Neutron star equation of state: QMF modeling and applications
Neutron star equation of state: QMF modeling and applications
Li, A.;Zhu, Z. -Y.;Zhou, E. -P.;Dong, J. -M.;Hu, J. -N.;Xia, C. -J.
2020-07-10 00:00:00
Because of the development of many-body theories of nuclear matter, the long- standing, open problem of the equation of state (EOS) of dense matter may be understood in the near future through the confrontation of theoretical cal- culations with laboratory measurements of nuclear properties & reactions and increasingly accurate observations in astronomy. In this review, we focus on the following six aspects: 1) providing a survey of the quark mean- eld (QMF) model, which consistently describes a nucleon and many-body nucleonic system from a quark potential; 2) applying QMF to both nuclear matter and neu- tron stars; 3) extending QMF formalism to the description of hypernuclei and hyperon matter, as well as hyperon stars; 4) exploring the hadron-quark phase transition and hybrid stars by combining the QMF model with the quark matter model characterized by the sound speed; 5) constraining interquark interactions through both the gravitational wave signals and electromagnetic signals of bi- nary merger event GW170817; and 6) discussing further opportunities to study dense matter EOS from compact objects, such as neutron star cooling and pulsar glitches. Preprint submitted to Elsevier July 13, 2020 arXiv:2007.05116v1 [nucl-th] 10 Jul 2020 1. Introduction The equation of state (EOS) of dense stellar matter is a problem for both nuclear physics and relativistic astrophysics and has been greatly promoted by the detection of gravitational waves from the GW170817 binary neutron star (NS) merger event [1] . Multimessenger observations of NS mergers [2] can provide information for determining the EOS of supranuclear matter [3, 4, 5] and that can possibly constrain the phase diagram of the quantum chromodynamics (QCD) [6, 7, 8, 9]. In NSs, nuclear matter is present in beta equilibrium from very low den- sity to several times the saturation density ( 0:16 fm ) and is extremely neutron-rich [10, 11, 12, 13]. One assumes that there is one theoretical model that can correctly explain the nuclear matter data of dierent physical situa- tions obtained in both laboratory nuclear experiments [e.g., 15, 16, 17, 18, 19] and astronomical observations [e.g., 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30]. However, this is a demanding task. It not only requires the theoretical models to extrapolate from lower density/temperature/isospin to unknown regions at high density/temperature/isospin [31, 32] but also depends on the relevant degrees of freedom of the problem, from nucleons to exotic particles [33, 34, 35, 71, 36], even dark matter particles [e.g., 37, 38]. In this paper, we follow a widely used relativistic mean- eld (RMF) ap- proach [39] based on an eective Lagrangian with meson elds mediating strong interactions between quarks, which we call the quark mean- eld (QMF) model [40, 41]. It self-consistently relates the internal quark structure of a nucleon and a hyperon to the RMFs arising in nuclear and hyperonic matter, respectively, and has been employed extensively in the calculations of nite (hyperon-)nuclei and in nite dense matter [40, 42, 45, 46, 47, 48, 49, 50, 43, 44]. We focus on the EOS that have been developed so far, testing the QMF predictions concerning the constraints from experiments. We also illustrate the developments of this arXiv page: http://blogs.cornell.edu/arxiv/2017/10/16/gw170817/ 2 approach for applications to open questions in the present multiscale multimes- senger gravitational wave era of astronomy. Another complementary approach for nuclear matter is the ab initio approach, such as the Brueckner theory [e.g., 51, 52], the chiral eective eld theory [e.g., 53, 54], the quantum Monte Carlo method [e.g., 55, 56], and the variational method [e.g., 57], which starts from microscopic nucleon-nucleon potentials explicitly including many-body forces. As a comparison, we include some results based on these ab initio many-body approaches. The paper is organized as follows. In Sec. 2, we introduce QMF models by introducing the con nement potential of the constituent quarks for a nucleon. Sec. 3 is then devoted to the NS properties based on the QMF EOSs. In Sec. 4, we demonstrate how strange baryons, e.g., hyperons, are incorporated in the QMF model and discuss the hyperon puzzle with the obtained hyperon star maximum mass. We also discuss hybrid stars and strange quark stars (QSs) by introducing quark matter models. This is followed by the discussions of the NS binary in Sec. 5. Other opportunities for studying EOS are given in Sec. 6, including NS cooling and pulsar glitches. Sec. 7 contains the main conclusions and future perspectives of this review. 2. EOS models from the quark level within QMF In 1988, Guichon [58] developed a novel model for nuclear matter to treat the changes in the nucleon properties of nuclear matter, i.e., the European Muon Collaboration (EMC) eects. This model is similar to the RMF model, but the scalar and the vector meson elds couple not with the nucleons but directly with the quarks. Then, the nucleon properties change according to the strengths of the mean elds acting on the quarks, and the nucleon is dealt with in terms of the MIT bag model [59]. The Guichon model was extended by Thomas and his collaborators under the name of the quark-meson coupling (QMC) model. Excellent reviews on the QMC model can be found in the literature [60, 61]; see also, e.g., [62, 63, 64, 65] for some of the latest improvements. Taking an alternative model for the nucleon, the quark potential model [66], Toki and his 3 collaborators constructed the QMF model [41]. For a more detailed comparison of these two models, we refer to [40, 42]. Brie
y, the bag model assumes the nucleon is constituted by bare quarks in the perturbative vacuum, i.e., current quarks, with a bag constant to account for the energy dierence between the perturbative vacuum and the nonperturbative vacuum, while in the potential model, the nucleon is described in terms of the constituent quarks, which couple with the mesons and gluons. We shall rst introduce the potential model and then introduce the QMF formalism. 2.1. Quark potential model In the MIT bag model, the quarks inside the nucleon are con ned by a bag, which ensures that the quarks can only move freely and independently inside the nucleon through an in nite potential well. In the potential model, quarks are con ned by a phenomenological con nement potential, where the polynomial forms are widely used. A harmonic oscillator potential is usually adopted, with which the Dirac equation can be solved analytically, 0 2 U (r) = (1 +
)(ar + V ); (1) where the scalar-vector form of the Dirac structure is chosen for the quark con- nement potential and the parameters a and V are determined from the vacuum nucleon properties. When the eect of the nuclear medium is considered, the quark eld (~r) satis es the following Dirac equation: [
( g ! g ~
p~ q !q 3q q (m g ) U (r)] (~r) = 0; (2) q q q where , !, and are the classic meson elds. g , g , and g are the cou- q !q q pling constants of ; ! and mesons with quarks, respectively. is the third 3q component of the isospin matrix, and m is the constitute quark mass at approx- imately 300 MeV. The nucleon mass in the nuclear medium can be expressed as the binding energy of three quarks, de ned by the zeroth-order term after solving the Dirac equation E = . The quarks are simply con ned in a N q q 4 two-body con nement potential. Three corrections are taken into account in the zeroth-order nucleon mass in the nuclear medium, including the contribution of the center-of-mass (c.m.) correction , pionic correction M and gluonic c:m: correction (E ) . The pion correction is generated by the chiral symmetry of N g QCD theory and the gluon correction by the short-range exchange interaction of quarks. Finally, the mass of the nucleon in the nuclear medium becomes M = E + M + (E ) : (3) c:m: N g N N N The nucleon radius is written as 0 0 11 + m q q hr i = : (4) 0 0 02 02 (3 + m )( m ) q q q q 0 0 where = V =2; m = m + V =2. The eective single quark energy is 0 0 q q q q given by = g ! g , and the eective quark mass is given by q q! 3q q m = m g . By reproducing the nucleon mass and radius (M ; r ) in free q q N N space, we determine the potential parameters (a and V ) in Eq. (1). We obtain V = 62:257187 MeV and a = 0:534296 fm with m = 300 MeV by tting 0 q M = 939 MeV and r = 0:87 fm [15]. N N 2.2. Nuclear matter from an RMF Lagrangian In the above section, we construct the nucleon at the quark level with the con nement potential and the pion and gluon corrections. Next, we would like to connect such nucleons in a nuclear medium with nuclear objects, such as nuclear matter and systems of nite nuclei. A good bridge is the RMF model at the hadron level, which is developed based on the one-boson exchange potential between two nucleons. The eective nucleon mass from the quark model is inserted into the RMF Lagrangian. The nucleon and meson elds are solved self-consistently, and then, the properties of the nuclear many-body system are obtained. We mention here that the nucleons are treated as point-like particles even though a quark model is used to describe the structure of the nucleon. In many-body calculations, the structure of the nucleon only modi es the eective mass of a nucleon, i.e., Eq. (3). 5 We consider the ; ! and mesons exchanging in the Lagrangian [42, 43, 44], and the cross-coupling from the ! meson and meson is introduced to achieve a reasonable slope of symmetry energy (see Sec. 2.3) [67], 0 0 L = i
@ M g !
g !N N 3 1 1 1 1 2 2 2 3 4 (r) m g g 2 3 2 2 3 4 1 1 1 2 2 2 2 2 2 2 + (r) + m + g g ! N !N 2 2 2 1 1 2 2 2 + (r!) + m ! ; (5) 2 2 where g and g are the nucleon coupling constants for ! and mesons. From !N N the simple quark counting rule, we obtain g = 3g and g = g . The calcu- !N !q N q lation of the con ned quarks gives the relation of the eective nucleon mass M as a function of the eld, g = @M =@, which de nes the coupling with nucleons (depending on the parameter g ). m = 510 MeV, m = 783 MeV, q ! and m = 770 MeV are the meson masses. In this Lagrangian, we already consider the static approximation on the mesons so that their time components are neglected. The spatial part of the ! meson disappears for the time rever- sal symmetry. The in nite nuclear matter has translational invariance, which further removes the partial part of the coordinate space. The equations of motion of nucleons and mesons can be generated by the Euler-Lagrangian equation from the Lagrangian, 0 0 (i
@ M g !
g
) = 0; (6) ! 3 @M 2 2 3 m + g + g = h i; (7) 2 3 2 2 2 2 0 m ! + g g ! = g h
i; (8) v !N ! !N N 2 2 2 2 0 m + g g ! = g h
i: (9) v N 3 N !N where Z i X F 1 M 2 N = h i = dpp p ; (10) 2 2 M + p N i i=n;p i 2 i E = M + (p ) ; (11) N F 6 2 2 2 2 2 2 2 2 2 2 m = m + g g ; m = m + g g ! : (12) v v ! ! !N N N !N n 0 p (p ) is the Fermi momentum for a neutron (proton), = h
i = + , p n F F and = h
i = , which equals 0 in symmetric nuclear matter. 3 3 p n Then, the energy density and pressure, with arbitrary isospin asymmetry = ( )=, can be generated by the energy-momentum tensor, n p Z i 2 2 " = k + M k dk i=n;p 1 1 1 2 2 3 4 + m + g + g 2 3 2 3 4 1 1 3 2 2 2 2 2 2 2 2 + m ! + m + g g ! ; (13) ! v N !N 2 2 2 Z i k 4 1 k P = dk 2 2 k + M i=n;p 1 1 1 2 2 3 4 m g g 2 3 2 3 4 1 1 1 2 2 2 2 2 2 2 2 + m ! + m + g g ! : (14) ! v N !N 2 2 2 where we have written the meson eld with their mean- eld values denoted by , !, and . 2.3. Symmetry energy We subtract the nucleon mass from the energy density (Eq. (13)) to study the binding energy per nucleon, E=A = "= M . The parabolic approximation is usually applicable, and the energy per nucleon can be written as E=A(; ) = E=A(; = 0) + E (n) + ::: (15) sym and it is sucient for performing the calculations only for symmetric nuclear matter and pure neutron matter. E=A(; = 0) can be expanded around the saturation density, E=A(; 0) = E=A( ) + K + ::: (16) where K is the incompressibility at the saturation point. The symmetry energy E () can be expressed in terms of the dierence between the energies per sym 7 particle of pure neutrons ( = 1) and symmetric ( = 0) matter, E () sym E=A(; 1) E=A(; 0). To characterize its density dependence, E () can be sym expanded around the saturation density as follows: E () = E ( ) sym sym 0 dE 1 d E sym sym + ( ) + ( ) + ::: (17) 0 0 d 2 d and the following parameters can be de ned, where all have an energy dimension (MeV), E = E ( ); (18) sym sym 0 dE sym L = 3 ( ) ; (19) d E sym K = 9 ( ) : (20) sym 0 0 E () can also be written as sym 1 1 0 0 E () = E + L + K ( ) + ::: (21) sym sym sym 3 18 0 0 In laboratory experiments, the symmetry energy E () can be studied sym by analyzing the neutron skin [e.g., 68], the dierent isovector nuclear excita- tions [e.g., 69], and the data on heavy-ion collisions such as isospin diusion and the isotopic distribution in multifragmentation processes [e.g., 70]. The large amount of novel exotic nuclei produced in the laboratory and the development of radioactive ion beams have greatly stimulated new research projects on sym- metry energy [71, 72, 73, 74]. We mention here that in the following discussion, we only discuss up to the second expansion terms in both the binding energy (Eq. 15) and the symmetry energy (Eq. 17); see, e.g., [75, 76] for detailed dis- cussions on the higher order terms and the suitability of a nuclear EOS for up to the high density matter possible in NSs. Some of the latest constraints on higher order terms are also discussed in, e.g., [77, 78, 79, 80]. 2.4. Results and discussion There are six parameters (g ; g ; g ; g ; c ; ) in this Lagrangian (Eq. (1)) q !q q 3 3 v to be determined by tting the saturation density and the corresponding val- ues at the saturation point of the binding energy E=A, the incompressibility 8 Table 1: Properties of nuclear matter at saturation predicted by the EOSs employed in this study, in a comparison with the empirical ranges. The BCPM EoS, named after the Barcelona- Catania-Paris-Madrid energy density functional [52], is based on the microscopic Brueckner- Hartree-Fock (BHF) theory [51]. The BSk20 and BSk21 EoS belong to the family of Skyrme nuclear eective forces derived by the Brussels-Montreal group [84]. The high-density part of the BSk20 EoS is adjusted to t the result of the neutron matter APR EOS [57], whereas the high-density part of the BSk21 EOS is adjusted to the result of the BHF calculations using the Argonne v18 potential plus a microscopic nucleonic three-body force. The TM1 EOS is based on a phenomenological nuclear RMF model with the TM1 parameter set [85], as well as the GM1 EOS, which uses a dierent parameter set [86]. The number density n is in fm . The energy per baryon E=A and the compressibility K, as well as the symmetry energy E sym and its slope L at saturation, are in MeV. The empirical values are taken from [17, 18, 71, 81]. E=A K E L 0 sym EoS (fm ) (MeV) (MeV) (MeV) (MeV) QMF 0.16 -16.00 240.00 31.00 40.0 BCPM 0.16 -16.00 213.75 31.92 53.0 TM1 0.145 -16.26 281.14 36.89 110.8 BSk20 0.159 -16.08 241.4 30.0 37.4 BSk21 0.158 -16.05 245.8 30.0 46.6 APR 0.16 -16.00 247.3 33.9 53.8 GM1 0.153 -16.32 299.2 32.4 93.9 Empirical 0:16 0:01 16:0 0:1 240 20 31:7 3:2 58:7 28:1 K , the symmetry energy E , the symmetry energy slope L and the eective sym (Landau) mass M ( 0:74M ). In particular, we use the most preferred values for (K; E ; L) as recently suggested by [71, 81], namely, K = 240 20 MeV, sym E = 31:7 3:2 MeV, and L = 58:7 28:1 MeV. A recent tting of nite sym nuclei data in the same model yielded K = 328 MeV [49], and we choose this case as well for a comparison. To study the eect of r , we varied this parame- ter from the intermediate value 0:87 fm [15] by approximately 10% according to our model capability: r = 0.80 fm, 0.87 fm, and 1.00 fm. This covers both of the most recent experimental analyses of the rms radius of the proton charge distribution: 0.879 0.009 fm [82] from electron-proton scattering and 0.8409 0.0004 fm [83] from the Lamb shift measurement in muonic hydrogen. For each nucleon radius, we rst determine the potential parameters (a and V ) by repro- ducing (m ; r ) and then determine QMF many-body parameters by reproduc- N N ing the saturation properties of nuclear matter ( ; E=A; E ; K; L; M =M ), 0 sym N which is shown in the rst line of Table 1. Six EOS models from other theo- 9 (a) SNM (b) PNM 0.80 fm 0.87 fm K [MeV] = 328 1.00 fm (c) SNM (d) PNM HIC HIC + sym_soft 200 260 HIC + sym_stiff 0 1 2 3 4 5 6 1 2 3 4 5 6 ρ /ρ ρ /ρ N 0 N 0 Figure 1: (Color online) Binding energy (B.E.) and pressure as a function of the number density for symmetric nuclear matter (SNM) and pure neutron matter (PNM). The calcula- tions are performed for xed symmetry parameters E = 31 and L = 60 MeV and dierent sym cases of incompressibility K at saturation: K = 240; 260; 328 MeV. The results with dierent nucleon radii of 0.80, 0.87, and 1.00 fm, chosen from the CODATA values and two recent experiments [15, 82, 83], are shown by the solid, dashed, and dotted curves, respectively. Heavy-ion collisions (HIC) are expected to go through a quarkCgluon plasma (QGP) phase, where matter is strongly interacting, resulting in the development of collective motion. The EOS results for SNM and PNM lie inside the boundaries obtained from the analysis of the collective
ow in HIC [16], which are shown with two density-dependent cases of symmetry energy (light blue for the sti case and dark blue for the soft case). The radius of the nucleon is shown to have limited eects on the nuclear matter EOSs even at high density. Taken from Zhu & Li [43]. retical frameworks are also listed, together with the empirical ranges in the last row. The binding energy and pressure from the QMF are displayed in Fig. 1 for symmetric nuclear matter and pure neutron matter with dierent nucleon radii. The EOS results within the QMF ful ll the
ow constraints from heavy-ion collisions for both symmetric nuclear matter and pure neutron matter. The nucleon radius has a weak eect on the nuclear matter even at high density. We address other important aspects before closing this section: Temperature: The above discussions are only for the zero-temperature case, below 1 MeV for cold NSs, lower than the characteristic nuclear P [MeV/fm ] B. E. [MeV] Fermi energy, while dense matter is usually hot in heavy-ion collisions and proto-neutron stars, with a temperature as high as 50 MeV. Although 22 24 the matter is expected to cool down on timescales of 10 10 seconds and 1 10 seconds, respectively, the thermal eects cannot be ignored, especially in the study of dynamic processes [e.g., 87, 88]. However, for the equilibrium con gurations of cold NSs, the EOSs are not aected much by nite temperature. For example, the temperature in
uence on the maximum mass is very limited, and there is an increase in the NS radius for a xed amount of gravitational mass [e.g., 89]. Meson-coupling parameters: The present calculations are structured to be renormalizable to x the coupling constants and the mass parameters by the empirical properties of nuclear matter at saturation. They can also be determined by tting the ground-state properties of closed-shell nuclei. In the latter case, a substantial sti EOS with an extremely high incompressibility is usually obtained, 328 MeV, which is not consistent with recent experimental results [81] (as seen in Table 1). An alternatively low compressibility usually cannot describe the nite nuclei with a proper spin-orbit coupling. Beyond mean-field: As a starting point, we choose the mean- eld ap- proximation, which should be reasonably good at very high densities (a few times the nuclear matter density). There have been studies that demon- strate that the isoscalar Fock terms could be important for the prediction of NS properties (see, e.g., Zhu et al. [90] for a study based on relativistic Hartree-Fock theory). In such models, the Lorentz covariant structure is kept in full rigor, which guarantees all well-conserved relativistic symme- tries. Additionally, the attractive Fock term introduced in the framework of QMC could eectively decrease the incompressibility at the saturation point [91]. 11 3. Neutron star NSs with typical masses M 1 3 M (where M is the mass of the sun, M = 1:99 10 g) and radii on the order of R 10 km have many extreme features that are unique in the universe [10, 11, 12, 13] and lie outside the realm of terrestrial laboratories, such as rapid rotation, extremely strong magnetic elds, superstrong gravitation, interior super
uidity and superconductivity, and superprecise spin period. These intriguing features have aroused much interest from researchers of many branches of contemporary physics as well as astron- omy because of their importance to fundamental physics. However, information regarding the NS interior has not yet been suciently revealed through the cur- rent observations due to the complexity of the NS system and many uncertain factors [92]. It is time to combine the eorts from dierent communities and discuss mutual interests and problems. In this section, we introduce the basic insights into NSs, in particular the global properties such as the mass, radius, and tidal deformability of the star, which have a one-to-one correspondence to its underlying EOS and are usually used as a tool to connect nuclear physics to astrophysics for the study of dense matter above the nuclear saturation density. [e.g., 44, 80, 64, 93, 94, 95, 96, 97, 98, 99, 100, 101, 102, 103, 104, 105, 106, 107, 108, 109, 110, 111]. A wide range of matter density from 0:1 g cm in the star atmosphere to 14 3 values larger than 10 g cm in the star core is encountered in these objects. Theoretically, the global properties are studied by using the overall EOSs as basic input and ignoring their thin atmosphere ( 0:1 10) cm, where hot X- rays originate. The observations of massive NSs [21, 22, 23, 24, 25] have already ruled out soft EOSs that cannot reach 2M . Here, this serves as a criterion for the selection of the NS (core) EOSs. The saturation properties of the employed core EOSs are collected in Table 1, with the empirical ranges listed in the last row. The determination of the EOS above the saturation density represents one of the main problems in NS study because rst principle QCD calculations are dicult to perform in such a many-body system. In most of the model 12 calculations available in the literature, a central density as high as (2 10) is found for the maximum mass, and one or more types of strangeness-driven phase transitions (hyperons, kaons, Delta isobars or quarks) may take place in the NSs' innermost parts, e.g., [47, 89, 90, 112, 113, 114, 115, 116, 158, 117, 118]. NSs with exotic phases are discussed in Sec. 4. In this section, we restrict ourselves to normal nuclear matter. 3.1. Neutron star crust 7 3 In the outer crust, at densities below 10 g cm , nuclei arrange themselves 56 7 in a Coulomb lattice mainly populated by Fe nuclei. At higher densities (10 3 11 3 g cm 4 10 g cm ), the nuclei are stabilized against beta decay by the lled Fermi sea of electrons and become increasingly neutron-rich. The composition of the outer crust is mainly determined by the nuclear masses, which are experimentally measured close to stability, whereas the masses of the very neutron-rich nuclei are not known, and they have to be calculated using nuclear models. The inner crust is a nonuniform system of more exotic neutron-rich nuclei, degenerate electrons, and super
uid neutrons. The density range extends from 11 3 14 3 4 10 g cm to the nuclear saturation density 2:8 10 g cm , at which point the nuclei begin to dissolve and merge together. Nonspherical nuclear structures, generically known as nuclear "pasta", may appear at the bottom layers of the inner crust. In fact, one of NSs' irregular behaviors, the glitch, is closely related to the inner crust EOS and the crust-core transition properties [e.g., 119, 120, 121, 122]. The crust is also crucial for NS cooling [119]. It may be necessary to calculate all EOS segments (outer crust, inner crust, and liquid core) using the same nuclear interaction, the so-called "uni ed" EOS [e.g., 123, 84, 52, 124], since matching problems in nonuni ed EOS could cause nontrivial con
icts in the predictions of the stars' properties [125]. Fig. 2 shows the crust EOS for the dierent theoretical approaches in Table 1. We observe that all outer crust EOSs display a similar pattern, with some dier- 13 Figure 2: Various EOSs for the outer crust (left) and inner crust (right). Among them, BCPM, TM1, BSk20, and BSk21 are uni ed NS EOSs, namely, all EOS segments (outer crust, inner crust, liquid core) are calculated using the same nuclear interaction. The BPS (NV) EOS for the outer (inner) crust part is indicated by the black dotted line. The BPS outer crust 7 3 11 EOS is based on a semi-empirical mass formula for matter from 10 g cm to 3.4 10 g cm [126], whereas the NV inner crust EOS is based on quantal Hartree-Fock calculations for spherical Wigner-Seitz cells [127]. ences around the densities where the composition changes from one nucleus to the next one. Only the TM1 EOS, based on an RMF model, shows a slightly dierent trend due to the semiclassic-type mass calculations, in which A and Z vary in a continuous way, without jumps at the densities associated with a change in the nucleus in the crust. A is the number of nucleons in the nucleus, and Z is the atomic number. On the other hand, the energy in the inner crust is largely determined by the properties of the neutron gas; hence, the neutron matter EOS plays an important role. Moreover, the treatment of complicated nuclear shapes, in a range of average baryon densities between the crust and the core, produces some uncertainties in the EOS of the inner crust, where some dierences are visible. In Fig. 3, we show the above discussed EOSs, with the full symbols indicating the transition point from the inner crust to the core for each chosen EOS. The APR and GM1 EOSs have to be matched with an inner crust EOS, which is at variance with the uni ed EOSs (BCPM, TM1, BSk20, and BSk21), and we achieve this by imposing that the pressure is an increasing function of the energy density. It is evident that the matching of the GM1 core (dotted black line) to 14 Figure 3: Various EOSs for the NS core from the low-density inner crust indicated with symbols. In addition to the models in Fig. 2, we include another model within the RMF, the GM1 EOS [86], as well as the present QMF model. The inner crust EOS of NV is also included. the TM1 crust (solid gray line) shows nonsmooth behavior in the dP=d slope compared to the matching to the BCPM and NV crust. Since the crust eects were shown to be more important for distorted fast-rotating stars than for static stars [128], later in Sec. 3.4 on rotating NSs, we discuss three widely used crust EOSs (TM1, BCPM, NV + BPS) that are matched with one core EOS (GM1). Note that the above crust is based on the ground state approximation for zero-temperature matter, which can only be applied to an isolated NS born in a core-collapse supernova explosion. It is assumed that during the process of cooling and crystallization, the plasma maintains nuclear equilibrium. Conse- quently, when the matter becomes strongly degenerate, the structure and EOS of the crust can be approximated well via cold-catalyzed matter. For an NS crust formed by accreted plasma from the companion star in a low-mass X-ray binary, the outermost layer of the accreted plasma undergoes thermonuclear ashes, observed as X-ray bursts, during the active stages. The layers deeper than a few meters are at T < 5 10 K, becoming increasingly neutron-rich due to electron capture and neutrino emissions and nally dissolving in the 15 liquid core. After the fully accreted crust is formed, the layered structure of the crust ceases to evolve and becomes quasistationary, with matter elements moving inwards due to compression and undergoing exothermic nuclear trans- formations [129]. There is a microscopic model for a fully accreted crust [130] that calculates the EOS and distribution of deep crustal heating sources by fol- lowing the nuclear evolution of an element of matter consisting initially of X-ray 7 3 14 3 ashes under quasistatic compression from 10 g cm to 10 g cm (crust-core interface). 3.2. Mass-radius relation To study the structure of NSs, we have to calculate the composition and EOS of cold, neutrino-free, catalyzed matter. We require that the NS contains charge-neutral matter consisting of neutrons, protons, and leptons (e , ) in beta equilibrium. Additionally, since we are looking at NSs after neutrinos have escaped, we set the neutrino chemical potentials equal to zero. The en- ergy density of NS matter can be written as a function of the dierent partial densities, "( ; ; ; ) = M + E ( ; ) =A n p e N n p 2 5=3 2 4=3 1 (3 ) (3 ) + m + + 2 2 2m 5 4 (22) where we use ultrarelativistic and nonrelativistic approximations for the elec- trons and muons, respectively, from textbooks [10]. Then, the various chemical potentials of the species (i = n; p; e; ) can be computed, = @"=@ ; (23) i i which ful lls beta-equilibrium, = b q (24) i i n i e (b and q denote the baryon number and charge of species i). Supplemented i i with the charge neutrality condition, q = 0 (25) i i 16 the equilibrium composition () can be determined at the given baryon density , and nally, the EOS is d("=) d" P () = = " = " (26) d d for the interior of NSs. The NS stable con guration in hydrostatic equilibrium can be obtained by solving the Tolman-Oppenheimer-Volko (TOV) equation [131, 132] for the pressure P and the enclosed mass m, h ih i P (r) 4r P (r) 1 + 1 + dP (r) Gm(r)"(r) "(r) m(r) = ; (27) 2Gm(r) dr r dm(r) = 4r "(r): (28) dr G is the gravitational constant. Starting with a central mass density "(r = 0) = " , we integrate out until the pressure on the surface equals that corresponding to the density of iron. This gives the stellar radius R, and the gravitational mass is then m(R) = 4 drr "(r) (29) For the description of the NS crust, we usually join the EOS P (") with the NV EOSs of Negele and Vautherin in the medium-density regime [127] and those of Baym-Pethick-Sutherland for the outer crust [126]. After solving the TOV equations, we can obtain the maximum mass M and the mass-radius TOV relation for comparison with astrophysical observations. 3.3. Symmetry energy eects on neutron star structure Currently, the EOS of SNM ( = 0) is constrained relatively well. Mat- ter with nonzero isospin asymmetry remains unknown, largely due to the un- certainty in the symmetry energy. Con
icts remain for the symmetry energy (especially its slope) despite signi cant progress in constraining the symmetry energy around and below the nuclear matter saturation density [72, 73]. The symmetry energy slope characterizes the density dependence of the symmetry 17 L = 20 MeV L = 40 MeV 1000 3.0 L = 60 MeV L = 25 MeV L = 80 MeV L = 30 MeV Λ > 800 2.5 L = 35 MeV Λ < 800 PSR J0740 + 6620 L = 40 MeV PSR J0348 + 0432 L = 60 MeV 2.0 PSR J1614 − 2230 L = 80 MeV Λ < 400 1.5 326 373 M < 2.0M max Λ = 331 PSR J0030 + 0451 1.0 0.5 100 1000 6 8 10 12 14 16 E [MeV/fm ] R [km] Figure 4: (Left) NS EOSs and (right) mass-radius relation within QMF with dierent values of symmetry energy slope L, with more L cases shown in the left panel than in the right panel. The shaded region is the favored region from ab initio calculations at the subsaturation density in chiral eective eld theory [53] and from [95]. They are causal and ful ll the two-solar- mass constraint of heavy pulsars (M > 2M ) and the tidal deformability constraint TOV of binary merger event GW170817 ( 800) for a 1:4M star. Also shown are the 1:4 latest NICER measurements from the pulse-pro le modeling of the accretion hot spots of the isolated millisecond pulsar PSR J0030+0451 [27, 28]. The general constraints from the black hole limit, the Buchdahl limit and the causality limit are also included. The gure shows that the radius sensitively depends on the symmetry energy slope with the maximum mass only slightly modi ed. A smaller L (softer symmetry energy) leads to a smaller radius. All cases of L = 20 80 MeV lie within the 800 boundary [1] and ful ll the updated limit = 190 [180] using the PhenomPNRT waveform model at a 90% con dence level. 1:4 Causality BH P < ∞ P [MeV/fm ] M [M ] energy and largely dominates the ambiguity and stiness of the EOS in NSs' high-density cores in the case of no strangeness phase transition. Fig. 4 shows our EOSs and the corresponding mass-radius relation under dierent symmetry energy slopes L in the range of 20 80 MeV. The QMF parameters are tted to reproduce the saturation properties in Table 1, with the other ve parameters ( ; E=A; K; E ) unchanged [44]. The TOV mass 0 sym of the star hardly changes with changing L and ful lls the recent observational constraints of three massive pulsars for which the masses are precisely mea- sured [21, 22, 23, 24, 25]. There is a strong positive correlation between the slope parameter and the radius of a 1:4M star [for more discussion, see, e.g., 96, 133]. However, a small dependence is found in Hornick et al. [133]. The cases of L 30 60 MeV in our QMF model may be more compatible with the neutron matter constraint [53]. Capano et al. [134] found that the radius of a +0:9 1:4M NS is R = 11:0 km (90% credible interval), assuming a descrip- 1:4 0:6 tion in terms of nuclear degrees of freedom remains valid up to 2 . The recent NICER measurements of PSR J0030+0451 [27, 28] might indicate L & 40MeV. The EOS governs not only the stable con guration of a single star but also the dynamics of NS mergers. During the inspiral phase, the in
uence of the EOS is evident on the tidal polarizability [135, 136]. In Fig. 4, we also include the calculated results of the tidal deformability and the constraining region from binary merger event GW170817, namely, 800 for a 1:4M star [1]. 1:4 The tidal deformability describes the magnitude of the induced mass quadrupole moment when reacting to a certain external tidal eld. It is zero in the black hole case. The dimensionless tidal deformability is related to the compactness 2 5 M=R and the Love number k through = k (M=R) (see more discussion 2 2 later in Sec. 5.1). To study the eects of the symmetry energy slope L in more detail, we present the resulting Love numbers (tidal deformabilities) as a function of the mass and the compactness for dierent L in Fig. 5 (Fig. 6). In Fig. 5, k rst increases and then decreases with mass and compactness. In Fig. 6, monotonously decreases with the mass and compactness. The increase in k 19 0.200 L = 20 MeV L = 20 MeV 0.175 0.175 L = 40 MeV L = 40 MeV L = 60 MeV L = 60 MeV 0.150 0.150 L = 80 MeV L = 80 MeV 0.125 0.125 0.100 0.100 0.075 0.075 0.050 0.050 0.025 0.025 0.000 0.0 0.5 1.0 1.5 2.0 2.5 0.05 0.10 0.15 0.20 0.25 M [M ] M/R Figure 5: Love numbers as a function of the mass (left) and compactness (right) for four EOSs with dierent values for the symmetry energy slope L (20, 40, 60, 80 MeV). k rst increases and then decreases with mass and compactness. The increase in k (below 1.0M ) is due to large radii and a large portion of soft crust matter. The vertical line and colorful dots indicate M = 1:4 M . Taken from Zhu et al. [44]. 3000 3000 L = 20 MeV L = 20 MeV L = 40 MeV L = 40 MeV 2500 2500 L = 60 MeV L = 60 MeV L = 80 MeV L = 80 MeV 2000 2000 Λ 1500 Λ 1500 1000 1000 500 500 0 0 0.5 1.0 1.5 2.0 0.05 0.10 0.15 0.20 0.25 M [M ] M/R Figure 6: Same as Fig. 5 but for the tidal deformabilities. monotonously decreases with the mass and compactness. Similar to Fig .5, the large values of for small masses (below 1.0M ) are due to large radii and a large portion of soft crust matter. Taken from Zhu et al. [44]. 2 and large values of for small masses (below 1.0M ) are due to large radii and a large portion of soft crust matter. If no crust is considered (e.g., an EOS described by a pure polytropic function), k still decreases monotonously with mass and compactness. Further loud gravitational-wave measurements from merging binary NSs would provide data with good precision for learning more about the slope parameter as well as the NS structure. Moreover, the nal fate of the merger, i.e., prompt or delayed collapse to a black hole or a single NS star, depends on the EOS, as well as the amount of ejected matter that undergoes nucleosynthesis of heavy elements. These discussions are presented in Sec. 5. 3.4. Rotating neutron star NSs are usually rotating, and the rotational periods P of rapidly rotating NSs (pulsars) could provide restrictions on the EOSs and their evolution pro- cesses when combined with the mass constraint. When rapidly rotating, an NS is
attened by the centrifugal force, and the TOV equation, suitable for a static and spherically symmetric situation, cannot correctly describe the rotating stel- lar structure. We assume NSs are steadily rotating and have an axisymmetric structure. Based on the axial symmetry, the space-time metric used to model a rotating star can be expressed as 2 2 2 2 2 ds = e dt + e dr + e r d 2 2 2 +e r sin (d !dt) ; (30) where ; ; and ! is the function of r; . The matter inside the star is approx- imated by a perfect
uid, and the energy-momentum tensor is given by T = (" + p)u u pg ; (31) where "; p and u are the energy density, pressure and four-velocity, respectively. To solve Einsteins eld equation for potentials ; ; and !, Komatsu et al. [137] transformed the Einstein equation from dierential equations to integrals by using the Green function method. In this form, the asymptotic
atness condition, which is the boundary condition of the Einstein equation, can be 21 Figure 7: NSs' masses as a function of central energy density (left) and radius (right) for three cases of crust EOSs (TM1, BCPM, NV + BPS) matching one GM1 core EOS, with the detailed EOS matching data shown in Table 1. The calculations are performed for both the static case and Keplerian rotating case. The maximum masses and radii, as well as the central densities, hardly depend on how the inner crusts are described for NSs heavier than 1:0 M . satis ed automatically. This method for solving the Einstein equation is written as a standard code. This is the well-tested RNS code . Using tabulated EOSs, the stationary and equilibrium sequences of rapidly rotating, relativistic stars can then be computed in general relativity [see more detail about the code in, e.g., 137, 138, 139]. The Keplerian (mass-shedding) frequency f is one of the most studied physical quantities for rotating stars. An EOS that predicts Kepler frequencies that are smaller than the observed rotational frequencies is to be rejected, as it is not compatible with observation. An empirical formula was proposed in Lattimer & Prakash [11], 1 3 2 2 M R f = f ; (32) K 0 M 10 km where M is the gravitational mass of the Keplerian con guration, R is the http://www.gravity.phys.uwm.edu/rns/ 22 radius of the nonrotating con guration of mass M , and f is a constant that does not depend on the EOS. An optimal prefactor f = 1080 Hz was found in [140, 141] for NSs as well as hybrid stars. See more discussion in [140, 141] regarding the justi cation of the functional form of Eq. (32) and its valid range. The calculated highest spin frequencies f are all higher than 1000 Hz, while the current observed maximum is f = 716 Hz [20] for PSR J1748-2446a in the globular cluster Terzan 5. A possible reason for this discrepancy is that the star uid is suering from r-mode instability [142]. A simple estimation showed that this would lower the maximum frequency by 37%, which might satisfactorily explain the observations to date. Fig. 7 shows the crust eects on the star's mass-radius relations in nonuni ed EOSs, where three widely used crust EOSs (TM1, BCPM, NV + BPS) are matched with one core EOS (GM1). It is clear that for both the static case and Keplerian rotating case, the results hardly depend on how the inner crusts are described. This is true not only for the maximum mass and central densities but also for the radii. For less massive stars, crust-core matching has a slightly larger eect on the radii, and the TM1 curve deviates slightly from the other two due to the relatively larger dierence in the crust-core interface for TM1 mentioned before. This deviation may be relevant only for NSs' masses smaller than 1:0 M . Generally, rotation increases both the gravitational mass and the radius. Based on the EOSs collected in Table 1, rotation can increase the star's gravi- tational mass up to 18 19%, and the star can be as massive as 2:61M in the APR case. Additionally, the star becomes
attened, and the corresponding circumferential radius is increased up to 3 4 km, i.e., 29 36%. For lighter stars such as 1.4 M , the radius increase is more pronounced, reaching 5 6 km, i.e., 41 43%. Additionally, rotation lowers the central density from 7 10 to 6 9 , which is due to the eect of the centrifugal 0 0 force, eectively stiening the EOS. We show in Fig. 8 the gravitational mass as a function of the central density at various xed rotation frequencies based on the QMF EOS. In static stars, QMF18 gives M = 2:08M at a central TOV 23 2.50 2.25 2.00 1/P=0.4 1.75 1/P=0.6 1.50 1/P=0.8 1/P=1.0 1.25 1/P=1.2 1.00 1/P=1.4 1/P=1.6 0.75 1/P=1.8 0.50 5 10 15 14 − 3 ε [× 10 g cm ] Figure 8: NSs' masses as a function of the central energy density with the QMF EOS at various xed rotation frequencies 1=P =0.4-1.8 kHz. The lower blue curve is the static case, and the upper red curve corresponds to the Keplerian frequencies at dierent rotating cases. The change in the maximum mass M with frequency is indicated with a dashed black crit curve. In static stars, QMF gives M = 2:08M at a central density = 6:92 with a TOV 0 corresponding radius 10:5 km. At Keplerian frequency f = 1699 Hz, the maximum mass and corresponding radius with QMF are 2:50M and 14:0 km at a central density = 8:21 . Curves with a xed baryonic mass of M = 2:2; 2:4; 2:6; 2:8M are also shown with nearly horizontal gray curves. density = 6:92 with a corresponding radius of 10:5 km. At Keplerian fre- c 0 quency f = 1699 Hz and with QMF18, the maximum mass is 2:50M with a corresponding radius of 14:0 km at a central density of = 8:21 . c 0 One of the most interesting rotating stars is the so-called \supramassive" star, which exists only by virtue of rotation. It is well known that the onset of the instability of the static sequence is determined by the condition dM=d = 0, i.e., the curve should stop at the maximum value of gravitational mass M . In the rotating case, the above criteria have to be generalized, i.e., a stellar con guration is stable if its mass M increases with increasing central density for a xed angular momentum J . Therefore, the onset of the instability, which is called the secular axisymmetric instability, is expressed by @M = 0: (33) Since rotation increases the mass M that a star of a given central density can TOV support, the static con guration with the baryon mass M > M (the baryon M [M ] 2.50 2.50 2.25 2.25 2.00 2.00 1.75 1.75 J=0.5 1.50 J=1.0 1.50 J=1.0 J=1.5 1.25 J=1.5 1.25 J=2.0 J=2.0 J=2.5 1.00 1.00 J=2.5 J=3.0 J=3.0 J=3.5 0.75 0.75 J=3.5 J=4.0 0.50 0.50 10 11 12 13 14 15 16 17 5 10 15 14 − 3 R [km] ε [× 10 g cm ] Figure 9: (Left) NS masses as a function of the central energy density and (right) mass- radius relations with the QMF EOS at various xed angular momenta J . The lower blue curve is the static case, and the upper red curve corresponds to the Keplerian frequencies in dierent rotating cases. In the left panel, the change in the maximum mass M with the crit angular momentum is indicated with a dashed black curve. A spin-down star, losing angular momentum over its evolution, follows the lines with xed baryonic mass M , shown by the nearly horizontal gray curves for M = 2:2; 2:4; 2:6; 2:8M . mass of a TOV mass star) does not exist. Such sequences are supramassive stars that are doomed to collapse as they lose energy and angular momentum during their spin-down, following the nearly horizontal line of xed baryonic mass M . We show in Fig. 9 the NS mass as a function of the central energy density as well as the mass-radius relations with the QMF EOS at various xed angular momenta J . The lower blue curve is the static case, and the upper red curve corresponds to the Keplerian frequencies in dierent rotating cases. In the left panel, the change in the maximum mass M with the angular momentum is crit indicated with a dashed black curve. There may be a universal relation between M =M and j=j [e.g., 143, 144, 145, 146, 147, 148] that does not depend crit TOV K on the speci c choice of EOS or the f value, 2 4 M j j crit = 1 + a + a (34) 2 4 M j j TOV K K where j = J=M is the dimensionless angular momentum and the coecients 1 2 are a = 1:316 10 and a = 7:111 10 [143]. 2 4 Note that the above discussions focus only on the case of rigid rotation, M [M ] M [M ] while dierential rotation can be much more ecient in increasing the maximum allowed mass. In dierentially rotating stars, the high-density inner part may rotate faster than the low-density outer part, so the inner part can be supported by rapid rotation without the equator having to exceed the Keplerian limit [e.g., 149]. While both rigid and dierential rotation follow axisymmetry, there are cases when a rotating NS breaks its axial symmetry if the rotational kinetic energy to gravitational binding energy ratio, T=jWj, exceeds a critical value. The abovementioned r-mode instability could also trigger NSs' motion with o- axis symmetry. It is presently unclear whether such con gurations of NSs can actually be realized in practice [e.g., 150]. Overall, it is especially important to calculate models of rotating stars to better understand the observations of binary merger events (see details in Sec. 5.2). 4. EOS with exotic particles 4.1. Hyperon star and hyperon puzzle While around the saturation densities = , the matter inside an NS con- sists only of nucleons and leptons, at higher densities, several other species of particles may appear due to the fast increase in the baryon chemical poten- tials with density [33, 35, 36], just because their appearance is able to lower the ground state energy of the dense nuclear matter phase. Among these new 0; 0; particles are strange baryons, namely, the ; ; hyperons. Other species (such as kaons and Delta isobars) might also appear in stellar matter, which the present paper does not cover. Generally, the presence of one species of strange particle is found to push the onset of other species of strange particles to higher densities, even out of the physically relevant density regime [e.g., 65, 112]. It is necessary to generalize the QMF study of the nuclear EOS with the inclusion of hyperons [e.g., 45, 46, 47, 49, 50, 48]. The density thresholds of hyperons are essentially determined by the masses and their interaction. The + 0 mass of (uds) is 1116 MeV. The masses of (uus), (uds), (dds) are 1189, 1193, and 1197 MeV, respectively. The masses of (uss) and (dss) 26 are 1315 and 1321 MeV, respectively. From hypernuclei experiments in the lab- oratory [19], we know that -nucleus and interactions are attractive, while -nucleus interactions are repulsive. Additionally, the nature of the -nucleus interaction has been suggested to be attractive [151]. Theoretically, any eective many-body theories should respect the available hypernuclei data before pro- ceeding with other sophisticated studies [35, 36]. The adopted hyperon-meson couplings need to at least reproduce unambiguous hypernuclear data. At the mean- eld level, the single , , potential well depths in symmetric (N ) nuclear matter are U 30, 30, 14 MeV at the saturation density, re- ;; spectively. In the extended QMF model [46, 47], we introduce dierent con ning strengths for the s quarks and the u; d quarks in the corresponding Dirac equa- tions (under the in
uence of the meson mean elds). The con ning strength of the u; d quarks is constrained by nite nuclei properties, and that of the s quarks (N ) is constrained by the well-established empirical value of U 30 MeV. The mass dierence among baryons is generated by taking into account the spin cor- relations E = e + E , and the spin correlations of the baryons are xed B i i spin by tting the baryon masses in free space. In addition, the spurious c.m. motion 2 2 is removed through the usual square root method as M = E hp i. The B B cm contribution of the meson is contained in the eective mass M , and the ;; ! and mesons couple to the baryons with the following coupling constants: q q q g = 3g ; g = cg = 2g ; g = g (35) !N ! ! ! ! ! ! q q q g = g ; g = 0; g = 2g ; g = g (36) where a factor c is introduced before g for a large ! coupling, in contrast with the ! coupling, to simulate the additional repulsion on the nucleon (N ) channel, and U = 30 MeV at the nuclear saturation density. The basic q q q parameters are the quark-meson couplings (g , g , and g ), the nonlinear self- coupling constants (g and c ), and the mass of the meson (m ) [for more 3 3 detail regarding the formalism and the parameters, see 40, 47]. With such a parameter set, the saturation properties of nuclear matter can be described [47]. (N ) The values of a single hyperon in nuclear matter are obtained as U = 12 27 Figure 10: (a) Single hyperon potential, (b) fractions of leptons and baryons, (c) EOS, and (d) mass-radius relations for NSs with hyperons within QMF. The cases without hyperons in the star's core are also shown in the lower panels. The maximum mass of QMF EOS without hyperons is slightly lower than 2M due to the absence of a high-order vector coupling term for eective nuclear interaction from an earlier work [47]. When hyperons are included, the mass is largely reduced and well below the observational 2-solar-mass limit. The hyperon puzzle is also present in many microscopic studies based on developed realistic baryon-baryon interactions [e.g., 55, 113, 153, 154]. Adapted from Hu et al. [47]. MeV, consistent with the BNL-E885 experiments [151]. The density depen- dences of the single hyperon (; ; ) potentials are shown in Fig. 10(a). Regarding the EOS of hyperonic matter, the baryon contributions can be obtained through the mean- eld ansatz from the Lagrangian (including hyper- ons) [40, 47]. Electrons are again treated as a free ultrarelativistic gas, whereas the muons are relativistic, as in Eq. (22). The total EOS can be calculated for a given composition of baryon components. This allows the determination of the chemical potentials of all species, which are the fundamental input for the 28 equations of chemical equilibrium: = = 0 = 0 (37) = (38) = = + (39) n e p + = = (40) n e The above equations must be supplemented with two other conditions, i.e., charge neutrality and baryon number conservation. These are + + = + + + ; (41) p e = + + + + + + + + : (42) n p Finally, the actual detailed fraction Y = () of the dense matter is determined i i for each xed baryon density , as shown in Fig. 10 (b). In the low-density region (until < 0:21 fm ), the proton fraction = is well below 1=9, which ful lls the astrophysical observation that direct URCA cooling might not occur at too of low densities [152]. With the properly chosen ; and hyperon potentials, is the rst hyperon appearing at 2 3 . Then, hyperons appear at 3 followed by hyperons at 7 . The fractions of hyperons increase with 0 0 density. At densities above 10 , the fractions of and are almost the same as the fractions of protons and neutrons. , however, does not appear until very high density of 2:0 fm . In Fig. 10 (c), we show the pressure of beta-equilibrated matter as a function of the energy density. The solid curve represents the EOS including the hyper- ons, and the dot-dashed curve is the EOS without hyperons. The EOS becomes softer with the presence of strangeness freedom. The NS properties are calcu- lated by using the EOSs with/without hyperons obtained from the QMF model, and the NS mass-radius relations are plotted in Fig. 10 (d). It is found that the maximum mass of the NSs including hyperons is approximately 1:6 M , much lower than that without hyperons, which is below the observational limit. This is the so-called hyperon puzzle, which is also found in many microscopic studies based on developed realistic baryon-baryon interactions [e.g., 55, 113, 153, 154]. 29 Since hyperons are not present in nuclear matter, they cannot be determined from the nuclear matter properties. Moreover, the analysis of experimental data on hypernuclei shows that we cannot x these parameters in a unique way. How can a suciently sti high-density EOS generate a heavy hyperon star with properly reproduced nuclear matter properties at the saturation density? There may be three approaches forward: 1. Three-body hyperon interactions can be introduced in microscopic studies or high-order meson elds in eective calculations. If they are repulsive, a sti enough hyperon star EOS can be obtained by increasing the repulsion as the density increases, and there is a possibility of massive hyperon stars with a central density > 5 , [e.g., 91, 155, 156, 157]. This is a natural solution based on the known importance of three-body nucleon forces in nuclear physics; 2. Larger maximum masses can be produced through a transition to another phase of dense (quark) matter in the stellar core at high densities [e.g., 117, 118, 158]. This approach is presented in Sec. 4.2; 3. A separate branch of pulsar-like objects can be introduced to account for the heavy ones, for example, QSs made of free quarks [e.g., 159, 160, 161]. Unlike NSs, which are bound by gravity, QSs are bound by strong interactions; therefore, they have opposite M-R dependence. This is the so-called two-branch scenario [e.g., 162, 163], which is discussed in Sec. 4.3. 4.2. Strange quark matter and hybrid stars The matter inside the NS core possesses densities ranging from a few times to one order of magnitude higher. At such densities, the hadronic matter might undergo a phase transition to quark matter, and a hybrid NS with a quark matter or mixed core can be formed. However, the exact value of the transition density to quark matter is unknown and still a matter of recent debate not only in astrophysics but also within the theory of high energy heavy-ion collisions. Additionally, it is not obvious whether the information on the nuclear EOS from high energy heavy-ion collisions can be related to the physics of NS 30 interiors. The possible quark-gluon plasma produced in heavy-ion collisions is expected to be characterized by low baryon density and high temperature, while the possible quark phase in NSs appears at high baryon density and low temperature. Nevertheless, we must be careful that the transition cannot occur at too low of density (below the nuclear saturation density ) to maintain consistency with the current experimental data of heavy-ion collisions. The possibility of the existence of strange quark matter in NS high-density cores is of special interest in the present era of GW astronomy [e.g., 9, 101, 164, 165, 166, 167, 168, 169, 170, 171, 172, 173, 174]. Presently, we have no uni ed models to address the hadron phase and the quark phase, and it is still not clear whether the change in the hadron phase corresponding to that in the quark phase is a crossover or a rst-order transition. Here, we analyze a speci c example in the context of a rst-order transition (at transition pressure P ) trans to express the experimental constraints in model-independent terms. For the hadronic sector, we use the above QMF model. For the high-density quark phase, we utilize the CSS parametrization [175], exploiting the fact that for a considerable class of microscopic quark matter models, the speed of sound is weakly density-dependent, e.g., [176, 177, 178]. The present scheme can only discuss the transition that occurs at a sharp interface (Maxwell construction) between bulk hadronic matter and quark matter, i.e., the quark-hadron surface tension is high enough to disfavor mixed phases (Gibbs construction). It has been shown that a strong rst-order phase transition with a sharp interface is the most promising scenario to be tested or distinguished from pure hadronic matter by future observations [9, 101, 171]. We tend to nd that the observation of a two-solar-mass star and the accurate measurement of the typical NS radius constrain the CSS parameters, including the squared speed of sound in the high- density phase c , the hadron-quark phase transition density , and the trans QM discontinuity in the energy density at the transition "=" , where trans trans (P ); " " (P ). NM trans trans NM trans 31 Figure 11: EOSs (left) and mass-radius relations (right) for hybrid stars (colorful curves) at a xed discontinuity in energy density at the transition "=" = 0:5 for dierent transition trans densities = = 1:25; 1:5; 2; 3; 4; 5; 6, with the symmetry energy slope 60 MeV. The trans 0 squared speed of sound is xed at c = 1 in the quark matter. The corresponding NS QM results within the QMF are shown for comparison (black curves). The shaded region is the favored region from the maximal model [54], with the underlying EOSs constrained at low densities from EFT, facilitating the complete allowed parameter space for the speed of sound above the saturation density n and enforcing the LIGO/Virgo constraint from GW170817 (70 720) [29]. For a given nuclear matter EOS " (P ), the full CSS EOS is NM " (P ) P < P NM trans "(P ) = " (P ) + " + c (P P ) P > P NM trans trans trans QM 2 2 We perform the calculation by varying c from the causality limit (c = QM QM 1) to the conformal limit (c = 1=3, the value for systems with conformal QM symmetry that may be applicable to relativistic quarks). It is worth mentioning that perturbative QCD calculations exhibit quark matter with c values of QM approximately 0.2 to 0.3 [179]. We use units where ~ = c = 1. In Fig 11, we show representative EOSs P (") for dense matter with a sharp rst-order phase transition and the corresponding mass-radius relations. The two-solar- mass lower limit for maximum gravitational mass is explicitly indicated in the mass vs. radius plot. We include the curves with increasing transition density from 1:25 to 6 at a xed energy density discontinuity and speed of sound 0 0 in quark matter, and the nuclear matter EOS is chosen to be the QMF model result with the symmetry energy slope L = 60 MeV. We mention here that L 30 60 MeV is the preferred range within QMF as indicated by the ab 32 initio calculations (shown in Fig. 4). We see that a lower transition density (pressure), therefore a stier EOS, leads to a heavier hybrid star. The smallest hybrid star is typically the heaviest. Systematically, we carry out calculations for the mass-radius of hybrid stars spanning the whole parameter space of the speed of sound in a domain with a transition density up to = 6 and an energy density discontinuity up trans 0 to " = 1:5 " . The calculations are performed using two values for the trans symmetry energy slope parameter of L = 30 MeV and L = 60 MeV, and the results are shown in Fig. 12. Figure 12 displays the correlation of the radius of a 1:4 M hybrid star R with the transition density = (upper panels) and the maximum 1:4 trans 0 mass M (lower panels). In general, there exists an anticorrelation between TOV R and = and a correlation between R and M . A conservative 1:4 trans 0 1:4 TOV upper limit of 13:6 km for R can be obtained with dierent analyses [e.g., 1:4 94, 95, 97, 125, 180, 181, 182, 183]. In the upper panel, the upper limit of 13:6 km for R corresponds to 1:3 for L = 30 MeV and 1:5 for L = 60 1:4 0 0 MeV. The possible onset of a rst-order phase transition below such densities might be strongly disfavored. From the mass measurement of heavy pulsars, we can put lower limits on R by making use of the R M correlation in 1:4 1:4 TOV the lower panel. The two-solar-mass constraint leads to a lower limit of 9:6 km, in good concurrence with other analyses in the literature based on X-ray observations or LIGO/Virgo measurements [e.g., 94, 97, 180, 181]. An upper limit on the maximum mass can also be indicated from R < 13:6 km of 1:4 M < 3:6M . TOV We conclude this section by further discussing the following aspects: Submillisecond rotation: It is commonly believed that only self-bound stable QSs may rotate rapidly with a submillisecond period [188]. How- ever, it is suggested that pulsars rotating with approximately half a mil- lisecond period could also be interpreted as hybrid stars [189], with NSs containing a metastable decon ned quark phase at their centers. This 33 Δε/ ε t r a n s 1 5 S o l i d : L = 6 0 M e V D a s h e d : L = 3 0 M e V 0 . 1 1 4 1 3 . 6 k m 0 . 2 1 3 0 . 4 0 . 0 R 0 . 0 0 . 2 1 2 0 . 1 0 . 3 0 . 3 0 . 4 1 1 0 . 5 0 . 5 1 0 1 . 0 1 . 5 2 . 0 2 . 5 3 . 0 ρ / ρ t r a n s 0 Q M 1 5 1 . 0 1 4 1 . 0 1 3 0 . 6 1 2 0 . 4 0 . 4 1 1 0 . 8 0 . 8 1 0 1 . 0 1 . 0 1 . 5 2 . 0 2 . 5 3 . 0 3 . 5 4 . 0 M [ M ] T O V 0 . 6 Figure 12: Upper: Radius of a 1:4M star vs. the transition density, with the energy density discontinuity explicitly indicated; Lower: Radius of a 1:4M star vs. the maximum mass, with the sound speed explicitly indicated. The results are shown for two kinds of symmetry energy slopes 60 MeV (solid lines) and L = 30 MeV (dashed lines). The horizontal line represents a conservative upper limit of 13:6 km for a 1:4M star with or without a phase transition [e.g., 94, 95, 97, 125, 180, 181, 182, 183]. The vertical line in the lower panel represents the two- solar-mass constraint. There are cases when no 1:4M star is possible, shown by breaks in the curves. Taken from Miao et al. [184]. conclusion does not depend on the quark matter EOS models. Therefore, very rapidly rotating pulsars may be interpreted as either QSs or NSs with decon ned quark matter interiors. Mixed phase: In the present work, we adopt the simple Maxwell con- struction. The Gibbs construction provides a realistic model of the phase transition between the hadronic and quark phases inside the star [190], yielding a range of baryon densities where both phases coexist, which pro- vides an EOS containing a pure hadronic phase, a mixed phase, and a pure QM region [e.g., 118, 191]. The pressure is the same in the two phases to ensure mechanical stability, while the chemical potentials of the dierent species are related to each other, satisfying chemical and beta stability. Both the hadron and quark phases are separately charged while preserving total charge neutrality [190]. As a consequence, the pressure is a monotonically increasing function of the density. The realization of the mixed phase depends on the nuclear surface tension, which is currently an unknown parameter [192]. The Gibbs treatment is the zero surface tension limit of the calculations, including nite-size eects. It was demonstrated that the in
uence of dierent constructs on the maximum mass value is rather small [191]. 4.3. Quark stars and two-branch scenario We now turn to the description of the bulk properties of uniform quark mat- ter. The strange quark matter is composed of up (u), down (d) and strange (s) quarks with charge neutrality maintained by the inclusion of electrons (hereafter muons as well if present): 2 1 1 = 0; (43) u d s e 3 3 3 The baryon number conservation, ( + + ) = ; (44) u d s 35 is also satis ed with n being the baryon number density. Due to the weak interactions between quarks and leptons, d ! u + e + ~ ; u + e ! d + ; e e s ! u + e + ~ ; u + e ! s + ; e e s + u $ d + u ; The -stable conditions = = + should be ful lled in neutrino-free s d u e matter. The energy density and pressure include both contributions from quarks and leptons, and those of leptons can be easily calculated by the model of an ideal Fermi gas such as in the NS matter case. In the density regime achieved inside compact stars, the dense matter prop- erties cannot be calculated directly from the rst principle lattice QCD or per- turbative QCD. The latter is only applicable at ultrahigh densities beyond the range of compact stars. The current theoretical description of quark matter is based on phenomenological models [e.g., 159, 160] and burdened with large uncertainties. In the following, we consider the nonperturbative contributions from perturbative QCD [193]. For simplicity, we use the pQCD thermodynamic potential density to the order of [194], pt ; (45) 0 1 s with g m i 2 i 4 = 6 ln + 4 [u v ln(u + v )] + 3 [u v ln(u + v )] 2v ; 1 i i i i i i i i 12 m i=u;d;s (46) where u =m and v u 1. The coupling constant and quark i i i i s masses m run with the energy scale and can be determined by [194], 1 ln L () = 1 ; (47) L L 1 1 0 m () = m ^ 1 + : (48) i i s s Here, L ln , and we take the MS renormalization point = 376:9 MS MS MeV based on the latest results for the strong coupling constant [195]. Following 36 Eq. (48), the invariant quark masses are m ^ = 3:8 MeV, m ^ = 8 MeV, and u d m ^ = 158 MeV. The parameters for the -function and
-function are = s 0 1 2 1 38 1 202 20 (11 N ), = (102 N ),
= 1=, and
= ( N ) [196] f 1 2 f 0 1 2 f 4 3 16 3 16 3 9 (The formulas are for arbitrary N , and in this study, N = 3). It is not clear f f how the renormalization scale evolves with the chemical potentials of quarks, and we use = , with C = 1 4 [193]. i 1 3 i We also introduce the bag mechanism to account for the energy dierence between the physical vacuum and perturbative vacuum, and the bag parameter is dynamically scaled [e.g., 197, 198]. The total thermodynamic potential density for strange quark matter can be written as [192], pt + B " # pt i + B + (B B ) exp QCD 0 QCD (49) where we take B = 40; 50 MeV=fm [59] for the calculations and = 1 indicates no medium eect for the bag parameter. If and m run with the S u;d;s energy scale as reported by the particle data group [195], the maximum mass of QSs does not reach 2M . In such cases, the dynamic rescaling of the bag con- stant with nite is essential, which basically originates from nonperturbative eects such as chiral symmetry breaking and color superconductivity [6, 7, 199]. B = 400 MeV=fm is demanded by the dynamic equilibrium condition at QCD the critical temperature of the decon nement phase transition and is obtained 2 4 by equating the pressures of the QGP ( B + 37 T =90) and pion gas QCD 2 4 ( T =30) at T = T ( 170 MeV). At given chemical potentials , the pressure P , particle number density , i i and energy density " are determined by: P = ; (50) g 3 @ i 1 2 2 = m + ; (51) i s 0 i i 6 @ " = + : (52) i i 37 Q M F p Q C D 1 0 0 0 m a x c / c Q M f 1 . 0 e 1 0 0 0 . 9 2 0 . 8 0 1 0 0 . 6 8 1 0 0 0 1 0 0 0 0 ε ( M e V / f m ) 1 . 0 0 . 9 0 . 8 0 . 7 0 . 6 C o n f o r m a l l i m i t 0 . 5 0 . 4 1 0 0 0 1 0 0 0 0 ε ( M e V / f m ) Figure 13: EOSs (left) and sound speed c (right) of SQM. The EOSs are generated in the QM perturbation model, ful lling the available astrophysical constraints of mass [21, 22, 23, 24, 25], radius [27, 28] and tidal deformability [29] for QSs. They are compared with the results of the perturbative QCD (red curve) without nonperturbative corrections. The EOS of the nuclear matter obtained with the QMF (blue curve) is shown in the left panel. The maximum sound max speeds c are explicitly indicated. The horizontal lines in the c plot show the conformal QM QM limit. To ensure a large mass for QSs above two solar mass, the sound speed is necessarily large, c =c > 0:68. Taken from [169]. QM The common term for the particle number density in Eq. (51) is obtained with C @ dm 1 0 1 i = + 0 s 3 @m @m d i i C @ C d @B 1 1 1 s + + : (53) s 1 3 3 @ @ d In Fig. 13, we show the EOSs generated in the perturbation model, which ful ll the available astrophysical constraints of mass [21, 22, 23, 24, 25], ra- dius [27, 28] and tidal deformability [29] for QSs. They are compared with the results of pQCD without nonperturbative corrections, taking C = 1 4 and B = 0. The EOSs for the nuclear matter obtained with the QMF models are also shown. [200] pointed out that if the two-solar-mass constraint is combined with the hadronic matter EOS below and around the nuclear saturation density, c might rst increase then decrease after reaching a maximum (maybe even QM up to 0:9c) and nally approach the conformal limit c= 3 from below. This peculiar shape resembles the analysis of the case of crossover EOS [201]. To en- max sure a large mass for QSs above two solar mass, the obtained peak value (c ) QM ranges from 0.68c to c, similar to previous results for NSs [193, 202, 203, 204]. We show in Fig. 14 the various properties of strange quark matter based on the perturbation model using the exemplary parameters of C = 3:5; B = 40 1 0 MeV/fm , and = 800 MeV. The composition, binding energy, pressure, sound velocity, and scaled bag parameter are shown as functions of the baryon density or chemical potential. Note that in the binding energy plot, the con- dition that the strange quark matter be the absolute stable strong-interaction system, requiring that at P = 0; E=A M ( Fe=56) =930 MeV, is ful lled. Fig. 15 shows the predicted properties of the QSs, including the mass, Love number, and tidal deformability. Previously, although the quark star EOS models could reach 2 solar mass, they required a too small surface density (not much larger than the nuclear satu- ration density) in some cases, and a larger maximum mass meant an even smaller surface density (because of the anticorrelation between the two [205]), for ex- ample, the CDDM2 model [206] and the PMQS3 model [205]. Then, the radius 39 0 . 6 1 2 0 0 7 0 0 6 0 0 0 . 5 1 1 5 0 5 0 0 d ) 0 . 4 1 1 0 0 u e V 4 0 0 / e i M 0 . 3 ( 1 0 5 0 3 0 0 s A 0 . 2 1 0 0 0 2 0 0 5 6 0 . 1 9 5 0 M ( F e ) / 5 6 1 0 0 0 . 0 9 0 0 0 0 1 2 3 0 . 0 0 . 4 0 . 8 1 . 2 1 2 0 0 1 6 0 0 - 3 - 3 μ ( M e V ) ρ ( f m ) ρ ( f m ) 1 2 0 0 1 . 0 4 0 0 1 0 0 0 0 . 8 3 0 0 8 0 0 f V c c o n f o r m a l l i m i t / 0 . 6 6 0 0 M M Q 2 0 0 c P 0 . 4 4 0 0 1 0 0 0 . 2 2 0 0 0 . 0 0 0 1 2 3 0 1 2 3 0 1 2 3 - 3 - 3 - 3 ρ ( f m ) ρ ( f m ) ρ ( f m ) Figure 14: Various properties of betastable strange quark matter, including the quark frac- tions, the binding energy, the pressure, the sound velocity, and the scaled bag parameter, which are plotted as a function of the baryon density or chemical potential. The calculations are performed based on the perturbation model using the parameters C = 3:5; B = 40 1 0 MeV/fm , and = 800 MeV. The horizontal line in the E=A plot indicates that of the most stable Fe nucleus. The horizontal line in the c plot shows the conformal limit. QM ρ ρ 2 . 5 2 . 5 2 . 0 2 . 0 1 . 5 1 . 5 1 . 0 1 . 0 0 . 5 0 . 5 0 . 0 0 . 0 0 . 0 0 . 5 1 . 0 1 . 5 4 6 8 1 0 1 2 - 3 R ( k m ) ρ ( f m ) 0 . 2 5 1 0 0 0 1 . 4 M s t a r 0 . 2 0 1 0 0 1 . 4 M s t a r 0 . 1 5 0 . 1 0 1 0 0 . 0 5 0 . 0 0 1 0 . 1 5 0 . 2 0 0 . 2 5 0 . 3 0 0 . 3 5 0 . 1 5 0 . 2 0 0 . 2 5 0 . 3 0 0 . 3 5 M / R M / R Figure 15: Various properties of QSs based on the EOS in Fig. 14, including the gravitational mass, the Love number, and the tidal deformability, which are plotted as a function of central density, radius, or compactness. The crosses in the upper two panels show where the maximum mass is reached. The locations of a 1:4M star are explicitly indicated in the lower two panels. (and the tidal deformability) exceeded the observational values [1, 27, 28, 29]. There models were not welcomed by particle physicists studying hadrons (for which experiments have been established studying the nonperturbative eects) because in such a density realm, the quarks are thought to be very dilute and are very possibly con ned inside hadrons. In the real world, we do obtain nu- clear matter rather than quark matter around the nuclear saturation density. The present perturbative model with an in-medium bag can achieve both a rea- sonable surface density and a maximum mass as large as 2:2M . The predicted properties of dense matter (c ; ) and quark stars (R; M ), as well as the EOS QM of Fig. 14, are collected in Table 2. We conclude this section by further discussing the following aspects: QS vs: hybrid stars: Although it is known that the degree of freedom is hadronic around the nuclear saturation density, the QCD phase state for cold, dense matter of intermediate densities is unfortunately unknown, and a great deal of eort is being applied in the communities of astrophysics, Table 2: QS EOS with proper sound velocity behavior and the predicted properties of dense matter (c ; ) and quark stars (R; M ). The calculations are done based on the perturbation QM model using the parameters C = 3:5; B = 40 MeV/fm , and = 800 MeV. See Sec. 4.3 1 0 for details. 3 3 3 (fm ) " (MeV/fm ) P (MeV/fm ) c =c R (km) M=M QM 0.277 254.68 3.22 0.5365 23.053 3.834 0.0530 0.298 274.56 8.98 0.5400 9.2107 6.241 0.2350 0.332 306.51 18.42 0.5481 5.2979 8.078 0.5320 0.366 340.66 28.88 0.5590 3.9983 9.123 0.7994 0.401 376.61 40.38 0.5725 3.3848 9.778 1.0253 0.436 413.89 52.94 0.5886 3.0548 10.205 1.2125 0.471 452.00 66.56 0.6073 2.8730 10.488 1.3669 0.506 490.42 81.22 0.6288 2.7826 10.675 1.4945 0.539 528.59 96.90 0.6535 2.7564 10.796 1.6004 0.571 566.01 113.56 0.6816 2.7802 10.871 1.6890 0.601 602.20 131.15 0.7134 2.8460 10.914 1.7638 0.629 636.77 149.60 0.7487 2.9465 10.933 1.8274 0.655 669.49 168.87 0.7865 3.0713 10.936 1.8821 0.663 679.97 175.46 0.7994 3.1152 10.934 1.8986 0.686 710.22 195.71 0.8364 3.2383 10.923 1.9439 0.708 739.02 216.62 0.8665 3.3126 10.904 1.9837 0.728 767.11 238.17 0.8818 3.2818 10.879 2.0186 0.749 795.72 260.33 0.8741 3.0995 10.851 2.0495 0.771 826.60 283.12 0.8405 2.7687 10.819 2.0766 0.795 862.04 306.60 0.7856 2.3523 10.785 2.1000 0.824 904.94 330.86 0.7197 1.9348 10.749 2.1199 0.859 958.68 356.09 0.6532 1.5756 10.710 2.1360 0.904 1027.09 382.51 0.5930 1.2958 10.668 2.1483 0.959 1114.25 410.42 0.5421 1.0915 10.623 2.1565 1.054 1266.67 450.58 0.4896 0.9136 10.558 2.1609 42 nuclear physics, and particle physics due to the crucial importance of this aspect. One key point is still not clear: Does the matter go through a phase transition from hadron matter to quark matter at some densities (which is relevant to compact star physics) or is quark matter the absolute ground state of strongly interacting matter (the conjecture of Bodmer- Witten [207, 208])?. Therefore, there remains the problem of how to verify QSs or distinguish them from NSs or hybrid stars [209, 210]. For a xed gravitational mass, hybrid stars are characterized by a smaller radius than their hadronic counterparts and could be as compact as QSs for masses above 1:0M . The similarity of the sound speed of the hadron-quark mixed phase with that of the pure quark matter in the intermediate density region of 3 8 , a particular shape with a peak, further complicates the distinguishing of QS from hybrid stars [169]. Two branch scenario: Because of the tension of a low tidal deformability +390 +0:20 (190 [180]) and a high maximum mass (2:14 M for the presently 120 0:18 heaviest pulsar [25] and 2:35M based on the numerical simulation studies on NS binary mergers [211, 212, 213]) for a certain EOS in the NS model, binary QSs have been proposed as the possible scenario for the GW170817 event [100, 214]. A binary QS merger for some binary con gu- rations could eject amounts of matter (comparable to the binary NS case) to account for the electromagnetic observations in the optical/infrared/UV bands (namely, kilonova) [215]. A magnetar with QS EOS is preferred as the post-merger remnant for explaining some groups of short gamma-ray burst (SGRB) observations [206, 205]. It has been suggested that because of this discrepancy (if con rmed), small and large stars of the same mass could coexist as hadronic and quark matter stars [162, 163]. Comments on the maximum mass of NS=QS : Various microscopic cal- culations of NS matter (without strangeness) indicate a possible upper limit of 2:3 2:4M for the NS maximum mass, for example, Brueckner theory calculations [153] and quantum Monte Carlo calculations [56]. Ex- 43 otic particles, if they are present, usually lower the limit as a result of the extra degrees of freedom during the phase transition (while the appearance of quarks might increase the limit in the case of crossover [104, 201]). The quark-hadron crossover EOS gives a maximum mass of 2:35M [201]. The bound of M . 2:3 2:4M may be applicable to QSs. For example, TOV the maximum mass of QSs is 2:18M (2:32M with color-
avor-locked su- per
uity [216]) within the MIT bag model [100]. The present perturbation max model yields 2:24M with the peak sound speed (c ) approaching the QM speed of sound (we expect an increase in the value including the uncertain super
uity of 0:1M ). These high theoretical limits on the maximum mass are higher than (but close to) the observational bound of pulsars of approximately 2:14M [25]. A looser upper limit based on extreme causal EOSs may be in the range of M < 3:6 4:8M [54, 185, 186, 187]. TOV The observations of accreting black holes, on the other hand, revealed a paucity of sources with masses below 5M [e.g., 217, 218, 219, 220]. Presently, binary mergers involving one or two companions have masses that fall into the so-called mass gap range (3 5M ) that are hard to distinguish [e.g., 221, 222, 223]. 5. Neutron star binary The gravitational waves (GWs) detected from binary neutron star (BNS) merger event GW170817 [224], as well as its electromagnetic (EM) counter- parts [225], announced the beginning of the multimessenger astronomy era. In addition to hinting at the origin of SGRB [1, 226] and revealing the site of r-process nucleosynthesis [2, 227], our knowledge of the EOS of cold dense matter at supranuclear densities has been greatly enriched. In the past year, various studies have been performed to constrain the EOS of dense matter, ei- ther by putting constraints on observable characteristics of NSs [i.e., radius or tidal deformability; see e.g., 95, 97, 180] or by connecting the constraint with model parameters in nuclear physics [i.e., symmetry energy slope or neutron skin parameter; see e.g., 44, 228]), which could be tested by nuclear physics 44 experiments. Some studies also go beyond the conventional NS scenario and put constraints on compact star models involving strong interaction phase tran- sitions [e.g., 97, 100, 171, 181]. Those constraints mainly come from 3 aspects from a BNS merger event. First, during the inspiral stage, unlike binary black hole (BBH) mergers, defor- mation is induced for each NS due to the tidal eld of the companion, providing additional dissipation of the orbital energy and angular momentum and hence accelerating coalescence [224]. This deformation therefore leaves a detectable signature in the GW signal of the late inspiral stage, from which we can learn about the tidal deformability of the NS EOSs [229]. Second, the detection of SGRB hints at a delayed collapse to a BH for the merger remnant [230, 231]. This interpretation of the SGRB observation provides information on the maxi- mum mass of a nonrotating con guration for the NS EOS (namely, M ). For TOV instance, the EOS should not be too sti; otherwise, the remnant supramassive NS lives much longer [232] in the magnetar central engine model [233, 234]. However, if the EOS is too soft, the merger might result in a prompt collapse to a BH. In such occasions, the magnetic eld might not be enhanced suciently (by a dierentially rotating NS remnant) and thus not able to launch a jet [212]. Third, the features of the transient optical/infrared/UV observations (namely, the kilonova) powered by the radio activity of the neutron-rich elements in the ejecta depend directly on the mass, velocity and electron fraction in the ejecta, which is related to the properties of the EOS for the merging NS. In this section, we brie
y review the information we have learned about the EOS of NSs from the BNS merger events GW170817 and GRB170817A as well as AT2017gfo. 5.1. GW170817 and tidal deformability The nite size eects of NSs alter the late inspiral GW signal compared with that of the BBH case [229, 235]. Through the leading order, the GW observations constrained a combination of the tidal deformability for each NS in the binary ( and ) [224]. 1 2 45 4 16 (12q + 1) + (12 + q)q 1 2 = ; (54) 13 (1 + q) in which q = M =M is the mass ratio of the binary. The dimensionless tidal 2 1 deformability of each star is 2 R = k ( ) ; (55) 3 M where k is the tidal Love number describing the fraction between the induced quadrupole moment of a star and the external tidal eld and R and M are the radius and mass of the star, respectively. On the other hand, k can be obtained for a given EOS for any given mass and hence can be tested with the observation of GW170817. Practically, the tidal deformability is tted to the GW observation together with other parameters [29]. For instance, in the Taylor F2 post-Newtonian aligned-spin model, 13 parameters need to be tted, including 7 extrinsic pa- rameters (sky location, distance of the source, polarization angle, inclination angle and coalescence phase and time) and 6 intrinsic parameters (mass of each NS, tidal deformability of each NS and the aligned spin of each NS). Therefore, the uncertainties in the estimation of other parameters weakly correlate with the determination of the tidal deformability. Hence, the constraint on the tidal deformability is normally made with certain assumptions. For instance, as seen in [224], the assumption in the spin parameter of the NS could signi cantly aect the interpretation of the mass of each NS, thus further aecting the constraint on the tidal deformability. Through the assumption that the of each NS vary independently, the rst constraint on and could 1 2 be obtained under dierent spin priors. The result favors the softer EOS, i.e., the EOS that predicts more compact stars. Another analysis assumes a uniform ~ ~ prior for , which sets an upper limit of 800 on in the low-spin case and 700 in the high spin case. Alternatively, through the expansion of (M ) around a certain M , constraints can be directly placed on the tidal deformability of a certain mass star. This constraint is (1:4) 1400 in the high-spin case and (1:4) 800 in the low-spin case. 46 Follow-up analysis further improves these constraints under more assump- tions. For example, in [180], instead of assuming an independent and uniform prior for the tidal deformability of each star, the EOS of each star in the merging binary is assumed to be the same. Consequently, the area of the 90% con dence region in the - parameter space shrinks by a factor of 3. This also im- 1 2 proves the determination of (1:4) to 190 . In [29], a lower cut-o frequency of 23 Hz is used instead of the 30 Hz in the original analysis. Although the tidal eects mainly aect the GW signals above several hundred Hz, a lower frequency cut o allows for the better determination of other parameters, hence improving the measurement of the tidal deformability. In [183], it was pointed out that under the assumption that two stars in a binary system have a common EOS, there is an approximate relation between the tidal deformability of each star, i.e., = = q , where q is the mass ratio. With the aid of this relation, 1 2 once the assumption in the mass ratio of the binary is made, the tidal deforma- bility can be further constrained. In [183], the improved analysis shows that +420 +453 is 222 for a uniform mass prior and 245 for a mass prior inferred from 138 151 observed double neutron star systems and 233 for a mass prior informed by all galactic neutron star masses within the 90% credibility level. We can directly test existing EOS models by simply calculating the tidal deformability and comparing it with the observational constraint. Nevertheless, more insight could be gained regarding the EOS of neutron-rich matter by more systematically studying the impact of the EOS (i.e., p as a function of ", where p and " are the pressure and rest mass density of the matter) on the tidal deformability. Such interpretations are presented in [e.g., 95, 97, 180]. For instance, in [180], the logarithm of the adiabatic index of the EOSs is treated as a polynomial of the pressure for the high density EOS, namely, = (p;
) and
= (
;
;
;
) are free parameters. For densities below half the nuclear i 0 1 2 3 saturation density, the EOS is connected to the SLy EOS [236]. The sampling of the EOS then consists of uniformly sampling
in certain intervals. For each of the EOS samples, the mass radius relation and tidal deformability could be theoretically obtained and constrained by the observation of both the tidal 47 145 150 (2) a = 0.61 2 flavor line 125 125 0.4 0.5 0.6 0.7 0.8 0.9 1.0 10 20 30 40 50 60 70 80 90 100 4 Δ [MeV] Figure 16: Parameter space for QS EOS models within the MIT bag model (B ; a ; ) for e 4 normal QSs (left) and super
uid QSs (right) obtained by combining the GW170817 constraint on (1.4), the two-solar-mass constraint on M and the stability window for quark matter. TOV With the constraint of the "2
avor" line, we ensure that normal atomic nuclei do not decay into nonstrange quark matter. With the constraint of the "3
avor" line, we ensure that strange quark matter is more stable than normal nuclear matter, namely, Bodmer-Witten's conjecture [207, 208]. The perturbative QCD correction parameter a characterizes the degree of the quark interaction correction, with a = 1 corresponding to no QCD corrections (Fermi gas approximation). a = 0:61 is chosen to be close to the calculated result with dierent choices of the renormalization scale [237]. The eective bag constant (B ) also includes a phe- nomenological representation of nonperturbative QCD eects. Due to the strong correlation between M and (1:4), a lower bound can be inferred for (1:4) from the two-solar-mass TOV limit, namely, 510 (380) MeV for normal (super
uid) QSs. Taken from Zhou et al. [100]. deformability and 2 solar mass pulsars [21, 22, 23, 24, 25]. The constraint on the +1:4 neutron star radius is R = 11:9 km for the merging binary of GW170817. 1:4 Similar analysis can be found in, e.g., [95], which shows that the radius of a 1.4 solar mass NS is in the range of [9:9; 13:8] km. However, it is worth noting that such analysis might be aected by the choice of EOS priors. In [97], it was pointed out that when the prior for possible twin star (for which there is a hadron-quark phase transition inside the star) branch EOSs is considered, the radius becomes less constrained, i.e., R 2 [8:53; 13:74] km. 1:4 In addition to systematic studies on parameterized EOS priors, phenomeno- logical models can be applied to interpret tidal deformability constraints. Ac- cording to [228], a better upper limit for neutron skin eects is obtained com- pared with that of the experiment done by PREX, and better results could be achieved with future GW observations and terrestrial nuclear physics experi- ments. Both the symmetry energy parameter and the symmetry energy slope Λ(1 .4) = 600 Λ(1 .4) = 800 Λ(1 .4) = 1400 2.01 M 2.2 M Λ(1 .4) = 600 Λ(1 .4) = 800 Λ(1 .4) = 1400 2.01 M 3 flavor line 2.2 M 2 flavor line 3 flavor line 1/4 B [MeV] eff 1/4 B [MeV] eff Figure 17: Postmerger product fractions for (a) stable star, (b) supermassive star and (c) black hole for the NS and QS EOS models, labelled with the star type plus the corresponding M : TOV Uni ed BSk21 (red line labelled NS2.28), nonuni ed GM1 (blue line labelled NS2.37), and MIT model (green and purple lines labelled QS2.08 and QS2.48). In panel (b), the observed 22% constraint for supermassive stars from Gao et al. [240] is shown by the horizontal line for comparison. The vertical dotted line in the same panel is for a typical initial period of 1 ms. Taken from Li et al. [205]. are better constrained with respect to previous nuclear physics studies [44]. Un- der the assumption that GW170817 originates from a binary quark star (BQS) merger, the quark interaction parameters are studied in [100]. Fig. 16 shows the results of both normal and super
uid QSs. It is worth noting that due to the nite surface density of QSs, a surface correction needs to be taken into account when calculating of a QS [238]. Therefore, a QS could have a larger TOV maximum mass without violating the tidal deformability constraint compared with those of NSs [239]. 5.2. GRB170817A and merger remnant BNS mergers have long been proposed as the central engine of SGRBs [226]. This suggestion has been veri ed by GRB170817A detected by Fermi/GBM and INTEGRAL/SPI-ACS, which accompanies the detection of GW170817. According to the time of the merger implied from the chirp signal, there is a 49 1:74 0:05 s delay for the onset of the SGRB [2]. The detection of GRB170817A not only helps the determination of the location of the source, which allows for abundant follow-up observations in other bands, but also provides useful in- formation about the post-merger evolution of the merger event, thus providing constraints on the EOSs [e.g., 240, 241, 242]. Exemplary fractions of the outcome of the binary are shown in Fig. 17 using both NS and QS EOSs. The dependence on the EOS, as well as the initial period, is evident [205]. It is found that the fraction of stable star (panel (a)) is determined by the static maximum mass M . The fractions of supermassive star (panel (b)) and black hole (panel (c)) TOV are further sensitive to the initial period since the fast-rotating con gurations of the star have to be taken into account for them [206]. Depending on the TOV maximum mass of the NS EOS and the total mass of the merging binary, there could be 4 dierent outcomes after the merger: if the total mass of the binary system (M ) is much larger than M , tot TOV the direct formation of a black hole (BH) on a dynamic time scale, namely, prompt collapse, occurs. The total binary mass, above which prompt collapse can occur, is denoted as the threshold mass (M ); thres if M is smaller than M but larger than the maximum mass that can tot thres be reached by uniformly rotating NSs (denoted M ), then a short-lived supra NS could exist as a remnant supported by dierential rotation . The dierential rotation dissipates due to magnetorotational instabilities as well as viscosity within a timescale of 100 ms, and then the NS collapses to a BH; if M is smaller than M but larger than M , the remnant is a tot supra TOV long-lived supramassive NS. The uniformly rotating NS could still lose angular momentum by magnetic dipole radiation, but it takes a much longer time to suciently reduce the angular momentum to induce the Such NSs are called hypermassive NSs (HMNS). NSs that can be supported by only uniform rotation are called supramassive NSs (SMNS) 50 collapse to a BH; if M is smaller than M , a stable NS remnant exists. tot TOV The GW observations do not provide any hint to which scenario applies to the case of GW170817 due to the lack of post-merger GW observations [224, 243]. The electromagnetic counterparts, on the other hand, can indicate what happens after the merger of the two NSs. A very robust interpretation is that scenario 1) should be excluded due to the SGRB detected. According to Ruiz et al. [212], in the BH central engine model for SGRBs, enhancement in the magnetic eld of the merger remnant due to the dierential rotation of the hypermassive NS is essential for jet formation. A prompt collapse results in a magnetic eld that is too weak to explain the SGRB observations. Hence, the detection of GRB170817A directly implies M < tot M for the case of GW170817. The total mass of the binary can be measured thres by the inspiral GW signal, which places a constraint on M . thres For a given EOS model, the threshold mass can be determined by perform- ing numerical simulations with dierent total binary masses. This can be ex- tremely time-consuming for full 3-dimensional general relativistic simulations. In Bauswein et al. [94], the so-called smooth particle hydrodynamics (SPH) method as well as a conformally
at assumption for the spacetime metric is used to reduce the simulation time to make it plausible. According to the re- sults, the threshold mass is related to M and R (i.e., the radius of the TOV TOV TOV maximum mass con guration) as TOV M = ( 3:38 + 2:43)M : (56) thres TOV TOV Alternatively, the results can be tted with similar accuracy in terms of the radius of a 1.6 solar mass star (R ) as 1:6 TOV M = ( 3:606 + 2:38)M : (57) thres TOV 1:6 Note that this reveals a quadratic relation between M and M once thres TOV R (or R ) is xed. Particularly, since the coecient of the M term is TOV 1:6 TOV 51 negative, there exists a maximum value for M . This maximum possible value thres of M must be larger than M in the case of GW170817; otherwise, there thres tot is no parameter space to prevent a prompt collapse. In other words, any choice of R (or R ) that results in a maximum possible value M smaller than TOV 1:6 thres 2.74 solar mass should be excluded by the observations. This requires R to TOV be larger than 9.26 km and R to be larger than 10.30 km. 1:6 For NS EOS models, scenario 4) can also be excluded. Scenario 4) requires a very large M , which results in a large tidal deformability, hence violating TOV the tidal deformability constraint . Distinguishing between scenarios 2) and 3) could indicate more on the TOV maximum mass of NSs; however, this is quite model dependent. Under dierent SGRB central engine model assumptions, totally opposite conclusions could be drawn. If the SGRB originates from a BH central engine, as assumed in [212], the delay collapse has to occur within 1.7 s after the merger. Therefore, scenario 2) most likely occurred for the case of GW170817. As shown by previous studies [143], the ratio between M supra and M is almost universal for various NS EOS models, and the value is TOV approximately 1.2. Combining the total mass of the merging binary, an upper limit [approximately 2.15-2.25 M according to dierent studies, e.g., 211, 212, 244, 245] could then be obtained for the TOV maximum mass. In addition, it has been pointed out that the merger remnant might collapse to a BH with angular momentum smaller than that of the Keplerian limit due to angular momentum transfer by post-merger GW emission, neutrino and viscous eects, which leads to a dierent constraint on the TOV maximum mass. With this in mind, Shibata et al. [213] performed an analysis by considering conservation laws of baryonic mass, energy and angular momentum, and the constraint on the TOV maximum mass is found to be 2:10 M < M < 2:35 M . It TOV is worth noting that this constraint is valid only under the BH central engine assumption. In fact, it has been pointed out that the multimessenger observation of GW170817 is consistent in a magnetar central engine model [246] and could Note that this possibility still remains for QSs, as discussed in [239] 52 even be favored by an X-ray activity detected very long time after the merger [247]. In such a magnetar central engine model, in contrast, scenario 3) is favored, and hence, the upper limit of M mentioned before becomes a lower TOV limit instead. 5.3. AT2017gfo and ejecta properties It has long been suggested that the BNS merger is an important site for the production of heavy elements in the Universe [227]. R-process nucleosynthesis is expected to occur in the neutron-rich matter ejected during the merger. The radioactive decay of such neutron-rich isotopes could power a transient in opti- cal/UV/IR, i.e., a kilonova [248, 249]. Such a transient event (AT2017gfo) was detected several hours after the merger time of GW170817 [2], the luminosity, spectrum and light curve of which are consistent with the prediction of the kilo- nova model. Such a kilonova detection not only enriches our knowledge about the abundance of heavy elements in the Universe but also greatly increases our understanding of NS EOSs. The observational properties, for example, the peak luminosity and peak time, of the kilonova are closely related to the ejecta properties (c.f. [250]). The abundance of lanthanides (atomic numbers from 58 to 71) is strongly related to the electron fraction (Y ) of the ejecta. On the other hand, the opacity of the ejecta is mainly determined by the lanthanides in it, and is hence related to the electron fraction of the ejecta. The overall luminosity is related to the amount of radioactive heavy elements and thus the total mass of the ejecta. The ejecta is expanding at a certain velocity and becomes translucent after a period of time. Therefore, the characteristic duration of a kilonova is related to the velocity of the ejecta. To summarize, the mass, velocity and electron fraction of the ejecta are key parameters for understanding the observations of AT2017gfo. The ejecta during a BNS merger mainly consists of two components. The rst component is the so-called dynamic ejecta, which is normally more neutron- rich (lower Y ). Part of the dynamic ejecta is due to the tidal torque during the inspiral [251, 252]; hence, it has a lower temperature and very low Y (smaller 53 than 0:1 0:2) and is spatially distributed in the equatorial plane of the binary. Another part of the dynamic ejecta results from shock during coalescence (also called shock-driven ejecta) [253, 254]. Due to the higher temperature at coales- cence, this part of the dynamic ejecta normally has a slightly higher electron fraction (Y > 0:25) [255, 256] and can expand in the polar direction. In addition to the dynamic ejecta, the neutrino emissions from the remnant before collaps- ing to BH as well as the viscosity could further drive more ejecta (wind-driven ejecta) from the disc surrounding the remnant [257, 258, 259]. Due to neutrino irradiation, this part of the ejecta has a broader distribution of Y , which could be as high as 0.5 [260, 261, 262]. Clearly, the amount of wind-driven ejecta is dependent on the lifetime of the remnant NS. For instance, in the case of a prompt collapse, the wind-driven ejecta could be signi cantly suppressed. The kilonova observation following GW170817 has shown clear evidence of two distinct ejecta components [263, 250, 264] , an early rising ( 2 days after the merger) and bluer component (which indicates a lower opacity and higher velocity) and a more extended redder component. The required amount of ejecta accounting for the "blue" component is approximately 0:01M with a relatively blue larger electron fraction Y > 0:25 and velocity v 0:2 0:3c. For the red component observed at later times, in total, approximately 0:05M lanthanide- red rich (Y < 0:25) ejecta is needed, with a lower velocity of v 0:1 0:2c. The inferred properties can be used to constrain the EOS of the merged NSs, although this constraint is quite model dependent. One property we can use to constrain NS EOS models is the mass of the ejecta, as it is related to the properties of the merging binary. As summarized in [265], stier EOS models (for which the tidal deformability is larger) typically have a smaller amount of tidal-induced dynamic ejecta than softer EOS models. However, softer EOSs normally eject more dynamic ejecta overall because of a more violent coalescence and hence eject more shock-driven ejecta. The amount of wind-driven ejecta depends on the lifetime of the merger remnant before Note that there are studies arguing a model with 3 components (c.f. [341]). 54 collapsing to a BH, which is determined again by the M of the NS EOS. TOV Ideally, the details of the ejecta property for BNS mergers with dierent EOS models could be determined through extensive numerical simulations. However, to fully resolve the wind-driven ejecta, a very long-term post-merger simulation with the implementation of viscosity and neutrino cooling is required, which is normally extremely time consuming and not aordable in practice. Neverthe- less, conservative estimations can still be made. In [266], it has been found that the total dynamic ejecta plus all the mass in the disc surrounding the remnant has a positive correlation with the parameter of the binary. As not all the matter in the remnant disc is ejected, the dynamic ejecta plus the remnant disc mass has to be larger than the inferred mass of the ejecta to explain the observation of AT2017gfo, which is approximately 0.05 solar mass. This sets up a lower limit for the dynamic ejecta mass plus the remnant disc mass and hence a lower limit for the binary tidal deformability. In [266], this conservative constraint is > 400. However, in the more systematic study of [267], which employs a set of more general parameterized EOS and considers unequal-mass binaries, a contradiction was found. In other words, it was shown that a binary system with < 400 could still be consistent with the luminosity of AT2017gfo in terms of the ejecta mass. Therefore, this lower limit must be considered with caution. Another implication of the observation of AT2017gfo is the fate of the merger remnant (the 4 scenarios mentioned in the previous subsection). This observed "blue" component of the ejecta clearly rules out the possibility of a prompt col- lapse, in which case there is a negligible amount of high Y shock-driven ejecta and wind-driven ejecta. In such a case, the kilonova observation should be dom- inated by the tidal ejecta and thus should be red. Distinguishing whether the remnant is long-lived is very uncertain and model dependent. In [244], it was suggested that if a long-lived SMNS is produced, then a signi cant amount of the rotational kinetic energy of the SMNS is injected into either a collimated relativistic jet or the ejecta within tens of seconds. This extra energy injec- tion is considered to be inconsistent with the observations of GRB170817A and 55 AT2017gfo. Therefore, the authors of [244] believe an HMNS or very short-lived SMNS is produced in the remnant and put a similar upper limit on M of TOV 2.17 M . Nevertheless, in [246], it was shown that with a long-lived SMNS as the merger remnant, both the early and late emission of AT2017gfo can be ex- plained without requiring an unrealistically low opacity and high ejecta mass. Similar arguments can also be found in [245]. A long-lived SMNS remnant is believed to be able to provide strong neutrino emissions to reduce the lanthanide contamination in our line of sight as well as to produce enough ejecta with rel- atively high speed in the post-merger phase. Particularly, a temporal feature observed 155 d after the merger in the X-ray afterglow provides a more direct hint supporting a long-lived remnant. Considering the possibility of dierent merger outcomes, the allowed range for the TOV maximum mass for NSs could actually be larger [268]. 5.4. NS vs QS in the multimessenger era We have summarized some of the many studies on NS EOSs in light of GW170817 and its EM counterparts. Nevertheless, it is worth noting that the phase diagram of strong interactions at supranuclear densities is not yet clearly understood due to the nonperturbative nature of QCD at low energy scales. Apart from conventional NS models, other models involving strong interaction phase transitions are suggested, e.g., twin stars or strange stars [175, 269, 270]. As is summarized below, a binary quark star (BQS) scenario could be totally consistent with the observation of GW170817 and its EM counterparts. Due to the self-bound nature, QSs have a very large surface density. This leads to many signi cant dierences between QSs and NSs. For example, when supported by uniform rotation, QSs can reach an even higher maximum mass with respect to their TOV maximum mass (40% more) than NSs (20% more) [271]. The nite surface density leads to a correction when calculating the tidal deformability [135]. QSs could reach a much higher T=jWj ratio when rotating, which could lead to more signi cant GW radiation in the post-merger phase [100]. As a result, the above analysis on NS models should not be directly applied to QSs. 56 It is interesting and useful to understand the constraints on QS models from what we have learned from GW170817 and its EM counterparts and how to distinguish between NS/QS models in the multimessenger era. Qualitatively, the tidal deformability measurement constrains the stiness of QSs, similar to the case of NSs. Stier EOSs normally reach higher M TOV but also have larger tidal deformability than a softer EOS due to the larger size of the star described by a stier EOS. There is an overall positive correlation between M and in NS models [95]. Investigating QS properties based on TOV the MIT bag model reveals a similar relation between M and (1:4) [150]. TOV However, the quantitative results are quite dierent. In [95], creating NS EOS models sti enough to reach M 2.8 M with (1:4) 800 was found to be TOV impossible. In [239], it was shown that a self-bound strange star model can be sti enough to reach M > 3 M without violating the tidal deformability TOV constraint. It was suggested that a BQS merger should result in a clean environment with little or no hadronic dynamic ejecta [272]. It is not easy to verify this argument with numerical simulations, as the density discontinuity on the QS surface is dicult to handle numerically. Nevertheless, in [215], several BQS merger simulations were performed with the SPH method, and it was shown that with some binary con gurations, a BQS merger could eject a comparable amount of (quark) matter in the case of BNS mergers. According to [273, 274], under certain conditions (i.e., if the baryon number of an ejected quark nugget is smaller than a critical value), quark matter could eciently evaporate into normal nucleon matter and contribute to the kilonova observation [214, 275]. In addition, as uniformly rotating QSs can reach a higher maximum mass, the post-merger remnant is likely to be longer lived than the case of the BNS merger. It has been found that a magnetar with QS EOS as the post-merger remnant is better for understanding the internal X-ray plateau observations following SGRBs [206]. In addition, both dierentially rotating and uniformly rotating triaxial QSs are found to be sucient GW emitters [100, 149], which could be tested by future GW observations. 57 2.4 2.2 2.0 APR 1.8 SFH0 DD2 1.6 ALF2 1.4 1.2 1.0 9 10 11 12 13 14 15 16 R [km] Figure 18: The combined constraints on NS EOSs from the multimessenger observations of GW170817, GRB170817A and AT2017gfo. The gray shaded regions are excluded by the tidal deformability measurement of GW170817 [243, 95]. The red shaded regions are forbidden because the GRB170817A and AT2017gfo observations exclude a prompt collapse after merger [94]. The solid horizontal line is the 2.01 solar mass lower limit for M according to the TOV observation of the massive pulsar [22], and the dotted horizontal line is the 2.17 solar mass constraint. Note that if the SGRB is powered by a BH central engine (or magnetar central engine), the dotted horizontal line is an upper limit (or lower limit). The M-R relation of several commonly used NS EOSs is shown in the gure. Another interesting possibility is a BNS merger that leads to a conversion of the merger remnant to a QS. In such a case, the inspiral GW signal and dynamic ejecta properties should be exactly the same as the case of a normal BNS merger, whereas the post-merger behavior could be quite dierent. On the one hand, if the phase transition occurs partially inside the star (i.e., only the high-density core part of the remnant), a softening of the EOS occurs, hence reducing the lifetime of the merger remnant as well as shifting the f peak in the post-merger GW signal to a higher frequency compared with those of the purely nucleonic merger remnant case [166, 168]. On the other hand, if the entire star could be converted to a QS after merger, which results in a stiening of the EOS, the lifetime of the remnant is longer, and the f frequency is smaller [276]. In this scenario, the time delay between the merger and the SGRB is believed to be the time needed for converting the surface of the remnant to quark matter, which signi cantly reduces the baryon load in the environment, thus helping M [M ] the formation of a collimated jet [277]. Overall, the possibilities of a BQS merger or BNS merger with a QS remnant are consistent with the multimessenger observations of GW170817. However, current knowledge of the details of mergers involving QSs is quite limited. More simulations need to be performed for a better and more complete understanding in the future. With the help of numerical results and more future observations, whether QS could be involved or formed in a merger event could be distin- guished, particularly according to the post-merger GW signals. 5.5. Conclusion To summarize, the multimessenger observation of BNS merger GW170817 has greatly increased our knowledge about the EOS of dense matter. The most robust constraint is from the tidal deformability encoded in the GW signal during the inspiral. Such a tidal deformability measurement could translate into a constraint on the radius of the NS at a given mass. The EM counterparts contain large amounts of information on the EOS models, the most reliable of which is to exclude the prompt collapse scenario. This provides independent constraints on the neutron star radius for a given mass. Other constraints on the lifetime of the remnant NS, however, depend on the central engine model of the SGRB and are not reliable. The constraints could be totally opposite in the BH central engine model and magnetar model, and the current observations could not reliably rule out either possibility for GW170817. We summarized all the constraints mentioned above, which is shown in Fig. 18. Nevertheless, GW170817 is just the beginning of the multimessenger era. As an increasing number of GW signals and EM counterparts from BNS mergers are detected in the future, our knowledge of NS and even QS EOS models will be enriched. In particular, if the time delay between the merger and the collapse to BH could be more robustly determined by future observations (either by post-merger GW signals or by more extensive EM counterpart observations), the TOV maximum mass can be crucially inferred, leading to a much better understanding of the EOS of dense neutron-rich matter. 59 6. Other opportunities from compact objects 6.1. Neutron star cooling With the ever-increasing accuracy of observational instruments, more details of the signals emitted by NSs can be quantitatively monitored. Apart from the measurements of NS masses, radii and tidal deformabilities, high-density NS models can be confronted with the surface temperatures of isolated NSs of known or estimated ages and the thermal photon luminosity of the X-ray transients in quiescence with an estimated time-averaged accretion rate on the NS [92, 152, 279, 278, 280]. The NS EOS determines the stellar structure, as well as the eective masses and super
uid gaps of baryons, and is therefore crucial for the heat capacity and neutrino emission rate [e.g., 152, 281, 282]. In the Newtonian framework, the energy balance equation for NS cooling is written as [278], dE dT th = C = L L + H; (58) dt dt where T and C are the stellar internal temperature and the total heat capac- ity, respectively. The loss of the thermal energy E occurs through neutrino th emission (total luminosity L ) and photon emission (total luminosity L ). H represents all possible energy sources to heat the star, such as the decay of the magnetic eld energy stored in stars. Current simulations of thermal evolution are usually based on a general relativistic formulation, and some robust codes have already been established , which comprises all the relevant cooling reac- tions: direct URCA (DU) (n ! p + l + ; p + l ! n + ), modi ed URCA l l (MU) (n + N ! p + N + l + ; p + N + l ! n + N + ), nucleon-nucleon l l bremsstrahlung (NNB) (N + N ! N + N + + ), Cooper pair breaking and formation (PBF) processes (N +N ! [NN ] + +). Exemplary cooling curves of a 1:4M NS [283] are displayed in Fig. 19. In the quiescent state of X-ray transients, the accreted matter sinks gradually in the interior of the NS and undergoes a series of nuclear reactions [129]. These http://www.astroscu.unam.mx/neutrones/ NSCool/ 60 Figure 19: Cooling curves of a canonically isolated NS within the minimal cooling paradigm [338], without including fast neutrino emissions, charged meson condensate, hy- perons, or con nement quarks in canonical NSs. The stellar structure is built with the APR EOS [57]. The calculations are carried out in three cases for a comparison: without any Z factors (the Fermi surface depletion due to the SRC), with Z factors only in super
uidity, with Z factors both in super
uidity and neutrino emission and with Z factors. Dong et al. [283] found that the SRC eect reduces the neutrino emissivity for the DU, MU, NNB and PBF processes, as well as the heat capacity of the stellar interior [283]. reactions release some heat, which propagates into the whole NS, inwardly heat- ing the core and outwardly emitted in the form of photons at the surface. This is the so-called deep crustal heating. The heating curves of X-ray transients can be derived, relating the L in quiescence to the estimated time-averaged accretion rate M [284]. The relevance of the pasta phase, which is beyond the neutron drip density, to explaining some X-ray transients (if con rmed) might be re- garded as smoking-gun evidence of the NS model for pulsar-like objects [e.g., 285] and disfavors the alternative QS model. Above all, a reliable theory for NS cooling, in combination with accurate observations, is indispensable for gaining important information about the stel- lar interior. The complexity of NS systems, such as anisotropic magnetic elds and the compositions of the stellar core and envelope, is not controlled well theoretically, currently rendering the task of providing reliable and quantitative predictions extremely dicult, and considerable eort might be achieved in the future. 61 6.2. Pulsar glitch and glitch crisis A glitch (sudden spin-up) is a common phenomenon in pulsar observations and was discovered during pulsar timing studies in the Vela pulsar [287]. Since then, the number of known glitches has greatly increased, with more than 555 glitches now known in more than 190 pulsars. The observed glitches are collected 8 9 by the Jodrell Bank Observatory and the ATNF Pulsar Catalogue . During glitches, the pulsar spin frequency suddenly increases over a very short time and then usually relaxes to its preglitch rate over a longer time. The glitch size, often de ned as the relative increase in the spin frequencies during glitches, 10 5 , has a bimodal distribution ranging from 10 to 10 with peaks at 9 6 10 and 10 [288, 289]. The glitches in young pulsars, including magnetars, are generally large [290]. However, the youngest pulsars, e.g., the Crab pulsar, tend to have more frequent and smaller glitches. Various authors have used observed glitch properties as a probe to investigate the pulsar inner structure, i.e., the EOS of dense matter [291]. The physical mechanism behind glitches is suggested to be a sudden de- crease in the NS moment of inertia, which could result from the coupling and decoupling between the (observed) charged component (rotating slower; labelled as index p) and the super
uid component (rotating faster; labelled as index n) [292]. The fractional moment of intertia I =I is related to the glitch activ- n p ity A = (1=T )( )= as follows, g p p 2 A . ; (59) c g where T is the total data span for glitch monitoring and ( ) is the sum of all observed glitch frequency jumps. = =2 is the characteristic c p p age of the pulsar. The glitch activity A can be estimated for systems that have exhibited at least two glitches of similar magnitude, like the Vela pul- sar. The glitch observations from the Vela pulsar place a constraint on the An \antiglitch", i.e., an abrupt spin-down, has also been detected [286]. http://www.jb.man.ac.uk/pulsar/glitches/gTable.html http://www.atnf.csiro.au/research/pulsar/psrcat/glitchTbl.html 62 Figure 20: Fractional moments of inertia as a function of the stellar mass for both the charged component I and the (crustal) super
uid component I , with four cases of NS EOSs (BCPM, p n BSk21, BSk20, Av18*). Av18* indicates the nonunifed EOS of "Av18 + NV + BPS". Taken from Li & Wang [331]. fractional moment of inertia, , which is I =I & 1:6% [293, 294]. It has been n p argued that \entrainment" of the neutron super
uid by the crystalline struc- ture of the crust greatly reduces its mobility, increasing the lower limit from 1:6% to 7% and making it very dicult for the nuclear EOSs to ful ll with a normal M > 1:0M NS [122, 294, 295]. This is clearly shown in Fig. 20. Consequently, the unpinning of the crustal super
uid is insucient to account for large glitches. This is the so-called glitch crisis problem. Other mechanisms, e.g., the unpinning of core super
uid neutrons, may be required. However, the mobility of super
uid neturons are related to the eective neutron mass, which has been discussed actively in the literature, see e.g., [296]. According to the calculation in Watanabe & Pethick [296], the constraint for the fractional mo- ment of inertia is I =I 2:5 2:3% Then, an NS of M . 1:55M NS would be acceptable, and there is no need to invoke the core super
uid. However, this is an open problem, and more detailed work has to be done. At present, we are still far from a thorough understanding of the general 63 picture of glitches; for example, is there a connection between the stellar interior and the magnetosphere of a star? How can various types of post-glitch behavior be explained? Is there an alternative model besides the super
uid model? The original starquake model [297] suggested that the change in the moment of inertia was due to relaxation in the NS (solid) crust to the current equilibrium oblateness but has diculty explaining the large glitches observed from the Vela pulsar. In a solid star model, the whole body of the star, rather than only the crust, is in a solid state. In such cases, the glitch amplitude could be explained [298, 299]. It is a challenge to quantitatively describe glitch behaviors since the related physical processes are complicated. There has been progress in the determination of NS properties in the literature [e.g., 300, 301, 302, 303]. 7. Summary and perspectives Although NSs were anticipated in the early thirties and discovered as pul- sars in the late sixties, the state of their liquid interiors remains unclear due to a lack of a good understanding of QCD at low energy scales. The current and upcoming multimessenger observatories [e.g., 304, 305, 306, 307, 308, 309] will continue improving the detection of pulsars together with the precise mea- surements of their masses and radii. Laboratory experiments will provide an emerging understanding of nuclear matter EOS and the transition to decon ned quark matter. Hopefully, the high density behavior of the NS EOS can be de- termined soon, shedding light on many unsolved problems in nuclear physics and high-energy astrophysics [e.g., 310, 311, 312, 313, 314, 315]. Since the compact star EOS is such a demanding problem, it is necessary to combine eorts from dierent communities and discuss mutual interests and problems [e.g., 134, 316, 317]. Additionally, it is important to establish new quantitative results testable by experiments/observations. In this work, the mi- croscopic physics of dense matter are modelled within the QMF, which connects the structure of a nucleon to the EOS of in nite nuclear matter, with a wide range of experimental and observational data available for use. The parameter space of the QMF has been constrained well for describing NSs, following the 64 present robust measurements of mass, radius, and tidal deformability. The pure NS maximum mass is approximately 2:1M , with a satisfying reproduction of the nuclear matter properties around the saturation density. The results have a modest dependence on the model parameters. Based on the available hyperonu- clei data, the hyperon puzzle is present, and we need to understand better how hyperon three-body interaction plays a role to understand more clearly whether hyperons are relevant in NSs (especially the heavy ones). The CSS parametriza- tion allows us to handle the high-density cores of NSs in a model-independent way. After demonstrating how the NSs' mass and radius depend on the CSS parameters for the phase transition of decon ned quark matter, we nd a safe upper limit for the hybrid star maximum mass at approximately 3:6M based on the extreme causality EOS, similar to previous studies. In particular, the NS/QS maximum mass is predicted to be approximately 2:3 2:4M from vari- ous model calculations, as well as analysis on the merger remnant of GW170817. Therefore, if more massive pulsars above 2:4M or fewer massive black holes be- low 5M are found, what is their nature? Is there a mass gap between NSs and black holes and why? Such problems are mysteries to be solved. The detailed feature of the sound speed in quark matter is explored in a perturbative model, and an enhancement in the sound speed is necessary to ful ll the two-solar-mass constraint of pulsars, located at intermediate densities, indicating that the pair of quarks starts to play a nontrivial role at such densities. Quark super
uids should not be ignored in the quark matter relevant to compact stars. This is also consistent with a lower bound for tidal deformability that is more consistent with the GW170817-like event in another MIT model than in the case without super
uid. In the future, even if we understand the stiness of the EOS, a further challenge is the particle degree of freedom in cold, dense matter. This could help us understand the nonperturbative properties of low-energy QCD (or the parameters of an eective model). This is work for the future. 65 Acknowledgements We are grateful to the members of the XMU neutron star group. We are thankful to the anonymous referee for his or her bene cial comments and Hong Shen, Jirina Stone, Anthony Thomas for insightful discussions. This work was supported by the National Natural Science Foundation of China (Grant Nos. 11873040, 11775276, 11705163, and 11775119), the Youth Innovation Promotion Association of the Chinese Academy of Sciences, the Natural Science Foundation of Tianjin, the China Scholarship Council (Grant No. 201906205013), and the Ningbo Natural Science Foundation (Grant No. 2019A610066). References [1] Abbott, B. P. et al., Oct. 2017. GW170817: Observation of Gravitational Waves from a Binary Neutron Star Inspiral. PhRvL, 119 (16), 161101. [2] Abbott, B. P. et al., Oct. 2017. Multi-messenger Observations of a Binary Neutron Star Merger. ApJ, 848 (2), L12. [3] Abbott, B. P. et al., Feb. 2017. Exploring the sensitivity of next generation gravitational wave detectors. CQGra, 34 (4), 044001. [4] Baiotti, L., Nov. 2019. Gravitational waves from neutron star mergers and their relation to the nuclear equation of state. PrPNP, 109, 103714. [5] Chirenti, C., Gold, R., Miller, M. C., Mar. 2017. Gravitational Waves from F-modes Excited by the Inspiral of Highly Eccentric Neutron Star Binaries. ApJ, 837 (1), 67. [6] Alford, M., Schmitt, A., Rajagopal, K., Sch afer, T., Oct. 2008. Color superconductivity in dense quark matter. RvMP, 80 (4), 1455{1515. [7] Baym, G., Hatsuda, T., Kojo, T., Powell, P. D., Song, Y., Takatsuka, T., May 2018. From hadrons to quarks in neutron stars: a review. RPPh, 81 (5), 056902. 66 [8] Paschalidis, V., Yagi, K., Alvarez-Castillo, D., Blaschke, D. B., Sedrakian, A., Apr. 2018. Implications from GW170817 and I-Love-Q relations for relativistic hybrid stars. PhRvD, 97 (8), 084038. [9] Han, S., Mamun, M. A. A., Lalit, S., Constantinou, C., Prakash, M., Nov. 2019. Treating quarks within neutron stars. PhRvD, 100 (10), 103022. [10] Shapiro, S. L., Teukolsky, S. A., 1983. Black holes, white dwarfs, and neutron stars : the physics of compact objects. bhwd.book [11] Lattimer, J. M., Prakash, M., Apr. 2004. The Physics of Neutron Stars. Sci, 304 (5670), 536{542. [12] Baldo, M., Burgio, G. F., Feb. 2012. Properties of the nuclear medium. RPPh, 75 (2), 026301. [13] Graber, V., Andersson, N., Hogg, M., Jan. 2017. Neutron stars in the laboratory. IJMPD, 26 (8), 1730015{347. [14] Guichon, P. A. M., Saito, K., Rodionov, E., Thomas, A. W., Feb. 1996. The role of nucleon structure in nite nuclei. NuPhA, 601 (3), 349{379. [15] Mohr, P. J., Newell, D. B., Taylor, B. N., Sep. 2016. CODATA rec- ommended values of the fundamental physical constants: 2014*. RvMP, 88 (3), 035009. [16] Danielewicz, P., Lacey, R., Lynch, W. G., Nov. 2002. Determination of the Equation of State of Dense Matter. Sci, 298 (5598), 1592{1596. [17] Wang, M., Audi, G., A. H., W., F. G., K., MacCormick, M., Xu, X., Pfeier, B., Dec. 2012. The Ame2012 atomic mass evaluation. ChPhC, 36 (12), 003. [18] Angeli, I., Marinova, K. P., Jan. 2013. Table of experimental nuclear ground state charge radii: An update. ADNDT, 99 (1), 69{95. 67 [19] Feliciello, A., Nagae, T., Sep. 2015. Experimental review of hypernu- clear physics: recent achievements and future perspectives. RPPh, 78 (9), [20] Hessels, J. W. T., Ransom, S. M., Stairs, I. H., Freire, P. C. C., Kaspi, V. M., Camilo, F., Mar. 2006. A Radio Pulsar Spinning at 716 Hz. Sci, 311 (5769), 1901{1904. [21] Demorest, P. B., Pennucci, T., Ransom, S. M., Roberts, M. S. E., Hes- sels, J. W. T., Oct. 2010. A two-solar-mass neutron star measured using Shapiro delay. Natur, 467 (7319), 1081{1083. [22] Antoniadis, J., et al., Apr. 2013. A Massive Pulsar in a Compact Rela- tivistic Binary. Sci, 340 (6131), 448. [23] Fonseca, E., et al., Dec. 2016. The NANOGrav Nine-year Data Set: Mass and Geometric Measurements of Binary Millisecond Pulsars. ApJ, 832 (2), [24] Arzoumanian, Z. et al., Apr. 2018. The NANOGrav 11-year Data Set: High-precision Timing of 45 Millisecond Pulsars. ApJS, 235 (2), 37. [25] Cromartie, H. T., et al., Jan. 2020. Relativistic Shapiro delay measure- ments of an extremely massive millisecond pulsar. NatAs, 4, 72{76. [26] Ozel, F., Freire, P., Sep. 2016. Masses, Radii, and the Equation of State of Neutron Stars. ARA&A, 54, 401{440. [27] Miller, M. C., et al., Dec. 2019. PSR J0030+0451 Mass and Radius from NICER Data and Implications for the Properties of Neutron Star Matter. ApJ, 887 (1), L24. [28] Riley, T. E., et al., Dec. 2019. A NICER View of PSR J0030+0451: Mil- lisecond Pulsar Parameter Estimation. ApJ, 887 (1), L21. [29] Abbott, B. P. et al., Jan. 2019. Properties of the Binary Neutron Star Merger GW170817. PhRvX, 9 (1), 011001. 68 [30] Abbott, B. P. et al., Feb. 2020. Model a comparison from LIGO-Virgo data on GW170817's binary components and consequences for the merger remnant. CQGra, 37 (4), 045006. [31] Prakash, M., Lattimer, J. M., Pons, J. A., Steiner, A. W., Reddy, S., 2001. Evolution of a Neutron Star from Its Birth to Old Age. LNP, 578, [32] Burgio, G. F., Fantina, A. F., 2018. Nuclear Equation of State for Compact Stars and Supernovae. ASSL, 457, 255. [33] Weber, F., Mar. 2005. Strange quark matter and compact stars. PrPNP, 54 (1), 193{288. [34] Li, A., Burgio, G. F., Lombardo, U., Peng, G. X., Jan. 2008. Exotic Phases in Neutron Stars. IJMPE, 17 (9), 1635{1647. [35] Gal, A., Hungerford, E. V., Millener, D. J., Aug. 2016. Strangeness in nuclear physics. RvMP, 88 (3), 035004. [36] Tolos, L., Fabbietti, L., Feb. 2020. Strangeness in Nuclei and Neutron Stars. preprint (arXiv:2002.09223). [37] Li, A., Huang, F., Xu, R.-X., Sep. 2012. Too massive neutron stars: The role of dark matter? APh, 37, 70{74. [38] Das, H. C., Kumar, A., Kumar, B., Biswal, S. K., Nakatsukasa, T., Li, A., Patra, S. K., Feb. 2020. Eects of Dark Matter on Nuclear and Neutron Star Matter. preprint (arXiv:2002.00594). [39] Nik si c, T., Vretenar, D., Ring, P., Jul. 2011. Relativistic nuclear energy density functionals: Mean- eld and beyond. PrPNP, 66 (3), 519{548. [40] Shen, H., Toki, H., Apr. 2000. Quark mean eld model for nuclear matter and nite nuclei. PhRvC, 61 (4), 045205. 69 [41] Toki, H., Meyer, U., Faessler, A., Brockmann, R., Dec. 1998. Quark mean eld model for nucleons in nuclei. PhRvC, 58 (6), 3749{3752. [42] Zhu, Z.-Y., Li, A., Hu, J.-N., Shen, H., Feb. 2019. Quark mean- eld model for nuclear matter with or without bag. PhRvC, 99 (2), 025804. [43] Zhu, Z.-Y., Li, A., Mar. 2018. Eects of the nucleon radius on neutron stars in a quark mean eld model. PhRvC, 97 (3), 035805. [44] Zhu, Z.-Y., Zhou, E.-P., Li, A., Aug. 2018. Neutron Star Equation of State from the Quark Level in Light of GW170817. ApJ, 862 (2), 98. [45] Shen, H., Toki, H., Sep. 2002. Study of hypernuclei in the quark mean- eld model. NuPhA, 707 (3), 469{476. [46] Hu, J. N., Li, A., Shen, H., Toki, H., Jan. 2014. Quark mean- eld model for single and double and hypernuclei. PTEP, 2014 (1), 013D02. [47] Hu, J. N., Li, A., Toki, H., Zuo, W., Feb. 2014. Extended quark mean- eld model for neutron stars. PhRvC, 89 (2), 025802. [48] Hu, J., Shen, H., Nov. 2017. Single- hypernuclei within a quark mean- eld model. PhRvC, 96 (5), 054304. [49] Xing, X., Hu, J., Shen, H., Oct. 2016. Quark mean eld model with pion and gluon corrections. PhRvC, 94 (4), 044308. [50] Xing, X., Hu, J., Shen, H., May 2017. Quark mean eld model with pion and gluon corrections for and hypernuclei and neutron stars. PhRvC, 95 (5), 054310. [51] Baldo, M., Nov. 1999. Nuclear Methods And The Nuclear Equation Of State. in: Baldo M. (Ed.), Series: International Review of Nuclear Physics. World Scienti c Publishing Co. Pte. Ltd., vol. 8. [52] Sharma, B. K., Centelles, M., Vinas, ~ X., Baldo, M., Burgio, G. F., Dec. 2015. Uni ed equation of state for neutron stars on a microscopic basis. A&A, 584, A103. 70 [53] Hebeler, K., Lattimer, J. M., Pethick, C. J., Schwenk, A., Aug. 2013. Equation of State and Neutron Star Properties Constrained by Nuclear Physics and Observation. ApJ, 773 (1), 11. [54] Tews, I., Margueron, J., Reddy, S., Jun. 2019. Confronting gravitational- wave observations with modern nuclear physics constraints. EPJA, 55 (6), [55] Lonardoni, D., Lovato, A., Gandol , S., Pederiva, F., Mar. 2015. Hyperon Puzzle: Hints from Quantum Monte Carlo Calculations. PhRvL, 114 (9), [56] Gandol , S., Lonardoni, D., Lovato, A., Piarulli, M., Jan. 2020. Atomic nuclei from quantum Monte Carlo calculations with chiral EFT interac- tions. preprint (arXiv:2001.01374). [57] Akmal, A., Pandharipande, V. R., Ravenhall, D. G., Sep. 1998. Equation of state of nucleon matter and neutron star structure. PhRvC, 58 (3), 1804{1828. [58] Guichon, P. A. M., Jan. 1988. A possible quark mechanism for the satu- ration of nuclear matter. PhLB, 200 (3), 235{240. [59] Degrand, T., Jae, R. L., Johnson, K., Kiskis, J., Oct. 1975. Masses and other parameters of the light hadrons. PhRvD, 12 (7), 2060{2076. [60] Guichon, P. A. M., Stone, J. R., Thomas, A. W., May 2018. Quark- Meson-Coupling (QMC) model for nite nuclei, nuclear matter and be- yond. PrPNP, 100, 262{297. [61] Saito, K., Tsushima, K., Thomas, A. W., Jan. 2007. Nucleon and hadron structure changes in the nuclear medium and the impact on observables. PrPNP, 58 (1), 1{167. [62] Bohr, H., Moszkowski, S. A., Panda, P. K., Provid^ encia, C., da Provid^ encia, J., Feb. 2016. QMC approach based on the Bogoliubov in- 71 dependent quark model of the nucleon. International Journal of Modern Physics E 25 (2), 1650007. [63] Stone, J. R., Guichon, P. A. M., Reinhard, P. G., Thomas, A. W., Mar. 2016. Finite Nuclei in the Quark-Meson Coupling Model. PhRvL, 116 (9), [64] Motta, T. F., Kalaitzis, A. M., Anti c, S., Guichon, P. A. M., Stone, J. R., Thomas, A. W., Jun. 2019. Isovector Eects in Neutron Stars, Radii, and the GW170817 Constraint. ApJ, 878 (2), 159. [65] Motta, T. F., Thomas, A. W., Guichon, P. A. M., Mar. 2020. Do Delta baryons play a role in neutron stars?. PhLB, 802, 135266. [66] Isgur, N., Karl, G., Dec. 1978. P-wave baryons in the quark model. PhRvD, 18 (11), 4187{4205. [67] Horowitz, C. J., Piekarewicz, J., Jun. 2001. Neutron Star Structure and the Neutron Radius of Pb. PhRvL, 86 (25), 5647{5650. [68] Tagami, S., Yasutake, N., Fukuda, M., Yahiro, M., Mar. 2020. Determina- tion of symmetry energy from experimental and observational constraints; prediction on CREX. preprint (arXiv:2003.06168). [69] Danielewicz, P., Lee, J., Feb. 2014. Symmetry energy II: Isobaric analog states. NuPhA, 922, 1{70. [70] Li, B.-A., Chen, L.-W., Ko, C. M., Aug. 2008. Recent progress and new challenges in isospin physics with heavy-ion reactions. PhR, 464 (4-6), 113{281. [71] Oertel, M., Hempel, M., Kl ahn, T., Typel, S., Jan. 2017. Equations of state for supernovae and compact stars. RvMP, 89 (1), 015007. [72] Li, B.-A., Ramos, A., Verde, G., Vidana, ~ I., Feb. 2014. Topical issue on nuclear symmetry energy. EPJA, 50, 72 [73] Baldo, M., Burgio, G. F., Nov. 2016. The nuclear symmetry energy. Progress in Particle and Nuclear Physics 91, 203{258. [74] Zhang, Y., Liu, M., Xia, C.-J., Li, Z., Biswal, S. K., Feb. 2020. Constraints on the symmetry energy and its associated parameters from nuclei to neutron stars. preprint(arXiv:2002.10884). [75] Dutra, M., Louren co, O., S a Martins, J. S., Del no, A., Stone, J. R., Stevenson, P. D., Mar. 2012. Skyrme interaction and nuclear matter con- straints. PhRvC, 85 (3), 035201. [76] Dutra, M., Louren co, O., Avancini, S. S., Carlson, B. V., Del no, A., Menezes, D. P., Provid^ encia, C., Typel, S., Stone, J. R., Nov. 2014. Relativistic mean- eld hadronic models under nuclear matter constraints. PhRvC, 90 (5), 055203. [77] Malik, T., Alam, N., Fortin, M., Provid^ encia, C., Agrawal, B. K., Jha, T. K., Kumar, B., Patra, S. K., Sep. 2018. GW170817: Constraining the nuclear matter equation of state from the neutron star tidal deformability. PhRvC, 98 (3), 035804. [78] Zhang, N.-B., Li, B.-A., Jul. 2019. Implications of the Mass M=f2.17g f- 0.10gf+0.11g M of PSR J0740+6620 on the Equation of State of Super- dense Neutron-rich Nuclear Matter. ApJ, 879 (2), 99. [79] Zhang, N.-B., Li, B.-A., Mar. 2019. Extracting nuclear symmetry ener- gies at high densities from observations of neutron stars and gravitational waves. European Physical Journal A 55 (3), 39. [80] Zimmerman, J., Carson, Z., Schumacher, K., Steiner, A. W., Yagi, K., Feb. 2020. Measuring Nuclear Matter Parameters with NICER and LIGO/Virgo. preprint (arXiv:2002.03210). [81] Shlomo, S., Kolomietz, V. M., Col o, G., Oct. 2006. Deducing the nuclear- matter incompressibility coecient from data on isoscalar compression modes. EPJA, 30 (1), 23{30. 73 [82] Arrington, J., Sick, I., Sep. 2015. Evaluation of the Proton Charge Ra- dius from Electron-Proton Scattering. Journal of Physical and Chemical Reference Data 44 (3), 031204. [83] Pohl, R., et al., Jul. 2010. The size of the proton. Natur, 466 (7303), 213{216. [84] Potekhin, A. Y., Fantina, A. F., Chamel, N., Pearson, J. M., Goriely, S., Dec. 2013. Analytical representations of uni ed equations of state for neutron-star matter. A&A, 560, A48. [85] Shen, H., Toki, H., Oyamatsu, K., Sumiyoshi, K., Jul. 1998. Relativis- tic equation of state of nuclear matter for supernova and neutron star. NuPhA, 637 (3), 435{450. [86] Glendenning, N. K., Moszkowski, S. A., Oct. 1991. Reconciliation of neutron-star masses and binding of the Lambda in hypernuclei. PhRvL, 67, 2414{1417. [87] Li, A., Liu, T., Jul. 2013. Revisiting the hot matter in the center of gamma- ray bursts and supernovae. A&A, 555, A129. [88] Li, A., Liu, T., Gubler, P., Xu, R.-X., Mar. 2015. Revisiting the boiling of primordial quark nuggets at nonzero chemical potential. APh, 62, 115{ [89] Li, A., Zhou, X. R., Burgio, G. F., Schulze, H. J., Feb. 2010. Protoneutron stars in the Brueckner-Hartree-Fock approach and nite-temperature kaon condensation. PhRvC, 81 (2), 025806. [90] Zhu, Z.-Y., Li, A., Hu, J.-N., Sagawa, H., Oct. 2016. (1232) eects in density-dependent relativistic Hartree-Fock theory and neutron stars. PhRvC, 94 (4), 045803. [91] Stone, J., Guichon, P. A. M., Matevosyan, H. H., Thomas, A. W., Aug. 2007. Cold uniform matter and neutron stars in the quark meson-coupling model. NuPhA, 792 (3-4), 341{369. 74 [92] Lattimer, J. M., Prakash, M., Apr. 2007. Neutron star observations: Prog- nosis for equation of state constraints. PhR, 442 (1-6), 109{165. [93] Lattimer, J. M., Nov. 2012. The Nuclear Equation of State and Neutron Star Masses. ARNPS, 62 (1), 485{515. [94] Bauswein, A., Just, O., Janka, H.-T., Stergioulas, N., Dec. 2017. Neutron- star Radius Constraints from GW170817 and Future Detections. ApJ, 850 (2), L34. [95] Annala, E., Gorda, T., Kurkela, A., Vuorinen, A., Apr. 2018. Gravitational-Wave Constraints on the Neutron-Star-Matter Equation of State. PhRvL, 120 (17), 172703. [96] Lim, Y., Holt, J. W., Aug. 2018. Neutron Star Tidal Deformabilities Con- strained by Nuclear Theory and Experiment. PhRvL, 121 (6), 062701. [97] Most, E. R., Weih, L. R., Rezzolla, L., Schaner-Bielich, J., Jun. 2018. New Constraints on Radii and Tidal Deformabilities of Neutron Stars from GW170817. PhRvL, 120 (26), 261103. [98] Nandi, R., Char, P., Apr. 2018. Hybrid Stars in the Light of GW170817. ApJ, 857 (1), 12. [99] Zhao, T., Lattimer, J. M., Sep. 2018. Tidal deformabilities and neutron star mergers. PhRvD, 98 (6), 063020. [100] Zhou, E.-P., Zhou, X., Li, A., Apr. 2018. Constraints on interquark in- teraction parameters with GW170817 in a binary strange star scenario. PhRvD, 97 (8), 083015. [101] Han, S., Steiner, A. W., Apr. 2019. Tidal deformability with sharp phase transitions in binary neutron stars. PhRvD, 99 (8), 083014. 2 [102] Raithel, C. A., Ozel, F., Psaltis, D., Apr. 2018. Tidal Deformability from GW170817 as a Direct Probe of the Neutron Star Radius. ApJ, 857 (2), L23. 75 [103] Tsang, C. Y., Tsang, M. B., Danielewicz, P., Fattoyev, F. J., Lynch, W. G., Sep. 2019. Insights on Skyrme parameters from GW170817. PhLB, 796, 1{5. [104] Wang, Q., Shi, C., Yan, Y., Zong, H.-S., Dec. 2019. Crossover hadron- quark transition with a modi ed interpolation method and constraints from tidal deformability of GW170817. preprint (arXiv:1912.02312). [105] Zhou, Y., Chen, L.-W., Nov. 2019. Ruling Out the Supersoft High-density Symmetry Energy from the Discovery of PSR J0740+6620 with Mass +0:10 2:14 M . ApJ, 886 (1), 52. 0:09 [106] Essick, R., Landry, P., Holz, D. E., Mar. 2020. Nonparametric inference of neutron star composition, equation of state, and maximum mass with GW170817. PhRvD, 101 (6), 063007. [107] Ferreira, M., Fortin, M., Malik, T., Agrawal, B. K., Provid^ encia, C., Feb. 2020. Empirical constraints on the high-density equation of state from multimessenger observables. PhRvD, 101 (4), 043021. [108] Guv en, H., Bozkurt, K., Khan, E., Margueron, J., Jan. 2020. Multi- messenger and multi-physics bayesian inference for GW170817 binary neu- tron star merger. preprint (arXiv:2001.10259). [109] Louren co, O., Bhuyan, M., Lenzi, C. H., Dutra, M., Gonzalez-Boquera, C., Centelles, M., Vinas, ~ X., Apr. 2020. GW170817 constraints analyzed with Gogny forces and momentum-dependent interactions. PhLB, 803, [110] Louren co, O., Dutra, M., Lenzi, C. H., Biswal, S. K., Bhuyan, M., Menezes, D. P., Feb. 2020. Consistent Skyrme parametrizations con- strained by GW170817. EPJA, 56 (2), 32. [111] Traversi, S., Char, P., Pagliara, G., Feb. 2020. Bayesian Inference of Dense Matter Equation of State within Relativistic Mean Field Models using Astrophysical Measurements. preprint (arXiv:2002.08951). 76 [112] Li, A., Zuo, W., Mi, A.-J., G, B., Jul. 2007. Hyperon hyperon interaction in relativistic mean eld model. ChPhy, 16 (7), 1934{1940. [113] Burgio, G. F., Schulze, H. J., Li, A., Feb. 2011. Hyperon stars at nite temperature in the Brueckner theory. PhRvC, 83 (2), 025804. [114] Zuo, W., Li, A., Li, Z. H., Lombardo, U., Nov. 2004. Nuclear three-body force eect on a kaon condensate in neutron star matter. PhRvC, 70 (5), [115] Li, A., Burgio, G. F., Lombardo, U., Zuo, W., Nov. 2006. Microscopic three-body forces and kaon condensation in cold neutrino-trapped matter. PhRvC, 74 (5), 055801. [116] Li, A., Zuo, W., Schulze J., H., Lombardo, U., Dec. 2008. Hot Nuclear Matter Equation of State and Finite Temperature Kaon Condensation. ChPhL, 25 (12), 4233{4236. [117] Li, A., Peng, G.-X., Lombardo, U., Mar. 2009. Decon nement phase tran- sition in neutron star matter. ChPhC, 33 (S1), 61{63. [118] Li, A., Zuo, W., Peng, G. X., Mar. 2015. Massive hybrid stars with a rst-order phase transition. PhRvC, 91 (3), 035803. [119] Chamel, N., Haensel, P., Dec. 2008. Physics of Neutron Star Crusts. LRR, 11 (1), 10. [120] Piekarewicz, J., Fattoyev, F. J., Horowitz, C. J., Jul. 2014. Pulsar glitches: The crust may be enough. PhRvC, 90 (1), 015803. [121] Li, A., Jul. 2015. Glitch Crisis or Not: a Microscopic Study. Chinese Physics Letters 32 (7), 079701. [122] Li, A., Dong, J. M., Wang, J. B., Xu, R. X., Mar. 2016. Structures of the Vela Pulsar and the Glitch Crisis from the Brueckner Theory. ApJS, 223 (1), 16. 77 [123] Fantina, A. F., Chamel, N., Pearson, J. M., Goriely, S., Nov. 2013. Neu- tron star properties with uni ed equations of state of dense matter. A&A, 559, A128. [124] Antic, S., Stone, J. R., Thomas, A. W., Dec. 2019. Neutron stars from crust to core within the Quark-meson coupling model. preprint (arXiv:1912.03815) [125] Fortin, M., Provid^ encia, C., Raduta, A. R., Gulminelli, F., Zdunik, J. L., Haensel, P., Bejger, M., Sep. 2016. Neutron star radii and crusts: Uncer- tainties and uni ed equations of state. PhRvC, 94 (3), 035804. [126] Baym, G., Pethick, C., Sutherland, P., Dec. 1971. The Ground State of Matter at High Densities: Equation of State and Stellar Models. ApJ, 170, 299. [127] Negele, J. W., Vautherin, D., Jun. 1973. Neutron star matter at sub- nuclear densities. NuPhA, 207 (2), 298{320. [128] Stergioulas, N., Jun. 2003. Rotating Stars in Relativity. LRR, 6 (1), 3. [129] Fortin, M., Taranto, G., Burgio, G. F., Haensel, P., Schulze, H. J., Zdunik, J. L., Apr. 2018. Thermal states of neutron stars with a consistent model of interior. MNRAS, 475 (4), 5010{5022. [130] Haensel, P., Zdunik, J. L., Mar. 2008. Models of crustal heating in accret- ing neutron stars. A&A, 480 (2), 459{464. [131] Tolman, R. C., Feb. 1939. Static Solutions of Einstein's Field Equations for Spheres of Fluid. PhRv, 55 (4), 364{373. [132] Oppenheimer, J. R., Volko, G. M., Feb. 1939. On Massive Neutron Cores. PhRv, 55 (4), 374{381. [133] Hornick, N., Tolos, L., Zacchi, A., Christian, J.-E., Schaner-Bielich, J., Dec. 2018. Relativistic parameterizations of neutron matter and implica- tions for neutron stars. PhRvC, 98 (6), 065804. 78 [134] Capano, C. D., Tews, I., Brown, S. M., Margalit, B., De, S., Kumar, S., Brown, D. A., Krishnan, B., Reddy, S., Mar. 2020. Stringent constraints on neutron-star radii from multimessenger observations and nuclear the- ory. NatAs, 1{8. [135] Postnikov, S., Prakash, M., Lattimer, J. M., Jul. 2010. Tidal Love numbers of neutron and self-bound quark stars. PhRvD, 82 (2), 024016. [136] Read, J. S., et al., Aug. 2013. Matter eects on binary neutron star wave- forms. PhRvD, 88 (4), 044042. [137] Komatsu, H., Eriguchi, Y., Hachisu, I., Mar. 1989. Rapidly rotating gen- eral relativistic stars. I - Numerical method and its application to uni- formly rotating polytropes. MNRAS, 237, 355{379. [138] Cook, G. B., Shapiro, S. L., Teukolsky, S. A., Feb. 1994. Rapidly Rotating Polytropes in General Relativity. ApJ, 422, 227. [139] Stergioulas, N., Friedman, J. L., May 1995. Comparing Models of Rapidly Rotating Relativistic Stars Constructed by Two Numerical Methods. ApJ, 444, 306. [140] Haensel, P., Zdunik, J. L., Bejger, M., Lattimer, J. M., Aug. 2009. Ke- plerian frequency of uniformly rotating neutron stars and strange stars. A&A, 502 (2), 605{610. [141] Wei, J. B., Chen, H., Burgio, G. F., Schulze, H. J., Aug. 2017. Rotat- ing hybrid stars with the Dyson-Schwinger quark model. PhRvD, 96 (4), [142] Andersson, N., Kokkotas, K. D., Jan. 2001. The R-Mode Instability in Rotating Neutron Stars. IJMPD, , 10 (4), 381{441. [143] Breu, C., Rezzolla, L., Jun. 2016. Maximum mass, moment of inertia and compactness of relativistic stars. MNRAS, 459 (1), 646{656. 79 [144] Urbanec, M., Miller, J. C., Stuchl k, Z., Aug. 2013. Quadrupole moments of rotating neutron stars and strange stars. MNRAS, 433 (3), 1903{1909. [145] Yagi, K., Yunes, N., Jul. 2013. I-Love-Q: Unexpected Universal Relations for Neutron Stars and Quark Stars. Sci, 341 (6144), 365{368. [146] Yagi, K., Yunes, N., Jul. 2013. I-Love-Q relations in neutron stars and their applications to astrophysics, gravitational waves, and fundamental physics. PhRvD, 88 (2), 023009. [147] Haskell, B., Ciol , R., Pannarale, F., Rezzolla, L., Feb. 2014. On the universality of I-Love-Q relations in magnetized neutron stars. MNRAS, 438 (1), L71{L75. [148] Pappas, G., Apostolatos, T. A., Mar. 2014. Eectively Universal Behav- ior of Rotating Neutron Stars in General Relativity Makes Them Even Simpler than Their Newtonian Counterparts. PhRvL, 112 (12), 121101. [149] Zhou, E., Tsokaros, A., Uryu, K., Xu, R., Shibata, M., Aug. 2019. Dier- entially rotating strange star in general relativity. PhRvD, 100 (4), 043015. [150] Zhou, E., Tsokaros, A., Rezzolla, L., Xu, R., Uryu, K., Jan. 2018. Uni- formly rotating, axisymmetric, and triaxial quark stars in general relativ- ity. PhRvD, 97 (2), 023013. [151] Khaustov, P. et al., May 2000. Evidence of hypernuclear production in 12 + 12 the C(K ,K ) Be reaction. PhRvC, 61 (5), 054603. [152] Yakovlev, D. G., Kaminker, A. D., Gnedin, O. Y., Haensel, P., Nov. 2001. Neutrino emission from neutron stars. PhR, 354 (1-2), 1{155. [153] Rijken, T. A., Schulze, H. J., Feb. 2016. Hyperon-hyperon interactions with the Nijmegen ESC08 model. EPJA, 52, 21. [154] Bombaci, I., Jan. 2017. The Hyperon Puzzle in Neutron Stars. hspp.conf, 80 [155] Yamamoto, Y., Furumoto, T., Yasutake, N., Rijken, T. A., Oct. 2014. Hy- peron mixing and universal many-body repulsion in neutron stars. PhRvC, 90 (4), 045805. [156] Haidenbauer, J., Meiner, U. G., Kaiser, N., Weise, W., Jun. 2017. Lambda-nuclear interactions and hyperon puzzle in neutron stars. EPJA, 53 (6), 121. [157] Logoteta, D., Vidana, ~ I., Bombaci, I., Nov. 2019. Impact of chiral hyper- onic three-body forces on neutron stars. EPJA, 55 (11), 207. [158] Peng, G. X., Li, A., Lombardo, U., Jun. 2008. Decon nement phase tran- sition in hybrid neutron stars from the Brueckner theory with three-body forces and a quark model with chiral mass scaling. PhRvC, 77 (6), 065807. [159] Li, A., Xu, R.-X., Lu, J.-F., Mar. 2010. Strange stars with dierent quark mass scalings. MNRAS, 402 (4), 2715{2719. [160] Li, A., Peng, G.-X., Lu, J.-F., Apr. 2011. Strange star candidates revised within a quark model with chiral mass scaling. RAA, 11 (4), 482{490. [161] Haensel, P., Zdunik, J. L., Schaefer, R., May 1986. Strange quark stars. A&A, 160 (1), 121{128. [162] Wiktorowicz, G., Drago, A., Pagliara, G., Popov, S. B., Sep. 2017. Strange Quark Stars in Binaries: Formation Rates, Mergers, and Explosive Phe- nomena. ApJ, 846 (2), 163. [163] Drago, A., Pagliara, G., Jan. 2018. Merger of Two Neutron Stars: Pre- dictions from the Two-families Scenario. ApJ, 852 (2), L32. [164] Alford, M., Han, S., Schwenzer, K., Nov. 2019. Signatures for quark matter from multi-messenger observations. JPhG, 46 (11), 114001. [165] Aloy, M. A., Ib anez, ~ J. M., Sanchis-Gual, N., Obergaulinger, M., Font, J. A., Serna, S., Marquina, A., Apr. 2019. Neutron star collapse and 81 gravitational waves with a non-convex equation of state. MNRAS, 484 (4), 4980{5008. [166] Bauswein, A., Bastian, N.-U. F., Blaschke, D. B., Chatziioannou, K., Clark, J. A., Fischer, T., Oertel, M., Feb. 2019. Identifying a First-Order Phase Transition in Neutron-Star Mergers through Gravitational Waves. PhRvL, 122 (6), 061102. [167] Most, E. R., Papenfort, L. J., Dexheimer, V., Hanauske, M., Schramm, S., St ocker, H., Rezzolla, L., Feb. 2019. Signatures of Quark-Hadron Phase Transitions in General-Relativistic Neutron-Star Mergers. PhRvL, 122 (6), [168] Weih, L. R., Most, E. R., Rezzolla, L., Aug. 2019. Optimal Neutron-star Mass Ranges to Constrain the Equation of State of Nuclear Matter with Electromagnetic and Gravitational-wave Observations. ApJ, 881 (1), 73. [169] Xia, C., Zhu, Z., Zhou, X., Li, A., Jun. 2019. Sound velocity in dense stellar matter with strangeness and compact stars. preprint (arXiv:1906.00826). [170] Gomes, R. O., Char, P., Schramm, S., Jun. 2019. Constraining Strangeness in Dense Matter with GW170817. ApJ, 877 (2), 139. [171] Chatziioannou, K., Han, S., Feb. 2020. Studying strong phase transitions in neutron stars with gravitational waves. PhRvD, 101 (4), 044019. [172] Fischer, T., Wu, M.-R., Wehmeyer, B., Bastian, N.-U. F., Mart nez- Pinedo, G., Thielemann, F.-K., Mar. 2020. Core-collapse supernova ex- plosions driven by the hadron-quark phase transition as rare r process site. preprint (arXiv:2003.00972). [173] Tonetto, L., Lugones, G., Mar. 2020. Discontinuity gravity modes in hy- brid stars: assessing the role of rapid and slow phase conversions. preprint (arXiv:2003.01259). 82 [174] Nunna, K. P., Banik, S., Chatterjee, D., Feb. 2020. Signatures of strangeness in neutron star merger remnants. preprint (arXiv:2002.07538). [175] Alford, M., Han, S., Prakash, M., Oct. 2013. Generic conditions for stable hybrid stars. PhRvD, 88 (8), 083013. [176] Zdunik, J. L., Haensel, P., Mar. 2013. Maximum mass of neutron stars and strange neutron-star cores. A&A, 551, A61. [177] Alford, M., Burgio, G. F., Han, S., Taranto, G., Zappal a, D., Oct. 2015. Constraining and applying a generic high-density equation of state. PhRvD, 92 (8), 083002. [178] Ranea-Sandoval, I. F., Han, S., Orsaria, M. G., Contrera, G. A., We- ber, F., Alford, M. G., Apr. 2016. Constant-sound-speed parametrization for Nambu-Jona-Lasinio models of quark matter in hybrid stars. PhRvC, 93 (4), 045812. [179] Kurkela, A., Romatschke, P., Vuorinen, A., May2010. Cold quark matter. PhRvD, 81 (10), 105021. [180] Abbott, B. P. et al., Oct. 2018. GW170817: Measurements of Neutron Star Radii and Equation of State. PhRvL, 121 (16), 161101. [181] Montana, ~ G., Tol os, L., Hanauske, M., Rezzolla, L., May 2019. Constrain- ing twin stars with GW170817. PhRvD, 99 (10), 103009. [182] Burgio, G. F., Drago, A., Pagliara, G., Schulze, H. J., Wei, J. B., Jun. 2018. Are Small Radii of Compact Stars Ruled out by GW170817/AT2017gfo? ApJ, 860 (2), 139. [183] De, S., Finstad, D., Lattimer, J. M., Brown, D. A., Berger, E., Biwer, C. M., Aug. 2018. Tidal Deformabilities and Radii of Neutron Stars from the Observation of GW170817. PhRvL, 121 (9), 091102. [184] Miao, Z. Q., Li, A., Zhu Z. Y. and Han, S., 2020. Constraining hadron- quark phase transition parameters within the quark-mean- eld model 83 using multi-messenger observations of neutron stars. preprint (arXiv: 2006.00839) [185] Kalogera, V., Baym, G., Oct. 1996. The Maximum Mass of a Neutron Star. ApJ, 470, L61. [186] Rhoades, C. E., Runi, R., Feb. 1974. Maximum Mass of a Neutron Star. PhRvL, 32 (6), 324{327. [187] Brecher, K., Caporaso, G., Feb. 1976. Obese `neutron' stars. Natur, 259 (5542), 377{378. [188] Glendenning, N. K., Jan. 1990. Fast Pulsars, Strange Stars:. AN Oppor- tunity in Radio Astronomy. MPLA, 5 (27), 2197{2207. [189] Burgio, G. F., Schulze, H. J., Weber, F., Sep. 2003. On the maximum rotational frequency of neutron and hybrid stars. A&A, 408, 675{680. [190] Glendenning, N. K., Aug. 1992. First-order phase transitions with more than one conserved charge: Consequences for neutron stars. PhRvD, 46 (4), 1274{1287. [191] Maruyama, T., Chiba, S., Schulze, H.-J., Tatsumi, T., Dec. 2007. Hadron- quark mixed phase in hyperon stars. PhRvD, 76 (12), 123015. [192] Xia, C.-J., Maruyama, T., Yasutake, N., Tatsumi, T., May 2019. Constraining quark-hadron interface tension in the multimessenger era. PhRvD, 99 (10), 103017. [193] Kurkela, A., Fraga, E. S., Schaner-Bielich, J., Vuorinen, A., Jul. 2014. Constraining Neutron Star Matter with Quantum Chromodynamics. ApJ, 789 (2), 127. [194] Fraga, E. S., Romatschke, P., May 2005. Role of quark mass in cold and dense perturbative QCD. PhRvD, 71 (10), 105014. 84 [195] Olive, K. A., Particle Data Group, Aug. 2014. Review of Particle Physics. ChPhC, 38 (9), 090001. [196] Vermaseren, J. A. M., Larin, S. A., van Ritbergen, T., Feb. 1997. The 4-loop quark mass anomalous dimension and the invariant quark mass. PhLB, 405, 327{333. [197] Burgio, G. F., Baldo, M., Sahu, P. K., Santra, A. B., Schulze, H. J., Jan. 2002. Maximum mass of neutron stars with a quark core. PhLB, 526 (1-2), 19{26. [198] Maieron, C., Baldo, M., Burgio, G. F., Schulze, H. J., Aug. 2004. Hybrid stars with the color dielectric and the MIT bag models. PhRvD, 70 (4), [199] Buballa, M., Feb. 2005. NJL-model analysis of dense quark matter [review article]. PhR, 407 (4-6), 205{376. [200] Bedaque, P., Steiner, A. W., Jan. 2015. Sound Velocity Bound and Neu- tron Stars. PhRvL, 114 (3), 031103. [201] Baym, G., Furusawa, S., Hatsuda, T., Kojo, T., Togashi, H., Nov. 2019. New Neutron Star Equation of State with Quark-Hadron Crossover. ApJ, 885 (1), 42. [202] Alsing, J., Silva, H. O., Berti, E., Jul. 2018. Evidence for a maximum mass cut-o in the neutron star mass distribution and constraints on the equation of state. MNRAS, 478 (1), 1377{1391. [203] Tews, I., Carlson, J., Gandol , S., Reddy, S., Jun. 2018. Constraining the Speed of Sound inside Neutron Stars with Chiral Eective Field Theory Interactions and Observations. ApJ, 860 (2), 149. [204] McLerran, L., Reddy, S., Mar. 2019. Quarkyonic Matter and Neutron Stars. PhRvL, 122 (12), 122701. 85 [205] Li, A., Zhu, Z.-Y., Zhou, X., Jul. 2017. New Equations of State for Post- merger Supramassive Quark Stars. ApJ, 844 (1), 41. [206] Li, A., Zhang, B., Zhang, N.-B., Gao, H., Qi, B., Liu, T., Oct. 2016. Internal x-ray plateau in short GRBs: Signature of supramassive fast- rotating quark stars? PhRvD, 94 (8), 083010. [207] Bodmer, A. R., Sep. 1971. Collapsed Nuclei. PhRvD, 4 (6), 1601{1606. [208] Witten, E., Jul. 1984. Cosmic separation of phases. PhRvD, 30 (2), 272{ [209] Alford, M., Braby, M., Paris, M., Reddy, S., Aug. 2005. Hybrid Stars that Masquerade as Neutron Stars. ApJ, 629 (2), 969{978. [210] Wei, W., Irving, B., Salinas, M., Kl ahn, T., Jaikumar, P., Dec. 2019. Camou
age of the Phase Transition to Quark Matter in Neutron Stars. ApJ, 887 (2), 151. [211] Rezzolla, L., Most, E. R., Weih, L. R., Jan. 2018. Using Gravitational- wave Observations and Quasi-universal Relations to Constrain the Maxi- mum Mass of Neutron Stars. ApJ, 852 (2), L25. [212] Ruiz, M., Shapiro, S. L., Tsokaros, A., Jan. 2018. GW170817, general rel- ativistic magnetohydrodynamic simulations, and the neutron star maxi- mum mass. PhRvD, 97 (2), 021501. [213] Shibata, M., Zhou, E., Kiuchi, K., Fujibayashi, S., Jul. 2019. Constraint on the maximum mass of neutron stars using GW170817 event. PhRvD, 100 (2), 023015. [214] Lai, X.-Y., Yu, Y.-W., Zhou, E.-P., Li, Y.-Y., Xu, R.-X., Feb. 2018. Merg- ing strangeon stars. Research in Astronomy and Astrophysics 18 (2), 024. [215] Bauswein, A., Janka, H. T., Oechslin, R., Pagliara, G., Sagert, I., Schaner-Bielich, J., Hohle, M. M., Neuh auser, R., Jul. 2009. Mass Ejec- 86 tion by Strange Star Mergers and Observational Implications. PhRvL, 103 (1), 011101. [216] Alford, M., Rajagopal, K., Wilczek, F., Jan. 1999. Color-
avor locking and chiral symmetry breaking in high density QCD. NuPhB, 537 (1), 443{458. [217] Bailyn, C. D., Jain, R. K., Coppi, P., Orosz, J. A., May 1998. The Mass Distribution of Stellar Black Holes. ApJ, 499 (1), 367{374. [218] Ozel, F., Psaltis, D., Narayan, R., McClintock, J. E., Dec. 2010. The Black Hole Mass Distribution in the Galaxy. ApJ, 725 (2), 1918{1927. [219] Farr, W. M., Sravan, N., Cantrell, A., Kreidberg, L., Bailyn, C. D., Man- del, I., Kalogera, V., Nov. 2011. The Mass Distribution of Stellar-mass Black Holes. ApJ, 741 (2), 103. [220] Kreidberg, L., Bailyn, C. D., Farr, W. M., Kalogera, V., Sep. 2012. Mass Measurements of Black Holes in X-Ray Transients: Is There a Mass Gap? ApJ, 757 (1), 36. [221] Wyrzykowski, L., Mandel, I., Apr. 2019. Constraining the masses of microlensing black holes and the mass gap with Gaia DR2. preprint (arXiv:1904.07789). [222] Tsokaros, A., Ruiz, M., Shapiro, S. L., Sun, L., Uryu, K., Feb. 2020. Great Impostors: Extremely Compact, Merging Binary Neutron Stars in the Mass Gap Posing as Binary Black Holes. PhRvL, 124 (7), 071101. [223] Abbott, B. P. et al., Jan. 2020. GW190425: Observation of a Compact Bi- nary Coalescence with Total Mass 3:4M . preprint (arXiv:2001.01761). [224] Abbott, B. P. et al., Dec. 2017. Estimating the Contribution of Dynamical Ejecta in the Kilonova Associated with GW170817. ApJ, 850 (2), L39. [225] Abbott, B. P. et al., Dec. 2017. Search for Post-merger Gravitational Waves from the Remnant of the Binary Neutron Star Merger GW170817. ApJ, 851 (1), L16. 87 [226] Narayan, R., Paczynski, B., Piran, T., Aug. 1992. Gamma-Ray Bursts as the Death Throes of Massive Binary Stars. ApJ, 395, L83. [227] Eichler, D., Livio, M., Piran, T., Schramm, D. N., Jul. 1989. Nucleosyn- thesis, neutrino bursts and
-rays from coalescing neutron stars. Natur, 340 (6229), 126{128. [228] Fattoyev, F. J., Piekarewicz, J., Horowitz, C. J., Apr. 2018. Neutron Skins and Neutron Stars in the Multimessenger Era. PhRvL, 120 (17), 172702. [229] Flanagan, E. E., Hinderer, T., Jan. 2008. Constraining neutron-star tidal Love numbers with gravitational-wave detectors. PhRvD, 77 (2), 021502. [230] Lawrence, S., Tervala, J. G., Bedaque, P. F., Miller, M. C., Aug. 2015. An Upper Bound on Neutron Star Masses from Models of Short Gamma-Ray Bursts. ApJ, 808 (2), 186. [231] Murguia-Berthier, A., Montes, G., Ramirez-Ruiz, E., De Colle, F., Lee, W. H., Jun. 2014. Necessary Conditions for Short Gamma-Ray Burst Pro- duction in Binary Neutron Star Mergers. ApJ, 788 (1), L8. [232] Ai, S., Gao, H., Dai, Z.-G., Wu, X.-F., Li, A., Zhang, B., Li, M.-Z., Jun. 2018. The Allowed Parameter Space of a Long-lived Neutron Star as the Merger Remnant of GW170817. ApJ, 860 (1), 57. [233] Dai, Z. G., Wang, X. Y., Wu, X. F., Zhang, B., Feb. 2006. X-ray Flares from Postmerger Millisecond Pulsars. Sci, 311 (5764), 1127{1129. [234] Metzger, B. D., Quataert, E., Thompson, T. A., Apr. 2008. Short-duration gamma-ray bursts with extended emission from protomagnetar spin-down. MNRAS, 385 (3), 1455{1460. [235] Hinderer, T., Apr. 2008. Tidal Love Numbers of Neutron Stars. ApJ, 677 (2), 1216{1220. [236] Douchin, F., Haensel, P., Dec. 2001. A uni ed equation of state of dense matter and neutron star structure. A&A, 380, 151{167. 88 [237] Fraga, E. S., Pisarski, R. D., Schaner-Bielich, J., Jun. 2001. Small, dense quark stars from perturbative QCD. PhRvD, 63 (12), 121702. [238] Damour, T., Nagar, A., Oct. 2009. Relativistic tidal properties of neutron stars. PhRvD, 80 (8), 084035. [239] Lai, X., Zhou, E., Xu, R., Apr. 2019. Strangeons constitute bulk strong matter: Test using GW 170817. EPJA, 55 (4), 60. [240] Gao, H., Zhang, B., Lu, H.-J., Feb. 2016. Constraints on binary neutron star merger product from short GRB observations. PhRvD, 93 (4), 044065. [241] Ma, P.-X., Jiang, J.-L., Wang, H., Jin, Z.-P., Fan, Y.-Z., Wei, D.-M., May 2018. GW170817 and the Prospect of Forming Supramassive Remnants in Neutron Star Mergers. ApJ, 858 (2), 74. [242] Gao, H., Ai, S.-K., Cao, Z.-J., Zhang, B., Zhu, Z.-Y., Li, A., Zhang, N.-B., Bauswein, A., Jan. 2020. Relation between gravitational mass and baryonic mass for non-rotating and rapidly rotating neutron stars. FrPhy, 15 (2), 24603. [243] Abbott, B. P. et al., Oct. 2017. Gravitational Waves and Gamma-Rays from a Binary Neutron Star Merger: GW170817 and GRB 170817A. ApJ, 848 (2), L13. [244] Margalit, B., Metzger, B. D., Dec. 2017. Constraining the Maximum Mass of Neutron Stars from Multi-messenger Observations of GW170817. ApJ, 850 (2), L19. [245] Shibata, M., Fujibayashi, S., Hotokezaka, K., Kiuchi, K., Kyutoku, K., Sekiguchi, Y., Tanaka, M., Dec. 2017. Modeling GW170817 based on nu- merical relativity and its implications. PhRvD, 96 (12), 123012. [246] Li, S.-Z., Liu, L.-D., Yu, Y.-W., Zhang, B., Jul. 2018. What Powered the Optical Transient AT2017gfo Associated with GW170817? ApJ, 861 (2), L12. 89 [247] Piro, L., et al., Feb. 2019. A long-lived neutron star merger remnant in GW170817: constraints and clues from X-ray observations. MNRAS, 483 (2), 1912{1921. [248] Li, L.-X., Paczynski, B., Nov. 1998. Transient Events from Neutron Star Mergers. ApJ, 507 (1), L59{L62. [249] Metzger, B. D., et al., Aug. 2010. Electromagnetic counterparts of com- pact object mergers powered by the radioactive decay of r-process nuclei. MNRAS, 406 (4), 2650{2662. [250] Kasen, D., Metzger, B., Barnes, J., Quataert, E., Ramirez-Ruiz, E., Nov. 2017. Origin of the heavy elements in binary neutron-star mergers from a gravitational-wave event. Natur, 551 (7678), 80{84. [251] Ruert, M., Janka, H. T., Takahashi, K., Schaefer, G., Mar. 1997. Coalesc- ing neutron stars - a step towards physical models. II. Neutrino emission, neutron tori, and gamma-ray bursts. A&A, 319, 122{153. [252] Rosswog, S., Liebend orfer, M., Thielemann, F. K., Davies, M. B., Benz, W., Piran, T., Jan. 1999. Mass ejection in neutron star mergers. A&A, 341, 499{526. [253] Oechslin, R., Janka, H. T., Marek, A., May 2007. Relativistic neutron star merger simulations with non-zero temperature equations of state. I. Variation of binary parameters and equation of state. A&A, 467 (2), 395{ [254] Hotokezaka, K., Kiuchi, K., Kyutoku, K., Okawa, H., Sekiguchi, Y.-i., Shibata, M., Taniguchi, K., Jan. 2013. Mass ejection from the merger of binary neutron stars. PhRvD, 87 (2), 024001. [255] Wanajo, S., Sekiguchi, Y., Nishimura, N., Kiuchi, K., Kyutoku, K., Shi- bata, M., Jul. 2014. Production of All the r-process Nuclides in the Dy- namical Ejecta of Neutron Star Mergers. ApJ, 789 (2), L39. 90 [256] Sekiguchi, Y., Kiuchi, K., Kyutoku, K., Shibata, M., Taniguchi, K., Jun. 2016. Dynamical mass ejection from the merger of asymmetric binary neu- tron stars: Radiation-hydrodynamics study in general relativity. PhRvD, 93 (12), 124046. [257] Metzger, B. D., Piro, A. L., Quataert, E., Oct. 2008. Time-dependent models of accretion discs formed from compact object mergers. MNRAS, 390 (2), 781{797. [258] Dessart, L., Ott, C. D., Burrows, A., Rosswog, S., Livne, E., Jan. 2009. Neutrino Signatures and the Neutrino-Driven Wind in Binary Neutron Star Mergers. ApJ, 690 (2), 1681{1705. [259] Fern andez, R., Metzger, B. D., Oct. 2013. Delayed out
ows from black hole accretion tori following neutron star binary coalescence. MNRAS, 435 (1), 502{517. [260] Metzger, B. D., Fern andez, R., Jul. 2014. Red or blue? A potential kilo- nova imprint of the delay until black hole formation following a neutron star merger. MNRAS, 441 (4), 3444{3453. [261] Perego, A., Rosswog, S., Cabez on, R. M., Korobkin, O., K appeli, R., Arcones, A., Liebend orfer, M., Oct. 2014. Neutrino-driven winds from neutron star merger remnants. MNRAS, 443 (4), 3134{3156. [262] Martin, D., Perego, A., Arcones, A., Thielemann, F. K., Korobkin, O., Rosswog, S., Nov. 2015. Neutrino-driven Winds in the Aftermath of a Neutron Star Merger: Nucleosynthesis and Electromagnetic Transients. ApJ, 813 (1), 2. [263] Cowperthwaite, P. S. et al., Oct. 2017. The Electromagnetic Counterpart of the Binary Neutron Star Merger LIGO/Virgo GW170817. II. UV, Opti- cal, and Near-infrared Light Curves and a comparison to Kilonova Models. ApJ, 848 (2), L17. 91 [264] Tanaka, M. et al., Dec. 2017. Kilonova from post-merger ejecta as an optical and near-Infrared counterpart of GW170817. PASJ, 69 (6), 102. [265] Radice, D., Perego, A., Zappa, F., Bernuzzi, S., Jan. 2018. GW170817: Joint Constraint on the Neutron Star Equation of State from Multimes- senger Observations. ApJ, 852 (2), L29. [266] Radice, D., Perego, A., Hotokezaka, K., Fromm, S. A., Bernuzzi, S., Roberts, L. F., Dec. 2018. Binary Neutron Star Mergers: Mass Ejection, Electromagnetic Counterparts, and Nucleosynthesis. ApJ, 869 (2), 130. [267] Kiuchi, K., Kyutoku, K., Shibata, M., Taniguchi, K., May 2019. Revisiting the Lower Bound on Tidal Deformability Derived by AT 2017gfo. ApJ, 876 (2), L31. [268] Ai, S., Gao, H., Zhang, B., Dec. 2019. What constraints on the neutron star maximum mass can one pose from GW170817 observations? preprint (arXiv:1912.06369). [269] Drago, A., Lavagno, A., Pagliara, G., Feb. 2014. Can very compact and very massive neutron stars both exist? PhRvD, 89 (4), 043014. [270] Alcock, C., Farhi, E., Olinto, A., Nov. 1986. Strange Stars. ApJ, 310, 261. [271] Gondek-Rosinsk a, D., Bulik, T., Zdunik, L., Gourgoulhon, E., Ray, S., Dey, J., Dey, M., Nov. 2000. Rapidly rotating compact strange stars. A&A, 363, 1005{1012. [272] Haensel, P., Paczynski, B., Amsterdamski, P., Jul. 1991. Gamma-Ray Bursts from Colliding Strange Stars. ApJ, 375, 209. [273] Alcock, C., Farhi, E., Sep. 1985. Evaporation of strange matter in the early Universe. PhRvD, 32 (6), 1273{1279. [274] Madsen, J., Heiselberg, H., Riisager, K., Nov. 1986. Does strange matter evaporate in the early Universen? PhRvD, 34 (10), 2947{2955. 92 [275] Paulucci, L., Horvath, J. E., Benvenuto, O., Jan. 2017. Nucleosynthesis in Strange Star Mergers. IJMPS, 45, 1760042. [276] De Pietri, R., Drago, A., Feo, A., Pagliara, G., Pasquali, M., Traversi, S., Wiktorowicz, G., Aug. 2019. Merger of Compact Stars in the Two-families Scenario. ApJ, 881 (2), 122. [277] Drago, A., Lavagno, A., Metzger, B. D., Pagliara, G., May 2016. Quark de- con nement and the duration of short gamma-ray bursts. PhRvD, 93 (10), [278] Page, D., Geppert, U., Weber, F., Oct. 2006. The cooling of compact stars. NuPhA, 777, 497{530. [279] Page, D., Reddy, S., Nov. 2006. Dense Matter in Compact Stars: The- oretical Developments and Observational Constraints. ARNPS, 56 (1), 327{374. [280] Potekhin, A. Y., Pons, J. A., Page, D., Oct. 2015. Neutron Stars|Cooling and Transport. SSRv, 191 (1-4), 239{291. [281] Li, A., Hu, J. N., Shang, X. L., Zuo, W., Jan. 2016. Nonrelativistic nucleon eective masses in nuclear matter: Brueckner-Hartree-Fock model versus relativistic Hartree-Fock model. PhRvC, 93 (1), 015803. [282] Shang, X. L., Li, A., Miao, Z. Q., Burgio, G. F., Schulze, H. J., Jun. 2020. Nucleon eective mass in hot dense matter. PhRvC, 101 (6), 065801. [283] Dong, J. M., Lombardo, U., Zhang, H. F., Zuo, W., Jan. 2016. Role of Nucleonic Fermi Surface Depletion in Neutron Star Cooling. ApJ, 817 (1), [284] Yakovlev, D. G., Leven sh, K. P., Haensel, P., Aug. 2003. Thermal state of transiently accreting neutron stars. A&A, 407, 265{271. 93 [285] Deibel, A., Cumming, A., Brown, E. F., Reddy, S., Apr. 2017. Late-time Cooling of Neutron Star Transients and the Physics of the Inner Crust. ApJ, 839 (2), 95. [286] Archibald, R. F., Kaspi, V. M., Ng, C. Y., Gourgouliatos, K. N., Tsang, D., Scholz, P., Beardmore, A. P., Gehrels, N., Kennea, J. A., May 2013. An anti-glitch in a magnetar. Natur, 497 (7451), 591{593. [287] Radhakrishnan, V., Manchester, R. N., Apr. 1969. Detection of a Change of State in the Pulsar PSR 0833-45. Natur, 222 (5190), 228{229. [288] Espinoza, C. M., Lyne, A. G., Stappers, B. W., Kramer, M., Jun. 2011. A study of 315 glitches in the rotation of 102 pulsars. MNRAS, 414 (2), 1679{1704. [289] Yu, M., Manchester, R. N., Hobbs, G., Johnston, S., Kaspi, V. M., Keith, M., Lyne, A. G., Qiao, G. J., Ravi, V., Sarkissian, J. M., Shannon, R., Xu, R. X., Feb. 2013. Detection of 107 glitches in 36 southern pulsars. MNRAS, 429 (1), 688{724. [290] Manchester, R. N., Jan. 2018. Pulsar glitches and their impact on neutron- star astrophysics. preprint (arXiv:1801.04332). [291] Haskell, B., Melatos, A., Jan. 2015. Models of pulsar glitches. IJMPD, 24 (3), 1530008. [292] Anderson, P. W., Itoh, N., Jul. 1975. Pulsar glitches and restlessness as a hard super
uidity phenomenon. Natur, 256 (5512), 25{27. [293] Link, B., Epstein, R. I., Lattimer, J. M., Oct. 1999. Pulsar Constraints on Neutron Star Structure and Equation of State. PhRvL, 83 (17), 3362{ [294] Andersson, N., Glampedakis, K., Ho, W. C. G., Espinoza, C. M., Dec. 2012. Pulsar Glitches: The Crust is not Enough. PhRvL, 109 (24), 241103. 94 [295] Chamel, N., Jan. 2013. Crustal Entrainment and Pulsar Glitches. PhRvL, 110 (1), 011101. [296] Watanabe, G., Pethick, C. J., Aug. 2017. Super
uid Density of Neutrons in the Inner Crust of Neutron Stars: New Life for Pulsar Glitch Models. PhRvL, 119 (6), 062701. [297] Baym, G., Pethick, C., Pines, D., Ruderman, M., Nov. 1969. Spin Up in Neutron Stars : The Future of the Vela Pulsar. Natur, 224 (5222), 872{874. [298] Zhou, E. P., Lu, J. G., Tong, H., Xu, R. X., Sep. 2014. Two types of glitches in a solid quark star model. MNRAS, 443 (3), 2705{2710. [299] Lai, X. Y., Yun, C. A., Lu, J. G., Lu, G. L., Wang, Z. J., Xu, R. X., May 2018. Pulsar glitches in a strangeon star model. MNRAS, 476 (3), 3303{3309. [300] Ho, W. C. G., Espinoza, C. M., Antonopoulou, D., Andersson, N., Oct. 2015. Pinning down the super
uid and measuring masses using pulsar glitches. SciA, 1 (9), e1500578{e1500578. [301] Pizzochero, P. M., Antonelli, M., Haskell, B., Seveso, S., Jul. 2017. Con- straints on pulsar masses from the maximum observed glitch. NatAs, 1, [302] Ashton, G., Lasky, P. D., Graber, V., Palfreyman, J., Aug. 2019. Ro- tational evolution of the Vela pulsar during the 2016 glitch. NatAs, 3, 1143{1148. [303] Ge, M. Y., Lu, F. J., Yan, L. L., Weng, S. S., Zhang, S. N., Wang, Q. D., Wang, L. J., Li, Z. J., Zhang, W., Aug. 2019. The brightening of the pulsar wind nebula of PSR B0540-69 after its spin-down-rate transition. NatAs, 3, 1122{1127. 95 [304] Wei, J. Y., et al., Oct. 2016. The Deep and Transient Universe in the SVOM Era: New Challenges and Opportunities - Scienti c prospects of the SVOM mission. preprint (arXiv:1610.06892). [305] Ozel, F., Psaltis, D., Arzoumanian, Z., Morsink, S., Baub ock, M., Nov. 2016. Measuring Neutron Star Radii via Pulse Pro le Modeling with NICER. ApJ, 832 (1), 92. [306] Li, T. et al., Mar. 2018. Insight-HXMT observations of the rst binary neutron star merger GW170817. SCPMA, 61 (3), 31011. [307] Li, D., et al., Apr. 2018. FAST in Space: Considerations for a Multibeam, Multipurpose Survey Using China's 500-m Aperture Spherical Radio Tele- scope (FAST). IMMag, 19 (3), 112{119. [308] Ray, P. S. a., Mar. 2019. STROBE-X: X-ray Timing and Spec- troscopy on Dynamical Timescales from Microseconds to Years. preprint (arXiv:1903.03035). [309] Watts, A. L. et al., Feb. 2019. Dense matter with eXTP. SCPMA, 62 (2), [310] Liu, T., Lin, C.-Y., Song, C.-Y., Li, A., Nov. 2017. a comparison of Grav- itational Waves from Central Engines of Gamma-Ray Bursts: Neutrino- dominated Accretion Flows, Blandford-Znajek Mechanisms, and Millisec- ond Magnetars. ApJ, 850 (1), 30. [311] Zhou, X., Tong, H., Zhu, C., Wang, N., Dec. 2017. Dependence of pulsar death line on the equation of state. MNRAS, 472 (2), 2403{2409. [312] Yasin, H., Sch afer, S., Arcones, A., Schwenk, A., Dec. 2018. Equation of state eects in core-collapse supernovae. preprint (arXiv:1812.02002). [313] Lau, S. Y., Leung, P. T., Lin, L. M., Jan. 2019. Two-layer compact stars with crystalline quark matter: Screening eect on the tidal deformability. PhRvD, 99 (2), 023018. 96 [314] Deng, Z.-L., Gao, Z.-F., Li, X.-D., Shao, Y., Mar. 2020. On the Formation of PSR J1640+2224: A Neutron Star Born Massive? ApJ, 892 (1), 4. [315] Shen, H., Ji, F., Hu, J., Sumiyoshi, K., Mar. 2020. Eects of Symmetry Energy on the Equation of State for Simulations of Core-collapse Super- novae and Neutron-star Mergers. ApJ, 891 (2), 148. [316] Zhang, N.-B., Li, B.-A., Xu, J., Jun. 2018. Combined Constraints on the Equation of State of Dense Neutron-rich Matter from Terrestrial Nuclear Experiments and Observations of Neutron Stars. ApJ, 859 (2), 90. [317] Dietrich, T., Coughlin, M. W., Pang, P. T. H., Bulla, M., Heinzel, J., Issa, L., Tews, I., Antier, S., Feb. 2020. New Constraints on the Supranuclear Equation of State and the Hubble Constant from Nuclear Physics { Multi- Messenger Astronomy. preprint (arXiv:2002.11355). [318] Baldo, M., Burgio, G. F., Centelles, M., Sharma, B. K., Vinas, ~ X., Sep. 2014. From the crust to the core of neutron stars on a microscopic basis. PAN, 77 (9), 1157{1165. [319] Centelles, M., Roca-Maza, X., Vinas, ~ X., Warda, M., Mar. 2009. Nuclear Symmetry Energy Probed by Neutron Skin Thickness of Nuclei. PhRvL, 102 (12), 122502. [320] Chrien, R. E., Dover, C. B., Jan. 1989. Nuclear systems with strangeness. ARNPS, 39 (39), 113{150. [321] Fraga, E. S., Kurkela, A., Vuorinen, A., Feb. 2014. Interacting Quark Matter Equation of State for Compact Stars. ApJ, 781 (2), L25. [322] Frederico, T., Carlson, B. V., Rego, R. A., Hussein, M. S., Mar. 1989. Quark structure of the nucleon and quantum hadrodynamics. JPhG, 15 (3), 297{302. [323] Gamba, R., Read, J. S., Wade, L. E., Jan. 2020. The impact of the crust equation of state on the analysis of GW170817. CQGra, 37 (2), 025008. 97 [324] Gittins, F., Andersson, N., Pereira, J. P., Mar. 2020. Tidal deformations of neutron stars with elastic crusts. preprint (arXiv:2003.05449). [325] Kl ahn, T., Fischer, T., Sep. 2015. Vector Interaction Enhanced Bag Model for Astrophysical Applications. ApJ, 810 (2), 134. [326] Kojo, T., Powell, P. D., Song, Y., Baym, G., Feb. 2015. Phenomenological QCD equation of state for massive neutron stars. PhRvD, 91 (4), 045003. [327] Lattimer, J. M., Lim, Y., Jul. 2013. Constraining the Symmetry Parame- ters of the Nuclear Interaction. ApJ, 771 (1), 51. [328] Lattimer, J. M., Prakash, M., Mar. 2001. Neutron Star Structure and the Equation of State. ApJ, 550 (1), 426{442. [329] Lattimer, J. M., Steiner, A. W., Feb. 2014. Constraints on the symmetry energy using the mass-radius relation of neutron stars. EPJA, 50, 40. [330] Lattimer, J. M., Steiner, A. W., Apr. 2014. Neutron Star Masses and Radii from Quiescent Low-mass X-Ray Binaries. ApJ, 784 (2), 123. [331] Li, A., Wang, R., Aug. 2018. Pulsar glitch and nuclear EoS: Applicability of super
uid model. In: Weltevrede, P., Perera, B. B. P., Preston, L. L., Sanidas, S. (Eds.), IAUS, 337, 360. [332] Li, B.-A., Han, X., Nov. 2013. Constraining the neutron-proton eective mass splitting using empirical constraints on the density dependence of nuclear symmetry energy around normal density. PhLB, 727 (1-3), 276{ 281. 9. [333] Li, B.-A., Steiner, A. W., Nov. 2006. Constraining the radii of neutron stars with terrestrial nuclear laboratory data. PhLB, 642 (5-6), 436{440. [334] Li, Z. H., Schulze, H. J., Aug. 2008. Neutron star structure with modern nucleonic three-body forces. PhRvC, 78 (2), 028801. 98 [335] Link, B., Jul. 2014. Thermally Activated Post-glitch Response of the Neu- tron Star Inner Crust and Core. I. Theory. ApJ, 789 (2), 141. [336] Margaritis, C., Koliogiannis, P. S., Moustakidis, C. C., Feb. 2020. Speed of sound constraints on maximally rotating neutron stars. PhRvD, 101 (4), [337] Muller, H., Serot, B. D., Feb. 1996. Relativistic mean- eld theory and the high-density nuclear equation of state. NuPhA, 606, 508{537. [338] Page, D., Lattimer, J. M., Prakash, M., Steiner, A. W., Dec. 2004. Mini- mal Cooling of Neutron Stars: A New Paradigm. ApJS, 155 (2), 623{650. [339] Patrignani, C. et al., Oct. 2016. Review of Particle Physics. ChPhC, 40 (10), 100001. [340] Pearson, J. M., Chamel, N., Goriely, S., Ducoin, C., Jun. 2012. Inner crust of neutron stars with mass- tted Skyrme functionals. PhRvC, 85 (6), [341] Perego, A., Radice, D., Bernuzzi, S., Dec. 2017. AT 2017gfo: An Anisotropic and Three-component Kilonova Counterpart of GW170817. ApJ, 850 (2), L37. [342] Raaijmakers, G., et al., Dec. 2019. A Nicer View of PSR J0030+0451: Implications for the Dense Matter Equation of State. ApJ, 887 (1), L22. [343] Samajdar, A., Dietrich, T., Feb. 2020. Constructing Love-Q-Relations with Gravitational Wave Detections. preprint (arXiv:2002.07918). [344] Schaeer, R., Zdunik, L., Haensel, P., Sep. 1983. Phase transitions in stellar cores. I - Equilibrium con gurations. A&A, 126 (1), 121{145. [345] Schaner-Bielich, J., Gal, A., Sep. 2000. Properties of strange hadronic matter in bulk and in nite systems. PhRvC, 62 (3), 034311. 99 [346] Schaner-Bielich, J., Hanauske, M., St ocker, H., Greiner, W., Jan. 2002. Phase Transition to Hyperon Matter in Neutron Stars. PhRvL, 89 (17), [347] Soares-Santos, M. et al., Oct. 2017. The Electromagnetic Counterpart of the Binary Neutron Star Merger LIGO/Virgo GW170817. I. Discovery of the Optical Counterpart Using the Dark Energy Camera. ApJ, 848 (2), L16. [348] Vidana, ~ I., Polls, A., Ramos, A., Engvik, L., Hjorth-Jensen, M., Sep. 2000. Hyperon-hyperon interactions and properties of neutron star matter. PhRvC, 62 (3), 035801. [349] Watts, A. L., et al., Apr. 2016. Colloquium: Measuring the neutron star equation of state using x-ray timing. RvMP, 88 (2), 021001. [350] Wei, J. B., Burgio, G. F., Schulze, H. J., Apr. 2019. Neutron star cooling with microscopic equations of state. MNRAS, 484 (4), 5162{5169. [351] Weissenborn, S., Chatterjee, D., Schaner-Bielich, J., Jun. 2012. Hyper- ons and massive neutron stars: Vector repulsion and SU(3) symmetry. PhRvC, 85 (6), 065802. [352] Weissenborn, S., Sagert, I., Pagliara, G., Hempel, M., Schaner-Bielich, J., Oct. 2011. Quark Matter in Massive Compact Stars. ApJ, 740 (1), L14. [353] Whittenbury, D. L., Carroll, J. D., Thomas, A. W., Tsushima, K., Stone, J. R., Jun. 2014. Quark-meson coupling model, nuclear matter constraints, and neutron star properties. PhRvC, 89 (6), 065801. [354] Yakovlev, D. G., Pethick, C. J., Sep. 2004. Neutron Star Cooling. ARA&A, 42 (1), 169{210. [355] Zhang, K., Hirayama, T., Luo, L.-W., Lin, F.-L., Feb. 2020. Compact star of holographic nuclear matter and GW170817. PhLB., 801, 135176. 100 [356] Zhang, K., Huang, G.-Z., Lin, F.-L., Feb. 2020. GW170817 and GW190425 as Hybrid Stars of Dark and Nuclear Matters. preprint (arXiv:2002.10961). [357] Zhang, Z., Chen, L.-W., Sep. 2015. Electric dipole polarizability in Pb as a probe of the symmetry energy and neutron matter around /3. PhRvC, 92 (3), 031301.
http://www.deepdyve.com/assets/images/DeepDyve-Logo-lg.pngNuclear TheoryarXiv (Cornell University)http://www.deepdyve.com/lp/arxiv-cornell-university/neutron-star-equation-of-state-qmf-modeling-and-applications-nxETgQa0eY
Neutron star equation of state: QMF modeling and applications