This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Exploring universal characteristics of neutron star matter with relativistic ab initio equations of state

Sibo Wang Department of Physics, Chongqing University, Chongqing 401331, China    Chencan Wang School of Physics, Nankai University, Tianjin 300071, China    Hui Tong [email protected] College of Physics and Materials Science, Tianjin Normal University, Tianjin 300387, China Strangeness Nuclear Physics Laboratory, RIKEN Nishina Center, Wako, 351-0198, Japan
Abstract

Starting from the relativistic realistic nucleon-nucleon (NNNN) interactions, the relativistic Brueckner-Hartree-Fock (RBHF) theory in the full Dirac space is employed to study neutron star properties. First, the one-to-one correspondence relation for gravitational redshift and mass is established and used to infer the masses of isolated neutron stars by combining gravitational redshift measurements. Next, the ratio of the moment of inertia II to mass times radius squared MR2MR^{2} as a function of the compactness M/RM/R is obtained, and is consistent with the universal relations in the literature. The moment of inertia for 1.338M1.338M_{\odot} pulsar PSR J0737-3039A I1.338MI_{1.338M_{\odot}} is predicted to be 1.356×1045\times 10^{45}, 1.381×1045\times 10^{45}, and 1.407×1045gcm21.407\times 10^{45}\ \mathrm{g~{}cm^{2}} by the RBHF theory in the full Dirac space with NNNN interactions Bonn A, B, and C, respectively. Finally, the quadrupole moment of neutron star is calculated under the slow-rotation and small-tidal-deformation approximation. The equations of state constructed by the RBHF theory in the full Dirac space, together with those by the projection method and momentum-independence approximation, conform to universal II-Love-QQ relations as well. By combing the tidal deformability from GW170817 and the universal relations from relativistic ab initio methods, the moment of inertia of a neutron star with 1.4 solar mass is also deduced as I1.4M=1.220.25+0.40×1045gcm2I_{1.4M_{\odot}}=1.22^{+0.40}_{-0.25}\times 10^{45}\mathrm{g\ cm^{2}}.

I Introduction

Neutron stars are one of the most compact objects in the universe: their central densities can reach as high as 5 to 10 times the saturation density of nuclear matter, ρ00.16fm3\rho_{0}\approx 0.16\ \text{fm}^{-3} Lattimer and Prakash (2004), which is far beyond what can be achieved in terrestrial laboratories. Therefore, neutron stars are ideal laboratories for studying ultradense matter, and have established close connections among nuclear physics, particle physics, and astrophysics.

The astrophysical observations of the global properties of neutron stars provide important constraints for the equation of state (EOS) of dense matter Lattimer and Prakash (2007); Özel and Freire (2016); Lattimer and Prakash (2016); Burgio et al. (2021), which is the only ingredient needed to unveil the structure of neutron stars theoretically. The high-precision mass measurements of massive neutron stars constitute nowadays one of the most stringent astrophysical constraints on the nuclear EOS; such measurements include PSR J1614-2230 (1.928±0.017M1.928\pm 0.017M_{\odot}Demorest et al. (2010); Fonseca et al. (2016), PSR J0348+0432 (2.01±0.04M2.01\pm 0.04M_{\odot}Antoniadis et al. (2013), and PSR J0740+6620 (2.08±0.07M2.08\pm 0.07M_{\odot}Cromartie et al. (2020); Fonseca et al. (2021). Recently, the Neutron star Interior Composition Explorer (NICER) mission reported two independent Bayesian parameter estimations of the mass and equatorial radius of the millisecond pulsar PSR J0030+0451: 1.340.16+0.15M1.34_{-0.16}^{+0.15}\ M_{\odot} and 12.711.19+1.14km12.71_{-1.19}^{+1.14}\ \text{km} Riley et al. (2019) as well as 1.440.14+0.15M1.44_{-0.14}^{+0.15}\ M_{\odot} and 13.021.06+1.24km13.02_{-1.06}^{+1.24}\ \text{km} Miller et al. (2019). In combination with constraints from radio timing, gravitational wave (GW) observations, and nuclear physics experiments, these posterior distributions have been used to infer the properties of the dense matter EOS (see Ref. Riley et al. (2021) and references therein). Moreover, two independent Bayesian estimations of the radius for the massive millisecond pulsar PSR J0740+6620 have also been reported Riley et al. (2021); Miller et al. (2021).

Another unique probe for studying the properties of dense matter was extracted from the recent observation of GW signals emitted from a binary neutron star merger, i.e., GW170817 Abbott and et al. (2017). The tidal deformability, which denotes the mass quadrupole moment response of a neutron star to the strong external gravitational field induced by its companion Damour et al. (1992); Hinderer (2008); Flanagan and Hinderer (2008); Damour and Nagar (2009); Postnikov et al. (2010), can be inferred from the GW signals. The limits on the tidal deformability have been widely used to constrain the neutron star radius Fattoyev et al. (2018); Annala et al. (2018); Most et al. (2018); Tews et al. (2018), the asymmetric nuclear matter EOS Malik et al. (2018); Zhang et al. (2018); Tong et al. (2020), and hence the neutron skin thickness of 208Pb Fattoyev et al. (2018).

Besides, as rotating objects, the internal structures of neutron stars are strongly constrained by the moment of inertia, which can be determined from the measurements of spin-orbit coupling in double pulsar systems Lyne et al. (2004). Such a measurement of the moment of inertia for neutron star would have crucial implications for delimiting the EOS significantly Lattimer and Schutz (2005) and can be used to distinguish neutron stars from quark stars Yagi and Yunes (2013a). Special attention has been attracted by the system PSR J0737-3039 Burgay et al. (2003); Lyne et al. (2004); Kramer and Wex (2009), which is the only currently known double pulsar system. It is hoped that the moment of inertia of the 1.338MM_{\odot} primary component in this system, i.e., PSR J0737-3039A, will be measured eventually to within 10% Kramer et al. (2021), and could be used to impose new constraints on the EOS Lattimer (2021).

The global properties of neutron stars, like masses, radii, tidal deformabilities, and moments of inertia are highly sensitive to the EOS for neutron star matter Lim et al. (2019); Li and Sedrakian (2019); Tong et al. (2022). Nevertheless, it has been shown Yagi and Yunes (2013a, b) that, for slowly rotating neutron stars, there exist universal relations between the moment of inertia II, the tidal deformability Λ\Lambda (or Love number), and the quadrupole moment QQ of neutron stars, i.e., the so-called II-Love-QQ relations, which are approximately independent of the internal composition and the EOS for neutron star matter. Wide attention has been attracted by these universal relations (see Ref. Yagi and Yunes (2017) for a review). Although so far the reasons for these universal behaviors are not well understood Yagi et al. (2014); Jiang and Yagi (2020), attempts have been made to combine the universal relations with GW detections to infer neutron star properties Landry and Kumar (2018); Kumar and Landry (2019).

The robustness of the II-Love-QQ relations has been extensively studied with EOSs constructed from a variety of nuclear models (see Refs. Yagi and Yunes (2017); Wei et al. (2019) and references therein), including the relativistic Brueckner-Hartree-Fock (RBHF) theory Müther et al. (1987); Brockmann and Machleidt (1990); Li et al. (1992); Engvik et al. (1996); Gross-Boelting et al. (1999); Katayama and Saito (2015). Since the 1980s the RBHF theory has played an important role in understanding the properties of dense nuclear matter from realistic nucleon-nucleon (NNNN) interactions Anastasio et al. (1980, 1981). In the RBHF theory, the single-particle motion of the nucleon in nuclear matter is described with the Dirac equation, where the medium effects are absorbed into the single-particle potentials. In principle, the scalar and the vector components of the single-particle potentials should be determined in the full Dirac space Nuppenau et al. (1989), i.e., by considering the positive-energy states (PESs) and negative-energy states (NESs) simultaneously. However, to avoid the difficulties induced by NESs, the RBHF calculations are primarily performed in the Dirac space without NESs Müther et al. (1987); Brockmann and Machleidt (1990); Li et al. (1992); Engvik et al. (1996); Horowitz and Serot (1984); Gross-Boelting et al. (1999).

Recently, a self-consistent RBHF calculation in the full Dirac space was achieved for symmetric nuclear matter (SNM) Wang et al. (2021, 2022a) and asymmetric nuclear matter (ANM) Wang et al. (2022b). By decomposing the matrix elements of single-particle potential operator in the full Dirac space, the momentum-dependent scalar and vector components of the single-particle potentials are determined uniquely Wang et al. (2021). The long-standing controversy about the isospin dependence of the effective Dirac mass in relativistic ab initio calculations of ANM is also clarified Wang et al. (2022b).

The RBHF theory in the full Dirac space has been applied to neutron stars Wang et al. (2022b); Tong et al. (2022), where the mass, radius, and tidal deformability are calculated with realistic NNNN interactions Bonn A, B, and C Machleidt (1989). The maximum mass of a neutron star is found less than 2.4M2.4M_{\odot} and the neutron star radius for 1.4M1.4M_{\odot} is predicted about 12 km, which are consistent with the astrophysical observations of massive neutron stars and simultaneous mass-radius estimations by NICER Miller et al. (2019). The tidal deformabilities for a 1.4M1.4M_{\odot} neutron star are predicted as 376, 473, and 459 for the three parametrizations of NNNN interactions respectively, and all lie in the region Λ1.4M=190120+390\Lambda_{1.4M_{\odot}}=190^{+390}_{-120} inferred from the revised analysis by LIGO and Virgo Collaborations Abbott and et al. (2018).

In this work, we employ the RBHF theory in the full Dirac space to study other global properties of neutron stars, including the gravitational redshift, moment of inertia, and quadrupole moment under the slow-rotation and small-tidal-deformation approximation. The main focus will be the relation between the moment of inertia and the compactness parameter, as well as the universal II-Love-QQ relations. This paper is organized as follows. In Sec. II, the theoretical framework of the RBHF theory and structure equations for neutron star properties are briefly described. The obtained results and discussions are presented in Sec. III. Finally, a summary is given in Sec. IV.

II Theoretical framework

II.1 The relativistic Brueckner-Hartree-Fock theory

In the RBHF calculations, one of the most important procedure is the self-consistent determination of the single-particle potential operator 𝒰\mathcal{U} of the nucleons, which is generally divided into scalar and vector components Serot and Walecka (1986)

𝒰(𝒑)=US(p)+γ0U0(p)+𝜸𝒑^UV(p).\mathcal{U}(\bm{p})=U_{S}(p)+\gamma^{0}U_{0}(p)+\bm{\gamma\cdot\hat{p}}\,U_{V}(p). (1)

Here 𝒑^=𝒑/p\hat{\bm{p}}=\bm{p}/p is the unit vector parallel to the momentum 𝒑\bm{p}. The quantities US(p)U_{S}(p), U0(p)U_{0}(p), and UV(p)U_{V}(p) are the scalar potential, the timelike part and the spacelike part of the vector potential.

In principle, the scalar and the vector components of the single-particle potentials can only be determined uniquely in the full Dirac space. However, to avoid the numerical difficulties in the full Dirac space, different approximations are proposed to extract the single-particle potentials in the Dirac space without NESs. The momentum-independence approximation Brockmann and Machleidt (1990) assumes that the single-particle potentials are independent of the momentum, and the spacelike component of the vector potential, UVU_{V}, is negligible. The scalar potential USU_{S} and the timelike part of the vector potential, U0U_{0}, are then extracted from the single-particle potential energies at two selected momenta. In the projection method Gross-Boelting et al. (1999), the effective NNNN interaction GG matrix, which is obtained by solving the in-medium scattering equation, is projected onto a complete set of five Lorentz invariant amplitudes, from which the single-particle potentials are calculated analytically. However, the choice of these Lorentz invariant amplitudes is not unique.

Only by decomposing the matrix elements of 𝒰\mathcal{U} in the full Dirac space, can the Lorentz structure and momentum dependence of single-particle potentials be uniquely determined Nuppenau et al. (1989). The theoretical framework for the RBHF theory in the full Dirac space has been described in detail in Ref. Wang et al. (2021) for SNM and Ref. Tong et al. (2022) for ANM. In this work this method is used to construct the EOS of neutron star matter, which is regarded as beta equilibrium nuclear matter consisting of protons, neutrons, electrons, and muons Krastev and Sammarruca (2006). Using the relativistic Bonn A potential Machleidt (1989), the RBHF theory in the full Dirac space for nuclear matter is applicable for density in the range 0.08-0.57 fm3\text{fm}^{-3}. For lower density in the crust of a neutron star, the EOS introduced with the Baym-Bethe-Pethick (BBP) Baym et al. (1971) and Baym-Pethick-Sutherland (BPS) model Baym et al. (1971) is used. For higher density, we follow the strategy proposed in Ref. Rhoades and Ruffini (1974) and applied in Refs. Gandolfi et al. (2012); Wang et al. (2022b), where the neutron-star matter EOS above a critical density ρc=0.57fm3\rho_{c}=0.57\ \text{fm}^{-3} is replaced with the maximally stiff or causal one, which predicts the most rapid increase of pressure with energy density without violating the causality limit.

II.2 Mass, radius, gravitational redshift, and tidal deformability

The stable configurations of a cold, spherically symmetric, and nonrotating neutron star can be obtained from the Tolman-Oppenheimer-Volkov (TOV) equations Oppenheimer and Volkoff (1939); Tolman (1939). Adopting natural units G=c=1G=c=1, the TOV equations are given by

dP(r)dr=\displaystyle\frac{dP(r)}{dr}= [P(r)+(r)][M(r)+4πr3P(r)]r[r2M(r)],\displaystyle\ -\frac{[P(r)+\mathcal{E}(r)][M(r)+4\pi r^{3}P(r)]}{r[r-2M(r)]}, (2a)
dM(r)dr=\displaystyle\frac{dM(r)}{dr}= 4πr2(r),\displaystyle\ 4\pi r^{2}\mathcal{E}(r), (2b)

where P(r)P(r) is the pressure at neutron star radius rr, M(r)M(r) is the total neutron star mass inside a sphere of radius rr, and (r)\mathcal{E}(r) is the total energy density. These differential equations can be solved numerically with a given central pressure PcP_{c} and M(0)=0M(0)=0. The quantity RR for P(R)=0P(R)=0 denotes the radius of the neutron star, and M(R)M(R) is its mass. The gravitational redshift which relates the mass of the neutron star to its radius is defined as

z=(12M/R)1/21.z=\left(1-2M/R\right)^{-1/2}-1. (3)

Since the radius of the neutron star is harder to observe relative to its mass, the simultaneous measurements of the mass and the gravitational redshift would provide a clear radius determination.

The tidal deformability is defined as

Λ=23k2C5.\Lambda=\frac{2}{3}k_{2}C^{-5}. (4)

where C=M/RC=M/R is the compactness parameter. The second Love number k2k_{2} Hinderer (2008); Hinderer et al. (2010) is calculated by

k2=8C55(12C)2[2yR+2C(yR1)]×{6C[2yR+C(5yR8)]+4C3[1311yR+C(3yR2)+2C2(1+yR)]+3(12C)2[2yR+2C(yR1)]ln(12C)}1,\begin{split}k_{2}=&\ \frac{8C^{5}}{5}(1-2C)^{2}[2-y_{R}+2C(y_{R}-1)]\times\left\{6C[2-y_{R}+C(5y_{R}-8)]\right.\\ &\ +4C^{3}[13-11y_{R}+C(3y_{R}-2)+2C^{2}(1+y_{R})]\\ &\ \left.+3(1-2C)^{2}[2-y_{R}+2C(y_{R}-1)]\ln(1-2C)\right\}^{-1},\end{split} (5)

where yR=y(R)y_{R}=y(R) is the solution of the following nonlinear, first-order differential equation

rdy(r)dr+y2(r)+F(r)y(r)+r2Q(r)=0.r\frac{dy(r)}{dr}+y^{2}(r)+F(r)y(r)+r^{2}Q(r)=0. (6)

Here the two functions F(r)F(r) and Q(r)Q(r) depend on the known mass, radius, pressure, and energy density profiles of the star:

F(r)=\displaystyle F(r)= [12M(r)r]1{14πr2[(r)P(r)]},\displaystyle\ \left[1-\frac{2M(r)}{r}\right]^{-1}\left\{1-4\pi r^{2}[\mathcal{E}(r)-P(r)]\right\}, (7a)
Q(r)=\displaystyle Q(r)= {4π[5(r)+9P(r)+(r)+P(r)P/]6r2}×[12M(r)r]1\displaystyle\ \left\{4\pi\left[5\mathcal{E}(r)+9P(r)+\frac{\mathcal{E}(r)+P(r)}{\partial P/\partial\mathcal{E}}\right]-\frac{6}{r^{2}}\right\}\times\left[1-\frac{2M(r)}{r}\right]^{-1}
[2M(r)r2+8πrP(r)]2×[12M(r)r]2\displaystyle\ -\left[\frac{2M(r)}{r^{2}}+8\pi rP(r)\right]^{2}\times\left[1-\frac{2M(r)}{r}\right]^{-2} (7b)

The differential equation (6) for k2k_{2} can be solved together with the TOV equations and the initial condition y(0)=2y(0)=2.

II.3 The moment of inertia

The moment of inertia is calculated under the slow-rotation approximation pioneered by Hartle and Thorne Hartle (1967); Hartle and Thorne (1968), where the frequency Ω\Omega of a uniformly rotating neutron star is far smaller than the Kepler frequency at the equator:

ΩΩmaxM/R3.\Omega\ll\Omega_{\text{max}}\simeq\sqrt{M/R^{3}}. (8)

In the slow-rotation approximation the moment of inertia of a uniformly rotating, axially symmetric neutron star is given by the following expression Fattoyev and Piekarewicz (2010):

I=8π30Rr4eν(r)ω¯(r)Ω(r)+P(r)12M(r)/r𝑑r.I=\frac{8\pi}{3}\int^{R}_{0}r^{4}e^{-\nu(r)}\frac{\bar{\omega}(r)}{\Omega}\frac{\mathcal{E}(r)+P(r)}{\sqrt{1-2M(r)/r}}dr. (9)

The quantity ν(r)\nu(r) is a radially dependent metric function and is defined as

ν(r)=12ln(12MR)rRM(x)+4πx3P(x)x2[12M(x)/x]𝑑x.\nu(r)=\frac{1}{2}\ln\left(1-\frac{2M}{R}\right)-\int^{R}_{r}\frac{M(x)+4\pi x^{3}P(x)}{x^{2}[1-2M(x)/x]}dx. (10)

The frame-dragging angular velocity ω¯\bar{\omega} is usually obtained by the dimensionless relative frequency w~ω¯/Ω\tilde{w}\equiv\bar{\omega}/\Omega, which satisfies the following second-order differential equation:

ddr[r4j(r)dω~(r)dr]+4r3dj(r)drω~(r)=0,\frac{d}{dr}\left[r^{4}j(r)\frac{d\tilde{\omega}(r)}{dr}\right]+4r^{3}\frac{dj(r)}{dr}\tilde{\omega}(r)=0, (11)

where j(r)=eν(r)12M(r)/rj(r)=e^{-\nu(r)}\sqrt{1-2M(r)/r} for rRr\leq R. The relative frequency ω~(r)\tilde{\omega}(r) is subject to the following two boundary conditions

ω~(0)=\displaystyle\tilde{\omega}^{\prime}(0)= 0,\displaystyle\ 0, (12a)
ω~(R)+R3ω~(R)=\displaystyle\tilde{\omega}(R)+\frac{R}{3}\tilde{\omega}^{\prime}(R)= 1.\displaystyle\ 1. (12b)

It should be noted that under the slow-rotation approximation the moment of inertia does not depend on the stellar frequency Ω\Omega.

II.4 The quadrupole moment

It has been shown Yagi and Yunes (2013a, b) that there exist universal relations between the moment of inertia, the Love number, and the quadrupole moment of neutron stars. Physically, the moment of inertia quantifies how fast a neutron star can spin for a fixed angular momentum, the quadrupole moment describes how much a neutron star is deformed away from sphericity due to rotation, and the Love number characterizes how easily a neutron star can be deformed due to an external tidal field. These quantities can be computed by numerically solving for the interior and exterior gravitational field of a neutron star in a slow-rotation Hartle (1967); Hartle and Thorne (1968) approximation and in a small-tidal-deformation approximation Hinderer (2008); Hinderer et al. (2010). In this work the quadrupole moment is calculated by following the detailed instructions described in Ref. Yagi and Yunes (2013b). In order to investigate the universal II-Love-QQ relations, the following dimensionless quantities are introduced:

I¯IM3,Q¯QM(IΩ)2.\bar{I}\equiv\frac{I}{M^{3}},\qquad\bar{Q}\equiv-\frac{QM}{(I\Omega)^{2}}. (13)

III Results and discussions

Refer to caption
Figure 1: (Color online) The speed of sound squared for neutron star matter as a function of density (left) and the neutron star mass-radius relation (right) obtained by the RBHF theory with the potential Bonn A in the full Dirac space (red solid line), in comparison with those obtained by the projection method (green dashed line) and the momentum-independence approximation (gray dotted line). The dark (light) blue and purple regions indicate the 68% (95%) confidence intervals constrained by the NICER analysis of PSR J0030+0451 Miller et al. (2019) and PSR J0740+6620 Miller et al. (2021). The red dotted line corresponds to the maximally stiff EOS.

In the left panel of Fig. 1, the speed of sound squared for neutron star matter from the RBHF theory with potential Bonn A is depicted as a function of density. The results calculated with the projection method and momentum-independence approximation show slowly increasing tendencies with the increase of density. In the full Dirac space, the speed of sound squared increases quickly with the density, and reaches 0.77 at 0.57fm30.57\ \text{fm}^{-3}. Above this density, the amplitude of scalar potential USU_{S} exceeds the nucleon rest mass, and the RBHF iteration in the full Dirac space is very difficult to achieve. This fact might be related to the extension of the Bonn potential to the full Dirac space, which was determined with PESs only. When the central density of a neutron star is fixed at ρc=0.57fm3\rho_{c}=0.57\ \text{fm}^{-3}, our calculations could support a neutron star with mass equal to 1.9M1.9M_{\odot}. To further explore the maximum mass of a neutron star from the RBHF theory in the full Dirac space, we continue with an EOS where the speed of sound is equal to the speed of light Rhoades and Ruffini (1974); Gandolfi et al. (2012); Wang et al. (2022b) (red dotted line). This would provide an upper bound on the maximum mass of a neutron star.

Based on the EOS from the RBHF theory, the mass-radius relations of a neutron star can be calculated from the TOV equations, which are shown in the right panel of Fig. 1. The 68% and 95% contours of the joint probability density distribution of the mass and radius of PSR J0030+0451 Miller et al. (2019) and PSR J0740+6620 Miller et al. (2021) from the NICER analysis are also shown. It can be found that the results obtained in the full Dirac space, with the projection method, and with the momentum-independence approximation are consistent with the recent constraints by NICER. The maximum masses MmaxM_{\text{max}} predicted by the three methods are 2.43M2.43M_{\odot}, 2.31M2.31M_{\odot}, and 2.18M2.18M_{\odot} respectively, which are consistent with the available astrophysical constraints from massive neutron star observations, such as PSR J1614-2230 (1.928±0.017M1.928\pm 0.017M_{\odot}Demorest et al. (2010); Fonseca et al. (2016), PSR J0348+0432 (2.01±0.04M2.01\pm 0.04M_{\odot}Antoniadis et al. (2013), and PSR J0740+6620 (2.08±0.07M2.08\pm 0.07M_{\odot}Cromartie et al. (2020); Fonseca et al. (2021). The radii R1.4MR_{1.4M_{\odot}} of a 1.4M1.4M_{\odot} neutron star from the three methods are 11.97km11.97\ \text{km}, 12.38km12.38\ \text{km}, and 12.35km12.35\ \text{km}, respectively.

Table 1: Neutron star properties and nuclear matter properties at saturation density calculated by the RBHF theory in the full Dirac space with potential Bonn A, in comparison with the results obtained by the RBHF calculation with the projection method and the momentum-independence approximation.
Model MmaxM_{\text{max}} R1.4MR_{1.4M_{\odot}} ρ1.4M\rho_{1.4M_{\odot}} ρ0\rho_{0} E/AE/A EsymE_{\mathrm{sym}} LL
(MM_{\odot}) (km) (fm-3) (fm-3) (MeV) (MeV) (MeV)
Full Dirac space 2.43 11.97 0.46 0.188 -15.40 33.1 65.2
Projection method 2.31 12.38 0.42 0.179 -16.15 34.7 68.8
Mom.-ind. app. 2.18 12.35 0.43 0.178 -15.36 33.2 67.3

In Table 1 we summarize the maximum mass of a neutron star and the radius and the central density for a 1.4M1.4M_{\odot} neutron star obtained by the RBHF theory in the full Dirac space, together with those obtained with the projection method and momentum-independence approximation. The smallest value for R1.4MR_{1.4M_{\odot}} found in the full Dirac space corresponds to the softest EOS for neutron star matter below a density of about 0.4 fm-3. This fact can be further related to nuclear matter properties, which are also summarized in Tab. 1, including the binding energy per nucleon E/AE/A, the symmetry energy EsymE_{\text{sym}}, and its slope LL. It is found that the symmetry energy and its slope at the saturation density obtained in the full Dirac space are the smallest among the three methods. This explains the reason for smallest R1.4MR_{1.4M_{\odot}} and shows how important it is to take both the PESs and NESs into account.

Refer to caption
Figure 2: (Color online) The relation between the gravitational redshift zz of a neutron star with its mass MM (left) and radius RR (right) obtained by the RBHF theory in the full Dirac space with the potentials Bonn A, B, and C. Astrophysical observation for the gravitational redshift of isolated neutron stars RBS 1223 Hambaryan et al. (2014), RX J1856.5-3754 Hambaryan et al. (2014), and RX J0720.4-3125 Hambaryan, V. et al. (2017) are also shown.

Figure 2 shows the relations between the gravitational redshift zz of a neutron star with its mass MM and radius RR obtained by the RBHF theory in the full Dirac space with the potentials Bonn A, B, and C. It can be seen that the zz-MM relation is not strongly affected by the NNNN interactions. With the increase of mass, the gravitational redshift shows a monotonically increasing behavior. A backbending phenomenon is found for the gravitational redshift with the decrease of radius, similar to the case for MM-RR relations as shown in Fig. 1.

Table 2: The masses of isolated neutron star RBS 1223 Hambaryan et al. (2014), RX J1856.5-3754 Hambaryan et al. (2014), and RX J0720.4-3125 Hambaryan, V. et al. (2017) predicted by combining gravitational redshift measurements and the RBHF theory in the full Dirac space with potential Bonn A. The radii are also given. The uncertainties in the last two columns are all from the gravitational redshift measurements.
System Gravitational redshift zz Mass M[M]M[M_{\odot}] Radius R[km]R[\text{km}]
RBS 1223 Hambaryan et al. (2014) 0.160.02+0.030.16^{+0.03}_{-0.02} 1.030.11+0.151.03^{+0.15}_{-0.11} 11.850.04+0.0511.85^{+0.05}_{-0.04}
RX J1856.5-3754 Hambaryan et al. (2014) 0.220.12+0.060.22^{+0.06}_{-0.12} 1.330.64+0.251.33^{+0.25}_{-0.64} 11.940.18+0.0311.94^{+0.03}_{-0.18}
RX J0720.4-3125 Hambaryan, V. et al. (2017) 0.2050.003+0.0060.205^{+0.006}_{-0.003} 1.2580.014+0.0281.258^{+0.028}_{-0.014} 11.9220.004+0.00811.922^{+0.008}_{-0.004}

The clear one-to-one correspondence relation for gravitational redshift and mass established in the left panel in Fig. 2 can be used to infer the mass of a isolated neutron star, when the observation of the gravitational redshift is provided. In Fig. 2, astrophysical observations of gravitational redshift for isolated neutron stars RX J0720.4-3125 Hambaryan, V. et al. (2017), RBS 1223 Hambaryan et al. (2014), and RX J1856.5-3754 Hambaryan et al. (2014) are shown as shadow bands. The predicted masses by combining these observations and the theoretical calculations from the RBHF theory in the full Dirac space with Bonn A potential are listed in the third column of Table 2. The uncertainties are from the gravitational redshift measurements. In Table 2 the corresponding radii are also shown in the last column.

Refer to caption
Figure 3: (Color online) The ratio of the moment of inertia II to MR2MR^{2} as a function of the compactness parameter M/RM/R obtained with the RBHF theory in the full Dirac space with NNNN interactions Bonn A, B, and C. The probability distribution (orange) from the Bayesian analysis together with the fitted (orange) line in Ref. Lim et al. (2019) are shown for comparison. The empirical (purple) band from Ref. Lattimer and Schutz (2005) is also shown. See the text for details.

In Fig. 3 we display the ratio of the moment of inertia II to MR2MR^{2} as a function of the compactness parameter M/RM/R obtained by the RBHF theory in the full Dirac space with NNNN interactions Bonn A, B, and C. Lattimer et al. Lattimer and Schutz (2005) showed that, in the absence of phase transition and other effects that strongly soften the EOS at supranuclear densities, there is a relatively unique relation between the quantity I/MR2I/MR^{2} and M/RM/R:

I/MR2(0.237±0.008)(1+2.844C+18.91C4).I/MR^{2}\simeq(0.237\pm 0.008)(1+2.844C+18.91C^{4}). (14)

This relation is shown as the purple band in Fig. 3. It is found that our results are consistent with the universal relations obtained in Ref. Lattimer and Schutz (2005) for the range where M/R>0.08M/kmM/R>0.08\ M_{\odot}/\text{km}. The derivation for smaller compactness is unimportant, since the observational evidence for neutron star masses and radii lie in the ranges of 1.2MM2.2M1.2\ M_{\odot}\leq M\leq 2.2\ M_{\odot} and 9kmR15km9\ \text{km}\leq R\leq 15\ \text{km} respectively, which leads to the range of compactness as 0.08M/km<M/R<0.24M/km0.08\ M_{\odot}/\text{km}<M/R<0.24\ M_{\odot}/\text{km}. Lim and collaborators Lim et al. (2019) have investigated neutron star moments of inertia from Bayesian posterior probability distributions of the nuclear EOSs that incorporate information from microscopic many-body theory and empirical data of finite nuclei. The probability distribution for I/MR2I/MR^{2} is shown as in Fig. 3. They found that over the entire range of neutron star compactness 0<C0.25M/km0<C\leq 0.25\ M_{\odot}/\text{km}, their results can be well fitted with the following formula:

I/MR2=C+27.178C40.0871+2.183C.I/MR^{2}=\frac{C+27.178C^{4}}{0.0871+2.183C}. (15)

This result is also shown as the orange dashed line in Fig. 3. It can be seen that our results are very close to that obtained by Lim et al., especially for the neutron stars with small compactness.

Table 3: The momenta of inertia for 1.338M1.338\ M_{\odot} pular PSR J0737-3039A predicted by the RBHF theory in the full Dirac space with potentials Bonn A, B, and C. The results obtained with the projection method Gross-Boelting et al. (1999); van Dalen et al. (2004) and momentum-independence approximation (Mom.-ind. approx.) Brockmann and Machleidt (1990) are also shown. For comparison, the values obtained with nonrelativistic ab initio calculation Morrison et al. (2004), predicted with the tidal deformability in GW170817 combing universal relations Landry and Kumar (2018) as well as that inferred from Bayesian analysis (95% credibility) Lim et al. (2019) are also shown. The last column is the corresponding radius.
Model Potential I1.338M[1045gcm2]I_{1.338M_{\odot}}~{}[10^{45}\ \mathrm{g~{}cm^{2}}] R1.338MR_{1.338M_{\odot}} [km]
A 1.356 11.94
Full Dirac space B 1.381 12.11
C 1.407 12.26
A 1.440 12.34
Projection method B 1.465 12.49
C 1.487 12.60
A 1.431 12.32
Mom.-ind. approx. B 1.452 12.46
C 1.471 12.57
Variational calculation (APR) Morrison et al. (2004) AV18 +δv+\ \delta v ++ UIX* 1.24 11.56
GW170817 ++ universal relations Landry and Kumar (2018) 1.150.24+0.381.15^{+0.38}_{-0.24}
Bayesian analysis Lim et al. (2019) 1.360.32+0.151.36^{+0.15}_{-0.32} 12.21.9+0.712.2^{+0.7}_{-1.9}

Although the ratio of the neutron star moment of inertia II to MR2MR^{2} has a universal function of the compactness parameter M/RM/R, the moment of inertia itself depends sensitively on the neutron star’s internal structure. It has been suggested Lattimer and Schutz (2005) that a measurement accuracy of 10% for II is sufficient to place strong constraints on the EOS.

In Table 3, we show the momenta of inertia I1.338MI_{1.338M_{\odot}} and radius R1.338MR_{1.338M_{\odot}} for PSR J0737-3039A predicted by the RBHF theory in the full Dirac space with three parametrizations for NNNN interactions. The results obtained by the projection method Gross-Boelting et al. (1999); van Dalen et al. (2004) and momentum-independence approximation Brockmann and Machleidt (1990) are also shown. The RBHF theory in the full Dirac space leads to minimum values compared to the approximations in the Dirac space without NESs. This is understandable since the RBHF theory in the full Dirac space gives the minimum radius of a neutron star for the fixed canonical mass, as shown in Table 1.

The moment of inertia for PSR J0737-3039A predicted by the RBHF theory in the full Dirac space with Bonn A is 1.356×1045gcm21.356\times 10^{45}\ \mathrm{g~{}cm^{2}}, which is very close to the most probable value 1.36×1045gcm21.36\times 10^{45}\ \mathrm{g~{}cm^{2}} obtained from Bayesian analysis (95% credibility) Lim et al. (2019). The result from the nonrelativistic ab initio variational calculations Akmal et al. (1998) is also shown in Table 3, where the Argonne v18v18 interaction (AV18) Wiringa et al. (1995) is used, together with the relativistic boost corrections to the two-nucleon interaction as well as three-nucleon interactions modeled with the Urbana force Pudliner et al. (1995). The nonrelativistic ab initio calculation leads to a moment of inertia smaller that what we obtain, similarly to the case for radius. In Ref. Landry and Kumar (2018), by using well-known universal relations among neutron star observables, the reported 90% credible bound on the tidal deformability Λ1.4M=190120+390\Lambda_{1.4M_{\odot}}=190^{+390}_{-120} from GW170817 Abbott and et al. (2018) has been translated into a direct constraint on the moment of inertia of PSR J0737-3039A, giving I1.338M=1.150.24+0.38×1045gcm2I_{1.338M_{\odot}}=1.15^{+0.38}_{-0.24}\times 10^{45}\ \mathrm{g~{}cm^{2}}. It can be seen that the results with the three methods for RBHF theory are consistent with this constraint.

Refer to caption
Refer to caption
Figure 4: (Color online) Top: The II-Love (left) and QQ-Love (right) relations for slowly-rotating neutron stars with EOSs obtained by the RBHF theory in the full Dirac space, projection method, and momentum-independence approximations with potentials Bonn A, B, and C. The solid curves show the fit by using Eq. (16). Bottom: Absolute fractional difference from the fit.

Let us now compare the EOSs obtained by the relativistic ab initio calculations, i.e., the RBHF theory in the full Dirac space, projection method, and momentum-independence approximation with Bonn potentials, to the universal II-Love-QQ relations. The II-Love as well as QQ-Love relations and II-QQ relations are shown in the top panels of Figs. 4 and 5, respectively. The dimensionless moment of inertia I¯\bar{I} and dimensionless quadrupole moment Q¯\bar{Q} are defined in Eq. (15). A single parameter along the curve is the mass or compactness, which increases to the left of the plots. Similarly to Ref. Yagi and Yunes (2017), we only show data with the mass of an isolated, nonrotating configuration in the range 1M<M<Mmax1M_{\odot}<M<M_{\text{max}} with MmaxM_{\text{max}} representing the maximum mass for such a configuration. One observes that the universal relations hold very well. Since the relations are insensitive to EOS, one can construct a single fit (black solid curves) given by Yagi and Yunes (2013b, 2017)

lnyi=ai+bilnxi+ci(lnxi)2+di(lnxi)3+ei(lnxi)4,\ln y_{i}=a_{i}+b_{i}\ln x_{i}+c_{i}(\ln x_{i})^{2}+d_{i}(\ln x_{i})^{3}+e_{i}(\ln x_{i})^{4}, (16)

where coefficients are listed in Table 4. These coefficients are very close to that in Ref. Yagi and Yunes (2017), where a large number of EOSs are considered. The bottom panels of Figs. 4 and 5 show the absolute fractional difference between all the data and the fit, which is less than 1%1\% in the whole range.

The universal relations between I¯\bar{I} and Λ\Lambda allows one to extract the momenta of inertia of a neutron star with 1.41.4 solar mass, I¯1.4M\bar{I}_{1.4M_{\odot}}, from the tidal deformability Λ1.4M\Lambda_{1.4M_{\odot}} from GW170817. The revised analysis from LIGO and Virgo Collaborations, Λ1.4M=190120+390\Lambda_{1.4M_{\odot}}=190^{+390}_{-120} Abbott and et al. (2018), corresponds to I¯1.4M=10.302.10+3.39\bar{I}_{1.4M_{\odot}}=10.30^{+3.39}_{-2.10} as shown in the left panel of Fig. 4. From I¯1.4M\bar{I}_{1.4M_{\odot}} and the relation I¯=I/M3\bar{I}=I/M^{3} we obtain I1.4M=1.220.25+0.40×1045gcm2I_{1.4M_{\odot}}=1.22^{+0.40}_{-0.25}\times 10^{45}\mathrm{g\ cm^{2}}. These values are consistent with the results I¯1.4M=11.102.28+3.64\bar{I}_{1.4M_{\odot}}=11.10^{+3.64}_{-2.28} and I1.4M=1.150.24+0.38×1045gcm2I_{1.4M_{\odot}}=1.15^{+0.38}_{-0.24}\times 10^{45}\mathrm{g\ cm^{2}} in Ref. Landry and Kumar (2018), where the II-Love relation is obtained by using a large set of candidate neutron star EOSs based on relativistic mean-field and Skyrme-Hartree-Fock theory.

Refer to caption
Figure 5: (Color online) The same as the left panel in Fig. 4, but for II-QQ relations.
Table 4: Numerical coefficients for the fitting formula of the II-Love, II-QQ, and QQ-Love relations given in Eq. (16).
yiy_{i} xix_{i} aia_{i} bib_{i} cic_{i} did_{i} eie_{i}
I¯\bar{I} Λ\Lambda 1.493 0.06409 0.02104 5.381×104-5.381\times 10^{-4} 1.957×1061.957\times 10^{-6}
I¯\bar{I} Q¯\bar{Q} 1.387 0.5722 0.01043 0.023420.02342 6.245×1046.245\times 10^{-4}
Q¯\bar{Q} Λ\Lambda 0.1899 0.09937 0.04380 3.430×103-3.430\times 10^{-3} 7.054×1057.054\times 10^{-5}

IV Summary

In summary, the RBHF theory in the full Dirac space has been employed to study the gravitational redshift, moment of inertia, and quadrupole moment of neutron star under the slow-rotation and small-tidal-deformation approximation. The one-to-one correspondence relation for gravitational redshift and mass is established and used to infer the masses of isolated neutron stars by combining gravitational redshift measurements. The ratio of the moment of inertia II to MR2MR^{2} as a function of the compactness M/RM/R is obtained, and is consistent with the universal relations shown by Lattimer et al. Lattimer and Schutz (2005) and that from Bayesian posterior probability distributions by Lim et al. Lim et al. (2019). Using NNNN interactions Bonn A, B, and C, the moment of inertia for 1.338M1.338M_{\odot} pulsar PSR J0737-3039A is predicted to be 1.356×10451.356\times 10^{45}, 1.381×10451.381\times 10^{45}, and 1.407×1045gcm21.407\times 10^{45}\ \mathrm{g~{}cm^{2}}, which are consistent with the constraint translated from the tidal deformability deduced from GW170817 with universal relations among neutron star observables. The EOSs constructed by the RBHF theory in the full Dirac space, together with those by the projection method and momentum-independence approximation, are compared successfully to universal II-Love-QQ relations. By combing the tidal deformability Λ1.4M\Lambda_{1.4M_{\odot}} from GW170817 and the numerical fitting for these universal relations from relativistic ab initio EOSs, the moment of inertia of neutron star with 1.4 solar mass is deduced as I1.4M=1.220.25+0.40×1045gcm2I_{1.4M_{\odot}}=1.22^{+0.40}_{-0.25}\times 10^{45}\mathrm{g\ cm^{2}}.

Acknowledgements.
We thank Yeunhwan Lim for providing the probability distribution data for the moment of inertia from the Bayesian analysis and Armen Sedrakian for carefully reading the manuscript. This work was partly supported by the National Natural Science Foundation of China (NSFC) under Grant No. 12147102, the Fundamental Research Funds for the Central Universities under Grants No. 2020CDJQY-Z003 and No. 2021CDJZYJH-003, and the MOST-RIKEN Joint Project “Ab initio investigation in nuclear physics.” Part of this work was achieved by using the supercomputer OCTOPUS at the Cybermedia Center, Osaka University under the support of the Research Center for Nuclear Physics of Osaka University.

References