Geoneutrino

Last updated

A geoneutrino is a neutrino or antineutrino emitted in decay of radionuclide naturally occurring in the Earth. Neutrinos, the lightest of the known subatomic particles, lack measurable electromagnetic properties and interact only via the weak nuclear force when ignoring gravity. Matter is virtually transparent to neutrinos and consequently they travel, unimpeded, at near light speed through the Earth from their point of emission. Collectively, geoneutrinos carry integrated information about the abundances of their radioactive sources inside the Earth. A major objective of the emerging field of neutrino geophysics involves extracting geologically useful information (e.g., abundances of individual geoneutrino-producing elements and their spatial distribution in Earth's interior) from geoneutrino measurements. Analysts from the Borexino collaboration have been able to get to 53 events of neutrinos originating from the interior of the Earth. [1]

Contents

Most geoneutrinos are electron antineutrinos originating in
β
decay branches of 40K, 232Th and 238U. Together these decay chains account for more than 99% of the present-day radiogenic heat generated inside the Earth. Only geoneutrinos from 232Th and 238U decay chains are detectable by the inverse beta-decay mechanism on the free proton because these have energies above the corresponding threshold (1.8 MeV). In neutrino experiments, large underground liquid scintillator detectors record the flashes of light generated from this interaction. As of 2016 geoneutrino measurements at two sites, as reported by the KamLAND and Borexino collaborations, have begun to place constraints on the amount of radiogenic heating in the Earth's interior. A third detector (SNO+) is expected to start collecting data in 2017. JUNO experiment is under construction in Southern China. Another geoneutrino detecting experiment is planned at the China Jinping Underground Laboratory.

History

The Feynman diagram for
b
decay of a neutron into a proton, electron, and electron antineutrino via an intermediate
W
boson. Beta Negative Decay.svg
The Feynman diagram for
β
decay
of a neutron into a proton, electron, and electron antineutrino via an intermediate
W
boson
.

Neutrinos were hypothesized in 1930 by Wolfgang Pauli. The first detection of antineutrinos generated in a nuclear reactor was confirmed in 1956. [2] The idea of studying geologically produced neutrinos to infer Earth's composition has been around since at least mid-1960s. [3] In a 1984 landmark paper Krauss, Glashow & Schramm presented calculations of the predicted geoneutrino flux and discussed the possibilities for detection. [4] First detection of geoneutrinos was reported in 2005 by the KamLAND experiment at the Kamioka Observatory in Japan. [5] [6] In 2010 the Borexino experiment at the Gran Sasso National Laboratory in Italy released their geoneutrino measurement. [7] [8] Updated results from KamLAND were published in 2011 [9] [10] and 2013, [11] and Borexino in 2013 [12] and 2015. [13]

Geological motivation

Geologically significant antineutrino- and heat-producing radioactive decays and decay chains [14]

The Earth's interior radiates heat at a rate of about 47 TW (terawatts), [15] which is less than 0.1% of the incoming solar energy. Part of this heat loss is accounted for by the heat generated upon decay of radioactive isotopes in the Earth interior. The remaining heat loss is due to the secular cooling of the Earth, growth of the Earth's inner core (gravitational energy and latent heat contributions), and other processes. The most important heat-producing elements are uranium (U), thorium (Th), and potassium (K). The debate about their abundances in the Earth has not concluded. Various compositional estimates exist where the total Earth's internal radiogenic heating rate ranges from as low as ~10 TW to as high as ~30 TW. [16] [17] [18] [19] [20] About 7 TW worth of heat-producing elements reside in the Earth's crust, [21] the remaining power is distributed in the Earth mantle; the amount of U, Th, and K in the Earth core is probably negligible. Radioactivity in the Earth mantle provides internal heating to power mantle convection, which is the driver of plate tectonics. The amount of mantle radioactivity and its spatial distribution—is the mantle compositionally uniform at large scale or composed of distinct reservoirs?—is of importance to geophysics.

The existing range of compositional estimates of the Earth reflects our lack of understanding of what were the processes and building blocks (chondritic meteorites) that contributed to its formation. More accurate knowledge of U, Th, and K abundances in the Earth interior would improve our understanding of present-day Earth dynamics and of Earth formation in early Solar System. Counting antineutrinos produced in the Earth can constrain the geological abundance models. The weakly interacting geoneutrinos carry information about their emitters’ abundances and location in the entire Earth volume, including the deep Earth. Extracting compositional information about the Earth mantle from geoneutrino measurements is difficult but possible. It requires a synthesis of geoneutrino experimental data with geochemical and geophysical models of the Earth. Existing geoneutrino data are a byproduct of antineutrino measurements with detectors designed primarily for fundamental neutrino physics research. Future experiments devised with a geophysical agenda in mind would benefit geoscience. Proposals for such detectors have been put forward. [22]

Geoneutrino prediction

Geoneutrino signal prediction at Earth's surface in terrestrial neutrino units (TNU). Geoneutrino signal prediction - rotating globe.gif
Geoneutrino signal prediction at Earth's surface in terrestrial neutrino units (TNU).
The radiogenic heat from the decay of Th (violet) is a major contributor to the earth's internal heat budget. The other major contributors are U (red), U (green), and K (yellow). Evolution of Earth's radiogenic heat.svg
The radiogenic heat from the decay of Th (violet) is a major contributor to the earth's internal heat budget. The other major contributors are U (red), U (green), and K (yellow).

Calculations of the expected geoneutrino signal predicted for various Earth reference models are an essential aspect of neutrino geophysics. In this context, "Earth reference model" means the estimate of heat producing element (U, Th, K) abundances and assumptions about their spatial distribution in the Earth, and a model of Earth's internal density structure. By far the largest variance exists in the abundance models where several estimates have been put forward. They predict a total radiogenic heat production as low as ~10 TW [16] [23] and as high as ~30 TW, [17] the commonly employed value being around 20 TW. [18] [19] [20] A density structure dependent only on the radius (such as the Preliminary Reference Earth Model or PREM) with a 3-D refinement for the emission from the Earth's crust is generally sufficient for geoneutrino predictions.

The geoneutrino signal predictions are crucial for two main reasons: 1) they are used to interpret geoneutrino measurements and test the various proposed Earth compositional models; 2) they can motivate the design of new geoneutrino detectors. The typical geoneutrino flux at Earth's surface is few × 106 cm−2⋅s−1. [24] As a consequence of (i) high enrichment of continental crust in heat producing elements (~7 TW of radiogenic power) and (ii) the dependence of the flux on 1/(distance from point of emission)2, the predicted geoneutrino signal pattern correlates well with the distribution of continents. [25] At continental sites, most geoneutrinos are produced locally in the crust. This calls for an accurate crustal model, both in terms of composition and density, a nontrivial task.

Antineutrino emission from a volume V is calculated for each radionuclide from the following equation:

where dφ(Eν,r)/dEν is the fully oscillated antineutrino flux energy spectrum (in cm−2⋅s−1⋅MeV−1) at position r (units of m) and Eν is the antineutrino energy (in MeV). On the right-hand side, ρ is rock density (in kg⋅m−3), A is elemental abundance (kg of element per kg of rock) and X is the natural isotopic fraction of the radionuclide (isotope/element), M is atomic mass (in g⋅mol−1), NA is the Avogadro constant (in mol−1), λ is decay constant (in s−1), dn(Eν)/dEν is the antineutrino intensity energy spectrum (in MeV−1, normalized to the number of antineutrinos nν produced in a decay chain when integrated over energy), and Pee(Eν,L) is the antineutrino survival probability after traveling a distance L. For an emission domain the size of the Earth, the fully oscillated energy-dependent survival probability Pee can be replaced with a simple factor ⟨Pee⟩ ≈ 0.55, [14] [26] the average survival probability. Integration over the energy yields the total antineutrino flux (in cm−2⋅s−1) from a given radionuclide:

The total geoneutrino flux is the sum of contributions from all antineutrino-producing radionuclides. The geological inputs—the density and particularly the elemental abundances—carry a large uncertainty. The uncertainty of the remaining nuclear and particle physics parameters is negligible compared to the geological inputs. At present it is presumed that uranium-238 and thorium-232 each produce about the same amount of heat in the Earth's mantle, and these are presently the main contributors to radiogenic heat. However, neutrino flux does not perfectly track heat from radioactive decay of primordial nuclides, because neutrinos do not carry off a constant fraction of the energy from the radiogenic decay chains of these primordial radionuclides.

Geoneutrino detection

Detection mechanism

Instruments that measure geoneutrinos are large scintillation detectors. They use the inverse beta decay reaction, a method proposed by Bruno Pontecorvo that Frederick Reines and Clyde Cowan employed in their pioneering experiments in 1950s. Inverse beta decay is a charged current weak interaction, where an electron antineutrino interacts with a proton, producing a positron and a neutron:

Only antineutrinos with energies above the kinematic threshold of 1.806 MeV—the difference between rest mass energies of neutron plus positron and proton—can participate in this interaction. After depositing its kinetic energy, the positron promptly annihilates with an electron:

With a delay of few tens to few hundred microseconds the neutron combines with a proton to form a deuteron:

The two light flashes associated with the positron and the neutron are coincident in time and in space, which provides a powerful method to reject single-flash (non-antineutrino) background events in the liquid scintillator. Antineutrinos produced in man-made nuclear reactors overlap in energy range with geologically produced antineutrinos and are also counted by these detectors. [25]

Because of the kinematic threshold of this antineutrino detection method, only the highest energy geoneutrinos from 232Th and 238U decay chains can be detected. Geoneutrinos from 40K decay have energies below the threshold and cannot be detected using inverse beta decay reaction. Experimental particle physicists are developing other detection methods, which are not limited by an energy threshold (e.g., antineutrino scattering on electrons) and thus would allow detection of geoneutrinos from potassium decay.

Geoneutrino measurements are often reported in Terrestrial Neutrino Units (TNU; analogy with Solar Neutrino Units) rather than in units of flux (cm−2 s−1). TNU is specific to the inverse beta decay detection mechanism with protons. 1 TNU corresponds to 1 geoneutrino event recorded over a year-long fully efficient exposure of 1032 free protons, which is approximately the number of free protons in a 1 kiloton liquid scintillation detector. The conversion between flux units and TNU depends on the thorium to uranium abundance ratio (Th/U) of the emitter. For Th/U=4.0 (a typical value for the Earth), a flux of 1.0 × 106 cm−2 s−1 corresponds to 8.9 TNU. [14]

Detectors and results

Schematic of the KamLAND antineutrino detector. KamLAND schematic.png
Schematic of the KamLAND antineutrino detector.

Existing detectors

KamLAND (Kamioka Liquid Scintillator Antineutrino Detector) is a 1.0 kiloton detector located at the Kamioka Observatory in Japan. Results based on a live-time of 749 days and presented in 2005 mark the first detection of geoneutrinos. The total number of antineutrino events was 152, of which 4.5 to 54.2 were geoneutrinos. This analysis put a 60 TW upper limit on the Earth's radiogenic power from 232Th and 238U. [5]

A 2011 update of KamLAND's result used data from 2135 days of detector time and benefited from improved purity of the scintillator as well as a reduced reactor background from the 21-month-long shutdown of the Kashiwazaki-Kariwa plant after Fukushima. Of 841 candidate antineutrino events, 106 were identified as geoneutrinos using unbinned maximum likelihood analysis. It was found that 232Th and 238U together generate 20.0 TW of radiogenic power. [9]

Borexino is a 0.3 kiloton detector at Laboratori Nazionali del Gran Sasso near L'Aquila, Italy. Results published in 2010 used data collected over live-time of 537 days. Of 15 candidate events, unbinned maximum likelihood analysis identified 9.9 as geoneutrinos. The geoneutrino null hypothesis was rejected at 99.997% confidence level (4.2σ). The data also rejected a hypothesis of an active georeactor in the Earth's core with power above 3 TW at 95% C.L. [7]

A 2013 measurement of 1353 days, detected 46 'golden' anti-neutrino candidates with 14.3±4.4 identified geoneutrinos, indicating a 14.1±8.1 TNU mantle signal, setting a 95% C.L limit of 4.5 TW on geo-reactor power and found the expected reactor signals. [27] In 2015, an updated spectral analysis of geoneutrinos was presented by Borexino based on 2056 days of measurement (from December 2007 to March 2015), with 77 candidate events; of them, only 24 are identified as geonetrinos, and the rest 53 events are originated from European nuclear reactors. The analysis shows that the Earth crust contains about the same amount of U and Th as the mantle, and that the total radiogenic heat flow from these elements and their daughters is 23–36 TW. [28]

SNO+ is a 0.8 kiloton detector located at SNOLAB near Sudbury, Ontario, Canada. SNO+ uses the original SNO experiment chamber. The detector is being refurbished and is expected to operate in late 2016 or 2017. [29]

Planned and proposed detectors

  • Ocean Bottom KamLAND-OBK OBK is a 50 kiloton liquid scintillation detector for deployment in the deep ocean.
  • JUNO (Jiangmen Underground Neutrino Observatory, website) is a 20 kiloton liquid scintillation detector currently under construction in Southern China. The JUNO detector is scheduled to become operational in 2023. [30]
  • Jinping Neutrino Experiment is a 4 kiloton liquid scintillation detector currently under construction in the China JinPing Underground Laboratory (CJPL) scheduled for completion in 2022. [31]
  • LENA (Low Energy Neutrino Astronomy, website) is a proposed 50 kiloton liquid scintillation detector of the LAGUNA project. Proposed sites include Centre for Underground Physics in Pyhäsalmi (CUPP), Finland (preferred) and Laboratoire Souterrain de Modane (LSM) in Fréjus, France. [32] This project seems to be cancelled.
  • at DUSEL (Deep Underground Science and Engineering Laboratory) at Homestake in Lead, South Dakota, USA [33]
  • at BNO (Baksan Neutrino Observatory) in Russia [34]
  • EARTH (Earth AntineutRino TomograpHy)
  • Hanohano (Hawaii Anti-Neutrino Observatory) is a proposed deep-ocean transportable detector. It is the only detector designed to operate away from the Earth's continental crust and from nuclear reactors in order to increase the sensitivity to geoneutrinos from the Earth's mantle. [22]

Desired future technologies

  • Directional antineutrino detection. Resolving the direction from which an antineutrino arrived would help discriminate between the crustal geoneutrino and reactor antineutrino signal (most antineutrinos arriving near horizontally) from mantle geoneutrinos (much wider range of incident dip angles).
  • Detection of antineutrinos from 40K decay. Since the energy spectrum of antineutrinos from 40K decay falls entirely below the threshold energy of inverse beta decay reaction (1.8 MeV), a different detection mechanism must be exploited, such as antineutrino scattering on electrons. Measurement of the abundance of 40K within the Earth would constrain Earth's volatile element budget. [24]

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">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.

<span class="mw-page-title-main">Cowan–Reines neutrino experiment</span> Institute of Technology Experimental confirmation of neutrinos

The Cowan–Reines neutrino experiment was conducted by physicists Clyde Cowan and Frederick Reines in 1956. The experiment confirmed the existence of neutrinos. Neutrinos, subatomic particles with no electric charge and very small mass, had been conjectured to be an essential particle in beta decay processes in the 1930s. With neither mass nor charge, such particles appeared to be impossible to detect. The experiment exploited a huge flux of electron antineutrinos emanating from a nearby nuclear reactor and a detector consisting of large tanks of water. Neutrino interactions with the protons of the water were observed, verifying the existence and basic properties of this particle for the first time.

<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">Kamioka Liquid Scintillator Antineutrino Detector</span> Neutrino oscillation experiment in Japan

The Kamioka Liquid Scintillator Antineutrino Detector (KamLAND) is an electron antineutrino detector at the Kamioka Observatory, an underground neutrino detection facility in Hida, Gifu, Japan. The device is situated in a drift mine shaft in the old KamiokaNDE cavity in the Japanese Alps. The site is surrounded by 53 Japanese commercial nuclear reactors. Nuclear reactors produce electron antineutrinos () during the decay of radioactive fission products in the nuclear fuel. Like the intensity of light from a light bulb or a distant star, the isotropically-emitted flux decreases at 1/R2 per increasing distance R from the reactor. The device is sensitive up to an estimated 25% of antineutrinos from nuclear reactors that exceed the threshold energy of 1.8 megaelectronvolts (MeV) and thus produces a signal in the detector.

<span class="mw-page-title-main">MINOS</span> Particle physics experiment

Main injector neutrino oscillation search (MINOS) was a particle physics experiment designed to study the phenomena of neutrino oscillations, first discovered by a Super-Kamiokande (Super-K) experiment in 1998. Neutrinos produced by the NuMI beamline at Fermilab near Chicago are observed at two detectors, one very close to where the beam is produced, and another much larger detector 735 km away in northern Minnesota.

<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".

Inverse beta decay, commonly abbreviated to IBD, is a nuclear reaction involving an electron antineutrino scattering off a proton, creating a positron and a neutron. This process is commonly used in the detection of electron antineutrinos in neutrino detectors, such as the first detection of antineutrinos in the Cowan–Reines neutrino experiment, or in neutrino experiments such as KamLAND and Borexino. It is an essential process to experiments involving low-energy neutrinos such as those studying neutrino oscillation, reactor neutrinos, sterile neutrinos, and geoneutrinos.

T2K is a particle physics experiment studying the oscillations of the accelerator neutrinos. The experiment is conducted in Japan by the international cooperation of about 500 physicists and engineers with over 60 research institutions from several countries from Europe, Asia and North America and it is a recognized CERN experiment (RE13). T2K collected data within its first phase of operation from 2010 till 2021. The second phase of data taking (T2K-II) is expected to start in 2023 and last until commencement of the successor of T2K – the Hyper-Kamiokande experiment in 2027.

The standard solar model (SSM) is a mathematical treatment of the Sun as a spherical ball of gas. This model, technically the spherically symmetric quasi-static model of a star, has stellar structure described by several differential equations derived from basic physical principles. The model is constrained by boundary conditions, namely the luminosity, radius, age and composition of the Sun, which are well determined. The age of the Sun cannot be measured directly; one way to estimate it is from the age of the oldest meteorites, and models of the evolution of the Solar System. The composition in the photosphere of the modern-day Sun, by mass, is 74.9% hydrogen and 23.8% helium. All heavier elements, called metals in astronomy, account for less than 2 percent of the mass. The SSM is used to test the validity of stellar evolution theory. In fact, the only way to determine the two free parameters of the stellar evolution model, the helium abundance and the mixing length parameter, are to adjust the SSM to "fit" the observed Sun.

<span class="mw-page-title-main">SNO+</span>

SNO+ is a physics experiment designed to search for neutrinoless double beta decay, with secondary measurements of proton–electron–proton (pep) solar neutrinos, geoneutrinos from radioactive decays in the Earth, and reactor neutrinos. It is under construction using the underground equipment already installed for the former Sudbury Neutrino Observatory (SNO) experiment at SNOLAB. It could also observe supernovae neutrinos if a supernova occurs in our galaxy.

<span class="mw-page-title-main">MINERνA</span> Neutrino scattering experiment at Fermilab in Illinois, USA

Main Injector Experiment for ν-A, or MINERνA, is a neutrino scattering experiment which uses the NuMI beamline at Fermilab. MINERνA seeks to measure low energy neutrino interactions both in support of neutrino oscillation experiments and also to study the strong dynamics of the nucleon and nucleus that affect these interactions.

<span class="mw-page-title-main">Borexino</span> Neutrino physics experiment in Italy

Borexino is a deep underground particle physics experiment to study low energy (sub-MeV) solar neutrinos. The detector is the world's most radio-pure liquid scintillator calorimeter and is protected by 3,800 meters of water-equivalent depth. The scintillator is pseudocumene and PPO which is held in place by a thin nylon sphere. It is placed within a stainless steel sphere which holds the photomultiplier tubes (PMTs) used as signal detectors and is shielded by a water tank to protect it against external radiation. Outward pointing PMT's look for any outward facing light flashes to tag incoming cosmic muons that manage to penetrate the overburden of the mountain above. Neutrino energy can be determined through the number of photoelectrons measured in the PMT's. While the position can be determined by extrapolating the difference in arrival times of photons at PMT's throughout the chamber.

<span class="mw-page-title-main">Double Chooz</span>

Double Chooz was a short-baseline neutrino oscillation experiment in Chooz, France. Its goal was to measure or set a limit on the θ13 mixing angle, a neutrino oscillation parameter responsible for changing electron neutrinos into other neutrinos. The experiment uses reactors of the Chooz Nuclear Power Plant as a neutrino source and measures the flux of neutrinos they receive. To accomplish this, Double Chooz has a set of two detectors situated 400 meters and 1050 meters from the reactors. Double Chooz was a successor to the Chooz experiment; one of its detectors occupies the same site as its predecessor. Until January 2015 all data has been collected using only the far detector. The near detector was completed in September 2014, after construction delays, and started taking data at the beginning of 2015. Both detectors stopped taking data in late December 2017.

The Nucifer Experiment is a proposed test of equipment and methodologies for using neutrino detection for the monitoring of nuclear reactor activity and the assessment of the isotopic composition of reactor fuels for non-proliferation treaty compliance monitoring. Based upon an idea proposed by L.A. Mikaélyan in 1977, the Nucifer Experiment was proposed to the IAEA in October 2008.

Measurements of neutrino speed have been conducted as tests of special relativity and for the determination of the mass of neutrinos. Astronomical searches investigate whether light and neutrinos emitted simultaneously from a distant source are arriving simultaneously on Earth. Terrestrial searches include time of flight measurements using synchronized clocks, and direct comparison of neutrino speed with the speed of other particles.

<span class="mw-page-title-main">Earth's internal heat budget</span> Accounting of the energy flows at and below the planets crust

Earth's internal heat budget is fundamental to the thermal history of the Earth. The flow of heat from Earth's interior to the surface is estimated at 47±2 terawatts (TW) and comes from two main sources in roughly equal amounts: the radiogenic heat produced by the radioactive decay of isotopes in the mantle and crust, and the primordial heat left over from the formation of Earth.

<span class="mw-page-title-main">Accelerator Neutrino Neutron Interaction Experiment</span> Water Cherenkov detector experiment

The Accelerator Neutrino Neutron Interaction Experiment (ANNIE) is a proposed water Cherenkov detector experiment designed to examine the nature of neutrino interactions. This experiment will study phenomena like proton decay, and neutrino oscillations, by analyzing neutrino interactions in gadolinium-loaded water and measuring their neutron yield. Neutron Tagging plays an important role in background rejection from atmospheric neutrinos. By implementing early prototypes of LAPPDs, high precision timing is possible. The suggested location for ANNIE is the SciBooNE hall on the Booster Neutrino Beam associated with the MiniBooNE experiment. The neutrino beam originates in Fermilab where The Booster delivers 8 GeV protons to a beryllium target producing secondary pions and kaons. These secondary mesons decay to produce a neutrino beam with an average energy of around 800 MeV. ANNIE will begin installation in the summer of 2015. Phase I of ANNIE, mapping the neutron background, completed in 2017. The detector is being upgraded for full science operation which is expected to begin late 2018.

The diffuse supernova neutrino background(DSNB) is a theoretical population of neutrinos (and anti-neutrinos) cumulatively originating from all core-collapse supernovae events throughout the history of the universe. Though it has not yet been directly detected, the DSNB is theorized to be isotropic and consists of neutrinos with typical energies on the scale of 107 eV. Current detection efforts are limited by the influence of background noise in the search for DSNB neutrinos and are therefore limited to placing limits on the parameters of the DSNB, namely the neutrino flux. Restrictions on these parameters have gotten more strict in recent years, but many researchers are looking to make direct observations in the near future with next generation detectors. The DSNB is not to be confused with the cosmic neutrino background (CNB), which is comprised by relic neutrinos that were produced during the Big Bang and have much lower energies (10−4 to 10−6 eV).

The STEREO experiment investigates the possible oscillation of neutrinos from a nuclear reactor into light so-called sterile neutrinos. It is located at the Institut Laue–Langevin (ILL) in Grenoble, France. The experiment started operating and taking data in November 2016.

References

  1. "Signals from Inside the Earth". Tech Explorist. 2020-01-23. Retrieved 2020-01-23.
  2. Cowan, C. L.; Reines, F.; Harrison, F. B.; Kruse, H. W.; McGuire, A. D. (1956). "Detection of the free neutrino: a confirmation". Science. 124 (3212): 103–662. Bibcode:1956Sci...124..103C. doi:10.1126/science.124.3212.103. PMID   17796274.
  3. Eder, G. (1966). "Terrestrial neutrinos". Nuclear Physics. 78 (3): 657–662. Bibcode:1966NucPh..78..657E. doi:10.1016/0029-5582(66)90903-5.
  4. Krauss, L. M.; Glashow, S. L.; Schramm, D. N. (1984). "Antineutrino astronomy and geophysics". Nature. 310 (5974): 191–198. Bibcode:1984Natur.310..191K. doi:10.1038/310191a0. S2CID   4235872.
  5. 1 2 Araki, T; et al. (2005). "Experimental investigation of geologically produced antineutrinos with KamLAND". Nature. 436 (7050): 499–503. Bibcode:2005Natur.436..499A. doi:10.1038/nature03980. PMID   16049478. S2CID   4367737.
  6. Overbye, D. (July 28, 2005). "Baby Oil and Benzene Provide Look at Earth's Radioactivity". New York Times. Retrieved 9 January 2013.
  7. 1 2 Borexino Collaboration (2010). "Observation of geo-neutrinos". Phys. Lett. B. 687 (4–5): 299–304. arXiv: 1003.0284 . Bibcode:2010PhLB..687..299B. doi:10.1016/j.physletb.2010.03.051.
  8. Edwards, L. (March 16, 2010). "Borexino experiment detects geo-neutrinos". PhysOrg.com. Retrieved 9 January 2013.
  9. 1 2 The KamLAND Collaboration (2011). "Partial radiogenic heat model for Earth revealed by geoneutrino measurements" (PDF). Nature Geoscience. 4 (9): 647–651. Bibcode:2011NatGe...4..647K. doi:10.1038/ngeo1205.
  10. "What Keeps Earth Cooking?". ScienceDaily. July 18, 2011. Retrieved 9 January 2013.
  11. KamLAND Collaboration; Gando, A.; Gando, Y.; Hanakago, H.; Ikeda, H.; Inoue, K.; Ishidoshiro, K.; Ishikawa, H.; Koga, M. (2013-08-02). "Reactor on-off antineutrino measurement with KamLAND". Physical Review D. 88 (3): 033001. arXiv: 1303.4667 . Bibcode:2013PhRvD..88c3001G. doi:10.1103/PhysRevD.88.033001. S2CID   55754667.
  12. Bellini, G.; Benziger, J.; Bick, D.; Bonfini, G.; Bravo, D.; Buizza Avanzini, M.; Caccianiga, B.; Cadonati, L.; Calaprice, F. (2013-05-24). "Measurement of geo-neutrinos from 1353 days of Borexino". Physics Letters B. 722 (4–5): 295–300. arXiv: 1303.2571 . Bibcode:2013PhLB..722..295B. doi:10.1016/j.physletb.2013.04.030. S2CID   55822151.
  13. Borexino Collaboration; Agostini, M.; Appel, S.; Bellini, G.; Benziger, J.; Bick, D.; Bonfini, G.; Bravo, D.; Caccianiga, B. (2015-08-07). "Spectroscopy of geoneutrinos from 2056 days of Borexino data". Physical Review D. 92 (3): 031101. arXiv: 1506.04610 . Bibcode:2015PhRvD..92c1101A. doi:10.1103/PhysRevD.92.031101. S2CID   55041121.
  14. 1 2 3 Dye, S. T. (2012). "Geoneutrinos and the radioactive power of the Earth". Rev. Geophys. 50 (3): RG3007. arXiv: 1111.6099 . Bibcode:2012RvGeo..50.3007D. doi:10.1029/2012RG000400. S2CID   118667366.
  15. Davies, J. H.; Davies, D. R. (2010). "Earth's surface heat flux" (PDF). Solid Earth. 1 (1): 5–24. Bibcode:2010SolE....1....5D. doi: 10.5194/se-1-5-2010 .
  16. 1 2 Javoy, M.; et al. (2010). "The chemical composition of the Earth: Enstatite chondrite models". Earth Planet. Sci. Lett. 293 (3–4): 259–268. Bibcode:2010E&PSL.293..259J. doi:10.1016/j.epsl.2010.02.033.
  17. 1 2 Turcotte, D. L.; Schubert, G. (2002). Geodynamics, Applications of Continuum Physics to Geological Problems. Cambridge University Press. ISBN   978-0521666244.
  18. 1 2 Palme, H.; O'Neill, H. St. C. (2003). "Cosmochemical estimates of mantle composition". Treatise on Geochemistry. 2 (ch. 2.01): 1–38. Bibcode:2003TrGeo...2....1P. doi:10.1016/B0-08-043751-6/02177-0.
  19. 1 2 Hart, S. R.; Zindler, A. (1986). "In search of a bulk-Earth composition". Chem. Geol. 57 (3–4): 247–267. Bibcode:1986ChGeo..57..247H. doi:10.1016/0009-2541(86)90053-7.
  20. 1 2 McDonough, W. F.; Sun, S.-s. (1995). "The composition of the Earth". Chem. Geol. 120 (3–4): 223–253. Bibcode:1995ChGeo.120..223M. doi:10.1016/0009-2541(94)00140-4.
  21. Huang, Y.; Chubakov, V.; Mantovani, M.; Rudnick, R. L.; McDonough, W. F. (2013). "A reference Earth model for the heat producing elements and associated geoneutrino flux". arXiv: 1301.0365 [physics.geo-ph].
  22. 1 2 Learned, J. G.; Dye, S. T.; Pakvasa, S. (2008). "Hanohano: A Deep Ocean Anti-Neutrino Detector for Unique Neutrino Physics and Geophysics Studies". Proceedings of the Twelfth International Workshop on Neutrino Telescopes, Venice, March 2007. arXiv: 0810.4975 . Bibcode:2008arXiv0810.4975L.
  23. O'Neill, H. St. C.; Palme, H. (2008). "Collisional erosion and the non-chondritic composition of the terrestrial planets". Phil. Trans. R. Soc. Lond. A. 366 (1883): 4205–4238. Bibcode:2008RSPTA.366.4205O. doi:10.1098/rsta.2008.0111. PMID   18826927. S2CID   14526775.
  24. 1 2 Bellini, G.; Ianni, A.; Ludhova, L.; Mantovani, F.; McDonough, W. F. (2013-11-01). "Geo-neutrinos". Progress in Particle and Nuclear Physics. 73: 1–34. arXiv: 1310.3732 . Bibcode:2013PrPNP..73....1B. doi:10.1016/j.ppnp.2013.07.001. S2CID   237116200.
  25. 1 2 Usman, S.; et al. (2015). "AGM2015: Antineutrino Global Map". Scientific Reports. 5: 13945. arXiv: 1509.03898 . Bibcode:2015NatSR...513945U. doi:10.1038/srep13945. PMC   4555106 . PMID   26323507.
  26. Fiorentini, G; Fogli, G. L.; Lisi, E.; Mantovani, F.; Rotunno, A. M. (2012). "Mantle geoneutrinos in KamLAND and Borexino". Phys. Rev. D. 86 (3): 033004. arXiv: 1204.1923 . Bibcode:2012PhRvD..86c3004F. doi:10.1103/PhysRevD.86.033004. S2CID   118437963.
  27. Borexino Collaboration (24 May 2013). "Measurement of geo-neutrinos from 1353 days of Borexino". Physics Letters B. 722 (4–5): 295–300. arXiv: 1303.2571 . Bibcode:2013PhLB..722..295B. doi:10.1016/j.physletb.2013.04.030. S2CID   55822151.
  28. Borexino Collaboration (7 August 2015). "Spectroscopy of geoneutrinos from 2056 days of Borexino data". Phys. Rev. D. 92 (3): 031101. arXiv: 1506.04610 . Bibcode:2015PhRvD..92c1101A. doi:10.1103/PhysRevD.92.031101. S2CID   55041121.
  29. Andringa, S.; et al. (SNO+ Collaboration) (2015-11-13). "Current Status and Future Prospects of the SNO+ Experiment". Advances in High Energy Physics. 2016: 6194250. arXiv: 1508.05759 . doi: 10.1155/2016/6194250 . S2CID   10721441.
  30. JUNO website, 2022-07-23
  31. Beacom, John F.; Chen, Shaomin; Cheng, Jianping; Doustimotlagh, Sayed N.; Gao, Yuanning; Ge, Shao-Feng; Gong, Guanghua; Gong, Hui; Guo, Lei (2016-02-04). "Letter of Intent: Jinping Neutrino Experiment". Chinese Physics C. 41 (2): 023002. arXiv: 1602.01733 . Bibcode:2017ChPhC..41b3002B. doi:10.1088/1674-1137/41/2/023002. S2CID   197514524.
  32. Wurm, M.; et al. (2012). "The next-generation liquid-scintillator neutrino observatory LENA". Astroparticle Physics. 35 (11): 685–732. arXiv: 1104.5620 . Bibcode:2012APh....35..685W. doi:10.1016/j.astropartphys.2012.02.011. S2CID   118456549.
  33. Tolich, N.; et al. (2006). "A Geoneutrino Experiment at Homestake". Earth, Moon, and Planets. 99 (1): 229–240. arXiv: physics/0607230 . Bibcode:2006EM&P...99..229T. doi:10.1007/s11038-006-9112-8. S2CID   54889933.
  34. Barabanov, I. R.; Novikova, G. Ya.; Sinev, V. V.; Yanovich, E. A. (2009). "Research of the natural neutrino fluxes by use of large volume scintillation detector at Baksan". arXiv: 0908.1466 [hep-ph].

Further reading