Investigation of decay modes of superheavy nuclei

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

Investigation of decay modes of superheavy nuclei

H.C. Manjunatha
N. Sowmya
P.S. Damodara Gupta
K.N. Sridhar
A.M. Nagaraja
L. Seenappa
S. Alfred Cecil Raj
Nuclear Science and TechniquesVol.32, No.11Article number 130Published in print 01 Nov 2021Available online 23 Nov 2021
6800

A detailed investigations of different decay modes, namely, alpha decay, beta decay, cluster decay, including heavy particle emission (Zc > 28), and spontaneous fission, was carried out, leading to the identification of new cluster and beta-plus emitters in superheavy nuclei with 104 ≤ Z ≤ 126. For the first time, we identified around 20 beta-plus emitters in superheavy nuclei. Heavy-particle radioactivity was observed in superheavy elements of atomic number in the range 116 ≤ Z ≤ 126. 292-293Og were identified as 86Kr emitters, and 298122 and 300122 were identified as 94Zr emitters, whereas heavy-particle radioactivity from 91Y was also observed in 299123. Furthermore, the nuclei 300124 and 306126 exhibit 96Mo radioactivity. The reported regions of beta-plus and heavy-particle radioactivity for superheavy nuclei are stronger than those for alpha decay. The identified decay modes for superheavy nuclei are presented in a chart. This study is intended to serve as a reference for identifying possible decay modes in the superheavy region.

Alpha decayBeta decayHeavy-particle radioactivityBranching ratios
1

Introduction

The most important unanswered questions in Nuclear Physics are to determine the heaviest superheavy nuclei that can exist, and to investigate whether very-long-lived superheavy nuclei exist in nature. The past ten years have been marked by remarkable progress in the science of superheavy elements and nuclei. The existence of superheavy nuclei above Z=103 can be studied in terms of whether they can occur naturally or can be synthesized in the laboratory. There are no definitive conclusions regarding the existence of superheavy nuclei in nature. In contrast, such superheavy nuclei, with half-lives ranging between days to μs, can be synthesized using cold and hot fusion reactions. Cold fusion reactions involve either lead or bismuth as targets [1], whereas hot fusion reactions include 48Ca beams on various actinide targets. [2,3]. Many theoretical predictions, such as microscopic–macroscopic [4] (single-particle potential) and self-consistent approaches, including nucleus–nucleus potential [5, 6], relativistic field models [7, 8], and multinucleon transfer reactions [9], provide information regarding the nucleus structure, shell closure location, and decay modes in heavy and superheavy nuclei.

The discovery of superheavy elements [10, 11] points to the island of stability. Boilley et al. [12] predicted the evaporation residue cross sections in superheavy elements and the influence of shell effects [13]. The entrance channel dynamics were studied using 48Ca as a projectile and 208Pb as target [14]. In 1966, two groups of researchers, namely, Mayers and Swiatecki, and Viola and Seaborg [15], separately predicted the presence of heavy nuclei near the island of stability. Later, Sobiczewski et al. [16] predicted that the nucleus Z=114 will be the center of the island of stability, with neutron number N=184. In 1955 Nilsson [17] proposed a shell model which includes deformation property of the nuclei. Bender et al. [18] used a Skyrme energy density functional model and studied the deformation properties of closed proton and neutron shells. The nuclear mass, radius, and spectroscopy far away from the valley of stability were experimentally analyzed earlier [19]. The investigation of isomers of the superheavy nucleus 254No is a stepping stone toward the island of stability [20]. Previous researchers [21] analyzed the nuclear shell structure and discovered additional stability near magic nuclei. The present scenario is almost near the center of the presumed island of stability, but the final landing is yet to be completed, and the intriguing question is how these superheavy nuclei are still accessible.

The identification of superheavy nuclei is based on observations of decay chains. Superheavy element region 114 ≤ Z ≤ 118 were observed owing to their consistent decay chains, which end in the isotopes of rutherfordium(Rf) and dubnium(Db). Spontaneous fission and α-decay are the dominant decay modes in superheavy nuclei, and limit their stability. Furthermore, newly synthesized superheavy elements are primarily identified by their decay chains from unknown nuclei to known daughter nuclei by using the parent-and-daughter correlation.

The competition between different decay modes, such as ternary fission, spontaneous fission, cluster decay, proton decay, β-decay, and α-decay, in the heavy and superheavy region, has been extensively studied using various theoretical models, such as Coulomb and proximity potential models, modified generalized liquid drop models, effective liquid drop models, and temperature-dependent proximity potential models [22-23]. The possible decay modes in the superheavy nuclei Z=119 and 120 were predicted in Ref. [34]. From Ref. [35], it is clearly observed that the isotopes of the superheavy nuclei Z=104–112 have α-decay and spontaneous fission as dominant decay modes. However, only α-decay is dominant in the isotopes of superheavy nuclei Z=113, 115–118. The isotopes of the superheavy nucleus Z=114 have spontaneous fission as the dominant decay mode in the nucleus 284Fl, α-decay is dominant in the nuclei 286-289Fl, and β+ is dominant in the nucleus 290Fl. Furthermore, the concept of heavy-particle radioactivity [36] in the superheavy region has important applications in the synthesis of superheavy nuclei. Despite the significant experimental and theoretical progress, there are many unanswered questions related to the decay modes of superheavy nuclei. Until now, only α-decay and spontaneous fission have been successfully observed in experiments.

Experimental results suggest a considerable increase in the lifetime of nuclei as they approach closed proton and neutron shells [37]. The lifetimes of most known superheavy nuclei are governed by the competition between α-decay and spontaneous fission. The existence of the island of stability has been confirmed experimentally in the previous decade [38]. Some theoretical studies reveal that superheavy elements with 114 and 164 protons are stable against fission as well as alpha and beta decay [39]. Various phenomenological and microscopic models, such as the fission model [40], the cluster model [41], generalized liquid drop model [42], and the unified model for alpha decay and alpha capture [43], are available to study the different decay modes of superheavy nuclei. In addition, many studies have been concerned with the alpha decay and spontaneous fission of superheavy nuclei [44-46]. Simple empirical formulas are also available for determining the decay half-lives [47]. The possible isotopes of new superheavy elements are identified by studying the competition between different probable decay modes, such as α-decay, β-decay, cluster decay, and spontaneous fission. This study focuses on the different decay modes of superheavy nuclei in the atomic number range 104 ≤ Z ≤ 126. After a detailed investigation of the competition between different decay modes, the possible isotopes and their decay modes with branching ratios are identified in the superheavy nuclei region. Hence, the contribution of this study is in the prediction of the most possible decay mode in superheavy nuclei, and in the identification of possible emitters in this superheavy region. The formalism is explained in Section 2. The analysis of different decay modes and possible emitters in the superheavy region is explained in Section 2.4. The paper is concluded in Section 3.

2

Theory

2.1
Alpha decay and cluster decay

In the effective liquid-drop model (ELDM), the α-decay half-life is computed using the relation

T1/2(s)=ln2ν0PPα, (1)

where ν0 is the assault frequency on the barrier, and ν0=1.8×1022s-1 [48]. is the preformation factor, which is closely related to the shell structure [49]. The empirical formula for is expressed as

logPα=p1+p2(ZZ1)(Z2Z)+p3(NN1)(N2N)+p4A, (2)

where N, Z, and A are the neutron, charge, and mass number of the parent nucleus, respectively, Z1 and Z2 are the proton magic numbers around Z (Z1ZZ2), and N1 and N2 are the neutron magic numbers around N (N1NN2). p1, p2, and p3 correspond to parameters in the region even(Z)-even(N), even(Z)-odd(N), odd(Z)-even(N), and odd(Z)-odd(N). They are presented in Table I of Ref. [50]. P is the Gamow penetrability factor, given by the expression

P=exp[2ζ0ζc2μ[V(ζ)Q]dζ], (3)

where μ is the inertial coefficient resulting from the Werner—Wheeler approximation [51]. The limits of integration ζ0 and ζc are the inner and outer turning points, expressed as ζ0=Rp-R¯1 and ζc=Z1Z2e2Q. Rp is the radius of the parent nucleus, and R¯1 is the final radius of the emitted cluster. In the ELDM, the total potential has been demonstrated to be the sum of Coulomb, proximity, and centrifugal potential [52, 53]. Hence, we can use the effective one-dimensional total potential energy as follows:

V=Vc+Vs+Vl. (4)

To evaluate the Coulomb contribution in terms of the deformation parameter, we used Vc as defined in Ref. [54]:

Vc(R)=e2Z1Z2R+3Z1Z2e2λ,i=1,2×Riλ(αi,T)(2λ+1)Yλ(0)(θi)[βλi+47βλi2Yλ(0)(θi)], (5)

with

Ri(αi)=Roi[1+λβλiYλ(0)(αi)], (6)

where βλi is the deformation parameter, (0) are the spherical harmonics, and Roi=1.28Ai1/30.76+0.8Ai1/3. The effective surface potential can be calculated by

Vs=σeff(S1+S2), (7)

where S1 and S2 are the surface areas of the spherical fragments. σeff is the effective surface tension, which is defined as

σeff=14(R2R12R22)(Q320πϵ0e2[Z2RZ12R1Z22R2]), (8)

where R2 is the final radius of the daughter fragment. The centrifugal potential energy is determined by

Vl=2l(l+1)2μζ2, (9)

where 𝓁 is the angular momentum of the emitted alpha/cluster and is calculated using the selection rules. In the case of alpha/cluster decay [55, 56], the selection rules follow the condition

|JpJd|lα|Jp+Jd|andπpπd=(1)la, (10)

where Jp, πp and Jd, πd are the spin and parity of the parent and daughter nuclei, respectively. μ=M1M2M1+M2 is the reduced mass of the fragments, where M1 and M2 denote their atomic masses.

In ELDM, a system with two intersecting spherical nuclei with different radii is considered [52]. A schematic diagram for the representation of four independent coordinates, namely R1, R2, ζ, and ξ, is shown in Fig. 1. Three constraints are used to reduce the 4-dimensional spherical problem to an equivalent 1-dimensional problem. The geometric constraint given below is introduced so that the spherical segments remain in contact:

Fig. 1.
Schematic presentation of molecular phase of the di-nuclear system (the daughter nucleus and the emitted (smaller) fragments). The distance between their geometrical centers and the distance between the center of the heavier fragment and the circular sharp neck of radius a are denoted by ζ and ξ, respectively.
pic
R12(ζξ)2=R2ξ2. (11)

The variables ζ and ξ represent the distance between the geometrical centers and the distance between the center of the heavier fragment and the circular sharp neck of the radius, respectively [53, 57]. Assuming that nuclear matter is incompressible, the constraint for the conservation of the total volume of the system is

2(R13+R23)+3[R12(ζξ)+R22ξ][(ζξ)2+ξ3]=4R3, (12)

where R=r0A1/3 is the radius of the parent nucleus (r0=1.34 fm is an adjustable parameter), with A being the mass number of the parent.

The radius of the α particle, R1, is assumed to be constant in the varying mass asymmetry shape description:

R1R¯1=0, (13)

where R¯1=(ZiZ)1/3R, i=1,2; R¯1 provides the final radius of the α particle. Here, Z1, Z2, and Z are the atomic numbers of the α particle, daughter nucleus, and parent nucleus, respectively.

2.2
Beta decay

For all types of β processes, the expression for the half-life is given by [58]

1Tβ=1Tβ++1Tβ+1EC. (14)

Here, EC is the electron capture. For a particular type of β-decay, the half-life is expressed as follows:

f0bTb=ln2[g2me5c42π37|Mif|2]1. (15)

Here, f0b is the Fermi function, b=β± or EC, me is the mass of the electron, and Mif is the transition matrix element between the initial and final states. The right-hand side of the above may be approximated by a constant for each type of β-decay [59]. This constant is different for allowed and forbidden cases of beta decay. For allowed β-decay, this constant has been determined as 5.7±1.1 [60]. Equation (15) is reduced to

log10[f0bTb(sec)]=5.7±1.1, (16) log10[f0bTb]=4.7. (17)
2.2.1
β± decay

The Fermi function for β-decay is expressed as

f0β±=1E0F(E,Z)P(E)(E0E)2dE. (18)

Here, P(E) is the momentum of the particle, and F(E,Z) can be computed at the nuclear surface using the magnitude of the radial electron/positron wave function. The first approximation of F(E,Z) is

F0(E,Z)=2(γ+1)(2pR)2(γ1)exp[πξp]|Γ(γ+iεp)|2Γ2(2γ+1). (19)

Here, γ=1α2Z2, ξα ZE (+ for β- decay and - for β+ decay), α=1/137 is the fine structure constant, R is the radius of the nucleus, and Γ is the gamma function.

At the surface of the nucleus (for β+ decay), the orbital electron screening effect has a significant impact on the β electron/positron wave function. Thus, F(E,Z) becomes

F(E,Z)=F0(EV0,Z)æ(EV0,Z)p(EV0)(EV0)p(E)E. (20)

Here, V0=1.81α2Z4/3, is the finite wavelength of the β particle, p(E)=E21 is the momentum of the β particle, E0=1+Qβ±/mec2 is the total limit energy of the β decay, E=1+ε/mec2, and ε is the kinetic energy of the β particle

The expression for the energy released in β+ decay is

Qβ+=M(A,Z)M(A,Z1)2mec2. (21)

Similarly, for β- decay,

Qβ=M(A,Z)M(A,Z1). (22)
2.2.2
Electron capture

The value of Q for electron capture is given by

QEC=M(A,Z)M(A,Z1)Be=Qβ++2mec2Be. (23)

Here, Be is the electron binding energy. Hence, even for the forbidden β+ decay, electron capture is allowed. The capture of electrons of the K-shell for lower Z, and of the L-shell for higher Z is the major contributor to electron capture. The contributions of the electrons of higher shells are negligible. Thus, the Fermi function becomes

f0EC=f0K+f0LI+f0LII. (24)

In general, for any shell,

f0X=π2[E(QEC)+EX]2[gX2(ZX)+fX2(ZX)]X=K,LI,LII. (25)

Here, EX is the total energy of the electron:

EK=γ,EL=(1+γ2)1/2, (26)

where ZX is the effective charge, which considers the screening of the Coulomb field of the nucleus by other electrons [61]:

ZK=Z0.35andZL=Z4.15. (27)

The non-zero components of the radial parts (gX & fX) of the wave function of the relativistic electron of orbit X are

gK2(Z)=4(1+γ)(2αZR)2(γ1)(αZ)3Γ(2γ+1), (28) gLI2(Z)=[(2γ+2)1/2+2](2γ+1)(2αZR)2(γ1)(αZ)3Γ(2γ+1)[(2γ+2)1/2+1](2γ+2)γ, (29) gLII2(Z)=316(αZ)2gLI2(Z). (30)
2.3
Spontaneous fission

Spontaneous fission decay is studied by employing the quantum tunneling effect through the potential barrier. The decay constant of spontaneous fission is expressed as

λ=ln2Tsf=νSPs, (31)

where ν, S, and Ps are model-dependent quantities, namely, assault frequency, preformation probability, and barrier penetrability, respectively. In the above equation, P = SPs and the spontaneous-fission half-lives are calculated as

T=ln2νP=hln221EνP, (32)

where h is the Planck constant, and Ev=/2 is the zero-point vibration energy. The penetration probability is evaluated using the action integral K:

P=exp(K), (33)

and hence the decimal logarithm of T(s) is given by

log10T=0.43429K20.8436log10Eν. (34)

If =0.5 MeV, then the above equation becomes log10T=-log10P-20.5426. The action integral K is evaluated as follows:

K=22mRaRb(B(r)[E(R)Q])1/2dR. (35)

The term E(R) is the macroscopic energy in terms of the surface, volume, Coulomb, proximity energy, shell correction, and pairing energy term [62], and m is the rest mass of the neutron. A few superheavies are spherical, the rest are deformed, primarily prolate or oblate. To include this effect, deformations are also involved in the calculation of E(r), which is adopted from Ref. [62]. In the above equation, R is the separation distance between the center of the fission fragments, and Ra and Rb are the turning points, which are evaluated using the boundary conditions E(Ra) and E(Rb)=Q. However, the term B(r) is the inertia with respect to r and is evaluated using the semi-empirical model for inertia [63]:

B(r)=μ(1+kexp[12851(rRsph/R0)]), (36)

where μ and k are the reduced mass of the fission fragments, and a semi-empirical constant (k=14.8), respectively. Rsph is the distance between the center of mass of the fission fragments, set as Rsph/R0=0.75 in the symmetric case. The decay constant (λ) and the total fission decay constant are evaluated as described in Ref. [62]

2.4
Results and discussion

The mass excess values play a major role in the prediction of the decay mode and the corresponding half-lives. The predicted half-lives are sensitive to the Q-values, and small changes in the Q-values result in a notable change in the half-lives, with a magnitude of order 101 to 102 [36]. Mass excess tables such as WS4 [64], EBW [65], HFB28 and HFB29 [66], DZ10 [67], KTUY [68], finite-range droplet model (FRDM) [69], and AME16 [70] are available in the literature. In the present study, we used the updated AME16 [70] mass excess values up to Z=118, and above Z>118, the mass excess values are taken from the FRDM [69]. The dominant decay mode is identified by studying the competition between different decay modes: α-decay, β-decay, cluster decay, and spontaneous fission in the superheavy nuclei region 104 ≤ Z ≤ 126 .

A detailed literature review indicates that there is no experimental evidence for cluster radioactivity in the superheavy region. Furthermore, experimental studies of cluster decay in the actinide region are available. To validate the present study, the cluster-decay half-lives obtained in the present study in the actinide region were compared with the experiments, and good agreement was observed. With this confidence, we studied cluster decay in the superheavy region, and the results are presented in Table 1. Similarly, Table 2 shows a comparison of the studied logarithmic half-lives (in years) of spontaneous fission from the present study with those from available experiments. It can be seen that the cluster-decay and spontaneous-fission half-lives obtained in the present study are close to those of the experiments.

Table 1:
Cluster-decay half-lives obtained from present study (PS) and available experiments (exp)
Decay QExp(MeV) log T1/2exp[71] logT1/2PS
221Fr→14C+207Tl 31.317 14.51 14.91
221Ra14C+207Pb 32.396 13.37 13.56
222Ra14C+208Pb 33.05 11.05 12.70
223Ra14C+209Pb 31.829 15.05 13.94
224Ra14C+210Pb 30.54 15.9 15.52
226Ra14C+212Pb 28.2 21.29 22.74
225Ac14C+211Bi 30.477 17.16 17.06
228Th20O+208Pb 44.72 20.73 22.04
230U22Ne+208Pb 61.4 19.56 20.21
230Th24Ne+206Hg 57.571 24.61 25.07
231Pa24Ne+207Tl 60.417 22.89 23.07
232U24Ne+208Pb 62.31 20.39 22.25
233U24Ne+209Pb 60.486 24.84 25.05
234U26Ne+208Pb 59.466 25.93 25.62
234U28Mg+206Hg 74.11 25.74 26.04
236Pu28Mg+208Pb 79.67 21.65 22.07
238Pu28Mg+210Pb 75.912 25.66 25.98
238Pu30Mg+208Pb 77 25.66 26.25
238Pu32Si+206Hg 91.19 25.3 26.05
242Cm34Si+208Pb 96.509 23.11 24.24
Show more
Table 2:
Comparison of logarithm half-lives (years) of spontaneous fission in the superheavy region 104 ≤ Z ≤ 114 from present study with those from available experiments.
Parent nuclei logSFExptyr [72] logSFThyr
254Rf -12.1 -10.91
256Rf -9.71 -8.48
258Rf -9.35 -7.06
260Rf -9.2 -6.35
262Rf -7.18 -6.36
258Sg -10 -11.33
260Sg -9.65 -10.17
262Sg -9.32 -8.722
264Sg -8.93 -7.98
266Sg -7.86 -7.96
264Hs -10.2 -11.02
270Ds -8.6 -9.46
282Cn -10.6 -9.39
284Cn -8.5 -7.98
286Fl -8.08 -7.58
Show more

As a part of this investigation, we studied the α-decay properties of superheavy nuclei using the formalism explained in the theory section. The predicted alpha-decay half-lives were validated by comparison with those from available experiments in the superheavy region. The results are given in Table 3.

Table 3:
Comparison of alpha-decay half-lives from the present study (PS) and those from available experimental (Exp) values
Parent nuclei (MeV) logT1/2 (Exp) logT1/2 (PS)
261Bh 8.649 1.515 1.86
260Db 9.379 -0.295 0.11
269Sg 8.8 2.27 2.12
265Sg 9.078 0.869 1.15
263Sg 9.391 -0.932 0.12
261Sg 9.803 -1.469 -1.21
272Bh 9.3 1.025 0.78
271Bh 9.5 0.176 0.18
270Bh 9.3 1.785 1.02
277Hs 8.4 -2.523 -1.02
273Hs 9.9 -0.119 -0.32
269Hs 9.629 0.851 0.65
274Hs 9.5 0.079 0.21
278Mt 9.1 0.653 1.65
276Mt 9.8 -0.284 0.05
274Mt 10.5 -0.357 -0.98
281Ds 8.958 1.104 1.45
282Rg 9.38 2 1.85
280Rg 9.98 0.623 0.55
279Rg 10.45 -1.046 -1.04
285Cn 8.793 1.447 2.85
283Cn 9.62 0.623 0.89
281Cn 10.28 -0.886 -0.68
284Cn 9.301 1.013 1.78
277Cn 11.622 -2.551 -2.65
286Nh 9.68 0.978 1.22
285Nh 10.02 0.623 0.76
284Nh 10.25 -0.013 0.08
283Nh 10.6 -1.125 -0.98
289Fl 9.847 0.279 0.96
288Fl 9.969 -0.18 -0.16
287Fl 10.436 -0.319 -0.28
286Fl 10.7 -0.921 -0.87
285Fl 11 -0.824 -1.89
290Mc 10.3 -0.187 0.18
289Mc 10.6 -0.481 -0.35
293Lv 8.886 -1.244 0.12
292Lv 10.707 -1.886 -0.96
291Lv 11 -1.721 -1.45
289Lv 11.7 -2.848 -2.97
294Ts 8.963 -1.292 0.06
294Og 8.47 -3.161 -2.45
295Og 9.056 -1.745 0.58
298120 13.355 -3.051 -4.68
299120 13.105 -3.15 -1.58
Show more

From the comparison, it is observed that the predicted half-lives are in good agreement with those of the experiments. With this confidence, we obtained the alpha-decay half-lives of superheavy nuclei in the region 104 ≤ Z ≤ 126. Figure 2

Fig. 2.
(Color online) Map of nuclei reflecting the logarithmic α-decay half-lives for the isotopes of elements from Z=104 to 126. The Q-values were estimated using AME16 and FRDM95. The vertical line on the right side of the figure shows an increase in the log T1/2 values from the navy-blue region to the brown region.
pic

shows a wide range of α-decay half-lives. For a given superheavy nucleus, the alpha decay half-lives increase as the neutron number of its isotopes increases. For instance, the α-decay half-lives are of the order of nanoseconds at N/Z=1.307692 for Rutherfordium, whereas for the same superheavy element, the α-decay half-lives are of the order of 102s at N/Z=1.504762. Similarly, all neutron-rich superheavy nuclei have comparably longer α-decay half-lives, which is in agreement with the report available in Ref. [73]. The obtained α-decay half-lives of all possible superheavy nuclei are presented in the heat map in Fig. 2. The right vertical bar shows the magnitude of the logT1/2 values. The color variation from navy blue to wine indicates values in the range 10-10–102 s. The contrast in the blue region lies between 10-10 s and 10-7 s, in the green region, it lies in the range 10-6–10-4 s, and the range 10-4-10-3 s is presented in the yellow region. Finally, the red-to-wine region shows higher half-lives in the range 10-2-102 s. The inset of Fig. 2 on the top left side provides information on the magnified portion of α-decay half-lives in the superheavy region Z=104-114, whereas the bottom-right inset provides information on the magnified portion of the superheavy region Z=115-126. After the detailed investigation of the α-decay, a search was made to identify the cluster emitters in the superheavy region. Cluster radioactivity is energetically favorable if the Q-values are positive. We studied the possibility of cluster decay with 3 ≤ Zc ≤ 45 in the superheavy region 104 ≤ Z ≤ 126. For a given parent nucleus, the half-lives corresponding to various cluster emission were evaluated, and the cluster corresponding to shorter half-lives was identified. Furthermore, the cluster emitters corresponding to shorter half-lives for different isotopes of a given superheavy element were also identified. Eventually, cluster emissions corresponding to the shortest half-lives Tc were identified; these are referred to as cluster-decay half-lives (Tc). The predicted cluster decay half-lives in the atomic number region 104 ≤ Z ≤ 126 correspond to all the studied cluster emissions, as shown in Fig. 3.

Fig. 3.
(Color online) Predicted cluster-decay logarithmic half-lives in the atomic number range Z=104–126 using AME16 and FRDM95 mass excess values. The hallow bin with different color in each panel shows the cluster emission corresponding to minimum half-lives.
pic

This figure enables us to identify the cluster emission corresponding to the shorter half-lives of a given superheavy element. The half-lives of superheavy nuclei with Z=115–120 against cluster radioactivity are shorter for 86Kr than those of the other studied clusters. The superheavy nuclei with Z=104, 106, 108, 110, 112, 114, 124, and 126 have shorter half-lives against 96Mo cluster emissions than those of the other studied clusters. The decay half-lives are shorter for the 91Y emission from superheavy nuclei with Z=109, 111, 113, 121, and 123. Similarly, the half-lives of superheavy nuclei with Z=105 and 107 against cluster radioactivity are shorter for 97Tc and 101Rh than those of the other studied clusters.

Cluster radioactivity in the superheavy nuclei region has shorter half-lives for cluster neutron numbers 44–48 from parent nuclei with neutron numbers 130—200, as shown in Fig. 4.

Fig. 4.
(Color online) Map of nuclei reflecting the logarithmic cluster-decay half-lives for neutron number of parent and cluster isotopes of elements with Z=104–126.
pic

The range of cluster decay half-lives for superheavy elements with 104 ≤ Z ≤ 126 is shown in Fig. 5.

Fig. 5.
(Color online) Heat map showing the variations of lowest logarithmic half lives of clusters with 104 < Z < 126
pic

Shorter half-lives are observed for N/Z > 1.37068, and larger half-lives are observed for N/Z < 1.37068. From the figure, it is clear that up to superheavy nuclei 104 ≤ Z ≤ 115, larger cluster-decay half-lives are observed, whereas shorter cluster-decay half-lives are observed in the superheavy region 116 ≤ Z ≤ 126. The inset of Fig. 5 on the top-left side shows a magnified portion of the logarithmic half-lives (Tc) in the superheavy region 104 ≤ Z ≤ 115, whereas the inset at the right bottom shows a magnified portion of the shorter logarithmic half-lives (Tc) in the superheavy region 116 ≤ Z ≤ 126. This figure also shows that some of the superheavy nuclei have lifetimes of the order of ns to μs, and exhibit cluster decay.

The other prominent decay mode that was studied is spontaneous fission, which is also energetically feasible in heavy and superheavy nuclei. It may occur in such nuclei owing to an increase in the Coulomb interactions. References [10, 11, 38, 74-77] report consistent α-decay chains from superheavy nuclei followed by spontaneous fission. The spontaneous fission half-lives are studied using the theory explained in Sect. 2.3. The variations of spontaneous fission half-lives in the superheavy region Z=104–126 are shown in Fig. 6.

Fig. 6.
(Color online) Heat map of the variations of logarithmic half-lives for spontaneous fission for 104 < Z < 126
pic

The log TSF values vary between -50(dark blue region) and 50 (dark-red region). For instance, at atomic number Z=104, for isotopes 245—275, the log T1/2(SF) values ranging from -50 to 5 are shown, whereas the half-lives with smaller values are indicated by the color range from navy blue to blue. The half-lives ranging from nanoseconds to 105 s are indicated by the color range from yellow to light orange. Similarly, in the atomic number range Z=119 and above, larger values of spontaneous-fission logarithmic half-lives are indicated by the red color range. Thus, on either side of Fig. 6, for isotopes corresponding to the atomic number range Z=104–126, smaller half-lives are observed, whereas in the middle region of the figure, larger values of logT1/2 are observed up to Z=116. In contrast, smaller half-lives are observed for higher isotopes (Z>116), and larger logT1/2 for lower isotopes (Z<116). A similar trend was also observed in a previous study [78], in which the half-lives of nuclei Z=92–104 were compared with experimentally available values.

A detailed investigation of the Q-values corresponding to β-decay in the superheavy region demonstrates that β+-decay is energetically possible with Z=105, 107, 113, 114, 115, 117, 119, 121, 123, 125, and 126, whereas β--decay is not energetically possible. Furthermore, we also studied β-decay half-lives using the formalism explained in Sect. 2.2.1.

The competition between different possible decay modes, namely, α-decay, cluster-decay, β-decay and spontaneous fission, enables us to identify the dominant decay mode for superheavy elements in the atomic number region 104 ≤ Z ≤ 126 of all possible isotopes. Figure 7

Fig. 7.
(Color online) Chart of spontaneous fission (purple), alpha decay (brown), β+-decay (cyan), and cluster emitters (yellow) with atomic numbers Z=104–126. The Q-values were calculated using the FRDM95 mass tables.
pic

shows the decay modes of the superheavy nuclei. In the studied superheavy region, we identified around 20 β+emitters, which are presented in Table 6. We also identified 35 cluster emitters, which are presented in Table 4.

Table 6:
Identified β+ emitters in the superheavy nuclei region
Parent nuclei Q (MeV) logT1/2
264Db 2.24 -0.04
268Bh 2.93 -0.83
290Fl 0.79 1.28
286Mc 4.53 -3.68
292Ts 4.96 -4.12
290119 7.20 -6.27
296119 5.75 -5.01
292121 8.29 -7.56
294121 8.06 -7.15
298121 6.83 -6.32
302121 5.12 -5.49
298123 8.42 -8.04
300123 8.27 -7.64
302123 6.72 -7.23
306123 5.73 -6.42
304125 7.81 -8.55
306125 7.61 -8.15
308125 6.99 -7.75
310125 6.47 -7.35
312125 5.79 -6.96
Show more
Table 4:
Identified cluster emitters in the superheavy nuclei region
Parent nuclei Q (MeV) logT1/2 Cluster
292Og 304.08 -5.08 86Kr
293Og 303.63 -4.63 86Kr
298122 338.25 -6.02 94Zr
300122 337.45 -6.21 94Zr
299123 338.66 -7.18 91Y
300124 356.06 -7.35 96Mo
306126 364.27 -8.78 96Mo
Show more

It was demonstrated that the majority of superheavy nuclei undergo α-decay and spontaneous fission. The α-emitting superheavy nuclei are listed in Table 5.

Table 5:
Identified alpha emitters in the superheavy nuclei region
Parent nuclei Q (MeV) logT1/2 Parent nuclei Q (MeV) logT1/2 Parent nuclei Q (MeV) logT1/2
256Rf 10.15 0.92 277Ds 10.34 -2.51 280Lv 13.59 -6.78
257Rf 10.05 0.78 271Rg 11.61 -3.72 281Lv 13.35 -6.98
258Rf 9.94 -0.16 273Rg 11.44 -3.56 282Lv 13.13 -5.14
259Rf 9.67 0.35 275Rg 11.37 -3.42 283Lv 12.91 -5.78
260Rf 9.4 -1.12 277Rg 10.88 -2.21 284Lv 12.7 -3.89
261Rf 9.14 1.56 279Rg 10.44 -1.23 285Lv 12.51 -3.99
262Rf 8.92 0.52 280Rg 10.24 0.62 286Lv 12.34 -4.25
258Db 10.45 0.55 282Rg 9.89 2.15 287Lv 12.19 -3.98
259Db 10.36 -0.35 271Cn 12.1 -4.89 288Lv 12.06 -3.56
260Db 10.08 0.26 272Cn 11.96 -4.65 290Lv 11.83 -1.75
261Db 9.81 0.65 273Cn 11.87 -4.33 291Lv 11.73 -2.36
263Db 9.34 1.52 274Cn 11.8 -4.16 292Lv 11.64 -1.88
270Db 8.45 3.25 275Cn 11.76 -3.98 293Lv 11.55 -1.23
259Sg 10.84 -0.15 276Cn 11.74 -2.98 279Ts 14.06 -7.36
260Sg 10.74 -2.16 277Cn 11.49 -3.15 281Ts 14.02 -8.25
261Sg 10.47 -0.48 278Cn 11.25 -2.47 283Ts 13.56 -5.36
262Sg 10.2 -1.86 279Cn 11.03 -2.36 285Ts 13.14 -5.12
263Sg 9.95 0.25 280Cn 10.81 -1.78 287Ts 12.78 -4.88
269Sg 9.16 2.56 281Cn 10.62 -0.99 289Ts 12.5 -4.65
260Bh 11.21 -1.42 285Cn 10 1.56 291Ts 12.28 -2.98
263Bh 10.59 -1.76 273Nh 12.4 -5.65 294Ts 12 -1.45
265Bh 10.12 0.12 275Nh 12.24 -4.79 281Og 14.44 -7.65
266Bh 9.94 0.22 276Nh 12.2 -4.78 282Og 14.43 -7.63
270Bh 9.56 1.89 277Nh 12.18 -4.52 283Og 14.19 -7.45
271Bh 9.53 0.18 279Nh 11.7 -2.89 284Og 13.97 -6.25
272Bh 9.27 1.12 281Nh 11.26 -2.12 285Og 13.76 -6.41
274Bh 8.78 1.23 282Nh 11.07 -1.69 286Og 13.56 -6.24
263Hs 11.27 -2.56 284Nh 10.73 -0.16 287Og 13.37 -5.25
265Hs 10.77 -4.56 285Nh 10.59 0.78 288Og 13.21 -5.98
266Hs 10.55 -1.85 286Nh 10.47 0.88 294Og 12.53 -3.88
267Hs 10.37 -1.42 287Nh 10.35 0.76 295Og 12.44 -1.25
268Hs 10.23 0.69 275Fl 12.78 -4.69 285119 14.37 -5.69
269Hs 10.13 1.42 276Fl 12.72 -5.12 287119 13.96 -4.25
270Hs 10.05 1.78 277Fl 12.68 -5.36 289119 13.62 -5.97
271Hs 10 0.45 278Fl 12.66 -5.46 292119 13.23 -5.28
273Hs 9.71 -0.56 279Fl 12.42 -4.12 297119 12.78 -3.97
275Hs 9.23 -0.15 280Fl 12.19 -4.36 287120 14.56 -6.58
266Mt 11.27 -1.93 281Fl 11.96 -3.78 288120 14.37 -6.25
267Mt 11.06 -2.52 282Fl 11.76 -3.15 290120 14.03 -6.46
269Mt 10.74 -2.32 283Fl 11.56 -2.99 292120 13.76 -5.85
271Mt 10.57 -1.85 288Fl 10.86 -0.25 298120 13.2 -3.87
273Mt 10.49 -1.23 289Fl 10.75 0.62 299120 13.11 -4.12
274Mt 10.23 -0.12 277Mc 13.19 -6.85 300120 13.02 -4.36
275Mt 9.99 -1.25 278Mc 13.16 -6.48 290121 14.59 -6.28
276Mt 9.75 -0.36 279Mc 13.14 -5.96 296121 13.87 -5.48
278Mt 9.33 1.36 280Mc 12.9 -5.12 300121 13.53 -5.22
268Ds 11.62 -3.56 281Mc 12.67 -5.36 303122 13.77 -4.98
269Ds 11.45 -3.24 283Mc 12.24 -4.25 304122 13.67 -4.99
270Ds 11.31 -3.68 285Mc 11.88 -3.12 304123 14.18 -5.12
271Ds 11.21 -0.58 288Mc 11.47 -0.89 306124 14.49 -5.66
272Ds 11.14 -3.09 289Mc 11.36 -0.52 308124 14.28 -5.98
273Ds 11.1 -2.96 290Mc 11.26 -0.25 310124 14.05 -5.87
274Ds 11.07 -2.45 278Lv 13.64 -6.75
275Ds 10.81 -2.12 279Lv 13.61 -6.12      
Show more

The identified alpha emitters have half-lives of approximately μs to 100 s in the superheavy region 104 ≤ Z ≤ 126. Table 4 lists the identified cluster emissions with the corresponding half-lives. The amount of energy released during cluster emission, cluster emitted, and logT1/2 values are presented in the table. The minimum cluster decay half-lives correspond to 86Kr, 94Zr, 91Y, and 96Mo for the nuclei 292-293Og, 298,300122, 299123, 300124, and 306126, respectively. From the available literature, it is also evident that the heavy particle radioactivity of 86Kr is observed in the superheavy nucleus Z=118 [36, 79]. In addition, Rb, Sr, Y, Zr, Nb, and Mo cluster emissions [80] were observed for Z=119–124 respectively. As in previous studies, in the present study, shorter half-lives in the superheavy region Z=118, 122-124, and 126 were observed, with 86Kr, 94Zr, 91Y, and 96Mo cluster emissions, respectively. Similarly, approximately 20 possible β+ emitters were identified in the superheavy region 105 ≤ Z ≤ 125, and they are presented in Table 6.

The information provided Table 7 regarding the half-lives and branching ratios presents ambiguities in terms of determining a single decay mode. The branching ratios relative to the minimum half-lives among the studied decay modes are obtained, and the second column of the table shows the logT1/2 values corresponding to spontaneous-fission, α-decay, β+-decay, and cluster-decay half-lives. For instance, the superheavy nucleus 263Rf exhibits shorter logT1/2 values for spontaneous fission and β+-decay than for other decay modes. The branching ratio of spontaneous fission and β+-decay was obtained, and it was found that the branching ratio corresponding to spontaneous fission and β+ was 55% and 45%, respectively. Similarly, we identified the branching ratios for the superheavy region 104 ≤ Z ≤ 126, which are presented in Table 7.

Table 7:
Superheavy nuclei with dual decay mode and branching ratios
Parent logT1/2 Decay mode   Parent nuclei logT1/2 Decay mode
nuclei Sf α β+ CD       Sf α β+ CD    
263Rf 1.25 1.98 1.65 36.6 Sf=57% β+=43% 296120 19.58 -4.87 -4.38 -5.45 CD=53% α=47%
262Db 2.31 -0.48 -0.51 34.76 β+=52% α=48% 291120 25.45 -6.21 -5.85 -4.69 α=51% β+=49%
264Bh 0.71 -1.65 -1.75 20.58 β+=51% α=49% 297120 15.36 -4.25 -4.6 -3.62 β+=52% α=48%
262Bh -3.25 -3.25 -2.22 17.73 Sf=50% Sf=50% 289120 21.36 -6.31 -6.27 -4.3 α=50% β+=50%
266Bh 1.62 -1.89 -1.29 23.6 α=59% β+=41% 293120 25.32 -5.18 -5.43 -5.06 β+=51% α=49%
264Hs -2.98 -3.15 -1.11 15.37 α=51% Sf=49% 293121 29.32 -6.98 -6.29 -6.47 α=52% CD=48%
270Mt 2.11 -2.45 -2.08 19.43 α=54% β+=46% 299121 18.56 -4.75 -5.05 -5.43 CD=52% β+=48%
272Mt 1.35 -1.65 -1.63 22.66 α=50% β+=50% 303121 -4.36 -3.52 -4.22 2.35 Sf=51% β+=49%
268Mt 0.99 -2.78 -2.54 16.53 α=52% β+=48% 291121 25.87 -5.94 -6.7 -6.65 β+=50% CD=50%
276Ds -1.72 -1.65 -0.09 23.61 Sf=51% α=49% 295121 30.36 -5.78 -5.88 -6.34 CD=52% β+=48%
272Rg 2.22 -3.15 -3.34 12.55 β+=51% α=49% 297121 26.9 -4.77 -5.46 -6.33 CD=54% β+=46%
276Rg 2.42 -2.16 -2.44 18.44 β+=53% α=47% 301121 8.25 -5.36 -4.64 -1.47 α=54% β+=46%
269Rg -3.98 -3.86 -2.95 9.04 Sf=51% α=49% 294122 30.65 -5.98 -6.51 -6.89 CD=51% β+=49%
270Rg -1.56 -3.87 -3.79 10.12 α=51% β+=49% 295122 34.25 -5.99 -6.73 -6.78 CD=50% β+=50%
278Rg -1.32 -1.54 -1.99 21.54 β+=56% α=44% 296122 36.89 -6.24 -6.1 -6.83 CD=52% α=48%
274Rg 3.56 -3.22 -2.89 15.4 α=53% β+=47% 297122 34.22 -6.96 -6.32 -6.77 α=51% CD=49%
283Nh -1.55 -1.89 -1.55 18.55 α=55% Sf=45% 305122 -0.58 -4.98 -4.68 -0.68 α=52% β+=48%
274Nh 0.65 -4.01 -4.6 6.88 β+=53% α=47% 302122 18.21 -5.27 -4.87 -6.31 CD=54% α=46%
280Nh 4.98 -2.78 -3.27 14.3 β+=54% α=46% 301122 23.56 -5.96 -5.5 -6.46 CD=52% α=48%
278Nh 5.88 -3.87 -3.72 11.51 α=51% β+=49% 299122 30.89 -6.14 -5.91 -6.7 CD=52% α=48%
274Fl -5.12 -4.89 -3.95 3.62 Sf=51% α=49% 303122 13.77 -5.11 -5.09 -4.33 α=50% β+=50%
285Fl -0.45 -2.11 -1.97 16.71 α=52% β+=48% 303123 27.89 -6.36 -5.96 -3.66 α=52% β+=48%
282Mc 11.02 -3.78 -4.55 0.52 β+=55% α=45% 297123 38.48 -6.17 -7.18 -7.51 CD=51% β+=49%
287Mc 1.32 -2.98 -2.4 0.5 α=55% β+=45% 301123 35.02 -6.18 -6.37 -6.86 CD=52% β+=48%
284Mc 8.98 -3.58 -4.12 0.37 β+=54% α=46% 307123 2.18 -4.98 -5.15 2.81 β+=51% α=49%
276Mc -3.12 -5.89 -5.86 1.45 α=50% β+=50% 305123 15.58 -5.48 -5.56 -0.65 β+=50% α=50%
277Lv -6.85 -5.78 -5.42 0.18 Sf=54% α=46% 308123 -4.89 -5.26 -6.01 3.61 β+=53% α=47%
289Lv 3.15 -3.05 -2.83 -1.32 α=52% β+=48% 303124 40.25 -6.25 -6.83 -6.77 β+=50% CD=50%
293Ts -2.98 -3.58 -2.84 -2.88 α=55% Sf=45% 305124 33.21 -5.12 -6.42 -4.29 β+=56% α=44%
284Ts 13.25 -5.25 -5.83 -2.5 β+=53% α=47% 302124 40.25 -5.96 -6.6 -7.13 CD=52% β+=48%
286Ts 14.55 -5.16 -5.4 -2.6 β+=51% α=49% 304124 35.85 -6.01 -6.19 -6.45 CD=51% β+=49%
290Ts 9.18 -4.36 -4.55 -3.09 β+=51% α=49% 307124 22.14 -5.97 -6.02 -0.09 β+=50% α=50%
288Ts 13.78 -4.87 -4.98 -2.91 β+=51% α=49% 301124 44.14 -6.04 -7.23 -7.53 CD=51% β+=49%
280Ts 0.68 -6.25 -6.69 -2.01 β+=52% α=48% 309124 6.87 -5.11 -5.62 3.17 β+=52% α=48%
282Ts 8.58 -5.85 -6.26 -2.19 β+=52% α=48% 309125 26.25 -6.51 -6.49 0.26 α=50% β+=50%
290Og 16.98 -3.78 -3.91 -4.72 CD=55% β+=45% 307125 38.21 -6.21 -6.89 -3.75 β+=53% α=47%
291Og 14.89 -3.78 -4.13 -4.9 CD=54% β+=46% 311125 11.58 -5.75 -6.09 2.43 β+=51% α=49%
289Og 18.74 -5.25 -4.56 -4.46 α=54% β+=46% 305125 45.35 -6.87 -7.29 -7.5 CD=51% β+=49%
284119 5.95 -6.68 -7.53 -3.17 β+=53% α=47% 303125 45.25 -6.96 -7.69 -8.07 CD=51% β+=49%
293119 18.29 -3.98 -4.57 -5.15 CD=53% β+=47% 315126 -1.58 -5.75 -6.17 2.82 β+=52% α=48%
286119 14.98 -7.25 -7.11 -3.79 α=50% β+=50% 313126 15.25 -6.24 -6.57 0.85 β+=51% α=49%
294119 15.98 -3.89 -5.43 -5.15 β+=51% CD=49% 308126 48.21 -6.87 -7.12 -5.37 β+=51% α=49%
288119 19.35 -5.85 -6.69 -4.15 β+=53% α=47% 311126 32.21 -5.12 -6.96 -0.97 β+=58% α=42%
298119 -4.85 -3.22 -4.59 0.23 Sf=51% β+=49% 310126 39.21 -6.32 -6.73 -1.99 β+=52% α=48%
291119 21.25 -4.58 -4.99 -4.74 β+=51% CD=49% 309126 44.58 -6.58 -7.36 -3.69 β+=53% α=47%
295119 11.35 -4.23 -4.15 -5.03 CD=54% α=46% 312126 26.12 -5.74 -6.33 0.01 β+=52% α=48%
295120 22.15 -4.78 -5.02 -5.46 CD=52% β+=48% 307126 51.32 -7.11 -7.75 -6.99 β+=52% α=48%
294120 23.58 -4.22 -4.79 -5.36 CD=53% β+=47% 314126 7.56 -6.21 -5.94 1.7 α=51% β+=49%
Show more

Finally, Fig. 8 shows the lifetimes of the superheavy elements after the competition between different decay modes was studied.

Fig. 8.
(Color online) Heat map showing the variations of atomic number, mass number of parent and logarithmic half-lives of different decay modes (life times) for 104 < Z < 126
pic

It can be seen that the lifetime varies from ns to min and decreases as the atomic number increases. For instance, the average lifetime of a superheavy element with Z=104 is approximately 10 min, whereas that of a hypothetical superheavy element with Z=126 is of the order of ms.

3

Conclusion

We systematically investigated all possible decay modes, namely, α-decay, β-decay, cluster decay, and spontaneous fission, in the superheavy region 104 ≤ Z ≤ 126. The findings of this study were validated by comparison with experiments. Approximately 20 β+ and 7 heavy particle emitters were found in the superheavy region. Furthermore, the nuclei with almost the same half-lives for the two decay modes were also reported, with the corresponding branching ratios. However, an experimental study is necessary to draw definite conclusions.

References
[1] Hebberger F. P., Hofmann S., Ackermann D. et al.,

Decay properties of neutron-deficient isotopes 256,257Db, 255Rf, 252,253Lr

. Europ. Phys. J. A 12, 57-67 (2001). doi: 10.1007/s100500170039
Baidu ScholarGoogle Scholar
[2] Oganessian Y.T., Utyonkov V.K., Lobanov Y.V. et al.,

Publisher's note: Measurements of cross sections and decay properties of the isotopes of elements 112, 114, and 116 produced in the fusion reactions 233,238 U, 242 Pu, and 248 Cm ? 48 Ca. Phys

. Rev. C 70, 064609 (2004). doi: 10.1103/PhysRevC.71.029902.
Baidu ScholarGoogle Scholar
[3] Oganessian Y. T., Utyonkov V. K., Lobanov Y. V. et al.

Publisher’s note: Measurements of cross sections and decay properties of the isotopes of elements 112, 114, and 116 produced in the fusion reactions 233,238 U, 242Pu, and 248 Cm +48 Ca [phys. rev. c 70, 064609 (2004)]

. Phys. Rev. C, 71, 029902 (2005). doi: 10.1103/PhysRevC.71.029902.
Baidu ScholarGoogle Scholar
[4] Adamian G.G., Antonenko N.V., Scheid W.,

High-spin isomers in some of the heaviest nuclei: Spectra, decays, and population

. Phys. Rev. C 81, 024320 (2010). doi: 10.1103/PhysRevC.81.024320
Baidu ScholarGoogle Scholar
[5] Kruppa A. T., Bender M., Nazarewicz W. et al.,

Shell corrections of superheavy nuclei in selfconsistent calculations

. Phys. Rev. C 61, 034313 (2000). doi: 10.1103/PhysRevC.61.034313
Baidu ScholarGoogle Scholar
[6] Shi Y., Ward D.E., Carlsson B.G. et al.,

Structure of superheavy nuclei along decay chains of element 115

. Phys. Rev. C 90, 014308 (2014). doi: 10.1103/PhysRevC.90.014308
Baidu ScholarGoogle Scholar
[7] Meng X., Lu B., Zhou S.,

Ground state properties and potential energy surfaces of 270Hs from multidimensionally-constrained relativistic mean field model

. Sci. ChinaI Phys. Mech. Astron. 63, 1-11 (2020). doi: 10.10072Fs11433-019-9422-1
Baidu ScholarGoogle Scholar
[8] Swain R., Sahu B..

Structure and reaction dynamics of she Z= 130

. Chinese Physics C 43 (2019) 104103. doi: 10.1088/1674-1137/43/10/104103
Baidu ScholarGoogle Scholar
[9] Niu F., Chen P.-H., Cheng H.-G. et al.,

Multinucleon transfer dynamics in nearly symmetric nuclear reactions

. Nucl. Sci. Tech. 31, 59 (2020). doi: 10.10072Fs41365-020-00770-1
Baidu ScholarGoogle Scholar
[10] Hofmann S., Ninov V., Heberger F. P. et al.,

Production and decay of 269110

. Zeitschrift fur Physik A Hadrons and Nuclei 350, 277-280 (1995). doi: 10.1007/BF01291181
Baidu ScholarGoogle Scholar
[11] Hofmann S., Munzenberg G.,

The discovery of the heaviest elements

. Rev. Mod. Phys. 72, 733-767 (2000). doi: 10.1103/RevModPhys.72.733
Baidu ScholarGoogle Scholar
[12] Boilley D., Cauchois B., Lv H. et al.,

How accurately can we predict synthesis cross sections of superheavy elements?

Nucl. Sci. Tech. 29, 172 (2018). doi: 10.1007/s41365-018-0509-7
Baidu ScholarGoogle Scholar
[13] Naderi D., Alavi S. A.,

Influence of the shell effects on evaporation residue cross section of superheavy nuclei

. Nucl. Sci. Tech. 29, 161 (2018). doi: 10.1007/s41365-018-0498-6
Baidu ScholarGoogle Scholar
[14] Li X., Wu Z., Guo L.,

Entrancechannel dynamics in the reaction 40Ca +208Pb

. Sci. China-Phy. Mech. Astron. 62, 12 (2019). doi: 10.10072Fs11433-019-9435-x
Baidu ScholarGoogle Scholar
[15] Oganessian Y. T. and Rykaczewski K. P.,

A beachhead on the island of stability

. Physics Today 68 (2015). doi: 10.1063/PT.3.2880
Baidu ScholarGoogle Scholar
[16] Sobiczewski A., Gareev F. A., Kalinkin B. N.,

Closed shells for Z > 82 and N > 126 in a diffuse potential well

. Phys. Lett. 22, 500-502 (1966). doi: 10.1016/0031-9163(66)91243-1
Baidu ScholarGoogle Scholar
[17] Nilsson S..

Binding states of individual nucleons in strongly deformed nuclei

. Dan. Mat. Fys. Medd. 29, 1-69 (1955). https://cds.cern.ch/record/212345
Baidu ScholarGoogle Scholar
[18] Bender M., Bennaceur K., Duguet T. et al.,

Tensor part of the skyrme energy density functional. ii. deformation properties of magic and semimagic nuclei

. Phys. Rev. C 80, 064302 (2009). doi: 10.1103/PhysRevC.80.064302.
Baidu ScholarGoogle Scholar
[19] Brown B. A.,

The nuclear shell model towards the drip lines

. Prog. Part. Nucl. Phys. 47, 517-599 (2001). doi: 10.1016/S0146-6410(01)00159-4.
Baidu ScholarGoogle Scholar
[20] Herzberg R. D., Greenlees P. T., Butler P. A. et al.,

Nuclear isomers in superheavy elements as stepping stones towards the island of stability

. Nature 442 896-899 (2006). doi: 10.1038/nature05069
Baidu ScholarGoogle Scholar
[21] Tománek D. and Schlüter M. A.,

Calculation of magic numbers and the stability of small si clusters

. Phys. Rev. Lett. 56, 1055-1058 (1986). doi: 10.1103/PhysRevLett.56.1055
Baidu ScholarGoogle Scholar
[22] Nagaraja A. M., Manjunatha H. C., Sowmya N. et al.,

Cluster radioactivity in superheavy nuclei 299-302120

. Indian J. Pure Appl. Phys. 58, 207-212 (2020). http://nopr.niscair.res.in/handle/123456789/54509
Baidu ScholarGoogle Scholar
[23] Manjunatha H. C., Sowmya N.,

Competition between spontaneous fission ternary fission cluster decay and alpha decay in the super heavy nuclei of Z = 126

. Nucl. Phys. A, 969, 68-82 (2018). doi: 10.1016/j.nuclphysa.2017.09.008.
Baidu ScholarGoogle Scholar
[24] Sowmya N., Manjunatha H. C., Dhananjaya N. et al.,

Competition between binary fission, ternary fission, cluster radioactivity and alpha decay of 281Ds

. J. Radioanalytical Nucl. Chem. 323, 1347-1351 (2020). doi: 10.1007/s10967-019-06706-3
Baidu ScholarGoogle Scholar
[25] sridhar G. R., Manjunatha H. C., Sowmya N. et al.,

Atlas of cluster radioactivity in actinide nuclei

. Euro. Phys. J. Plus 135, 1-28 (2020). doi: 10.1140/epjp/s13360-020-00302-1
Baidu ScholarGoogle Scholar
[26] Manjunatha H. C., Sridhar K. N., Sowmya N.,

Investigations of the synthesis of the superheavy element Z = 122

. Phys. Rev. C 98, 024308 (2018) . doi: 10.1103/PhysRevC.98.024308
Baidu ScholarGoogle Scholar
[27] Manjunatha H.C., Sowmya N.,

Decay modes of superheavy nuclei Z = 124

. Intern. J. Mod. Phys. E 27, 1850041 (2018). doi: 10.1142/S0218301318500416
Baidu ScholarGoogle Scholar
[28] Sowmya N., Manjunatha H.C.,

Investigations on different decay modes of darmstadtium

. Phys. Particles Nuclei Lett. 17, 370-378 (2020). doi: 10.1134/S1547477120030140
Baidu ScholarGoogle Scholar
[29] Sowmya N., Manjunatha H.C.,

Competition between different decay modes of superheavy element Z = 116 and synthesis of possible isotopes

. Brazilian J. Physics 49, 874-886 (2019) doi: 10.1007/s13538-019-00710-4
Baidu ScholarGoogle Scholar
[30] Sowmya N., Manjunatha H. C., Gupta P. S. D.,

Competition between decay modes of superheavy nuclei 281-310Og

. Intern. J. Mod. Phys. E 29 (2020). doi: 10.1142/S0218301320500871
Baidu ScholarGoogle Scholar
[31] Sowmya N., Manjunatha H.C.,

Investigations on the synthesis and decay properties of roentgenium

. Brazilian J. Physics 50, 317-330 (2020). doi: 10.1007/s13538-020-00747-w
Baidu ScholarGoogle Scholar
[32] Sowmya N., Manjunatha H. C., Gupta P. S. D et al.,

Competition between cluster and alpha decay in odd Z superheavy nuclei 111 ≤ Z ≤ 125

. Brazilian J. Phys. 51, 99-135 (2021) . doi: 10.1007/s13538-020-00801-7.
Baidu ScholarGoogle Scholar
[33] Sridhar K. N., Manjunatha H. C., Ramalingam H. B.,

Search for possible fusion reactions to synthesize the superheavy element Z = 121

. Phys. Rev. C 98, 064605 (2018). doi: 10.1103/PhysRevC.98.064605.
Baidu ScholarGoogle Scholar
[34] Soylu A.,

Search for decay modes of heavy and superheavy nuclei

. Chinese Phys. C 43, 074102 (2019). doi: 10.1088/1674-1137/43/7/074102
Baidu ScholarGoogle Scholar
[35]

Nuclear live chart

. https://www-nds.iaea.org/relnsd/vcharthtml/VChartHTML.html
Baidu ScholarGoogle Scholar
[36] Poenaru D.N., Gherghescu R.A., Greiner W.,

Heavy-particle radioactivity of superheavy nuclei

. Phys. Rev. Lett. 107, 062503 (2011) . doi: 10.1103/PhysRevLett.107.062503.
Baidu ScholarGoogle Scholar
[37] Oganessian Y. T.,

Route to islands of stability of superheavy elements

. Phys. Atom. Nuclei 63, 1315-1336 (2000). doi: 10.1134/1.1307456.
Baidu ScholarGoogle Scholar
[38] Oganessian Y..

Heaviest nuclei from 48Ca induced reactions

. J. Phys. G 34, R165 (2007). doi: 10.1016/j.nuclphysa.2015.07.003
Baidu ScholarGoogle Scholar
[39] Grumann J., Mosel U., Fink B. et al.

Investigation of the stability of superheavy nuclei around Z= 114 and Z= 164

. Zeitschrift fur Physik 228, 371-386 (1969). doi: 10.1007/BF01406719
Baidu ScholarGoogle Scholar
[40] Poenaru D. N., Iva¸scu M., Sndulescu A. et al.,

Atomic nuclei decay modes by spontaneous emission of heavy ions

. Phys. Rev. C 32, 572-581 (1985). doi: 10.1103/PhysRevC.32.572
Baidu ScholarGoogle Scholar
[41] Buck B., Merchant A. C., Perez S. M.,

α decay calculations with a realistic potential

. Phys. Rev. C 45, 2247-2253 (1992) . doi: 10.1103/PhysRevC.45.2247
Baidu ScholarGoogle Scholar
[42] Zhang H.F., Royer G.,

Theoretical and experimental α decay half-lives of the heaviest odd-Z elements and general predictions

. Phys. Rev. C 76, 047304 (2007). doi: 10.1103/PhysRevC.76.047304
Baidu ScholarGoogle Scholar
[43] Denisov V. Y., Ikezoe H..

α-nucleus potential for α-decay and sub-barrier fusion

. Phys. Rev. C 72, 064613 (2005). doi: 10.1103/PhysRevC.72.064613
Baidu ScholarGoogle Scholar
[44] Möller P., Leander G. A., Nix J.R.,

On the stability of the transeinsteinium elements

. Zeitschrift fr Physik A Atomic Nuclei 323, 41-45 (1986). doi: 10.1007/BF01294553
Baidu ScholarGoogle Scholar
[45] Xu F.R., Zhao E.G., Wyss R. et al.,

Enhanced stability of superheavy nuclei due to high-spin isomerism

. Phys. Rev. Lett. 92, 252501 (2004) . doi: 10.1103/PhysRevLett.92.252501
Baidu ScholarGoogle Scholar
[46] Wong C.,

Additional evidence of stability of the superheavy element 310126 according to the shell model

. Physics Letters 21, 688-690 (1966). doi: 10.1016/0031-9163(66)90127-2.
Baidu ScholarGoogle Scholar
[47] Poenaru D.N., Gherghescu R. A., Greiner W.,

Single universal curve for cluster radioactivities and α decay

. Phys. Rev. C 83, 014601 (2011). doi: 10.1103/PhysRevC.83.014601
Baidu ScholarGoogle Scholar
[48] Duarte S. B., OAP T F. G., Dimarco A.,

Half-lives for proton emission, alpha decay, cluster radioactivity, and cold fission processes calculated in a unified theoretical framework

. Atomic Data and Nuclear Data Tables, 80, 235 (2002). doi: 10.1006/adnd.2002.0881
Baidu ScholarGoogle Scholar
[49] Zhang H. F., Royer G., Wang Y. J., et al.

Analytic expressions for α particle preformation in heavy nuclei

. Phys. Rev. C 80, 057301 (2009). doi: 10.1103/PhysRevC.80.057301
Baidu ScholarGoogle Scholar
[50] Shan Z., Yanli Z., Jianpo C. et al.,

Improved semi-empirical relationship for α-decay half-lives

. Phys. Rev. C 95, 014311 (2017). doi: 10.1103/PhysRevC.95.014311.
Baidu ScholarGoogle Scholar
[51] Poenaru D.N., Maruhn J.A., Greiner W., et al.

Inertia and fission paths in a wide range of mass asymmetry

. Zeitschrift fur Physik A Atomic Nuclei 333, 291 (1989). doi: 10.1007/BF01294517
Baidu ScholarGoogle Scholar
[52] Gongalves M. and Duarte S. B.,

Effective liquid drop description for the exotic decay of nuclei

. Phys. Rev. C 48, 2409-2414 (1993). doi: 10.1103/PhysRevC.48.2409.
Baidu ScholarGoogle Scholar
[53] Gonçalves M., Teruya N., Tavares O. et al.,

Twoproton emission half-lives in the effective liquid drop model

. Phys. Lett. B 774, 14-19 (2017). doi: 10.1016/j.physletb.2017.09.032
Baidu ScholarGoogle Scholar
[54] Santhosh K. and Biju R..

Stability of 248, 254Cf isotopes against alpha and cluster radioactivity

. Ann. Phys. 334, 280-287 (2013). doi: 10.1016/j.aop.2013.04.008.
Baidu ScholarGoogle Scholar
[55] Manjunatha H. C., Sridhar G. R., Sowmya N. et al.,

A systematic study of alpha decay in actinide nuclei using modified generalized liquid drop model

. Intern. J. Mod. Phys. E 30, 2150013 (2021). doi: 10.1142/S0218301321500130.
Baidu ScholarGoogle Scholar
[56] Maroufi N., Dehghani V., Alavi S. A.,

Alpha and cluster decay of some deformed heavy and superheavy nuclei

. Nucl. Phys. A 983, 77-89 (2019). doi: 10.1016/j.nuclphysa.2018.12.023.
Baidu ScholarGoogle Scholar
[57] Gonçalves M., Duarte S. B.,

Effective liquid drop description for the exotic decay of nuclei

. Phys. Rev. C 48, 2409-2414 (1993). doi: 10.1103/PhysRevC.48.2409.
Baidu ScholarGoogle Scholar
[58] Dzhelepov B. S., Zyryanova L. N., Suslov Y. P.,

Beta processes functions for the analysis of beta spectra and electron capture

. 373 (1972).
Baidu ScholarGoogle Scholar
[59] Karpov A. V., Zagrebaev V. I., Martinez Palenzuela Y., et al.

Decay properties and stability of heaviest elements

. Intern. J. Mod. Phys. E 21, 1250013 (2012). doi: 10.1142/S0218301312500139.
Baidu ScholarGoogle Scholar
[60] Wu C. S. and Moszkowski S. A.. Beta decay. 183 (1966)
[61] Preston M. A.. Addison-Wesley Publishing Company, Inc., (1962).
[62] Bao X., Zhang H., Royer G. et al.,

Spontaneous fission half-lives of heavy and superheavy nuclei within a generalized liquid drop model

. Nucl. Phys. A 906, 1-13 (2013). doi: 10.1016/j.nuclphysa.2013.03.002
Baidu ScholarGoogle Scholar
[63] Randrup J., Larsson S.E., Moller P. et al.,

Spontaneous-fission half-lives for even nuclei with Z ≥ 92

. Phys. Rev. C 13, 229-239 (1976). doi: 10.1103/PhysRevC.13.229
Baidu ScholarGoogle Scholar
[64] Wang N., Liu M., Wu X. et al.,

Surface diffuseness correction in global mass formula

. Phys. Lett. B 734, 215-219 (2014). doi: 10.1016/j.physletb.2014.05.049
Baidu ScholarGoogle Scholar
[65] Kirson M. W..

Mutual influence of terms in a semi-empirical mass formula

. Nucl. Phys. A 798, 29-60 (2008). doi: 10.1016/j.nuclphysa.2007.10.011
Baidu ScholarGoogle Scholar
[66] Goriely S.,

Further explorations of skyrme–hartree– fock–bogoliubov mass formulas. XV: The spin–orbit coupling

. Nucl. Phys. A 933, 68-81 (2015). doi: 10.1016/j.nuclphysa.2014.09.045
Baidu ScholarGoogle Scholar
[67] Duflo J., Zuker A.,

Microscopic mass formulas

. Phys. Rev. C 52, R23-R27 (1995). doi: 10.1103/PhysRevC.52.R23
Baidu ScholarGoogle Scholar
[68] Koura H., Tachibana T., Uno M. et al.,

Nuclidic mass formula on a spherical basis with an improved even-odd term

. Prog. Theor. Phys. 113, 305-325 (2005). doi: 10.1143/PTP.113.305
Baidu ScholarGoogle Scholar
[69]

Reference input parameter library (ripl-3)

. https://www-nds.iaea.org/RIPL-3
Baidu ScholarGoogle Scholar
[70] Wang M., Audi G., Kondev F. G. et al.,

The ame2016 atomic mass evaluation (ii). tables, graphs and references

. Chinese Physics C 41, 030003 (2017). doi: 10.1088/1674-1137/41/3/030003
Baidu ScholarGoogle Scholar
[71] Santhosh K.P., Biju R. K., Joseph A.,

A semi-empirical model for alpha and cluster radioactivity

. J. Phys. G 35, 085102 (2008). doi: 10.1088/0954-3899/35/8/085102.
Baidu ScholarGoogle Scholar
[72] Xu C., Ren Z., Guo Y.,

Competition between α decay and spontaneous fission for heavy and superheavy nuclei

. Phys. Rev. C 78, 044329 (2008). doi: 10.1103/PhysRevC.78.044329.
Baidu ScholarGoogle Scholar
[73] Wieleczko J. P., Bonnet E., Gomez del Campo J. et al.,

Influence of the n/z ratio on disintegration modes of compound nuclei

. In AIP Conference Proceedings 1098, 64-69 (2009). doi: 10.1063/1.3108862.
Baidu ScholarGoogle Scholar
[74] Oganessian Y. T., Utyonkov V. K., Lobanov Y. V., et al.,

Synthesis of superheavy nuclei in the 48Ca+244Pu reaction: 288114

. Phys. Rev. C 62, 041604 (2000). doi: 10.1103/PhysRevC.62.041604.
Baidu ScholarGoogle Scholar
[75] Oganessian Y. T., Utyonkov V. K., Lobanov Y. V., et al.,

Observation of the decay of 292116

. Phys. Rev. C 63, 011301 (2000). doi: 10.1103/PhysRevC.63.011301
Baidu ScholarGoogle Scholar
[76] Oganessian Y. T., Utyonkov V. K., Lobanov Y. V. et al.,

Measurements of cross sections for the fusion-evaporation reactions 244Pu(48Ca, xn)292-x114 and 245Cm(48Ca,xn)293-x116

. Phys. Rev. C 69, 054607 (2004). doi: 10.1103/PhysRevC.69.054607
Baidu ScholarGoogle Scholar
[77] Morita K., Morimoto K., Kaji D. et al.,

Experiment on the synthesis of element 113 in the reaction 209Bi(70Zn, n) 278113

. J. Phys. Soc. Japan 73, 2593-2596 (2004). doi: 10.1143/JPSJ.73.2593
Baidu ScholarGoogle Scholar
[78] Baran A., Pomorski K., Lukasiak A. et al.,

A dynamic analysis of spontaneous-fission halflives

. Nucl. Phys. A 361, 83-101 (1981). doi: 10.1016/0375-9474(81)90471-1
Baidu ScholarGoogle Scholar
[79] Matheson Z., Giuliani S. A., Nazarewicz W. et al.,

Cluster radioactivity of 118294 Og176

. Phys. Rev. C 99, 041304 (2019). doi: 10.1103/PhysRevC.99.041304
Baidu ScholarGoogle Scholar
[80] Poenaru D. N., Gherghescu R. A., Greiner W.,

Heavy-particle radioactivity

. J. Physics Conference Series 436, 012056 (2013). doi: 10.1088/1742-6596/436/1/012056
Baidu ScholarGoogle Scholar