This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Odd-even mass differences of well and rigidly deformed nuclei in the rare earth region: A test of a newly proposed fit of average pairing matrix elements

T. V. Nhan Hao [email protected] Faculty of Physics, University of Education, Hue University, 34 Le Loi Street, Hue City, Vietnam Center for Theoretical and Computational Physics, University of Education, Hue University, 34 Le Loi Street, Hue City, Vietnam    N. N. Bao Nguyen Faculty of Physics, University of Education, Hue University, 34 Le Loi Street, Hue City, Vietnam Center for Theoretical and Computational Physics, University of Education, Hue University, 34 Le Loi Street, Hue City, Vietnam    D. Quang Tam Faculty of Physics, University of Education, Hue University, 34 Le Loi Street, Hue City, Vietnam Faculty of Basic Sciences, University of Medicine and Pharmacy, Hue University, 06 Ngo Quyen Street, Hue City, Vietnam    P. Quentin [email protected] LP2I, UMR 5797, Université de Bordeaux, CNRS, F-33170, Gradignan, France    Meng-Hock Koh (辜明福) [email protected] Department of Physics, Faculty of Science, Universiti Teknologi Malaysia, 81310 Johor Bahru, Johor, Malaysia. UTM Centre for Industrial and Applied Mathematics, 81310 Johor Bahru, Johor, Malaysia    L. Bonneau LP2I, UMR 5797, Université de Bordeaux, CNRS, F-33170, Gradignan, France
(April 5, 2025)
Abstract

We discuss a test of a recently proposed approach to determine average pairing matrix elements within a given interval of single-particle states (sp) around the Fermi level λ\lambda as obtained in the so-called uniform gap method (UGM). It takes stock of the crucial role played by the averaged sp level density ρ~(e)\tilde{\rho}(e). These matrix elements are deduced within the UGM approach, from microscopically calculated ρ~(e)\tilde{\rho}(e) and gaps obtained from analytical formulae of a semi-classical nature. Two effects generally ignored in similar fits have been taken care of. They are: (a) the correction for a systematic bias in choosing to fit pairing gaps corresponding to equilibrium deformation solutions as discussed by Möller and Nix [Nucl. Phys. A 476, 1 (1992)] and (b) the correction for a systematic spurious enhancement of ρ~(e)\tilde{\rho}(e) for protons in the vicinity of λ\lambda, because of the local Slater approximation used for the treatment of the Coulomb exchange terms in most calculations (see e.g. [Phys. Rev C 84, 014310 (2011)]). This approach has been deemed to be very efficient upon performing Hartree-Fock + BCS (with seniority force and self-consistent blocking when dealing with odd nuclei) calculations of a large sample of well and rigidly deformed even-even rare-earth nuclei. The reproduction of their experimental moments of inertia has been found to be at least of the same quality as what has been obtained in a direct fit of these data [Phys. Rev C 99, 064306 (2019)]. We extend here the test of our approach to the reproduction, in the same region, of three-point odd-even mass differences centered on odd-NN or odd-ZZ nuclei. The agreement with the data is again roughly of the same quality as what has been obtained in a direct fit, as performed in [Phys. Rev C 99, 064306 (2019)].

I Introduction

A simple prescription to determine average pairing matrix elements from averaged single-particle (sp) level densities for the ground states of well and rigidly deformed nuclei has been proposed recently in [1] and deemed to be rather successful. These matrix elements VqV_{q} (where qq stands for the charge state i.e. neutron or proton) correspond to their average values around the Fermi energies λq\lambda_{q}. They are to be used in a microscopic approach of the Hartree-Fock-plus-BCS (HF + BCS) type. The pairing correlations are treated within the so-called seniority force approximation. It consists in solving the BCS variational equations in a restricted sp space around the Fermi energy, assuming the constancy of the pairing matrix element within this interval.

It takes stock of the strong dependence of the VqV_{q} values upon the sp level densities at the Fermi energies gq(λq)g_{q}(\lambda_{q}) averaged à la Strutinsky considered as a sound ersatz of a semi-classical approximation. It relies also on the consideration that any fitting of nuclear energies (here any energy related to pairing correlations) in terms of nucleon numbers correspond in an effective way to such a semi-classical approximation. Therefore the input of experimental data is naturally performed through some standard analytical formula expressing the nucleon number dependence of odd-even mass differences defined in different ways noted here as δE\delta E (see, e.g., Refs. [2, 3]) taking into account the important corrective contribution of [4].

After the pionneering work of Ref. [5], it has been a current practice to adjust the intensity of pairing correlations to reproduce the data on δE\delta E. To a lesser extent, for this purpose, one has also considered moments of inertia noted as 𝒥\mathcal{J}, deduced from the first 2+2^{+} excitation energies of well and rigidly deformed even-even heavy nuclei. In a recent paper [6], it has been demonstrated in the region of (or close to) rare-earth deformed nuclei (loosely dubbed as rare-earth nuclei in what follows) that making separate fits on δE\delta E and 𝒥\mathcal{J} yielded similarly good results for the parameters in use to calculate pairing correlations within HF + (seniority) BCS calculations. This was a confirmation that both pieces of data were highly contingent upon a good description of pairing properties and thus asserting their relevance for such fits.

In both cases (with δE\delta E or 𝒥\mathcal{J}), these fitting approaches possibly suffer from accidental local deficiencies of the sp distribution of levels in the vicinity of the Fermi energy. A priori, resorting to semi-classical quantities associated to a given microscopic theoretical approach one gets rid of such a practical problem. What remains of course to be assessed is the relevance of the analytical expressions for the (N,Z)(N,Z) dependence of the average δE\delta E differences as well as the quality of the averaged sp level density of the canonical basis states.

The method proposed in [1] has been validated by comparing its results for moments of inertia of well and rigidly deformed rare-earth nuclei with those obtained through a specific fit on the 𝒥\mathcal{J} data as performed in Ref. [6]. A similar quality in the reproduction of such spectroscopic data was obtained.

It is the aim of the present paper to make a similar comparison for the δE\delta E data now, between the results obtained within the method developed in [1] and the direct fit of these differences performed in the same nuclear region, in Ref. [6].

In Section II, we will briefly recall the approach developed in [1], whereas Section III will provide some details about the calculations performed here. Our results will be discussed in Section IV ; Section V will be devoted to some conclusions and perspectives.

II Brief survey of the approach

The method in use has been presented and discussed in detail in [1]. We will thus restrict here to sketchily describe its main characteristics.

Our starting point is a sp spectrum for a charge state qq, obtained a priori through any microscopic approach. In what follows it will result from self-consistent HF+BCS calculations within the seniority force simple ansatz for the pairing matrix elements.

We calculate an approximate semi-classically averaged sp level density ρ~(e)\tilde{\rho}(e) as a function of the sp energy ee through a standard Strutinsky’s energy averaging (see Refs. [7, 8]) using the equation:

ρ~q(e)=1γρ(e)f(eeγ)𝑑e.\tilde{\rho}_{q}(e)=\frac{1}{\gamma}\int_{-\infty}^{\infty}\rho(e^{\prime})f\Big{(}\frac{e^{\prime}-e}{\gamma}\Big{)}de^{\prime}. (1)

As discussed in [1] for the considered nuclei located far enough from the neutron drip line, the averaging width is taken as γ=1.2ω\gamma=1.2\>\hbar\omega where the energy scale as a function of the nucleon number AA is the usual expression ω=41A1/3\hbar\omega=41A^{-1/3} MeV [9]. The f(x)f(x) term corresponding to the so called curvature correction, defined as

f(x)=P(x)w(x)f(x)=P(x)\>w(x) (2)

is a product of the weight factor w(x)w(x)

w(x)=1πex2.w(x)=\frac{1}{\sqrt{\pi}}e^{-x^{2}}. (3)

and a polynomial P(x)P(x) taken here to be the generalized Laguerre polynomials LM(α)L_{M}^{(\alpha)} for the variable x2x^{2} of order M=2M=2 written as

P(x)=LM1/2(x2)=n=0Ma2nx2n.P(x)=L_{M}^{1/2}(x^{2})=\sum_{n=0}^{M}a_{2n}\>x^{2n}. (4)

with coefficients α2n\alpha_{2n} given e.g. in Table II of [1].

For NqN_{q} nucleons of the charge state qq, we then compute the Fermi energy λ~q\tilde{\lambda}_{q} as

Nq=λ~qρ~q(e)𝑑e.N_{q}=\int_{-\infty}^{\tilde{\lambda}_{q}}\tilde{\rho}_{q}(e)de. (5)

Within the so-called uniform gap method of Refs. [7] [10], given a suitably average gap Δ~q\tilde{\Delta}_{q} (see below), the average pairing matrix element VqV_{q} to be used in our HF+BCS approach is given by

1Vq=λ~qΩλ~q+Ωρ~q(e)(eλ~q)2+Δ~q2𝑑e.\frac{1}{V_{q}}=\int_{\tilde{\lambda}_{q}-\Omega}^{\tilde{\lambda}_{q}+\Omega}\frac{\tilde{\rho}_{q}(e)}{\sqrt{(e-\tilde{\lambda}_{q})^{2}+{\tilde{\Delta}}_{q}^{2}}}de. (6)

As well-known, the value of this matrix element depends on the choice of the energy interval 2Ω2\Omega centered at the Fermi energy and including the sp states active in the BCS variational determination of occupation probabilities. In this study as in [1] we take Ω=6\Omega=6 MeV.

The nucleon-number NqN_{q} dependence of average gaps Δ~q\tilde{\Delta}_{q} could be taken a priori from standard formulas (as, e.g., those of Refs. [2], [3]). However Möller and Nix [4] have pointed out that there is a bias in such estimates due to the mere selection of nuclei at equilibrium deformation, corresponding systematically to lower than average quantal sp level densities. As a result of their study, they proposed the following parametrisation of the average gaps

Δ~q=rBsNq1/3,\tilde{\Delta}_{q}=\frac{rB_{s}}{N_{q}^{1/3}}\,, (7)

where BsB_{s} is set to 11 and r=4.8r=4.8 MeV.

While we stick to this value for neutrons, we have remarked in [1] that this must not be the case for protons. In most of microscopic approaches of the Hartree-Fock plus BCS, Hartree-Fock-Bogoliubov (HFB) or relativistic mean-field type, one simplifies the treatment of exchange terms of the infinite range Coulomb interaction by having recourse to the local Slater approximation [11]. It has been found long ago [12] and further studied in Refs. [13, 14] that the Slater approximation systematically and significantly overestimates the quantal sp level density near the Fermi energy (especially at the equilibrium deformation).

To ensure a safe use of the Slater approximation in our approach, we must thus quench the parameter rr of Eq (7) by a factor RpR_{p} which is the ratio of the pairing gaps obtained in two separate quantal BCS calculations (exact and approximated à la Slater). It has been shown in Ref. [1] that the dependence of this ratio with respect to the intensity of pairing correlations may be approximatively given by

Rp=0.0181Econdp+0.781,R_{p}=0.0181\>E_{cond}^{p}+0.781\,, (8)

where EcondpE_{cond}^{p} is the average pairing condensation energy (defined as the part of the total energy involving explicitely the abnormal BCS density) given in MeV.

It is estimated using the quantal gap obtained within HF+BCS calculations with some starting ansatz for the proton matrix element VpV_{p} such that (see Appendix A of [1]):

Econdp=Δp2VpE_{cond}^{p}=\frac{\Delta_{p}^{2}}{V_{p}} (9)

where Δp\Delta_{p} is the BCS proton pairing gap (in MeV).

In principle this should imply an iterative process to define new Möller-Nix average gaps and accordingly a new VqV_{q} value. However, it has been shown in [1] that choosing some starting VpV_{p} which is approximated by a constant matrix element with an initial pairing strength Gq=19G_{q}=19 MeV (see Sec. III.1) range of values (as it turns out easily delineated), these iterations do not bring significant modifications of the resulting matrix elements given the rough character of the above corrective process. As such, we limit ourselves herein to a non-iterative process of estimating the pairing matrix elements.

III Some details of the calculations

III.1 The canonical basis

In our approach, the sp canonical basis states are obtained within self-consistent HF+BCS calculations. Let us recall that throughout this paper the BCS correlations are determined within the seniority force simple approximation. As an alternative presentation for the values of the the average matrix elements VqV_{q} we provide GqG_{q} parameters introduced in Ref. [15] to the effect of removing approximately the dependence of these matrix elements from NqN_{q} values

Vq=Gq11+Nq.V_{q}=\frac{G_{q}}{11+N_{q}}. (10)

The particle-hole interaction in use is of the standard Skyrme type. Calculations are performed with the SIII parametrisation [16] which has been shown in many instances to provide a rather good description of the spectroscopic properties of well and rigidly deformed nuclei (see, e.g., for a recent account Ref. [17]). The averaged pairing matrix elements VqV_{q} are given for each nucleus according to the method summarized in Section II for even-even nuclei. For odd-ZZ (odd-NN respectively) nuclei the retained values of the matrix elements are interpolated between the values found for the neighbouring isotones (isotopes respectively).

The axial and intrinsic parity symmetries have been imposed to the solutions. The nuclei are thus specified by their projection KK of their angular momentum on the symmetry axis and their parity π\pi. To solve the Schrödinger equation determining the canonical basis states, we have expanded these states on eigenstates of an axially symmetrical harmonic oscillator Hamiltonian. It is defined by a basis size parameter N0=16N_{0}=16 defining a deformation-dependent selection of relevant basis states corresponding at sphericity to the consideration of 17 shells. This basis is truncated according to (see Ref. [18])

ω(n+1)+ωz(nz+12)ω0(N0+2).\hbar\omega_{\bot}(n_{\bot}+1)+\hbar\omega_{z}(n_{z}+\frac{1}{2})\leq\hbar\omega_{0}(N_{0}+2). (11)

where nzn_{z} and nn_{\bot} are the number of oscillator quanta in the symmetry-axis direction (zz-axis) and in the perpendicular direction and N0+1N_{0}+1 corresponds to the number of spherical shells.

The inverse-length bb and deformation qq parameters are defined, mm being the average nucleonic mass, such that [18]

b=mω0;q=ωωz.b=\sqrt{\frac{m\omega_{0}}{\hbar}}\;\;\;;\;\;\;q=\frac{\omega_{\bot}}{\omega_{z}}. (12)

These two parameters are related to the angular frequencies on the (x,y)(x,y) plane, ω\omega_{\bot}, and zz-axis, ωz\omega_{z}, while ω03=ω2ωz\omega_{0}^{3}=\omega_{\bot}^{2}\omega_{z} is the angular frequency at sphericity. The two parameters bb and qq have been optimized for even-even nuclei to give the lowest energy for each equilibrium solution, whereas for odd-AA nuclei they have been interpolated from those obtained for even-even nuclei. Integrals involving the local densities are performed using the Gauss-Hermite and Gauss-Laguerre approximate integration methods with 50 points along the symmetry axis and 16 points in the perpendicular direction, respectively.

III.2 Odd-even mass differences

The data on odd-even mass differences are extracted through a 3-point formula. As discussed in Refs. [19, 20] such differences δq(3)\delta_{q}^{(3)} centered on an odd-neutron or odd-proton nucleus are good markers on the degree of pairing correlations. They are, to a large extent, free from single-particle filling effects and are given, e.g., for an isotopic series by

δn3(N)\displaystyle\delta^{3}_{n}(N) =(1)N2[E(N+1,Z)2E(N,Z)+E(N1,Z)]\displaystyle=\frac{(-1)^{N}}{2}\Bigg{[}E(N+1,Z)-2E(N,Z)+E(N-1,Z)\Bigg{]}
=(1)N2[Sn(N,Z)Sn(N+1,Z)]\displaystyle=\frac{(-1)^{N}}{2}\Bigg{[}S_{n}(N,Z)-S_{n}(N+1,Z)\Bigg{]} (13)

where NN is odd and Sn(N,Z)S_{n}(N,Z) is the neutron separation energy of a nucleus composed of NN neutrons and ZZ protons whose total energies are denoted as E(N,Z)E(N,Z). Similar expressions are easily deduced from the above for odd-proton nuclei.

These energies are compared with those extracted directly from calculated binding energies within the HF+BCS approach. In the case of odd-AA nuclei, we have performed selfconsistent blocking calculations (i.e. placing a nucleon in the relevant sp orbit specified by the KπK^{\pi} quantum numbers). The breaking of the time-reversal symmetry implies the presence of new (time-odd) local densities in the expression of the Hamiltonian density resulting in new terms in the corresponding Hartree-Fock potential. As explicitely detailed in Ref. [21], when using the SIII interaction we have considered a restricted choice of time-odd potential fields, yet preserving the Galilean invariance (namely the vector spin field 𝕊(𝕣)\mathbb{S}(\mathbb{r}) and the vector current field 𝔸(𝕣)\mathbb{A}(\mathbb{r}), with usual notation). This choice is dubbed as the minimal scheme in Ref. [21].

In the spirit of the Bohr-Mottelson unified model, well suited for these deformed nuclei, we assimilate the nuclear angular momentum and parity quantum numbers IπI^{\pi} to those KπK^{\pi} of the blocked blocked-nucleon sp state. This is of course not free for any perturbation of the low-energy nuclear spectra from possible Coriolis coupling which will be ignored here. Particular cases concern solutions where K=1/2K=1/2 with a decoupling parameter aa outside the range 1a4-1\leq a\leq 4. They will be specifically discussed in Section IV and shown to be unable to perturb the natural rotational band ordering of states. The decoupling parameter aa is defined, as well known [22], through the relation

a=i|J^+|i~a=-\langle i|\widehat{J}_{+}|\widetilde{i}\rangle (14)

where J^+\widehat{J}_{+} is the usual angular momentum ladder operator (sum of orbital and spin angular momenta) in \hbar unit and |i~|\widetilde{i}\rangle is the single-particle state canonically conjugate of the blocked state |i|i\rangle. These two states are such that J^z|i=Ωi|i\widehat{J}_{z}|i\rangle=\Omega_{i}|i\rangle and J^z|i~=Ωi|i~\widehat{J}_{z}|\widetilde{i}\rangle=-\Omega_{i}|\widetilde{i}\rangle. In our HFBCS code with selfconsistent blocking, the time-reversal symmetry is broken in the one-body sector and the definition of the two sp states forming the equivalent of Cooper pairs in our BCS wavefunction is provided in Appendix A of [23].

We have systematically calculated solutions for odd-AA nuclei corresponding to the ground state IπI^{\pi} experimental values [24] as well as those where the calculated energies lie below the former or in some cases (see the discussions of Section IV) above generally up to a couple of 100 keV.

III.3 Choice of sample nuclei

We have chosen odd-AA nuclei belonging to the rare-earth region plus Hafnium isotopes (loosely called below rare-earth nuclei), i.e. some odd-proton isotopes from Europium to Lutetium and odd-neutron isotopes from Samarium to Hafnium. They are chosen to be well and rigidly deformed and, moreover, far enough from a region of transition between deformed and soft nuclei. The first criterion is retained to be able to reduce approximately the collective dynamics to a pure rigid rotation within the Bohr-Mottelson unified model, allowing in particular to limit ourselves to a single BCS state, i.e. ignoring quantal shape fluctuations. The second criterion is retained to avoid wide shape variations (and therefore large sp spectrum changes) between the three isotopes (isotones respectively) entering the calculation of the energy differences δn(3)\delta_{n}^{(3)} (δp(3)\delta_{p}^{(3)} respectively).

Table 1 displays the ratio R42R_{42} of the excitation energies of the first 4+4^{+} and 2+2^{+} states [24] of the 22 even-even nuclei bracketing the odd-AA nuclei whose energy differences δq(3)\delta_{q}^{(3)} are evaluated in our calculations. It appears that they satisfy reasonably well the first criterion since for all of them one has R423.29R_{42}\geq 3.29.

On this table also, the intrinsic axial charge quadrupole moments Q20intQ^{\rm int}_{20} obtained in our calculations are compared, when available, with the corresponding experimental values deduced either from reduced B(E2) data [25] or those deduced from the spectroscopic moments Q20spectQ^{\rm spect}_{20} of the first 2+2^{+} state [26], upon using the unified model relation for I=2,K=0I=2,K=0, namely Q20int=3.5Q^{\rm int}_{20}=-3.5 Q20spectQ^{\rm spect}_{20}. One notices a rather good reproduction of the Q20intQ^{\rm int}_{20} data with the interaction SIII in use, as shown long ago [27] (somewhat less good however for the 156,158Sm isotopes).

Table 1: Some static properties of the 22 even-even nuclei considered in this paper. The calculated total binding energies EtheoryE^{\rm theory} are given in MeV. The experimental intrinsic axial quadrupole moments for the charge distribution are deduced (within the unified model relations for rotational band states) from Refs. [26] ([25] respectively) for spectroscopic first 2+2^{+} states data noted as Q20spectQ_{20}^{\rm spect} moments (for moments deduced from reduced B(E2)B(E2) data noted as Q20BE2Q_{20}^{\rm BE2} respectively). They are compared with the corresponding calculated moments Q20theoryQ_{20}^{\rm theory}. All these moments are given in barn.
Nucleus Etheory{}^{\text{theory}} R42 Q20spect{}^{\text{spect}}_{20} Q20BE2{}^{\text{BE2}}_{20} Q20theory{}^{\text{theory}}_{20}
156Sm -1276.496 3.290 5.85 (7) - 6.81
158Sm -1288.582 3.301 6.55 (14) - 6.99
160Sm -1299.953 3.292 - - 7.11
160Gd -1305.533 3.302 7.28 (14) 7.265 (42) 7.25
162Gd -1318.251 3.302 - - 7.40
164Gd -1330.079 3 295 - - 7.51
166Gd -1341.067 3.300 - - 7.58
162Dy -1319.955 3.294 - 7.33 (8) 7.38
164Dy -1333.969 3.301 7.28 (53) 7.503 (33) 7.56
166Dy -1347.128 3.310 - - 7.69
168Dy -1359.411 3.313 - - 7.76
168Er -1361.733 3.309 - 7.63 (7) 7.84
170Er -1375.231 3.310 6.65 (70) 7.65 (7) 7.93
172Er -1387.702 3.314 - - 7.72
170Yb -1374.077 3.293 7.63 (11) 7.63 (9) 7.90
172Yb -1388.724 3.305 7.77 (14) 7.792 (45) 7.98
174Yb -1402.460 3.310 7.63 (18) 7.727 (39) 7.77
176Yb -1415.595 3.310 7.98 (21) 7.30 (13) 7.58
178Yb -1427.957 3.310 - - 7.46
178Hf -1429.290 3.291 7.07 (7) 6.961 (43) 7.22
180Hf -1442.851 3.307 7.00 (7) 6.85 (9) 7.08
182Hf -1455.193 3.295 - - 6.85

IV Results

While this paper is concerned with a test of calculated odd-even mass differences δq(3)\delta_{q}^{(3)}, we will first discuss the nature and relevance of the configurations retained in our comparison with experimental data. By configuration we mean the nuclear spin (assumed as we have seen, to be equal to the projection of the angular momentum on the quantification axis) and parity (well defined in our solution since we impose an intrinsic reflection symmetry). Obviously the choice of the configuration determining the location in the sp spectrum of the unpaired nucleon has a direct effect on the relevant separation energies. As a rule, however, to avoid overestimating unduly the predictive value of our approach, we will retain in our comparison with the data, theoretical solutions possessing the experimental ground state spin and parity values even though they do not correspond to the lowest calculated total energy. It was nevertheless interesting to see how well the spin and parity of the calculated ground states match with the data. In the same way, since the ordering of sp states around the Fermi depends significantly on the deformation of the mean field, we checked the agreement of our calculated axial moments with the intrinsic moments extracted from two pieces of experimental data (reduced E2 transition and spectroscopic moment data).

IV.1 Discussion of the ground-state configurations obtained in our calculations for odd-Z nuclei

We compare here the lowest-energy configurations as obtained in our calculations with the experimental values of the angular momentum and parity quantum numbers IπI^{\pi} given in the current version of the NUDAT compilation [24].

As seen on Table 2, for 8 nuclei out of 13 (161Tb, 163Tb, 167Ho, 169Ho, 169Tm, 171Tm, 177Lu, 179Lu), our theoretical assignments agree with the data.

We confirm the suggested assignments for 3 nuclei, namely: 165Tb and 167Tb as 3/2+3/2^{+} and 173Tm as 1/2+1/2^{+}. No assignment is proposed in Ref. [24] for the 161Eu nucleus. We suggest a 5/25/2^{-} configuration noting however that we have obtained a 5/2+5/2^{+} solution 179179 keV above the 5/25/2^{-} configuration. In one case (159Eu) the experimental lowest configuration 5/2+5/2^{+} is obtained at 145 keV above a 5/25/2^{-} state. Finally, we remark that the decoupling constant values corresponding to the 1/2+1/2^{+} state considered in the three calculated isotopes of Thulium belong to the [0.64,0.60][-0.64,-0.60] interval ensuring that the band head spin assignment as 1/21/2 is correct.

Table 2: Comparison of some spectroscopic properties of the 13 odd-ZZ nuclei considered in this paper. Along with the calculated total binding energies EtheoryE^{\rm theory} given in MeV, the intrinsic configuration spins and parities KπK^{\pi} of our solutions are reported. The corresponding experimental ground state values IπI^{\pi} (proposed or suggested - in brackets - when available) for a given nucleus are displayed for the sake of comparison, assuming the validity of the unified model assumption of I=KI=K for the band head states. Both calculated δcalc(3)\delta_{\rm calc}^{(3)} and experimental δexp(3)\delta_{\rm exp}^{(3)} odd-even mass differences (given in keV) are also reported.
Nucleus Iπ(exp)\text{I}^{\pi}(\text{exp}) Iπ(calc)\text{I}^{\pi}(\text{calc}) Etheory\text{E}^{\text{theory}} δcalc(3)\delta^{(3)}_{\text{calc}} δexp(3)\delta^{(3)}_{\text{exp}}
159Eu 5/2+ 5/2+ -1296.280 777 554
5/2- -1296.425 633 -
161Eu - 5/2- -1308.524 579 -
5/2+ -1308.345 757 466
161Tb 3/2+ 3/2+ -1312.161 583 600
163Tb 3/2+ 3/2+ -1325.597 513 529
165Tb (3/2+) 3/2+ -1338.149 455 550
167Tb (3/2+) 3/2+ -1349.801 438 579
167Ho 7/2- 7/2- -1353.783 648 508
169Ho 7/2- 7/2- -1366.734 588 535
169Tm 1/2+ 1/2+ -1367.174 732 602
171Tm 1/2+ 1/2+ -1381.392 585 471
173Tm (1/2+) 1/2+ -1394.585 496 458
177Lu 7/2+ 7/2+ -1422.045 397 579
179Lu 7/2+ 7/2+ -1435.055 348 669
Table 3: Same as Table 2 for the 16 odd-NN nuclei considered in this paper. Note that for the 165Dy and 171Yb nuclei, the experimental values of the spin and parity of a low-lying isomeric state have been also displayed.
Nucleus Iπ(exp)\text{I}^{\pi}(\text{exp}) Iπ(calc)\text{I}^{\pi}(\text{calc}) Etheory\text{E}^{\text{theory}} δcalc(3)\delta^{(3)}_{\text{calc}} δexp(3)\delta^{(3)}_{\text{exp}}
157Sm (3/2)(3/2^{-}) 3/2- -1281.764 775 629
159Sm 5/2- 5/2- -1293.550 718 535
161Gd 5/2- 5/2- -1311.129 763 605
163 Gd (5/2,7/2+5/2^{-},7/2^{+}) 7/2+ -1323.398 766 599
5/2- -1322.819 1346
1/2- -1323.457 707
165Gd - 7/2+ -1334.916 657 507
163Dy 5/2- 5/2- -1326.097 864 694
165Dy 7/2+,1/2m7/2^{+},1/2^{-}_{m} 7/2+ -1339.788 760 664
1/2- -1339.836 712 -
167Dy (1/2)(1/2^{-}) 1/2- -1352.563 707 661
7/2+ -1352.620 650 -
169Er 1/2- 1/2- -1367.745 737 627
7/2+ -1367.817 665 -
171Er 5/2- 5/2- -1380.826 640 577
171Yb 1/2,7/2m+1/2^{-},7/2^{+}_{m} 1/2- -1380.625 776 703
7/2+ -1380.698 703 -
173Yb 5/2- 5/2- -1394.879 713 549
175Yb 7/2- 7/2- -1408.299 728 522
177Yb 9/2+ 9/2+ -1421.235 541 598
179Hf 9/2+ 9/2+ -1435.547 524 644
181Hf 1/2- 1/2- -1448.366 655 512

IV.2 Discussion of the ground-state configurations obtained in our calculations for odd-NN nuclei

As seen on Table 3, for 9 nuclei out of 16 (159Sm, 161Gd, 163Dy, 171Er, 173Yb, 175Yb, 177Yb, 179Hf, 181Hf), our theoretical assignments agree with the data.

We confirm the suggested assignment for one nucleus, namely: 157Sm as 3/23/2^{-}. No assignment is proposed in Ref. [24] for the 165Gd nucleus. We suggest a 7/2+7/2^{+} configuration. In 167Dy the suggested ground state configuration 1/21/2^{-} suggested in Ref. [24] is calculated only 58 keV above a 7/2+7/2^{+} state. In one nucleus (169Er) the experimental lowest configuration 1/21/2^{-} is obtained only 72 keV above a 7/2+7/2^{+} state.

Three cases deserve a particular attention. In Ref. [24] two assignments (5/2,7/2+5/2^{-},7/2^{+}) are suggested for 163Gd. We found the latter (7/2+7/2^{+}) as the ground state with the former (5/25/2^{-}) lying 579 keV higher with a 1/21/2^{-} configuration located in between at 59 keV. In 165Dy one has found experimentally a 7/2+7/2^{+} ground state with a 1/21/2^{-} isomeric state with an excitation energy of 108 keV. We found both states as the lowest ones but with an inversion, the 1/21/2^{-} being found 48 keV lower in energy. A similar situation is also present in 171Yb with an experimental 1/21/2^{-} ground state and a 7/2+7/2^{+} isomeric state 95 keV above. Here, too, we found these states as the lowest ones but with an inversion whereby the 1/21/2 being found 73 keV lower.

IV.3 Comments on our assignments of spin and parity

In our calculations for both odd-ZZ and odd-NN nuclei, we thus found that we are in agreement in 21 out 29 cases where an assignment of spin and parity has been reported or suggested in Ref. [24]. We have proposed an assignment for two nuclei. In two instances where an isomeric state has been experimentally found at excitation energies in the 100-150 keV range, we have obtained them, yet with an inversion in their respective ordering, corresponding to an error of 50-70 keV. From all these results implying a sample of 29 odd-AA nuclei, yet limiting ourselves to the consideration of intrinsic states within the unified model, we can reasonably hint that our estimates of the low lying bandhead spectra provides a good reproduction of relative energies within not much more than about 150 keV.

Finally we discuss the decoupling constant values corresponding to the 1/21/2 states considered in our calculations of these odd-NN nuclei. It is found to be equal to 0.61-0.61 for the 1/21/2^{-} state of 163Gd, 0.64-0.64 and 0.65-0.65 for the 1/21/2^{-} states of 165Dy and 167Dy, 0.73-0.73 for the 1/21/2^{-} state of 169Er, 0.76-0.76 for the 1/21/2^{-} state of 171Yb and +0.24+0.24 for the 1/21/2^{-} state of 181Hf. In all cases, such values warrants that the bandhead spin assignment as 1/21/2 is correct.

IV.4 Discussion of the odd-even mass differences obtained in our calculations

Tables 2 and 3 display the three-points odd-even mass differences δq(3)\delta_{q}^{(3)} for protons and neutrons. The rms differences between experimental (as deduced from the separation energies given in Ref. [24]) and calculated in all cases for the experimental IπI^{\pi} assignments are 165 keV for protons and 141 keV for neutrons. We can compare these values with what has been obtained in Ref. [6] where direct fits of the pairing matrix elements VnV_{n} and VpV_{p} have been performed on the odd-even mass differences δn(3)\delta_{n}^{(3)} and δp(3)\delta_{p}^{(3)} for a similar sampling of nuclei. There, the rms deviations were calculated as 182 keV for protons and 78 keV for neutrons. Given the expected accuracy on energies of our approach, we can deem the quality of our current approach to be comparable to those of Ref. [24].

To conclude this Subsection, it is instructive to assess numerically on two test cases, the effect of the two improvements with respect to the standard fitting procedure which have been considered in this paper (the Möller-Nix prescription and the correction for the deficiency of the Slater approximation)

We display in Table 4 the average matrix elements VqV_{q} and the resulting odd-even mass-differences with and without taking the RpR_{p} corrective factor for two odd-mass nuclei and consider two types of empirical formula, namely that of Jensen and that of Möller-Nix.

We first compare the results obtained using the Jensen and Möller-Nix formulas without Slater correction. Using the Jensen pairing gaps, the absolute values of the estimated neutron and proton pairing matrix elements are always lower than in the Möller-Nix case. This decrease of about 10 keV in both neutron and proton pairing matrix elements is translated into a decrease of the odd-even mass differences of about 200 keV.

For the odd-neutron 177Yb nucleus and the neighbouring even-even nuclei, using the Möller-Nix gaps, inclusion of the RpR_{p} corrective factor decreases the absolute value of the proton pairing matrix elements (i.e. less pairing) by about 10 keV, having no significant effect, as expected, on the neutron odd-even mass difference around the 177Yb nucleus.

For the odd-proton 177Lu nucleus, including the Slater correction yields also a decrease of about 10 keV of the average pairing matrix element resulting in a decrease of the calculated δp3\delta_{p}^{3} of about 250 keV.

Table 4: Absolute values of the average neutron (VnV_{n}) and proton (VpV_{p}) pairing matrix elements together with the calculated binding energies (BE) and odd-even mass differences δq3\delta^{3}_{q} obtained with four different approaches in determining the pairing gap Δq\Delta_{q} for the 177Yb (odd-neutron with Kπ=9/2+K^{\pi}=9/2^{+}) and 177Lu (odd-proton with Kπ=7/2+K^{\pi}=7/2^{+}) nuclei. Two types of empirical formulas are considered namely the Jensen and Möller-Nix in which the formula are either used as they are, or in which the Slater correction is taken into account via equation (8).
Nucleus Gap formula ZZ NN AA VnV_{n} [MeV] VpV_{p} [MeV] BE [MeV] δq3\delta^{3}_{q} [keV]
177Yb Jensen 70 106 176 0.1582 0.2221 -1415.378
70 107 177 0.1553 0.2205 -1421.230 351
70 108 178 0.1524 0.2188 -1427.784
Jensen + Slater corr. 70 106 176 0.1582 0.2093 -1415.237
70 107 177 0.1553 0.2075 -1421.110 401
70 108 178 0.1524 0.2057 -1427.784
Möller-Nix 70 106 176 0.1681 0.2373 -1415.848
70 107 177 0.1664 0.2364 -1421.462 544
70 108 178 0.1646 0.2355 -1428.164
Möller-Nix + Slater corr. 70 106 176 0.1681 0.2215 -1415.595
70 107 177 0.1664 0.2210 -1421.235 541
70 108 178 0.1646 0.2204 -1427.957
177Lu Jensen 70 106 176 0.1582 0.2221 -1415.378
71 106 177 0.1603 0.2210 -1421.858 432
72 106 178 0.1623 0.2199 -1429.201
Jensen + Slater corr. 70 106 176 0.1582 0.2093 -1415.237
71 106 177 0.1603 0.2088 -1421.858 270
72 106 178 0.1623 0.2082 -1429.017
Möller-Nix 70 106 176 0.1681 0.2373 -1415.848
71 106 177 0.1683 0.2333 -1422.065 648
72 106 178 0.1685 0.2292 -1429.579
Möller-Nix + Slater corr. 70 106 176 0.1681 0.2215 -1415.595
71 106 177 0.1683 0.2190 -1422.045 397
72 106 178 0.1685 0.2164 -1429.290

V Concluding remarks

The present study complements in the rare-earth region and for the SIII Skyrme interaction what had been found in Ref. [1]. There it was shown that the rms deviation between calculated and experimental moments of inertia deduced from the first 2+2^{+} level energy found in 11 well and rigidly deformed rare-earth nuclei was equal to 1.77 2MeV1\hbar^{2}\mbox{MeV}^{-1} corresponding to about 5%5\%. It was almost equal (1.75 2MeV1\hbar^{2}\mbox{MeV}^{-1}) with what had been found from a direct fit of these moments of inertia (in the same region with the same interaction) in Ref. [6].

Combining these results with ours, we can conclude that the method proposed in [1] allows for a treatment of pairing correlations of the same quality as and simpler than a lengthy and more or less localized fit process to be performed for each particle-hole interaction. However, it is a priori suited to describe pairing correlations only at the equilibrium deformation of well and rigidly deformed nuclei. In practice, all the calculations of the HF+BCS type within the seniority forces approach have assumed that the matrix elements obtained in these particular situations could be widely used in, e.g., computing potential-energy curves or surfaces as functions of some deformation parameters or multipole moments. Using the current approach as a starting point to determine the intensities of pairing residual interactions would free corresponding HF+BCS or HFB calculations from such ambiguities. This improvement is currently under study.

Acknowledgements.
T. V. Nhan Hao, P. Quentin and D. Quang Tam acknowledge the support by the Hue University under the Core Research Program, Grant no. NCM.DHH.2018.09. Another co-author (M.H. Koh) would also like to acknowlege Universiti Teknologi Malaysia for its UTMShine grant (grant number Q.J130000.2454.09G96).

References

  • Koh and Quentin [2024] M.-H. Koh and P. Quentin, Phys. Rev. C 110, 024311 (2024).
  • Jensen and Hansen [1984] A. S. Jensen and P. G. Hansen, Nucl. Phys. A 431, 393 (1984).
  • Madland and Nix [1988] D. Madland and J. Nix, Nucl. Phys. A 476, 1 (1988).
  • Möller and Nix [1992] P. Möller and J. Nix, Nucl. Phys. A 536, 20 (1992).
  • Bohr et al. [1958] A. Bohr, B. R. Mottelson, and D. Pines, Phys. Rev. 110, 936 (1958).
  • Nor et al. [2019] N. M. Nor, N.-A. Rezle, K.-W. Kelvin-Lee, M.-H. Koh, L. Bonneau, and P. Quentin, Phys. Rev. C 99, 064306 (2019).
  • Brack et al. [1972] M. Brack, J. Damgaard, and A. J. et al., Rev. Mod. Phys. 44, 320 (1972).
  • Brack and Pauli [1973] M. Brack and H. Pauli, Nucl. Phys. A 207, 401 (1973).
  • Moszkowski [1957] S. A. Moszkowski, Handbuch der Physik XXXIX, 469 (1957).
  • Ring and Schuck [1980] P. Ring and P. Schuck, The Nuclear Many-body Problem (Springer-Verlag, 1980) p. 240.
  • Slater [1951] J. Slater, Phys. Rev. 81, 385 (1951).
  • Titin-Schnaider and Quentin [1974] C. Titin-Schnaider and P. Quentin, Phys. Lett. B 49, 397 (1974).
  • Skalski [2001] J. Skalski, Phys. Rev. C 63, 024312 (2001).
  • Bloas et al. [2011] J. L. Bloas, M.-H. Koh, P. Quentin, L. Bonneau, and J. Ithnin, Phys. Rev. C 84, 0143310 (2011).
  • Bonche et al. [1985] P. Bonche, H. Flocard, P. Heenen, S. Krieger, and M. Weiss, Nucl. Phys. A 443, 39 (1985).
  • Beiner et al. [1975] M. Beiner, H. Flocard, N. V. Giai, and P. Quentin, Nucl. Phys. A 238, 29 (1975).
  • Minkov et al. [2022] N. Minkov, L. Bonneau, P. Quentin, J. Bartel, H. Molique, and D. Ivanova, Phys. Rev. C 105, 044329 (2022).
  • Flocard et al. [1973a] H. Flocard, P. Quentin, A. Kerman, and D. Vautherin, Nucl. Phys. A 203, 433 (1973a).
  • Dobaczewski et al. [2001] J. Dobaczewski, P. Magierski, W. Nazarewicz, W. Satula, and Z. Szymanski, Phys. Rev. C 63, 024308 (2001).
  • Duguet et al. [2001] T. Duguet, P. Bonche, P.-H. Heenen, and J. Meyer, Phys. Rev. C 65, 014310 (2001).
  • Bonneau et al. [2015] L. Bonneau, N. Minkov, D. D. Duc, P. Quentin, and J. Bartel, Phys. Rev. C 91, 054307 (2015).
  • Å Bohr and Mottelson [1998] Å Bohr and B. Mottelson, Nuclear Structure Vol. II (World Scientific, Singapore, 1998) p. 32.
  • Koh et al. [2016] M. Koh, D. Duc, T. N. Hao, H. Long, P. Quentin, and L. Bonneau, Eur. Phys. J. A 52, 3 (2016).
  • [24] NNDC, https://www.nndc.bnl.gov/nudat3/ .
  • Raman et al. [2001] S. Raman, C. N. Jr., and P. Tikkanen, At. Data Nucl. Tables 78, 1 (2001).
  • [26] NUMOR, https://magneticmoments.info/numor/in_search.php (cut-off date 2019.03.31) and references quoted therein .
  • Flocard et al. [1973b] H. Flocard, P. Quentin, and D. Vautherin, Phys. Lett. B 46, 304 (1973b).