logo

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

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

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

Yong-Hang Yang
He Liu
Peng-Cheng Chu
Nuclear Science and TechniquesVol.35, No.10Article number 166Published in print Oct 2024Available online 18 Sep 2024
25808

We investigated the properties of the phase diagram of high-order susceptibilities, speed of sound, and polytropic index based on an extended Nambu-Jona-Lasinio model with an eight-quark scalar-vector interaction. Non-monotonic behavior was observed in all these quantities around the phase transition boundary, which also revealed the properties of the critical point. Further, this study indicated that the chiral phase transition boundary and critical point could vary depending on the scalar-vector coupling constant GSV. At finite densities and temperatures, the negative GSV term exhibited attractive interactions, which enhanced the critical point temperature and reduced the chemical potential. The GSV term also affected the properties of the high-order susceptibilities, speed of sound, and polytropic index near the critical point. The nonmonotonic (peak or dip) structures of these quantities shifted to a low baryon chemical potential (and high temperature) with a negative GSV. GSV also changed the amplitude and range of the nonmonotonic regions. Therefore, the scalar-vector interaction was useful for locating the phase boundary and critical point in QCD phase diagram by comparing the experimental data. The study of the non-monotonic behavior of high-order susceptibilities, speed of sound, and polytropic index is of great interest, and further observations related to high-order susceptibilities, speed of sound, and polytropic index being found and applied to the search for critical points in heavy-ion collisions and the study of compact stars are eagerly awaited.

QCD phase diagramHigh-order susceptibilitiesSpeed of soundPolytropic indexNJL model
1

Introduction

The search for a first-order phase transition and critical point (CP) in the Quantum Chromodynamics (QCD) phase diagram is a fundamental goal of relativistic heavy-ion collision experiments [1-4]. Lattice QCD (LQCD) simulations have shown that the phase transition between the quark-gluon plasma (QGP) and hadronic matter is a smooth crossover at nearly zero baryon chemical potential (μB~0) [5-7]. Based on investigations from various effective models, such as the Nambu-Jona-Lasinio (NJL) model and quark-meson (QM) model, as well as advanced functional methods including the Dyson-Schwinger equation (DSE) and functional renormalization group (FRG) [8-20], the transition can be of the first-order in the large μB region, resulting in a critical end point on the first-order phase transition line. However, different theoretical models and methods exhibit significant differences in the predicted CP location, even for the same model with different parameter sets [21]. To further explore the QCD phase diagram and search for a possible CP signal, experimental programs such as the Beam-Energy Scan (BES) are currently underway at the Relativistic Heavy-Ion Collider (RHIC). These involve varying the collision energy of heavy-ion collisions to cover a wide range of temperatures T and μB in the QCD phase diagram [22-24]. In the BES program, physical observations sensitive to CP and/or first-order phase transitions have been made, such as the pion Hanbury-Brown Twiss (HBT) radii [25, 26], baryon directional flow [27, 28], net proton fluctuation [29, 30], light nuclear yield ratio [31-33], and interaction of charged hadrons [34]. Although nonmonotonic dependencies of these observations on collision energy have been observed, significant uncertainties remain in the nonmonotonic energy range.

Fluctuations in the conserved charges such as net charge, net baryon number, and net strangeness are predicted to depend on the nonequilibrium correlation length (ξ), which diverges at the singular critical point in the ideal system. However, they are typically limited by the finite size or the finite time of the system, and thus serve as possible signals of CP of the strongly interacting QCD matter created in relativistic heavy-ion collisions [35, 36]. Recent reports from first phase of the beam energy scan program (BES-I) of the RHIC [29, 30] demonstrated the non-monotonic variation of high-order moments of proton multiplicity (replacing the net baryon number) with collision energy in the range 7.7 GeV<sNN<27 GeV. The measured higher-order moments refer to the skewness (S=(δN)3/σ3) and kurtosis (κ=(δN)4/σ43), where δN=NM, M is the mean and σ is the quadratic variance. Furthermore, the cumulant ratio, C4/C2=κσ2, of proton multiplicity distribution from sNN=3.0 GeV Au + Au collisions was reported [24]. Compared with the Poisson baseline, a significant suppression in protons C4/C2=-0.85 was observed in the most central 0%-5% collisions at 3 GeV. This can be explained by fluctuations driven by baryon number conservation in a region of high baryon density, where hadronic interactions are dominant. These results imply that the QCD critical region, if created in heavy-ion collisions, can only exist at energies higher than 3 GeV. Light nuclei production has been suggested for exploring the existence and location of the QCD first-order phase-transition region with a corresponding critical point. In the 0%-10% most central Au + Au collisions at sNN=19.6 GeV and 27 GeV, Nt×Np/Nd2 shows enhancements relative to the coalescence baseline, with a combined significance of 4.1σ [33]. Further studies on heavy-ion collisions are required to confirm whether these nonmonotonic dependencies are due to large fluctuations near the CP and to narrow the energy range at which CP may occur.

Hydrodynamics simulations provide another option for exploring QCD phase transitions. The spatiotemporal evolution of QCD matter can be successfully described using relativistic dissipative hydrodynamics [37-39]. The speed of sound is a crucial physical quantity in hydrodynamics, which carries important information for describing the evolution of strongly interacting matter and final observables in heavy-ion collision experiments. According to Ref. [40-42], the speed of sound as a function of the charged particle multiplicity dNch/dη can be extracted from heavy-ion collision data. A recent study Ref. [43] estimated the value of cs as well as its logarithmic derivative with respect to the baryon number density and attempted to build a connection with the cumulants of the net baryon number in matter created in heavy-ion collisions to aid in detecting the QCD CP. Furthermore, the behavior of cs as a function of the baryon number density influences the equation of state (EOS) of neutron star matter. An EOS with a pronounced peak in the speed of sound was found to leave a clear and unique signature in the main frequency of the postmerger gravitational wave (GW) spectrum [44], which can sensitively probe the hadron-quark phase transition in the dense core. However, the polytropic index, defined as γlnP/lnε, is potentially important for determining the presence of quark matter in the core of massive neutron stars [45-48]. The approximate rule that γ < 1.75 [45] or γ ≤ 1.6 and cs20.7 [49] can be used as a good criterion for separating quark matter from hadronic matter in a massive neutron star core. A recent study [50] introduced a new conformality criterion quantity dcΔ2+(Δ')2, where Δ(ε3P)/(3ε) and Δ'dΔ/d ln ε are the normalized trace anomaly and its logarithmic rate of change with respect to the energy density, respectively, and compared dc with the speed of sound, polytropic index, and normalized trace anomaly. The results indicate that the conformality criterion dc<0.2 is considerably more restrictive than the criterion γ <1.75 for quark matter in compact stars. Therefore, the speed of sound and polytropic index can serve as new probes for QCD CP or first-order phase transitions in future experimental exploration.

We utilized the significant findings of phase diagrams and critical points in the experimental and theoretical research mentioned above in recent years to explore the properties of QCD phase diagrams in the present work by calculating high-order moments, speed of sound, and polytropic index in an extended 3-flavor NJL model including an eight-quark scalar-vector interaction. Scalar-vector interactions are important for reproducing nuclear saturation properties when using the NJL-type model for nuclear matter [51]. When applied to the hadron-quark phase transition, the strength of the scalar-vector interactions can change the critical temperature Tc, thereby facilitating the study of the effect of CP in heavy-ion collisions [52, 53]. The remainder of this paper is organized as follows. In Sect. 2, we briefly review the calculation process of the extended 3-flavor NJL model, as well as the calculation formulas for the high-order moments, speed of sound, and polytropic index. The phase diagram of the high-order moments, speed of sound, and polytropic index, as well as the effects of the scalar-vector interaction on the phase diagram, are discussed in Sect. 3. Finally, the summary is presented in Sect. 4.

2

Theoretical model

The Lagrangian density of the extended three-flavor NJL model with an eight-quark scalar-vector interaction is given by [53] LNJLSU(3)=ψ¯(iγμμm^)ψ+GSa=08[(ψ¯λaψ)2+(ψ¯iγ5λaψ)2]K[detψ¯(1+γ5)ψ+detψ¯(1γ5)ψ]+GSVa=13[(ψ¯λaψ)2+(ψ¯iγ5λaψ)2]×a=13[(ψ¯λμλaψ)2+(ψ¯γ5λμλaψ)2], (1) where ψ=(u,d,s)T represents the three flavor quark fields and m^=diag(mu,md,ms) is the corresponding current quark mass matrix. In the above formula, λa(a=1,,8), where λ0=2/3 are the Gell-Mann matrices; GS and GSV are the strengths of the scalar coupling and scalar-vector coupling, respectively; and the K term is the six-quark interaction (also known as the Kobayashi-Maskawa-t’Hooft interaction) that breaks the axial UA(1) symmetry [54, 55]. This term gives rise to six-point interactions in the three flavors and is responsible for the flavor mixing effect.

In the mean-field approximation [56], quarks can be treated as quasiparticles whose constituent mass Mi is determined by the gap equation Mu=mu2GSϕu+2Kϕdϕs2GSV(ρu+ρd)2(ϕu+ϕd),Md=md2GSϕd+2Kϕuϕs2GSV(ρu+ρd)2(ϕu+ϕd),Ms=ms2GSϕs+2Kϕuϕd, (2) where ϕi=i¯i and ρi denote the quark condensate and net quark number densities, respectively. The thermodynamic properties of the three-flavor quark matter are determined by the partition function Z=Tr[exp[β(H^μN^)]]. Here, β =1/T and μ are the reciprocals of the temperature and chemical potential, respectively, while H^ and N^ represent the Hamiltonian and the corresponding conserved charge number operators, respectively. The thermodynamic potential of the system can be calculated as ΩNJLSU(3)=GS(ϕu2+ϕd2+ϕs2)+3GSV(ϕu+ϕd)2(ρu+ρd)24Kϕuϕdϕs2Nci=u,d,s0Λd3p(2π)3Ei2Ti=u,d,s0Λd3p(2π)3(z+(Ei)+z(Ei)), (3) where Nc =3 is the number of colors and z±(Ei)=Ncln[1+eβ(Eiμ˜i)], (4) with an effective chemical potential μ˜u=μu+2GSV(ρu+ρd)(ϕu+ϕd)2,μ˜d=μd+2GSV(ρu+ρd)(ϕu+ϕd)2,μ˜s=μs. (5) The energy Ei is expressed as Ei=Mi2+p2, where Mi is the mass of the constituent quarks. The quark condensate ϕi and net quark number density ρi can be derived by minimizing the thermodynamic potential as follows: ϕi=2Ncd3p(2π)3MiEi(fi+f¯i1),ρi=2Ncd3p(2π)3(fif¯i), (6) where fi=11+eβ(Eiμ˜i),f¯i=11+eβ(Ei+μ˜i). (7) In the present numerical calculation, we employ the following values for the parameters [57] as mu=md=5.5 MeV, ms=135.7 MeV, GSΛ2=3.67, KΛ5=9.29, and the cutoff value in the momentum integral Λ=631.4 MeV. For the coupling constant GSV, Eqs. (2) and (5) show that its value can affect the effective masses of the quarks and their effective chemical potentials. However, the entire term, including GSV does not affect the description of QCD vacuum properties at zero baryon density. Furthermore, treating the value of GSV as a free parameter allows us to obtain the different properties of quark matter.

In the following section, we consider the fluctuation moments of conserved quantities, such as the net baryon number, from the above 3-flavor NJL model. The experimentally measured proton multiplicity (net baryon number) distributions are described by the moments or ratios of the cumulants. The relationships between the cumulants Cn and moments are defined as C1=M, C2=σ2, C3=Sσ3, and C4=κσ4. Here, M and σ are the mean and quadratic variance, respectively, while the skewness S and kurtosis κ refer to higher-order moments. The ratios of the cumulants are often used to reduce the volume dependence, which can be read as C2/C1=σ2/M, C3/C2=Sσ, and C4/C2=κσ2. Moreover, the ratios of cumulants in the theoretical calculations are related to the ratios of the baryon-number susceptibilities. The nth-order susceptibility of the net baryon number in the grand ensemble can be expressed as the derivative of the thermodynamic potential [58] χnB=n(Ω/T)(μB/T)n. (8) The ratios of the baryon-number susceptibilities related to the skewness and kurtosis κσ2 measured experimentally in relativistic heavy-ion collisions can be calculated using [58] Sσ(B)=C3BC2B=χ3Bχ2B,κσ2(B)=C4BC2B=χ4Bχ2B. (9)

In heavy-ion collision experiments, the speed of sound cs is a crucial thermodynamic observable, which provides important information for describing the evolution of the fireball. Recently, the authors in Ref. [43] attempted to build a connection between cs and the cumulants of the baryon number distribution in matter created in heavy-ion collisions to aid in detecting QCD CP. The general definition of the speed of sound is cs2=P/ε, and its calculation requires specifying the thermodynamic variables that are kept constant. For the QGP created in relativistic heavy-ion collisions, it evolves with a constant entropy per baryon s/ρB, which justifies using the adiabatic speed of sound cs/ρB. In contrast, the isothermal speed of sound, cT, is widely used in neutron star matter. However, the values of cs/ρB2 and cT2 largely coincide near the critical point [59, 60], implying that the cumulants of the baryon number distribution in heavy-ion collisions can be used to estimate the isothermal squared speed of sound [43]. Therefore, in this study, we mainly discuss the phase diagram properties of the isothermal speed of sound. The pressure and entropy density in the NJL model can be derived using the thermodynamic relations in the grand canonical ensemble as follows: P=ΩNJL,s=ΩNJLT, (10) and the energy density can be calculated as ε=P+Ts+μBρB. (11) Using the Jacobi method and thermodynamic relations, cT2 can be written in terms of T and μB as cT2=(Pε)T=ρBT(sμB)T+μB(ρBμB)T. (12) Meanwhile, the polytropic index introduced in [45] γ ln P/ ln ε serves as a criterion for separating hadronic and quark matter [45, 47-49]. The isothermal polytropic index can be derived using the expression for the speed of sound: γT=(Pε)T/Pε=εPcT2. (13) The above expression shows that the speed of sound is an important intermediate quantity for calculating the polytropic index. The differences between these two quantities may lead to a polytropic index that includes more intricate features near the phase boundary, which could potentially provide a new probe for exploring QCD CP.

3

Results and Discussions

First, we discuss the higher-order susceptibilities of the net baryon number in the μB-T plane based on the 3-flavor NJL model with different scalar-vector coupling constants GSV. As shown in Fig. 1, we present the skewness (B) and kurtosis κσ2(B) of the baryon number fluctuations as functions of temperature and baryon chemical potential. The red and blue regions represent positive and negative values, respectively, and the green regions represent the values of (B) and κσ2(B) that are approximately zero. The black dashed-dotted and solid lines represent the chiral crossover and first-order phase transition, respectively, while the white dots connecting the chiral crossover and first-order phase transition represent the CPs. The chiral phase transition boundaries separate the red and blue regions for skewness, whereas the phase boundaries for kurtosis pass through the blue regions and divide the red regions with the CP located at the ends of the blue regions. In addition, the phase transition boundaries and CPs vary depending on the scalar-vector coupling constants GSV. Although decreasing the value of GSV effectively enhances the temperature of the CP, its baryon chemical potential decreases. The effects of GSV can be understood using Eq. (2) and (5). According to Eq. (5), the negative GSV resembles a vector interaction in the NJL model, which induces a repulsive interaction among the quarks. Meanwhile, compared with the scalar term GS in the NJL model, a negative GSV counteracts the repulsive interaction, as shown in Eq. (2), which reduces the constituent quark mass. For quark matter at a low baryon density (or chemical potential), the repulsive quark interaction with a negative GSV is stronger than the attractive quark interaction [53, 57], resulting in a net repulsive effect of the negative GSV. For quark matter at intermediate densities, because the quadratic of GSV term in Eq. (2) depends on the quark density, the effect of the GSV term on constituent quark masses is significantly enhanced with increasing quark density, which generates an effectively attractive interaction among quarks. Consequently, when the coupling constant GSV=200Λ8, the CP temperature increases to 104 MeV, while the chemical potential decreases to 750 MeV. For a positive GSV, the effects on the properties of quark matter are opposite to those of a negative GSV leading to a lower temperature and higher baryon chemical potential CP. Therefore, the range of the CP location obtained by varying GSV is sufficiently large to cover the region that can be probed during heavy-ion collisions. In quark matter at very high densities, where the chiral symmetry is largely restored and the quark condensates are close to zero, the effects of the GSV term become less important. This differs from the usual vector interaction in the NJL model that becomes stronger at high densities.

Fig. 1
(Color online) Skewness (B) and kurtosis κσ2(B) of the baryon number fluctuations as functions of the temperature and baryon chemical potential in the NJL model with different scalar-vector coupling constants GSV. The black dash-dotted and solid lines represent the chiral crossover and first-order phase transition, respectively, while the white dots connecting the chiral crossover and first-order phase transition represent the CPs
pic

In relativistic heavy-ion collision experiments, fluctuations in the net baryon number and net charge number were measured at the chemical freeze-out lines. However, the location of chemical freeze-out could not be accurately determined at the RHIC-BES energies. Several criteria exist for chemical freeze-out, such as fixed energy per particle at approximately 1 GeV, fixed total density of baryons and anti-baryons, fixed entropy density over T3, and the percolation model. Recently, the hypothetical freeze-out lines derived by fitting the experimental data have been expressed as T=abμB2cμB4 [2, 61]. However, the exact correspondence between the experimental parameters and location on the phase diagram remains unclear. Therefore, for simplicity and better elucidation of the properties of the phase diagram near the critical point, the hypothetical freeze-out lines shown in Fig. 2 were obtained by rescaling μB of the phase boundaries with factors of 0.90, 0.95, 0.98, 1.02, and 1.05, which can also be found in Refs. [16, 20, 62]. Figure 3 shows the skewness (B) and kurtosis κσ2(B) of the net baryon number fluctuations as functions of the baryon chemical potential along the hypothetical chemical freeze-out lines with different scalar-vector coupling constants GSV. Along the hypothetical chemical freeze-out lines, nonmonotonic dependencies in both (B) and κσ2(B) occur, and peak or dip structures appear around the critical points. In contrast, κσ2(B) features both a positive peak and negative dip, whereas (B) has either a single peak or dip, whose nature (peak or dip) depends on whether the hypothetical freeze-out lines are below the chiral phase transition boundaries. The amplitude of the peak or dip structures increases significantly as the hypothetical chemical freeze-out lines approach the CPs, for example, along the hypothetical lines with factors of 0.98 and 1.02. The nonmonotonic behavior of κσ2(B) is qualitatively consistent with the experimental results, while the experimental results for (B) seem to be positive above sNN=7.7 GeV from both STAR and PHENIX measurements [27]. Further experiments on collisions of different heavy ions at different energies are required, so that the freeze-out lines are close to or even above the phase transition boundaries. In addition, the peak and dip structures of the net-baryon susceptibilities shift to the low-baryon chemical potential (and high-temperature) side with a negative scalar-vector coupling constant GSV, which is consistent with our previous discussions and expectations. Furthermore, the amplitudes of variation for both (B) and κσ2(B) values significantly increase as GSV decreases, which may be because GSV is a quartic term in the thermodynamic potential. Thus, fluctuations in the density and quark condensation near the CP are greatly amplified.

Fig. 2
(Color online) Hypothetical chemical freeze-out lines with different scalar-vector coupling constants GSV by rescaling μB of the chiral phase transition boundary with factors of 0.90, 0.95, 0.98, 1.02, and 1.05
pic
Fig. 3
(Color online) Skewness (B) and kurtosis κσ2(B) of the net baryon number fluctuations as functions of the baryon chemical potential along the hypothetical chemical freeze-out lines with different scalar-vector coupling constants GSV
pic

The speed of sound, which is a fundamental property of a substance, provides important information for describing the evolution of strongly interacting matter and the final observables in heavy-ion collision experiments. It can be obtained from heavy-ion collision data through the charged particle multiplicity dNch/dη [40-42]. Furthermore, the net baryon fluctuations in heavy-ion collisions can be used to estimate the isothermal speed of sound and its logarithmic derivative [43]. The numerical results from effective models [59, 60, 67] suggest that the speed of sound at the CP is the global minimum of the full phase diagram and is also considered a sensitive probe of the hadron-quark phase transition in the compact star core. Fig. 4 presents the contour maps of the squared speed of sound cT2 in the μB-T plane based on the NJL model with different scalar-vector coupling constants GSV. The values of cT2 in the chiral breaking regions corresponding to the blue regions are mostly less than 0.2. For the different scalar-vector coupling constants GSV, changes in the range of chiral breaking (blue) regions and the location of the critical points are evident. After passing through the chiral phase transition boundary, cT2 rapidly increases with the restoration of chiral symmetry. In Fig. 4, a dip structure appears on the low-temperature side of the chiral phase transition boundary, and a minimum of cT2 appears around the critical point. For the different scalar-vector coupling constants, the value of cT2 is less than 1 in the entire μB-T plane, satisfying the constraint of causality. However, in the NJL model, cT2 does not approach the value cs2=1/3 of conformal matter limit at the high temperatures and baryon chemical potentials, because of the cutoff parameter set for the momentum integral in the NJL model. For the Polyakov-looped Nambu-Jona-Lasinio (PNJL) model, because momentum integration is finite and does not require an ultraviolet cutoff [11, 12], the squared speed of sound at high temperatures and baryon chemical potentials approaches 1/3 [60]. To further explore the properties of the speed of sound near the critical points, we present the results of the squared speed of sound cT2 along the hypothetical chemical freeze-out lines with different scalar-vector coupling constants GSV in Fig. 5. For the hypothetical freeze-out lines over the chiral phase transition boundaries, the cT2 curves rapidly increase after softening near the CPs. However, the speed of sound along the hypothetical freeze-out lines below the chiral phase transition boundary varies considerably: cT2 rapidly decreases near the CPs, especially as it approaches the local minimum along the curve by a factor of 0.98, followed by a spinodal behavior that eventually continues to decrease. Additionally, the spinodal regions are influenced by GSV. As GSV decreases, the span of the spinodal regions expands accordingly; conversely, it decreases or approaches zero with increasing GSV. This is also observed in Fig. 1, where the positive GSV reduces the amplitude of the fluctuations around the critical points.

Fig. 4
(Color online) Contour maps of the squared speed of sound cT2 in the μB-T plane based on the NJL model with different scalar-vector coupling constants GSV. The black dash-dotted and solid lines indicate the chiral crossover and first-order phase transition, respectively, while the white dots connecting the chiral crossover and first-order phase transition represent the critical points (CPs)
pic
Fig. 5
(Color online) Squared speed of sound cT2 as a function of baryon chemical potential along the hypothetical chemical freeze-out lines with different scalar-vector coupling constants GSV
pic

Subsequently, we investigated the contour maps of the polytropic index γT in the μB-T plane with different GSV values, as shown in Fig. 6. The values of γT are almost less than 1.5 in the chiral breaking regions with different GSV. On either side of the phase boundary, particularly around the first-order phase transition, non-monotonic behavior, including dip (blue region) and peak (red region) structures in γT are clearly observed. Following an isotherm that passes through a first-order phase transition, the polytropic index rapidly decreases to a minimum value, increases rapidly to the maximum with chiral restoration, and subsequently decreases again, eventually stabilizing at high temperatures and baryon chemical potentials. This can be explained by Eq. (13), where the polytropic index is equal to the squared speed of sound divided by P/ε. Around the first-order phase transition, the squared speed of sound (the derivative of P with respect to ε) and P/ε are not synchronized; therefore, a local minimum and maximum of the polytropic index appear at different chemical potentials. A more detailed explanation is provided in Ref. [60]. In Fig. 6, the amplitudes and spans of the dip and peak structures are influenced by the coupling constants GSV. The amplitude and span of the peak structures (red regions) are effectively enhanced by decreasing the value of GSV; however, those of the dip structures (blue regions) decrease. Similar to the speed of sound, we investigate the properties of γT along the hypothetical chemical freeze-out lines in Fig. 7. All curves of γT first decrease and attain a minimum near the CP, and closer critical points correspond to smaller minimum values, such as the hypothetical freeze-out lines with 0.98 or 1.02. Consistent with the results for (B) and κσ2(B), the mininum value of γT shifts to a low baryon chemical potential with a negative GSV. Further, a negative GSV leads to a larger minimum value of γT because the negative GSV, which serves as a repulsive interaction at a low chemical potential, stiffens the equation of stateand increases the speed of sound and polytropic index. Compared to cT2 as shown in Fig. 5, the polytropic index exhibits a more pronounced dip structure near the CP, which could provide a more sensitive probe for the QCD CP in future experiments.

Fig. 6
(Color online) Contour maps of the polytropic index γT in the μB-T plane based on the NJL model with different scalar-vector coupling constants GSV. The black dash-dotted and solid represent the chiral crossover and first-order phase transition, respectively, and the white dots connecting the chiral crossover and first-order phase transition represent the critical points (CPs)
pic
Fig. 7
(Color online) Polytropic index γT as a function of the baryon chemical potential along the hypothetical chemical freeze-out lines for different scalar-vector coupling constants GSV
pic
4

Summary and Outlook

In conclusion, we investigated the properties of the phase diagrams of higher-order susceptibilities, speed of sound, and polytropic index based on an extended NJL model with an eight-quark scalar-vector interaction. We found that all the aforementioned quantities exhibited non-monotonic behavior around the phase transition boundaries, which also revealed the location of the critical point. Specifically, the baryon susceptibility κσ2(B) featured both a positive peak and negative dip, whereas (B) had either a single peak or dip, whose nature depended on whether the hypothetical freeze-out lines were below the phase-transition boundaries. Meanwhile, for cT2, a dip structure appeared on the low-temperature side of the phase transition boundary, and a minimum appeared around the critical point. Along the hypothetical freeze-out lines below the phase transition boundary, the speed of sound rapidly decreased near the CP, followed by spinodal behavior, whereas the polytropic index γT exhibited a more pronounced dip as it approached the CP. Compared to the speed of sound, the polytropic index could provide a more sensitive probe for QCD CP in future experiments. In this study, we explored the effect of the eight-quark scalar-vector interaction on the phase diagram. The chiral phase transition boundary and critical point varied depending on the scalar-vector coupling constants GSV. At finite densities and temperatures, the negative GSV term exhibited attractive interactions, which enhanced the critical point temperature and reduced the chemical potential. The GSV term also affected the high-order susceptibilities, speed of sound, and polytropic index. The nonmonotonic (peak or dip) structures of these quantities shifted to a low baryon chemical potential (and high temperature) with a negative GSV. GSV also changed the amplitude and range of the nonmonotonic regions. Therefore, the scalar-vector interaction is useful for locating the phase boundary and critical point in the QCD phase diagram by comparing the experimental data. Although significant uncertainty remains regarding the properties of the QCD phase diagram and the precise location of the critical point, the study of the nonmonotonic behavior of these observables is of great interest. A more accurate BES-II measurement in the near future will provide further information regarding the QCD phase diagram. Moreover, in the study of massive neutron stars, the approximate rule γ < 1.75 [45] or γ ≤ 1.6 and cs20.7 may be used as a good criterion for separating quark matter from hadronic matter in a massive neutron star core [49]. Recently, a new conformality criterion dc<0.2, defined as comprising γ and cs, was introduced and compared with the speed of sound, polytropic index, and normalized trace anomaly [50]. The results indicate that the conformality criterion dc<0.2 is considerably more restrictive than the criterion γ <1.75 for quark matter in compact stars [50]. In recent years, machine learning (ML)and high-energy nuclear physics have begun to merge, yielding interesting results [67]. ML methods have also been applied to phase transition studies, enabling finer characterization of the QCD phase diagram and deeper understanding of quark matter properties [68]. Therefore, it is worth looking forward to new methods and observations related to high-order susceptibilities, speed of sound, and polytropic index and their application to the search for critical points in heavy-ion collisions and the study of compact stars.

References
1. P. Braun-Munzinger, V. Koch, T. Schäfer et al.,

Properties of hot and dense matter from relativistic heavy ion collisions

. Phys. Rep. 621, 76 (2016). https://doi.org/10.1016/j.physrep.2015.12.003
Baidu ScholarGoogle Scholar
2. X. Luo and 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
3. Y. Zhang, D.W. Zhang, X.F. Luo,

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

. Nucl. Tech. 46, 040001 (2023). https://doi.org/10.11889/j.0253-3219.2023.hjs.46.040001
Baidu ScholarGoogle Scholar
4. A. Bzdak, S. Esumi, V. Koch et al.,

Mapping the phases of quantum chromodynamics with beam energy scan

. Phys. Rep. 853, 187 (2020). https://doi.org/10.1016/j.physrep.2020.01.005
Baidu ScholarGoogle Scholar
5. C. Bernard, T. Burch, E. B. Gregory et al.,

QCD thermodynamics with three flavors of improved staggered quarks

. Phys. Rev. D 71 034504 (2005). https://doi.org/10.1103/PhysRevD.71.034504
Baidu ScholarGoogle Scholar
6. Y. Aoki, G. Endrődi, Z. Fodor et al.,

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

. Nature 443 675 (2006). https://doi.org/10.1038/nature05120
Baidu ScholarGoogle Scholar
7. A. Bazavov, T. Bhattacharya, M. Cheng et al.,

Chiral and deconfinement aspects of the QCD transition

. Phys. Rev. D 85 054503 (2012). https://doi.org/10.1103/PhysRevD.85.054503
Baidu ScholarGoogle Scholar
8. T.M. Schwarz, S.P. Klevansky, and G. Papp,

Phase diagram and bulk thermodynamical quantities in the Nambu-Jona-Lasinio model at finite temperature and density

. Phys. Rev. C 60, 055205 (1999). https://doi.org/10.1103/PhysRevC.60.055205
Baidu ScholarGoogle Scholar
9. P. Zhuang, M. Huang, and Z. Yang,

Density effect on hadronization of a quark plasma

. Phys. Rev. C 62, 054901 (2000). https://doi.org/10.1103/PhysRevC.62.054901
Baidu ScholarGoogle Scholar
10. W.J. Fu, Z. Zhang, and 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
11. K. Fukushima,

Phase diagrams in the three-flavor Nambu–Jona-Lasinio model with the Polyakov loop

. Phys. Rev. D 77, 114028 (2008). https://doi.org/10.1103/PhysRevD.77.114028
Baidu ScholarGoogle Scholar
12. K. Fukushima, Phys. Rev. D 78, 039902 (2008). https://doi.org/10.1103/PhysRevD.77.114028
13. 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
14. X.Y. Xin, S.X. Qin, Y.X. Liu et al.,

Quark number fluctuations at finite temperature and finite chemical potential via the Dyson-Schwinger equation approach

. Phys. Rev. D 90, 076006 (2014). https://doi.org/10.1103/PhysRevD.90.076006
Baidu ScholarGoogle Scholar
15. F. Gao, Y.X. Liu.

QCD phase transitions using the QCD Dyson-Schwinger equation Nucl

. Tech. 46, 040015 (2023). https://doi.org/10.11889/j.0253-3219.2023.hjs.46.040015
Baidu ScholarGoogle Scholar
16. J.W. Chen, J. Deng, H. Kohyama et al.,

Robust characteristics of non-Gaussian fluctuations from the NJL model

. Phys. Rev. D 93, 034037 (2016). https://doi.org/10.1103/PhysRevD.93.034037
Baidu ScholarGoogle Scholar
17. H. Liu, J. Xu, L.W. Chen et al.,

Isospin properties of quark matter from a 3-flavor NJL model

. Phys. Rev. D 94, 065032 (2016). https://doi.org/10.1103/PhysRevD.94.065032
Baidu ScholarGoogle Scholar
18. W.J. Fu, J.M. Pawlowski, and 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
19. S. Yin, Y.Y Tan, W.J. Fu,

Critical phenomena and functional renormalization group

. Nucl. Tech. 46, 040002 (2023). https://doi.org/10.11889/j.0253-3219.2023.hjs.46.040002
Baidu ScholarGoogle Scholar
20. H. Liu and J. Xu,

Isospin effect on baryon and charge fluctuations from the pNJL model

. Universe 7, 6 (2021). https://doi.org/10.3390/universe7010006
Baidu ScholarGoogle Scholar
21. V. Vovchenko, J. Steinheimer, O. Philipsen et al.,

Cluster expansion model for QCD baryon number fluctuations: No phase transition at μB/π<T

. Phys. Rev. D 97 (2018) 114030. https://doi.org/10.1103/PhysRevD.97.114030
Baidu ScholarGoogle Scholar
22. M. M. Aggarwal 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
23. L. Adamczyk et al., (STAR Collaboration),

Beam Energy Dependence of Moments of the Net-Charge Multiplicity Distributions in Au + Au Collisions at RHIC

. Phys. Rev. Lett. 112, 032302 (2014). https://doi.org/10.1103/PhysRevLett.113.092301
Baidu ScholarGoogle Scholar
24. M. Abdallah et al., (STAR Collaboration),

Measurements of Proton High-Order Cumulants in 3 GeV Au + Au Collisions and Implications for the QCD Critical Point

. Phys. Rev. Lett. 128, 202303 (2022). https://doi.org/10.1103/PhysRevLett.128.202303
Baidu ScholarGoogle Scholar
25. L. Adamczyk et al., (STAR Collaboration),

Beam-energy-dependent two-pion interferometry and the freeze-out eccentricity of pions measured in heavy ion collisions at the STAR detector

. Phys. Rev. C 92, 014904 (2015). https://doi.org/10.1103/PhysRevC.92.014904
Baidu ScholarGoogle Scholar
26. J. Adam et al., (STAR Collaboration),

Flow and interferometry results from Au + Au collisions at 4.5 GeV

. Phys. Rev. C 103, 034908 (2021). https://doi.org/10.1103/PhysRevC.103.034908
Baidu ScholarGoogle Scholar
27. L. Adamczyk et al., (STAR Collaboration),

Beam-energy dependence of the directed flow of protons, antiprotons, and pions in Au + Au collisions

. Phys. Rev. Lett. 112, 162301 (2014). https://doi.org/10.1103/PhysRevLett.112.162301
Baidu ScholarGoogle Scholar
28. L. Adamczyk et al., (STAR Collaboration),

Beam-Energy Dependence of Directed Flow of Λ, Λ¯, K±, Ks0, and ϕ in Collisions

. Phys. Rev. Lett. 120, 062301 (2018). https://doi.org/10.1103/PhysRevLett.120.062301
Baidu ScholarGoogle Scholar
29. J. Adam et al., (STAR Collaboration),

Observation Ds±/D0 of Enhancement in Au + Au Collisions at =200 GeV

. Phys. Rev. Lett. 126, 092301 (2021). https://doi.org/10.1103/PhysRevLett.127.092301
Baidu ScholarGoogle Scholar
30. M. Abdallah et al., (STAR Collaboration),

Cumulants and correlation functions of net-proton, proton, and antiproton multiplicity distributions in Au +Au collisions at energies available at the BNL Relativistic Heavy Ion Collider

. Phys. Rev. C 104, 024902 (2021). https://doi.org/10.1103/PhysRevC.104.024902
Baidu ScholarGoogle Scholar
31. K.J. Sun, L.W. Chen, C.M. Ko et al.,

Light nuclei production as a probe of the QCD phase diagram

. Phys. Lett. B 781, 499 (2018). https://doi.org/10.1016/j.physletb.2018.04.035
Baidu ScholarGoogle Scholar
32. K.J. Sun, L.W. Chen, C.M. Ko et al.,

Light nuclei production and QCD phase transition in heavy-ion collisions

. Nucl. Tech. 46, 040012 (2023). https://doi.org/10.11889/j.0253-3219.2023.hjs.46.040012
Baidu ScholarGoogle Scholar
33. M.I. Abdulhamid et al., (STAR Collaboration),

Beam Energy Dependence of Triton Production and Yield Ratio (Nt×Np/Nd2) in Au + AuCollisions at RHIC

. Phys. Rev. Lett. 130, 202301 (2023). https://doi.org/10.1103/PhysRevLett.130.202301
Baidu ScholarGoogle Scholar
34. M.I. Abdulhamid et al., (STAR Collaboration),

Energy dependence of intermittency for charged hadrons in Au+ Au collisions at RHIC

. Phys. Lett. B 845, 138165 (2023). https://doi.org/10.1016/j.physletb.2023.138165
Baidu ScholarGoogle Scholar
35. M. Asakawa, U. Heinz, and B. Müller,

Fluctuation probes of quark deconfinement

. Phys. Rev. Lett. 85, 2072 (2000). https://doi.org/10.1103/PhysRevLett.85.2072
Baidu ScholarGoogle Scholar
36. S. Jeon and V. Koch,

Event-by-event fluctuations

. Quark–Gluon Plasma 3, 430 (2004). https://doi.org/10.48550/arXiv.hep-ph/0304012
Baidu ScholarGoogle Scholar
37. H.C. Song, S.A. Bass, U. Heinz et al.,

200 A GeV Au + Au Collisions Serve a Nearly Perfect Quark-Gluon Liquid

. Phys. Rev. Lett. 106, 192301 (2011). https://doi.org/10.1103/PhysRevLett.106.192301
Baidu ScholarGoogle Scholar
38. H.C. Song, S.A. Bass, and U. Heinz,

Viscous QCD matter in a hybrid hydrodynamic + Boltzmann approach

. Phys. Rev. C 83, 024912 (2011). https://doi.org/10.1103/PhysRevC.83.024912
Baidu ScholarGoogle Scholar
39. P. Deb, G. P. Kadam, and H. Mishra,

Estimating transport coefficients in hot and dense quark matter

. Phys. Rev. D 94, 094002 (2016). https://doi.org/10.1103/PhysRevD.94.094002
Baidu ScholarGoogle Scholar
40. R. Campanini and G. Ferri,

Experimental equation of state in pp and pp collisions and phase transition to quark gluon plasma

. Phys. Lett. B 703, 237 (2011). https://doi.org/10.1016/j.physletb.2011.08.009
Baidu ScholarGoogle Scholar
41. F.G. Gardim, G. Giacalone, M. Luzum et al.,

Revealing QCD thermodynamics in ultrarelativistic nuclear collisions

. Nat. Phys. 16, 615 (2020). https://doi.org/10.1038/s41567-020-0846-4
Baidu ScholarGoogle Scholar
42. D. Sahu, S. Tripathy, R. Sahoo et al.,

Multiplicity dependence of shear viscosity, isothermal compressibility and speed of sound in pp collisions at 7 TeV

. Eur. Phys. J. A 56, 187 (2020). https://doi.org/10.1140/epja/s10050-020-00197-7
Baidu ScholarGoogle Scholar
43. A. Sorensen, D. Oliinychenko, V. Koch et al.,

Speed of sound and baryon cumulants in heavy-ion collisions

. Phys. Rev. Lett. 127, 042303 (2021). https://doi.org/10.1103/PhysRevLett.127.042303
Baidu ScholarGoogle Scholar
44. Y.J. Huang, L. Baiotti, T. Kojo et al.,

Merger and Postmerger of Binary Neutron Stars with a Quark-Hadron Crossover Equation of State

. Phys. Rev. Lett. 129, 181101 (2022). https://doi.org/10.1103/PhysRevLett.129.181101
Baidu ScholarGoogle Scholar
45. E. Annala, T. Gorda, A. Kurkela et al.,

Evidence for quark-matter cores in massive neutron stars

. Nat. Phys. 16, 907 (2020). https://doi.org/10.1038/s41567-020-0914-9
Baidu ScholarGoogle Scholar
46. H. Liu, J. Xu, and P. C. Chu,

Symmetry energy effects on the properties of hybrid stars

. Phy. Rev. D 105, 043015 (2022). https://doi.org/10.1103/PhysRevD.105.043015
Baidu ScholarGoogle Scholar
47. H. Liu, X.M. Zhang, and P.C. Chu,

Properties of quark-matter cores in massive hybrid stars

. Phys. Rev. D 107, 094032 (2023). https://doi.org/10.1103/PhysRevD.107.094032
Baidu ScholarGoogle Scholar
48. H. Liu, Y.H. Yang, Y. Han et al.,

Properties of quark matter and hybrid stars from a quasiparticle model

. Phys. Rev. D 108, 034004 (2023). https://doi.org/10.1103/PhysRevD.108.034004
Baidu ScholarGoogle Scholar
49. M.Z. Han, Y.J. Huang, S.P. Tang et al.,

Plausible presence of quark matter in neutron stars with masses above 0.97MTOV

. Sci. Bull. 68, 913 (2023). https://doi.org/10.1016/j.scib.2023.04.007
Baidu ScholarGoogle Scholar
50. E. Annala, T. Gorda, J. Hirvonen et al.,

Strongly interacting matter exhibits deconfined behavior in massive neutron stars

. Nat. Commun. 14, 8451 (2023). https://doi.org/10.1038/s41467-023-44051-y
Baidu ScholarGoogle Scholar
51. V. Koch, T.S. Biro, J. Kunz et al.,

A chirally invariant fermionic field theory for nuclear matter

. Phys. Lett. B 185, 15 (1987). https://doi.org/10.1016/0370-2693(87)91517-6
Baidu ScholarGoogle Scholar
52. T.G. Lee, Y. Tsue, J. da Providencia et al.,

Quark-Hadron Phase-Transition in an Extended NJL Model with Scalar-Vector Interaction

. Prog. Theor. Exp. Phys. 2013, 013D02 (2013). https://doi.org/10.1093/ptep/pts055
Baidu ScholarGoogle Scholar
53. K.J. Sun, C.M. Ko, S. Cao et al.,

QCD critical point from the Nambu-Jona-Lasino model with a scalar-vector interaction

. Phys. Rev. D 103, 014006 (2021). https://doi.org/10.1103/PhysRevD.103.014006
Baidu ScholarGoogle Scholar
54. G. t’Hooft,

Computation of the quantum effects due to a four-dimensional pseudo-particle

. Phys. Rev. D 14, 3432 (1976). https://doi.org/10.1103/PhysRevD.14.3432
Baidu ScholarGoogle Scholar
55. G. t’Hooft,

Erratum: Computation of the quantum effects due to a four-dimensional pseudoparticle

. Phys. Rev. D 18, 2199 (1978). https://doi.org/10.1103/PhysRevD.14.3432
Baidu ScholarGoogle Scholar
56. U. Vogl and W. Weise,

The Nambu and Jona-Lasinio model: Its implications for Hadrons and Nuclei

. Prog. Part. Nucl. Phys. 27, 195 (1991). https://doi.org/10.1016/0146-6410(91)90005-9
Baidu ScholarGoogle Scholar
57. M. Buballa,

NJL-model analysis of dense quark matter

. Phys. Rept. 407, 205 (2005). https://doi.org/10.1016/j.physrep.2004.11.004
Baidu ScholarGoogle Scholar
58. M. Asakawa and M. Kitazawa,

Fluctuations of conserved charges in relativistic heavy ion collisions: An introduction

. Prog. Part. Nucl. Phys. 90, 299 (2016). https://doi.org/10.1016/j.ppnp.2016.04.002
Baidu ScholarGoogle Scholar
59. W.B. He, G.Y. Shao, X.Y. Gao et al.,

Speed of sound in QCD matter

. Phys. Rev. D 105, 094024 (2022). https://doi.org/10.1103/PhysRevD.105.094024
Baidu ScholarGoogle Scholar
60. H. Liu, Y.H. Yang, C. Yuan et al.,

Speed of sound and polytropic index in the Polyakov-loop Nambu-Jona-Lasinio model

. Phys. Rev. D 109, 074037 (2024). https://doi.org/10.1103/PhysRevD.109.074037
Baidu ScholarGoogle Scholar
61. V.V. Begun, V. Vovchenko, and M.I. Gorenstein,

Updates to the p + p and A + A chemical freeze-out lines from the new experimental data

. J. Phys. Conf. Ser. 779, 012080 (2017). https://doi.org/10.1088/1742-6596/779/1/012080
Baidu ScholarGoogle Scholar
62. J.W. Chen, J. Deng, H. Kohyama et al.,

Universal relations between non-Gaussian fluctuations in heavy-ion collisions

. Phys. Rev. D 95, 014038 (2017). https://doi.org/10.1103/PhysRevD.95.014038
Baidu ScholarGoogle Scholar
63. W.B. He, G.Y. Shao, and 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
64. S.K. Ghosh, T.K. Mukherjee, M.G. Mustafa et al.,

Susceptibilities and speed of sound from the Polyakov-Nambu-Jona-Lasinio model

. Phys. Rev. D 73, 114007 (2006). https://doi.org/10.1103/PhysRevD.73.114007
Baidu ScholarGoogle Scholar
65. R. Marty, E. Bratkovskaya, W. Cassing et al.,

Transport coefficients from the Nambu-Jona-Lasinio model for SU(3)f

. Phys. Rev. C 88, 045204 (2013). https://doi.org/10.1103/PhysRevC.88.045204
Baidu ScholarGoogle Scholar
66. K. Saha, S. Ghosh, S. Upadhaya et al.,

Transport coefficients in a finite volume Polyakov-Nambu-Jona-Lasinio model

. Phys. Rev. D 97, 116020 (2018). https://doi.org/10.1103/PhysRevD.97.116020
Baidu ScholarGoogle Scholar
67. W.B. He, Y.G. Ma, L.G. Pang et al.,

High‐energy nuclear physics meets machine learning

. Nucl. Sci. Tech. 34, 88 (2023). https://doi.org/10.1007/s41365-023-01233-z
Baidu ScholarGoogle Scholar
68. 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/10.1088/0256-307X/40/12/122101
Baidu ScholarGoogle Scholar
Footnote

The authors declare that they have no competing interests.