Introduction
Nuclear level density (NLD) is the basic physical input for nuclear reactions. This is the key component for the calculation of reaction cross-sections relevant to nucleosynthesis [1-6]. The study of NLDs dates back to the 1930s, with Bethe’s pioneering work [7]. Since then, many theoretical models such as the back-shifted Fermi gas model (BFM) [8], composite constant temperature model (CTM) [9] and generalized superfluid model (GSM) [10] have been adopted for NLD studies. These phenomenological models are widely used for nuclear reaction calculations. Phenomenological models rely on experimental data for adjusting the parameters; however, experimental data are limited, especially for nuclei far from the β-stability line [11]. To address these difficulties, many microscopic methods have been developed.
Over the past decades, various microscopic approaches for NLD have been proposed, including the equidistant spacing model [12-15], the shell-model Monte Carlo method [16-20], the spectral distribution calculation [21-23], the independent particle model at finite temperature [24-27], the micro-statistical methods [28-31] and the random matrix method [32]. Recently, a stochastic estimation method for the level density within the framework of the configuration-interaction shell model (CISM) was proposed [33] and applied in the calculation of NLDs of fission products 133-137Xe and 135-138Ba [34]. Over the past two decades, microscopic methods based on the Hartree–Fock–Bogoliubov (HFB) combinatorial model [11] have been developed. The idea of using a combinatorial method to calculate the level densities was derived from the calculation of excitation-state densities [35]. After successfully describing the excitation state densities using combinations of nucleons occupying single-particle levels at the mean field, it was natural to further describe the level densities by considering the collective effects [36]. This combinatorial approach is on par with statistical methods with respect to reproducing experimental data and can provide energy-, spin-, and parity-dependent NLDs beyond the scope of statistical methods [37]. The nonrelativistic HFB combinatorial methods based on Skyrme and Gogny effective interactions have successfully reproduced NLDs for various nuclei [37, 38] and have been applied to astrophysical reactions. The accuracy of NLD is related to the basic information about the nuclear structure, such as single-particle levels, deformation, and binding energy. In recent years, the covariant (relativistic) density functional theory has attracted considerable attention in the field of nuclear physics owing to its successful description of complex nuclear structures and reaction dynamics [39-45]. For instance, it can satisfactorily reproduce the isotopic shifts in Pb isotopes [46] and can naturally explain the origin of the pseudospin and spin symmetries in the antinucleon spectrum [47], as well as provide a good description of the nuclear magnetic moments [48, 49]. Recently, a microstatistical method based on the CDFT was developed to describe NLDs [50]. The method was applied to the calculation of NLDs of 94,96,98Mo and 106,108Pd as well as the odd-A nuclei at saddle point [51], which were in good agreement with experimental data over the entire energy range of the measured values [50]. While the microstatistical method can only calculate energy-dependent NLDs, the combinatorial method can calculate energy-, spin-, and parity-dependent NLDs. Therefore, it is meaningful to calculate the NLDs using the CDFT combinatorial method.
The theoretical framework and methods are introduced in Sect. 2. The NLDs calculated using the CDFT combinatorial method are compared with other NLD predictions and experimental results in Sect. 3. Finally, the conclusions and future prospects are presented in Sect. 4.
Theoretical Framework
Covariant density functional theory starts from a Lagrangian, and the corresponding Kohn-Sham equations have the form of a Dirac equation with effective fields S and V derived from this Lagrangian [39, 41, 49, 52]. Specifically, the nucleons in the nucleus are described as Dirac particles moving in the average potential field provided by the meson and photon fields, interacting with each other through the exchange of mesons and photons. The Dirac equation is written as follows:
Upon obtaining the energy ε, spin projection m, and parity p of the single-particle levels using CDFT, the level information is substituted into the generating function defined in the combinatorial method [11] to obtain the particle-hole state density ρi; the generating function 𝒵 is expressed as:
Summing all the
Finally, a phenomenological damping function is introduced to avoid sharp transitions between the spherical and deformed level densities that affect the NLD predictions. The expression for the NLD after introduction of the damping function is [31]:
Results and Discussion
In this section, we present our results for the NLDs obtained using the combinatorial method based on the CDFT and compare them with the results obtained using other NLD models [57] and experimental data [58, 59]. The effective meson-exchange interaction parameter PK1 [60] is adopted throughout the CDFT calculations. For spherical CDFT calculations, we fixed the box size as Rbox = 20 fm and the step size as Δr = 0.1 fm. In the present deformed CDFT calculations, the Dirac equation for nucleons and the Klein–Gordon equations for mesons are solved using an isotropic harmonic oscillator, and a basis for the 18 major oscillator shells is adopted.
The combinatorial results rely on the properties of single-particle levels. To understand the effect of using different effective interactions on the NLDs, the NLDs of 112Cd and 162Dy obtained with different CDFT effective interactions, namely, PK1 [60], NL3 [61], DD-ME2 [62], and DD-PC1 [63] are presented for comparison in Fig. 1. For nuclei with small deformations (112Cd), the NLDs for the four effective interactions are close for excitation energies above 4 MeV. However, significant differences are observed for excitation energies below 4 MeV. This difference arises from the fact that the single-particle energies near the Fermi level are exceptionally sensitive to the choice of effective interaction, especially when the Fermi level is near the proton or neutron shells. Meanwhile, the CDFT calculations with the chosen interactions do not adequately reproduce the single-particle levels of the magic nuclei 132Sn [64], which explains the deviations of our results from the measurements presented in Fig.1. For the well-deformed 162Dy, the NLDs calculated with the four effective interactions deviate from each other in the entire region of the excitation energy, although the overall trend is consistent. This deviation at low excitation energies is mainly caused by differences in the single-particle energies around the Fermi level. Meanwhile, the entropy S obtained from the four effective interactions is significantly different for nuclei with large deformations, which ultimately leads to deviation of the NLDs at high excitation energies. Compared with the experimental data of 112Cd [58] and 162Dy [59] extracted by the Oslo group based on the analysis of particle-γ coincidence in the (3He, αγ) and (3He, 3He’γ) reactions, the deviation between the NLDs obtained by using PK1 effective interaction and the results obtained by using other effective interactions is within a reasonable range.
The nonrelativistic HFB combinatorial methods, including Skyrme and Gogny interactions, have been widely used in NLD predictions [55, 58]; the NLDs of spherical nuclei (55Co, 132Sn, and 208Pb) and deformed nuclei (94Mo, 95Mo, and 166Er) calculated using the CDFT combinatorial method are compared with the corresponding experimental data, as illustrated in Fig. 2.
For 55Co, the CDFT combinatorial method yields a lower NLD than the experimental value, whereas the Skyrme combinatorial method yields a higher NLD. Although the results of the Gogny combinatorial method are closer to the experimental data to a certain extent, a strong oscillation exists. For 132Sn and 208Pb, the results of the CDFT combinatorial method are similar to those of the Skyrme combinatorial method; however, the results of the Gogny combinatorial method are significantly lower than those of the two methods. Compared to the experimental data, the CDFT combinatorial method can better describe the NLD of 208Pb. The NLDs of 55Co, 132Sn, and 208Pb obtained using the three combinatorial methods are oscillatory to a certain extent because the three nuclei are spherical with highly degenerate single-particle levels with a strong shell effect.
For 94Mo, the results of both the CDFT and Skyrme combinatorial methods are closer to the experimental data than those of the Gogny combinatorial method. For 95Mo and 166Er, the NLDs obtained using the three combinatorial methods are similar and slightly higher than the experimental data. It should be emphasized that the deformed nuclei 94Mo, 95Mo, and 166Er break the single-particle level degeneracy, yielding smoother NLDs, as shown in Fig. 2. It is noteworthy that the CDFT combinatorial method yields higher NLDs for 94Mo and 95Mo. One possible reason for this is that the CDFT combinatorial method uses a rigid-body value when considering the rotational effect, which is inappropriate for soft nuclei [69].
In addition, the NLDs of 55Co and 94Mo can be adequately described by the microstatistical model [31] relative to the CDFT combinatorial method, while both models provide similar descriptions for the NLDs of 95Mo and 166Er. In comparison, the microstatistical model yields relatively smooth results for spherical and deformed nuclei, which do not satisfactorily reflect the shell effect. Therefore, it can be concluded that the CDFT combinatorial method is as capable of describing NLDs as the other two combinatorial methods and the microstatistical model.
Furthermore, it is necessary to compare the NLDs obtained in the study with those obtained using phenomenological models (BFM, CTM, and GSM) [57]. Phenomenological models can adequately describe the experimental data at neutron separation by fitting the experimental data. The level densities at the neutron separation energy are the most commonly used experimental data and are obtained using the mean distance D0 of the s-wave resonance. The NLDs of the even-even nuclei (112Cd and 162Dy), odd-A nuclei (51V, 97Mo, and 119Sn), and odd-odd nucleus (60Co) calculated using the CDFT combinatorial method are compared with those obtained using phenomenological models and the available experimental data, as demonstrated in Fig. 3. As shown in Fig. 3, the CDFT combinatorial method is as capable of reproducing the experimental data on the neutron separation energy as the phenomenological models; in particular, the CDFT combinatorial method can adequately describe the NLD of 51V. In addition, the overall description of the experimental data below the neutron separation energy provided by the CDFT combinatorial method is similar to that of the phenomenological methods. In conclusion, although there are few differences in the results for each NLD model, the overall ability to describe the experimental data is similar.
The most extensive and reliable sources of experimental information on NLD are the s-wave neutron resonance spacing D0 [56] and observed low-energy excited levels [56]. To measure the dispersion between the theoretical and experimental D0, the frms factor is defined as:
-202309/1001-8042-34-09-012/alternativeImage/1001-8042-34-09-012-F004.jpg)
When phenomenological NLDs are used in nuclear physics applications, such as nuclear data evaluation or accurate and reliable estimation of reaction cross-sections, a few parameters are used considering the dependence of the phenomenological expressions [71]. The results of the combinatorial method can also normalize both the experimental-level scheme at a low excitation energy and the neutron resonance spacing at U=Sn in a manner similar to what is typically used for analytical formulas. More precisely, the normalized level density can be obtained as: [37]
As an illustration, the variation in the NLDs obtained using the CDFT combinatorial method before and after normalization is shown in Fig. 5. When normalization is applied, the NLDs pass through the experimental data at the neutron separation energy Sn. As shown in Fig. 5(a), if the theoretically calculated value is close to the experimental value at the neutron separation energy Sn, the results before and after normalization do not change significantly. As shown in Fig. 5(b), the results after normalization are in better agreement with the experimental data [70], particularly above 2 MeV. This is because the coefficient
-202309/1001-8042-34-09-012/alternativeImage/1001-8042-34-09-012-F006.jpg)
Finally, the NLDs calculated using the CDFT combinatorial method are compared with the observed low-energy excited levels, which constitute the most extensive and reliable source of experimental information on NLDs [56]. The cumulative number of nuclear levels N(U) indicates the sum of the number of levels below the excitation energy U (including U)
-202309/1001-8042-34-09-012/alternativeImage/1001-8042-34-09-012-F007.jpg)
In Fig. 8, the CDFT combinatorial method predictions after normalization are compared with the experimental data extracted using the Oslo method [58, 59, 66-68, 70, 72, 73] and the particle evaporation spectrum [74, 75]. The Oslo method is model-dependent. To extract the absolute value of the total level density from the measured data, the experimental NLDs must be normalized by the total level density at the neutron binding energy, which in turn is derived from the neutron resonance spacing. For a meaningful comparison between the CDFT combinatorial predictions and data obtained using the Oslo method, it is important to normalize the NLDs obtained using the CDFT combinatorial method to the level density value at U=Sn considered by the Oslo group. As shown in Fig. 8, the results of the CDFT combinatorial method after normalization agree well with the experimental data below Sn, except for the small NLDs of 111Cd and 161Dy at low excitation energies. This low result is due to the larger energy spacing of the theoretically calculated single-particle levels near the Fermi level. Overall, the results obtained using the CDFT combinatorial method are reliable.
Summary and Prospects
A combinatorial method was adopted to describe the nuclear level densities for nuclear reaction calculations. The particle-hole state density was obtained using a combinatorial method with the single-particle level provided by the PK1 effective interaction based on CDFT. Energy-, spin-, and parity-dependent NLDs were obtained after accounting for collective effects, including vibration and rotation. Our results were compared with those obtained using other NLD models, including phenomenological, microstatistical, and non-relativistic HFB combinatorial models. The comparison suggests that bar some small deviations among the different NLD models, the general trends among these models are essentially the same. In conclusion, the CDFT combination method can reproduce experimental data at or below the neutron separation energy. This implies that the CDFT combinatorial method is as reliable as the other models at describing NLDs. Finally, the comparison of the NLDs of the CDFT combinatorial method with normalization with the experimental data exhibited excellent agreement between the observed cumulative number of levels at low excitation energies and the measured NLDs below the neutron separation energy.
Our results demonstrate the predictive power of the CDFT combinatorial method. However, in our approach, pairing correlations are not considered and the collective effects are empirical. In our future work, we aim to improve upon the CDFT combinatorial method by considering the inclusion of energy-dependent pairing correlations and adopting a partition function approach to the treatment of vibrational enhancement. These results can potentially help our study of important neutron capture processes, such as the r-process.
Nuclear physics A. Stationary states of nuclei
. Rev. Mod. Phys. 8, 82-229 (1936). doi: 10.1103/RevModPhys.8.82Nuclear ground-state masses and deformations: Frdm (2012)
. At. Data Nucl. Data Tables 109–110, 1-204 (2016). doi: 10.1016/j.adt.2015.10.002Nuclear physics B. Nuclear dynamics, theoretical
. Rev. Mod. Phys. 9, 69-244 (1937). doi: 10.1103/RevModPhys.9.69The cross-section calculation of 112Sn(α,γ)116Te reaction with different nuclear models at the astrophysical energy range
. Nucl. Sci. Tech 28, 113 (2017). doi: 10.1007/s41365-017-0267-yMeasurement of 134xe(n,2n)133m, gxe reaction cross sections in 14-mev region with detailed uncertainty quantification
. Nucl. Sci. Tech. 34, 4 (2023). doi: 10.1007/s41365-022-01158-zPrediction of synthesis cross sections of new moscovium isotopes in fusion-evaporation reactions
. Nucl. Sci. Tech. 34, 7 (2023). doi: 10.1007/s41365-022-01157-0An attempt to calculate the number of energy levels of a heavy nucleus
. Phys. Rev. 50, 332-341 (1936). doi: 10.1103/PhysRev.50.332Level density parameters for the back-shifted Fermi gas model in the mass range 40 < A < 250
. Nucl. Phys. A 217, 269-298 (1973). doi: 10.1016/0375-9474(73)90196-6A composite nuclear-level density formula with shell corrections
. Can. J. Phys. 43, 1446-1496 (1965). doi: 10.1139/p65-139Global and local level density models
. Nucl. Phys. A 810, 13-76 (2008). doi: 10.1016/j.nuclphysa.2008.06.005Combinatorial nuclear level densities based on the Gogny nucleon–nucleon effective interaction
. Eur. Phys. J. A 12, 169-184 (2001). doi: 10.1007/s100500170025Particle-hole state density in the uniform spacing model
. Nucl. Phys. A 166, 231-240 (1971). doi: 10.1016/0375-9474(71)90426-XThe finite depth of the nuclear potential well in the exciton model of preequilibrium decay
. Z. Phys. A 279, 319-324 (1976). doi: 10.1007/BF01408305Particle-hole state densities for statistical multistep compound reactions
. Nucl. Phys. A 453, 127-140 (1986). doi: 10.1016/0375-9474(86)90033-3Generalized particle-hole state densities within the equidistant spacing model
. Nucl. Phys. A 632, 417-441 (1998). doi: 10.1016/S0375-9474(98)00003-7Particle-number reprojection in the shell model Monte Carlo method: application to nuclear level densities
. Phys. Rev. Lett. 83, 4265-4268 (1999). doi: 10.1103/PhysRevLett.83.4265Estimating the nuclear level density with the Monte Carlo shell model
. Phys. Rev. C 56, R1678-R1682 (1997). doi: 10.1103/PhysRevC.56.R1678Shell model Monte Carlo investigation of rare earth nuclei
. Phys. Rev. C 61, 034303 (2000). doi: 10.1103/PhysRevC.61.034303Combinatorial nuclear level density by a Monte Carlo method
. Phys. Rev. C 49, 852-866 (1994). doi: 10.1103/PhysRevC.49.852Realistic microscopic level densities for spherical nuclei
. Phys. Rev. C 50, 836-844 (1994). doi: 10.1103/PhysRevC.50.836Spectral distribution calculations of the level density of 24Mg
. Phys. Rev. C 36, 1604-1610 (1987). doi: 10.1103/PhysRevC.36.1604New expansion technique for spectral distribution calculations
. Phys. Rev. C 51, 606-610 (1995). doi: 10.1103/PhysRevC.51.606Spectral distributions in nuclei
. Phys. Rev. C 3, 94-117 (1971). doi: 10.1103/PhysRevC.3.94Testing the constant-temperature approach for the nuclear level density
. Phys. Rev. C 96, 054321 (2017). doi: 10.1103/PhysRevC.96.054321Simultaneous microscopic description of nuclear level density and radiative strength function
. Phys. Rev. Lett. 118, 022502 (2017). doi: 10.1103/PhysRevLett.118.022502Level density and thermodynamics in the hot rotating 96Tc nucleus
. Phys. Rev. C 96, 054326 (2017). doi: 10.1103/PhysRevC.96.054326S-shaped heat capacity in an odd–odd deformed nucleus
. Phys. Lett. B 789, 634-638 (2019). doi: 10.1016/j.physletb.2018.12.007Excitation energy and angular momentum dependence of nuclear level densities and spin cut-off factor in spa and spa + rpa approaches
. Nucl. Phys. A 640, 362-374 (1998).doi: 10.1016/S0375-9474(98)00462-XA new nuclear level density formula including shell and pairing correction in the light of a microscopic model calculation
. Nucl. Phys. A 605, 28-60 (1996). doi: 10.1016/0375-9474(96)00162-5On superconductivity effects in nuclear level density
. Nucl. Phys. A 110, 129-141 (1968). doi: 10.1016/0375-9474(68)90687-8Microscopic nuclear level densities for practical applications
. Nucl. Phys. A 695, 95-108 (2001). doi: 10.1016/S0375-9474(01)01095-8Colloquium: random matrices and chaos in nuclear spectra
. Rev. Mod. Phys. 79, 997-1013 (2007). doi: 10.1103/RevModPhys.79.997Stochastic estimation of nuclear level density in the nuclear shell model: an application to parity-dependent level density in 58ni
. Phys. Lett. B 753, 13-17 (2016). doi: 10.1016/j.physletb.2015.12.005Shell-model-based investigation on level density of xe and ba isotopes
. Phys. Rev. C 107, 054306 (2023). doi: 10.1103/PhysRevC.107.054306Shell effects on state densities with given numbers of excited protons and neutrons
. Nucl. Phys. A 226, 391-412 (1974). doi: 10.1016/0375-9474(74)90491-6Spectroscopy of neutronrich nickel isotopes: experimental results and microscopic interpretation
. Phys. Rev. C 37, 2600-2612 (1988). doi: 10.1103/PhysRevC.37.2600Improved microscopic nuclear level densities within the Hartree–Fock–Bogoliubov plus combinatorial method
. Phys. Rev. C 78, 064307 (2008). doi: 10.1103/PhysRevC.78.064307Temperature-dependent combinatorial level densities with the D1M Gogny force
. Phys. Rev. C 86, 064317 (2012). doi: 10.1103/PhysRevC.86.064317Relativistic mean field theory in finite nuclei
. Prog. Part. Nucl. Phys. 37, 193-263 (1996). doi: 10.1016/0146-6410(96)00054-3Relativistic continuum Hartree Bogoliubov theory for ground-state properties of exotic nuclei
. Prog. Part. Nucl. Phys. 57, 470-563 (2006). doi: 10.1016/j.ppnp.2005.06.001Relativistic Hartree–Bogoliubov theory: static and dynamic aspects of exotic nuclear structure
. Phys. Rep. 409, 101-259 (2005). doi: 10.1016/j.physrep.2004.10.001Relativistic nuclear energy density functionals: Mean-field and beyond
. Prog. Part. Nucl. Phys. 66, 519-548 (2011). doi: 10.1016/j.ppnp.2011.01.055Progress on tilted axis cranking covariant density functional theory for nuclear magnetic and antimagnetic rotation
. Front. Phys. 8, 55-79 (2013). doi: 10.1007/s11467-013-0287-yTowards an ab initio covariant density functional theory for nuclear structure
. Prog. Part. Nucl. Phys 109, 103713 (2019). doi: 10.1016/j.ppnp.2019.103713Anomaly in the charge radii of Pb isotopes
. Phys. Lett. B 317, 9-13 (1993). doi: 10.1016/0370-2693(93)91561-ZSpin symmetry in the antinucleon spectrum
. Phys. Rev. Lett. 91, 262501 (2003). doi: 10.1103/PhysRevLett.91.262501Relativistic description of magnetic moments in nuclei with doubly closed shells plus or minus one nucleon
. Phys. Rev. C 88, 064307 (2013). doi: 10.1103/PhysRevC.88.064307Magnetic moments of 33Mg in the time-odd relativistic mean field approach
. Sci. China G 52, 1586-1592 (2009). doi: 10.1007/s11433-009-0194-yMicroscopic model for the collective enhancement of nuclear level densities
. Phys. Rev. C 102, 054606 (2020). doi: 10.1103/PhysRevC.102.054606Level density of odd—a nuclei at saddle point
. Nucl. Sci. Tech. 34, 124 (2023). doi: 10.1007/s41365-023-01270-8The significance of shell corrections in the parameterization of numerical state density calculations
. Nucl. Phys. A 187, 225-248 (1972). doi: 10.1016/0375-9474(72)90576-3Nuclear level densities with collective rotations included
. Nucl. Phys. A 222, 493-511 (1974). doi: 10.1016/0375-9474(74)90334-0Global microscopic nuclear level densities within the HFB plus combinatorial method for practical applications
. Nucl. Phys. A 779, 63-81 (2006). doi: 10.1016/j.nuclphysa.2006.08.014RIPL: reference input parameter library for calculation of nuclear reactions and nuclear data evaluations
. Nucl. Data Sheets 110, 3107-3214 (2009).doi: 10.1016/j.nds.2009.10.004.(Special Issue on Nuclear Reaction Data)Modern nuclear data evaluation with the Talys code system
. Nucl. Data Sheets 113, 2841-2934 (2012). doi: 10.1016/j.nds.2012.11.002.(Special Issue on Nuclear Reaction Data)Transitional γ strength in cd isotopes
. Phys. Rev. C 87, 014319 (2013). doi: 10.1103/PhysRevC.87.014319Thermal properties and radiative strengths in 160,161,162Dy
. Phys. Rev. C 68, 064306 (2003). doi: 10.1103/PhysRevC.68.064306New effective interactions in relativistic mean field theory with nonlinear terms and densitydependent meson–nucleon coupling
. Phys. Rev. C 69, 034319 (2004). doi: 10.1103/PhysRevC.69.034319The effective force NL3 revisited
. Phys. Lett. B 671, 36-41 (2009). doi: 10.1016/j.physletb.2008.11.070Dirhb: a relativistic selfconsistent mean-field framework for atomic nuclei
. Comput. Phys. Commun. 185, 1808-1821 (2014). doi: 10.1016/j.cpc.2014.02.027New relativistic mean-field interaction with density-dependent meson–nucleon couplings
. Phys. Rev. C 71, 024312 (2005). doi: 10.1103/PhysRevC.71.024312Dynamics of nuclear single-particle structure in covariant theory of particle-vibration coupling: From light to superheavy nuclei
. Phys. Rev. C 84, 014305 (2011).doi: 10.1103/PhysRevC.84.014305Nuclear level densities of 64,66Zn from neutron evaporation
. Phys. Rev. C 88, 064324 (2013). doi: 10.1103/PhysRevC.88.064324Level densities and thermodynamical quantities of heated 93−98Mo isotopes
. Phys. Rev. C 73, 034311 (2006). doi: 10.1103/PhysRevC.73.034311Level density and γ-decay properties of closed shell Pb nuclei
. Phys. Rev. C 79, 024316 (2009). doi: 10.1103/PhysRevC.79.024316Thermal and electromagnetic properties of 166Er and 167Er
. Phys. Rev. C 63, 044309 (2001). doi: 10.1103/PhysRevC.63.044309Collective enhancements in the level densities of Dy and Mo isotopes
. Phys. Rev. C 101, 054315 (2020). doi: 10.1103/PhysRevC.101.054315Microcanonical entropies and radiative strength functions of 50,51V
. Phys. Rev. C 73, 064301 (2006). doi: 10.1103/PhysRevC.73.064301Iterative Bayesian Monte Carlo for nuclear data evaluation
. Nucl. Sci. Tech. 33, 50 (2022). doi: 10.1007/s41365-022-01034-wLevel densities and γ-ray strength functions in sn isotopes
. Phys. Rev. C 81, 064311 (2010). doi: 10.1103/PhysRevC.81.064311Level densities and γ-ray strength functions in 170,171,172Yb
. Phys. Rev. C 70, 054611 (2004). doi: 10.1103/PhysRevC.70.054611Level density and mechanism of deuteron-induced reactions on 54,56,58Fe
. Phys. Rev. C 92, 014303 (2015). doi: 10.1103/PhysRevC.92.014303Test of level density models from reactions of 6Li on 58Fe and 7Li on 57Fe
. Phys. Rev. C 80, 034305 (2009). doi: 10.1103/PhysRevC.80.034305The authors declare that they have no competing interests.