PSR J0030+0451, GW170817, and the Nuclear Data: Joint Constraints on Equation of State and Bulk Properties of Neutron Stars

, , , , and

Published 2020 March 27 © 2020. The American Astronomical Society. All rights reserved.
, , Citation Jin-Liang Jiang et al 2020 ApJ 892 55 DOI 10.3847/1538-4357/ab77cf

Download Article PDF
DownloadArticle ePub

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

0004-637X/892/1/55

Abstract

Very recently the NICER collaboration published the first-ever accurate measurement of mass and radius together for PSR J0030+0451, a nearby isolated quickly rotating neutron star (NS). In this work we set the joint constraints on the equation of state (EoS) and some bulk properties of NSs with the data of PSR J0030+0451, GW170817, and some nuclear experiments. The piecewise polytropic expansion method and the spectral decomposition method have been adopted to parameterize the EoS. The resulting constraints are consistent with each other. Assuming the maximal gravitational mass of nonrotating NS MTOV lies between 2.04M and 2.4M, with the piecewise method the pressure at twice nuclear saturation density is measured to be ${3.19}_{-1.35}^{+2.63}\times {10}^{34}\,\mathrm{dyn}\,{\mathrm{cm}}^{-2}$ at the 90% level. For an NS with canonical mass of 1.4M, we have the moment of inertia ${I}_{1.4}={1.43}_{-0.13}^{+0.30}\times {10}^{38}\,\mathrm{kg}\cdot {{\rm{m}}}^{2}$, tidal deformability ${{\rm{\Lambda }}}_{1.4}\,={370}_{-130}^{+360}$, radius ${R}_{1.4}={12.1}_{-0.8}^{+1.2}\,\mathrm{km}$, and binding energy ${\mathrm{BE}}_{1.4}={0.16}_{-0.02}^{+0.01}{M}_{\odot }$ at the 90% level, which are improved in comparison to the constraints with the sole data of GW170817. These conclusions are drawn for the mass/radius measurements of PSR J0030+0451 by Riley et al. For the measurements of Miller et al., the results are rather similar.

Export citation and abstract BibTeX RIS

1. Introduction

As the most compact directly observable objects in the universe, neutron stars (NSs) are made of material mainly at supranuclear densities. The equation of state (EoS) of NSs, i.e., the relation between pressure and energy density, describes the general properties of such dense matter. Numerous theoretical EoSs have been developed in the literature (see Oertel et al. 2017 for a review) and the observational data are highly needed to distinguish between them.

The masses and radii of the NSs, the nuclear experiment data, and the gravitational-wave data of neutron star mergers are widely known as powerful probes (see Lattimer 2012; Özel & Freire 2016; Baiotti 2019 for comprehensive reviews). As witnessed in these two years, the discovery of the binary neutron star merger event GW170817 in the O2 run of advanced LIGO/Virgo detectors (Abbott et al. 2017), together with some reasonable assumptions and empirical relationships, have significantly boosted research on the EoS of neutron stars (e.g., Abbott et al. 2018; Annala et al. 2018; De et al. 2018; Most et al. 2018; Jiang et al. 2019). The masses of a small fraction of Galactic NSs in the binary systems have been precisely measured and the record of the most massive NS was broken over and over. The latest record is held by PSR J0740+6620, a millisecond pulsar with a mass of ${2.14}_{-0.09}^{+0.10}{M}_{\odot }$ (Cromartie et al. 2020). Much more massive NSs might exist, as indicated in the updated constraints on the mass of PSR J1748-2021B (Clifford & Ransom 2019, by assuming random inclinations a median pulsar mass of ${2.548}_{-0.078}^{+0.047}{M}_{\odot }$ is inferred with 11 years of continued observation data of the Green Bank Telescope). The radii of some NSs have been previously inferred in a few ways (see Özel & Freire 2016 for a recent review). However, these results are much more model dependent than the masses because the radius "measurements" are usually indirect and suffer from some systematic uncertainties, including for instance the composition of the atmosphere, the distance of the source, the interstellar extinction, and the brightness (see Miller & Lamb 2016 for the detailed discussion). Pulse profile modeling (also known as waveform modeling; Psaltis & Özel 2014) exploits the effects of general and special relativity on rotationally modulated emission from neutron star surface hot spots (see Watts et al. 2016 for a review) and does not suffer from these limitations. In principle, the pulse profile modeling can deliver simultaneous measurements of mass and radius at an unprecedented level of a few percent (e.g., Psaltis et al. 2014; Özel et al. 2016). Such precise measurements are the primary scientific goal of the Neutron Star Interior Composition Explorer (NICER; Gendreau et al. 2016), a pioneering soft X-ray telescope installed on the International Space Station in 2017. Thanks to the successful performance of NICER, very recently the first-ever accurate measurement of mass and radius together for PSR J0030+0451, a nearby isolated quickly rotating NS, has been achieved (Riley et al. 2019a; Miller et al. 2019a) and has far-reaching implications on the EoS of NSs (e.g., Miller et al. 2019a; Raaijmakers et al. 2019; Sieniawska et al. 2018).

In this work, we aim to set the joint constraints on the EoS and some bulk properties of NSs with the data of PSR J0030+0451, GW170817, and some nuclear experiments. This work is organized as follows. We describe our parameterization methods, data sets, and models in Section 2. The results of joint constraints on the EoS and some bulk properties of NSs are presented in Section 3. We summarize our conclusion with a discussion in Section 4.

2. Method

2.1. Parameterizing Methods

Parameterized representations of the EoS can reasonably describe the main properties of the dense matters in NSs (see Baiotti 2019 for a review). Currently there are two widely adopted methods to parameterize the EoS, namely, the piecewise polytropic expansion method (Vuille & Ipser 1999; Read et al. 2009; Lackey & Wade 2015) and the spectral decomposition method (Lindblom 2010; Carney et al. 2018).

In this work we take the same piecewise polytropic expansion method and parameter settings used in Jiang et al. (2019), along with the same pressure-based spectral decomposition method used in Annala et al. (2018) to parameterize the EoS. The piecewise polytropic expansion method uses four pressures $\{{P}_{1{{\rm{n}}}_{{\rm{s}}}},{P}_{1.85{{\rm{n}}}_{{\rm{s}}}},{P}_{3.7{{\rm{n}}}_{{\rm{s}}}},{P}_{7.4{{\rm{n}}}_{{\rm{s}}}}\}$ located, respectively, at four different rest mass densities $\{{\rho }_{1{{\rm{n}}}_{{\rm{s}}}},{\rho }_{1.85{{\rm{n}}}_{{\rm{s}}}},{\rho }_{3.7{{\rm{n}}}_{{\rm{s}}}},{\rho }_{7.4{{\rm{n}}}_{{\rm{s}}}}\}$ to parameterize the EoS, where ${n}_{{\rm{s}}}$ is the nuclear saturation density. Between each pair of adjoining bounds of densities, we approximately take a polytropic form. While the spectral decomposition method uses four expansion parameters {γ0, γ1, γ2, γ3} to describe the relation between the pressure and the adiabatic index, and thus uniquely determine an EoS.

2.2. Constraints of the EoS

Nuclear experiments and gravitational-wave data have been proved to be essential in constraining the EoS (e.g., Lattimer & Steiner 2014; Tews et al. 2017; Abbott et al. 2018; Annala et al. 2018; De et al. 2018; Most et al. 2018). As summarized in Section 2.4 of Jiang et al. (2019), the current nuclear data have set interesting constraints on both ${P}_{1{{\rm{n}}}_{{\rm{s}}}}$ and ${P}_{1.85{{\rm{n}}}_{{\rm{s}}}}$, which are $4.70\geqslant {P}_{1{{\rm{n}}}_{{\rm{s}}}}/({10}^{33}\mathrm{dyn}\,{\mathrm{cm}}^{-2})\geqslant 3.12$ and ${P}_{1.85{{\rm{n}}}_{{\rm{s}}}}$/$({10}^{34}\mathrm{dyn}\,{\mathrm{cm}}^{-2})\geqslant 1.21$, respectively. We also follow that paper (Section 2.5 therein) to take into account the constraints set by the gravitational-wave data of GW170817. However, in this work we do not consider the low-mass X-ray binary data further because the latest NICER measurement of PSR J0030+0451 is much more direct and suffers from significantly fewer systematic uncertainties.

The NICER teams have measured the mass and the radius of PSR J0030+0451 through pulse profile modeling methods. Riley et al. (2019a) reported a mass (radius) ${1.34}_{-0.16}^{+0.15}{M}_{\odot }$ $({12.71}_{-1.19}^{+1.14}\mathrm{km})$, while Miller et al. (2019a) reported a mass (radius) ${1.44}_{-0.14}^{+0.15}{M}_{\odot }$ $({13.02}_{-1.06}^{+1.24}\mathrm{km})$, which are consistent with each other. We mimic the mass–radius posterior distribution of this source by a two-dimensional Gaussian kernel density estimation (KDE) to constrain the EoS:

Equation (1)

where ${\boldsymbol{S}}$ is posterior samples taken from the best fitting three ovals case of Miller et al. (2019b) and the ST+PST case of Riley et al. (2019b). Unless specified, we only show the results with the data of Riley et al. (2019b), because the data of Miller et al. (2019b) yield rather similar results, as shown in Table 2. Besides, we do not consider the impact of spin on the radius because Raaijmakers et al. (2019) have shown that the spin effect of this source is small in comparison to the rather large systematic uncertainties.

In addition to satisfying the above bounds, the EoS should meet the following general requests: (i) the causality condition, i.e., the speed of sound of the dense matter can never exceed the speed of light c; (ii) the microscopical stability condition, i.e., the pressure cannot be smaller in denser matters; (iii) the maximal gravitational mass of nonrotating NSs (MTOV) should be above all those accurately measured. Note that for the rapidly rotating NSs, the gravitational mass has been enhanced. For PSR J0740+6620, the enhancement of the gravitational mass is about 0.01M (Breu & Rezzolla 2016; Ma et al. 2018). Thus we have a robust lower limit of ${M}_{\mathrm{TOV}}\geqslant 2.04{M}_{\odot }$ based on the mass measurement of PSR J0740+6620 (Cromartie et al. 2020). Additionally, the adiabatic index Γ(p) for the spectral decomposition method are limited in the range [0.6, 4.5] (Abbott et al. 2018).

2.3. Models

As mentioned in Section 2.1, two parameterization methods are adopted in this work. To investigate the impact of MTOV on constraining the EoS, we consider the possible regions of (2.04, 2.40) M (i.e., the low MTOV case) and (2.40, 2.90) M (i.e., the high MTOV case), respectively.3 With these considerations we carry out four tests:

  • (i)  
    Using the piecewise method to parameterize the EoS, assuming ${M}_{\mathrm{TOV}}\in (2.04,2.40){M}_{\odot }$.
  • (ii)  
    The same as (i), except assuming ${M}_{\mathrm{TOV}}\in (2.40,2.90){M}_{\odot }$.
  • (iii)  
    Using the spectral method to parameterize the EoS, assuming the low MTOV case.
  • (iv)  
    The same as (iii), except assuming the high MTOV case.

To investigate the role of prior in each test, we also construct the prior for piecewise method and spectral method, where we just take into account the causality condition, the microscopical stability condition, and the lower limit MTOV > 2.04M, while in the spectral case the adiabatic index is assumed to be within the range [0.6, 4.5].

Since we combine the data of GW170817 and PSR J0030+0451 to do the analyis, the likelihood in each test will take the form

Equation (2)

where ${{\boldsymbol{\theta }}}_{\mathrm{gw}}=\{{{ \mathcal M }}_{{\rm{c}}},q,{\chi }_{1},{\chi }_{2},{\theta }_{\mathrm{jn}},{t}_{{\rm{c}}},{\rm{\Psi }},{{\rm{\Lambda }}}_{1},{{\rm{\Lambda }}}_{2}\}$ are the gravitational-wave parameters. The ${{ \mathcal M }}_{{\rm{c}}}$, q, χi, θjn, tc, Ψ, and Λi are chirp mass, mass ratio, spin of the ith neutron star, inclination angle, coalescence time, polarization, and tidal deformability of the ith neutron star, respectively. The $\tilde{d}(f)$, $\tilde{h}(f)$, and Sn(f) are frequency domain gravitational-wave data of GW170817, frequency domain waveform, and power spectral density of the GW170817 data, respectively. The two tidal deformabilities of source NSs of GW170817 and the mass/radius of PSR J0030+0451 are determined by

Equation (3)

where ${{\boldsymbol{\theta }}}_{\mathrm{eos}}$ are parameters needed to describe an EoS, in the case of the piecewise method, ${{\boldsymbol{\theta }}}_{\mathrm{eos}}=\{{P}_{1{{\rm{n}}}_{{\rm{s}}}},{P}_{1.85{{\rm{n}}}_{{\rm{s}}}},{P}_{3.7{{\rm{n}}}_{{\rm{s}}}},{P}_{7.4{{\rm{n}}}_{{\rm{s}}}}\}$, whereas in the case of spectral method, ${{\boldsymbol{\theta }}}_{\mathrm{eos}}=\{{\gamma }_{0},{\gamma }_{1},{\gamma }_{2},{\gamma }_{3}\}$. The pc is the central pressure of PSR J0030+0451. The M1 and M2 are evaluated from ${{ \mathcal M }}_{{\rm{c}}}$ and q. Since we marginalize the distance and phase in analyzing the gravitational wave, there are only 12 parameters in total, namely, ${\boldsymbol{\theta }}={{\boldsymbol{\theta }}}_{\mathrm{gw}}\cup {{\boldsymbol{\theta }}}_{\mathrm{eos}}\cup {p}_{c}$, and we sample these parameters using the PyMultiNest code (Buchner et al. 2014; Buchner 2016) implemented in Bilby (Ashton et al. 2019a, 2019b), while the likelihood contribution of the gravitational wave is calculated using PyCBC (The PyCBC Team 2018; Biwer et al. 2019).

3. Constraint Results

Following the implement of methods described in Section 2.1, all the tests are carried out to get the posterior distributions of ${\boldsymbol{\theta }}$ using Bayesian parameter estimation with the gravitational-wave data, the M − R data of PSR J0030+0451, as well as the constraints from nuclear experiments and the limits on MTOV. The 90% uncertainties of the EoS parameters are summarized in Table 1. Besides, although the spectral method in all the cases yield slightly higher pressure, larger redshift, and smaller radius than the piecewise method given the same conditions, these two methods yield results highly consistent with each other (as shown in Figures 13, and Table 2). It is also found that the harder region of EoS in the prior is largely excluded by the gravitational-wave data and the M − R measurement results of PSR J0030+0451.

Figure 1.

Figure 1. Panels (a)–(d) show the 90% (EoS, sound speed, mass-gravitational redshift, and mass–radius) regions constrained by the data sets described in Section 2.2. The red (cyan) regions show results from the piecewise method with a constraint from the low (high) MTOV range. The blue (yellow) regions show results from the spectral method with a constraint from the low (high) MTOV range. The black dashed line and the black dotted line show the prior of the piecewise method and the spectral method, respectively. The horizontal purple dashed line in (b) represents the limit of ${c}_{{\rm{s}}}/c=1/\sqrt{3}$. The megenta regions in (b) represent the 68% and 90% regions of central rest mass density of PSR J0030+0451 obtained in the low MTOV case of piecewise analysis. The red and cyan data points in (d) are the mass and radius measurement results of PSR J0030+0451 obtained from Riley et al. (2019a) and Miller et al. (2019a), respectively.

Standard image High-resolution image
Figure 2.

Figure 2. Canonical properties of the 1.4M NS and $\tilde{{\rm{\Lambda }}}$ of GW170817. Panels (a)–(c) show the canonical tidal deformability, radius, and gravitational redshift, respectively. Panel (d) presents the combined dimensionless tidal deformability of GW170817. The red lines and the blue lines show properties obtained from the piecewise method and the spectral method, respectively. The solid, dashed–dotted, and dashed lines represent the cases of ${M}_{\mathrm{TOV}}\in (2.04,2.40){M}_{\odot }$, ${M}_{\mathrm{TOV}}\in (2.40,2.90){M}_{\odot }$, and prior, respectively.

Standard image High-resolution image
Figure 3.

Figure 3. Same as Figure 1 but for bulk properties of NSs inferred from universal relations. Panels (a)–(c) show the possible regions of, respectively, tidal deformability, moment of inertia, and binding energy in the mass range of (1.1–1.7) M.

Standard image High-resolution image

Table 1.  Priors and Posteriors of ${{\boldsymbol{\theta }}}_{\mathrm{eos}}$ for Different Parameterizing Methods

Methods Parameters Prior Distributions 90% Range (low MTOV) 90% Range (high MTOV)
Spectral γ0 Ua(0.2, 2.0) ${0.80}_{-0.44}^{+0.49}$ ${0.69}_{-0.32}^{+0.42}$
  ${\gamma }_{1}$ U(-1.6, 1.7) ${0.18}_{-0.62}^{+0.68}$ ${0.32}_{-0.53}^{+0.51}$
  γ2 U(-0.6, 0.6) $-{0.02}_{-0.20}^{+0.16}$ $-{0.04}_{-0.16}^{+0.14}$
  γ3 U(-0.02, 0.02) $-{0.00}_{-0.01}^{+0.02}$ ${0.00}_{-0.01}^{+0.01}$
Piecewise ${P}_{1{{\rm{n}}}_{{\rm{s}}}}/({10}^{33}\mathrm{dyn}\,{\mathrm{cm}}^{-2})$ U(3.12, 4.7) ${3.96}_{-0.72}^{+0.66}$ ${3.93}_{-0.72}^{+0.69}$
  ${P}_{1.85{{\rm{n}}}_{{\rm{s}}}}/({10}^{34}\mathrm{dyn}\,{\mathrm{cm}}^{-2})$ U(1.21, 8.0) ${2.42}_{-1.12}^{+2.45}$ ${2.85}_{-1.51}^{+2.05}$
  ${P}_{3.7{{\rm{n}}}_{{\rm{s}}}}/({10}^{35}\mathrm{dyn}\,{\mathrm{cm}}^{-2})$ U(0.6, 7.0) ${2.90}_{-0.88}^{+0.97}$ ${5.15}_{-1.41}^{+1.59}$
  ${P}_{7.4{{\rm{n}}}_{{\rm{s}}}}/({10}^{36}\mathrm{dyn}\,{\mathrm{cm}}^{-2})$ U(0.3, 4.0) ${2.01}_{-1.36}^{+1.65}$ ${2.30}_{-1.49}^{+1.41}$

Note.

aUniform distribution.

Download table as:  ASCIITypeset image

Table 2.  90% Intervals of Canonical Properties

Tests/Properties ${R}_{1.4}/\mathrm{km}$ Λ1.4 zg,1.4 ${I}_{1.4}/{10}^{38}\mathrm{kg}\cdot {{\rm{m}}}^{2}$ BE1.4/ M
Piecewise prior ${13.6}_{-1.9}^{+1.8}$ ${800}_{-510}^{+810}$ ${0.20}_{-0.03}^{+0.05}$ ${1.77}_{-0.42}^{+0.37}$ ${0.14}_{-0.02}^{+0.03}$
Piecewise, Low ${M}_{\mathrm{TOV}}$, RWBa ${12.1}_{-0.8}^{+1.2}$ ${370}_{-130}^{+360}$ ${0.23}_{-0.03}^{+0.02}$ ${1.43}_{-0.13}^{+0.30}$ ${0.16}_{-0.02}^{+0.01}$
Piecewise, High MTOV, RWB ${12.5}_{-0.9}^{+0.8}$ ${480}_{-180}^{+270}$ ${0.22}_{-0.02}^{+0.02}$ ${1.54}_{-0.17}^{+0.20}$ ${0.15}_{-0.01}^{+0.01}$
Piecewise, Low MTOV, MLDb ${12.4}_{-1.0}^{+0.8}$ ${450}_{-190}^{+260}$ ${0.22}_{-0.02}^{+0.03}$ ${1.52}_{-0.20}^{+0.20}$ ${0.16}_{-0.01}^{+0.02}$
Piecewise, High MTOV, MLD ${12.6}_{-0.9}^{+0.8}$ ${510}_{-190}^{+270}$ ${0.22}_{-0.02}^{+0.02}$ ${1.57}_{-0.18}^{+0.19}$ ${0.15}_{-0.01}^{+0.01}$
Spectral prior ${14.0}_{-2.5}^{+1.4}$ ${1010}_{-730}^{+830}$ ${0.19}_{-0.02}^{+0.06}$ ${1.90}_{-0.54}^{+0.35}$ ${0.13}_{-0.01}^{+0.03}$
Spectral, Low ${M}_{\mathrm{TOV}}$, RWB ${12.0}_{-1.1}^{+1.0}$ ${360}_{-180}^{+300}$ ${0.23}_{-0.03}^{+0.03}$ ${1.42}_{-0.22}^{+0.26}$ ${0.16}_{-0.02}^{+0.02}$
Spectral, High MTOV, RWB ${12.6}_{-0.9}^{+0.6}$ ${540}_{-220}^{+210}$ ${0.22}_{-0.01}^{+0.02}$ ${1.59}_{-0.21}^{+0.15}$ ${0.15}_{-0.01}^{+0.02}$
Spectral, Low ${M}_{\mathrm{TOV}}$, MLD ${12.3}_{-1.2}^{+0.8}$ ${430}_{-230}^{+260}$ ${0.23}_{-0.02}^{+0.04}$ ${1.50}_{-0.28}^{+0.20}$ ${0.16}_{-0.01}^{+0.02}$
Spectral, High MTOV, MLD ${12.6}_{-0.8}^{+0.6}$ ${530}_{-210}^{+210}$ ${0.22}_{-0.01}^{+0.02}$ ${1.59}_{-0.19}^{+0.15}$ ${0.15}_{-0.01}^{+0.01}$

Notes.

aPosterior sample of Riley et al. (2019b) used. bPosterior sample of Miller et al. (2019b) used.

Download table as:  ASCIITypeset image

3.1. The EoS and the Sound Speeds of Neutron Stars

We construct the P − ρ relation for each posterior sample of ${{\boldsymbol{\theta }}}_{\mathrm{eos}}$ (Figure 1(a)). The narrow pressure uncertainty at ρsat is mainly governed by the nuclear constraint, and the higher density part between ρsat and 3ρsat is mainly determined by the gravitational-wave data and the M − R data of PSR J0030+0451, while the highest density region is mainly constrained by the limits on MTOV. Meanwhile, for the two investigated MTOV regions, the EoSs at high densities are slightly different, because the more massive the compact star, the stiffer the EoS for the dense matter.

The character of dense matter can also be described by the sound speed ${c}_{{\rm{s}}}=\sqrt{{dp}/d\epsilon }$, for which there are two interesting limits ${c}_{{\rm{s}}}\leqslant c$ and ${c}_{{\rm{s}}}\leqslant c/\sqrt{3}$ predicted by causality and asymptotically free theories like QCD, respectively. Our results are presented in Figure 1(b), where the megenta region shows the central density of PSR J0030+0451, and the dashed purple line represents the asymptotic limit of ${c}_{{\rm{s}}}=c/\sqrt{3}$. It is unclear whether such a limit has been reached by PSR J0030+0451 because of the relatively large uncertainties of both ρc and cs (see also Greif et al. 2019; Raaijmakers et al. 2019). Future observation of NSs by NICER and Advanced LIGO/Virgo are necessary to robustly solve this issue.

The M − R and Mzg relations are also constructed (Figures 1(c) and (d)), and the credible regions are consistent with that of Raaijmakers et al. (2019). Interestingly, the measured gravitational redshift of the isolated NS RX J0720.4-3125 (${z}_{g}={0.205}_{-0.003}^{+0.006};$ Hambaryan et al. 2017) is nearly the same as that of PSR J0030+0451 (${z}_{{\rm{g}}}\simeq {0.206}_{-0.017}^{+0.014};$ Riley et al. 2019a). If the NSs share the same EoS and the spin effect is neglected, then the masses of these two isolated objects are expected to be nearly the same, because the radii of NSs are almost unchanged in a relative narrow mass range. For zg ≃ 0.206, using results shown in Figure 1(c) we find that the mass of PSR J0030+0451 is consistent with that of binary neutron star (BNS) systems (Kiziltan et al. 2013), suggesting no evidence for experiencing significant accretion of these isolated objects (see also Tang et al. 2020).

3.2. Bulk Properties of Neutron Stars

For a given EoS parameter ${{\boldsymbol{\theta }}}_{\mathrm{eos}}$, the canonical properties of the NSs (or the bulk properties at M = 1.4M) can be obtained through optimizing the central pressure pc to reach a gravitational mass M = 1.4M. With a group of posterior samples of ${{\boldsymbol{\theta }}}_{\mathrm{eos}}$, we can deduce the probability distributions of bulk properties, such as Λ1.4, R1.4, and ${z}_{{\rm{g}},1.4}$. As shown in Figure 2, large ${M}_{\mathrm{TOV}}$ will boost the ${{\rm{\Lambda }}}_{1.4}$ and R1.4 to higher values in both parameterization methods. This is understandable, because larger MTOV can lead to stiffening of EoS. We also notice that the spectral method is more easily affected by the MTOV condition than the piecewise method, because smooth parameterization models (e.g., spectral method) usually couple the high-density EoS, which sets MTOV, from the low-density EoS, that determines R1.4 (Capano et al. 2019; Tews et al. 2019). The canonical gravitational redshift ${z}_{{\rm{g}},1.4}$, which is connected to the R1.4 by one-to-one mapping, mainly reflects the variation contrary to the R1.4. While the combined dimensionless tidal deformability $\tilde{{\rm{\Lambda }}}$ of GW170817 simply keeps the same trend as Λ1.4. This is mainly because both of the masses of NSs in GW170817 are near the region of 1.4M (Abbott et al. 2019). Besides, in comparison to using the data of Riley et al. (2019b), the incorporation of the data of Miller et al. (2019b) yields slightly harder EoS in the low MTOV case, while in high MTOV case the results are remarkably consistent.

With the posterior samples of three pairs of M − Λ, namely, $\{{M}_{1},{{\rm{\Lambda }}}_{1},{M}_{2},{{\rm{\Lambda }}}_{2},{M}_{3},{{\rm{\Lambda }}}_{3}\}$, it is possible to deduce some bulk properties of NSs, as done in Landry & Kumar (2018) and Abbott et al. (2018). Here we adopt three universal relations to transform our posterior samples into constraints of tidal deformability, moment of inertia, and binding energy of NSs in the mass range (1.1–1.7)M. These relations (see Yagi & Yunes 2017, and references therein) read

Equation (4)

where $\lambda {(M)\equiv {\rm{\Lambda }}(M)({GM}/{c}^{2})}^{5}$ is the tidal deformability (its dimensionless form is Λ), $\bar{I}\equiv {c}^{4}I/{G}^{2}{M}^{3}$ is the dimensionless moment of inertia (its dimensional form is I), BE is the binding energy, and G is the Newton's gravitational constant. We take the coefficients an and bn from Landry & Kumar (2018) and Steiner et al. (2016), respectively. For each reference mass Mref, we can obtain a best fit of λref and λ1 with a single posterior sample $\{{M}_{1},{{\rm{\Lambda }}}_{1},{M}_{2},{{\rm{\Lambda }}}_{2},{M}_{3},{{\rm{\Lambda }}}_{3}\}$. Repeating this process for a population of posterior samples, we get a sample of λref and λ1. The λref is inserted into the $\bar{I}-{\rm{\Lambda }}$ relation to get $\bar{I}$. Then, the resulting samples of $\bar{I}$ are inserted into the ${BE}-\bar{I}$ relation to obtain samples of BE. We finally get constraints on Λ, I, and BE for each reference mass in the range of (1.1–1.7) M. We find that all four tests give results that are consistent with each other and shrink the prior of EoS to a softer region, with high MTOV cases showing slightly larger Λ, larger I, and lower BE due to the stiffening of EoS in these cases (as shown in Figure 1(a) and Table 2). For PSR J0737-3039A, which has an accurately measured mass of 1.338M and is expected to have a precise determination of moment of inertia in the next few years via radio observations, our Test (i) predicts its moment of inertia $I={1.35}_{-0.14}^{+0.26}\times {10}^{38}\mathrm{kg}\cdot {{\rm{m}}}^{2}$. This value is higher than that of Landry & Kumar (2018), which is mainly caused by the fact that they adopted a group of Λ1.4 smaller than ours.

4. Conclusion and Discussion

In this work, we perform parameter estimation of EoS using two kinds of parameterization methods, with the information obtained from the combination of latest M − R measurements of PSR J0030+0451 from NICER, strain data of GW170817, and constraints from nuclear experiments/theories. Our results show that, with the additional inclusion of the robust measurements of M − R of PSR J0030+0451, the uncertainty region of the P − ρ diagram is reduced compared to previous works. Steiner et al. (2016) and Bedaque & Steiner (2015) have shown that the sound velocity at ultradense matter can exceed $c/\sqrt{3}$, considering the nuclear theories and the maximum observed NS mass. As for PSR J0030+0451, the situation is unclear because of the relatively large uncertainties of both ρc and cs (see Figure 1(c)).

Meanwhile, bulk properties of NS, e.g., Λ1.4, R1.4, have been better determined. For PSR J0737-3039A with the mass of 1.338M we predict a moment of inertia $I={1.35}_{-0.14}^{+0.26}\,\times {10}^{38}\mathrm{kg}\cdot {{\rm{m}}}^{2}$. Using the zg − M relation, the isolated NS RX J0720.4-3125 (${z}_{g}={0.205}_{-0.003}^{+0.006};$ Hambaryan et al. 2017) is evaluated to have nearly the same mass as PSR J0030+0451, which favors BNS mass distribution (Kiziltan et al. 2013), consistent with our previous work (Tang et al. 2020). We also notice the difference between the two parameterization methods and the impact of different choices of MTOV ranges. For both methods, higher MTOV presents the stiffening of EoS above 2ρsat, along with larger R1.4 and Λ1.4. Besides, smooth EoS models (e.g., spectral method), which couple the EoS in the whole ranges of density, are more sensitive to MTOV constraints (Capano et al. 2019; Tews et al. 2019). Though the results are consistent with each other if the uncertainties have been taken into account, these phenomena caution that a reliable region of MTOV and the choice of parameterizing methods are important for constraining the EoS.

With the accumulated observation data of NICER, the M − R measurements of more targeted NSs (Guillot et al. 2019), such as PSR J1614-2230 and PSR J0740+6620, will be obtained with unprecedented accuracy thanks to the remarkable performance of NICER. Therefore, the EoS can be effectively determined with the radii measurements of NSs with various masses as shown in Weih et al. (2019). Moreover, reducing the uncertainty of radius from 10% to 5%, Sieniawska et al. (2018) showed that ∼10% and ∼40% accuracy in central parameter estimation can be obtained for low-mass and high-mass NSs, respectively. Thus, the behavior of sound velocity at ultra-dense matter and whether the phase transition occurs in NS can be reliably probed. Meanwhile, with the upgrade of Advanced LIGO/Virgo detectors, more and more BNS and neutron star black hole (NSBH) merger events will be caught. For instance, quite a few BNS/NSBH candidates have been reported in the LIGO/Virgo O3 public alerts (GraceDB4 ). A sample of ∼100 BNS merger events is expected to tightly constraint the pressure of neutron star matter in a wide density region (Forbes et al. 2019). Therefore, it is feasible to robustly determine the EoS with the gravitational-wave data and the M − R measurements from NICER in the near future.

We thank the anonymous referee for helpful suggestions. This work was supported in part by NSFC under grants of No. 11525313 (i.e., Funds for Distinguished Young Scholars) and No. 11921003, the Chinese Academy of Sciences via the Strategic Priority Research Program (grant No. XDB23040000), Key Research Program of Frontier Sciences (No. QYZDJ-SSW-SYS024). This research has made use of data and software obtained from the Gravitational Wave Open Science Center https://www.gw-openscience.org, a service of LIGO Laboratory, the LIGO Scientific Collaboration and the Virgo Collaboration. LIGO is funded by the U.S. National Science Foundation. Virgo is funded by the French Centre National de Recherche Scientifique (CNRS), the Italian Istituto Nazionale della Fisica Nucleare (INFN) and the Dutch Nikhef, with contributions by Polish and Hungarian institutes.

Software: Bilby (Ashton et al. 2019b,  https://git.ligo.org/lscsoft/bilby/), LALSuite (LIGO Scientific Collaboration 2018, https://git.ligo.org/lscsoft/lalsuite), PyCBC (The PyCBC Team 2018,  https://pycbc.org, PyMultiNest (Buchner 2016https://github.com/JohannesBuchner/PyMultiNest).

Footnotes

  • The exact value of MTOV is still unknown. In a recent study incorporating an explicitly isospin-dependent parametric EoS of the neutron star matter, MTOV is found to be ≤2.40M (Zhang & Li 2019). A higher MTOV, however, may have been suggested by the updated mass measurement/estimate of PSR J1748-2021B (Clifford & Ransom 2019). Therefore, in this work we investigate both the low and high MTOV cases.

Please wait… references are loading.
10.3847/1538-4357/ab77cf