logo

Correlations of the net baryon number and electric charge in nuclear matter

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

Correlations of the net baryon number and electric charge in nuclear matter

Xin-Ran Yang
Guo-Yun Shao
Chong-Long Xie
Zhi-Peng Li
Nuclear Science and TechniquesVol.36, No.8Article number 138Published in print Aug 2025Available online 03 Jun 2025
14400

We investigated the correlations between the net baryon number and electric charge up to the sixth order related to the interactions of nuclear matter at low temperature and explored their relationship with the nuclear liquid-gas phase transition (LGPT) within the framework of the nonlinear Walecka model. The calculations showed that strong correlations between the baryon number and electric charge existed near the LGPT, and higher-order correlations were more sensitive than the lower-order correlations near the phase transition. However, in the high-temperature region away from the LGPT, the rescaled lower-order correlations were relatively larger than most of the higher-order correlations. In addition, some of the fifth- and sixth-order correlations possibly changed sign from negative to positive along the chemical freeze-out line with decreasing temperature. In combination with future experimental projects at lower collision energies, the derived results can be used to study the phase structure of strongly interacting matter and analyze the related experimental signals.

Correlations of conserved chargesNuclear matterNuclear liquid-gas phase transitionHeavy-ion collision
1

Introduction

Mapping the phase diagram of quantum chromodynamics (QCD)is a primary objective in nuclear physics, which involves chiral and deconfinement phase transitions related to the transformation of quark-gluon plasma to hadronic matter [1]. The calculations from lattice QCD and hadron resonance gas (HRG) model indicate that a smooth crossover tranformation occurs at high temperatures and small chemical potentials [2-8]. Furthermore, studies on effective quark models [9-23], the Dyson-Schwinger equation approach [24-29], the functional renormalization group theory [30-32] and machine learning [33] suggest that a first-order chiral phase transition occurs at large chemical potentials.

The fluctuations and correlations of conserved charges (baryon number B, electric charge Q, and strangeness S) are sensitive observables for studying the phase transitions of strongly interacting matter [34, 35]. The net proton (proxy for net baryon) cumulants measured in the beam energy scan (BES) program at the Relativistic Heavy Ion Collider (RHIC) [36-42] has inspired extensive studies on QCD phase transition, particularly the QCD critical endpoint (CEP). More impressively, the distributions of the net proton number at the center-of-mass energy sNN=3 GeV and 2.4 GeV are different from those at 7.7 GeV and above because the fluctuation distributions of the net proton number are primarily dominated by the interaction among hadrons [40].

The experimental results at 3 GeV and below necessitate studies on the effect of hadronic interactions on the fluctuations of conserved charges at lower energies [43-46]. The nuclear liquid-gas phase transition (LGPT) may be involved at lower collision energies [47-63].

A van der Waals model was used to study the high-order distributions of the net baryon number in both pure and mixed phases of the LGPT[64-66]. The second-order susceptibility of the net baryon number for positive- and negative-parity nucleons was examined near the chiral and nuclear liquid-gas phase transitions using a double-parity model, in which both the chiral phase transition and nuclear LGPT are effectively included [45]. The net baryon kurtosis and skewness were considered in the nonlinear Walecka model to analyze the experimental signals at lower collision energies [55, 56]. The hyperskewness and hyperkurtosis of the net baryon number were recently calculated to explore the relationship between nuclear LGPT and experimental observables [67].

Because the interactions among hadrons dominate the density fluctuations in lower-energy regimes (below 3 GeV), the BES program at collision energies lower than 7.7 GeV is expected to provide detailed information on the phase structure of strongly interacting matter. Additionally, relevant experiments have been planned at the High Intensity Heavy-ion Accelerator Facility (HIAF). Meanwhile, the HADES collaboration at the GSI Helmholtzzentrum für Schwerionenforschung planned to measure the higher-order net proton and net charge fluctuations in the central Au + Au reactions at collision energies ranging from 0.2 A GeV to 1.0 A GeV to probe the LGPT region [68]. These experiments are significant for investigating nuclear liquid-gas and chiral phase transitions through density fluctuations.

In addition to the fluctuations in conserved charges, the correlations between different conserved charges provide important information for exploring phase transitions. The correlations of conserved charges or off-diagonal susceptibilities have been calculated to study the chiral and deconfinement phase transitions at high temperatures in lattice QCD and some effective quark models (e.g., [69-75]). However, correlations between the net baryon number and electric charge in nuclear matter and their relationship with nuclear LGPT, which are useful for diagnosing the phase diagram of strongly interacting matter at low temperatures, have not yet been explored. In this study, we explored the correlations between the net baryon number and electric charge up to the sixth order in nuclear matter using the nonlinear Walecka model. The characteristic behaviors of correlations evoked by the nucleon-nucleon interaction, both near and far away from the nuclear LGPT, were obtained. These results are expected to aid future analyses of chiral phase transitions, nuclear LGPT, and related experimental signals.

The remainder of this paper is organized as follows. In Sect. 2, we introduce formulas to describe the correlations between conserved charges and the nonlinear Walecka model. In Sect. 3, we illustrate the numerical results for the correlation between the net baryon number and electric charge. Finally, a summary is presented in Sect. 4.

2

Theoretical descriptions

The fluctuations and correlations of conserved charges are related to the equation of state of a thermodynamic system. In the grand canonical ensemble of strongly interacting matter, the pressure is the logarithm of the partition function [76]: P=TVlnZ(V,T,μB,μQ,μS), (1) where μB,μQ,μS are the chemical potentials of the conserved charges, that is, the baryon number, electric charge, and strangeness in a strong interaction, respectively. The generalized susceptibilities are derived by taking the partial derivatives of the pressure with respect to the corresponding chemical potentials [39] χijkBQS=i+j+k[P/T4](μB/T)i(μQ/T)j(μS/T)k. (2) The cumulants of the multiplicity distributions of the conserved charges are usually measured experimentally. These are related to the generalized susceptibilities by CijkBQS=i+j+kln[Z(V,T,μB,μQ,μS)](μB/T)i(μQ/T)j(μS/T)k=VT3χijkBQS. (3) To eliminate the volume dependence in heavy-ion collision experiments, observables are usually constructed using the ratios of cumulants and then compared with the theoretical calculations of the generalized susceptibilities with CijkBQSClmnBQS=χijkBQSχlmnBQS. (4) In this study, the nonlinear Walecka model was used to calculate the correlations between the net baryon number and electric charge in nuclear matter at low temperatures. This model describes the properties of finite nuclei and the equation of state of nuclear matter. The approximate equivalence of this model with the HRG model at low temperatures and densities is also indicated in Ref. [77]. This model was recently used to explore fluctuations in the net baryon number in nuclear matter, for example, kurtosis and skewness [55, 56] and hyperskewness and hyperkurtosis [67].

The Lagrangian density for the nucleon–meson system in the nonlinear Walecka model [54, 78] is L=Nψ¯N[iγμμ(mNgσσ)gωγμωμgργμτρμ]ψN +12(μσμσmσ2σ2)13bmN(gσσ)314c(gσσ)4 +12mω2ωμωμ14ωμνωμν +12mρ2ρμρμ14ρμνρμν, (5) where ωμν=μωννωμ, ρμν=∂μρν-∂νρμ and mN is the nucleon mass in vacuum. The interactions between nucleons are mediated by σ, ω, ρ mesons..

The thermodynamic potential can be derived in the mean-field approximation as Ω=β1N2d3k(2π)3[ln(1+eβ(EN*(k)μN*)) +ln(1+eβ(EN*(k)+μN*))]+12mσ2σ2+13bmN(gσσ)3 +14c(gσσ)412mω2ω212mρ2ρ32, (6) where β=1/T, EN*=k2+mN*2, and ρ3 denotes the third component of the ρ meson field. The effective nucleon mass mN*=mNgσσ and effective chemical potential μN* is defined as μN*=μNgωωτ3Ngρρ3 (τ3N=1/2 for proton, -1/2 for neutron).

Minimizing the thermodynamical potential Ωσ=Ωω=Ωρ3=0, (7) the meson field equations can be derived as: gσσ=(gσmσ)2[ρps+ρnsbmN(gσσ)2c(gσσ)3], (8) gωω=(gωmω)2(ρp+ρn), (9) gρρ3=12(gρmρ)2(ρpρn). (10) In Eqs.(8)(10), the nucleon number density ρi=2d3k(2π)3[f(Ei*μi*)f¯(Ei*+μi*)], (11) and scalar density ρis=2d3k(2π)3mi*Ei*[f(Ei*μi*)+f¯(Ei*+μi*)], (12) where f(Ei*μi*) and f¯(Ei*+μi*) are the fermion and antifermion distribution functions, respectively, with f(Ei*μi*)=11+exp{[Ei*μi*]/T}, (13) and f(Ei*+μi*)=11+exp{[Ei*+μi*]/T}. (14) The mesonfield equations can be solved for a given temperature and chemical potential (or baryon number density). The model parameters, gσ,gω,gρ,b and c, are listed in Table 1. They were fitted with the compression modulus K=240 MeV, symmetric energy asym=31.3 MeV, effective nucleon mass mN*=mNgσσ=0.75 mN, and binding energy B/A=16.0 MeV at nuclear saturation density with ρ0=0.16 fm3.

Table 1
Parameters in the nonlinear Walecka model
(gσ/mσ)2 (fm2) (gω/mω)2 (fm2) (gρ/mρ)2 (fm2) b c
10.329 5.423 0.95 0.00692 -0.0048
Show more
3

Results and discussion

In this section, we present the numerical results for the correlation between the net baryon number and electric charge in the nonlinear Walecka model. To simulate the physical conditions in the BES program at RHIC STAR, the isospin asymmetric nuclear matter was considered in the calculation with the constraint ρQ/ρB=0.4. In the present Walecka model, strange baryons were not included; thus, the strangeness condition ρS=0 was automatically satisfied. ρQ/ρB = 0.4 might deviate slightly owing to isospin dynamics. The influence of different isospin asymmetries on the fluctuations and correlations of conserved charges will be explored in detail in a separate study.

The correlations between the baryon number and electric charge are related to the baryon (μB) and isospin μQ (μQ=μpμn) chemical potentials. In Fig. 1, we demonstrate the value of μQ as a function of the temperature and baryon chemical potential by first plotting the contour map of μQ in the TμB plane derived under the constraint of ρQ/ρB=0.4. Additionally, the corresponding liquid-gas phase transition line with a CEP located at T=13 MeV and μB=919 MeV is plotted in this figure. To compare with the chiral crossover phase transition of quarks, the dashed “Line A” in Fig. 1 was derived under the condition that the value of σ/μB was maximum for each given temperature. To a certain degree, this line is analogous to a chiral crossover transformation, although it is not a true phase transition in nuclear matter. It indicates the location at which the dynamic nucleon mass changes most rapidly with an increase in the chemical potential. The reason for plotting “Line A” is to emphasize that both the σ field in nuclear matter and the quark condensate in quark matter are associated with the dynamic mass of fermions, and therefore, the rapid change of mass might have a universal effect on the fluctuation distributions of conserved charges. As indicated in our previous studies [54, 55, 67], the location of line A helps us understand the behavior of the interaction measurement (trace anomaly) and the fluctuations of conserved charges near the phase transition [55, 67].

Fig. 1
(Color online) Contour of μQ in the TμB plane derived in the nonlinear Walecka model with the constraint of ρQ/ρB=0.4. The solid line is the liquid-gas transition line with a CEP located at T=13 MeV and μB=919 MeV. The blue line is the chemical freeze-out line fitted in Ref. [79]. The dash-dotted line corresponds to the temperature and chemical potential for ρB=0.1ρ_0. “Line A” is derived using the maximum value of σ/μB for each given temperature
pic

Additionally, “Line A” can be defined by the maximum point of ω/μB or nB/μB, because the density can be considered as the order parameter for liquid-gas phase transition. Under this definition, the results obtained using the quark model do not correspond to a chiral crossover phase transition. However, this was not the purpose of the present study. Our aim was to identify common properties related to the dynamical fermion mass near the critical region of a first-order phase transition. However, the calculation indicates that the curves (“Line A”) under the two definitions coincide near the critical region and gradually deviate at higher temperatures away from the critical region.

For convenience, in the subsequent discussion of the experimental observables, we include a plot of the chemical freeze-out line fitted to the experimental data at high energies in Fig. 1 [79], which can be described by T(μB)=abμB2cμB4, (15) where a=0.166 GeV, b=0.139 GeV1 and c=0.053 GeV3.

It should be noted that the trajectories of the present relativistic heavy-ion collisions do not pass through TC of nuclear LGPT. It is still not known how far the realistic chemical freeze-out line is from the critical region at the present time. However, similar to the chiral phase transition of quarks, the existence of nuclear LGPT affects the fluctuation and correlation of the net baryon and electric charge numbers in the region not adjacent to the critical endpoint in intermediate-energy heavy-ion collision experiments. The numerical results for the parameterized chemical freeze-out line in this study can be used as a reference. The realistic chemical freeze-out conditions at intermediate and low energies will be extracted in future heavy-ion collision experiments. The contribution from LGPT needs to be considered when analyzing the experimental data.

Figure 1 shows that the value of |μQ| is less than 40 MeV in the red area. In this region, the baryon number density is very small, which is approximately indicated by the temperature and chemical potential curves for ρB=0.1ρ_0(dash-dotted line). The value of |μQ| increases with the baryon density(corresponding to a larger chemical potential). This trend in |μQ| is illustrated in Fig. 1. Along the chemical freeze-out line(solid blue line), changes in μQ during freeze-out with decreasing temperature or collision energy are observed.

Figure 2 shows the second-order correlation between the baryon number and electric charge, χ11BQ/χ2Q, as a function of the baryon chemical potential for T=75, 50, 25, MeV. To derive a physical quantity comparable with future experimental data, the correlated susceptibility was divided by χ2Q to eliminates volume dependence. For each temperature, the rescaled second-order correlation χ11BQ/χ2Q shown in Fig. 2 displays nonmonotonic behavior with a peak at a certain chemical potential. The values of these peaks increase with decreasing temperature, which indicates that the correlation between the baryon number and electric charge is enhanced near the phase transition region. The solid dots in Fig. 2 show the values for the chemical freeze-out described by Eq. (15), which illustrates that the value of χ11BQ/χ2Q increases along the freeze-out line when moving from the high-temperature region to the critical region.

Fig. 2
(Color online) Second order correlation between the baryon number and electric charge as a function of the chemical potential for different temperatures. The solid dots indicate the values on the chemical freeze-out line given in Fig. 1
pic

Figure 3 shows the third-order correlations χ12BQ/χ2Q and χ21BQ/χ2Q as functions of the chemical potential at several temperatures. Compared with χ12BQ/χ2Q, χ21BQ/χ2Q exhibits relatively larger fluctuations at the same temperature. The solid dots on the chemical freeze-out line show the same trend. This means that the measurement of χ21BQ/χ2Q is more sensitive than χ12BQ/χ2Q in heavy-ion collision experiments. Figure 3 also indicates that the correlations between the baryon number and electric charge intensify with decreasing temperatures. Evident oscillations of χ12BQ/χ2Q and χ21BQ/χ2Q appear for T=25 MeV, accompanied by alternating positives and negatives. As the temperature decreases, divergent behavior occurs at the CEP of LGPT. These features can be used to determine the signal for the phase transition in the experiments.

Fig. 3
(Color online) Third order correlations between the baryon number and electric charge as functions of the chemical potential at different temperatures. The solid dots indicate the values on the chemical freeze-out line plotted in Fig. 1
pic

In Fig. 4, we plot the fourth-order correlations between the baryon number and electric charge: χ13BQ/χ2Q, χ22BQ/χ2Q, and χ31BQ/χ2Q. Compared to the second- and third-order correlations, Figs. Fig. 2, Fig. 3, 4 shows that the fourth-order correlations rescaled by χ2Q are weaker at higher temperatures such as T=75 MeV. However, the correlations are much stronger at T=25 MeV, near the critical region of LGPT. Correspondingly, with an increase in the chemical potential at lower temperatures, a bimodal structure evidently exists for all three correlations. In addition, the maximum values of χ13BQ/χ2Q, χ22BQ/χ2Q and χ31BQ/χ2Q increase sequentially. Furthermore, the solid dots demonstrate that the value of each correlation during freeze-out increases with the decreasing temperatures. Moreover, χ13BQ/χ2Q<χ22BQ/χ2Q<χ31BQ/χ2Q during chemical freeze-out at each temperature, which implies that χ31BQ/χ2Q is the most sensitive among the three fourth-order correlations.

Fig. 4
(Color online) Fourth order correlations between the baryon number and electric charge as functions of the chemical potential for different temperatures. The solid dots indicate the values on the chemical freeze-out line given in Fig. 1
pic

Figure 5 presents the fifth-order correlations between the baryon number and electric charge, χ14BQ/χ2Q, χ23BQ/χ2Q, χ32BQ/χ2Q, and χ41BQ/χ2Q for T=75, 50, 25, MeV. At T=75 MeV, the values of the four rescaled correlations are small; however, they become significant at T=25 MeV. Combined with the phase diagrams shown in Fig. 1, it is seen that the high-order correlated fluctuations strengthen as they approach the liquid-gas transition. Similar to the fourth-order correlations, the rescaled fifth correlations fulfill the relationships |χ14BQ/χ2Q|<|χ23BQ/χ2Q|<|χ32BQ/χ2Q|<|χ41BQ/χ2Q| at chemical freeze-out. Moreover, all the four fifth-order correlation fluctuations are negative at chemical freeze-out for T=75 and 50 MeV; however, they are positive at T=25 MeV, near the liquid-gas transition. This is a prominent feature for exploring the interactions and phase transitions of nuclear matter.

Fig. 5
(Color online) Fifth order correlations between the baryon number and electric charge as functions of the chemical potential for different temperatures. The solid dots indicate the values on the chemical freeze-out line given in Fig. 1
pic

Figure 6 shows the sixth order correlations of the baryon number and electric charge, that is, χ15BQ/χ2Q, χ24BQ/χ2Q, χ33BQ/χ2Q, χ42BQ/χ2Q, and χ51BQ/χ2Q. Each sixth-order correlation exhibits a double peak and double valley structure, although one of the two peaks is not prominent. The oscillating behavior intensifies when moving towards the phase transition region from high to low temperatures. Similarly, the intensity of the oscillations increases from χ15BQ/χ2Q, χ24BQ/χ2Q, χ33BQ/χ2Q, χ42BQ/χ2Q to χ51BQ/χ2Q.

Fig. 6
(Color online) Sixth order correlations between the baryon number and electric charge as functions of the chemical potential for different temperatures. The solid dots demonstrate the values on the chemical freeze-out line given in Fig. 1
pic

For a given order of correlations, the numerical results shown in Fig. 2, 3, 4, 5, and 6 indicates that the signals become stronger when taking the higher-order partial derivatives of the baryon chemical potential. Additionally, we examined the pure baryon number fluctuation and found that its highest sensitivity was of the same order as the LGPT critical endpoint, possibly because the baryon number fluctuation includes both proton and neutron contributions. However, the electric charge fluctuation involves the isospin density ρNρP. The baryon number density is always larger than the isospin density, which is associated with stronger fluctuations when there are more derivatives with respect to the baryon chemical potential than that with to electric chemical potential for a given order of correlations.

In addition, a comparison of the results shown in Fig. 2, 3, 4, 5, and 6 shows that the rescaled higher-order correlations fluctuate more strongly near the phase transition region, whereas the lower-order correlations at high temperatures are larger than most of the higher-order correlations away from the phase transition region. A similar phenomenon occurs in the correlations of conserved charges in quark matter [74]. According to the fluctuations of net baryon number [55, 67] and the correlations between net baryon number and electric charge in this study, the fluctuations and correlations of conserved charges have similar organizational structures for nuclear and quark matter. This is primarily attributed to the fact that the two phase transitions belong to the same universal class, and both describe the interaction of matter with temperature- and chemical-potential-dependent fermion masses.

Because the QCD phase transition and nuclear LGPT possibly occur sequentially from high to low temperatures (even if LGPT is not triggered), the energy-dependent behaviors of the fluctuations and correlations can be referenced to determine the phase transition signals of the strongly interacting matter. Although the latest reported BES II high-precision data at 7.7-3.9 GeV do not display a drastic change in the net baryon number kurtosis, stronger fluctuation signals may appear in heavy-ion experiments with collision energies lower than 7.7 GeV. Furthermore, in the hadronic interaction dominant evolution with collision energies lower than the threshold of the generation of QGP, the nuclear interaction and phase structure of LGPT will dominate over the behavior of fluctuations and correlations of conserved charges. The nature of the changes in fluctuations and correlations with decreasing collision energy during experiments requires investigation.

4

Summary

Fluctuations and correlations between conserved charges are sensitive probes for investigating the phase structure of strongly interacting matter. In this study, we used the non-linear Walecka model to calculate the correlations between the net baryon number and electric charge up to the sixth order, which originated from hadronic interactions in nuclear matter and explored their relationship with the nuclear liquid-gas phase transition.

The calculation indicated that the correlations between the net baryon number and electric charge gradually became stronger from the high-temperature region to the critical region of the nuclear LGPT. In particular, the correlations were significant at the location where the σ field or nucleon mass changed rapidly near the critical region. Similar behavior was observed for the chiral crossover phase transition of quark matter, primarily because of the similar dynamic mass evolution and same universal class of the chiral phase transition of quark matter and the liquid-gas phase transition of nuclear matter.

Compared to the lower-order correlations, the higher-order correlations fluctuated more strongly near the phase transition region, whereas the rescaled lower-order correlations were relatively stronger than most of the higher-order correlations away from the phase transition region at high temperatures. At the chemical freeze-out for each temperature, the calculation indicated that χ13BQ/χ2Q<χ22BQ/χ2Q<χ31BQ/χ2Q for the fourth order correlation, |χ14BQ/χ2Q|<|χ23BQ/χ2Q|<|χ32BQ/χ2Q|<|χ41BQ/χ2Q| for the fifth order correlations, and |χ15BQ/χ2Q|<|χ24BQ/χ2Q|<|χ33BQ/χ2Q|<|χ42BQ/χ2Q|<|χ51BQ/χ2Q| for the sixth order correlations. In particular, the values of the fifth- and sixth-order correlations changed from negative to positive when approaching the critical region of LGPT from the high-temperature side along the extrapolated chemical freeze-out line. With the availability of more precise data from experiments below 7.7 GeV in the future, realistic chemical freeze-out conditions can be fitted, and the results obtained in this study can be used to analyze the signals of QCD phase transitions and the influence of the nuclear liquid-gas phase transition.

References
1. Q.Y. Shou, Y.G. Ma, S. Zhang et al.,

Properties of QCD matter: a review of selected results from ALICE experiment

. Nucl. Sci. Tech. 35, 219 (2024). https://doi.org/10.1007/s41365-024-01583-2
Baidu ScholarGoogle Scholar
2. Y. Aoki, G. Endrodi, Z. Fodor et al.,

The order of the quantum chromodynamics transition predicted by the standard model of particle physics

. Nature 443, 675-678 (2006). https://doi.org/10.1038/nature05120
Baidu ScholarGoogle Scholar
3. A. Bazavov, H.T. Ding, P. Hegde et al., (HotQCD Collaboration),

Chiral crossover in QCD at zero and non-zero chemical potentials

. Phys. Lett. B 795, 15 (2019). https://doi.org/10.1016/j.physletb.2019.05.013
Baidu ScholarGoogle Scholar
4. S. Borsányi, Z. Fodor, S.D. Katz et al.,

Freeze-Out Parameters: Lattice Meets Experiment

. Phys. Rev. Lett. 111, 062005 (2013). https://doi.org/10.1103/PhysRevLett.111.062005
Baidu ScholarGoogle Scholar
5. A. Bazavov, T. Bhattacharya, C. DeTar et al., (HotQCD Collaboration),

Equation of state in (2+1)-flavor QCD

. Phys. Rev. D 90, 094503 (2014). https://doi.org/10.1103/PhysRevD.90.094503
Baidu ScholarGoogle Scholar
6. A. Bazavov, H.T. Ding, P. Hegde et al., (HotQCD Collaboration),

Skewness and kurtosis of net baryon-number distributions at small values of the baryon chemical potential

. Phys. Rev. D 96, 074510 (2017). https://doi.org/10.1103/PhysRevD.96.074510
Baidu ScholarGoogle Scholar
7. S. Borsányi, Z. Fodor, C. Hoelbling et al.,

Full result for the QCD equation of state with flavors

. Phys. Lett. B 730, 99 (2014). https://doi.org/10.1016/j.physletb.2014.01.007
Baidu ScholarGoogle Scholar
8. S. Borsányi, Z. Fodor, J.N. Guenther et al.,

QCD Crossover at Finite Chemical Potential from Lattice Simulations

. Phys. Rev. Lett. 125, 052001 (2020). https://doi.org/10.1103/PhysRevLett.125.052001
Baidu ScholarGoogle Scholar
9. K. Fukushima,

Chiral effective model with the Polyakov loop

. Phys. Lett. B 591, 277 (2004). https://doi.org/10.1016/j.physletb.2004.04.027
Baidu ScholarGoogle Scholar
10. C. Ratti, M.A. Thaler, W. Weise,

Phases of QCD: Lattice thermodynamics and a field theoretical model

. Phys. Rev. D 73, 014019 (2006). https://doi.org/10.1103/PhysRevD.73.014019
Baidu ScholarGoogle Scholar
11. P. Costa, M.C. Ruivo, C.A. De Sousa et al.,

Phase diagram and critical properties within an effective model of QCD: the Nambu-Jona-Lasinio model coupled to the Polyakov loop

. Symmetry 2, 1338-1374 (2010). https://doi.org/10.3390/sym2031338
Baidu ScholarGoogle Scholar
12. W.J. Fu, Z. Zhang, Y.X. Liu,

2+1 flavor Polyakov-Nambu-Jona-Lasinio model at finite temperature and nonzero chemical potential

. Phys. Rev. D 77, 014006 (2008). https://doi.org/10.1103/PhysRevD.77.014006
Baidu ScholarGoogle Scholar
13. T. Sasaki, J. Takahashi, Y. Sakai et al.,

Theta vacuum and entanglement interaction in the three-flavor Polyakov-loop extended Nambu-Jona-Lasinio model

. Phys. Rev. D 85, 056009 (2012). https://doi.org/10.1103/PhysRevD.85.056009
Baidu ScholarGoogle Scholar
14. M. Ferreira, P. Costa, C. Providência,

Deconfinement, chiral symmetry restoration and thermodynamics of (2+1)-flavor hot QCD matter in an external magnetic field

. Phys. Rev. D 89, 036006 (2014). https://doi.org/10.1103/PhysRevD.89.036006
Baidu ScholarGoogle Scholar
15. Y.Q. Zhao, S. He, D.F. Hou et al.,

Phase diagram of holographic thermal dense QCD matter with rotation

. J. High Energy Phys. 04, 115 (2023). https://doi.org/10.1007/JHEP04(2023)115
Baidu ScholarGoogle Scholar
16. G.Y. Shao, Z.D. Tang, X.Y. Gao et al.,

Baryon number fluctuations and the phase structure in the PNJL model

. Eur. Phys. J. C 78, 138 (2018). https://doi.org/10.1140/epjc/s10052-018-5636-0
Baidu ScholarGoogle Scholar
17. B.J. Schaefer, M. Wagner, J. Wambach,

Thermodynamics of (2+1)-flavor QCD: Confronting models with lattice studies

. Phys. Rev. D 81, 074013 (2010). https://doi.org/10.1103/PhysRevD.81.074013
Baidu ScholarGoogle Scholar
18. V. Skokov, B. Friman, K. Redlich,

Quark number fluctuations in the Polyakov loop-extended quark-meson model at finite baryon density

. Phys. Rev. C 83, 054904 (2011). https://doi.org/10.1103/PhysRevC.83.054904
Baidu ScholarGoogle Scholar
19. L.M. Liu, J. Xu, G.X. Peng,

Three-dimensional QCD phase diagram with a pion condensate in the NJL model

. Phys. Rev. D 104, 076009 (2021). https://doi.org/10.1103/PhysRevD.104.076009
Baidu ScholarGoogle Scholar
20. X. Chen, D.N. Li, D.F. Hou et al.,

Quarkyonic phase from quenched dynamical holographic QCD model

. J. High Energy Phys. 2020, 73 (2020). https://doi.org/10.1007/JHEP03(2020)073
Baidu ScholarGoogle Scholar
21. M. Ferreira, P. Costa, C. Providência,

Presence of a critical endpoint in the QCD phase diagram from the net-baryon number fluctuations

. Phys. Rev. D 98, 034006 (2018). https://doi.org/10.1103/PhysRevD.98.034006
Baidu ScholarGoogle Scholar
22. Y.H. Yang, L. He, P.C. Chu,

Properties of the phase diagram from the Nambu-Jona-Lasino model with a scalar-vector interaction

. Nucl. Sci. Tech. 35, 166 (2024). https://doi.org/10.1007/s41365-024-01559-2
Baidu ScholarGoogle Scholar
23. Y.D. Chen, D.N. Li, M. Huang,

Inhomogeneous chiral condensation under rotation in the holographic QCD

. Phys.Rev. D 106, 106002 (2022). https://doi.org/10.1103/PhysRevD.106.106002
Baidu ScholarGoogle Scholar
24. F. Gao, Y.X. Liu,

QCD phase transitions using the QCD Dyson-Schwinger equation approach

. Nucl. Tech. (in Chinese) 46 040015 (2024). https://doi.org/10.11889/j.02533219.2023.hjs.46.040015
Baidu ScholarGoogle Scholar
25. S.X. Qin, L. Chang, H. Chen et al.,

Phase Diagram and Critical End Point for Strongly Interacting Quarks

. Phys. Rev. Lett. 106, 172301 (2011). https://doi.org/10.1103/PhysRevLett.106.172301
Baidu ScholarGoogle Scholar
26. F. Gao, J. Chen, Y.X. Liu et al.,

Phase diagram and thermal properties of strong-interaction matter

. Phys. Rev. D 93, 094019 (2016). https://doi.org/10.1103/PhysRevD.93.094019
Baidu ScholarGoogle Scholar
27. F. Gao, J.M. Pawlowski,

QCD phase structure from functional methods

. Phys. Rev. D 102, 034027 (2020). https://doi.org/10.1103/PhysRevD.102.034027
Baidu ScholarGoogle Scholar
28. C.S. Fischer, J. Luecker, C.A. Welzbacher,

Phase structure of three and four flavor QCD

. Phys. Rev. D 90, 034022 (2014). https://doi.org/10.1103/PhysRevD.90.034022
Baidu ScholarGoogle Scholar
29. C. Shi, Y.L. Wang, Y. Jiang et al.,

Locate QCD critical end point in a continuum model study

. J. High Energy Phys. 1407, 014 (2014). https://doi.org/10.1007/JHEP07(2014)014
Baidu ScholarGoogle Scholar
30. W.J. Fu, J.M. Pawlowski, F. Rennecke,

QCD phase structure at finite temperature and density

. Phys. Rev. D 101, 054032 (2020). https://doi.org/10.1103/PhysRevD.101.054032
Baidu ScholarGoogle Scholar
31. F. Rennecke, B.J. Schaefer,

Fluctuation-induced modifications of the phase structure in (2+1)-flavor QCD

. Phys. Rev. D 96, 016009 (2017). https://doi.org/10.1103/PhysRevD.96.016009
Baidu ScholarGoogle Scholar
32. W.J. Fu, X.F. Luo, J.M. Pawlowski et al.,

Hyper-order baryon number fluctuations at finite temperature and density

. Phys. Rev. D 104, 094047 (2021). https://doi.org/10.1103/PhysRevD.104.094047
Baidu ScholarGoogle Scholar
33. Y.G. Ma, L.G. Pang, R. Wang et al.,

Phase Transition Study Meets Machine Learning

. Chin. Phys. Lett. 40, 122101 (2023). https://doi.org/10.1088/0256307X/40/12/122101
Baidu ScholarGoogle Scholar
34. M.A. Stephanov,

Non-Gaussian Fluctuations near the QCD Critical Point

. Phys. Rev. Lett. 102, 032301 (2009). https://doi.org/10.1103/PhysRevLett.102.032301
Baidu ScholarGoogle Scholar
35. M.A. Stephanov,

Sign of Kurtosis near the QCD Critical Point

. Phys. Rev. Lett. 107, 052301 (2011). https://doi.org/10.1103/PhysRevLett.107.052301
Baidu ScholarGoogle Scholar
36. M.M. Aggarwal, Z. Ahammed, A.V. Alakhverdyants et al., (STAR Collaboration),

Higher Moments of Net Proton Multiplicity Distributions at RHIC

. Phys. Rev. Lett. 105, 022302 (2010). https://doi.org/10.1103/PhysRevLett.105.022302
Baidu ScholarGoogle Scholar
37. L. Adamczyk, J.K. Adkins, G. Agakishiev et al., (STAR Collaboration),

Energy Dependence of Moments of Net-Proton Multiplicity Distributions at RHIC

. Phys. Rev. Lett. 112, 032302 (2014). https://doi.org/10.1103/PhysRevLett.112.032302
Baidu ScholarGoogle Scholar
38. J.H. Chen, X. Dong,

X.H. Properties of QCD matter: review of selected results from the relativistic heavy ion collider beam energy scan (RHIC BES) program

. Nucl. Sci. Tech. 35, 214 (2024). https://doi.org/10.1007/s41365-024-01591-2
Baidu ScholarGoogle Scholar
39. X.F. Luo, N. Xu,

Search for the QCD critical point with fluctuations of conserved quantities in relativistic heavy-ion collisions at RHIC: an overview

. Nucl. Sci. Tech. 28, 112 (2017). https://doi.org/10.1007/s41365-017-0257-0
Baidu ScholarGoogle Scholar
40. B.E. Aboona, J. Adam, L. Adamczyk et al., (STAR Collaboration),

Measurement of Sequential γ Suppression in Au+Au Collisions at sNN=200GeV with the STAR Experiment

. Phys. Rev. Lett. 130, 082301 (2023). https://doi.org/10.1103/PhysRevLett.130.112301
Baidu ScholarGoogle Scholar
41. Y. Zhang, D.W. Zhang, X.F. Luo,

Experimental study of the QCD phase diagram in relativistic heavy-ion collisions

. Nucl. Tech. (in Chinese), 46, 040001 (2024). https://doi.org/10.11889/j.02533219.2023.hjs.46.040001
Baidu ScholarGoogle Scholar
42. Q. Chen, G.L. Ma, J.H. Chen,

Transport model study of conserved charge fluctuations and QCD phase transition in heavy-ion collisions

. Nucl. Tech. (In Chinese) 46 040013 (2023). https://doi.org/10.11889/j.02533219.2023.hjs.46.040013
Baidu ScholarGoogle Scholar
43. C. Huang, Y.Y. Tan, R. Wen et al.,

Reconstruction of baryon number distributions

. Chin. Phys. C 47, 104106 (2023). https://doi.org/10.1088/1674-1137/aceee1
Baidu ScholarGoogle Scholar
44. J.X. Shao, W.J. Fu, Y.X. Liu,

Chemical freeze-out parameters via a functional renormalization group approach

. Phys. Rev. D 109, 034019 (2024). https://doi.org/10.1103/PhysRevD.109.034019
Baidu ScholarGoogle Scholar
45. M. Marczenko, K. Redlich, C. Sasaki,

Fluctuations near the liquid-gas and chiral phase transitions in hadronic matter

. Phys. Rev. D 107, 054046 (2023). https://doi.org/10.1103/PhysRevD.107.054046
Baidu ScholarGoogle Scholar
46. C. Liu, X.G. Deng, Y.G. Ma,

Density fluctuations in intermediate-energy heavy-ion collisions

. Nucl. Sci. Tech. 33, 52 (2022). https://doi.org/10.1007/s41365-02201040-y
Baidu ScholarGoogle Scholar
47. P. Chomaz, M. Colonna, J. Randrup,

Nuclear spinodal fragmentation

. Phys. Rep. 389, 263 (2004). https://doi.org/10.1016/j.physrep.2003.09.006
Baidu ScholarGoogle Scholar
48. J. Pochodzalla, T. Mohlenkamp, T. Rubehn et al., (ALADIN Collaboration),

Probing the Nuclear Liquid-Gas Phase Transition

. Phys. Rev. Lett. 75, 1040 (1995). https://doi.org/10.1103/PhysRevLett.75.1040
Baidu ScholarGoogle Scholar
49. B. Borderie, G. Tabacaru, P. Chomaz et al., (NDRA Collaboration),

Evidence for Spinodal Decomposition in Nuclear Multifragmentation

. Phys. Rev. Lett. 86, 3252 (2001). https://doi.org/10.1103/PhysRevLett.86.3252
Baidu ScholarGoogle Scholar
50. A.S. Botvina, I.N. Mishustin, M. Begemann-Blaich et al.,

Multifragmentation of spectators in relativistic heavy-ion reactions

. Nucl. Phys. A 584, 737-756 (1995). https://doi.org/10.1016/0375-9474(94)00621-S
Baidu ScholarGoogle Scholar
51. M. D’Agostino, A.S. Botvina, M. Bruno et al.,

Thermodynamical features of multifragmentation in peripheral Au + Au collisions at 35AMeV

. Nucl. Phys. A 650, 329-357 (1999). https://doi.org/10.1016/S03759474(99)00097-4
Baidu ScholarGoogle Scholar
52. B.K. Srivastava, R.P. Scharenberg, S. Albergo et al., (EOS Collaboration),

Multifragmentation and the phase transition: A systematic study of the multifragmentation of 1AGeV Au, La, and Kr

. Phys. Rev. C 65, 054617 (2002). https://doi.org/10.1103/PhysRevC.65.054617
Baidu ScholarGoogle Scholar
53. J.B. Elliott, L.G. Moretto, L. Phair et al., (ISiS Collaboration),

Liquid to Vapor Phase Transition in Excited Nuclei

. Phys. Rev. Lett. 88, 042701 (2002). https://doi.org/10.1103/PhysRevLett.88.042701
Baidu ScholarGoogle Scholar
54. W.B. He, G.Y. Shao, C.L. Xie,

Speed of sound and liquid-gas phase transition in nuclear matter

. Phys. Rev. C 107, 014903 (2023). https://doi.org/10.1103/PhysRevC.107.014903
Baidu ScholarGoogle Scholar
55. G.Y. Shao, X.Y. Gao, W.B. He,

Baryon number fluctuations induced by hadronic interactions at low temperature and large chemical potential

. Phys. Rev. D 101, 074029 (2020). https://doi.org/10.1103/PhysRevD.101.074029
Baidu ScholarGoogle Scholar
56. K. Xu, M. Huang,

The baryon number fluctuation κσ2 as a probe of nuclear matter phase transition at high baryon density

. arXiv:2307.12600v1. https://doi.org/10.48550/arXiv.2307.12600
Baidu ScholarGoogle Scholar
57. Y.G. Ma,

Application of Information Theory in Nuclear Liquid Gas Phase Transition

. Phys. Rev. Lett. 83, 3617 (1999). https://doi.org/10.1103/PhysRevLett.83.3617
Baidu ScholarGoogle Scholar
58. X.G. Deng, P. Danielewicz, Y.G. Ma et al.,

Impact of fragment formation on shear viscosity in the nuclear liquid-gas phase transition region

. Phys. Rev. C 105, 064613 (2022). https://doi.org/10.1103/PhysRevC.105.064613
Baidu ScholarGoogle Scholar
59. M. Hempel, V. Dexheimer, S. Schramm et al.,

Noncongruence of the nuclear liquid-gas and deconfinement phase transitions

. Phys. Rev. C 88, 014906 (2013). https://doi.org/10.1103/PhysRevC.88.014906
Baidu ScholarGoogle Scholar
60. A. Mukherjee, J. Steinheimer, S. Schramm,

Higher-order baryon number susceptibilities: Interplay between the chiral and the nuclear liquid-gas transitions

. Phys. Rev. C 96, 025205 (2017). https://doi.org/10.1103/PhysRevC.96.025205
Baidu ScholarGoogle Scholar
61. J. Xu, L.W. Chen, C.M. Ko et al.,

Shear viscosity of neutron-rich nucleonic matter near its liquid-gas phase transition

. Phys. Lett. B 727, 244 (2013). https://doi.org/10.1016/j.physletb.2013.10.051
Baidu ScholarGoogle Scholar
62. O. Savchuk, V. Vovchenko, R.V. Poberezhnyuk et al.,

Traces of the nuclear liquid-gas phase transition in the analytic properties of hot QCD

. Phys. Rev. C 101, 035205 (2020). https://doi.org/10.1103/PhysRevC.101.035205
Baidu ScholarGoogle Scholar
63. R. Wang, Y.G. Ma, R. Wada et al.,

Nuclear liquid-gas phase transition with machine learning

. Phys. Rev. Res. 2, 043202 (2020). https://doi.org/10.1103/PhysRevResearch.2.043202
Baidu ScholarGoogle Scholar
64. R.V. Poberezhnyuk, V. Vovchenko, A. Motornenko et al.,

Chemical freeze-out conditions and fluctuations of conserved charges in heavy-ion collisions within a quantum van der Waals model

. Phys. Rev. C 100, 054904 (2019). https://doi.org/10.1103/PhysRevC.100.054904
Baidu ScholarGoogle Scholar
65. V. Vovchenko, M.I. Gorenstein, H. Stöecker,

van der Waals Interactions in Hadron Resonance Gas: From Nuclear Matter to Lattice QCD

. Phys. Rev. Lett. 118, 182301 (2017). https://doi.org/10.1103/PhysRevLett.118.182301
Baidu ScholarGoogle Scholar
66. R.V. Poberezhnyuk, O. Savchuk, M.I. Gorenstein et al.,

Higher order conserved charge fluctuations inside the mixed phase

. Phys. Rev. C 103, 024912 (2021). https://doi.org/10.1103/PhysRevC.103.024912
Baidu ScholarGoogle Scholar
67. X.R. Yang, G.Y. Shao, W.B. He,

Fifth- and sixth-order net baryon number fluctuations in nuclear matter at low temperature

. Phys. Rev. C 109, 014908 (2024). https://doi.org/10.1103/PhysRevC.109.014908
Baidu ScholarGoogle Scholar
68. M. Bluhm, A. Kalweit, M. Nahrgang et al.,

Dynamics of critical fluctuations: Theory - phenomenology - heavy-ion collisions

. Nucl. Phys. A 1003, 122016 (2020). https://doi.org/10.1016/j.nuclphysa.2020.122016
Baidu ScholarGoogle Scholar
69. R. Bellwied, S. Borsányi, Z. Fodor et al.,

Fluctuations and correlations in high temperature QCD

. Phys. Rev. D 92, 114505 (2015). https://doi.org/10.1103/PhysRevD.92.114505
Baidu ScholarGoogle Scholar
70. R. Wen, W.J. Fu,

Correlations of conserved charges and QCD phase structure

. Chin. Phys. C 45, 044112 (2021). https://doi.org/10.1088/1674-1137/abe199
Baidu ScholarGoogle Scholar
71. H.T. Ding, S. Mukherjee, H. Ohno et al.,

Diagonal and off-diagonal quark number susceptibilities at high temperatures

. Phys. Rev. D 92, 074043 (2015). https://doi.org/10.1103/PhysRevD.92.074043
Baidu ScholarGoogle Scholar
72. S. Borsányi, Z. Fodor, J.N. Guenther et al.,

Higher order fluctuations and correlations of conserved charges from lattice QCD

. J. High Energy Phys. 10, 205 (2018). https://doi.org/10.1007/JHEP10(2018)205
Baidu ScholarGoogle Scholar
73. R. Bellwied, S. Borsányi, Z. Fodor et al.,

Off-diagonal correlators of conserved charges from lattice QCD and how to relate them to experiment

. Phys. Rev. D 101, 034506 (2020). https://doi.org/10.1103/PhysRevD.101.034506
Baidu ScholarGoogle Scholar
74. W.J. Fu, Y.L. Wu,

Fluctuations and correlations of conserved charges near the QCD critical point

. Phys. Rev. D 82, 074013 (2010). https://doi.org/10.1103/PhysRevD.82.074013
Baidu ScholarGoogle Scholar
75. A. Bhattacharyya, P. Deb, A. Lahiri et al.,

Correlation between conserved charges in Polyakov–Nambu–Jona-Lasinio model with multiquark interactions

. Phys. Rev. D 83, 014011 (2011). https://doi.org/10.1103/PhysRevD.83.014011
Baidu ScholarGoogle Scholar
76. H.T. Ding, F. Karsch, S. Mukherjee,

Thermodynamics of strong-interaction matter from lattice QCD

. Int. J. Mod. Phys. E 24, 530007 (2015). https://doi.org/10.1142/S0218301315300076
Baidu ScholarGoogle Scholar
77. K. Fukushima,

Hadron resonance gas and mean-field nuclear matter for baryon number fluctuations

. Phys. Rev. C 91, 044910 (2015). https://doi.org/10.1103/PhysRevC.91.044910
Baidu ScholarGoogle Scholar
78. N.K. Glendenning, COMPACT STARS, Nuclear Physics, Particle Physics, and General Relativity 2nd edn. (Springer, New York, 2000), pp.189
79. J. Cleymans, H. Oeschler, K. Redlich et al.,

Comparison of chemical freeze-out criteria in heavy-ion collisions

. Phys. Rev. C 73, 034905 (2006). https://doi.org/10.1103/PhysRevC.73.034905
Baidu ScholarGoogle Scholar
Footnote

The authors declare that they have no competing interests.