Extragalactic cosmic ray

Last updated
The energy spectrum for cosmic rays. Cosmic ray flux versus particle energy.svg
The energy spectrum for cosmic rays.

Extragalactic cosmic rays are very-high-energy particles that flow into the Solar System from beyond the Milky Way galaxy. While at low energies, the majority of cosmic rays originate within the Galaxy (such as from supernova remnants), at high energies the cosmic ray spectrum is dominated by these extragalactic cosmic rays. The exact energy at which the transition from galactic to extragalactic cosmic rays occurs is not clear, but it is in the range 1017 to 1018 eV. [1]

Contents

Observation

A 3D simulation of the air shower created by a 1 TeV proton hitting the atmosphere, from the COSMUS group at the University of Chicago. The ground shown is an 8 km x 8 km area. Protonshower.jpg
A 3D simulation of the air shower created by a 1 TeV proton hitting the atmosphere, from the COSMUS group at the University of Chicago. The ground shown is an 8 km x 8 km area.

The observation of extragalactic cosmic rays requires detectors with an extremely large surface area, due to the very limited flux. As a result, extragalactic cosmic rays are generally detected with ground-based observatories, by means of the extensive air showers they create. These ground based observatories can be either surface detectors, which observe the air shower particles which reach the ground, or air fluorescence detectors (also called 'fly's eye' detectors [2] ), which observe the fluorescence caused by the interaction of the charged air shower particles with the atmosphere. In either case, the ultimate aim is to find the mass and energy of the primary cosmic ray which created the shower. Surface detectors accomplish this by measuring the density of particles at the ground, while fluorescence detectors do so by measuring the depth of shower maximum (the depth from the top of the atmosphere at which the maximum number of particles are present in the shower). [3] The two currently operating high energy cosmic ray observatories, the Pierre Auger Observatory and the Telescope Array, are hybrid detectors which use both of these methods. This hybrid methodology allows for a full three-dimensional reconstruction of the air shower, and gives much better directional information as well as more accurate determination of the type and energy of the primary cosmic ray than either technique on its own. [4]

Pierre Auger Observatory

The Pierre Auger Observatory, located in the Mendoza province in Argentina, consists of 1660 surface detectors, each separated by 1.5 km and covering a total area of 3000 km2, and 27 fluorescence detectors at 4 different locations overlooking the surface detectors. [5] [6] The observatory has been in operation since 2004, and began operating at full capacity in 2008 once construction was completed. The surface detectors are water Cherenkov detectors, each detector being a tank 3.6 m in diameter. One of the Pierre Auger Observatory's most notable results is the detection of a dipole anisotropy in the arrival directions of cosmic rays with energy greater than 8 x 1018 eV, which was the first conclusive indication of their extragalactic origin. [7] [8]

Telescope Array

The Telescope Array is located in the state of Utah in the United States of America, and consists of 507 surface detectors separated by 1.2 km and covering a total area of 700 km2 [9] , and 3 fluorescence detector stations with 12-14 fluorescence detectors at each station. [10] The Telescope Array was constructed by a collaboration between the teams formerly operating the Akeno Giant Air Shower Array (AGASA), which was a surface detector array in Japan, and the High Resolution Fly's Eye (HiRes), which was an air fluorescence detector also located in Utah. [11] The Telescope Array was initially designed to detect cosmic rays with energy above 1019 eV, but an extension to the project, the Telescope Array Low Energy extension (TALE) is currently underway and will allow observation of cosmic rays with energies above 3 x 1016 eV [12]

Spectrum and Composition

Energy spectrum of cosmic rays with energy greater than 2.5 x 10 eV from data observed by the Pierre Auger Observatory Pierre Auger cosmic ray spectrum 2020.png
Energy spectrum of cosmic rays with energy greater than 2.5 x 10 eV from data observed by the Pierre Auger Observatory

Two clear and long-known features of the spectrum of extragalactic cosmic rays are the 'ankle', which is a flattening of the spectrum at around 5 x 1018 eV, [14] and suppression of the cosmic ray flux at high energies (above about 4 x 1019 eV). [15] [16] More recently the Pierre Auger Observatory also observed a steepening of the cosmic ray spectrum above the ankle, [17] before the steep cutoff above than 1019 eV (see figure). The spectrum measured by the Pierre Auger Observatory does not appear to depend on the arrival direction of the cosmic rays. [18] However, there are some discrepancies between the spectrum (specifically the energy at which the suppression of flux occurs) measured by the Pierre Auger Observatory in the Southern hemisphere and the Telescope Array in the Northern hemisphere. [19] It is unclear whether this is the result of an unknown systematic error or a true difference between the cosmic rays arriving at the Northern and Southern hemispheres.

The interpretation of these features of the cosmic ray spectrum depends on the details of the model assumed.Historically the ankle is interpreted as the energy at which the steep Galactic cosmic ray spectrum transitions to a flat extragalactic spectrum. [20] However diffusive shock acceleration in supernova remnants, which is the predominant source of cosmic rays below 1015 eV, can accelerate protons only up to 3 x 1015 eV and iron up to 8 x 1016 eV. [20] [21] Thus there must be an additional source of Galactic cosmic rays up to around 1018 eV. On the other hand, the 'dip' model assumes that the transition between Galactic and extragalactic cosmic rays occurs at about 1017 eV. This model assumes that extragalactic cosmic rays are composed purely of protons, and the ankle is interpreted as being due to pair production arising from interactions of cosmic rays with the Cosmic Microwave Background (CMB). [22] This suppresses the cosmic ray flux and thus causes a flattening of the spectrum. Older data, as well as more recent data from the Telescope Array [23] [24] do favour a pure proton composition. However recent Auger data suggests a composition which is dominated by light elements to 2 x 1018 eV, but becomes increasingly dominated by heavier elements with increasing energy. [25] In this case a source of the protons below 2 x 1018 eV is needed.

The suppression of flux at high energies is generally assumed to be due to the Greisen–Zatsepin–Kuz'min (GZK) effect in the case of protons, or due to photodisintegration by the CMB (the Gerasimova-Rozental or GR effect) in the case of heavy nuclei. However it could also be because of the nature of the sources, that is because of the maximum energy to which sources can accelerate cosmic rays. [26]

As mentioned above the Telescope Array and the Pierre Auger Observatory give different results for the most likely composition. However the data used to infer composition from these two observatories is consistent once all systematic effects are taken into account. [19] The composition of extragalactic cosmic rays is thus still ambiguous

Origin

Unlike solar or galactic cosmic rays, little is known about the origins of extragalactic cosmic rays. This is largely due to a lack of statistics: only about 1 extragalactic cosmic ray particle per square kilometer per year reaches the Earth's surface (see figure). The possible sources of these cosmic rays must satisfy the Hillas criterion, [27]

where E is the energy of the particle, q its electric charge, B is the magnetic field in the source and R the size of the source. This criterion comes from the fact that for a particle to be accelerated to a given energy, its Larmor radius must be less than the size of the accelerating region. Once the Larmor radius of the particle is greater than the size of the accelerating region, it escapes and does not gain any more energy. As a consequence of this, heavier nuclei (with a greater number of protons), if present, can be accelerated to higher energies than protons within the same source.

Active galactic nuclei

Image of an active galactic nucleus of the active galaxy M87. M87 jet.jpg
Image of an active galactic nucleus of the active galaxy M87.

Active galactic nuclei (AGNs) are well known to be some of the most energetic objects in the universe, and are therefore often considered as candidates for the production of extragalactic cosmic rays. Given their extremely high luminosity, AGNs can accelerate cosmic rays to the required energies even if only 1/1000 of their energy is used for this acceleration. There is some observational support for this hypothesis. Analysis of cosmic ray measurements with the Pierre Auger Observatory suggests a correlation between the arrival directions of cosmic rays of the highest energies of more than 5×1019  eV and the positions of nearby active galaxies. [28] In 2017, IceCube detected a high energy neutrino with energy 290 TeV whose direction was consistent with a flaring blazar, TXS 0506-056, [29] which strengthened the case for AGNs as a source of extragalactic cosmic rays. Since high-energy neutrinos are assumed to come from the decay of pions produced by the interaction of correspondingly high-energy protons with the Cosmic Microwave Background (CMB) (photo-pion production), or from the photodisintegration of energetic nuclei, and since neutrinos travel essentially unimpeded through the universe, they can be traced back to the source of high-energy cosmic rays.

Clusters of galaxies

A multiwavelength image of the galaxy cluster Abell 1689, with X-ray (purple) and optical (yellow) data. The diffuse X-ray emission arises from the hot intracluster medium Abell 1689- A Galaxy Cluster Makes Its Mark (A galaxy cluster at a distance of about 2.3 billion light years from Earth.) (2941478428).jpg
A multiwavelength image of the galaxy cluster Abell 1689, with X-ray (purple) and optical (yellow) data. The diffuse X-ray emission arises from the hot intracluster medium

Galaxy clusters continuously accrete gas and galaxies from filaments of the cosmic web. As the cold gas which is accreted falls into the hot intracluster medium, it gives rise to shocks at the outskirts of the cluster, which could accelerate cosmic rays through the diffusive shock acceleration mechanism. [30] Large scale radio halos and radio relics, which are expected to be due to synchrotron emission from relativistic electrons, [31] show that clusters do host high energy particles. [32] Studies have found that shocks in clusters can accelerate iron nuclei to 1020 eV, [33] which is nearly as much as the most energetic cosmic rays observed by the Pierre Auger Observatory. [18] However, if clusters do accelerate protons or nuclei to such high energies, they should also produce gamma ray emission due to the interaction of the high-energy particles with the intracluster medium. [34] This gamma ray emission has not yet been observed, [35] which is difficult to explain.

Gamma ray bursts

Gamma ray bursts (GRBs) were originally proposed as a possible source of extragalactic cosmic rays because the energy required to produce the observed flux of cosmic rays was similar their typical luminosity in γ-rays, and because they could accelerate protons to energies of 1020 eV through diffusive shock acceleration. [36] Long gamma ray bursts (GRBs) are especially interesting as possible sources of extragalactic cosmic rays in light of the evidence for a heavier composition at higher energies. Long GRBs are associated with the death of massive stars, [37] which are well known to produce heavy elements. However, in this case many of the heavy nuclei would be photo-disintegrated, leading to considerable neutrino emission also associated with GRBs, which has not been observed. [38] Some studies have suggested that a specific population of GRBs known as low-luminosity GRBs might resolve this, as the lower luminosity would lead to less photo-dissociation and neutrino production. [39] These low luminosity GRBs could also simultaneously account for the observed high-energy neutrinos. [40] However it has also been argued that these low-luminosity GRBs are not energetic enough to be a major source of high energy cosmic rays. [41]

Neutron stars

Neutron stars are formed from the core collapse of massive stars, and as with GRBs can be a source of heavy nuclei. In models with neutron stars - specifically young pulsars or magnetars - as the source of extragalactic cosmic rays, heavy elements (mainly iron) are stripped from the surface of the object by the electric field created by the magnetized neutron star's rapid rotation. [42] This same electric field can accelerate iron nucleii up to 1020 eV. [42] The photodisintegration of the heavy nucleii would produce lighter elements with lower energies, matching the observations of the Pierre Auger Observatory. [43] In this scenario, the cosmic rays accelerated by neutron stars within the Milky Way could fill in the 'transition region' between Galactic cosmic rays produced in supernova remnants, and extragalactic cosmic rays. [44]

See also

Related Research Articles

<span class="mw-page-title-main">Neutrino</span> Elementary particle with extremely low mass

A neutrino is a fermion that interacts only via the weak interaction and gravity. The neutrino is so named because it is electrically neutral and because its rest mass is so small (-ino) that it was long thought to be zero. The rest mass of the neutrino is much smaller than that of the other known elementary particles. The weak force has a very short range, the gravitational interaction is extremely weak due to the very small mass of the neutrino, and neutrinos do not participate in the electromagnetic interaction or the strong interaction. Thus, neutrinos typically pass through normal matter unimpeded and undetected.

<span class="mw-page-title-main">Cosmic ray</span> High-energy particle, mainly originating outside the Solar system

Cosmic rays or astroparticles are high-energy particles or clusters of particles that move through space at nearly the speed of light. They originate from the Sun, from outside of the Solar System in our own galaxy, and from distant galaxies. Upon impact with Earth's atmosphere, cosmic rays produce showers of secondary particles, some of which reach the surface, although the bulk are deflected off into space by the magnetosphere or the heliosphere.

The Greisen–Zatsepin–Kuzmin limit (GZK limit or GZK cutoff) is a theoretical upper limit on the energy of cosmic ray protons traveling from other galaxies through the intergalactic medium to our galaxy. The limit is 5×1019 eV (50 EeV), or about 8 joules (the energy of a proton travelling at ≈ 99.99999999999999999998% the speed of light). The limit is set by the slowing effect of interactions of the protons with the microwave background radiation over long distances (≈ 160 million light-years). The limit is at the same order of magnitude as the upper limit for energy at which cosmic rays have experimentally been detected, although indeed some detections appear to have exceeded the limit, as noted below. For example, one extreme-energy cosmic ray, the Oh-My-God Particle, which has been found to possess a record-breaking 3.12×1020 eV (50 joules) of energy (about the same as the kinetic energy of a 95 km/h baseball).

In astroparticle physics, an ultra-high-energy cosmic ray (UHECR) is a cosmic ray with an energy greater than 1 EeV (1018 electronvolts, approximately 0.16 joules), far beyond both the rest mass and energies typical of other cosmic ray particles.

<span class="mw-page-title-main">Neutrino astronomy</span> Observing low-mass stellar particles

Neutrino astronomy is the branch of astronomy that observes astronomical objects with neutrino detectors in special observatories. Neutrinos are created as a result of certain types of radioactive decay, nuclear reactions such as those that take place in the Sun or high energy astrophysical phenomena, in nuclear reactors, or when cosmic rays hit atoms in the atmosphere. Neutrinos rarely interact with matter, meaning that it is unlikely for them to scatter along their trajectory, unlike photons. Therefore, neutrinos offer a unique opportunity to observe processes that are inaccessible to optical telescopes, such as reactions in the Sun's core. Neutrinos can also offer a very strong pointing direction compared to charged particle cosmic rays.

The Oh-My-God particle was an ultra-high-energy cosmic ray detected on 15 October 1991 by the Fly's Eye camera in Dugway Proving Ground, Utah, United States. Still, as of 2023, it is the highest-energy cosmic ray ever observed. Its energy was estimated as (3.2±0.9)×1020 eV (320 exa-eV). The particle's energy was unexpected and called into question prevailing theories about the origin and propagation of cosmic rays.

<span class="mw-page-title-main">Solar neutrino</span> Extremely light particle produced by the Sun

A solar neutrino is a neutrino originating from nuclear fusion in the Sun's core, and is the most common type of neutrino passing through any source observed on Earth at any particular moment. Neutrinos are elementary particles with extremely small rest mass and a neutral electric charge. They only interact with matter via the weak interaction and gravity, making their detection very difficult. This has led to the now-resolved solar neutrino problem. Much is now known about solar neutrinos, but the research in this field is ongoing.

<span class="mw-page-title-main">Pierre Auger Observatory</span> International cosmic ray observatory in Argentina

The Pierre Auger Observatory is an international cosmic ray observatory in Argentina designed to detect ultra-high-energy cosmic rays: sub-atomic particles traveling nearly at the speed of light and each with energies beyond 1018 eV. In Earth's atmosphere such particles interact with air nuclei and produce various other particles. These effect particles (called an "air shower") can be detected and measured. But since these high energy particles have an estimated arrival rate of just 1 per km2 per century, the Auger Observatory has created a detection area of 3,000 km2 (1,200 sq mi)—the size of Rhode Island, or Luxembourg—in order to record a large number of these events. It is located in the western Mendoza Province, Argentina, near the Andes.

<span class="mw-page-title-main">IceCube Neutrino Observatory</span> Neutrino detector at the South Pole

The IceCube Neutrino Observatory is a neutrino observatory constructed at the Amundsen–Scott South Pole Station in Antarctica. The project is a recognized CERN experiment (RE10). Its thousands of sensors are located under the Antarctic ice, distributed over a cubic kilometre.

<span class="mw-page-title-main">Antarctic Impulsive Transient Antenna</span>

The Antarctic Impulsive Transient Antenna (ANITA) experiment has been designed to study ultra-high-energy (UHE) cosmic neutrinos by detecting the radio pulses emitted by their interactions with the Antarctic ice sheet. This is to be accomplished using an array of radio antennas suspended from a helium balloon flying at a height of about 37,000 meters.

<span class="mw-page-title-main">Neutrino detector</span> Physics apparatus which is designed to study neutrinos

A neutrino detector is a physics apparatus which is designed to study neutrinos. Because neutrinos only weakly interact with other particles of matter, neutrino detectors must be very large to detect a significant number of neutrinos. Neutrino detectors are often built underground, to isolate the detector from cosmic rays and other background radiation. The field of neutrino astronomy is still very much in its infancy – the only confirmed extraterrestrial sources as of 2018 are the Sun and the supernova 1987A in the nearby Large Magellanic Cloud. Another likely source is the blazar TXS 0506+056 about 3.7 billion light years away. Neutrino observatories will "give astronomers fresh eyes with which to study the universe".

<span class="mw-page-title-main">Cosmic-ray observatory</span> Installation built to detect high-energy-particles coming from space

A cosmic-ray observatory is a scientific installation built to detect high-energy-particles coming from space called cosmic rays. This typically includes photons, electrons, protons, and some heavier nuclei, as well as antimatter particles. About 90% of cosmic rays are protons, 9% are alpha particles, and the remaining ~1% are other particles.

The Baikal Deep Underwater Neutrino Telescope (BDUNT) is a neutrino detector conducting research below the surface of Lake Baikal (Russia) since 2003. The first detector was started in 1990 and completed in 1998. It was upgraded in 2005 and again starting in 2015 to build the Baikal Gigaton Volume Detector (Baikal-GVD.) BDUNT has studied neutrinos coming through the Earth with results on atmospheric muon flux. BDUNT picks up many atmospheric neutrinos created by cosmic rays interacting with the atmosphere – as opposed to cosmic neutrinos which give clues to cosmic events and are therefore of greater interest to physicists.

<span class="mw-page-title-main">Cosmology Large Angular Scale Surveyor</span> Microwave telescope array in Chile

The Cosmology Large Angular Scale Surveyor (CLASS) is an array of microwave telescopes at a high-altitude site in the Atacama Desert of Chile as part of the Parque Astronómico de Atacama. The CLASS experiment aims to improve our understanding of cosmic dawn when the first stars turned on, test the theory of cosmic inflation, and distinguish between inflationary models of the very early universe by making precise measurements of the polarization of the Cosmic Microwave Background (CMB) over 65% of the sky at multiple frequencies in the microwave region of the electromagnetic spectrum.

<span class="mw-page-title-main">High Altitude Water Cherenkov Experiment</span>

The High Altitude Water Cherenkov Experiment or High Altitude Water Cherenkov Observatory is a gamma-ray and cosmic ray observatory located on the flanks of the Sierra Negra volcano in the Mexican state of Puebla at an altitude of 4100 meters, at 18°59′41″N97°18′30.6″W. HAWC is the successor to the Milagro gamma-ray observatory in New Mexico, which was also a gamma-ray observatory based around the principle of detecting gamma-rays indirectly using the water Cherenkov method.

Multi-messenger astronomy is astronomy based on the coordinated observation and interpretation of signals carried by disparate "messengers": electromagnetic radiation, gravitational waves, neutrinos, and cosmic rays. They are created by different astrophysical processes, and thus reveal different information about their sources.

<span class="mw-page-title-main">GW170817</span> Gravitational-wave signal detected in 2017

GW 170817 was a gravitational wave (GW) signal observed by the LIGO and Virgo detectors on 17 August 2017, originating from the shell elliptical galaxy NGC 4993. The signal was produced by the last minutes of a binary pair of neutron stars' inspiral process, ending with a merger. It is the first GW observation that has been confirmed by non-gravitational means. Unlike the five previous GW detections, which were of merging black holes not expected to produce a detectable electromagnetic signal, the aftermath of this merger was also seen by 70 observatories on 7 continents and in space, across the electromagnetic spectrum, marking a significant breakthrough for multi-messenger astronomy. The discovery and subsequent observations of GW 170817 were given the Breakthrough of the Year award for 2017 by the journal Science.

TXS 0506+056 is a very high energy blazar – a quasar with a relativistic jet pointing directly towards Earth – of BL Lac-type. With a redshift of 0.3365 ± 0.0010, it is about 1.75 gigaparsecs from Earth. Its approximate location on the sky is off the left shoulder of the constellation Orion. Discovered as a radio source in 1983, the blazar has since been observed across the entire electromagnetic spectrum.

The Waxman-Bahcall bound is a computed upper limit on the observed flux of high energy neutrinos based on the observed flux of high energy cosmic rays. Since the highest energy neutrinos are produced in the same interactions as high energy cosmic rays, the observed rate of production of the latter places a limit on the former. It is named for John Bahcall and Eli Waxman.

<span class="mw-page-title-main">GRB 221009A</span> Gamma-ray burst

GRB 221009A also known as Swift J1913.1+1946 was an unusually bright and long-lasting gamma-ray burst (GRB) jointly discovered by the Neil Gehrels Swift Observatory and the Fermi Gamma-ray Space Telescope on October 9, 2022. The gamma-ray burst was around seven minutes long, but was detectable for more than ten hours following initial detection, and for several hours was bright enough in visible frequencies to be observable by amateur astronomers. Despite being over 2 billion light-years away, it was powerful enough to affect Earth's atmosphere, having the strongest effect ever recorded by a gamma-ray burst on the planet. The peak luminosity of GRB 221009A was measured by Konus-Wind to be ∼ 2.1 × 1047 J/s and by Fermi-GBM to be ∼ 1.0 × 1047 J/s over the 1.024s interval. A burst as energetic and as close to Earth as 221009A is thought to be a once-in-10,000-year event. It was the brightest and most energetic gamma-ray burst ever recorded, being deemed the "BOAT", or brightest of all time.

References

  1. Aloisio, R.; Berezinsky, V.; Gazizov, A. (December 2012). "Transition from galactic to extragalactic cosmic rays". Astroparticle Physics. 39–40: 129–143. arXiv: 1211.0494 . Bibcode:2012APh....39..129A. doi:10.1016/j.astropartphys.2012.09.007. S2CID   9266571.
  2. "HiRes - The High Resolution Fly's Eye Ultra High Energy Cosmic Ray Observatory". www.cosmic-ray.org. Retrieved 2021-04-28.
  3. Kampert, Karl-Heinz; Unger, Michael (May 2012). "Measurements of the Cosmic Ray Composition with Air Shower Experiments". Astroparticle Physics. 35 (10): 660–678. arXiv: 1201.0018 . Bibcode:2012APh....35..660K. doi:10.1016/j.astropartphys.2012.02.004. S2CID   118540292.
  4. Sommers, Paul (1995-08-01). "Capabilities of a giant hybrid air shower detector". Astroparticle Physics. 3 (4): 349–360. Bibcode:1995APh.....3..349S. doi:10.1016/0927-6505(95)00013-7. ISSN   0927-6505.
  5. "Auger Hybrid Detector". www.auger.org. Retrieved 2021-04-28.
  6. The Pierre Auger Collaboration (October 2015). "The Pierre Auger Cosmic Ray Observatory". Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment. 798: 172–213. arXiv: 1502.01323 . Bibcode:2015NIMPA.798..172P. doi: 10.1016/j.nima.2015.06.058 .
  7. The Pierre Auger Collaboration; Aab, A.; Abreu, P.; Aglietta, M.; Samarai, I. Al; Albuquerque, I. F. M.; Allekotte, I.; Almela, A.; Castillo, J. Alvarez; Alvarez-Muñiz, J.; Anastasi, G. A. (2017-09-22). "Observation of a Large-scale Anisotropy in the Arrival Directions of Cosmic Rays above $8 \times 10^{18}$ eV". Science. 357 (6357): 1266–1270. arXiv: 1709.07321 . Bibcode:2017Sci...357.1266P. doi: 10.1126/science.aan4338 . ISSN   0036-8075. PMID   28935800.
  8. The Pierre Auger Collaboration; Aab, A.; Abreu, P.; Aglietta, M.; Albuquerque, I. F. M.; Albury, J. M.; Allekotte, I.; Almela, A.; Castillo, J. Alvarez; Alvarez-Muñiz, J.; Anastasi, G. A. (2018-11-13). "Large-scale cosmic-ray anisotropies above 4 EeV measured by the Pierre Auger Observatory". The Astrophysical Journal. 868 (1): 4. arXiv: 1808.03579 . Bibcode:2018ApJ...868....4A. doi: 10.3847/1538-4357/aae689 . hdl: 2434/605925 . ISSN   1538-4357. S2CID   84836470.
  9. Abu-Zayyad, T.; et al. (2012-10-11). "The surface detector array of the Telescope Array experiment". Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment. 689: 87–97. arXiv: 1201.4964 . Bibcode:2012NIMPA.689...87A. doi: 10.1016/j.nima.2012.05.079 . ISSN   0168-9002.
  10. Kawai, H.; Yoshida, S.; Yoshii, H.; Tanaka, K.; Cohen, F.; Fukushima, M.; Hayashida, N.; Hiyama, K.; Ikeda, D.; Kido, E.; Kondo, Y. (January 2008). "Telescope Array Experiment". Nuclear Physics B - Proceedings Supplements. 175–176: 221–226. Bibcode:2008NuPhS.175..221K. doi:10.1016/j.nuclphysbps.2007.11.002. S2CID   53604164.
  11. "Telescope Array". www.telescopearray.org. Retrieved 2021-04-28.
  12. Ogio, Shoichi (2018-01-18). "Telescope Array Low energy Extension: TALE". Proceedings of 2016 International Conference on Ultra-High Energy Cosmic Rays (UHECR2016). JPS Conference Proceedings. Vol. 19. Kyoto, Japan: Journal of the Physical Society of Japan. p. 011026. Bibcode:2018uhec.confa1026O. doi: 10.7566/JPSCP.19.011026 . ISBN   978-4-89027-126-9.
  13. The Pierre Auger Collaboration; Aab, A.; Abreu, P.; Aglietta, M.; Albury, J. M.; Allekotte, I.; Almela, A.; Alvarez Castillo, J.; Alvarez-Muñiz, J.; Alves Batista, R.; Anastasi, G. A. (2020-09-16). "Features of the Energy Spectrum of Cosmic Rays above $2.5\ifmmode\times\else\texttimes\fi{}{10}^{18}\text{ }\text{ }\mathrm{eV}$ Using the Pierre Auger Observatory". Physical Review Letters. 125 (12): 121106. arXiv: 2008.06488 . Bibcode:2020PhRvL.125l1106A. doi: 10.1103/PhysRevLett.125.121106 . PMID   33016715.
  14. Abbasi, R.U.; et al. (2005-07-21). "Observation of the ankle and evidence for a high-energy break in the cosmic ray spectrum". Physics Letters B. 619 (3–4): 271–280. arXiv: astro-ph/0501317 . Bibcode:2005PhLB..619..271A. doi: 10.1016/j.physletb.2005.05.064 . ISSN   0370-2693.
  15. HiRes Collaboration (2008-03-10). "First Observation of the Greisen-Zatsepin-Kuzmin Suppression". Physical Review Letters. 100 (10): 101101. arXiv: astro-ph/0703099 . Bibcode:2008PhRvL.100j1101A. doi:10.1103/PhysRevLett.100.101101. ISSN   0031-9007. PMID   18352170. S2CID   118960558.
  16. The Pierre Auger Collaboration (2008-08-04). "Observation of the suppression of the flux of cosmic rays above 4x10^19eV". Physical Review Letters. 101 (6): 061101. arXiv: 0806.4302 . doi:10.1103/PhysRevLett.101.061101. hdl: 2440/47607 . ISSN   0031-9007. PMID   18764444. S2CID   118478479.
  17. Aab, A.; Abreu, P.; Aglietta, M.; Albury, J. M.; Allekotte, I.; Almela, A.; Alvarez Castillo, J.; Alvarez-Muñiz, J.; Alves Batista, R.; Anastasi, G. A.; Anchordoqui, L. (2020-09-16). "Measurement of the cosmic-ray energy spectrum above 2.5 × 10 18 eV using the Pierre Auger Observatory". Physical Review D. 102 (6): 062005. arXiv: 2008.06486 . Bibcode:2020PhRvD.102f2005A. doi: 10.1103/PhysRevD.102.062005 . ISSN   2470-0010.
  18. 1 2 Aab, A.; Abreu, P.; Aglietta, M.; Albury, J. M.; Allekotte, I.; Almela, A.; Alvarez Castillo, J.; Alvarez-Muñiz, J.; Alves Batista, R.; Anastasi, G. A.; Anchordoqui, L. (2020-09-16). "Features of the Energy Spectrum of Cosmic Rays above 2.5 × 10 18 eV Using the Pierre Auger Observatory". Physical Review Letters. 125 (12): 121106. arXiv: 2008.06488 . Bibcode:2020PhRvL.125l1106A. doi: 10.1103/PhysRevLett.125.121106 . ISSN   0031-9007. PMID   33016715.
  19. 1 2 Abbasi, R.; Bellido, J.; Belz, J.; de Souza, V.; Hanlon, W.; Ikeda, D.; Lundquist, J. P.; Sokolsky, P.; Stroman, T.; Tameda, Y.; Tsunesada, Y. (2016-04-06). "Report of the Working Group on the Composition of Ultra High Energy Cosmic Rays". Proceedings of International Symposium for Ultra-High Energy Cosmic Rays (UHECR2014): 010016. arXiv: 1503.07540 . Bibcode:2016uhec.confa0016A. doi: 10.7566/JPSCP.9.010016 . ISBN   978-4-89027-113-9.
  20. 1 2 Aloisio, R.; Berezinsky, V.; Gazizov, A. (December 2012). "Transition from galactic to extragalactic cosmic rays". Astroparticle Physics. 39–40: 129–143. arXiv: 1211.0494 . Bibcode:2012APh....39..129A. doi:10.1016/j.astropartphys.2012.09.007. S2CID   9266571.
  21. De Marco, Daniel; Stanev, Todor (2005-10-13). "On the shape of the UHE cosmic ray spectrum". Physical Review D. 72 (8): 081301. arXiv: astro-ph/0506318 . doi:10.1103/PhysRevD.72.081301. ISSN   1550-7998. S2CID   118149419.
  22. Berezinsky, V.; Gazizov, A.Z.; Grigorieva, S.I. (2005-04-21). "Dip in UHECR spectrum as signature of proton interaction with CMB". Physics Letters B. 612 (3–4): 147–153. arXiv: astro-ph/0502550 . doi: 10.1016/j.physletb.2005.02.058 . ISSN   0370-2693.
  23. Rachen, Joerg P.; Stanev, Todor; Biermann, Peter L. (1993-02-04). "Extragalactic ultra high energy cosmic rays, II. Comparison with experimental data". Astron. Astrophys. 273: 377. arXiv: astro-ph/9302005 . Bibcode:1993A&A...273..377R.
  24. Hanlon, William (2019-08-04). "Telescope Array 10 Year Composition". Proceedings of 36th International Cosmic Ray Conference — PoS(ICRC2019). Vol. 36. p. 280. arXiv: 1908.01356 . Bibcode:2019ICRC...36..280H. doi: 10.22323/1.358.0280 . S2CID   199442393.{{cite book}}: |journal= ignored (help)
  25. Pierre Auger Collaboration (2014-12-31). "Depth of Maximum of Air-Shower Profiles at the Pierre Auger Observatory: Measurements at Energies above 10^17.8 eV". Physical Review D. 90 (12): 122005. arXiv: 1409.4809 . Bibcode:2014PhRvD..90l2005A. doi: 10.1103/PhysRevD.90.122005 . ISSN   1550-7998.
  26. Watson, A A (2014-02-19). "High-energy cosmic rays and the Greisen–Zatsepin–Kuz'min effect". Reports on Progress in Physics. 77 (3): 036901. arXiv: 1310.0325 . Bibcode:2014RPPh...77c6901W. doi:10.1088/0034-4885/77/3/036901. ISSN   0034-4885. PMID   24552650. S2CID   20408181.
  27. Hillas, A. M. (September 1984). "The Origin of Ultra-High-Energy Cosmic Rays". Annual Review of Astronomy and Astrophysics. 22 (1): 425–444. Bibcode:1984ARA&A..22..425H. doi:10.1146/annurev.aa.22.090184.002233. ISSN   0066-4146.
  28. The Pierre Auger Collaboration; Abraham, J.; Abreu, P.; Aglietta, M.; Aguirre, C.; Allard, D.; Allekotte, I.; Allen, J.; Allison, P.; Alvarez, C.; Alvarez-Muniz, J. (2007-11-09). "Correlation of the Highest-Energy Cosmic Rays with Nearby Extragalactic Objects". Science. 318 (5852): 938–943. arXiv: 0711.2256 . Bibcode:2007Sci...318..938P. doi:10.1126/science.1151124. ISSN   0036-8075. PMID   17991855. S2CID   118376969.
  29. The IceCube Collaboration; Fermi-LAT; MAGIC; AGILE; ASAS-SN; HAWC; H.E.S.S.; INTEGRAL; Kanata; Kiso; Kapteyn (2018-07-13). "Multimessenger observations of a flaring blazar coincident with high-energy neutrino IceCube-170922A". Science. 361 (6398): eaat1378. arXiv: 1807.08816 . Bibcode:2018Sci...361.1378I. doi: 10.1126/science.aat1378 . ISSN   0036-8075. PMID   30002226.
  30. Hong, Sungwook E.; Ryu, Dongsu; Kang, Hyesung; Cen, Renyue (2014-04-03). "Shock Waves and Cosmic Ray Acceleration in the Outskirts of Galaxy Clusters". The Astrophysical Journal. 785 (2): 133. arXiv: 1403.1420 . Bibcode:2014ApJ...785..133H. doi:10.1088/0004-637X/785/2/133. ISSN   0004-637X. S2CID   73590389.
  31. Ferrari, C.; Govoni, F.; Schindler, S.; Bykov, A. M.; Rephaeli, Y. (February 2008). "Observations of extended radio emission in clusters". Space Science Reviews. 134 (1–4): 93–118. arXiv: 0801.0985 . Bibcode:2008SSRv..134...93F. doi:10.1007/s11214-008-9311-x. ISSN   0038-6308. S2CID   18428157.
  32. Brunetti, G.; Jones, T. W. (April 2014). "Cosmic rays in galaxy clusters and their non-thermal emission". International Journal of Modern Physics D. 23 (4): 1430007–1430098. arXiv: 1401.7519 . Bibcode:2014IJMPD..2330007B. doi:10.1142/S0218271814300079. ISSN   0218-2718. S2CID   119308380.
  33. Vannoni, G.; Aharonian, F. A.; Gabici, S.; Kelner, S. R.; Prosekin, A. (2011-12-01). "Acceleration and radiation of ultra-high energy protons in galaxy clusters". Astronomy & Astrophysics. 536: A56. arXiv: 0910.5715 . Bibcode:2011A&A...536A..56V. doi: 10.1051/0004-6361/200913568 . ISSN   0004-6361.
  34. Berezinsky, V. S.; Blasi, P.; Ptuskin, V. S. (October 1997). "Clusters of Galaxies as a Storage Room for Cosmic Rays". The Astrophysical Journal. 487 (2): 529–535. arXiv: astro-ph/9609048 . Bibcode:1997ApJ...487..529B. doi:10.1086/304622. ISSN   0004-637X. S2CID   12525472.
  35. Wittor, Denis (May 2021). "On the Challenges of Cosmic-Ray Proton Shock Acceleration in the Intracluster Medium". New Astronomy. 85: 101550. arXiv: 2102.08059 . Bibcode:2021NewA...8501550W. doi:10.1016/j.newast.2020.101550. ISSN   1384-1076. S2CID   229413947.
  36. Waxman, Eli (1995-07-17). "Cosmological Gamma-Ray Bursts and the Highest Energy Cosmic Rays". Physical Review Letters. 75 (3): 386–389. arXiv: astro-ph/9505082 . Bibcode:1995PhRvL..75..386W. doi:10.1103/PhysRevLett.75.386. ISSN   0031-9007. PMID   10060008. S2CID   9827099.
  37. Hjorth, Jens; Bloom, Joshua S. (2012). "The Gamma-Ray Burst - Supernova Connection". In C. Kouveliotou; R. A. M. J. Wijers; S. E. Woosley (eds.). Gamma-Ray Bursts. Cambridge Astrophysics Series. Vol. 51. Cambridge University Press. pp. 169–190. arXiv: 1104.2274 .
  38. IceCube Collaboration; Abbasi, R.; Abdou, Y.; Abu-Zayyad, T.; Ackermann, M.; Adams, J.; Aguilar, J. A.; Ahlers, M.; Altmann, D.; Andeen, K.; Auffenberg, J. (April 2012). "An Absence of Neutrinos Associated with Cosmic Ray Acceleration in Gamma-Ray Bursts". Nature. 484 (7394): 351–354. arXiv: 1204.4219 . doi:10.1038/nature11068. ISSN   0028-0836. PMID   22517161. S2CID   205228690.
  39. Boncioli, Denise; Biehl, Daniel; Winter, Walter (2019-02-14). "On the Common Origin of Cosmic Rays across the Ankle and Diffuse Neutrinos at the Highest Energies from Low-luminosity Gamma-Ray Bursts". The Astrophysical Journal. 872 (1): 110. arXiv: 1808.07481 . Bibcode:2019ApJ...872..110B. doi: 10.3847/1538-4357/aafda7 . ISSN   1538-4357.
  40. Yoshida, Shigeru; Murase, Kohta (2020-10-21). "Constraining photohadronic scenarios for the unified origin of IceCube neutrinos and ultrahigh-energy cosmic rays". Physical Review D. 102 (8): 083023. arXiv: 2007.09276 . Bibcode:2020PhRvD.102h3023Y. doi:10.1103/PhysRevD.102.083023. ISSN   2470-0010. S2CID   220646878.
  41. Samuelsson, Filip; Bégué, Damien; Ryde, Felix; Pe'er, Asaf; Murase, Kohta (2020-10-23). "Constraining Low-luminosity Gamma-Ray Bursts as Ultra-high-energy Cosmic Ray Sources Using GRB 060218 as a Proxy". The Astrophysical Journal. 902 (2): 148. arXiv: 2005.02417 . Bibcode:2020ApJ...902..148S. doi: 10.3847/1538-4357/abb60c . ISSN   1538-4357. S2CID   218516915.
  42. 1 2 Blasi, P.; Epstein, R. I.; Olinto, A. V. (2000-04-20). "Ultra-High Energy Cosmic Rays from Young Neutron Star Winds". The Astrophysical Journal. 533 (2): L123–L126. arXiv: astro-ph/9912240 . Bibcode:2000ApJ...533L.123B. doi:10.1086/312626. PMID   10770705. S2CID   6026463.
  43. Fang, Ke; Kotera, Kumiko; Olinto, Angela V. (April 2012). "Newly Born Pulsars as Sources of Ultrahigh Energy Cosmic Rays". The Astrophysical Journal. 750 (2): 118. arXiv: 1201.5197 . Bibcode:2012ApJ...750..118F. doi:10.1088/0004-637X/750/2/118. ISSN   0004-637X. S2CID   9129110.
  44. Fang, Ke; Kotera, Kumiko; Olinto, Angela V. (2013-03-11). "Ultrahigh Energy Cosmic Ray Nuclei from Extragalactic Pulsars and the effect of their Galactic counterparts". Journal of Cosmology and Astroparticle Physics. 2013 (3): 010. arXiv: 1302.4482 . Bibcode:2013JCAP...03..010F. doi:10.1088/1475-7516/2013/03/010. ISSN   1475-7516. S2CID   118721122.