 | 152 | |||||
J2019+368 | 1.8 | 120 | |||||
LHAASO | J2226+6057 | 280.7 | 2021arXiv210507927C | ||||
J1908+0621 | 2.37 | 370.5 | 2021arXiv210507927C | ||||
J1825–1326 | 1.55 | 169.9 | 2021arXiv210507927C | ||||
Crab Nebula | 2 | 750 | 2021arXiv210612350T | ||||
J2032+4102 | 1.4 | 1140 | 2021arXiv210612350T |
aThe distance of possible celestial objects associated with J2226+6057 was given.
Table 3 lists a summary of 95% confidence-level lower limits on and the inferred LIV energy scales obtained through photon decay and photon splitting. As illustrated in Table 3, the highest-energy source LHAASO J2032+4102 provides the strongest constraints on the energy scales of the superluminal LIV among experimental results with the same method. The linear LIV energy scale is constrained to be higher than about times the Planck energy scale , and the quadratic LIV energy scales is over .
4.4 Comparison with different methods
The comparison with the resulting constraints on the superluminal LIV energy scales obtained from different experiments is shown in Figure 4. We show strong limits on photon decay and photon splitting using ultra-high-energy gamma-ray observations from HEGRA 2017PhRvD..95f3001M ; 2019JCAP…04..054A , Tibet 2019EPJC…79.1011S , HAWC 2020PhRvL.124m1101A , and LHAASO 2021arXiv210612350T . We also show limits due to LIV energy-dependent time delay searches with the Fermi/LAT 2013PhRvD..87l2001V . Obviously, the limits on and from photon decay and photon splitting are at least four orders of magnitude stronger than those from the time delay method.
Additionally, observations of linear polarization from astrophysical sources have been widely used to place limits on the vacuum birefringent parameter . By comparing Eqs. (1) and (11), we can derive the conversion from to the limit on the linear LIV energy scale, i.e., . At this point, it is interesting to make a comparison of recent achievements in sensitivity of polarization measurements versus photon decay measurements. The detection of linear polarization in the prompt gamma-ray emission of GRB 061122 yielded the strictest limit on the birefringent parameter, , which corresponds to 2016MNRAS.463..375L ; 2019MNRAS.485.2401W . Obviously, this polarization constraint is about 11 orders of magnitude stronger than the constraint from photon decay () 2021arXiv210612350T . However, both time-of-flight constraints and photon decay constraints are essential in an extensive search for nonbirefringent Lorentz-violating effects.

5 Summary and outlook
Many tight constraints on violations of Lorentz invariance have been achieved by observing tiny changes in light that has propagated over astrophysical distances. Some of these search for a frequency-dependent velocity resulting from vacuum dispersion. Some works seek a change in polarization arising from vacuum birefringence. Others look for a hard cutoff in the gamma-ray spectra due to photon decay and photon splitting.
For vacuum dispersion studies, in order to tightly constrain the LIV effects one should choose those explosive astrophysical sources with shorter spectral lags, higher energy emissions, and longer propagation distances. As the most energetic bursting events occurring at cosmological sources, GRBs are excellent astrophysical probes for LIV constraints in the dispersive photon sector. In the past, emission from GRBs had been detected only at energies below 100 GeV, and LIV limits from time-of-flight measurements of GRBs had also been restricted in the relatively low energy range. Now that photons of energies above 100 GeV have been detected from two bright GRBs (GRB 190114C 2019Natur.575..455M and GRB 180720B 2019Natur.575..464A ), opening a new spectral window in GRB research. It is expected that such detections will become routine in the future 2019Natur.575..448Z , especially with the operations of facilities such as HAWC, LHAASO, and the future international Cherenkov Telescope Array. Vacuum dispersion constraints will greatly benefit from the observations of extremely high-energy emission of more GRBs. For vacuum birefringence studies, in order to obtain tighter constraints on the LIV effects one should choose those astrophysical sources with longer distances and polarization measurements at higher energies. In this respect, the ideal sources to seek LIV-induced vacuum birefringence would be polarized gamma-rays from bright GRBs at deep cosmological distances. The technology for measuring polarization in the 5 to 100 MeV energy range has been constantly improved and developed. As more and more gamma-ray polarimeters (such as POLAR-II, TSUBAME, NCT/COSI, GRAPE; see 2017NewAR..76….1M for a review) enter service, more polarized astrophysical sources at higher gamma-ray energies are expected to be detected. Polarization measurements of cosmological sources such as GRBs at such energies will further improve LIV constraints through the vacuum birefringent effect. For photon decay studies, in order to obtain more stringent bounds on the LIV effects one should choose those astrophysical sources with abundant ultra-high-energy photons. High energy resolution of spectrum measurements are desired to determine whether there is a hard cutoff consistent with the prediction of photon decay in the observed spectra of astrophysical sources. When using ultra-high-energy photons for LIV limits, we should be careful with the uncertainty for failure to reject a cosmic-ray event. Owing to its unprecedented capability of cosmic ray background rejection and high energy resolution 2021Natur.594…33C , LHAASO has competitive advantage in constraining LIV-induced photon decay.
References
- (1) V.A. Kostelecký, S. Samuel, Phys. Rev. D39, 683 (1989). DOI 10.1103/PhysRevD.39.683
- (2) V.A. Kostelecký, R. Potting, Nuclear Physics B 359, 545 (1991)
- (3) V.A. Kostelecký, R. Potting, Phys. Rev. D51, 3923 (1995). DOI 10.1103/PhysRevD.51.3923
- (4) D. Mattingly, Living Reviews in Relativity 8, 5 (2005). DOI 10.12942/lrr-2005-5
- (5) R. Bluhm, Lecture Notes in Physics 702, 191 (2006). DOI 10.1007/3-540-34523-X“˙8
- (6) G. Amelino-Camelia, Living Reviews in Relativity 16, 5 (2013). DOI 10.12942/lrr-2013-5
- (7) J.D. Tasson, Reports on Progress in Physics 77(6), 062901 (2014). DOI 10.1088/0034-4885/77/6/062901
- (8) J.J. Wei, X.F. Wu, Frontiers of Physics 16(4), 44300 (2021). DOI 10.1007/s11467-021-1049-x
- (9) V.A. Kostelecký, N. Russell, Reviews of Modern Physics 83, 11 (2011). DOI 10.1103/RevModPhys.83.11
- (10) V.A. Kostelecký, M. Mewes, Astrophys. J. Lett.689, L1 (2008). DOI 10.1086/595815
- (11) F.W. Stecker, S.L. Glashow, Astroparticle Physics 16(1), 97 (2001). DOI 10.1016/S0927-6505(01)00137-2
- (12) G. Amelino-Camelia, J. Ellis, N.E. Mavromatos, D.V. Nanopoulos, S. Sarkar, Nature393, 763 (1998). DOI 10.1038/31647
- (13) J. Ellis, N.E. Mavromatos, D.V. Nanopoulos, A.S. Sakharov, E.K.G. Sarkisyan, Astroparticle Physics 25, 402 (2006). DOI 10.1016/j.astropartphys.2006.04.001
- (14) U. Jacob, T. Piran, JCAP1, 031 (2008). DOI 10.1088/1475-7516/2008/01/031
- (15) V.A. Kostelecký, M. Mewes, Phys. Rev. D80(1), 015020 (2009). DOI 10.1103/PhysRevD.80.015020
- (16) A.A. Abdo, et al., Science 323(5922), 1688 (2009). DOI 10.1126/science.1169101
- (17) A.A. Abdo, et al., Nature462(7271), 331 (2009). DOI 10.1038/nature08574
- (18) Z. Chang, Y. Jiang, H.N. Lin, Astroparticle Physics 36(1), 47 (2012). DOI 10.1016/j.astropartphys.2012.04.006
- (19) R.J. Nemiroff, R. Connolly, J. Holmes, A.B. Kostinski, Phys. Rev. Lett.108(23), 231103 (2012). DOI 10.1103/PhysRevLett.108.231103
- (20) V. Vasileiou, A. Jacholkowska, F. Piron, J. Bolmont, C. Couturier, J. Granot, F.W. Stecker, J. Cohen-Tanugi, F. Longo, Phys. Rev. D87(12), 122001 (2013). DOI 10.1103/PhysRevD.87.122001
- (21) J. Ellis, N.E. Mavromatos, Astroparticle Physics 43, 50 (2013). DOI 10.1016/j.astropartphys.2012.05.004
- (22) F. Kislat, H. Krawczynski, Phys. Rev. D92(4), 045016 (2015). DOI 10.1103/PhysRevD.92.045016
- (23) S. Zhang, B.Q. Ma, Astroparticle Physics 61, 108 (2015). DOI 10.1016/j.astropartphys.2014.04.008
- (24) J.J. Wei, B.B. Zhang, L. Shao, X.F. Wu, P. Mészáros, Astrophys. J. Lett.834(2), L13 (2017). DOI 10.3847/2041-8213/834/2/L13
- (25) J.J. Wei, X.F. Wu, B.B. Zhang, L. Shao, P. Mészáros, V.A. Kostelecký, Astrophys. J.842(2), 115 (2017). DOI 10.3847/1538-4357/aa7630
- (26) J.J. Wei, X.F. Wu, Astrophys. J.851(2), 127 (2017). DOI 10.3847/1538-4357/aa9d8d
- (27) J. Ellis, R. Konoplich, N.E. Mavromatos, L. Nguyen, A.S. Sakharov, E.K. Sarkisyan-Grinbaum, Phys. Rev. D99(8), 083009 (2019). DOI 10.1103/PhysRevD.99.083009
- (28) V.A. Acciari, et al., Phys. Rev. Lett. 125, 021301 (2020). DOI 10.1103/PhysRevLett.125.021301.
- (29) S.D. Biller, A.C. Breslin, J. Buckley, M. Catanese, M. Carson, D.A. Carter-Lewis, M.F. Cawley, D.J. Fegan, J.P. Finley, J.A. Gaidos, A.M. Hillas, F. Krennrich, R.C. Lamb, R. Lessard, C. Masterson, J.E. McEnery, B. McKernan, P. Moriarty, J. Quinn, H.J. Rose, F. Samuelson, G. Sembroski, P. Skelton, T.C. Weekes, Phys. Rev. Lett.83(11), 2108 (1999). DOI 10.1103/PhysRevLett.83.2108
- (30) P. Kaaret, Astron. Astroph.345, L32 (1999)
- (31) S.M. Carroll, G.B. Field, R. Jackiw, Phys. Rev. D41(4), 1231 (1990). DOI 10.1103/PhysRevD.41.1231
- (32) D. Colladay, V.A. Kostelecký, Phys. Rev. D58(11), 116002 (1998). DOI 10.1103/PhysRevD.58.116002
- (33) R.J. Gleiser, C.N. Kozameh, Phys. Rev. D64(8), 083007 (2001). DOI 10.1103/PhysRevD.64.083007
- (34) V.A. Kostelecký, M. Mewes, Physical Review Letters 87(25), 251304 (2001). DOI 10.1103/PhysRevLett.87.251304
- (35) V.A. Kostelecký, M. Mewes, Physical Review Letters 97(14), 140401 (2006). DOI 10.1103/PhysRevLett.97.140401
- (36) V.A. Kostelecký, M. Mewes, Physical Review Letters 99(1), 011601 (2007). DOI 10.1103/PhysRevLett.99.011601
- (37) V.A. Kostelecký, M. Mewes, Physical Review Letters 110(20), 201601 (2013). DOI 10.1103/PhysRevLett.110.201601
- (38) I.G. Mitrofanov, Nature426, 139 (2003). DOI 10.1038/426139a
- (39) T. Jacobson, S. Liberati, D. Mattingly, F.W. Stecker, Physical Review Letters 93(2), 021101 (2004). DOI 10.1103/PhysRevLett.93.021101
- (40) Y.Z. Fan, D.M. Wei, D. Xu, Mon. Not. R. Astron. Soc.376, 1857 (2007). DOI 10.1111/j.1365-2966.2007.11576.x
- (41) G. Gubitosi, L. Pagano, G. Amelino-Camelia, A. Melchiorri, A. Cooray, JCAP8, 021 (2009). DOI 10.1088/1475-7516/2009/08/021
- (42) P. Laurent, D. Götz, P. Binétruy, S. Covino, A. Fernandez-Soto, Phys. Rev. D83(12), 121301 (2011). DOI 10.1103/PhysRevD.83.121301
- (43) F.W. Stecker, Astroparticle Physics 35, 95 (2011). DOI 10.1016/j.astropartphys.2011.06.007
- (44) K. Toma, S. Mukohyama, D. Yonetoku, T. Murakami, S. Gunji, T. Mihara, Y. Morihara, T. Sakashita, T. Takahashi, Y. Wakashima, H. Yonemochi, N. Toukairin, Physical Review Letters 109(24), 241104 (2012). DOI 10.1103/PhysRevLett.109.241104
- (45) D. Götz, S. Covino, A. Fernández-Soto, P. Laurent, Ž. Bošnjak, Mon. Not. R. Astron. Soc.431, 3550 (2013). DOI 10.1093/mnras/stt439
- (46) D. Götz, P. Laurent, S. Antier, S. Covino, P. D’Avanzo, V. D’Elia, A. Melandri, Mon. Not. R. Astron. Soc.444, 2776 (2014). DOI 10.1093/mnras/stu1634
- (47) H.N. Lin, X. Li, Z. Chang, Mon. Not. R. Astron. Soc.463, 375 (2016). DOI 10.1093/mnras/stw2007
- (48) F. Kislat, H. Krawczynski, Phys. Rev. D95(8), 083013 (2017). DOI 10.1103/PhysRevD.95.083013
- (49) A.S. Friedman, D. Leon, K.D. Crowley, D. Johnson, G. Teply, D. Tytler, B.G. Keating, G.M. Cole, Phys. Rev. D99(3), 035045 (2019). DOI 10.1103/PhysRevD.99.035045
- (50) J.J. Wei, Mon. Not. R. Astron. Soc.485(2), 2401 (2019). DOI 10.1093/mnras/stz594
- (51) H. Martínez-Huerta, A. Pérez-Lorenzana, Phys. Rev. D95(6), 063001 (2017). DOI 10.1103/PhysRevD.95.063001
- (52) K. Astapov, D. Kirpichnikov, P. Satunin, JCAP2019(4), 054 (2019). DOI 10.1088/1475-7516/2019/04/054
- (53) P. Satunin, European Physical Journal C 79(12), 1011 (2019). DOI 10.1140/epjc/s10052-019-7520-y
- (54) A. Albert, et al., Phys. Rev. Lett.124(13), 131101 (2020). DOI 10.1103/PhysRevLett.124.131101
- (55) L. Chen, Z. Xiong, C. Li, S. Chen, H. He, arXiv e-prints arXiv:2105.07927 (2021)
- (56) The LHAASO Collaboration, arXiv e-prints arXiv:2106.12350 (2021)
- (57) J. Ellis, N.E. Mavromatos, D.V. Nanopoulos, A.S. Sakharov, Astron. Astroph.402, 409 (2003). DOI 10.1051/0004-6361:20030263
- (58) S.E. Boggs, C.B. Wunderer, K. Hurley, W. Coburn, Astrophys. J. Lett.611(2), L77 (2004). DOI 10.1086/423933
- (59) J. Bolmont, A. Jacholkowska, J.L. Atteia, F. Piron, G. Pizzichini, Astrophys. J.676(1), 532 (2008). DOI 10.1086/527524
- (60) R. Lamon, N. Produit, F. Steiner, General Relativity and Gravitation 40(8), 1731 (2008). DOI 10.1007/s10714-007-0580-6
- (61) M. Rodríguez Martínez, T. Piran, Y. Oren, JCAP2006(5), 017 (2006). DOI 10.1088/1475-7516/2006/05/017
- (62) J. Ellis, N.E. Mavromatos, D.V. Nanopoulos, A.S. Sakharov, E.K.G. Sarkisyan, Astroparticle Physics 29(2), 158 (2008). DOI 10.1016/j.astropartphys.2007.12.003
- (63) Z. Xiao, B.Q. Ma, Phys. Rev. D80(11), 116005 (2009). DOI 10.1103/PhysRevD.80.116005
- (64) L. Shao, Z. Xiao, B.Q. Ma, Astroparticle Physics 33(5-6), 312 (2010). DOI 10.1016/j.astropartphys.2010.03.003
- (65) MAGIC Collaboration, et al., Nature575(7783), 455 (2019). DOI 10.1038/s41586-019-1750-x
- (66) M. Biesiada, A. Piórkowska, Classical and Quantum Gravity 26(12), 125007 (2009). DOI 10.1088/0264-9381/26/12/125007
- (67) Y. Pan, Y. Gong, S. Cao, H. Gao, Z.H. Zhu, Astrophys. J.808(1), 78 (2015). DOI 10.1088/0004-637X/808/1/78
- (68) X.B. Zou, H.K. Deng, Z.Y. Yin, H. Wei, Physics Letters B 776, 284 (2018). DOI 10.1016/j.physletb.2017.11.053
- (69) Y. Pan, J. Qi, S. Cao, T. Liu, Y. Liu, S. Geng, Y. Lian, Z.H. Zhu, Astrophys. J.890(2), 169 (2020). DOI 10.3847/1538-4357/ab6ef5
- (70) T.N. Ukwatta, K.S. Dhuga, M. Stamatikos, C.D. Dermer, T. Sakamoto, E. Sonbas, W.C. Parke, L.C. Maximon, J.T. Linnemann, P.N. Bhat, A. Eskand arian, N. Gehrels, A.U. Abeysekara, K. Tollefson, J.P. Norris, Mon. Not. R. Astron. Soc.419(1), 614 (2012). DOI 10.1111/j.1365-2966.2011.19723.x
- (71) M.G. Bernardini, G. Ghirlanda, S. Campana, S. Covino, R. Salvaterra, J.L. Atteia, D. Burlon, G. Calderone, P. D’Avanzo, V. D’Elia, G. Ghisellini, V. Heussaff, D. Lazzati, A. Meland ri, L. Nava, S.D. Vergani, G. Tagliaferri, Mon. Not. R. Astron. Soc.446(2), 1129 (2015). DOI 10.1093/mnras/stu2153
- (72) Z. Chang, X. Li, H.N. Lin, Y. Sang, P. Wang, S. Wang, Chinese Physics C 40(4), 045102 (2016). DOI 10.1088/1674-1137/40/4/045102
- (73) L. Shao, B.B. Zhang, F.R. Wang, X.F. Wu, Y.H. Cheng, X. Zhang, B.Y. Yu, B.J. Xi, X. Wang, H.X. Feng, M. Zhang, D. Xu, Astrophys. J.844(2), 126 (2017). DOI 10.3847/1538-4357/aa7d01
- (74) R.J. Lu, Y.F. Liang, D.B. Lin, J. Lü, X.G. Wang, H.J. Lü, H.B. Liu, E.W. Liang, B. Zhang, Astrophys. J.865(2), 153 (2018). DOI 10.3847/1538-4357/aada16
- (75) S.S. Du, L. Lan, J.J. Wei, Z.M. Zhou, H. Gao, L.Y. Jiang, B.B. Zhang, Z.K. Liu, X.F. Wu, E.W. Liang, Z.H. Zhu, Astrophys. J.906(1), 8 (2021). DOI 10.3847/1538-4357/abc624
- (76) F. Aharonian, et al., Physical Review Letters 101(17), 170402 (2008). DOI 10.1103/PhysRevLett.101.170402
- (77) H. E. S. S. Collaboration, et al., Astroparticle Physics 34(9), 738 (2011). DOI 10.1016/j.astropartphys.2011.01.007
- (78) A. Abramowski, et al., Astrophys. J.802(1), 65 (2015). DOI 10.1088/0004-637X/802/1/65
- (79) MAGIC Collaboration, et al., Physics Letters B 668, 253 (2008). DOI 10.1016/j.physletb.2008.08.053
- (80) M. Martínez, M. Errando, Astroparticle Physics 31(3), 226 (2009). DOI 10.1016/j.astropartphys.2009.01.005
- (81) H. Abdalla, et al., Astrophys. J.870(2), 93 (2019). DOI 10.3847/1538-4357/aaf1c4
- (82) N. OTTE, in International Cosmic Ray Conference, International Cosmic Ray Conference, vol. 7 (2011), International Cosmic Ray Conference, vol. 7, p. 256. DOI 10.7529/ICRC2011/V07/1302
- (83) B. Zitzer, VERITAS Collaboration, in International Cosmic Ray Conference, International Cosmic Ray Conference, vol. 33 (2013), International Cosmic Ray Conference, vol. 33, p. 2768
- (84) MAGIC Collaboration, et al., Astrophys. J. Suppl.232(1), 9 (2017). DOI 10.3847/1538-4365/aa8404
- (85) R.C. Myers, M. Pospelov, Physical Review Letters 90(21), 211601 (2003). DOI 10.1103/PhysRevLett.90.211601
- (86) L. Maccione, S. Liberati, A. Celotti, J.G. Kirk, P. Ubertini, Phys. Rev. D78(10), 103003 (2008). DOI 10.1103/PhysRevD.78.103003
- (87) W. Coburn, S.E. Boggs, Nature423, 415 (2003). DOI 10.1038/nature01612
- (88) R.E. Rutledge, D.B. Fox, Mon. Not. R. Astron. Soc.350, 1288 (2004). DOI 10.1111/j.1365-2966.2004.07665.x
- (89) C. Wigger, W. Hajdas, K. Arzner, M. Güdel, A. Zehnder, Astrophys. J.613, 1088 (2004). DOI 10.1086/423163
- (90) E. Kalemci, S.E. Boggs, C. Kouveliotou, M. Finger, M.G. Baring, Astrophys. J. Suppl.169, 75 (2007). DOI 10.1086/510676
- (91) S. McGlynn, D.J. Clark, A.J. Dean, L. Hanlon, S. McBreen, D.R. Willis, B. McBreen, A.J. Bird, S. Foley, Astron. Astroph.466, 895 (2007). DOI 10.1051/0004-6361:20066179
- (92) D. Götz, P. Laurent, F. Lebrun, F. Daigne, Ž. Bošnjak, Astrophys. J. Lett.695, L208 (2009). DOI 10.1088/0004-637X/695/2/L208
- (93) D. Yonetoku, T. Murakami, S. Gunji, T. Mihara, K. Toma, T. Sakashita, Y. Morihara, T. Takahashi, N. Toukairin, H. Fujimoto, Y. Kodama, S. Kubo, IKAROS Demonstration Team, Astrophys. J. Lett.743, L30 (2011). DOI 10.1088/2041-8205/743/2/L30
- (94) D. Yonetoku, T. Murakami, S. Gunji, T. Mihara, K. Toma, Y. Morihara, T. Takahashi, Y. Wakashima, H. Yonemochi, T. Sakashita, N. Toukairin, H. Fujimoto, Y. Kodama, Astrophys. J. Lett.758, L1 (2012). DOI 10.1088/2041-8205/758/1/L1
- (95) J.J. Wei, X.F. Wu, European Physical Journal Plus 135(6), 527 (2020). DOI 10.1140/epjp/s13360-020-00554-x
- (96) Q.Q. Zhou, S.X. Yi, J.J. Wei, X.F. Wu, Galaxies 9(2), 44 (2021). DOI 10.3390/galaxies9020044
- (97) Z. Cao, et al., Nature594(7861), 33 (2021). DOI 10.1038/s41586-021-03498-z
- (98) H. Abdalla, et al., Nature575(7783), 464 (2019). DOI 10.1038/s41586-019-1743-9
- (99) B. Zhang, Nature575(7783), 448 (2019). DOI 10.1038/d41586-019-03503-6
- (100) M.L. McConnell, New Astron. Rev.76, 1 (2017). DOI 10.1016/j.newar.2016.11.001