logo

Constraint on the symmetry energy at high densities from neutron star observations using relativistic mean-field models

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

Constraint on the symmetry energy at high densities from neutron star observations using relativistic mean-field models

Ying Cui
Yuan Tian
Cheng-Jun Xia
Ying-Xun Zhang
Zhu-Xia Li
Nuclear Science and TechniquesVol.36, No.8Article number 141Published in print Aug 2025Available online 03 Jun 2025
14001

In this paper, we constrain the symmetry energy at high densities in nuclear matter using recent observations of neutron stars based on the calculations of relativistic mean-field (RMF) models. Using the observations of the neutron stars, we obtain the constraint on the symmetry energy at high densities, S(2ρ0) =40.54±12.47 MeV, and S(3ρ0) =44.12±29.38 MeV.

Equation of stateSymmetry energyNeutron star
1

Introduction

The nuclear equation of state (EoS), particularly the symmetry energy, is essential in both nuclear physics and astrophysics [1-7]. In nuclear physics, the symmetry energy significantly affects the structure of finite nuclei, including the neutron-skin thickness of heavy nuclei [8]. In astrophysics, it influences the key properties of neutron stars, such as their masses and radii [3, 4], as well as their cooling processes [9]. Consequently, constraining the symmetry energy is an important challenge in nuclear physics.

Many experimental attempts have been conducted to constrain the symmetry energy around the saturation density ρ0 [10-23]. These include measurements of the dipole polarizabilities of 208Pb [14, 15], giant dipole resonance energies [13], isospin diffusion in heavy-ion collisions [11], isobaric analog states [19], and neutron skin thickness [12, 24]. Because of the extensive research on determining the symmetry energy, the constraint on the symmetry energy at saturation density is relatively precise, with a commonly accepted value of J=30±4 MeV [10, 11, 13-22].

In contrast, the symmetry energy at suprasaturation remains unclear. Many terrestrial experiments have been performed to extract information on symmetry energy at suprasaturation. In studies on heavy-ion collisions, the constraints on the suprasaturation density dependence of the symmetry energy have been obtained from analyses of the π-/π+ ratio [25-31] and n/p elliptic flows ratio [32, 33]. However, the symmetry energy at supurasaturation densities exhibit a large model dependence. This is caused by the difficulties in solving the transport models and extrapolating the finite excited nuclear system to infinite nuclear matter at zero temperature. Consequently, further progress in understanding and constraining the equations of state for nuclear matter at high densities is expected from analyses that incorporate the properties of neutron stars.

The mass measurement of the pulsar PSR J0740+6620 from the Neutron Star Interior Composition Explorer [34, 35] revealed a neutron star with a mass of 2.080.07+0.07M, and GW190814 from the LIGO/Virgo collaboration [36] observed a 2.5-2.67 M compact star because of the neutron stars impose tight constraints on the EoS [37-40]. For the observations of the mass around the 1.4 M neutron star, the analysis of the tidal deformability from GW170817 by LIGO/Virgo [41, 42] along with radius R1.4=13.021.06+1.24 km and mass–radius posterior distributions from millisecond-pulsar PSR J0030+0451 via the (NICER) mission [43] have significantly enhanced the ability to constrain the symmetry energy of nuclear matter [44, 45]. Moreover, constraints on the symmetry energy have been derived from multiple observations of neutron stars in combination with nuclear theory [10, 44-47]. Among these studies, Skyme interactions [45, 48, 49], chiral effective field theory [10] and empirical local density functional metamodels [50] were developed under the non-relativistic framework. Relativistic mean-field models have been widely used to describe the properties of infinite nuclear matter, finite nuclei, and stellar matter [51-54]. A systematically analysis of the influence of all relativistic mean-field (RMF) sets on the properties of neutron stars. In this paper, we utilize RMF models to describe the properties of neutron stars with a focus on exploring the constraints imposed by these models on the EoS and symmetry energy at both saturation and suprasaturation densities.

The remainder of this paper is organized as follows. In Sect. 2, we review the theoretical aspects of RMF models, the EoS of a neutron star, and the Tolman–Oppenheimer–Volkov (TOV) equation. In Sect. 3, we present the results of the constraint on the symmetry energy obtained from the neutron star observations. Finally, the summary is presented in Sect. 4.

2

Model

2.1
Relativistic mean field

In this paper, we employ the framework of RMF models to extract information on the symmetry energy in symmetric nuclear matter (SNM) systems. RMF models provide a comprehensive description of both nuclear matter and finite nuclei and are widely used to study neutron star properties. To better categorize the parametrizations associated with RMF models, Ref. [53] defined three distinct types: (i) nonlinear, (ii) density-dependent, and (iii) point-coupling models.

The Lagrangians for the different RMF models are

1. Nonlinear (NL) model LNL=Ψ¯[ iγμμmN ]Ψ+12(μσμσmσ2σ2)U(σ)14ωμνωμν+12mω2ωμωμ+14ζ4(ωμωμ)214ρμνρμν+12mρ2ρμρμ+12(μδμδmδ2δ2)+gσNNΨ¯ΨσgωNNΨ¯γμΨωμgρNNΨ¯γμτΨρμ+gδNNΨ¯τΨδ+gσgω2σωμωμ(α1+12α1'gσ)+gσgρ2σρμρμ(α2+12α2'gσ)+12α3'gω2gρ2ωμωμρμρμ , (1) 2. Density-dependence (DD) model LDD=Ψ¯[iγμμmN]Ψ+12(μσμσmσ2σ2)14ωμνωμν+12mω2ωμωμ14ρμνρμν+12mρ2ρμρμ+12(μδμδmδ2δ2)+Γσ(ρ)Ψ¯ΨσΓω(ρ)Ψ¯γμΨωμΓρ(ρ)Ψ¯γμτΨρμ+Γδ(ρ)Ψ¯τΨδ, (2) 3. Point-coupling (PC) model LPC=Ψ¯[ iγμμmN ]ΨαS2(Ψ¯Ψ)2αV2(Ψ¯γμΨ)(Ψ¯γμΨ)αTV2(Ψ¯γμτΨ)(Ψ¯γμτΨ)αTS2(Ψ¯τΨ)(Ψ¯τΨ)βS3(Ψ¯Ψ)3γS4(Ψ¯Ψ)4γV4(Ψ¯γμΨΨ¯γμΨ)2αTV4(Ψ¯γμτΨΨ¯γμτΨ)2+[ η1+η2(Ψ¯Ψ) ](Ψ¯Ψ)(Ψ¯γμΨ)(Ψ¯γμΨ)η3(Ψ¯Ψ)(Ψ¯γμτΨ)(Ψ¯γμτΨ), (3) where ωμν and ρμν are defined by μωννωμ and μρννρμ, respectively. In Eq. 1, U(σ)=13g2σ3+14g3σ4 is the nonlinear potential of the σ field.

As an example, we use the nonlinear RMF model to present the expressions of relevant quantities [51-53]. In the RMF model, meson fields can be replaced with their expectation values: σ<σ>=σ¯, ωμ<ωμ>=ω¯0ρμ<ρμ>=ρ¯30, δμ<δμ>=δ¯30 (4) The equation of motions (EOMs) for nucleons and mesons were derived from the Lagrangian density: [γμiμgωω¯0gρρ¯30τ3(mNgσσ¯gδδ¯3τ3)]Ψ=0 (5) mσ2σ¯=gσρsg2σ¯2g3σ¯3+gσgω2(ω¯0)2(α1+α1'gσσ¯)+gσgρ2(ρ¯30)2(α2+α2'gσσ¯) (6) mω2ω¯0=gωρζgω4(ω¯0)3gσgω2σ¯ω¯0(2α1+α1'gσσ¯)α3'gω2gρ2(ρ¯30)2ω¯0 (7) mρ2ρ¯30=gρρ3gσgρ2σ¯ρ¯30(2α2+α2'gσσ¯)α3'gω2gρ2ρ¯30(ω¯0)2 (8) mδ2δ¯3=gδρs3 , (9) where the different densities are defined as ρs=Ψ¯Ψ=ρsn+ρsp, (10) ρ=Ψ¯γ0Ψ=ρn+ρp, (11) ρs3=Ψ¯τ3Ψ=ρspρsn, (12) ρ3=Ψ¯γ0τ3Ψ=ρpρn, (13) Filling up to the Fermi momenta kF,i for i = n or p in the nuclear matter, the neutron (n) and proton (p) scalar and baryon densities are given by ρsi=C(i)(2π)3k<kFid3kmi*k2+mi*2=mi*2π2[ kFiEFi*mi*2lnkFi+EFi*mi*2 ], (14) ρi=C(i)(2π)3k<kFid3k=(kFi)33π2 , (15) where the degeneracy factor C(i=n,p)=2, and EFi*=kFi2+mi*2 is the Fermi energy of neutrons and protons. Here, mi* is the Dirac effective mass of nucleons, given by the relation mn*=mNgσσ¯+gδδ¯3, (16) mp*=mNgσσ¯gδδ¯3. (17) The eigenvalues of the neutrons and protons from the Dirac equation are en=gωω¯0gρρ¯30+kn2*+mn*2, (18) ep=gωω¯0+gρρ¯30+kp2*+mp*2. (19) The expressions for the energy density and pressure are obtained from the given Lagrangian using the energy-momentum tensor relation given by Tμν=iL(μϕi)νϕigμνL, (20) where ϕi runs over all the possible fields. The energy density ϵ and pressure P can be obtained from the energy-momentum tensor: ϵNL= T00 =12mσ2σ¯2+13g2σ¯3+14g3σ¯412mω2(ω¯0)2ζ4gω4(ω¯0)4+gωω¯0ρ12mρ2(ρ¯30)2+gρρ¯30ρ3+12mδ2δ¯32gσgω2σ¯(ω¯0)2(α1+12α1'gσσ¯)gσgρ2σ¯(ρ¯30)2(α2+12α2'gσσ¯)12α3'gω2gρ2(ρ¯30)2(ω¯0)2+14[ 3EFn*ρn+mn*ρs n ]+14[ 3EF p*ρp+mp*ρs p ], (21) and PNL=13i=13 Tii =12mσ2σ¯213g2σ¯314g3σ¯4+12mω2(ω¯0)2+ζ4gω4(ω¯0)4+12mρ2(ρ¯30)212mδ2δ¯32+gσgω2σ¯(ω¯0)2(α1+12α1'gσσ¯)+gσgρ2σ¯(ρ¯30)2(α2+12α2'gσσ¯)+12α3'gω2gρ2(ρ¯30)2(ω¯0)2+14[ EFn*ρnmn*ρsn ]+14[ EFp*ρpmp*ρsp ]. (22) The same calculations for the density-dependence and point-coupling models are given in Refs. [55-59].

The binding energy per particle in asymmetric nuclear matter can be expressed as follows: E(ρ,α)=ϵρmN=E0(ρ)+S(ρ)α2+O(α4), (23) Here, the isoscalar term E0(ρ)=E(ρ,α=0) is the binding energy per nucleon in SNM, and the isovector term S(ρ) is the symmetry energy. ρ=ρn+ρp is the nuclear matter density, and α=(ρnρp)/(ρn+ρp) is the isospin asymmetry. The nuclear symmetry energy S(ρ) is defined as S(ρ)=122E(ρ,α)α2|α=0. (24) The symmetry energy is expanded in terms of (ρρ0)/3ρ0 as S(ρ)=J+L3ρ0(ρρ0)+ (25) where J=S(ρ0) is the symmetry energy, and L=3ρ0Sρ|ρ=ρ0 is the slope of the symmetry energy at the saturation density.

For SNM, mn*=mp*=mN* because δ¯3 vanishes. The symmetry energies of the RMF models are expressed as follows: S(ρ)NL=kF26EF*+12ρgρ2mρ*212ρ(gδ2mδ2mN*2EF*2[1+gδ2mδ2A(ρ,mN*)]), (26) S(ρ)DD=kF26EF*+12ρΓρ2mρ212ρ(Γδ2mδ2mN*2EF*2[1+Γδ2mδ2A(ρ,mN*)]), (27) S(ρ)PC=kF26EF*+12αVρ+η3ρsρ +12αTSρ(mN*2EF*2[1αTSA(ρ,mN*)]), (28) where mρ*2=mρ2+gσgρ2σ¯(2α2+α2'gσσ¯)+α3'gω2gρ2(ω¯0)2, and A(ρ,mN*)=3(ρsmN*ρEF*). (29) The expressions of the slope of symmetry energy (L) of the various RMF models are LNL=kF23EF*(1kF22EF*2kF3mN*EF*2π2mN*ρ)+3gρ22mρ*2ρ(11mρ*2mρ*2ρρ)12ρ(gδ2mδ2mN*2EF*2[ 1+gδ2mδ2A(ρ,mN*) ])×{ 32kF2EF*2+6(1mN*2EF*2)ρmN*mN*ρ3gδ2mδ211+gδ2mδ2A [ 2A(ρmN*mN*ρ)+ρkF2EF*3(13ρmN*mN*ρ) ] }, (30) LDD=kF23EF*(1kF22EF*2kF3mN*EF*2π2mN*ρ)+3Γρ22mρ2ρ(1+6ρΓρΓρρ)12ρ(Γδ2mδ2mN*2EF*2[ 1+Γδ2mδ2A(ρ,mN*) ])×{ 3+6ρΓδΓδρ2kF2EF*2+6(1mN*2EF*2)ρmN*mN*ρ3Γδ2mδ211+Γδ2mδ2A [ 2A(ρΓδΓδρ+ρmN*mN*ρ)+ρkF2EF*3(13ρmN*mN*ρ) ] }, (31) LPC=kF23EF*(1kF22EF*2kF3mN*EF*2π2mN*ρ)+32αVρ+3η3ρsρ+3η3ρ2ρsρ+12αTSρ(mN*2EF*2[ 1αTSA(ρ,mN*) ])×{ 32kF2EF*2+6(1mN*2EF*2)ρmN*mN*ρ+3αTS11αTSA [ 2A(ρmN*mN*ρ)+ρkF2EF*3(13ρmN*mN*ρ) ] }. (32) In this paper, we utilize 180 interaction parameter sets selected from those used in RMF models [53]. These parameter sets satisfy the incompressibility at saturation densities K0 within the range of 200 to 300 A MeV by computing the distribution of isoscalar monopole strength in 208Pb with relativistic models [60]. The interaction parameter sets are listed below, and detailed information is available in Ref. [53]:

(i)165 nonlinear RMF models (E [61], ER [61], NL1 [51], NL3 [62], NL3-II [62], NL3* [63], NL4 [64], NLC [54], NLB1 [51], NLB2 [51], NLRA1 [65], NLS [66], P-067 [67], P-070 [67], P-075 [67], P-080 [67], GL1 [68], GL2 [68], GL3 [68], GL4 [68], GL5 [68], GL6 [68], GL7 [68], GL8 [68], GL82 [69], GL9 [68], GM1 [70], GM2 [70], GM3 [70], GPSa [71], GPSb [71], NLρA [72], NLρB [72], RMF301 [73], RMF302 [73], RMF303 [73], RMF304 [73], RMF305 [73], RMF306 [73], RMF307 [73], RMF308 [73], RMF309 [73], RMF310 [73], RMF311 [73], RMF312 [73], RMF313 [73], RMF314 [73], RMF315 [73], RMF316 [73], RMF317 [73], RMF401 [73], RMF402 [73], RMF403 [73], RMF404 [73], RMF405 [73], RMF406 [73], RMF407 [73], RMF408 [73], RMF409 [73], RMF410 [73], RMF411 [73], RMF412 [73], RMF413 [73], RMF414 [73], RMF415 [73], RMF416 [73], RMF417 [73], RMF418 [73], RMF419 [73], RMF420 [73], RMF421 [73], RMF422 [73], RMF423 [73], RMF424 [73], RMF425 [73], RMF426 [73], RMF427 [73], RMF428 [73], RMF429 [73], RMF430 [73], RMF431 [73], RMF432 [73], RMF433 [73], RMF434 [73], Q1 [74], G1 [74], G2 [74], SMFT2 [75], DJM [75], S271 [76], Z271 [76], SRK3M5 [77], HD [78],HC [78], MS1 [79], MS3 [80], XS [80], NLSV1 [81], NLSV2 [81], TM1 [82], PK1 [83], hybrid [84], Z271* [85], G2* [85], BKA20 [86], BKA22 [86], BKA24 [86], FSUGOLD [9], FSUGOLD4 [87], FSUGOLD5 [87], FSUGZ00 [88], FSUGZ03 [88], FSUGZ06 [88], IU-FSU [89], NL3V1 [90], NL3V2 [90], NL3V3 [90], NL3V4 [90], NL3V5 [90], NL3V6 [90], S271V1 [90], S271V2 [90], S271V3 [90], S271V4 [90], S271V5 [90], S271V6 [90], Z271S1 [90], Z271S2 [90], Z271S3 [90], Z271S4 [90], Z271S5 [90], Z271S6 [90], Z271V1 [90], Z271V2 [90], Z271V3 [90], Z271V4 [90], Z271V5 [90], Z271V6 [90], TM1* [91], BSR1 [92], BSR2 [92], BSR3 [92], BSR4 [92], BSR5 [92], BSR6 [92], BSR7 [92], BSR8 [92], BSR9 [92], BSR10 [92], BSR11 [92], BSR12 [92], BSR13 [92], BSR14 [92], BSR15 [92], BSR16 [92], BSR17 [92], BSR18 [92], BSR19 [92], BSR20 [92], BSR21 [92], SVI-1 [93], SVI-2 [93], SIG-OM [94], NLρδ A [72], NLρδ B [72]);

(ii) Nine density-dependent RMF models (TW99 [95], DD-ME1 [96], PKDD [83], DD-ME2 [58], DD [97], DD-F [98], DD2 [99], DDMEδ [59], DDRHρδ [59]);

(iii)6 point-coupling RMF models (FA3 [57], FA4 [57], FZ3 [57], VZ3 [57], PC-F1 [55], PC-F3 [55]).

2.2
Equation of state of a neutron star

For different density regions of a neutron star, different forms of the EoS are employed in this paper as follows.

(i) For the outer crust of a neutron star, the EoS provided by Baym, Pethick, and Sutherland (BPS) [100] is adopted for densities in the range ρminρouter. Here, ρmin is the minimum density (ρmin =4.73×10-15 fm-3), with the corresponding energy density ϵmin=4.38×10-12 MeV, fm-3 and pressure Pmin=6.3×10-25 MeV fm-3. The outer crust density is ρouter=2.57×10-4 fm-3 with an energy density ϵouter=0.24 MeV fm-3 and pressure Pouter=4.86×10-4 MeV, fm-3. In this region, because of the heavy nuclei, primarily around the iron mass number, a Coulomb lattice coexists in β-equilibrium (i.e., equilibrium with respect to weak interaction processes) with an electron gas in the neutron star outer crust [100].

(ii) For the inner crust, ρouter<ρρT, the EoS takes the form P=A+Bϵ4/3 from Refs. [101, 102], where two constants A and B are adjusted to match the outer crust EoS to that of a liquid core at the crust-core transition density ρT (ρT=0.5ρ0 [103]).

(iii) In the liquid core region, also referred to as the outer core (OC) region, the density range is ρT<ρρOC (ρOC=3ρ0), where the EOS is obtained using the RMF. This region requires the system to be in β-equilibrium and is composed of protons, neutrons, electrons, and muons. For β equilibrium to be attained in nuclear matter, the following electroweak processes must occur: np+e+ν¯e, p+en+νe, eμ+νe+ν¯μ, p+μn+νμ, and np+μ+νμ. The β-equilibrated matter must satisfy the following conditions for the chemical potentials: μnμp=μe,μe=μμ, (33) where μνi=0, i=e,μ. At zero temperature, the neutron and proton chemical potentials are μn=en and μp=ep, respectively. For relativistic degenerate electrons, μe=me2+kFe2=me2+(3π2xeρ)2/3 (34) μμ=mμ2+kFμ2=mμ2+(3π2xμρ)2/3, (35) where me=0.511 MeV, and mμ=0.105 GeV. With the particle fraction xi=ρi/ρ (i=n,p,e,μ), xp=xe+xμ satisfies charge neutrality.

(iv) For the neutron star inner core (IC) region at a high density (ρ > ρOC), studies have incorporated exotic particles such as hyperons, Δs, quarks, and dark matter into the core of massive neutron stars at densities exceeding 3ρ0 [104-115]. However, owing to the significant uncertainties in the interactions between nucleons and exotic particles (e.g., hyperons, Δs, and quarks), and the unclear composition of the neutron star core, a polytropic EoS is employed instead of explicitly modeling all possible exotic particles. In this paper, a piecewise polytropic EoS of the form Pi(ρ)=κiργi [116-118] is used to smoothly extend the EoS to the density region of the IC of the neutron star (ρ > ρOC). The polytropic EoS is constructed as follows: Pi=κiργi,withκiκi1 (36) γi=ln(Pi/Pi1)ln(ρi/ρi1),Pi=Pi1(ρiρi1)γi, (37) ϵi=(ϵi1Pi1γi1)(PiPi1)1/γi+Piγi1, (38) where the set of dividing densities ρOC=ρ1<ρ2<, and ρi+1ρi=Δρ=0.05ρ0.

In summary, the EoS of neutron star matter is P(ϵ)={PBPS(ϵ)for ρminρρouterA+Bϵ4/3for ρouter<ρρTPRMF(ϵ)for ρT<ρρOCPIC(ϵ)for ρ>ρOC , where P=PN+Pe+Pμ, ϵ=ϵN+ϵe+ϵμ, and the total pressure and energy density should include the leptons. The thermodynamically stable EoS must satisfy Pϵ0. The adiabatic speed of sound can be expressed as cs2=(vsc)2=Pϵ<1, (39) where ϵ is the energy density of the β-stable nuclear matter. For the causality condition, i.e., the speed of sound is always less than that of light vs<c.

2.3
Tolman–Oppenheimer–Volkov equation

The structure of a neutron star is obtained by solving the TOV equation derived from General Relativity [119-121]. The TOV equations are dPdr=GMϵr2(1+P/ϵ)(1+4πr3P/M)12GM/r, (40) dMdr=4πr2ϵ, (41) where G=6.707×10-45 MeV-2 is the gravitational constant, r is the distance from the core of the star, P=P(r) is the pressure, and M=M(r) is the mass with radius r.

The in-spiral phase of the two merging neutron stars creates strong tidal gravitational fields, resulting in the deformation of the multipolar structure of the star. The deforming effects are quantified through the tidal deformability parameter Λ, which relates the induced mass quadruple moment Qij to the time-independent external tidal field Eij through the relation [122-124]: Qij=k22R53GEij. (42) Here, k2 is the Love number, which can be obtained from the solution of the first-order differential equation [124] dydr=y2ryr(r4πGr3(ϵp))r2GMrQ, (43) Q=4πr[G(5ϵ+9p+(ϵ+p)/cs2)321πr2]12GM/r [2G(M+4πpr3)r(r2GM)]2, (44) k2=85β5(12β)2[2yR+2β(yR1)]×{2β[63yR+3β(5yR8)]+4β3[1311yR+β(3yR2)+2β2(1+yR)]+3(12β)2[2yR+2β(yR1)]ln(12β)}1, (45) where yR=y(R), and β=GM/R. Eqs. 40, 41, 43 are solved using the boundary conditions at the center of the star, M(0)=0, P(0)=Pc, and y(0)=2, where Pc is the central pressure. Varying Pc yields all possible stars for a given EoS. Thus, P(R)=0 (vacuum pressure being set to zero) defines the radius of star R and the total gravitational mass of the star is M(R), which is simply denoted by M in the following.

The dimensionless deformability Λ is defined as follows: Λ=2k23(RGM)5 . (46)

3

Results and discussions

Generally, most RMF models are adjusted to describe nuclei and nuclear matter in the density region from near subsaturation density ρ2/3ρ0 (average between central and surface densities [74, 103, 125-128]) to saturation density. Moreover, the symmetry energy has large variations at high densities obtained using different parameters of RMF models. The symmetry energy as a function of density obtained by RMF models with 180 parameter sets is shown in Fig. 1. We observe that various RMF models predict very different density behaviors of the symmetry energy, particularly at suprasaturation densities ρ>ρ0. For example, the magnitude of the symmetry energy varies from 29.5 to 114.8 MeV at 2ρ0 and from about 14.7 to 188.7 MeV at a density of 3ρ0. In Fig. 1, the symmetry energy behaviors are classified into three types: hard, linear, and soft, represented by the green, gray, and orange lines, respectively. We observe from the figure that most of the symmetry energy occurs near the linear types in RMF models.

Fig. 1
(Color online) Symmetry energy for 180 RMF models as a function of density ρ/ρ0; the green, gray, and orange lines represent hard, linear, and soft symmetry energies, respectively
pic

In this paper, we use the observations of neutron stars to constrain the nuclear EoS at high densities ρρ0 because neutron star properties are strongly correlated with the nuclear EoS, as mentioned in Ref. [44, 50, 129]. This constraint can be obtained from the measurements of neutron stars for a mass of M = 1.4 M, such as the tidal deformability from the analysis of gravitational wave data, Λ1.4190120+390 from GW170817 [41], and the radius–mass relation from both GW170817 and PSRJ0030+0451 [41, 43]. For the maximum masses of neutron stars, the maximum neutron star mass MMax2M is used from Refs. [35, 42], because a compact star with a mass of 2.590.09+0.08M may be the lightest black hole [36].

Our results for tidal deformability with the empirical values from GW170817 (red point) [41] are plotted in panel (a) of Fig. 2. The mass–radius relationship of neutron stars is shown in panel (b) of Fig. 2, where the pink shaded areas represent the region of the posterior distributions at 90% confidence the analysis from PSRJ0030+0451 [43], the blue shaded region is the posterior distribution at 95.4% confidence from PSR J0740+6620 [39], and the purple and green shaded areas denote the region of the posterior distributions at 90% confidence for GW170817’s lighter and heavier neutron stars [42], respectively. In Fig. 2, the gray lines represent the results for all the selected 180 RMF parameter sets, which are models without the constraint of the neutron star. With the constraint from the observables of neutron stars such as tidal deformability [41] and mass–radius relation around 1.4 M [42, 43] combined with the maximum masses of neutron stars above 2 M, the results of restricted RMF models are indicated by black lines in Fig. 2. Because of the neutron star multi-observables, the analysis of empirical values excludes most RMF parameter sets, and only nine sets remain: HC, FSUGZ03, IU-FSU, G2*, BSR8, BSR9, DD-F, FA3, and FZ3, which can simultaneously describe both the tidal deformability [40] and the overlap of the two mass–radius relation regions [42, 43]. The symmetry energy for RMF models HC, FSUGZ03, IU-FSU, G2*, BSR8, BSR9 is described by Eq. 28, that for DD-F by Eq. 25, and that for FA3, FZ3 by Eq. 28.

Fig. 2
(Color online) (a) Tidal deformability Λ as a function of the neutron star mass M, where the gray and black lines are the results from the RMF models without and with constraint of the neutron star respectively. (b) Neutron star mass M as a function of neutron star radius R. The green, purple, and pink shaded regions are the posterior distributions at 90% confidence for GW170817’s lighter neutron star, GW170817’s heavier neutron star [42], and J0030+0451 [43], respectively. The blue shade region represents the posterior distributions at 95.4% confidence from PSR J0740+6620 [39]
pic

Figure 3 shows a comparison of the pressure–density relations between the empirical values from measurements of neutron stars and the RMF model calculations. The green shaded regions enclose the empirical pressure given by the “spectral” EoS inferred from the Bayesian analysis of the GW170817 data at a 90% confidence level, maintaining the lower limit of the maximum neutron star mass at 2 M. Our results of the pressure–density relation (black lines) for neutron-rich matter with β stability are mostly under the limit of the empirical region at the OC (0.5ρ0<ρ3ρ0), where the EoS is described by RMF models. For the IC area, ρ>3ρ0, the order-by-order polytropes EoS in Eqs. (36, 37, 38) occur primarily in the region of neutron star measurement (except the EoS from the HC model).

Fig. 3
(Color online) Pressure as a function of density for the neutron-rich matter with β stability. The green shaded area represents data from GW170817 [42]
pic

Figure 4 depicts the constraints on the J-L relation, which were compiled in [10, 130, 131]. In this paper, the symmetry energy J-L constraint from the neutron star observables based on the RMF models is shown as scattering points (cyan stars) with the range of symmetry energy at saturation J=30.66±0.96 MeV, and slope L=49.47±20.39 MeV. Figure 4 includes the J-L constraints obtained in the analysis of the finite nuclei (neutron skin thicknesses of Sn isotopes and 209Pb, isobaric analog states and isovector skins, and the dipole polarizability of 209Pb) and nuclear matter (heavy-ion collisions and chiral effective field theory calculations of nuclear matter). The enclosed overlap region [131] from constraints obtained from experimental measurements of the neutron skin thicknesses of Sn, dipole polarizabilities, giant dipole resonances, heavy-ion collisions, and nuclear mass fitting correspond to J of approximately 29.0-32.7 MeV, and L is approximately 40.5-61.9 MeV [130, 131]. In this paper, the constraints for the symmetry energy and its slope at saturation, which are obtained from neutron star observables, are almost in this overlap region, as shown in Fig. 4. Our results can reproduce the properties of neutron stars but cannot reproduce the PREX-II neutron skin using the RMF model. Compared with the research by Chen et al. [132], where the slope parameter L=80±25 MeV was obtained from a study of 23 RMF parameter sets, the result in this paper, L=49.47±20.39 MeV, is relatively soft.

Fig. 4
(Color online) Constraints on the J-L correlation. The cyan stars are our result with the constraint from the RMF models using the neutron star observables. The enclosed white area is the overlap region obtained from heavy-ion collisions (HIC) [11], neutron skin thicknesses of Sn isotopes [12], giant dipole resonances (GDR) [13], the dipole polarizability of 208Pb [14, 15], and the energy-density functionals for nuclear masses (masses) [16]. Experimental constraints are obtained from the isobaric analog states and isovector skins (IAS+ΔR) [19], unitary gas (UG) limit by Tews et al. [22], and the neutron skin thicknesses of 208Pb [24]. The microscopic calculations of neutron matter have shown to be χEFT are from chiral effective field theory via the Gaussian Process–BUQEYE collaboration [10], Hebeler et al. (H) [20], and Gandolfi et al. (G) [21]
pic

The symmetry energy at ρ03ρ0 constrained using the neutron star multi-observables is displayed as a cyan shaded region in Fig. 5. Analysis of doubly magic nuclei and masses of neutron-rich nuclei [133] (black square), isobaric analog states (IAS) [134] (red region), and isospin diffusion in heavy-ion collisions [11] (olive region), the electric dipole polarizability in 209Pb [135] (red point), the multi-observables (isospin diffusion, neutron skin and neutron star) [45] (blue dash line), χEFT (magenta hatched) based on Gaussian Process–BUQEYE (χEFT-GPB) Collaboration [10, 136, 137], the constraint from neutron star observations and nuclear matter experiments (gray hatched shaded area) [47], and neutron skin thicknesses of 209Pb by PREX-II [24] (up triangle), are also included for comparison. To obtain full information on the symmetry energy, we also present the range of symmetry energies at suprasaturation S(2ρ0)=28.07-53.00 MeV, and S(3ρ0)=14.74-73.49 MeV in Fig.5. Our result for the symmetry energy constraint is similar to that from the χEFT-GPB Collaboration, which is a microscopic calculation. The symmetry energy at a high density in this paper is softer than the result from Tsang et al. [47] (gray hatched area), which used parametric priors based on an expansion that is widely used in nuclear physics. This discrepancy suggests that further constraints on the symmetry energy should be achieved by reducing the uncertainties in the HIC experiments. Additionally, the pressure of the SNM is presented in Appendix.5.

Fig. 5
(Color online) Symmetry energy as a function of density. The blue shadowed region represents the symmetry energy constraint by the observables of the neutron star in this paper. The black square shows the properties of doubly magic nuclei (DMN) and masses of neutron-rich nuclei [133], the red region represents results of IAS [134], the olive region is from HIC [11], the red point is from the electric dipole polarizability (EDP) in 209Pb [135], the blue dash line represents data obtained using multi-observables (isospin diffusion, neutron skin and neutron star) [45], the magenta (hatched) contours represent the calculations from χEFT based on the Gaussian process from the BUQEYE Collaboration [10, 136, 137], the gray (hatched) shaded area shows the constraint from neutron star observations and nuclear matter (NSNM) experiments [47], and, for the neutron, the triangle shows constraints from neutron skin thicknesses of 209Pb by PREX-II
pic
4

Summary

We have extracted information on the symmetry energy at suprasaturation densities from astronomical observations using relativistic mean-field models. In this paper, we have employed 180 RMF parameter sets with incompressibility at the saturation density K0=200-300 MeV, which are suitable for describing the isoscalar monopole distribution strength in 209Pb. By combining the measurements of the 1.4 solar-mass neutron stars, such as tidal deformability (Λ1.4190120+390), the mass–radius relation [41, 43], and the maximum massive at least 2 M neutron stars, we derive constraints on the symmetry energy in the density region ρ03ρ0. At the saturation density, the symmetry energy is J=30.66±0.96 0.96 MeV and the slope is L=49.47±20.39 MeV, which are consistent with the overlap region of J-L constraints from some territory experiments. The symmetry energy constraints at 2ρ0 and 3ρ0 are S(2ρ0) =40.54±12.47 MeV and S(3ρ0) =44.12±29.38 MeV, as shown in Fig. 5.

In the next step, we will explore the entire parameter space of relativistic mean-field models using Bayesian inference with neutron star observational data, which can refine the parameter constraints and provide quantitative constraints on the EoS. Furthermore, the combination of constraints on the EoS from heavy-ion collision analyses (e.g., K/K+ data), neutron star cooling properties such as luminosity data, and measurements of neutron skin thickness in finite nuclei (209Pb and 48Ca) could also reduce the uncertainties of the constraints in future research.

References
1. B.A. Li, C.M. Ko, and W. Bauer,

Isospin physics in heavy ion collisions at intermediate-energies

. Int. J. Mod. Phys. E 7, 147 (1998). https://doi.org/10.1142/S0218301398000087
Baidu ScholarGoogle Scholar
2. 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
3. J.M. Lattimer and M. Prakash,

The physics of neutron stars

. Science 304, 536 (2004). https://doi.org/10.1126/science.1090720
Baidu ScholarGoogle Scholar
4. J.M. Lattimer and M. Prakash,

Neutron Star Observations: Prognosis for Equation of State Constraints

. Phys. Rep. 442, 109 (2007). https://doi.org/10.1016/j.physrep.2007.02.003
Baidu ScholarGoogle Scholar
5. A.W. Steiner, M. Prakash, J.M. Lattimer et al.,

Isospin asymmetry in nuclei and neutron stars

. Phys. Rep. 411, 325 (2005). https://doi.org/10.1016/j.physrep.2005.02.004
Baidu ScholarGoogle Scholar
6. V. Baran, M. Colonna, V. Greco et al.,

Reaction dynamics with exotic beams

. Phys. Rep. 410, 335 (2005). https://doi.org/10.1016/j.physrep.2004.12.004
Baidu ScholarGoogle Scholar
7. B.A. Li, L.W. Chen, and C.M. Ko,

Recent progress and new challenges in isospin physics with heavy-ion reactions

. Phys. Rep. 464, 113 (2008). https://doi.org/10.1016/j.physrep.2008.04.005
Baidu ScholarGoogle Scholar
8. P.G. Reinhard and W. Nazarewicz,

Nuclear charge and neutron radii and nuclear matter: Trend analysis in Skyrme density-functional-theory approach

. Phys. Rev. C 93, 051303 (2016). https://doi.org/10.1103/PhysRevC.93.051303
Baidu ScholarGoogle Scholar
9. B.G. Todd-Rutel and J. Piekarewicz,

Neutron-rich nuclei and neutron stars: A new accurately calibrated interaction for the study of neutron-rich matter

. Phys. Rev. Lett. 95, 122501 (2005). https://doi.org/10.1103/PhysRevLett.95.122501
Baidu ScholarGoogle Scholar
10. C. Drischler, R.J. Furnstahl, J.A. Melendez et al.,

How well do we know the neutron-matter equation of state at the densities inside neutron stars? A bayesian approach with correlated uncertainties

. Phys. Rev. Lett. 125, 202702 (2020). https://doi.org/10.1103/PhysRevLett.125.202702
Baidu ScholarGoogle Scholar
11. M.B. Tsang, Y. Zhang, P. Danielewicz et al.,

Constraints on the density dependence of the symmetry energy

. Phys. Rev. Lett. 102, 122701 (2009). https://doi.org/10.1103/PhysRevLett.102.122701
Baidu ScholarGoogle Scholar
12. L.W. Chen, C.M. Ko, B.A. Li et al.,

Density slope of the nuclear symmetry energy from the neutron skin thickness of heavy nuclei

. Phys. Rev. C 82, 024321 (2010). https://doi.org/10.1103/PhysRevC.82.024321
Baidu ScholarGoogle Scholar
13. L. Trippa, G. Colo, and E. Vigezzi,

The Giant Dipole Resonance as a quantitative constraint on the symmetry energy

. Phys. Rev. C 77, 061304 (2008). https://doi.org/10.1103/PhysRevC.77.061304
Baidu ScholarGoogle Scholar
14. A. Tamii, I. Poltoratska, P. von Neumann-Cosel et al.,

Complete electric dipole response and the neutron skin in 208Pb

. Phys. Rev. Lett. 107, 062502 (2011). https://doi.org/10.1103/PhysRevLett.107.062502
Baidu ScholarGoogle Scholar
15. X. Roca-Maza, M. Centelles, X. Viñas et al.,

Electric dipole polarizability in 208Pb: Insights from the droplet model

. Phys. Rev. C 88, 024316 (2013). https://doi.org/10.1103/PhysRevC.88.024316
Baidu ScholarGoogle Scholar
16. T. Kortelainen, M. Lesinski, J. Moré et al.,

Nuclear energy density optimization

. Phys. Rev. C 82, 024313 (2010). https://doi.org/10.1103/PhysRevC.82.024313
Baidu ScholarGoogle Scholar
17. A.W. Steiner, J.M. Lattimer, and E.F. Brown,

The equation of state from observed masses and radii of neutron stars

. Astrophys. J. 722, 33 (2010). https://doi.org/10.1088/0004-637X/722/1/33
Baidu ScholarGoogle Scholar
18. A.W. Steiner, J.M. Lattimer, and E.F. Brown,

The neutron star mass-radius relation and the equation of state of dense matter

. Astrophys. J. 765, L5 (2013). https://doi.org/10.1088/2041-8205/765/1/L5
Baidu ScholarGoogle Scholar
19. P. Danielewicz, P. Singh, and J. Lee,

Symmetry energy III: Isovector skins

. Nucl. Phys. A 958, 147 (2017). https://doi.org/10.1016/j.nuclphysa.2016.11.008
Baidu ScholarGoogle Scholar
20. K. Hebeler, J.M. Lattimer, C.J. Pethick et al.,

Constraints on neutron star radii based on chiral effective field theory interactions

. Phys. Rev. Lett. 105, 161102 (2010). https://doi.org/10.1103/PhysRevLett.105.161102
Baidu ScholarGoogle Scholar
21. S. Gandolfi, J. Carlson, and S. Reddy,

The maximum mass and radius of neutron stars and the nuclear symmetry energy

. Phys. Rev. C 85, 032801(R) (2012). https://doi.org/10.1103/PhysRevC.85.032801
Baidu ScholarGoogle Scholar
22. I. Tews, J.M. Lattimer, A. Ohnishi et al.,

Symmetry parameter constraints from a lower bound on neutron-matter energy

. Astrophys. J. 848, 105 (2017). https://doi.org/10.3847/1538-4357/aa8db9
Baidu ScholarGoogle Scholar
23. M. Thiel, C. Sfienti, J. Piekarewicz et al.,

Neutron skins of atomic nuclei: per aspera ad astra

. J. Phys. G 46, 093003 (2019). https://doi.org/10.1088/1361-6471/ab2c6d
Baidu ScholarGoogle Scholar
24. B.T. Reed, F.J. Fattoyev, C.J. Horowitz et al.,

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
25. Z. Xiao, B.A. Li et al.,

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
26. M. Zhang, Z.G. Xiao, B.A. Li et al.,

A Systematic study of the pi-/pi+ ratio in heavy-ion collisions with the same neutron/proton ratio but different masses

. Phys. Rev. C 80, 034616 (2009). https://doi.org/10.1103/PhysRevC.80.034616
Baidu ScholarGoogle Scholar
27. W.J. Xie, J. Su, L. Zhu et al.,

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
28. Z.Q. Feng, 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
29. J. Xu, C.M. Ko, Y. Oh,

Isospin-dependent pion in-medium effects on charged pion ratio in heavy ion collisions

. Phys. Rev. C 81, 024910 (2010). https://doi.org/10.1103/PhysRevC.81.024910
Baidu ScholarGoogle Scholar
30. 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 (2016). https://doi.org/10.1016/j.physletb.2015.12.015
Baidu ScholarGoogle Scholar
31. T. Song and 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
32. 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
33. M.D. Cozma, Y. Leifels, W. Trautmann et al.,

Toward a model-independent constraint of the high-density dependence of the symmetry energy

. Phys. Rev. C 88, 044912 (2013). https://doi.org/10.1103/PhysRevC.88.044912
Baidu ScholarGoogle Scholar
34. H.T. Cromartie, E. Fonseca, S.M. Ransom et al.,

Relativistic Shapiro delay measurements of an extremely massive millisecond pulsar

. Nature Astronomy 4, 72-76 (2019). https://doi.org/10.1038/s41550-019-0880-2
Baidu ScholarGoogle Scholar
35. E. Fonseca, H.T. Cromartie, T.T. Pennucci et al.,

Refined mass and geometric measurements of the high-mass PSR J0740+6620

. Astrophys. J. Lett. 915, L12 (2021). https://doi.org/10.3847/2041-8213/ac03b8
Baidu ScholarGoogle Scholar
36. R. Abbott et al. (LIGO Scientifc, Virgo),

GW190814: Gravitational waves from the coalescence of a 23 solar mass black hole with a 2.6 solar mass compact object

. Astrophys. J. Lett. 896, L44 (2020). https://doi.org/10.3847/2041-8213/ab960f
Baidu ScholarGoogle Scholar
37. B.A. Li, P.G. Krastev, D.H. Wen et al.,

Towards understanding astrophysical effects of nuclear symmetry energy

. Euro. Phys. J. A 55, 117 (2019). https://doi.org/10.1140/epja/i2019-12780-8
Baidu ScholarGoogle Scholar
38. F.J. Fattoyev, C.J. Horowitz, J. Piekarewicz et al.,

GW190814: Impact of a 2.6 solar mass neutron star on the nucleonic equations of state

. Phys.Rev.C 102, 065805 (2020). https://doi.org/10.1103/PhysRevC.102.065805
Baidu ScholarGoogle Scholar
39. M. Miller, F.K. Lamb, A.J. Dittmann et al.,

The radius of PSR J0740+6620 from NICER and XMM-Newton Data

. Astrophys. J. 918, L28 (2021). https://doi.org/10.3847/2041-8213/ac089b
Baidu ScholarGoogle Scholar
40. S.P. Tang, J.L. Jiang, M.Z. Han et al.,

Constraints on the phase transition and nuclear symmetry parameters from PSR J0740+6620 and multimessenger data of other neutron stars

. Phys. Rev. D 104, 063032. (2021). https://doi.org/10.1103/PhysRevD.104.063032
Baidu ScholarGoogle Scholar
41. B.P. Abbott, R. Abbott, T.D. Abbott et al.,

Properties of the binary neutron star merger GW170817

. Phys. Rev. X 9, 011001 (2019). https://doi.org/10.1103/PhysRevX.9.011001
Baidu ScholarGoogle Scholar
42. B.P. Abbott et al,

GW170817: 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
43. M.C. Miller, F.K. Lamb, A.J. Dittmann et al.,

Astrophys. PSR J0030+0451 Mass and Radius from NICER Data and Implications for the Properties of Neutron Star Matter

. Astrophys. J. Lett. 887, L24 (2019). https://doi.org/10.3847/2041-8213/ab50c5
Baidu ScholarGoogle Scholar
44. W.J. Xie and B.A. Li,

Bayesian inference of the symmetry energy of superdense neutron-rich matter from future radius measurements of massive neutron stars

. Astrophys. J. Lett. 899, 4 (2020). https://doi.org/10.3847/1538-4357/aba271
Baidu ScholarGoogle Scholar
45. Y.X. Zhang, M. Liu, C.J. Xia et al.,

Constraints on the symmetry energy and its associated parameters from nuclei to neutron stars

. Phys. Rev. C 101, 034303 (2020). https://doi.org/10.1103/PhysRevC.101.034303
Baidu ScholarGoogle Scholar
46. S.K. Greif, K. Hebeler, J.M. Lattimer et al.,

Equation of state constraints from nuclear physics, neutron star masses, and future moment of inertia measurements

. Astrophys. J. Lett. 901, 155(2020). https://doi.org/10.3847/1538-4357/abaf55
Baidu ScholarGoogle Scholar
47. C.Y. Tsang, M.B. Tsang, W.G. Lynch et al.,

Determination of the equation of state from nuclear experiments and neutron star observations

. Nature Astronomy 8, 328 (2024). https://doi.org/10.1038/s41550-023-02161-z
Baidu ScholarGoogle Scholar
48. C.Y. Tsang, M.B. Tsang, P. Danielewicz et al.,

Insights on Skyrme parameters from GW170817

. Phys. Lett. B 796, 1-5 (2019). https://doi.org/10.1016/j.physletb.2019.05.055
Baidu ScholarGoogle Scholar
49. B.A Li, B.J. Cai, W.J. Xie et al.,

Progress in constraining nuclear symmetry energy using neutron star observables since GW170817

. Universe 7, 182 (2021). https://doi.org/10.3390/universe7060182
Baidu ScholarGoogle Scholar
50. C.Y. Tsang, M.B. Tsang, Pawel Danielewicz et al.,

Impact of the neutron-star deformability on equation of state parameters

. Phys. Rev. C 102, 045808 (2020). https://doi.org/10.1103/PhysRevC.102.045808
Baidu ScholarGoogle Scholar
51. P.G. Reinhard,

The relativistic mean field description of nuclei and nuclear dynamics

. Rep. Prog. Phys. 52, 439 (1989). https://doi.org/10.1088/0034-4885/52/4/002
Baidu ScholarGoogle Scholar
52. P. Ring,

Relativistic mean field in finite nuclei

. Prog. Part. Nucl. Phys 37, 193 (1996). https://doi.org/10.1016/0146-6410(96)00054-3
Baidu ScholarGoogle Scholar
53. M. Dutra, O. Lourenço, S.S. Avancini et al.,

Relativistic mean-field hadronic models under nuclear matter constraints

. Phys. Rev. C 90, 055203 (2014). https://doi.org/10.1103/PhysRevC.90.055203
Baidu ScholarGoogle Scholar
54. B.D. Serot and J.D. Walecka,

Recent progress in quantum hadrodynamics

. Int. J. Mod. Phys. E 6, 515 (1997). https://doi.org/10.1142/S0218301397000299
Baidu ScholarGoogle Scholar
55. T. Bürvenich, D.G. Madland, J.A. Maruhn et al.,

Nuclear ground state observables and QCD scaling in a refined relativistic point coupling model

. Phys. Rev. C 65, 044308 (2002). https://doi.org/10.1103/PhysRevC.65.044308
Baidu ScholarGoogle Scholar
56. C. Fuchs, H. Lenske and H.H. Wolter,

Density dependent hadron field theory

. Phys. Rev. C 52, 3043 (1995). https://doi.org/10.1103/PhysRevC.52.3043
Baidu ScholarGoogle Scholar
57. J.J. Rusnak and R.J. Furnstahl,

Relativistic point coupling models as effective theories of nuclei

. Nucl. Phys. A 627, 495 (1997). https://doi.org/10.1016/S0375-9474(97)00598-8
Baidu ScholarGoogle Scholar
58. G.A. Lalazissis, T. Nikšić, D. Vretenar et al.,

New relativistic mean-field interaction with density-dependent meson-nucleon couplings

. Phys. Rev. C 71, 024312 (2005). https://doi.org/10.1103/PhysRevC.71.024312
Baidu ScholarGoogle Scholar
59. X. Roca-Maza, X. Viñas, M. Centelles et al.,

Relativistic mean field interaction with density dependent meson-nucleon vertices based on microscopical calculations

. Phys. Rev. C 84, 054309 (2011). https://doi.org/10.1103/PhysRevC.84.054309
Baidu ScholarGoogle Scholar
60. J. Piekarewicz,

Correlating the giant monopole resonance to the nuclear matter incompressibility

. Phys. Rev. C. 66, 034305 (2002). https://doi.org/10.1103/PhysRevC.66.034305
Baidu ScholarGoogle Scholar
61. M. Rufa, P.G. Reinhard, J.A. Maruhn et al.,

Optimal parametrization for the relativistic mean-field model of the nucleus

. Phys. Rev. C 38, 390 (1988). https://doi.org/10.1103/PhysRevC.38.390
Baidu ScholarGoogle Scholar
62. G.A. Lalazissis, J. Karatzikos, and P. Ring,

A New parametrization for the Lagrangian density of relativistic mean field theory

. Phys. Rev. C 55, 540 (1997). https://doi.org/10.1103/PhysRevC.55.540
Baidu ScholarGoogle Scholar
63. G.A. Lalazissis S. Karatzikos, R. Fossion et al.,

The effective force NL3 revisited

. Phys. Lett. B 671, 36 (2009). https://doi.org/10.1016/j.physletb.2008.11.070
Baidu ScholarGoogle Scholar
64. B. Nerlo-Pomoroska and J. Sykut,

A New parameter set for the relativistic mean field theory

. Int. J. Mod. Phys. E 13, 75 (2004). https://doi.org/10.1142/S0218301304001758
Baidu ScholarGoogle Scholar
65. M. Rashdan,

Structure of exotic nuclei and superheavy elements in a relativistic shell model

. Phys. Rev. C 63, 044303 (2001). https://doi.org/10.1103/PhysRevC.63.044303
Baidu ScholarGoogle Scholar
66. P.G. Reinhard,

The nonlinearity of the scalar field in a relativistic mean-field theory of the nucleus

. Z. Phys. A 329, 257 (1988). https://doi.org/10.1007/BF01290231
Baidu ScholarGoogle Scholar
67. A. Sulaksono, T. Mart, and C. Bahri,

Nilsson parameters kappa and mu in the relativistic mean field models

. Phys. Rev. C 71, 034312 (2005). https://doi.org/10.1103/PhysRevC.71.034312
Baidu ScholarGoogle Scholar
68. N.K. Glendenning, Compact stars: nuclear physics, particle physics, and general relativity, 2nd ed. (Springer, New York, 2000).
69. N.K. Glendenning,

The hyperon composition Of neutron stars

. Phys. Lett. B 114, 392 (1982). https://doi.org/10.1016/0370-2693(82)90078-8
Baidu ScholarGoogle Scholar
70. N.K. Glendenning and S.A. Moszkowski,

Reconciliation of neutron star masses and binding of the lambda in hypernuclei

. Phys. Rev. Lett. 67, 2414 (1991). https://doi.org/10.1103/PhysRevLett.67.2414
Baidu ScholarGoogle Scholar
71. S.K. Ghosh, S.C. Phatak, and P.K. Sahu,

Quark hadron phase transition and hybrid stars

. Z. Phys. A 352, 457 (1995). https://doi.org/10.1007/BF01299764
Baidu ScholarGoogle Scholar
72. B. Liu, H. Guo, M. Di Toro et al.,

Neutron stars with the isovector scalar field

. Eur. Phys. J. A 25, 293 (2005). https://doi.org/10.1140/epja/i2005-10095-1
Baidu ScholarGoogle Scholar
73. A.A. Dadi,

Parametrization of the relativistic (σ−ω) model for nuclear matter

Phys. Rev. C 82, 025203 (2010). https://doi.org/10.1103/PhysRevC.82.025203
Baidu ScholarGoogle Scholar
74. R.J. Furnstahl, B.D. Serot, and H.B. Tang,

A chiral effective lagrangian for nuclei

. Nucl. Phys. A 615, 441 (1997). https://doi.org/10.1016/S0375-9474(96)00472-1
Baidu ScholarGoogle Scholar
75. S. Gmuca,

Finite-nuclei calculations based on relativistic mean-field effective interactions

. Nucl. Phys. A 547, 447 (1992). https://doi.org/10.1016/0375-9474(92)90032-F
Baidu ScholarGoogle Scholar
76. C.J. Horowitz and J. Piekarewicz,

Neutron star structure and the neutron radius of Pb-208

. Phys. Rev. Lett. 86, 5647 (2001). https://doi.org/10.1103/PhysRevLett.86.5647
Baidu ScholarGoogle Scholar
77. M. Centelles X. Vinas, M. Barranco et al.,

Semiclassical approximations in nonlinear sigma omega models

. Nucl. Phys. A 537, 486 (1992). https://doi.org/10.1016/0375-9474(92)90365-Q
Baidu ScholarGoogle Scholar
78. J.K. Bunta and S. Gmuca,

Asymmetric nuclear matter in the relativistic mean field approach with vector cross interaction

. Phys. Rev. C 68, 054318 (2003). https://doi.org/10.1103/PhysRevC.68.054318
Baidu ScholarGoogle Scholar
79. H. Müller and B.D. Serot,

Phase transitions in warm, asymmetric nuclear matter

. Phys. Rev. C 52, 2072 (1995). https://doi.org/10.1103/PhysRevC.52.2072
Baidu ScholarGoogle Scholar
80. F.J. Fattoyev and J. Piekarewicz,

Relativistic models of the neutron-star matter equation of state

. Phys. Rev. C 82, 025805 (2010). https://doi.org/10.1103/PhysRevC.82.025805
Baidu ScholarGoogle Scholar
81. M.M. Sharma, A.R. Farhan, and S. Mythili,

Shell effects in nuclei with vector selfcoupling of omega meson in relativistic Hartree-Bogolyubov theory

. Phys. Rev. C 61, 054306 (2000). https://doi.org/10.1103/PhysRevC.61.054306
Baidu ScholarGoogle Scholar
82. Y. Sugahara and H. Toki,

Relativistic mean field theory for unstable nuclei with nonlinear sigma and omega terms

. Nucl. Phys. A 579, 557 (1994). https://doi.org/10.1016/0375-9474(94)90923-7
Baidu ScholarGoogle Scholar
83. W. Long, J. Meng, N. VanGiai et al.,

New effective interactions in RMF theory with nonlinear terms and density dependent meson nucleon coupling

. Phys. Rev. C 69, 034319 (2004). https://doi.org/10.1103/PhysRevC.69.034319
Baidu ScholarGoogle Scholar
84. J. Piekarewicz and M. Centelles,

Incompressibility of neutron-rich matter

. Phys. Rev. C 79, 054311 (2009). https://doi.org/10.1103/PhysRevC.79.054311
Baidu ScholarGoogle Scholar
85. A. Sulaksono and T. Mart,

Low densities instability of relativistic mean field models

. Phys. Rev. C 74, 045806 (2006). https://doi.org/10.1103/PhysRevC.74.045806
Baidu ScholarGoogle Scholar
86. B.K. Agrawal,

Asymmetric nuclear matter and neutron-skin in extended relativistic mean field model

. Phys. Rev. C 81, 034323 (2010). https://doi.org/10.1103/PhysRevC.81.034323
Baidu ScholarGoogle Scholar
87. J. Piekarewicz and S.P. Weppner,

Insensitivity of the elastic proton-nucleus reaction to the neutron radius of Pb-208

. Nucl. Phys.A 778, 10 (2006). https://doi.org/10.1016/j.nuclphysa.2006.08.004
Baidu ScholarGoogle Scholar
88. R. Kumar, B.K. Agrawal, and S.K. Dhiman,

Effects of omega meson self-coupling on the properties of finite nuclei and neutron stars

. Phys. Rev. C 74, 034323 (2006). https://doi.org/10.1103/PhysRevC.74.034323
Baidu ScholarGoogle Scholar
89. F.J. Fattoyev, C.J. Horowitz, J. Piekarewicz et al.,

Relativistic effective interaction for nuclei, giant resonances, and neutron stars

. Phys. Rev. C 82, 055803 (2010). https://doi.org/10.1103/PhysRevC.82.055803
Baidu ScholarGoogle Scholar
90. C.J. Horowitz and J. Piekarewicz,

Constraining URCA cooling of neutron stars from the neutron radius of Pb-208

. Phys. Rev. C 66, 055803 (2002). https://doi.org/10.1103/PhysRevC.66.055803
Baidu ScholarGoogle Scholar
91. M. DelEstal, M. Centelles, X. Viñas et al.,

Effects of new nonlinear couplings in relativistic effective field theory

. Phys. Rev. C 63, 024314 (2001). https://doi.org/10.1103/PhysRevC.63.024314
Baidu ScholarGoogle Scholar
92. S.K. Dhiman, R. Kumar, and B.K. Agrawal,

Non-rotating and rotating neutron stars in the extended field theoretical model

. Phys. Rev. C 76, 045801 (2007). https://doi.org/10.1103/PhysRevC.76.045801
Baidu ScholarGoogle Scholar
93. M.M. Sharma,

Scalar-vector Lagrangian without nonlinear self-interactions of bosonic fields in the relativistic mean-field theory

. Phys. Lett. B 666, 140 (2008). https://doi.org/10.1016/j.physletb.2008.07.005
Baidu ScholarGoogle Scholar
94. M.M. Haidari and M.M. Sharma,

Sigma-omega meson coupling and properties of nuclei and nuclear matter

. Nucl. Phys. A 803, 159 (2008). https://doi.org/10.1016/j.nuclphysa.2008.02.296
Baidu ScholarGoogle Scholar
95. S. Typel and H.H. Wolter,

Relativistic mean field calculations with density dependent meson nucleon coupling

. Nucl. Phys. A 656, 331 (1999). https://doi.org/10.1016/S0375-9474(99)00310-3
Baidu ScholarGoogle Scholar
96. T. Nikšić, D. Vretenar, P. Finelli et al.,

Relativistic Hartree-Bogolyubov model with density dependent meson nucleon couplings

. Phys. Rev. C 66, 024306 (2002). https://doi.org/10.1103/PhysRevC.66.024306
Baidu ScholarGoogle Scholar
97. S. Typel,

Relativistic model for nuclear matter and atomic nuclei with momentum-dependent self-energies

. Phys. Rev. C 71, 064301 (2005). https://doi.org/10.1103/PhysRevC.71.064301
Baidu ScholarGoogle Scholar
98. T. Klähn, D. Blaschke, S. Typel et al.,

Constraints on the high-density nuclear equation of state from the phenomenology of compact stars and heavy-ion collisions

. Phys. Rev. C 74, 035802 (2006). https://doi.org/10.1103/PhysRevC.74.035802
Baidu ScholarGoogle Scholar
99. S. Typel, G. Röpke, T. Klähn et al.,

Composition and thermodynamics of nuclear matter with light clusters

. Phys. Rev. C 81, 015803 (2010). https://doi.org/10.1103/PhysRevC.81.015803
Baidu ScholarGoogle Scholar
100. G. Baym, C. Pethick, and P. Sutherland,

The Ground state of matter at high densities: Equation of state and stellar models

. Astrophys. J. 170, 299 (1971). https://doi.org/10.1086/151216
Baidu ScholarGoogle Scholar
101. B. Link, R.I. Epstein, J.M. Lattimer,

Pulsar constraints on neutron star structure and equation of state

. Phys. Rev. Lett. 83, 3362 (1999). https://doi.org/10.1103/PhysRevLett.83.3362
Baidu ScholarGoogle Scholar
102. J. Carriere, C.J. Horowitz, J. Peikarewicz,

Low mass neutron stars and the equation of state of dense matter

. Astrophys. J. 593, 463 (2003). https://doi.org/10.1086/376515
Baidu ScholarGoogle Scholar
103. C. Ducoin, J. Marugueron, C. Providência et al.,

Core-crust transition in neutron stars: predictivity of density developments

. Phys. Rev. C 83, 045810 (2011). https://doi.org/10.1103/PhysRevC.83.045810
Baidu ScholarGoogle Scholar
104. F. Weber,

Strange quark matter and compact stars

. Prog. Part. Nucl. Phys. 54, 193 (2005). https://doi.org/10.1016/j.ppnp.2004.07.001
Baidu ScholarGoogle Scholar
105. A. Li, G.F. Burgio, U. Lombardo et al.,

Exotic phases in neutron stars

. Int. J. Mod. Phys. E 17, 1635 (2008). https://doi.org/10.1142/S0218301308010659
Baidu ScholarGoogle Scholar
106. M. Orsaria, H. Rodrigues, F. Weber et al.,

Quark deconfinement in high-mass neutron stars

. Phys. Rev. C 89, 015806(2014). https://doi.org/10.1103/PhysRevC.89.015806
Baidu ScholarGoogle Scholar
107. A. Gal, E.V. Hungerford, D.J. Millener,

Strangeness in nuclear physics

. Rev. Mod. Phys. 88(3), 035004 (2016). https://doi.org/10.1103/RevModPhys.88.035004
Baidu ScholarGoogle Scholar
108. L. Tolos, L. Fabbietti,

Strangeness in nuclei and neutron stars

. Prog. Part. Nucl. Phys. 112, 103770 (2020). https://doi.org/10.1016/j.ppnp.2020.103770
Baidu ScholarGoogle Scholar
109. M. Oertel, M. Hempel, T. Klähn et al.,

Equations of state for supernovae and compact stars

. Rev. Mod. Phys. 89(1), 015007 (2017). https://doi.org/10.1103/RevModPhys.89.015007
Baidu ScholarGoogle Scholar
110. A. Li, F. Huang, R.X. Xu,

Too massive neutron stars: The role of dark matter?

. Astropart.Phys. 37, 70 (2012). https://doi.org/10.1016/j.astropartphys.2012.07.006
Baidu ScholarGoogle Scholar
111. H.S. Sahoo, G. Mitra, R. Mishra et al.,

Neutron star matter with Δ isobars in a relativistic quark model

. Phys. Rev. C 98, 045801 (2018). https://doi.org/10.1103/PhysRevC.98.045801
Baidu ScholarGoogle Scholar
112. H.C. Das, A. Kumar, B. Kumar et al.,

Effects of dark matter on the nuclear and neutron star matter

. Mon. Not. Roy. Astron. Soc., 495, 4893 (2020). https://doi.org/10.1093/mnras/staa1435
Baidu ScholarGoogle Scholar
113. D. Lonardoni, A. Lovato, S. Gandolfi et al.,

Hyperon puzzle: hints from quantum Monte Carlo calculations

. Phys. Rev. Lett. 114, 092301 (2015). https://doi.org/10.1103/PhysRevLett.114.092301
Baidu ScholarGoogle Scholar
114. Z.Q. Feng,

Extracting the hyperon-nucleon interaction via collective flows in heavy-ion collisions

. Phys. Lett. B 851, 138580(2024). https://doi.org/10.1016/j.physletb.2024.138580
Baidu ScholarGoogle Scholar
115. S.N. Wei, Z.Q. Feng, W.Z. Jiang,

Correlation of the hyperon potential stiffness with hyperon constituents in neutron stars and heavy-ion collisions

. Phys. Lett. B 853 138658(2024). https://doi.org/10.1016/j.physletb.2024.138658
Baidu ScholarGoogle Scholar
116. J.S. Read, B.D. Lackey, B.J. Owen et al.,

Constraints on a phenomenologically parameterized neutron-star equation of state

. Phys. Rev. D. 79, 124032 (2009). https://doi.org/10.1103/PhysRevD.79.124032
Baidu ScholarGoogle Scholar
117. J.M. Lattimer, 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
118. F.J. Fattoyev, J. Carvajal, W.G. Newton et al.,

Constraining the high-density behavior of the nuclear symmetry energy with the tidal polarizability of neutron stars

. Phys. Rev. C 87 015806 (2013). https://doi.org/10.1103/PhysRevC.87.015806
Baidu ScholarGoogle Scholar
119. R.C. Tolman, Relativity, thermodynamics and cosmology (Oxford University Press, 1934).
120. R.C. Tolman,

Static solutions of Einstein’s field equations for spheres of fluid

. Phys. Rev. 55, 364 (1939). https://doi.org/10.1103/PhysRev.55.364
Baidu ScholarGoogle Scholar
121. J.R. Oppenheimer and G.M. Volkoff,

On massive neutron cores

. Phys. Rev. 55, 374 (1939). https://doi.org/10.1103/PhysRev.55.374
Baidu ScholarGoogle Scholar
122. T. Damour and A. Nagar,

Relativistic tidal properties of neutron stars

. Phys. Rev. D 80, 084035 (2009) https://doi.org/10.1103/PhysRevD.80.084035
Baidu ScholarGoogle Scholar
123. T. Hinderer, B.D. Lackey, R.N. Lang et al.,

Tidal deformability of neutron stars with realistic equations of state and their gravitational wave signatures in binary inspiral

. Phys. Rev. D 81, 123016 (2010). https://doi.org/10.1103/PhysRevD.81.123016
Baidu ScholarGoogle Scholar
124. S. Postnikov, M. Prakash, J.M. Lattimer,

Tidal love numbers of neutron and self-bound quark stars

. Phys. Rev. D 82, 024016(2010). https://doi.org/10.1103/PhysRevD.82.024016
Baidu ScholarGoogle Scholar
125. R.J. Furnstahl,

Neutron radii in mean field models

. Nucl. Phys. A 706, 85 (2002). https://doi.org/10.1016/S0375-9474(02)00867-9
Baidu ScholarGoogle Scholar
126. C. Ducoin, J. Margueron, and C. Providencia,

Nuclear symmetry energy and core-crust transition in neutron stars: a critical study

. Europhys. Lett. 91, 32001 (2010). https://doi.org/10.1209/0295-5075/91/32001
Baidu ScholarGoogle Scholar
127. Z. Zhang and L.W. Chen,

Constraining the symmetry energy at subsaturation densities using isotope binding energy difference and neutron skin thickness

. Phys. Lett. B 726, 234 (2013). https://doi.org/10.1016/j.physletb.2013.08.002
Baidu ScholarGoogle Scholar
128. C.J. Horowitz, E.F. Brown, Y. Kim et al.,

A way forward in the study of the symmetry energy: experiment, theory, and observation

. J. Phys. G 41, 093001 (2014). https://doi.org/10.1088/0954-3899/41/9/093001
Baidu ScholarGoogle Scholar
129. T. Malik, M. Ferreira, B.K. Agrawal et al.,

Relativistic description of dense matter equation of state and compatibility with neutron star observables: a bayesian approach

. Astrophys.J. 930, 17 (2022). https://doi.org/10.3847/1538-4357/ac5d3c
Baidu ScholarGoogle Scholar
130. J. M. Lattimer, A.W. Steiner,

Constraints on the symmetry energy using the mass-radius relation of neutron stars

. Eur. Phys. J. A 50, 40 (2014). https://doi.org/10.1140/epja/i2014-14040-y
Baidu ScholarGoogle Scholar
131. J.M. Lattimer and Y. Lim,

Constraining the symmetry parameters of the nuclear interaction

. Astrophys. J. 771, 51 (2013). https://doi.org/10.1088/0004-637X/771/1/51
Baidu ScholarGoogle Scholar
132. 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
133. B.A. Brown,

Constraints on the skyrme equations of state from properties of doubly magic nuclei

. Phys. Rev. Lett. 111, 232502 (2013). https://doi.org/10.1103/PhysRevLett.111.232502
Baidu ScholarGoogle Scholar
134. P. Danielewicz, and J. Lee,

Symmetry energy II: isobaric analog states

. Nucl. Phys. A 922, 1-70 (2014). https://doi.org/10.1016/j.nuclphysa.2013.11.005
Baidu ScholarGoogle Scholar
135. Z. Zhang and L.W. Chen,

Constraining the density slope of nuclear symmetry energy at subsaturation densities using electric dipole polarizability in 208Pb

. Phys. Rev. C 90, 064317 (2014). https://doi.org/10.1103/PhysRevC.90.064317
Baidu ScholarGoogle Scholar
136. C. Drischler, K. Hebeler, and A. Schwenk,

Chiral interactions up to next-to-next-to-next-to-leading order and nuclear saturation

. Phys. Rev. Lett. 122, 042501 (2019). https://doi.org/10.1103/PhysRevLett.122.042501
Baidu ScholarGoogle Scholar
137. C. Drischler, J.A. Melendez, R.J. Furnstahl et al.,

Quantifying uncertainties and correlations in the nuclear-matter equation of state

. Phys. Rev. C 102, 054315 (2020). https://doi.org/10.1103/PhysRevC.102.054315
Baidu ScholarGoogle Scholar
138. W.G. Lynch, M.B. Tsang, Y. Zhang et al.,

Probing the Symmetry Energy with Heavy Ions

. Prog. Part. Nucl. Phys. 62, 427 (2009). https://doi.org/10.1016/j.ppnp.2009.01.001
Baidu ScholarGoogle Scholar
139. D.H. Youngblood, H.L. Clark, Y.W. Lui,

Incompressibility of Nuclear Matter from the Giant Monopole Resonance

. Phys. Rev. Lett. 82, 691 (1999). https://doi.org/10.1103/PhysRevLett.82.691
Baidu ScholarGoogle Scholar
Footnote

The authors declare that they have no competing interests.