Combined Constraints on the Equation of State of Dense Neutron-rich Matter from Terrestrial Nuclear Experiments and Observations of Neutron Stars

, , and

Published 2018 May 29 © 2018. The American Astronomical Society. All rights reserved.
, , Citation Nai-Bo Zhang et al 2018 ApJ 859 90 DOI 10.3847/1538-4357/aac027

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/859/2/90

Abstract

Within the parameter space of the equation of state (EOS) of dense neutron-rich matter limited by existing constraints mainly from terrestrial nuclear experiments, we investigate how the neutron star maximum mass Mmax > 2.01 ± 0.04 M, radius 10.62 km < R1.4 < 12.83 km and tidal deformability Λ1.4 ≤ 800 of canonical neutron stars together constrain the EOS of dense neutron-rich nucleonic matter. While the 3D parameter space of Ksym (curvature of nuclear symmetry energy), Jsym, and J0 (skewness of the symmetry energy and EOS of symmetric nuclear matter, respectively) is narrowed down significantly by the observational constraints, more data are needed to pin down the individual values of Ksym, Jsym, and J0. The J0 largely controls the maximum mass of neutron stars. While the EOS with J0 = 0 is sufficiently stiff to support neutron stars as massive as 2.37 M, supporting the hypothetical ones as massive as 2.74 M (composite mass of GW170817) requires J0 to be larger than its currently known maximum value of about 400 MeV and beyond the causality limit. The upper limit on the tidal deformability of Λ1.4 = 800 from the recent observation of GW170817 is found to provide upper limits on some EOS parameters consistent with but far less restrictive than the existing constraints of other observables studied.

Export citation and abstract BibTeX RIS

1. Introduction

What is the nature of neutron stars and dense nuclear matter? Answering this question has been a long-standing and shared goal of both astrophysics and nuclear physics. The fundamental importance and broad impacts of answering this question have been well documented in the literature. It is a major scientific thrust for many major research facilities in astrophysics (National Academies 2011) and nuclear physics (National Academies 2012), such as various advanced X-ray satellites and Earth-based large telescopes, the Neutron Star Interior Composition Explorer (NICER), various gravitational wave detectors, and all advanced radioactive beam facilities being built around the world. In particular, answering this question has been identified as a major goal in both the U.S. 2015 Long Range Plan for Nuclear Sciences (U.S. LRP 2015) and the Nuclear Physics European Collaboration Committee (NuPECC) 2017 Long Range Plan (NuPECC LRP 2017). However, answering this question accurately is very challenging. It is well known that properties of neutron stars are determined by the equation of state (EOS) of neutron-rich matter over a large density range from zero to about 10 times normal nuclear matter density. Unfortunately, even within the simplest model considering npeμ particles only, the corresponding EOS of neutron star matter is still poorly known, not to mention possibly other particles and various phase transitions predicted to occur in the core of neutron stars.

To summarize what terrestrial experiments and nuclear theories have taught us about the EOS of neutron-rich matter near the saturation density ρ0 of symmetric nuclear matter (SNM), we first recall that the specific energy E(ρ, δ) in asymmetric nucleonic matter (ANM) can be approximated parabolically in isospin asymmetry δ = (ρn − ρp)/ρ as

Equation (1)

where E0(ρ) is the specific energy in SNM and Esym(ρ) is the symmetry energy. Around ρ0, the E0(ρ) and Esym(ρ) can be Taylor expanded up to ${[(\rho -{\rho }_{0})/3{\rho }_{0}]}^{3}$ as

Equation (2)

Equation (3)

in terms of several EOS characteristic parameters: the incompressibility K0 and skewness J0 of SNM, as well as the slope L, curvature Ksym, and skewness Jsym of the symmetry energy. While the Taylor expansions are not used in our study for neutron stars, they provide asymptotic boundary conditions at ρ0 for the parameterizations we shall use to describe the EOS of neutron star matter.

Generally speaking, the EOS of neutron-rich nucleonic matter remains very uncertain mostly because of the poorly known nuclear symmetry energy especially at supra-saturation densities (Li et al. 2014). Nevertheless, thanks to the hard work of many people in both astrophysics and nuclear physics over many years, much progress has been made in constraining the EOS of neutron-rich nucleonic matter. In particular, various analyses of terrestrial nuclear experiments and astrophysical observations have constrained the K0, Esym(ρ0), and L in reasonably small ranges around K0 ≈ 240 ± 20 MeV (Shlomo et al. 2006; Piekarewicz 2010), Esym(ρ0) = 31.7 ± 3.2 MeV, and L ≈ 58.7 ± 28.1 MeV. Especially worth noting, the quoted most probable values of Esym(ρ0) and L are based on surveys of 53 analyses of different kinds of terrestrial and astrophysical data available up to 2016 October (Li & Han 2013; Oertel et al. 2017). Moreover, extensive analyses of heavy-ion reactions at intermediate and relativistic energies, especially various forms of nucleon collective flow and kaon production, have provided a reasonably tight constraining band for the EOS of SNM up to about 4.5ρ0 (see, e.g., Danielewicz et al. 2002). However, the coefficients characterizing the high-density behavior of neutron-rich matter, such as the Ksym, Jsym, and J0, are only loosely known to be in the range of −400 MeV ≤ Ksym ≤ 100 MeV, −200 MeV ≤ Jsym ≤ 800 MeV, and −800 MeV ≤ J0 ≤ 400 MeV, respectively, mostly based on analyses of terrestrial nuclear experiments and energy density functionals (Tews et al. 2017; Zhang et al. 2017).

While continuous efforts have been made to constrain the EOS of dense neutron-rich matter using heavy-ion reactions that may involve rare isotopes with large neutron/proton ratios, currently no consensus has been reached yet from analyzing the limited data available (Li 2017). On the other hand, properties of neutron stars and events involving them, such as the mass, radius, moment of inertia, quadrupole deformation, pulsing frequency (Hessels et al. 2006), cooling curve (Yakovlev et al. 2001; Page et al. 2006), frequencies and damping times of various oscillating modes, spin parameter of pulsars, and the strain amplitude and phase evolution of gravitational waves from inspiraling neutron star binaries, all depend significantly on the EOS of neutron-rich matter (see, e.g., Newton et al. 2014; Grigorian et al. 2016; Lattimer & Prakash 2016; Özel & Freire 2016; Watts et al. 2016; Oertel et al. 2017, for recent reviews). While the observational data of neutron star properties are relatively limited so far, they also provide stringent constraints on the EOS and guide theories of dense neutron-rich matter. In particular, the observed masses around 2 M for the two most massive pulsars J1614-2230 (Demorest et al. 2010) and J0348+0432 (Antoniadis et al. 2013) restrict mostly the stiffness of dense SNM EOS, while the radii of neutron stars are known to be more sensitive to the symmetry energy around 2ρ0 (Lattimer & Prakash 2000, 2001). While much progress has been made in measuring the radii of neutron stars, because of the great difficulties involved especially in determining accurately the distance and modeling reliably the spectrum absorptions with different atmosphere models, the reported radii normally suffer from relatively large uncertainties. In fact, some radii extracted from different analyses and observations are still controversial (see, e.g., Miller & Lamb 2016, for a recent review). Since we are not in a position to make any judgment on the reliability of any astrophysical observations, in our analysis here we use as an example the radius of 1.4 M canonical neutron stars (R1.4) in the range of 10.62 km < R1.4 < 12.83 km inferred from analyzing quiescent low-mass X-ray binaries in Lattimer & Steiner (2014). Interestingly though, this radius range is the same as that predicted using the MDI EOS constrained by terrestrial laboratory data in Li & Steiner (2006). Moreover, we also use the upper limit of the dimensionless tidal deformation Λ1.4 ≤ 800 from the recent observation of GW170817 by the LIGO+Virgo Collaborations (Abbott et al. 2017; LIGO & Virgo 2017).

The first discovery of a neutron star merger using multiple messengers further signifies and sets an excellent example of combining and cross-checking multiple probes of dense neutron-rich matter on Earth and in space. Given the aforementioned constraints on the EOS parameters of neutron-rich matter, mostly based on terrestrial nuclear laboratory experiments, here we study how the astrophysical observations of Mmax > 2.01 ± 0.04 M, 10.62 km < R1.4 < 12.83 km, and Λ1.4 ≤ 800 together constrain the high-density EOS in a way that is consistent naturally with the existing constraints from terrestrial nuclear experiments. For this purpose, fixing the K0, Esym(ρ0), and L at their most probable values mentioned earlier, we explore the intersections of constant surfaces with Mmax = 2.01 M, R1.4 = 10.62 km, R1.4 = 12.83 km, and Λ1.4 = 800, respectively, in the 3D parameter space of Ksym, Jsym, and J0. The 3D parameter space allowed by all three observational constraints is identified. Moreover, in constructing the EOS of neutron star matter, the crust–core transition density ρt and pressure Pt have to be calculated consistently. While effects of the magnitude Esym(ρ0) and slope L of symmetry energy at ρ0 on the crust–core transition have been extensively studied in the literature, effects of the curvature Ksym are less known. We therefore will also explore contours of the ρt and Pt in the Ksym versus L plane (2D). The significant role of Ksym is clearly revealed. Furthermore, within the currently known uncertainty ranges of J0 and Jsym, by setting J0 = Jsym = 0 in Equations (2) and (3) as often done in the literature, we also explore how/what the same three astrophysical constraints may teach us about the EOS of dense neutron-rich matter in the LKsym parameter plane. In both the 3D and 2D model frameworks, the upper limit of the tidal deformability Λ1.4 = 800 from the recent observation of GW170817 is found to provide upper limits on some EOS parameters consistent with but less restrictive than the existing constraints on them. Overall, while combing existing constraints from both terrestrial nuclear experiments and astrophysical observations allows us to limit significantly the EOS parameter space of high-density neutron-rich matter, data of more independent observables are needed to pin down the individual values of Ksym, Jsym, and J0.

This paper is organized as follows. The construction of the EOS of neutron star matter is presented in Section 2. Section 3 is devoted to constraining the EOS of dense neutron-rich nucleonic matter with astrophysical observations first in the 3D space of Ksym, Jsym, and J0 and then in the 2D plane of LKsym with Jsym = J0 = 0. Finally, a summary is given in Section 4.

2. Modeling the Equation of State of Neutron Star Matter

The focus of this study is on the EOS of dense neutron-rich matter. We shall thus adopt the NV EOS (Negele & Vautherin 1973) for the inner crust and the BPS EOS (Baym et al. 1971b) for the outer crust while focusing on the EOS of the dense core in a large 3D and 2D parameter space. In this section, we shall first investigate the crust–core transition density and pressure using a thermal dynamical approach. Our main goal is to examine effects of the symmetry energy, especially its curvature Ksym on the transition point. We then discuss how we construct the EOS for the core. For the purposes of this work, it is sufficient to use the simplest model for nonrotating and charge-neutral neutron stars consisting of only npeμ particles at β equilibrium. As we mentioned earlier, essentially all available many-body theories using various interactions have been used to predict both the EOS of SNM and symmetry energy from low to supra-saturation densities. Given their widely different predictions, especially at supra-saturation densities, we try to extract parameters characterizing the EOS of dense neutron-rich matter from the astrophysical observations using as little as possible predictions of any particular many-body theory and/or interaction. However, we must ensure that the EOS parameters asymptotically become naturally the ones extracted from terrestrial nuclear experiments near ρ0.

To avoid misleading the reader, here we discuss in more detail the relationship between the Taylor expansions of Equations (1)–(3) used to characterize the EOS near the saturation density ρ0 both in nuclear theory and in analyzing terrestrial experiments and the parameterizations that we shall use in modeling the EOS of neutron star matter. We also address the question why we cannot just adopt the widely used isospin-independent multiparameter polytropic EOSs (see, e.g., Topper 1964; Butterworth 1976; Read et al. 2009), for the purposes of our work here. Near ρ0 and δ = 0, any nuclear energy density functional E(ρ, δ) can be Taylor expanded using Equations (1)–(3). The relevant expansion coefficients are determined by the E(ρ, δ) via ${E}_{\mathrm{sym}}(\rho )=\tfrac{1}{2}{[{\partial }^{2}{E}_{b}(\rho ,\delta )/\partial {\delta }^{2}]}_{\delta =0}$ for the symmetry energy, $L=3{\rho }_{0}[\partial {E}_{\mathrm{sym}}(\rho )/\partial \rho ]{}_{\rho ={\rho }_{0}}$ for the slope parameter of Esym(ρ), ${K}_{0}\,=9{\rho }_{0}^{2}[{\partial }^{2}{E}_{0}(\rho )/\partial {\rho }^{2}]{}_{\rho ={\rho }_{0}}$ and ${K}_{\mathrm{sym}}=9{\rho }_{0}^{2}[{\partial }^{2}{E}_{\mathrm{sym}}(\rho )/\partial {\rho }^{2}]{}_{\rho ={\rho }_{0}}$ for the incompressibility of SNM and the curvature of Esym(ρ), and ${J}_{0}=27{\rho }_{0}^{3}[{\partial }^{3}{E}_{0}(\rho )/\partial {\rho }^{3}]{}_{\rho ={\rho }_{0}}$ and ${J}_{\mathrm{sym}}=27{\rho }_{0}^{3}[{\partial }^{3}{E}_{\mathrm{sym}}(\rho )/\partial {\rho }^{3}{]}_{\rho ={\rho }_{0}}$ for the skewness of E0(ρ) and Esym(ρ), respectively. The Taylor expansions become progressively inaccurate for large densities and do not converge when ρ > 1.5ρ0. Therefore, for describing neutron star matter, we parameterize the Esym(ρ) and E0(ρ) in exactly the same form as in Equations (1)–(3), but their coefficients are no longer given by the above expressions from any energy density functional E(ρ, δ). Instead, we treat them as unknown parameters to be extracted from astrophysical observations. More specifically, while the NV+BPS EOSs are used for the crust, starting from the crust–core transition density we parameterize the E0(ρ) and Esym(ρ) as

Equation (4)

Equation (5)

using the parameters ${K}^{\mathrm{NS}},{J}^{\mathrm{NS}},{L}^{\mathrm{NS}},{K}_{\mathrm{sym}}^{\mathrm{NS}}$, and ${J}_{\mathrm{sym}}^{\mathrm{NS}}$. These parameterizations naturally become the Taylor expansions in the limit of $\rho \to {\rho }_{0}$. As parameterizations, mathematically they can be used at any density without the convergence issue associated with the Taylor expansion. Any parameterized EOS has to satisfy all the constraints (asymptotic boundary conditions) for the EOS near ρ0 from terrestrial experiments. The above parameterizations facilitate comparisons with the terrestrial constraints around ρ0. When the above parameterizations are used at high densities for modeling the core EOS of neutron stars, the parameters in the above two equations are not necessarily the same as the Taylor expansion coefficients in Equations (1)–(3). While both mathematically and physically L and LNS are expected to be very close, the high-order coefficients (parameters) in the Taylor expansions (parameterizations) may be very different. Unfortunately, very little is known about these parameters from experiments/observations, and the model predictions for the EOS of dense neutron-rich matter diverge. Since the uncertainty ranges of the Taylor expansion coefficients L, Ksym, Jsym, and J0 are already so large, it is then meaningful to use their uncertainty ranges as the starting/reference ranges in our search for the high-density EOS parameters KNS, JNS, LNS, KNSsym, and ${J}_{\mathrm{sym}}^{\mathrm{NS}}$ by studying properties of neutron stars.

With the above definitions and explanations, we now introduce/discuss a convention/simplification that we shall use in the following. It is consistent with the existing convention widely used in the literature for Equation (1). The latter is often referred to as the empirical parabolic law/approximation of the EOS of isospin asymmetric nuclear matter (Bombaci & Lombardo 1991). Namely, it is well known in nuclear physics that Equation (1) has the dual meanings of being a Taylor expansion in δ in the limit of $\delta \to 0$ on one hand and being a parameterization when it is used in very neutron-rich matter where $\delta \to 1$, on the other hand. Adopting this convention and being consistent with that used for Equation (1), albeit confusing without the above explanations, it is really not necessary to write both the Taylor expansions of Equations (1)–(3) and the parameterizations of Equations (4) and (5), as they have the same form although different meanings for the coefficients/parameters involved. In the following discussions, when Equations (1)–(3) are used in describing terrestrial nuclear EOS or predictions of nuclear energy density functional theories, they are Taylor expansions near ρ0 and δ = 0. On the other hand, when they are used in describing the EOS of neutron stars, they are parameterizations to be constrained by astrophysical observations. Then, all three equations in Equations (1)–(3) have the dual meanings in describing the EOS of isospin asymmetric matter encountered in both terrestrial experiments and neutron stars within broad ranges of δ and ρ without the convergence problem.

In the sense that because there is no reliable theory for the EOS of neutron-rich matter significantly above the saturation density, one has to use parameterizations to describe the EOS of neutron star matter; the spirit of the parameterizations of Equations (4) and (5) is essentially the same as many other parameterizations used in the literature (see, e.g., Gandolfi et al. 2009, 2012; Steiner et al. 2016). However, instead of requiring indirectly the involved parameters to reproduce the known properties of symmetric nuclear matter and the symmetry energy near ρ0 through sometimes complicated equations, we use directly the coefficients of Equations (2) and (3). These coefficients as the asymptotic boundary conditions of the EOS are themselves known near ρ0 and δ = 0 through either experiments or converged predictions of many reliable models (see, e.g., Danielewicz et al. 2002; Shlomo et al. 2006; Chen et al. 2009; Piekarewicz 2010; Steiner et al. 2010; Dutra et al. 2012, 2014; Khan et al. 2012; Li & Han 2013; Colò et al. 2014; Cai & Chen 2017; Zhang et al. 2017). In fact, parameterizations similar to Equations (2) and (3), albeit often truncated at first order in density for Esym(ρ) and second order for E0(ρ), i.e., using L and K0 only, have already been used successfully in studying various properties of neutron stars (see, e.g., Oyamatsu & Iida 2007; Sotani et al. 2012). Also, a similar approach of describing approximately the EOS of dense neutron-rich matter was recently proposed in Margueron et al. (2017a) and successfully used in studying properties of both neutron stars (Margueron et al. 2017b) and finite nuclei (Chatterjee et al. 2017) with Bayesian perspectives. With the above cautions, conventions, and justifications, we shall use the parameterized EOS for the core of neutron stars as described in more detail in Section 2.3.

The multiparameter polytropic EOSs (see, e.g., Topper 1964; Butterworth 1976; Read et al. 2009) are widely used in modeling the core EOS of neutron stars at supra-saturation densities. Why do we not simply follow this popular approach? This is mainly because the polytropes used so far depend only on the nucleon number (energy) density, but explicitly independent of the isospin asymmetry δ. To extract information about the high-density symmetry energy from neutron stars, we need to know explicitly the underlying isospin composition of the pressure/density. For example, to the best of our knowledge, all Bayesian analyses done so far using masses and radii of neutron stars have adopted the polytropes. They infer only the total pressures at a few fiducial high densities without any information about the isospin content of the matter at those densities. Indeed, to invert mathematically the Tolman–Oppenheimer–Volkov (TOV) equation, one only needs to know the total pressure as a function of density regardless of its microphysics composition. However, to understand the entire structure, composition, and cooling mechanism of neutron stars, it is necessary to know the isospin dependence of its EOS from the crust to the dense core consistently. Since the pressure has a term proportional to ${\delta }^{2}\cdot {{dE}}_{\mathrm{sym}}(\rho )/d\rho $, where the isospin asymmetry profile δ(ρ) in neutron stars at β equilibrium is determined uniquely by the Esym(ρ), as we shall discuss in more detail below, by directly parameterizing the E0(ρ) and Esym(ρ) separately and then evaluating the resulting pressure as a function of both ρ and δ(ρ), we shall obtain information about the high-density Esym(ρ) underlying the pressure–density relation P(ρ). On the other hand, directly parameterizing the P(ρ) does not facilitate the extraction of such information about the high-density symmetry energy from astrophysical observations.

2.1. Sampling the Density Dependence of Nuclear Symmetry Energy

First of all, we illustrate the broad variation of Esym(ρ) by varying independently the coefficients in Equation (3) within their known uncertainty ranges. It is well known that the isospin asymmetry δ(ρ) in neutron stars at a given baryon density ρ is uniquely determined by the Esym(ρ) in Equation (1) through the charge neutrality and β equilibrium conditions as we shall recall more formally in Section 2.3. Generally speaking, because of the ${E}_{\mathrm{sym}}(\rho )\cdot {\delta }^{2}$ term in the EOS, a higher value of Esym(ρ) will lead to a smaller δ(ρ) at β equilibrium. As quantitative examples, shown in Figure 1 are the Esym(ρ) and δ(ρ) as functions of reduced density by varying only one coefficient each time while fixing all others: (a) L = 40, 50, 60, 70, and 80 MeV; (b) Ksym = −400, −300, −200, −100, 0, and 100 MeV; and (c) Jsym = −200, 0, 200, and 400 MeV. As their names indicate, the slope L, curvature Ksym, and skewness Jsym of symmetry energy play different roles and in order become increasingly more important at higher densities. Obviously, variations of them within their currently known uncertainty ranges allow us to sample very different behaviors of the Esym(ρ) and the corresponding δ(ρ).

Figure 1.

Figure 1. Symmetry energy Esym(ρ) and isospin asymmetry δ(ρ) in neutron star matter at β equilibrium as a function of the reduced density ρ/ρ0 for (a) L = 40, 50, 60, 70, and 80 MeV; (b) Ksym = −400, −300, −200, −100, 0, and 100 MeV; and (c) Jsym = −200, 0, 200, and 400 MeV. All parameters are in units of MeV.

Standard image High-resolution image

It is worth noting that some combinations of the parameters lead to a decreasing Esym(ρ) that may even become negative at high densities. As summarized earlier in Szmaglinski et al. (2006) and reviewed very recently in Li et al. (2018), such a supersoft Esym(ρ) at high densities was predicted in a number of theoretical calculations using various interactions. At very high densities, when the short-range repulsive tensor force due to the ρ-meson exchange makes the EOS of SNM increase faster with density than that of pure neutron matter where the tensor force is much weaker, the Esym(ρ) decreases or even becomes negative at high densities (Pandharipande et al. 1972; Wiringa et al. 1988; Li et al. 2008). To the best of our knowledge, such a seemingly unusual high-density behavior of the Esym(ρ) is not excluded by either any known fundamental physics principle or experiments/observations so far (Tews et al. 2017). Possible consequences of such symmetry energies are discussed in more detail in Szmaglinski et al. (2006), Li et al. (2018), and references therein. In fact, EOSs with a supersoft Esym(ρ) can still support massive neutron stars if the SNM parts of the EOSs are sufficiently stiff even without the help from the new light mesons proposed and/or possible modified strong-field gravity for massive objects (see, e.g., Krivoruchenko et al. 2009; Wen et al. 2009; Lin et al. 2014; Jiang et al. 2015). Interestingly, while not completely settled yet (Li 2017), there are indeed some circumstantial evidences from intermediate-relativistic energy heavy-ion collisions indicating that the Esym(ρ) may become supersoft above about 2ρ0 (Xiao et al. 2009). Currently, devoted efforts are being made by the intermediate-relativistic heavy-ion reaction community to pin down the high-density behavior of nuclear symmetry energy (see, e.g., Li et al. 2014; Russotto et al. 2016; Xu et al. 2016; Trautmann & Wolter 2017; Tsang et al. 2017).

As we shall discuss in detail next, the Esym(ρ) around the crust–core transition is mostly controlled by the L and Ksym parameters when the Esym(ρ0) is fixed at its most probable empirical value of 31.6 MeV. The variation of L from 40 to 80 MeV and of Ksym from −400 to 100 MeV allows us to sample the usual behavior of Esym(ρ0) predicted by various nuclear many-body theories in the sub-saturation density region and within their known empirical constraints at ρ0.

2.2. Imprints of the Density Dependence of Nuclear Symmetry Energy on the Crust–Core Transition in Neutron Stars

Although the crust of neutron stars contributes only a small fraction of the total mass and radius, it plays an important role in various astrophysical phenomena (Baym et al. 1971a, 1971b; Pethick & Ravenhall 1995; Link et al. 1999; Lattimer & Prakash 2000, 2001, 2007; Steiner et al. 2005; Chamel & Haensel 2008; Sotani et al. 2012; Newton et al. 2013; Pons et al. 2013; Piekarewicz et al. 2014; Horowitz et al. 2015). Critical to many effects of the crust is the transition density ρt between the inner crust and the outer core of neutron stars. Previous studies have found consistently that the transition density is very sensitive to the density dependence of nuclear symmetry energy. In particular, the role of the slope parameter L has been extensively studied (see, e.g., Douchin & Haensel 2000; Kubis 2004, 2007; Lattimer & Prakash 2007; Oyamatsu & Iida 2007). Often the studies employ predictions of a particular nuclear many-body theory where the values of L and Ksym are normally correlated. Here we shall study first the individual roles of the L and Ksym in determining the core–crust transition properties and then contours of the transition density and pressure in the L versus Ksym plane. Finally, effects of the LKsym correlation based on the systematics from analyzing over 500 nuclear energy density functions are examined.

In the present study, we employ the thermodynamical approach (Kubis 2004, 2007; Lattimer & Prakash 2007) to estimate the crust–core transition point where the uniform matter becomes unstable against being separated into a mixture of single nucleons and their clusters. The method is known to slightly overestimate the transition density compared to the dynamical approach (Xu et al. 2009; Ducoin et al. 2011; Providência et al. 2014), but it is sufficiently good for the purposes of this work. Specifically, the transition density is determined by the vanishing effective incompressibility of neutron star matter at β equilibrium under the charge neutrality condition (Kubis 2004, 2007; Lattimer & Prakash 2007), i.e.,

Equation (6)

This approach has been used extensively in the literature to locate the transition point using various EOSs (see, e.g., Xu et al. 2009; Ducoin et al. 2011; Providência et al. 2014; Routray et al. 2016). Enclosed in the brackets of the above expression for Kμ are the first-order and second-order derivatives of the symmetry energy, i.e., quantities directly determining the L and Ksym. It is thus necessary and interesting to first explore separate roles of the latter on the transition density. Shown in panels (a) and (b) of Figure 2 are the transition density ρt as functions of L with different values of Ksym and Ksym with different values of L, respectively. It is clearly seen that the ρt changes more dramatically with the variation of Ksym than L in their respective uncertainty ranges. This is mainly because the last two terms in the expression for Kμ largely cancel out, leaving the curvature of Esym(ρ) dominant. In addition, the value of L is already relatively well constrained in a smaller range than the Ksym, making the variation of ρt with L look weaker.

Figure 2.

Figure 2. Crust–core transition density ρt as a function of L (left panel) with Ksym fixed at −400, −300, −200, −100, 0, and 100 MeV, and Ksym (right panel) with L fixed at 40, 50, 60, 70, and 80 MeV.

Standard image High-resolution image

Next, we examine the transition density ρt and pressure Pt by varying both the Ksym and L within their uncertainty ranges continuously. Shown in the two panels of Figure 3 are contours of constant transition densities ρt and pressures Pt in the LKsym plane, respectively. For transition densities larger than ρt = 0.07 fm−3, the required Ksym increases monotonically with L, while different behaviors are observed for lower values of ρt. The lowest transition density about ρt = 0.0549 fm−3 appears around the boundary corner at L = 77 MeV and Ksym = 100 MeV. Different from the contours of constant ρt, for a fixed Pt, the required Ksym always increases linearly with L before reaching the Pt = 0 boundary along the line Ksym = 3.64 L −163.96 (MeV). The latter is used as a limit in exploring properties of neutron stars in the EOS parameter space.

Figure 3.

Figure 3. Contours of (a) the crust–core transition density ρt in fm−3 and (b) the corresponding pressure Pt in MeV fm−3 in the LKsym plane. Lines with fixed values of transition densities and pressures are labeled. The white and red dashed lines are the correlations between Ksym and L from Tews et al. (2017) and Mondal et al. (2017) (see text), respectively. The white region in panel (b) is where the transition pressure vanishes.

Standard image High-resolution image

As indicated earlier, in using Equations (2) and (3) we assume that the coefficients are independent and intend to constrain them directly from observations using as little as possible predictions of any particular many-body theory. Thus, in determining the crust–core transition point for constructing the EOS of neutron stars, we freely vary the Ksym and L within their uncertainty ranges specified earlier. Nevertheless, theoretically predicted values of the Ksym and L are often correlated when model ingredients and/or interactions are varied. Thus, for a consistency check we also study how the predicted correlation between Ksym and L may limit the transition point. As discussed by many people in the literature (see, e.g., Farine et al. 1978; Chen et al. 2009; Danielewicz & Lee 2009; Vidaña et al. 2009; Ducoin et al. 2011; Pearson et al. 2014; Providência et al. 2014; Mondal et al. 2017; Tews et al. 2017), based on the systematics of many predictions using various many-body theories and interactions, an approximately universal and linear correlation exists between the Ksym and L. For example, using a total of over 500 energy density functions, including 263 relativistic mean field (RMF) models and Hartree–Fok calculations using 240 Skyrme (Dutra et al. 2012, 2014) and some realistic and Gogny interactions, Mondal et al. (2017) found the following KsymL correlation:

Equation (7)

Using essentially the same sets of energy density functionals but requiring 0.149 fm−3 < ρ0 < 0.17 fm−3, −17 MeV < E0(ρ0) < −15 MeV, 25 MeV < Esym(ρ0) < 36 MeV, and 180 MeV < K0 < 275 MeV, Tews et al. (2017) rejected some of the energy functionals. They found the following KsymL correlation using the remaining 188 Skyrme and 73 RMF interactions:

Equation (8)

where the ±24.26 and ±56.59 are error bars including 68.3% and 95.4% of the accepted EOSs, respectively. In Figure 3, the above two KsymL correlation functions are shown with the white and red dashed lines, separately. The 68.3% uncertainty range (±24.26 MeV) in Equation (8) is indicated by the red dotted lines, while that of Equation (7) is too small to be shown here. While the parameterizations in Equations (7) and (8) are consistent, their mean values have slightly different slopes because of the different selection criteria used. Nevertheless, if the uncertainty range of Equation (8) is enlarged from 68.3% to 95.4% of the accepted EOSs, the parameterization in Equation (7) can then be fully covered by that in Equation (8). Using the above two correlations, the transition density and pressure are then restricted to be about ρt = 0.08 fm−3 and Pt = 0.40 MeV fm−3, respectively. These values are consistent with the crust–core transition properties often used in the literature. To this end, especially since some of the apparent correlations among EOS parameters from model calculations may be spurious (Margueron et al. 2017a), it is worth noting that while the KsymL correlation from the systematics of over 500 energy density functions is very useful for the consistency check, it is still necessary and important to determine the individual values of Ksym and L from experiments/observations. In our following calculations, we thus consistently use the crust–core transition density and pressure by varying independently the Ksym and L values within their respective uncertain ranges without using any of the above two correlation functions.

2.3. The Core EOS of Neutron Stars

For completeness and the ease of our discussions in the following, we first recall here the formalism for calculating the EOS in the cores of neutron stars. The total energy density epsilon(ρ, δ) of charge-neutral npeμ matter at β equilibrium can be written as

Equation (9)

where epsilonb(ρ, δ) and epsilonl(ρ, δ) are the energy density of baryons and leptons, respectively. The epsilonb(ρ, δ) can be calculated from

Equation (10)

where the specific energy E(ρ, δ) of baryons is given in Equation (1) and MN is the average rest mass of nucleons. The epsilonl(ρ, δ) from the noninteracting Fermi gas model can be expressed as (${\hslash }=c=1$) (Oppenheimer & Volkoff 1939)

Equation (11)

with

Equation (12)

and

Equation (13)

The chemical potential of particle i can be calculated from

Equation (14)

The isospin asymmetry δ(ρ) and relative particle fractions at different densities in neutron stars are obtained through the β equilibrium condition ${\mu }_{n}-{\mu }_{p}={\mu }_{e}={\mu }_{\mu }\approx 4\delta {E}_{\mathrm{sym}}(\rho )$ and the charge neutrality condition ρp = ρe + ρμ. The pressure of the system can be calculated numerically from

Equation (15)

The above expressions allow us to calculate the core EOS, which is connected smoothly at the core–crust transition point to the NV EOS (Negele & Vautherin 1973) for the inner crust followed by the BPS EOS (Baym et al. 1971b) for the outer crust.

As mentioned earlier, the EOS parameters K0, Esym(ρ0), and L near ρ0 are relatively well determined. To investigate how/what high-density EOS parameters are constrained by the three astrophysical observations considered in this work, we construct the EOS of neutron star matter by varying the poorly known J0, Ksym, and Jsym characterizing the EOS of dense neutron-rich nucleonic matter. In principle, all coefficients used in Equations (2) and (3) should be varied simultaneously within a multivariant Bayesian inference (Steiner et al. 2010; Raithel et al. 2016; Margueron et al. 2017b). Such a study is in progress. In the present work, we shall perform traditional analyses first in the 3D parameter space spanned by J0, Ksym, and Jsym while fixing all other parameters at their currently known most probable values. Equivalent to re-parameterizing the EOS of SNM and Esym(ρ) with fewer parameters as often done in the literature, or expanding Equations (2) and (3) only up to ${[(\rho -{\rho }_{0})/3{\rho }_{0}]}^{2}$, we shall also explore the EOS in the 2D parameter space of L and Ksym by setting J0 = Jsym = 0 while keeping other parameters at their known most probable values. The two cases studied here are similar in spirit to using different numbers of piecewise polytropes or parameters to model the EOS of dense neutron-rich matter. Naturally, the values of the parameters involved may be different in the two cases, but they should all asymptotically approach the same existing constraints on them near ρ0.

Certainly, there have been continuous efforts in both astrophysics and nuclear physics to constrain the EOS parameters in both cases. For example, from the pressure of SNM constrained by nucleon collective flow data in relativistic heavy-ion collisions (Danielewicz et al. 2002), a constraint of −1280 MeV ≤ J0 ≤ −10 MeV was obtained by Cai & Chen (2017). Combining it with the mass of neutron star PSR J0348+0432, they further narrowed it down to −494 MeV ≤ J0 ≤ −10 MeV. By analyzing X-ray bursts, Steiner et al. (2010) extracted a value of −690 MeV ≤ J0 ≤ −208 MeV or −790 MeV ≤ J0 ≤ −330 MeV assuming a photospheric radius of rphR or rph = R, respectively. All of these constraints on J0 overlap but have different uncertainty ranges. Similarly, the Ksym and Jsym have not been well constrained either by any experiments/observations so far. Nevertheless, the systematics of over 500 RMF and SHF energy density functionals indicates the following range: −800 MeV ≤ J0 ≤ 400 MeV (Dutra et al. 2012, 2014), −400 MeV ≤ Ksym ≤ 100 MeV, and −200 MeV ≤ Jsym ≤ 800 MeV, respectively (see, e.g., Chen et al. 2009; Dutra et al. 2012, 2014; Colò et al. 2014; Zhang et al. 2017). Thus, we adopt these ranges for the parameters J0, Ksym, and Jsym to be consistent with both existing experimental and theoretical findings.

Within the above uncertainty ranges of the EOS parameters, the pressure in neutron stars can be varied within the shaded band shown in the left panel of Figure 4. Its upper and lower limits are obtained by using the parameter set of L = 80 MeV, Ksym = 100 MeV, Jsym = 800 MeV, and J0 = 400 MeV and the set of L = 40 MeV, Ksym = −400 MeV, Jsym = 134 MeV, and J0 = 400 MeV, respectively. The individual roles of these parameters are examined by varying them independently in the right panel. As one expects, the variation of J0 is most effective in modifying the pressure at supra-saturation densities.

Figure 4.

Figure 4. Pressure of neutron star matter as a function of reduced nucleon density ρ/ρ0. (a) Region of pressure covered by the EOS parameters considered in the present work. (b) Effects of some EOS parameters on the pressure. All parameters are in units of MeV.

Standard image High-resolution image

3. Constraining the EOS of Dense Neutron-rich Matter with Observed Properties of Neutron Stars

With the EOSs prepared in the way described above, the mass (M)–radius (R) relationship of neutron stars is obtained by solving the TOV equations (Tolman 1934; Oppenheimer & Volkoff 1939)

Equation (16)

Equation (17)

where G is the gravitation constant, c is the light speed, and m(r) is the gravitational mass enclosed within a radius r. The dimensionless tidal deformability Λ is related to the Love number k2, neutron star mass M and radius R via

Equation (18)

The k2 is determined by the EOS thorough a differential equation coupled to the TOV equation (Hinderer 2008; Hinderer et al. 2010). More details about the formalism and code used in this work to calculate the k2 can be found in, e.g., Fattoyev et al. (2013, 2014).

Within the 3D parameter space of J0, Ksym, and Jsym and the 2D parameter space of L and Ksym under the conditions discussed in the previous section, using the observational data of Mmax = 2.01 M, 10.62 km ≤ R1.4 ≤ 12.83 km, and Λ1.4 ≤ 800, we study how/what high-density EOS parameters are constrained in the following two subsections, separately.

3.1. Observational Constraints in the J0–Ksym–Jsym EOS Parameter Space

To be clear, we first emphasize again that in this case the following parameters are fixed at their currently known most probable values based on previous systematic surveys as discussed earlier: K0 = 240 MeV, Esym(ρ0) = 31.7 MeV, and L = 58.9 MeV. We explore constraints on the EOS in the 3D KsymJsymJ0 parameter space with the crust–core transition density consistently determined and the condition that Pt ≥ 0. Technically, in exploring the 3D parameter space in three loops, we change the J0 to reproduce specific observational data at given values of Ksym and Jsym that are varied independently. Shown in Figure 5 are the constant surfaces of neutron stars' maximum mass of Mmax = 2.01 M (green), radius of R1.4 = 12.83 km (magenta) and 10.62 km (yellow), and the upper limit of the dimensionless tidal deformability Λ1.4 = 800 (orange) of canonical neutron stars, respectively. For clarity, a causality surface limiting M ≤ 2.4 M is not shown.

Figure 5.

Figure 5. Observational constraints of the maximum mass of neutron stars and the radius of canonical neutron stars on the EOS of dense neutron-rich matter in the Ksym, Jsym, and J0 parameter space. The green, magenta, yellow, blue, and orange surfaces represent Mmax = 2.01 M, R1.4 = 12.83 km, R1.4 = 10.62 km, M = 2.74 M, and Λ1.4 = 800, respectively.

Standard image High-resolution image

For ease of the following discussions, it is worth recalling first that, as shown in Equation (1), the EOS of SNM E0(ρ), the symmetry energy Esym(ρ), and the isospin asymmetry δ(ρ) at β equilibrium are the three quantities together determining the total pressure in neutron stars. More specifically, the total pressure is proportional to ${{dE}}_{0}(\rho )/d\rho +{\delta }^{2}\cdot {{dE}}_{\mathrm{sym}}(\rho )/d\rho $. While the dE0(ρ)/ term is controlled by the J0 parameter with a fixed K0, the dEsym(ρ)/ and δ(ρ) are determined by the Ksym and Jsym parameters when L is fixed. Moreover, the symmetry energy contribution to the total pressure is weighted by δ2(ρ). When the Esym(ρ) is softer with smaller or negative Ksym and Jsym values, the system is more neutron-rich as shown in Figure 1. In particular, for extremely small Ksym and Jsym values (e.g., Ksym = −400 MeV and Jsym = −200 MeV), the Esym(ρ) becomes negative and the δ(ρ) reaches its maximum of 1 at high densities. Then, the necessary contribution to the pressure from the E0(ρ) term will require a large J0 value to support massive neutron stars. At the other extreme, however, when both the Ksym and Jsym are strongly positive (e.g., Ksym = 100 MeV and Jsym = 800 MeV at the bottom left corner), the symmetry energy Esym(ρ) is superstiff and the δ(ρ) is very small as shown in Figure 1. The required J0 to support massive neutron stars is then very small.

It is interesting to note several major features in this rather information-rich 3D plot summarizing very extensive calculations. Let us first focus on the constant surface of Mmax = 2.01 M. From the top right to the bottom left corner, the required J0 first decreases quickly and then remains almost constant with the increasing values of Ksym and Jsym from negative to positive. This feature is completely understandable based on the discussions in the previous paragraph. Namely, near the upper right corner, the symmetry energy is supersoft and the resulting δ(ρ) is close to 1. The weight δ2(ρ) of the symmetry energy contribution to the pressure is significant, while the dEsym(ρ)/ value is small and may even be largely negative. To support neutron stars with the maximum mass of M = 2.01 M, the value of J0 has to be highly positive to compensate the small contribution or overcome the possibly negative contribution from the symmetry energy. However, near the bottom left corner where the symmetry energy is superstiff, the resulting δ2 at β equilibrium becomes so small that the symmetry energy contribution to the pressure is strongly suppressed. The necessary value of J0 is therefore small, and the constant surface of Mmax = 2.01 M becomes rather flat at small J0 values. More quantitatively, the required minimum value of J0 is about −243 MeV at Ksym = −400 MeV and Jsym = 800 MeV.

For a comparison and to see more clearly the role of J0 in determining the masses of neutron stars, a constant surface of Mmax = 2.74 M (sky blue) corresponding to the composite mass of GW170817 is also shown. It is seen that the overall shapes of the constant surfaces of Mmax = 2.01 and 2.74 M are rather similar. To support such a hypothetical neutron star would require a J0 value above 400 MeV beyond its current uncertain range and the causality limit. While the fate of GW170817 is not completely determined yet, several analyses using observations of GW170817 combined with theories/simulations and the causality limit under some caveats have placed the upper bounds on neutron star masses in the range of 2.16–2.28 M (Margalit & Metzger 2017; Shibata et al. 2017; Rezzolla et al. 2018; Ruiz et al. 2018; see also Radice et al. 2018, for a very recent review). For instance, Margalit & Metzger (2017) used electromagnetic constraints on the remnant imposed by the kilonova observations after the merger and the gravitational wave information predicted a maximum mass of Mmax ≤ 2.17 M with 90% confidence. To support neutron stars with such masses, the J0 just needs to be slightly positive well within its current uncertain range. Thus, once the maximum mass is pinned down, it will put a stringent upper limit on the J0 parameter.

In addition, it is also known that quadrupole deformations of neutron stars depend on the symmetry energy (Krastev et al. 2008a, 2008b; Fattoyev et al. 2013, 2014, 2017; Krastev & Li 2018). Interestingly, Abbott et al. (2017) inferred that the dimensionless tidal deformability of GW170817 has an upper limit of Λ1.4 ≤ 800 at the 90% confidence level for the low-spin prior. However, it is seen in Figure 5 that the constant surface of Λ1.4 = 800 (orange) locates far outside the constant surface of R1.4 = 12.83 km. Thus, limits on the high-density EOS parameters from the Λ1.4 ≤ 800 constraint alone are currently much looser than the radius constraint extracted from analyzing the X-ray data. Nevertheless, the expected detection of gravitational waves from a large number of neutron star mergers has the potential to improve the situation.

Now, let us move to the EOS constraints from the radii of neutron stars. The radius of canonical neutron stars is known to depend strongly (weakly) on the nuclear symmetry energy (EOS of SNM; Li & Steiner 2006). It is thus not surprising that the two constant surfaces of radius at R1.4 = 10.62 km and R1.4 = 12.83 km for canonical neutron stars are essentially vertical in Figure 5, indicating a weak dependence on the J0 as one expects. Indeed, they have significant dependences on both the Ksym and Jsym as indicated by the separation between the two constant-radius surfaces. More quantitatively, the required values of J0 in the constant surfaces of R1.4 = 10.62 km and R1.4 = 12.83 km decrease continuously with increasing Ksym and Jsym. The two surfaces can be approximately described by J0 = −585.64 (MeV) − 2.86 Ksym − 1.00 Jsym and J0 = 182.54 (MeV) − 3.19 Ksym − 0.60 Jsym, respectively. With a fixed value of L, the nuclear pressure becomes stronger with increasing values of Ksym and Jsym. Thus, the constant surface of R1.4 = 12.83 km is on the left (having stiffer symmetry energies) of the R1.4 = 10.62 km surface.

As indicated by the arrows in Figure 5, the space enclosed by the three constant surfaces of Mmax ≥ 2.01 M and 10.62 km ≤ R1.4 ≤ 12.83 km is the EOS parameter space allowed by the astrophysical observations of the maximum mass and radii of neutron stars. The space is constrained from the bottom by the maximum mass, and its intersections with the two limits on the radii set the boundary lines restricting the Ksym and Jsym at small J0 values around −200 MeV. At higher J0 values, the EOS parameter space is mainly bounded by the two radius constraints. To show the accepted EOS parameters more clearly, the allowed regions in the Ksym versus Jsym planes with J0 = 400, 200, 0, and −200 MeV, respectively, are shown in Figure 6. The boundaries of these allowed regions in the KsymJsym planes are shown in Figure 7. The regions inside the lines are allowed for a given J0. The shadowed regions are excluded values of Ksym and Jsym for all J0 values. More specifically, the boundary of the right shadowed region can be divided into two parts based on the slopes. The upper one with a negative slope is obtained by requiring the crust–core transition pressure to always stay positive, i.e., Pt ≥ 0. It can be described approximately by the expression Jsym = −11.00 Ksym + 457.71 MeV. The lower one with a positive slope is obtained by the intersection line between the surfaces of Mmax = 2.01 M and R1.4 = 12.83 km. It can be fitted by the expression Jsym = 7.68 Ksym − 504.90 MeV. This boundary sets an upper limit for Ksym at about 68 MeV. The left shadowed region is excluded by the intersection line of R1.4 = 10.62 km and J0 = 400 MeV in Figure 5. It can be fitted by the expression Jsym = −3.07 Ksym − 1054.89 MeV.

Figure 6.

Figure 6. Allowed regions in the KsymJsym planes with J0 = 400, 200, 0, and −200 MeV without considering the causality bounds, respectively.

Standard image High-resolution image
Figure 7.

Figure 7. Constraining boundaries in the KsymJsym plane for a given J0 value of −200, 0, 200, and 400 MeV. The regions inside the lines are allowed for a given J0. The shadowed regions are excluded values of Ksym and Jsym for all J0 values (see text).

Standard image High-resolution image

Overall, within the framework of our analyses, using the maximum mass of neutron stars and the upper and lower limits of the radii of canonical neutron stars, the three EOS parameters are only limited in a space shown in Figure 5 not completely closed, with its boundaries partially given in Figure 7. Obviously, data of more independent observables from either or both astrophysics and nuclear physics are needed to determine the individual values of Ksym, Jsym, and J0.

3.2. Observational Constraints in the L–Ksym EOS Parameter Plane

Within the currently known uncertainty ranges of J0 and Jsym, one can parameterize the EOS with fewer parameters by setting J0 = Jsym = 0 in Equations (2) and (3) as often done in the literature. Then, the two most poorly known parameters are Ksym and L. The latter is known to be around L ≈ 58.7 ± 28.1 MeV as mentioned earlier. In this 2D model framework, here we explore how/what the same three astrophysical constraints may teach us about the EOS of dense neutron-rich matter.

Shown in Figure 8 are contours of constant maximum masses of neutron stars in the LKsym plane. The red, magenta, blue, and cerulean shadows represent regions where Mmax ≤ 2.01 M, R1.4 ≥ 12.83, R1.4 ≤ 10.62 km, and Λ1.4 ≥ 800, respectively. In the white excluded region, the crust–core transition pressure Pt ≤ 0. It is seen that the maximum mass of neutron stars increases with increasing Ksym as one expects. However, it is rather insensitive to the variation of L in the region considered. Again and obviously, the tidal polarizability Λ1.4 ≤ 800 is much less restrictive than the radius constraint of R1.4 ≤ 12.83 km. It is also seen that the R1.4 ≥ 10.62 km constraint is covered by the constraint Mmax ≥ 2.01 M. Consequently, the area bounded by the curves of Mmax ≥ 2.01 M and R1.4 ≤ 12.83 km is the region allowed by the astrophysical observations considered. In this region, the maximum value of Ksym is about 52 MeV, consistent with that found in the 3D analyses in the previous subsection. Also in this region, the upper limit of the maximum mass is about 2.37 M, reached at L ≈ 60 MeV. Probably incidentally, the latter is also currently the most probable value of L based on the surveys of 53 analyses of existing data (Li & Han 2013; Oertel et al. 2017). Interestingly, the observed upper limit of the maximum mass is in good agreement with the findings by Fryer et al. (2015), Lawrence et al. (2015), and Alsing et al. (2017). They found that the upper limit of the maximum mass of neutron stars is between 2.0 and 2.5 M. However, it is necessary to caution here that we have taken J0 = Jsym = 0 in the 2D study. As we have discussed in the previous subsection, the value of J0 affects significantly the maximum mass of neutron stars. Thus, without more precise knowledge about the J0 parameter, the absolute maximum mass of neutron stars cannot be pinned down. Based on our 2D model analyses here, neutron stars more massive than 2.37 M would require a positive value of J0.

Figure 8.

Figure 8. Observational constraints of the maximum mass and radius of neutron stars on the EOS in the LKsym plane. The red, magenta, blue, and cerulean shadows represent regions where Mmax ≤ 2.01 M, R1.4 ≥ 12.83 km, R1.4 ≤ 10.62 km, and Λ1.4 ≤ 800, respectively. In the excluded white region the crust–core transition pressure Pt ≤ 0.

Standard image High-resolution image

4. Summary and Outlook

In summary, within both the 2D and 3D EOS parameter spaces limited by the existing constraints from terrestrial nuclear experiments, we studied how the astrophysical observations of Mmax > 2.01 ± 0.04 M, 10.62 km < R1.4 < 12.83 km, and Λ1.4 ≤ 800 together constrain the EOS parameters of dense neutron-rich nucleonic matter. We also investigated effects of the curvature Ksym of nuclear symmetry energy on the crust–core transition in neutron stars. The consistently calculated transition density in the LKsym plane is used in constructing the EOS of neutron star matter from the surface to the core. The Ksym is found to affect significantly the crust–core transition density and pressure. Fixing K0, Esym(ρ0), and L at their most probable values, determined mainly by terrestrial nuclear experiments, we explored the intersections of constant surfaces with Mmax = 2.01 M, R1.4 = 10.62 km, R1.4 = 12.83 km, and Λ1.4 = 800, respectively, in the 3D parameter space of Ksym, Jsym, and J0. The 3D parameter space narrowed down significantly by the observational constraints is clearly identified. This helps guide nuclear theories for dense neutron-rich matter and related studies in terrestrial experiments. However, to pin down the individual values of Ksym, Jsym, and J0, data of additional independent observables from astrophysical observations and/or laboratory experiments are needed. In particular, the skewness parameter J0 of SNM largely controls the maximum mass of neutron stars. The 2D EOS with J0 = 0 is found sufficiently stiff to support neutron stars as massive as 2.37 M, while to support a hypothetical neutron star as massive as 2.74 M (composite mass of GW170817) would require J0 to be larger than about 400 MeV beyond its known limit. In both the 2D and 3D model frameworks considered in this work, the upper limit of the tidal deformability Λ1.4 = 800 from the recent observation of GW170817 is found to provide upper limits on some EOS parameters consistent with but less restrictive than the existing constraints. In particular, its constraints on the symmetry energy parameters are far less restrictive than the observation of 10.62 km < R1.4 < 12.83 km from analyzing the X-ray data.

While the analyses and results presented here are useful in their own rights, some aspects of our present work should be improved in future studies. In particular, more data and/or a better approach are necessary to determine precisely individual values of the high-density EOS parameters with quantified uncertainties. In the era of gravitational wave astronomy accompanied by the planned new experiments using advanced radioactive beam facilities around the world, we are very hopeful that more data will come soon. Moreover, even with the limited data available, more quantitative information about the high-density EOS parameters may be obtained by using a more robust statistical approach. For example, we fixed the parameters describing the EOS near the saturation density at their most probable values mostly from terrestrial experiments. By doing so, we eliminated any possible correlation between these (fixed) parameters and the high-density parameters Ksym, Jsym, and J0 that we want to determine from astrophysical observations. A Bayesian analysis where the prior distribution of parameters encodes what is already known and the new observational data refine such prior knowledge is particularly well suited for inferring the posterior probability density distributions of all the EOS parameters. Such a study using the masses and radii of 14 neutron stars extracted from Chandra X-ray observations, the maximum masses of neutron stars from GW170817, and earlier observations of neutron stars, as well as all the information from terrestrial experiments about the low-order EOS parameters, will be carried out and reported elsewhere.

We thank F. J. Fattoyev for providing us with the code to calculate the dimensionless tidal deformability. We would also like to thank Alessandra Corsi, Benjamin Owen, and Renxin Xu for very helpful discussions, as well as Ang Li and Bing Zhang for useful communications. N.-B.Z. is supported in part by the China Scholarship Council. B.-A.L. acknowledges the U.S. Department of Energy, Office of Science, under award no. DE-SC0013702, the CUSTIPEN (China-U.S. Theory Institute for Physics with Exotic Nuclei) under the US Department of Energy grant no. DE-SC0009971, and the National Natural Science Foundation of China under grant no. 11320101004. J.X. is supported in part by the Major State Basic Research Development Program (973 Program) of China under contract nos. 2015CB856904 and 2014CB845401, the National Natural Science Foundation of China under grant nos. 11475243 and 11421505, the "100-talent plan" of Shanghai Institute of Applied Physics under grant nos. Y290061011 and Y526011011 from the Chinese Academy of Sciences, and the Shanghai Key Laboratory of Particle Physics and Cosmology under grant no. 15DZ2272100.

Please wait… references are loading.
10.3847/1538-4357/aac027