Double beta decay

Last updated

In nuclear physics, double beta decay is a type of radioactive decay in which two neutrons are simultaneously transformed into two protons, or vice versa, inside an atomic nucleus. As in single beta decay, this process allows the atom to move closer to the optimal ratio of protons and neutrons. As a result of this transformation, the nucleus emits two detectable beta particles, which are electrons or positrons.

Contents

The literature distinguishes between two types of double beta decay: ordinary double beta decay and neutrinoless double beta decay. In ordinary double beta decay, which has been observed in several isotopes, two electrons and two electron antineutrinos are emitted from the decaying nucleus. In neutrinoless double beta decay, a hypothesized process that has never been observed, only electrons would be emitted.

History

The idea of double beta decay was first proposed by Maria Goeppert Mayer in 1935. [1] [2] In 1937, Ettore Majorana demonstrated that all results of beta decay theory remain unchanged if the neutrino were its own antiparticle, now known as a Majorana particle . [3] In 1939, Wendell H. Furry proposed that if neutrinos are Majorana particles, then double beta decay can proceed without the emission of any neutrinos, via the process now called neutrinoless double beta decay . [4] It is not yet known whether the neutrino is a Majorana particle, and, relatedly, whether neutrinoless double beta decay exists in nature. [5]

As parity violation in weak interactions would not be discovered until 1956, earlier calculations showed that neutrinoless double beta decay should be much more likely to occur than ordinary double beta decay, if neutrinos were Majorana particles. The predicted half-lives were on the order of 1015~1016 years. [5] Efforts to observe the process in laboratory date back to at least 1948 when E.L. Fireman made the first attempt to directly measure the half-life of the 124
Sn
isotope with a Geiger counter. [6] Radiometric experiments through about 1960 produced negative results or false positives, not confirmed by later experiments. In 1950, for the first time the double beta decay half-life of 130
Te
was measured by geochemical methods to be 1.4×1021 years, [7] reasonably close to the modern value. This involved detecting the concentration in minerals of the xenon produced by the decay.

In 1956, after the V − A nature of weak interactions was established, it became clear that the half-life of neutrinoless double beta decay would significantly exceed that of ordinary double beta decay. Despite significant progress in experimental techniques in 1960–1970s, double beta decay was not observed in a laboratory until the 1980s. Experiments had only been able to establish the lower bound for the half-life – about 1021 years. At the same time, geochemical experiments detected the double beta decay of 82
Se
and 128
Te
. [5]

Double beta decay was first observed in a laboratory in 1987 by the group of Michael Moe at UC Irvine in 82
Se
. [8] Since then, many experiments have observed ordinary double beta decay in other isotopes. None of those experiments have produced positive results for the neutrinoless process, raising the half-life lower bound to approximately 1025 years. Geochemical experiments continued through the 1990s, producing positive results for several isotopes. [5] Double beta decay is the rarest known kind of radioactive decay; as of 2019 it has been observed in only 14 isotopes (including double electron capture in 130
Ba
observed in 2001, 78
Kr
observed in 2013, and 124
Xe
observed in 2019), and all have a mean lifetime over 1018 yr (table below). [5]

Ordinary double beta decay

In a typical double beta decay, two neutrons in the nucleus are converted to protons, and two electrons and two electron antineutrinos are emitted. The process can be thought as two simultaneous beta minus decays. In order for (double) beta decay to be possible, the final nucleus must have a larger binding energy than the original nucleus. For some nuclei, such as germanium-76, the isobar one atomic number higher (arsenic-76) has a smaller binding energy, preventing single beta decay. However, the isobar with atomic number two higher, selenium-76, has a larger binding energy, so double beta decay is allowed.

The emission spectrum of the two electrons can be computed in a similar way to beta emission spectrum using Fermi's golden rule. The differential rate is given by

where the subscripts refer to each electron, T is kinetic energy, w is total energy, F(Z, T) is the Fermi function with Z the charge of the final-state nucleus, p is momentum, v is velocity in units of c, is the angle between the electrons, and Q is the Q value of the decay.

For some nuclei, the process occurs as conversion of two protons to neutrons, emitting two electron neutrinos and absorbing two orbital electrons (double electron capture). If the mass difference between the parent and daughter atoms is more than 1.022 MeV/c2 (two electron masses), another decay is accessible, capture of one orbital electron and emission of one positron. When the mass difference is more than 2.044 MeV/c2 (four electron masses), emission of two positrons is possible. These theoretical decay branches have not been observed.

Known double beta decay isotopes

There are 35 naturally occurring isotopes capable of double beta decay. [9] In practice, the decay can be observed when the single beta decay is forbidden by energy conservation. This happens for elements with an even atomic number and even neutron number, which are more stable due to spin-coupling. When single beta decay or alpha decay also occur, the double beta decay rate is generally too low to observe. However, the double beta decay of 238
U
(also an alpha emitter) has been measured radiochemically. Two other nuclides in which double beta decay has been observed, 48
Ca
and 96
Zr
, can also theoretically single beta decay, but this decay is extremely suppressed and has never been observed. Similar suppression of energetically barely possible single beta decay occurs for 148Gd and 222Rn, but both these nuclides are rather short-lived alpha emitters.

Fourteen isotopes have been experimentally observed undergoing two-neutrino double beta decay (ββ) or double electron capture (εε). [10] The table below contains nuclides with the latest experimentally measured half-lives, as of December 2016, except for 124Xe (for which double electron capture was first observed in 2019). Where two uncertainties are specified, the first one is statistical uncertainty and the second is systematic.

NuclideHalf-life, 1021 yearsModeTransitionMethodExperiment
48
Ca
0.064+0.007
0.006
± +0.012
0.009
ββdirect NEMO-3 [11]
76
Ge
1.926 ±0.094ββdirect GERDA [10]
78
Kr
9.2 +5.5
2.6
±1.3
εε direct BAKSAN [10]
82
Se
0.096 ± 0.003 ± 0.010ββdirectNEMO-3 [10]
96
Zr
0.0235 ± 0.0014 ± 0.0016ββdirectNEMO-3 [10]
100
Mo
0.00693 ± 0.00004ββdirectNEMO-3 [10]
0.69+0.10
0.08
± 0.07
ββ0+→ 0+1Ge coincidence [10]
116
Cd
0.028 ± 0.001 ± 0.003
0.026+0.009
0.005
ββdirectNEMO-3 [10]
ELEGANT IV [10]
128
Te
7200 ± 400
1800 ± 700
ββgeochemical [10]
130
Te
0.82 ± 0.02 ± 0.06ββdirect CUORE-0 [12]
124
Xe
18 ± 5 ± 1 εε direct XENON1T [13]
136
Xe
2.165 ± 0.016 ± 0.059ββdirect EXO-200 [10]
130
Ba
(0.5 – 2.7) εε geochemical [14] [15]
150
Nd
0.00911+0.00025
0.00022
± 0.00063
ββdirectNEMO-3 [10]
0.107+0.046
0.026
ββ0+→ 0+1Ge coincidence [10]
238
U
2.0 ± 0.6ββradiochemical [10]

Searches for double beta decay in isotopes that present significantly greater experimental challenges are ongoing. One such isotope is 134
Xe
. [16]

The following known beta-stable (or almost beta-stable in the cases 48Ca, 96Zr, and 222Rn) [17] nuclides with A ≤ 260 are theoretically capable of double beta decay, where red are isotopes that have a double-beta rate measured experimentally and black have yet to be measured experimentally: 46Ca, 48Ca, 70Zn, 76Ge, 80Se, 82Se, 86Kr, 94Zr, 96Zr, 98Mo, 100Mo, 104Ru, 110Pd, 114Cd, 116Cd, 122Sn, 124Sn, 128Te, 130Te, 134Xe, 136Xe, 142Ce, 146Nd, 148Nd, 150Nd, 154Sm, 160Gd, 170Er, 176Yb, 186W, 192Os, 198Pt, 204Hg, 216Po, 220Rn, 222Rn, 226Ra, 232Th, 238U, 244Pu, 248Cm, 254Cf, 256Cf, and 260Fm. [9]

The following known beta-stable (or almost beta-stable in the case 148Gd) nuclides with A ≤ 260 are theoretically capable of double electron capture, where red are isotopes that have a double-electron capture rate measured and black have yet to be measured experimentally: 36Ar, 40Ca, 50Cr, 54Fe, 58Ni, 64Zn, 74Se, 78Kr, 84Sr, 92Mo, 96Ru, 102Pd, 106Cd, 108Cd, 112Sn, 120Te, 124Xe, 126Xe, 130Ba, 132Ba, 136Ce, 138Ce, 144Sm, 148Gd, 150Gd, 152Gd, 154Dy, 156Dy, 158Dy, 162Er, 164Er, 168Yb, 174Hf, 180W, 184Os, 190Pt, 196Hg, 212Rn, 214Rn, 218Ra, 224Th, 230U, 236Pu, 242Cm, 252Fm, and 258No. [9]

In particular, 36Ar is the lightest observationally stable nuclide whose decay is energetically possible.

Neutrinoless double beta decay

Feynman diagram of neutrinoless double beta decay, with two neutrons decaying to two protons. The only emitted products in this process are two electrons, which can occur if the neutrino and antineutrino are the same particle (i.e. Majorana neutrinos) so the same neutrino can be emitted and absorbed within the nucleus. In conventional double beta decay, two antineutrinos -- one arising from each W vertex -- are emitted from the nucleus, in addition to the two electrons. The detection of neutrinoless double beta decay is thus a sensitive test of whether neutrinos are Majorana particles. Double beta decay feynman.svg
Feynman diagram of neutrinoless double beta decay, with two neutrons decaying to two protons. The only emitted products in this process are two electrons, which can occur if the neutrino and antineutrino are the same particle (i.e. Majorana neutrinos) so the same neutrino can be emitted and absorbed within the nucleus. In conventional double beta decay, two antineutrinos — one arising from each W vertex — are emitted from the nucleus, in addition to the two electrons. The detection of neutrinoless double beta decay is thus a sensitive test of whether neutrinos are Majorana particles.

If the neutrino is a Majorana particle (i.e., the antineutrino and the neutrino are actually the same particle), and at least one type of neutrino has non-zero mass (which has been established by the neutrino oscillation experiments), then it is possible for neutrinoless double beta decay to occur. Neutrinoless double beta decay is a lepton number violating process. In the simplest theoretical treatment, known as light neutrino exchange, a nucleon absorbs the neutrino emitted by another nucleon. The exchanged neutrinos are virtual particles.

With only two electrons in the final state, the electrons' total kinetic energy would be approximately the binding energy difference of the initial and final nuclei, with the nuclear recoil accounting for the rest. Because of momentum conservation, electrons are generally emitted back-to-back. The decay rate for this process is given by

where G is the two-body phase-space factor, M is the nuclear matrix element, and mββ is the effective Majorana mass of the electron neutrino. In the context of light Majorana neutrino exchange, mββ is given by

where mi are the neutrino masses and the Uei are elements of the Pontecorvo–Maki–Nakagawa–Sakata (PMNS) matrix. Therefore, observing neutrinoless double beta decay, in addition to confirming the Majorana neutrino nature, can give information on the absolute neutrino mass scale and Majorana phases in the PMNS matrix, subject to interpretation through theoretical models of the nucleus, which determine the nuclear matrix elements, and models of the decay. [18] [19]

The observation of neutrinoless double beta decay would require that at least one neutrino is a Majorana particle, irrespective of whether the process is engendered by neutrino exchange. [20]

Experiments

Numerous experiments have searched for neutrinoless double beta decay. The best-performing experiments have a high mass of the decaying isotope and low backgrounds, with some experiments able to perform particle discrimination and electron tracking. In order to remove backgrounds from cosmic rays, most experiments are located in underground laboratories around the world.

Recent and proposed experiments include:

Status

While some experiments have claimed a discovery of neutrinoless double beta decay, modern searches have found no evidence for the decay.

Heidelberg-Moscow Controversy

Some members of the Heidelberg-Moscow collaboration claimed a detection of neutrinoless beta decay in 76Ge in 2001. [26] This claim was criticized by outside physicists [1] [27] [28] [29] as well as other members of the collaboration. [30] In 2006, a refined estimate by the same authors stated the half-life was 2.3×1025 years. [31] This half-life has been excluded at high confidence by other experiments, including in 76Ge by GERDA. [32]

Current results

As of 2017, the strongest limits on neutrinoless double beta decay have come from GERDA in 76Ge, CUORE in 130Te, and EXO-200 and KamLAND-Zen in 136Xe.

Higher order simultaneous beta decay

For mass numbers with more than two beta-stable isobars, quadruple beta decay and its inverse, quadruple electron capture, have been proposed as alternatives to double beta decay in the isobars with the greatest energy excess. These decays are energetically possible in eight nuclei, though partial half-lives compared to single or double beta decay are predicted to be very long; hence, quadruple beta decay is unlikely to be observed. The seven candidate nuclei for quadruple beta decay include 96Zr, 136Xe, and 150Nd capable of quadruple beta-minus decay, and 124Xe, 130Ba, 148Gd, and 154Dy capable of quadruple beta-plus decay or electron capture (though 148Gd and 154Dy are non-primordial alpha-emitters with geologically short half-lives). In theory, quadruple beta decay may be experimentally observable in three of these nuclei – 96Zr, 136Xe, and 150Nd – with the most promising candidate being 150Nd. Triple beta decay is also possible for 48Ca, 96Zr, and 150Nd. [33]

Moreover, such a decay mode could also be neutrinoless in physics beyond the standard model. [34] Neutrinoless quadruple beta decay would violate lepton number in 4 units, as opposed to a lepton number breaking of two units in the case of neutrinoless double beta decay. Therefore, there is no 'black-box theorem' and neutrinos could be Dirac particles while allowing these type of processes. In particular, if neutrinoless quadruple beta decay is found before neutrinoless double beta decay then the expectation is that neutrinos will be Dirac particles. [35]

So far, searches for triple and quadruple beta decay in 150Nd have remained unsuccessful. [33]

See also

Related Research Articles

<span class="mw-page-title-main">Beta decay</span> Type of radioactive decay

In nuclear physics, beta decay (β-decay) is a type of radioactive decay in which an atomic nucleus emits a beta particle, transforming into an isobar of that nuclide. For example, beta decay of a neutron transforms it into a proton by the emission of an electron accompanied by an antineutrino; or, conversely a proton is converted into a neutron by the emission of a positron with a neutrino in so-called positron emission. Neither the beta particle nor its associated (anti-)neutrino exist within the nucleus prior to beta decay, but are created in the decay process. By this process, unstable atoms obtain a more stable ratio of protons to neutrons. The probability of a nuclide decaying due to beta and other forms of decay is determined by its nuclear binding energy. The binding energies of all existing nuclides form what is called the nuclear band or valley of stability. For either electron or positron emission to be energetically possible, the energy release or Q value must be positive.

<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.

In particle physics, majorons are a hypothetical type of Goldstone boson that are conjectured to mediate the neutrino mass violation of lepton number or BL in certain high energy collisions such as

<span class="mw-page-title-main">Neutrinoless double beta decay</span> A nuclear physics process that has yet to observed

Neutrinoless double beta decay (0νββ) is a commonly proposed and experimentally pursued theoretical radioactive decay process that would prove a Majorana nature of the neutrino particle. To this day, it has not been found.

<span class="mw-page-title-main">Double electron capture</span> Mode of radioactive decay

Double electron capture is a decay mode of an atomic nucleus. For a nuclide (A, Z) with a number of nucleons A and atomic number Z, double electron capture is only possible if the mass of the nuclide (A, Z−2) is lower.

<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.

Sterile neutrinos are hypothetical particles that interact only via gravity and not via any of the other fundamental interactions of the Standard Model. The term sterile neutrino is used to distinguish them from the known, ordinary active neutrinos in the Standard Model, which carry an isospin charge of ±+1/ 2  and engage in the weak interaction. The term typically refers to neutrinos with right-handed chirality, which may be inserted into the Standard Model. Particles that possess the quantum numbers of sterile neutrinos and masses great enough such that they do not interfere with the current theory of Big Bang nucleosynthesis are often called neutral heavy leptons (NHLs) or heavy neutral leptons (HNLs).

There are 39 known isotopes and 17 nuclear isomers of tellurium (52Te), with atomic masses that range from 104 to 142. These are listed in the table below.

Germanium (32Ge) has five naturally occurring isotopes, 70Ge, 72Ge, 73Ge, 74Ge, and 76Ge. Of these, 76Ge is very slightly radioactive, decaying by double beta decay with a half-life of 1.78 × 1021 years (130 billion times the age of the universe).

<span class="mw-page-title-main">Majorana fermion</span> Fermion that is its own antiparticle

A Majorana fermion, also referred to as a Majorana particle, is a fermion that is its own antiparticle. They were hypothesised by Ettore Majorana in 1937. The term is sometimes used in opposition to a Dirac fermion, which describes fermions that are not their own antiparticles.

The Enriched Xenon Observatory (EXO) is a particle physics experiment searching for neutrinoless double beta decay of xenon-136 at WIPP near Carlsbad, New Mexico, U.S.

The Neutrino Ettore Majorana Observatory is an international collaboration of scientists searching for neutrinoless double beta decay (0νββ). The collaboration has been active since 1989. Observation of 0νββ would indicate neutrinos are Majorana particles and could be used to measure the neutrino mass. It is located in the Modane Underground Laboratory (LSM) in the Fréjus Road Tunnel. The experiment has had 3 detectors, NEMO-1, NEMO-2, NEMO-3 and is planning to construct a new detector SuperNEMO. The NEMO-1 and NEMO-2 prototype detectors were used until 1997. Latest experiment NEMO-3 was under design and construction from 1994 onwards, took data from January 2003 to January 2011 and the final data analysis was published in 2018. The NEMO-2 and NEMO-3 detectors produced measurements for double neutrino decays and limits for neutrinoless double-beta decay for a number of elements, such as molybdenum-100 and selenium-82. These double beta decay times are important contributions to understanding the nucleus and are needed inputs for neutrinoless decay studies, which constrain neutrino mass.

<span class="mw-page-title-main">Nuclear drip line</span> Atomic nuclei decay delimiter

The nuclear drip line is the boundary beyond which atomic nuclei are unbound with respect to the emission of a proton or neutron.

Hans Volker Klapdor-Kleingrothaus is a German physicist who works in nuclear physics, particle physics and astrophysics.

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

The MAJORANA project is an international effort to search for neutrinoless double-beta (0νββ) decay in 76Ge. The project builds upon the work of previous experiments, notably those performed by the Heidelberg–Moscow and IGEX collaborations, which used high-purity germanium (HPGe) detectors, to study neutrinoless double-beta decay.

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

The Cryogenic Underground Observatory for Rare Events (CUORE) – also cuore (Italian for 'heart'; ) – is a particle physics facility located underground at the Laboratori Nazionali del Gran Sasso in Assergi, Italy. CUORE was designed primarily as a search for neutrinoless double beta decay in 130Te, a process that has never been observed. It uses tellurium dioxide (TeO2) crystals as both the source of the decay and as bolometers to detect the resulting electrons. CUORE searches for the characteristic signal of neutrinoless double beta decay, a small peak in the observed energy spectrum around the known decay energy; for 130Te, this is Q = 2527.518 ± 0.013 keV. CUORE can also search for signals from dark matter candidates, such as axions and WIMPs.

The Germanium Detector Array experiment was searching for neutrinoless double beta decay (0νββ) in Ge-76 at the underground Laboratori Nazionali del Gran Sasso (LNGS). Neutrinoless beta decay is expected to be a very rare process if it occurs. The collaboration predicted less than one event each year per kilogram of material, appearing as a narrow spike around the 0νββ Q-value in the observed energy spectrum. This meant background shielding was required to detect any rare decays. The LNGS facility has 1400 meters of rock overburden, equivalent to 3000 meters of water shielding, reducing cosmic radiation background. The GERDA experiment was operated from 2011 onwards at LNGS.

The Cadmium Zinc Telluride 0-Neutrino Double-Beta (COBRA) experiment is a large array of cadmium zinc telluride (CdZnTe) semiconductors searching for evidence of neutrinoless double beta decay and to measure its half-life. COBRA is located underground, within the Gran Sasso National Laboratory. The experiment was proposed in 2001, and installation of a large prototype began in 2006.

<span class="mw-page-title-main">Yuri G. Zdesenko</span>

Yuri G. Zdesenko ; 6 October 1943 – 1 September 2004, was a Ukrainian nuclear physicist known for a significant contribution to investigations of double beta decay.

Ettore Fiorini was an Italian experimental particle physicist. He studied the physics of the weak interaction and was a pioneer in the field of double beta decay. He served as a professor of nuclear and subnuclear physics at the University of Milano-Bicocca.

References

  1. 1 2 Giuliani, A.; Poves, A. (2012). "Neutrinoless double-beta decay" (PDF). Advances in High Energy Physics . 2012: 1–38. doi: 10.1155/2012/857016 .
  2. Goeppert-Mayer, M. (1935). "Double beta-disintegration". Physical Review . 48 (6): 512–516. Bibcode:1935PhRv...48..512G. doi:10.1103/PhysRev.48.512.
  3. Majorana, E. (1937). "Teoria simmetrica dell'elettrone e del positrone". Il Nuovo Cimento (in Italian). 14 (4): 171–184. Bibcode:1937NCim...14..171M. doi:10.1007/BF02961314. S2CID   18973190.
  4. Furry, W.H. (1939). "On Transition Probabilities in Double Beta-Disintegration". Physical Review . 56 (12): 1184–1193. Bibcode:1939PhRv...56.1184F. doi:10.1103/PhysRev.56.1184.
  5. 1 2 3 4 5 Barabash, A.S. (2011). "Experiment double beta decay: Historical review of 75 years of research". Physics of Atomic Nuclei. 74 (4): 603–613. arXiv: 1104.2714 . Bibcode:2011PAN....74..603B. doi:10.1134/S1063778811030070. S2CID   118716672.
  6. Fireman, E. (1948). "Double beta decay". Physical Review . 74 (9): 1201–1253. Bibcode:1948PhRv...74.1201.. doi:10.1103/PhysRev.74.1201.
  7. Inghram, M.G.; Reynolds, J.H. (1950). "Double Beta-Decay of 130Te". Physical Review . 78 (6): 822–823. Bibcode:1950PhRv...78..822I. doi:10.1103/PhysRev.78.822.2.
  8. Elliott, S. R.; Hahn, A. A.; Moe; M. K. (1987). "Direct evidence for two-neutrino double-beta decay in 82Se". Physical Review Letters . 59 (18): 2020–2023. Bibcode:1987PhRvL..59.2020E. doi:10.1103/PhysRevLett.59.2020. PMID   10035397.
  9. 1 2 3 Tretyak, V.I.; Zdesenko, Yu.G. (2002). "Tables of Double Beta Decay Data — An Update". At. Data Nucl. Data Tables . 80 (1): 83–116. Bibcode:2002ADNDT..80...83T. doi:10.1006/adnd.2001.0873.
  10. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 Patrignani, C.; et al. (Particle Data Group) (2016). "Review of Particle Physics" (PDF). Chinese Physics C . 40 (10): 100001. Bibcode:2016ChPhC..40j0001P. doi:10.1088/1674-1137/40/10/100001. S2CID   125766528. See p. 768
  11. Arnold, R.; et al. (NEMO-3 Collaboration) (2016). "Measurement of the double-beta decay half-life and search for the neutrinoless double-beta decay of 48Ca with the NEMO-3 detector". Physical Review D . 93 (11): 112008. arXiv: 1604.01710 . Bibcode:2016PhRvD..93k2008A. doi:10.1103/PhysRevD.93.112008. S2CID   55485404.
  12. Alduino, C.; et al. (CUORE-0 Collaboration) (2016). "Measurement of the Two-Neutrino Double Beta Decay Half-life of 130Te with the CUORE-0 Experiment". The European Physical Journal C. 77 (1): 13. arXiv: 1609.01666 . Bibcode:2017EPJC...77...13A. doi:10.1140/epjc/s10052-016-4498-6. S2CID   73575079.
  13. Aprile, E.; et al. (2019). "Observation of two-neutrino double electron capture in 124Xe with XENON1T". Nature. 568 (7753): 532–535. arXiv: 1904.11002 . Bibcode:2019Natur.568..532X. doi:10.1038/s41586-019-1124-4. PMID   31019319. S2CID   129948831.
  14. A. P. Meshik; C. M. Hohenberg; O. V. Pravdivtseva; Ya. S. Kapusta (2001). "Weak decay of 130Ba and 132Ba: Geochemical measurements". Physical Review C . 64 (3): 035205 [6 pages]. Bibcode:2001PhRvC..64c5205M. doi:10.1103/PhysRevC.64.035205.
  15. M. Pujol; B. Marty; P. Burnard; P. Philippot (2009). "Xenon in Archean barite: Weak decay of 130Ba, mass-dependent isotopic fractionation and implication for barite formation". Geochimica et Cosmochimica Acta . 73 (22): 6834–6846. Bibcode:2009GeCoA..73.6834P. doi:10.1016/j.gca.2009.08.002.
  16. Albert, J. B.; et al. (EXO-200 Collaboration) (3 November 2017). "Searches for Double Beta Decay of 134Xe with EXO-200". Physical Review D. 96 (9): 092001. arXiv: 1704.05042 . Bibcode:2017PhRvD..96i2001A. doi:10.1103/PhysRevD.96.092001. S2CID   28537166.
  17. Belli, P.; Bernabei, R.; Danevich, F. A.; et al. (2019). "Experimental searches for rare alpha and beta decays". European Physical Journal A. 55 (8): 140–1–140–7. arXiv: 1908.11458 . Bibcode:2019EPJA...55..140B. doi:10.1140/epja/i2019-12823-2. ISSN   1434-601X. S2CID   201664098.
  18. Grotz, K.; Klapdor, H. V. (1990). The Weak Interaction in Nuclear, Particle and Astrophysics . CRC Press. ISBN   978-0-85274-313-3.
  19. Klapdor-Kleingrothaus, H. V.; Staudt, A. (1998). Non-accelerator Particle Physics (PDF) (Reprint ed.). IOP Publishing. ISBN   978-0-7503-0305-7.
  20. Schechter, J.; Valle, J. W. F. (1982). "Neutrinoless double-β decay in SU(2)×U(1) theories". Physical Review D . 25 (11): 2951–2954. Bibcode:1982PhRvD..25.2951S. doi:10.1103/PhysRevD.25.2951. hdl: 10550/47205 .
  21. Aalseth, C. E.; et al. (2000). "Recent Results of the IGEX 76Ge Double-Beta Decay Experiment". Physics of Atomic Nuclei . 63 (7): 1225–1228. Bibcode:2000PAN....63.1225A. doi:10.1134/1.855774. S2CID   123335600.
  22. 1 2 Schwingenheuer, B. (2013). "Status and prospects of searches for neutrinoless double beta decay". Annalen der Physik . 525 (4): 269–280. arXiv: 1210.7432 . Bibcode:2013AnP...525..269S. CiteSeerX   10.1.1.760.5635 . doi:10.1002/andp.201200222. S2CID   117129820.
  23. Xu, W.; et al. (2015). "The Majorana Demonstrator: A Search for Neutrinoless Double-beta Decay of 76Ge". Journal of Physics: Conference Series . 606 (1): 012004. arXiv: 1501.03089 . Bibcode:2015JPhCS.606a2004X. doi:10.1088/1742-6596/606/1/012004. S2CID   119301804.
  24. Khanbekov, N. D. (2013). "AMoRE: Collaboration for searches for the neutrinoless double-beta decay of the isotope of 100Mo with the aid of 40Ca100MoO4 as a cryogenic scintillation detector". Physics of Atomic Nuclei . 76 (9): 1086–1089. Bibcode:2013PAN....76.1086K. doi:10.1134/S1063778813090093. S2CID   123287005.
  25. Albert, J. B.; et al. (nEXO Collaboration) (2018). "Sensitivity and Discovery Potential of nEXO to Neutrinoless Double Beta Decay". Physical Review C. 97 (6): 065503. arXiv: 1710.05075 . Bibcode:2018PhRvC..97f5503A. doi:10.1103/PhysRevC.97.065503. S2CID   67854591.
  26. Klapdor-Kleingrothaus, H. V.; Dietz, A.; Harney, H. L.; Krivosheina, I. V. (2001). "Evidence for Neutrinoless Double Beta Decay". Modern Physics Letters A . 16 (37): 2409–2420. arXiv: hep-ph/0201231 . Bibcode:2001MPLA...16.2409K. doi:10.1142/S0217732301005825. S2CID   18771906.
  27. Feruglio, F.; Strumia, A.; Vissani, F. (2002). "Neutrino oscillations and signals in beta and 0nu2beta experiments". Nuclear Physics . 637 (1): 345–377. arXiv: hep-ph/0201291 . Bibcode:2002NuPhB.637..345F. doi:10.1016/S0550-3213(02)00345-0. S2CID   15814788.
  28. Aalseth, C. E.; et al. (2002). "Comment on "evidence for Neutrinoless Double Beta Decay"". Modern Physics Letters A . 17 (22): 1475–1478. arXiv: hep-ex/0202018 . Bibcode:2002MPLA...17.1475A. doi:10.1142/S0217732302007715. S2CID   27406915.
  29. Zdesenko, Y. G.; Danevich, F. A.; Tretyak, V. I. (2002). "Has neutrinoless double β decay of 76Ge been really observed?". Physics Letters B . 546 (3–4): 206. Bibcode:2002PhLB..546..206Z. doi: 10.1016/S0370-2693(02)02705-3 .
  30. Bakalyarov, A. M.; Balysh, A. Y.; Belyaev, S. T.; Lebedev, V. I.; Zhukov, S. V. (2005). "Results of the experiment on investigation of Germanium-76 double beta decay". Physics of Particles and Nuclei Letters . 2 (2005): 77–81. arXiv: hep-ex/0309016 . Bibcode:2003hep.ex....9016B.
  31. Klapdor-Kleingrothaus, H. V.; Krivosheina, I. V. (2006). "The Evidence for the Observation of 0νββ Decay: The Identification of 0νββ Events from the Full Spectra". Modern Physics Letters A . 21 (20): 1547. Bibcode:2006MPLA...21.1547K. doi:10.1142/S0217732306020937.
  32. Agostini, M.; et al. (GERDA Collaboration) (2017). "Background-free search for neutrinoless double-β decay of 76Ge with GERDA". Nature . 544 (7648): 47–52. arXiv: 1703.00570 . Bibcode:2017Natur.544...47A. doi:10.1038/nature21717. PMID   28382980. S2CID   4456764.
  33. 1 2 Barabash, A. S.; Hubert, Ph.; Nachab, A.; Umatov, V. I. (2019). "Search for triple and quadruple β decay of Nd150". Physical Review C. 100 (4): 045502. arXiv: 1906.07180 . doi:10.1103/PhysRevC.100.045502. S2CID   189999159.
  34. Heeck, J.; Rodejohann, W. (2013). "Neutrinoless Quadruple Beta Decay". Europhysics Letters. 103 (3): 32001. arXiv: 1306.0580 . Bibcode:2013EL....10332001H. doi:10.1209/0295-5075/103/32001. S2CID   118632700.
  35. Hirsch, M.; Srivastava, R.; Valle, JWF. (2018). "Can one ever prove that neutrinos are Dirac particles?". Physics Letters B. 781: 302–305. arXiv: 1711.06181 . Bibcode:2018PhLB..781..302H. doi: 10.1016/j.physletb.2018.03.073 .