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

Quasi-elastic electron scattering with KIDS nuclear energy density functional

Hana Gil Center for Extreme Nuclear Matter, Korea University, Seoul 02841, Korea    Chang Ho Hyun [email protected] Department of Physics Education, Daegu University, Gyeongsan 38453, Korea    Kyungsik Kim [email protected] School of Liberal Arts and Science, Korea Aerospace University, Goyang 10540, Korea
Abstract

Isoscalar and isovector effective masses of the nucleon in nuclear medium are explored in the quasi-elastic electron scattering off nuclei with KIDS (Korea-IBS-Daegu-SKKU) density functional model. Effective masses are varied in the range (0.71.0)M(0.7-1.0)M where MM is the mass of the nucleon in free space. Parameters in the KIDS functional are adjusted to nuclear matter equation of state, energy and radius of selected nuclei, and effective mass of nucleons. Hartree-Fock equation is solved to obtain the wave functions of the nucleon in target nuclei, and they are plugged in the calculation of electron-nucleus scattering cross sections at the energies of incident electrons 300 MeV – 2.5 GeV. Theoretical prediction agrees well with measurement. Dependence on the effective mass is evident: cross section tends to increase with small isoscalar effective masses. However, effect of isovector effective mass is negligible. Spectroscopic factors are estimated for the protons in the outermost shells of 16O, 40Ca, and 208Pb. Results are consistent with the values in the literature.

I Introduction

Quasi-elastic electron-nucleus scatterings are acknowledged as a useful tool for studying the structure of nuclei and the change of properties of nucleon in nuclear medium. In the conventional approaches, theoretical models are calibrated to accurate data of nuclei, so the basic properties of nuclei such as binding energy and charge radii agree to data with differences less than 1% in most models. However, model dependence becomes manifest in the structural details. Representative example is the single particle level: SLy4 sly4 and UNEDF unedf0 models which are well-known non-relativistic models show discriminate results of the single particle levels in light and heavy nuclei. Since the protons in specific levels contribute to the scattering with electrons, model dependence could have effect and appear in the result of cross section.

Effective mass of the nucleon in nuclear medium has long been a controversial issue in the physics of nuclear structure and dense nuclear matter. It is related to diverse phenomena such as the restoration of chiral symmetry at high densities, density distribution in nuclei, and single particle energy levels. Even though many experimental and theoretical efforts have been accumulated, value of the effective mass is not constrained precisely: generally accepted range is (0.71.0)M(0.7-1.0)M where MM is the nucleon mass in free space.

Exclusive (e,ep)(e,e^{\prime}p) reaction furnishes a tool useful for studying single particle properties of target nucleus. It is well known that the process is sensitive to individual orbits and energy levels. The response functions that give valuable information about the nucleons inside a nucleus can be extracted from a given orbit with respect to transfers of momenum and energy by comparing with experimental data. A scale factor called spectroscopic factor (SF) should be determined when theoretical results are compared with data. Spectroscopic factor contains information about the probability to occupy a given orbit, so the SF takes a value between 0 and 1 in a nuclear model. Since the SF represents the structural characteristic of a nuclei, it is useful to estimate the value of the SF for testing a nuclear model.

In this work, we investigate the sensitivity of effective mass to the structure and scattering of nuclei with electrons. In the KIDS (Korea-IBS-Daegu-SKKU) density functional model, one can fix effective masses to specific values without altering the nuclear matter equation of state and static properties of nuclei kidsnuclei1 . Effective masses are considered in the range (0.71.0)M(0.7-1.0)M. After model parameters are determined from the nuclear matter properties, nuclear data, and assumed effective masses, non-relativistic wave equation is solved, and the resulting wave function is transformed to a form adaptable in the relativistic formalism. Scattering cross sections with protons in the outermost shells are calculated and compared with data. To include the final state interaction, the wave functions of the continuum nucleon are obtained from the relativistic optical model clark for a knocked-out proton solving the Dirac equation. The incident and outgoing electron Coulomb distortions are treated by the same method as the Ohio group kim .

In the result, we find that density distribution around surface shows rare dependence on the effective mass. However, the dependence becomes evident in the core of heavy nuclei. Single particle levels also depend on the effective mass. In the light nuclei, small effective mass gives better agreement to data, but for 208Pb, data are reproduced well with effective mass close to the free space mass. Cross section is in good agreement with experiment. Dependence on the effective mass is manifest: smaller effective mass tends to give larger cross sections. Spectroscopic factors are extracted by adjusting the cross sections calculated from theory to data. Results are comparable with values of the literature.

Remaining part of the manuscript presents the following contents. In Section II, we introduce the basic formalism. Results of the density distributions, single particle levels, and the cross sections are displayed and discussed in Section III. In Section IV, we summarize the work.

II Formalism

II.1 Relativistic wave function from a non-relativisitc nuclear model

Electrons in the process we consider are highly relativistic, so the available code is constructed in the relativistic formalism in which both electrons and nucleons are treated relativistically. Many and well-known nuclear structure models such as Skyrme force models and Gogny force models are formulated on the non-relativistic ground. KIDS model is also based on the non-relativistic phenomenology, so it is necessary to bridge the two approaches.

In the relativistic formalism, single particle wave function with angular momentum 𝐉2\mathbf{J}^{2} and JzJ_{z}, and in good quantum state of parity and time reversal takes the form

Ψ(𝐫)=(f(r)χκμ(𝐫^)ig(r)χκμ(𝐫^)).\displaystyle\Psi(\mathbf{r})=\left(\begin{array}[]{c}f(r)\chi^{\mu}_{\kappa}(\hat{\mathbf{r}})\\ ig(r)\chi^{\mu}_{-\kappa}(\hat{\mathbf{r}})\end{array}\right). (3)

Orbital and spin states are represented by

χκμ(𝐫^)=m,slm12s|jμYlm(𝐫^)χs,\chi^{\mu}_{\kappa}(\hat{\mathbf{r}})=\sum_{m,s}\left<lm\left.\frac{1}{2}s\right|j\mu\right>Y_{lm}(\hat{\mathbf{r}})\chi_{s}, (4)

where χs\chi_{s} and YlmY_{lm} are the Pauli spinor and the spherical harmonics, respectively. κ\kappa is the eigenvalue of the operator (σ𝐋+1)(\mathbf{\sigma}\cdot\mathbf{L}+1) given by

κ=(l;j=l12,l1;j=l+12.\kappa=\left(\begin{array}[]{c l}l&;\,j=l-\frac{1}{2},\\ -l-1&;\,j=l+\frac{1}{2}.\end{array}\right. (5)

Solving the non-relativistic wave equation

d2Fdr2+2rdFdrκ(κ+1)r2F2M[Vcen(r)+VSO(r)(κ1)]F+p2F=0,\frac{d^{2}F}{dr^{2}}+\frac{2}{r}\frac{dF}{dr}-\frac{\kappa(\kappa+1)}{r^{2}}F-2M\left[V_{\rm cen}(r)+V_{\rm SO}(r)(-\kappa-1)\right]F+p^{2}F=0, (6)

one can determine F(r)F(r). Radial functions f(r)f(r) and g(r)g(r) can be calculated from the relation

f(r)=D1/2(r)F(r),g(r)=D1/2(r)G(r),f(r)=D^{1/2}(r)F(r),\,\,\,g(r)=D^{-1/2}(r)G(r), (7)

where Darwin factor D(r)D(r) is defined as

D(r)=exp[2r𝑑rMrVSO(r)].D(r)=\exp\left[-2\int^{\infty}_{r}drMrV_{\rm SO}(r)\right]. (8)

Lower component function G(r)G(r) is obtained from the relation with F(r)F(r)

G(r)=1E+M[dFdr+(κ+1rT(r))F],G(r)=\frac{1}{E+M}\left[\frac{dF}{dr}+\left(\frac{\kappa+1}{r}-T(r)\right)F\right], (9)

where

T(r)=MrVSO(r).T(r)=MrV_{\rm SO}(r). (10)

With non-relativistic nuclear potentials VcenV_{\rm cen} and VSOV_{\rm SO} given from a model, Hartree-Fock equation is solved and F(r)F(r) is obtained. G(r)G(r) is calculated from Eq. (9), and finally the relativistic wave functions f(r)f(r) and g(r)g(r) are obtained by calculating Eq. (7).

II.2 KIDS model

One purpose of the electron scattering is to get better knowledge about the structure of nuclei. At the energy of quasi-elastic scattering, the distribution of nucleons in the interior and the gradient of density around the surface could have effects on the cross section. In terms of the Skyrme force, derivative terms account for the contribution of density gradient. These terms also contribute to determining the effective mass of the nucleon, so the structural uncertainty around the surface as well as the core could be explored by probing the dependence on the effective mass.

Role of the effective mass could be singled out when other conditions (e.g. binding energy, charge radius) are unchanged. Independent control of the nuclear properties is accessible with the KIDS energy density functional kidsnm . In the KIDS framework several rules are assumed to develop a nuclear model. At first energy per particle in homogeneous nuclear matter is expanded in powers of Fermi momentum (equally ρ1/3\rho^{1/3}) as

(ρ,δ)=𝒯+i=0(αi+βiδ2)ρ1+i/3.\displaystyle{\cal E}(\rho,\,\delta)={\cal T}+\sum_{i=0}(\alpha_{i}+\beta_{i}\delta^{2})\rho^{1+i/3}. (11)

𝒯{\cal T} is the kinetic energy, and δ=(ρnρp)/ρ\delta=(\rho_{n}-\rho_{p})/\rho. Model parameters αi\alpha_{i} and βi\beta_{i} are fixed or fit to symmetric and asymmetric nuclear matter properties. In this work we fix α0\alpha_{0}, α1\alpha_{1} and α2\alpha_{2} to three saturation properties: saturation density ρ0=0.16\rho_{0}=0.16 fm-3, bindind energy per particle EB=16.0E_{B}=16.0 MeV, and incompressibility K0=240K_{0}=240 MeV kidsnuclei2 . Several ways have been tried to fix the βi\beta_{i} values. It has been proved that four βi\beta_{i}’s are good and enough to reproduce the neutron matter equation of state obtained from microscopic calculations kidsnuclei1 , static properties of neutron rich nuclei kidsnd , and neutron star observations kidsk0 ; kidssymene . In this work we use the values of βi\beta_{i} fit to the pure neutron matter of equation of state calculated by Akmal, Pandharipande and Ravenhall apr . The model thus fixed is labeled KIDS0.

   KIDS0 kidsnuclei1    KIDS0-m*99    KIDS-m*77    SLy4 sly4
t0t_{0} 1772.04-1772.04 1772.04-1772.04 1772.04-1772.04 2488.91-2488.91
y0y_{0} 127.52-127.52 127.52-127.52 127.52-127.52 2075.75-2075.75
t1t_{1} 275.72275.72 318.92318.92 441.99441.99 486.82486.82
y1y_{1} 0 361.17-361.17 109.03-109.03 167.37-167.37
t2t_{2} 161.50-161.50 26.8226.82 295.06-295.06 546.39-546.39
y2y_{2} 0 215.11-215.11 259.50259.50 546.39546.39
t31t_{31} 12216.7312216.73 12216.7312216.73 12216.7312216.73 13777.0013777.00
y31y_{31} 11969.99-11969.99 11969.99-11969.99 11969.99-11969.99 18654.0618654.06
t32t_{32} 571.07571.07 191.34-191.34 2572.65-2572.65 -
y32y_{32} 29485.4929485.49 34304.5734304.57 37593.4037593.40 -
t33t_{33} 0 0 0 -
y33y_{33} 22955-22955 22955-22955 22955-22955 -
W0W_{0} 108.35108.35 129.96129.96 115.28115.28 123.0
μS\mu_{S} 0.99 0.90 0.70 0.70
μV\mu_{V} 0.81 0.90 0.70 0.80
Table 1: Parameters in the single particle potential Eqs. (13, 14). We use the simplified notation yi=tixiy_{i}=t_{i}x_{i}. Units of the parameters are MeV\cdotfm3 for t0t_{0}, y0y_{0}, MeV\cdotfm4 for t31t_{31}, y31y_{31}, MeV\cdotfm5 for t1t_{1}, y1y_{1}, t2t_{2}, y2y_{2}, t32t_{32}, y32y_{32}, W0W_{0}, and MeV\cdotfm6 for t33t_{33}, y33y_{33}. Isoscalar and isovector effective masses in each model are shown in the last two rows.

When nuclei are described in the KIDS framework, energy density functional is transformed to the form of Skyrme force kidsnpsm2017 . Terms accounting for the density gradient and spin-orbit interactions are added In the notation of Skyrme force, they correspond to parameters t1t_{1}, t2t_{2}, x1x_{1}, x2x_{2} and W0W_{0}. We assume t1=t2t_{1}=t_{2}, x1=x2=0x_{1}=x_{2}=0, and t1t_{1} and W0W_{0} are fit to the data of binding energy and charge radius of 40Ca, 48Ca and 208Pb. In the KIDS0 model, x1x_{1} and x2x_{2} are assumed to be 0, so the effective masses are obtained as results of determining t1t_{1}. By adjusting x1x_{1} and x2x_{2}, one can produce specific values of isoscalar and isovector effective masses μS\mu_{S} and μV\mu_{V} defined by

μS\displaystyle\mu_{S} =\displaystyle= mSM=[1+M82ρ(3t1+5t2+4y2)]1,\displaystyle\frac{m^{*}_{S}}{M}=\left[1+\frac{M}{8\hbar^{2}}\rho(3t_{1}+5t_{2}+4y_{2})\right]^{-1},
μV\displaystyle\mu_{V} =\displaystyle= mVM=[1+M42ρ(2t1+2t2+y1+y2)]1,\displaystyle\frac{m^{*}_{V}}{M}=\left[1+\frac{M}{4\hbar^{2}}\rho(2t_{1}+2t_{2}+y_{1}+y_{2})\right]^{-1}, (12)

where yi=tixiy_{i}=t_{i}x_{i}. In this work we consider two cases (μS,μV)=(0.7,0.7)(\mu_{S},\,\mu_{V})=(0.7,0.7) and (0.9,0.9)(0.9,0.9). Each model is labeled KIDS0-m*77 and KIDS0-m*99, respectively.

Central and spin-orbit potentials entering Eq. (6) are obtained from KIDS functional as

Vcen\displaystyle V_{\text{cen}} =\displaystyle= Uq+Ucoul,\displaystyle U_{q}+U_{\text{coul}},
Uq\displaystyle U_{q} =\displaystyle= (t0+12y0)ρ(12t0+y0)ρq\displaystyle\left(t_{0}+\frac{1}{2}y_{0}\right)\rho-\left(\frac{1}{2}t_{0}+y_{0}\right)\rho_{q} (13)
+\displaystyle+ 18[(2t1+y1)+(2t2+y2)]τ18[(t1+2y1)+(t2+2y2)]τq\displaystyle\frac{1}{8}[(2t_{1}+y_{1})+(2t_{2}+y_{2})]\tau-\frac{1}{8}[(t_{1}+2y_{1})+(t_{2}+2y_{2})]\tau_{q}
\displaystyle- 116[(6t1+y1)(2t2+y2)]2ρ+116[(3t1+6y1)+(t2+2y2)]2ρq\displaystyle\frac{1}{16}[(6t_{1}+y_{1})-(2t_{2}+y_{2})]\nabla^{2}\rho+\frac{1}{16}[(3t_{1}+6y_{1})+(t_{2}+2y_{2})]\nabla^{2}\rho_{q}
+\displaystyle+ 112k=13ρk/3[(2+k3)(t3k+12y3k)ρ(t3k+2y3k)ρqk3(12t3k+y3k)ρp2+ρn2ρ2]\displaystyle\frac{1}{12}\sum_{k=1}^{3}\rho^{k/3}\left[\left(2+\frac{k}{3}\right)\left(t_{3k}+\frac{1}{2}y_{3k}\right)\rho-(t_{3k}+2y_{3k})\rho_{q}-\frac{k}{3}\left(\frac{1}{2}t_{3k}+y_{3k}\right)\frac{\rho^{2}_{p}+\rho^{2}_{n}}{\rho^{2}}\right]
\displaystyle- W02(r+2r)(J+Jq),\displaystyle\frac{W_{0}}{2}\left(\frac{\partial}{\partial r}+\frac{2}{r}\right)(J+J_{q}),
VSO\displaystyle V_{\text{SO}} =\displaystyle= 12W0(r+2r)(ρ+ρq)+18(t1t2)Jq18(y1+y2)J,\displaystyle\frac{1}{2}W_{0}\left(\frac{\partial}{\partial r}+\frac{2}{r}\right)(\rho+\rho_{q})+\frac{1}{8}(t_{1}-t_{2})J_{q}-\frac{1}{8}(y_{1}+y_{2})J, (14)

where τ\tau is the non-relativistic kinetic energy, JqJ_{q} denotes the spin current of the neutron and the proton, and J=Jn+JpJ=J_{n}+J_{p}. Parameters in the potential of each model are summarized in Tab. 1. We consider the SLy4 model for a comparison with a standard Skyrme force model. With these potentials, the transformed relativistic wave functions are generated and applied into the exclusive (e,ep)(e,e^{\prime}p) reaction in quasi-elastic region.

III Result and Discussion

III.1 Density distribution

Charge and neutron distributions in 16O, 40Ca and 208Pb are displayed in Figs. 1, 2, and 3, respectively. Measured values are denoted with gray bands.

Refer to caption
Figure 1: Density profile of charge (left) and neutron (right) for 16O.

In the result of 16O, KIDS models show weak dependence on the effective mass, and the results of models agree to each other in both charge and neutron distributions. SLy4 model agrees well with KIDS model at r>2r>2 fm, but in the interior region (r<2r<2 fm) densities are slightly suppressed compared to the KIDS model. All the models reproduce the data of charge distribution o16density well over r>r> 2.5 fm. Visual discrepancy is found in 0.5<r<2.50.5<r<2.5 fm, but the difference from experiment is less than 10%. Data for the neutron are not available so only the theory results are presented.

Refer to caption
Figure 2: Density profile of charge (left) and neutron (right) for 40Ca.

For 40Ca, model dependence is weak again, and the four models predict similar distributions of both protons and neutrons. In the comparison with experiment sick1981 ; ray1979 , distribution of the proton is reproduced well by the theory. In case of the neutron, theoretical results are within the errors of experiment in the core region (r<2r<2 fm), but on the surface (2<r<42<r<4 fm) where the density drops rapidly, theory exceeds experiment.

Refer to caption
Figure 3: Density profile of charge (left) and neutron (right) for 208Pb.

We have seen that the distribution of the proton and the neutron are insensitive to the effective mass in the light nuclei 16O and 40Ca. In the result of 208Pb, on the other hand, the effect of effective mass appears clear. For the proton, dependence on the effective mass is negligible at r>1r>1 fm. In this region model dependence is weak and theory agrees well with experiment pb208density . At r<1r<1 fm, theories are divided into two groups: one with KIDS0, KIDS0-m*99 (Group1), and the other with KIDS0-m*77, SLy4 (Group2). Group2 shows charge density large than Group1 as r0r\rightarrow 0. Models in each group have similar isoscalar effective mass, μS1.0\mu_{S}\sim 1.0 for Group1 and μS=0.7\mu_{S}=0.7 for Group2. Therefore different behavior of charge density at r<1r<1 fm could be originated from the value of isoscalar effective mass.

Neutron distribution shows pattern of agreement and disagreement to experiment saito2007 similar to the neutron distribution of 40Ca. Dependence on the effective mass is divided into three categories: no dependence at r>2.5r>2.5 fm, models are classified into two sets Group1 and Group2 at 1<r<2.51<r<2.5 fm, and four models behave independently at r<1r<1 fm. In the inner core r<2r<2 fm, theory results are within the range of experimental uncertainty. In the outer core region 2<r<52<r<5 fm, all the models obtain neutron densities less than experiment. The discrepancy does not exceed 10%. In the surface region 5<r<85<r<8 fm, theory overwhelms experiment slightly, and it is reversed in the tail r>8r>8 fm. There seems to be a pattern in the discrepancy of the neutron distribution. On the other hand, distribution of the proton which is relevant to the scattering with electrons agrees well with experiment.

III.2 Single particel levels

16O 40Ca 208Pb
1p1/2 1p3/2 2s1/2 1d3/2 3s1/2 2d3/2
Exp. volya2007 12.13-12.13 18.40-18.40 10.92-10.92 8.33-8.33 8.01-8.01 8.36-8.36
KIDS0 9.67-9.67 14.76-14.76 8.90-8.90 7.26-7.26 8.21-8.21 8.96-8.96
KIDS0-m*99 9.42-9.42 15.42-15.42 9.10-9.10 6.95-6.95 8.47-8.47 9.24-9.24
KIDS0-m*77 10.99-10.99 16.53-16.53 9.81-9.81 8.49-8.49 8.64-8.64 9.64-9.64
SLy4 10.53-10.53 15.99-15.99 9.82-9.82 8.08-8.08 8.79-8.79 9.57-9.57
Table 2: Single particle levels in MeV.

Electron-nucleus scattering data provide cross sections from protons at specific states. Table 2 collects the single particle levels of the proton for which cross sections calculated from theory will be compared with experiment. For light nuclei 16O and 40Ca, models with smaller isoscalar effective mass (Group2) obtain results closer to experiment than the models in Group1. For 16O, Group2 models differ from experiment by 9.4–13.2%, and Group1 models by 16.6–18.5%. For 40Ca, Group1 models give deviations from experiment 16.6–18.5% and 12.8–16.6% in the 2s1/2 and 1d3/2 states, respectively. With the models in Group2, we have 10.1–10.2% and 2–3% for 2s1/2 and 1d3/2 states, respectively.

For 208Pb Group1 models show better agreement to data. In the 3s1/2 state Group1 models give differences 2.5–5.7% and Group2 models 7.9–9.7%. In the 2d3/2 state, Group1 and Group2 give differences 7.2–10.5% and 14.5–15.3%, respectively. A similar pattern is reported in the UNEDF model unedf0 . Isoscalar effective mass is 0.9M0.9M in the UNEDF model, and the model shows agreement better for heavy nuclei than light ones.

As far as single particle levels are concerned, light nuclei favor small isoscalar effective mass, but heavy nuclei support effective masses close to 1.

III.3 Cross section

Refer to caption
Figure 4: 40Ca cross section with parallel kinematics. The experimental data taken from NIKHEF kramer .

In the present work, we calculate the reduced cross section ρ(pm)\rho(p_{m}) at one particular shell, which is related to the probability that a bound nucleon at a given orbit with the missing momentum 𝐩m\mathbf{p}_{m} can be knocked out of the nucleus with asymptotic momentum 𝐩\mathbf{p}. The reduced cross section as a function of pmp_{m} is commonly defined by

ρ(pm)=1pEσepd3σdEfdΩfdΩp,\displaystyle\rho(p_{m})={\frac{1}{pE\sigma_{ep}}}{\frac{d^{3}\sigma}{dE_{f}d\Omega_{f}d\Omega_{p}}}, (15)

where the missing momentum is determined by the kinematics 𝐩m=𝐩𝐪\mathbf{p}_{m}=\mathbf{p}-\mathbf{q} with 𝐪\mathbf{q} the momentum of virtual photon mediating EM interactions between electron and proton. The off-shell electron-proton cross section, σep\sigma_{ep} is not uniquely defined. We use the form CC1 σep\sigma_{ep} given by Ref. npa1983 . In all the calculations SFs are calibrated with the KIDS0 model.

Figure 4 shows the cross section of 40Ca in the parallel kinematics where 𝐩𝐪\mathbf{p}\parallel\mathbf{q}. The incident electron energy is Ei=412E_{i}=412 MeV and the energy transfer to proton is ω=100\omega=100 MeV. Theory results are represented with lines, and data from NIKHEF kramer are noted with red circles. In the 2s1/2 state, theory results are similar to each other. Spectroscopic factors are adjusted to reproduce the cross section data around peak at pm=0p_{m}=0 MeV/cc with KIDS0 model. Theory agrees well with data not only at p0p\simeq 0 MeV/cc, but over pm<200p_{m}<200 MeV/cc. In the 1d3/2 state, experiment at peaks around pm150p_{m}\simeq-150 MeV/cc and 100 MeV/cc are reproduced with good accracy, and agreement is extended over the range pm<200p_{m}<200 MeV/cc. Model dependence is weak, but a small gap between Group1 and Group2 is seen at pm<0p_{m}<0. Group2 models predict the cross sections slightly larger than the Group1 models.

Refer to caption
Figure 5: 208Pb cross section with parallel kinematics. The experimental data were measured from NIKHEF bobeldijk .

Figure 5 presents the result of 208Pb in the parallel kinematics. The incident electron energy is Ei=412E_{i}=412 MeV and the kinetic energy of knocked-out proton is Tp=100T_{p}=100 MeV. Predictions are divided into two groups around the peaks in both 3s1/2 and 2d3/2 states. Since the SFs are calibrated with KIDS0 model, models in Group1 are in good agreement with the NIKHEF data bobeldijk . Models in Group2 that have μS=0.7\mu_{S}=0.7 give cross sections larger than the Group1 models by 15–30%. Isovector effective mass does not follow the grouping of isoscalar effective mass: KIDS0 and SLy4 models have μV=0.8\mu_{V}=0.8, KIDS0-m*77 has μV=0.7\mu_{V}=0.7, and μV=0.9\mu_{V}=0.9 for the KIDS0-m*99 model. Therefore it is likely that the model dependence of the cross section is dominated by the isoscalar effective mass, and the effect of the isovector effective mass is, if ever, quite limited. Result of μS=0.9\mu_{S}=0.9 (KIDS0-m*99) can hardly be differentiated from that of μS=1.0\mu_{S}=1.0 (KIDS0) in the 3s1/2 state, but in the 2d3/2 state, result of the former is slightly enhanced over the latter. This difference is, though small, in accordance with the correpondence of small effective mass to large cross section.

Refer to caption
Refer to caption
Figure 6: 16O cross section with perpendicular kinematics. The experimental data in the upper panels were measured from Saclay chinitz and in the lower panels from JLab gao .

Figure 6 depicts the result of 16O in the perpendicular kinematics. In the perpendicular kinematics |𝐩|=|𝐪||\mathbf{p}|=|\mathbf{q}|, so 𝐩m\mathbf{p}_{m} is almost perpendicular to 𝐪\mathbf{q}. Upper panels compare the result with data from Saclay chinitz , and the lower panels with data from JLab gao . Energies of the incident electrons are 580 MeV in Saclay and 2441 MeV in JLab. The kinetic energies of the knocked-out protons are Tp=159T_{p}=159 MeV and Tp=427T_{p}=427 MeV, respectively. As a result, even though the ranges of pmp_{m} are similar, both theoretical results and experimental data are different in the upper and lower panels. For the same reason, SFs are different depending on the incident electron energies. Results obtained from the models are not sensitive to the effective mass, so they agree well with experiment at both energies. Looking into the details around peaks, cross sections are increasing in the order of KIDS0, KIDS0-m*99 and KIDS0-m*77. Results of SLy4 are indistinguishable from those of KIDS0-m*77. Again the isoscalar effective masses are decreasing in the order of increasing cross sections. This behavior is consistent with what has been observed in the result of 208Pb. The result supports that the correlation between the isoscalar effective mass and the scattering cross section is not limited to specific nuclei, but valid over a wide range of mass number.

Refer to caption
Figure 7: 40Ca cross section with perpendicular kinematics. The experimental data were measured from NIKHEF kramer .

Figure 7 reports the cross section of 40Ca in the perpendicular kinematics with the incident electron energy and the energy transfer to proton the same with those in Fig. 4. Similar to the parallel kinematics, theory reproduces the data with accracy, and the dependence on the effective mass is marginal. Spectroscopic factors are the same with that of the parallel kinematics in the 2s1/2 state, but we have a reduced value in the 1d3/2 state. In the result of 16O, we saw that the kinematic conditions have effect to the value of SF. Different SF values in the 1d3/2 state could be attributed to the kinematic conditions.

   This work    Ref. jin1992    Ref. udias1993    Ref. volya2007    Ref. gnez2014
40Ca 2s1/2 0.72 0.75 0.44 – 0.51 0.87 0.825 – 0.931
1d3/2 0.65 – 0.80 0.80 0.60 – 0.76 0.93 0.848 – 0.966
208Pb 3s1/2 0.95 0.71 0.65 – 0.70 0.85 0.787 – 0.929
2d3/2 0.90 - 0.66 – 0.73 0.90 0.783 – 0.937
Table 3: Spectroscopic factors calculated from theory jin1992 ; udias1993 ; gnez2014 and extracted from experiment volya2007 .

Spectroscopic factors provide understanding of the structural details of nuclei, and allow the estimation of the contribution of many-body correlations that could be missed in the mean field approximation. Therefore SFs is a measure to figure out the validity, accuracy and limit of shell description of a model. Table III collects SFs from literature, and compares them with the result of this work. Calculations based on relativistic formalism jin1992 ; udias1993 obtain results smaller than the values evaluated from experiment volya2007 and a phonon-coupling calculation gnez2014 . Results of the latter two volya2007 ; gnez2014 agree to each other. Result of 208Pb in this work is consistent with Refs. volya2007 ; gnez2014 , but those of 40Ca agree with relativistic calculations jin1992 ; udias1993 . Consequently our work predicts SFs of 40Ca smaller than those of 208Pb. It could be interpreted that the mean field approximation works better in heavy nuclei. Result of 16O is consistent with this tendency because the SFs are 0.62 – 0.75 in 1p1/2 and 0.61 – 0.70 in 1p3/2.

III.4 Response function

Refer to caption
Figure 8: The fourth response functions from 16O with the kinematics same as the JLab.

In the laboratory frame, the quasi-elastic cross section for the (e,ep)(e,e^{\prime}p) reaction is simply written as

d3σdEfdΩfdΩp\displaystyle{\frac{d^{3}\sigma}{dE_{f}\,d\Omega_{f}\,d\Omega_{p}}} =\displaystyle= K[vLRL+vTRT+cos2ϕpvTTRTT\displaystyle K\Big{[}v_{L}R_{L}+v_{T}R_{T}+\cos 2\phi_{p}\,v_{TT}R_{TT} (16)
+cosϕpvLTRLT+hsinϕpvLTRLT],\displaystyle\qquad\qquad\qquad+\cos\phi_{p}\,v_{LT}R_{LT}+h\sin\phi_{p}\,v_{LT^{\prime}}R_{LT^{\prime}}\Big{]}~{},

where the kinematic factor KK is given by K=pE(2π)3σMK={\frac{p\,E}{(2\pi)^{3}}}\,\sigma_{M} with the Mott cross section σM=α24Ei2cos2(θe/2)sin4(θe/2)\sigma_{M}={\frac{\alpha^{2}}{4E^{2}_{i}}}\,{\frac{\cos^{2}({\theta_{e}}/2)}{\sin^{4}({\theta_{e}}/2)}}. In Eq. (16), RLR_{L}, RTR_{T}, RTTR_{TT}, and RLTR_{LT} are referred to as the longitudinal, transverse, longitudinal-transverse, and transverse-transverse interferences response functions, respectively. The fifth one, RLTR_{LT^{\prime}}, indicates the polarized longitudinal-transverse interference, which is directly proportional to the electron beam asymmetry. The four-momenta of the incoming and outgoing electrons are labeled piμ=(Ei,𝐩i)p_{i}^{\mu}=(E_{i},{\bf p}_{i}) and pfμ=(Ef,𝐩f)p_{f}^{\mu}=(E_{f},{\bf p}_{f}). In the parallel kinematics, three interference terms disappear, but they can be extracted in the perpendicular kinematics. The detailed discussions are in Ref. kimepja01 . The electron kinematic factors in Eq. (16) are given in terms of the four-momentum transfer, q=(ω,𝐪)q=(\omega,\mathbf{q}), and the electron scattering angle θe\theta_{e}:

vL=q4𝐪4,vT=tan2θe2q22𝐪2,vTT=q22𝐪2,\displaystyle v_{L}=\frac{q^{4}}{\mathbf{q}^{4}}~{},\qquad v_{T}=\tan^{2}\frac{\theta_{e}}{2}-\frac{q^{2}}{2\mathbf{q}^{2}}~{},\qquad v_{TT}=-\frac{q^{2}}{2\mathbf{q}^{2}}~{},
vLT=q2𝐪2(tan2θe2q22𝐪2),vLT=q2𝐪2tan2θe2.\displaystyle v_{LT}=-\frac{q^{2}}{\mathbf{q}^{2}}\,\left(\tan^{2}\frac{\theta_{e}}{2}-\frac{q^{2}}{2\mathbf{q}^{2}}\right)~{},\qquad v_{LT^{\prime}}=-\frac{q^{2}}{\mathbf{q}^{2}}\,\tan^{2}\frac{\theta_{e}}{2}. (17)

In Eq. (16), the fourth structure function could be obtained by subtracting the cross sections at azimuthal angles of the outgoing proton ϕp=0\phi_{p}=0 and ϕp=π\phi_{p}=\pi and keeping the other electron and outgoing proton kinematics variables fixed. The fourth structure function is a function of the missing momentum given by

RLT=σRσL2KvLT,R_{LT}={\frac{\sigma^{R}-\sigma^{L}}{2Kv_{LT}}}, (18)

where LL (left) and RR (right) indicate the left side at ϕp=0\phi_{p}=0 and the right side at ϕp=π\phi_{p}=\pi of the cross section in Eq. (16), respectively.

Refer to caption
Figure 9: The left-right asymmetry from 16O with the kinematics same as the JLab.

Figure 8 shows the fourth response functions from 16O with the same kinematics as the JLab experiment. The explanations of the curves are the same as the previous results. The difference between the effective masses is at most about 45 % and 10 % around the position of peak at 1p1/2 and 1p3/2 orbits, respectively. The sensitivity of the effective masses is clearly exhibited on the fourth response function. We learn that the values of the SF for the fourth response function and for the reduced cross section may be different for a good agreement with the data.

Finally, we also calculate another left-right asymmetry, ALTA_{LT}, defined as

ALT=σRσLσR+σL.A_{LT}={\frac{\sigma^{R}-\sigma^{L}}{\sigma^{R}+\sigma^{L}}}. (19)

The kinematics in Fig. 9 are the same as in Fig. 8. Since this asymmetry does not require any SF, it is possible to compare the theoretical result with experimental data directly. The role of the effective masses is not large compared to the fourth response function but our theoretical results describe the experimental data relatively well except one point at pm350p_{m}\simeq 350 MeV/cc in the 1p3/2 orbit.

IV Summary

In the present work, we have explored the ground state properties of spherical nuclei with KIDS model by considering the quasi-elastic scattering of electrons from 16O, 40Ca, and 208Pb. Wave functions of the nucleon in nuclei were obtained by solving non-relativistic Hartree-Fock equations. They were transformed to a relativistic form, and the cross sections were calculated in the relativistic formalism. The effects of the isoscalar and isovector effective masses were investigated by calculating charge and neutron distributions, single particle energies for a few orbits in 16O, 40Ca, 208Pb, and cross sections in (e,ep)(e,e^{\prime}p), and by comparing the result with experimental data.

In light nuclei 16O and 40Ca, the distributions of proton and neutron are insensitive to the effective masses, but in 208Pb, the effect of the effective mass appears clearly. In particular, the difference of charge distribution at r<1r<1 fm is due to the isoscalar effective mass. For the single particle energies, our theoretical results deviate at most about 18 % from the experimental data. From the energy levels, the small isoscalar effective masses describe the experimental values well in light nuclei, but for 208Pb the isoscalar effecitve masses that are close to 1 reproduce the experiment better than the light effective masses.

Reduced cross sections of exclusive (e,ep)(e,e^{\prime}p) reaction from 16O, 40Ca, and 208Pb agree well with experimental data. In the light nuclei 16O and 40Ca, contribution of the effective mass is negligible. For 208Pb, however, cross section depends clearly on the isoscalar effective mass: cross sections become large with smaller isoscalar effective mass. Although the left-right asymmetry is insensitive to the effective mass, the contribution of the mass to RLTR_{LT} tends to be large with smaller mass. Value of SF to reproduce RLTR_{LT} could be different from that for reduced cross section. On the other hand, contribution of the isovector effective mass is vanishingly small.

Spectroscopic factors have been adjusted so that the KIDS0 model reproduces the cross section data. We obtained SFs in the range 0.6–0.8 for light nuclei and 0.9–0.95 for 208Pb. Our results are compatible with experiment and other theory. We showed that SF could be dependent on the kinematic conditions such as energy and angle.

Response functions were calculated for 16O in the longitudinal-transverse channel, and compared with data from JLab. Response function in the 1p1/2 orbit depends on the effective mass sensitively, so even the difference between μS=0.9\mu_{S}=0.9 and μS=1.0\mu_{S}=1.0 could be identified clearly. Left-right asymmetry, on the other hand, does not show notable dependence on the effective mass, and the theory result agrees well with data.

In conclusion, we have confirmed that quasi-electron scattering provides a useful tool to study the effective mass of the nucleon in nuclear medium.

Acknowledgments

This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea govenment (No. 2018R1A5A1025563 and No. 2020R1F1A1052495).

References

  • (1) E. Schanat, P. Boncher, P. Haensel, J. Meyer, and R. Schaeffer, Nucl. Phys. A 635, 231 (1998).
  • (2) M. Kortelainen, T. Lesinski, J. Moré, W. Nazarewicz, J. Sarich, N. Schunck, M. V. Stoitsov, and S. Wild, Phys. Rev. C 82, 024313 (2010).
  • (3) H. Gil, P. Papakonstantinou, C. H. Hyun, and Y. Oh, Phys. Rev. C 99, 064319 (2019).
  • (4) E. D. Cooper, S. Hama, B. Clark, and R. L. Mercer, Phys. Rev. C 47, 297 (1993).
  • (5) K. S. Kim, and L. E. Wright, Phys. Rev. C 56, 302 (1997); K. S. Kim, and L. E. Wright, ibid. 60, 067604 (1997).
  • (6) P. Papakonstantinou, T.-S. Park, Y. Lim, and C. H. Hyun, Phys. Rev. C 97, 014312 (2018).
  • (7) H. Gil, Y.-M. Kim, C. H. Hyun, P. Papakonstantinou, and Y. Oh, Phys. Rev. C 100, 014312 (2019).
  • (8) H. Gil, N. Hinohara, C. H. Hyun, and K. Yoshida, arXiv:2013.16135 [nucl-th].
  • (9) H. Gil, and C. H. Hyun, New Physics: Sae Mulli 71, 242 (2021).
  • (10) H. Gil, Y.-M. Kim, P. Papakonstantinou, and C. H. Hyun, Phys. Rev. C 103, 034330 (2021).
  • (11) A. Akmal, V. R. Pandharipande, and D. G. Ravenhall, Phys. Rev. C 58, 1804 (1998).
  • (12) H. Gil, Y. Oh, C. H. Hyun, and P. Papakonstantinou, New Physics: Sae Mulli 67, 456 (2017).
  • (13) I. Sick et al., Nucl. Phys. A 150, 631 (1970).
  • (14) I. Sick et al., Nucl. Phys. A 354, 37 (1981).
  • (15) L. Ray, and P. E. Hodgson, Phys. Rev. C 20, 2403 (1979).
  • (16) J. L. Faiar et al., Nucl. Phys. A 212, 93 (1973).
  • (17) K. Saito, K. Tsushima, and A. W. Thomas, Prog. Part. Nucl. Phys. 58, 1 (2007).
  • (18) N. Schwierz, I. Wiedenhöver, and A. Volya, arXiv:0709.3521 [nucl-th].
  • (19) T. De Forest, Nucl. Phys. A 392, 232 (1983).
  • (20) G. J. Kramer et al., Phys. Lett. B 227, 199 (1989).
  • (21) I. Bobeldijk et al., Phys. Rev. Lett. 73, 2684 (1994).
  • (22) L. Chinitz et al., Phys. Rev. Lett. 67, 568 (1991).
  • (23) J. Gao, et al., Phys. Rev. Lett. 84, 3265 (2000).
  • (24) Y. Jin, D. S. Onley, and L. E. Wright, Phys. Rev. C 45, 1311 (1992).
  • (25) J. M. Udias, P. Sarriguren, E. Moya de Guerra, E. Garrido, and J. A. Caballero, Phys. Rev. C 48, 2731 (1993).
  • (26) N. V. Genzdilov, E. E. Saperstein, and S. V. Tolokonnikov, EPL 107, 62001 (2014).
  • (27) K. S. Kim, M. K. Cheoun, Y. Chung, and H. J. Nam, Eur. Phys. J. A 11, 147 (2001).