Fact-checked by Grok 2 weeks ago

Double beta decay

Double beta decay is a rare nuclear process in which two neutrons within simultaneously transform into two protons, or vice versa, resulting in the emission of two electrons (or positrons) and, in the observed mode, two antineutrinos. This second-order occurs in certain even-even isotopes that are stable against single but can undergo this double transition due to , with observed half-lives ranging from $10^{18} to $10^{24} years in nuclei such as ^{76}Ge, ^{100}Mo, and ^{136}Xe. There are two primary modes: the standard two-neutrino double beta decay (2νββ), which is permitted by the of particle physics and produces a continuous energy spectrum shared among the emitted particles, and the hypothetical , which violates lepton number conservation by emitting only two electrons with a monochromatic energy peak at the decay's Q-value (typically 2–3 MeV). The 2νββ mode has been experimentally observed in 13 nuclides since its theoretical prediction by in 1935, confirming the process as a higher-order manifestation of the . In contrast, 0νββ remains undetected, with current lower limits on half-lives (as of 2025) exceeding $10^{25}–$10^{26} years from experiments like , KamLAND-Zen, and CUORE, which use ultra-low-background detectors to search for the signature electron sum energy peak. If observed, 0νββ would establish that neutrinos are massive Majorana particles— their own antiparticles—providing crucial insights into neutrino masses (effective Majorana mass m_{\beta\beta} estimated at 15–50 meV for inverted hierarchy), the origin of matter-antimatter asymmetry, and potential extensions beyond the . Ongoing and future experiments, such as nEXO and , aim to probe half-lives up to $10^{28} years to test these fundamental properties.

Principles of Double Beta Decay

Single Beta Decay Recap

is a type of in which the of the changes while the remains the same, resulting from the weak nuclear interaction. , including beta rays, was first discovered in 1896 by through the observation of spontaneous emissions from salts that could penetrate materials and expose photographic plates. In the early 20th century, advanced the understanding of through her studies of the continuous energy spectrum of emitted particles. This spectrum posed a puzzle, as it violated unless an undetected neutral particle carried away the missing energy; in 1930, proposed the existence of such a particle, later named the , to resolve this issue. In beta minus (β⁻) decay, a in the is transformed into a proton, accompanied by the emission of an (β⁻ particle) and an electron antineutrino. This process, mediated by the charged-current via W⁻ exchange, increases the by one and is represented by the reaction n \to p + e^- + \bar{\nu}_e, where the antineutrino ensures conservation of and . Conversely, in beta plus (β⁺) decay, a proton converts to a , emitting a (β⁺ particle) and an , decreasing the by one: p \to n + e^+ + \nu_e. This mode is also weak-mediated via W⁺ and occurs in proton-rich nuclei. The energy released in beta decay, known as the Q-value, is the difference in rest mass energy between the initial and final atomic states, converted into shared among the (or ), , and recoiling . For β⁻ decay, Q_{\beta^-} = [m(^{A}_{Z}X) - m(^{A}_{Z+1}Y)] c^2, where masses are ; positive Q enables the decay, with the spectrum ranging from 0 to Q. The decay rate is calculated using from time-dependent , which states that the transition probability per unit time is \Gamma = \frac{2\pi}{\hbar} | \langle f | H_w | i \rangle |^2 \rho(E_f), where H_w is the weak Hamiltonian matrix element and \rho(E_f) is the density of final states for the three-body decay products. Selection rules dictate allowed transitions: for Fermi (spin-singlet) decays, ΔJ = 0 with no parity change (π_i π_f = +1); for Gamow-Teller (spin-triplet) decays, ΔJ = 0, ±1 (no 0→0) with no parity change, conserving total and in the non-relativistic limit.

Double Beta Decay Modes

Double beta decay is a rare process in which two neutrons in an simultaneously transform into two protons, or vice versa in the mode, thereby increasing (or decreasing) the Z by 2 while preserving the A. This decay arises in certain even-even nuclei where single to an intermediate odd-A nucleus is energetically forbidden or significantly suppressed due to effects and stability, making the direct transition to the ground or low-lying excited states of the daughter the dominant pathway. As a second-order process, double decay proceeds through virtual intermediate nuclear states, rendering it extremely slow with half-lives typically exceeding $10^{18} years. The process manifests in two primary s, distinguished by their particle emissions and implications for symmetries. The two-neutrino , denoted $2\nu\beta\beta, involves the emission of two electrons and two electron antineutrinos, conserving total . The general reaction is given by (A,Z) \to (A,Z+2) + 2e^- + 2\bar{\nu}_e, where the antineutrinos carry away excess energy and momentum. In the neutrinoless mode, $0\nu\beta\beta, only two electrons are emitted, with no neutrinos involved, resulting in a violation of lepton number conservation by two units. The corresponding reaction is (A,Z) \to (A,Z+2) + 2e^-. The energetics of both modes are governed by the Q-value, calculated as the difference between the atomic masses of the parent and daughter nuclei (accounting for the two emitted electrons), which typically ranges from 0.1 to 3 MeV for viable candidate nuclei. This available energy determines the maximum shared among the decay products.

Historical Development

Theoretical Prediction

The theoretical prediction of double beta decay emerged in the mid-1930s as an extension of the Fermi theory of single , positing it as a second-order process involving the simultaneous emission of two electrons and two antineutrinos from a . In 1935, first proposed this process, calculating its rate using and estimating a lower limit on the of greater than $10^{17} years for candidate nuclei, highlighting its extreme rarity compared to single . Building on Mayer's work, G. Racah performed a more detailed calculation in 1937, applying perturbation theory to derive explicit decay rates for the two-neutrino mode and introducing the possibility of a neutrinoless variant. Racah's analysis framed the neutrinoless mode as a potential probe of neutrino properties, suggesting it could test theories where neutrinos are their own antiparticles, as proposed by Ettore Majorana earlier that year. During the 1930s and 1940s, the neutrinoless double beta decay was increasingly recognized as a sensitive test for lepton number conservation, with Wendell H. Furry's 1939 computation showing that its rate would be suppressed by a factor proportional to the square of the neutrino mass relative to the electron mass, yielding half-lives far exceeding those of the two-neutrino mode—potentially by orders of magnitude if the neutrino mass is small. These early theoretical efforts connected double beta decay to limits on neutrino mass before the advent of the Standard Model, providing a framework for understanding neutrino identity and weak interaction symmetries through non-observation of the process.

First Observations and Milestones

The earliest experimental hint of double beta decay came from geochemical evidence in the late 1940s and early 1950s, where excess isotopes in ancient minerals suggested the process had occurred over geological timescales. In 1950, Mark G. Inghram and John H. Reynolds reported the first positive indication of two-neutrino double beta (2νββ) decay in ^{130}Te through of natural samples, estimating a on the order of 10^{21} years. This geochemical approach, which integrated decay rates over millions of years, provided indirect proof of the process but lacked direct particle detection. Direct laboratory observations proved more challenging due to the rarity of the decay, but progress accelerated in the with improved detector technologies. Early searches using Geiger counters and scintillation detectors on isotopes like ^{48} and ^{124} yielded limits but no signals. By , Ettore Fiorini and collaborators conducted the first experiment with a germanium lithium-drifted (Ge(Li)) detector on enriched ^{76}Ge, setting stringent limits exceeding 10^{20} years and paving the way for semiconductor-based searches. Geochemical confirmations followed, such as Thomas Kirsten's measurement of 2νββ in ^{82} using from selenium minerals. These efforts in the established the involvement of neutrinos in the decay process by aligning observed rates with theoretical predictions for the two-neutrino mode. The 1980s marked a shift to underground laboratories to suppress cosmic-ray backgrounds, enabling higher sensitivities. Facilities like the Gran Sasso National Laboratory in Italy, operational from 1987, hosted early low-background experiments, including prototypes for tracking detectors. The first direct real-time observation of 2νββ came in 1987 with Steven R. Elliott and colleagues using a time projection chamber on ^{82}Se, confirming the mode with a measured half-life of (1.1^{+0.8}_{-0.3}) \times 10^{20} years. Confirmations extended to other isotopes by the early 1990s, such as ^{76}Ge by Frank T. Avignone III's group in 1991 and ^{100}Mo in the NEMO-1 experiment starting in 1989. Meanwhile, the Heidelberg-Moscow experiment, initiated in 1990 at Gran Sasso with enriched ^{76}Ge high-purity germanium detectors, reported a controversial claim in 2001 for neutrinoless double beta (0νββ) decay, suggesting a half-life of 0.35-0.8 \times 10^{25} years; this result was later disputed by subsequent analyses and experiments. Key milestones in the 2000s and 2010s reflected scaling to larger detectors for greater statistical power. The CUORICINO experiment, a bolometric array of 62 tellurium oxide crystals targeting ^{130}Te, began data-taking in 2003 at Gran Sasso, achieving background rates low enough to set competitive limits on both 2νββ and 0νββ modes after over 20 kg·year exposure. This paved the way for tonne-scale efforts in the 2010s, such as the NEMO-3 tracking calorimeter (operational 2001-2010, \sim 10 kg of multiple isotopes) and the CUORE successor (launched 2017 with 988 TeO_2 crystals totaling 741 kg), which pushed sensitivities to half-lives beyond 10^{26} years while confirming 2νββ rates in several nuclei. These advancements underscored the transition from exploratory searches to precision measurements probing neutrino properties.

Standard Two-Neutrino Mode

Mechanism and Kinematics

The two-neutrino double beta decay (2νββ) proceeds as a second-order process, in which two neutrons in an even-even transform into two protons via the exchange of two virtual bosons, passing through an odd-odd virtual nuclear state. This mechanism involves two successive virtual beta decays, each governed by the charged-current , with the antineutrinos emitted in the final state ensuring conservation. Kinematically, the total energy release Q, determined by the mass difference between the parent and daughter nuclei, is shared among the two electrons and two antineutrinos, resulting in a continuous spectrum for the summed kinetic energies of the two electrons, extending from near zero up to the full Q-value. The antineutrinos carry away variable portions of the energy, broadening the electron sum-energy distribution into a continuum shape characteristic of allowed second-order weak processes. In contrast, the neutrinoless mode yields a discrete monochromatic peak at the Q-value due to the absence of neutrinos. The factor G^{2\nu}, which encodes the available kinematic volume for the decay, is given by G^{2\nu} \propto \int p_1 E_1 p_2 E_2 F(Z,E_1) F(Z,E_2) \, dE_1 \, dE_2, where p_i and E_i are the momenta and total energies of the , and F(Z,E_i) are the Fermi functions accounting for distortion of the wave functions by the charge Z. This is evaluated numerically, incorporating relativistic and screening effects for precision. The corresponding half-life for the process is expressed as \left( T_{1/2}^{2\nu} \right)^{-1} = G^{2\nu} |M^{2\nu}|^2, where M^{2\nu} is the nuclear matrix element that captures the overlap of initial, intermediate, and final nuclear wave functions, modulated by operators. This formulation originates from the foundational treatment of the decay rate as a product of and matrix element factors. The matrix element M^{2\nu} is predominantly composed of Gamow-Teller contributions, involving spin-flip \sigma \tau^\pm operators, with smaller Fermi components from non-spin-flip \tau^\pm transitions, reflecting the axial-vector dominance in weak interactions.

Measured Half-Lives and Isotopes

Double beta decay in the two-neutrino mode (2νββ) has been experimentally observed in eleven even-even nuclei, providing crucial validation of the standard framework. These measurements span a wide range of half-lives, from approximately 10^{18} years for short-lived candidates to over 10^{24} years for the longest, reflecting variations in and factors. The isotopes studied include ^{48}Ca, ^{76}Ge, ^{82}Se, ^{96}Zr, ^{100}Mo, ^{116}Cd, ^{124}Xe, ^{128}Te, ^{130}Te, ^{136}Xe, and ^{150}Nd, selected for their stability, natural abundance, and favorable Q-values—the available energy release for the decay. Measurements employ two primary techniques: direct detection, which counts the summed kinetic energies of the two emitted electrons using low-background detectors such as high-purity diodes, scintillators, or time projection chambers; and geochemical methods, which integrate decay rates over geological timescales by measuring the accumulation of isotopes in ancient minerals. Direct methods offer higher for shorter half-lives but are limited by exposure time, while geochemical approaches excel for ultra-long half-lives like that of ^{128}, though they introduce larger systematic uncertainties from sample age assumptions. Recent advances in cryogenic bolometers and enriched sources have improved direct measurements, achieving percent-level in several cases. The following table summarizes the measured 2νββ half-lives to the , along with Q-values and methods. Half-lives are weighted averages or recommended values from recent evaluations, with statistical and systematic uncertainties where available, as of 2025.
IsotopeQ-value (MeV)Half-life (yr)MethodReference
^{48}Ca4.272(5.96^{+1.39}_{-1.08}) \times 10^{19}
^{76}2.039(2.022 \pm 0.040) \times 10^{21}
^{82}2.995(8.69 \pm 0.07) \times 10^{19}
^{96}Zr3.350(2.335 \pm 0.210) \times 10^{19}/Geochemical
^{100}Mo3.034(7.08 \pm 0.14) \times 10^{18}
^{116}Cd2.802(2.738 \pm 0.120) \times 10^{19}
^{124}Xe2.857(1.10 \pm 0.13) \times 10^{22}
^{128}Te0.868(2.167 \pm 0.200) \times 10^{24}Geochemical
^{130}Te2.527(8.69 \pm 0.18) \times 10^{20}/Geochemical
^{136}Xe2.458(2.240 \pm 0.061) \times 10^{21}
^{150}Nd3.367(1.160 \pm 0.370) \times 10^{19}
Uncertainties in half-lives arise from statistical counting errors, background subtraction, , and—for geochemical results—geological , typically ranging from 1% to 20% relative precision. These experimental values generally align with theoretical predictions from and quasi-particle random phase approximation calculations, which incorporate nuclear matrix elements and integrals, confirming the second-order weak process nature of 2νββ within 1-2 standard deviations for most isotopes. : Barabash, A. S. (2025). Comprehensive Review of 2β Decay Half-Lives. Nuclear Physics B, 1005, 118789. https://arxiv.org/abs/2503.19130 : Particle Data Group (2025). Review of Particle Physics: Double-β Decay. https://pdg.lbl.gov/2025/listings/rpp2025-list-double-beta-decay.pdf

Neutrinoless Mode

Physical Implications

The observation of neutrinoless double beta decay (0νββ) would have profound implications for fundamental physics, serving as a unique probe into the nature of neutrinos and potential violations of lepton number conservation beyond the Standard Model. Unlike the standard two-neutrino double beta decay (2νββ), which conserves lepton number, 0νββ would indicate a process where two neutrons transform into two protons and two electrons without emitting neutrinos, directly testing mechanisms that generate small neutrino masses. A primary physical implication of 0νββ is its sensitivity to the absolute neutrino mass scale through the effective electron neutrino Majorana mass, defined as m_{ee} = \left| \sum_i U_{ei}^2 m_i \right|, where U_{ei} are elements of the Pontecorvo-Maki-Nakagawa-Sakata (PMNS) mixing matrix and m_i are the mass eigenvalues. This parameter encapsulates the contribution of all mass states to the decay amplitude, providing a direct measure inaccessible to experiments, which only constrain mass-squared differences. Furthermore, 0νββ rates depend on the mass hierarchy—normal or inverted—and the PMNS matrix mixing angles, allowing constraints on these parameters that refine our understanding of flavor mixing and could resolve ambiguities in data. The half-life of 0νββ is related to these neutrino properties via the expression T_{1/2}^{0\nu} \propto \frac{1}{G^{0\nu} |M^{0\nu}|^2 m_{ee}^2}, where G^{0\nu} is the kinematic phase-space factor and |M^{0\nu}| is the nuclear matrix element encoding the nuclear structure effects. This relation underscores how 0νββ bridges particle and nuclear physics, with the decay rate scaling quadratically with m_{ee}, making it a powerful tool to bound or detect sub-eV neutrino masses. In , 0νββ complements indirect probes of masses, such as those from anisotropies and large-scale structure surveys, which constrain the of masses but not the contributions or Majorana phases entering m_{ee}. While oscillation experiments provide mass-squared differences like \Delta m_{21}^2 \approx 7.5 \times 10^{-5} eV² and |\Delta m_{31}^2| \approx 2.5 \times 10^{-3} eV², 0νββ offers a direct, flavor-specific probe that could reveal tensions between hierarchical spectra inferred from and the effective mass. Finally, 0νββ connects to grand unified theories (GUTs), where violation is a natural feature arising from the unification of fundamental forces at high energies, often predicting Majorana neutrinos through mechanisms like the seesaw model to explain the small observed masses. An observed decay would support GUT frameworks by providing evidence for such violations at low energies, potentially linking to supersymmetric extensions or other beyond-Standard-Model scenarios.

Lepton Number Violation and Majorana Neutrinos

The neutrinoless double beta decay (0νββ) process was first theoretically proposed by Wendell H. Furry in 1939, who recognized that it could occur without the emission of neutrinos if the neutrino is identical to its antiparticle, thereby allowing the process to proceed via a second-order weak interaction without lepton number conservation. This decay mode, in which two neutrons transform into two protons and two electrons, violates the total lepton number by two units (ΔL = 2), a conservation law strictly upheld in the Standard Model of particle physics but potentially broken in beyond-Standard-Model extensions. In the canonical mechanism for 0νββ mediated by light Majorana neutrinos, the two neutrons decay through left-handed weak currents, with a virtual neutrino exchanged between the decay vertices; the Majorana nature of the neutrino enables it to act as both particle and antiparticle, allowing absorption at the second vertex and closing the propagation loop without a net neutrino. Alternative mechanisms can involve right-handed weak currents, such as those from a nonzero neutrino magnetic moment or heavy sterile neutrinos, but the light Majorana exchange remains the dominant paradigm linking 0νββ to lepton number violation. The corresponding Feynman diagram depicts two negatively charged W bosons exchanged between the quark lines, with the Majorana neutrino propagator connecting the two charged-lepton vertices, effectively forming a loop that enforces the ΔL = 2 transition. The Majorana condition, under which the neutrino field satisfies \nu = \bar{\nu}, is essential for this process, as it permits the neutrino to couple to both left-handed creation and right-handed annihilation operators in the weak interaction. Such a self-conjugate nature can arise from beyond-Standard-Model origins, including the type-I seesaw mechanism, where heavy right-handed s generate light Majorana masses for the active s at low energies. At the effective field theory level, the 0νββ process is described by a dimension-9 of the form (udc)^2 / \Lambda^5, where u and d denote up- and down-quark fields, c the charged lepton field, and \Lambda a high-energy scale suppressing the ΔL = 2 violation. This Majorana character ties directly to the small neutrino masses inferred from oscillation experiments, providing a testable link between nuclear-scale observations and fundamental lepton physics.

Experimental Approaches

Detector Technologies

Double beta decay experiments require specialized detector technologies to achieve the extreme sensitivity needed for detecting rare events, primarily through the use of high-purity isotopes as both source and detector material. Key isotopes include ^{76}Ge, ^{136}Xe, and ^{130}Te, selected for their favorable decay kinematics and availability for enrichment. These detectors operate in deep underground laboratories, such as the (LNGS) in and the in , to shield against cosmic ray-induced backgrounds using overburden rock equivalent to thousands of meters of water. High-purity materials are essential, with rigorous screening for radioactive contaminants to minimize internal backgrounds. For ^{76}Ge, high-purity (HPGe) diodes serve as the primary detection method, where the is enriched to over 85% and fabricated into n-type detectors that act as both source and calorimeter. Experiments like GERDA and its successor employ arrays of these broad-energy germanium (BEGe) detectors, which provide excellent energy resolution of approximately 2-4 keV full width at half maximum () at the Q-value of the . Recent results from LEGEND-200, as of 2025, confirm sustained performance with resolutions around 2.5 keV . This calorimetric approach measures the total energy deposited by the two electrons, enabling identification of the sum energy peak relevant to both two-neutrino (2νββ) and neutrinoless (0νββ) modes. In contrast, ^{136}Xe-based detectors utilize in either liquid or gaseous form, often enriched to 90% or more to increase the fiducial mass. KamLAND-Zen employs a liquid scintillator doped with enriched Xe, where scintillation light and are detected to reconstruct energy with resolutions around 5-10% at the Q-value, though recent upgrades as of aim for ~2% σ (~4.7% FWHM) through improved light yield. For tracking capabilities, the NEXT experiment uses a high-pressure gaseous time projection chamber (TPC), combining electroluminescent amplification with photomultipliers to track trajectories and distinguish signal topologies, achieving sub-1% FWHM resolution. This hybrid calorimetric-tracking method enhances background rejection while maintaining high efficiency. ^{130}Te detectors, such as those in CUORE, rely on cryogenic bolometers made from TeO_2 crystals, which operate at millikelvin to measure temperature rises from deposits with high . These natural-abundance (34%) crystals, housed in a large array of 988 cubic crystals totaling 741 kg, deliver resolutions of around 5 keV FWHM at the Q-value, leveraging the calorimetric for full containment. The benefits from the of and the isotope's high Q-value, supporting searches in both decay modes, with recent 2025 analyses confirming stable performance. Overall, these technologies provide resolutions typically ranging from a few keV FWHM for solid-state detectors to around 5-10% for liquid systems at the Q-values (typically 2-3 MeV) to resolve the monochromatic 0νββ peak from continuum backgrounds, with ongoing advancements in material purification and detector design driving next-generation sensitivities.

Background Discrimination and Sensitivity

In double beta decay experiments, the primary background sources that challenge the search for the neutrinoless mode include the two-neutrino double beta decay (2νββ) continuum, which produces a broad overlapping the (ROI) for the neutrinoless signal; alpha decays from radioactive contaminants, often concentrated on detector surfaces; emanation from materials and air, leading to daughter products that deposit via beta and gamma emissions; and cosmogenic , where high-energy muons and neutrons induce long-lived isotopes in detector components during transport or operation. To discriminate these backgrounds from potential neutrinoless signals, experiments employ techniques tailored to detector types, such as pulse shape analysis in high-purity (HPGe) detectors, which distinguishes single-site events (like double beta decay) from multi-site interactions typical of gamma rays or betas by analyzing the time profile of the charge collection signal. In xenon time projection chambers (TPCs), topological track reconstruction identifies the number and spatial configuration of ionization tracks, rejecting single-electron backgrounds from beta or gamma events while favoring the paired topology expected for double beta decay electrons. To suppress pile-up events or uncorrelated backgrounds such as single beta decays, time correlation analysis requires coincident signals within short timescales, though it cannot reject the 2νββ continuum. The of searches, which determines the limit achievable, depends critically on levels and is quantified by the for the -limited : T_{1/2}^{0\nu} \propto \frac{\sqrt{M t / \mathcal{B}}}{\epsilon}, where M is the fiducial mass, t is the live time, \mathcal{B} is the index (in counts per keV per kg per year), and \epsilon is the signal detection ; this relation highlights how reducing \mathcal{B} improves quadratically. The analysis focuses on the ROI centered at the Q-value of the (typically a few keV wide to match energy resolution), where are modeled and subtracted to isolate any excess consistent with a monoenergetic . Bayesian statistical methods are commonly used to set limits, incorporating priors on rates and systematic uncertainties to compute credible intervals for the , providing robust upper bounds in the absence of signal. Key improvements in background discrimination include active veto systems, such as muon vetoes using scintillators or water Cherenkov detectors to tag cosmogenic events in real time, and rigorous material assays via low-background to screen components for radiopurity before assembly, minimizing internal contamination from the outset. These strategies, applied in and xenon-based detectors, enhance overall sensitivity by orders of magnitude compared to unvetoed setups.

Key Results and Status

Historical Controversies

The Heidelberg-Moscow experiment, conducted from 1990 to 2003 at the Gran Sasso National Laboratory using high-purity detectors enriched in ^{76}Ge, provided the most sensitive search for (0νββ) at the time, accumulating over 35 kg·yr of exposure. In 2001, a subgroup led by H. V. Klapdor-Kleingrothaus analyzed a portion of the data and reported a broad peak at 2039 keV in the summed , interpreted as for 0νββ with approximately 3σ significance and a half-life estimate of T_{1/2}^{0\nu} \approx 1.5 \times 10^{25} years. This claim sparked significant within the and the broader physics , as it stemmed from a selective reanalysis by only part of the team, excluding input from other members who favored a conservative of T_{1/2}^{0\nu} > 1.9 \times 10^{25} years at 90% confidence level. Critics highlighted statistical issues, including potential of background models, low event counts (around 29 excess events), and the failure to meet the conventional 5σ threshold for discovery in , arguing that the signal could arise from random fluctuations or unaccounted systematics. The dispute was resolved by subsequent experiments designed specifically to test the claim. The GERDA collaboration, using refurbished ^{76}Ge detectors from Heidelberg-Moscow and IGEX in a liquid argon cryostat for improved background rejection, collected from 2011 to 2013 and observed no excess events consistent with 0νββ, setting a lower limit of T_{1/2}^{0\nu} > 2.1 \times 10^{25} years at 90% confidence level and excluding the Heidelberg-Moscow signal at over 5σ in a model-independent manner. Earlier in the , searches for double beta decay modes in isotopes like ^{86}Kr also produced hints of anomalies, such as excess events in low-background counters that suggested possible 0νββ signals, but these were later dismissed as background artifacts upon refined and higher-statistics measurements that yielded only conservative limits exceeding $10^{21} years. These historical disputes underscored the challenges of rare-event detection and prompted stricter guidelines in , including requirements for blinded analyses, multi-experiment corroboration, and significance levels beyond 3σ for extraordinary claims, influencing the design of modern searches.

Current Limits and Recent Findings

Recent measurements of the two-neutrino double beta decay (2νββ) mode have refined half-lives for several isotopes, providing benchmarks for nuclear matrix element calculations. For ^{100}Mo, the NEMO-3 experiment reported a half-life of T_{1/2}^{2\nu} = [6.81 \pm 0.01_{\mathrm{stat}} ^{+0.38}{-0.40}{\mathrm{syst}}] \times 10^{18} , \mathrm{yr} based on the full dataset with ~7 kg of enriched material observed over ~5 years. For ^{130}Te, the CUORE experiment's 2025 analysis, utilizing 1038 kg·yr of exposure, yielded T_{1/2}^{2\nu} = (9.32^{+0.05}{-0.04} \pm 0.07) \times 10^{20} , \mathrm{yr}, improving precision by a factor of two through optimized signal modeling and background reduction. Similarly, KamLAND-Zen measured the ^{136}Xe half-life as T{1/2}^{2\nu} = (2.38 \pm 0.07_{\mathrm{stat}} \pm 0.14_{\mathrm{sys}}) \times 10^{21} , \mathrm{yr}, the most precise to date for this isotope. No evidence for (0νββ) has been observed in ongoing tonne-scale experiments, setting increasingly stringent lower limits on half-lives and constraining the effective Majorana m_{\beta\beta}. The combined analysis from GERDA, MAJORANA Demonstrator, and LEGEND-200 establishes a limit of T_{1/2}^{0\nu} > 2.8 \times 10^{26} , \mathrm{yr} (90% \mathrm{C.L.}) for ^{76}, incorporating LEGEND-200's initial dataset of 200 kg·yr exposure. For ^{130}Te, CUORE's October 2025 update reports T_{1/2}^{0\nu} > 3.5 \times 10^{25} , \mathrm{yr} (90% \mathrm{C.L.}), the tightest limit for this based on full bolometric data. KamLAND-Zen provides T_{1/2}^{0\nu} > 2.3 \times 10^{26} , \mathrm{yr} (90% \mathrm{C.L.}) for ^{136}Xe, derived from full Phase II data with enhanced xenon loading and background mitigation. These limits translate to upper bounds on m_{\beta\beta} of approximately 0.03--0.1 eV, depending on nuclear matrix element evaluations, tightening constraints on quasi-degenerate neutrino mass hierarchies. A March 2025 arXiv review confirms 2νββ half-life measurements for 14 nuclei, aligning with standard model expectations and aiding validation of phase-space factors. The 2025 Nuclear Matrix Element workshop highlighted advancements in ab initio calculations, reducing uncertainties in 0νββ interpretations by up to 20% for key isotopes. Preliminary data from NEXT-HD, post-2022, suggest potential sensitivity improvements for ^{136}Xe beyond current bounds, though full results await publication.

Theoretical Frameworks

Nuclear Matrix Element Calculations

The nuclear matrix element (NME) for neutrinoless double beta decay is defined as M = \langle f | \hat{O} | i \rangle, where |i\rangle and |f\rangle represent the initial and final nuclear wave functions, and \hat{O} is the two-body transition operator incorporating Gamow-Teller (GT) and Fermi components, often expressed as M_{0\nu} = M_{\mathrm{GT}}^{0\nu} + M_{\mathrm{F}}^{0\nu} with the axial-vector coupling g_A entering quadratically in the GT term. These NMEs quantify the overlap of nuclear structure effects in the decay amplitude and are essential for relating observed half-lives to fundamental parameters. Several computational methods are employed to evaluate NMEs, each suited to different nuclear mass regions. The shell model uses exact diagonalization within a truncated valence space, providing reliable results for lighter nuclei like ^{48}\mathrm{Ca} and ^{76}\mathrm{Ge} but becoming computationally intensive for heavier systems. The quasiparticle random phase approximation (QRPA) extends to medium-heavy nuclei such as ^{136}\mathrm{Xe}, incorporating particle-hole excitations and proton-neutron pairing interactions tuned to two-neutrino double beta decay data. The interacting boson model (IBM) approximates the nucleus as a system of s- and d-bosons representing nucleon pairs, yielding NMEs that interpolate between shell model and QRPA predictions for transitional nuclei. Uncertainties in NME calculations arise from model dependencies and nuclear structure approximations, typically varying by 20-50% across methods, with differences up to a factor of three for key isotopes. A significant source of uncertainty is the quenching of the axial coupling constant g_A, where the effective value in nuclei is reduced to about 0.7-0.8 times the free-nucleon value due to missing meson-exchange currents or configuration mixing, impacting GT-dominated transitions. Recent advances include ab initio approaches using chiral effective field theory (EFT), which systematically include short-range correlations and improve convergence for NMEs in light-to-medium mass nuclei. The 2025 international workshop on NMEs at Osaka University emphasized collaborations between theorists and experimentalists to refine these ab initio chiral EFT calculations, aiming to reduce uncertainties through better handling of two-body currents and higher-order terms. In extracting the effective Majorana neutrino mass m_{\beta\beta}, NMEs play a central role via the phase-space relation m_{\beta\beta} \propto 1 / \sqrt{T_{1/2} |M_{0\nu}|^2 }, where T_{1/2} is the half-life, underscoring the need for precise computations to interpret experimental limits.

Connections to Beyond-Standard-Model Physics

Neutrinoless double beta decay (0νββ) serves as a sensitive probe for by enabling violation through mechanisms that extend past the light Majorana neutrino exchange, such as right-handed weak currents and heavy particle mediation. These processes introduce additional contributions to the decay amplitude, parameterized by effective couplings that can be constrained by experimental limits. In particular, right-handed currents arise in extensions like left-right symmetric models, where the parameter η quantifies the coupling between a right-handed current and a left-handed current, while λ describes the coupling involving two right-handed quark currents. Heavy neutrino exchange, mediated by right-handed Majorana neutrinos with masses around the TeV scale, contributes inversely proportional to the square of the heavy neutrino mass (1/M_N^2), providing another avenue for violation. The generalized expression for the 0νββ half-life incorporates these beyond-Standard-Model effects alongside the effective electron neutrino Majorana mass m_ee, typically written as [T_{1/2}^{-1}]{0\nu} \propto |M{0\nu} (m_{ee}/\langle m_e \rangle + \eta + \lambda + \sum (U_{ei}^2 / M_{N_i}) )|^2, where M_{0\nu} is the nuclear matrix element, and the phase-space factor is implicit. This form highlights how experimental limits on the half-life translate into bounds on the BSM parameters, with current data from experiments like GERDA and CUORE yielding η < 10^{-6} and λ < 10^{-6} at 90% confidence level (as of 2024), depending on the nuclear matrix element evaluation. These constraints impact models such as supersymmetric grand unified theories (SUSY GUTs), where right-handed currents or R-parity violation can enhance the decay rate, requiring the right-handed gauge boson mass m_{WR} > 3 TeV in SO(10) SUSY GUTs to evade limits. Similarly, leptoquarks in extended models are bounded, with coupling products constrained to very small values (typically <10^{-8} for TeV-scale scalar leptoquarks) as their tree-level exchange contributes to the short-range part of the amplitude. In left-right symmetric models, the presence of V+A currents alongside the Standard Model V-A currents alters the kinematics of the decay, particularly the angular correlation between the two emitted electrons, which deviates from the standard (1 - \beta \cos\theta) form to include terms sensitive to η and λ. Measurements of these correlations in tracking detectors could distinguish V+A contributions, with projected sensitivities in future experiments like SuperNEMO potentially probing η down to 10^{-7}. Regarding future prospects, advances in daughter ion detection—such as barium tagging in xenon-based detectors like nEXO—enable precise event reconstruction and background rejection, facilitating the separation of BSM mechanisms from the m_ee-dominated scenario through improved spectral and kinematic analysis. This technique, combined with high-resolution tracking, promises to disentangle contributions by identifying unique signatures in the daughter ion momentum or charge state.

Double Electron Capture

Double electron capture (2EC) is a second-order weak interaction process in which two protons within an atomic nucleus simultaneously capture two orbital electrons, converting into two neutrons and emitting two electron antineutrinos in the two-neutrino mode: (A,Z) + 2e^- \to (A,Z-2) + 2\bar{\nu}_e. This decay mode competes with double beta decay in certain even-even nuclei where the intermediate odd-odd states suppress single beta transitions. The available energy, or Q-value, for 2EC is given by the difference between the atomic masses of the parent and daughter nuclei, reduced by the binding energies of the captured electrons (typically from the K-shell). The process can occur resonantly if this Q-value nearly matches the excitation energy of a low-lying state in the daughter nucleus, leading to partial or full degeneracy that enhances the decay rate by orders of magnitude compared to non-resonant cases. Such resonance conditions are particularly favorable in isotopes with precisely measured mass differences, often determined using Penning trap spectrometry. Promising candidates for 2EC include the isotopes ^{78}Kr, ^{106}Cd, ^{124}Xe, and ^{136}Xe, which share some overlap with searches due to their nuclear structure. Theoretical predictions for the two-neutrino 2EC half-lives in these nuclides range from approximately 10^{21} to 10^{24} years, reflecting the suppressed phase space and nuclear matrix elements involved. The two-neutrino 2EC process has been experimentally observed in ^{78}Kr and ^{124}Xe; for ^{78}Kr, low-temperature bolometers and gas detectors have provided evidence for the two-neutrino double K-shell capture mode, yielding a half-life of (1.9^{+1.3}_{-0.7} \pm 0.3) \times 10^{22} years at 90% confidence level. In 2025, the two-neutrino double electron capture in ^{124}Xe was observed with a half-life of (1.8 \pm 0.5) \times 10^{22} years at 4.4σ significance. The COBRA experiment, employing CdZnTe semiconductor detectors, is designed to probe electron capture modes in ^{106}Cd alongside other double beta processes, leveraging the self-contained source nature of the material. Experimental efforts to observe or limit 2EC rates have utilized low-background detectors tailored to the atomic signatures, such as X-ray emissions from atomic relaxation following electron capture. Recent searches have also set limits on neutrinoless 2EC; for instance, the CRESST-II experiment established a lower limit of T_{1/2} > 9.4 \times 10^{18} years at 90% confidence level for ^{180}W. A key advantage of 2EC over electron-emitting double beta modes is the absence of radiation from charged particles, which reduces background in the energy spectrum. In the neutrinoless 2EC case, the decay produces a sharp monochromatic line at the Q-value (minus binding energies), facilitating background discrimination and improving sensitivity to lepton-number-violating signals analogous to .

Higher-Order Beta Decays

Higher-order beta decays encompass rare nuclear processes that extend beyond the standard double beta decay modes, involving additional beta emissions or alternative charge-changing transitions. These include neutrinoless double positron emission (0νβ⁺β⁺), where two protons transform into two neutrons with the emission of two positrons and no neutrinos, and neutrinoless positron-electron capture (0νβ⁺EC), which combines positron emission with electron capture from the atomic shell. Neutrinoless double electron capture (0νECEC) represents the capture-only variant, while quadruple beta decay involves four simultaneous neutron-to-proton transitions (or the inverse), resulting in a change of atomic number by four units. These processes violate lepton number and are sensitive to physics beyond the Standard Model, such as right-handed currents or scalar interactions, distinct from the left-handed currents dominating standard 0νβ⁻β⁻ decay. The rates for these higher-order decays are highly suppressed relative to single beta decay due to the involvement of multiple vertices, scaling with G_F^4 / M_W^4 for double modes and even higher powers for quadruple processes, leading to predicted half-lives exceeding 10^{30} years under typical beyond-Standard-Model scenarios with light masses around 10-50 meV. Theoretical calculations employ matrix elements (NMEs) analogous to those in standard double beta decay, computed using methods like the quasiparticle random-phase approximation (QRPA) or interacting , to estimate transition strengths. For instance, in 0νβ⁺EC decay of ^{48}, NME evaluations indicate comparable magnitudes to β⁻β⁻ modes but with reduced phase space due to the 1.022 MeV per positron-electron pair, yielding half-lives on the order of 10^{28}-10^{32} years. Similarly, for quadruple β decay in candidates like ^{150}, the NMEs are derived from extensions of double beta frameworks, predicting half-lives around 10^{31} years or longer for effective masses in the millielectronvolt range, though light mediators could enhance rates. Experimental searches for these modes remain limited, primarily due to smaller phase spaces, lower isotopic abundances, and challenging detection signatures like X-rays from or positron gammas. The CUORE experiment, using TeO_2 bolometers enriched in ^{120}, has conducted searches for 0νβ⁺EC, setting a lower limit on the of T_{1/2} > 2.9 \times 10^{22} years at 90% confidence level with no observed signal. For 0νβ⁺β⁺, efforts in nuclei like ^{78}Kr and ^{106}Cd yield limits around 10^{21} years from scintillators and low-background detectors. Quadruple beta decay searches, such as in ^{136}Xe by XMASS-I, report limits of T_{1/2} > 3.7 \times 10^{24} years, with no observations, underscoring the extreme rarity. These probes complement standard double beta searches by accessing different BSM parameters, including scalar-mediated interactions that could enhance β⁺ channels or ΔL=4 violations in quadruple modes.

References

  1. [1]
    DOE Explains...Beta Decay - Department of Energy
    One example is two-neutrino double-beta decay. This is a transition inside a nucleus where two neutrons simultaneously undergo beta decay. In other words, two ...
  2. [2]
    [1810.12828] Neutrinoless Double Beta Decay Overview - arXiv
    Oct 30, 2018 · Neutrinoless Double Beta Decay is a hypothesised nuclear process in which two neutrons simultaneously decay into protons with no neutrino emission.<|control11|><|separator|>
  3. [3]
    Neutrinoless Double Beta Decay - Argonne National Laboratory
    Double beta decay is more exotic and occurs in just a few naturally occurring radioactive isotopes. Here, two electrons and two neutrinos are emitted by the ...
  4. [4]
    March 1, 1896: Henri Becquerel Discovers Radioactivity
    Feb 25, 2008 · On an overcast day in March 1896, French physicist Henri Becquerel opened a drawer and discovered spontaneous radioactivity.
  5. [5]
    Otto Hahn, Lise Meitner, and Fritz Strassmann
    In 1938 Hahn, Meitner, and Strassmann became the first to recognize that the uranium atom, when bombarded by neutrons, actually split.Hahn And Meitner Collaborate · Studies In Radioactivity · Nuclear Fission AnnouncedMissing: beta | Show results with:beta<|separator|>
  6. [6]
    This Month in Physics History | American Physical Society
    Physics Hx web ... Wolfgang Pauli first proposed the existence of neutrinos in 1930 while investigating the conundrum of radioactive beta decay, in which some of ...
  7. [7]
    Beta Decay
    Aug 9, 2000 · In beta minus decay, a neutron decays into a proton, an electron, and an antineutrino: n Æ p + e - +. In beta plus decay, a proton decays ...
  8. [8]
    Beta Decay - Q-value | nuclear-power.com
    Beta particles can therefore be emitted with any kinetic energy ranging from 0 to Q. After an alpha or beta decay, the daughter nucleus is often left in an ...
  9. [9]
    Fermi's Theory of Beta Decay | American Journal of Physics
    A complete English translation is given of the classic Enrico Fermi paper on beta decay published in Zeitschrift für Physik in 1934.
  10. [10]
    On the Selection Rules in Beta-Decay | Phys. Rev.
    On the Selection Rules in Beta-Decay. JR Oppenheimer Physics Department, University of California, Berkeley, California. PDF Share.
  11. [11]
    The search for neutrinoless double-beta decay
    Double-beta ( ) decay is a very rare nuclear transition in which a nucleus with Z protons decays into a nucleus with protons and the same mass number A.
  12. [12]
  13. [13]
    Double Beta-Disintegration | Phys. Rev.
    From the Fermi theory of β -disintegration the probability of simultaneous emission of two electrons (and two neutrinos) has been calculated.Missing: prediction | Show results with:prediction
  14. [14]
    On Transition Probabilities in Double Beta-Disintegration | Phys. Rev.
    In the older theory double 𝛽 -disintegration involves the emission of four particles, two electrons (or positrons) and two antineutrinos (or neutrinos), and the ...
  15. [15]
    Double beta decay, Majorana neutrinos, and neutrino mass
    Apr 9, 2008 · Neutrinoless double beta decay [ β β ( 0 ν ) ] is a very slow lepton-number-violating nuclear transition that occurs if neutrinos have mass ( ...Abstract · Article Text · Rate of Double Beta Decay · Double Beta Decay and New...
  16. [16]
    Two-Neutrino Double-Beta Decay - Annual Reviews
    Oct 19, 2013 · Two-neutrino double-β decay is a radioactive process with the longest lifetime ever observed. It has been a subject of experimental research for more than 60 ...Missing: kinematics | Show results with:kinematics
  17. [17]
    Phase Space Factors for Double-Beta Decays - Frontiers
    In this paper we review the progress made in the computation of the phase space factors for double beta decay. We begin with the non-relativistic approaches, ...Missing: kinematics seminal
  18. [18]
    The calculation of double-β decay half-lives with an improved ...
    Aug 9, 2025 · The calculation of double-β decay half-lives with an improved Primakoff-Rosen formula. May 2021; Chinese Physics C 45(8). DOI:10.1088/1674-1137/ ...Missing: seminal | Show results with:seminal
  19. [19]
    [2503.19130] Comprehensive Review of 2$β$ Decay Half-Lives
    Mar 24, 2025 · Double-beta decay was observed, and its half-life was measured in 14 parent nuclei using direct, radiochemical, and geochemical methods.
  20. [20]
    [PDF] Double-β Decay - Particle Data Group
    Half-life 0ν double-β decay. In most cases the transitions (Z,A) → (Z+2,A) + 2e− to the 0+ground state of the final nucleus are listed.
  21. [21]
    [PDF] Improving the Physics Impact of Ge Neutrinoless Double-Beta ...
    Some of the proposed 0νββ-decay experiments have additional physics reach into dark matter, neutrino astron- omy, solar axion, electron lifetime, coherent ...
  22. [22]
    Double Beta Decay, Majorana Neutrinos, and Neutrino Mass - arXiv
    Aug 7, 2007 · Abstract: The theoretical and experimental issues relevant to neutrinoless double-beta decay are reviewed.
  23. [23]
    [1902.04097] Neutrinoless Double-Beta Decay: Status and Prospects
    Feb 11, 2019 · In this review, we summarize the theoretical progress to understand this process, the expectations and implications under various particle ...
  24. [24]
  25. [25]
    [2004.02510] Neutrinoless double beta decay experiment - arXiv
    Apr 6, 2020 · This report is a brief review describing the 0\nu\beta\beta process and its significance. The detector technologies used in the present and ...
  26. [26]
    [PDF] Neutrinoless Double-Beta Decay: Status and Prospects - arXiv
    Feb 11, 2019 · The discussion will focus on the association of 0νββ decay with lepton-number violation and neutrino mass, as well as the mechanisms that could.
  27. [27]
    [PDF] Cosmogenic activation in double beta decay experiments - arXiv
    Oct 21, 2020 · The effect of cosmogenic activation in present and future double beta decay projects based on different types of detectors will be analyzed too.
  28. [28]
    [PDF] Neutrinoless Double Beta Decay
    Background sources for 0νββ experiments. • α background. • Short range (a few MeV), mostly surface events. • β background. • Can be both external and internal ...
  29. [29]
    [PDF] The Majorana Zero-Neutrino Double-Beta Decay Experiment
    Feb 20, 2002 · Radon daughter products are the most commonly-identifiable background source in low- background experiments. While elimination of radon from ...
  30. [30]
    [PDF] Neutrinoless Double Beta Decay with Germanium Detectors - arXiv
    Sep 15, 2021 · Both HDM and IGEX experiments applied pulse shape discrimination. (PSD) analysis to further reduce the background. A lower limit on the half- ...
  31. [31]
    Topological signature in the NEXT high pressure xenon TPC - arXiv
    Oct 10, 2016 · The NEXT experiment aims to observe the neutrinoless double beta decay of Xe-136 in a high-pressure xenon gas TPC using electroluminescence to ...Missing: tracks | Show results with:tracks
  32. [32]
    Latest Results from the Heidelberg-Moscow Double Beta Decay ...
    Mar 6, 2001 · The two neutrino accompanied double beta decay is evaluated for the first time for all five detectors with a statistical significance of 47.7 ...Missing: milestones Kohman Ingelman CUORICINO
  33. [33]
  34. [34]
    Rare decay claim stirs controversy - Physics World
    Feb 11, 2002 · Physicists from the Max Planck Institute for Nuclear Physics in Heidelberg claim to have observed the rarest known nuclear decay for the first ...
  35. [35]
    [2503.24137] Half-life and precision shape measurement of 2νββ ...
    Mar 31, 2025 · We present a new measurement of the 2nbb half-life of 130Te (T1/2) using the first complete model of the CUORE data, based on 1038 kg yr of collected exposure.
  36. [36]
    First results from LEGEND-200 on the search for neutrinoless ... - arXiv
    Sep 25, 2025 · The performed statistical analysis finds no evidence for a 0\nu\beta\beta signal and sets a lower limit on its half life to T_{1/2}^{0\nu} > 0.5 ...Missing: 0νββ | Show results with:0νββ
  37. [37]
    New limits on neutrinoless double beta decay in tellurium from ...
    Oct 17, 2025 · New limits on neutrinoless double beta decay in tellurium from CUORE experiment. 17 October 2025. New results from the CUORE (Cryogenic ...
  38. [38]
    [PDF] Status and future of nuclear matrix elements for neutrinoless double ...
    In single-β decay QRPA excitations of the initial state produce final states. In ββ decay, they produce intermediate states in equations (7) and (14) and one ...
  39. [39]
    Ab Initio Uncertainty Quantification of Neutrinoless Double-Beta ...
    We report the first comprehensive ab initio uncertainty quantification of the -decay NME, in the key nucleus.
  40. [40]
  41. [41]
    None
    ### Summary of Right-Handed Parameters λ and η in Neutrinoless Double-Beta Decay
  42. [42]
    [PDF] Neutrinoless Double Beta Decay and Physics Beyond The Standard ...
    Neutrinoless double beta decay is a sensitive probe of new physics beyond the standard model. In this review, we begin by describing the various mechanisms.
  43. [43]
    [PDF] arXiv:1505.06121v2 [hep-ph] 23 Jul 2015
    Jul 23, 2015 · The short-range part of the 0νββ decay rate [2] is generated via the exchange of exotic particles, such as leptoquarks and/or diquarks plus ...
  44. [44]
    Searching for New Physics in Two-Neutrino Double Beta Decay
    Oct 19, 2020 · In this Letter, we show how the presence of right-handed leptonic currents would affect the energy distribution and angular correlation of the outgoing ...
  45. [45]
    [PDF] Ion-tagging in neutrinoless double beta decay experiments
    Jun 22, 2023 · nEXO Signal and Background. • nEXO measures multiple parameters for each event to be able to robustly identify a 0νββ signal.Missing: future m_ee
  46. [46]
    [PDF] arXiv:2304.07198v1 [hep-ex] 14 Apr 2023
    Apr 14, 2023 · The detection of an excess of 130Xe was proof of the double-β decay of the initial nucleus and allowed a first determination of its half-life [4] ...
  47. [47]
    [2007.14908] Neutrinoless Double-Electron Capture - arXiv
    Jul 29, 2020 · An unprecedented precision in the determination of decay energies with Penning traps has allowed one to refine the values of the degeneracy ...
  48. [48]
    A Systematic Study of Two-Neutrino Double Electron Capture - MDPI
    Oct 2, 1997 · In this paper, we update the phase-space factors for all two-neutrino double electron capture processes.
  49. [49]
    [PDF] Indications of 2ν2K capture in 78Kr - INR RAS
    Mar 4, 2013 · The half-life of 78Kr relative to 2ν2K capture is [9.2 +5.5 −2.6(stat) ± 1.3(syst)] × 10^21 y. The statistical significance is 2.5σ.Missing: IGEX | Show results with:IGEX
  50. [50]
    COBRA—double beta decay searches using CdTe detectors
    Oct 25, 2001 · A new approach (called COBRA) for investigating double beta decay using CdTe semiconductor detectors is proposed.
  51. [51]
    New Limits on Double Electron Capture of $^{40}$Ca and $^{180}$W
    Apr 28, 2016 · We established new limits for ^{40}Ca with T_{1/2}^{2v2K}>9.9\times10^{21} y and T_{1/2}^{0v2EC}>1.4\times10^{22} y and for ^{180}W with T_{1/2 ...Missing: 2023 | Show results with:2023
  52. [52]
    Calculation of double beta plus decay matrix elements - ScienceDirect
    Half lives for neutrino accompanied double positron emission (2νβ+β+), positron emission/electron capture (2νβ+/EC) and double electron capture (2νEC/EC) ...
  53. [53]
    [1306.0580] Neutrinoless Quadruple Beta Decay - arXiv
    Jun 3, 2013 · As a consequence, neutrinoless double beta decay is forbidden, but neutrinoless quadruple beta decay is possible: (A,Z) \to (A,Z+4) + 4 e^-. We ...
  54. [54]
    Neutrinoless Double β+/EC Decays - Maalampi - Wiley Online Library
    Dec 1, 2013 · The 0νβ+EC decay to the ground state of 124Te is the fastest with a half-life of (1.2–4.2) × 1027 years, being in the range of the corresponding ...
  55. [55]
    Search for Neutrinoless $β^+EC$ Decay of $^{120}$Te with CUORE
    Mar 16, 2022 · Bayesian lower limit of 2.9 \cdot 10^{22} yr on ^{120}Te half-life. This result improves by an order of magnitude the existing limit from the ...
  56. [56]
    Search for neutrinoless quadruple beta decay of 136 Xe in XMASS-I
    Oct 10, 2022 · The XMASS-I experiment searched for neutrinoless quadruple beta decay of 136 Xe, setting a lower limit of 3.7 x 10^24 years for its half-life. ...
  57. [57]
    [PDF] Double-β Decay - Particle Data Group
    However, we also list transitions that increase the nuclear charge (2e+, e+/EC and ECEC) and transitions to excited states of the final nuclei (0+i , 2+, and 2+ ...