logo

Bρ-defined isochronous mass spectrometry at the storage ring CSRe

Special Issue: Dedicated to Professor Wenqing Shen in Honour of his 80th Birthday

Bρ-defined isochronous mass spectrometry at the storage ring CSRe

Xu Zhou
Meng Wang
Yu-Hu Zhang
Xiao-Hong Zhou
Xin-Liang Yan
Yuan-Ming Xing
Nuclear Science and TechniquesVol.35, No.12Article number 213Published in print Dec 2024Available online 25 Nov 2024
17203

A novel technique of isochronous mass spectrometry (IMS), termed -defined IMS, was developed at the experimental cooler-storage ring CSRe in Lanzhou for the first time. Two time-of-flight detectors were installed in a straight section of the CSRe, thereby enabling simultaneous measurements of the velocity and revolution time of each stored short-lived ion. This technique boosts the broadband precision, efficiency, sensitivity, and accuracy of mass measurements of short-lived exotic nuclides. Using -defined IMS, the masses of 22Al, 62Ge, 64As, 66Se, and 70Kr were measured for the first time, and the masses of 65As, 67Se, and other 21 nuclides were redetermined with improved accuracy. Mass data have been used in studies of relevant issues regarding nuclear structures and nuclear astrophysics. Herein, we review the development of experimental techniques and main physical results and outline plans for future experiments.

Nuclear massStorage ringBρ-defined isochronous mass spectrometryNuclear structureNuclear astrophysics
1

Introduction

The mass of an atomic nucleus is a fundamental property that reflects the total energy of this quantum many-body system, which is composed of two types of fermions: protons and neutrons. Systematic and accurate knowledge of nuclear masses has wide applications in many areas of subatomic physics, ranging from nuclear structures to nuclear astrophysics and fundamental interactions and symmetries, depending on the mass precision achieved [1, 2]. For example, based on nuclear masses, well-known shell structure and pairing correlations were discovered in stable nuclei [3], and the disappearance of the magic neutron number N = 20 [4] and a new shell closure at N = 32 [5] were revealed in exotic neutron-rich nuclides. In addition to the mapping of the nuclear-mass surface [6-8], considerable attention has been paid to the precise mass measurements of exotic nuclei in specific mass regions, such as in the vicinity of doubly magic nuclei far from stability and the waiting-point nuclei in the rapid proton and rapid neutron capture processes.

Owing to more than a century of effort, the masses of approximately 2550 nuclides [9, 10] have been measured, as shown in Fig. 1. Currently, nuclides with unknown masses lie far from the valley of βstability, close to the borders of nuclear existence. Such nuclides are inevitably short-lived and have very low production yields, making their mass measurements extremely challenging. The performance of mass spectrometry of short-lived nuclei must improve in terms of sensitivity and precision.

Fig. 1
(Color online) Nuclear chart displaying the mass-excess uncertainties in the latest atomic-mass evaluation AME2020 [10] for all nuclei in their ground state
pic

Penning-trap mass spectrometry is widely used to deliver precise nuclear masses. Penning traps have produced numerous important results [11, 12]. However, in this application, a certain restriction threshold exists for the half-lives and/or production rates of the investigated nuclides. In recent years, multireflection time-of-flight spectrometry has emerged as a powerful technique for studying nuclei with short half-lives and low production rates [5, 13, 14]. However, in addition to the measurement time itself, applying these devices requires preparation steps such as cooling and bunching of low-energy radioactive species [15]. At high energies, mass spectrometers coupled directly to in-flight separators do not require much time for ion preparation. Magnetic-rigidity time-of-flight (-TOF) mass spectrometry, implemented at radioactive-ion beamlines, has produced the masses of nuclides farthest away from the stability valley [16, 17], albeit with modest precision.

Isochronous mass spectrometry (IMS) in heavy-ion storage rings is an efficient and fast experimental technique [18] that is well suited for mass measurements of exotic nuclei with short lifetimes down to several tens of microseconds. Since the pioneering experiments conducted at the ESR in GSI, Darmstadt [19, 20], IMS has been established at the experimental cooler storage ring (CSRe) at the Institute of Modern Physics (IMP), Lanzhou [21, 22] and at the Rare RI Ring (R3) at the RIKEN Nishina Center in Wako/Japan [23].

IMS is typically used for mass measurements of heavy neutron-rich nuclides produced by the in-flight fission of uranium beams at the ESR [24, 25] and R3 [26-28]; see the review articles and references cited in [29, 15].

In the experiments, the isochronous mode of the storage ring was applied to achieve a relatively high resolving power for ions fulfilling the isochronous condition. However, for other ions that do not fulfill isochronicity, the mass resolution deteriorates rapidly. To improve the mass-resolving power over a broad m/q range, magnetic-rigidity tagging (-tagging) IMS was realized at the FRS-ESR facility at GSI [30, 31, 24] by inserting metal slits at the second dispersive focal plane of the FRS [32]. In the CSRe, an in-ring slit is used to restrict the magnetic-rigidity () acceptance of the stored fragments [33]. However, both approaches have the significant limitation of losing transmission efficiency, which is not tolerable for mass measurements of exotic nuclei with very low production yields.

Recently, a new technique, termed -defined IMS, has been developed [34, 35], and a mass-resolving power of 1.3×105 (full width at half maximum; FWHM) can be achieved at the edges of the mass-to-charge ratio spectrum without losing any ions of interest [34]. Using this technique, the masses of 22Al, 62Ge, 64As, 66Se, and 70Kr were measured for the first time, and the masses of 65As, 67Se, and other 21 nuclides were redetermined with improved accuracy. Based on these newly measured masses, investigations on different issues regarding nuclear structures and nuclear astrophysics based on these newly measured masses are also presented in this article.

2

-defined IMS

2.1
Principle

In conventional IMS, the ion-revolution times Texp are measured using a single TOF detector [36, 37]. The mass-to-charge ratios of the ions, m/q, are derived according to [19, 20]: mq=Bργv=BρTexp2C21vc2, (1) where , v, γ, and C are the magnetic rigidity, velocity, Lorenzt factor, and orbit length of the stored ions, respectively. vc is the speed of light. and C are assumed to be the same for all ions and the masses of aimed ions can be deduced using ions with known masses as calibrants.

In real experiments, the momentum acceptance of the storage ring should be considered, and the dispersed and C of the stored ions should follow the (C) function. The relative slope of (C) ΔBρ/BρΔC/C=γt2, (2) introduces a new quantity, γt, called the transition point of the storage ring [38]. γt is determined solely by the machine lattice settings and is independent of the ion species.

Owing to the momentum dispersion of the stored ions, the variation in the revolution times is as follows: ΔTexpTexp=ΔCCΔvv=(1γt21γ2)ΔBρBρ. (3) In the experiments, the storage ring was tuned to be isochronous, that is, γγt, for the ions of interest, and a high mass-resolving power could be achieved in a small region of the m/q spectrum. For other ions with γγt, the mass-resolving power decreased rapidly.

To further reduce the influence of momentum dispersion and achieve broadband high mass-resolving power, two TOF detectors are installed in one of the straight sections of the ring such that both the revolution times Texp and velocities vexp of the stored ions [39] can be measured simultaneously. Consequently, the orbital length of each ion Cexp=Texp×vexp can be deduced directly.

Using ions with known masses as calibrants, a one-to-one correspondence of BρC is obtained, providing more information on the momentum dispersion. By fitting the BρC data, we obtain the Bρfit(Cexp) function, and the m/q value of any ion can be deduced from the fitted function as follows: mq=Bρfit(Cexp)(γv)exp. (4) In the simulations, the uncertainties of m/q values are mainly due to the uncertainties of the velocity measurements and instability of γt [40-42]. A data-analysis method was developed to achieve -defined IMS at the storage ring CSRe in Lanzhou.

2.2
Technique details

The experiments were conducted at the accelerator complex of the IMP in Lanzhou, China. The nuclides of interest were produced by fragmenting the primary beams at a speed of >0.7× vc on 1015 mm 9Be production targets. The produced nuclides were selected using the in-flight fragment separator, the second Radiactive Isotope Beam Line in Lanzhou (RIBLL2) [43]. They were then injected into and stored in the CSRe.

Figure 2 presents a schematic of the CSRe with the three TOF detectors noted in this figure. The detector TOF0 was used in the conventional IMS. The detectors TOF1 and TOF2 were used in -defined IMS. Each detector consists of a thin carbon foil (ϕ=40 mm, 18 μg/cm2 thick) and a set of microchannel plates (MCP) [36, 37].

Fig. 2
(Color online) Schematic of CSRe with the arrangement of three TOF detectors [34]
pic

When an ion passes through the carbon foil, secondary electrons are released from the foil surface and guided to the MCP by perpendicularly arranged electric and magnetic fields. Compared to the previous detector TOF0, the electric-field strengths in TOF1 and TOF2 increased from 130 to 180 V/mm to improve the time resolution of the detector, as shown in Fig. 3 (a). A time resolution of 18.5±2 ps was achieved in the offline tests of TOF1 and TOF2 [37]. Fast timing signals from the two MCPs were recorded using a single oscilloscope to avoid the jitter effects of two or more electronic instruments [45]. The sampling rate was set to 50 GHz.

Fig. 3
(Color online) (a) Time resolution of the TOF detector as a function of the electric-field strength. [37]. (b) Dispersion functions used in conventional IMS (black solid line) and -defined IMS (red dashed line) as a function of the longitudinal position s of CSRe [44]. The positions of the three TOF detectors are denoted using dotted lines. (c) The transition energy γt as a function of orbital length due to isochronous corrections of quadrupole magnets and sextupole magnets [44]
pic

To achieve -defined IMS, a new isochronous mode setting was specially designed at the CSRe [44]. Figure 3 (b) shows the dispersion function Dx=xΔp/p as a function of the longitudinal position s of the CSRe. The dispersion function at the detector positions was designed to be close to 0 m for a high transition efficiency of the ions owing to the limited size of the carbon foils (ϕ=40 mm). To ensure the accuracy of the detector installation, a calibration system using a short-pulsed ultraviolet laser was also applied at the CSRe [46].

In real experiments, γt is not constant for all orbit lengths owing to high-order magnetic fields from imperfect storage-ring facilities [47, 44, 48]. A minor modification of the quadrupole and sextupole fields was performed to reduce the variation in γt as a function of the orbit length C [44], as shown in Fig 3 (c). However, the variation in γt within the momentum acceptance cannot be eliminated completely in a real experiment. The residual nonlinear effects in the γt(C) curve should be considered in data analysis.

2.3
Velocity measurement of stored ions

Figure 4 shows the passage times texp of a single 45V23+ ion as a function of revolution number N for TOF1 and TOF2 [39]. The slope texp/N is the revolution time Texp, and the distance between the two curves is the time of flight (TOF) Δtexp between the two detectors. The mean velocity of an ion circulating in the ring v is obtained by v=LΔtexpΔtd, (5) where L is the distance between the two TOF detectors in the straight section of the CSRe and Δtd=tdTOF2tdTOF1 is the time-delay difference between the two timing signals transmitted from the detectors to the oscilloscope. The values of L and Δtd were measured in a dedicated study using a laser-calibration system [46] and were redetermined in each experiment using stored ions with known masses as calibrants [35]. The precise velocity measurement of a stored ion critically depends on the precision determination of Δtexp between TOF1 and TOF2.

Fig. 4
(Color online) The passage times texp of a single 45V23+ measured by detectors TOF1 and TOF2 as a function of the revolution number N [39]. For illustration purposes, the time stamps from TOF2 are shifted upwards by 10 μs. The inset illustrates an enlarged region at 299≤N≤309 showing the missing time stamps in some revolutions. The error bar of each data point is within the symbol size
pic

Data analyses of conventional IMS usually show that the flight time of an ion is a smooth function of the revolution number. Considering the energy loss in a thin carbon foil, texp(N) can be described by a third-order polynomial function: tfit(N)=a0+a1×N+a2×N2+a3×N3. (6) In addition, the influence of the betatron oscillations of the ions should be considered [39], and the fit function of texp(N) curves should be modified as follows: tfit(N)=a0+a1×N+a2×N2+a3×N3+Ax×[2π(Qx0×N+Qx1×N2)+ϕx]+Ay×[2π(Qy0×N+Qy1×N2)+ϕy]. (7) The polynomial function in Eq. (7) describes ion motion with a mean orbital length, whereas the sinelike terms describe the periodic time fluctuations owing to betatron oscillations.

A straightforward method to deduce Δtexp is to use the existing coincident timestamps from both TOF detectors in the same revolution. Alternatively, a fitting procedure using Eq. (6) and Eq. (7) have been performed, and Fig. 5 shows the average uncertainties of Δtexp as a function of the atomic number Z using these three methods.

Fig. 5
(Color online) Average uncertainty of Δtexp as a function of atomic number Z, extracted from coincident time stamps (black filled circles) and by using the third-order polynomial fit function with [Eq. (7)] (blue filled triangles) and without [Eq. (6)] (red opened circles) the sinelike terms
pic

The smallest mean value of σtexp) has been obtained using Eq. (7). This indicates that considering and addressing the effects of betatron oscillations can significantly improve the precision of the Δtexp values as well as the ion velocities. The mean uncertainties exhibited a decreasing trend from lighter to heavier ions. This can be understood if the Z-dependence of the detection efficiency of the detectors is considered; that is, a higher detection efficiency yields lower statistical errors in Δtexp. The lowest mean uncertainties of Δtexp were within the range of 2.0-6.4 ps, corresponding to a relative precision of (2.2-7.2)×10-5 with respect to the average TOF between the two TOF detectors of Δtexp89 ns. Compared with the momentum acceptance ±0.2% of the CSRe [49], the precision of Δtexp is sufficient for new IMS development.

2.4
Determination of mass values

Using ions with known masses and the measured revolution times Texp and velocities vexp, the magnetic rigidity exp=m/q× (γ v)exp and orbit length Cexp=Texp×vexp can be easily extracted. Fig. 6 shows a scatterplot of exp versus Cexp [34, 35].

Fig. 6
(Color online) Plot of exp versus Cexp with the fit results (solid line) [35]. The fitted expression is noted in this figure
pic

For an ideal storage ring with a completely constant γt within the momentum accpetance, the one-to-one correspondence of C should follow the simple form: Bρ(C)=Bρ0(CC0)γt2, (8) where 0 and C0 are reference parameters.

In real experiments, the nonlinear effects in the γt(C) curve shown in Fig. 3 (c) should be considered. Because these residual nonlinear effects may differ in each experiment, Eq. (8) is modified by including additional terms. For example, Bρ(C)=Bρ0(CC0)K+a1ea2(CC0) (9) was used in the experiments described in Refs. [34, 35]. The fit result of Eq. (9) is shown in Fig. 6. The m/q spectrum was deduced from the fitted results.

The instability of magnetic fields is one of the most significant challenges in precision mass measurements at the CSRe storage ring. Over the past decade, extensive efforts have been dedicated to both hardware improvements [50] and data-analysis enhancements [51, 52], aimed at mitigating the impact of magnetic-field drift. Using data from the two TOF detectors, we developed a more precise correction method [35].

The changes in the dipole magnetic fields led to vertical drifts in the measured (C) curve. This phenomenon can be clearly observed in Fig. 7 (a), where the fit residuals BρexpiBρ(Cexpi) are presented as functions of the injection number. To achieve the high mass-resolving power, the originally determined quantities {Cexpi,Texpi,vexpi, i=1,2,...,Ns} need to be corrected to {Ccori,Tcori,vcori, i=1,2,...,Ns} corresponding to a reference magnetic-field setting with B0 assuming the magnetic field is stable. To obtain the latter, Cexpi=Ccori must be confined (equivalent to a common radius ρexpi=ρcori). Consequently, the magnetic rigidity and velocity of each ion in the same injection are scaled by a constant ratio: Bρcori=1RMBρexpior(γv)cori=1RM(γv)expi, (10) where RM=(BB0)CSRe=(BρexpiBρ(Cexpi))ion. (11) The ratio RM can be deduced from the experimental data of reference ions with known masses for each injection. Figure 7 (b) shows that the fitted residuals after correction produce BρcoriBρ(Cexpi) as a function of the injection number. The slow variations in the magnetic-field drifts were almost completely removed. After the field-drift correction, the obtained values {Bρcori,vcori,Cexpi} are available for all ion species. Their m/q values were then calculated from the BρcorCexp curve, using Eq. (4) following the same procedure as that shown in Fig. 6.

Fig. 7
(Color online) (a) Scatter plot of fit residuals, BρexpiBρ(Cexpi), as a function of injection number illustrating slow variations of the magnetic fields of the CSRe [35]. (b) Similar to (a), but with the corrected cor
pic
2.5
Improvements compared with the conventional IMS

Figure 8 shows part of the m/q spectrum of Ref. [53]. The peaks of the nuclei with nearly identical m/q ratios may overlap. To obtain the m/q values from the overlapping peaks, we introduce a Z-dependent parameter, U=εH¯, extracted from the timing signals of the TOF detectors [54, 55]. Here, ε is the detection efficiency for a specific ion and H¯ is the average signal amplitude of that ion.

Fig. 8
(a) Part of the m/q spectrum and (b) the corresponding scatter plot of U versus m/q in Ref. [53]. The ion species are also indicated. Note that the unresolved 62Ge32+, 66Se34+, and 70Kr36+ in the m/q spectrum are completely separated from 31S16+, 33Cl37+, and 35Ar18+, respectively, in the plot of U versus m/q
pic

To demonstrate the increasing resolving power of the new approach compared to the conventional IMS, we transformed the m/q spectrum into a new revolution-time spectrum [34, 35], Tfix, at a fixed magnetic rigidity fix = 5.4758 Tm (the corresponding orbit length is Cfix = 128.86 m), as follows: Tfix=Cfix1(Bρ)fix2(mq)2+(1vc)2. (12) For example, scatter plots of Texp and Tfix versus Cexp for 24Al13+ ions are shown in Fig. 9 [34, 35].

Fig. 9
Scatter plots of Texp and Tfix versus Cexp for 24Al13+ ions [34, 35]
pic

Clearly, the two states in 24Al, separated by a mass difference of 425.8(1) keV [9], cannot be resolved in the Texp spectrum, whereas the two peaks can be separated in the Tfix spectrum. The standard deviations of the T peaks derived from the corresponding spectra are shown in Fig. 10.

Fig. 10
Standard deviations of the T peaks (left scale) derived from the original revolution-time spectrum (black filled squares) and from the newly constructed spectrum (blue circles). The corresponding absolute accuracies of mass-to-charge ratios are given on the right scale. [34, 35]
pic

The standard deviations, σT, of the T peaks in the original Texp spectrum exhibit a parabolic dependence on m/q. σT approaches a minimum at approximately 2 ps for a limited number of nuclides (isochronicity window). In the spectrum obtained using the new -defined IMS technique, σT=0.5 ps was achieved in the isochronicity window, corresponding to a mass-resolving power of 3.3 × 105 (FWHM). At the edges of the spectrum, a mass-resolving power of 1.3× 105 (FWHM) can be achieved, which is an improvement by a factor of approximately 8 compared to the conventional method. We emphasize that this was done without reducing the acceptance of either the ring or transfer line. The right scale in Fig. 10 shows the corresponding absolute m/q precision. We emphasize that an m/q precision of 5 keV is obtained for a single stored ion.

Using the Tz=-1/2 nuclides to calibrate the spectrum, the redetermined masses of the Tz=-1 nuclides are compared with well-known literature values in Fig. 11[35]. The results obtained using Texp as in conventional IMS are shown in Fig. 11 (a), where the systematic deviations for the Tz=-1 nuclides can be clearly observed. Such systematic deviations are caused mainly by the different momentum distributions and energy losses of the two series of nuclides with Tz=-1/2 and -1 [49]. Not only did the statistical uncertainties significantly decrease, but the systematic deviations demonstrated in Fig. 11 (a) are completely removed over a wide range of revolution times.

Fig. 11
Comparison of redetermined mass excesses of Tz=-1 nuclides (filled red squares), deduced from (a) Texp and (b) Tfix, with literature values [9] using the Tz=-1/2 nuclides (filled black circles) as calibrants. Masses are determined from the revolution-time spectrum following the procedures described in [56] and χn is the normalized Chi-square for the reference nuclides used in the calibration [34, 35]
pic
3

Recent results

Several experiments on -defined IMS using 58Ni, 86Kr, 78Kr, and 36Ar as primary beams were performed. Some newly determined masses have also been reported [57, 53, 34, 35, 58]. Analysis of the data on neutron-rich 86Kr fragments and part of the data on neutron-deficient 36Ar fragments is still in progress. Figure 12 summarizes the published results in the chart of nuclides. The masses of five short-lived neutron-deficient nuclides were measured for the first time, and 23 nuclides were remeasured with higher precision. In the following section, we discuss the application of our measurements to the study of nuclear astrophysics and nuclear structures.

Fig. 12
(Color online) Partial chart of nuclides. Stable nuclides are shown in black. Nuclides with known masses in AME2020 [9] are shown in gray. The nuclei whose masses were measured for the first time are indicated in red. Blue represents the remeasured nuclei with improved precision
pic
3.1
GS 1826-24 X-ray Burst and Neutron Stars

Type-I X-ray bursts occur on the surfaces of neutron stars, accreting hydrogen- and helium-rich matter from companion stars in a stellar binary system [59]. The burst is powered by a sequence of nuclear reactions, termed the rapid-proton-capture nucleosynthesis process (rp process) [60], which is a sequence of proton captures, and β decays along the proton drip line. Figure 13 shows a nuclear chart around the rp-process waiting point 64Ge. Measurements of neutron-deficient 78Kr projectile fragments at the CSRe provided the masses for 63Ge, 64,65As, and 66,67Se using -defined IMS, and all the relevant separation energies around the waiting point 64Ge were determined [57].

Fig. 13
Nuclear chart around the rp-process waiting point 64Ge [57]. Nuclides whose masses were obtained from AME2020 [9], whose masses were experimentally determined, or whose mass uncertainties were improved in Ref. [57] are indicated with black, red, and blue, respectively. The one- (Sp) and two-proton (S2p) separation energies (values expressed in keV) follow the same color code. The pathway of rp-process nucleosynthesis is shown with black arrows. Refer to the legend for more details
pic

The impact of these new masses was investigated using state-of-the-art multizone X-ray burst simulations. Simulated X-ray burst light curves are shown in Fig. 14. The new nuclear masses, particularly the less-bound 65As and more-bound 66Se, result in a stalled rp-process at the 64Ge waiting point. The new distributions of the elements produced through the rp-nucleosynthesis (“ashes”) are also modified, as shown in the inset of Fig. 14. The ash abundances of Urca nuclides are particularly notable [66, 67]. A 17% increase in the A=64 ash mass fraction results in increased electron-capture heating, and a 14% decrease in the A=65 ash mass fraction resulted in reduced Urca cooling, implying a somewhat warmer accreted neutron-star crust.

Fig. 14
(Color online) Calculated X-ray-burst light curves [57]. The baseline nuclear-physics input is labelled as AME2020 (dashed-dotted line and gray band). The result including the nuclear masses determined in Ref. [57] is marked as Updated (solid-line and red band). The light-curve bands correspond to 68% confidence intervals. The inset shows the calculated mass fractions X(A) as a function of mass number
pic

The X-ray-burst luminosity measured by a telescope is directly proportional to the emitted light flux and inversely proportional to the square of the distance d2 and surface gravitational redshift (1+z)2 of the burst corrected by the electromagnetic-wave transport efficiency [68-70]. Our new light curve enables the setting of new constraints on the optimal d and (1+z) parameters that fit the observational data [57]. The results are presented in Fig. 15(a). The increased peak luminosity requires distance d to be increased by 0.4 kpc from approximately 5.8 kpc to approximately 6.2 kpc. The rp-process stalled at 64Ge leads to extended hydrogen burning in the light curve, thereby extending the burst tail. Consequently, the modelled light curve must be less time-dilated, thereby reducing (1+z).

Fig. 15
(Color online) (a) Surface gravitational redshift versus the distance of the source GS 1826-24 obtained through a comparison of the modelled light curves shown in Fig. 14 and the observational data from the year 2007 bursting epoch [57]. The colors indicate 68% (red), 90% (yellow), and 95% (gray) confidence intervals. Two regions correspond to results using previously known masses (AME2020) and those including our new masses (Updated). (b) Neutron-star mass (MNS) versus radius (RNS) constraints calculated from the (1+z) 95% confidence intervals shown in (a) [57]. Limits for neutron-star compactness determined using other astrophysical observables [61-65] are also shown in this figure
pic

The constraints on (1+z) can be further converted into limits on the neutron-star compactness (MNS/RNS) following the approach introduced in [71]. The general relativistic neutron-star mass MGR and radius RGR are determined using (1+z)=1/12GMGR/(RGRvc2), where G is the gravitational constant. If the Newtonian mass is equal to the general relativistic mass MNS=MGR, then RGR=(1+z)RNS. The compactness constraints corresponding to the 95% confidence intervals shown in Fig. 15(a) are shown in Fig. 15(b) along with the mass and radius constraints from other observational probes and theoretical limits.

Recently, the newly determined compactness ξ=MNS/RNS was incorporated into a comprehensive Bayesian statistical framework to investigate its impact on the equation of state (EOS) of supradense neutron-rich matter and the spin frequency required for GW 190814’s minor m2 with mass 2.59±0.05 M to be a rotationally stable pulsar. The EOS of high-density symmetric nuclear matter must be softened significantly, while the symmetry energy at supersaturation densities is stiffened compared to our prior knowledge from earlier analyses using data from both astrophysical observations and terrestrial nuclear experiments [72, 73].

3.2
Thermonuclear reaction rate of 57Cu(p,γ)58Zn in the rp process

In addition to research around the waiting-point 64Ge, the newly determined mass of the neutron-deficient nuclide 58Zn was used to re-evaluate the thermonuclear rate of the 57Cu(p,γ)58Zn reaction around the waiting-point 56Ni [74]. The re-evaluated thermonuclear rate was higher than the most recently published rate by a factor of up to three in the temperature range of 0.2 GK≤T≤1.5 GK. One-zone post-processing type-I X-ray-burst calculations were performed. Figure 16 shows the fractional difference in abundance as a function of mass number. The updated rate and new mass of 58Zn resulted in noticeable abundance variations for nuclei with A = 56-59 and a reduction in A = 57 abundance by up to 20.7%.

Fig. 16
Fractional difference of abundance by mass number of Ref. [74] (present) compared to those using the Langer et al. rate [75] (red filled circles) and Lam et al. [70] rates (blue open triangles). The black dashed line represents the result with initial mass fractions [76]
pic
3.3
Residual Proton–Neutron Interactions in the N=Z Nuclei

The binding energy of a nucleus, B(Z,N), is derived directly from the atomic masses and embodies the sum of the overall interactions inside the nucleus. Binding-energy differences can isolate specific classes of interactions and provide insights into nuclear structural modifications [1, 15]. An important mass filter, the double binding-energy difference denoted as δVpn, has been used to isolate the residual proton-neutron (pn) interactions [77-79] and to sensitively probe a variety of structure phenomena, such as the onset of collectivity and deformation [80-83], changes to the underlying shell structure [84], and phase-transitional behavior [81, 85]. Conventionally, δVpn values are derived as follows [79]: δVpnee(Z,N)=14[B(Z,N)B(Z,N2)B(Z2,N)+B(Z2,N2)], (13) δVpnoo(Z,N)=[B(Z,N)B(Z,N1)B(Z1,N)+B(Z1,N1)], (14) δVpnoe(Z,N)=12[B(Z,N)B(Z,N2)B(Z1,N)+B(Z1,N2)], (15) δVpneo(Z,N)=12[B(Z,N)B(Z,N1)B(Z2,N)+B(Z2,N1)], (16) and are termed the empirical residual proton–neutron interactions. Equations (13)(16) are suitable for nuclei with (Z, N) =even-even, odd-even, even-odd, and odd-odd, respectively.

Measurements of neutron-deficient 78Kr projectile fragments at the CSRe provided masses for 58Zn, 60Ga, 62Ge, 64As, 66Se, and 70Kr Tz=-1 nuclides as well as 61Ga, 63Ge, 65As, 67Se, 71Kr, and 75Sr Tz=-1 nuclides [53]. The newly measured masses provide the δVpn values using Eq. (13) for nuclei with N=Z=even(δVpnee) and Eq. (14) for nuclei with N=Z=odd (δVpnoo). The results are presented in Fig. 17 (a) together with the δVpn values extracted using the currently available mass models [86-94].

Fig. 17
(a) Experimental δVpn for N=Z nuclei beyond A=56 and comparison with different mass-model predictions [86-94]. Red lines are included to guide the eye. Red symbols indicate that one of the new masses in Ref. [53] was used. δVpn of 58Cu using the binding energy of the T=1, Jπ=0+ excited state [9] is marked with open diamond shapes. ME(70Br)=-52030(80) keV is obtained from [95, 96]. δVpn values obtained using the extrapolated masses in AME2020 [9, 9] are connected by the black solid line. (b) and (c) show experimental δVpn values for N=Z and N=Z+2 nuclei beyond A=56 together with the ab initio calculations [53]. Data uncertainties are within the size of the symbols. δVpn values from ab initio calculations using 2NF+3NF and 2NF only are plotted as red and blue lines (solid lines for even-even and dashed lines for odd-odd), respectively
pic

Our results show that an increasing trend in δVpnoo beyond Z=28 (red dashed line) is established, which has been suggested as an indication of the restoration of the pseudo-SU(4) symmetry in the fp shell [97, 98]. δVpnee follows a decreasing trend (red solid line), as reported in the lower-mass region [78, 99]. This contrary to the expectation of pseudo-SU(4) symmetry [79, 100]. We extracted the δVpn values using the predicted masses of frequently used mass models [86-94] and observed that none of the models can reproduce the bifurcation of δVpn (see Fig. 17(a)).

To understand the bifurcation of δVpn, we performed an ab initio valence-space in-medium similarity renormalization group (VS-IMSRG) calculation using a chiral interaction with a two-nucleon force (2NF) at N3LO and a three-nucleon force (3NF) at N2LO, named EM1.8/2.0 [101].

As shown in Fig. 17(c), our calculations with 3NF excellently reproduced the experimental δVpn with N = Z + 2 nuclei, demonstrating the capability of the ab initio approach. For N = Z nuclei, the calculations reproduced the δVpnee data from 56Ni to 72Kr, which decreased marginally with increasing A. As both T = 0 and T = 1 pn correlations are naturally included in the ab initio calculation, reasonable agreement with the experimental δVpnoo values is also obtained from 58Cu to 70Br. In particular, the increasing trend of δVpnoo with changing A is well reproduced.

Recently, a shell-model-like approach based on relativistic density-functional theory was established [102] by simultaneously treating the neutron-neutron, proton-neutron, and proton-proton pairing correlations both microscopically and self-consistently. The calculated δVpn values reproduced the observed bifurcation in Fig. 17 well. The mechanism for this abnormal bifurcation was owing to the enhanced proton–neutron pairing correlations in the odd-odd N=Z nuclei compared with the even-even ones.

3.4
Mirror Symmetry of Residual Proton–Neutron Interactions

In addition to the δVpn of N = Z nuclei, mirror symmetry of the residual p-n interactions on both sides of the Z = N line [56] were observed. Such mirror symmetry was observed many years ago by Jänecke [103] and was recently used by Zong et al. [104] to make high-precision mass predictions for neutron-deficient nuclei.

Figure 18 shows the energy differences were defined as Δ(δVpn)=δVpn(Tz<,A)δVpn(Tz>,A) [35]. Here, Tz</Tz> represents the negative/positive value of Tz for the mirror nuclei. The δVpn values are calculated using Eqs. (13)(16), using mass data from AME2020 [9]. Masses of 49Fe and 55Cu Tz=-3/2 nuclides as well as 54Ni Tz=-1 deduced from measurements of neutron-deficient 58Ni projectile fragments at the CSRe are also used in these calculations, as shown by the blue dots.

Fig. 18
(Color online) Differences of δVpn for mirror nuclei [35]. The gray shadow indicates an error band of 50 keV. The neutron-deficient partners are indicated for the mirror pairs with A≤20
pic

Δ(δVpn) values scatter around zero within an error band of ±50 keV for the A > 20 mirror pairs, indicating that the mirror symmetry of δVpn holds well in this mass region. Notably, the mass of 55Cu was redetermined with significantly improved precision, and its value was 172 keV more bound than the previous value [105]. Using this new mass, the calculated δVpn value of 56Cu was approximately equal to that of 56Co. Consequently, Δ(δVpn)=δVpn(56Cu)δVpn(56Co) fits general systematics well with a high level of accuracy (see Fig. 18). This result further confirmed the reliability of the new mass value for 55Cu.

In the lighter-mass region with A20, the so-called mirror symmetry of δVpn is broken into a few mirror pairs (see Fig. 18). That is, the δVpn values of the neutron-deficient nuclei are systematically smaller than those of the corresponding neutron-rich partners. Such mirror-symmetry breaking is simply attributed to the binding-energy effect [103]. Here, we note that one of the four nuclei is particle-unbound, which may lead to the mirror-symmetry breaking of δVpn. Thus, the mirror symmetry of δVpn cannot be used to predict masses with A20.

3.5
Ground-state mass of 22Al and test of ab initio calculations

The level structure of mirror nuclei is commonly addressed and discussed based on the isospin symmetry, which is a basic assumption in the fields of particle and nuclear physics. However, this symmetry is approximate and the corresponding deviation is called isospin symmetry breaking (ISB) [106-110]. Studies on mirror nuclei offer profound insights into the origin of ISB and further information about nuclear structures, as well as facilitate the evaluation of nuclear models [106-113]. A key quantity in the investigation of ISB in mirror nuclei is the mirror-energy difference (MED) [114, 106, 115], which is defined as follows: MED=Ex(J,T,Tz=T)Ex(J,T,Tz=T), (17) where Ex(J,T,Tz) denotes the excitation energy of a state with spin J, isospin T, and z-projection Tz.

22Al is the lightest-bound Al isotope with Tz=-2. Two low-lying 1+ states in odd-odd 22Al were identified via β-delayed one-proton emissions from 22Si [112]. However, their excitation energies have not been determined precisely because of the lack of an experimental ground-state mass of 22Al. Recently, the ground-state mass excess of 22Al was measured for the first time using -defined IMS as 18103(10) keV [58], which is 97(400) keV smaller than the extrapolated ME=18200(400) keV in AME2020 [9]. The new mass excess value allowed the determination of the excitation energies of the two low-lying states, Ex(11+)=1002(51) keV and Ex(12+)=2242(51) keV in 22Al, with a significantly reduced uncertainty of 51 keV. The experimental MEDs of 11,2+ states in 22Al-22F mirror nuclei are shown in Fig. 19 [58].

Fig. 19
(color online) MEDs of the 11,2+ states in 22Al/22F mirror nuclei by employing ab initio VS-IMSRG calculations with four sets of nuclear interactions. The experimental MED values are plotted with error bars, whereas the calculated values are not
pic

The MEDs of 22Al-22F mirror nuclei were calculated using an ab initio VS-IMSRG approach, employing several sets of nuclear forces derived from chiral effective-field theory [58]. The large uncertainties in the MEDs obtained in Ref. [112] prevented the benchmarking of the theoretical calculations. The uncertainties of the MEDs obtained in this study are significantly reduced, and the results show that the MED of the 11+ states is larger than that of the 12+ states in 22Al-22F mirror nuclei. Based on the ab initio calculations, we observe that the MEDs calculated with 1.8/2.0(EM), NNLOopt, and NN+3N(lnl) interactions are in good agreement with the experimental data, especially the result from NNLOopt, whereas the predicted MED values with NNLOsat for the 11,2+ states are inverted when compared with the experimental data. Therefore, the MED of the mirror nuclei serves as a more sensitive observable tool for testing theoretical calculations when comparing the excitation spectra. Moreover, mirror systems 22Al-22F were investigated using the ab initio Gamow shell model based on the same EM1.8/2.0 interaction [116]. Within the Gamow shell-model calculations, the effective Hamiltonian is derived through many-body perturbation theory, and continuum coupling is well accounted for using the Berggren basis. The results obtained from ab initio GSM calculations were consistent with those obtained from the VS-IMSRG calculations based on the EM1.8/2.0 interaction.

3.6
Test of Isospin Multiplet Mass Equation

The mass of a set of isobaric analog states (IASs) can be described by the well-known quadratic isospin multiplet mass equation (IMME) [117]: ME(A,T,Tz)=a(A,T)+b(A,T)×Tz+c(A,T)×Tz2, (18) where the MEs are the mass excesses of the IASs of a multiplet with a fixed mass number A and total isospin T. T is equal to or greater than the projection of T, Tz=(NZ)/2, for a specific nucleus. The coefficients a, b, and c depend on A, T, and other quantum numbers, such as spin and parity , but are independent of Tz. The quadratic form of the IMME, that is, Eq. (18), is commonly considered accurate within uncertainties of a few tens of keV. In this context, precise mass measurements can be used to test its validity (see Ref. [118] and references therein). Typically, we add extra terms to Eq. (18) such as dTz3 and/or eTz4, which provide a measure of the breakdown of the quadratic form of the IMME. Numerous measurements were performed to investigate the validity of the IMME. Reviews and compilations of existing data can be found in Refs. [119, 120] and the references cited therein.

Measurements of neutron-deficient 58Ni projectile fragments at the CSRe provided the masses for 41Ti, 45Cr, 49Fe, 53Ni, and 55Cu using -defined IMS [35]. Using these new masses, the four experimental masses of the T = 3/2 IASs were completed; thus, the validity of the quadratic form of the IMME could be tested, reaching the heaviest A = 55 isospin quartet. The mass data were fitted to the cubic form of the IMME by adding the dTz3 term to Eq. (18) and the fitted d coefficients are shown in Fig. 20.

Fig. 20
 d coefficients of the cubic form of IMME for the T = 3/2, A = 41, 45, 49, 53, and 55 isospin quartets [35]. The solid line connects the predicted d values from theoretical calculations [121]
pic

Compared with previous results in Ref. [118], we conclude that the trend of a gradual increase in d with A in the fp shell [118] was not confirmed, at least at the present level of accuracy. Given that all the extracted d coefficients are compatible with zero, the quadratic form of the IMME is valid for the cases investigated here. Figure 20 shows the d values obtained from the theoretical calculations [121]. The predicted nonzero d coefficients for these T = 3/2 isospin quartets cannot be discarded owing to large experimental uncertainties.

3.7
Ground-state mass of 70Br

The masses of the Tz=-1 nuclides 58Zn, 60Ga, 62Ge, 64As, 66Se, and 70Kr deduced from measurements of neutron-deficient 78Kr projectile fragments at the CSRe are used to investigate the systematics of b and c coefficients of the quadratic form of the IMME up to the upper fp-shell nuclei [122].

Using the new mass results reported in Ref. [53] and the literature values in the latest atomic-mass evaluation (AME2020) [9], the b and c coefficients are deduced according to Eqs. (18) up to A = 70, as shown in Fig. 21 [122].

Fig. 21
 b and c coefficients of the quadratic IMME [122] extracted using the AME2020 mass data [9] (black circles) and the new masses from the CSRe [53] (red circles). The c coefficient using ME(70Br =-51934(16) keV is indicated by the blue triangle [122]. The solid lines denote a semiempirical formula. Refer to details in Ref. [122]
pic

Figure 21 (a) shows that the deduced b coefficients follow the smooth trend established in the lighter-mass region. The obtained c coefficients exhibit regular zig-zag staggering with changes in A, as shown in Fig. 21 (b). However, the regular staggering trend breaks at A = 70 and the derived c coefficient becomes negative. This anomaly was analyzed and attributed to the presently adopted mass of 70Br in the latest atomic-mass evaluation (AME2020).

The smooth trend of the b coefficients and the regular staggering pattern of the c coefficients can be deduced using the mass formula given in Ref. [123] for the members of an isobaric multiplet characterized by A, T, and Tz: M(A,T,Tz)=M0(A,T)+Ec(A,T,Tz)+Tz×ΔmnH, (19) where Ec(A,T,Tz) represents the total charge-dependent energy in the nucleus and ΔmnH is the neutron–hydrogen mass difference. The semi-classical approach [124, 125], Ec=0.6Z20.46Z4/30.15[1(1)Z]e2r0A1/3, (20) can be used to calculate the theoretical values of parameters b, c. The calculated b and c coefficients are shown in Fig. 21 as solid lines.

As reported in Refs. [126, 127], a combination of the theoretical calculations of Coulomb-energy differences (CEDs) and experimental masses of neutron-rich nuclei provides reliable masses for proton-rich nuclei. The CED between the Tz = 0 and Tz =+1 isobaric pairs was extracted using CED=M(A,Tz=0)M(A,Tz=+1)+ΔmnH. (21) Figure 22 shows the CDE difference ΔCED=CEDexpCEDth as a function of the mass number A [122].

Fig. 22
Plot of ΔCED=CEDexpCEDth and the least-squares fits with linear functions [122]. ΔCED=152(16) keV at A = 70 is thereby deduced. Data uncertainties are within the size of the symbols. Orbits for major shells and subshells are indicated
pic

The ΔCEDs follow very good linear behavior within each shell closure, while exhibiting a large deviation at A = 70 between A = 58 and A = 74.

Under the assumption of a smooth variation in the CED with changing mass number A, the ground-state mass excess of 70Br is deduced to be -51934(16) keV, which is 508(22) keV more bound than the adopted value [122]. Using the new ME of 70Br, the recalculated c coefficient is obtained using Eq. (18), and the results are shown in Fig. 21 (b). The new c coefficient now fits well into the systematics and is consistent with a simple theoretical estimate. This supports our recommended ME value of 70Br obtained from the systematics of the Coulomb-energy differences.

4

Summary and outlook

In summary, significantly improved isochronous mass spectrometry, the -defined IMS, has been pioneered at the experimental cooler-storage ring CSRe. Owing to the simultaneous measurement of the revolution time and velocity of every stored short-lived ion, the sensitivity and precision of the mass measurements significantly increased. The time sequences of the two TOF detectors were unique for each ion. Only a few tens of signals were sufficient for unambiguous ion identification. This unparalleled property of the -defined IMS makes it, in principle, a background-free technique. The overall measurement time was less than 1 ms, indicating that all β-decaying nuclei could be studied without lifetime restrictions. High mass-resolving power was achieved over the entire acceptance of the storage ring, implying that a large range of m/q values can be covered in a single-machine setting. This is a remarkable achievement, which indicates that the storage of a single short-lived (T1/2≥100 μs) ion is sufficient for mass determination with ≈5q keV precision. Thus, -defined IMS is an ideal technique for high-precision mass measurements of the most exotic nuclides, which have the shortest half-lives and lowest production yields.

Using -defined IMS, the masses of 22Al, 62Ge, 64As, 66Se, and 70Kr were measured for the first time, and the masses of 41Ti, 43V, 45Cr, 47Mn, 49Fe, 51Co, 53Ni, 55Cu, 44V, 46Cr, 48Mn, 50Fe, 52Co, 54Ni, 56Cu, 58Zn, 60Ga, 61Ga, 63Ge, 66As, 67Se, 71Kr, and 75Sr were re-determined with improved accuracy. Given these newly determined masses, interesting issues in nuclear astrophysics and nuclear structures have been investigated.

The merits of the novel IMS are demonstrated by the considerably increased sensitivity and accuracy of its measurements. A Spectrometer Ring (SRing) [129] at the high-intensity Heavy-Ion Accelerator Facility (HIAF) [130, 131, 128] is under construction, as shown in Fig. 23. -defined IMS is planned for application in the SRing, and the distance between the two TOF detectors in the straight section will be 39 m, which is 2.0 times the distance in the CSRe. Assuming that the time resolution of the TOF detectors is the same as that used in the CSRe, in principle, the precision of the velocity measurement would be increased by a factor of approximately 2.0, and the mass precision would be further improved. Two isochronous mode settings with γt=1.43,1.67 were designed for the mass-measurement experiments in the SRing. In the future, both neutron-deficient and -rich nuclei will be covered in experiments at the SRing.

Fig. 23
Layout of HIAF [128]. The major sub-systems are sketched, including the ion sources, superconducting ion linear accelerator, high-energy synchrotron booster, high-energy fragment separator, and experimental spectrometer ring. The low-energy and high-energy experimental stations are also indicated
pic
References
1K. Blaum,

High-accuracy mass spectrometry with stored ions

. Physics Reports 425, 1-78 (2006). https://doi.org/10.1016/j.physrep.2005.10.011
Baidu ScholarGoogle Scholar
2D. Lunney, J.M. Pearson, C. Thibault,

Recent trends in the determination of nuclear masses

. Rev. Mod. Phys. 75, 1021-1082 (2003). https://doi.org/10.1103/RevModPhys.75.1021
Baidu ScholarGoogle Scholar
3A.N. Bohr, B.R. Mottelson, Nuclear Structure (in 2 volumes), (World Scientific Publishing Company, 1998)
4C. Thibault, R. Klapisch, C. Rigaud, et al.,

Direct measurement of the masses of 11Li and 26-32Na with an on-line mass spectrometer

. Phys. Rev. C 12, 644-657 (1975). https://doi.org/10.1103/PhysRevC.12.644
Baidu ScholarGoogle Scholar
5F. Wienholtz, D. Beck, K. Blaum, et al.,

Masses of exotic calcium isotopes pin down nuclear forces

. Nature 498, 346-349 (2013). https://doi.org/10.1038/nature12226
Baidu ScholarGoogle Scholar
6T. Radon, H. Geissel, G. Münzenberg, et al.,

Schottky mass measurements of stored and cooled neutron-deficient projectile fragments in the element range of 57≤Z≤84

. Nuclear Physics A 677, 75-99 (2000). https://doi.org/10.1016/S0375-9474(00)00304-3
Baidu ScholarGoogle Scholar
7Y. Novikov, F. Attallah, F. Bosch, et al.,

Mass mapping of a new area of neutron-deficient suburanium nuclides

. Nuclear Physics A 697, 92-106 (2002). https://doi.org/10.1016/S0375-9474(01)01233-7
Baidu ScholarGoogle Scholar
8J. Van Schelt, D. Lascar, G. Savard, et al.,

First results from the CARIBU facility: Mass measurements on the r-process path

. Phys. Rev. Lett. 111, 061102 (2013). https://doi.org/10.1103/PhysRevLett.111.061102
Baidu ScholarGoogle Scholar
9M. Wang, W. Huang, F. Kondev, et al.,

The AME 2020 atomic mass evaluation (II)

. tables, graphs and references*. Chinese Physics C 45, 030003 (2021). https://doi.org/10.1088/1674-1137/abddaf
Baidu ScholarGoogle Scholar
10F. Kondev, M. Wang, W. Huang, et al.,

The NUBASE2020 evaluation of nuclear physics properties

. Chinese Physics C 45, 030001 (2021). https://doi.org/10.1088/1674-1137/abddae
Baidu ScholarGoogle Scholar
11T. Eronen, A. Kankainen, J. Äystö,

Ion traps in nuclear physics-recent results and achievements

. Progress in Particle and Nuclear Physics 91, 259-293 (2016). https://doi.org/10.1016/j.ppnp.2016.08.001
Baidu ScholarGoogle Scholar
12J. Dilling, K. Blaum, M. Brodeur, et al.,

Penning-trap mass measurements in atomic and nuclear physics

. Annual Review of Nuclear and Particle Science 68, 45-74 (2018). https://doi.org/10.1146/annurev-nucl-102711-094939
Baidu ScholarGoogle Scholar
13M. Mougeot, D. Atanasov, J. Karthein, et al.,

Mass measurements of 99-101In challenge ab initio nuclear theory of the nuclide 100Sn

. Nature Physics 17, 1099-1103 (2021). https://doi.org/10.1038/s41567-021-01326-9
Baidu ScholarGoogle Scholar
14S. Beck, B. Kootte, I. Dedes, et al.,

Mass measurements of neutron-deficient Yb isotopes and nuclear structure at the extreme proton-rich side of the N=82 shell

. Phys. Rev. Lett. 127, 112501 (2021). https://doi.org/10.1103/PhysRevLett.127.112501
Baidu ScholarGoogle Scholar
15T. Yamaguchi, H. Koura, Y. Litvinov, et al.,

Masses of exotic nuclei

. Progress in Particle and Nuclear Physics 120, 103882 (2021). https://doi.org/10.1016/j.ppnp.2021.103882
Baidu ScholarGoogle Scholar
16Z. Meisel, S. George, S. Ahn, et al.,

Nuclear mass measurements map the structure of atomic nuclei and accreting neutron stars

. Phys. Rev. C 101, 052801 (2020). https://doi.org/10.1103/PhysRevC.101.052801
Baidu ScholarGoogle Scholar
17S. Michimasa, M. Kobayashi, Y. Kiyokawa, et al.,

Mapping of a new deformation region around 62Ti

. Phys. Rev. Lett. 125, 122501 (2020). https://doi.org/10.1103/PhysRevLett.125.122501
Baidu ScholarGoogle Scholar
18Y.H. Zhang, Y.A. Litvinov, T. Uesaka, et al.,

Storage ring mass spectrometry for nuclear structure and astrophysics research

. Physica Scripta 91, 073002 (2016). https://doi.org/10.1088/0031-8949/91/7/073002
Baidu ScholarGoogle Scholar
19M. Hausmann, F. Attallah, K. Beckert, et al.,

First isochronous mass spectrometry at the experimental storage ring ESR

. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 446, 569-580 (2000). https://doi.org/10.1016/S0168-9002(99)01192-4
Baidu ScholarGoogle Scholar
20J. Stadlmann, M. Hausmann, F. Attallah, et al.,

Direct mass measurement of bare short-lived 44V, 48Mn, 41Ti and 45Cr ions with isochronous mass spectrometry

. Physics Letters B 586, 27-33 (2004). https://doi.org/10.1016/j.physletb.2004.02.014
Baidu ScholarGoogle Scholar
21H.S. Xu, Y.H. Zhang, Y.A. Litvinov,

Accurate mass measurements of exotic nuclei with the CSRe in Lanzhou

. International Journal of Mass Spectrometry 349-350, 162-171 (2013). https://doi.org/10.1016/j.ijms.2013.04.029
Baidu ScholarGoogle Scholar
22M.Z. Sun, X.H. Zhou, M. Wang, et al.,

Precision mass measurements of short-lived nuclides at HIRFL-CSR in lanzhou

. Frontiers of Physics 13, 132112 (2018). https://doi.org/10.1007/s11467-018-0844-5
Baidu ScholarGoogle Scholar
23D. Nagae, S. Omika, Y. Abe, et al.,

First demonstration of mass measurements for exotic nuclei using Rare-RI Ring

. JPS Conf. Proc. 35, 011014. https://doi.org/10.7566/JPSCP.35.011014
Baidu ScholarGoogle Scholar
24B. Sun, R. Knöbel, Y. Litvinov, et al.,

Nuclear structure studies of short-lived neutron-rich nuclei with the novel large-scale isochronous mass spectrometry at the FRS-ESR facility

. Nuclear Physics A 812, 1-12 (2008). https://doi.org/10.1016/j.nuclphysa.2008.08.013
Baidu ScholarGoogle Scholar
25R. Knöbel, M. Diwisch, H. Geissel, et al.,

New results from isochronous mass measurements of neutron-rich uranium fission fragments with the FRS-ESR-facility at GSI

. The European Physical Journal A 52, 138 (2016). https://doi.org/10.1140/epja/i2016-16138-6
Baidu ScholarGoogle Scholar
26H.F. Li, S. Naimi, T.M. Sprouse, et al.,

First application of mass measurements with the Rare-RI Ring reveals the solar r-process abundance trend at A=122 and A=123

. Phys. Rev. Lett. 128, 152701 (2022). https://doi.org/10.1103/PhysRevLett.128.152701
Baidu ScholarGoogle Scholar
27S. Naimi, Y. Yamaguchi, T. Yamaguchi, et al.,

Recent achievements at the Rare-RI Ring, a unique mass spectrometer at the RIBF/RIKEN

. The European Physical Journal A 59, 90 (2023). https://doi.org/10.1140/epja/s10050-023-01009-4
Baidu ScholarGoogle Scholar
28D. Nagae, S. Omika, Y. Abe, et al.,

Isochronous mass spectrometry at the riken rare-ri ring facility

. Phys. Rev. C 110, 014310 (2024). https://doi.org/10.1103/PhysRevC.110.014310
Baidu ScholarGoogle Scholar
29B. Franzke, H. Geissel, G. Münzenberg,

Mass and lifetime measurements of exotic nuclei in storage rings

. Mass Spectrometry Reviews 27, 428-469 (2008). https://doi.org/10.1002/mas.20173
Baidu ScholarGoogle Scholar
30H. Geissel, Y.A. Litvinov,

Precision experiments with relativistic exotic nuclei at GSI

. Journal of Physics G: Nuclear and Particle Physics 31, S1779-S1783 (2005). https://doi.org/10.1088/0954-3899/31/10/072
Baidu ScholarGoogle Scholar
31H. Geissel, R. Knöbel, Y.A. Litvinov, et al.,

A new experimental approach for isochronous mass measurements of short-lived exotic nuclei with the FRS-ESR facility

. Hyperfine Interactions 173, 49-54 (2006). https://doi.org/10.1007/s10751-007-9541-4
Baidu ScholarGoogle Scholar
32H. Geissel, G. Munzenberg, K. Riisager,

Secondary exotic nuclear beams

. Annual Review of Nuclear and Particle Science 45, 163-203 (1995).
Baidu ScholarGoogle Scholar
33J. Liu, X. Xu, P. Zhang, et al.,

Improving the resolving power of Isochronous Mass Spectrometry by employing an in-ring mechanical slit

. Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms 463, 138-142 (2020). https://doi.org/10.1016/j.nimb.2019.06.007
Baidu ScholarGoogle Scholar
34M. Wang, M. Zhang, X. Zhou, et al.,

Bρ-defined isochronous mass spectrometry: An approach for high-precision mass measurements of short-lived nuclei

. Phys. Rev. C 106, L051301 (2022). https://doi.org/10.1103/PhysRevC.106.L051301
Baidu ScholarGoogle Scholar
35M. Zhang, X. Zhou, M. Wang, et al.,

Bρ-defined isochronous mass spectrometry and mass measurements of 58Ni fragments

. The European Physical Journal A 59, 27 (2023). https://doi.org/10.1140/epja/s10050-023-00928-6
Baidu ScholarGoogle Scholar
36B. Mei, X. Lin Tu, M. Wang, et al.,

A high performance Time-of-Flight detector applied to isochronous mass measurement at CSRe

. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 624, 109-113 (2010). https://doi.org/10.1016/j.nima.2010.09.001
Baidu ScholarGoogle Scholar
37W. Zhang, X.L. Tu, M. Wang, et al.,

Time-of-flight detectors with improved timing performance for isochronous mass measurements at the CSRe

. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 756, 1-5 (2014). https://doi.org/10.1016/j.nima.2014.04.051
Baidu ScholarGoogle Scholar
38S.Y. Lee, Accelerator Physics (Fourth Edition), (World Scientific Publishing Company, 2018). https://doi.org/10.1142/11111
39X. Zhou, M. Zhang, M. Wang, et al.,

In-ring velocity measurement for isochronous mass spectrometry

. Phys. Rev. Accel. Beams 24, 042802 (2021). https://doi.org/10.1103/PhysRevAccelBeams.24.042802
Baidu ScholarGoogle Scholar
40R.J. Chen, Y.J. Yuan, M. Wang, et al.,

Simulations of the isochronous mass spectrometry at the HIRFL-CSR

. Physica Scripta T166, 014044 (2015). https://doi.org/10.1088/0031-8949/2015/t166/014044
Baidu ScholarGoogle Scholar
41X. Xu, W. Meng, P. Shuai, et al.,

A data analysis method for isochronous mass spectrometry using two time-of-flight detectors at CSRe

. Chinese Physics C 39, 106201 (2015). https://doi.org/10.1088/1674-1137/39/10/106201
Baidu ScholarGoogle Scholar
42P. Shuai, X. Xu, Y.H. Zhang, et al.,

An improvement of isochronous mass spectrometry: Velocity measurements using two time-of-flight detectors

. Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms 376, 311-315 (2016). https://doi.org/10.1016/j.nimb.2016.02.006
Baidu ScholarGoogle Scholar
43J. Xia, W. Zhan, B. Wei, et al.,

The heavy ion cooler-storage-ring project (HIRFL-CSR) at Lanzhou

. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 488, 11-25 (2002). https://doi.org/10.1016/S0168-9002(02)00475-8
Baidu ScholarGoogle Scholar
44W.W. Ge, Y.J. Yuan, J.C. Yang, et al.,

Experimental investigation of the transition energy γt in the isochronous mode of the HIRFL-CSRe

. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 908, 388-393 (2018). https://doi.org/10.1016/j.nima.2018.08.059
Baidu ScholarGoogle Scholar
45Y.M. Xing, M. Wang, Y.H. Zhang, et al.,

First isochronous mass measurements with two time-of-flight detectors at CSRe

. Physica Scripta T166, 014010 (2015). https://doi.org/10.1088/0031-8949/2015/t166/014010
Baidu ScholarGoogle Scholar
46X.L. Yan, R.J. Chen, M. Wang, et al.,

Characterization of a double Time-of-Flight detector system for accurate velocity measurement in a storage ring using laser beams

. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 931, 52-59 (2019). https://doi.org/10.1016/j.nima.2019.03.058
Baidu ScholarGoogle Scholar
47R.J. Chen, X.L. Yan, W.W. Ge, et al.,

A method to measure the transition energy γt of the isochronously tuned storage ring

. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 898, 111-116 (2018). https://doi.org/10.1016/j.nima.2018.04.056
Baidu ScholarGoogle Scholar
48M. Zhang, Y. Zhang, M. Wang, et al.,

Precision measurement of the transition energy t versus magnetic rigidity for storage-ring isochronous mass spectrometry

. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 1027, 166329 (2022). https://doi.org/10.1016/j.nima.2022.166329
Baidu ScholarGoogle Scholar
49M. Zhang, Y. Zhang, M. Wang, et al.,

Experimental Study on the Systematic Deviation in the Conventional Isochronous Mass Spectrometry

. Nuclear Physics Review 40, 181-187 (2023). https://doi.org/10.11804/NuclPhysRev.40.2022078
Baidu ScholarGoogle Scholar
50Y. Li, X. Wang, H. Zhang, et al.,

Implementation of a dipole magnet power supply control system to improve magnetic field stability at the csre storage ring facility for precision mass measurement

. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 1049, 168108 (2023). https://doi.org/10.1016/j.nima.2023.168108
Baidu ScholarGoogle Scholar
51Y. Xing, Y. Zhang, M. Wang, et al.,

Particle identification and revolution time corrections for the isochronous mass spectrometry in storage rings

. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 941, 162331 (2019). https://doi.org/10.1016/j.nima.2019.06.072
Baidu ScholarGoogle Scholar
52X.L. Tu, M. Wang, Y.A. Litvinov, et al.,

Precision isochronous mass measurements at the storage ring CSRe in Lanzhou

. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 654, 213-218 (2011). https://doi.org/10.1016/j.nima.2011.07.018
Baidu ScholarGoogle Scholar
53M. Wang, Y.H. Zhang, X. Zhou, et al.,

Mass Measurement of Upper fp-Shell N=Z-2 and N=Z-1 Nuclei and the Importance of Three-Nucleon Force along the N=Z Line

. Phys. Rev. Lett. 130, 192501 (2023). https://doi.org/10.1103/PhysRevLett.130.192501
Baidu ScholarGoogle Scholar
54X. Zhou, M. Wang, Y.H. Zhang, et al.,

Charge resolution in the isochronous mass spectrometry and the mass of 51Co

. Nuclear Science and Techniques 32, 37 (2021). https://doi.org/10.1007/s41365-021-00876-0
Baidu ScholarGoogle Scholar
55P. Shuai, H.S. Xu, X.L. Tu, et al.,

Charge and frequency resolved isochronous mass spectrometry and the mass of 51Co

. Physics Letters B 735, 327-331 (2014). https://doi.org/10.1016/j.physletb.2014.06.046
Baidu ScholarGoogle Scholar
56Y.H. Zhang, P. Zhang, X.H. Zhou, et al.,

Isochronous mass measurements of Tz=-1 fp-shell nuclei from projectile fragmentation of 58Ni

. Phys. Rev. C 98, 014319 (2018). https://doi.org/10.1103/PhysRevC.98.014319
Baidu ScholarGoogle Scholar
57X. Zhou, M. Wang, Y.H. Zhang, et al.,

Mass measurements show slowdown of rapid proton capture process at waiting-point nucleus 64Ge

. Nature physics 19, 1091-1097 (2023). https://doi.org/10.1038/s41567-023-02034-2
Baidu ScholarGoogle Scholar
58M. Sun, Y. Yu, X. Wang, et al.,

Ground-state mass of 22al and test of state-of-the-art ab initio calculations

. Chinese Physics C 48, 034002 (2024). https://doi.org/10.1088/1674-1137/ad1a0a
Baidu ScholarGoogle Scholar
59P. Joss,

X-ray bursts and neutron-star thermonuclear flashes

. Nature 270, 310-314 (1977). https://doi.org/10.1038/270310a0
Baidu ScholarGoogle Scholar
60R.K. Wallace, S.E. Woosley,

Explosive hydrogen burning

. The Astrophysical Journal Supplement Series 45, 389-420 (1981). https://doi.org/10.1086/190717
Baidu ScholarGoogle Scholar
61A. Bauswein, O. Just, H.T. Janka, et al.,

Neutron-star Radius Constraints from GW 170817 and Future Detections

. Astrophys. J. Lett 850, L34 (2017). https://doi.org/10.3847/2041-8213/aa9994
Baidu ScholarGoogle Scholar
62F.J. Fattoyev, J. Piekarewicz, C.J. Horowitz,

Neutron Skins and Neutron Stars in the Multimessenger Era

. Phys. Rev. Lett. 120, 172702 (2018). https://doi.org/10.1103/PhysRevLett.120.172702
Baidu ScholarGoogle Scholar
63A.W. Steiner, C.O. Heinke, S. Bogdanov, et al.,

Constraining the Mass and Radius of Neutron Stars in Globular Clusters

. Mon. Not. R. Astron. Soc. 476, 421 (2018). https://doi.org/10.1093/mnras/sty215
Baidu ScholarGoogle Scholar
64T.E. Riley, A.L. Watts, P.S. Ray, et al.,

A NICER View of the Massive Pulsar PSR J0740+6620 Informed by Radio Timing and XMM-Newton Spectroscopy

. Astrophys. J. Lett. 918, L27 (2021). arXiv:2105.06980, https://doi.org/10.3847/2041-8213/ac0a81
Baidu ScholarGoogle Scholar
65S. Koranda, N. Stergioulas, J.L. Friedman,

Upper Limits Set By Causality on the Rotation and Mass of Uniformly Rotating Relativistic Stars

. Astrophys. J. 488, 799 (1997). https://doi.org/10.1086/304714
Baidu ScholarGoogle Scholar
66H. Schatz, S. Gupta, P. Møller, et al.,

Strong neutrino cooling by cycles of electron capture and β-decay in neutron star crusts

. Nature 505, 62-65 (2014). https://doi.org/10.1038/nature12757
Baidu ScholarGoogle Scholar
67R. Lau, M. Beard, S.S. Gupta, et al.,

Nuclear reactions in the crusts of accreting neutron stars

. The Astrophysical Journal 859, 62 (2018). https://doi.org/10.3847/1538-4357/aabfe0
Baidu ScholarGoogle Scholar
68M.Y. Fujimoto,

Angular Distribution of Radiation from Low-Mass X-Ray Binaries

. Astroph. J 324, 995 (1988). https://doi.org/10.1086/165955
Baidu ScholarGoogle Scholar
69C.C. He, L. Keek,

Anisotropy of X-Ray Bursts from Neutron Stars with Concave Accretion Disks

. Astroph. J. 819, 47 (2016). https://doi.org/10.3847/0004-637X/819/1/47
Baidu ScholarGoogle Scholar
70Y.H. Lam, N. Lu, A. Heger, et al.,

The Regulated NiCu Cycles with the New 57Cu(p,γ)58Zn Reaction Rate and Its Influence on Type I X-Ray Bursts: the GS 1826-24 Clocked Burster

. The Astrophysical Journal 929, 73 (2022). https://doi.org/10.3847/1538-4357/ac4d89
Baidu ScholarGoogle Scholar
71Z. Meisel, G. Merz, S. Medvid,

Influence of Nuclear Reaction Rate Uncertainties on Neutron Star Properties Extracted from X-Ray Burst Model–Observation Comparisons

. The Astrophysical Journal 872, 84 (2019). https://doi.org/10.3847/1538-4357/aafede
Baidu ScholarGoogle Scholar
72W.J. Xie, B.A. Li, N.B. Zhang,

Impact of the newly revised gravitational redshift of x-ray burster gs 1826-24 on the equation of state of supradense neutron-rich matter

. (2024). arXiv:2404.01989
Baidu ScholarGoogle Scholar
73B.J. Cai, B.A. Li, Z. Zhang,

Central speed of sound, the trace anomaly, and observables of neutron stars from a perturbative analysis of scaled Tolman-Oppenheimer-Volkoff equations

. Phys. Rev. D 108, 103041 (2023). https://doi.org/10.1103/PhysRevD.108.103041
Baidu ScholarGoogle Scholar
74M. Zhang, X. Xu, Y.M. Xing, et al.,

Thermonuclear reaction rate of 57Cu(p,γ)58Zn in the rp process

. Phys. Rev. C 109, 045802 (2024). https://doi.org/10.1103/PhysRevC.109.045802
Baidu ScholarGoogle Scholar
75C. Langer, F. Montes, A. Aprahamian, et al.,

Determining the rp-process flow through 56Ni: Resonances in 57Cu(p,γ)58Zn identified with GRETINA

. Phys. Rev. Lett. 113, 032502 (2014). https://doi.org/10.1103/PhysRevLett.113.032502
Baidu ScholarGoogle Scholar
76H. Schatz, A. Aprahamian, V. Barnard, et al.,

End Point of the rp Process on Accreting Neutron Stars

. Phys. Rev. Lett. 86, 3471-3474 (2001). https://doi.org/10.1103/PhysRevLett.86.3471
Baidu ScholarGoogle Scholar
77J.Y. Zhang, R. Casten, D. Brenner,

Empirical proton-neutron interaction energies. linearity and saturation phenomena

. Physics Letters B 227, 1-5 (1989). https://doi.org/10.1016/0370-2693(89)91273-2
Baidu ScholarGoogle Scholar
78D. Brenner, C. Wesselborg, R. Casten, et al.,

Empirical p-n interactions: global trends, configuration sensitivity and N=Z enhancements

. Physics Letters B 243, 1-6 (1990). https://doi.org/10.1016/0370-2693(90)90945-3
Baidu ScholarGoogle Scholar
79P. Van Isacker, D.D. Warner, D.S. Brenner,

Test of Wigner’s Spin-Isospin Symmetry from Double Binding Energy Differences

. Phys. Rev. Lett. 74, 4607-4610 (1995). https://doi.org/10.1103/PhysRevLett.74.4607
Baidu ScholarGoogle Scholar
80I. Talmi,

Effective Interactions and Coupling Schemes in Nuclei

. Rev. Mod. Phys. 34, 704-722 (1962). https://doi.org/10.1103/RevModPhys.34.704
Baidu ScholarGoogle Scholar
81P. Federman, S. Pittel,

Towards a unified microscopic description of nuclear deformation

. Physics Letters B 69, 385-388 (1977). https://doi.org/10.1016/0370-2693(77)90825-5
Baidu ScholarGoogle Scholar
82R.F. Casten,

Possible unified interpretation of heavy nuclei

. Phys. Rev. Lett. 54, 1991-1994 (1985). https://doi.org/10.1103/PhysRevLett.54.1991
Baidu ScholarGoogle Scholar
83R.B. Cakirli, R.F. Casten,

Direct Empirical Correlation between Proton-Neutron Interaction Strengths and the Growth of Collectivity in Nuclei

. Phys. Rev. Lett. 96, 132501 (2006). https://doi.org/10.1103/PhysRevLett.96.132501
Baidu ScholarGoogle Scholar
84K. Heyde, P. Van Isacker, R. Casten, et al.,

A shell-model interpretation of intruder states and the onset of deformation in even-even nuclei

. Physics Letters B 155, 303-308 (1985). https://doi.org/10.1016/0370-2693(85)91575-8
Baidu ScholarGoogle Scholar
85P. Federman, S. Pittel,

Hartree-fock-bogolyubov study of deformation in the Zr Mo region

. Physics Letters B 77, 29-32 (1978). https://doi.org/10.1016/0370-2693(78)90192-2
Baidu ScholarGoogle Scholar
86M. Hasegawa, K. Kaneko,

Microscopic analysis of T=1 and T=0 proton-neutron correlations in N=Z nuclei

. Nuclear Physics A 748, 393-401 (2005). https://doi.org/10.1016/j.nuclphysa.2004.11.014
Baidu ScholarGoogle Scholar
87S. Goriely, N. Chamel, J.M. Pearson,

Hartree-fock-bogoliubov nuclear mass model with 0.50 mev accuracy based on standard forms of skyrme and pairing functionals

. Phys. Rev. C 88, 061302 (2013). https://doi.org/10.1103/PhysRevC.88.061302
Baidu ScholarGoogle Scholar
88J. Duflo, A. Zuker,

Microscopic mass formulas

. Phys. Rev. C 52, R23-R27 (1995). https://doi.org/10.1103/PhysRevC.52.R23
Baidu ScholarGoogle Scholar
89M. Kortelainen, J. McDonnell, W. Nazarewicz, et al.,

Nuclear energy density optimization: Large deformations

. Phys. Rev. C 85, 024304 (2012). https://doi.org/10.1103/PhysRevC.85.024304
Baidu ScholarGoogle Scholar
90N. Wang, M. Liu, X. Wu, et al.,

Surface diffuseness correction in global mass formula

. Physics Letters B 734, 215-219 (2014). https://doi.org/10.1016/j.physletb.2014.05.049
Baidu ScholarGoogle Scholar
91T. KAWANO, S. CHIBA, H. KOURA,

Phenomenological nuclear level densities using the ktuy05 nuclear mass formula for applications off-stability

. Journal of Nuclear Science and Technology 43, 1-8 (2006). https://doi.org/10.1080/18811248.2006.9711062
Baidu ScholarGoogle Scholar
92P. Möller, A. Sierk, T. Ichikawa, et al.,

Nuclear ground-state masses and deformations: FRDM(2012)

. Atomic Data and Nuclear Data Tables 109-110, 1-204 (2016). https://doi.org/10.1016/j.adt.2015.10.002
Baidu ScholarGoogle Scholar
93L. Geng, H. Toki, J. Meng, Masses ,

Deformations and Charge Radii-Nuclear Ground-State Properties in the Relativistic Mean Field Model

. Progress of Theoretical Physics 113, 785-800 (2005). https://doi.org/10.1143/PTP.113.785
Baidu ScholarGoogle Scholar
94P. Moller, J. Nix, W. Myers, et al.,

Nuclear Ground-State Masses and Deformations

. Atomic Data and Nuclear Data Tables 59, 185-381 (1995). https://doi.org/10.1006/adnd.1995.1002
Baidu ScholarGoogle Scholar
95M. Karny, L. Batist, D. Jenkins, et al.,

Excitation energy of the T=0β-decaying 9+ isomer in 70Br

. Phys. Rev. C 70, 014310 (2004). https://doi.org/10.1103/PhysRevC.70.014310
Baidu ScholarGoogle Scholar
96D.G. Jenkins, N.S. Kelsall, C.J. Lister, et al.,

T=0 and T=1 states in the odd-odd N=Z nucleus, 3570Br35

. Phys. Rev. C 65, 064307 (2002). https://doi.org/10.1103/PhysRevC.65.064307
Baidu ScholarGoogle Scholar
97P. Schury, C. Bachelet, M. Block, et al.,

Precision mass measurements of rare isotopes near N=Z=33 produced by fast beam fragmentation

. Phys. Rev. C 75, 055801 (2007). https://doi.org/10.1103/PhysRevC.75.055801
Baidu ScholarGoogle Scholar
98I. Mardor, S.A.S. Andrés, T. Dickel, et al.,

Mass measurements of As, Se, and Br nuclei, and their implication on the proton-neutron interaction strength toward the N=Z line

. Phys. Rev. C 103, 034319 (2021). https://doi.org/10.1103/PhysRevC.103.034319
Baidu ScholarGoogle Scholar
99D.S. Brenner, R.B. Cakirli, R.F. Casten,

Valence proton-neutron interactions throughout the mass surface

. Phys. Rev. C 73, 034315 (2006). https://doi.org/10.1103/PhysRevC.73.034315
Baidu ScholarGoogle Scholar
100P. Van Isacker, O. Juillet, F. Nowacki,

Pseudo-SU(4) Symmetry in pf-Shell Nuclei

. Phys. Rev. Lett. 82, 2060-2063 (1999). https://doi.org/10.1103/PhysRevLett.82.2060
Baidu ScholarGoogle Scholar
101K. Hebeler, S.K. Bogner, R.J. Furnstahl, et al.,

Improved nuclear matter calculations from chiral low-momentum interactions

. Phys. Rev. C 83, 031301 (2011). https://doi.org/10.1103/PhysRevC.83.031301
Baidu ScholarGoogle Scholar
102Y.P. Wang, Y.K. Wang, F.F. Xu, et al.,

Abnormal bifurcation of the double binding energy differences and proton-neutron pairing: Nuclei close to N=Z line from ni to rb

. Phys. Rev. Lett. 132, 232501 (2024). https://doi.org/10.1103/PhysRevLett.132.232501
Baidu ScholarGoogle Scholar
103J. Jänecke,

Neutron-proton interaction in mirror nuclei

. Phys. Rev. C 6, 467-468 (1972). https://doi.org/10.1103/PhysRevC.6.467
Baidu ScholarGoogle Scholar
104Y.Y. Zong, C. Ma, Y.M. Zhao, et al.,

Mass relations of mirror nuclei

. Phys. Rev. C 102, 024302 (2020). https://doi.org/10.1103/PhysRevC.102.024302
Baidu ScholarGoogle Scholar
105X.L. Yan, H.S. Xu, Y.A. Litvinov, et al.,

MASS MEASUREMENT OF 45Cr AND ITS IMPACT ON THE Ca-Sc CYCLE IN X-RAY BURSTS

. The Astrophysical Journal 766, L8 (2013). https://doi.org/10.1088/2041-8205/766/1/l8
Baidu ScholarGoogle Scholar
106A.P. Zuker, S.M. Lenzi, G. Martínez-Pinedo, et al.,

Isobaric multiplet yrast energies and isospin nonconserving forces

. Phys. Rev. Lett. 89, 142502 (2002). https://doi.org/10.1103/PhysRevLett.89.142502
Baidu ScholarGoogle Scholar
107J. Ekman, D. Rudolph, C. Fahlander, et al.,

Unusual isospin-breaking and isospin-mixing effects in the A=35 mirror nuclei

. Phys. Rev. Lett. 92, 132502 (2004). https://doi.org/10.1103/PhysRevLett.92.132502
Baidu ScholarGoogle Scholar
108M.A. Bentley, C. Chandler, M.J. Taylor, et al.,

Isospin symmetry of odd-odd mirror nuclei: Identification of excited states in N=Z-248Mn

. Phys. Rev. Lett. 97, 132501 (2006). https://doi.org/10.1103/PhysRevLett.97.132501
Baidu ScholarGoogle Scholar
109A. Gadea, S.M. Lenzi, S. Lunardi, et al.,

Observation of 54Ni: Cross-conjugate symmetry in f7/2 mirror energy differences

. Phys. Rev. Lett. 97, 152501 (2006). https://doi.org/10.1103/PhysRevLett.97.152501
Baidu ScholarGoogle Scholar
110K. Kaneko, Y. Sun, T. Mizusaki, et al.,

Variation in displacement energies due to isospin-nonconserving forces

. Phys. Rev. Lett. 110, 172505 (2013). https://doi.org/10.1103/PhysRevLett.110.172505
Baidu ScholarGoogle Scholar
111H.H. Li, Q. Yuan, J.G. Li, et al.,

Investigation of isospin-symmetry breaking in mirror energy difference and nuclear mass with ab initio calculations

. Phys. Rev. C 107, 014302 (2023). https://doi.org/10.1103/PhysRevC.107.014302
Baidu ScholarGoogle Scholar
112J. Lee, X.X. Xu, K. Kaneko, et al.,

Large isospin asymmetry in 22Si/22O mirror gamow-teller transitions reveals the halo structure of 22Al

. Phys. Rev. Lett. 125, 192503 (2020). https://doi.org/10.1103/PhysRevLett.125.192503
Baidu ScholarGoogle Scholar
113H.H. Li, J.G. Li, M.R. Xie, et al.,

Ab initio calculations of mirror energy difference in sd-shell nuclei*

. Chinese Physics C 47, 124101 (2023). https://doi.org/10.1088/1674-1137/acf035
Baidu ScholarGoogle Scholar
114J. Li, H. Li, S. Zhang, et al.,

Double-magicity of proton drip-line nucleus 22Si with ab initio calculation

. Physics Letters B 846, 138197 (2023). https://doi.org/10.1016/j.physletb.2023.138197
Baidu ScholarGoogle Scholar
115G.A. Miller, A.K. Opper, E.J. Stephenson,

Charge symmetry breaking and QCD

. Annual Review of Nuclear and Particle Science 56, 253-292 (2006). https://doi.org/10.1146/annurev.nucl.56.080805.140446
Baidu ScholarGoogle Scholar
116Z.C. Xu, S. Zhang, J.G. Li, et al.,

Complex valence-space effective operators for observables: The Gamow-Teller transition

. Phys. Rev. C 108, L031301 (2023). https://doi.org/10.1103/PhysRevC.108.L031301
Baidu ScholarGoogle Scholar
117W. Benenson, E. Kashy,

Isobaric quartets in nuclei

. Rev. Mod. Phys. 51, 527-540 (1979). https://doi.org/10.1103/RevModPhys.51.527
Baidu ScholarGoogle Scholar
118Y.H. Zhang, H.S. Xu, Y.A. Litvinov, et al.,

Mass Measurements of the Neutron-Deficient 41Ti, 45Cr, 49Fe, and 53Ni Nuclides: First Test of the Isobaric Multiplet Mass Equation in fp-Shell Nuclei

. Phys. Rev. Lett. 109, 102501 (2012). https://doi.org/10.1103/PhysRevLett.109.102501
Baidu ScholarGoogle Scholar
119M. MacCormick, G. Audi,

Evaluated experimental isobaric analogue states from T=1/2 to T=3 and associated IMME coefficients

. Nuclear Physics A 925, 61-95 (2014). https://doi.org/10.1016/j.nuclphysa.2014.01.007
Baidu ScholarGoogle Scholar
120Y.H. Lam, B. Blank, N.A. Smirnova, et al.,

The isobaric multiplet mass equation for A≤71 revisited

. Atomic Data and Nuclear Data Tables 99, 680-703 (2013). https://doi.org/10.1016/j.adt.2012.11.002
Baidu ScholarGoogle Scholar
121J.M. Dong, J.Z. Gu, Y.H. Zhang, et al.,

Beyond Wigner’s isobaric multiplet mass equation: Effect of charge-symmetry-breaking interaction and Coulomb polarization

. Phys. Rev. C 99, 014319 (2019). https://doi.org/10.1103/PhysRevC.99.014319
Baidu ScholarGoogle Scholar
122W.J. Huang, X. Zhou, Y.H. Zhang, et al.,

Ground-state mass of the odd-odd N=Z nuclide 70Br

. Phys. Rev. C 108, 034301 (2023). https://doi.org/10.1103/PhysRevC.108.034301
Baidu ScholarGoogle Scholar
123J. Jänecke,

Vector and Tensor Coulomb Energies

. Phys. Rev. 147, 735-742 (1966). https://doi.org/10.1103/PhysRev.147.735
Baidu ScholarGoogle Scholar
124J.A. Nolen, J.P. Schiffer,

Coulomb Energies

. Annual Review of Nuclear Science 19, 471-526 (1969). https://doi.org/10.1146/annurev.ns.19.120169.002351
Baidu ScholarGoogle Scholar
125S. Shlomo,

Nuclear Coulomb energies

. Reports on Progress in Physics 41, 957-1026 (1978). https://doi.org/10.1088/0034-4885/41/7/001
Baidu ScholarGoogle Scholar
126W.E. Ormand,

Mapping the proton drip line up to A=70

. Phys. Rev. C 55, 2407-2417 (1997). https://doi.org/10.1103/PhysRevC.55.2407
Baidu ScholarGoogle Scholar
127B.A. Brown, R.R.C. Clement, H. Schatz, et al.,

Proton drip-line calculations and the rp process

. Phys. Rev. C 65, 045802 (2002). https://doi.org/10.1103/PhysRevC.65.045802
Baidu ScholarGoogle Scholar
128X. Zhou, J. Yang, the HIAF project team,

Status of the high-intensity heavy-ion accelerator facility in China

. AAPPS Bulletin 32, 35 (2022). https://doi.org/10.1007/s43673-022-00064-1
Baidu ScholarGoogle Scholar
129B. Wu, J. Yang, J. Xia, et al.,

The design of the Spectrometer Ring at the HIAF

. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 881, 27-35 (2018). https://doi.org/10.1016/j.nima.2017.08.017
Baidu ScholarGoogle Scholar
130J. Yang, J. Xia, G. Xiao, et al.,

High intensity heavy ion accelerator facility (HIAF) in China

. Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms 317, 263-265 (2013). XVIth International Conference on ElectroMagnetic Isotope Separators and Techniques Related to their Applications, December 2-7, 2012 at Matsue, Japan. https://doi.org/10.1016/j.nimb.2013.08.046
Baidu ScholarGoogle Scholar
131J. Liu, J.C. Yang, J.W. Xia, et al.,

Transverse impedances and collective instabilities in a heavy ion accelerator

. Phys. Rev. Accel. Beams 21, 064403 (2018). https://doi.org/10.1103/PhysRevAccelBeams.21.064403
Baidu ScholarGoogle Scholar
Footnote

Dedicated to Professor Wenqing Shen in honour of his 80th birthday