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

Realistic evaluation of Coulomb potential in spherical nuclei and test of the traditional approach

L. Xayavong xayavong.latsamy@yonsei.ac.kr    Y. Lim ylim@yonsei.ac.kr Department of Physics, Yonsei University, Seoul 03722, South Korea
Abstract

A realistic evaluation of Coulomb potential has been made for some selected nuclei using the available model-independent data for the charge density and the recent development of Coulomb energy-density functional. Within the Woods-Saxon potential as a nuclear component, we are able to quantify the differences in proton single-particle energies due to the differences from the model-independent data of the uniform distribution, the two-parameter Fermi function, as well as the charge density obtained from a microscopic Hartree-Fock calculation using the effective Skyrme interaction. The obtained energy differences are generally small in magnitude, namely about 100 keV or less if the parameters of the charge density models are appropriately determined. Considerable larger differences appear when the last occupied state is highly filled and, at the same time, has a small orbital angular momentum. Sulfur isotopes (Z=16Z=16) are a perfect example of these nuclei. Unfortunately, despite its simplicity, the uniform distribution cannot be used for evaluating the Coulomb exchange term within a well-established method because it is not differentiable at the surface of a nucleus. Traditionally, the missing of exchange term is corrected for by simply excluding the last-proton contribution to the direct term. We also investigate this approach and find that its effect is simply an introduction of the factor (Z1)/Z(Z-1)/Z into the Coulomb direct term. From medium to heavy nuclei (typically beyond the sdsd shell) the resulting proton levels are 300-800 keV higher than those obtained with the exact Fock term. The result for lighter nuclei tends to be opposite because the factor (Z1)/Z(Z-1)/Z decreases rapidly towards the limit of Z1Z\to 1. Therefore, this traditional approach should be avoided for a precision nuclear structure calculation.

I Introduction

The Coulomb repulsion between protons is the best known part of nuclear Hamiltonian. Its action results in a large class of nuclear properties, including the departure from the N=ZN=Z line of the stability valley or the fission phenomenon which yields a clean-cut limitation to the nuclear size, and some specific properties of the β\beta, proton and α\alpha radioactivities. As the main actor in the violation of the isospin symmetry, the Coulomb interaction is also responsible for various nuclear structure phenomena, in particular it induces isospin mixing in nuclear states, displacement energy between members of isobaric multiplet and isospin forbidden transitions [1, 2, 3, 4]. In addition, the correction due to the isospin-symmetry breaking has became a limiting factor in the low-energy tests of the electroweak sector of the Standard Model via nuclear β\beta decays [5]. Therefore, in a microscopic model of the nuclear structure, it is of paramount importance to properly take the Coulomb interaction into account.

Within the independent-particle framework, the AA-nucleon problem is simplified to AA single-nucleon problems starting either from an effective nucleon-nucleon interaction [6, 7] or an effective potential [8, 9]. Therefore, the resulting nuclear wave function is simply an antisymmetric product of individual wave functions or the so-called Slater determinant. As is well known, this simplified model itself has a limited application. In principle, the self-consistent spherical Hartree-Fock (HF) mean field is only appropriate for a closed-shell system, whereas the phenomenological one is usually optimized for single-particle or single-hole states such as low-lying states of nuclei in the vicinity of a closed-shell core. Nevertheless, the independent-particle potential serves as the basic part of the full nuclear Hamiltonian and therefore is the key for success for all nuclear many-body approaches. In some specific applications, it is mandatory to include the Coulomb contribution within the one-body part, especially in the shell model for which the valence spaces are too small to produce all significant configurations induced by the Coulomb interaction. As an example, the shell-model description of the superallowed 0+0+0^{+}\to 0^{+} Fermi β\beta decay would not agree with the Standard-Model predictions, unless the harmonic oscillator basis is replaced with realistic radial wave functions [10, 11, 12]. In particular, the radial mismatch between the initial and final states is greatly enhanced when the mother nucleus is weakly bound.

In nearly all calculations within the phenomenological Woods-Saxon (WS) potential [8, 13, 14, 9, 15], the Coulomb repulsion is accounted for using the approximation of a uniformly charge distributed sphere of radius RCR_{C}:

VC(r)=(Z1)e2{1r,r>RC1RC(32r22RC2),otherwise,V_{C}(r)=(Z-1)e^{2}\left\{\begin{array}[]{ll}\displaystyle\frac{1}{r},&r>R_{C}\\[7.22743pt] \displaystyle\frac{1}{R_{C}}\left(\frac{3}{2}-\frac{r^{2}}{2R_{C}^{2}}\right),&\text{otherwise,}\end{array}\right. (1)

where RCR_{C} is usually taken as RC=r0(A1)13R_{C}=r_{0}(A-1)^{\frac{1}{3}} with r01.26r_{0}\approx 1.26 fm [9]. Alternatively RCR_{C} can be extracted from the charge radius RchR_{ch} via [16]

RC2=53Rch252i=13θiri254(mc)2+52b2A\displaystyle R_{C}^{2}=\frac{5}{3}R_{ch}^{2}-\frac{5}{2}\sum_{i=1}^{3}\theta_{i}r_{i}^{2}-\frac{5}{4}\left(\frac{\hbar}{mc}\right)^{2}+\frac{5}{2}\frac{b^{2}}{A} (2)

where the nuclear oscillator length parameter is given by b2A13b^{2}\approx A^{\frac{1}{3}} fm2. An improved parameterization of b2b^{2} can be found in Ref. [17]. The last three terms on the right-hand-side of Eq. (2) account for the internal structure of proton where iθiri2=0.518\sum_{i}\theta_{i}r_{i}^{2}=0.518 fm2 [18], Darwin-Foldy term (/mc=0.21\hbar/mc=0.21 fm) and center-of-mass (COM) motion, respectively. It should be noted that a further modification was made in the construction of Eq. (1), namely the contribution of the last proton was excluded so that the potential in Eq. (1) is proportional to Z1Z-1 instead of ZZ as seen in classical electromagnetism. This exclusion of the last proton can also be considered as the correction of missing of the Coulomb exchange potential in the uniform charge approximation. However, this approach for the self-interaction correction has not been rigorously checked for a microscopic nuclear structure calculation.

In general, the Coulomb potential can be derived from the two-body Coulomb interaction using the variational principle. Its direct part is basically a known functional of charge density, while its exchange counterpart can be treated with a great precision using a local density approximation [19]. Therefore, the Coulomb potential can be determined in a self-consistent manner by minimizing the total energy as in the HF theory [6], or evaluated independently from nuclear component using charge density data from external sources.

In the present work, Eq. (1) is investigated in various aspects. The validity of the uniform charge distribution which is the central pillar of Eq. (1) is checked within the framework of the phenomenological WS mean field and the model-independent data deduced from electron scattering experiments [20]. The errors due to the above-mentioned approach for the self-interaction correction are separately quantified through the comparison to the results obtained with the exact Coulomb exchange functional. Moreover, we discuss the difficulty in evaluating the Coulomb exchange term when a uniform distribution is assumed or when the model-independent data are used instead. As a sensitivity study, we also consider the two-parameter Fermi (2pF) function and the microscopic Skyrme HF calculation as an alternative model for the charge-density distribution. Our calculations are performed for a wide variety of nuclei and covers a mass range between A=16A=16 and 209209. There includes two closed-shell N=ZN=Z nuclei (16O and 40Ca), two closed-shell NZN\neq Z nuclei (48Ca and 208Pb), two closed-subshell nuclei (28Si and 32S) and four opened-shell nuclei (58Ni, 205Tl, 206Pb and 209Bi).

The paper is organized as follows. The standard parameterization structure of the WS potential is described in the section II. Section III reviews the density functional forms of the Coulomb direct and exchange terms as employed in the self-consistent mean-field theory. The selection of input charge densities and discussion of their properties are given in the section IV In the section V, a comparative test of Coulomb potential for different charge density models and different functional forms is carried out. The summary and conclusion are given in the last section VI.

II Nuclear potential

In this work, the phenomenological WS potential is selected for the nuclear component of our independent-particle Hamiltonian. As a standard parameterization structure, this potential consists of a central, spin-orbit, isospin-dependent and Coulomb term, namely

V(r)=V0f0(r)Vs(rs)21rddrfs(r)lσ+Vsym(r)+(12tz)VC(r),\begin{array}[]{ll}V(r)=&\displaystyle V_{0}f_{0}(r)-V_{s}\left(\frac{r_{s}}{\hbar}\right)^{2}\frac{1}{r}\frac{d}{dr}f_{s}(r)\braket{\vec{l}\cdot\vec{\sigma}}\\[7.22743pt] &+V_{sym}(r)+(\frac{1}{2}-t_{z})V_{C}(r),\end{array} (3)

where tzt_{z} is the isospin projection of the nucleon with the convention of tz=12t_{z}=\frac{1}{2} for neutron and 12-\frac{1}{2} for proton. The functions fi(r)f_{i}(r) are defined as

fi(r)=11+exp(rRiai),f_{i}(r)=\frac{1}{1+\exp\left(\frac{r-R_{i}}{a_{i}}\right)}, (4)

with i=0i=0 or ss denoting either the central or spin-orbit terms. The restriction of a0=asa_{0}=a_{s} is usually adopted because of lack of experimental constraints. On the contrary to the assumption a0=asa_{0}=a_{s}, a smaller spin-orbit radius (Rs<R0)(R_{s}<R_{0}) was suggested because the two-body spin-orbit interaction has a shorter range [15]. For example, Rs/R0=0.921R_{s}/R_{0}=0.921 was obtained for the Seminole parameterization given in Ref. [9].

The expectation value lσ\braket{\vec{l}\cdot\vec{\sigma}} appearing in the spin-orbit term can be written as

lσ={lifj=l+12(l+1)ifj=l12.\braket{\vec{l}\cdot\vec{\sigma}}=\left\{\begin{array}[]{lll}l&\text{if}&j=l+\frac{1}{2}\\[7.22743pt] -(l+1)&\text{if}&j=l-\frac{1}{2}.\end{array}\right. (5)

In order to preserve the fundamental symmetries, the phenomenological effective potential such as WS is normally treated as a nuclear mean field created by the core of (A1)(A-1) nucleons. The exclusion of last nucleon contribution also serves as a self-interaction correction due to the missing of exchange terms as discussed for the Coulomb potential in the introduction. For these reasons, the WS radii are usually parameterized as a function of (A1)(A-1) instead of AA namely Ri=ri(A1)13R_{i}=r_{i}(A-1)^{\frac{1}{3}}. Furthermore, if we neglect for instance the internal structure of the core of (A1)(A-1) nucleons, the nucleus can be regarded as a system of two point-like particles. Within this simplified picture, the COM correction for the WS Hamiltonian can be easily implemented by replacing the nucleon mass mm in the kinetic energy term of the radial Schrödinger’s equation with the reduced mass μ\mu defined below,

μ=m(A1)A.\mu=m\frac{(A-1)}{A}. (6)

This COM correction is fully validated at large separation where the core structure contribution is negligible. A further discussion of the COM corrections is given in Refs. [21, 12].

Most nuclei have a different number of protons and neutrons. Apart from the Coulomb repulsion, different number of neutrons and protons in nuclei causes an extra shift between neutron and proton potential depth. As an experimental evidence, nuclei with N=ZN=Z tend to have a greatest binding energy comparing with the other possible configurations of protons and neutrons. In practice, this effect is commonly accounted for by adding the symmetry term expressed below,

Vsym(r)=2tzV1(NZ)Af0(r).V_{sym}(r)=2t_{z}V_{1}\frac{(N-Z)}{A}f_{0}(r). (7)

In this work, we neglect the symmetry term contribution to the spin-orbit coupling [22] for simplicity. Some further additional terms or slightly different parameterization structure can also be found in literature. For example, a more fundamental form of the isospin-dependent term was proposed by Lane in Ref. [23]. Study of such variations of the nuclear component is outside the scope of this work.

The Coulomb potential VC(r)V_{C}(r) can be evaluated within various different methods, it will be separately described in the following sections.

Basically, the WS potential cannot be used for the total binding energy since it is not based upon a specific effective two-body interaction. Normally the WS parameters (V0V_{0}, VsV_{s}, r0r_{0}, rsr_{s}, a0a_{0}, asa_{s}, V1V_{1}, RCR_{C}) are chosen for a best fit of nuclear single-particle energies and nuclear charge radii. The set of WS parameter values named BMm as listed in Table I of Ref. [11] will be used in the present study.

III Coulomb potential as a charge density functional

Before discussing the details of our study, it is instructive to describe in this section the basic formulas and some general properties of the Coulomb potential. We recall that, according to the self-consistent HF theory [6], the Coulomb contribution to the mean field consists of a direct and an exchange term which can be symbolically written as

VC(r)=Vdir(r)+Vexc(r).V_{C}(r)=V_{dir}(r)+V_{exc}(r). (8)

Note that the spherical symmetry is assumed throughout this paper. The Coulomb direct term for a spherically symmetric nucleus, after integrating out the angular variables, is reduced to

Vdir(r)=4πe2[1r0rx2ρch(x)𝑑x+rxρch(x)𝑑x].V_{dir}(r)=4\pi e^{2}\left[\frac{1}{r}\int_{0}^{r}x^{2}\rho_{ch}(x)\,dx+\int_{r}^{\infty}x\rho_{ch}(x)\,dx\right]. (9)

One can observe here that if the charge density ρch(r)\rho_{ch}(r) is constant inside the radius RCR_{C} and vanishes elsewhere, the expression (9) will return the potential in Eq. (1) except that it would be proportional to ZZ instead of Z1Z-1 for the reason discussed in the introduction. More details on the uniform charge distribution are given in subsection IV.2.

Since the Coulomb force has an infinite range, the Coulomb exchange term is nonlocal in coordinate space and thus much harder to calculate especially within the self-consistent mean-field framework. To avoid this inconvenience, a local density approximation is usually employed. The popular one is that invented by Slater [24] with which the Coulomb exchange term is given by a function of the charge density

VexcSla(r)=e2[3πρch(r)]13.V_{exc}^{Sla}(r)=-e^{2}\left[\frac{3}{\pi}\rho_{ch}(r)\right]^{\frac{1}{3}}. (10)

The quality of this approximation was checked against the exact calculation [25] for a number of spherical nuclei from 16O to 126310{}^{310}_{126}Ubh. The proton energy levels found in Ref. [25] were underbound by 100 to 550 keV for occupied states and overbound by 100 to 200 keV for unoccupied states, compared with those obtained with the exact Fock term.

It was recently demonstrated that the Coulomb energy density functional built by using the generalized gradient approximation (GGA) [19] produces almost the same accuracy for the total energy as the exact treatment of the Fock term while the numerical price is still the same as that of the Slater approximation. The Coulomb exchange term derived from the GGA depends not only on the charge density but its gradient with respect to radial distance, namely

VexcGGA(r)=VexcSla(r){F(s)[s+34kFr]F(s)+[s23ρch′′(r)8ρch(r)kF2]F′′(s)}.V_{exc}^{GGA}(r)=V_{exc}^{Sla}(r)\left\{F(s)-\left[s+\frac{3}{4k_{F}r}\right]F^{\prime}(s)+\left[s^{2}-\frac{3\rho^{\prime\prime}_{ch}(r)}{8\rho_{ch}(r)k_{F}^{2}}\right]F^{\prime\prime}(s)\right\}. (11)

where ρch′′(r)\rho^{\prime\prime}_{ch}(r) denotes the second derivative of ρch(r)\rho_{ch}(r) with respect to rr, whereas F(s)F^{\prime}(s) and F′′(s)F^{\prime\prime}(s) denote the first and second derivatives of F(s)F(s) with respect to ss, respectively. Note that F(s)F(s) is the enhancement factor introduced for the Coulomb exchange potential in GGA. Following Perdew-Burke-Ernzerhof [26], F(s)F(s) is parameterized as

F(s)=1+κκ1+μs2/κ,\displaystyle F(s)=1+\kappa-\frac{\kappa}{1+\mu s^{2}/\kappa}, (12)

where the two parameters κ\kappa and μ\mu have recently been revised for nuclear physics applications by Naito et al. [19] and the best-fit values were found to be 0.804 and 0.274, respectively. The function ss denotes the dimensionless density gradient

s=|ρch(r)|2kFρch(r),s=\frac{|\boldsymbol{\nabla}\rho_{ch}(r)|}{2k_{F}\rho_{ch}(r)}, (13)

where |ρch(r)||\boldsymbol{\nabla}\rho_{ch}(r)| is the norm of ρch(r)\boldsymbol{\nabla}\rho_{ch}(r) and the Fermi momentum is defined as kF=[3π2ρch(r)]13k_{F}=[3\pi^{2}\rho_{ch}(r)]^{\frac{1}{3}}. We remark here that if ρch(r)\rho_{ch}(r) is a slowly varying function of rr, its gradient will be vanished (s=0s=0) and thus F(0)=1F(0)=1, then the Coulomb exchange term in Eq. (11) will be reduced to the Slater approximation (10). Considering the general behavior of charge density in a nucleus, the Coulomb exchange potential from the GGA is mostly affected in the nuclear surface region where the charge density gradient is peaked.

The contribution of higher order electromagnetic effects such as the vacuum polarization and the Coulomb spin-orbit can also be included. However, both of them were found to be completely negligible [12].

IV Input charge densities

According to the formalism reviewed in the previous section, the charge density is regarded as the fundamental ingredient with which the Coulomb potential is determined. In the following list, we briefly describe the conventional methods for deducing charge density from electron scattering or the so-called model-independent analyses. The data obtained from this source is used as the reference for our comparative study. The frequently-used hypothetical and phenomenological models as well as the microscopic self-consistent mean-field calculation of the charge density are also discussed. The validity of these theoretical models as the input of the Coulomb potential is investigated in the next section.

IV.1 Model-independent analyses

The charge form factors for some stable nuclei were accurately measured by electron scattering. These data are usually analyzed within two different model-independent approaches to extract numerical values for the charge density as a function of rr namely the Fourier-Bessel (FB) [27] and the sum-of-Gaussians (SOG) [28] expansion. In the former approach, the charge density is expanded in terms of the spherical Bessel function of order zero (j0j_{0}) namely,

ρFB(r)={ν=1nmaxaνj0(νπrRcut),rRcut0,otherwise,\rho_{\text{FB}}(r)=\left\{\begin{array}[]{ll}\displaystyle\sum_{\nu=1}^{n_{max}}a_{\nu}j_{0}\left(\frac{\nu\pi r}{R_{cut}}\right),&r\leq R_{cut}\\[10.84006pt] 0,&\text{otherwise},\end{array}\right. (14)

where aνa_{\nu} are the expansion coefficients and RcutR_{cut} is a cut-off radius beyond which the charge density is sufficiently small and is equated to zero. The first nmaxn_{max} coefficients of this series expansion are obtained directly from the experimental data [20]. Here nmax=Rcutqmax/πn_{max}=R_{cut}q_{max}/\pi where qmaxq_{max} is the maximum momentum transfer up to which the charge form factor data are determined. For the normalization, the integral of ρFB(r)\rho_{\text{FB}}(r) over all spaces must be equal to the total nuclear charge +Ze+Ze.

In the sum-of-Gaussians approach, the charge density is expressed as

ρSOG(r)=i=1mmaxAi{exp[xi(r)2]+exp[yi(r)2]},\rho_{\text{SOG}}(r)=\sum_{i=1}^{m_{max}}A_{i}\left\{\exp\left[-x_{i}(r)^{2}\right]+\exp\left[-y_{i}(r)^{2}\right]\right\}, (15)

where yi(r)=(r+Ri)/γy_{i}(r)=(r+R_{i})/\gamma and xi(r)=(rRi)/γx_{i}(r)=(r-R_{i})/\gamma. The expansion coefficients AiA_{i} are given by

Ai=ZeQi2π32γ3(1+2Ri2/γ2),A_{i}=\frac{ZeQ_{i}}{2\pi^{\frac{3}{2}}\gamma^{3}(1+2R_{i}^{2}/\gamma^{2})}, (16)

where γ\gamma is the width of the Gaussians and is related to the root-mean-square radius (RMS) through Rg=γ3/2R_{g}=\gamma\sqrt{3/2}. According to Sick [28], γ\gamma is chosen equal to the smallest width of the peaks in the nuclear radial wave functions calculated using the HF method. The author reported that the γ\gamma values extracted from the harmonic oscillator and WS radial wave functions yield almost identical results. The charge fraction QiQ_{i} must be normalized such that iQi=1\sum_{i}Q_{i}=1.

In practice, only first few terms are included within the sum in Eq. (14) and Eq. (15). For example, a truncation with nmax17n_{max}\leq 17 and mmax12m_{max}\leq 12 was imposed for the analyses carried out in Ref. [20]. The parameters of ρFB(r)\rho_{\text{FB}}(r) and ρSOG(r)\rho_{\text{SOG}}(r) deduced from experiments are given in the data compilation [20].

Refer to caption
Figure 1: (Color online) Illustration of oscillations in the charge densities deduced from electron scattering experiments using the FB (right) and SOG (left) analyses [20].

We remark that the charge density obtained from the FB approach has an undesired property–it contains an oscillatory component and, sometimes, has a negative value near the cut-off radius RcutR_{cut} as illustrated in the right panel of Fig. 1. Although these oscillations are extremely small in magnitude, they are greatly amplified by the first and second derivatives, then leading to a large error, if the GGA is employed for treating the Coulomb exchange term. Oscillations in the asymptotics of charge density obtained from the SOG approach are also visible for some nuclei, however their magnitude are generally much smaller (larger width) as shown in the left panel of Fig. 1. This unphysical property originates from the incompleteness in the expansions and too small cut-off radius in the FB analysis. Fundamentally, the tail of charge density should decrease monotonically since the asymptotic radial wave functions decay exponentially.

In order to avoid this problem, we replace both FB and SOG data from a point just before the oscillations occur (denoted as RxR_{x}) towards infinity with the 2pF function (see subsection IV.3 for the definition of 2pF). We select for both data sets Rx=4.5R_{x}=4.5 fm for A58A\leq 58 and 8.58.5 fm for heavier nuclei. Instead of performing a global fit, the 2pF function parameters are here determined by matching it as well as its first and second derivatives with those of the data at RxR_{x}. Therefore, the resulting 2pF function might not be optimal for the whole available data in the domain of [Rx,+][R_{x},+\infty]. Nevertheless, it is the only way to ensure the continuity of the charge density at RxR_{x}. This replacement would not cause a serious error since the charge density in this domain is small.

IV.2 Uniform charge distribution

For a sphere of radius RCR_{C} containing a total charge of +Ze+Ze uniformly distributed throughout its volume, the charge density is written as

ρUnif(r)={ρ0,rRC0,otherwise,\rho_{\text{Unif}}(r)=\left\{\begin{array}[]{ll}\displaystyle\rho_{0},&r\leq R_{C}\\[7.22743pt] 0,&\text{otherwise},\end{array}\right. (17)

and the normalization condition implies that

ρ0=3Ze4πRC3.\rho_{0}=\frac{3Ze}{4\pi R_{C}^{3}}. (18)

Here we follow Ref. [29] wherein ρUnif(r)\rho_{\text{Unif}}(r) is regarded as a pointed proton distribution and is normalized to ZZ instead of +Ze+Ze. Therefore, its RMS radius (denoted as RUnifR_{\text{Unif}}) can be calculated analytically, namely

RUnif=RC35.\displaystyle R_{\text{Unif}}=R_{C}\sqrt{\frac{3}{5}}. (19)

In this uniform distribution, the corrections for the finite size, Darwin-Foldy term and COM motion must be introduced in order to convert RUnifR_{\text{Unif}} into the charge radius. Eq. (19) provides an experimental constraint for RCR_{C}, at least, for the cases where the experimental data are available. Otherwise it can be parameterized as usual, namely RC=rC(A1)13R_{C}=r_{C}(A-1)^{\frac{1}{3}} where rC1.26r_{C}\approx 1.26 fm [9]. The validity of different methods for determining RCR_{C} can also be tested within the framework of the present study. More discussions on this point are given in the next section.

Because of its simplicity, the uniform charge distribution is widely used in nuclear physics, especially within the nuclear optical model. However, one should note that it is an assumption of classical electromagnetism and has no quantum mechanical equivalent since wave functions, which are the building-blocks of charge density, are required to be continuous in coordinate space. Furthermore, the two aforementioned approaches to the Coulomb exchange term (Slater and GGA) are evidently inapplicable for this distribution because of its discontinuity and nondifferentiability at the surface of a nucleus.

IV.3 Two-parameter Fermi function

We also consider a realistic phenomenological model for the charge density distribution namely the two-parameter Fermi function,

ρ2pF(r)=ρ~01+exp(rcz).\displaystyle\rho_{2pF}(r)=\frac{\tilde{\rho}_{0}}{1+\exp(\frac{r-c}{z})}. (20)

Unlike the previous models, the 2pF is continuous and differentiable, and moreover it decays monotonically towards large distances. Although the 2pF function looks similar to the fi(r)f_{i}(r) functions of the WS potential in Eq. (4), in general cc and zz are smaller than the length and the surface diffuseness parameter of the WS potential. It is also evident from the Skyrme HF theory that the mean field potential is not a linear function of densities. Thus it is not necessary that their geometrical characterizing parameters a the same with the WS potential parameters. More details on this point are given in Ref. [30].

As its name indicates, the 2pF is determined by two parameters because ρ~0\tilde{\rho}_{0} is obtained via the normalization condition,

ρ~0=Ze4πz3F2(c/z),\tilde{\rho}_{0}=\frac{Ze}{4\pi z^{3}F_{2}(c/z)}, (21)

where F2(c/z)F_{2}(c/z) is the second order Fermi integral (see Appendix C of Ref. [16]). Similar to the previous subsection, we employ the convention that ρ2pF(r)\rho_{2pF}(r) is normalized to ZZ, therefore the parameter cc can be extracted from the charge radius by solving the following equation [16]

Rch2=4πρ~0z5ZF4(cz)+32i=13θiri2+34(mc)232b2A.R_{ch}^{2}=\frac{4\pi\tilde{\rho}_{0}z^{5}}{Z}F_{4}\left(\frac{c}{z}\right)+\frac{3}{2}\sum_{i=1}^{3}\theta_{i}r_{i}^{2}+\frac{3}{4}\left(\frac{\hbar}{mc}\right)^{2}-\frac{3}{2}\frac{b^{2}}{A}. (22)

As before, the last three terms account for the finite size of proton, Darwin-Foldy term and center-of-mass motion, respectively. F4(c/z)F_{4}(c/z) is the fourth order Fermi integral (see Appendix C of Ref. [16]).

Recently, Horiuchi [31] proposed a new method for determining the surface thickness of charge density. Implementing the Taylor expansion of ρ2pF(r)\rho_{2pF}(r) at r=cr=c and retaining up to the first-order term, he derived the following relation,

z=ρ~04[ρ2pF(c)]1,z=-\frac{\tilde{\rho}_{0}}{4}[\rho_{2pF}^{\prime}(c)]^{-1}, (23)

where ρ2pF(c)\rho_{2pF}^{\prime}(c) denotes the first derivative of ρ2pF(r)\rho_{2pF}(r) at r=cr=c. By matching ρ2pF\rho_{2pF} on the right-hand-side of Eq. (23) with the charge density constructed from eigenfunctions of the WS potential whose Coulomb term is, in turn, a function of ρ2pF\rho_{2pF}, Eq. (23) can be solved in a self-consistent manner. This offers an alternative means for constraining zz when experimental data are unavailable.

Although this distribution is nicely representative for the diffuseness at the nuclear surface region, it is not able to describe the oscillations of the charge density observed in the nucleus interior. It has been shown that this difficulty can be overcome by extending Eq. (20) to the three-parameter Fermi (3pF) function [32]. However, the 3pF is not a good choice for a general application because of lack of experimental data for constraining the third parameter.

IV.4 Hartree-Fock calculations

Within a microscopic nuclear structure model, the charge density is basically decomposed into three components

ρch(r)=ρchp(r)+ρchn(r)+ρchls(r)\rho_{ch}(r)=\rho_{ch}^{p}(r)+\rho_{ch}^{n}(r)+\rho_{ch}^{ls}(r) (24)

where ρchp(r)\rho_{ch}^{p}(r)/ρchn(r)\rho_{ch}^{n}(r) come from the finite charge distribution of the proton/neutron folded with the point-like proton/neutron density, and ρchls(r)\rho_{ch}^{ls}(r) is the relativistic electromagnetic correction which depends on the spin-orbit coupling. However, the shape of the charge-density distribution is mainly determined by the shape of the point-like proton density. The contributions ρchn(r)\rho_{ch}^{n}(r) and ρchls(r)\rho_{ch}^{ls}(r) were found to be negligible [12]. Furthermore, they tend to cancel out each other, thus will not be considered here.

By definition ρchp(r)\rho_{ch}^{p}(r) is given by

ρchp(r)=𝑑𝒓ρp(r)Gp(𝒓𝒓),\rho_{ch}^{p}(r)=\int d\boldsymbol{r}^{\prime}\rho_{p}(r^{\prime})G_{p}(\boldsymbol{r}-\boldsymbol{r}^{\prime}), (25)

where 𝒓\boldsymbol{r} is the position vector in 3\mathbb{R}^{3}. The effective electromagnetic form factor GpG_{p} is taken as a sum of three Gaussians as described in Ref. [18]. The point-like proton density ρp(r)\rho_{p}(r) can be defined in terms of proton radial wave functions Rα(r)R_{\alpha}(r) namely

ρp(r)=14παnα|Rα(r)|2,\rho_{p}(r)=\frac{1}{4\pi}\sum_{\alpha}n_{\alpha}|R_{\alpha}(r)|^{2}, (26)

where α\alpha stands for the spherical quantum numbers nljnlj and the sum runs over all occupied states. The proton occupation number nαn_{\alpha} is obtained with the so-called equal-filling approximation. Therefore, for a closed-shell configuration, nα=(2j+1)n_{\alpha}=(2j+1) for the occupied orbits and 0 for the unoccupied orbits.

The radial wave functions Rα(r)R_{\alpha}(r) are taken as the eigenfunctions of the Skyrme HF mean field. For the Skyrme Hamiltonian, we select the SLY5 parameter set [33] which is invariant under rotation in the isospin space, while the Coulomb exchange term is treated using the Slater approximation. The charge-symmetry and charge-independence breaking forces [34] are neglected. We use the HFBRAD program [7] to solve the spherical Skyrme HF equation with which the Coulomb terms are evaluated using the point-like proton density. The finite size effect is corrected for by using Eq. (25) within an external program after the HF variation is terminated.

Refer to caption
Figure 2: (Color online) Averaged proton single-particle energies between the refined FB and SOG data for charge density. For this calculation, the Coulomb exchange term is evaluated with the GGA and the spin-orbit interaction is included.

V Comparative study of Coulomb potential

This section starts with an inspection of the uncertainties on the experimental data for charge density as well as their impact on proton single-particle energies. As a next step, the validity of the charge density models listed in section IV as an input for Coulomb potential will be checked, through the comparison of their prediction for proton energy levels to that of the model-independent data. Likewise, the approximate Coulomb functionals including that of Eq. (1) and those reviewed in section III will be tested by comparing their prediction with the result obtained using an exact treatment of the Coulomb exchange term.

V.1 Inspection of experimental charge density data

Before performing a comparative study of Coulomb potential, it is necessary to figure out the impact of the differences between the FB and SOG data for charge density on proton single-particle energies. For simplicity, we neglect the uncertainties on the coefficients of FB and SOG expansions even though they are available from the data compilations. Noted that the quantity which will be considered throughout this section is the proton single-particle energies. The other observables such as the RMS radii and neutron skin thicknesses were found to be much less sensitive to small variation in the Coulomb terms.

Furthermore, instead of a point-by-point comparison, we define the following radial mismatch factor,

Λ=Ω¯ΩΩ¯,\Lambda=\frac{\bar{\Omega}-\Omega}{\bar{\Omega}}, (27)

as an effective measure for the charge density differences, where Ω\Omega is the overlap integral between the FB and SOG data:

Ω=4π0ρFB(r)ρSOG(r)r2𝑑r,\Omega=4\pi\int_{0}^{\infty}\rho_{\text{FB}}(r)\rho_{\text{SOG}}(r)r^{2}dr, (28)

whereas Ω¯\bar{\Omega} is the integral of their average squared:

Ω¯=4π0[ρFB(r)+ρSOG(r)2]2r2𝑑r.\displaystyle\bar{\Omega}=4\pi\int_{0}^{\infty}\left[\frac{\rho_{\text{FB}}(r)+\rho_{\text{SOG}}(r)}{2}\right]^{2}r^{2}dr. (29)

The resulting Λ\Lambda values in % are 3.441×103\times 10^{-3} (16O), 5.659×103\times 10^{-3} (28Si), 0.209 (32S), 2.601×102\times 10^{-2} (40Ca), 4.126×102\times 10^{-2} (48Ca), 3.866×103\times 10^{-3} (58Ni), 3.641×104\times 10^{-4} (205Tl), 5.048×104\times 10^{-4} (206Pb) and 1.459×102\times 10^{-2} (208Pb).

We recall that the exact Coulomb exchange term is nonlocal in coordinate spaces and cannot be written as a function of charge density, it is thus inappropriate for the present study which replaces the self-consistent HF mean field with the phenomenological WS potential. In order to demonstrate whether or not the selection of Coulomb exchange functional is matter to the impact of the FB-SOG charge density differences on proton single-particle energies, we consider three different approximations for Coulomb potential: one of them consists purely of the direct term whereas the other two include also an exchange term employing either the Slater approximation or the GGA. Then together with the nuclear component described in Eq. (3), the single-particle energies and wave functions can be obtained by solving the radial Schrödinger’s equation. The splits in proton energies due to the differences in the charge density data are listed in Table 1. As a convention, the negative sign indicates that the FB data yields a lower proton energy level relative to that yielded by the SOG data i.e., ΔESOGFB=EFBESOG\Delta E^{\rm FB}_{\rm SOG}=E_{\rm FB}-E_{\rm SOG}. The meaning of the positive sign is opposite. For completeness, the averaged proton energies between the FB and SOG results calculated using the GGA functional [denoted as E¯SOGFB=(EFB+ESOG)/2\bar{E}^{\rm FB}_{\rm SOG}=(E_{\rm FB}+E_{\rm SOG})/2] are graphically illustrated in Fig. 2. It is clearly observed from Table 1 that the splits in proton energies induced by the charge density differences are insensitive to the Coulomb exchange term. This witnesses that any existing Coulomb functionals can be employed for our test of the charge density models which will be discussed in the following subsection.

We remark that although the amount of energy splits is mainly determined by the size of Λ\Lambda, it can be greatly enhanced by the shell-structure effect. As a general feature, the solution of the radial Schrödinger equation for protons will be particularly sensitive to the Coulomb terms when the last occupied orbit is highly filled and at the same time, has a low centrifugal barrier (low orbital angular momentum). In the remainder of this paper, this phenomenon will be referred to as the “weakly bound effect”. This is the main reason behind the large energy split of about 500-500 keV for 32S. We notice that this number is comparable to the energy splits due to the use of the Slater approximation within the Skyrme HF framework studied in Ref. [25]. Therefore, a special attention must be paid to the inclusion of this nucleus for our analyses in the following subsection. The impact for the two Calcium (Ca40{}^{40}\rm{Ca},Ca48{}^{48}\rm{Ca}) isotopes is also quite large namely about 100-100 keV, while the impact for the other cases is less than 50 keV in magnitude.

V.2 Test of charge density models

Because of its discontinuity and nondifferentiability, the uniform distribution is inappropriate to be used as an input for calculating the Coulomb exchange term within any existing approximations. In order to evaluate the Coulomb potential for this hypothetical distribution as well as the other charge density models listed in section IV on an equal footing, the Coulomb exchange term will be omitted for our study in this subsection. It was shown in the previous subsection that the absence of the Coulomb exchange term does not significantly affect the energy splits induced by the differences in the charge density data (ΔESOGFB\Delta E^{\rm FB}_{\rm SOG}), even for 32S where Λ\Lambda is as large as 0.209 %.

Here we use the averaged proton energy between FB and SOG values as a reference (denoted as E¯R\bar{E}_{R}) relative to which the result obtained for a given charge density model is compared. As an exception, only the FB data are used for 209Bi because a SOG analysis has not been carried out for this nucleus. The proton energies evaluated with the uniform distribution, the 2pF function and the microscopic Skyrme-HF charge density are denoted as EUnifE_{\rm Unif}, E2pFE_{2pF} and EHFE_{HF}111EHFE_{HF} denotes the proton single-particle energies evaluated with the WS potential as a nuclear component and the charge density obtained from the microscopic Skyrme HF calculation as an input for the Coulomb potential. It should not be confused with the eigenvalues of the HF mean field., respectively. Our results for the proton energy differences including ΔERUnif=EUnifE¯R\Delta E^{\rm Unif}_{R}=E_{\rm Unif}-\bar{E}_{R}, ΔER2pF=E2pFE¯R\Delta E^{2pF}_{R}=E_{2pF}-\bar{E}_{R} and ΔERHF=EHFE¯R\Delta E^{HF}_{R}=E_{HF}-\bar{E}_{R} are given in Table 2.

We remark that the energy differences obtained in this subsection for all charge-density models are generally small in magnitude. For most cases, their magnitude is only of a few tens keV larger than those induced by the experimental uncertainties obtained in the previous subsection (ΔESOGFB\Delta E^{\rm FB}_{\rm SOG}). The values obtained for 32S are somewhat larger compared with the other nuclei and positive for all bound orbitals, because of large uncertainties in the model-independent data for charge density and an additional enhancement due to the weakly bound effect. It is also seen that all the selected charge density models provide a similar accuracy for proton single-particle energies. In particular, the microscopic Skyrme HF model for charge density works adequately regardless of the various deficiencies related to the isospin-symmetry breaking discussed in Ref. [12]. We notice that, in general the calculations using the uniform distribution or the 2pF function are strongly parameter-dependent. For example, if the Coulomb radius is parameterized as RC1.26fm×(A1)13R_{C}\approx 1.26~{}\text{fm}\times(A-1)^{\frac{1}{3}} as usual, considerably larger proton single-particle energy differences are obtained in many cases. Similar problem was also observed within the 2pF function. Therefore, it is necessary to constrain their parameters using the relevant experimental data if these distribution functions are selected as a charge-density model.

V.3 Test of Coulomb functionals

The present subsection is concerned with the compensation for the omission of Coulomb exchange term which is unavoidable when using the uniform distribution for charge densities. We are interested specifically in the method employed in Eq. (1) by excluding the last proton contribution to Coulomb direct term. In order to verify this traditional idea of self-interaction correction, it is instructive to perform an analytic analysis of its impact before discussing the numerical results.

Since the Coulomb direct term (9) is linear in charge density, it can be written as

Vdir[ρchZ]=Vdir[ρchZ1]+Vdir[ρch1]V_{dir}[\rho_{ch}^{Z}]=V_{dir}[\rho_{ch}^{Z-1}]+V_{dir}[\rho_{ch}^{1}] (30)

where ρchZ\rho_{ch}^{Z} is the total charge density which can be decomposed as

ρchZ=ρchZ1+ρch1\rho_{ch}^{Z}=\rho_{ch}^{Z-1}+\rho_{ch}^{1} (31)

with ρch1\rho_{ch}^{1} being the last proton contribution and ρchZ1\rho_{ch}^{Z-1} the contribution of the remaining Z1Z-1 protons. In order to gain a better insight into the functional structure of the Coulomb direct term, we further assume that ρchZ\rho_{ch}^{Z} and ρchZ1\rho_{ch}^{Z-1} have identical radial form and they differ from each other only by the normalization condition, namely ρchZ\rho_{ch}^{Z} is normalized to ZZ whereas ρchZ1\rho_{ch}^{Z-1} is normalized to Z1Z-1 (this is equivalent to the assumption that each proton has an equal contribution to the total charge density). Subsequently, the following relations can be derived

ρchZ1=(Z1Z)ρchZ=(Z1)ρch1.\rho_{ch}^{Z-1}=\left(\frac{Z-1}{Z}\right)\rho_{ch}^{Z}=(Z-1)\rho_{ch}^{1}. (32)

Because of the linearity, the corresponding Coulomb direct terms can be evaluated as

Vdir[ρchZ1]=(Z1Z)Vdir[ρchZ]=(Z1)Vdir[ρch1].V_{dir}[\rho_{ch}^{Z-1}]=\left(\frac{Z-1}{Z}\right)V_{dir}[\rho_{ch}^{Z}]=(Z-1)V_{dir}[\rho_{ch}^{1}]. (33)

Based on this assumption, the self-interaction correction discussed above can be done simply by replacing Vdir[ρchZ]V_{dir}[\rho_{ch}^{Z}] with Vdir[ρchZ]×(Z1)/ZV_{dir}[\rho_{ch}^{Z}]\times(Z-1)/Z. Note that the Coulomb potential in Eq. (1) fully satisfies these properties with ρchZ\rho_{ch}^{Z} being the uniform distribution. It is interesting to remark that the factor (Z1)/Z(Z-1)/Z goes to 1 as ZZ goes to infinity, indicating that this correction has a greatest effect in light nuclei. We found that this replacement leads to lower proton energy levels, namely by about 500 keV in 16O and 200 keV in 208Pb, relative to the levels obtained with the pure Coulomb direct term Vdir[ρchZ]V_{dir}[\rho_{ch}^{Z}].

Throughout this subsection, the Coulomb functional obtained in this way will be referred to as “traditional functional”. In order to make a numerical test this traditional approach, it is useful to decompose the proton energy difference for a given orbital as

ΔEFT=ΔEGT+ΔESG+ΔEFS,\Delta E^{T}_{F}=\Delta E^{T}_{G}+\Delta E^{G}_{S}+\Delta E^{S}_{F}, (34)

where ΔEFT\Delta E^{T}_{F} is the single-particle energy difference between the traditional functional and the exact Fock term. On the right-hand-side, ΔEGT\Delta E^{T}_{G}, ΔESG\Delta E^{G}_{S} and ΔEFS\Delta E^{S}_{F} are, respectively, the single-particle energy difference of the traditional treatment relative to GGA, GGA relative to Slater approximation, and Slater relative to the exact Fock term. Explicitly, ΔEFT=ETEF\Delta E^{T}_{F}=E^{T}-E^{F}, ΔEGT=ETEG\Delta E^{T}_{G}=E^{T}-E^{G}, ΔESG=EGES\Delta E^{G}_{S}=E^{G}-E^{S} and ΔEFS=ESEF\Delta E^{S}_{F}=E^{S}-E^{F}. As the uniform distribution is inappropriate for the evaluation of Coulomb exchange term, the 2pF function will be selected instead for this test.

We notice that no exact treatment of Coulomb exchange term is performed in the present work, the data of ΔEFS\Delta E^{S}_{F} for 16O, 40Ca, 48Ni and 208Pb are taken from Ref. [25]. Since the mass dependence of ΔEFS\Delta E^{S}_{F} is not so strong, it is reasonable to use the values obtained for 40Ca for 28Si, 32S and 48Ca which were not considered in Ref. [25] because they reside in the neighbor in the nuclear chart. For the same reason, we use the ΔEFS\Delta E^{S}_{F} values obtained for 208Pb for the remaining nuclei. One may argue here that the proton energy differences may depend on the nuclear component as well as the method for solving the Schrödinger equation, or they may vary significantly when transferring from a self-consistent to a phenomenological mean field. We have checked such dependence by looking at the term ΔESG\Delta E^{G}_{S}, we found that its values obtained for 208Pb using WS potential are in the range between -1 and 19 keV, which are in remarkably good agreement with those calculated within the self-consistent Skyrme HF method.

Our numerical results for ΔEFT\Delta E^{T}_{F} as well as for ΔEGT\Delta E^{T}_{G} and ΔESG\Delta E^{G}_{S} are given in Table 3. It is seen that in light nuclei (around Z=8Z=8) negative ΔEGT\Delta E^{T}_{G} values are obtained because the traditional self-interaction correction is stronger than the GGA exchange term. After Z=8Z=8, ΔEGT\Delta E^{T}_{G} increases gradually with the atomic number and finally reaches a saturation at around Z=80Z=80. The saturated value of ΔEGT\Delta E^{T}_{G} is about 300 keV. We notice also that the functional-driven energy differences are insensitive to the weakly bound effect, because the values obtained for 32S do not differ significantly from those of the neighboring nuclei as observed in the two previous subsections. Combining these values with those of ΔEFS\Delta E^{S}_{F} from the above-cited self-consistent calculations, we obtained that the total energy differences between the traditional correction and the exact treatment are in the range between 130-130 and 620 keV for light nuclei, and 136 and 800 keV for heavier nuclei. A larger value is expected if those induced by the charge density are added together. Therefore, the expression Eq. (1) for Coulomb potential should not be applied for a high precision calculation such as the shell-model description of the isospin-symmetry breaking.

VI Conclusion

We have performed a comparative study of Coulomb potential for various charge-density models, including the uniform distribution, the 2pF function, and the microscopic Skyrme HF calculation. The phenomenological WS potential was selected as a nuclear component of the mean-field Hamiltonian. We found that the differences between the proton single-particle energies produced with these charge-density models and those yielded by the model-independent data are relatively small. Generally, these differences are less than 100 keV in magnitude, except for 32S because of high sensitivity to small variations in the potential occurred when the last occupied state is fully filled and has no centrifugal barrier. Although the proton single-particle energies obtained in this work using the uniform distribution and the 2pF function are remarkably accurate, in general this observable is strongly parameter-dependent. Thus, it is highly recommended to constraint the charge-density model parameters case-by-case using the available experimental data on charge radii as illustrated in the present paper. Otherwise the microscopic Skyrme-HF model should be employed instead.

In addition, we have also studied the traditional approach to compensate for the omission of the exchange term by excluding the last proton contribution. This approach is commonly employed when the uniform distribution is selected as a charge-density model. We found that such compensation can be implemented by simply multiplying the Coulomb direct term with the factor (Z1)/Z(Z-1)/Z. Therefore, it can be regarded as a variation in Coulomb functional. Combining our numerical result with that of Ref. [25], the proton energy levels obtained with this traditional functional are underbound by 100 to 800 keV for nuclei with Z28Z\geq 28, relative to the those obtained with an exact treatment. The opposite pattern trends to appear in the lighter ZZ region where the factor (Z1)/Z(Z-1)/Z reduces significantly from unity, for example the 1d5/21d_{5/2} level in 16O obtained with the traditional functional is overbound by 254 keV. It is also seen from our results that the functional-driven energy differences are rather insensitive to the weakly-bound effect since the values obtained for 32S where this effect is expected to be strongest do not alter significantly from those of its neighboring nuclei.

As a final conclusion, the use of Eq. (1) for a precision calculation is at risk. Instead of this traditional formula, it is desired to evaluate the Coulomb potential using realistic charge density models such as the 2pF function or Skyrme HF, as well as more fundamental approaches for Coulomb-exchange functional such as the Slater or the GGA.

Acknowledgements.
This work is supported by the National Research Foundation of Korea(NRF) grant funded by the Korea government(MSIT)(No. 2021R1A2C2094378).

References

*

Appendix A Numerical results

Table 1: Differences (ΔESOGFB=EFBESOG\Delta E^{\rm FB}_{\rm SOG}=E_{\rm FB}-E_{\rm SOG}) in proton single-particle energies induced by the differences between the FB and SOG charge-density data. The calculation without Coulomb exchange term is labeled with ‘Direct’, whereas those performed using the GGA and the Slater approximation are labeled with ‘GGA’ and ‘Slater’, respectively. The listed numbers are in keV unit.
Orbital Direct Slater GGA Direct Slater GGA Direct Slater GGA Direct Slater GGA Direct Slater GGA
16O 28Si 32S 40Ca 48Ca
2p3/22p_{3/2} -78 -80 -71
1f7/21f_{7/2} -59 -58 -53 -74 -73 -66
1d3/21d_{3/2} -25 -26 -20 -452 -452 -452 -75 -72 -70 -94 -91 -86
2s1/22s_{1/2} -28 -28 -20 -462 -462 -462 -82 -81 -76 -102 -100 -95
1d5/21d_{5/2} 2 3 0 -28 -27 -23 -460 -460 -460 -75 -73 -70 -92 -88 -84
1p1/21p_{1/2} 4 3 3 -36 -34 -33 -514 -514 -514 -96 -91 -91 -115 -110 -108
1p3/21p_{3/2} 4 4 3 -36 -34 -32 -511 -511 -511 -94 -90 -89 -112 -107 -105
1s1/21s_{1/2} 8 7 6 -46 -43 -42 -570 -570 -570 -118 -112 -111 -138 -130 -130
54Ni 205Tl 206Pb 208Pb 209Bi
2f5/22f_{5/2} -13 -14 -14 -16 -18 -16 -18 -19
2f7/22f_{7/2} -13 -14 -13 -17 -19 -19 -18 -19 -19
1h9/21h_{9/2} -12 -13 -12 -16 -17 -18 -16 -17 -17
3s1/23s_{1/2} -17 -17 -18 -21 -22 -23 -21 -22 -22
2d3/22d_{3/2} -16 -16 -16 -20 -21 -21 -20 -21 -21
1g7/21g_{7/2} -13 -14 -13 -18 -19 -19 -18 -18 -19
2d5/22d_{5/2} -16 -15 -16 -20 -21 -21 -20 -21 -21
1h11/21h_{11/2} -11 -11 -12 -15 -16 -17 -16 -16 -16
1g9/21g_{9/2} -12 -12 -13 -17 -17 -18 -17 -18 -18
1f5/21f_{5/2} -14 -14 -15 -20 -20 -21 -20 -20 -21
2p1/22p_{1/2} 29 30 27 -18 -18 -19 -23 -23 -24 -22 -23 -24
2p3/22p_{3/2} 31 32 29 -18 -18 -18 -22 -24 -24 -23 -24 -23
1f7/21f_{7/2} 30 29 27 -13 -14 -14 -19 -19 -27 -23 -19 -19
1d3/21d_{3/2} 37 35 34 -16 -16 -16 -22 -22 -23 -22 -22 -22
2s1/22s_{1/2} 40 39 38 -20 -20 -20 -25 -26 -27 -25 -25 -27
1d5/21d_{5/2} 36 34 34 -16 -15 -16 -21 -21 -22 -21 -21 -21
1p1/21p_{1/2} 44 42 42 -17 -18 -18 -24 -25 -25 -23 -24 -24
1p3/21p_{3/2} 43 41 41 -18 -17 -17 -23 -24 -24 -23 -23 -24
1s1/21s_{1/2} 52 50 49 -20 -20 -20 -26 -27 -27 -27 -26 -27
Table 2: Proton single-particle energy differences induced by the deviations of the charge-density models relative to the model-independent data. Where ΔERUnif\Delta E^{\rm Unif}_{R}, ΔER2pF\Delta E^{2pF}_{R} and ΔERHF\Delta E^{HF}_{R} correspond to the uniform distribution, the 2pF function and the Skyrme-HF calculation of charge density, respectively (see subsection V.2 for detailed description). The Coulomb exchange potential is excluded for these calculations. The listed numbers are in keV unit.
Orbital ΔERUnif\Delta E^{\rm Unif}_{R} ΔER2pF\Delta E^{2pF}_{R} ΔERHF\Delta E^{HF}_{R} ΔERUnif\Delta E^{\rm Unif}_{R} ΔER2pF\Delta E^{2pF}_{R} ΔERHF\Delta E^{HF}_{R} ΔERUnif\Delta E^{\rm Unif}_{R} ΔER2pF\Delta E^{2pF}_{R} ΔERHF\Delta E^{HF}_{R} ΔERUnif\Delta E^{\rm Unif}_{R} ΔER2pF\Delta E^{2pF}_{R} ΔERHF\Delta E^{HF}_{R} ΔERUnif\Delta E^{\rm Unif}_{R} ΔER2pF\Delta E^{2pF}_{R} ΔERHF\Delta E^{HF}_{R}
16O 28Si 32S 40Ca 48Ca
2p3/22p_{3/2} -4 7 11
1f7/21f_{7/2} 73 6 46 113 13 13
1d3/21d_{3/2} 29 -18 23 667 631 693 35 8 47 85 24 5
2s1/22s_{1/2} -69 -34 3 540 599 694 -72 -17 48 -6 19 8
1d5/21d_{5/2} -1 -22 -10 33 -16 24 689 651 712 42 8 47 93 22 5
1p1/21p_{1/2} -40 -21 -30 -38 -22 13 682 726 804 -41 3 47 34 34 -5
1p3/21p_{3/2} -42 -20 -32 -32 -21 15 689 725 801 -31 4 47 43 32 -3
1s1/21s_{1/2} -144 -17 -60 -159 -37 -8 618 783 894 -158 -12 48 -50 40 -10
54Ni 205Tl 206Pb 208Pb 209Bi
2f5/22f_{5/2} -70 -48 -15 -37 -72 -9 -55 -42 20
2f7/22f_{7/2} -78 -50 -18 -62 -46 2 -46 -74 -11 -64 -44 18
1h9/21h_{9/2} 41 -46 -8 63 -39 3 84 -59 -9 58 -38 15
3s1/23s_{1/2} -122 -31 -33 -120 -40 5 -109 -74 -9 -116 -33 30
2d3/22d_{3/2} -83 -42 -23 -69 -42 1 -56 -72 -12 -70 -38 20
1g7/21g_{7/2} 20 -47 -15 41 -39 -2 59 -63 -15 35 -40 10
2d5/22d_{5/2} -91 -44 -25 -78 -41 -1 -63 -73 -13 -78 -39 20
1h11/21h_{11/2} 59 -45 -2 81 -38 8 105 -56 -3 76 -38 18
1g9/21g_{9/2} 41 -46 -9 63 -39 3 84 -60 -10 57 -39 14
1f5/21f_{5/2} -10 -47 -23 10 -41 -7 26 -68 -21 5 -40 9
2p1/22p_{1/2} -36 -8 37 -104 -34 -32 -94 -36 0 -84 -73 -16 -92 -31 24
2p3/22p_{3/2} -42 -6 36 -108 -35 -33 -99 -37 -1 -88 -73 -15 -97 -32 24
1f7/21f_{7/2} 90 5 49 14 -46 -16 38 -35 2 53 -64 -17 29 -40 10
1d3/21d_{3/2} 45 9 47 -48 -48 -30 -28 -41 -11 -16 -73 -27 -34 -42 6
2s1/22s_{1/2} -52 3 40 -130 -24 -43 -128 -32 0 -122 -73 -17 -124 -25 29
1d5/21d_{5/2} 57 9 46 -26 -47 -26 -6 -41 -8 9 -70 -23 -11 -40 8
1p1/21p_{1/2} -23 11 41 -97 -47 -40 -79 -41 -14 -70 -79 -31 -82 -40 7
1p3/21p_{3/2} -9 11 43 -79 -46 -35 -62 -41 -12 -52 -77 -29 -66 -40 7
1s1/21s_{1/2} -120 11 35 -152 -40 -49 -141 -39 -14 -136 -84 -32 -141 -37 11
Table 3: Numerical results for ΔEGT\Delta E^{T}_{G}, ΔESG\Delta E^{G}_{S} and ΔEFT\Delta E^{T}_{F} (see subsection V.3 for detailed description). The total energy differences, ΔEFT\Delta E^{T}_{F} is deduced via Eq. (34) using the data for ΔEFS\Delta E^{S}_{F} taken from Ref. [25]. All these calculations employ the 2pF function for the charge density. Units are in keV

. Orbital ΔEGT\Delta E^{T}_{G} ΔESG\Delta E^{G}_{S} ΔEFT\Delta E^{T}_{F} ΔEGT\Delta E^{T}_{G} ΔESG\Delta E^{G}_{S} ΔEFT\Delta E^{T}_{F} ΔEGT\Delta E^{T}_{G} ΔESG\Delta E^{G}_{S} ΔEFT\Delta E^{T}_{F} ΔEGT\Delta E^{T}_{G} ΔESG\Delta E^{G}_{S} ΔEFT\Delta E^{T}_{F} ΔEGT\Delta E^{T}_{G} ΔESG\Delta E^{G}_{S} ΔEFT\Delta E^{T}_{F} 16O 28Si 32S 40Ca 48Ca 2p3/22p_{3/2} 22 14 -128 1f7/21f_{7/2} 87 -13 -84 95 -14 -77 1d3/21d_{3/2} 34 -11 225 70 -16 256 119 -20 301 126 -23 305 2s1/22s_{1/2} -14 8 188 20 4 218 71 -3 262 79 -5 268 1d5/21d_{5/2} -82 2 -254 52 -16 242 83 -20 269 125 -22 309 126 -23 309 1p1/21p_{1/2} -35 -17 178 81 -24 432 106 -25 456 141 -23 493 141 -23 493 1p3/21p_{3/2} -26 -21 193 84 -25 437 107 -25 460 141 -24 495 142 -25 495 1s1/21s_{1/2} -5 -31 395 86 -23 561 105 -21 582 135 -17 616 137 -18 617 48Ni 205Tl 206Pb 208Pb 209Bi 1h9/21h_{9/2} 278 -18 136 280 -18 138 280 -19 137 281 -18 139 3s1/23s_{1/2} 228 -1 366 231 -1 369 231 -2 368 232 -1 370 2d3/22d_{3/2} 247 -7 382 248 -6 384 249 -7 384 250 -7 385 1g7/21g_{7/2} 289 -18 427 290 -18 428 290 -18 428 291 -18 429 2d5/22d_{5/2} 244 -5 395 245 -5 396 246 -6 396 247 -6 397 1h11/21h_{11/2} 271 -19 393 273 -19 395 273 -19 395 275 -20 396 1g9/21g_{9/2} 283 -20 530 284 -19 532 284 -19 532 286 -20 533 1f5/21f_{5/2} 294 -17 572 295 -16 574 294 -16 573 296 -16 575 2p1/22p_{1/2} 263 -11 528 265 -10 531 265 -11 530 267 -11 532 2p3/22p_{3/2} 74 12 228 261 -10 537 262 -9 539 262 -10 538 264 -10 540 1f7/21f_{7/2} 148 -14 309 290 -18 625 291 -18 626 291 -18 626 292 -18 627 1d3/21d_{3/2} 177 -20 355 294 -14 675 296 -14 677 295 -14 676 296 -14 677 2s1/22s_{1/2} 133 -6 330 272 -11 643 273 -11 644 272 -11 643 274 -11 645 1d5/21d_{5/2} 176 -21 479 293 -15 699 295 -16 700 294 -15 700 296 -16 701 1p1/21p_{1/2} 189 -20 554 291 -10 750 293 -10 752 292 -10 751 294 -10 753 1p3/21p_{3/2} 188 -21 594 292 -12 757 294 -12 759 292 -11 758 293 -11 759 1s1/21s_{1/2} 183 -14 676 285 -8 797 287 -8 799 286 -7 799 287 -7 800