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

Special Section on the Celebration for 55 Years’ Dedicated Research on Heavy-Ion Physics of Natowitz and his 80th Birthday

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

Bao-Jun Cai
Lie-Wen Chen
Nuclear Science and TechniquesVol.28, No.12Article number 185Published in print 01 Dec 2017Available online 05 Dec 2017
7902

Within the nonlinear relativistic mean field (NL-RMF) model, we show that both the pressure of symmetric nuclear matter at supra-saturation densities and the maximum mass of neutron stars are sensitive to the skewness coefficient, J0, of symmetric nuclear matter. Using experimental constraints on the pressure of symmetric nuclear matter at supra-saturation densities from flow data in heavy ion collisions and the astrophysical observation of a large mass neutron star PSR J0348+0432, with the former favoring a smaller J0 while the latter favors a larger J0, we extract a constraint of -494 MeV≤J0≤-10 MeV based on the NL-RMF model. This constraint is compared with the results obtained in other analyses.

Equation of state of nuclear matterHeavy-ion collisionsNeutron stars

1 Introduction

Determination of the equation of state (EOS) of asymmetric nuclear matter (ANM) is one of fundamental questions in contemporary nuclear physics and astrophysics. The exact knowledge on the EOS of ANM provides important information on the in-medium nuclear effective interactions which play a central role in understanding the structure and decay properties of finite nuclei as well as the related dynamical problems in nuclear reactions ([1-16]). The EOS of ANM also plays a decisive role in understanding a number of important issues in astrophysics including the structure and evolution of neutron stars as well as the mechanism of supernova explosion ([17-23]). Conventionally, the EOS of ANM is given by the binding energy per nucleon as functions of nucleon density, ρ, and isospin asymmetry, δ, i.e., E(ρ, δ), and some bulk characteristic parameters defined at the saturation density, ρ0, of symmetric nuclear matter (SNM) are usually introduced to quantitatively characterize the EOS of ANM. For example, the energy, E0(ρ0), and incompressibility, K0, of SNM, as well as the symmetry energy, Esym(ρ0), and its slope parameter, L, are the four famous lower-order bulk characteristic parameters of EOS of ANM. These bulk parameters defined at ρ0 provide important information on both sub- and supra-saturation density behaviors of the EOS of ANM ([24, 25]).

Based on the empirical liquid-drop-like model analyses of high precision data about nuclear masses, the E0(ρ0) is well known to be about -16 MeV. The incompressibility has been determined to be K0 = 240 ± 40 MeV from analyzing experimental data of nuclear giant monopole resonances (GMR) ([1, 26-30]). For Esym(ρ0) and L, the existing constraints extracted from terrestrial laboratory measurements and astrophysical observations are found to be essentially consistent with Esym(ρ0) = 32.5 ± 2.5 MeV and L = 55 ± 25 MeV (see, e.g., Refs. [31, 32]). While these lower-order bulk characteristic parameters have been relatively well determined or in significant progress, our knowledge on the higher-order bulk characteristic parameters remains very limited. Following E0(ρ0), K0, Esym(ρ0), and L, the next bulk characteristic parameter should be the skewness coefficient, J0, (also denoted as K’ or Q0 in some literature) of SNM, which is related to the third-order density derivative of the binding energy per nucleon of SNM at ρ0. The higher-order bulk characteristic parameter J0 is expected to be important for the high density behaviors of nuclear matter EOS and thus may play an essential role in heavy ion collisions (HIC), the structure and evolution of neutron stars, supernova explosion, and gravitational wave radiation from merging of compact stars. To our best knowledge, there is very little experimental information on the J0 parameter, and it is thus of great interest and critical importance to constrain the J0 parameter, which is the main motivation of the present work.

Within the nonlinear relativistic mean field (RMF) model, we demonstrate in this work that the pressure of SNM at supra-saturation densities and the maximum mass of neutron stars provide good probes of the skewness coefficient, J0. In particular, combining the experimental constraints on the pressure of SNM at supra-saturation densities from flow data in HIC and the recent astrophysical observation of a large mass neutron star, PSR J0348+0432, one can obtain a strong constraint on the J0 parameter.

2 The skewness coefficient J0 in nonlinear RMF model

2.1 Nuclear matter characteristic parameters

The EOS of isospin asymmetric nuclear matter, namely E(ρ,δ), can be expanded as a power series of even-order terms in δ as

E(ρ,δ)E0(ρ)+Esym(ρ)δ2+O(δ4), (1)

where E0(ρ)=E(ρ,δ =0) is the EOS of symmetric nuclear matter, and the symmetry energy is expressed as

Esym(ρ)=122E(ρ,δ)δ2|δ=0. (2)

Around the saturation density, ρ0, the E0(ρ) can be expanded, e.g., up to 3rd-order in density, as,

E0(ρ)=E0(ρ0)+K02!χ2+J03!χ3+O(χ4), (3)

where χ =(ρ -ρ0)/3ρ0 is a dimensionless variable characterizing the deviations of the density from the saturation density, ρ0. The first term E0(ρ0) on the right-hand-side of Eq. (3) is the binding energy per nucleon in SNM at ρ0 and the coefficients of other terms are

K0=9ρ02d2E0(ρ)dρ2|ρ=ρ0, (4) J0=27ρ03d3E0(ρ)dρ3|ρ=ρ0, (5)

where K0 is the well-known incompressibility coefficient of SNM and J0 is the skewness coefficient of SNM, i.e. the 3rd-order incompressibility coefficient of SNM ([24, 25]).

Similarly, one can expand the Esym(ρ) around an arbitrary reference density, ρr, as

Esym(ρ)=Esym(ρr)+L(ρr)χr+O(χr2), (6)

with χr=(ρ -ρr)/3ρr, and the slope parameter of the symmetry energy at ρr is expressed as [33]

L(ρr)=3ρrdEsym(ρ)dρ|ρ=ρr. (7)

For ρr = ρ0, the L(ρr) is reduced to the conventional slope parameter L3ρ0dEsym(ρ)/dρ|ρ=ρ0.

If δ and χ are assumed to be small quantities on the same order, nuclear matter bulk characteristic parameters can then be classified accordingly in different orders. For example, L and J0 are on the same order-3, i.e., δ2χ for L and δ0χ3 for J0. In this sense, E0(ρ0) is on the order-0, and, K0 and Esym(ρ0) are on the order-2. To see the role of J0 in the EOS of SNM, one can re-write Eq. (3) in a slightly different form as

E0(ρ)E0(ρ0)+12K0χ2(1+χJ03K0). (8)

Assuming J0 has roughly the same magnitude as K0, one can see that the contribution from the J0 term to the EOS of SNM becomes comparable with that from the K0 term if the baryon density is larger than about 3ρ0, corresponding to the typical densities inside a neutron star. On the other hand, the J0 term plays a minor role for the EOS of SNM at subsaturation densities relevant for nuclear structure properties. As we will see later, the pressure of SNM at supra-saturation densities and the maximum mass of neutron stars indeed display strong sensitivity on the J0 parameter.

2.2 Nuclear matter characteristic parameters in nonlinear RMF model

The nonlinear RMF model has made great success during the last decades in describing many nuclear phenomena (see, e.g., [34-46]). In the following, we briefly describe the nonlinear RMF model that we shall adopt in this work and present some useful expressions of nuclear matter characteristic parameters, especially the skewness coefficient, J0. The interacting Lagrangian of the nonlinear RMF model supplemented with couplings between the isoscalar and the isovector mesons reads ([47-51])

L=ψ¯[γμ(iμgωωμgρρμτ)(Mgσσ)]ψ12mσ2σ2+12μσμσU(σ)+12mω2ωμωμ14ωμνωμν+14cω(gωωμωμ)2+12mρ2ρμρμ14ρμνρμν+12gρ2ρμρμΛVgω2ωμωμ, (9)

where ωμνμωννωμ and ρμνμρννρμ are are strength tensors for ω field and ρ field, respectively. ψ, σ, ωμ, ρμ are nucleon field, isoscalar-scalar field, isoscalar-vector field, and isovector-vector field, respectively, and the arrows denote the vector in isospin space; U(σ)=bσM(gσσ)3/3+(gσσ)4/4 is the self interaction term for σ field. ΛV represents the coupling constant between the isovector ρ meson and the isoscalar ω meson, and it is important for the description of the density dependence of the symmetry energy. In addition, M is the nucleon mass and , , are masses of mesons.

In the mean field approximation, after neglecting effects of fluctuation and correlation, meson fields are replaced by their expectation values, i.e., σ¯σ, ω¯0ωμ, ρ¯0(3)ρμ, where subscript “0" indicates zeroth component of the four-vector, superscript “ (3)" indicates third component of the isospin. Furthermore, we also use in this work the non-sea approximation which neglects the effect due to negative energy states in the Dirac sea. The mean field equations are then expressed as

mσ2σ¯=gσ[ρSbσM(gσσ¯)2cσ(gσσ¯)3], (10) mω2ω¯0=gω[ρcω(gωω¯0)3ΛVgωω¯0(gρρ¯0(3))2], (11) mρ2ρ¯0(3)=gρ[ρpρnΛVgρρ¯0(3)(gωω¯0)2], (12)

where

ρ=ψ¯γ0ψ=ρn+ρp, ρS=ψ¯ψ=ρS,n+ρS,p, (13)

are the baryon density and scalar density, respectively, with the latter given by

ρS,J=2(2π)30kFJdkM|k|2+M2=M2π2[kFJEFJM2ln(kFJ+EFJM)],J=p,n. (14)

In the above expression, we have EFJ=kFJ2+M*2 and the nucleon Dirac mass is defined as

M*Mdirac=Mgσσ¯. (15)

kFJ=kF(1+τ3Jδ)1/3 is the Fermi momentum with τ3n=+1 for neutrons and τ3p=1 for protons, and kF=(3π2ρ /2)1/3 is the Fermi momentum for SNM at ρ.

The energy-momentum density tensor for the interacting Lagrangian density in Eq. (9) can be written as

Tμν=ψ¯iγμνψ+μσνσωμηνωηρμηνρηLgμν, (16)

where gμν=(+,-,-,-) is the Minkowski metric. In the mean field approximation, the mean value of the time (zero) component of the energy-momentum density tensor is the energy density of the nuclear matter system, i.e.,

ε=T00=εkinn+εkinp+12[mσ2σ¯2+mω2ω¯02+mρ2(ρ¯0(3))2]+13bσ(gσσ¯)3+14cσ(gσσ¯)4+34cω(gωω¯0)4+32(gρρ¯0(3))2ΛV(gωω¯0)2, (17)

where

εkinJ=2(2π)30kFJdk|k|2+M2=1π20kFJk2dkk2+M2=14[3EFJρJ+MρS,J], J=p,n, (18)

is the kinetic part of the energy density. Similarly, the mean value of space components of the energy-momentum density tensor corresponds to the pressure of the system, i.e.,

P=13j=13Tjj=Pkinn+Pkinp12[mσ2σ¯2mω2ω¯02mρ2(ρ¯0(3))2]13bσ(gσσ¯)314cσ(gσσ¯)4+14cω(gωω¯0)4+12(gρρ¯0(3))2ΛV(gωω¯0)2, (19)

where the kinetic part of pressure is given by

PkinJ=13π20kFJdkk4k2+M2, J=p,n. (20)

The EOS of ANM can be calculated through the energy density, ε(ρ,δ), by

E(ρ,δ)=ε(ρ,δ)ρM. (21)

The EOS of SNM is just E0(ρ)E(ρ,δ=0), and the characteristic parameters K0 and J0 can be obtained from the following expressions

K0(ρ)9ρ2d2E0dρ2=9ρgσ2M2QσEF2+9ρgω2Qω+3kF2EF6L0(ρ), (22) J0(ρ)27ρ3d3E0dρ3=3kF2EF3kF4EF3+27gσ2M2ρ2QσEF3×(3π22kFEF+2gσ2QσgσMηQσ22gσ2M2QσEF2+EFϕ2Qσ)162cωgω7ω¯0ρ2Qω39K0(ρ), (23)

with

L0(ρ)3ρdE0dρ=3[EF4MρS4ρ+gωω¯01ρ(12mσ2σ¯2+U(σ¯)+12mω2ω¯02+34cωgω4ω¯04)]. (24)

In the above expressions, we have

Qσ=mσ2+gσ2(3ρSM3ρEF)+2bσMgσ3σ¯+3cσgσ4σ¯2, (25) Qω=mω2+3cωgω4ω¯02, (26)

and

η=3gσ3(2ρSM23ρMEF+MρEF3)2bσMgσ36cσgσ4σ¯, (27) ϕ=2gσ2kF2EF3, (28)

with EF=(kF2+M2)1/2. To our best knowledge, Eq. (23) gives, for the first time [52], the analytical expression of the J0 parameter in the nonlinear RMF model. In addition, we would like to point out that the general expression for Esym(ρ) and L(ρ) in the nonlinear RMF model has been derived in Ref. [53].

3 Results and discussions

For the Lagrangian in Eq. (9), the properties of infinite nuclear matter is uniquely determined by =, , , =/, , =/, ΛV, and M. If the nucleon mass in vacuum is set to be M = 939 MeV, one then has seven total parameters to determine the properties of infinite nuclear matter in the nonlinear RMF model. Following the correlation analysis method proposed in Ref. [54] within the Skyrme-Hartree-Fock (SHF) approach, instead of directly using the seven microscopic parameters, i.e., , , , , , , and ΛV, one can determine their values explicitly in terms of seven macroscopic quantities, i.e., ρ0, E0(ρ0), K0, J0, Mdirac0Mdirac(ρ0), Esym(ρc), and L(ρc) where ρc is the cross density whose value is fixed in this work at 0.11 fm-3 ([33]). Then, by varying individually these macroscopic quantities within their known ranges, one can examine transparently the correlation of nuclear matter properties with each individual macroscopic quantity. Recently, this simple correlation analysis method has been successfully applied to study the neutron skin ([54, 33]) and giant resonances of finite nuclei ([30, 55]), the higher-order bulk characteristic parameters of ANM ([25]), and the relationship between the nuclear matter symmetry energy and the symmetry energy coefficient in the mass formula ([56]). We would like to point out although the seven macroscopic quantities defined above coherently act on the maximum mass of neutron stars and the pressure of the SNM, they are independent with each other in our analysis since we vary one quantity by keeping other six quantities fixed. This is one of the main advantages of our approach since the physics of these macroscopic quantities is different.

To examine the correlation of pressure of SNM at supra-saturation densities with each macroscopic quantity, we show in Fig. 1 the pressure of SNM P(ρ) at ρ=3ρ0 from the nonlinear RMF model based on the FSUGold interaction ([49]) by varying individually ρ0, E0(ρ0), Mdirac0, K0, and J0 within their empirical uncertain ranges, namely, varying one quantity at a time while keeping all others at their default values in FSUGold for which we have ρ0=0.148 fm-3, E0(ρ0)=-16.3 MeV, Mdirac0=0.61M, K0=230 MeV, J0=-522.6 MeV, Esym(ρc)=27.11 MeV, and L(ρc)=49.97 MeV. It should be mentioned that the pressure of SNM is independent of the values of Esym(ρc) and L(ρc). It is seen from Fig. 1 that the pressure of SNM P(ρ) at ρ=3ρ0 increases with ρ0, Mdirac0, K0, and J0 while it decreases with E0(ρ0). In particular, the pressure of SNM P(ρ) at ρ=3ρ0 displays a specially strong correlation with J0. We note that the pressure of SNM P(ρ) at other supra-saturation densities exhibits similar correlations with ρ0, E0(ρ0), Mdirac0, K0, and J0. These features indicate that the pressure of SNM at supra-saturation densities is sensitive to the J0 value, and thus the experimental constraints on the pressure of SNM at supra-saturation densities may provide important information on the J0 value.

Figure 1:
Pressure of SNM at ρ=3ρ0 from the nonlinear RMF model based on the FSUGold interaction by varying individually ρ0 (a), E0(ρ0) (b), Mdirac0 (c), K0 (d), and J0 (e).
pic

Since the pressure of SNM at supra-saturation densities is sensitive to the J0 value, the maximum mass, Mmax, of static neutron stars is also expected to be sensitive to the J0 value. The mass and radius of static neutron stars can be obtained from solving the Tolman-Oppenheimer-Volkoff (TOV) equations with a given neutron star matter EOS. A neutron star generally contains core, inner crust, and outer crust from the center to surface. In this work, for the core where the baryon density is larger than the core-curst transition density, ρt, we use the EOS of β-stable and charge neutral, npeμ matter obtained from the nonlinear RMF model. In the inner crust with densities between ρout and ρt where the nuclear pastas may exist, we construct its EOS (pressure, P, as a function of energy density, 𝒠) according to P=a+bE4/3 because of our poor knowledge about its EOS from first principle ([57, 58]). The ρout=2.46× 10-4 fm-3 is the density separating the inner from the outer crust. The constants a and b are then easily determined by the pressure and energy density at ρt and ρout [58]. In this work, the ρt is determined self-consistently within the nonlinear RMF model using the thermodynamical method (see, e.g., [51] for the details). In the outer crust with 6.93× 10-13 fm-3<ρ <ρout, we use the EOS of BPS ([59, 60]), and in the region of 4.73×10-15 fm-3<ρ <6.93× 10-13 fm-3, we use the EOS of FMT ([59]).

Similarly, as in Fig. 1, we plot in Fig. 2 the maximum mass, Mmax, of static neutron stars from the nonlinear RMF model based on the FSUGold interaction by varying individually ρ0, E0(ρ0), Mdirac0, K0, J0, Esym(ρc), and L(ρc) within their empirical uncertain ranges. Indeed, one can see that the Mmax displays a very strong positive correlation with the J0 parameter. In addition, the Mmax exhibits weak positive correlation with the Mdirac0 and K0, and weak negative correlation with the ρ0 and E0(ρ0). It is interesting to see that the Mmax is essentially independent of the values of Esym(ρc) and L(ρc), implying that, in the nonlinear RMF model, the Mmax is basically determined by the isoscalar part of the nuclear matter EOS. Since the seven microscopic parameters change with the macroscopic quantities, it is thus not surprising to see that the maximum mass of a neutron star based on the FSUGold interaction by varying macroscopic quantities may be totally different from the default one from FSUGold, which is about 1.74M. These features indicate that the observed largest mass of neutron stars may put important constraint on the J0 value. We would like to point out that the interaction FSUGold is only used in Figs. 1 and 2 as a reference for the correlation analyses and using other RMF interactions will not change our conclusion.

Figure 2:
(Color online) Maximum mass of static neutron stars from the nonlinear RMF model based on the FSUGold interaction by varying individually ρ0 (a), E0(ρ0) (b), Mdirac0 (c), K0 (d), J0 (e), Esym(ρc) (f), and L(ρc) (g).
pic

Experimentally, the pressure of SNM at supra-saturation densities (from 2ρ0 to about 5ρ0) has been constrained by measurements of collective flows in HIC ([3]), which is shown as a band in the left window of Fig. 3. In the nonlinear RMF model, if one only changes the J0 value while the other 6 macroscopic quantities are kept at their values in the FSUGold interaction, one can find that the J0 value should be in the range of -985 MeV ≤ J0≤ -327 MeV to be consistent with the flow data in HIC ([3]). However, keeping the other 6 macroscopic quantities at their values in the FSUGold interaction is obviously a strong assumption because the extraction of the J0 value from the flow data in HIC will also depends on the values of ρ0, E0(ρ0), Mdirac0, and K0, which can be varied within their empirical uncertain ranges. For the nonlinear RMF model, we use, in this work, the following empirical uncertain ranges for these macroscopic quantities, i.e., ρ0=0.153±0.008 fm-3, E0(ρ0)=-16.2±0.3 MeV, Mdirac*0/M=0.61±0.04, and K0=230±20 MeV, which represent the typical uncertain ranges known or predicted from different interactions in the nonlinear RMF model ([50]).

Figure 3:
(Color online) Left window: Pressure of SNM as a function of baryon density. The solid (dashed) line is the prediction from the nonlinear RMF model with J0=-10 (-1280) MeV and the set “S (H)" for ρ0, E0(ρ0), Mdirac0, and K0. The band represents the constraints from flow data in HIC [3]. Right window: The maximum mass of static neutron stars as a function of J0 in the nonlinear RMF model with the set “NS-H" for ρ0, E0(ρ0), Mdirac0, K0, Esym(ρc), and L(ρc). The band represents mass 2.01± 0.04M for PSR J0348+0432 [61].
pic

Based on the pressure of SNM constrained by flow data in HIC ([3]), to extract the upper limit of the J0 value, one should use the values of ρ0, E0(ρ0), Mdirac0, and K0 that make the resulting pressure of SNM as small as possible when J0 is fixed. This can be obtained by using ρ0=0.145 fm-3, E0(ρ0)=-15.9 MeV, Mdirac*/M=0.57, and K0=210 MeV, denoted as set “S", since the pressure of SNM P(ρ / ρ0) at supra-saturation densities increases with ρ0, Mdirac0, and K0 while decreases with E0(ρ0), as shown in Fig. 1. With the set “S" for ρ0, E0(ρ0), Mdirac0, and K0, one can find the upper limit of J0= -10 MeV for the J0 value, which is indicated by a solid line in the left window of Fig. 3. For J0 > -10 MeV, the model would over-predict the pressure of SNM constrained by flow data in HIC ([3]). Similarly, one can obtain the lower limit of the J0 value by using the values of ρ0, E0(ρ0), Mdirac0, and K0 that make the resulting pressure of SNM as large as possible when J0 is fixed, and this can be obtained with ρ0=0.161 fm-3, E0(ρ0)=-16.5 MeV, Mdirac0/M=0.65, and K0=250 MeV, denoted as set “H". Using the set “H" for ρ0, E0(ρ0), Mdirac0, and K0, one can extract the lower limit of J0= -1280 MeV, which is indicated by dashed line in the left window of Fig. 3. The model would under-predict the pressure of SNM constrained by flow data in HIC ([3]) if J0 < -1280 MeV. Therefore, from the pressure of SNM constrained by flow data in HIC ([3]), one can extract the constraint of -1280 MeV ≤ J0 ≤ -10 MeV.

Recently, a new neutron star, PSR J0348+0432, with a mass of 2.01±0.04M was discovered ([61]), and this neutron star is only the second pulsar with a precisely determined mass around 2M after PSR J1614-2230 ([62]) and sets a new record of the maximum mass of neutron stars. The lower mass limit of 1.97M for PSR J0348+0432 thus may set a lower limit of the J0 value below which the model cannot predict a neutron star with mass equal or above 1.97M. To extract the lower limit of the J0 value from the observed heaviest neutron star, PSR J0348+0432, one can use the values of ρ0, E0(ρ0), Mdirac0, K0, Esym(ρc), and L(ρc) that make the resulting maximum mass of neutron stars as large as possible when J0 is fixed, and from Fig. 2 this can be obtained with ρ0=0.145 fm-3, E0(ρ0)=-16.5 MeV, K0=250 MeV, Mdirac0/M=0.65, Esym(ρc)=26 MeV, and L(ρc)=60 MeV, denoted as set “NS-H". This leads to a lower limit of J0= -494 MeV for the J0 value as shown in the right window of Fig. 3 where the maximum mass of neutron stars is plotted as a function of J0 when the other 6 macroscopic quantities are fixed at their values as in set “NS-H". For J0 < -494 MeV, the maximum mass of static neutron stars predicted in the nonlinear RMF model would be always smaller than 1.97M. It should be pointed out that here the interior of neutron stars has been assumed to be npeμ matter. New degrees of freedom, such as hyperons or/and quark matter that could be present in the interior of neutron stars, usually soften the EOS of neutron star matter and thus a larger J0 value would be necessary to obtain a neutron star with mass of 1.97M. Therefore, including the new degrees of freedom in neutron stars will be consistent with the constraint of J0≥ -494 MeV.

Combining the constraint of -1280 MeV ≤ J0 ≤ -10 MeV from the pressure of SNM constrained by flow data in HIC ([3]) which favors a smaller J0 value and the constraint of J0≥ -494 MeV from the recently discovered heaviest neutron star PSR J0348+0432 ([61]) which favors a larger J0 value, one can extract the following constraint for the J0 parameter

494MeVJ010MeV. (29)

It should be emphasized that the constraint -494 MeV≤J0≤-10 MeV represents a conservative extraction based on flow data in HIC ([3]) and the recently discovered heaviest neutron star, PSR J0348+0432 ([61]) in the nonlinear RMF model. This is because we have not considered the possible correlations that existed among the ρ0, E0(ρ0), Mdirac0, K0, Esym(ρc), and L(ρc), and have simply set simultaneously their values in the boundary of their empirical uncertain ranges. Considering the correlations possibly existed among the ρ0, E0(ρ0), Mdirac0, K0, Esym(ρc) and L(ρc) should further narrow the constraint -494 MeV≤J0≤-10 MeV, and it will be interesting to see the quantitative constraint on the J0 parameter based on the data of finite nuclei, neutron stars, and heavy-ion collisions using the exhaustive statistical analysis method although this is beyond the scope of the present work. The conservative constraint -494 MeV≤J0≤-10 MeV obtained in the present work indicates that, if the J0 value is out of the region -494 MeV≤J0≤-10 MeV, the nonlinear RMF model either cannot predict the pressure of SNM constrained by flow data in HIC ([3]) or cannot describe the recently discovered heaviest neutron star PSR J0348+0432 ([61]). It is worth mentioning that the constraint on the J0 depends on our knowledge of the other six quantities. Any improvement on these six macroscopic quantities will make the range for the J0 constraint narrower. In addition, the extracted constraint on the J0 could depend on the form of the energy density functional, and it will be interesting to see how the constraint changes if other energy density functionals are used.

In Fig. 4, we show the comparison of the J0 constraint obtained in our analysis with those obtained with other analyses and/or other methods ([22, 25, 63, 64]), including the constraint of J0 = -700 ± 500 MeV obtained by Ref. [63] from the analysis of nuclear GMR, the constraint of J0=280410+72 (500290+170) MeV obtained by Ref. [64] from analyzing a heterogeneous data set of six neutron stars using a Markov chain Monte Carlo algorithm within a Bayesian framework by assuming rph>> R (rph= R) where rph is the photospheric radius at the time the flux is evaluated and R is the stellar radius, the constraint of J0 = -355 ± 95 MeV deduced by Ref. [25] based on a correlation analysis method within SHF energy density functional, and the constraint of J0 = -390 ± 90 MeV deduced by Ref. [22] who used a similar method as Ref. [25]. It is seen that the constrained region of J0 obtained in the present work has a remarkable overlap with those existing in the literature. In particular, our present constraint from the relativistic model is nicely consistent with the constraints deduced from the non-relativistic SHF approach, and they all indicate that the J0 parameter should be larger than about -500 MeV.

Figure 4:
(Color online) Comparison between the constraint of J0 extracted in the present work and those from the existing literature [22, 25, 63, 64].
pic

4 Summary

Within the nonlinear relativistic mean field model, using macroscopic nuclear matter characteristic parameters instead of the microscopic coupling constants as direct input quantities, we have demonstrated that the pressure of symmetric nuclear matter at supra-saturation densities and the maximum mass of neutron stars provide useful probes for the skewness coefficient, J0, of symmetric nuclear matter. In particular, using the existing experimental constraints on the pressure of symmetric nuclear matter at supra-saturation densities from flow data in heavy ion collisions and the astrophysical observation of recently discovered heaviest neutron star PSR J0348+0432, with the former requiring a smaller J0 while the latter requires a larger J0, we have extracted a constraint of -494 MeV≤J0≤-10 MeV.

We have compared the present constraint with the results obtained in other analyses, and found they are nicely in agreement. In particular, our present constraint from the relativistic model is nicely consistent with the constraints deduced from the non-relativistic Skyrme-Hartree-Fock approach, and they all indicate that the J0 parameter cannot be too small, namely, it should be larger than about -500 MeV. The present constraint on the J0 parameter provides important information on the high density behaviors of the EOS of symmetric nuclear matter and also may be potentially useful for the determination of the high density behaviors of the EOS of asymmetric nuclear matter, especially the high density symmetry energy.

References
[1] J.P. Blaizot,

Nuclear compressibilities

. Phys. Rep. 64, 171-248 (1980). doi: 10.1016/0370-1573(80)90001-0
Baidu ScholarGoogle Scholar
[2] B.A. Li, C.M. Ko, W. Bauer,

Isospin physics in heavy-Ion collisions at intermediate energies

. Int. J. Mod. Phys. E 7, 147 (1998). doi: 10.1142/S0218301398000087
Baidu ScholarGoogle Scholar
[3] P. Danielewicz, R. Lacey, W.G. Lynch,

Determination of the equation of state of dense matter

science, 298, 1592-1596 (2002). doi: 10.1126/science.1078070
Baidu ScholarGoogle Scholar
[4] V. Baran, M. Colonna, V. Greco, et al.,

Reaction dynamics with exotic nuclei

, Phys. Rep. 410, 335-466 (2005). doi: 10.1016/j.physrep.2004.12.004
Baidu ScholarGoogle Scholar
[5] A.W. Steiner, M. Prakash, J.M. Lattimer, et al.,

Isospin asymmetry in nuclei and neutron stars

. Phys. Rep. 411, 325-375 (2005). doi: 10.1016/j.physrep.2005.02.004
Baidu ScholarGoogle Scholar
[6] L.W. Chen, C.M. Ko, B.A. Li, et al.,

Probing the nucle- ar symmetry energy with heavy-ion reactions induced by neutron-rich nuclei

. Front. Phys. China 2, 327-357 (2007). doi: 10.1007/s11467-007-0037-0
Baidu ScholarGoogle Scholar
[7] B.A. Li, L.W. Chen, C.M. Ko,

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

. Phys. Rep. 464, 113-281 (2008). doi: 10.1016/j.physrep.2008.04.005
Baidu ScholarGoogle Scholar
[8] J. B. Natowitz, G. Röpke, S. Typel, et al.,

Symmetry energy of dilute warm nuclear matter

. Phys. Rev. Lett. 104, 202501 (2010). doi: 10.1103/PhysRevLett.104.202501
Baidu ScholarGoogle Scholar
[9] B.M. Tsang, J. R. Stone, F. Camera, et al.,

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

. Phys. Rev. C 86, 015803 (2012). doi: 10.1103/PhysRevC.86.015803
Baidu ScholarGoogle Scholar
[10] W. Trautmann, H.H. Wolter,

Elliptic flow and the symmetry energy at supra-saturation density

. Int. J. Mod. Phys. E 21, 1230003 (2012). doi: 10.1142/S0218301312300032
Baidu ScholarGoogle Scholar
[11] L.W. Chen, C.M. Ko, B.A. Li, et al.,

Probing isospin- and momentum-dependent nuclear effective interactions in neutron-rich matter

. Eur. Phys. J. A 50, 29 (2014). doi: 10.1140/epja/i2014-14029-6
Baidu ScholarGoogle Scholar
[12] 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). doi: 10.1088/0954-3899/41/9/093001
Baidu ScholarGoogle Scholar
[13] B.A. Li, A. Ramos, G. Verde, et al.,

Topical issue on nuclear symmetry energy

. Eur. Phys. J. A 50, 9 (2014). doi: 10.1140/epja/i2014-14009-x
Baidu ScholarGoogle Scholar
[14] X.Q. Liu, M.R. Huang, R. Wada, et al.,

Symmetry energy extraction from primary fragments in intermediate heavy-ion collisions

. Nucl. Sci. Tech. 26, S20508 (2015) doi: 10.13538/j.1001-8042/nst.26.S20508;
F. F. Duan, X.Q. Liu, W.P. Lin, et al.,

Investigation on symmetry and characteristic properties of the fragmenting source in heavy-ion reactions through reconstructed primary isotope yields

. Nucl. Sci. Tech. 27, 131 (2016). doi: 10.1007/s41365-016-0138-y
Baidu ScholarGoogle Scholar
[15] M. Baldo, G.F. Burgio,

The nuclear symmetry energy

. Prog. Part. Nucl. Phys. 91, 203-258 (2016). doi: 10.1016/j.ppnp.2016.06.006
Baidu ScholarGoogle Scholar
[16] B.A. Li,

Nucl. Phys. News, in press

, (2017) [arXiv:1701.03564]
Baidu ScholarGoogle Scholar
[17] N.K. Glendenning, Compact Stars, 2nd edition, Spinger-Verlag New York, Inc., 2000.
[18] J.M. Lattimer and M. Prakash,

The physics of neutron stars

. Science 304, 536-542 (2004). doi: 10.1126/science.1090720;
J.M. Lattimer and M. Prakash,

Neutron star observations: Prognosis for equation of state constraints

. Phys. Rep. 442, 109-165 (2007). doi: 10.1016/j.physrep.2007.02.003
Baidu ScholarGoogle Scholar
[19] J.M. Lattimer,

The nuclear equation of state and neutron star masses

. Annu. Rev. Nucl. Part. Sci. 62, 485-515 (2012). doi: 10.1146/annurev-nucl-102711-095018
Baidu ScholarGoogle Scholar
[20] K. Kotake, K. Sato, K. Takahashi,

Explosion mechanism, neutrino burst and gravitational wave in core-collapse supernovae

. Rep. Prog. Phys. 69, 971-1143 (2006) doi: 10.1088/0034-4885/69/4/R03;
H.-Th. Janka, K. Langanke, A. Marek, et al.,

Theory of core-collapse supernovae

. Phys. Rep. 442, 38-74 (2007). doi: 10.1016/j.physrep.2007.02.002
Baidu ScholarGoogle Scholar
[21] M. Hempel, T. Fischer, J. Schaffner-Bielich, et al.,

New equations of state in simulations of core-collapse supernovae

. Astrophys. J. 748, 70 (2012). doi: 10.1088/0004-637X/748/1/70
Baidu ScholarGoogle Scholar
[22] M. Meixner, J.P. Olson, G. Mathews, et al.,

The NDL equation of state for supernova simulations

. arXiv:1303.0064, (2013).
Baidu ScholarGoogle Scholar
[23] M. Oertel, M. Hempel, T. Klähn, et al.,

Equations of state for supernovae and compact stars

. Rev. Mod. Phys. 89, 015007 (2017). doi: 10.1103/RevModPhys.89.015007
Baidu ScholarGoogle Scholar
[24] L.W. Chen, B.J. Cai, C.M. Ko, et al.,

Higher-order effects on the incompressibility of isospin asymmetric nuclear matter

. Phys. Rev. C 80, 014322 (2009).doi: 10.1103/PhysRevC.80.014322
Baidu ScholarGoogle Scholar
[25] L.W. Chen,

Higher order bulk characteristic parameters of asymmetric nuclear matter

. Sci. China: Phys. Mech. Astron. 54, s124-s129 (2011). doi: 10.1007/s11433-011-4415-9
Baidu ScholarGoogle Scholar
[26] D.H. Youngblood, H.L. Clark, Y.-W. Lui,

Incompressibility of nuclear matter from the giant monopole resonance

. Phys. Rev. Lett. 82, 691-694 (1999). doi: 10.1103/PhysRevLett.82.691
Baidu ScholarGoogle Scholar
[27] S. Shlomo, V.M. Kolomietz, and G Colò,

Deducing the nuclear-matter incompressibility coefficient from data on isoscalar compression modes

. Eur. Phys. J. A 30, 23-30 (2006). doi: 10.1140/epja/i2006-10100-3
Baidu ScholarGoogle Scholar
[28] G. Colò,

Constraints, limits and extensions for nuclear energy functionals

. AIP Conf. Proc. 1128, 59 (2009). doi: 10.1063/1.3146221
Baidu ScholarGoogle Scholar
[29] J. Piekarewicz,

Do we understand the incompressibility of neutron-rich matter?

J. Phys. G 37, 064038 (2010). doi: 10.1088/0954-3899/37/6/064038
Baidu ScholarGoogle Scholar
[30] L.W. Chen, J.Z. Gu,

Correlations between the nuclear breathing mode energy and properties of asymmetric nuclear matter

. J. Phys. G 39, 035104 (2012). doi: 10.1088/0954-3899/39/3/035104
Baidu ScholarGoogle Scholar
[31] L.W. Chen,

Recent progress on the determination of the symmetry energy

. Nuclear Structure in China 2012, 43-54 (2013). doi: 10.1142/9789814447485_0007 arXiv:1212.0284.
Baidu ScholarGoogle Scholar
[32] B.A. Li, L.W. Chen, F.J. Fattoyev, et al.,

Probing nuclear symmetry energy and its imprints on properties of nuclei, nuclear reactions, neutron stars and gravitational waves

. J. Phys.: Conf. Ser. 413, 012021 (2013) doi: 10.1088/1742-6596/413/1/012021
Baidu ScholarGoogle Scholar
[33] Z. Zhang, L.W. Chen,

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

. Phys. Lett. B 726, 234-238 (2013). doi: 10.1016/j.physletb.2013.08.002
Baidu ScholarGoogle Scholar
[34] B.D. Serot and J.D. Walecka, Advances in Nuclear Physics. Vol. 16, J.W. Negele, E. Vogt, Eds., Plenum, New York (1986); B.D. Serot, J.D. Walecka,

Recent progress in quantum hadrodynamics

. Int. J. Mod. Phys. E 6, 515 (1997). doi: 10.1142/S0218301397000299
Baidu ScholarGoogle Scholar
[35] P.-G. Reinhard,

The relativistic mean-field description of nuclei and nuclear dynamics

. Rep. Prog. Phys. 52, 439-514 (1989). doi: 10.1088/0034-4885/52/4/002
Baidu ScholarGoogle Scholar
[36] P. Ring,

Relativistic mean field theory in finite nuclei

. Prog. Part. Nucl. Phys. 37, 193-263 (1996). doi: 10.1016/0146-6410(96)00054-3
Baidu ScholarGoogle Scholar
[37] J. Meng, H. Toki, S.G. Zhou, et al.,

Relativistic continuum Hartree Bogoliubov theory for ground-state properties of exotic nuclei

. Prog. Part. Nucl. Phys. 57, 470-563 (2006). doi: 10.1016/j.ppnp.2005.06.001
Baidu ScholarGoogle Scholar
[38] Y. Sugahara, H. Toki,

Relativistic mean-field theory for unstable nuclei with non-linear σ and ω terms

. Nucl. Phys. A 579, 557-572 (1994). doi: 10.1016/0375-9474(94)90923-7
Baidu ScholarGoogle Scholar
[39] Z.Z. Ren, Z.Y. Zhu, Y.H. Cai, et al.,

Relativistic mean-field study of Mg isotopes

. Phys. Lett. B 380, 241-246 (1996). doi: 10.1016/0370-2693(96)00462-5
Baidu ScholarGoogle Scholar
[40] G.A. Lalazissis, J. König, P. Ring,

New parametrization for the Lagrangian density of relativistic mean field theory

. Phys. Rev. C 55, 540 (1997). doi: 10.1103/PhysRevC.55.540
Baidu ScholarGoogle Scholar
[41] W.H. Long, J. Meng, N. Van Giai, et al.,

New effective interactions in relativistic mean field theory with nonlinear terms and density-dependent meson-nucleon coupling

. Phys. Rev. C 69, 034319 (2004). doi: 10.1103/PhysRevC.69.034319
Baidu ScholarGoogle Scholar
[42] W.Z. Jiang, Z.Z. Ren, T.T. Wang, et al.,

Relativistic mean-field study for Zn isotopes

. Eur. Phys. J. A 25, 29-39 (2005). doi: 10.1140/epja/i2004-10235-1
Baidu ScholarGoogle Scholar
[43] W.Z. Jiang,

Effects of the density dependence of the nuclear symmetry energy on the properties of superheavy nuclei

. Phys. Rev. C 81, 044306 (2010). doi: 10.1103/PhysRevC.81.044306
Baidu ScholarGoogle Scholar
[44] 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). doi: 10.1103/PhysRevC.82.055803
Baidu ScholarGoogle Scholar
[45] B.K. Agrawal, A. Sulaksono, P.-G. Reinhard,

Optimization of relativistic mean field model for finite nuclei to neutron star matter

. Nucl. Phys. A 882, 1-20 (2012). doi: 10.1016/j.nuclphysa.2012.03.004
Baidu ScholarGoogle Scholar
[46] 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). doi: 10.1103/PhysRevC.87.015806
Baidu ScholarGoogle Scholar
[47] H. Müller, B. D. Serot,

Relativistic mean-field theory and the high-density nuclear equation of state

. Nucl. Phys. A 606, 508-537 (1996). doi: 10.1016/0375-9474(96)00187-X
Baidu ScholarGoogle Scholar
[48] C.J. Horowitz, J. Piekarewicz,

Neutron star structure and the neutron radius of 208Pb

. Phys. Rev. Lett 86, 5647 (2001). doi: 10.1103/PhysRevLett.86.5647;
C.J. Horowitz, J. Piekarewicz,

Neutron radii of 208Pb and neutron stars

. Phys. Rev. C 64, 062802 (R) (2001). doi: 10.1103/PhysRevC.64.062802;
C. J. Horowitz, J. Piekarewicz,

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

. Phys. Rev. C 66, 055803 (2002). doi: 10.1103/PhysRevC.66.055803
Baidu ScholarGoogle Scholar
[49] B.G. Todd-Rutel, 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). doi: 10.1103/PhysRevLett.95.122501
Baidu ScholarGoogle Scholar
[50] L.W. Chen, C.M. Ko, B.A. Li,

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

. Phys. Rev. C 76, 054316 (2007). doi: 10.1103/PhysRevC.76.054316
Baidu ScholarGoogle Scholar
[51] B.J. Cai, L.W. Chen,

Nuclear matter fourth-order symmetry energy in the relativistic mean field models

. Phys. Rev. C 85, 024302 (2012). doi: 10.1103/PhysRevC.85.024302
Baidu ScholarGoogle Scholar
[52]

Equation (19) was given in the first version of the present paper, i.e.

, arXiv:1402.4242v1 [nucl-th], in February, 2014.
Baidu ScholarGoogle Scholar
[53] B.J. Cai, L.W. Chen,

Lorentz covariant nucleon self-energy decomposition of the nuclear symmetry energy

. Phys. Lett. B 711, 104-108 (2012). doi: 10.1016/j.physletb.2012.03.058
Baidu ScholarGoogle Scholar
[54] 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). doi: 10.1103/PhysRevC.82.024321
Baidu ScholarGoogle Scholar
[55] Z. Zhang, 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). doi: 10.1103/PhysRevC.90.064317
Baidu ScholarGoogle Scholar
[56] L.W. Chen,

Nuclear matter symmetry energy and the symmetry energy coefficient in the mass formula

. Phys. Rev. C 83, 044308 (2011). doi: 10.1103/PhysRevC.83.044308
Baidu ScholarGoogle Scholar
[57] J. Carriere, C.J. Horowitz, J. Piekarewicz,

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

. Astrophys. J. 593, 463-471 (2003). doi: 10.1086/376515
Baidu ScholarGoogle Scholar
[58] J. Xu, L.W. Chen, B.A. Li, et al.,

Locating the inner edge of the neutron star crust using terrestrial nuclear laboratory data

. Phys. Rev. C 79, 035802 (2009). doi: 10.1103/PhysRevC.79.035802;
J. Xu, L.W. Chen, B.A. Li, et al.

Nuclear constraints on properties of neutron star crusts

. Astrophys. J. 697, 1549-1568 (2009). doi: 10.1088/0004-637X/697/2/1549
Baidu ScholarGoogle Scholar
[59] G. Baym, C. Pethick, P. Sutherland,

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

. Astrophys. J. 170, 299 (1971). doi: 10.1086/151216
Baidu ScholarGoogle Scholar
[60] K. Iida, K. Sato,

Spin-down of neutron stars and compositional transitions in the cold crustal matter

. Astrophys. J. 477, 294-312 (1997). doi: 10.1017/S0074180900115451
Baidu ScholarGoogle Scholar
[61] J. Antoniadis, P.C.C. Freire, N. Wex, et al.,

A massive pulsar in a compact relativistic binary

. Science 340, 1233232 (2013). doi: 10.1126/science.1233232
Baidu ScholarGoogle Scholar
[62] P. Demorest, T. Pennucci, S. Ransom, et al.,

A two-solar-mass neutron star measured using Shapiro delay

. Nature 467, 1081-1083 (2010). doi: 10.1038/nature09466
Baidu ScholarGoogle Scholar
[63] M. Farine, J.M. Pearson, F. Tondeur,

Nuclear-matter incompressibility from fits of generalized Skyrme force to breathing-mode energies

. Nucl. Phys. A 615, 135-161 (1997). doi: 10.1016/S0375-9474(96)00453-8
Baidu ScholarGoogle Scholar
[64] A.W. Steiner, J.M. Lattimer, E.F. Brown,

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

. Astrophys. J. 722, 33-54 (2010). doi: 10.1088/0004-637X/722/1/33
Baidu ScholarGoogle Scholar