logo

Properties of collective flow and pion production in intermediate-energy heavy-ion collisions with a relativistic quantum molecular dynamics model

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

Properties of collective flow and pion production in intermediate-energy heavy-ion collisions with a relativistic quantum molecular dynamics model

Si-Na Wei
Zhao-Qing Feng
Nuclear Science and TechniquesVol.35, No.1Article number 15Published in print Jan 2024Available online 01 Feb 2024
390010

The relativistic mean-field approach was implemented in the Lanzhou quantum molecular dynamics transport model (LQMD.RMF). Using the LQMD.RMF, the properties of collective flow and pion production were investigated systematically for nuclear reactions with various isospin asymmetries. The directed and elliptic flows of the LQMD.RMF are able to describe the experimental data of STAR Collaboration. The directed flow difference between free neutrons and protons was associated with the stiffness of the symmetry energy, that is, a softer symmetry energy led to a larger flow difference. For various collision energies, the ratio between the π- and π+ yields increased with a decrease in the slope parameter of the symmetry energy. When the collision energy was 270 MeV/nucleon, the single ratio of the pion transverse momentum spectra also increased with decreasing slope parameter of the symmetry energy in both nearly symmetric and neutron-rich systems. However, it is difficult to constrain the stiffness of the symmetry energy with the double ratio because of the lack of threshold energy correction on the pion production.

Heavy-ion collisionCollective flowPion productionSymmetry energyRelativistic mean field
1

Introduction

The equation of state (EOS) of nuclear matter, which originates from the interactions between nuclear matter, plays an important role in nuclear physics and astrophysics. Heavy-ion collisions, the properties of nuclei, and neutron stars (NSs) have been widely studied to extract the nuclear EOS. Because nuclear many-body problems are highly nonlinear and the EOS is not a directly observable quantity in experiments, there are still some uncertainties in the EOS despite great efforts [1-6]. For instance, the EOS extracted from the GW170817 event has uncertainties at high nuclear densities [2]. Although the EOS can be extracted from the properties of NSs, the internal composition of NSs is still poorly understood. The core of an NS may contain exotic materials such as hyperons, kaons, pions, and deconfined quark matter [7-12]. Heavy-ion collisions in terrestrial laboratories provide a unique opportunity to study both the EOS and exotic materials.

Collective flows of heavy-ion collisions were proposed in the 1970s and first detected in experiments at Bevalac [13-16]. Because collective flows are associated with nucleon-nucleon interactions, nucleon-nucleon scattering, etc., collective flows have been used to extract the nuclear EOS [1]. Collective flows are also helpful for understanding the phase transition between hadronic and quark matter. Generally, when a phase transition between hadronic and quark matter occurs, the collective flows of heavy-ion collisions indicate a soft EOS [17-21]. In addition, the ratios of the isospin particles in heavy-ion collisions, such as π-+, K0/K+, and ∑-/∑+, are thought to be sensitive to the stiffness of the EOS [22-28]. The production of pions and kaons has been experimentally measured in 197Au+197Au collisions. The K+ production predicted by various transport models favors a soft EOS of isospin-symmetric nuclear matter at high baryon densities [29-33]. The π-+ ratio predicted by various transport models is still model-dependent [34-37]. Based on the FOPI data for the π-+ ratio [38], some results favor a stiff symmetry energy (isospin asymmetric part of the EOS) [34, 35], whereas others imply a soft symmetry energy [36, 37]. Recently, by analyzing the ratios of charged pions in 132Sn + 124Sn, 112Sn +124Sn, and 108Sn + 112Sn collisions [39], a slope of the symmetry energy ranging from 42 to 117 MeV was suggested [40, 41]. The collective flows and ratios of charged pions are still worth studying to find the sources of the difference in various transport models and to extract information about the EOS from heavy-ion collisions.

Quantum molecular dynamics (QMD) is a popular transport model that has been developed into many versions and used to successfully describe the properties of nuclear matter, nuclei, mesons, and hyperons [33, 34, 42-55]. In high-energy heavy-ion collisions, the relativistic effects should be considered in QMD because they become significant. The RQMD approach has been proposed for this purpose [42, 43]. Recently, relativistic mean field theory (RMF) was implemented in RQMD (RQMD.RMF) [44-46]. The RQMD.RMF has been successfully applied to investigate the collective flows of hadrons [44-46]. In this study, we implemented RMF theory with isovector-vector and isovector-scalar fields in the Lanzhou quantum molecular dynamics model (LQMD.RMF). The channels for the generation and decay of resonances (Δ(1232), N*(1440), N*(1535), etc.), hyperons, and mesons were included in the previous LQMD model [33, 34, 47-50]. Using the LQMD.RMF, we explored the relationship between the symmetry energy and the properties of the collective flow and pion production.

The remainder of this paper is organized as follows. In Sect. 2, we briefly introduce the formulas and approaches used in this study. The formulas include RMF theory, the dispersion relation, and the production of pions. The results and discussion are presented in Sect. 3. Finally, a summary is presented in Sect. 4.

2

Formalism

2.1
Relativistic mean field approach

The RMF interaction is achieved by exchanging mesons. Scalar and vector mesons provide medium-range attraction and short-range repulsion between nucleons, respectively [56]. The nonlinear self-interaction of the σ meson is introduced to reduce the incompressibility to a reasonable domain [57]. To investigate the properties of symmetry energy, we also consider the isovector-vector ρ [58] and isovector-scalar δ mesons [59]. The Lagrangian density is expressed as [59, 60] L=ψ¯[γμ(iμgωωμgρτbμ)(MNgσσgδτδ)]ψ+12(μσμσmσ2σ2)13g2σ314g3σ4+12mω2ωμωμ14FμνFμν+12mρ2bμbμ14BμνBμν+12(μδμδmδ2δ2), (1) where MN=938 MeV is the nucleon mass in free space, gi with i=σ,ω,ρ,δ is the coupling constant between nucleons, mi with i=σ,ω,ρ,δ denotes the meson mass, g2 and g3 are the coupling constants for the nonlinear self-interaction of the σ meson, and Fμν=μωννωμ and Bμν=μbννbμ are the strength tensors of the ω and ρ mesons, respectively. The equations of motion for the nucleon and meson are obtained from the Euler–Lagrange equations and are written as: [iγμμgωγ0ω0gργ0b0τ3    (MNgσσgδτ3δ3)]ψ=0, (2) mσ2σ+g2σ2+g3σ3=gσψ¯ψ=gσρS, (3) mω2ω0=gωψ¯γ0ψ=gωρ, (4) mρ2b0=gρψ¯γ0τ3ψ=gρρ3, (5) mδ2δ3=gδψ¯τ3ψ=gδρS3, (6) where ρ and ρS are the baryon and scalar densities, respectively, ρ3=ρpρn is the difference between the proton and neutron densities, and ρS3=ρSpρSn is the difference between the proton and neutron scalar densities.

In the RMF approximation, the energy density is given by ϵ=i=n,p20pFd3p(2π)3p2+Mi*2+12mσ2σ2+13g2σ3+14g3σ4+12mω2ω02+12mρ2b02+12mδ2δ32, (7) where pF denotes the nucleon Fermi momentum, and Mi*=MNgσσgδδ3 (- proton, + neutron) denotes the effective nucleon mass. The isospin splitting of the effective nucleon mass Mn*Mp* still has a large uncertainty at this point. Analyses of nucleon-nucleus scattering data based on the optical model favor Mn*Mp*>0 [61, 62]. Calculations based on Brueckner theory [63-65] and density-dependent relativistic Hartree-Fock [66] also indicate Mn*Mp*>0. However, Mn*Mp*<0 is predicted by the transport model for heavy-ion collisions [67, 68] and nonlinear RMF models [59, 60, 69]. In addition, both Mn*Mp*<0 and Mn*Mp*>0 can be predicted by the point-coupling RMF [69] and Skyrme and Gogny forces [70-76]. Because the Lagrangian density in this study is the same as that in Ref. [59] and [60] except for the parameter settings, as shown in Fig. 1, the negative isospin splitting of the effective nucleon mass Mn*Mp*<0 is consistent with that in Ref. [59] and [60]. In the nonlinear RMF model, the isospin splitting of the effective nucleon mass is primarily determined by the coupling strength of the δ meson. The large coupling strength of the δ meson results in large isospin splitting of the effective nucleon mass. When the coupling strength of the δ meson is zero (set1), there is no isospin splitting of the effective nucleon mass.

Fig. 1
(Color online) Difference between neutron and proton effective masses as a function of the baryon density
pic

Using the isospin asymmetry parameter α=(ρnρp)/(ρn+ρp), the symmetry energy can be written as [59] Esym=122E(ρ,α)α2|α=0=122[ϵ(ρ,α)/ρ]α2|α=0=16pF2EF*+12fρρfδ2M*2ρEF*2[1+fδA(pF,M*)], (8) where figi2mi2, i=ρ,δ, and EF*=pF2+M*2 and M*=MNgσσ are the effective nucleon masses of the symmetric nuclear matter. The integral A(pF,M*) is defined as A(pF,M*)=4(2π)3d3pp2(p2+M*2)3/2=3(ρSM*ρEF*). (9) Based on the symmetry energy Esym, the slope L and curvature Ksym of the symmetry energy at the saturation density ρ0 can be obtained as L=3ρ0(Esymρ)|ρ=ρ0Ksym=9ρ02(2Esymρ2)|ρ=ρ0. (10) In this study, we set the saturation density to ρ0=0.16fm3, and the binding energy per particle of the symmetric nuclear matter was set to E/A-MN=-16 MeV. For symmetric nuclear matter, we set set1, set2, and set3 models to be the same as the result of vanishing isospin asymmetry. As shown in Table 1 and Fig. 2, the symmetry energy of set1, set2, and set3 was set to be 31.6 MeV at the saturation density. Set1 contained only ρ mesons; however, set2 and set3 contained both the ρ and δ mesons. For set1, when the symmetry energy was set to be 31.6 MeV at the saturation density, the coupling parameter was fixed and the slope of symmetry was fixed at L=85.3 MeV. For set2 and set3, when the symmetry energy was fixed at 31.6 MeV, the slope of the symmetry energy was obtained by varying the coupling parameters and . Because the effective mass M* of the above models for symmetric nuclear matter was the same, the symmetry energy with both the ρ and δ mesons could not be softer than that of set1 containing only ρ mesons. To broaden the range of the slope parameters, we set the slope parameter of set2 and set3 to be 109.3 and 145.0 MeV by varying the coupling parameters and , respectively. A broader range of slope parameters would be helpful for understanding the relationship between the properties of symmetry energy and the observables of heavy-ion collisions. The curvature of the symmetry energy Ksym, which is a higher-order expansion coefficient of the symmetry energy compared to the slope parameter L, may also affect the properties of the symmetry energy and the observables of heavy-ion collisions at densities far beyond the saturation density. Ksym of set1, set2, and set3 is obtained as -15, 141, and 391 MeV, respectively.

Table 1
Parameter sets for RMF
Model gσ g2 (fm-1) g3 K (MeV) Esym(ρ0) (MeV) L (ρ0) (MeV) M*(ρ0)/MN Ksym (MeV)
Set1 8.145 7.570 31.820 28.100 4.049 - 230 31.6 85.3 0.81 -15
Set2 8.145 7.570 31.820 28.100 8.673 5.347 230 31.6 109.3 0.81 141
Set3 8.145 7.570 31.820 28.100 11.768 7.752 230 31.6 145.0 0.81 391
Show more
The saturation density ρ0 is set to 0.16 fm-3. The binding energy of the saturation density is E/A-MN=-16 MeV. The isoscalar-vector ω and isovector-vector ρ masses are fixed at their physical values, = 783 MeV and = 763 MeV, respectively. The remaining meson masses, mσ and , are set to be 550 and 500 MeV, respectively
Fig. 2
(Color online) Symmetry energy as a function of the baryon density
pic
2.2
Relativistic quantum molecular dynamics approach

To investigate high-energy heavy-ion collisions, RQMD was proposed [42, 43]. Recently, RMF was implemented in RQMD [44-46]. In RQMD, for an N-body system, there are 4N position coordinates qiμ and 4N momentum coordinates piμ (i=1,...,N). However, the physical trajectories (qi and pi) are 6N for an N-body system. 2N constraints are required to reduce the number of dimensions from 8N to physical trajectories 6N [42-46, 77-79]. ϕi0(i=1,...,2N), (11) where 2N constraints satisfy the physical 6N phase space. The sign indicates Dirac’s weak equality. The on-mass shell conditions can reduce the phase space from 8N to 7N dimensions. ϕipi*2Mi*2=(piVi)2(MNSi)2=0, (12) where i=1,...,N. The remaining N constraints are time fixation constraints. A simple choice of time fixation constraints that obey the worldline condition is written as [43, 77, 79, 80] ϕi+Na^(qiqN)=0,(i=1,...,N1),ϕ2Na^qNτ=0, (13) where a^=(1,0) denotes a four-dimensional unit-vector [42-46, 77]. In a two-body center-of-mass system, a^ is defined as pijμ/pij2 with pijμ=piμ+pjμ. We observe that only the constraint i=2N depends on τ. With the above 2N constraints, the number of dimensions from 8N will reduce to 6N. These 2N constraints are conserved over time. dϕidτ=ϕiτ+k2NCik1λk=0. (14) Because only the constraint i=2N depends on τ, λ is written as[77] λi=C2N,iϕ2Nτ,(i=1,...,2N1), (15) with Cij1=[ϕi,ϕj]. Following previous studies, the Hamiltonian of the N-body system was constructed as a linear combination of 2N-1 constraints [77, 79, 80]: H=i=12N1λi(τ)ϕi, (16) Assuming [ϕi,ϕj]=0, λi=0 for N+1<i<2N [77]. The equations of motion are then obtained as dqidτ=[H,qi]=jNλjϕjpi,dpidτ=[H,pi]=jNλjϕjqi, (17) with the on-mass shell conditions (Eq. (12)) as inputs, the equations of motion can be obtained as r˙i=pi*pi*0+j=1N(Mj*pj*0Mj*pi+zj*μVjμpi),p˙i=j=1N(Mj*pj*0Mj*ri+zj*μVjμri), (18) where zi*μ=pi*μ/pi*0 and Mi*=MNSi. The scalar potential Si and vector potential Viμ in RQMD are defined as Si=gσσi+gδtiδi,Vi,μ=Bigωωi,μ+Bitigρbi,μ, (19) where ti=1 for protons, and ti=-1 for neutrons, and Bi is the baryon number of the ith particle. The meson field can be obtained from the RMF: mσ2σi+g2σi2+g3σi3=gσρS,i,mω2ωiμ=gωJiμ,mδ2δi=gδ(ρSp,iρSn,i)=gδρS3,i,mρ2biμ=gρ(ρp,iρn,i)=gρRiμ. (20) Substituting Eq. (20) into Eq. (19), the scalar potential Si and vector potential Viμ of the nucleons can be obtained. For other hadrons, such as Δ resonances, similar to other transport models [34, 81], the Δ optical potential is estimated using the nucleon potentials and square of the Clebsch–Gordan coefficient. In the RQMD approach, the scalar density, isovector-scalar density, baryon current, and isovetor baryon current are expressed as ρS,i=jiMjpj0ρij,ρS3,i=jitjMjpj0ρij,Jiμ=jiBjpjμpj0ρij,Riμ=jitjBjpjμpj0ρij. (21) Because the difference between the numerical results obtained using the effective mass Mj* and kinetic momentum pjμ* in the density and current and those obtained using a free mass Mj=MN=938 MeV and canonical momentum pjμ in the density and current is small, a free mass Mj=MN=938 MeV and canonical momentum pjμ have been used in the above density and current [45]. The interaction density ρij is given by a Gaussian ρij=γij4πLG±3/2exp(qT,ij24LG), (22) where qT,ij2=qij2[qij,σ(piσ+pjσ)(pi+pj)2]2 is a distance squared [77], and γij is a Lorentz factor that ensures the correct normalization of the Gaussian [82] and is equal to (pi0+pj0)/(pi+pj) in the two-body center-of-mass frame. In this study, we set the Gaussian width to LG=2.0 fm2.

2.3
Dispersion relation and production of pions

The Hamiltonian of the mesons is defined as [48, 83-85] HM=i=1NM[ViC+ω(pi,ρi)], (23) where ViC is the Coulomb potential, which is expressed as ViC=j=1NBeiejrij, (24) and NM and NB are the total numbers of mesons and baryons, including charged resonances, respectively. The pion potential in the medium, which contains isoscalar and isovector contributions, is defined as ω(pi,ρi)=ωisoscalar(pi,ρi)+Cπτzα(ρ/ρ0)γπ, (25) where α denotes the isospin asymmetry parameter, the coefficient is 36 MeV, the isospin quantity τ is 1, 0, and -1 for π-, π0, and π+, respectively, and γπ determines the isospin splitting of the pion potential and is set to two. In this study, the isoscalar part of pion potential ωisoscalar was chosen as the Δ-hole model. The pion potential, which contains a pion branch (smaller value) and Δ-hole (larger value) branch, is defined as ωisoscalar(pi,ρi)=Sπ(pi,ρi)ωπlike(pi,ρi)+SΔ(pi,ρi)ωΔlike(pi,ρi). (26) The probabilities of the pion and Δ-hole branches satisfy the following equation: Sπ(pi,ρi)+SΔ(pi,ρi)=1. (27) The probability of both the pion and Δ-hole branches is defined as [85] S(pi,ρi)=11Π(ω)/ω2, (28) where ω denotes ωπ-like and ωΔ-like. The eigenvalues of ωπ-like and ωΔ-like are generated from the pion dispersion relation ω2=pi2+mπ2+Π(ω), (29) where denotes the pion self-energy. Including the short-range Δ-hole interaction, the pion self-energy is defined as Π=pi2χ1gχ, (30) where denotes pion mass. The Migdal parameter, g’, was set to 0.6. χ is defined as χ=89(fΔmπ)2ωΔρ3ωΔ2ω2exp(2pi2/b2), (31) where ωΔ=mΔ2+pi2MN and is the delta mass. In this study, the π NΔ coupling constant was 2, and the cutoff factor b was 7mπ.

Both the Coulomb and pion potentials contribute to the decay of resonances and the reabsorption of pions. For instance, the energy balance of Δ0 in the decay of resonances and the reabsorption of pions can be written as mR2+pR2=MN2+(pRpπ)2+ωπ(pπ,ρ)+VπC, (32) where pR and pπ are the momenta of the resonances and pions, respectively, and mR is the mass of the resonances. Because Δ0 is uncharged, the Coulomb potential exists only for charged particles after the decay of Δ0.

The pion is generated from direct nucleon-nucleon collision and decay of the resonances Δ(1232) and N* (1440). The reaction channels of the resonances and pions, which are the same as those in the LQMD model, are as follows [33, 48, 86, 87]: NNNΔ,  NNNN*,  NNΔΔ,  ΔNπ,N*Nπ,  NNNNπ(sstate). (33) For the production of the Δ(1232) and N* (1440) resonances in nucleon-nucleon scattering, the parameterized cross-sections calculated using the one-boson exchange model were employed [88].

The decay width of Δ(1232) and N* (1440), which originates from the p-wave resonances, is momentum-dependent and is expressed as [88] Γ(|p|)=a1|p|3(1+a2|p|2)(a3+|p|2)Γ0, (34) where |p| is the momentum of the created pion (in GeV/c). The parameters a1, a2, and a3 were taken as 22.48 (17.22) c/GeV, 39.69 (39.69)c2/GeV2, and 0.04(0.09) GeV2/c2, respectively, for Δ (N*). The bare decay width of Δ (N*) was given by Γ0=0.12 (0.2) GeV. With the momentum-dependent decay width, the cross-section of pion-nucleon scattering has the Breit–Wigner form: σπN(s)=σmax(pmp)20.25Γ2(p)0.25Γ2(p)+(sm0)2, (35) where p and pm are the three momenta of the pions at energies of s and m0, respectively. The maximum cross-section σmax of Δ and N* resonances was obtained by fitting the total cross-sections of the experimental data in pion-nucleon scattering using the Breit–Wigner formula [89]. For instance, the maximum cross-section σmax of Δ resonance was 200, 133.33, and 66.7 mb for π+pΔ++ (πnΔ), π0pΔ+ (π0nΔ0) and πpΔ0 (π+nΔ+), respectively [87].

Note that the threshold effect was neglected in this study. The threshold effect mainly refers to the Δ production threshold energy and incident energy of two colliding nucleons modified by the medium. The incident and threshold energies in the medium are defined as sin=(E1*+10+E2*+20)2(1+2)2 and sth=(m3*+30+m4*+40)2(3+4)2, respectively [81, 90, 91]. Because i=0 and pi*0 for static nuclear matter, the difference between the incident and threshold energies is sinsth=E1*+10+E2*+20m3*30m4*40 [90]. The difference between the incident and threshold energies, which is isospin dependent, may result in an enhancement in the nnpΔ channel and suppression of the ppnΔ++ channel.

3

Results and discussions

The directed and elliptic flows were derived from the Fourier expansion of the azimuthal distribution: dNdϕ(y,pT)=N0[1+2V1(y,pT)cos(ϕ)+2V2(y,pT)cos(2ϕ)], (36) where the azimuthal angle of the emitted particle ϕ was measured from the reaction plane. pT=px2+py2 denotes the transverse momentum, and the directed flow V1 and elliptic flow V2 are expressed as follows: V1cos(ϕ)=pxpT,V2cos(2ϕ)=px2py2pT2. (37) The directed flow provides information on the azimuthal anisotropy of the transverse emission. The elliptic flow describes the competition between the in-plane (V2>0) and out-of-plane (V2<0) emissions.

Firstly, the collective flows of LQMD.RMF in the 197Au + 197Au collisions have been investigated at an incident energy of 2.92 GeV/nucleon (the corresponding nucleon-nucleon center-of-mass energy is SNN=3 GeV). The collective flows of LQMD with Skyrme interaction and without the momentum-dependent interaction have also been investigated. The Skyrme interaction of symmetric nuclear matter is taken to be the same as that of SLy6, with an incompressibility of K=230 MeV at ρ0=0.16 fm3[92, 93]. The symmetry energy of Skyrme interaction is defined as Esym=1322MN(3π2ρ2)2/3+Csym3(ρ/ρ0)γs. When the Csym and γs are set to be 38.6 MeV and 1.049, respectively, the symmetry energy and the slope parameter of symmetry energy are 31.6 MeV and 85.3 MeV, respectively. The symmetry energy and the slope parameter of symmetry energy of the Skyrme interaction are as same as those of set1. As shown in Fig. 3, we have compared the collective flows of LQMD.RMF and LQMD with Skyrme interaction with recent experimental data from STAR Collaboration [94]. The collective follows of LQMD.RMF and LQMD with Skyrme interaction can describe the STAR data at an impact parameter b=4 fm and b=7 fm, respectively. The directed flows of LQMD.RMF are almost consistent with the STAR data across the entire rapidity. However, The directed flows of LQMD with Skyrme interaction are weaker than the STAR data across the entire rapidity. This result may be due to the fact that the value of directed transverse momentum with the Lorentz effect is larger than that without the Lorentz effect [43]. The elliptic flows of both LQMD.RMF and LQMD with Skyrme interaction are consistent with the STAR data at rapidities smaller than 0.5. However, the elliptic flows of both LQMD.RMF and LQMD with Skyrme interaction are weaker than the STAR data at large rapidity. At an incident energy of 2.92 GeV/nucleon, the LQMD.RMF can better describe the experimental data compared to the LQMD with Skyrme interaction and without the momentum-dependent interaction. Based on the above analyses, the RMF models have been implemented into the LQMD model successfully.

Fig. 3
(Color online) Rapidity distribution of the collective flows of free protons in the 197Au + 197Au reaction at an incident energy of 2.92 GeV/nucleon. The open circles correspond to the STAR data [94]
pic

With this LQMD.RMF model, the 108Sn + 112Sn and 132Sn + 124Sn collisions in this study were investigated at an incident energy of 270 A MeV and an impact parameter b=3 fm. At an incident energy of 270 A MeV, the nuclear matter of the collision center can be compressed to densities approaching 2ρ0. In this dense region, collective flows, which reflect nucleon-nucleon interactions, can be used to extract the high-density behavior of the EOS [1, 44, 79, 86, 95]. The collective flows of free protons in the 108Sn + 112Sn and 132Sn + 124Sn collisions are shown in Fig. 4 and Fig. 5, respectively. It is reasonable that the maximum value of the directed flow V1 was significantly larger than that of the elliptic flow V2. In the same reaction system, the difference in the directed flows with various slopes of symmetry energy (set1, set2, and set3) was small. The difference in the elliptic flows with various slopes of symmetry energy was also small. To determine the relationship between the slope of the symmetry energy and the collective flow, we must process the collective flow data.

Fig. 4
(Color online) Rapidity distribution of the collective flows of free protons in the 108Sn + 112Sn reaction at an incident energy of 270 MeV/nucleon
pic
Fig. 5
(Color online) Rapidity distribution of the collective flows of free protons in the 132Sn + 124Sn reaction at an incident energy of 270 MeV/nucleon
pic

The difference between neutron and proton directed flows emitted from heavy-ion collisions can be used to extract the density dependence of the symmetry energy [86, 95]. The difference between the neutron and proton directed flows is defined as V1n-V1p, as shown in Fig. 6. The trend and shape of the difference between the neutron and proton directed flows were similar to those of nonrelativistic LQMD [86]. For a given reaction system (nearly symmetric 108Sn + 112Sn system or neutron-rich 132Sn + 124Sn system), the absolute value of the difference between the neutron and proton directed flows with a soft symmetry energy was higher than that with a stiff symmetry energy. Interestingly, the stiffness of the symmetry energy can be reflected through the difference between the neutron and proton directed flows. The relationship between the curvature of the symmetry energy Ksym and the collective flows was then investigated. The difference between Ksym of set1 and Ksym of set3 was 406 MeV, which was significantly larger than the 59.7 MeV difference between L of set1 and L of set3. Although the curvature of the symmetry energy Ksym also affected the difference between the neutron and proton directed flows, because the nuclear matter of the collision center could only approach 2ρ0 at an incident energy of 270 A MeV, the effects of Ksym were not significant compared to the effects of the slope parameter L.

Fig. 6
(Color online) Difference between neutron and proton directed flows in the 108Sn + 112Sn and 132Sn + 124Sn reactions at an incident energy of 270 MeV/nucleon
pic

In addition to collective flows, the production of isospin exotic particles such as hyperons, kaons, and pions can also be used to extract the symmetry energy [22-28]. Because the thresholds of hyperons and kaons were not reached at the incident energies in this study, the isospin exotic particles were pions. First, we calculated the ratio between the π- and π+ yields of the neutron-rich 132Sn + 124Sn system as a function of the collision energy at the impact parameter b=3fm and θcm<90°. Because set1 had the softest symmetry energy, it had the highest neutron density. Consequently, there were more neutron-neutron scatterings in set1, resulting in the production of more Δ- and π-. As shown in Fig. 7, the π-+ ratio of set1 was the highest, and the π-+ ratio of set2 was higher than that of set3. Specifically, at a collision energy of 270 MeV/nucleon, the π-/π+ ratio without (with) the π potential changed from 2.71 (2.54) to 2.23 (2.06) when the slope parameter was varied from L=85.3 to 145.0 MeV, that is, from set1 to set3. In other words, the π-+ ratio as a function of collision energy depends on the stiffness of the symmetry energy. This result was consistent with the predictions of most transport models [28, 39, 90]. When the π potential was considered, the interaction between π and the nucleon became attractive, resulting in a decrease in both π- and π+ via the absorbed channels πNΔ(1232) and Δ(1232)NNN. However, with the π potential, because there were more neutrons in the neutron-rich 132Sn + 124Sn system, π- was more easily absorbed than π+. Consequently, the π-+ ratio without the π potential was higher than that with the π potential in the same RMF model. Moreover, it is worth mentioning that the threshold effect neglected in this study may cause the π-+ ratio to be reversed. In other words, with the threshold effect [81, 90, 91], π-+ of set3 may be the highest, and the π-+ ratio of set2 may be higher than that of set1.

Fig. 7
(Color online) Ratio between the π- and π+ yields as a function of the incident energy in the 132Sn + 124Sn reaction
pic

Next, the properties of π were predicted as a function of the transverse momentum. As shown in Fig. 8, the left and right panels are the transverse momentum spectra of pions for the nearly symmetric 108Sn + 112Sn and neutron rich 132Sn + 124Sn reactions at θcm<90°, respectively. In collisions between isotopes, π+ is mainly generated from collisions between protons and π- is mainly generated from collisions between neutrons. Theoretically, a stiffer symmetry energy would have a stronger repulsive force to push out neutrons and a stronger attractive force to squeeze protons, resulting in a decrease in the π- yield and an increase in the π+ yield, respectively. As shown in panels (b) and (d) of Fig. 8, the stiffer symmetry energy indeed led to larger transverse momentum spectra for π+. However, the relationship between the transverse momentum spectra of π- and the stiffness of the symmetry energy could not be directly explained. Compared with the stiffness of the symmetry energy, the π potential had a more significant impact on the transverse momentum spectrum of π. For the neutron-rich 132Sn + 124Sn system, the transverse momentum spectra of both π+ and π- without the π potential were lower than those of the SπRIT data at pT≲200 MeV. When the π potential was considered, the transverse momentum spectra of both π+ and π- increased at pT≲200 MeV but decreased at pT≳200 MeV. Consequently, the transverse momentum spectra of π+ with the π potential were almost consistent with the SπRIT data [40]; however, the transverse momentum spectra of π- were still lower than the SπRIT data for the entire pT domain. The lower transverse momentum spectra of π- may be due to the absence of a threshold effect. The threshold effect, which was not considered in this study, may enhance the production of π- [81, 90, 91].

Fig. 8
(Color online) Transverse momentum spectra of pions as functions of transverse momentum at an incident energy of 270 MeV/nucleon. The left two panels [(a) and (b)] are the results of the 108Sn + 112Sn reaction, and the right two panels [(c) and (d)] are the results of the 132Sn + 124Sn reaction. The full circles correspond to the SπRIT data [40]
pic

Because a stiffer symmetry energy would have a stronger repulsive force to push out neutrons and a stronger attractive force to squeeze protons, resulting in a decrease in the π- yield and an increase in the π+ yield, respectively, the single ratio SR(π/π+)=[dM(π)/dpT]/[dM(π+)/dpT] may depend on the stiffness of the symmetry energy and the reaction system. As shown in Fig. 9, for both the nearly symmetric system and neutron-rich system, a softer symmetry energy led to a larger single ratio. In addition, for the same stiffness of the symmetry energy, because the neutron-neutron scattering of the neutron-rich 132Sn + 124Sn system was greater than that of the nearly symmetric 108Sn + 112Sn system, the single ratio of the neutron-rich system was higher than that of the nearly symmetric system. It is worth mentioning that the single ratio of 108Sn + 112Sn was almost consistent with the experimental data. However, the single ratio of 132Sn + 124Sn was lower than that of the experimental data for the entire pT domain. This was because the transverse momentum spectra π- of 132Sn + 124Sn were lower than the experimental data.

Fig. 9
(Color online) Single spectral ratios of pions as functions of transverse momentum for the 108Sn + 112Sn and 132Sn + 124Sn reactions at an incident energy of 270 MeV/nucleon. The full circles correspond to the SπRIT data [40]
pic

The double ratio between the neutron rich system and the nearly symmetric system DR(π/π+)=SR(π/π+)132+124/SR(π/π+)108+112, which can cancel out most of the systematic errors caused by Coulomb and isoscalar interactions, was suggested to extract the properties of the symmetry energy [40]. However, because a lower symmetry energy will lead to a larger single ratio for both the nearly symmetric system and neutron-rich system, as shown in Fig. 10, it is still difficult to understand the dependence of the double ratio on the stiffness of symmetry energy. In addition, the double ratio without the π potential decreased slightly as a function of the transverse momentum, whereas it increased slightly as the transverse momentum increased when the π potential was considered. However, the increasing trend was still considerably weaker than that of the experimental results. The lower double ratio originated from the lower single ratio of the neutron-rich 132Sn + 124Sn system compared with the experimental data. The threshold effect may enhance the production of π- [81, 90, 91] and the single ratio of the neutron-rich system, resulting in a larger double ratio.

Fig. 10
(Color online) Double ratio of pions as a function of transverse momentum at an incident energy of 270 MeV/nucleon. The full circles correspond to the SπRIT data [40]
pic
4

Conclusion

An RMF with various symmetry energies, namely set1, set2, and set3, was implemented in LQMD. The collective flows of the nearly symmetric 108Sn + 112Sn and neutron-rich 132Sn + 124Sn systems were successfully generated from the LQMD.RMF. It has been observed that the directed flow V1 was an order of magnitude larger than the elliptic flow V2. However, the difference between the directed flows V1 with various slopes of symmetry energy was small. The difference in the directed flows V2 with various slopes of symmetry energy was also small. To explore the relationship between the collective flow and the stiffness of the symmetry energy, we defined the difference between neutron and proton directed flows V1n-V1p. For a given nearly symmetric system or neutron-rich system, the absolute value of V1n-V1p increased with decreasing slope of the symmetry energy.

We also investigated the relationship between isospin exotic particles and the stiffness of the symmetry energy. The ratio between π- yield and π+ yield of the neutron-rich 132Sn + 124Sn system as a function of the collision energy increased with a decrease in the slope parameter of the symmetry energy. At an incident energy of 270 MeV/nucleon, the properties of π were predicted as a function of the transverse momentum. For the nearly symmetric 108Sn + 112Sn system, the single ratio of the nearly symmetric system was consistent with the experimental data. However, for the neutron-rich 132Sn + 124Sn system, the single ratio was lower than the experimental data, resulting in a double ratio lower than the experimental data. The π potential did not explain the lower transverse momentum spectra of π- in the neutron-rich system. Considering the π potential, the double ratio increased slightly with increasing transverse momentum. However, the increasing trend was still considerably weaker than that observed in the experimental results. The single ratio of the neutron-rich system and the double ratio may be lower than the experimental data because of the lack of a threshold effect. The threshold effect, which can enhance the production of π-, could be a candidate for enhancing the single ratio of the neutron-rich system to a double ratio. Moreover, because a softer symmetry energy led to a larger single ratio for both nearly symmetric and neutron-rich systems, the dependence of the double ratio on the stiffness of the symmetry energy was not significant. The sensitivity of the double ratio to the stiffness of the symmetry energy may also be due to the lack of a threshold effect. When the threshold effect is considered, the production of π- in a neutron-rich system may be more significant than that in a nearly symmetric system. In the future, when collective flows, the single ratio of the neutron-rich system and the double ratio of the LQMD.RMF are consistent with the experimental data, V1nV1p and the single ratio of the neutron-rich system π-+, which are sensitive to the stiffness of the symmetry energy, may be used to extract the slope parameter of the symmetry energy.

References
1. P. Danielewicz, R. Lacey, and W.G. Lynch,

Determination of the Equation of State of Dense Matter

. Science, 298, 1592(2002). https://doi.org/10.1126/science.1078070
Baidu ScholarGoogle Scholar
2. B.P. Abbott, R. Abbott, T.D. Abbott et al.,

Measurements of Neutron Star Radii and Equation of State

. Phys. Rev. Lett. 121, 161101 (2018). https://doi.org/10.1103/PhysRevLett.121.161101
Baidu ScholarGoogle Scholar
3. S. Huth, P.T.H. Pang, I. Tews et al.,

Constraining neutron-star matter with microscopic and macroscopic collisions

. Nature 606, 276-280 (2022). https://doi.org/10.1038/s41586-022-04750-w
Baidu ScholarGoogle Scholar
4. L.W. Chen,

Nucl. Symmetry Energy in Nucleon and Quark Matter

. Phys. Rev.,34, 20-28(2017). https://doi.org/10.11804/NuclPhysRev.34.01.020
Baidu ScholarGoogle Scholar
5. B.A. Li, P.G. Krastev, D.H. Wen et al.,

Towards understanding astrophysical effects of nuclear symmetry energy

. Eur. Phys. J. A, 55, 117(2019). https://doi.org/10.1140/epja/i2019-12780-8
Baidu ScholarGoogle Scholar
6. B.T. Reed, F.J. Fattoyev, C.J. Horowitz, and J. Piekarewicz,

Implications of PREX-2 on the Equation of State of Neutron-Rich Matter

. Phys. Rev. Lett. 126, 172503(2021). https://doi.org/10.1103/PhysRevLett.126.172503
Baidu ScholarGoogle Scholar
7. J.M. Lattimer, and M. Prakash,

Neutron star observations: Prognosis for equation of state Constraints

. Science, 442, 109-165 (2007). https://doi.org/10.1016/j.physrep.2007.02.003
Baidu ScholarGoogle Scholar
8. A. Drago, A. Lavagno, G. Pagliara, and D. Pigato,

Early appearance of Δ isobars in neutron stars

. Phys. Rev. C 90, 065809 (2014). https://doi.org/10.1103/PhysRevC.90.065809
Baidu ScholarGoogle Scholar
9. J.M. Lattimer and M. Prakash,

The equation of state of hot, dense matter and neutron stars, Phys

. Rep. 621, 127 (2016). https://doi.org/10.1016/j.physrep.2015.12.005
Baidu ScholarGoogle Scholar
10. M. Tsang, W. Lynch, P. Danielewicz, and C. Tsang,

Symmetry energy constraints from GW170817 and laboratory experiments

. Phys. Lett. B 795, 533 (2019). https://doi.org/10.1016/j.physletb.2019.06.059
Baidu ScholarGoogle Scholar
11. B. Fore and S. Reddy,

Pions in hot dense matter and their astrophysical implications

. Phys. Rev. C 101, 035809 (2020). https://doi.org/10.1103/PhysRevC.101.035809
Baidu ScholarGoogle Scholar
12. W.B. He, Y..G.. Ma. L. G. Pang, H..C.. Song, and K. Zhou,

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
13. W. Scheid, H. Müller, and W. Greiner,

Nuclear Shock Waves in Heavy-Ion Collisions

. Phys. Rev. Lett. 32, 741 (1974). https://doi.org/10.1103/PhysRevLett.32.741
Baidu ScholarGoogle Scholar
14. J. Kapusta and D. Strottman,

Global analysis of relativistic heavy-ion collisions

. Phys. Lett. B 106, 33 (1981). https://doi.org/10.1016/0370-2693(81)91074-1
Baidu ScholarGoogle Scholar
15. H. Stocker, L.P. Csernai, G. Graebner, G. Buchwald, H. Kruse, R.Y. Cusson, J.A. Maruhn, and W. Greiner,

Jets of nuclear matter from high energy heavy ion collisions

. Phys. Rev. C 25, 1873 (1982). https://doi.org/10.1103/PhysRevC.25.1873
Baidu ScholarGoogle Scholar
16. H.A. Gustafsson, H.H. Gutbrod, B. Kolb et al.,

Collective Flow Observed in Relativistic Nuclear Collisions

. Phys. Rev. Lett. 52, 1590 (1984). https://doi.org/10.1103/PhysRevLett.52.1590
Baidu ScholarGoogle Scholar
17. D.H. Rischke, Y. Pursun, J.A. Maruhn et al.,

The phase transition to the quark-gluon plasma and its effect on hydrodynamic flow

. Acta Phys. Hung. A 1, 309 (1995). https://doi.org/10.1007/BF03053749
Baidu ScholarGoogle Scholar
18. J. Brachmann, S. Soff, A. Dumitru et al.,

Antiflow of nucleons at the softest point of the equation of state

. Phys. Rev. C 61, 024909 (2000). https://doi.org/10.1103/PhysRevC.61.024909
Baidu ScholarGoogle Scholar
19. L.P. Csernai and D. Rohrich,

Third flow component as QGP signal

. Phys. Lett. B 458, 454 (1999). https://doi.org/10.1016/S0370-2693(99)00615-2
Baidu ScholarGoogle Scholar
20. B.A. Li and C.M. Ko,

Probing the softest region of the nuclear equation of state

. Phys. Rev. C 58, R1382 (1998). https://doi.org/10.1103/PhysRevC.58.R1382
Baidu ScholarGoogle Scholar
21. Y.G. Ma,

The Collective Flow from the Degree of Freedom of Nucleons to Quarks

. Journal of Fudan University (Natural Science)62, 273-292 (2023). https://doi.org/0427-7104(2023)03-0273-20
Baidu ScholarGoogle Scholar
22. L. Scalone, M. Colonna, and M. Di Toro,

Transverse flows in charge asymmetric collisions

. Phys. Lett. B 461, 9 (1999). https://doi.org/10.1016/S0370-2693(99)00835-7
Baidu ScholarGoogle Scholar
23. B.A. Li, A.T. Sustich, and B. Zhang,

Proton differential elliptic flow and the isospin dependence of the nuclear equation of state

. Phys. Rev. C 64, 054604(2001). https://doi.org/10.1103/PhysRevC.64.054604
Baidu ScholarGoogle Scholar
24. Q.F. Li, C.W. Shen, C.C. Guo et al.,

Nonequilibrium dynamics in heavy-ion collisions at low energies available at the GSI Schwerionen Synchrotron

. Phys. Rev. C 83, 044617 (2011). https://doi.org/10.1103/PhysRevC.83.044617
Baidu ScholarGoogle Scholar
25. M.D. Cozma,

Neutron-proton elliptic flow difference as a probe for the high density dependence of the symmetry energy

. Phys. Lett. B 700, 139 (2011). https://doi.org/10.1016/j.physletb.2011.05.002
Baidu ScholarGoogle Scholar
26. G. Ferini, T. Gaitanos, M. Colonna et al.,

Isospin Effects on Subthreshold Kaon Production at Intermediate Energies

. Phys. Rev. Lett. 97, 202301 (2006). https://doi.org/10.1103/PhysRevLett.97.202301
Baidu ScholarGoogle Scholar
27. B.A. Li,

Probing the High Density Behavior of the Nuclear Symmetry Energy with High Energy Heavy-Ion Collisions

. Phys. Rev. Lett. 88, 192701 (2002). https://doi.org/10.1103/PhysRevLett.88.192701
Baidu ScholarGoogle Scholar
28. T. Gaitanos, M. Di Toro, S. Typel et al.,

On the Lorentz structure of the symmetry energy

. Nucl. Phys. A 732, 24 (2004). https://doi.org/10.1016/j.nuclphysa.2003.12.001
Baidu ScholarGoogle Scholar
29. C. Sturm, I. Bottcher, M. Debowski, A. Forster, E. Grosse, P. Koczon, and B. Kohlmeyer, et al.,

(KaoS Collaboration), Evidence for a Soft Nuclear Equation-of-State from Kaon Production in Heavy-Ion Collisions

. Phys. Rev. Lett.86, 39 (2001). https://doi.org/10.1103/PhysRevLett.86.39
Baidu ScholarGoogle Scholar
30. G.Q. Li and C.M. Ko,

Subthreshold kaon production and the nuclear equation of state

. Phys. Lett. B 349, 405 (1995). https://doi.org/10.1016/0370-2693(95)00301-Z
Baidu ScholarGoogle Scholar
31. C. Fuchs, A. Faessler, E. Zabrodin et al.,

Probing the Nuclear Equation of State by K+ Production in Heavy-Ion Collisions

. Phys. Rev. Lett. 86, 1974 (2001). https://doi.org/10.1103/PhysRevLett.86.1974
Baidu ScholarGoogle Scholar
32. C. Hartnack, H Oeschler, and J. Aichelin,

Hadronic Matter Is Soft

. Phys. Rev. Lett. 96, 012302 (2006). https://doi.org/10.1103/PhysRevLett.96.012302
Baidu ScholarGoogle Scholar
33. Z.Q. Feng,

Constraining the high-density behavior of the nuclear equation of state from strangeness production in heavy-ion collisions

. Phys. Rev. C 83, 067604 (2011). https://doi.org/10.1103/PhysRevC.83.067604
Baidu ScholarGoogle Scholar
34. Z.Q. Feng and G.M. Jin,

Probing high-density behavior of symmetry energy from pion emission in heavy-ion collisions

. Phys. Lett. B 683, 140 (2010). https://doi.org/10.1016/j.physletb.2009.12.006
Baidu ScholarGoogle Scholar
35. P. Russotto, P.Z. Wu, M. Zoric et al.,

Symmetry energy from elliptic flow in 197Au+ 197Au

. Phys. Lett. B 697, 471 (2011). https://doi.org/10.1016/j.physletb.2011.02.033
Baidu ScholarGoogle Scholar
36. Z.G. Xiao, B.A. Li, L.W. Chen, G.C. Yong, and M. Zhang,

Circumstantial Evidence for a Soft Nuclear Symmetry Energy at Suprasaturation Densities

. Phys. Rev. Lett. 102, 062502 (2009). https://doi.org/10.1103/PhysRevLett.102.062502
Baidu ScholarGoogle Scholar
37. W.J. Xie, J. Su, L. Zhu, and F.S. Zhang,

Symmetry energy and pion production in the Boltzmann-Langevin approach

. Phys. Lett. B 718,1510 (2013). https://doi.org/10.1016/j.physletb.2012.12.021
Baidu ScholarGoogle Scholar
38. W. Reisdorf, M. Stockmeier, A. Andronic, M.L. Benabderrahmane, O.N. Hartmann, and N. Herrmann, et al.,

Systematics of pion emission in heavy ion collisions in the 1 A GeV regime

. Nucl. Phys. A 781, 459 (2007). https://doi.org/10.1016/j.nuclphysa.2006.10.085
Baidu ScholarGoogle Scholar
39. G. Jhang, J. Estee, J. Barney, G. Cerizza, and M. Kaneko, et al.,

Symmetry energy investigation with pion production from Sn+ Sn systems

. Phys. Lett. B 813, 136016 (2021). https://doi.org/10.1016/j.physletb.2020.136016
Baidu ScholarGoogle Scholar
40. J. Estee, W.G. Lynch, C.Y. Tsang, J. Barney, G. Jhang, M.B. Tsang, and R. Wang, et al.,

Probing the Symmetry Energy with the Spectral Pion Ratio

. Phys. Rev. Lett. 126, 162701 (2021). https://doi.org/10.1103/PhysRevLett.126.162701
Baidu ScholarGoogle Scholar
41. G.F. Wei, X. Huang, Q.J. Zhi, A.J. Dong, C.G. Peng, and Z.W. Long,

Effects of the momentum dependence of nuclear symmetry potential on pion observables in Sn plus Sn collisions at 270 MeV/nucleon

. Nucl. Sci. Tech. 33, 163 (2022). https://doi.org/10.1007/s41365-022-01146-3
Baidu ScholarGoogle Scholar
42. H. Sorge, H. Stöcker, and W. Greiner,

Poincare invariant Hamiltonian dynamics: modelling multi-hadronic interactions in a phase space approach

. Ann. Phys. 192, 266 (1989). https://doi.org/10.1016/0003-4916(89)90136-X
Baidu ScholarGoogle Scholar
43. T. Maruyama, S.W. Huang, N. Ohtsuka, G. Li, and A. Faessler,

Lorentz-covariant description of intermediate energy heavy-ion reactions in relativistic quantum molecular dynamics

. Nucl. Phys. A 534, 720 (1991). https://doi.org/10.1016/0375-9474(91)90468-L
Baidu ScholarGoogle Scholar
44. Y. Nara and H. Stoecker,

Sensitivity of the excitation functions of collective flow to relativistic scalar and vector meson interactions in the relativistic quantum molecular dynamics model RQMD.RMF

. Phys. Rev. C 100, 054902 (2019). https://doi.org/10.1103/PhysRevC.100.054902
Baidu ScholarGoogle Scholar
45. Y. Nara, T. Maruyama, and H. Stoecker,

Momentum-dependent potential and collective flows within the relativistic quantum molecular dynamics approach based on relativistic mean-field theory

. Phys. Rev. C 102, 024913 (2020). https://doi.org/10.1103/PhysRevC.102.024913
Baidu ScholarGoogle Scholar
46. Y. Nara, A. Jinno, K. Murase, and A. Ohnishi,

Directed flow of Λ in high-energy heavy-ion collisions andnΛ potential in dense nuclear matter

. Phys. Rev. C 106, 044902 (2022).
Baidu ScholarGoogle Scholar
47. Z.Q. Feng and G.M. Jin,

Pion production in heavy-ion collisions in the 1 A GeV region

. Chin. Phys. Lett. 26, 062501(2009). https://doi.org/10.1088/0256-307X/26/6/062501
Baidu ScholarGoogle Scholar
48. Z.Q. Feng, W.J. Xie, P.H. Chen, J. Chen, and G.M. Jin,

In-medium and isospin effects on particle production near threshold energies in heavy-ion collisions

. Phys. Rev. C 92, 044604 (2015). https://doi.org/10.1103/PhysRevC.92.044604
Baidu ScholarGoogle Scholar
49. Z.Q. Feng,

Nuclear dynamics and particle production near threshold energies in heavy-ion collisions

. Nucl. Sci. Tech. 29, 40 (2018). https://doi.org/10.1007/s41365-018-0379-z
Baidu ScholarGoogle Scholar
50. Z.Q. Feng, H.J. Liu, H.G. Cheng, and S.N. Wei,

Progress in strange particle production and hypernuclear physics in intermediate and high-energy heavy-ion collisions

. Nucl. Tech. 46, 103-113(2023). https://doi.org/10.11889/j.0253-3219.2023.hjs.46.080010
Baidu ScholarGoogle Scholar
51. M.H. Zhang, Y.H. Zhang, J.J. Li, N. Tang, S. Sun, and F.S. Zhang,

Progress in transport models of heavy-ion collisions for the synthesis of superheavy nuclei

. Nucl. Tech. 46, 137-145(2023). https://doi.org/10.11889/j.0253-3219.2023.hjs.46.080014
Baidu ScholarGoogle Scholar
52. Y.G. Ma,

Effects of α-clustering structure on nuclear reaction and relativistic heavy-ion collisions

. Nucl. Tech. 46, 8-29 (2023). https://doi.org/10.11889/j.0253-3219.2023.hjs.46.080001
Baidu ScholarGoogle Scholar
53. P.C. Li, Y.J. Wang, Q.F. Li, and H.F. Zhang,

Transport model analysis of the pion interferometry in Au+Au collisions at Ebeam =1.23 GeV/nucleon

. Sci. China-Phys. Mech. Astron. 66, 222011(2023). https://doi.org/10.1007/s11433-022-2026-5
Baidu ScholarGoogle Scholar
54. P.C. Li, J. Steinheimer, T. Reichert, A. Kittiratpattana, M. Bleicher, and Q.F. Li,

Effects of a phase transition on two-pion interferometry in heavy ion collisions at SNN=2.4-7.7 GeV

. Sci. China-Phys. Mech. Astron. 66, 232011(2023). https://doi.org/10.1007/s11433-022-2041-8
Baidu ScholarGoogle Scholar
55. K. Xiao, P.C. Li, Y.J. Wang, F.H. Liu, and Q.F. Li,

Effects of sequential decay on collective flows and nuclear stopping power in heavy-ion collisions at intermediate energies

. Nucl. Sci. Tech. 34, 62(2023). https://doi.org/10.1007/s41365-023-01205-3
Baidu ScholarGoogle Scholar
56. J.D. Walecka,

A theory of highly condensed matter

. Ann. Phys 83, 491(1974). https://doi.org/10.1016/0003-4916(74)90208-5
Baidu ScholarGoogle Scholar
57. J. Boguta, and A.R. Bodmer,

Relativistic calculation of nuclear matter and the nuclear surface

. Nucl. Phys. A 292, 413(1977). https://doi.org/10.1016/0375-9474(77)90626-1
Baidu ScholarGoogle Scholar
58. B.D. Serot,

A relativistic nuclear field theory with pi and rho mesons

. Phys. Lett. B 86B,146 (1979). https://doi.org/10.1016/0370-2693(79)90804-9
Baidu ScholarGoogle Scholar
59. S. Kubis, and M. Kutschera,

Nuclear matter in relativistic mean field theory with isovector scalar meson

. Phys. Lett. B 399, 191 (1997). https://doi.org/10.1016/S0370-2693(97)00306-7
Baidu ScholarGoogle Scholar
60. B. Liu, V. Greco, V. Baran, M. Colonna, and M. Di Toro,

Asymmetric nuclear matter: The role of the isovector scalar channel

. Phys.Rev. C 65, 045201 (2002). https://doi.org/10.1103/PhysRevC.65.045201
Baidu ScholarGoogle Scholar
61. C. Xu, B.A. Li, and L.W. Chen,

Symmetry energy, its density slope, and neutron-proton effective mass splitting at normal density extracted from global nucleon optical potentials.Phys

. Rev. C 82, 054607 (2010). https://doi.org/10.1103/PhysRevC.82.054607
Baidu ScholarGoogle Scholar
62. X.H. Li, W.J. Guo, B.A. Li, L.W. Chen, F.J. Fattoyev, and W.G. Newton,

Neutron-proton effective mass splitting in neutron-rich matter at normal density from analyzing nucleon-nucleus scattering data withinan isospin dependent optical model

. Phys. Lett. B 743, 408 (2015). https://doi.org/10.1016/j.physletb.2015.03.005
Baidu ScholarGoogle Scholar
63. W. Zuo, I. Bombaci, and U. Lombardo,

Asymmetric nuclear matter from an extended Brueckner-Hartree-Fock approach

. Phys. Rev. C 60,024605 (1999). https://doi.org/10.1103/PhysRevC.60.024605
Baidu ScholarGoogle Scholar
64. Z.Y. Ma, J. Rong, B.Q. Chen, Z.Y. Zhu, and H.Q. Song,

Isospin dependence of nucleon effective mass in Dirac Brueckner-Hartree-Fock approach

. Phys. Lett. B 604, 170 (2004). https://doi.org/10.1016/j.physletb.2004.11.004
Baidu ScholarGoogle Scholar
65. E.N.E. van Dalen, C. Fuchs, and A. Faessler,

Effective Nucleon Masses in Symmetric and Asymmetric Nuclear Matter

. Phys. Rev. Lett.95, 022302 (2005). https://doi.org/10.1103/PhysRevLett.95.022302
Baidu ScholarGoogle Scholar
66. W.H. Long, N. Van Giai, and J. Meng,

Density-dependent relativistic Hartree-Fock approach

. Phys. Lett. B 640, 150 (2006). https://doi.org/10.1016/j.physletb.2006.07.064
Baidu ScholarGoogle Scholar
67. D.D.S. Coupland, M. Youngs, Z. Chajecki, et al.,

Probing effective nucleon masses with heavy-ion collisions

. Phys. Rev. C 94, 011601 (2016). https://doi.org/10.1103/PhysRevC.94.011601
Baidu ScholarGoogle Scholar
68. P. Morfouace, C.Y. Tsang, Y. Zhang, et al.,

Constraining the symmetry energy with heavy-ion collisions and Bayesian analyses

. Phys. Lett. B 799, 135045 (2019). https://doi.org/10.1016/j.physletb.2019.135045
Baidu ScholarGoogle Scholar
69. L.W. Chen, C.M. Ko, and B.A. Li,

Isospin-dependent properties of asymmetric nuclear matter in relativistic mean field models

. Phys. Rev. C 76, 054316 (2007). https://doi.org/10.1103/PhysRevC.76.054316
Baidu ScholarGoogle Scholar
70. V. Baran, M. Colonna, V. Greco, and M. Di Toro,

Reaction dynamics with exotic nuclei

. Phys. Rep. 410, 335 (2005). https://doi.org/10.1016/j.physrep.2004.12.004
Baidu ScholarGoogle Scholar
71. M. Dutra, O. Lourenco, J.S.S. Martins, A. Delfino, J.R. Stone, and P.D. Stevenson,

Skyrme interaction and nuclear matter constraints

. Phys. Rev. C 85, 035201 (2012). https://doi.org/10.1103/PhysRevC.85.035201
Baidu ScholarGoogle Scholar
72. R. Sellahewa and A. Rios,

Isovector properties of the Gogny interaction

. Phys. Rev. C 90, 054327 (2014). https://doi.org/10.1103/PhysRevC.90.054327
Baidu ScholarGoogle Scholar
73. L. Ou, Z.X. Li, Y.X. Zhang, and M. Liu,

Effect of the splitting of the neutron and proton effective masses on the nuclear symmetry energy at finite temperatures

. Phys. Lett. B 697, 246 (2011). https://doi.org/10.1016/j.physletb.2011.01.062
Baidu ScholarGoogle Scholar
74. S. Goriely, S. Hilaire, M. Girod, and S. Pru,

First Gogny-Hartree-Fock-Bogoliubov Nuclear Mass Model

. Phys. Rev. Lett. 102, 242501 (2009). https://doi.org/10.1103/PhysRevLett.102.242501
Baidu ScholarGoogle Scholar
75. Z. Zhang, and L.w. Chen,

Isospin splitting of the nucleon effective mass from giant resonances in 208Pb

. Phys. Rev. C 93, 034335 (2016). https://doi.org/10.1103/PhysRevC.93.034335
Baidu ScholarGoogle Scholar
76. Z. Zhang, X.B. Feng, and L.W. Chen,

Bayesian inference on isospin splitting of nucleon effective mass from giant resonances in 208Pb

. Chin. Phys. C 45, 064104 (2021). https://doi.org/10.1088/1674-1137/abf428
Baidu ScholarGoogle Scholar
77. R. Marty, and J. Aichelin,

Molecular dynamics description of an expanding q/q¯ plasma with the Nambu-Jona-Lasinio model and applications to heavy ion collisions at energies available at the BNL Relativistic Heavy Ion Collider and the CERN Large Hadron Collider

. Phys. Rev. C 87, 034912 (2013). https://doi.org/10.1103/PhysRevC.87.034912
Baidu ScholarGoogle Scholar
78. A. Komar,

Constraint formalism of classical mechanics

. Phys. Rev. D 18, 1881 (1978). https://doi.org/10.1103/PhysRevD.18.1881
Baidu ScholarGoogle Scholar
79. M. Isse, A. Ohnishi, N. Otuka, P.K. Sahu, and Y. Nara,

Mean-field effects on collective flow in high-energy heavy-ion collisions at 2-158A GeV energies

. Phys. Rev. C72, 064908 (2005). https://doi.org/10.1103/PhysRevC.72.064908
Baidu ScholarGoogle Scholar
80. E.C.G. Sudarshan, N. Mukunda, and J.N. Goldberg,

Constraint dynamics of particle world lines

. Phys. Rev. D 23, 2218 (1981). https://doi.org/10.1103/PhysRevD.23.2218
Baidu ScholarGoogle Scholar
81. G. Ferini, M. Colonna, T. Gaitanos, M. Di Toro,

Aspects of particle production in isospin-asymmetric matter

. Nucl. Phys. A 762, 147 (2005). https://doi.org/10.1016/j.nuclphysa.2005.08.007
Baidu ScholarGoogle Scholar
82. D. Oliinychenko, and H. Petersen,

Deviations of the energy-momentum tensor from equilibrium in the initial state for hydrodynamics from transport approaches

. Phys. Rev. C 93, 034905 (2016). https://doi.org/10.1103/PhysRevC.93.034905
Baidu ScholarGoogle Scholar
83. G.E. Brown and W. Weise,

Pion scattering and isobars in nuclei

. Phys. Rep. 22, 279 (1975). https://doi.org/10.1016/0370-1573(75)90026-5
Baidu ScholarGoogle Scholar
84. B. Friemann, V.P. Pandharipande, and Q.N. Usmani,

Very hot nuclear matter and pion production in relativistic heavy-ion collisions

. Nucl. Phys. A 372, 483 (1981). https://doi.org/10.1016/0375-9474(81)90048-8
Baidu ScholarGoogle Scholar
85. L. Xiong, C.M. Ko, and V. Koch,

Transport model with quasipions

. Phys. Rev. C 47, 788 (1993). https://doi.org/10.1103/PhysRevC.47.788
Baidu ScholarGoogle Scholar
86. Z.Q. Feng,

Nuclear in-medium effects and collective flows in heavy-ion collisions at intermediate energies

. Phys. Rev. C 85, 014604 (2012). https://doi.org/10.1103/PhysRevC.85.014604
Baidu ScholarGoogle Scholar
87. Z.Q. Feng,

Nuclear fragmentation and charge-exchange reactions induced by pions in the Δ-resonance region

. Phys. Rev. C 94, 054617 (2016). https://doi.org/10.1103/PhysRevC.94.054617
Baidu ScholarGoogle Scholar
88. S. Huber, J. Aichelin,

Production of Δ - and N∗-resonances in the one-boson exchange model

. Nucl. Phys. A 573, 587 (1994). https://doi.org/10.1016/0375-9474(94)90232-1
Baidu ScholarGoogle Scholar
89. B.A. Li, A.T. Sustich, B. Zhang, C.M. Ko,

Studies of superdense hadronic matter in a relativistic transport model

. Int. J. Mod. Phys. E 10, 267 (2001). https://doi.org/10.1142/S0218301301000575
Baidu ScholarGoogle Scholar
90. T. Song, C.M. Ko,

Modifications of the pion-production threshold in the nuclear medium in heavy ion collisions and the nuclear symmetry energy

. Phys. Rev. C 91, 014901 (2015). https://doi.org/10.1103/PhysRevC.91.014901
Baidu ScholarGoogle Scholar
91. M.D. Cozma,

The impact of energy conservation in transport models on the π−∕π+ multiplicity ratio in heavy-ion collisions and the symmetry energy

. Phys. Lett. B 753, 166-172 (2016). https://doi.org/10.1016/j.physletb.2015.12.015
Baidu ScholarGoogle Scholar
92. E. Chabanat, R. Bonche, R. Haensel, J. Meyer, R. Schaeffer,

A Skyrme parametrization from subnuclear to neutron star densities Part II. Nuclei far from stabilities

. Nucl. Phys. A 635, 231-256 (1998). https://doi.org/10.1016/S0375-9474(98)00180-8
Baidu ScholarGoogle Scholar
93. Z.Q. Feng, G.M. Jin, F.S. Zhang,

Dynamical analysis on heavy-ion fusion reactions near Coulomb barrier

. Nucl. Phys. A 802, 91-106 (2008). https://doi.org/10.1016/j.nuclphysa.2008.01.022
Baidu ScholarGoogle Scholar
94. M.S. Abdallah et al.,

(STAR Collaboration), Disappearance of partonic collectivity in SNN=3 GeV Au+Au collisions at RHIC

. Phys. Lett. B 827, 137003 (2022). https://doi.org/10.1016/j.physletb.2022.137003
Baidu ScholarGoogle Scholar
95. V. Giordano, M. Colonna, M. Di Toro, V. Greco, J. Rizzo,

Isospin emission and flow at high baryon density: a test of the symmetry potential

. Phys. Rev. C 81, 044611 (2010). https://doi.org/10.1103/PhysRevC.81.044611
Baidu ScholarGoogle Scholar
Footnote

The authors declare that they have no competing interests.