Skip to main content

ORIGINAL RESEARCH article

Front. Astron. Space Sci., 14 July 2021
Sec. High-Energy and Astroparticle Physics
This article is part of the Research Topic Neutrino Nuclear Responses for Astro-Particle Physics by Nuclear Reactions and Nuclear Decays View all 11 articles

Quenching of Isovector and Isoscalar Spin-M1 Excitation Strengths in N = Z Nuclei

  • 1Research Center for Nuclear Physics (RCNP), Osaka University, Ibaraki, Japan
  • 2Department of Radiology, Kyoto Prefectural University of Medicine, Kyoto, Japan

Spin-M1 excitations of nuclei are important for describing neutrino reactions in supernovae or in neutrino detectors since they are allowed transitions mediated by neutral current neutrino interactions. The spin-M1 excitation strength distributions in self-conjugate N=Z nuclei were studied by proton inelastic scattering at forward angles for each of isovector and isoscalar excitations as reported in H. Matsubara et al., Phys. Rev. Lett. 115, 102501 (2015). The experiment was carried out at the Research Center for Nuclear Physics, Osaka University, employing a proton beam at 295 MeV and the high-resolution spectrometer Grand Raiden. The measured cross-section of each excited state was converted to the squared nuclear matrix elements of spin-M1 transitions by applying a unit cross-section method. Comparison with predictions by a shell-model has revealed that isoscalar spin-M1 strengths are not quenched from the prediction although isovector spin-M1 strengths are quenched similarly with Gamow-Teller strengths in charged-current reactions. This finding hints at an important origin of the quenching of the strength relevant to neutrino scattering, that is, the proton-neutron spin-spin correlation in the ground state of the target nucleus. In this manuscript we present the details of the unit cross-section method used in the data analysis and discuss the consistency between the quenching of the isoscalar magnetic moments and that of the isoscalar spin-M1 strengths.

1 Introduction

Response of nuclei to incoming neutrinos is categorized into two types of reactions: charged-current (CC) and neutral-current (NC). Gamow-Teller (GT) transition of nuclei belongs to the CC neutrino reaction, while the isovector (IV) spin magnetic-dipole (M1) transition to NC. The GT (ΔJπ=1+, ΔT=1 and ΔTz=±1) transitions are analogous to the IV spin-M1 (ΔJπ=1+, ΔT=1 and ΔTz=0) transitions under isospin symmetry. Relevant transition rates are predicted by theoretical models such as the shell-model. The experimentally observed transition rates are, however, quenched compared to the model predictions by employing bare transition operators. Quenching is a basic property of nuclear structure and influences the neutrino reaction rates in astrophysical processes and terrestrial neutrino detectors. The nuclear spin responses and their quenching have strong effects on the mean free path of neutrinos in dense nuclear matter, the size of the neutrino sphere formed in the center of a core-collapsing star, and the cooling process of proton-neutron stars.

The quenching has been extensively studied for the GT transitions. The GT transition strength contained in the GT giant resonances studied by (p,n) reactions was found to be ∼60% of the prediction by the Ikeda-Fujii-Fujita sum-rule (Ikeda et al., 1963) consistently for a large variety of nuclei. Two mechanisms were presented to explain GT quenching in the context of the mixing of higher-order configurations with the fundamental one-particle-one-hole nature of the GT excitation. One is the Δ-hole mixing originating from the quark degree of freedom. The other is the two-particle-two-hole as well as the higher-order particle-hole excitations within the nucleonic degree of freedom. Detailed study of the GT strength distribution embedded in the continuum located above the GT giant resonance revealed that the major part of quenching is caused by mixing in the nucleonic degree of freedom (Ichimura et al., 2006). The observed GT transition strengths studied by beta-decay are also quenched from shell-model predictions using the bare g-factor. Recent ab initio calculations using the chiral effective field theory (Gysbers et al., 2019) indicated that quenching was resolved by introducing two-body currents and nuclear many-body correlations. Quenching in the analogous spin-M1 transitions was studied by proton inelastic scattering (Anantaraman et al., 1984; Crawley et al., 1989). The result was unclear due to the poor quality of the experimental data and to the ambiguity of the transition matrix element relying on the reaction calculation. This problem has been overcome by the achievement of a high-precision measurement (Matsubara et al., 2015) using high-resolution proton scattering at forward scattering angles including zero degrees that reported the quenching of the nuclear matrix elements for the IV spin-M1 transitions, similar to the analogous GT transitions, but no-quenching for the isoscalar (IS) spin-M1. Exhaustion of the sum-rule of the spin-M1 strengths is relevant to the spin susceptibility of asymmetric nuclear matter, its response to the strong magnetic field in magnetars, and possible phase-transition of nuclear medium in a neutron star to the ferromagnetic state.

The experimental work gave new insight, by the use of the non-energy-weighted sum-rule, that the underlying quenching mechanism is embedded in the ground state property as spin-spin correlation of neutron(n)-proton(p) pairs. The expectation value of the correlation in the ground state is equivalent to the difference of the quenching of nuclear matrix elements between the IS and IV spin-M1 transitions. Thus, the quenching of the IS and IV spin-M1 transitions needs to be described simultaneously and mutually-consistently. The details are described in Section 5.4.

The n-p correlation in the nuclear ground state is one of the recent topics in nuclear physics. The short range correlation as well as the tensor correlation are considered to be the origin of the neutron-proton correlation and are relevant to the IS spin-triplet n-p pairing. The high-momentum nature of the correlation has been studied by the knockout reaction by electron scattering (Subedi et al., 2008; Hen et al., 2014; Hen et al., 2017) or by high-momentum transfer reaction by using (p,d) scattering (Ong et al., 2013; Terashima et al., 2018). In both cases, dominant contribution of the n-p pairs is reported rather than the identical pairs, n-n or p-p, to the high-momentum component in the ground state.

In a different approach, the spin-aligned IS n-p coupling of the valence particles around the Fermi surface was studied from a level structure determined by gamma spectroscopy (Cederwall et al., 2011). It is interesting to observe how those n-p paring components and the IS spin-triplet n-p pairing are related to the n-p spin-spin correlation in the ground state.

In the work of Matsubara et al. (2015), IS and IV spin-M1 excitation strength distributions were individually determined by a high-resolution proton inelastic scattering experiment at zero degrees and forward angles for self-conjugate even-even nuclei from 12C to 36Ar. The squared nuclear matrix element of each transition was extracted from the observed differential cross-section by using the unit cross-section method. The summed strength up to the excitation energy of 16 MeV was compared with the prediction by a shell-model for the discussion of quenching of the IS and IV transitions. In this article we describe in greater detail the unit cross-section method in the data analysis. Also, the observed no-quenching of the IS spin-M1 transitions is compared with the historical knowledge of the quenching of the IS magnetic moment. The difference of IS and IV quenching is discussed in terms of the n-p spin-spin correlation in the ground state.

In Section 2, formalism is presented for discussion of the nuclear matrix elements of the IS and IV spin-M1 transitions, the IS and IV magnetic moments of the nuclear ground state, sum-rules, and the n-p spin-spin correlation function. The experimental methods and the data analysis are described in Section 3. The nuclear matrix elements are determined from the experimental data by using the unit cross-section method as described in Section 4. The quenching of the IS and IV spin-M1 nuclear matrix elements, IS magnetic moment, and the n-p spin-spin correlation are discussed in Section 5. Summary and prospects are given in Section 6.

2 Formalism

2.1 Nuclear Magnetic Moment

The M1 operator O^(M1) for magnetic dipole moments and M1 transitions consists of an orbital part (l) and a spin part (s).

O^(M1)=[k=1Z(glπlk+gsπsk)+k=Z+1A(glνlk+gsνsk)]μN(1)
=[k=1A{(glISlk+gsISσk2)+(glIVlk+gsIVσk2)τz,k}]μN,(2)

where μN is the nuclear magneton, σ is the Pauli spin matrix, the operator τz,k is the third component of the isospin operator τ acting on the k-th nucleon and its eigen value is +1 for neutrons (ν) and −1 for protons (π). The gyromagnetic factors (g-factors) of glIS, gsIS, glIV, and gsIV are taken as glIS=12(glπ+glν)=0.5, gsIS=12(gsπ+gsν)=0.880, glIV=12(glπglν)=0.5, and gsIV=12(gsπgsν)=4.706, where the g-factors in the free space, glπ=1, glν=0, gsπ=5.586, and gsν=3.826, are employed. The suffixes of IS and IV denote isoscalar and isovector, respectively. Thus, a magnetic moment is expressed as

μ=i|O^(M1)|iM=J,(3)

where |i denotes an initial state. Here, magnetic moments can be divided into IS and IV parts by corresponding analogous magnetic moments in mirror nuclei (Tz=±T) as

μIS=12(μ++μ)(4)
μIV=12(μ+μ),(5)

respectively, where μ+() is a magnetic moment in the case of Tz=+T(T). Because total spin has a relation J=l+s, the following can be obtained

J=i|Lz|iM=J+i|Sz|iM=J,(6)

where L=kAlk and S=kAsk (Brown and Wildenthal, 1983). Applying the above relation to IS magnetic moment (μIS=i|glISL+gsISS|iM=J), one gets

Si|Sz|iM=J=μISglIS×JgsISglIS,(7)

where we denote the function as S according to Brown and Wildenthal (1983), and it will be discussed in Section 5.3.

2.2 M1 Transition Strength

The reduced transition probability (transition strength) for M1 excitation is written as

B(M1)=34π|f||O^(M1)i|2=34π|glISMf(l)+gsIS2Mf(σ)+glIVMf(lτz)+gsIV2Mf(στz)|2μN2,(8)

where |f denotes a final state in a transition. Following the convention of Edmonds (Edmonds, 1960; Brown and Wildenthal, 1987), the reduced nuclear matrix element in spin from an initial state to a final state is defined as

Mf(O^)=12Ji+1f||k=1AO^k||i,(9)

where O^(k) denotes l(k), σ(k), l(k)τz(,k), and σ(k)τz(,k). Expressing corresponding analogous M1 transitions in mirror nuclei (Tz=±T) as B(M1)±, IS and IV parts of a transition strength B(M1) can be written using B(M1)± (Fujita et al., 2000; Fujita et al., 2011) or using the reduced matrix elements as

B(M1)IS=14[B(M1)+B(M1)]2=34π|glISMf(l)+gsIS2Mf(σ)|2μN2(10)
B(M1)IV=14[B(M1)++B(M1)]2=34π|glIVMf(lτz)+gsIV2Mf(στz)|2μN2,(11)

respectively. If only spin parts are extracted as IS and IV spin-M1 transition strengths, they are expressed as

B(M1)σ=34π|gsIS2Mf(σ)|2μN2(12)
B(M1)στ=34π|gsIV2Mf(στz)|2μN2(13)

respectively.

Here, we focus on an IS part, B(M1)IS. Because total angular momentum operator (j=l+σ/2) gives a good quantum number, taking a ground state as |g.s., j|g.s is proportional to |g.s but is orthogonal to any other eigenstates. Thus, the following restriction (Bernabéu et al., 1992; Kawabata et al., 2004) can be obtained

f||k=1A(lk+12σk)||g.s.=0(14)

Since the above restriction leads to Mf(l)=12Mf(σ), the right-hand side of Eq. 10 is rewritten as

B(M1)IS=34π|gsISglIS2Mf(σ)|2μN2=(gsISglISgsIS)2B(M1)σ(15)

2.3 Squared Nuclear Matrix Element

Nuclear excitation ΔJπ=1+ with ΔTz=0 is an M1 transition. When nuclear excitation at low momentum transfer is considered, spin-parts of the M1 transition are probed due to the local nature of the nucleon-nucleon (NN) interaction (Petrovich and Love, 1981). Thus, the 1+ excitation by (p,p) reaction at forward angles is spin-M1 transition. Since spin-M1 transition is not probed by electromagnetic interaction but by nuclear interaction, its transition strength does not relate to g-factors. Therefore, transition strengths of IS and IV spin-M1 transitions from the ground state |g.s. to an excited state |f are expressed by squared nuclear matrix element (SNME) as

|Mf(σ)|2=|12Ji+1f||k=1Aσk||g.s.|2(16)
|Mf(στz)|2=|12Ji+1f||k=1Aσkτz,k||g.s.|2(17)

respectively. The factor 1/2Ji+1 is unity for a 0+ ground state.

2.4 Relation to Gamow-Teller Excitation

Next, reduced nuclear matrix element and transition strength in GT excitation are defined as

Mf(στ±)=12Ji+1f||12k=1Aσkτ±,k||g.s.(18)
B(GT±)=|Mf(στ±)|2,(19)

respectively, where τ±=12(τx±iτy). Applying the Wigner-Eckart theorem in the isospin space, transition strengths of GT and IV spin-M1 excitations are obtained as

B(GT±)=|12Ji+1Ti,Tiz,1,±1|Tf,Tfz>2Tf+112M(GT±)|2,(20)
B(M1)στ=34π|gsIV212Ji+1Ti,Tiz,1,0|Tf,Tfz2Tf+1M(M1στ)|2μN2,(21)

respectively, where M(GT±) and M(M1στ) are reduced nuclear matrix elements in spin and isospin expressed as

M(GT±)=f|||k=1Aσkτ±,k|||g.s.(22)
M(M1στ)=f|||k=1Aσkτz,k|||g.s.(23)

respectively. Because |M(GT±)|2=|M(M1στ)|2 is realized under the assumption of isospin symmetry, the following relationship between B(GT±) and B(M1)στ can be obtained as

B(GT±)B(M1)στ/μN2=8π31(gsIV)2Ti,Tiz,1,±1|Tf,Tfz2Ti,Tiz,1,0|Tf,Tfz2(24)

Here, it should be noted that the isospin symmetry is reasonably assumed within the accuracy of the data in the present study although the meson exchange current contribution can be different between an IV M1 transition measured by electron scattering and the analogous GT transition by charge-exchange reaction (Richter et al., 1990; Lüttge et al., 1996).

2.5 Total Spin Correlation in Ground State

For the discussion of total spin correlation in a ground state (Matsubara et al., 2015), the difference between the sums of the IS and IV spin-M1 SNMEs integrated up to the excitation energy of Ex, Δspin(Ex), is defined as

Δspin(Ex)=116{Ef<Ex|Mf(σ)|2Ef<Ex|Mf(στz)|2}

With the proton (neutron) total spin operator Sp(n) defined as

Sp(n)=12k=1Z(N)σk,(26)

the sum is taken for all the protons (neutron).When a ground state is Jπ=0+, the IS and IV spin-M1 nuclear matrix elements are represented by

Mf(σ)=2f|Sn+Sn|0(27)
Mf(στz)=2f|SnSp|0(28)

respectively, where |0 denotes the ground state. In the limit of Ex, the completeness of the final state, |f, yields

(Sn+Sp)2=fg.s.|Sn+Sp|ff|Sn+Sp|g.s.=limEx14Ef<Ex|Mf(σ)|2(29)
(SnSp)2=limEx14Ef<Ex|Mf(στz)|2(30)

Here the expectation values of the left side of the equations are taken for the 0+ ground state. We then derive

ΔspinlimExΔspin(Ex)=14{(Sn+Sp)2(SnSp)2}=SpSn

which represents the expectation value of the proton-neutron spin-spin correlation in the ground state.

3 Experiment

In this section we briefly describe the experimental method and the assignment of the spin-M1 excitations. Details can be found in former publications (Tamii et al., 2009; Matsubara et al., 2015; von Neumann-Cosel and Tamii, 2019).

3.1 Measurement of the (p,p’) Reactions

The experiment was performed at the cyclotron facility of the Research Center for Nuclear Physics (RCNP), Osaka University. A proton beam was accelerated by a cascade of two cyclotrons to Ep = 295 MeV. The beam was transported to the West-South (WS) beam line (Wakasa et al., 2002), where a high-dispersion on target was created. An excitation-energy resolution of 18 keV (FWHM) was achieved by applying dispersion matching (Fujita et al., 1997; Fujita H. et al., 2002; Fujita et al., 2011) between the WS beam line and the Grand Raiden (GR) spectrometer (Fujiwara et al., 1999). The scattered protons by the target were momentum-analyzed and were detected by two sets of multi-wire drift-chambers and two plastic scintillation counters at the focal plane by GR spectrometer. A scattering angle range of 0–14° was covered by placing the GR spectrometer at 0, 2.5, 4.5, 6, 8, 10, 12, and 14°. The details of the experimental technique are described in Ref. (Tamii et al., 2009; von Neumann-Cosel and Tamii, 2019).

Self-conjugate even-even nuclei, 12C, 24Mg, 28Si, 32S and 36Ar, were used as the target. Areal densities of 1.0–2.5 mg/cm2 were prepared for 12C, 24Mg, and 28Si. Magnesium and argon targets were isotopically enriched to 100%, while the others were in natural abundance. The 32S target was kept at the liquid nitrogen temperature for preventing sublimation due to heat by charged particle irradiation (Matsubara et al., 2009) with an areal density of 15 mg/cm2. The 36Ar target was kept at 1.0 atm in a gas cell at room temperature (Matsubara et al., 2012) sealed by aramid foils with a thickness of 6 μm on one side.

3.2 Assignment of Spin-M1 Excitations

Figure 1A shows an excitation energy spectrum of the 28Si(p,p) reaction at a scattering angle of 0.0–0.5°. The excited states below Ex = 16 MeV are well isolated from the others. Excited state with a spin-parity of 1+ state was identified by comparing the shape of the measured angular distribution of the differential cross-section with Distorted-Wave Impulse Approximation (DWIA) calculations by using the code DWBA07 (Raynal, 2007). One-body transition densities were obtained by shell-model calculations with the code Nushell@MSU (Brown and Rae, 2014) and the USD interactions (Brown and Wildenthal, 1983). No sizable difference in the angular distribution was observed depending on the choice of the effective interaction from USD, USDA, or USDB (Brown and Richter, 2006; Richter et al., 2008). The effective NN interaction parametrized at 325 MeV was used after conversion to a beam energy of 295 MeV as indicated in (Love and Franey, 1981; Franey and Love, 1985). Optical potential parameters were determined by fitting the angular distribution of the differential cross-section of the elastic scattering measured in the same experiment (Matsubara, 2010). Harmonic oscillator parameters were taken from a global analysis (Kirson, 2007).

FIGURE 1
www.frontiersin.org

FIGURE 1. (A) Excitation energy spectrum of the 28Si(p,p') reaction at Ep=295 MeV and θlab=00.5. The arrows in the figure indicate conditions that are definitely determined as 1+. Angular distributions of (B) IS and (C) IV spin-M1 excitations marked by red and blue arrows in (A), respectively, in comparison with theoretical angular distributions for several multipolarities. The figures were taken from Matsubara et al. (2015).

The measured angular distributions of the differential cross-section for the excited states at Ex = 9.495 (1+: T = 0) and 11.447 MeV (1+: T = 1) in 28Si are shown in Figures 1B,C, respectively, by the solid circles. They are compared with the predictions of the DWIA calculation shown by the curves for Jπ = 0+, 1, and 2+, and for each of the IS and IV 1+ transitions. The predicted cross-sections are normalized to the experimental data at the smallest measured angle. The angular distribution of the IS 1+ excitation is predicted to be flatter than the IV 1+ excitation at the forward angles smaller than 5° due to the relatively stronger contribution of the exchange tensor component compared to the central component in the effective NN interaction (Franey and Love, 1985; Tamii et al., 1999). Thus, the measured angular distribution allowed determination of the transferred isospin, ΔT = 0 (IS) or 1 (IV), for the 1+ transitions.

The IS and IV 1+ excited states were assigned from the observed discrete excited states by the following method. First the angular distribution of the differential cross-sections for each of the observed discrete excited states was deduced from the data analysis. The excited states were selected according to the angular distribution below 5° for those having an almost flat distribution as a signature of an IS 1+ state or a quickly dropping distribution of an IV 1+ state as shown in Figures 1B,C, respectively. Second, the most appropriate assignment was chosen from IS 1+, IV 1+, 0+, 1, and 2+ distributions by comparing the angular distribution up to 14° with the theoretical predictions. Third, the assignment for the IS or IV 1+ states with a reduced χ2 value close to unity was taken as confident and that for the rests as less confident. The confident assignments of the IS and IV 1+ states were in good agreement for the conditions studied by electron scattering (Bendel et al., 1971; Richter et al., 1990; Foltz et al., 1994; Lüttge et al., 1996; Hofmann et al., 2002). The high energy-resolution measurement allowed us to observe several new states including the less-confident assignments as shown in Section 5.1.

4 Unit Cross-Section Method

4.1 Definition

4.1.1 Unit Cross-Section

The differential cross-section of the 1+ excitations by proton inelastic scattering at 0° is considered to be approximately proportional to SNME in the intermediate energy region of 100–400 MeV. Unit cross-sections (UCSs), σ^IS and σ^IV, are introduced in analogy to the study of Gamow-Teller excitations by (p,n) reactions (Taddeucci et al., 1987; Sasano et al., 2009). The differential cross-section at 0° is written as

dσ(0)=σ^TFT(q,Ex)|Mf(OT)|2(32)

where T stands for IS or IV. OT is the operator, σ or στz, for IS or IV transitions, respectively. FT(q,Ex) is the kinematic factor that accounts for dependence on the momentum transfer (q) and the excitation energy (Ex), and was determined by the DWIA calculation as explained in Section 4.1.2. The target mass (A) dependence of UCSs is parameterized as (Taddeucci et al., 1987)

σ^T(A)=NTexp(xTA1/3)(33)

where NT and xT are the normalization and mass-dependence parameters. The mass-dependence parameter xT essentially originates from the distortion effect of the reaction that is common between the IS and IV transitions. Thus we assume xIS = xIV, which is held within an accuracy of 5% in DWIA calculations. Details will be discussed in Section 4.4.2.

4.1.2 Kinematic Factor

The kinematic factors FIS(q,Ex) and FIV(q,Ex) were determined by DWIA calculation using the USD interaction as shown in Figure 2 as a case of 28Si, where only OBTDs above the experimental detection limit were employed. The distributions were obtained from a ratio of differential cross-section at 0° at Ex0 to that at Ex=0. As seen in Figure 2B, IV distributions of the kinematic factor overlap. Since the result suggests that FIV(q,Ex) does not depend on wavefunction, FIV(q,Ex) was expressed as a smooth function of Ex and A by fitting. As shown in Figure 2A, however, IS distributions of the kinematic factor depend on wavefunction as they do not overlap. Because the result suggests that FIS(q,Ex) cannot be expressed as a function of Ex owing to dependence of wavefunction, FIS(q,Ex)=1.00±0.10 was simply assumed in order to cover the variation of wavefunctions below Ex=15 MeV.

FIGURE 2
www.frontiersin.org

FIGURE 2. Kinematic correction factors FIS(q,Ex) and FIV(q,Ex) in the case of 28Si obtained from DWIA calculation using USD interaction. (A) The strongest excitation and others are drawn in red bold and black dashed curves, respectively. (B) Distributions of several IV OBTDs are overlapped.

4.2 Derivation From Experiment

4.2.1 Cases of 12-Carbon, 26-Magnesium, 58-Nickel

Isovector UCSs of 12C, 26Mg, and 58Ni were obtained using the data summarized in Table 1 by assuming isospin symmetry. The differential cross-section of (p,p) reaction at Ep=295 MeV at scattering angles 0.4° was taken from Tamii et al. (2009), where 0.4° is the most forward angle by selecting scattering angles between 0.0 and 0.5°. The cross-section was extrapolated to Ex=0 MeV using FIV(q,EX).

TABLE 1
www.frontiersin.org

TABLE 1. Data used for obtaining IV UCSs.

Log ft-value of β-decay and (3He, t) data (Alburger and Nathan, 1978; Fujita Y. et al., 2002; Zegers et al., 2006; Fujita et al., 2007; Fujita et al., 2011) were used to obtain GT strength (B(GT±)) from ground state to an excited state corresponding to the (p,p) cross-section under the isospin symmetry. After B(GT±) was converted to IV spin-M1 SNME following Eqs. 13, 24, σ^IV was obtained. Here, the (p,n) data (Sasano et al., 2009) were also employed for the calibration of B(GT) in 58Ni.

4.2.2 Case of 11-Boron

The γ-decay widths of the mirror states in 11B and 11C from the first excited states (Firestone, 1996), corresponding to B(M1)11B and B(M1)11C, respectively, were employed to obtain B(M1)IS and B(M1)IV following Eqs. 10, 11. Then, B(M1)σ and B(M1)στ were obtained from B(M1)IS and B(GT)=0.402±0.031 (Taddeucci et al., 1990; Kawabata et al., 2004) using Eqs. 15, 24, respectively. Finally, IS and IV spin-M1 SNMEs in the case of B11 were obtained using Eqs. 12, 13, respectively.

The differential cross-section data of the 11B(p,p) reaction at Ep=392 MeV and at Ex=2.12 MeV was taken from Kawabata et al. (2004). The angular distribution of the differential cross-section was decomposed into IS 1+, IV 1+, and 2+ transitions using the DWIA calculation, where an incoherent (a coherent) sum was assumed between 1+ and 2+ (IS 1+ and IV 1+) excitations. The OBTDs based on CKPOT interaction (Cohen and Kurath, 1965), the effective NN interaction derived at 325 MeV (Love and Franey, 1981; Franey and Love, 1985), and the global optical potential parameters (Cooper et al., 2009) were employed in this calculation. The OBTDs were normalized to reproduce the experimental values of IS and IV SNMEs, and B(E2) (Firestone, 1996) in IS 1+, IV 1+, and 2+ excitations, respectively. Because the normalized OBTDs did not reproduce the experimental angular distribution of the differential cross-section, additional normalization factors of 1.1 and 1.3 were applied to the 1+ and 2+ excitations, respectively, as shown in Figure 3. The differential cross-section data of total, IS 1+, and IV 1+ at Ep=392 MeV and at 0° were decomposed as summarized in Table 2, where the experimental uncertainty was assumed to be negligible because the error bars were invisible in Kawabata et al. (2004). Here, choice of an interaction with OBTDs did not change the final result because differences among CKPOT, SFO (Suzuki et al., 2003), and MK3w (Warburton and Millener, 1989) interactions in the model spaces of p, psd, and spsdpf, respectively, were within 0.5%. Next, the cross-section data at Ep=392 MeV were converted to those at Ep=295 MeV by making use of a ratio of the 12C(p,p) reaction at 0° to the states Ex=12.71 and 15.11 MeV known as IS 1+ and IV 1+ excitations, respectively, where the experimental data were taken from Tamii et al. (1999), Tamii et al. (2009). Thus, the data of 11B(p,p) excitation to Ex=2.12 MeV at Ep=295 MeV and at 0° were obtained.

FIGURE 3
www.frontiersin.org

FIGURE 3. Decomposition of differential cross-section of 11B(p,p) reaction at Ep=392 MeV compared to the state at Ex=2.12 MeV. The experimental data were taken from Kawabata et al. (2004).

TABLE 2
www.frontiersin.org

TABLE 2. Data used for obtaining UCSs of 11B. The upper table is expressed in unit of μN2.

Combining the SNMEs with the differential cross-section after correction of FT(q,Ex), IS and IV UCSs in the case of 11B were obtained as summarized in Table 2. Here, σ^IS does not include uncertainty owing to FIS(q,Ex).

4.3 Results of Unit Cross-Section

The experimentally obtained IS and IV UCSs were plotted as a function of the mass number in Figures 4A,B, respectively. The mass dependence of the IV UCS data was fitted with the functional form of Eq. 33 having free parameters of NIV and xIV. The result is shown by the solid line in Figure 4B with the one-sigma uncertainty band. The IS normalization parameter, NIS, was determined from the data at A = 11. The IS slope parameter, xIS, was taken to be the same as xIV as discussed in Section 4.4.2. The IS uncertainty band includes contribution of 10% from the FIS(q,Ex) (Section 4.1.2). The obtained parameters are summarized in Table 3.

FIGURE 4
www.frontiersin.org

FIGURE 4. Empirically determined IS (left) and IV (right) UCSs. The mass dependence of the data were fit with the functional form of Eq. 33 as shown by the solid lines in the IV panel. The band shows the uncertainty of one sigma. The IS slope parameter xIS is taken to be the same as xIV. The figures were adapted from Matsubara et al. (2015).

TABLE 3
www.frontiersin.org

TABLE 3. Empirically determined UCS parameters from the fit shown in Figure 4.

4.4 Model Study

4.4.1 Proportionality of Squared Nuclear Matrix Elements to the Differential Cross-Section

Validity of the application of the UCS method has been theoretically examined by the use of shell-model target wave functions as USD interaction and DWIA calculation. Since B(M1)σ and B(M1)στ were calculated using the code Nushell@MSU (Brown and Rae, 2014), SNMEs were obtained from the relations in Eqs. 12, 13. Each differential cross-section at q=0 and at Ex=0 MeV was calculated using corresponding OBTD of the USD interaction. Here the following correction for mass dependence to be normalized to A=28 was applied as

dσdΩ*=dσdΩ×e0.38×(281/3A1/3)(34)

where dσdΩ* is the corrected cross-section and 0.38 comes from xT determined in Table 3. As shown in Figure 5, good linearity was seen especially above the experimental detection limit shown by the lines for each target nucleus.

FIGURE 5
www.frontiersin.org

FIGURE 5. Proportionality of UCS. Horizontal lines indicate experimental detection limit for 24Mg, 28Si and 32S (dotted), 20Ne (short-dashed) and 36Ar (long-dashed). Cross-section is corrected in terms of mass dependence as in Eq. 34.

4.4.2 Distortion Effect

The mass-dependence term in Eq. 32, xT, essentially originates from the distortion effect that is common between the IS and IV transitions. A slope for each nucleus in Figure 5 corresponds to UCSs theoretically obtained, where the weighting sum within the experimental detection limit was taken. Those ratios of IS to IV UCSs for each nucleus are plotted as a function of mass number in Figure 6, where the suffix of “cal” indicates that the value of UCS is obtained from the calculation. The result for 12C is added in Figure 6. The SFO interaction (Suzuki et al., 2003), which is applicable to the p-shell nuclei, is used instead of the USD interaction. The results suggest that the assumption is supported, as the ratios are constant within 5%.

FIGURE 6
www.frontiersin.org

FIGURE 6. Mass dependence of ratio of IS to IV UCSs studied by calculation.

5 Results and Discussion

5.1 Strength Distribution

The SNMEs of the transitions to the excited states assigned as T=0 (IS) or T=1 (IV) 1+ states were determined by using Eq. 32 from the measured differential cross-section at 0°. The results are plotted in Figure 7 , where the figure was taken from Matsubara et al. (2015) with addition of the 12C data. The strengths of 12C were observed to be centered to single state, which was consistent with previous work (von Neumann-Cosel et al., 2000). For 24Mg, 28Si, 32S, and 36Ar, we identified 1–4 (4–8) states in each target nucleus corresponding to the IS (IV) spin-M1 transitions.

FIGURE 7
www.frontiersin.org

FIGURE 7. Observed distributions of spin-M1 SNMEs. The bars labeled + indicate states with a less confident spin assignment. The figures were taken from Matsubara et al. (2015) with 12C data added.

Additionally, 1–3 (1–7) states were assigned as IS (IV) spin-M1 transitions with less confidence, and they are marked with ”+.” We reassigned 1–6 states as 0+, which were claimed as 1+ in previous studies (Anantaraman et al., 1984; Crawley et al., 1989).

5.2 Quenching of Squared Nuclear Matrix Elements of Isoscalar and Isovector Spin-M1 Transitions

The integrated values of the SNMEs up to Ex=16 MeV were plotted as a function of the target mass in Figure 8 for each of the IS (left panel) and IV (right panel) transitions. Error bars show the full experimental uncertainties, while gray bands show the partial uncertainties originating from the less-confident spin-parity assignment of the transitions (see Section 5.1). Predicted SNMEs by shell-model calculation employing USD interaction (Brown and Wildenthal, 1987) integrated up to 16 MeV is shown by solid lines. Altering the effective interaction to USDA or USDB (Brown and Richter, 2006; Richter et al., 2008) gave only a small change in the prediction (<10%), as shown in Figure 8 for USDB (dashed line). It has been found that the measured values are significantly smaller than the model prediction for the integrated IV SNMEs, while the values are consistent with the model prediction of the integrated IS SNMEs within the experimental uncertainties. We defined the quenching factor as the ratio of the experimentally observed SNMEs integrated up to the experimental limit of 16 MeV to the theoretical predictions integrated to the same excitation energy. The numbers are 1.01(9) and 0.61(6), compared with predictions by the shell model using the USD interaction for the IS and IV spin-M1 transitions, respectively, when the averages of the measured nuclei are used.

FIGURE 8
www.frontiersin.org

FIGURE 8. Integrated values of spin-M1 SNMEs for IS (A) and IV (B) transitions up to Ex=16 MeV as a function of the target mass. Error bars indicate the total experimental uncertainties, and gray bands show partial uncertainties from indefinite spin assignments. Solid and long-dashed lines indicate the prediction of a shell-model calculation using bare USD and bare USDB interactions, respectively. Dotted and dashed-dotted lines applied empirical quenching factors (see text for more details) to USD and USDB interactions, respectively. The short-dashed lines (green) are calculated using the USDB interaction modification in the IS spin-triplet interaction and in the IV quenching factor from Sagawa and Suzuki (2018). The figures were taken from Matsubara et al. (2015) with some updates using the modified interactions (Sagawa and Suzuki, 2018).

The observed quenching factor of the IV spin-M1 transitions is similar to the study of the quenching of the Gamow Teller transitions that is analogous to the IV spin-M1 transitions in terms of isospin symmetry (Anderson et al., 1987). Quenching of the reduced transition probability B(M1) is, in a conventional prescription, implemented by modification of the g-factors (Brown and Wildenthal, 1987) that is multiplied to the SNMEs to obtain B(M1). Here, we applied the same empirical quenching factor as used in the modification of g-factors to the SNMEs calculated by the shell model. The results are shown by the dotted lines (USDBeff) in Figure 8. The empirical quenching factors are similar between IS and IV transitions. SNMEs of the IV spin-M1 transitions with the empirical quenching factor turned out to be compatible with the data.

In contrast, description of the IS transitions became worse by introducing the empirical quenching factor. The present result shows that the widely-used effective g-factors lead to an over-quenching of the IS spin component of the M1transition in the sd-shell. It should be noted that the observed B(M1) strength of the 1+ excited state at 10.23 MeV in 48Ca by (p,p') scattering is more consistent with electron scattering data when no quenching is assumed for the IS part of the transition strength (Birkhan et al., 2016; Mathy et al., 2017). A recent theoretical work (Sagawa et al., 2016; Sagawa and Suzuki, 2018) reported that both the IS and IV SNMEs can be reproduced (short dashed line) by enhancing the IS spin-triplet pairing matrices by a factor of 1.1 in addition to applying the empirical quenching factor to the IV spin-M1 operator but not the IS spin-M1 operator (USDB*qiv). Note that in (Sagawa and Suzuki, 2018) the USDB*qiv result for the IS spin-M1 excitations is not shown, but it is the same as USDB*. The theoretical work implies that the present description of the shell model using USD interactions and the empirical quenching factors may have room to be improved in the IS spin-triplet interaction channel and in the IS quenching factor.

5.3 Isoscalar Magnetic Moments

In this section, we discuss how the new finding of quenching of the IS and IV spin-M1 SNMEs is understood in relation to the IS and IV magnetic moment studied in the past. The magnetic moment is described by the diagonal component of the nuclear matrix element with the relevant M1 operator while the M1 transitions correspond to the off-diagonal components as described in Section 2.

The experimental S data are plotted in the left panel of Figure 9, and the squared values in the right panel. The S values calculated by the shell model with the effective IS g-factors (Brown and Wildenthal, 1983; Brown and Wildenthal, 1987) using the USD interaction are indicated by open circles and crosses for the bare and effective g factors, respectively. The results show clear discrepancies between the experimental data and the prediction with the bare g-factor, corresponding to the quenching, in the edge regions of the sd-shell (A = 17–19 and 35–39) close to the magic numbers of A = 16 and 40. The difference is, however, not obvious in the mid-shell region (A = 21–31). The discrepancies in the edge regions are reduced by introduction of the effective IS g-factors. The IS magnetic moments in the mid-shell region are reasonably reproduced without introducing the effective IS g-factors, which is consistent with our finding of no quenching of the IS spin-M1 SNMEs.

FIGURE 9
www.frontiersin.org

FIGURE 9. The diagonal spin matrix element S obtained from the IS magnetic moments of mirror nuclei in the sd-shell region. The data were taken and calculated from Brown and Wildenthal (1983), Brown and Wildenthal (1987). The lines are for guiding the viewer’s eyes.

5.4 Proton-Neutron Spin-Spin Correlation in the Ground State

Figure 10 shows the Δspin(Ex) data summed up to Ex=16 MeV by the solid bars. Electron scattering data (von Neumann-Cosel et al., 2000) is plotted at A = 12 by the black cross mark with the error bar, which is consistent with the proton scattering data within experimental uncertainty. It is interesting to note that all the experimental data show positive numbers. The Δspin values at A = 4 predicted by Correlated Gaussian (CG) method (Suzuki et al., 2008), an ab initio approach, are plotted for each of the NN interactions of AV8’ (red solid square), G3RS (red open square) and Minnesota (red plus). The predictions by the No-Core Shell-Model (NCSM) (Roth et al., 2010) are plotted for chiral NN (Entem and Machleidt, 2003) (blue open circle) and Minnesota (blue cross) interactions. Both predictions using the Minnesota interaction are consistent with each other and are slightly negative. The Minnesota interaction does not contain tensor interactions. It is illuminating to see that the predictions are positive when more realistic NN interactions are used: AV8’ (Pudliner et al., 1997) and G3RS (Tamagaki, 1968) interactions for the CG method and chiral NN interaction (Entem and Machleidt, 2003) for NCSM. The shell model predictions of Δspin at A = 12 with SFO interaction (solid horizontal line) and A = 20–36 with USD interaction are slightly negative or close to zero, which is significantly smaller than the experimental data of Δspin. The trend of the USDB interaction (dashed line) is similar. The predicted values increase a little but are still much smaller than the experimental results when the IS and IV effective quenching factors determined in the study of the g-factors (Brown and Wildenthal, 1987) are applied to the SNME predicted by the shell model (dot-dashed line). The situation is similar for the other studies of effective g-factors (Towner and Khanna, 1983; Arima et al., 1987; Towner, 1987). The experimental data are reproduced when IS spin-triple pairing matrices are enhanced by a factor of 1.1, and only the IV part of SNME is quenched by effective quenching factors (dotted line) (Sagawa and Suzuki, 2018).

FIGURE 10
www.frontiersin.org

FIGURE 10. Experimental data of Δspin(Ex) summed up to Ex=16 MeV are plotted for proton (electron) scattering by the horizontal bars (cross mark) with the error bars. The definitions of error bars and gray bands are the same as in Figure 8. Theoretical predictions of Δspin=SpSn are plotted for the shell models, Correlated Gaussian (CG) method (Suzuki et al., 2008), and No-Core Shell-Model (NCSM) (Roth et al., 2010). The figure was taken from Matsubara et al. (2015) with updated predictions from Sagawa and Suzuki (2018).

NCSM with the chiral NN interaction predicts Δspin values for N20e (NMAX=4) and M24g (NMAX=2), as indicated by the open blue circles in Figure 10. Here, NMAX defines the maximal allowed harmonic-oscillator excitation energy above the unperturbed ground state (Roth et al., 2010), hence representing a measure of the model space. Variation of the values depending on NMAX shows a clear trend toward positive when NMAX increases: −0.007, 0.028, and 0.072 for NMAX = 0, 2 and 4 for N20e and −0.018 and 0.011 for NMAX=0 and 2 for 24Mg, respectively. Although the values are not converged yet, they are taken as a lower limit in the plot (expressed by arrows). The increase of Δspin with increasing NMAX implies that mixing of higher-lying orbits is important for reproducing the Δspin>0 values.

The observed positive Δspin value implies that deuteron-like correlated np pairs are formed in the ground state of the target nuclei. Note that Δspin takes a value of 1/4 for the IS np pair like a deuteron, −3/4 for the IV np pair, and zero for uncorrelated np pairs. Thus the IS np pairs are favored over the IV np pairs. It would be interesting to see how Δspin values are predicated by ab initio calculations (Gysbers et al., 2019) that reproduce the GT transition strengths studied by beta-decay without a quenching factor by introducing the contribution from the two-body current and many-body correlations. The finding would have relevance to the observed np pair dominance in the correlated NN pairs with high relative momentum in nuclei observed by electron scattering (Subedi et al., 2008; Hen et al., 2014; Hen et al., 2017) or by O16(p,d) reactions (Ong et al., 2013; Terashima et al., 2018). The electron scattering data probe all the components of the correlated NN pairs due to the high incident energy of the electrons without limitation placed on the excitation energy of the residual nucleus after knockout of a correlated NN pair. In contrast, the O16(p,d) data would be relevant to the NN pairs at around the Fermi surface of the target nucleus since the excitation energy of the residual nucleus is limited in the region of several MeV. The present experimental data of Δspin(Ex) are also limited to 16 MeV and thus are considered to be sensitive to the spin-spin correlation in the np pairs at around the Fermi surface. The spin-aligned IS np coupling of the valence particles in 92Pd was studied from the level structure determined by gamma spectroscopy (Cederwall et al., 2011), which also indicates the effect of IS np pairs at around the Fermi surface. It would be interesting to extend the study of Δspin(Ex) to higher-excitation energies to observe how the np spin-spin correlation changes in deeper single particle orbits.

6 Summary

In summary, spin-M1 excitation in nuclei is important for the study of NC neutrino reactions in astrophysical phenomena and in neutrino detectors. Quenching of the IS and IV spin-M1 SNMEs for N=Z sd-shell nuclei has been studied by high energy-resolution measurement of proton inelastic scattering up to the excitation energy of 16 MeV. No quenching of the IS spin-M1 SNMEs has been observed in the measured nuclei, while the IV spin-M1 SNMEs are quenched by an amount comparable with the analogous GT transitions. Consistency with the study of the IS and IV magnetic moment in the same mass region has been discussed. It has been shown by applying the sum rule values that the difference of the IS and IV spin-M1 SNMEs is relevant to the np spin-spin correlation in the ground state. Thus, quenching of the IS and IV spin-M1 SNMEs needs to be described in a mutually consistent manner by theoretical models. The np spin-spin correlation would have a relevance to the correlated np pair with high relative momentum studied by electron scattering and (p,d) reactions as well as to the spin-aligned IS np coupling. It would be of interest to extend the present study to higher excitation energies and to mass and isotope dependencies. For example, decomposition of the spin-M1 strength in the continuum might be applicable by the multipole decomposition analysis of the angular distribution in combination with an isoscalar probe like deuteron scattering or with a pure isovector probe like (p,n). It is worth considering the measurement of (12C,12C*) reactions for studying each of the IS and IV spin-flip excitations of the target nucleus by tagging the IS and IV 1+ states of the ejectile with the coincidence detection of the α or γ emission. Also a measurement in inverse kinematics with an active target based on a time projection chamber would be able to extend the study to larger masses than 40 by employing radioactive secondary beams.

We note that a theoretical work (Isacker and Macchiavelli, 2021) appeared during the review process of this article. That study reported that the positive Δspin values, referred as SnSp in their paper, were not reproduced by Hamiltonian for all the possible parameter values describing neutrons and protons interacting in a single-l shell through a surface delta interaction. Theoretical interpretation of the positive values is still an open question.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors, without undue reservation.

Author Contributions

All authors listed have made a substantial, direct, and intellectual contribution to the work and approved it for publication.

Funding

This work was supported by the JSPS International Training Program (ITP) and was partially supported by JSPS (No. 14740154, and 25105509).

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

The handling editor declared a past co-authorship with one of the authors AT.

Acknowledgments

The authors gratefully acknowledge the collaborators of the E249 and E299 experiments at the RCNP. They are indebted to the RCNP cyclotron staff for providing us with the excellent beam. We are grateful to Sagawa, Nakada, and Ichimura for valuable discussions.

References

Alburger, D. E., and Nathan, A. M. (1978). Beta-Ray Branching and Half-Lives of 12B and 12N. Phys. Rev. C 17, 280–286. doi:10.1103/PhysRevC.17.280

CrossRef Full Text | Google Scholar

Anantaraman, N., Brown, B. A., Crawley, G. M., Galonsky, A., Djalali, C., Marty, N., et al. (1984). Observation of Quenching in Isoscalar and Isovector 0+ → 1+ Transitions in Si28(p,p′). Phys. Rev. Lett. 52, 1409–1412. doi:10.1103/PhysRevLett.52.1409

CrossRef Full Text | Google Scholar

Anderson, B. D., Chittrakarn, T., Baldwin, A. R., Lebo, C., Madey, R., Tandy, P. C., et al. (1987). Gamow-Teller and M1 Strength in the 32S(p,n)32Cl Reaction at 135 Mev. Phys. Rev. C. Nucl. Phys. 36, 2195–2205. doi:10.1103/physrevc.36.2195

PubMed Abstract | CrossRef Full Text | Google Scholar

Arima, A., Shimizu, K., Bentz, W., and Hyuga, H. (1987). Nuclear Magnetic Properties and Gamow-Teller Transitions. Adv. Nucl. Phys. 18, 1–106.

Google Scholar

Bendel, W. L., Fagg, L. W., Numrich, S. K., Jones, E. C., and Kaiser, H. F. (1971). Excitation of by 180 Electron Scattering. Phys. Rev. C 3, 1821–1827. doi:10.1103/PhysRevC.3.1821

CrossRef Full Text | Google Scholar

Bernabéu, J., Ericson, T., Hernández, E., and Ros, J. (1992). Effects of the Axial Isoscalar Neutral-Current for Solar Neutrino Detection. Nucl. Phys. B 378, 131–149. doi:10.1016/0550-3213(92)90006-W

CrossRef Full Text | Google Scholar

Birkhan, J., Matsubara, H., von Neumann-Cosel, P., Pietralla, N., Ponomarev, V. Y., Richter, A., et al. (2016). Electromagnetic M 1 Transition Strengths from Inelastic Proton Scattering: The Cases of ca 48 and pb 208. Phys. Rev. C 93, 041302. doi:10.1103/PhysRevC.93.041302

CrossRef Full Text | Google Scholar

Brown, B., and Rae, W. (2014). The Shell-Model Code Nushellx@msu. Nucl. Data Sheets 120, 115–118. doi:10.1016/j.nds.2014.07.022

CrossRef Full Text | Google Scholar

Brown, B., and Wildenthal, B. (1987). Empirically Optimum m1 Operator for sd-Shell Nuclei. Nucl. Phys. A 474, 290–306. doi:10.1016/0375-9474(87)90619-1

CrossRef Full Text | Google Scholar

Brown, B. A., and Richter, W. A. (2006). New “usd” Hamiltonians for the sd Shell. Phys. Rev. C 74, 034315. doi:10.1103/PhysRevC.74.034315

CrossRef Full Text | Google Scholar

Brown, B. A., and Wildenthal, B. H. (1983). Corrections to the Free-Nucleon Values of the Single-Particle Matrix Elements of the m1 and Gamow-Teller Operators, From a Comparison of Shell-Model Predictions With -Shell Data. Phys. Rev. C 28, 2397–2413. doi:10.1103/PhysRevC.28.2397

CrossRef Full Text | Google Scholar

Cederwall, B., Moradi, F. G., Bäck, T., Johnson, A., Blomqvist, J., Clément, E., et al. (2011). Evidence for a Spin-Aligned Neutron–Proton Paired Phase From the Level Structure of 92 pd. Nature 469, 68–71. doi:10.1038/nature09644

PubMed Abstract | CrossRef Full Text | Google Scholar

Cohen, S., and Kurath, D. (1965). Effective Interactions for the 1p Shell. Nucl. Phys. 73, 1–24. doi:10.1016/0029-5582(65)90148-3

CrossRef Full Text | Google Scholar

Cooper, E. D., Hama, S., and Clark, B. C. (2009). Global Dirac Optical Potential from Helium to lead. Phys. Rev. C 80, 034605. doi:10.1103/PhysRevC.80.034605

CrossRef Full Text | Google Scholar

Crawley, G. M., Djalali, C., Marty, N., Morlet, M., Willis, A., Anantaraman, N., et al. (1989). Isovector and Isoscalar Spin-Flip Excitations in Even-Even s-d Shell Nuclei Excited by Inelastic Proton Scattering. Phys. Rev. C 39, 311–323. doi:10.1103/PhysRevC.39.311

CrossRef Full Text | Google Scholar

Edmonds, A. (1960). Angular Momentum in Quantum Mechanics. Princeton, NJ: Princeton University Press.

Entem, D. R., and Machleidt, R. (2003). Accurate Charge-Dependent Nucleon-Nucleon Potential at Fourth Order of Chiral Perturbation Theory. Phys. Rev. C 68, 041001. doi:10.1103/PhysRevC.68.041001

CrossRef Full Text | Google Scholar

Firestone, R. (1996). Table of Isotopes. 8th Edn. New York: John Wiley & Sons.

Foltz, C. W., Sober, D. I., Fagg, L. W., Gräf, H. D., Richter, A., Spamer, E., et al. (1994). Electroexcitation of Low-Multipolarity Magnetic Transitions in 36Ar and 38Ar. Phys. Rev. C 49, 1359–1371. doi:10.1103/PhysRevC.49.1359

CrossRef Full Text | Google Scholar

Franey, M. A., and Love, W. G. (1985). Nucleon-Nucleon t-Matrix Interaction for Scattering at Intermediate Energies. Phys. Rev. C 31, 488–498. doi:10.1103/PhysRevC.31.488

CrossRef Full Text | Google Scholar

Fujita, H., Fujita, Y., Adachi, T., Bacher, A. D., Berg, G. P. A., Black, T., et al. (2007). Isospin Structure of States in 58Ni and 58Cu Studied by 58Ni(p, p′) and 58Ni(3He,t)58Cu Measurements. Phys. Rev. C 75, 034310. doi:10.1103/PhysRevC.75.034310

CrossRef Full Text | Google Scholar

Fujita, H., Fujita, Y., Berg, G., Bacher, A., Foster, C., Hara, K., et al. (2002). Realization of Matching Conditions for High-Resolution Spectrometers. Nucl. Instr. Methods Phys. Res. A: Acc. Spectr. Detect. Assoc. Equip. 484, 17–26. doi:10.1016/S0168-9002(01)01970-2

CrossRef Full Text | Google Scholar

Fujita, Y., Brown, B. A., Ejiri, H., Katori, K., Mizutori, S., and Ueno, H. (2000). Separation of Isoscalar, Isovector, Orbital, and Spin Contributions in Transitions in Mirror Nuclei. Phys. Rev. C 62, 044314. doi:10.1103/PhysRevC.62.044314

CrossRef Full Text | Google Scholar

Fujita, Y., Fujita, H., Adachi, T., Berg, G. P. A., Caurier, E., Fujimura, H., et al. (2002). Gamow-Teller Transitions From 58Ni to Discrete States of 58Cu - The Study of Isospin Symmetry in Atomic Nuclei. Eur. Phys. J. A. 13, 411–418. doi:10.1140/epja/iepja1344

CrossRef Full Text | Google Scholar

Fujita, Y., Hatanaka, K., Berg, G., Hosono, K., Matsuoka, N., Morinobu, S., et al. (1997). Matching of a Beam Line and a Spectrometer New Beam Line Project at rcnp. Nucl. Instr. Methods Phys. Res. B: Beam Inter. Mater. Atoms 126, 274–278. doi:10.1016/S0168-583X(96)01008-7

CrossRef Full Text | Google Scholar

Fujita, Y., Rubio, B., and Gelletly, W. (2011). Spin-Isospin Excitations Probed by Strong, Weak and Electro-Magnetic Interactions. Prog. Part. Nucl. Phys. 66, 549–606. doi:10.1016/j.ppnp.2011.01.056

CrossRef Full Text | Google Scholar

Fujiwara, M., Akimune, H., Daito, I., Fujimura, H., Fujita, Y., Hatanaka, K., et al. (1999). Magnetic Spectrometer Grand Raiden. Nucl. Instr. Methods Phys. Res. A 422, 484–488. doi:10.1016/S0168-9002(98)01009-2

CrossRef Full Text | Google Scholar

Gysbers, P., Hagen, G., Holt, J., Jansen, G. R., Morris, T. D., Navrátil, P., et al. (2019). Discrepancy Between Experimental and Theoretical β-Decay Rates Resolved From First Principles. Nat. Phys. 15, 428–431. doi:10.1038/s41567-019-0450-7

CrossRef Full Text | Google Scholar

Hen, O., Miller, G. A., Piasetzky, E., and Weinstein, L. B. (2017). Nucleon-Nucleon Correlations, Short-Lived Excitations, and the Quarks Within. Rev. Mod. Phys. 89, 045002. doi:10.1103/RevModPhys.89.045002

CrossRef Full Text | Google Scholar

Hen, O., Sargsian, M., Weinstein, L. B., Piasetzky, E., Hakobyan, H., Higinbotham, D. W., et al. (2014). Momentum Sharing in Imbalanced Fermi Systems. Science 346, 614–617. doi:10.1126/science.1256785

PubMed Abstract | CrossRef Full Text | Google Scholar

Hofmann, F., von Neumann-Cosel, P., Neumeyer, F., Rangacharyulu, C., Reitz, B., Richter, A., et al. (2002). Magnetic Dipole Transitions in 32S From Electron Scattering at 180. Phys. Rev. C 65, 024311. doi:10.1103/PhysRevC.65.024311

CrossRef Full Text | Google Scholar

Ichimura, M., Sakai, H., and Wakasa, T. (2006). Spin-Isospin Responses via (p,n) and (n,p) Reactions. Prog. Part. Nucl. Phys. 56, 446–531. doi:10.1016/j.ppnp.2005.09.001

CrossRef Full Text | Google Scholar

Ikeda, K., Fujii, S., and Fujita, J. (1963). The (p,n) Reactions and Beta Decays. Phys. Lett. 3, 271–272. doi:10.1016/0031-9163(63)90255-5

CrossRef Full Text | Google Scholar

Isacker, P. V., and Macchiavelli, A. (2021). Neutron-proton Spin-Spin Correlations in the Ground States of Nuclei. Eur. Phys. J. 57, 178. doi:10.1140/epja/s10050-021-00489-6

CrossRef Full Text | Google Scholar

Kawabata, T., Akimune, H., Fujimura, H., Fujita, H., Fujita, Y., Fujiwara, M., et al. (2004). Isovector and Isoscalar Spin-Flip Strengths in 11B. Phys. Rev. C 70, 034318. doi:10.1103/PhysRevC.70.034318

CrossRef Full Text | Google Scholar

Kirson, M. W. (2007). Oscillator Parameters in Nuclei. Nucl. Phys. A 781, 350–362. doi:10.1016/j.nuclphysa.2006.10.077

CrossRef Full Text | Google Scholar

Love, W. G., and Franey, M. A. (1981). Effective Nucleon-Nucleon Interaction for Scattering at Intermediate Energies. Phys. Rev. C 24, 1073–1094. doi:10.1103/PhysRevC.24.1073

CrossRef Full Text | Google Scholar

Lüttge, C., Neumann-Cosel, Pv., Neumeyer, F., Rangacharyulu, C., Richter, A., Schrieder, G., et al. (1996). Isovector m1 Transitions in and the Role of Meson Exchange Currents. Phys. Rev. C 53, 127–130. doi:10.1103/PhysRevC.53.127

CrossRef Full Text | Google Scholar

Mathy, M., Birkhan, J., Matsubara, H., von Neumann-Cosel, P., Pietralla, N., Ponomarev, V. Y., et al. (2017). Search for Weak m 1 Transitions in ca 48 With Inelastic Proton Scattering. Phys. Rev. C 95, 054316. doi:10.1103/PhysRevC.95.054316

CrossRef Full Text | Google Scholar

Matsubara, H. (2010). Isoscalar and isovector spin-M1 transitions from the even-even, N = Z nuclei across the sd-shell region. PhD thesis. Toyonaka: Osaka University.

Matsubara, H., Sakaguchi, H., Kishi, T., and Tamii, A. (2009). Self-supporting Elemental Sulfur Target for Charged Particle Irradiation. Nucl. Instr. Methods Phys. Res. B 267, 3682–3687. doi:10.1016/j.nimb.2009.09.002

CrossRef Full Text | Google Scholar

Matsubara, H., Tamii, A., Nakada, H., Adachi, T., Carter, J., Dozono, M., et al. (2015). Nonquenched Isoscalar Spin- Excitations in -Shell Nuclei. Phys. Rev. Lett. 115, 102501. doi:10.1103/PhysRevLett.115.102501

PubMed Abstract | CrossRef Full Text | Google Scholar

Matsubara, H., Tamii, A., Shimizu, Y., Suda, K., Tameshige, Y., and Zenihiro, J. (2012). Wide-Window Gas Target System for High Resolution experiment with Magnetic Spectrometer. Nucl. Instr. Methods Phys. Res. A 678, 122–129. doi:10.1016/j.nima.2012.03.005

CrossRef Full Text | Google Scholar

Ong, H., Tanihata, I., Tamii, A., Myo, T., Ogata, K., Fukuda, M., et al. (2013). Probing Effect of Tensor Interactions in 16o via (p,d) Reaction. Phys. Lett. B 725, 277–281. doi:10.1016/j.physletb.2013.07.038

CrossRef Full Text | Google Scholar

Petrovich, F., and Love, W. (1981). The Scattering of Elementary Probes From Nuclei at Medium Energy: A New Look at the Nucleus. Nucl. Phys. A 354, 499–534. doi:10.1016/0375-9474(81)90613-8

CrossRef Full Text | Google Scholar

Pudliner, B. S., Pandharipande, V. R., Carlson, J., Pieper, S. C., and Wiringa, R. B. (1997). Quantum Monte Carlo Calculations of Nuclei With A <= 7. Phys. Rev. C 56, 1720–1750. doi:10.1103/PhysRevC.56.1720

CrossRef Full Text | Google Scholar

Raynal, J. (2007). Computer Code dwba07. Report No. NEA-1209/008.

Google Scholar

Richter, A., Weiss, A., Haüsser, O., and Brown, B. A. (1990). New Evidence for Meson-Exchange-Current Enhancement of Isovector M1 Strength. Phys. Rev. Lett. 65, 2519–2522. doi:10.1103/PhysRevLett.65.2519

PubMed Abstract | CrossRef Full Text | Google Scholar

Richter, W. A., Mkhize, S., and Brown, B. A. (2008). sd-Shell Observables for the usda and usdb Hamiltonians. Phys. Rev. C 78–064302. doi:10.1103/PhysRevC.78.064302

CrossRef Full Text | Google Scholar

Roth, R., Neff, T., and Feldmeier, H. (2010). Nuclear Structure in the Framework of the Unitary Correlation Operator Method. Prog. Part. Nucl. Phys. 65, 50–93. doi:10.1016/j.ppnp.2010.02.003

CrossRef Full Text | Google Scholar

Sagawa, H., and Suzuki, T. (2018). Isoscalar and Isovector Spin Response in -Shell Nuclei. Phys. Rev. C 97, 054333. doi:10.1103/PhysRevC.97.054333

CrossRef Full Text | Google Scholar

Sagawa, H., Suzuki, T., and Sasano, M. (2016). Effect of Isoscalar Spin-Triplet Pairings on Spin-Isospin Responses in s d-Shell Nuclei. Phys. Rev. C 94, 041303. doi:10.1103/PhysRevC.94.041303

CrossRef Full Text | Google Scholar

Sasano, M., Sakai, H., Yako, K., Wakasa, T., Asaji, S., Fujita, K., et al. (2009). Gamow-teller Unit Cross Sections of the Reaction at 198 and 297 mev on Medium-Heavy Nuclei. Phys. Rev. C 79, 024602. doi:10.1103/PhysRevC.79.024602

CrossRef Full Text | Google Scholar

Subedi, R., Shneor, R., Monaghan, P., Anderson, B., Aniol, K., Annand, J., et al. (2008). Probing Cold Dense Nuclear Matter. Science 320, 1476–1478. doi:10.1126/science.1156675

PubMed Abstract | CrossRef Full Text | Google Scholar

Suzuki, T., Fujimoto, R., and Otsuka, T. (2003). Gamow-Teller Transitions and Magnetic Properties of Nuclei and Shell Evolution. Phys. Rev. C 67, 044302. doi:10.1103/PhysRevC.67.044302

CrossRef Full Text | Google Scholar

Suzuki, Y., Horiuchi, W., Orabi, M., and Arai, K. (2008). Global-vector Representation of the Angular Motion of Few-Particle Systems ii. Few-Body Syst. 42. doi:10.1007/s00601-008-0200-3

CrossRef Full Text | Google Scholar

Taddeucci, T., Goulding, C., Carey, T., Byrd, R., Goodman, C., Gaarde, C., et al. (1987). The (p,n) Reaction as a Probe of Beta Decay Strength. Nucl. Phys. A 469, 125–172. doi:10.1016/0375-9474(87)90089-3

CrossRef Full Text | Google Scholar

Taddeucci, T. N., Byrd, R. C., Carey, T. A., Ciskowski, D. E., Foster, C. C., Gaarde, C., et al. (1990). Gamow-teller Transition Strengths from the 11B(p,nc) Reaction in the Energy Range 160–795 mev. Phys. Rev. C 42, 935–946. doi:10.1103/PhysRevC.42.935

CrossRef Full Text | Google Scholar

Tamagaki, R. (1968). Potential Models of Nuclear Forces at Small Distances. Prog. Theor. Phys. 39, 91–107. doi:10.1143/PTP.39.91

CrossRef Full Text | Google Scholar

Tamii, A., Akimune, H., Daito, I., Fujita, Y., Fujiwara, M., Hatanaka, K., et al. (1999). Polarization Transfer Observables for Proton Inelastic Scattering from 12c at 0. Phys. Lett. B 459, 61–66. doi:10.1016/S0370-2693(99)00654-1

CrossRef Full Text | Google Scholar

Tamii, A., Fujita, Y., Matsubara, H., Adachi, T., Carter, J., Dozono, M., et al. (2009). Measurement of High Energy Resolution Inelastic Proton Scattering at and Close to Zero Degrees. Nucl. Instrum.Methods Phys. Res. Sect. A. 605, 326–338. doi:10.1016/j.nima.2009.03.248

CrossRef Full Text | Google Scholar

Terashima, S., Yu, L., Ong, H. J., Tanihata, I., Adachi, S., Aoi, N., et al. (2018). Dominance of Tensor Correlations in High-Momentum Nucleon Pairs Studied by Reaction. Phys. Rev. Lett. 121, 242501. doi:10.1103/PhysRevLett.121.242501

PubMed Abstract | CrossRef Full Text | Google Scholar

Towner, I., and Khanna, F. (1983). Corrections to the Single-Particle m1 and Gamow-Teller Matrix Elements. Nucl. Phys. A 399, 334–364. doi:10.1016/0375-9474(83)90252-X

CrossRef Full Text | Google Scholar

Towner, I. (1987). Quenching of Spin Matrix Elements in Nuclei. Phys. Rep. 155, 263–377. doi:10.1016/0370-1573(87)90138-4

CrossRef Full Text | Google Scholar

von Neumann-Cosel, P., and Tamii, A. (2019). Electric and Magnetic Dipole Modes in High-Resolution Inelastic Proton Scattering at 0. Eur. Phys. J. A 55. doi:10.1140/epja/i2019-12781-7

CrossRef Full Text | Google Scholar

von Neumann-Cosel, P., Gräf, H. D., Krämer, U., Richter, A., and Spamer, E. (2000). Electroexcitation of Isoscalar and Isovector Magnetic Dipole Transitions in 12c and Isospin Mixing. Nucl. Phys. A 669 3–13. doi:10.1016/S0375-9474(99)00564-3

CrossRef Full Text | Google Scholar

Wakasa, T., Hatanaka, K., Fujita, Y., Berg, G., Fujimura, H., Fujita, H., et al. (2002). High Resolution Beam Line for the Grand Raiden Spectrometer. Nucl. Instr. Methods Phys. Res. A 482, 79–93. doi:10.1016/S0168-9002(01)01686-2

CrossRef Full Text | Google Scholar

Warburton, E. K., and Millener, D. J. (1989). Structure of 17C and 17N. Phys. Rev. C 39, 1120–1129. doi:10.1103/PhysRevC.39.1120

CrossRef Full Text | Google Scholar

Zegers, R. G. T., Akimune, H., Austin, S. M., Bazin, D., van der Berg, A. M., Berg, G. P. A., et al. (2006). The (t,3He) and (3He,t) Reactions as Probes of Gamow-Teller Strength. Phys. Rev. C 74, 024309. doi:10.1103/PhysRevC.74.024309

CrossRef Full Text | Google Scholar

Keywords: Sd-shell nuclei, shell-model calculation, proton inelastic scattering, spin-M1 transition, GT-transition

Citation: Matsubara H and Tamii A (2021) Quenching of Isovector and Isoscalar Spin-M1 Excitation Strengths in N = Z Nuclei. Front. Astron. Space Sci. 8:667058. doi: 10.3389/fspas.2021.667058

Received: 11 February 2021; Accepted: 14 June 2021;
Published: 14 July 2021.

Edited by:

Francesco Cappuzzello, University of Catania, Italy

Reviewed by:

Marco Martini, Institut Polytechnique des Sciences Avancées, France
Zhenbin Wu, University of Illinois at Chicago, United States

Copyright © 2021 Matsubara and Tamii. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Hiroaki Matsubara, mats.hiroaki@gmail.com

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.