Next Article in Journal
The Classification of Blazar Candidates of Uncertain Types
Next Article in Special Issue
Simulation of the Isotropic Ultra-High Energy Photon Flux in the Solar Magnetic Field
Previous Article in Journal
Properties and Dynamics of Neutron Stars and Proto-Neutron Stars
Previous Article in Special Issue
Analysis of the Capability of Detection of Extensive Air Showers by Simple Scintillator Detectors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Ultra-High-Energy Astroparticles as Probes for Lorentz Invariance Violation

by
Rodrigo Guedes Lang
1,*,
Humberto Martínez-Huerta
2 and
Vitor de Souza
3
1
Erlangen Centre for Astroparticle Physics, Friedrich-Alexander-Universität Erlangen-Nürnberg, Erwin-Rommel-Str. 1, D-91058 Erlangen, Germany
2
Departmento de Física y Matemáticas, Universidad de Monterrey, Avenida Morones Prieto 4500, San Pedro Garza García 66238, NL, Mexico
3
Instituto de Física de São Carlos, Universidade de São Paulo, São Carlos 13566-590, SP, Brazil
*
Author to whom correspondence should be addressed.
Universe 2022, 8(8), 435; https://doi.org/10.3390/universe8080435
Submission received: 9 June 2022 / Revised: 7 August 2022 / Accepted: 18 August 2022 / Published: 22 August 2022
(This article belongs to the Special Issue Ultra High Energy Photons)

Abstract

:
Compelling evidence for Lorentz invariance violation (LIV) would demand a complete revision of modern physics. Therefore, searching for a signal or extending the validity of the invariance is fundamental for building our understanding of the extreme phenomena in the Universe. In this paper, we review the potential of ultra-high-energy astroparticles in setting limits on LIV. The standard framework of LIV studies in astroparticle physics is reviewed and its use on the electromagnetic and hadronic sectors are discussed. In particular, the current status of LIV tests using experimental data on ultra-high-energy photons and cosmic rays is addressed. A detailed discussion with improved argumentation about the LIV kinematics of the relevant interactions is shown. The main previous results are presented together with new calculations based on recently published astrophysical models.

1. Introduction

Lorentz invariance is one of the fundamental symmetries in the standard model (SM) of elementary particles. There is renewed interest in the community in testing the limits of its validity given that grand unified theories, such as quantum gravity and string theory, speculate about a certain degree of Lorentz invariance violation (LIV) [1,2,3,4,5,6,7,8,9,10].
Most such LIV effects are expected to be small. However, the incredibly high energies of ultra-high-energy (UHE, E > 10 18  eV) astroparticles, the extreme distances they travel to Earth, and the current level of experimental precision and of the high exposure of observatories in operation makes them promising probes for LIV.
Cosmic and gamma-ray observations have been used to set strong constraints to LIV [11]. Examples of the most explored channels are: (a) changes in the kinematics of processes such as photo-pair production [12,13,14,15,16,17,18,19], (b) processes forbidden in the SM, such as photon splitting and decay [20,21], and Cherenkov radiation in vacuum [22,23,24], and (c) energy-dependent time delay in photons [25,26].
In this review, we summarize these efforts, discussing the current state-of-the-art analyses and their limitations. First, in Section 2, we present the most common phenomenological framework used in astrophysical tests of LIV. Next, in Section 3 and Section 4, we discuss the most up-to-date results and limits obtained with UHE photons and cosmic rays, respectively. The dependence of these analyses on astrophysical models is also explored. Finally, in Section 5, we address our final remarks and prospects for future LIV studies using UHE astroparticles.

2. LIV Framework

The astrophysical corrections of LIV can be addressed via a family of modified dispersion relations (MDR) for each type of particle (see, for instance, [1] and Refs. therein),
E a 2 = m a 2 + p a 2 1 + n = 0 , 1 , 2 , δ a , n E a n ,
where E a and p a are the energy and momenta, and δ is the LIV parameter for the a-particle.
In the special case of photons ( m γ = 0 ), the MDR is,
E γ 2 = p γ 2 1 + n = 0 , 1 , 2 , δ γ , n E a n ,
where E γ and p γ is the energy and momenta and δ n , γ stands for the photon LIV parameter. The positive sign corresponds to the so-called superluminal phenomena and the minus to the subluminal. Using one leading order n at a time, when  n = 0 , the LIV correction corresponds the so-called maximal velocity of photons. If n = 1 or n = 2 , the LIV scale of the correction can be expressed by E LIV ( n ) = ( | 1 / δ a , n n | ) .
Astrophysical tests had settled strong constraints on the different LIV parameters [1]. For instance, for the subluminal side, the LIV modified pair production interaction on EBL ( γ γ E B L e + e ) bounded δ γ , 1 8.3 × 10 30 eV 1 E LIV ( 1 ) 1.2 × 10 29 eV and δ γ , 2 1.74 × 10 43 eV 2 E LIV ( 2 ) 2.4 × 10 21 eV , using TeV gamma rays [13], and ref. [27] constrained δ γ , 2 > 3.46 × 10 45 eV 2 E LIV ( 2 ) 1.7 × 10 22 eV through LIV air shower suppression. On the other hand, astrophysical superluminal tests usually derive stronger limits, see, for instance, [28], which by the absence of LIV photon decay ( γ e + e ) effects on gamma rays reports δ γ , 0 7.1 × 10 19 , as well as δ γ , 1 3.46 × 10 67 eV 1 E LIV ( 1 ) 1.7 × 10 33 eV , and δ γ , 2 1.6 × 10 51 eV 2 E LIV ( 2 ) 2.5 × 10 25 eV by the lack of photon splitting ( γ 3 γ ). Furthermore, for a collection of Lorentz violating limits in the context of the standard model extension [5] including other SM sectors and dimensions, see the tables in [29].

3. Testing LIV with UHE Photons

No UHE photon has been detected so far [30,31,32]. The identification of a UHE photon source would lead to very precise and strong limits on LIV. UHE cosmic rays (UHECR) interact with the photon background producing pions. The neutral ones rapidly decay, producing UHE photons which could, in principle, be detected on Earth [33,34] as a secondary flux of UHE photons.
As first proposed in [17,35] and further developed by in [14], if subluminal LIV ( δ γ , n < 0 ) is considered, the interaction rate of UHECRs with background photons decreases, leading to an increase in the expected number of UHE photons arriving on Earth.

3.1. Propagation of UHE Photons Considering LIV

Photons may be absorbed on their way to Earth, producing an electron–positron pair when interacting with the photon background ( γ γ bg e e + ) [36]. The mean free path of the interaction is determined by the energy distribution of the background photons and the kinematics of the pair production. In an LI scenario, the mean free path is such that most extragalactic UHE photons are expected to be absorbed. If LIV in the electromagnetic sector is considered, on the other hand, the kinematics of the interaction is expected to be modified [11]. The energy threshold for the pair production considering LIV effects is given by [12,14,15,16,17,20]
ϵ th E γ = m e 2 E γ δ n , γ E γ n + 1 4 ,
where E γ and ϵ are, respectively, the energies of the UHE and background photons, m e is the electron mass, and δ n , γ is the LIV coefficient. The LIV energy threshold ( ϵ th ) can be larger or smaller than the Lorentz-invariant (LI) case, depending on the sign of δ n , γ in the last term of the equation.
Figure 1 shows the threshold energy for the background photon as a function of the energy of the UHE photon for different subluminal LIV coefficients of order n = 0 . Similar results are found for higher orders. In a LI scenario, the dominant fields are the extragalactic background light (EBL) for E 10 14  eV, the cosmic microwave background (CMB) for 10 14 E / eV 10 17 , and the radio background for E 10 17  eV. When subluminal (superluminal) LIV is considered, the threshold increases (decreases) with respect to the LI scenario and, thus, fewer (more) interactions are expected. The energy above which this effect becomes significant decreases with increasing absolute values of the LIV coefficient. The survival probability, P surv , of a photon emitted at redshift z, with energy E γ is obtained by the following expression,
P surv = exp 0 z d z c H 0 ( 1 + z ) h ( z ) 1 1 d cos θ 1 cos θ 2 ϵ th d ϵ n ( ϵ , z ) σ ( E γ , ϵ , z ) ,
where H 0 and h ( z ) are the cosmological terms, σ ( E γ , ϵ , z ) is the Breit–Wheeler cross-section [37], and n ( ϵ , z ) is the distribution of background photons.
Figure 2 shows the survival probability as a function of the energy for UHE photons emitted at z = 0.001 , z = 0.01 and z = 0.1 under different subluminal LIV assumptions. For large LIV coefficients and high energies, the survival probability is increased up to the point where no photons are absorbed. The limiting assumptions, LI ( δ = 0 ) and the maximum LIV ( δ ), in which no interaction happens at all energies, are shown by the black and red lines, respectively.

3.2. Searching for LIV in UHE Photon Data

When superluminal LIV in considered in photon propagation, the survival probability of the UHE photon is reduced. Since no UHE photon event has yet been detected, it is not straightforward to use current UHE data to search for a decrease in the expected flux.
When subluminal LIV is considered, the survival probability of the UHE photon increases allowing the calculation of upper limits on the LIV coefficients based on the measured upper limits of UHE photons. Several works have used this technique [14,17,35,40]. Since no UHE photon source was detected, only secondary photons produced in the propagation of UHECR are considered in the calculation. As such, the flux of UHE photons arriving on Earth is highly dependent on the UHECR astrophysical model: spacial distribution of the sources, emission energy spectra and chemical composition [41]. Thereby, as first discussed in [14], LIV limits imposed by this technique are highly dependent on the astrophysical assumptions about the UHECR sources.
In phenomenological studies [42], homogeneously distributed sources of UHECRs are usually considered to emit cosmic rays isotropically with energy spectra described by
J UHECR A ( E ) = f A E Γ e E Z R max ,
where A and Z denote, respectively, the atomic mass and charge for each species of emitted nuclei. The fraction of a given species, f A , the spectral index, Γ , and the maximum rigidity attainable by the sources, R max , are free parameters and vary between different astrophysical models. Chemical composition is usually divided into five major groups almost equally separated in ln A , with five representative nuclei, 1 H , 4 He , 14 N , 28 Si and 56 Fe . Such a simplistic description is justified by the lack of knowledge about the sources of UHECR.
Following the procedure proposed in [14], we calculated LIV limits for several astrophysical models covering the phase space of Equation (5) to illustrate their influence on the LIV limits. UHECR propagation is simulated with the state-of-the-art Monte Carlo package CRPropa 3 [43]. The propagation of the secondary UHE photons is then performed using the package EleCa [44], with an implementation of the LIV corrections described in Section 3.1. The simulated energy spectrum of UHECR is normalized to the energy spectrum measured by the Pierre Auger Observatory [45] at E = 10 18.75  eV. Finally, the LIV limit is obtained as the smallest absolute value of the LIV coefficient for which the predicted flux is more intense than the upper limits imposed by the Pierre Auger Observatory [30,31,32].
Figure 3 shows the logarithmic order of the LIV limit imposed for each combination of maximum attainable rigidity at the sources and spectral index for the two extreme compositions at the sources, pure proton and pure iron. The results cover all the possibilities, ranging from not being sensitive to LIV up to the case where the model would be inconsistent with data even under a LI assumption. The largest dependence comes from R max . For small maximum rigidities, no cosmic ray will be accelerated up to enough energies to produce UHE photons and, thus, modifying the propagation of UHE photons will have no effect. For large maximum rigidities, on the other hand, too many photons are produced, surpassing the upper limits on the UHE photon flux measured by the Pierre Auger Observatory even without increasing the threshold of the pair production. The composition also plays an important role. For heavy compositions a larger fraction of the phase space of models is allowed under an LI assumption. On the other hand, a larger fraction of the phase space is also not sensitive to LIV.

LIV Limits from Different UHECR Astrophysical Models

UHECR astrophysical models are expected to describe current energy spectrum and composition measurements [45,46,47,48,49], reducing the allowed phase space of Figure 3. In Table 1, we present a collection of phenomenological models for the UHECR sources which can either be described by Equation (5) or be approximated by it.
Figure 4, Figure 5 and Figure 6 show the limits on LIV in photon sector for orders n = 0 , 1 , 2 considering each model from Table 1. The range of limits is quite broad, covering from complete insensitivity to LIV up to δ γ , 0 > 10 22 , δ γ , 1 > 10 42 eV 1 E LIV ( 1 ) > 10 42 eV and δ γ , 2 > 10 60 eV 2 E LIV ( 2 ) > 10 30 eV , which, for n = 1 , would be 13 orders of magnitude more restrictive than current subluminal LIV limits obtained in the photon sector using TeV gamma rays [11,13]. This emphasizes once more the strong model dependence of these results. A study comparing the capability of these models to describe UHECR data is outside the scope of this work.
A further reduction of the allowed phase space of UHECR models is expected in the near future with the ongoing upgrade of current facilities, in particular, the AugerPrime [57,58] at the Pierre Auger Observatory and the TAx4 [59] at the Telescope Array experiment. Future data obtained with the upgraded experiments will improve current statistics, allowing a better description of the spectral features, specially the suppression, which has a direct impact on the possible values for the spectral index and for the maximum power of acceleration of the sources. AugerPrime will also present the capability of investigating the presence of a proton component at the highest energies. If this is confirmed, models which predict a larger flux of UHE photons will be preferred and, consequently, the possibility of imposing restrictive LIV limits is favored. Future experiments such as the Probe of Extreme Multi-Messenger Astrophysics (POEMMA) [60] and Global Cosmic Ray Observatory (GCOS) [61] are expected to shrink the phase space even further in the years to come, softening the model dependence of current LIV studies using UHECR.

4. Testing LIV with UHECR

As first proposed independently by Greisen and by Zatsepin & Kuzmin [62,63], the propagation of UHECR at energies of a few tens of EeV is limited by interactions with the photon background. A sharp decrease in the attenuation length of protons is expected above E 10 19.5 eV due to photopion production ( p + γ bg Δ + p / n + π 0 / + ). Furthermore, at these energies, the propagation of heavier nuclei is restricted to a few tens of Mpc by the photodisintegration, through which the nucleus emits a nucleon and becomes a lighter chemical species ( A N + γ bg A 1 N + p / n ). The emission of two or more nucleons is also possible, but rates are considerably smaller. At lower energies, 10 18 eV E 10 19.5 eV, the propagation is dominated by pair production ( p + γ bg p + e + e + ).
These interactions create a propagation horizon, which, combined with the maximum power of acceleration of the sources, leads to suppression at the highest energy end of the spectrum [42] and increases the relative contribution of local sources located up to a few tens of Mpc [64,65]. The relative contribution to the suppression coming from the propagation horizon and from the acceleration limitations is yet to be clarified and differs for each UHECR phenomenological models (see, e.g., Table 1).
If LIV is considered in the hadronic sector, the kinematics of these interactions is modified, leaving imprints that could be detected in UHECR data. Below, we develop further the previous calculations [15,66] of photopion production and photodisintegration under LIV assumption.

4.1. Kinematics of Interactions of UHECR with the Photon Background Considering LIV

The kinematics of any interaction in the form of ( a + b c + d ) can be solved by imposing energy-momentum conservation such as
s init = ( E a + E b ) 2 p a + p b 2 = s final = ( E c + E d ) 2 p c + p d 2 ,
where s is the square of the total rest energy in the center-of-mass reference frame (CMS). Following the MDR in Equation (1), we also define the square of the energy of the particle i in the reference frame where its momentum is zero as
s i = E i 2 p i 2 = m i 2 + δ i , n E i ( n + 2 ) .
The last term incorporates the LIV effects and can be seen as a shift in the effective mass of the particle. For the cases treated here, b represents the background photon, with energy ϵ , for which no LIV is considered and, thus, m b = 0 and δ b = 0 . The energies of the outgoing particles, E c and E d , are unknown and can be expressed as a function of the inelasticity, K, of the interaction,
E c = K E a E d = ( 1 K ) E a .
In order to calculate s = s init = s final , we use the reference frame of the propagating UHECR (a), denoted with ,
s = s init = E a + ϵ 2 + ϵ 2 = s a + 2 s a ϵ .
Joining Equations (6)–(9) leads to
1 K = s + s d ( K ) s c ( K ) 2 s + cos θ s + s d ( K ) s c ( K ) 2 s 2 s d ( K ) s ,
where θ is the angle between the outgoing particles, c and d.
For the photodisintegration, the bound energy of the nucleus also needs to be taken into account. This can be achieved by shifting Equation (9) [66],
s = s a + 2 s a ϵ s = s a + 2 s a ϵ ϵ thr LI ,
where ϵ thr LI is the energy threshold in the NRF in a LI scenario.
Finally, the phase space of the interaction can be obtained as the combinations of E a , ϵ , δ a , n , δ c , n and δ d , n for which Equation (10) has at least one real solution with 0 < K < 1 .
While LIV could be introduced independently for each particle species motivated by the underlying theory, kinematically and in a phenomenological approach for astroparticles [15], one can consider a MDR to hadronic particles ( δ had , n ), such as protons (p), nuclei (A: the atomic mass), and pions ( π ), then, Equation (1) reads,
E p 2 = p p 2 + m p 2 + δ p , n E p n + 2 ,
E A 2 = p A 2 + m A 2 + δ A , n E A n + 2 ,
E π 2 = p π 2 + m π 2 + δ π , n E π n + 2 .
Considering heavier nuclei as a superposition of A-protons,
E A 2 = p A 2 + m A 2 + δ A , n E A n + 2 , A 2 E p 2 = A 2 p p 2 + A 2 m p 2 + δ A , n A n + 2 E p n + 2 ,
Then, using Equation (12) in Equation (15),
δ p , n = A n δ A , n .
Following that LIV effects on pion production process relies on δ π , n δ p , n [15], and using the latest derivation in Equation (16), we consider,
δ p , n = A n δ A , n = δ π , n / 2 : = δ had , n .
Figure 7 and Figure 8 show the energy threshold in the nucleus reference frame (NRF) for the pion production ( a p , c p , d π ) and for the photodisintegration ( a A N , c A 1 N , d p ) for different positive LIV hadronic coefficients ( δ had , n ). Similar to the case of UHE photons, the main effect of LIV is a pronounced increase in the energy threshold (and consequent decrease in the phase space), which becomes significant above given energy that depends on δ had , n . Nevertheless, the effect is saturated for the highest energies differently from the pair production. This is because considered particles affected by LIV are present on both sides of the interaction. The effect of negative values of δ had , n is the opposite, decreasing the energy threshold. It is, however, suppressed by a decrease in the cross-section for lower background photon energies.
For the case of the photodisintegration, [66] presents a parameterization for the energy threshold in the NRF in the form of
ϵ thr LIV = ϵ thr LI + a A , δ had , n log E A / eV 1 + exp c A , δ had , n log E A / eV b A , δ had , n ,
where a, b, and c have their own functional forms. UHECR propagation is usually solved via Monte Carlo simulations [43,67] and, thus, can be computationally costly. For that reason, parameterizing the calculation of LIV effects is a useful approach for speeding up the process.

4.2. Searching for LIV in UHECR Data

The pioneering work [70] on testing LIV with UHECR was motivated by the lack of the flux suppression at the highest energies as published by the Akeno Giant Air Shower Array (AGASA) [71]. The absence of a cut-off in the energy spectrum was interpreted as a hint of LIV in the highest energies which would cause the decrease in the interaction rate of UHECRs. However, the data measured by the High-Resolution Fly’s Eye Cosmic Ray Detector (HiRes) [72] and by the Pierre Auger Observatory [73] have shown a suppression at the expected energy range. More recent results confirmed the suppression of the energy spectrum at 46 EeV with more than 5 σ confidence level [45,46]. Accordingly, LIV studies [15,74,75] have changed their approach and focused on estimating the LIV scale for which the lack of suppression would be in contradiction with measurements and, thus, LIV limits could be imposed. These works considered a pure proton composition and found that the data at that time could be well described up to some small LIV coefficients, and limits ranging from δ p , 0 10 22 were imposed.
The continuous operation of the Pierre Auger Observatory imposed another challenge to the LIV studies. The distributions of the depth in which the shower has its maximum energy deposit, X max , cannot be described by a pure proton composition and point towards an intermediate composition [47,48,76].
Over the last years, the Pierre Auger Collaboration has worked on searching for LIV imprints in their data [40,77,78] taking into account the measured energy spectrum and X max distributions. A combined fit of both energy spectrum and X max distributions was performed under different LIV assumptions. The effects of LIV on the photodisintegration were also considered to treat the chemical composition properly. The combined fit follows the procedure in [42], where a homogeneous distribution of UHECR sources emitting cosmic rays isotropically is considered, and the propagation is simulated with Monte Carlo methods.
The parameters of the energy spectrum used in the fit were shown to be sensitive to the LIV assumption. In particular, for intermediate to strong LIV coefficients, the spectral index shifts towards 2 as predicted in mechanisms based on Fermi acceleration [79,80,81,82,83]. The deviance of the fit decreases for LIV coefficients of the order δ had , 0 = 10 20 to 21 . Nevertheless, the absolute value of the deviance of the fit is quite large, with a reduced value of 300/101, indicating overall poor description of the data by the considered model. Therefore, even though the results point to a preference for models with fewer interactions during propagation, it was not possible to claim whether that could be caused by LIV or by other astrophysical features not covered in the simplistic model considered, such as a larger contribution of local sources, effects of structured extragalactic magnetic fields or differences in emission from different sources.

5. Conclusions and Future Prospects

In this review, we explore the potential of UHE astroparticles as probes for LIV, presenting the current status of phenomenological research in this area. In particular, we discuss LIV in two sectors, photonic and hadronic, using the latest data of UHE photons and UHECR, respectively.
For the case of UHE photons, we review a commonly used analytical approach for the shift in the energy threshold of the pair production due to LIV. This interaction drives the photon propagation, leading to an absorption of the initial flux that modulates the expected measurements on Earth. When LIV is considered, an increase in the energy threshold is predicted, resulting in fewer interactions and, thus, a higher expected flux on Earth. For some astrophysical and subluminal LIV scenarios, the predicted flux is stronger than current upper limits on the photon flux, and, consequently, limits on LIV can be imposed.
We discuss the model dependence of these results. We show that, within a range of reasonable astrophysical assumptions for the UHECR sources, the results can vary from complete insensitivity to LIV to the most restrictive LIV limits in the literature, δ γ , 0 > 10 22 , δ γ , 1 > 10 42 eV 1 E LIV ( 1 ) > 10 42 eV and δ γ , 2 > 10 60 eV 2 E LIV ( 2 ) > 10 30 eV . We demonstrate that such a dependence is dominated by the maximum attainable energy at the sources, followed by the chemical composition. Low maximum power of acceleration of the sources contributes to insensitivity to LIV since most UHECR do not achieve enough energy to produce pions and, therefore, fewer UHE photons are emitted. Heavier compositions tend to lead to weaker LIV limits in the same direction, as the photodisintegration dominates over the pion production. In Figure 3, we present the LIV limits imposed by a collection of astrophysical models (Table 1). A comparison between the capabilities of each model to describe UHECR data is beyond the scope of this work, and, thus, neither of the limits (or absence of limits) are favored.
For the case of UHECR, we review the calculation of the effects of LIV in the kinematics of interactions which can be expressed in the form a + b c + d . In particular, we discuss the cases of the pion production and the photodisintegration, the leading energy losses for the highest energies. The inelasticity of the interaction can be reduced to a transcendental equation. For the photodisintegration, the energy threshold can be described by the parameterization proposed in [66], significantly speeding up the simulation of the UHECR propagation with LIV. Once more, the main LIV effect is an increase in the energy threshold, which leads to fewer interactions.
We briefly summarize the efforts of the community to describe the UHECR data considering LIV. Pioneer works have considered LIV in the proton sector and proposed to describe the energy spectrum measured by the AGASA, HiRes, and Pierre Auger Observatory experiments with a pure proton composition. Restrictive limits on the proton LIV coefficient were imposed. Nevertheless, a pure proton composition conflicts with current X max measurements performed by the Pierre Auger Observatory. Recently, the Pierre Auger Collaboration has published a fit of the energy spectrum and X max distributions under LIV assumptions. The results indicate a smaller deviance for a LIV scenario, however, a bad fit performance is seen overall and a more realistic model for the UHECR sources is needed for conclusive claims.
In summary, UHE astroparticles have proven to be a crucial tool in the search for a possible violation of Lorentz invariance. Current experiments provide precise and high statistics measurements, and a robust and widely spread framework was developed by many authors for the calculation of expected signatures of LIV in UHE data. Nevertheless, modeling UHECR is intricate due to their charged nature, the lack of knowledge about galactic and extragalactic magnetic fields, and the intrinsic fluctuations of extensive air-showers, which prevents an event-by-event composition measurement. Results on LIV strongly differ for different classes of models that can describe current UHE data. Planned upgrades of current experiments, such as AugerPrime [57,58] for the Pierre Auger Observatory and TAx4 [59] for the Telescope Array experiment, as well as future experiments such as the Probe of Extreme Multi-Messenger Astrophysics (POEMMA) [60] and Global Cosmic Ray Observatory (GCOS) [61] will be able to shrink the phase space of UHECR models, by constraining a possible proton component and better measuring the energy spectrum beyond the suppression. We are, thus, yet to see the full potential of UHE astroparticles as probes on LIV in the upcoming years.

Author Contributions

Conceptualization, R.G.L., H.M.-H. and V.d.S.; writing—original draft preparation, R.G.L.; writing—review and editing, H.M.-H. and V.d.S. All authors have read and agreed to the published version of the manuscript.

Funding

VdS acknowledges FAPESP support No. 2021/01089-1 and CNPq. VdS also acknowledges the National Laboratory for Scientific Computing (LNCC/MCTI, Brazil) for providing HPC resources of the SDumont supercomputer, which have contributed to the research results reported within this paper (http://sdumont.lncc.br, accessed on 8 June 2022).

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Abbreviations

The following abbreviations are used in this manuscript:
AGASAAkeno Giant Air Shower Array
CMBCosmic microwave background
CRCosmic rays
EBLExtragalactic background bight
HiResHigh Resolution Fly’s Eye Cosmic Ray Detector
LILorentz invariant
LIVLorentz invariance violation
MDRModified dispersion relation
NRFNucleus reference frame
SMStandard Model
UHEUltra-high energy
UHECRUltra-high-energy cosmic rays

References

  1. Addazi, A.; Alvarez-Muniz, J.; Alves Batista, R.; Amelino-Camelia, G.; Antonelli, V.; Arzano, M.; Asorey, M.; Atteia, J.L.; Bahamonde, S.; Bajardi, F.; et al. Quantum gravity phenomenology at the dawn of the multi-messenger era—A review. Prog. Part. Nucl. Phys. 2022, 125, 103948. [Google Scholar] [CrossRef]
  2. Alfaro, J. Quantum Gravity and Lorentz Invariance Violation in the Standard Model. Phys. Rev. Lett. 2005, 94, 221302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Amelino-Camelia, G. A phenomenological description of space-time noise in quantum gravity. Nature 2001, 410, 1065–1067. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Bluhm, R. Observational Constraints on Local Lorentz Invariance. In Springer Handbook of Spacetime; Springer: Berlin, Germany, 2014; pp. 485–507. [Google Scholar] [CrossRef]
  5. Colladay, D.; Kostelecky, V.A. Lorentz violating extension of the standard model. Phys. Rev. 1998, D58, 116002. [Google Scholar] [CrossRef] [Green Version]
  6. Ellis, J.; Mavromatos, N.E.; Nanopoulos, D.V. A Microscopic Recoil Model for Light-Cone Fluctuations in Quantum Gravity. Phys. Rev. D 1999, 61, 027503. [Google Scholar] [CrossRef] [Green Version]
  7. Gambini, R.; Pullin, J. Nonstandard optics from quantum spacetime. Phys. Rev. 1999, 59, 124021. [Google Scholar]
  8. Kostelecky, V.A.; Samuel, S. Spontaneous Breaking of Lorentz Symmetry in String Theory. Phys. Rev. 1989, D39, 683. [Google Scholar] [CrossRef] [Green Version]
  9. Nambu, Y. Quantum Electrodynamics in Nonlinear Gauge. Suppl. Prog. Theor. Phys. 1968, E68, 190–195. [Google Scholar] [CrossRef] [Green Version]
  10. Potting, R. Lorentz and CPT violation. J. Phys. Conf. Ser. 2013, 447, 012009. [Google Scholar] [CrossRef]
  11. Martínez-Huerta, H.; Lang, R.G.; de Souza, V. Lorentz Invariance Violation Tests in Astroparticle Physics. Symmetry 2020, 12, 1232. [Google Scholar] [CrossRef]
  12. Biteau, J.; Williams, D.A. The extragalactic background light, the Hubble constant, and anomalies: Conclusions from 20 years of TeV gamma-ray observations. Astrophys. J. 2015, 812, 60. [Google Scholar] [CrossRef] [Green Version]
  13. Lang, R.G.; Martínez-Huerta, H.; de Souza, V. Improved limits on Lorentz invariance violation from astrophysical gamma-ray sources. Phys. Rev. D 2019, 99, 043015. [Google Scholar] [CrossRef] [Green Version]
  14. Guedes Lang, R.; Martínez-Huerta, H.; de Souza, V. Limits on the Lorentz Invariance Violation from UHECR astrophysics. Astrophys. J. 2018, 853, 23. [Google Scholar] [CrossRef]
  15. Scully, S.T.; Stecker, F.W. Lorentz Invariance Violation and the Observed Spectrum of Ultrahigh Energy Cosmic Rays. Astropart. Phys. 2009, 31, 220–225. [Google Scholar] [CrossRef] [Green Version]
  16. Stecker, F.W.; Scully, S.T. Searching for New Physics with Ultrahigh Energy Cosmic Rays. New J. Phys. 2009, 11, 085003. [Google Scholar] [CrossRef]
  17. Galaverni, M.; Sigl, G. Lorentz Violation in the Photon Sector and Ultra-High Energy Cosmic Rays. Phys. Rev. Lett. 2008, 100, 021102. [Google Scholar] [CrossRef] [Green Version]
  18. Stecker, F.W.; Glashow, S.L. New tests of Lorentz invariance following from observations of the highest energy cosmic gamma-rays. Astropart. Phys. 2001, 16, 97–99. [Google Scholar] [CrossRef] [Green Version]
  19. Stecker, F.W. Constraints on Lorentz invariance violating quantum gravity and large extra dimensions models using high energy gamma-ray observations. Astropart. Phys. 2003, 20, 85–90. [Google Scholar] [CrossRef] [Green Version]
  20. Martínez-Huerta, H.; Pérez-Lorenzana, A. Restrictions from Lorentz invariance violation on cosmic ray propagation. Phys. Rev. 2017, D95, 063001. [Google Scholar] [CrossRef] [Green Version]
  21. Astapov, K.; Kirpichnikov, D.; Satunin, P. Photon splitting constraint on Lorentz Invariance Violation from Crab Nebula spectrum. JCAP 2019, 1904, 054. [Google Scholar] [CrossRef] [Green Version]
  22. Schreck, M. Vacuum Cherenkov radiation for Lorentz-violating fermions. Phys. Rev. D 2017, 96, 095026. [Google Scholar] [CrossRef] [Green Version]
  23. Schreck, M. (Gravitational) Vacuum Cherenkov Radiation. Symmetry 2018, 10, 424. [Google Scholar] [CrossRef] [Green Version]
  24. Klinkhamer, F.R.; Schreck, M. New two-sided bound on the isotropic Lorentz-violating parameter of modified Maxwell theory. Phys. Rev. D 2008, 78, 085026. [Google Scholar] [CrossRef] [Green Version]
  25. Amelino-Camelia, G.; Ellis, J.R.; Mavromatos, N.E.; Nanopoulos, D.V.; Sarkar, S. Tests of quantum gravity from observations of gamma-ray bursts. Nature 1998, 393, 763–765. [Google Scholar] [CrossRef] [Green Version]
  26. Ellis, J.R.; 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]
  27. Satunin, P. Two-sided constraints on Lorentz invariance violation from Tibet-ASγ and LHAASO very-high-energy photon observations. Eur. Phys. J. C 2021, 81, 750. [Google Scholar] [CrossRef]
  28. Chen, L.; Xiong, Z.; Li, C.; Chen, S.; He, H. Strong constraints on Lorentz violation using new γ-ray observations around PeV. Chin. Phys. C 2021, 45, 105105. [Google Scholar] [CrossRef]
  29. Kostelecky, V.A.; Russell, N. Data Tables for Lorentz and CPT Violation. arXiv 2008, arXiv:hep-ph/0801.0287. [Google Scholar]
  30. Aab, A. et al. [Pierre Auger Collaboration] Search for photons with energies above 1018 eV using the hybrid detector of the Pierre Auger Observatory. JCAP 2017, 04, 009, Erratum in JCAP 2020, 9, E02. [Google Scholar] [CrossRef]
  31. Rautenberg, J. et al. [Pierre Auger Collaboration] Limits on ultra-high energy photons with the Pierre Auger Observatory. PoS 2019, ICRC2019, 398. [Google Scholar]
  32. Savina, P. [Pierre Auger Collaboration]. A search for ultra-high-energy photons at the Pierre Auger Observatory exploiting air-shower Universality. PoS 2021, ICRC2021, 373. [Google Scholar]
  33. Hooper, D.; Sarkar, S.; Taylor, A.M. The intergalactic propagation of ultrahigh energy cosmic ray nuclei. Astropart. Phys. 2007, 27, 199–212. [Google Scholar] [CrossRef] [Green Version]
  34. Allard, D. Extragalactic propagation of ultrahigh energy cosmic-rays. Astropart. Phys. 2012, 39–40, 33–43. [Google Scholar] [CrossRef] [Green Version]
  35. Galaverni, M.; Sigl, G. Lorentz Violation and Ultrahigh-Energy Photons. Phys. Rev. D 2008, 78, 063003. [Google Scholar] [CrossRef] [Green Version]
  36. De Angelis, A.; Galanti, G.; Roncadelli, M. Transparency of the Universe to gamma rays. Mon. Not. R. Astron. Soc. 2013. [Google Scholar] [CrossRef] [Green Version]
  37. Breit, G.; Wheeler, J.A. Collision of Two Light Quanta. Phys. Rev. 1934, 46, 1087–1091. [Google Scholar] [CrossRef]
  38. Gilmore, R.; Somerville, R.; Primack, J.; Dominguez, A. Semi-analytic modeling of the EBL and consequences for extragalactic gamma-ray spectra. Mon. Not. Roy. Astron. Soc. 2012, 422, 3189. [Google Scholar] [CrossRef] [Green Version]
  39. Gervasi, M.; Tartari, A.; Zannoni, M.; Boella, G.; Sironi, G. The Contribution of the Unresolved Extragalactic Radio Sources to the Brightness Temperature of the Sky. Astrophys. J. 2008, 682, 223–230. [Google Scholar] [CrossRef] [Green Version]
  40. Abreu, P.; Aglietta, M.; Albury, J.; Allekotte, I.; Cheminant, K.A.; Almela, A.; Alvarez-Muñiz, J.; Batista, R.A.; Anastasi, G.; Anchordoqui, L.; et al. Testing effects of Lorentz invariance violation in the propagation of astroparticles with the Pierre Auger Observatory. J. Cosmol. Astropart. Phys. 2022, 2022, 023. [Google Scholar] [CrossRef]
  41. Kampert, K.H.; Sarkar, B. Ultra-High Energy Photon and Neutrino Fluxes in Realistic Astrophysical Scenarios. In Proceedings of the International Cosmic Ray Conference, Beijing, China, 11–18 August 2011; Volume 2, p. 198. [Google Scholar] [CrossRef]
  42. Aab, A. et al. [Pierre Auger Collaboration] Combined fit of spectrum and composition data as measured by the Pierre Auger Observatory. JCAP 2017, 04, 038, Erratum in JCAP 2018, 03, E02. [Google Scholar] [CrossRef]
  43. Batista, R.; Dundovic, A.; Erdmann, M.; Kampert, K.H.; Kuempel, D.; Müller, G.; Sigl, G.; van Vliet, A.; Walz, D.; Winchen, T. CRPropa 3—A Public Astrophysical Simulation Framework for Propagating Extraterrestrial Ultra-High Energy Particles. JCAP 2016, 05, 038. [Google Scholar] [CrossRef]
  44. Settimo, M.; De Domenico, M.; Lyberis, H. EleCa: A Monte Carlo code for the propagation of extragalactic photons at ultra-high energy. Nucl. Phys. B Proc. Suppl. 2013, 239–240, 279–282. [Google Scholar] [CrossRef] [Green Version]
  45. Aab, A. et al. [Pierre Auger Collaboration] Measurement of the cosmic-ray energy spectrum above 2.5 × 1018 eV using the Pierre Auger Observatory. Phys. Rev. D 2020, 102, 062005. [Google Scholar] [CrossRef]
  46. Aab, A. et al. [Pierre Auger Collaboration] Features of the Energy Spectrum of Cosmic Rays above 2.5 × 1018 eV using the Pierre Auger Observatory. Phys. Rev. Lett. 2020, 125, 121106. [Google Scholar] [CrossRef]
  47. Aab, A. et al. [Pierre Auger Collaboration] Depth of maximum of air-shower profiles at the Pierre Auger Observatory. II. Composition implications. Phys. Rev. D 2014, 90, 122006. [Google Scholar] [CrossRef] [Green Version]
  48. Aab, A. et al. [Pierre Auger Collaboration] Depth of Maximum of Air-Shower Profiles at the Pierre Auger Observatory: Measurements at Energies above 1017.8 eV. Phys. Rev. D 2014, 90, 122005. [Google Scholar] [CrossRef] [Green Version]
  49. Abbasi, R. et al. [Pierre Auger Collaboration] Joint analysis of the energy spectrum of ultra-high-energy cosmic rays as measured at the Pierre Auger Observatory and the Telescope Array. PoS 2021, ICRC2021, 337. [Google Scholar] [CrossRef]
  50. Berezinsky, V.; Gazizov, A.Z.; Grigorieva, S.I. On astrophysical solution to ultrahigh-energy cosmic rays. Phys. Rev. D 2006, 74, 043005. [Google Scholar] [CrossRef] [Green Version]
  51. Allard, D.; Busca, N.G.; Decerprit, G.; Olinto, A.V.; Parizot, E. Implications of the cosmic ray spectrum for the mass composition at the highest energies. JCAP 2008, 10, 033. [Google Scholar] [CrossRef]
  52. Aloisio, R.; Berezinsky, V.; Blasi, P. Ultra high energy cosmic rays: Implications of Auger data for source spectra and chemical composition. JCAP 2014, 10, 020. [Google Scholar] [CrossRef] [Green Version]
  53. Taylor, A.M.; Ahlers, M.; Hooper, D. Indications of Negative Evolution for the Sources of the Highest Energy Cosmic Rays. Phys. Rev. D 2015, 92, 063011. [Google Scholar] [CrossRef] [Green Version]
  54. Unger, M.; Farrar, G.R.; Anchordoqui, L.A. Origin of the ankle in the ultrahigh energy cosmic ray spectrum, and of the extragalactic protons below it. Phys. Rev. D 2015, 92, 123001. [Google Scholar] [CrossRef] [Green Version]
  55. Muzio, M.S.; Unger, M.; Farrar, G.R. Progress towards characterizing ultrahigh energy cosmic ray sources. Phys. Rev. D 2019, 100, 103008. [Google Scholar] [CrossRef] [Green Version]
  56. Luce, Q.; Marafico, S.; Biteau, J.; Condorelli, A.; Deligny, O. Observational constraints on cosmic-ray escape from UHE accelerators. arXiv 2022, arXiv:astro-ph.HE/2207.08092. [Google Scholar]
  57. Martello, D. et al. [Pierre Auger Collaboration] The Pierre Auger Observatory Upgrade. PoS 2017, ICRC2017, 383. [Google Scholar]
  58. Cataldi, G. et al. [Pierre Auger Collaboration] The upgrade of the Pierre Auger Observatory with the Scintillator Surface Detector. PoS 2021, ICRC2021, 251. [Google Scholar]
  59. Abbasi, R.U. et al. [Pierre Auger Collaboration] Surface detectors of the TAx4 experiment. Nucl. Instrum. Meth. A 2021, 1019, 165726. [Google Scholar] [CrossRef]
  60. Olinto, A.V. et al. [Pierre Auger Collaboration] The POEMMA (Probe of Extreme Multi-Messenger Astrophysics) observatory. JCAP 2021, 06, 007. [Google Scholar] [CrossRef]
  61. Hörandel, J.R. GCOS-The Global Cosmic Ray Observatory. PoS 2021, ICRC2021, 027. [Google Scholar] [CrossRef]
  62. Greisen, K. End to the cosmic ray spectrum? Phys. Rev. Lett. 1966, 16, 748–750. [Google Scholar] [CrossRef]
  63. Zatsepin, G.; Kuzmin, V. Upper limit of the spectrum of cosmic rays. JETP Lett. 1966, 4, 78–80. [Google Scholar]
  64. Taylor, A.M.; Ahlers, M.; Aharonian, F.A. The need for a local source of UHE CR nuclei. Phys. Rev. D 2011, 84, 105007. [Google Scholar] [CrossRef] [Green Version]
  65. Lang, R.G.; Taylor, A.M.; Ahlers, M.; de Souza, V. Revisiting the distance to the nearest ultrahigh energy cosmic ray source: Effects of extragalactic magnetic fields. Phys. Rev. D 2020, 102, 063012. [Google Scholar] [CrossRef]
  66. Lang, R.G. Effects of Lorentz Invariance Violation on the Ultra-High Energy Cosmic Rays Spectrum. Master’s Thesis, University of São Paulo, São Paulo, Brazil, 2017. [Google Scholar] [CrossRef] [Green Version]
  67. Aloisio, R.; Boncioli, D.; Di Matteo, A.; Grillo, A.F.; Petrera, S.; Salamida, F. SimProp v2r4: Monte Carlo simulation code for UHECR propagation. JCAP 2017, 11, 009. [Google Scholar] [CrossRef] [Green Version]
  68. Mucke, A.; Engel, R.; Rachen, J.P.; Protheroe, R.J.; Stanev, T. SOPHIA: Monte Carlo simulations of photohadronic processes in astrophysics. Comput. Phys. Commun. 2000, 124, 290–314. [Google Scholar] [CrossRef] [Green Version]
  69. Rachen, J.P. Interaction Processes and Statistical Properties of the Propagation of Cosmic Rays in Photon Backgrounds. Ph.D. Thesis, Max-Planck-Institute for Radioastronomy, Bonn, Germany, 1996. [Google Scholar]
  70. Stecker, F.W.; Scully, S.T. Lorentz invariance violation and the spectrum and source power of ultrahigh energy cosmic rays. Astropart. Phys. 2005, 23, 203–209. [Google Scholar] [CrossRef] [Green Version]
  71. Takeda, M.; Sakaki, N.; Honda, K.; Chikawa, M.; Fukushima, M.; Hayashida, N.; Inoue, N.; Kadota, K.; Kakimoto, F.; Kamata, K.; et al. Energy determination in the Akeno Giant Air Shower Array experiment. Astropart. Phys. 2003, 19, 447–462. [Google Scholar] [CrossRef] [Green Version]
  72. Abbasi, R.U.; Abu-Zayyad, T.; Allen, M.; Amman, J.F.; Archbold, G.; Belov, K.; Belz, J.W.; Zvi, S.B.; Bergman, D.R.; Blake, S.A.; et al. First observation of the Greisen-Zatsepin-Kuzmin suppression. Phys. Rev. Lett. 2008, 100, 101101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Abraham, J.; Abreu, P.; Aglietta, M.; Aguirre, C.; Allard, D.; Allekotte, I.; Allen, J.; Allison, P.; Alvarez-Muniz, J.; Ambrosio, M.; et al. Observation of the suppression of the flux of cosmic rays above 4 × 1019eV. Phys. Rev. Lett. 2008, 101, 061101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Bi, X.J.; Cao, Z.; Li, Y.; Yuan, Q. Testing Lorentz Invariance with Ultra High Energy Cosmic Ray Spectrum. Phys. Rev. D 2009, 79, 083015. [Google Scholar] [CrossRef] [Green Version]
  75. Maccione, L.; Taylor, A.M.; Mattingly, D.M.; Liberati, S. Planck-scale Lorentz violation constrained by Ultra-High-Energy Cosmic Rays. JCAP 2009, 04, 022. [Google Scholar] [CrossRef]
  76. De Souza, V. Testing the agreement between the Xmax distributions measured by the Pierre Auger and Telescope Array Observatories. PoS 2018, ICRC2017, 522. [Google Scholar] [CrossRef]
  77. Boncioli, D. et al. [Pierre Auger Collaboration] Probing Lorentz symmetry with the Pierre Auger Observatory. PoS 2018, ICRC2017, 561. [Google Scholar] [CrossRef]
  78. Lang, R.G. et al. [Pierre Auger Collaboration] Testing Lorentz Invariance Violation at the Pierre Auger Observatory. PoS 2019, ICRC2019, 450. [Google Scholar] [CrossRef]
  79. Fermi, E. On the origin of the Cosmic Radiation. Phys. Rev. 1949, 75, 1169–1174. [Google Scholar] [CrossRef]
  80. Axford, W. Acceleration of Cosmic Rays by Shock Waves. Ann. N. Y. Acad. Sci. 2006, 375, 297–313. [Google Scholar] [CrossRef]
  81. Krymskii, G.F. A regular mechanism for the acceleration of charged particles on the front of a shock wave. Akad. Nauk SSSR Dokl. 1977, 234, 1306–1308. Available online: https://ui.adsabs.harvard.edu/abs/1977DoSSR.234.1306K (accessed on 8 December 2020).
  82. Bell, A.R. The acceleration of cosmic rays in shock fronts. I. Mon. Not. R. Astron. Soc. 1978, 182, 147–156. [Google Scholar] [CrossRef] [Green Version]
  83. Blandford, R.; Ostriker, J. Particle Acceleration by Astrophysical Shocks. Astrophys. J. Lett. 1978, 221, L29–L32. [Google Scholar] [CrossRef]
Figure 1. Energy threshold for the pair production as a function of the energy of the propagating photon. The black line shows the results for a LI assumption, while the different colored lines show the results for different subluminal LIV coefficients. The right panel shows the background distribution for that given energy. The dashed lines indicate the region in which each background is dominant. For the EBL, the model in [38] was used as an example. For the RB, the parameterization from [39] was used.
Figure 1. Energy threshold for the pair production as a function of the energy of the propagating photon. The black line shows the results for a LI assumption, while the different colored lines show the results for different subluminal LIV coefficients. The right panel shows the background distribution for that given energy. The dashed lines indicate the region in which each background is dominant. For the EBL, the model in [38] was used as an example. For the RB, the parameterization from [39] was used.
Universe 08 00435 g001
Figure 2. Survival probability as a function of the propagating photon energy. The black line and red line show the limiting cases, i.e., LI and maximum LIV, while the different colored lines show the results for different subluminal LIV coefficients. The three panels correspond to different values of redshift.
Figure 2. Survival probability as a function of the propagating photon energy. The black line and red line show the limiting cases, i.e., LI and maximum LIV, while the different colored lines show the results for different subluminal LIV coefficients. The three panels correspond to different values of redshift.
Universe 08 00435 g002
Figure 3. Limits on the subluminal LIV coefficient for the phase space of generic UHECR models. The left and right panels are, respectively, for a pure composition of proton and iron. The shaded red area shows the region of the phase space for which the models are not sensitive to LIV. The shaded black area shows the region already ruled out even in a LI scenario. The different colors show the regions in which the imposed limits would be of a given logarithmic order.
Figure 3. Limits on the subluminal LIV coefficient for the phase space of generic UHECR models. The left and right panels are, respectively, for a pure composition of proton and iron. The shaded red area shows the region of the phase space for which the models are not sensitive to LIV. The shaded black area shows the region already ruled out even in a LI scenario. The different colors show the regions in which the imposed limits would be of a given logarithmic order.
Universe 08 00435 g003
Figure 4. Logarithmic order of the limits on subluminal LIV of order n = 0 imposed for each UHECR model. The models below the dashed lines are insensitive to LIV using this technique and, thus, no limits could be imposed. The red texts show a brief description about the two leading features of the models: its composition and maximum attainable rigidity at the sources.
Figure 4. Logarithmic order of the limits on subluminal LIV of order n = 0 imposed for each UHECR model. The models below the dashed lines are insensitive to LIV using this technique and, thus, no limits could be imposed. The red texts show a brief description about the two leading features of the models: its composition and maximum attainable rigidity at the sources.
Universe 08 00435 g004
Figure 5. Same as Figure 4, but for n = 1 .
Figure 5. Same as Figure 4, but for n = 1 .
Universe 08 00435 g005
Figure 6. Same as Figure 4, but for n = 2 .
Figure 6. Same as Figure 4, but for n = 2 .
Universe 08 00435 g006
Figure 7. Energy threshold in the proton reference frame for the pion production. The black line shows the results for a LI assumption, while the different colored lines show the results for different LIV coefficients. The right panel shows the cross-section for that given energy, based on the parameterizations from [68].
Figure 7. Energy threshold in the proton reference frame for the pion production. The black line shows the results for a LI assumption, while the different colored lines show the results for different LIV coefficients. The right panel shows the cross-section for that given energy, based on the parameterizations from [68].
Universe 08 00435 g007
Figure 8. Energy threshold in the NRF for the photodisintegration. The black line shows the results for a LI assumption, while the different colored lines show the results for different LIV coefficients. The right panel shows the cross section for that given energy, based on the parameterizations from [69]. Top and bottom panels correspond to an initial nuclues of helium and iron, respectively.
Figure 8. Energy threshold in the NRF for the photodisintegration. The black line shows the results for a LI assumption, while the different colored lines show the results for different LIV coefficients. The right panel shows the cross section for that given energy, based on the parameterizations from [69]. Top and bottom panels correspond to an initial nuclues of helium and iron, respectively.
Universe 08 00435 g008
Table 1. List of UHECR models used. The middle column specifies the model when multiple ones come from the same reference.
Table 1. List of UHECR models used. The middle column specifies the model when multiple ones come from the same reference.
ModelDetailsReference
Berezinsky et al. ’02[50]
Allard et al. ’08 (a)pure He[51]
Allard et al. ’08 (b)pure CNO[51]
Allard et al. ’08 (c)pure Si[51]
Aloisio et al. ’15[52]
Taylor et al. ’15 (a) n = 0 [53]
Unger et al. ’15 (a)fiducial model[54]
Unger et al. ’15 (a)galactic abundance[54]
Pierre Auger Collaboration ’17 (a)SPG[42]
Pierre Auger Collaboration ’17 (a)STG[42]
Pierre Auger Collaboration ’17 (a)SPD[42]
Pierre Auger Collaboration ’17 (a)CTG ( γ = 1.03 )[42]
Pierre Auger Collaboration ’17 (a)CTG ( γ = + 0.87 )[42]
Pierre Auger Collaboration ’17 (a)CTD[42]
Pierre Auger Collaboration ’17 (a)CGD[42]
Muzio et al. ’20[55]
Luce et al. ’22[56]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lang, R.G.; Martínez-Huerta, H.; de Souza, V. Ultra-High-Energy Astroparticles as Probes for Lorentz Invariance Violation. Universe 2022, 8, 435. https://doi.org/10.3390/universe8080435

AMA Style

Lang RG, Martínez-Huerta H, de Souza V. Ultra-High-Energy Astroparticles as Probes for Lorentz Invariance Violation. Universe. 2022; 8(8):435. https://doi.org/10.3390/universe8080435

Chicago/Turabian Style

Lang, Rodrigo Guedes, Humberto Martínez-Huerta, and Vitor de Souza. 2022. "Ultra-High-Energy Astroparticles as Probes for Lorentz Invariance Violation" Universe 8, no. 8: 435. https://doi.org/10.3390/universe8080435

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