logo

Impact of initial fluctuations and nuclear deformations in isobar collisions

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

Impact of initial fluctuations and nuclear deformations in isobar collisions

Jian-Fei Wang
Hao-Jie Xu
Fu-Qiang Wang
Nuclear Science and TechniquesVol.35, No.6Article number 108Published in print Jun 2024Available online 04 Jul 2024
61202

Relativistic isobar (4496Ru+4496Ru and 4096Zr+4096Zr) collisions have revealed intricate differences in their nuclear size and shape, inspiring unconventional studies of nuclear structure using relativistic heavy ion collisions. In this study, we investigate the relative differences in the mean multiplicity (RNch) and the second- (Rϵ2) and third-order eccentricity (Rϵ3) between isobar collisions using initial state Glauber models. It is found that initial fluctuations and nuclear deformations have negligible effects on RNch in most central collisions, while both are important for the Rϵ2 and Rϵ3, the degree of which is sensitive to the underlying nucleonic or sub-nucleonic degree of freedom. These features, compared to real data, may probe the particle production mechanism and the physics underlying nuclear structure.

Relativistic isobar collisionsInitial fluctuationsNuclear deformation
1

Introduction

The recent isobar data [1, 2] stimulated a wide interest from the relativistic heavy ion community in the physics of nuclear structures of isobars (nuclei with equal number of nucleons but different numbers of protons and neutrons) [3-12]. The isobar collisions of 4496Ru+4496Ru and 4096Zr+4096Zr were initially proposed to search for the chiral magnetic effect (CME) [13-16]. The measurements in isobar collisions at sNN=200 GeV by the STAR Collaborations showed sizable differences in the multiplicity distribution, elliptic flow, and triangular flow [1], indicating differences in the density profiles of the colliding nuclei [3, 6]. Nuclear density is usually obtained from theory calculations, such as energy density functional theory (DFT) [17-20], incorporating information from experimental measurements. In relativistic heavy ion collision simulations, it is common to use the parameterized nuclear density in the Woods-Saxon formula, ρ(r)=ρ0exp[(rR(1+β2Y20+β3Y30+...)/a]+1, (1) with the radius parameter R, the diffuseness parameter a, and the quadrupole (octupole) deformation parameter β2 (β3). A large a value in 96Zr is suggested by the STAR data on multiplicity distribution and elliptic flow ratios in non-central collisions [1], consistent with a halo-type neutron skin thickness as predicted by the DFT calculations [21]. The flow ratios observed in the most central collisions suggest that the 96Ru has a larger β2, while 96Zr has a larger β3 [6].

Much progress has been made in quantitatively assessing the differences in size and shape of isobars by using relativistic isobar collisions [5, 7-9, 12, 22-26]. These differences can place stringent constraints on key parameters in theoretical modeling of nuclear interactions, which is the backbone of all theoretical calculations of nuclear structure including DFT. One of such parameters is the slope parameter L(ρ) of the nuclear symmetry energy as a function of nuclear density, which cannot yet be well determined by low-energy nuclear experiments [27-35]. In addition, nuclear deformations are largely uncertain from nuclear structure calculations [36, 37, 15]. On the other hand, the STAR experiment has collected approximately 2 billion minimum-bias events for each species of 4496Ru+4496Ru and 4096Zr+4096Zr collisions at sNN=200 GeV, with good control of systematic uncertainties [1]. These provide an unique opportunity to study the density slope parameter of the symmetry energy and the deformation parameters of the isobars. Previous studies suggest that ratio observables in isobar collisions are less sensitive to final-state interactions [5, 21, 9]. In this work, we investigate the effect of initial-state fluctuations on those ratio observables in relativistic isobar collisions.

We use the Optical Glauber model and Monte Carlo Glauber model [38-40] to investigate the effect of nuclear deformations and initial fluctuations on the relative differences RNch, Rϵ2, and Rϵ3 in relativistic isobar collisions, where RX2XRuRuXZrZrXRuRu+XZrZr. (2) The Glauber model is widely used to determine the centrality in experiments, including the blind analysis of the STAR isobar data [1]. Due to symmetry, the elliptic flow v2 with impact parameter b=0 and the triangular flow v3 can not be generated in spherical nucleus-nucleus collisions with smooth initial conditions in Optical Glauber simulation. The collision configuration of deformed nuclei will contribute to spatial anisotropy, while event-by-event fluctuation contributions are also important. We further investigate the effect of initial fluctuations by comparing the results under the scenarios of nucleon participants and quark participants.

The rest of the paper is organized as follows. Section 2 gives a brief description of the models and the nuclear density profiles used in this work. Section 3 discusses the effects of nuclear deformation and initial fluctuations on RNch, Rϵ2, and Rϵ3. A summary is given in Sect. 4.

2

Glauber model and isobar nuclear densities

Particle production in relativistic heavy ion collisions is usually parameterized by the two-component model, Nch=npp[(1x)Npart/2+xNcoll], (3) with the number of participants Npart and number of binary collisions Ncoll obtained from the Glauber model [39]. Here the x is the relative contribution of hard processes and npp=Nch in nucleon-nucleon collisions. For the case of x=0, the particle production is independent of hard processes and it can alternatively be described by a variant of the Glauber model, the so-called Trento model [41, 42], Nchs(r)d2r(TAp(r)+TBp(r)2)1/pd2r (4) with p=1 (we note that previous studies favor p~0 [42]). Here s is the entropy density, r is the transverse radius, and TA(B) is the reduced thickness function for the colliding nuclei A(B) with (TA+TB)d2r=Npart. In this work, we focus on the centrality dependence of the ratio observables in isobar collisions, so the parameter npp and the normalization factor hidden in the above equation are not relevant to our study.

The elliptic flow (v2), triangular flow (v3) are the Fourier coefficients of the azimuthal angle (ϕ) distribution vn=cos[n(ϕΨn)], where ... denotes the average over an event and Ψn is the plane angle in momentum phase. The large anisotropic flow can be successfully described by relativistic hydrodynamics with a large spatial eccentricity [43], ϵneinΦn=rneinφw(r)d2rrnw(r)d2r. (5) Here w(r) is the weight factor, often taken to be the entropy density s or the energy density e. Φn is the nth order plane angle in the configuration space. The anisotropic flow is approximately proportional to the initial eccentricity for small amplitude [44]. In this work, we use the participant density to calculate ϵ2 and ϵ3, however, our main conclusions do not change if alternative w is used.

For spherical nucleus-nucleus collisions, the multiplicity and eccentricities can be calculated analytically in the Optical Glauber model for a given impact parameter b. For deformed nucleus-nucleus collisions, we calculate the ratio observables at a given shape orientation with a specific Euler rotation and then take the average over all orientations X(b)=1(4π)2X(ΩA,ΩB;b)dΩAdΩB. (6) Here X is the observable of interest and ΩA(B) is the orientation solid angle for colliding nuclei A(B).

The event-by-event fluctuations are not included in the Optical Glauber model [39], and such an effect is crucial for the anisotropic flow observables, especially the odd-order ones. We, therefore, use the Monte Carlo Glauber model to include the fluctuation effect, where the finite number of nucleons (96 for the Ru and Zr isobars) are sampled in each colliding nucleus. The event-by-event fluctuations depend on the number of nucleons in the colliding nuclei, so these effects are expected to be larger in the isobar collisions than in Au+Au collisions. After sampling the nucleons with the appropriate nuclear density, the multiplicity is calculated based on Eq. (3), and the eccentricity is calculated from the transverse position of those finite participating nucleons. Here the participants and binary collisions are determined from nucleon-nucleon interactions which happens when the distance between the two nucleons, each from the incoming and outgoing nuclei, is less than σNN/π [39], with the inelastic cross section set to σNN=42 mb for sNN=200 GeV. Poisson-type multiplicity fluctuations are applied in the Monte Carlo simulations.

The density profiles of 96Ru and 96Zr nuclei are required in the Glauber simulations. In this study, the parameters for the Woods-Saxon density distributions of 96Ru and 96Zr are taken from Ref. [22, 45] (see Tab. 1), which are extracted from the isobar densities obtained by DFT calculations with the symmetry energy Esym(ρc)=26.65 MeV and its density slope parameter L(ρc)=47.3 MeV at subsaturation cross density ρc0.11 fm-3 [20]. Nuclear deformation is not included in the DFT calculations due to its strong model dependence. Following our previous work [22, 45], to introduce the nuclear deformation, the WS parameters R and a for the given β2 (or β3) are calculated to match the volume and root-mean-square (RMS) of the corresponding nucleus calculated by DFT, keeping the normalization factor ρ0 fixed.

Table 1
Woods-Saxon parameterizations of the 96Ru and 96Zr nuclear density distributions. The quoted values for R0 and a are in fm and that for ρ0 is in 1/fm3
96Ru 96Zr
ρ0 R a β2 ρ0 R a β3
0.159 5.093 0.487 0.00 0.163 5.022 0.538 0.00
0.159 5.093 0.471 0.16 0.163 5.021 0.517 0.20
Show more

Centrality is determined by the multiplicity distributions in the Monte Carlo Glauber model, similar to what was done in heavy ion experiments [1]. For the optical Glauber model and optical Trento model in spherical nucleus-nucleus collisions, the centrality is related to the differential cross-section in impact parameter [39]. In this study, we use the 16 equal-size centrality bins from 0-80% centrality range. The results obtained with the fixed impact parameter corresponding to the centre of each centrality bin are used to mimic the observables in such a centrality bin. The impact parameters for 96Ru and 96Zr are listed in Tab. 2. The impact parameters are larger in 4096Zr+4096Zr collisions than in 4496Ru+4496Ru collisions for a given centrality bin because the total cross section for 4096Zr+4096Zr collisions (σXZrZr=4.551 barn) is larger than that for 4496Ru+4496Ru collisions (σXRuRu=4.369 barn) based on the nuclear density parameters listed in Tab. 1. For simplicity, the impact parameter list is also used to determine the centrality for deformed nucleus-nucleus collisions in Optical Glauber simulations. We note that determining centrality in the deformed case is in principle more complicated, but this does not affect our conclusion (see discussion in the next section).

Table 2
The fixed impact parameter used to mimic the 16 equal-size centrality bins from 0-80% centrality range
Centrality (%) 2.5 7.5 12.5 17.5 22.5 27.5 32.5 37.5 42.5 47.5 52.5 57.5 62.5 67.5 72.5 77.5
bRu (fm) 1.865 3.230 4.170 4.933 5.594 6.184 6.723 7.222 7.688 8.128 8.545 8.943 9.323 9.689 10.042 10.384
bZr (fm) 1.903 3.296 4.255 5.035 5.709 6.312 6.862 7.370 7.846 8.295 8.721 9.127 9.515 9.888 10.249 10.600
Show more

In order to achieve high precision of calculations in our study, the GPU parallel computing technology [46] is used in our Glauber model simulations.

3

Results and discussions

The RNch obtained from the Optical Glauber model with x=0 and x=0.15, as well as those from the Optical Trento model with p=0 and p=-1, are shown in Fig 1. The case of the Optical Glauber simulation with x=0 is nothing but the Optical Trento simulation with p=1, and the multiplicity, in this case, is only proportional to the number of participants. The RNch distribution depends on the model parameters x and p. The magnitudes increase with x and p over the whole centrality range. For the most central collisions, the RNch increases with x in Optical Glauber simulations, but is independent of p in Optical Trento simulations. This is shown in the top 5% centrality range in Fig. 1(b). The magnitudes of RNch converge to the same value from Optical Trento simulations with different values of p, but the trends in centrality differ.

Fig. 1
(Color online). The centrality dependence of RNch from Optical Glauber (Opt.Glb) simulations with x=0.15 and x=0, and from Optical Trento (Opt.Trento) simulations with p=0 and p=-1 for (a) 0-80% centrality range and (b) top 5% centrality
pic

The parameters x and p are usually constrained by the individual multiplicity distributions (in this case, of isobar collisions or their mean). The RNch in the most central collisions has been proposed to probe the neutron skin and symmetry energy. However, the particle production mechanisms will introduce uncertainties as shown in previous study [5]. The results shown in Fig. 1 imply that the model uncertainty of RNch is negligible if the particle production follows the Trento model. They also suggest that the precision measurements of the mean multiplicity distribution of 4496Ru+4496Ru and 4096Zr+4096Zr collisions and their difference may offer insights on the particle production mechanisms in relativistic heavy ion collisions.

To study effects from event-by-event fluctuations, we present the RNch distributions from Monte Carlo Glauber simulations with x=0.15 and x=0 in Fig. 2. The apparent large fluctuation effect (i.e. difference in RNch between the Optical and the Monte Carlo Glauber) is because of the different total cross sections predicted by these two models [39]. For instance, the impact parameter cut would be smaller for a given centrality when a smaller total cross-section is used. Based on the nuclear density profiles in our study, the total cross sections for 4496Ru+4496Ru and 4096Zr+4096Zr collisions are σXMC,RuRu=4.206 barn and σXMC,ZrZr=4.348 barn, which are 4.46% and 3.73% smaller than those in the Optical cases, respectively. To remove this trivial difference, we calculate the optical model using the total cross sections from the Monte Carlo Glauber. The results are shown in Fig. 2; they become similar now. Our results indicate that both the data of centrality dependent RNch and the model prediction of total cross section are important for the model to data comparison in relativistic isobar collisions [1].

Fig. 2
(Color online). The centrality dependence of RNch from Optical Glauber (Opt.Glb, solid curves) and Monte Carlo Glauber (MC.Glb, symbols) simulations with x=0.15 and x=0. The dashed curves are the results from Optical Glauber calculations with the total cross section from Mont Carlo Glauber simulations
pic

The 4496Ru and 4096Zr nuclei are deformed [47, 48], as also suggested by recent studies [1, 6]. The quadrupole deformation parameter for 96Ru is β2,Ru=0.16 and is negligible for 96Zr. The octupole deformation parameter is negligible for 96Ru and is β3,Zr=0.20 for 96Zr. Figure 3 shows the RNch from deformed isobar collisions, where the volume and RMS radius of the colliding nuclei are constrained to match those of spherical nuclei [22, 45]. For both the Optical and Monte Carlo calculations, the results from deformed isobar collisions are similar to those from spherical ones. Here we have used the same impact parameters listed in Tab 2 for deformed isobar collisions. As mentioned above, the centrality definition in the deformed case may be biased by multiplicity variations at a given impact parameter. However, the centrality is well defined in the Monte Carlo Glauber simulation, and the consistency in the corresponding results between spherical and deformed cases indicates that our centrality treatment in the Optical Glauber simulation is appropriate. The RNch is considerably larger in the Optical case because of the aforementioned larger total cross section. The differences are ~10% between with and without nuclear deformation. Based on previous study [5], this would result in an uncertainty on the extracted slope parameter of symmetry energy of about 5 MeV. Our results indicate that, with appropriate magnitudes of deformation parameters, the effect of nuclear deformations on the uncertainties of RNch distributions are small.

Fig. 3
(Color online). The centrality dependence of RNch with spherical (Sph) and deformed (Def) colliding nuclei, calculated with (a) Optical Glauber (Opt.Glb) (b) Monte Carlo (MC.Glb) Glauber models
pic

This is not the case for the anisotropic flow observables, where previous studies suggest that both nuclear deformation and event-by-event fluctuations are important [39, 49, 50]. Apart from nuclear deformation, most of the anisotropic flow produced in the most central heavy-ion collisions is due to the position fluctuations of nucleons in the colliding nuclei. In this paper, we study the effect of nuclear deformations and initial fluctuations on geometry eccentricities Rϵn (n=2,3). The results are shown in Fig. 4. Note that, for the Optical Glauber simulations, ϵ3 vanishes for the spherical case due to symmetry, and for the deformed case the Rϵ3<1 (because ϵ3ZrZrϵ3RuRu) is outside the frame of Fig. 4(b). Comparing the results between deformed and spherical nuclei, the general features are similar for both. This is consistent with the results shown in previous studies [6, 8], where the large quadrupole deformation β2 in Ru increases the Rϵ2 in most central collisions while the large octupole deformation β3 suppresses the values in mid-central collisions.

Fig. 4
(Color online). The centrality dependence of (a) Rϵ2 and (b) Rϵ3. The results are calculated with spherical (Sph) and deformed (Def) colliding nuclei using Optical Glauber (Opt.Glb) and Monte Carlo Glauber (MC.Glb) models. The results from Monte Carlo Glauber model with quark participant assumption (MC.qGlb) are presented as open symbols. From Optical Glauber simulations, ϵ3 vanishes due to symmetry in spherical cases, and the Rϵ3<1 in deformed cases are outside the frame of the plot
pic

The magnitudes of Rϵ2 show significant differences between the Optical and Monte Carlo Glauber models. The Rϵ2 caused by the nuclear geometry is significantly diluted by the event-by-event fluctuations. The reason is rather straightforward: event-by-event fluctuations contribute to a large fraction of the eccentricities in Monte Carlo simulations, and these contributions are common between 4496Ru+4496Ru and 4096Zr+4096Zr collisions, suppressing the differences caused by nuclear deformation and/or collision geometry.

In the Monte Carlo Glauber simulation, the participating nucleons are usually treated as point-like particles as we did in the simulations above. However, nucleons have substructures [51-53]. More substructure constituents are expected to suppress the fluctuations. We use the quark participant assumption as it was done in Ref. [51], increasing the density by three times and decreasing the inelastic cross-section by a factor of 9. The results are also shown in Fig. 4. The Monte Carlo Glauber simulations under quark participant scenario significantly increase the Rϵ2 and decrease Rϵ3, bringing the values closer to the Optical limit. We note that despite of the cross-section differences, the Rϵn calculated in the Optical Glauber model are similar to those calculated in the Monte Carlo Glauber model with respect to reaction plane instead of participant plane [3], which is more related to the elliptic flow measurement with respect to zero-degree calorimeter (ZDC) instead of time projection chamber (TPC) in experiment [1].

There are more sources of initial fluctuations in heavy ion collisions, for example, the correlations between the quarks in a nucleon, as well as the correlations between the nucleons in the colliding nuclei. We find that the minimum distance between two nucleons in simulations, as one of the sources of nucleon-nucleon correlations, does not change our results. We note that the final-state effects, which are not the focus of the present study, are also important for the accurate extraction of the symmetry energy slope parameter. We postpone such investigation to a future work.

4

Summary

In this work, the effects of nuclear deformations and initial fluctuations on the ratio observables in relativistic isobar collisions are studied using the Optical Glauber and Monte Carlo Glauber models. The GPU parallel computing technology is used in our calculations to achieve high precision, especially for the deformed colliding nucleus. Our main findings are as follow.

RNch depends on particle production mechanisms and model parameters, while it converges to a single value in most central collisions in the Trento model (see Fig. 1).

• Initial fluctuations affect the total cross section and thus the RNch, however the effect is negligible for most central collisions (see Fig. 2).

• Because of the constraints of the total volumes and the measured RMS radii of the isobar nuclei, their deformations are found to have negligible effects on RNch (see Fig. 3).

• The Rϵ2 and Rϵ3, driven by the nuclear diffusenesses and deformations, are suppressed by the initial finite number fluctuations, dependent on the degree of freedom of nucleons or constituent quarks (see Fig. 4).

Our results indicate unique (in)sensitivities of the ratio observables of RNch and Rϵn to nuclear deformations and initial fluctuations in relativistic isobar collisions. These features, compared to data, may potentially probe the particle production mechanism and the physics underlying nuclear structure.

References
1. M. Abdallah, et al.,

Search for the chiral magnetic effect with isobar collisions at sNN=200 GeV by the STAR Collaboration at the BNL Relativistic Heavy Ion Collider

. Phys. Rev. C 105, 014901 (2022). arXiv:2109.00131, https://doi.org/10.1103/PhysRevC.105.014901
Baidu ScholarGoogle Scholar
2. J. Adam, et al.,

Methods for a blind analysis of isobar data collected by the STAR collaboration

. Nucl. Sci. Tech. 32, 48 (2021). arXiv:1911.00596, https://doi.org/10.1007/s41365-021-00878-y
Baidu ScholarGoogle Scholar
3. H.J. Xu, X. Wang, H. Li, et al.,

Importance of isobar density distributions on the chiral magnetic effect search

. Phys. Rev. Lett. 121, 022301 (2018). arXiv:1710.03086, https://doi.org/10.1103/PhysRevLett.121.022301
Baidu ScholarGoogle Scholar
4. H. Li, H.j. Xu, J. Zhao, et al.,

Multiphase transport model predictions of isobaric collisions with nuclear structure from density functional theory

. Phys. Rev. C98, 054907 (2018). arXiv:1808.06711, https://doi.org/10.1103/PhysRevC.98.054907
Baidu ScholarGoogle Scholar
5. H. Li, H.j. Xu, Y. Zhou, et al.,

Probing the neutron skin with ultrarelativistic isobaric collisions

. Phys. Rev. Lett. 125, 222301 (2020). arXiv:1910.06170, https://doi.org/10.1103/PhysRevLett.125.222301
Baidu ScholarGoogle Scholar
6. C. Zhang, J. Jia,

Evidence of Quadrupole and Octupole Deformations in 96Zr+96Zr and 96Ru+96Ru Collisions at Ultrarelativistic Energies

. Phys. Rev. Lett. 128, 022301 (2022). arXiv:2109.01631, https://doi.org/10.1103/PhysRevLett.128.022301
Baidu ScholarGoogle Scholar
7. G. Nijs, W. van der Schee,

Inferring nuclear structure from heavy isobar collisions using Trajectum

. SciPost Phys. 15, 041 (2023). arXiv:2112.13771, https://doi.org/10.21468/SciPostPhys.15.2.041
Baidu ScholarGoogle Scholar
8. J. Jia, G. Giacalone, C. Zhang,

Separating the Impact of Nuclear Skin and Nuclear Deformation in High-Energy Isobar Collisions

. Phys. Rev. Lett. 131, 022301 (2023). arXiv:2206.10449, https://doi.org/10.1103/PhysRevLett.131.022301
Baidu ScholarGoogle Scholar
9. C. Zhang, S. Bhatta, J. Jia,

Ratios of collective flow observables in high-energy isobar collisions are insensitive to final-state interactions

. Phys. Rev. C 106, L031901 (2022). arXiv:2206.01943, https://doi.org/10.1103/PhysRevC.106.L031901
Baidu ScholarGoogle Scholar
10. M. Nie, C. Zhang, Z. Chen, et al.,

Impact of nuclear structure on longitudinal flow decorrelations in high-energy isobar collisions

. Phys. Lett. B 845, 138177 (2023). arXiv:2208.05416, https://doi.org/10.1016/j.physletb.2023.138177
Baidu ScholarGoogle Scholar
11. B.S. Xi, X.G. Deng, S. Zhang, et al.,

Vorticity in isobar collisions of 4496Ru+4496Ru and 4096Zr+4096Zr at sNN = 200 GeV

. Eur. Phys. J. A 59, 33 (2023). https://doi.org/10.1140/epja/s10050-023-00932-w
Baidu ScholarGoogle Scholar
12. Y.L. Cheng, S. Shi, Y.G. Ma, et al.,

Examination of nucleon distribution with Bayesian imaging for isobar collisions

. Phys. Rev. C 107, 064909 (2023). arXiv:2301.03910, https://doi.org/10.1103/PhysRevC.107.064909
Baidu ScholarGoogle Scholar
13. D.E. Kharzeev, L.D. McLerran, H.J. Warringa,

The Effects of topological charge change in heavy ion collisions: ’Event by event P and CP violation’

. Nucl. Phys. A 803, 227-253 (2008). arXiv:0711.0950, https://doi.org/10.1016/j.nuclphysa.2008.02.298
Baidu ScholarGoogle Scholar
14. S.A. Voloshin,

Testing the Chiral Magnetic Effect with Central U+U collisions

. Phys. Rev. Lett. 105, 172301 (2010). arXiv:1006.1020, https://doi.org/10.1103/PhysRevLett.105.172301
Baidu ScholarGoogle Scholar
15. W.T. Deng, X.G. Huang, G.L. Ma, et al.,

Test the chiral magnetic effect with isobaric collisions

. Phys. Rev. C 94, 041901 (2016). arXiv:1607.04697, https://doi.org/10.1103/PhysRevC.94.041901
Baidu ScholarGoogle Scholar
16. F.Q. Wang, J. Zhao,

Search for the chiral magnetic effect in heavy ion collisions

. Nucl. Sci. Tech. 29, 179 (2018). https://doi.org/10.1007/s41365-018-0520-z
Baidu ScholarGoogle Scholar
17. E. Chabanat, J. Meyer, P. Bonche, et al.,

A Skyrme parametrization from subnuclear to neutron star densities

. Nucl. Phys. A 627, 710-746 (1997). https://doi.org/10.1016/S0375-9474(97)00596-4
Baidu ScholarGoogle Scholar
18. N. Chamel, S. Goriely, J.M. Pearson,

Further explorations of Skyrme-Hartree-Fock-Bogoliubov mass formulas. XI. Stabilizing neutron stars against a ferromagnetic collapse

. Phys. Rev. C 80, 065804 (2009). arXiv:0911.3346, https://doi.org/10.1103/PhysRevC.80.065804
Baidu ScholarGoogle Scholar
19. X.B. Wang, J.L. Friar, A.C. Hayes,

Nuclear Zemach moments and finite-size corrections to allowed β decay

. Phys. Rev. C 94, 034314 (2016). arXiv:1607.02149, https://doi.org/10.1103/PhysRevC.94.034314
Baidu ScholarGoogle Scholar
20. Z. Zhang, L.W. Chen,

Extended Skyrme interactions for nuclear matter, finite nuclei and neutron stars

. Phys. Rev. C 94, 064326 (2016). arXiv:1510.06459, https://doi.org/10.1103/PhysRevC.94.064326
Baidu ScholarGoogle Scholar
21. H.j. Xu, H. Li, X. Wang, et al.,

Determine the neutron skin type by relativistic isobaric collisions

. Phys. Lett. B 819, 136453 (2021). arXiv:2103.05595, https://doi.org/10.1016/j.physletb.2021.136453
Baidu ScholarGoogle Scholar
22. H.j. Xu, H. Li, Y. Zhou, et al.,

Measuring neutron skin by grazing isobaric collisions

. Phys. Rev. C 105, L011901 (2022). arXiv:2105.04052, https://doi.org/10.1103/PhysRevC.105.L011901
Baidu ScholarGoogle Scholar
23. H.j. Xu, W. Zhao, H. Li, et al.,

Probing nuclear structure with mean transverse momentum in relativistic isobar collisions

. Phys. Rev. C 108, L011902 (2023). arXiv:2111.14812, https://doi.org/10.1103/PhysRevC.108.L011902
Baidu ScholarGoogle Scholar
24. H. Xu,

Constraints on Neutron Skin Thickness and Nuclear Deformations Using Relativistic Heavy-ion Collisions from STAR

. Acta Phys. Polon. Supp. 16, 30 (2023). arXiv:2208.06149, https://doi.org/10.5506/APhysPolBSupp.16.1-A30
Baidu ScholarGoogle Scholar
25. L.M. Liu, C.J. Zhang, J. Zhou, et al.,

Probing neutron-skin thickness with free spectator neutrons in ultracentral high-energy isobaric collisions

. Phys. Lett. B 834, 137441 (2022). arXiv:2203.09924, https://doi.org/10.1016/j.physletb.2022.137441
Baidu ScholarGoogle Scholar
26. M. Luzum, M. Hippert, J.Y. Ollitrault,

Methods for systematic study of nuclear structure in high-energy collisions

. Eur. Phys. J. A 59, 110 (2023). arXiv:2302.14026, https://doi.org/10.1140/epja/s10050-023-01021-8
Baidu ScholarGoogle Scholar
27. L.W. Chen, C.M. Ko, B.A. Li,

Nuclear matter symmetry energy and the neutron skin thickness of heavy nuclei

. Phys. Rev. C 72, 064309 (2005). arXiv:nucl-th/0509009, https://doi.org/10.1103/PhysRevC.72.064309
Baidu ScholarGoogle Scholar
28. M.B. Tsang, et al.,

Constraints on the symmetry energy and neutron skins from experiments and theory

. Phys. Rev. C 86, 015803 (2012). arXiv:1204.0466, https://doi.org/10.1103/PhysRevC.86.015803
Baidu ScholarGoogle Scholar
29. B.J. Cai, L.W. Chen,

Constraints on the skewness coefficient of symmetric nuclear matter within the nonlinear relativistic mean field model

. Nucl. Sci. Tech. 28, 185 (2017). arXiv:1402.4242, https://doi.org/10.1007/s41365-017-0329-1
Baidu ScholarGoogle Scholar
30. X.F. Li, D.Q. Fang, Y.G. Ma,

Determination of the neutron skin thickness from interaction cross section and charge-changing cross section for B, C, N, O, F isotopes

. Nucl. Sci. Tech. 27, 71 (2016). https://doi.org/10.1007/s41365-016-0064-z
Baidu ScholarGoogle Scholar
31. H. Yu, D.Q. Fang, Y.G. Ma,

Investigation of the symmetry energy of nuclear matter using isospin-dependent quantum molecular dynamics

. Nucl. Sci. Tech. 31, 61 (2020). https://doi.org/10.1007/s41365-020-00766-x
Baidu ScholarGoogle Scholar
32. B.A. Li, B.J. Cai, W.J. Xie, et al.,

Progress in Constraining Nuclear Symmetry Energy Using Neutron Star Observables Since GW170817

. Universe 7, 182 (2021). arXiv:2105.04629, https://doi.org/10.3390/universe7060182
Baidu ScholarGoogle Scholar
33. Y. Wang, Z. Gao, H. , et al.,

Decoding the nuclear symmetry energy event-by-event in heavy-ion collisions with machine learning

. Phys. Lett. B 835, 137508 (2022). arXiv:2208.10681, https://doi.org/10.1016/j.physletb.2022.137508
Baidu ScholarGoogle Scholar
34. C.W. Ma, Y.P. Liu, H. Wei, et al.,

Determination of neutron-skin thickness using configurational information entropy

. Nucl. Sci. Tech. 33, 6 (2022). arXiv:2201.01442, https://doi.org/10.1007/s41365-022-00997-0
Baidu ScholarGoogle Scholar
35. R. An, S. Sun, L.G. Cao, et al.,

Constraining nuclear symmetry energy with the charge radii of mirror-pair nuclei

. Nucl. Sci. Tech. 34, 119 (2023). arXiv:2303.14667, https://doi.org/10.1007/s41365-023-01269-1
Baidu ScholarGoogle Scholar
36. P. Möller, A.J. Sierk, T. Ichikawa, et al.,

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

. Atom. Data Nucl. Data Tabl. 109-110, 1-204 (2016). arXiv:1508.06294, https://doi.org/10.1016/j.adt.2015.10.002
Baidu ScholarGoogle Scholar
37. K. Zhang, et al.,

Nuclear mass table in deformed relativistic Hartree–Bogoliubov theory in continuum, I: Even–even nuclei

. Atom. Data Nucl. Data Tabl. 144, 101488 (2022). arXiv:2201.03216, https://doi.org/10.1016/j.adt.2022.101488
Baidu ScholarGoogle Scholar
38. P.F. Kolb, J. Sollfrank, U.W. Heinz,

Anisotropic transverse flow and the quark hadron phase transition

. Phys. Rev. C 62, 054909 (2000). arXiv:hep-ph/0006129, https://doi.org/10.1103/PhysRevC.62.054909
Baidu ScholarGoogle Scholar
39. M.L. Miller, K. Reygers, S.J. Sanders, et al.,

Glauber modeling in high energy nuclear collisions

. Ann. Rev. Nucl. Part. Sci. 57, 205-243 (2007). arXiv:nucl-ex/0701025, https://doi.org/10.1146/annurev.nucl.57.090506.123020
Baidu ScholarGoogle Scholar
40. C. Loizides, J. Kamin, D. d’Enterria,

Improved Monte Carlo Glauber predictions at present and future nuclear colliders

. Phys. Rev. C 97, 054910 (2018). [Erratum: Phys.Rev.C 99, 019901 (2019)]. arXiv:1710.07098, https://doi.org/10.1103/PhysRevC.97.054910
Baidu ScholarGoogle Scholar
41. J.S. Moreland, J.E. Bernhard, S.A. Bass,

Alternative ansatz to wounded nucleon and binary collision scaling in high-energy nuclear collisions

. Phys. Rev. C 92, 011901 (2015). arXiv:1412.4708, https://doi.org/10.1103/PhysRevC.92.011901
Baidu ScholarGoogle Scholar
42. J.E. Bernhard, J.S. Moreland, S.A. Bass, et al.,

Applying Bayesian parameter estimation to relativistic heavy-ion collisions: simultaneous characterization of the initial state and quark-gluon plasma medium

. Phys. Rev. C 94, 024907 (2016). arXiv:1605.03954, https://doi.org/10.1103/PhysRevC.94.024907
Baidu ScholarGoogle Scholar
43. A.M. Poskanzer, S.A. Voloshin,

Methods for analyzing anisotropic flow in relativistic nuclear collisions

. Phys. Rev. C 58, 1671-1678 (1998). arXiv:nucl-ex/9805001, https://doi.org/10.1103/PhysRevC.58.1671
Baidu ScholarGoogle Scholar
44. J. Noronha-Hostler, L. Yan, F.G. Gardim, et al.,

Linear and cubic response to the initial eccentricity in heavy-ion collisions

. Phys. Rev. C 93, 014909 (2016). arXiv:1511.03896, https://doi.org/10.1103/PhysRevC.93.014909
Baidu ScholarGoogle Scholar
45. S. Zhao, H.j. Xu, Y.X. Liu, et al.,

Probing the nuclear deformation with three-particle asymmetric cumulant in RHIC isobar runs

. Phys. Lett. B 839, 137838 (2023). arXiv:2204.02387, https://doi.org/10.1016/j.physletb.2023.137838
Baidu ScholarGoogle Scholar
46. H.Z. Wu, J.J. Zhang, L.G. Pang, et al.,

ZMCintegral: a Package for Multi-Dimensional Monte Carlo Integration on Multi-GPUs

. Comput. Phys. Commun. 248, 106962 (2020). arXiv:1902.07916, https://doi.org/10.1016/j.cpc.2019.106962
Baidu ScholarGoogle Scholar
47. B. Pritychenko, M. Birch, B. Singh, et al.,

Tables of E2 Transition Probabilities from the first 2+ States in Even-Even Nuclei

. Atom. Data Nucl. Data Tabl. 107, 1-139 (2016). [Erratum: Atom.Data Nucl.Data Tabl. 114, 371–374 (2017)]. arXiv:1312.5975, https://doi.org/10.1016/j.adt.2015.10.001
Baidu ScholarGoogle Scholar
48. T. KIBÉDI, R.H. SPEAR,

REDUCED ELECTRIC-OCTUPOLE TRANSITION PROBABILITIES, B (E 3;01+ 31-)—AN UPDATE

. Atom. Data Nucl. Data Tabl. 80, 35-82 (2002). https://doi.org/10.1006/adnd.2001.0871
Baidu ScholarGoogle Scholar
49. B. Alver, G. Roland,

Collision geometry fluctuations and triangular flow in heavy-ion collisions

. Phys. Rev. C 81, 054905 (2010). [Erratum: Phys.Rev.C 82, 039903 (2010)]. arXiv:1003.0194, https://doi.org/10.1103/PhysRevC.82.039903
Baidu ScholarGoogle Scholar
50. B. Schenke, S. Jeon, C. Gale,

Elliptic and triangular flow in event-by-event (3+1)D viscous hydrodynamics

. Phys. Rev. Lett. 106, 042301 (2011). arXiv:1009.3244, https://doi.org/10.1103/PhysRevLett.106.042301
Baidu ScholarGoogle Scholar
51. S. Eremin, S. Voloshin,

Nucleon participants or quark participants

? Phys. Rev. C 67, 064905 (2003). arXiv:nucl-th/0302071, https://doi.org/10.1103/PhysRevC.67.064905
Baidu ScholarGoogle Scholar
52. C. Loizides,

Glauber modeling of high-energy nuclear collisions at the subnucleon level

. Phys. Rev. C 94, 024914 (2016). arXiv:1603.07375, https://doi.org/10.1103/PhysRevC.94.024914
Baidu ScholarGoogle Scholar
53. J.L. Albacete, H. Petersen, A. Soto-Ontoso,

Symmetric cumulants as a probe of the proton substructure at LHC energies

. Phys. Lett. B 778, 128-136 (2018). arXiv:1707.05592, https://doi.org/10.1016/j.physletb.2018.01.011
Baidu ScholarGoogle Scholar
Footnote

The authors declare that they have no competing interests.