logo

Nuclear deformation effects on α-decay half-lives with empirical formula and machine learning

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

Nuclear deformation effects on α-decay half-lives with empirical formula and machine learning

Hong-Qiang You
Xiao-Tao He
Ren-Hang Wu
Shuang-Shuang Zhang
Jing-Jing Li
Qing-Hua He
Hai-Qian Zhang
Nuclear Science and TechniquesVol.36, No.10Article number 191Published in print Oct 2025Available online 24 Jul 2025
13601

An improved formula considering the deformation effect for the α-decay half-lives is proposed based on WKB barrier penetrability. Using the quadrupole deformation values of the daughter nuclei obtained from the WS4 and FRDM models in the improved formula, the root mean square deviation (RMSD) between the calculated results and experimental data decreased from 0.456 to 0.413 and 0.415, respectively. Although the improved formula did not significantly reduce the overall RMSD, it produced results that better matched the experimental values for nuclei with larger deformations. Additionally, eXtreme Gradient Boosting (XGBoost) models were employed to further reduce the deviations between the calculated α-decay half-lives and experimental data, with the corresponding RMSDs decreasing from 0.413 to 0.295 and from 0.415 to 0.302, respectively. Furthermore, the improved empirical formula and XGBoost models were used to predict the α-decay half-lives of nuclei with Z = 117, 118, 119, and 120. The results suggest that N = 184 is the magic number.

α-decaynuclear deformation effectsmagic numbersSuperheavy nuclei
1

Introduction

In 1928, α-decay was first described as a quantum mechanical tunneling effect by Gurney and Condon [1] and Gamow [2]. Within Gamow’s picture, the α cluster is preformed in the parent nucleus before it travels through a potential barrier. α-decay has been used as an important tool for understanding nuclear structure. Experimentally, because α-decay is the dominant decay mode of superheavy nuclei, detecting α-decay chains of synthesized superheavy nuclei (SHN) is an important method for identifying them. Consequently, α-decay remains a prominent and active area of study in nuclear physics [3-5].

Many empirical formulas have been proposed to study α-decay half-live. The earliest law for α-decay half-lives was formulated by Geiger and Nutall [6], which states that the logarithm of α-decay half-lives log log10T1/2 is linearly correlated with the reciprocal of the square of α-decay energy . Subsequently, many additional empirical formulas have been introduced. Royer developed an analytical formula for α-decay under the framework of the generalized liquid drop model (GLDM) [7, 8]. According to the Viola and Seaborg (VSS) formula, certain parameters are used to describe α-decay half-lives by considering the unpaired nucleon blocking effect [9]. A unified description (UD) formula for the half-life and decay energy of complex cluster radioactivity was presented by Ni et al. [10]. Ren et al. improved the Geiger–Nuttall (G–N) law for α-decay half-lives by incorporating a parameter representing the shell effect in 2012 [11]. Qi proposed a universal decay law (UDL) using the structure of the R-matrix and microscopic mechanism of charged particle emission [12]. Additionally, improved formulas were proposed by adding asymmetry terms or angular momentum effects in Refs. [13, 14, 14-22].

In Refs. [23-30], the authors demonstrated that the deformation of nuclei plays an essential role in α-decay. The total energy, fission barrier, and Coulomb interaction of nuclei have been recalculated by incorporating deformation effects. The penetration probability of α particles is affected by the deformation effects [31]. The deformation and orientation of the daughter nucleus can change the slope and intercept of the linear relationship between log10T1/2 and Q1/2 [32]. For SHN, the effects of nuclear deformation on α-decay half-lives were explored [33]. Studies investigating the impact of deformation effects on α-decay half-lives have been conducted based on macro-micro and microscopic models. The results indicate that considering deformation effects in the theoretical model can better describe the experimental values of α-decay half-lives. In this study, the effect of deformation on α-decay half-lives was investigated using an empirical formula. Within the Wentzel-Kramers-Brillouin (WKB) framework, the Coulomb potential containing the quadrupole deformation of daughter nuclei is considered as the total potential energy of the radioactive system. With some approximations and simplifications, the relationship between the deformation term and α-decay half-life is derived. By incorporating the deformation and angular momentum effects [13, 19, 20], an improved empirical formula was used to investigate α-decay half-lives of even Z-even N, even Z-odd N, odd Z-even N, and odd Z-odd N nuclei with 62Z118.

Recently, machine learning (ML) algorithms have been used as powerful alternative tools for studying and predicting complex data in nuclear physics [34, 35]. In nuclear structure, ML is used to predict nuclear masses, binding energies, charge radii, etc. [36-49]. ML is a valuable tool for constructing predictive models for nuclear reactions. These models help describe and infer important nuclear reaction data, such as refining the description of reaction data, exploring the initial state of nuclei and reaction geometry, and understanding the reaction mechanism and phase transition [50-53]. Additionally, ML has several advantages in nuclear experiments and the study of dense matter properties [54-59]. Motivated by these advancements, we applied ML to investigate α-decay half-lives of known and unknown nuclei, demonstrating the advantage of this approach in studying α-decay half-lives. In this study, the XGBoost model, which is a powerful and efficient ML algorithm based on gradient boosting decision trees, was employed to further reduce the deviations between experimental and calculated α-decay half-lives of nuclei with 62Z118. Additionally, we employed XGBoost models along with an improved formula to predict the behavior of unknown superheavy nuclei.

The remainder of this paper is organized as follows. In Sect. 2, a theoretical method for deducing the improved formula and XGBoost model is presented. The results and discussion are presented in Sect. 3. Finally, a brief summary is presented in the final section.

2

Theoretical framework

2.1
Improved formula incorporating deformation effect

In 1928, α-decay was described as a quantum tunneling phenomenon through the potential barrier. In the framework of the semiclassical approximation, the α-decay width (or decay constant) is given by Γln2/T1/2=P0FP, (1) where P0 is the preformation probability of the α cluster in the parent nucleus. F is the frequency of the α cluster within the barrier, and P is the probability of transmission through the barrier, which is given by the WKB approximation as follows: P=exp(2RinRout2μ|V(r)Q|dr). (2) Here, Rin is the touching radius, given by Rin = + Rd, where and Rd are the hard-sphere radii of the α cluster and daughter nuclei, respectively. Rout is the classical turning point and is given by Rout=ZαZde2/Q [10]. μ=AαAd/(Aα+Ad) is the reduced mass of the α-core system. Aα and Ad are the mass numbers of the α cluster and daughter nucleus, respectively.

The deformed Coulomb potential [25, 60] is given by V(r)=ZαZde2r[1+3R025r2β2Y20(θ)], (3) where R0=1.15Ad1/3 (fm). Zα and Zd represent the atomic numbers of the α cluster and daughter nucleus, respectively. β2 is the quadrupole deformation parameter of the daughter nucleus, and Y20(θ) is a harmonic function, where θ is the orientation angle of the emitted alpha particle relative to the symmetric axis of the deformed daughter nucleus. Regarding angle θ, previous studies have shown that the penetration probability P=0πP(θ)sin(θ)dθ/2 is predominantly concentrated around θ=0, which is a pole-to-pole state [61-63]. This alignment minimizes the potential barrier, thereby maximizing penetration probability. Therefore, in this study, adopting θ=0 simplifies the model while maintaining its accuracy. Combining the above results, probability P can be expressed as follows: P=exp(2RinRout2μ|(ZαZde2rQ)(1+0.162ZαZde2Ad2/3β2r3(ZαZde2rQ))|dr), (4) By expanding the right-hand term under the square root of Eq. (4) to the first order, we obtain log10P=2ln102μ[RinRout(ZαZde2rQ)1/2dr+0.081ZαZde2Ad2/3β2RinRout1r3(ZαZde2rQ)1/2dr]=F1+F2, (5) where F1 and F2 are F1=2ln102μRinRout(ZαZde2rQ)1/2dr,F2=812μZαZde2Ad2/3β2500ln10RinRout1r3(ZαZde2rQ)1/2dr. (6) As Rout=ZαZde2/Q, F1 can be obtained as follows: F1=22e2ln10μZαZdQ1/2×[acrcos(Rin/Rout)(Rin/Rout)1(Rin/Rout)2]. (7) As the first approximation of the last part of Eq. (7), we obtain F1=2πe2ln10μZαZdQ1/2+4e2Rinln10μ(ZαZd)1/2. (8) For α-decay, the variation in Rin is minimal. Therefore, Rin and Rin are treated as constants here. Equation (8) can be simplified to F1=c1μZαZdQ1/2+c2μ(ZαZd)1/2, (9) where c1 and c2 are constant coefficients.

Similarly, F2 can be expressed as F2=812μAd2/3β21500ln10ZαZde2[(ZαZde2RoutQ)3/2+(ZαZde2RinQ)3/2]+812μAd2/3β2Q500ln10ZαZde2(ZαZde2RoutQ+ZαZde2RinQ). (10) As Rout=ZαZde2/QZαZde2RoutQ=0, Eq. (10) can be simplified as follows: F2=812μAd2/3β21500ln10ZαZde2(ZαZde2Rin+2Q) ×(ZαZde2Rin)1/2(1QRinZαZde2)1/2. (11) Because QRinZαZde21 and considering the first approximation of the expression in the last part of Eq. (11), we obtain F2=81μ1500(Rin)3/2Ad2/3(ZαZde2)1/2β2+81μ1000(Rin)1/2Ad2/3(ZαZde2)1/2β2Q+161μ(Rin)1/21500Ad2/3(ZαZde2)3/2β2Q2=c3μ(ZαZd)1/2Ad2/3β2+c4μ(ZαZd)1/2QAd2/3β2+c5μ(ZαZd)3/2Q2Ad2/3β2, (12) where c3, c4, and c5 are constants. Similarly, as (ZαZd)3/2Q2Ad2/3(ZαZd)1/2QAd2/3, Eq. (12) can be simplified to F2=c3μ(ZαZd)1/2Ad2/3β2+c4μ(ZαZd)1/2QAd2/3β2. (13) According to Eq. (1), we have log10T1/2=log10(ln2/P0F)log10P. (14) In Ref. [10], P0=10c6μ(ZαZd)1/2+c7, where c6 and c7 are constants based on experimental results. By combining the results of Eqs. (5), (8), and (13), the right-hand side of Eq. (14) can be expressed as the sum of the following terms: μZαZdQ1/2, μ(ZαZd)1/2, μ(ZαZd)1/2Ad2/3β2, μ(ZαZd)1/2QAd2/3β2, and a constant. The equation for α-decay half-lives can be written as log10T1/2=aμZαZdQ1/2+bμ(ZαZd)1/2+c1μ(ZαZd)1/2Ad2/3β2+c2μ(ZαZd)1/2QAd2/3β2+h, (15) where a, b, c1, c2, and h are constants to be determined. Additionally, the influence of angular momentum on α-decay half-lives was studied in Refs. [13, 19, 20]. In this paper, we propose an improved empirical formula to investigate the deformation effect on α-decay half-lives, which is given as log10T1/2=aμZαZdQ1/2+bμ(ZαZd)1/2+c1μ(ZαZd)1/2Ad2/3β2+c2μ(ZαZd)1/2QAd2/3β2+dl(l+1)+h, (16) where T1/2 (s) is the half-life of α-decay, and Q (MeV) is the α-decay energy. β2 denotes quadrupole deformation of the daughter nucleus. l is the minimum angular momentum carried away by an α particle [20]. a, b, c1, c2, d, and h are constants to be determined.

To compare the calculated results from Eq. (16) with those obtained when the quadrupole deformation of the daughter nuclei is not taken into account, Eq. (16) becomes log10T1/2=aμZαZdQ1/2+bμ(ZαZd)1/2+dl(l+1)+h. (17)

2.2
Methodology of eXtreme Gradient Boosting (XGBoost) for α-decay half-lives

The XGBoost algorithm is an ensemble ML algorithm that operates within a gradient boosting framework. It utilizes gradient-boosted decision trees, a technique that significantly enhances performance and speed compared with traditional methods. This algorithm efficiently handles classification, regression, and ranked objective functions and offers a more effective solution than its counterparts, such as Decision Trees (DTs) and Random Forests (RFs). In gradient boosting, a series of weak learners or models are combined to create a robust final model, which is derived by considering the weighted sum of all learned models. XGBoost excels in handling small-to-medium-sized, structured, or low-dimensional datasets, making it suitable for various ML problems.

In contrast to the regularized greedy forest model, XGBoost introduces a unique regularized learning objective. This objective is notable for its simplicity, which enhances the efficiency and effectiveness of the model. The predictive function of XGBoost, a key component of this approach, is expressed as follows: y^i=k=1Kfk(x1),fkΓ. (18) In this expression, y^i represents the predicted output value for the i-th instance, k denotes the number of regression trees used, and fk represents the structure of each tree. The feature vector for the i-th sample is denoted by xi, and Γ represents the space of all possible regression trees.

To address overfitting, a penalty function is introduced to regularize the learning weights as follows: Obj(t)=[l(yty^t(t1))+gtft(xt)+12htft2(xt)]+Ω(ft), (19) where gi=y^(t1)l(yi,y^(t1)) and hi=y^(t1)2l(yiy^(k1)) are the first- and second-order gradient statistics of the loss function, respectively. The constant term is eliminated, yielding the objective function for the i-th step as follows: Obj(t)=t=1n[gift(xl)+12htft2(xl)]+Ω(ft). (20) The parameters in Eq. (20) can be updated continuously until the stopping criteria are satisfied. Further details on XGBoost can be found in Ref. [64].

In the context of α-decay half-lives, XGBoost considers the experimental α-decay half-lives as the ground truth and learns from the residuals between these data and the predictions made by the empirical formulas in Eqs. (17) and (16). This residual-based learning allows the model to implicitly account for missing physical factors, such as shell effects. The input features include the quadrupole deformation values (β2) of the daughter nucleus derived from the WS4 and FRDM models, along with the reduced mass (μ) of the α-core system, proton numbers (Zd, Zc) of the daughter nucleus and α cluster, α-decay energy, and other relevant nuclear properties. These features allow the model to identify complex correlations, such as those related to magic and sub-magic numbers, that are not encoded in empirical formulas. Owing to these advantages, XGBoost enhances the ability to capture nonlinear relationships and systematic discrepancies, particularly in regions where empirical formulas typically fail, such as near-shell closures. The XGBoost model is formulated as Rn=T1/2exp(xn)T1/2cal(xn)=k=1Kfk(x1),fkΓ, (21) where Rn represents the residuals for the n-th nucleus, T1/2exp(xn) is the experimental α-decay half-life, and T1/2cal(xn) is the calculated α-decay half-life obtained using the empirical formula. The XGBoost model, through its sophisticated learning mechanism, is trained to minimize these residuals, thereby improving the alignment between theoretical predictions and experimental data. The optimization of model parameters, as described in Eq. (20), focuses on minimizing the residuals to achieve the highest possible prediction accuracy.

3

Results and discussion

3.1
Nuclear deformation effect on α-decay half-lives with improved formula

In this study, the α-decay half-lives of 675 nuclei were extracted, including 183 even-even nuclei, 188 even Z-odd N nuclei, 178 odd Z-even N nuclei, and 126 odd Z-odd N nuclei. The experimental α-decay half-lives, spins, and parities were obtained from the evaluated properties table NUBASE2020 [65], and the α-decay energy was obtained from the evaluated atomic mass table AME2020 [66]. The quadrupole deformation values of the daughter nuclei in Eq. (16) were obtained from the WS4 [67] and FRDM [68] mass models. For different nuclei types, the parameters in Eqs. (17) and (16) were obtained by fitting the experimental α-decay half-lives. The fitting coefficients are listed in Tables 1, 2, and 3. Note that the parameters in Table 1 are fitted by the experimental α-decay half-lives of the nuclei with an absolute quadrupole deformation of less than 0.1. The parameters in Tables 2 and 3 were fitted using the experimental α-decay half-lives of all nuclei.

Table 1
Parameter values of Eq. (17)
Nuclei a b d h
Even Z-even N 0.4052 -1.5073 0 -12.1265
Even Z-odd N 0.4101 -1.5008 0.04556 -12.6194
Odd Z-even N 0.4181 -1.4863 0.05270 -14.0310
Odd Z-odd N 0.4163 -1.4782 0.04998 -13.6438
Show more
Table 2
Parameter values of Eq. (16) obtained by fitting the experimental α-decay half-lives, where the values of the quadrupole deformation of the daughter nuclei were taken from the WS4 model [67]
Nuclei a b c1 c2 d h
Even Z-even N 0.4124 -1.5354 -0.0072 0.1536 0 -12.2610
Even Z-odd N 0.4138 -1.4461 -0.0032 0.0693 0.0453 -14.4611
Odd Z-even N 0.4148 -1.4657 -0.0059 0.1316 0.0507 -14.1217
Odd Z-odd N 0.4145 -1.4866 0.0074 -0.0178 0.0511 -13.2610
Show more
Table 3
Same as Table 2, but the values of the quadrupole deformation of the daughter nuclei were taken from the FRDM model [68]
Nuclei a b c1 c2 d h
Even Z-even N 0.4117 -1.5500 -0.0067 0.1527 0 -11.8213
Even Z-odd N 0.4140 -1.4474 -0.0031 0.0647 0.0453 -14.4561
Odd Z-even N 0.4146 -1.4723 -0.0061 0.1420 0.0511 -13.9411
Odd Z-odd N 0.4176 -1.4695 -0.0020 0.0946 0.0516 -14.0118
Show more

To precisely compare the different formulas, the root mean square deviation (RMSD) of the nuclei can be calculated as follows: σ=i=1n(log10T1/2callog10T1/2exp)i2/n, (22) where T1/2cal is the calculated α-decay half-life and T1/2exp is the experimental α-decay half-life. The RMSDs for the different formulas are listed in Table 4. Using the quadrupole deformation values of the daughter nuclei in Eq. (16), taken from the WS4 and FRDM models, the corresponding σ values are denoted as σ2 and σ3, respectively. Additionally, σ1, σ4, and σ5 represent the σ values between the experimental data and calculated results obtained from Eq. (17) and other formulas [8, 13], respectively.

Table 4
Root mean square deviations between experimental α-decay half-lives and the calculated results obtained by Eq. (17), Eq. (16), and other formulas [8, 13]. σ1, σ4, and σ5 represent the σ deviations between the experimental data and calculated results from Eq. (17) and other formulas [8, 13], respectively. With the quadrupole deformation values of the daughter nuclei in Eq. (16) taken from the WS4 and FRDM models, the corresponding σ deviations are denoted as σ2 and σ3, respectively
  Eq. (17) Eq. (16) Ref. [8] Ref. [13]  
Nuclei σ1 σ2 σ3 σ4 σ5 No. of nuclei
Even Z-even N 0.406 0.363 0.368 0.376 0.375 183
Even Z-odd N 0.475 0.445 0.445 0.668 0.457 188
Odd Z-even N 0.469 0.412 0.409 0.653 0.425 178
Odd Z-odd N 0.472 0.432 0.439 0.634 0.475 126
Total 0.456 0.413 0.415 0.589 0.428 675
Show more

When the quadrupole deformation of the daughter nuclei in Eq. (16) is taken from the WS4 mass model, the σ values decrease from 0.406 to 0.363 for even Z-even N nuclei, from 0.475 to 0.445 for even Z-odd N nuclei, from 0.469 to 0.412 for odd Z-even N nuclei, and from 0.472 to 0.432 for odd Z-odd N nuclei. A similar trend was observed when the quadrupole deformation values were taken from the FRDM model. Among all subsets, the σ values for even Z-even N nuclei were the lowest, with σ1, σ2, and σ3 values of 0.406, 0.363, and 0.368, respectively. This can be attributed to the fact that in the α-decay of even Z-even N nuclei, the angular momentum l carried away by the α particle is always zero, eliminating the ambiguity in determining l, which is not true for other subsets. For the total dataset, the σ2 and σ3 values, when compared with the σ1 value from Eq. (17), decrease from 0.456 to 0.413 and 0.415, respectively. From the σ values, including deformation effects in the empirical formula does not significantly improve describing experimental α-decay half-lives.

Additionally, by fixing the parameters a, b, d, and h in Eq. (17), the deformation-related parameters c1 and c2 are independently fitted by the experimental α-decay half-lives. Using this approach, Eq. (16) produces σ= 0.419 and 0.421 with the quadrupole deformation values obtained from WS4 and FRDM, respectively. Compared with the values of σ2 and σ3 shown in Table 4, their σ values are higher. Hence, the results calculated from Eq. (16), where the parameters of Eq. (16) are fitted using the experimental α-decay half-lives of all nuclei, will be used in the subsequent discussions on XGBoost optimization and the prediction of α-decay half-lives for superheavy nuclei.

In Fig. 1, the σ deviations of Eqs. (17) and (16) across different quadrupole deformation regions are presented, with the deformation values derived from the WS4 and FRDM models, respectively. For nuclei with small deformations (|β2|0.05), the difference between the two formulas (Eqs. (17) and (16)) is negligible. This suggests that the deformation effect is minimal in this range. In the intermediate deformation range (0.05<|β2|0.15), the σ values in Eq. (16) are slightly lower than those in Eq. (17), indicating a better consistency with the experimental data in this range. For nuclei with large deformation (|β2| > 0.15), the deformation-inclusive equation (Eq. (16)) shows a further reduction in the σ values, indicating that considering deformation improves the agreement with experimental data for highly deformed nuclei. Overall, the σ deviations in Eq. (16) are consistently smaller than those in Eq. (17) across all deformation ranges, as shown in Figs. 1(a) and 1(b). Although deformation has little impact on near-spherical nuclei, it becomes increasingly relevant as deformation increases. Although the improved formula does not significantly reduce the overall σ value, it achieves better consistency with experimental data for highly deformed nuclei, with minimal effect on less deformed nuclei.

Fig. 1
(Color online) Root mean square deviations of Eqs. (17) and (16) for nuclei across different deformation regions of the daughter nuclei, with the quadrupole deformation values of daughter nuclei taken from WS4 (a) and FRDM (b) models. The horizontal axis represents the range of deformation values for the daughter nuclei, whereas the vertical axis indicates the root mean square values for nuclei within each deformation range. n represents the number of nuclei present in each deformation interval
pic

Using the RMSD as an indicator of the accuracy of the formula provides only a rough estimate of its performance. For a more detailed insight, Fig. 2 presents the logarithmic deviations between the experimental data and the results calculated using Eqs. (17) and (16). In Fig. 2(a), the black circles and red stars represent the deviations for Eqs. (17) and (16), respectively, using the deformation values from the WS4 model [67]. Similarly, Fig. 2(b) presents the same analysis using the deformation values from the FRDM model [68]. Points closer to zero indicate better agreement with experimental data, and evidently, the deviations of Eq. (16) (red stars) are more concentrated near zero than those of Eq. (17) (black circles) for both models. A notable feature in Fig. 2 is the large deviations observed near neutron numbers N = 126 and N = 152, where both formulas exhibit significant discrepancies from the experimental data. Previous studies identified N = 126 and N = 152 as magic and sub-magic numbers, respectively [69-71]. These deviations can be attributed to shell effects associated with these neutron numbers, which are not explicitly accounted for in either Eq. (17) or (16). To address this issue, in the following sections, we explore using the XGBoost algorithm to further reduce the σ values of empirical formulas, including nuclei near the magic and sub-magic numbers.

Fig. 2
(Color online) Logarithmic differences between the experimental α-decay half-lives and calculated results obtained by Eqs. (17) and (16), with values of the quadrupole deformation of the daughter nuclei taken from the WS4 (a) and FRDM (b) models, respectively. The dashed lines represent the neutron numbers N = 126, 152, and the deviations of the decimal logarithms of T1/2cal/T1/2exp are 0, -1, and 1, respectively
pic

To gain further insight into the deformation effect on α-decay half-lives, we compared the improved formula of Eq. (16) with other formulas [8, 13]. The experimental data used in this study differ from those in Refs. [8] and [13]. The parameters of the empirical formulas in Refs. [8, 13] were recalibrated using the experimental data obtained in this study. This recalibration resulted in σ values of 0.589 and 0.428, respectively, which were lower than the original σ values of 0.779 and 0.429 with the parameters obtained from Refs. [8] and Ref. [13], respectively. The recalibrated σ values for the four subsets and total dataset are displayed in the fifth and sixth columns of Table 4, respectively. As indicated in the table, the improved formula of Eq. (16) exhibits smaller σ values for the entire set of data than those of the other formulas [8, 13]. This indicates that the improved formula, which incorporates the deformation effect, can be used to study α-decay half-lives.

3.2
Study on α-decay half-lives with eXtreme Gradient Boosting

In this study, we built three XGBoost models to further improve the prediction accuracy of α-decay half-lives, denoted as XGBoost1, XGBoost2, and XGBoost3. For the XGBoost2 and XGBoost3 models, the quadrupole deformation values of the daughter nuclei were obtained from the WS4 and FRDM models, respectively. Additionally, in the XGBoost models, the physical terms of Eqs. (17) and (16) were used as inputs, and the residuals between the calculated and experimental α-decay half-lives served as the output.

The experimental data for 675 nuclei were extracted and divided into two sets: the training set (540 nuclei) and testing set (135 nuclei). The corresponding σ values calculated using the formulas and XGBoost1, XGBoost2, and XGBoost3 models are presented in Table 5. The σ values of the XGBoost1, XGBoost2, and XGBoost3 models show an obvious decrease compared with those of Eqs. (17) and (16) for the training and testing sets. For the entire set, the σ values of XGBoost1, XGBoost2, and XGBoost3 models reduced from 0.456 to 0.316, from 0.413 to 0.295, and from 0.415 to 0.302, respectively. This indicates that XGBoost models can further reduce the deviation between the calculated and experimental α-decay half-lives.

Table 5
Root mean square deviations between experimental α-decay half-lives and calculated results using Eq. (17), Eq. (16), XGBoost1, XGBoost2, and XGBoost3 for training, testing, and entire sets
  inputs σ (training set) σ (testing set) σ (entire set)
Eq. (17) - 0.462 0.418 0.456
XGBoost1 (, μ, Zc, Zd, l) 0.310 0.346 0.316
(Eq. (16), β2WS4) - 0.427 0.345 0.413
XGBoost2 (, μ, Zc, Zd, l, β2WS4) 0.288 0.318 0.295
(Eq. (16), β2FRDM) - 0.430 0.345 0.415
XGBoost3 (, μ, Zc, Zd, l, β2FRDM) 0.301 0.308 0.302
Show more

The deviations between the experimental data and results calculated using the empirical formula (blue circles) and XGBoost models (red stars) are illustrated in Fig. 3. As shown in Fig. 3, the majority of blue circles are scattered in the range of ± 1, suggesting larger deviations. By contrast, most red stars fall within the range of ± 0.5, indicating a closer alignment to the zero line. For nuclei near the neutron magic number N = 126 and sub-magic number N = 152, the deviations obtained by XGBoost models are smaller than those from the empirical formulas in Eqs. (17) and (16). To further quantify these differences, Fig. 4 shows the σ values in Eq. (17), Eq. (16), and the XGBoost models (XGBoost1, XGBoost2, and XGBoost3) for nuclei with N=125, 126, 127, 151, 152, and 153. The comparison shows that the σ values for the XGBoost models were reduced from 0.623 to 0.562, from 0.685 to 0.533, and from 0.687 to 0.543, respectively, indicating better consistency with the experimental data than with the empirical formulas. For these nuclei, the larger deviations of the empirical formulas can be attributed to their lack of consideration of shell effects. By contrast, the XGBoost models effectively capture missing physical effects, such as shell effects, by learning from the residuals between the empirical formulas and experimental data. Consequently, the XGBoost models further reduced the deviations for nuclei near the neutron magic number N = 126 and sub-magic number N = 152, providing more accurate predictions.

Fig. 3
(Color online) Logarithmic differences between the experimental data and calculated α-decay half-lives using (a) Eq. (17) and XGBoost1, (b) Eq. (16) and XGBoost2, and (c) Eq. (16) and XGBoost3. The values of quadrupole deformation of the daughter nuclei in Eq. (16) are taken from the WS4 and FRDM models, respectively. The dashed lines represent the deviation of the decimal logarithms of T1/2cal/T1/2exp of 0, -1, and 1, respectively
pic
Fig. 4
(Color online) Root mean square deviations between the experimental data and calculated α-decay half-lives using Eq. (17), Eq. (16), XGBoost1, XGBoost2, and XGBoost3 models for the 53 nuclei at N= 125, 126, 127, 151, 152, and 153
pic
3.3
Predictions of α-decay half-lives of nuclei with Z = 117, 118, 119, and 120

Using the improved formula of Eq. (16), along with the XGBoost2 and XGBoost3 models, we predict the α-decay half-lives of SHN with Z = 117, 118, 119, and 120. In Table 6, the first column lists the α-decay, followed by the α-decay energy and quadrupole deformation of the daughter nuclei (columns 2-5), obtained from WS4 [67] and FRDM [68] mass models, respectively. The spin and parity values of the parent and daughter nuclei in the nuclear ground state were taken from Ref. [72]. Because of challenges in determining the spin and parity of doubly odd SHN, we assume that the minimum angular momentum l carried away by the α particle is zero. As shown in the eighth to last columns, the decimal logarithms of α-decay half-lives, calculated using the improved formula Eq. (16), are shown as log10T1/2Cal1 and log10T1/2Cal2, with the values of the quadrupole deformation taken from WS4 and FRDM models, respectively. Additionally, log10T1/2XGBoost1 and log10T1/2XGBoost2 were calculated using the XGBoost2 and XGBoost3 models, respectively.

Table 6
α-decay half-lives of nuclei with Z = 117, 118, 119, and 120, with the α-decay energy and quadrupole deformation obtained by the WS4 [67] and FRDM [68] mass models, respectively. The logarithms of α-decay half-lives were calculated using Eq. (16), denoted by log10T1/2Cal1 and log10T1/2Cal2, with the values of the quadrupole deformation obtained from the WS4 and FRDM models, respectively. Additionally, log10T1/2XGBoost1 and log10T1/2XGBoost2 were calculated using the XGBoost2 and XGBoost3 models, respectively
α transition QαWS4 (MeV) QαFRDM (MeV) β2WS4 β2FRDM l log10T1/2Cal1 log10T1/2XGBoost1 log10T1/2Cal2 log10T1/2XGBoost2
279Ts→275Mc 13.36 12.97 0.1792 0.198 2 -5.734 -5.391 -4.904 -4.306
280Ts→276Mc 13.42 12.79 0.1825 0.198 0 -4.876 -5.454 -4.097 -4.388
281Ts→277Mc 13.21 12.32 0.1822 0.198 2 -5.440 -5.619 -3.633 -3.525
282Ts→278Mc 13.00 12.59 0.1803 0.032 0 -4.039 -4.495 -4.376 -4.512
283Ts→279Mc 12.86 12.96 -0.0141 -0.011 2 -5.150 -4.800 -5.402 -4.909
284Ts→280Mc 12.66 12.65 -0.0239 0.021 0 -5.023 -5.008 -4.548 -4.447
285Ts→281Mc 12.42 12.45 0.0404 0.032 2 -4.123 -4.319 -4.240 -4.287
286Ts→282Mc 12.24 12.18 -0.0573 0.064 0 -4.385 -4.455 -3.357 -3.201
287Ts→283Mc 12.02 12.11 -0.0719 0.064 2 -3.403 -3.743 -3.441 -3.446
288Ts→284Mc 11.95 12.01 0.0618 0.064 0 -2.743 -3.121 -2.976 -3.170
289Ts→285Mc 11.96 11.98 0.0707 0.064 2 -3.076 -2.888 -3.152 -2.743
290Ts→286Mc 11.81 11.85 0.0676 0.075 0 -2.364 -2.774 -2.567 -2.752
291Ts→287Mc 11.69 11.75 -0.1016 0.064 2 -2.647 -2.806 -2.630 -2.549
292Ts→288Mc 11.72 11.71 0.0654 0.064 0 -2.178 -2.141 -2.282 -2.021
293Ts→289Mc 11.60 11.40 0.062 0.064 2 -2.255 -2.294 -1.805 -1.565
294Ts→290Mc 11.35 11.29 0.0544 0.053 0 -1.363 -1.204 -1.306 -0.965
295Ts→291Mc 11.27 11.55 -0.0729 -0.042 2 -1.570 -1.741 -2.312 -2.218
296Ts→292Mc 11.48 11.64 -0.062 -0.042 0 -2.678 -2.359 -2.511 -2.007
297Ts→293Mc 11.59 11.76 -0.0555 -0.021 2 -2.363 -2.621 -2.786 -2.712
298Ts→294Mc 11.49 11.86 -0.0392 -0.021 0 -2.523 -2.811 -2.956 -3.001
299Ts→295Mc 11.43 11.86 -0.0334 -0.021 2 -1.951 -1.686 -3.021 -2.512
300Ts→296Mc 11.53 11.88 -0.0319 -0.011 0 -2.536 -2.401 -2.964 -2.679
301Ts→297Mc 11.59 11.93 -0.0276 -0.011 2 -2.323 -1.957 -3.166 -2.612
302Ts→298Mc 12.20 12.75 -0.0154 -0.011 0 -3.959 -4.070 -4.891 -4.898
303Ts→299Mc 12.75 12.74 -0.0087 0 2 -4.918 -4.509 -4.918 -4.417
304Ts→300Mc 12.52 12.82 -0.0145 0 0 -4.642 -5.088 -4.988 -5.366
305Ts→301Mc 12.06 12.69 -0.0214 0 2 -3.413 -3.229 -4.813 -4.475
306Ts→302Mc 11.58 12.40 -0.0277 0.011 0 -2.629 -2.864 -4.046 -4.195
307Ts→303Mc 10.90 12.24 -0.0466 0.011 1 -0.785 -0.473 -4.026 -3.480
308Ts→304Mc 10.33 11.49 -0.0575 0.021 0 0.375 0.186 -1.908 -1.883
309Ts→305Mc 9.99 10.80 -0.1822 -0.011 3 2.346 2.788 -0.048 0.579
310Ts→306Mc 9.73 10.37 -0.1855 0.044 0 0.974 1.273 1.115 1.415
311Ts→307Mc 9.42 10.15 -0.2 0.054 2 3.924 3.842 1.488 1.311
312Ts→308Mc 9.05 9.78 -0.2111 0.065 0 2.968 2.905 2.923 2.805
281Og→277Lv 13.77 13.15 0.1787 0.186 5 -5.274 -5.254 -4.129 -3.825
282Og→278Lv 13.49 13.12 0.1799 0 0 -6.582 -6.431 -6.335 -5.864
283Og→279Lv 13.32 13.20 0.1792 0.032 7 -3.262 -3.395 -3.171 -3.062
284Og→280Lv 13.21 13.56 -0.0262 0 0 -6.475 -6.502 -7.191 -7.133
285Og→281Lv 13.05 13.24 0.0363 0.032 0 -5.399 -5.362 -5.788 -5.692
286Og→282Lv 12.89 13.05 0.0479 0.053 0 -5.698 -5.463 -6.066 -5.741
287Og→283Lv 12.77 12.96 0.056 0.064 0 -4.820 -4.884 -5.204 -5.084
288Og→284Lv 12.59 12.86 -0.0765 0.064 0 -5.257 -5.026 -5.662 -5.215
289Og→285Lv 12.56 12.76 0.0711 0.075 0 -4.370 -4.664 -4.787 -4.914
290Og→286Lv 12.57 12.68 0.075 0.075 0 -5.008 -4.771 -5.273 -4.896
291Og→287Lv 12.39 12.55 0.075 0.075 2 -3.728 -3.781 -4.075 -3.934
292Og→288Lv 12.21 12.39 0.0735 0.075 0 -4.248 -4.392 -4.671 -4.676
293Og→289Lv 12.21 12.34 0.0711 0.075 2 -3.344 -3.807 -3.624 -3.867
294Og→290Lv 12.17 12.37 0.0687 0.064 0 -4.163 -3.963 -4.649 -4.223
295Og→291Lv 11.88 11.92 0.0617 0.064 0 -2.857 -3.270 -2.962 -3.192
296Og→292Lv 11.73 12.28 0.0532 -0.073 0 -3.176 -3.102 -4.703 -4.420
297Og→293Lv 12.08 12.38 -0.0789 -0.063 0 -3.385 -3.801 -4.055 -4.241
298Og→294Lv 12.16 12.49 -0.0651 -0.042 0 -4.279 -3.902 -5.112 -4.549
299Og→295Lv 12.02 12.51 -0.0507 -0.021 2 -2.968 -2.939 -4.044 -3.709
300Og→296Lv 11.93 12.51 -0.0441 -0.011 0 -3.733 -3.926 -5.093 -5.066
301Og→297Lv 12.00 12.56 -0.0396 0 2 -2.903 -3.012 -4.138 -4.000
302Og→298Lv 12.02 12.62 -0.0336 0 0 -3.923 -3.913 -5.305 -5.175
303Og→299Lv 12.58 13.38 -0.0197 0 4 -3.566 -3.666 -5.179 -5.227
304Og→300Lv 13.10 13.39 -0.0136 0 0 -6.232 -5.888 -6.863 -6.450
305Og→301Lv 12.89 13.45 -0.0188 0 2 -4.837 -4.617 -5.950 -5.621
306Og→302Lv 12.46 13.35 -0.0264 0 0 -4.906 -5.248 -6.785 -7.042
307Og→303Lv 11.90 12.57 -0.0397 0.011 2 -2.690 -2.802 -4.152 -4.016
308Og→304Lv 11.18 12.10 -0.0484 0 0 -1.893 -1.810 -4.168 -3.900
309Og→305Lv 10.70 11.07 -0.0584 0.022 2 0.359 0.422 -0.626 -0.288
310Og→306Lv 10.41 10.74 -0.1812 0.044 0 0.268 0.534 -0.795 -0.415
311Og→307Lv 10.08 9.98 -0.2049 -0.042 2 2.177 2.561 2.441 2.917
312Og→308Lv 9.74 10.05 -0.207 0.055 0 2.406 1.999 1.131 0.706
313Og→309Lv 8.63 9.72 -0.2135 0.066 0 6.890 7.432 2.925 3.570
285119→281Ts 13.60 14.06 0.0304 0 2 -5.998 -5.813 -6.979 -6.712
286119→282Ts 13.42 13.75 0.0401 0.043 0 -5.494 -5.914 -6.112 -6.442
287119→283Ts 13.26 13.37 0.0524 0.064 2 -5.288 -5.535 -5.497 -5.466
288119→284Ts 13.20 13.57 -0.0751 0.075 0 -6.040 -5.900 -5.620 -5.264
289119→285Ts 13.13 13.47 -0.0845 0.075 2 -5.310 -5.027 -5.656 -5.025
290119→286Ts 13.04 13.31 0.0721 0.075 0 -4.465 -4.804 -5.117 -5.316
291119→287Ts 13.02 13.24 0.0773 0.075 2 -4.773 -4.409 -5.215 -4.657
292119→288Ts 12.87 13.08 0.0776 0.086 0 -4.073 -3.611 -4.610 -4.013
293119→289Ts 12.69 12.92 0.0746 0.075 2 -4.099 -3.580 -4.581 -3.869
294119→290Ts 12.70 12.85 0.0739 0.075 0 -3.731 -3.474 -4.187 -3.790
295119→291Ts 12.73 12.94 0.0732 0.075 2 -4.191 -4.541 -4.620 -4.776
296119→292Ts 12.45 12.98 0.0733 0.075 0 -3.191 -3.550 -4.452 -4.671
297119→293Ts 12.40 12.90 -0.0856 -0.073 2 -3.716 -3.561 -4.890 -4.418
298119→294Ts 12.69 13.09 -0.0806 -0.073 0 -5.037 -5.387 -5.335 -5.430
299119→295Ts 12.74 13.08 -0.0721 -0.052 2 -4.453 -4.813 -5.218 -5.312
300119→296Ts 12.55 13.04 -0.0601 -0.032 0 -4.559 -4.939 -5.048 -5.185
301119→297Ts 12.40 13.08 -0.0494 -0.032 0 -3.977 -4.229 -5.474 -5.483
302119→298Ts 12.40 13.05 -0.0431 0 0 -4.096 -4.639 -4.925 -5.282
303119→299Ts 12.39 13.11 -0.0355 0 2 -3.632 -3.256 -5.147 -4.523
304119→300Ts 12.91 13.86 -0.02 0 0 -4.972 -4.737 -6.519 -6.215
305119→301Ts 13.40 13.86 0.0147 0 2 -5.650 -5.384 -6.607 -6.250
306119→302Ts 13.18 13.92 -0.0198 0 0 -5.525 -5.464 -6.631 -6.501
307119→303Ts 12.76 13.39 -0.0287 0 0 -4.721 -5.152 -6.012 -6.358
308119→304Ts 12.04 12.17 -0.0396 0.011 0 -3.247 -2.919 -2.970 -2.456
309119→305Ts 11.35 11.70 -0.0508 0 3 -0.836 -1.338 -1.722 -2.005
310119→306Ts 10.86 10.67 -0.0616 -0.032 0 -0.517 -0.651 0.666 0.752
311119→307Ts 10.76 10.32 -0.1895 0 0 0.098 0.171 1.315 1.434
312119→308Ts 10.56 9.90 -0.1936 -0.397 0 -0.908 -0.512 1.916 2.463
313119→309Ts 9.37 10.06 -0.2055 0.066 2 4.825 5.333 2.398 2.812
314119→310Ts 8.97 9.55 -0.2136 -0.407 0 3.896 4.091 3.085 3.233
315119→311Ts 8.65 9.36 -0.2137 -0.406 1 7.318 7.005 4.635 4.535
316119→312Ts 8.67 9.20 -0.2369 -0.406 0 7.050 6.877 5.404 5.308
317119→313Ts 9.20 9.06 -0.4273 -0.416 3 5.859 5.518 6.462 6.180
287120→283Og 13.84 13.96 -0.0531 0.043 4 -5.595 -5.706 -5.729 -5.590
288120→284Og 13.71 13.85 -0.0663 0.064 0 -7.039 -7.317 -7.068 -7.147
289120→285Og 13.69 13.75 -0.0797 0.075 0 -6.253 -6.497 -6.213 -6.227
290120→286Og 13.68 13.75 -0.0872 0.075 0 -7.028 -6.996 -6.852 -6.641
291120→287Og 13.48 13.87 -0.096 0.075 2 -5.586 -6.056 -6.164 -6.394
292120→288Og 13.44 13.78 0.0788 0.086 0 -6.222 -6.226 -6.874 -6.743
293120→289Og 13.37 13.65 0.0786 0.086 2 -5.206 -4.958 -5.741 -5.304
294120→290Og 13.22 13.49 -0.1098 0.086 0 -6.132 -5.759 -6.338 -5.857
295120→291Og 13.25 13.46 0.0762 0.075 2 -4.954 -4.862 -5.386 -5.101
296120→292Og 13.32 13.59 0.0752 0.075 0 -5.989 -6.199 -6.554 -6.624
297120→293Og 13.12 13.65 0.0751 0.075 0 -4.970 -5.153 -6.021 -6.065
298120→294Og 12.98 13.24 -0.0863 0.064 0 -5.595 -5.896 -5.915 -6.019
299120→295Og 13.23 13.73 -0.0849 -0.084 0 -5.343 -5.652 -6.325 -6.379
300120→296Og 13.29 13.70 -0.075 -0.063 0 -6.233 -6.400 -7.155 -7.105
301120→297Og 13.04 13.62 -0.065 -0.042 2 -4.654 -4.944 -5.800 -5.842
302120→298Og 12.87 13.55 0.0382 -0.032 0 -5.155 -5.582 -6.773 -7.067
303120→299Og 12.79 13.51 -0.0448 -0.011 2 -4.104 -3.690 -5.556 -4.860
304120→300Og 12.74 13.55 -0.0355 0 0 -4.999 -5.520 -6.684 -7.019
305120→301Og 13.26 14.26 -0.0214 0 2 -5.060 -4.988 -6.954 -6.729
306120→302Og 13.77 14.28 0.0139 0 0 -6.971 -6.621 -8.042 -7.641
307120→303Og 13.50 13.62 0.0191 0 4 -4.873 -4.879 -5.125 -5.078
308120→304Og 12.95 12.96 -0.0297 0 0 -5.426 -5.655 -5.503 -5.616
309120→305Og 12.14 11.77 -0.0416 -0.021 0 -2.932 -3.279 -2.053 -2.157
310120→306Og 11.48 11.29 -0.0527 0 0 -2.063 -1.779 -1.673 -1.204
311120→307Og 11.18 10.76 -0.1882 -0.397 2 -0.300 -0.385 0.877 1.003
312120→308Og 11.20 10.71 -0.1933 -0.397 0 -1.344 -1.316 -0.251 -0.074
313120→309Og 11.01 10.50 -0.194 -0.407 2 0.160 0.431 1.648 2.098
314120→310Og 10.74 10.33 -0.2019 -0.407 0 -0.043 0.109 0.961 1.208
315120→311Og 9.41 10.16 -0.2323 -0.407 0 4.792 5.516 2.425 2.993
316120→312Og 9.17 9.94 0.3893 -0.416 0 4.035 3.735 2.277 2.114
317120→313Og 9.91 9.83 -0.4157 -0.416 3 3.757 4.187 4.041 4.669
318120→314Og 9.91 9.66 -0.425 -0.416 0 2.772 2.722 3.266 3.408
Show more

The calculated α-decay half-lives of the nuclei with Z = 117, 118, 119, and 120 are plotted in Fig. 5, respectively. As shown in Fig. 5, the calculated α-decay half-lives for 293Ts, obtained using the improved formula (log10T1/2Cal1 and log10T1/2Cal2) and the XGBoost models(log10T1/2XGBoost1 and log10T1/2XGBoost2), are in good agreement with the experimental α-decay half-lives. For 294Og, the calculated α-decay half-life obtained by the XGBoost2 model was much closer to the experimental α-decay half-lives than those obtained by other models. Fig. 5 reveals that when neutron numbers surpass N = 184, the α-decay half-lives of nuclei Z = 117, 118, 119, and 120 exhibit a rapid decrease of more than 1.8 orders of magnitude. This phenomenon reflects strong shell effects, suggesting that N = 184 is a neutron magic number. Additionally, the decimal logarithms of α-decay half-life values for nuclei with Z = 117, 118, 119, and 120 ranged from -8 to 10. If these nuclei could be synthesized in a laboratory, the majority could be detected experimentally.

Fig. 5
(Color online) Logarithms of α-decay half-lives of nuclei with Z = 117, 118, 119, and 120 using Eq. (16), XGBoost2, and XGBoost3 with obtained by the WS4 [67] and FRDM [68] nuclear mass table. The black dashed lines denote the neutron number N = 184
pic
4

Summary

In summary, this study investigated the impact of deformation effects on α-decay half-lives using an improved formula within the WKB framework. By incorporating the quadrupole deformation of the daughter nucleus into the Coulomb potential, the improved formula provides a refined description of the α-decay half-lives of nuclei with 62Z118. Although the improved formula does not significantly reduce the overall σ value, it provides calculations that align more closely with the experimental values for nuclei with larger deformations. Furthermore, XGBoost models have been employed to further reduce the deviations between experimental and calculated α-decay half-lives using the improved formula. The results suggest that the XGBoost model can effectively reduce the RMSDs between the calculated and experimental values across all nuclei, including those near magic and sub-magic numbers. Additionally, the improved formula and XGBoost models that yield low σ values were extended to predict α-decay half-lives for nuclei with Z = 117, 118, 119, and 120. When neutron numbers surpass N = 184, the α-decay half-lives of nuclei Z = 117, 118, 119, and 120 exhibit a rapid decrease of more than 1.8 orders of magnitude. These results indicate that N = 184 is a magic number.

References
1.R.W. Gurney, E.U. Condon,

Wave mechanics and radioactive disintegration

. Nature 122, 1476-4687 (1928). https://doi.org/10.1038/122439a0
Baidu ScholarGoogle Scholar
2.G. Gamow,

Zur quantentheorie des atomkernes

. Z. Phys. 51, 204-212 (1928). https://doi.org/10.1007/BF01343196
Baidu ScholarGoogle Scholar
3.K. Wei, Y.L. Ye, Z.H. Yang,

Clustering in nuclei: progress and perspectives

. Nucl. Sci. Tech. 35, 216 (2024). https://doi.org/10.1007/s41365-024-01588-x
Baidu ScholarGoogle Scholar
4.M.H. Zhang, Y.H. Zhang, Y. Zou et al.,

Possibilities for the synthesis of superheavy element Z = 121 in fusion reactions

. Nucl. Sci. Tech. 35, 95 (2024). https://doi.org/10.1007/s41365-024-01452-y
Baidu ScholarGoogle Scholar
5.P.H. Chen, C. Geng, Z.X. Yang et al.,

Production of neutron-rich actinide isotopes in isobaric collisions via multinucleon transfer reactions

. Nucl. Sci. Tech. 34, 160 (2023). https://doi.org/10.1007/s41365-023-01314-z
Baidu ScholarGoogle Scholar
6.H. Geiger, J.M. Nuttall,

LVII. the ranges of the α particles from various radioactive substances and a relation between range and period of transformation

. Philos. Mag. 22, 613-621 (1911). https://doi.org/10.1080/14786441008637156
Baidu ScholarGoogle Scholar
7.G. Royer,

Alpha emission and spontaneous fission through quasi-molecular shapes

. J. Phys. G: Nucl. Part. Phys. 26, 1149-1170 (2000). https://doi.org/10.1088/0954-3899/26/8/305
Baidu ScholarGoogle Scholar
8.G. Royer,

Analytic expressions for alpha-decay half-lives and potential barriers

. Nucl. Phys. A 848, 279-291 (2010). https://doi.org/10.1016/j.nuclphysa.2010.09.009
Baidu ScholarGoogle Scholar
9.V.E. Viola, G.T. Seaborg,

Nuclear systematics of the heavy elements-II lifetimes for alpha, beta and spontaneous fission decay

. J. Ino. Nucl. Chem. 28, 741-761 (1966). https://doi.org/10.1016/0022-1902(66)80412-8
Baidu ScholarGoogle Scholar
10.D.D. Ni, Z.Z. Ren, T.K. Dong et al.,

Unified formula of half-lives forα decay and cluster radioactivity

. Phys. Rev. C 78, 044310 (2008). https://doi.org/10.1103/PhysRevC.78.044310
Baidu ScholarGoogle Scholar
11.Y.J. Ren, Z.Z. Ren,

New geiger-nuttall law for α decay of heavy nuclei

. Phys. Rev. C 85, 044608 (2012). https://doi.org/10.1103/PhysRevC.85.044608
Baidu ScholarGoogle Scholar
12.C. Qi, F.R. Xu, R.J. Liotta et al.,

Universal decay law in charged-particle emission and exotic cluster radioactivity

. Phys. Rev. Lett. 103, 072501 (2009). https://doi.org/10.1103/PhysRevLett.103.072501
Baidu ScholarGoogle Scholar
13.J.G. Deng, H.F. Zhang, G. Royer,

Improved empirical formula forα-decay half-lives

. Phys. Rev. C 101, 034307 (2020). https://doi.org/10.1103/PhysRevC.101.034307
Baidu ScholarGoogle Scholar
14.V.Y. Denisov, A.A. Khudenko,

α-decay half-lives: Empirical relations

. Phys. Rev. C 79, 054614 (2009). https://doi.org/10.1103/PhysRevC.79.054614
Baidu ScholarGoogle Scholar
15.D.N. Poenaru, D.T. Akrawy,

Alpha decay calculations with a new formula

. J. Phys. G: Nucl. Part. Phys. 44, 105105 (2017). https://doi.org/10.1088/1361-6471/aa8527
Baidu ScholarGoogle Scholar
16.D.T. Akrawy, A.H. Ahmed,

New empirical formula forα-decay calculations

. Int. J. Mod. Phys. E 27, 1850068 (2018). https://doi.org/10.1142/s0218301318500684
Baidu ScholarGoogle Scholar
17.E. Shin, Y. Lim, C.H. Hyun et al.,

Nuclear isospin asymmetry inα decay of heavy nuclei

. Phys. Rev. C 94, 024320 (2016). https://doi.org/10.1103/PhysRevC.94.024320
Baidu ScholarGoogle Scholar
18.J.M. Dong, H.F. Zhang, Y.Z. Wang et al.,

Alpha-decay for heavy nuclei in the ground and isomeric states

. Nucl. Phys. A 832, 198-208 (2010). https://doi.org/10.1016/j.nuclphysa.2009.10.082
Baidu ScholarGoogle Scholar
19.Z.Y. Wang, Z.M. Niu, Q. Liu et al.,

Systematic calculations ofα-decay half-lives with an improved empirical formula

. J. Phys. G: Nucl. Part. Phys. 42, 055112 (2015). https://doi.org/10.1088/0954-3899/42/5/055112
Baidu ScholarGoogle Scholar
20.D.T. Akrawy, H. Hassanabadi, Y. Qian et al.,

Influence of nuclear isospin and angular momentum onα-decay half-lives

. Nucl. Phys. A 983, 310-320 (2019). https://doi.org/10.1016/j.nuclphysa.2018.10.091
Baidu ScholarGoogle Scholar
21.D.T. Akrawy, A.H. Ahmed,

α-decay systematics for superheavy nuclei

. Phys. Rev. C 100, 044618 (2019). https://doi.org/10.1103/PhysRevC.100.044618
Baidu ScholarGoogle Scholar
22.Y.B. Qian, Z.Z. Ren,

Unfavoredα decay from ground state to ground state in the range 53 ≤ Z ≤91

. Phys. Rev. C 85, 027306 (2012). https://doi.org/10.1103/PhysRevC.85.027306
Baidu ScholarGoogle Scholar
23.S.Q. Guo, X.J. Bao, Y. Gao et al.,

The nuclear deformation and the preformation factor in the α-decay of heavy and superheavy nuclei

. Nucl. Phys. A 934, 110-120 (2015). https://doi.org/10.1016/j.nuclphysa.2014.12.001
Baidu ScholarGoogle Scholar
24.K.M.K. Sharma,

Cluster radioactivity within the collective fragmentation approach using different mass tables and related deformations

. Eur. Phys. J. A 56, 35 (2020). https://doi.org/10.1140/epja/s10050-020-00023-0
Baidu ScholarGoogle Scholar
25.Y.B. Qian, Z.Z. Ren, D.D. Ni,

Calculations of α-decay half-lives for heavy and superheavy nuclei

. Phys. Rev. C 83, 044317 (2011). https://doi.org/10.1103/PhysRevC.83.044317
Baidu ScholarGoogle Scholar
26.C. Xu, Z.Z. Ren,

Global calculation of alpha-decay half-lives with a deformed density-dependent cluster model

. Phys. Rev. C 74, 014304 (2006). https://doi.org/10.1103/PhysRevC.74.014304
Baidu ScholarGoogle Scholar
27.W. Seif, A. Adel, N. Antonenko et al.,

Enhanced α decays to negative-parity states in even-even nuclei with octupole deformation

. Phys. Rev. C 107, 044601 (2023). https://doi.org/10.1103/PhysRevC.107.044601
Baidu ScholarGoogle Scholar
28.W. Yahya, O. Azeez, J. Majekodunmi et al.,

Density-dependent parametrizations in B3Y-Fetal NN interaction: Application to alpha decay

. Int. J. Theo. Phys 54, 74 (2024). https://doi.org/10.1007/s13538-024-01453-7
Baidu ScholarGoogle Scholar
29.H.F. Gui, X.J. Liu, H. Mand Wu, P.C. Chu et al.,

Systematic study of αdecay half-lives for even–even nuclei within a deformed two-potential approach

. Commun. Theor. Phys. 74, 055301 (2022). https://doi.org/10.1088/1572-9494/ac6576
Baidu ScholarGoogle Scholar
30.Y. Xin, N.N. Ma, J.G. Deng et al.,

Properties of Z = 114 super-heavy nuclei

. Nucl. Sci. Tech. 32, 55 (2021). https://doi.org/10.1007/s41365-021-00899-7
Baidu ScholarGoogle Scholar
31.S.H. Cheng, Z.S. Ge, L.G. Cao et al.,

Theoretical calculations of the nuclear deformation effects on α-decay half-lives for heavy and super-heavy nuclei

. J. Phys. G 48, 095106 (2021). https://doi.org/10.1088/1361-6471/ac165f
Baidu ScholarGoogle Scholar
32.D. Naderi, M. Zargooshi,

Study of alpha-decay half-lives with deformed, oriented daughter nuclei

. Int. J. Mod. Phys. E 24, 1550010 (2015). https://doi.org/10.1142/S021830131550010X
Baidu ScholarGoogle Scholar
33.M. Pahlavani, S. Alavi, N. Tahanipour,

Effect of nuclear deformation on the potential barrier and alpha-decay half-lives of superheavy nuclei

. Mod. Phys. Lett. A 28, 1350065 (2013). https://doi.org/10.1142/S021773231350065X
Baidu ScholarGoogle Scholar
34.A. Boehnlein, M. Diefenthaler, N. Sato et al.,

Colloquium: Machine learning in nuclear physics

. Rev. Mod. Phys. 94, 031003 (2022). https://doi.org/10.1103/RevModPhys.94.031003
Baidu ScholarGoogle Scholar
35.W.B. He, Q.F. Li, Y.G. Ma et al.,

Machine learning in nuclear physics at low and intermediate energies

. Sci. China. Phys. Mech. 66, 282001 (2023). https://doi.org/10.1007/s11433-023-2116-0
Baidu ScholarGoogle Scholar
36.B.S. Cai, C. Yuan,

Random forest-based prediction of decay modes and half-lives of superheavy nuclei

. Nucl. Sci. Tech. 34, 204 (2023). https://doi.org/10.1007/s41365-023-01354-5
Baidu ScholarGoogle Scholar
37.S. Akkoyun, T. Bayram, S.O. Kara et al.,

An artificial neural network application on nuclear charge radii

. J. Phys. G: Nucl. Part. Phys. 40, 055106 (2013). https://doi.org/10.1088/0954-3899/40/5/055106
Baidu ScholarGoogle Scholar
38.T. Bayram, S. Akkoyun, S.O. Kara,

A study on ground-state energies of nuclei by using neural networks

. Ann. Nucl. Energy 63, 172-175 (2014). https://doi.org/10.1016/j.anucene.2013.07.039
Baidu ScholarGoogle Scholar
39.W.C. Chen, R. Utama, J. Piekarewicz,

Nuclear charge radii: density functional theory meets bayesian neural networks

. J. Phys. G 43, 114002 (2016). https://doi.org/10.1088/0954-3899/43/11/114002
Baidu ScholarGoogle Scholar
40.L. Neufcourt, Y. Cao, W. Nazarewicz et al.,

Bayesian approach to model-based extrapolation of nuclear observables

. Phys. Rev. C 98, 034318 (2018). https://doi.org/10.1103/PhysRevC.98.034318
Baidu ScholarGoogle Scholar
41.Z.M. Niu, H.Z. Liang,

Nuclear mass predictions based on bayesian neural network approach with pairing and shell effects

. Phys. Lett. B 778, 48-53 (2018). https://doi.org/10.1016/j.physletb.2018.01.002
Baidu ScholarGoogle Scholar
42.Z.M. Niu, H.Z. Liang, B.H. Sun et al.,

Predictions of nuclear β-decay half-lives with machine learning and their impact on r-process nucleosynthesis

. Phys. Rev. C 99, 064307 (2019). https://doi.org/10.1103/PhysRevC.99.064307
Baidu ScholarGoogle Scholar
43.G. Saxena, P.K. Sharma, P. Saxena,

Modified empirical formulas and machine learning for α-decay systematics

. J. Phys. G: Nucl. Part. Phys. 48, 055103 (2021). https://doi.org/10.1088/1361-6471/abcd1c
Baidu ScholarGoogle Scholar
44.Z.M. Niu, H.Z. Liang,

Nuclear mass predictions with machine learning reaching the accuracy required by r-process studies

. Phys. Rev. C 106, L021303 (2022). https://doi.org/10.1103/PhysRevC.106.L021303
Baidu ScholarGoogle Scholar
45.H.Q. You, Z.Z. Qu, R.H. Wu et al.,

Study on nuclear α-decay energy by an artificial neural network with pairing and shell effects

. Symmetry 5, 1006 (2022). https://doi.org/10.3390/sym14051006
Baidu ScholarGoogle Scholar
46.C.Q. Li, C.N. Tong, H.J. Du et al.,

Deep learning approach to nuclear masses and α-decay half-lives

. Phys. Rev. C 105, 064306 (2022). https://doi.org/10.1103/PhysRevC.105.064306
Baidu ScholarGoogle Scholar
47.A.E. Lovell, A.T. Mohan, T.M. Sprouse et al.,

Nuclear masses learned from a probabilistic neural network

. Phys. Rev. C 106, 014305 (2022). https://doi.org/10.1103/PhysRevC.106.014305
Baidu ScholarGoogle Scholar
48.X.X. Dong, R. An, J.X. Lu et al.,

Novel Bayesian neural network based approach for nuclear charge radii

. Phys. Rev. C 105, 014308 (2022). https://doi.org/10.1103/PhysRevC.105.014308
Baidu ScholarGoogle Scholar
49.Z. Yuan, D. Bai, Z. Wang et al.,

Reliable calculations of nuclear binding energies by the gaussian process of machine learning

. Nucl. Sci. Tech. 35, 105 (2024). https://doi.org/10.1007/s41365-024-01463-9
Baidu ScholarGoogle Scholar
50.Q.F. Song, L. Zhu, J. Su,

Target dependence of isotopic cross sections in the spallation reactions 238U + p and 9Be at 1 AGeV *

. Chin. Phys. C 46, 074108 (2022). https://doi.org/10.1088/1674-1137/ac6249
Baidu ScholarGoogle Scholar
51.T.X. Huang, X.H. Wu, P.W. Zhao,

Application of kernel ridge regression in predicting neutron-capture reaction cross-sections

. Commun. Theor. Phys. 74, 095302 (2022). https://doi.org/10.1088/1572-9494/ac763b
Baidu ScholarGoogle Scholar
52.F. Li, Y. Wang, Z. Gao et al.,

Application of machine learning in the determination of impact parameter in the 132Sn+124Sn system

. Phys. Rev. C 104, 034608 (2021). https://doi.org/10.1103/PhysRevC.104.034608
Baidu ScholarGoogle Scholar
53.X. Zhang, X. Liu, Y. Huang et al.,

Determining impact parameters of heavy-ion collisions at low-intermediate incident energies using deep learning with convolutional neural networks

. Phys. Rev. C 105, 034611 (2022). https://doi.org/10.1103/PhysRevC.105.034611
Baidu ScholarGoogle Scholar
54.L. Yang, C.J. Lin, Y.X. Zhang et al.,

Bayesian analysis on interactions of exotic nuclear systems

. Phys. Lett. B 807, 135540 (2020). https://doi.org/10.1016/j.physletb.2020.135540
Baidu ScholarGoogle Scholar
55.Y.D. Song, R. Wang, Y.G. Ma et al.,

Determining the temperature in heavy-ion collisions with multiplicity distribution

. Phys. Lett. B 814, 136084 (2021). https://doi.org/10.1016/j.physletb.2021.136084
Baidu ScholarGoogle Scholar
56.M. Omana Kuttan, J. Steinheimer, K. Zhou et al.,

The QCD EoS of dense nuclear matter from Bayesian analysis of heavy ion collision data

. arXiv e-prints arXiv:2211.11670 (2022)., https://doi.org/10.48550/arXiv.2211.11670
Baidu ScholarGoogle Scholar
57.W.J. Xie, B.A. Li,

Bayesian inference of the symmetry energy of superdense neutron-rich matter from future radius measurements of massive neutron stars

. Astrophys. J 899, 4 (2020). https://doi.org/10.3847/1538-4357/aba271
Baidu ScholarGoogle Scholar
58.S.L. Chen, T.X. Wang, Z. Zhang et al.,

Linear regression and machine learning for nuclear forensics of spent fuel from six types of nuclear reactors

. Phys. Rev. Appl. 19, 034028 (2023). https://doi.org/10.1103/PhysRevApplied.19.034028
Baidu ScholarGoogle Scholar
59.T. Li, Y. Chen, S.B. Wang et al.,

Reconstruction of the event vertex in the PandaX-III experiment with convolution neural network

. J. High. Engergy. Phys. 2023, 200 (2023). arXiv:2211.14992, https://doi.org/10.1007/JHEP05(2023)200
Baidu ScholarGoogle Scholar
60.V.Y. Denisov, A. Khudenko,

α-decay half-lives, α-capture, and α-nucleus potential

. At. Data Nucl. Data Tables 95, 815-835 (2009). https://doi.org/10.1016/j.adt.2009.06.003
Baidu ScholarGoogle Scholar
61.M. Ismail, A. Ellithi, M. Botros et al.,

Penetration factor in deformed potentials: Application to αdecay with deformed nuclei

. Phys. Rev. C 86, 044317 (2012). https://doi.org/10.1103/PhysRevC.86.044317
Baidu ScholarGoogle Scholar
62.R.K. Gupta, A. Sǎndulescu, W. Greiner,

Interaction barriers, nuclear shapes and the optimum choice of a compound nucleus reaction for producing super-heavy elements

. Phys. Lett. B 67, 257-261 (1977). https://doi.org/10.1016/0370-2693(77)90364-1
Baidu ScholarGoogle Scholar
63.N. Malhotra, R.K. Gupta,

Proximity potential for deformed, oriented collisions and its application to 238U + 238U

. Phys. Rev. C 31, 1179 (1985). https://doi.org/10.1103/PhysRevC.31.1179
Baidu ScholarGoogle Scholar
64.T. Chen, C. Guestrin, in Proceedings of the 22nd ACM SIGKDD International Conference on Knowledge Discovery and Data Mining,

Xgboost: A scalable tree boosting system. KDD ’16

, (Association for Computing Machinery, New York, NY, USA, 2016), p. 785-794. https://doi.org/10.1145/2939672.2939785
Baidu ScholarGoogle Scholar
65.F. Kondev, M. Wang, W. Huang et al.,

The NUBASE2020 evaluation of nuclear physics properties *

. Chin. Phys. C 45, 030001 (2021). https://doi.org/10.1088/1674-1137/abddae
Baidu ScholarGoogle Scholar
66.M. Wang, W. Huang, F. Kondev et al.,

The AME 2020 atomic mass evaluation (ii). tables, graphs and references*

. Chin. Phys. C 45, 030003 (2021). https://doi.org/10.1088/1674-1137/abddaf
Baidu ScholarGoogle Scholar
67.N. Wang, M. Liu, X. Wu et al.,

Surface diffuseness correction in global mass formula

. Phys. Lett. B 734, 215-219 (2014). https://doi.org/10.1016/j.physletb.2014.05.049
Baidu ScholarGoogle Scholar
68.P. Möller, A. Sierk, T. Ichikawa et al.,

Nuclear ground-state masses and deformations: Frdm(2012)

. At. Data Nucl. Data Tables 109-110, 1-204 (2016). https://doi.org/10.1006/adnd.1995.1002
Baidu ScholarGoogle Scholar
69.X.J. Bao, H.F. Zhang, G. Royer et al.,

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

. Nucl. Phys. A 906, 1-13 (2013). https://doi.org/10.1016/j.nuclphysa.2013.03.002
Baidu ScholarGoogle Scholar
70.G. Lalazissis, M. Sharma, P. Ring et al.,

Superheavy nuclei in the relativistic mean-field theory

. Nucl. Phys. A 608, 202-226 (1996). https://doi.org/10.1016/0375-9474(96)00273-4
Baidu ScholarGoogle Scholar
71.R. Smola n ńczuk, J. Skalski, A. Sobiczewski,

Spontaneous-fission half-lives of deformed superheavy nuclei

. Phys. Rev. C 52, 1871-1880 (1995). https://doi.org/10.1103/PhysRevC.52.1871
Baidu ScholarGoogle Scholar
72.https://t2.lanl.gov/nis/data/astro/molnix96/spidat.html.
Footnote

The authors declare that they have no competing interests.