Introduction
As a fundamental nuclear property, nuclear level density (NLD) plays a crucial role in many applications, such as the calculation of reaction cross sections with nucleosynthesis, the nuclear reaction calculation program TALYS [1], and the Hauser–Feshbach model for compound nucleus calculations [2, 3]. Owing to the complexity of nucleon interactions and the fact that the level density increases exponentially with an increase in the excitation energy, the accurate calculation of the NLD has long been a theoretical challenge.
Many methods for estimating NLD have been developed. The most common method is the Bethe formula based on the zero-order approximation of the partition function of the Fermi gas model [4, 5]. In attempting to reproduce the experimental data, various phenomenological modifications to Bethe’s original analytical formulation have been proposed—particularly to account for shell, pairing, and deformation effects—which led first to the constant-temperature formulation, then to the shifted Fermi gas model, and later to the popular back-shifted Fermi gas model [6, 2, 7].
There are many microscopic methods for calculating NLD, including the shell-model Monte Carlo method [8-10], the moments method derived from random matrix theory and statistical spectroscopy [11, 12], the stochastic estimation method [13], the Lanczos method using realistic nuclear Hamiltonians [14], the self-consistent mean-field approach based on the extended Thomas–Fermi approximation with Skyrme forces [15], and the exact pairing plus independent particle model at a finite temperature [16-19]. On the basis of the Hartree–Fock–Bogoliubov (HFB) model, S. Hilaire and S. Goriely developed a microscopic approach to describe NLD with great success [20-22]. Microscopic methods based on the self-consistent Hartree–Fock (HF) plus BCS model [23-25] have also been developed.
Recently, on the basis of the relativistic Hartree–Bogoliubov model [26-29], J. Zhao et al. developed a method for describing NLD [30]. In this model, the partition function is determined using the same two-body interaction employed in the HF plus BCS and HFB mean-field models [25], which includes shell, pairing, and deformation effects self-consistently. The total level densities are the convolution of the intrinsic level density and the collective level density. The intrinsic level density is obtained by an inverse Laplace transform of the partition function with the saddle-point approximation [31, 32]. Previously, the collective enhancement is considered via a phenomenological or semi-empirical multiplicative factor for rotational and vibrational degrees of freedom [24, 33-36] or more microscopically via a combinatorial method using single-particle level schemes obtained through HF plus BCS or HFB calculations [20, 22]. In Ref. [30], the collective enhancement is determined from the eigenstates of a corresponding collective Hamiltonian that considers quadrupole or octupole degrees of freedom. Both the intrinsic level density and the collective enhancement are determined by the same global energy density functional and pairing interaction.
The success of the microscopic description of NLD in even–even nuclei prompts us to extend it to odd-A nuclei. The core–quasiparticle coupling (CQC) model introduces a quasiparticle to the neighboring even–even core nuclei within the same covariant density functionals and achieved progress for describing the quantum phase transition in odd-A nuclei. It is based on the covariant density functional theory (CDFT), which has achieved considerable success in describing ground-state properties of both spherical and deformed nuclei all over the nuclear chart [26, 27, 37-40]. Its successful applications include superheavy nuclei [41-45], pseudospin symmetry [40, 46, 47], single-particle resonances [48-50], hypernuclei [51-56], thermal shape transition [57-59], and shell correction [60-64].
In this study, the CQC model is applied to the calculation of collective levels in even–odd uranium isotopes. For even–even isotopes, the collective levels are attained via the five-dimensional collective Hamiltonian (5DCH) model [65]. Similar to Ref. [30], the intrinsic level density is obtained using the finite-temperature CDFT [66, 59, 58]. Because the level density of the saddle point plays an important role in the compound nuclei reactions [67, 68], the level density of the saddle point
The remainder of this paper is organized as follows. The theoretical framework is introduced in Sect. 2. The results for
Theoretical framework
With the assumption of decoupling between intrinsic and collective degrees of freedom, the excitation energy of a nucleus can be written as
The intrinsic level density
According to the ideas presented in Ref. [32], the determinant
Entropy is extracted in the finite-temperature covariant density functional theory. In the covariant density functional theory, the Dirac equation for single nucleons is
respectively. The isoscalar density
For odd-
In this study, the 5DCH and CQC models are based on the CDFT calculation with the harmonic oscillator basis
Results and Discussion
The parameter sets of covariant density functional theory used in this study are PC-PK1 [70] and DD-LZ1 [71]. PC-PK1 is one of the most widely used point-coupling parameter sets, and DD-LZ1 is a density-dependent parameter set that aims to alleviate the spurious shell closure. The pairing effect is considered by the separable pairing force [72]. The nuclei considered are even–even
In the first step, large-scale finite-temperature CDFT calculations are performed for
-202308/1001-8042-34-08-010/alternativeImage/1001-8042-34-08-010-F001.jpg)
It is shown that the intrinsic level density has an exponential relationship with the entropy in Eq. (3). To study the dependence of the entropy on the nuclear deformations, Fig. 2 shows the entropy surfaces at
-202308/1001-8042-34-08-010/alternativeImage/1001-8042-34-08-010-F002.jpg)
In the nuclear reactions, the level density of the saddle point is critical. Fig. 3 shows several properties of the saddle points with respect to the temperature for
-202308/1001-8042-34-08-010/alternativeImage/1001-8042-34-08-010-F003.jpg)
For comparison, the corresponding properties of the global minimum of
-202308/1001-8042-34-08-010/alternativeImage/1001-8042-34-08-010-F004.jpg)
In addition, the result for the parameter set DD-LZ1 is obtained. The specific heat
In the second step, the collective level densities of odd-
Density functional | Nucleus | Parity | Parity | ||||
---|---|---|---|---|---|---|---|
235U | + | -5.7 | 4.0 | - | -6.9 | 11.5 | |
PC-PK1 | 237U | + | -7.1 | 4.0 | - | -6.7 | 11.5 |
239U | + | -6.1 | 4.0 | - | -5.2 | 8.5 | |
235U | + | -7.9 | 4.0 | - | -7.2 | 11.0 | |
DD-LZ1 | 237U | + | -5.0 | 7.0 | - | -6.9 | 11.0 |
239U | + | -6.0 | 4.0 | - | -8.1 | 8.0 |
Take 235U as an example. Its low excitation spectra obtained using PC-PK1 and DD-LZ1 together with the experimental data from the NNDC [73] are shown in Fig. 5. The coupling strength and Fermi surface are finely tuned to reflect the collective enhancement induced by negative parity according to the corresponding experimental data of the low excitation spectrum. The calculated levels exhibit good qualitative agreement with the experimental results.
-202308/1001-8042-34-08-010/alternativeImage/1001-8042-34-08-010-F005.jpg)
In the final step, when performing the convolution using the Eq. (1) to obtain the total level density, it is found that the total level density
In Fig. 6, the total level density and the intrinsic level density at the global minima and saddle points are compared with available experimental data below 5 MeV [75], available calculations in the TALYS-1.95 package [1] directory density/ground/hilaire or density/fission/hilaire/Max1 and directory density/ground/goriely or density/fission/goriely/inner, or the Reference Input Parameter Library (RIPL) [20, 22]. Clearly, the intrinsic level densities deviate from the experimental data, indicating the necessity of collective degrees of freedom. For excitation energies up to 10 MeV, the total level density calculated via the two parameter sets PC-PK1 and DD-LZ1 agrees well with the two sets of theoretical calculations performed by S. Goriely and S. Hilaire in TALYS. However, for energies above 10 MeV, the total level density obtained via the relativistic density functional agrees well with Hilaire’s calculation, whereas Goriely’s calculation exhibits a sharp increase. The ratio of the total level density to Hilaire’s calculation result
-202308/1001-8042-34-08-010/alternativeImage/1001-8042-34-08-010-F006.jpg)
Furthermore, we compare the total level densities at the saddle points
-202308/1001-8042-34-08-010/alternativeImage/1001-8042-34-08-010-F007.jpg)
Finally, it is convenient to study the ratio of the total level density to the intrinsic level density
-202308/1001-8042-34-08-010/alternativeImage/1001-8042-34-08-010-F008.jpg)
Summary
In this study, according to the finite-temperature covariant density functional theory, the total level densities were obtained by convolving the intrinsic level densities and the collective levels achieved using the parameter sets PC-PK1 and DD-LZ1. For saddle points on the free energy surface in the
A 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)00162-5.Collective effects in deuteron induced reactions of aluminum
. Nucl. Instrum. Methods Phys. Res. Sect. B 391, 73-77 (2017). doi: 10.1016/j.nimb.2016.11.006.Nuclear Physics B. Nuclear dynamics, theoretical
. Rev. Mod. Phys. 9, 69-244 (1937). doi: 10.1103/RevModPhys.9.69.Estimations of level density parameters by using artificial neural network for phenomenological level density models
. Appl. Radiat. Isot. 169, 109583 (2021). doi: 10.1016/j.apradiso.2020.109583.A composite nuclear-level density formula with shell corrections
. Can. J. Phys. 43, 1446-1496 (1965). doi: 10.1139/p65-139.Global and local level density models
. Nucl. Phys. A 810, 13-76 (2008). doi: 10.1016/j.nuclphysa.2008.06.005.Particle-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.83.4265.Spin projection in the shell model Monte Carlo method and the spin distribution of nuclear level densities
. Phys. Rev. Lett. 99, 162504 (2007). doi: 10.1103/PhysRevLett.99.162504.Direct microscopic calculation of nuclear level densities in the shell model Monte Carlo approach
. Phys. Rev. C 92, 024307 (2015). doi: 10.1103/PhysRevC.92.024307.High-performance algorithm to calculate spin- and parity-dependent nuclear level densities
. Phys. Rev. C 82, 024304 (2010). doi: 10.1103/PhysRevC.82.024304.Constant temperature model for nuclear level density
. Phys. Lett. B 783, 428-433 (2018). doi: 10.1016/j.physletb.2018.07.023.Stochastic 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.2015.12.005.Microscopic calculations of nuclear level densities with the lanczos method
. Phys. Rev. C 102, 014315 (2020). doi: 10.1103/PhysRevC.102.014315.Self-consistent mean-field approach to the statistical level density in spherical nuclei
. Phys. Rev. C 97, 064302 (2018). doi: 10.1103/PhysRevC.97.064302.Simultaneous microscopic description of nuclear level density and radiative strength function
. Phys. Rev. Lett. 118, 022502 (2017). doi: 10.1103/PhysRevLett.118.022502.Testing the constant-temperature approach for the nuclear level density
, Phys. Rev. C 96, 054321 (2017). doi: 10.1103/PhysRevC.96.054321.Level density and thermodynamics in the hot rotating 96Tc nucleus
. Phys. Rev. C 96, 054326 (2017). doi: 10.1103/PhysRevC.96.054326.S-shaped heat capacity in an odd–odd deformed nucleus
. Phys. Lett. B 789, 634-638 (2019). doi: 10.1016/j.physletb.2018.12.007.Combinatorial nuclear level densities based on the Gogny nucleon-nucleon effective interaction
. Eur. Phys. J. A bf 12, 196-184 (2001). doi: 10.1007/s100500170025.Global 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.2006.08.014.Improved microscopic nuclear level densities within the Hartree-Fock-Bogoliubov plus combinatorial method
. Phys. Rev. C 78, 064307 (2008). doi: 10.1103/PhysRevC.78.064307.Nuclear level density with realistic interactions
. Phys. Rev. C 16, 757-766 (1977). doi: 10.1103/PhysRevC.16.757.Microscopic nuclear level densities for practical applications
. Nucl. Phys. A 695, 95-108 (2001). doi: 10.1016/S0375-9474(01)01095-8.Nuclear level densities with microscopic statistical method using a consistent residual interaction
. J. Nucl. Sci. Technol. 48, 984-992 (2011). doi: 10.1080/18811248.2011.9711785.Relativistic Hartree-Bogoliubov theory: static and dynamic aspects of exotic nuclear structure
. Phys. Rep. 409,101-259 (2005). doi: 10.1016/j.physrep.2004.10.001.Relativistic continuum Hartree Bogoliubov theory for ground-state properties of exotic nuclei
. Prog. Part. Nucl. Phys. 57, 470-563 (2006). doi: 10.1016/j.ppnp.2005.06.001.Multidimensionally constrained covariant density functional theories—nuclear shapes and potential energy surfaces
. Phys. Scr. 91, 063008 (2016). doi: 10.1088/0031-8949/91/6/063008.Microscopic model for the collective enhancement of nuclear level densities
. Phys. Rev. C 102, 054606 (2020). doi: 10.1103/PhysRevC.102.054606.Projectile-fragment yields as a probe for the collective enhancement in the nuclear level density
. Nucl. Phys. A 629(3), 635-655 (1998). doi: 10.1016/S0375-9474(98)00658-7.Pairing correlations and thermodynamical quantities in 96,97Mo
. Phys. Rev. C 75, 064319 (2007). doi: 10.1103/PhysRevC.75.064319.Level density rotational enhancement factor
. Phys. Rev. C 99, 064331 (2019). doi: 10.1103/PhysRevC.99.064331.Collective enhancements in the level densities of Dy and Mo isotopes
. Phys. Rev. C 101, 054315 (2020). doi: 10.1103/PhysRevC.101.054315.Relativistic mean field theory in finite nuclei
. Prog. Part. Nucl. Phys. 37, 193-263 (1996). doi: 10.1016/0146-6410(96)00054-3.Spin symmetry in the antinucleon spectrum
. Phys. Rev. Lett. 91, 262501 (2003). doi: 10.1103/PhysRevLett.91.262501.Possible existence of multiple chiral doublets in 106Rh
. Phys. Rev. C 73, 037303 (2006). doi: 10.1103/PhysRevC.73.037303.Hidden pseudospin and spin symmetries and their origins in atomic nuclei
. Phys. Rep. 570, 1-84 (2015). doi: 10.1016/j.physrep.2014.12.005.Magic numbers for superheavy nuclei in relativistic continuum Hartree–Bogoliubov theory
. Nucl. Phys. A 753, 106-135 (2005). doi: 10.1016/j.nuclphysa.2005.02.086.Description of structure and properties of superheavy nuclei
. Prog. Part. Nucl. Phys. 58, 292-349 (2007). doi: 10.1016/j.ppnp.2006.05.001.Theoretical study of the synthesis of superheavy nuclei with z=119 and 120 in heavy-ion reactions with trans-uranium targets
. Phys. Rev. C 85, 041601 (2012). doi: 10.1103/PhysRevC.85.041601.Description of α-decay chains for 293,294117 within covariant density functional theory
. Phys. Rev. C 88, 054324 (2013). doi: 10.1103/PhysRevC.88.054324.Multidimensionally-constrained relativistic mean-field models and potential-energy surfaces of actinide nuclei
, Phys. Rev. C 89, 014323 (2014). doi: 10.1103/PhysRevC.89.014323.Spin and pseudospin symmetries in the single-Λ spectrum
, Phys. Rev. C 96, 044312 (2017). doi: 10.1103/PhysRevC.96.044312.Green’s function method for the spin and pseudospin symmetries in the single-particle resonant states
. Phys. Rev. C 99, 034310 (2019). doi: 10.1103/PhysRevC.99.034310.Green’s function method for the single-particle resonances in a deformed dirac equation
. Phys. Rev. C 101, 014321 (2020). doi: 10.1103/PhysRevC.101.014321.Properties of titanium isotopes in complex momentum representation within relativistic mean-field theory
. Nucl. Sci. Tech. 33, 117 (2022). doi: 10.1007/s41365-022-01098-8.Research on the infuence of quadrupole deformation and continuum efects on the exotic properties of 15,17,19B with the complex momentum representation method
. Nucl. Sci. Tech. 34, 25 (2023). doi: 10.1007/s41365-023-01177-4.Quadrupole deformation (β,γ) of light Λ hypernuclei in a constrained relativistic mean field model: Shape evolution and shape polarization effect of the Λ hyperon
. Phys. Rev. C 84, 014328 (2011). doi: 10.1103/PhysRevC.84.014328.Superdeformed Λ hypernuclei within relativistic mean field models
. Phys. Rev. C 89, 044307 (2014). doi: 10.1103/PhysRevC.89.044307.Mean-field approaches for Ξ− hypernuclei and current experimental data
, Phys. Rev. C 94, 064319 (2016). doi: 10.1103/PhysRevC.94.064319.Green’s function relativistic mean field theory for Λ hypernuclei
. Phys. Rev. C 95, 054318 (2017). doi: 10.1103/PhysRevC.95.054318.Relativistic mean-field approach for Λ,Ξ, and Σ hypernuclei
. Phys. Rev. C 98, 024316 (2018). doi: 10.1103/PhysRevC.98.024316.Possible shape coexistence in ne isotopes and the impurity effect of Λ hyperon
. Sci. China-Phys. Mech. Astron. 64, 282011 (2021). doi: 10.1007/s11433-021-1721-1.Shape evolution of 72,74Kr with temperature in covariant density functional theory*
. Chin. Phys. C 41, 094102 (2017). doi: 10.1088/1674-1137/41/9/094102.Shape transition with temperature of the pear-shaped nuclei in covariant density functional theory
. Phys. Rev. C 96, 054308 (2017). doi: 10.1103/PhysRevC.96.054308.Critical temperature for shape transition in hot nuclei within covariant density functional theory
. Phys. Rev. C 97, 054302 (2018). doi: 10.1103/PhysRevC.97.054302.Shell correction at the saddle point for superheavy nucleus
. Chin. Phys. Lett. 20, 1694-1697 (2003). http://cpl.iphy.ac.cn/EN/abstract/article_33959.shtml.Stability of strutinsky shell correction energy in relativistic mean field theory
. Chin. Phys. Lett. 26, 032103 (2009). doi: 10.1088/0256-307X/26/3/032103.Strutinsky shell correction energies in relativistic Hartree-Fock theory
. Phys. Rev. C 98, 064323 (2018). doi: 10.1103/PhysRevC.98.064323.Shell corrections with finite temperature covariant density functional theory
. Chin. Phys. C 45, 024107 (2021). doi: 10.1088/1674-1137/abce12.A global weizsäcker mass model with relativistic mean field shell correction
. Chin. Phys. C 46, 104105 (2022). doi: 10.1088/1674-1137/ac7b18.Beyond the relativistic mean-field approximation. III. collective hamiltonian in five dimensions,
Phys. Rev. C 79, 034303 (2009). doi: 10.1103/PhysRevC.79.034303.Pairing transitions in finite-temperature relativistic Hartree-Bogoliubov theory
. Phys. Rev. C 88, 034308 (2013). doi: 10.1103/PhysRevC.88.034308.Phenomenological statistical analysis of level densities, decay widths and lifetimes of excited nuclei
. Nucl. Phys. A 543, 517-557 (1992). doi: 10.1016/0375-9474(92)90278-R.Microscopic core-quasiparticle coupling model for spectroscopy of odd-mass nuclei
. Phys. Rev. C 96, 054309 (2017). doi: 10.1103/PhysRevC.96.054309.New parametrization for the nuclear covariant energy density functional with a point-coupling interaction
. Phys. Rev. C 82, 054319 (2010). doi: 10.1103/PhysRevC.82.054319.Novel relativistic mean field lagrangian guided by pseudo-spin symmetry restoration
. Chin. Phys. C 44, 074107 (2020). doi: 10.1088/1674-1137/44/7/074107.A finite range pairing force for density functional theory in superfluid nuclei
. Phys. Lett. B 676, 44-50 (2009). doi: 10.1016/j.physletb.2009.04.067.Collective enhancement of nuclear level density in the interacting boson model
. Phys. Rev. C 42, 988-992 (1990). doi: 10.1103/PhysRevC.42.988.Constant-temperature level densities in the quasicontinuum of Th and U isotopes
. Phys. Rev. C 88, 024307 (2013). doi: 10.1103/PhysRevC.88.024307.Energy dependent ratios of level-density parameters in superheavy nuclei
. Phys. Rev. C 105, 044328 (2022). doi: 10.1103/PhysRevC.105.044328.The authors declare that they have no competing interests.