Next Article in Journal
Interaction of Protons with Noble-Gas Atoms: Total and Differential Cross Sections
Previous Article in Journal
Atomic Models of Dense Plasmas, Applications, and Current Challenges
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

External Radiation Assistance of Neutrinoless Double Electron Capture

by
Vladimir N. Kondratyev
1,* and
Feodor F. Karpeshin
2
1
Bogolubov Laboratory of Theoretical Physics, JINR, 141980-RU Dubna, Russia
2
D.I. Mendeleyev Institute for Metrology, 190005-RU Saint-Petersburg, Russia
*
Author to whom correspondence should be addressed.
Atoms 2024, 12(5), 27; https://doi.org/10.3390/atoms12050027
Submission received: 3 February 2024 / Revised: 19 April 2024 / Accepted: 24 April 2024 / Published: 28 April 2024
(This article belongs to the Section Nuclear Theory and Experiments)

Abstract

:
The influence of electromagnetic radiation on nuclear processes is applied to an example of a neutrinoless double electron capture (0ν2ec). For cases with X-ray free-electron lasers (X-ray FELs) and/or inverse Compton X-ray sources, it was shown that such a decay can be significantly enhanced by tuning the system to the resonant conditions through the absorption and/or emission of a photon with the decay resonance defect energy Δ. In this case, the 0v2ec decay rate Γ2e of nuclide Z grew linearly with field intensity (S/Sz) up to the X-ray flux power Sm~Z6, while Sz~Z6 (Γ/Δ)2 with decay width Γ of a daughter atom. For the case of 78Kr → 78Se − 0ν2eL1L1 capture we find Sz~109 W cm−2 and Sm~1017 W cm−2 which indicate a possibility of increasing decay rate to eight orders of magnitude or even larger.

Graphical Abstract

1. Introduction

The evidence for physics beyond the Standard Model (SM) includes recent discoveries such as neutrino oscillations, adiabatic lepton flavor transformations and neutrino masses [1,2,3,4,5]. The direct hard probe for the lepton number violation and the Majorana nature of neutrinos [6] is provided by the neutrinoless double-beta (0ν2β) decay [7] and/or double electron capture (0ν2ec) processes that are forbidden within the SM. The breaking of global SM laws is illustrated by such phenomena that compel searches for experimental evidence and theoretical outlines of these effects. The double-beta (2β) decay represents an isobaric transition from a parent nucleus (A,Z) to a daughter nucleus (A,Z + 2), adding two charges. In the two-neutrino, double-beta (2ν2β) decay mode, (A, Z) → (A, Z + 2) + 2 e + 2 νe + Qββ, with the energy released represented by the Q-value Qββ. These SM-allowed, second-order weak decays correspond to a typical half-life of >1019 years.
The SM-forbidden 0ν2β decay (A, Z) → (A, Z + 2)++ + e + e, with a doubly ionized atom (A, Z + 2)++ in the final state, was first considered by Furry [8]. In this process of 0ν2β decay, the nucleus with mass number A and charge Z is accompanied by the Majorana neutrino exchange between the nucleons (see Figure 1a in which Z changes by Z + 2 and Z − 2 changes by Z). There are many models beyond the SM that provide alternative mechanisms of te 0ν2β decay, e.g., [9,10] and refs. therein.
Winter suggested [11] a related 0ν2ec process, e b + e b + (A, Z) → (A, Z − 2)∗∗, with bound electrons e b , excited states of the daughter nucleus and the electron shells of the neutral atom (A, Z − 2)∗∗. An example of a mechanism related to the Majorana neutrino exchange is shown in Figure 1a. Subsequent de-excitation of the daughter nucleus (A, Z − 2) ∗∗ proceeds via the Auger electron and/or gamma ray emission cascades with filling electronic shell vacancies.
The 0ν2ec process is several orders of magnitude less sensitive to the Majorana neutrino mass than the 0ν2β decay. Winter pointed out [12] that the energy degeneracy of parent (A, Z) and excited daughter (A, Z − 2)∗∗ atoms gives rise to resonant enhancement of the decay. Although such a possibility of resonant behavior was confirmed by numerous studies of the 0ν2ec process, e.g., [13,14], the resonance-like coincidence in 2ec is considered, as a doubtful option. Given the decay resonance defect ∆ = MA,Z M A , Z 2 * * , the difference in mass between the parent and daughter atoms in the decay probability is determined by the Breit–Wigner factor (see Section 2.1).
The near-resonant 0ν2ec process was analyzed in detail in ref. [15]. The authors developed a non-relativistic formalism of the resonant 0ν2ec in atoms and specified a dozen nuclide pairs for which degeneracy is not excluded. Detailed theoretical predictions for a list of near-resonant 0ν2ec nuclide pairs are provided in refs. [16,17,18]. Evidently, the proper fit of the difference between the masses of the parent and daughter atoms, i.e., Q-values, and the transition energy can result in a significant enhancement of the decay rate. Possibilities for an additional tuning of the degeneracy parameter ∆ have fundamental importance.
The manipulation of nuclear processes using irradiation from powerful X-ray sources, e.g., inverse Compton scattering X-ray sources [19,20], X-ray free-electron lasers (FELs) [21,22,23] and/or others [24], represents an extremely intriguing task of modern physics. For the purposes of these applications, the respective solutions can be used to trigger the release of enormous amounts of energy contained in nuclear isomers. In terms of fundamental research, this task is also of principal interest. The implementation of such a method is possible only through the involvement of the electronic shell, since the field of the strongest lasers is much smaller than the Coulomb field of the nucleus near its surface. The investigation of this fundamental process is of great interest from the point of view of understanding the nature of neutrinos. The discovery of a phenomenon such as 0ν2ec would unambiguously indicate the Majorana nature of neutrinos (e.g., [9,10,11,12,13,14,15,16,17,18]). However, these investigations encounter great difficulties due to the very long lifetimes of the samples, so it would be difficult to overestimate the possibility of radically accelerating this process by several orders of magnitude.
This work aims to analyse potential tools to adjust the decay resonance defect ∆ to the correct zero value by applying an external field. In particular, we consider the effect of strong X-ray pulses that are relevant for inverse Compton scattering X-ray sources [19,20], X-ray FELs [21,22,23] and/or others [24].

2. Environmental Effect of the Neutrinoless Double Electron Capture (0ν2ec) Process

Figure 1b represents a schematic diagram of the neutrinoless double electron capture (0ν2ec) decay, which accounts for coupling to external radiation or the environment. Such an electromagnetic interaction is shown by the line γ, indicating a real or virtual photon. Hereafter, we consider external X-ray radiation.

2.1. Resonant Character of the Neutrinoless Double Electron Capture (0ν2ec)

In the case of 0ν2e capture, the atom remains generally neutral. Therefore, the energy release is determined by the difference in the masses of neutral atoms, the initial M1 and the daughter M2, including its excitation energy EA, as given by
Q = M1M2EA.
Let us recall the formula for the traditional resonant fluorescent mechanism corresponding to the pole approximation. The capture of 0ν2ec leads to the formation of a doorway state. Due to the absence of degeneracy, it is located outside the mass surface, since its energy differs from the value EA by the magnitude of the resonance defect Δ = Q. Due to the uncertainty principle, a violation of the law of conservation of energy is possible for a time τ = 1/Q; hereafter, we use units with ћ = 1 to indicate otherwise. Then, a fluorescent or Auger transition occurs, resulting in a return of the atom to the mass surface. The energy of Δ is added to the usual energy of this transition. This whole sequence of transformations forms a single, fast process. Formally, it is described by multiplying the square of the amplitude of the capture itself, Γ 2 e , by the Breit–Wigner resonance factor BW:
Γ 2 e i = Γ 2 e   B W ,   B W = Γ / 2 π Δ 2 + Γ 2 / 4 .
In Equation (2), Γ denotes the sum of the widths of the doorway state of a daughter atom with two holes before and after its decay. As seen in Table 1, for typical cases, the value Δ varies from a few keV to hundreds keV, while the Breit–Wigner factor drops six or seven orders of magnitude. Here we draw attention to the way in which we can compensate the decay resonance defect ∆.
Γ 2 e = 2 π |   F 2 e   | 2 ,   F 2 e = ψ 2 H 2 e ψ 1 .
Here, H 2 e and F 2 e are the Hamiltonian and matrix elements of the double electron capture.

2.2. Resonance Tuning Using X-ray Radiation

The method of compensation of the resonance defect in the field of an intense electromagnetic wave was proposed in ref. [32] and described in detail in refs. [33,34]. Such a tool is completely general and can be applied in cases of both positive and negative values of Δ. The excess energy of the decay can be transferred to the external field, while the missing energy can be drawn from the radiation. Such a method is particularly effective in the presence of an electric dipole junction in the system, preferably close to resonance.
In more detail, we consider the example of 0ν2eL1L1 capture in 78Kr at the 2+ 2838.49 keV level 78Se (see Table 1). In an experiment by Gavrilyuk et al. [28], which gives an example of the calorimetric approach, an indication of the 2ν2eK capture in 78Kr was obtained with a proportional chamber filled with enriched 78Kr. This result was confirmed recently [29] with better statistical accuracy in a study that showed that the half-life of T 1 / 2 2 ν 2 K = 1.9 0.7 1.3 s t a t ± 0.3 s y s t × 10 22 yr. A diagram of the atomic levels is shown in Figure 2. It is convenient to measure the energy using the value for the selenium atom, assuming that the capture occurs in the ground state of the atomic shell according to Equation (1) (see Figure 2). In this case, we obtain Q = 9.15 keV. Two remaining L1 holes, according to Figure 2, reduce this value by 3.32 keV to Qeff = 5.84 keV. Therefore, the resonance defect Δ = Qeff = 5.84 keV (see Table 1). The state 2p3/2 in the wave field acquires satellites in the wave function equidistant from the main level by an amount ± l, n = 1, 2, 3, …, and the amplitudes decrease with the growth of n. For our purpose, the component with n = 1 is essential. In this approximation, the L1 level forms two satellites, L 1 ( + ) and L 1 ( ) , with the L3 state, with energies E L 3 ± ωl, respectively. Here, E L 3 denotes the energy of the L3 level or the 2p3/2 state (Figure 2). At a photon energy of ωl = 6.06 keV, it falls exactly into resonance with the ground level of the parent atom, as shown in Figure 2. Evidently, such experimental conditions solve the tuning problem, i.e., the Breit–Wigner factor turns into one, when integrated over energy. The gain reaches a unitary maximum. We estimate the required radiation power.
The amplitude of the admixture of electronic states in a field reads
β = eE ⟨2p|r|2s⟩/δ,
where E denotes the intensity vector of the electric field of the laser, e is the charge of the electron, r represents the spatial coordinate vector, and |2s⟩ and |2p⟩ indicate eigenfunctions for the 2s and 2p states, respectively. Here, δ is equal to the difference between the energy of the additional component we need and the respective component itself.
Taking into account the admixture value in Equation (4), the wave function of the L3 electron in the exit channel (state after capture) can be written as
ψ 2 = ( ψ L 3 + β   ψ L 1   exp { i ω l t } )   exp { i   E L 3 h   t } ,
where E L 3 h is the energy of the L3 holes. It is worth noting here that this linear approximation for state mixing corresponds to the condition β < 1. These limits define the upper values of the electromagnetic field strength and the photon flux intensity power Sm for reliable implementation of the considered approach. Given that in a hydrogen-like model with nuclide effective charge Z, the dependence for β~Z−3, and the photon flux power linear approximation limit Sm grows with Z as Sm~Z6.
Substituting the full wave function into Equation (5), taking into account the time factor in Equation (3) and leaving only the component L1, we obtain
F 2 e ( x ) = ψ 2 H 2 e ψ 1   exp { i   ( E 1 E 2 ω l )   t } .
The index 2 here is associated with the state ψL1. Then, squaring the amplitude in Equation (3) in the usual way (e.g., [35]), we move on to consider the probability of the process per unit of time:
Γ 2 e x ( ω l ) = β 2   Γ 2 e   Γ / 2 π ( E 1 E 2 ω l ) 2 + Γ 2 / 4   δ ( E 1 E 2 ω l )
Taking into account Equation (7), we write the law of energy conservation in the form
M1 = M2EL1 + EL3 + ωr,
which provide conditions for the resonant frequency of the X-ray (laser) irradiation:
ωr = Δ + EL1EL3.
Thus, in the electromagnetic field, the capture of the second electron still occurs from the 2s subshell, but not just a hole is formed in its orbit: a hole mixed with the 2p3/2 same hole state, with energy EL3 + ωl, was formed. Under the condition given by Equation (9), this energy just resonates with the mass of the parent atom. Accordingly, from Equation (7), we obtain
Γ 2 e x = Γ 2 e Γ / 2 π ( ω l ω r ) 2 + Γ 2 / 4 S ( ω l )   d ω l ,
where S(E) is the spectral density of the X-ray source or laser radiation. Integrating Equation (10) over the profile of the laser line, and assuming that the spectral width is close to Γ (S(E)~1/Γ), we write the decay rate in an electromagnetic field as
Γ 2 e x = Γ 2 e β 2   Γ .
Incorporating the ratio in Equation (11) into Equation (2), we obtain a gain in the probability of this process in the X-ray radiation field in the following form:
G = Γ 2 e x Γ 2 e = 2 π   N   β 2 ( Δ Γ ) 2
where N = 8 is determined by a statistical weight that takes into account the presence of two 2s holes and four 2p3/2 electrons.
From the condition G ≈ 1, we can estimate the value of β at which the probability of stimulated decay is compared with the natural one: βx ≈ 10−4. Bearing in mind the parameters of the gamma factory at CERN [25] for estimation, we set the photon flux power with energy ωl = 5 keV equal to 4 MW cm−2. This power corresponds to a field strength of 60 kV/cm. The calculated value of the matrix element ⟨2p|r|2s⟩ ≈ 22 keV−1 ≈ 4 × 10−7 s. From Figure 2 on the right (b), we find that Δ ≈ 5.84 keV. Using Equation (4), we estimate the real admixture that occurs in such a field: β ≈ 0.5 × 10−5. Thus, the entire order of intensity is missing, or three orders of magnitude in the power of the gamma ray flux of Swiss X-ray FELs.

3. Discussion

We analyse the effect of an intense electromagnetic field relevant for X-ray FELs, inverse Compton X-ray sources, etc., in the 0ν2ec radioactive decay process. As was illustrated, the emission and/or absorption of X-ray photons can result in significant increases in the decay rate. Such strengthening of the beta-decay channel associated with the double electron capture process originates from tuning the resonance defect Δ to zero and the system to the resonance conditions. A crucial requirement for an efficient photon absorption/emission is associated with a significant dipole coupling of substates in the electronic shell. The L-shell is well suited for this case. As was seen in Equations (4) and (5) and the discussion therein, the treatment of state mixing within perturbation theory is well justified up to an X-ray flux intensity of Sm ≈ Z6 × 108 W cm−2. We employ here a hydrogen-like description of electronic states with an effective charge Z, which is rather reliable for the internal shells. In this range, the decay rate grows linearly with the electromagnetic field intensity G ≈ S/Sz with SzZ6 (Γ/Δ)2 × 105.5 W cm−2. At the upper limit of the application of perturbation theory with Sm, the enhancement factor is GmSm/Sz ≈ (Δ/Γ)2 × 102.5. Such a feature shows that in the case of relatively weak X-ray flux power, the maximum factor for an increasing 0ν2ec decay rate is determined by the ratio Δ/Γ and amounts to about eight orders of magnitude.
To this end, let us estimate the rate of induced 0ν2eL1L1 capture in 78Kr at the 2+ 2838.49 keV level 78S when the sample is irradiated by the X-ray flux with resonance frequency and an effective intensity Sef close to the value Sm. The total number of decays per unit time is dN/dtNn Гef, with the total number of atoms in a sample Nn and an effective X-ray assisted transition width Гef = Sef Г2e ≈ Гres ≈ 10−20 yr−1. For a case in which Nn ≈ 1028.5, i.e., 3 tons of the material (e.g., the Gran Sasso experiment), the decay rate rises to 10 per second. Alternatively, to enhance the effect, one can imagine an ampoule of liquid 78Kr, 1 m long with a 1 cm2 cross-section, positioned along the beam. In this case, the counting rate would be four orders of magnitude lower, which, however, is quite sufficient for experimental observations. Evidently, additional evidence for the considered events are provided by subsequent Г decays of the daughter nucleus and radiative transitions due to relaxation of the L-shell holes. Such an accompanying emission of photons and Auger electrons is very favorable for detection and ensures a registration probability of tens percent.
It is worth noting that although the intensity electromagnetic field, Sm, considered here is much less than the upper power limit due to, e.g., the vacuum short circuit, the present intensities of gamma ray sources are still very small compared to the Sm value. However, very strong X-ray fluxes arise during astrophysical phenomena, e.g., supernova explosions, which have an important role in neutrino nuclear reactions [36].

4. Conclusions

In this study, we performed evaluations of the electromagnetic field effects for the possible acceleration of the neutrinoless double electron capture process. Unfortunately, the example discussed above suggests that the power of the gamma factory flow is not yet sufficient to radically manipulate this process. In conjunction with the present status of X-ray source intensities, we employed perturbation theory for the evaluation of the respective effects of the decay rate acceleration. As a result, we found that the decay rate grew linearly with increasing X-ray flux intensity up to ten orders of magnitude. Evidently, the design of X-ray fabrics such as X-ray FELs, inverse Compton X-ray sources, etc., has developed rapidly over the last two decades (see refs. [19,20,21,22,23,24]), which gives hope for a reassessment of possibilities in the future. The results obtained above should be perceived as an assessment of the parameters necessary to achieve the goals stated in this work.

Author Contributions

Conceptualization, V.N.K. and F.F.K.; methodology, V.N.K. and F.F.K.; software, V.N.K. and F.F.K.; validation, V.N.K. and F.F.K.; formal analysis, V.N.K. and F.F.K.; investigation, V.N.K. and F.F.K.; data curation, V.N.K. and F.F.K.; writing—original draft preparation, V.N.K. and F.F.K.; writing—review and editing, V.N.K. and F.F.K.; visualization, V.N.K. and F.F.K.; supervision, F.F.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Available on a request.

Acknowledgments

The authors are grateful to D. Budker for useful discussions of the available experimental installations with intense sources of electromagnetic fields.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Fukuda, Y.; Hayakawa, T.; Ichihara, E.; Inoue, K.; Ishihara, K.; Ishino, H.; Itow, Y.; Kajita, T.; Kameda, J.; Kasuga, S.; et al. Evidence for Oscillation of Atmospheric Neutrinos. Phys. Rev. Lett. 1998, 81, 1562. [Google Scholar] [CrossRef]
  2. Ahmad, Q.R.; Allen, R.C.; Andersen, T.C.; Anglin, J.D.; Barton, J.C.; Beier, E.W.; Bercovitch, M.; Bigu, J.; Biller, S.D.; Black, R.A.; et al. Direct Evidence for Neutrino Flavor Transformation from Neutral-Current Interactions in the Sudbury Neutrino Observatory. Phys. Rev. Lett. 2002, 89, 011301. [Google Scholar] [CrossRef]
  3. Eguchi, K.; Enomoto, S.; Furuno, K.; Goldman, J.; Hanada, H.; Ikeda, H.; Ikeda, K.; Inoue, K.; Ishihara, K.; Itoh, W.; et al. First Results from KamLAND: Evidence for Reactor Antineutrino Disappearance. Phys. Rev. Lett. 2003, 90, 021802. [Google Scholar] [CrossRef] [PubMed]
  4. Kajita, T. Nobel Lecture: Discovery of atmospheric neutrino oscillations. Rev. Mod. Phys. 2016, 88, 030501. [Google Scholar] [CrossRef]
  5. McDonald, A.B. Nobel Lecture: The Sudbury Neutrino Observatory: Observation of flavor change for solar neutrinos *. Rev. Mod. Phys. 2016, 88, 030502. [Google Scholar] [CrossRef]
  6. Majorana, E. Teoria simmetrica dell’elettrone e del positrone. Nuovo Cim. 1937, 14, 171–184. [Google Scholar] [CrossRef]
  7. Schechter, J.; Valle, J.W.F. Neutrinoless Double beta Decay in SU(2)×U(1) Theories. Phys. Rev. D 1982, 25, 2951. [Google Scholar] [CrossRef]
  8. Furry, W.H. On Transition Probabilities in Double Beta-Disintegration. Phys. Rev. 1939, 56, 1184–1193. [Google Scholar] [CrossRef]
  9. Dolinski, M.J.; Poon, A.W.; Rodejohann, W. Neutrinoless double-beta decay: Status and prospects. Annu. Rev. Nucl. Part. Sci. 2019, 69, 219–251. [Google Scholar] [CrossRef]
  10. Blaum, K.; Eliseev, S.; Danevich, F.A.; Tretyak, V.I.; Kovalenko, S.; Krivoruchenko, M.I.; Novikov, Y.N.; Suhonen, J. Neutrinoless double-electron capture. Rev. Mod. Phys. 2020, 92, 045007. [Google Scholar] [CrossRef]
  11. Winter, R.G.; Double, K. Capture and Single K Capture with Positron Emission. Phys. Rev. 1955, 100, 142–144. [Google Scholar] [CrossRef]
  12. Winter, R.G. Search for Double Beta Decay in Cadmium and Molybdenum. Phys. Rev. 1955, 99, 88. [Google Scholar] [CrossRef]
  13. Georgi, H.M.; Glashow, S.L.; Nussinov, S. Unconventional model of neutrino masses. Nucl. Phys. B 1981, 193, 297–316. [Google Scholar] [CrossRef]
  14. Voloshin, M.B.; Mitsel’makher, G.V.; Eramzhyan, R.A. Conversion of an atomic electron into a positron and double β + decay. JETP Lett. 1982, 35, 656–659. [Google Scholar]
  15. Bernabeu, J.; De Rujula, A.; Jarlskog, C. Neutrinoless double electron capture as a tool to measure the electron neutrino mass. Nucl. Phys. B 1983, 223, 15–28. [Google Scholar] [CrossRef]
  16. Sujkowski, Z.; Wycech, S. Neutrinoless double electron capture: A tool to search for Majorana neutrino. Phys. Rev. C 2004, 70, 052501. [Google Scholar] [CrossRef]
  17. Karpeshin, F.F. On neutrinoless double e capture problem. Phys. Part. Nucl. Lett. 2008, 5, 379–382. [Google Scholar] [CrossRef]
  18. Karpeshin, F.F.; Trzhaskovskaya, M.B. Shake-off in the 164Er neutrinoless double electronic capture and the dark matter puzzle. Phys. Rev. C 2023, 107, 045502. [Google Scholar] [CrossRef]
  19. Brümmer, T.; Bohlen, S.; Grüner, F.; Osterhoff, J.; Põder, K. Compact all-optical precision-tunable narrowband hard Compton X-ray source. Sci. Rep. 2022, 12, 16017. [Google Scholar] [CrossRef] [PubMed]
  20. Zheng, X. Inverse Compton gamma-ray source driven by a plasma flying mirror. JOSA B 2023, 4, 3262–3268. [Google Scholar] [CrossRef]
  21. Geloni, G.; Saldin, E.; Samoylova, L.; Schneidmiller, E.; Sinn, H.; Tschentscher, T.; Yurkov, M. Coherence properties of the European XFEL. New J. Phys. 2010, 12, 035021. [Google Scholar] [CrossRef]
  22. Milne, C.J.; Schietinger, T.; Aiba, M.; Alarcon, A.; Alex, J.; Anghel, A.; Arsov, V.; Beard, C.; Beaud, P.; Bettoni, S.; et al. SwissFEL: The Swiss X-ray free electron laser. Appl. Sci. 2017, 7, 720. [Google Scholar] [CrossRef]
  23. Duris, J.; Li, S.; Driver, T.; Champenois, E.G.; MacArthur, J.P.; Lutman, A.A.; Zhang, Z.; Rosenberger, P.; Aldrich, J.W.; Coffee, R.; et al. Tunable isolated attosecond X-ray pulses with gigawatt peak power from a free-electron laser. Nat. Photonics 2020, 14, 30–36. [Google Scholar] [CrossRef]
  24. Budker, D.; Berengut, J.C.; Flambaum, V.V.; Gorchtein, M.; Jin, J.; Karbstein, F.; Krasny, M.W.; Litvinov, Y.A.; Pálffy, A.; Pascalutsa, V.; et al. Expanding Nuclear Physics Horizons with the Gamma Factory (Ann. Phys. 3/2022). Ann. Phys. 2022, 534, 2270005. [Google Scholar] [CrossRef]
  25. Agostini, M.; GERDA Collaboration; Allardt, M.; Bakalyarov, A.M.; Balata, M.; Barabanov, I.; Barros, N.; Baudis, L.; Bauer, C.; Bellotti, E.; et al. Limit on the radiative neutrinoless double electron capture of 36Ar from GERDA Phase I. Eur. Phys. J. C 2016, 76, 652. [Google Scholar] [CrossRef]
  26. Angloher, G.; Bauer, M.; Bauer, P.; Bavykina, I.; Bento, A.; Bucci, C.; Canonica, L.; Ciemniak, C.; Defay, X.; Deuter, G.; et al. New limits on double electron capture of 40Ca and 180W. J. Phys. G 2016, 43, 095202. [Google Scholar] [CrossRef]
  27. Bikit, I.; Krmar, M.; Slivka, J.; Vesković, M.; Čonkić, L.; Aničin, I. New results on the double β decay of iron. Phys. Rev. C 1998, 58, 2566. [Google Scholar] [CrossRef]
  28. Gavrilyuk, Y.M.; Gangapshev, A.M.; Kazalov, V.V.; Kuzminov, V.V.; Panasenko, S.I.; Ratkevich, S.S. Indications of 2ν2K capture in 78Kr. Phys. Rev. C 2013, 87, 035501. [Google Scholar] [CrossRef]
  29. Ratkevich, S.S.; Gangapshev, A.M.; Gavrilyuk, Y.M.; Karpeshin, F.F.; Kazalov, V.V.; Kuzminov, V.V.; Yakimenko, S.P. Comparative study of the double-K-shell-vacancy production in single-and double-electron-capture decay. Phys. Rev. C 2017, 96, 065502. [Google Scholar] [CrossRef]
  30. Bukhner, E.; Vishnevskij, I.N.; Danevich, F.A. Rare decays of mercury nuclei. Sov. J. Nucl. Phys. 1990, 52, 193–197. [Google Scholar]
  31. Meshik, A.P.; Hohenberg, C.M.; Pravdivtseva, O.V.; Kapusta, Y.S. Weak decay of 130Ba and 132Ba: Geochemical measurements. Phys. Rev. C 2001, 64, 035205. [Google Scholar] [CrossRef]
  32. Zon, B.A.; Karpeshin, F.F. Acceleration of the decay of 235mU by laser-induced resonant internal conversion. Sov. Phys. JETP 1990, 70, 224. [Google Scholar]
  33. Karpeshin, F.F. Instantaneous Fission in Muonic Atoms and Resonance Conversion; Nauka: St. Petersburg, Russia, 2006. (In Russian) [Google Scholar]
  34. Karpeshin, F.F. Resonance Internal Conversion as the way of accelerating nuclear processes. Phys. Part. Nucl. 2006, 37, 522. [Google Scholar] [CrossRef]
  35. Landau, L.D.; Lifshitz, E.M. Quantum Mechanics: Non-Relativistic Theory. In Course of Theoretical Physics; Pergamon: New York, NY, USA, 1977; Volume 3. [Google Scholar]
  36. Kondratyev, V.N. Magnetorotational supernova neutrino emission spectra and prospects for observations by large-size underwater telescopes. Phys. At. Nucl. 2023, 86, 1083–1089. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of neutrinoless double electron capture (a) for an isolated nuclide and (b) under the influence of X-ray flux.
Figure 1. Schematic representation of neutrinoless double electron capture (a) for an isolated nuclide and (b) under the influence of X-ray flux.
Atoms 12 00027 g001
Figure 2. (a, left) Atomic level diagram of 78Se (in keV). (b, right) Compensation circuit for a resonance defect in the laser field. The energies of the two-hole states shown in the figure on the right are set equal to the sum of the energies of each of the holes shown in panel a. Energies are shown in keV.
Figure 2. (a, left) Atomic level diagram of 78Se (in keV). (b, right) Compensation circuit for a resonance defect in the laser field. The energies of the two-hole states shown in the figure on the right are set equal to the sum of the energies of each of the holes shown in panel a. Energies are shown in keV.
Atoms 12 00027 g002
Table 1. Experimental data for nuclides associated with 0ν2ec decay. The resonance defect Δ and the de-excitation width of the daughter nuclide electron shells Γ are relevant for the possible resonant enhancement value Δ/Γ.
Table 1. Experimental data for nuclides associated with 0ν2ec decay. The resonance defect Δ and the de-excitation width of the daughter nuclide electron shells Γ are relevant for the possible resonant enhancement value Δ/Γ.
Transition [Ref.]Δ (keV)Γ (eV)Δ/Γ 103
36Ar→36S [25]427.651.04411.2
40Ca→40Ar [26] 187.101.32141.74
50Cr→50Ti [27] 1159.71.78651.5
78Kr→78Se [28,29] 5.837.60.773
196Hg→196Pt [30] 661.8 ± 3.0996.7
37.6 ± 3.058.30.645
132Ba→132Xe [31] 774.823.033.7
54Fe→54Cr [27]668.32.04327.6
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kondratyev, V.N.; Karpeshin, F.F. External Radiation Assistance of Neutrinoless Double Electron Capture. Atoms 2024, 12, 27. https://doi.org/10.3390/atoms12050027

AMA Style

Kondratyev VN, Karpeshin FF. External Radiation Assistance of Neutrinoless Double Electron Capture. Atoms. 2024; 12(5):27. https://doi.org/10.3390/atoms12050027

Chicago/Turabian Style

Kondratyev, Vladimir N., and Feodor F. Karpeshin. 2024. "External Radiation Assistance of Neutrinoless Double Electron Capture" Atoms 12, no. 5: 27. https://doi.org/10.3390/atoms12050027

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop