Next Article in Journal
Cross Sections for Coherent Elastic and Inelastic Neutrino-Nucleus Scattering
Previous Article in Journal
Quasibound States, Stability and Wave Functions of the Test Fields in the Consistent 4D Einstein–Gauss–Bonnet Gravity
Previous Article in Special Issue
Pauli Exclusion Classical Potential for Intermediate-Energy Heavy-Ion Collisions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Probing the Neutron Skin of Unstable Nuclei with Heavy-Ion Collisions

1
China Institute of Atomic Energy, Beijing 102413, China
2
Guangxi Key Laboratory of Nuclear Physics and Technology, Guangxi Normal University, Guilin 541004, China
*
Author to whom correspondence should be addressed.
Universe 2023, 9(5), 206; https://doi.org/10.3390/universe9050206
Submission received: 13 March 2023 / Revised: 31 March 2023 / Accepted: 18 April 2023 / Published: 25 April 2023

Abstract

:
To improve the constraints of symmetry energy at subsaturation density, measuring and accumulating more neutron skin data for neutron-rich unstable nuclei is naturally required. Aiming to probe the neutron skin of unstable nuclei by using low-intermediate-energy heavy-ion collisions, we develop a new version of an improved quantum molecular dynamics model, in which the neutron skin of the initial nucleus and the mean-field potential in nucleon propagation are consistently treated. Our calculations show that the three observables, such as the cross-sections of the primary projectile-like residues with A > 100 ( σ A > 100 ), the difference of σ A > 100 between 132 Sn + 124 Sn and 124 Sn + 124 Sn systems ( δ σ A > 100 ), and the neutron-to-proton yield ratio ( R ( n / p ) ) in the transverse direction, could be used to measure the neutron skin of the unstable nuclei and to constrain the slope of the symmetry energy in the future.
PACS:
21.60.Jz; 21.65.Ef; 24.10.Lx; 25.70.-z

1. Introduction

The thickness of the neutron skin of a nucleus is characterized as
Δ r n p = r n 2 1 / 2 r p 2 1 / 2 ,
which reflects the difference between the root-mean-square (rms) radii of the neutron and proton density distributions in the nucleus and is strongly correlated with the slope of the density dependence of the symmetry energy [1,2,3,4,5,6]. Thus, accurate measurements of the neutron skin of a nucleus can be used to constrain the symmetry energy at subsaturation density.
For the determination of the neutron skin of a nucleus, measurements of the proton and neutron density distributions are needed. The proton density distribution can be accurately determined by electron elastic scattering experiments or isotope shift measurements [7,8,9], but the neutron density distribution is difficult to measure accurately. The reason is that the neutron is neutral and interacts mainly with hadronic probes. Therefore, the neutron density is mainly probed by proton elastic scattering [10,11,12,13,14,15,16], inelastic α scattering [17,18], coherent pion photoproduction scattering [19], antiprotonic atoms [20,21,22], and relativistic energy heavy-ion collisions (HICs) [23,24,25,26]. Another method to measure neutron density is to use weak electric probes, such as the parity-violating e-A scattering method [27,28,29,30,31] of the parity radius experiment at the Jefferson Laboratory, PREX-I [29], PREX-II [30] and CREX [31], or through coherent elastic neutrino–nucleus scattering [32,33]. All of the above methods are mainly used to measure the neutron skin of stable nuclei on the nuclear chart.
A great deal of theoretical analysis on the neutron skin of stable nuclei has been performed to constrain the symmetry energy at a subsaturation density [4,6,34,35,36,37,38,39,40,41,42], and the extrapolated values of the symmetry energy coefficient S 0 and the slope of the symmetry energy L are in 29–35 MeV and 5–80 MeV [4,35,38,39,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57], respectively. However, tension appeared after PREX-II published high-accuracy 208 Pb data since the analysis, with a special class of a relativistic mean-field approach, favors a very stiff symmetry energy, i.e., the constrained value of the symmetry energy coefficient S 0 = 38.1 ± 4.7 MeV and the slope of the symmetry energy L = 106 ± 37 MeV [58] are much larger than previously obtained.
To understand the tension and improve the constraints of the symmetry energy at subsaturation density, two aspects should be investigated. One is to understand the influence of cluster mechanisms in the nucleus [59,60], and we will not touch upon this in this work. Another is to enhance the reliability of constraints by using as much data as possible. The neutron skin data for stable nuclei has been analyzed in Refs. [4,6,34,35,36,37,38,39,40,41,42] to constrain the symmetry energy. For further constraints, accumulating more neutron skin data of neutron-rich unstable nuclei is necessary, and it naturally requires developing a method for measuring the neutron skin of unstable neutron-rich nuclei.
There have been some efforts to measure the neutron skin or the neutron density distribution for unstable nuclei. For example, total reaction cross-sections [61], cross-sections of the isovector spin-dipole resonances (SDR) excited by the ( 3 He, t) charge-exchange reaction [62], the strength of pygmy dipole resonances [63], neutron-removal cross-sections in high-energy nuclear collisions [64], and charged pion multiplicity ( π / π + ) ratios in peripheral heavy-ion collisions [65,66] have been used or proposed. Among these methods, using heavy-ion collisions is most suitable to measure the neutron skin of a very neutron-rich nucleus since the unstable neutron-rich nuclei in a wide range of isospin asymmetry are mainly produced by the projectile fragmentation mechanism in next-generation isotope facilities [67].
However, the measurement of the neutron skin of unstable nuclei via HICs depends on the transport models. In the pioneer transport model calculations [65,66], the slope of the symmetry energy and the thickness of the neutron skin were treated separately. The separate treatment of the neutron skin in the initial nuclei and the mean-field potential in the nucleon propagation increases the theoretical uncertainties for probing the neutron skin of unstable nuclei. Thus, a consistent treatment of the neutron skin in the initialization and the isospin-dependent mean-field potential in nucleon propagation is highly desired.
In this work, the neutron skin of the initial nucleus is correlated to the mean-field potential more consistently using the same Skyrme energy density functional in the updated version of the improved quantum molecular dynamics model (ImQMD-L) [68]. Based on the updated ImQMD-L model, the effects of neutron skin on the collision of 124 , 132 Sn + 124 Sn at 200 MeV/u are investigated. Our calculations show that the cross-sections of primary projectile-like residues with A > 100 can be used to distinguish the thickness of the neutron skin. Furthermore, the energy spectra of the yield ratios of emitted neutrons to protons are also analyzed, which can be used for a complementary understanding of the neutron skin effects and the reaction mechanism.

2. Theoretical Framework

In this part, we briefly introduce the form of the potential energy density in the ImQMD-L model and how we correlate the neutron skin of an initial nucleus to the mean-field potential in nucleon propagation with the same Skyrme energy density functional applied.

2.1. Potential Energy Density

In the ImQMD-L model [68], the Skyrme-type nucleonic potential energy density without the spin-orbit term is used,
u ( r ) s k y = α 2 ρ 2 ρ 0 + β η + 1 ρ η + 1 ρ 0 η + g s u r 2 ρ 0 ( ρ ) 2 + g s u r , i s o ρ 0 [ ( ρ n ρ p ) ] 2 + A s y m ρ 2 ρ 0 δ 2 + B s y m ρ η + 1 ρ 0 η δ 2 + u m d .
Here, ρ is the number density of nucleons, which is the summation of neutron density and proton denisity, i.e., ρ = ρ n + ρ p . δ is the isospin asymmetry, which is defined as δ = ρ n ρ p ρ n + ρ p . α is the parameter related to the two-body term, β and η are related to the three-body term, g s u r and g s u r , i s o are related to the surface terms, and A s y m and B s y m are the coefficients of the symmetry potential and come from the two- and the three-body interaction terms [69]. u m d is the Skyrme-type momentum-dependent energy density functional, and it is obtained based on its interaction form δ ( r 1 r 2 ) ( p 1 p 2 ) 2 [70], i.e.,
u m d ( r , { p i p j } ) = C 0 i j d 3 p d 3 p f i ( r , p ) f j ( r , p ) ( p p ) 2 + D 0 i j n d 3 p d 3 p f i ( r , p ) f j ( r , p ) ( p p ) 2 + D 0 i j p d 3 p d 3 p f i ( r , p ) f j ( r , p ) ( p p ) 2 .
C 0 and D 0 are the parameters related to the momentum-dependent interaction. More details about it can be found in Ref. [68].
The parameters in Equations (2) and (3) are obtained from the standard Skyrme interaction parameters as in Refs. [43,71]. The connection between the seven parameters, α , β , η , A s y m , B s y m , C 0 , and D 0 , used in the ImQMD-L model and the seven nuclear matter parameters, including the saturation density ρ 0 , the binding energy at the saturation density E 0 , the incompressibility K 0 , the symmetry energy coefficient S 0 , the slope of the symmetry energy L, the isoscalar effective mass m s * , and the isovector effective mass m v * , are given in Ref. [43]. In the following calculations, the g s u r and g s u r , i s o are 24.5 and −4.99 MeVfm 2 , respectively. Thus, one can alternatively use ρ 0 , E 0 , K 0 , S 0 , L, m s * , and m v * as input to study the influence of different nuclear matter parameters. In this work, we vary only the L to change the thickness of the neutron skin of the nucleus. All the parameters we used are listed in Table 1.

2.2. Initialization with Neutron Skin

To consistently correlate the neutron skin of initial nuclei with the mean-field potential in nucleon propagation, one has to know the neutron and proton density distributions first and then find a way to approximate the density distributions in the ImQMD-L model.
As discussed in Ref. [68], the Woods–Saxon density profile can be reproduced by sampling the centroids of the wave packets within the hard sphere with a radius that equals the half-density radius of the Woods–Saxon density profile. The relationship of the width of the wave packet σ r and the diffuseness of the nucleus a is σ r = f ( a ) = ( 1.71217 ± 0.01548 ) a + ( 0.01564 ± 0.01047 ) fm. Thus, a model that can be used to calculate the Woods–Saxon-type density distribution under the same Skyrme energy density functional is needed.
The restricted density variational (RDV) method [72] meets this criteria. In the RDV method, the density distributions
ρ i = ρ 0 i 1 1 + exp ( r R i a i ) , i = n , p ,
are adopted, where R p , a p , R n , and a n are the radius and diffuseness of the proton and neutron density distributions, respectively. ρ 0 i is the central density of neutrons or protons in the nucleus. These parameters are obtained by minimizing the total energy of the system, which is given by
E = H d 3 r = { 2 2 m [ τ n ( r ) + τ p ( r ) ] + u s k y + u c o u l } d 3 r ,
under the condition of the conservation of the number of particles in the system, i.e., N = ρ n ( r ) d 3 r and Z = ρ p ( r ) d 3 r . u c o u l is the Coulomb energy density, and τ n and τ p are the kinetic energy densities of neutrons and protons, respectively. The kinetic energy density in the RDV method is given by the extended Thomas–Fermi (ETF) approach, which includes all terms up to the second order (ETF2) and fourth order (ETF4), as in Ref. [73]. The same semiclassical expression of the Skyrme energy density functional as in ImQMD-L is used to calculate u s k y .
In Table 2, we show the RDV results of the binding energy B, the diffuseness parameters a p and a n , the half-density radius R p and R n , the rms radius for neutrons and protons, and the thickness of the neutron skin, respectively. Five Skyrme parameter sets, which are represented by different L values, are used for varying the thickness of the neutron skin. The upper part is for 124 Sn, and the bottom part is for 132 Sn.
Then, we prepare the initialization in the same manner as in Ref. [69], but with different treatments in the following two aspects. The first one is that the centroids of the wave packets for neutrons and protons are sampled within the half-density radius of the neutron and the proton obtained in the RDV method, i.e., R n and R p . Once the positions of all the nucleons have been determined, the density distribution is obtained. Then, the momenta of the nucleons are sampled according to the local density approach. The second one is that the binding energy of the sampled nucleus falling into the range of B ± 0.2 MeV is also required, where B is obtained by the RDV method.

3. Results and Discussions

For peripheral collisions at intermediate-energy HICs, there are two characteristics associated with the neutron skin of the nuclei. One is the size of the projectile-like and target-like residues; another is the isospin content of the nucleons and light particles, which are emitted in the transverse direction. The mechanism of the above two characteristics for peripheral HICs is shown in the schematic diagram in Figure 1. Panel (a) is the initial stage of the reaction, panel (b) is at the reaction stage, and panel (c) is at the later stage of the reaction where the projectile and target like fragments are formed and the light articles are emitted from the neck region. Thus, one can expect that a larger neutron skin could lead to a larger reaction cross-section or larger production cross-sections for projectile/target-like residues, as well as the emission of more neutrons and neutron-rich light particles.
To quantitatively understand the effects of neutron skin on heavy-ion collision observables, we simulate the collisions of 124 , 132 Sn + 124 Sn at a beam energy of 200 MeV/u and with the impact parameter b ranging from b = 5 fm to b m a x fm with Δ b = 0.5 fm. b m a x is calculated as
b m a x = R p r o j m a x + R t a r m a x + 2.2 ( a p r o j m a x + a t a r m a x ) .
For each impact parameter, 5000 events were simulated. The values of the half-density radius R p r o j / t a r m a x are the maximum values between the neutron half-density radius and the proton half-density radius of the projectile or target, and the values of a p r o j / t a r m a x have a similar meaning. The values of R n / p and a n / p are obtained with RDV and are listed in Table 2. The term with 2.2 a m a x in Equation (6) is used to consider the surface thickness of the nucleus.
In Figure 2a, we present P A > 100 ( b ) , which means the probability of observing a primary heavy residue with mass number A > 100 in the forward region, i.e., a projectile-like residue. The red lines are the results of 132 Sn + 124 Sn, and the black lines are the results of 124 Sn + 124 Sn. The P A > 100 ( b ) quickly increases from zero to one from semi-peripheral to peripheral collisions, i.e., in the range of b = 6.2–9.2 fm (b = 6.6–10.0 fm), for 132 Sn + 124 Sn ( 124 Sn + 124 Sn). This behavior is determined by the reaction mechanism. When b < 6 fm, multifragmentation occurs, and there are no fragments with A > 100 . At b > 10 fm, the distance between the projectile and the target is large enough to produce a heavy projectile-like residue with A > 100 in each event, and then P A > 100 ( b ) = 1 . The interesting point is that curves of P A > 100 ( b ) show a sensitivity to the thickness of the neutron skin or to the slope of the symmetry energy. In the range of b = 6–10 fm, the larger the values of L, the greater the P A > 100 . This is because the calculations with smaller L can reach a higher density in the overlap region, and thus, more compressional energy is stored than for the calculations with large L. Thus, the reaction system simulated with smaller L disintegrates into more light fragments compared to the case of the calculations with large L. Owing to the conservation of the nucleon number in the reaction system, the production of projectile-like residues with A > 100 calculated with larger L is higher than with a smaller L. It is consistent with the calculations in Refs. [74,75].
To understand where the projectile-like residues can be detected, we also present the θ c . m . distribution of projectile-like residues in Figure 2b. The projectile-like residues deflect in a small direction along the beam direction and are distributed within θ c . m . < 4 .
The cross-section for the projectile-like residues with A > 100 is calculated as
σ A > 100 = 2 π 0 b m a x P A > 100 ( b ) b d b .
In Figure 3, σ A > 100 as a function of the neutron skin thickness of a system, i.e.,
Δ R = Δ r n p p r o j + Δ r n p t a r g ,
is presented. Figure 3a,b are for 124 Sn + 124 Sn and 132 Sn + 124 Sn, respectively. The calculations illustrate that σ A > 100 increases with Δ R or the slope of the symmetry energy for both systems. For 124 Sn + 124 Sn, σ A > 100 is enhanced by a factor of ∼5.2% as Δ R increases from 0.33 fm to 0.526 fm or as L varies from 30 to 110 MeV. For 132 Sn + 124 Sn, σ A > 100 is enhanced by a factor of ∼5.4% as Δ R increases from 0.397 fm to 0.624 fm. Thus, measuring σ A > 100 could be used to obtain the neutron skin of the nucleus and the slope of the symmetry energy.
Nevertheless, the calculated results of σ A > 100 may be model-dependent or biased. One needs to seek a way to avoid or at least suppress the possible systematic uncertainty caused by the model. Ideally, the difference of σ A > 100 between system A s y s and B s y s , i.e.,
δ σ A > 100 = σ A > 100 ( A s y s ) σ A > 100 ( B s y s ) ,
where A s y s = 132 Sn + 124 Sn and B s y s = 124 Sn + 124 Sn, can be used. It is based on a situation in which the systematic uncertainty caused by the same model is similar for the two systems. In Figure 3c, we present δ σ A > 100 as a function of the difference of the neutron skins between A s y s and B s y s , i.e.,
δ R = Δ R ( A s y s ) Δ R ( B s y s ) .
Our calculations show that δ σ A > 100 keeps the sensitivity to δ R and L. Therefore, δ σ A > 100 could be used to probe the neutron skin of unstable nuclei and constrain the symmetry energy.
As charge size of fragments can be more easily measured than the mass of fragments, we also studied the sensitivity of the cross-section for the projectile-like residues with the charge number Z > 40 to L. Our calculations show that σ Z > 40 is also sensitive to L, but the sensitivity becomes weaker than σ A > 100 . This is because the cross-section measured by σ Z > 40 loses the information of the neutron number of fragments, so the sensitivity of σ Z > 40 to L is weaker than σ A > 100 .
Next, we analyze the neutron-to-proton yield ratio in the transverse direction. The transverse direction in this work corresponds to 70 < θ c . m . < 110 in the center-of-mass frame. This observable was first proposed to probe the strength of the symmetry potential in Ref. [76] and has been studied extensively for constraining the symmetry energy and effective mass splitting [51,74,77,78,79,80,81,82,83,84].
In this work, R ( n / p ) is obtained for peripheral collisions, with the impact parameter ranging from 5 fm to 11 fm as
R ( n / p ) = b = 5 11 d Y n ( b ) d E k 2 π b d b / b = 5 11 d Y p ( b ) d E k 2 π b d b .
This mainly reflects the information of the isospin contents of the overlap region, which is strongly correlated with the thickness of the neutron skin and the slope of the symmetry energy. The calculated results for R ( n / p ) are presented in Figure 4a,b. The black, red, and blue regions are the results for three values of Δ R , corresponding to L = 30, 70, and 110 MeV, respectively. Figure 4a is for 124 Sn + 124 Sn, and Figure 4b is for the 132 Sn + 124 Sn system. For the emitted nucleons with kinetic energy E k < 80 MeV, the R ( n / p ) values are greater for a thin neutron skin than that for a thick neutron skin. This corresponds to the R ( n / p ) values being greater for the symmetry energy with a small L case than for the symmetry energy with a large L. The reason is that the emitted nucleons with lower kinetic energy mainly come from the subsaturation density region, where the symmetry energy obtained with small L is larger than that with a large L. In addition, stronger effects are observed for 132 Sn + 124 Sn than for 124 Sn + 124 Sn.
For the emitted nucleons a kinetic energy E k 80 MeV, the R ( n / p ) values are greater for the symmetry energy with a large L case than for the symmetry energy with a small L. The reason is that the emitted nucleons with high kinetic energy mainly come from the high-density region, where the symmetry energy obtained with a large L is larger than with a small L. The calculations show that R ( n / p ) increases with Δ R but with large uncertainties. To further reduce the uncertainty and distinguish Δ R with a higher accuracy, the statistics in simulations must be increased. Another way is to integrate the yields of neutrons and protons above 80 MeV and then calculate the neutron-to-proton ratio as
R 80 ( n / p ) = b = 5 11 Y n ( b , E k 80 ) 2 π b d b b = 5 11 Y p ( b , E k 80 ) 2 π b d b .
In Figure 4c,d, we present R 80 ( n / p ) as a function of Δ R for 124 Sn + 124 Sn and 132 Sn + 124 Sn, respectively. According to the absolute statistical uncertainty of R 80 ( n / p ) in the current calculation, i.e., e r r ( R 80 ) = 0.01 , one can distinguish the neutron skin of the unstable nucleus with an accuracy of ∼0.02 fm.

4. Summary and Outlook

In summary, we consistently correlate the neutron skin of nuclei in the initialization and the isospin dependent mean-field potential in nucleon propagation with the same energy density functional in the improved quantum molecular dynamics for probing the neutron skin of an unstable nucleus. The unstable nucleus of 132 Sn on the target 124 Sn at peripheral collisions and the beam energy of 200 MeV per nucleon is simulated. Our calculations show that the cross-section of projectile-like residues σ A > 100 are correlated with the neutron skin of the system. To avoid the possible systematic deviations from the model, we also construct an observable δ σ A > 100 = σ A > 100 ( A s y s ) σ A > 100 ( B s y s ) , which reflects the difference of σ A > 100 between two systems A s y s = 132 Sn + 124 Sn and B s y s = 124 Sn + 124 Sn. Our calculations illustrate that δ σ A > 100 keeps the sensitivity to the thickness of the neutron skin of the unstable nucleus and the slope of the symmetry energy.
In addition, the neutron-to-proton yield ratios, i.e., R ( n / p ) , are also sensitive to the thickness of the neutron skin. In the low-kinetic-energy region, R ( n / p ) is negatively correlated with the thickness of the neutron skin. In the high-kinetic-energy region, R ( n / p ) is positively correlated with the thickness of the neutron skin. Thus, the combination analysis on the δ σ A > 100 and R ( n / p ) in different kinetic energy regions could improve the reliability and accuracy of the measurements of neutron skins using heavy-ion collisions.
However, one should note that the probe of δ σ A > 100 requires experimentalists to develop a method for reconstructing primary projectile-like residues from the emitted light particles and cold fragments. Currently, a kinematical focusing method has been developed for the reconstruction of intermediate-mass fragments in Ref. [85]. Even though there will be many difficulties in reconstructing the primary projectile-like residues, it still implies that the observables constructed from primary projectile-like residues may be in practical use in the future, and the probe of δ σ A > 100 and R ( n / p ) should be considered in the potential wish list of experimenters.

Author Contributions

Conceptualization, Y.Z.; methodology, J.Y., X.C. and Y.Z.; software, Y.Z.; validation, J.Y., X.C., Y.C. and Y.Z.; formal analysis, J.Y. and Y.Z.; investigation, J.Y. and Y.Z.; resources, Y.Z.; data curation, J.Y.; writing—original draft preparation, J.Y. and Y.Z.; writing—review and editing, J.Y., X.C., Y.C., Z.L. and Y.Z.; visualization, J.Y.; supervision, Y.Z.; project administration, Y.Z.; funding acquisition, Y.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partly inspired by the transport code comparison project, and it was supported by the National Natural Science Foundation of China Nos. 12275359, 11875323, and 11961141003, the National Key R&D Program of China under Grant No. 2018YFA0404404, the Continuous Basic Scientific Research Project (No. WDJC-2019-13, BJ20002501), and the funding of the China Institute of Atomic Energy (No. YZ222407001301). The work was carried out at the National Supercomputer Center in Tianjin, and the calculations were performed on TianHe-1 (A).

Data Availability Statement

The data presented in this study (i.e., data from simulations) are available on request from the corresponding author.

Acknowledgments

The authors are thankful for the useful discussion with Z.G. Xiao and W.P. Lin.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Alex Brown, B. Neutron Radii in Nuclei and the Neutron Equation of State. Phys. Rev. Lett. 2000, 85, 5296–5299. [Google Scholar] [CrossRef] [PubMed]
  2. Typel, S.; Brown, B.A. Neutron radii and the neutron equation of state in relativistic models. Phys. Rev. C 2001, 64, 027302. [Google Scholar] [CrossRef]
  3. Furnstahl, R. Neutron radii in mean-field models. Nucl. Phys. A 2002, 706, 85–110. [Google Scholar] [CrossRef]
  4. Zhang, Z.; Chen, L.W. Constraining the symmetry energy at subsaturation densities using isotope binding energy difference and neutron skin thickness. Phys. Lett. B 2013, 726, 234–238. [Google Scholar] [CrossRef]
  5. Horowitz, C.J.; Piekarewicz, J. Neutron Star Structure and the Neutron Radius of 208Pb. Phys. Rev. Lett. 2001, 86, 5647–5650. [Google Scholar] [CrossRef]
  6. Roca-Maza, X.; Centelles, M.; Viñas, X.; Warda, M. Neutron Skin of 208Pb, Nuclear Symmetry Energy, and the Parity Radius Experiment. Phys. Rev. Lett. 2011, 106, 252501. [Google Scholar] [CrossRef]
  7. Fricke, G.; Bernhardt, C.; Heilig, K.; Schaller, L.; Schellenberg, L.; Shera, E.; Dejager, C. Nuclear Ground State Charge Radii from Electromagnetic Interactions. At. Data Nucl. Data Tables 1995, 60, 177–285. [Google Scholar] [CrossRef]
  8. Angeli, I.; Gangrsky, Y.P.; Marinova, K.P.; Boboshin, I.N.; Komarov, S.Y.; Ishkhanov, B.S.; Varlamov, V.V. N and Z dependence of nuclear charge radii. J. Phys. G Nucl. Part. Phys. 2009, 36, 085102. [Google Scholar] [CrossRef]
  9. Angeli, I.; Marinova, K. Table of experimental nuclear ground state charge radii: An update. At. Data Nucl. Data Tables 2013, 99, 69–95. [Google Scholar] [CrossRef]
  10. Ray, L. Proton-nucleus total cross sections in the intermediate energy range. Phys. Rev. C 1979, 20, 1857–1872. [Google Scholar] [CrossRef]
  11. Hoffmann, G.W.; Ray, L.; Barlett, M.; McGill, J.; Adams, G.S.; Igo, G.J.; Irom, F.; Wang, A.T.M.; Whitten, C.A.; Boudrie, R.L.; et al. 0.8 GeV p+208Pb elastic scattering and the quantity Δrnp. Phys. Rev. C 1980, 21, 1488–1494. [Google Scholar] [CrossRef]
  12. Starodubsky, V.E.; Hintz, N.M. Extraction of neutron densities from elastic proton scattering by 206,207,208Pb at 650 MeV. Phys. Rev. C 1994, 49, 2118–2135. [Google Scholar] [CrossRef] [PubMed]
  13. Kanada-En’yo, Y. Deformation effects on the surface neutron densities of stable S and Ni isotopes probed by proton elastic scattering via isotopic analysis. arXiv 2022, arXiv:2208.00590. [Google Scholar]
  14. Zenihiro, J.; Sakaguchi, H.; Murakami, T.; Yosoi, M.; Yasuda, Y.; Terashima, S.; Iwao, Y.; Takeda, H.; Itoh, M.; Yoshida, H.P.; et al. Neutron density distributions of 204,206,208Pb deduced via proton elastic scattering at Ep=295 MeV. Phys. Rev. C 2010, 82, 044611. [Google Scholar] [CrossRef]
  15. Terashima, S.; Sakaguchi, H.; Takeda, H.; Ishikawa, T.; Itoh, M.; Kawabata, T.; Murakami, T.; Uchida, M.; Yasuda, Y.; Yosoi, M.; et al. Proton elastic scattering from tin isotopes at 295 MeV and systematic change of neutron density distributions. Phys. Rev. C 2008, 77, 024317. [Google Scholar] [CrossRef]
  16. Kłos, B.; Trzcińska, A.; Jastrzębski, J.; Czosnyka, T.; Kisieliński, M.; Lubiński, P.; Napiorkowski, P.; Pieńkowski, L.; Hartmann, F.J.; Ketzer, B.; et al. Neutron density distributions from antiprotonic 208Pb and 209Bi atoms. Phys. Rev. C 2007, 76, 014311. [Google Scholar] [CrossRef]
  17. Krasznahorkay, A.; Akimune, H.; van den Berg, A.; Blasi, N.; Brandenburg, S.; Csatlos, M.; Fujiwara, M.; Gulyas, J.; Harakeh, M.; Hunyadi, M.; et al. Neutron-skin thickness in neutron-rich isotopes. Nucl. Phys. A 2004, 731, 224–234. [Google Scholar] [CrossRef]
  18. Krasznahorkay, A.; Balanda, A.; Bordewijk, J.; Brandenburg, S.; Harakeh, M.; Kalantar-Nayestanaki, N.; Nyako, B.; Timar, J.; van der Woude, A. Excitation of the isovector GDR by inelastic α-scattering as a measure of the neutron skin of nuclei. Nucl. Phys. A 1994, 567, 521–540. [Google Scholar] [CrossRef]
  19. Tarbert, C.M.; Watts, D.P.; Glazier, D.I.; Aguar, P.; Ahrens, J.; Annand, J.R.M.; Arends, H.J.; Beck, R.; Bekrenev, V.; Boillat, B.; et al. Neutron Skin of 208Pb from Coherent Pion Photoproduction. Phys. Rev. Lett. 2014, 112, 242502. [Google Scholar] [CrossRef]
  20. Trzcińska, A.; Jastrzȩbski, J.; Lubiński, P.; Hartmann, F.J.; Schmidt, R.; von Egidy, T.; Kłos, B. Neutron Density Distributions Deduced from Antiprotonic Atoms. Phys. Rev. Lett. 2001, 87, 082501. [Google Scholar] [CrossRef]
  21. Jastrzbski, J.; Trzcinska, A.; Lubinski, P.; Klos, B.; Hartmann, F.J.; von Egidy, T.; Wycech, S. Neutron density distributions from antiprotonic atoms compared with hadron scattering data. Int. J. Mod. Phys. E 2004, 13, 343–351. [Google Scholar] [CrossRef]
  22. Brown, B.A.; Shen, G.; Hillhouse, G.C.; Meng, J.; Trzcinska, A. Neutron skin deduced from antiprotonic atom data. Phys. Rev. C 2007, 76, 034305. [Google Scholar] [CrossRef]
  23. Xu, H. Constraints on neutron skin thickness and nuclear deformations using relativistic heavy-ion collisions from STAR. arXiv 2022, arXiv:2208.06149. [Google Scholar] [CrossRef]
  24. Liu, L.M.; Xu, J.; Peng, G.X. Measuring deformed neutron skin with free spectator nucleons in relativistic heavy-ion collisions. Phys. Lett. B 2023, 838, 137701. [Google Scholar] [CrossRef]
  25. Xu, H.; Li, H.; Wang, X.; Shen, C.; Wang, F. Determine the neutron skin type by relativistic isobaric collisions. Phys. Lett. B 2021, 819, 136453. [Google Scholar] [CrossRef]
  26. Li, H.; Xu, H.-j.; Zhou, Y.; Wang, X.; Zhao, J.; Chen, L.W.; Wang, F. Probing the Neutron Skin with Ultrarelativistic Isobaric Collisions. Phys. Rev. Lett. 2020, 125, 222301. [Google Scholar] [CrossRef]
  27. Horowitz, C.J.; Pollock, S.J.; Souder, P.A.; Michaels, R. Parity violating measurements of neutron densities. Phys. Rev. C 2001, 63, 025501. [Google Scholar] [CrossRef]
  28. Horowitz, C.J.; Ahmed, Z.; Jen, C.M.; Rakhman, A.; Souder, P.A.; Dalton, M.M.; Liyanage, N.; Paschke, K.D.; Saenboonruang, K.; Silwal, R.; et al. Weak charge form factor and radius of 208Pb through parity violation in electron scattering. Phys. Rev. C 2012, 85, 032501. [Google Scholar] [CrossRef]
  29. Abrahamyan, S.; Ahmed, Z.; Albataineh, H.; Aniol, K.; Armstrong, D.S.; Armstrong, W.; Averett, T.; Babineau, B.; Barbieri, A.; Bellini, V.; et al. Measurement of the Neutron Radius of 208Pb through Parity Violation in Electron Scattering. Phys. Rev. Lett. 2012, 108, 112502. [Google Scholar] [CrossRef]
  30. Adhikari, D.; Albataineh, H.; Androic, D.; Aniol, K.; Armstrong, D.S.; Averett, T.; Ayerbe Gayoso, C.; Barcus, S.; Bellini, V.; Beminiwattha, R.S.; et al. Accurate Determination of the Neutron Skin Thickness of 208Pb through Parity-Violation in Electron Scattering. Phys. Rev. Lett. 2021, 126, 172502. [Google Scholar] [CrossRef]
  31. Adhikari, D.; Albataineh, H.; Androic, D.; Aniol, K.A.; Armstrong, D.S.; Averett, T.; Ayerbe Gayoso, C.; Barcus, S.K.; Bellini, V.; Beminiwattha, R.S.; et al. Precision Determination of the Neutral Weak Form Factor of 48Ca. Phys. Rev. Lett. 2022, 129, 042501. [Google Scholar] [CrossRef] [PubMed]
  32. Akimov, D.; AIbert, J.; An, P.; Awe, C.; Barbean, P.; Becker, B.; Belov, V.; Brown, A.; Bolozdynya, A.; Cabrera-Palmer, B.; et al. Observation of coherent elastic neutrino-nucleus scattering. Science 2017, 357, 1123–1126. [Google Scholar] [CrossRef] [PubMed]
  33. Cadeddu, M.; Giunti, C.; Li, Y.F.; Zhang, Y.Y. Average CsI Neutron Density Distribution from COHERENT Data. Phys. Rev. Lett. 2018, 120, 072501. [Google Scholar] [CrossRef]
  34. Centelles, M.; Roca-Maza, X.; Viñas, X.; Warda, M. Nuclear Symmetry Energy Probed by Neutron Skin Thickness of Nuclei. Phys. Rev. Lett. 2009, 102, 122502. [Google Scholar] [CrossRef]
  35. Chen, L.W.; Ko, C.M.; Li, B.A.; Xu, J. Density slope of the nuclear symmetry energy from the neutron skin thickness of heavy nuclei. Phys. Rev. C 2010, 82, 024321. [Google Scholar] [CrossRef]
  36. Xu, J.; Xie, W.J.; Li, B.A. Bayesian inference of nuclear symmetry energy from measured and imagined neutron skin thickness in 116,118,120,122,124,130,132Sn,208Pb, and 48Ca. Phys. Rev. C 2020, 102, 044316. [Google Scholar] [CrossRef]
  37. Zhang, Z.; Chen, L.W. Constraining the density slope of nuclear symmetry energy at subsaturation densities using electric dipole polarizability in 208Pb. Phys. Rev. C 2014, 90, 064317. [Google Scholar] [CrossRef]
  38. Oertel, M.; Hempel, M.; Klähn, T.; Typel, S. Equations of state for supernovae and compact stars. Rev. Mod. Phys. 2017, 89, 015007. [Google Scholar] [CrossRef]
  39. Li, B.A.; Han, X. Constraining the neutron-proton effective mass splitting using empirical constraints on the density dependence of nuclear symmetry energy around normal density. Phys. Lett. B 2013, 727, 276–281. [Google Scholar] [CrossRef]
  40. Gaidarov, M.K.; Antonov, A.N.; Sarriguren, P.; de Guerra, E.M. Symmetry energy of deformed neutron-rich nuclei. Phys. Rev. C 2012, 85, 064319. [Google Scholar] [CrossRef]
  41. Tsang, M.B.; Stone, J.R.; Camera, F.; Danielewicz, P.; Gandolfi, S.; Hebeler, K.; Horowitz, C.J.; Lee, J.; Lynch, W.G.; Kohley, Z.; et al. Constraints on the symmetry energy and neutron skins from experiments and theory. Phys. Rev. C 2012, 86, 015803. [Google Scholar] [CrossRef]
  42. Vi<i>n</i>˜as, X.; Centelles, M.; Roca-Maza, X.; Warda, M. Density dependence of the symmetry energy from neutron skin thickness in finite nuclei. Eur. Phys. J. A 2014, 50, 27. [Google Scholar] [CrossRef]
  43. Zhang, Y.; Liu, M.; Xia, C.J.; Li, Z.; Biswal, S.K. Constraints on the symmetry energy and its associated parameters from nuclei to neutron stars. Phys. Rev. C 2020, 101, 034303. [Google Scholar] [CrossRef]
  44. Estee, J.; Lynch, W.G.; Tsang, C.Y.; Barney, J.; Jhang, G.; Tsang, M.B.; Wang, R.; Kaneko, M.; Lee, J.W.; Isobe, T.; et al. Probing the Symmetry Energy with the Spectral Pion Ratio. Phys. Rev. Lett. 2021, 126, 162701. [Google Scholar] [CrossRef]
  45. Li, B.A.; Cai, B.J.; Xie, W.J.; Zhang, N.B. Progress in Constraining Nuclear Symmetry Energy Using Neutron Star Observables Since GW170817. Universe 2021, 7, 182. [Google Scholar] [CrossRef]
  46. Essick, R.; Tews, I.; Landry, P.; Schwenk, A. Astrophysical Constraints on the Symmetry Energy and the Neutron Skin of 208Pb with Minimal Modeling Assumptions. Phys. Rev. Lett. 2021, 127, 192701. [Google Scholar] [CrossRef]
  47. Danielewicz, P.; Singh, P.; Lee, J. Symmetry energy III: Isovector skins. Nucl. Phys. A 2017, 958, 147–186. [Google Scholar] [CrossRef]
  48. Lattimer, J.M.; Steiner, A.W. Constraints on the symmetry energy using the mass-radius relation of neutron stars. Eur. Phys. J. A 2014, 50, 40. [Google Scholar] [CrossRef]
  49. Roca-Maza, X.; Brenna, M.; Colò, G.; Centelles, M.; Viñas, X.; Agrawal, B.K.; Paar, N.; Vretenar, D.; Piekarewicz, J. Electric dipole polarizability in 208Pb: Insights from the droplet model. Phys. Rev. C 2013, 88, 024316. [Google Scholar] [CrossRef]
  50. Lattimer, J.M.; Lim, Y. Constraining the symmetry parameters of the nuclear interaction. Astrophys. J. 2013, 771, 51. [Google Scholar] [CrossRef]
  51. Tsang, M.B.; Zhang, Y.; Danielewicz, P.; Famiano, M.; Li, Z.; Lynch, W.G.; Steiner, A.W. Constraints on the Density Dependence of the Symmetry Energy. Phys. Rev. Lett. 2009, 102, 122701. [Google Scholar] [CrossRef] [PubMed]
  52. Trippa, L.; Colò, G.; Vigezzi, E. Giant dipole resonance as a quantitative constraint on the symmetry energy. Phys. Rev. C 2008, 77, 061304. [Google Scholar] [CrossRef]
  53. Tamii, A.; Poltoratska, I.; von Neumann-Cosel, P.; Fujita, Y.; Adachi, T.; Bertulani, C.A.; Carter, J.; Dozono, M.; Fujita, H.; Fujita, K.; et al. Complete Electric Dipole Response and the Neutron Skin in 208Pb. Phys. Rev. Lett. 2011, 107, 062502. [Google Scholar] [CrossRef] [PubMed]
  54. Kortelainen, M.; Lesinski, T.; Moré, J.; Nazarewicz, W.; Sarich, J.; Schunck, N.; Stoitsov, M.V.; Wild, S. Nuclear energy density optimization. Phys. Rev. C 2010, 82, 024313. [Google Scholar] [CrossRef]
  55. Wang, N.; Ou, L.; Liu, M. Nuclear symmetry energy from the Fermi-energy difference in nuclei. Phys. Rev. C 2013, 87, 034327. [Google Scholar] [CrossRef]
  56. Brown, B.A. Constraints on the Skyrme Equations of State from Properties of Doubly Magic Nuclei. Phys. Rev. Lett. 2013, 111, 232502. [Google Scholar] [CrossRef] [PubMed]
  57. Zhang, Z.; Chen, L.W. Bayesian Inference of the Symmetry Energy and the Neutron Skin in 48Ca and 208Pb from CREX and PREX-2. arXiv 2022, arXiv:2207.03328. [Google Scholar]
  58. Reed, B.T.; Fattoyev, F.J.; Horowitz, C.J.; Piekarewicz, J. Implications of PREX-2 on the Equation of State of Neutron-Rich Matter. Phys. Rev. Lett. 2021, 126, 172503. [Google Scholar] [CrossRef]
  59. Typel, S. Neutron skin thickness of heavy nuclei with α-particle correlations and the slope of the nuclear symmetry energy. Phys. Rev. C 2014, 89, 064321. [Google Scholar] [CrossRef]
  60. Tanaka, J.; Yang, Z.; Typel, S.; Adachi, S.; Bai, S.; van Beek, P.; Beaumel, D.; Fujikawa, Y.; Han, J.; Huang, S.H.S.; et al. Formation of α clusters in dilute neutron-rich matter. Science 2021, 371, 5. [Google Scholar] [CrossRef]
  61. Horiuchi, W.; Suzuki, Y.; Inakura, T. Probing neutron-skin thickness with total reaction cross sections. Phys. Rev. C 2014, 89, 011601. [Google Scholar] [CrossRef]
  62. Krasznahorkay, A.; Fujiwara, M.; van Aarle, P.; Akimune, H.; Daito, I.; Fujimura, H.; Fujita, Y.; Harakeh, M.N.; Inomata, T.; Jänecke, J.; et al. Excitation of Isovector Spin-Dipole Resonances and Neutron Skin of Nuclei. Phys. Rev. Lett. 1999, 82, 3216–3219. [Google Scholar] [CrossRef]
  63. Klimkiewicz, A.; Paar, N.; Adrich, P.; Fallot, M.; Boretzky, K.; Aumann, T.; Cortina-Gil, D.; Pramanik, U.D.; Elze, T.W.; Emling, H.; et al. Nuclear symmetry energy and neutron skins derived from pygmy dipole resonances. Phys. Rev. C 2007, 76, 051603. [Google Scholar] [CrossRef]
  64. Aumann, T.; Bertulani, C.A.; Schindler, F.; Typel, S. Peeling Off Neutron Skins from Neutron-Rich Nuclei: Constraints on the Symmetry Energy from Neutron-Removal Cross Sections. Phys. Rev. Lett. 2017, 119, 262501. [Google Scholar] [CrossRef] [PubMed]
  65. Hartnack, C.; Fevre, A.L.; Leifels, Y.; Aichelin, J. The influence of the neutron skin and the asymmetry energy on the π/π+ ratio. arXiv 2018, arXiv:1808.09868. [Google Scholar]
  66. Wei, G.F.; Li, B.A.; Xu, J.; Chen, L.W. Influence of neutron-skin thickness on the π/π+ ratio in Pb + Pb collisions. Phys. Rev. C 2014, 90, 014610. [Google Scholar] [CrossRef]
  67. Ma, C.W.; Wei, H.L.; Liu, X.Q.; Su, J.; Zheng, H.; Lin, W.P.; Zhang, Y.X. Nuclear fragments in projectile fragmentation reactions. Prog. Part. Nucl. Phys. 2021, 121, 103911. [Google Scholar] [CrossRef]
  68. Yang, J.; Zhang, Y.; Wang, N.; Li, Z. Influence of the treatment of initialization and mean-field potential on the neutron to proton yield ratios. Phys. Rev. C 2021, 104, 024605. [Google Scholar] [CrossRef]
  69. Zhang, Y.; Wang, N.; Li, Q.F.; Ou, L.; Tian, J.L.; Liu, M.; Zhao, K.; Wu, X.Z.; Li, Z.X. Progress of quantum molecular dynamics model and its applications in heavy ion collisions. Front. Phys. 2020, 15, 54301. [Google Scholar] [CrossRef]
  70. Skyrme, T.H.R. The nuclear surface. Philos. Mag. 1956, 1, 1043–1054. [Google Scholar] [CrossRef]
  71. Zhang, Y.; Li, Z. Elliptic flow and system size dependence of transition energies at intermediate energies. Phys. Rev. C 2006, 74, 014602. [Google Scholar] [CrossRef]
  72. Liu, M.; Wang, N.; Li, Z.-X.; Wu, X.-Z. Neutron Skin Thickness of Nuclei and Effective Nucleon–Nucleon Interactions. Chin. Phys. Lett. 2006, 23, 804. [Google Scholar] [CrossRef]
  73. Brack, M.; Guet, C.; Håkansson, H.B. Selfconsistent semiclassical description of average nuclear properties a link between microscopic and macroscopic models. Phys. Rep. 1985, 123, 275–364. [Google Scholar] [CrossRef]
  74. Zhang, Y.; Coupland, D.D.S.; Danielewicz, P.; Li, Z.; Liu, H.; Lu, F.; Lynch, W.G.; Tsang, M.B. Influence of in-medium NN cross sections, symmetry potential, and impact parameter on isospin observables. Phys. Rev. C 2012, 85, 024602. [Google Scholar] [CrossRef]
  75. Li, L.; Wang, F.; Zhang, Y. Isospin effects on intermediate mass fragments at intermediate energy-heavy ion collisions. Nucl. Sci. Tech. 2022, 33, 58. [Google Scholar] [CrossRef]
  76. Li, B.A.; Ko, C.M.; Ren, Z. Equation of State of Asymmetric Nuclear Matter and Collisions of Neutron-Rich Nuclei. Phys. Rev. Lett. 1997, 78, 1644–1647. [Google Scholar] [CrossRef]
  77. Zhang, Y.; Danielewicz, P.; Famiano, M.; Li, Z.; Lynch, W.; Tsang, M. The influence of cluster emission and the symmetry energy on neutron–proton spectral double ratios. Phys. Lett. B 2008, 664, 145–148. [Google Scholar] [CrossRef]
  78. Su, J.; Zhu, L.; Huang, C.Y.; Xie, W.J.; Zhang, F.S. Correlation between symmetry energy and effective k-mass splitting with an improved isospin- and momentum-dependent interaction. Phys. Rev. C 2016, 94, 034619. [Google Scholar] [CrossRef]
  79. Feng, Z.Q. Transverse emission of isospin ratios as a probe of high-density symmetry energy in isotopic nuclear reactions. Phys. Lett. B 2012, 707, 83–87. [Google Scholar] [CrossRef]
  80. Ono, A.; Danielewicz, P.; Friedman, W.A.; Lynch, W.G.; Tsang, M.B. Isospin fractionation and isoscaling in dynamical simulations of nuclear collisions. Phys. Rev. C 2003, 68, 051601. [Google Scholar] [CrossRef]
  81. Li, B.A.; Chen, L.W.; Yong, G.C.; Zuo, W. Double neutron/proton ratio of nucleon emissions in isotopic reaction systems as a robust probe of nuclear symmetry energy. Phys. Lett. B 2006, 634, 378–382. [Google Scholar] [CrossRef]
  82. Li, Q.; Li, Z.; Stöcker, H. Probing the symmetry energy and the degree of isospin equilibrium. Phys. Rev. C 2006, 73, 051601. [Google Scholar] [CrossRef]
  83. Zhang, Y.; Tsang, M.; Li, Z.; Liu, H. Constraints on nucleon effective mass splitting with heavy ion collisions. Phys. Lett. B 2014, 732, 186–190. [Google Scholar] [CrossRef]
  84. Coupland, D.D.S.; Youngs, M.; Chajecki, Z.; Lynch, W.G.; Tsang, M.B.; Zhang, Y.X.; Famiano, M.A.; Ghosh, T.K.; Giacherio, B.; Kilburn, M.A.; et al. Probing effective nucleon masses with heavy-ion collisions. Phys. Rev. C 2016, 94, 011601. [Google Scholar] [CrossRef]
  85. Lin, W.; Liu, X.; Rodrigues, M.R.D.; Kowalski, S.; Wada, R.; Huang, M.; Zhang, S.; Chen, Z.; Wang, J.; Xiao, G.Q.; et al. Experimental reconstruction of primary hot isotopes and characteristic properties of the fragmenting source in heavy-ion reactions near the Fermi energy. Phys. Rev. C 2014, 90, 044603. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of peripheral heavy-ion collisions. (a) Initial stage; (b) Reaction stage; (c) Light particle emitting stage.
Figure 1. Schematic diagram of peripheral heavy-ion collisions. (a) Initial stage; (b) Reaction stage; (c) Light particle emitting stage.
Universe 09 00206 g001
Figure 2. Panel (a) is the impact parameter dependence of the probability of the primary projectile-like residues with A > 100 . Panel (b) is the θ c . m . distribution of the projectile-like residues with mass number A.
Figure 2. Panel (a) is the impact parameter dependence of the probability of the primary projectile-like residues with A > 100 . Panel (b) is the θ c . m . distribution of the projectile-like residues with mass number A.
Universe 09 00206 g002
Figure 3. Panels (a) and (b) are the cross-section of the projectile-like residues with A > 100 as the function of the neutron skin of the system Δ R for 124 Sn+ 124 Sn and 132 Sn+ 124 Sn, respectively. Panel (c) is δ σ A > 100 as a function of δ R .
Figure 3. Panels (a) and (b) are the cross-section of the projectile-like residues with A > 100 as the function of the neutron skin of the system Δ R for 124 Sn+ 124 Sn and 132 Sn+ 124 Sn, respectively. Panel (c) is δ σ A > 100 as a function of δ R .
Universe 09 00206 g003
Figure 4. Panel (a,b) show R ( n / p ) as a function of the kinetic energy of the emitted nucleons; (c,d) are R 80 ( n / p ) as a function of Δ R . R 80 ( n / p ) is the ratio of all nucleons above 80 MeV, as shown in Equation (12). The left panels are for 124 Sn + 124 Sn, and the right are for 132 Sn + 124 Sn, respectively. To illustrate the results of two systems in a similar scale, the value of R 80 ( n / p ) for 132 Sn + 124 Sn is shifted down by 0.1 in Panel (d).
Figure 4. Panel (a,b) show R ( n / p ) as a function of the kinetic energy of the emitted nucleons; (c,d) are R 80 ( n / p ) as a function of Δ R . R 80 ( n / p ) is the ratio of all nucleons above 80 MeV, as shown in Equation (12). The left panels are for 124 Sn + 124 Sn, and the right are for 132 Sn + 124 Sn, respectively. To illustrate the results of two systems in a similar scale, the value of R 80 ( n / p ) for 132 Sn + 124 Sn is shifted down by 0.1 in Panel (d).
Universe 09 00206 g004
Table 1. The values of the nuclear matter parameters used in the ImQMD-L. m is the free nucleon mass. m v * , m s * , and m are in MeV, ρ 0 is in fm 3 , E 0 , K 0 , S 0 , and L are in MeV, and g s u r and g s u r , i s o are in MeVfm 2 .
Table 1. The values of the nuclear matter parameters used in the ImQMD-L. m is the free nucleon mass. m v * , m s * , and m are in MeV, ρ 0 is in fm 3 , E 0 , K 0 , S 0 , and L are in MeV, and g s u r and g s u r , i s o are in MeVfm 2 .
K 0 S 0 E 0 ρ 0 m v * / m m s * / m g sur g sur , iso L
24030−160.160.70.824.5−4.9930, 50, 70, 90, 110
Table 2. a p , R p , a n , R n , binding energy B, and rms radius of neutron and proton density distributions for 124 Sn and 132 Sn obtained with RDV method. L and B are in MeV; a p , R p , a n , R n , r p 2 1 / 2 , r n 2 1 / 2 , and r n p are in fm.
Table 2. a p , R p , a n , R n , binding energy B, and rms radius of neutron and proton density distributions for 124 Sn and 132 Sn obtained with RDV method. L and B are in MeV; a p , R p , a n , R n , r p 2 1 / 2 , r n 2 1 / 2 , and r n p are in fm.
124 Sn
L B a p R p a n R n r p 2 1 / 2 r n 2 1 / 2 r np
30−7.9710.4145.7330.5145.7774.6704.8650.165
50−8.0210.4155.7290.5075.8114.6984.8810.183
70−8.0730.4195.7070.5035.8384.6874.8930.206
90−8.1290.4225.6860.4965.8724.6764.9070.232
110−8.1910.4265.6560.4875.9094.6594.9220.263
132 Sn
L B a p R p a n R n r p 2 1 / 2 r n 2 1 / 2 r np
30−7.8100.4085.8270.5395.9064.7624.9940.232
50−7.8890.4105.8160.5325.9484.7565.0130.257
70−7.9750.4135.7980.5245.9904.7465.0320.286
90−8.0670.4165.7710.5146.0354.7305.0500.320
110−8.1660.4195.7330.5026.0844.7075.0680.361
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yang, J.; Chen, X.; Cui, Y.; Li, Z.; Zhang, Y. Probing the Neutron Skin of Unstable Nuclei with Heavy-Ion Collisions. Universe 2023, 9, 206. https://doi.org/10.3390/universe9050206

AMA Style

Yang J, Chen X, Cui Y, Li Z, Zhang Y. Probing the Neutron Skin of Unstable Nuclei with Heavy-Ion Collisions. Universe. 2023; 9(5):206. https://doi.org/10.3390/universe9050206

Chicago/Turabian Style

Yang, Junping, Xiang Chen, Ying Cui, Zhuxia Li, and Yingxun Zhang. 2023. "Probing the Neutron Skin of Unstable Nuclei with Heavy-Ion Collisions" Universe 9, no. 5: 206. https://doi.org/10.3390/universe9050206

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop