<<

Gravitational waves from mergers and their relation to the nuclear

L. Baiotti,1

1Department of and Space , Graduate School of Science, Osaka University, Japan

September 30, 2019

Abstract

In this article, I introduce ideas and techniques to extract information about the equation of state of at very high from gravitational waves emitted before, during and after the merger of binary neutron . I also review current work and results on the actual use of the first gravitational-wave observation of a neutron-star merger to set constraints on properties of such equation of state. In passing, I also touch on the possibility that what we observe in gravitational waves are not neutron stars, but something more exotic. In order to make this article more accessible, I also review the dynamics and gravitational-wave emission of neutron-star mergers in general, with focus on numerical simulations and on which representations of the equation of state are used for studies on binary systems.

Contents

1 Opening words 2

2 Introduction 4 2.1 Status of numerical-relativity simulations for binary neutron-star mergers ...... 4 2.2 The dynamics of neutron-star binaries ...... 6

arXiv:1907.08534v3 [astro-ph.HE] 26 Sep 2019 2.3 Gravitational-wave emission from binary–neutron-star mergers ...... 8 2.4 Describing equations of state with generic parameterizations ...... 8 2.5 Other compact objects possibly similar to neutron stars ...... 12 2.6 Universal relations for neutron stars ...... 13

3 Extracting information on the equation of state from gravitational waves emitted before the merger 15 3.1 The basic idea ...... 15 3.2 Applications to data analysis ...... 19 3.3 Estimates about inspiral waveforms ...... 21 3.4 Eccentric mergers ...... 23

1 4 Extracting information on the equation of state from gravitational waves emitted after the merger 24 4.1 Basic ideas ...... 25 4.2 Applications to data analysis ...... 29 4.3 Estimates about future observations ...... 30 4.4 Investigating phase transitions with post-merger waveforms ...... 31

5 Combining analyses of waveforms emitted before and after the merger 33

6 The merger of binaries of stars not made of ordinary matter 34

7 Constraints on stellar radius and equation-of-state parameters deduced from GW170817 35 7.1 Hadronic equations of state ...... 36 7.1.1 Correlations of nuclear-matter parameters of the equation of state ...... 39 7.2 stars ...... 39 7.3 Hybrid stars ...... 40 7.4 Possible influence of ...... 45 7.5 Summary of radius measurements ...... 47

8 Conclusion 47

9 Acknowledgements 49

1 Opening words

For decades hope has grown that measurements of gravitational waves (GWs), and related electromag- netic radiation, from mergers of binaries composed of compact stars1 may set more stringent constraints on the properties of the equation of state (EoS) of matter at densities higher and much higher than nuclear . Such hope has become reality since August 2017, when the first measurement of gravita- tional radiation interpreted as the coalescence of a binary–neutron-star (BNS) merger has been recorded as GW170817 [1]. There have been numerous follow-up observations (triggered by the GW detection) of electromagnetic radiation [4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28] from what are thought to be the material ejecta of the merger. While concluding this review, the third observational run of the LIGO-Virgo Collaborations has started and the observation of some candidate BNS waveforms has been announced. However, details have not been released and therefore for analysis purposes GW170817 is currently the only observation of the kind. And this one

1I have used the term compact stars, because it does not contain any hints on their composition: It just differentiates them from black holes or less dense stars. Indeed, what is inside compact stars like those in the first GW detection (GW170817) of objects thought not to be black holes [1] (but see [2] for the possibility that GW170817 was a merger of a with a ) is still largely unknown and it is indeed one of the scopes of this article to describe the status of research about the composition of compact stars (see Ref. [3] for a recent review), in view of GW observations. However, in most of the literature the term neutron stars is employed, without assuming that they are made only or mostly of . Exceptions to this routine use are works that explicitly study non-nucleonic equations of state for compact stars (see Sects. 4.4, 7.2 and 7.3). These works often distinguish compact stars into neutron stars, quark stars, strange stars (quark stars made of matter), hybrid stars (with a core of quark matter and an outer region of nucleonic matter). Furthermore, the term compact stars may be used to refer to objects that are not made of ordinary matter, such as boson stars (see Sect. 2.5). Having presented these notes of caution on nomenclature, in the rest of this article, I will mostly use the term neutron stars to refer to compact stars made of ordinary matter in general. Only where specification may be necessary, or at least useful, I will employ the term compact stars or one with further particularization.

2 observation has indeed already been used to gain knowledge2 on the EoS, in addition to several other findings, such as more stringent limits on the difference between the speed of and that of [32], on the equivalence principle (through Shapiro-delay [33] measurements), on Lorentz invariance [32] and on cosmologically modified gravity [34, 35, 36, 37, 38, 39, 40]. In this review I assume that the reader is familiar with nuclear EoSs. Therefore, I will not enter into details of the various proposals for describing and computing the EoS from fundamental principles with various levels of approximations, or the details of any of the results of such computations. For reviews on the current status of research in computations of EoS with NSs in mind see, e.g., Refs. [41, 3]. As is well known, a large number of experiments carried out in laboratories on Earth [42, 43, 44, 45, 46, 47, 48, 49, 50, 51], especially heavy-ion reactions with radioactive beams, constrain the EoS of matter in various ways and to various degrees up to around . Considerable attempts have been made to constrain the EoS at densities higher than the nuclear density, but it has turned out to be difficult to put stringent constraints on the EoS parameters [42, 52, 53, 54, 55, 56, 57, 47, 58, 59, 60]. Astrophysical observations (like those involving binary and x-ray binaries [50, 49, 61, 62, 63, 64, 65]) have been giving additional constraints on the EoS of ultra-high density matter, by trying to measure the and radii of neutron stars (NSs). Further information on the EoS of ultra-high density matter can be obtained through BNS mergers observations by determining (see further below for definitions): (i) the tidal deformability (sometimes also called tidal polarizability) Λ for a NS of a given ; (ii) the amount of material ejected from the merger (ejecta), which gives rise to a macronova and to radio emissions (see, e.g., Ref. [66]); (iii) the maximum mass Mmax of a non-rotating compact star that is stable against collapse to a black hole. The tidal deformability Λ is an EoS-dependent, dimensionless function of the NS mass that correlates with the gradients inside the star, namely with the stiffness of the EoS. Tidal deformations cause a phase shift (the phasing of the GW signal is significantly more important for parameter estimation than its amplitude [67]) in the waveform relative to the merger of point-particles, which is measurable with existing GW detectors [68, 69, 70, 71]. More details will be given in Sect. 3.1. A macronova [72] is an astronomical source of electromagnetic radiation about one to three orders of magnitudes brighter than a regular nova3, hence the name macronova [73] or [74] (the term mergernova has also been recently used [75]). In the standard scenario, macronovae shine hours to days after the merger in ultra-violet, optical, and bands and their power source is thought to be, with support by the observation of the macronova associated with GW170817 [7, 22, 24], the of r-process elements [72, 73, 74] (however, see also [76] for an alternative explanation). Macronovae are particularly promising electromagnetic counterparts to BNS mergers because their emission is relatively isotropic, contrary to gamma-ray bursts, which are thought to be highly beamed and thus observable only in some cases. In addition to the macronova associated with GW170817, named also AT2017gfo, a few other candidates have been identified (without a GW counterpart, of course) [77, 78, 79, 80, 81, 82, 83, 84]. Observational constraints on the quantities mentioned above (tidal deformability, maximum mass, amount of ejecta) can be obtained from BNS mergers through (i) GWs from the late inspiral, (ii) GWs from post-merger oscillations and (iii) electromagnetic emissions (macronova and radio emission). The rest of this article is organized as follows. In Sect. 2, I will review the dynamics and GW emission of BNS mergers in general, with focus on numerical simulations and on which representations of the EoS are used for studies on BNS systems. Then, in Sect. 3, I will introduce ideas and techniques to find information about the EoS from GWs emitted before the merger of the NSs (coalescence or

2There is, naturally, much excitement about this first detection, but several experts in the field think that GW170817 has actually not provided new insights about the EoS that cannot be obtained from knowledge already available from nuclear physics theory and experiments, and from preceding astrophysical observations. However, to my knowledge, statements of this kind have appeared only in a few articles [29, 30, 31]. See the rest of this review for more details. 3Novae are caused by fusion explosions on a accreting from a larger companion star.

3 inspiral phase), while in Section 4 I will focus on analyses of gravitational radiation emitted during and after the merger (merger and post-merger phases) and in Sect. 5 on ideas for combining pre- and post-merger waveforms. Finally, before concluding, in Sect. 7 I will review current work and results on the actual use of the one GW observation of BNS merger currently available to set constraints on EoS properties. In passing, in Sect. 6, I will also touch on the possibility that what we observe in GWs are not NSs, but something more exotic.

2 Introduction

For the convenience of the reader, I present here some of the generalities of BNS systems and of popular ways used to advance our knowledge on them, in particular on their EoS.

2.1 Status of numerical-relativity simulations for binary neutron-star merg- ers In order to help and interpret observations, we need solutions of the general-relativistic equations describing spacetime4, matter and radiation (in particular, magnetic fields). As everyone knows, analytic solutions of astrophysical relevance for BNS systems are not available, therefore the field of – the science of simulating (solving) general-relativistic dynamics on computers – has seen lots of efforts being put into it and a reasonable amount of reliable results. Numerical relativity is now mature, but at the beginning, a couple of decades ago, it has struggled to get decent results because straightforward discretization of the Einstein equations just does not work. In addition to the standard problems of any numerical simulation, there are multiple reasons for the increased difficulty inherent to general-relativistic simulations: (i) The formulation of the equations is not self-evident; e.g. time is not simply defined and very careful variables definitions are needed to obtain a system that is strongly hyperbolic; (ii) Physical singularities may be present and need special treatment; (iii) While not carrying physical information, gauges play an important role in numerical stability, for example in countering grid stretching. For explanations on all these issues, well-written textbooks are nowadays available [98, 99, 100, 101, 102]. Despite such difficulties, nowadays the number of research groups with their own independent codes capable of performing (at least in some respects) state-of-the-art BNS simulations is of order ten. The current status of the capabilities of codes used for simulating BNS can be summarized as follows. (i) All codes can robustly compute the matter and dynamics (including long-term evo- lutions of the formed black hole and disc), even if improvements are being constantly made. The selection of appropriate gauges and the extraction of GW signals from the dynamics is nowadays routinely done by everyone in the field. The EoSs used in simulations are based on published work from specialists in nuclear theory and so on and are either piecewise polytropes5 [103] (plus a thermal part, at times) or tabulated [104, 105, 106, 107]. Also the first general-relativistic simulations of merging NSs including at finite temperatures have been performed recently [108, 109]. See Sect. 4.4 for more details on these works on phase transitions. (ii) In addition to the theme of this review, namely the linking of GW observations to physical properties of the emitting system and in particular the EoS, there is a lot of ongoing work and some robust results also on heavy-element production [110, 111, 112, 113, 114, 115, 116, 117, 118, 119, 120,

4Except for some side remarks and Sect. 6, in this review I will assume that the theory of gravity is , because the prospects for deviations from general relativity in stellar-mass objects (see Refs. [85, 86] for reviews) are severely limited, also by observations of GWs from mergers of binary black holes [87] and NSs [32], and because numerical simulations of BNS mergers in alternative theories of gravity are very few [88, 89, 90, 91, 92, 93, 94, 95, 96, 97]. 5See Sect. 2.4.

4 105, 106, 121, 122, 123, 124, 125, 126] and macronovae [110, 127, 128, 116, 117, 106, 129, 130, 131, 132], on improved initial data, including better computations of BNS with non-negligibly spinning stars [133, 134, 135, 136, 137, 138, 139, 140, 141, 142, 143, 144, 145, 146, 147, 148, 149] and BNS with eccentricity [150, 151, 111, 106, 152, 153, 154, 155], and on the treatment of physical viscosity [156, 157, 158, 159, 160]. (iii) Some other issues, instead, are still very open, in particular the effects of magnetic fields and of radiation transport, especially radiation transport. Simulating magnetic fields is challeng- ing because of physical instabilities that require very high resolutions to be resolved and because of limitations in the modelling of electromagnetic interactions. Most simulations, in fact, are carried out in the approximation, which does not capture all the physical effects, like up- per limits to the growth of instabilities. Resistive-magnetohydrodynamics simulations exist in small numbers [161, 162, 163], but they are limited by our lack of knowledge about the resistivity of matter in and around NSs. The open problems with magnetic fields in BNS mergers apply especially to the post-merger phase, where magnetic fields may have a huge importance for the dynamics itself, for the ejecta, and for electromagnetic emissions from the vicinity of the merged object (like those thought to produce short gamma-ray bursts). Before the merger, magnetic fields are not relevant for the global dynamics [164], but may produce observable electromagnetic radiation, as found in works employing resistive magnetohydrodynamics [161, 162]. Effects of and in general of radiation transport are also important for the production of ejecta and direct electromagnetic emissions and advances are progressively being made [165, 104, 105, 123, 157, 126, 166].

Figure 1: Isodensity contours on the equatorial plane for the of a BNS with the polytropic EoS. The thick dashed line in the last panel represents the apparent horizon. (From Ref. [167])

5 Figure 2: Isodensity contours on a meridional plane, highlighting the formation of a torus surrounding the central black hole, whose apparent horizon is indicated with a thick dashed line. The data refer to a low-mass binary evolved with a polytropic EoS (cf. Fig. 1). (From Ref. [167])

2.2 The dynamics of neutron-star binaries

For a more in-depth overview on the state of the art of numerical studies on BNS systems, the reader is referred to recent review articles, like Refs. [168, 169, 170]. Here, I will only give a brief overview of their dynamics, also by showing images from a representative simulation taken from Ref. [167]. Fig. 1 collects some representative isodensity contours (i.e. contours of equal rest-mass density) on the equatorial plane. The initial coordinate separation between the maxima in the rest-mass density (defining the stellar centres) is 45 km. As the stars inspiral with increasing angular velocity, each one becomes more and more tidally deformed by the gravity of its companion. This to an increase in the inspiral rate [68], which also depends, of course, on the total of the system, the orbital one and the spins of the stars. During the merger, when regions of the two stars with density around a factor of a few less than their maximum density come into contact, a noticeable vortex sheet (or shear interface) develops, where the tangential components of the velocity exhibit a discontinuity. This condition is known to be unstable to very small perturbations and it can develop a -Helmholtz instability, which will curl the interface forming a series of vortices [171]. This is indeed what is observed in all simulations of this kind, with features that are not much dependent on the mass or on the EoS used. Even if this instability is purely hydrodynamical, it can have strong consequences when studying the dynamics of BNS systems in the presence of magnetic fields, because it leads to an exponential growth of the toroidal component of the magnetic field even if the initial magnetic field is a mostly or purely poloidal one6 [175, 176, 167]. Through this mechanism, the instability can to an overall amplification of the magnetic field of about three orders of [177, 178, 179, 104, 180, 181, 159]. Such high magnetic fields are presumed to be behind the phenomenology of [182, 183] and short hard gamma-ray bursts [184, 185, 186, 187]. After the merger, the cores of the two NSs coalesce. During their rapid infall they experience a considerable decompression of 15%, for about 1 ms. Given the choice of mass of the initial configu- ration and of the EoS, in the simulation≈ shown in Fig. 1 the merged object is initially a hypermassive (HMNS), i.e. a NS with mass above the upper limit for uniformly rotating NSs and that is temporarily supported against collapse by differential rotation and thermal gradients [188]. The HMNS undergoes a number of violent non-axisymmetric oscillations, with a dominant overall m = 2 defor-

6Realistic descriptions of NSs, isolated or in binaries, probably require poloidal-toroidal mixed field configurations [172, 173, 174].

6 mation7, i.e. a bar, as the system moves towards an energetically favourable configuration through the rearrangement of the angular-momentum distribution [190, 191, 192, 193, 194, 195, 196, 180, 197, 198]. As the bar rotates, it also loses large amounts of angular momentum through gravitational radiation. m = 1 deformations have also been reported [199, 200, 176, 140, 141, 201, 202, 144, 203, 204, 152, 205]. Together with these oscillations, a secular increase of the central rest-mass density is also observed. Then, about 15 ms after the merger, the maximum rest-mass density is seen again to increase rapidly, exponentially, and the object collapses to a (Kerr black hole). This is expected from the fact that the HMNS born after the merger, while initially not beyond the stability limit for gravi- tational collapse, has lost its original differential rotation8. The collapse is marked by the appearance of an apparent horizon9. When an apparent horizon appears, a large amount of high–angular-momentum matter remains outside of it in the form of an accretion torus. In the example presented here, it has an average density between 1012 and 1013 g/cm3, a vertical size of a few km and a horizontal extension of a few tens of km. The initial rest mass of the torus is an important quantity to be determined in simulations, since it is related to the ejecta and to electromagnetic emission in general. In particular, the existence of a massive torus around the newly formed rotating black hole is a key ingredient in the modelling of short gamma-ray bursts. The dynamics of the torus are summarized for the representative case adopted here in Fig. 2, which shows the isodensity contours on a meridional plane. Note that the panels refer to times between the times of the last two panels of Fig. 1. Overall, the torus has a dominant m = 0 (axisymmetric) structure but, because of its violent birth, it is very far from an equilibrium. As a result, it is subject to large oscillations, mostly in the radial direction. In the example simulation presented above, the merged object is a HMNS that collapses to a rotating black hole in a few tens of milliseconds. More in general, a HMNS can exist for up to 1 s, during which time cooling through neutrino emission, angular-momentum transport associated with∼ magnetic-field effects (such as the magneto-rotational instability and magnetic braking), and the gravitational torque resulting from its non-axisymmetric structure lead the remnant to collapse [212, 179, 213]. As it can be easily imagined, depending on the initial mass of the system and on the EoS, other outcomes are possible for the merger. The merger of BNS with higher masses and/or softer EoSs ends in a prompt collapse to a black hole, while binaries with lower rest masses and/or stiffer EoSs produce a merged object that does not collapse for a longer time. This would likely be a supramassive NS [214], namely an axisymmetric, uniformly rotating NS with mass exceeding the upper limit for nonrotating NSs. After losing angular momentum through secular mechanisms likely related to electromagnetic emission, it would then collapse to a black hole on timescales of 10 104 s [215]. Finally, in possibly marginal cases, the merged object may not collapse at all, namely∼ it− may become a stable NS. See Fig. 1 of Ref.

7As widely known, stellar deformations can be described decomposing the linear perturbations of the or rest- mass density as a sum of quasi-normal modes that are characterized by the indices (l,m) of the spherical harmonic functions. Then the m mentioned in the text is the dominant term of such expansion. m = 0 is a spherical perturbation, m = 1 is a one-lobed (or one-armed) perturbation, m = 2 is a bar-shaped perturbation. See Ref. [189] for a review. 8Note that the physical non-axisymmetry of the HMNS and coordinate effects due to the choice of gauges (gauges are usually not fixed in numerical relativistic simulations, but are evolved together with the other spacetime quantities [98, 99, 100, 101, 102]) make it difficult to provide a unique measure of the degree of differential rotation (but see Refs. [141, 206]). On average, however, the angular velocity decreases of about one order of magnitude between the rotation axis and the surface. 9Black holes are defined through their event horizons, which are surfaces bounding a region in spacetime inside which events cannot affect an outside observer. In numerical computations it is possible [207] but not easy to find event horizons, because information from all times are necessary for this and so it can be computed only at the end of the simulation. Instead, apparent horizons are usually searched for in numerical simulations. These are surfaces that are the boundary between light rays that are directed outwards and move outwards, and those that are directed outwards but move inwards. Apparent horizons are a local concept and thus they can be computed immediately for each time step of a simulation, with efficient techniques [208, 209, 210, 211].

7 [168] for a schematic view of what said in this paragraph about the possible outcomes of the merger.

2.3 Gravitational-wave emission from binary–neutron-star mergers Representative waveforms from the merger of different BNS systems are shown in Fig. 3, taken from Ref. [216]. Reflecting the dynamics of matter as described above, gravitational waveforms increase in amplitude and in during the inspiral (the so-called chirp signal10), while the waveforms after the merger, reflecting the oscillations of the merged object, are much more varied and in many cases terminate with the ringdown11 signal of the black hole, during which the distortions with respect to a Kerr black hole are damped in a characteristic GW signal. The ringdown signal for black holes formed in BNS mergers is at of several kHz and so not easily measurable by current detectors. The post-merger signal from before the collapse is at lower frequencies than the ringdown, but still so high that detection is probably limited to close sources [222, 223, 224, 225, 226, 227, 228]. The inspiral signal, instead, can be better measured in current detectors, because of its duration (tens of seconds or even more) and its frequency range of up to 1 kHz. Detectors that are currently active (Advanced LIGO [229], Advanced Virgo [230]), under construction∼ (KAGRA [231], Indigo/LIGO India [232]) and most of those that are planned (LIGO Voyager [233], [234, 235], Cosmic Explorer [233]), in fact, are projected to have maximum sensitivity around 1 kHz (see also Ref. [236] for a recent proposal on further detector developments).

2.4 Describing equations of state with generic parameterizations Given the complexity of the interactions and the structure of NSs, all of the (supra)nuclear EoSs proposed so far use certain approximations and involve large numbers of fundamental parameters. Researchers have indeed been checking some of the proposed EoSs against the data we could gather from GW170817, but another more general informative approach is to parameterize possible EoSs with a few phenomenological parameters and to analyse which constraints are imposed by observations on these parameters. These results are less model dependent, since they do not make use of a specific EoS among those proposed. However, the choice of parameters for the parameterization is itself a model and actually the form of the parameterization chosen has consequences on the Bayesian inference of EoS parameters made using astrophysical data [238, 239, 240]. A few parameterizations have been proposed and used, with different degrees of complexity and fidelity to realistic EoS models [241]. One is the parameterization with piecewise [103, 242, 243, 244, 245, 246, 247], which, as the name says, consists in stitching together polytropic EoSs with different adiabatic indexes Γ(ρ) for different ranges of density. It has been found that the optimal number of polytropic pieces is five, each describing a density interval between one and several (eight) times the nuclear saturation density [246, 248]. It is a rather straightforward method, which is used in most numerical-relativity simulations because of its ease and efficiency of implementation, and which allows to parameterize a large portion of the space of possible EoSs with a few parameters (usually only three or four pieces are used and necessary to get satisfactory agreement with all EoSs computed from fundamental principles), but it has drawbacks as well. One of them is that when comparing parameterized EoSs with specific EoSs computed from fundamental principles, the accuracy drops near the fixed joining densities [245]. Another drawback is that the piecewise-polytropic parameterizations

3/5 10 (mAmB ) From the chirp signal, the Mchirp = 1/5 (the dominant parameter of the inspiral) of the binary (mA+mB ) system can be measured accurately [217, 218, 219], while the component masses mA and mB have much larger errors. For GW170817, the LIGO-Virgo Collaborations determined Mchirp = 1.186 0.001M [220]. 11At its birth from collapse (or merger), the black hole rings, namely oscillates± in shape. Since this ringing is rapidly damped through the emission of GWs, it is called ringdown. Its frequencies are given by its quasinormal modes of oscillation, calculable through perturbation theory (see Ref. [221] for a review).

8 22 10− 4 × IF q10 IF q08 H4 q10 H4 q08

3

2

1

0 at 100 MPc + 22

h 1 −

2 −

3 −

4 − 6 5 4

[kHz] 3

f 2 1 0 5 0 5 10 15 10 5 0 5 10 5 0 5 10 15 10 0 10 20 − − − − − − − (t r) [ms] (t r) [ms] (t r) [ms] (t r) [ms] − − − −

Figure 3: Some representative plots of the GW strain (l = 2; m = 2 mode only) taken from Ref. [216]. Signals are for (from left to right): an equal-mass binary evolved with an ideal-fluid EoS, a binary with a mass ratio of the components of q = 0.8 evolved with an ideal-fluid EoS, an equal-mass binary evolved with the H4 EoS [237], a binary with a mass ratio of the components of q = 0.8 evolved with the H4 EoS. The top panels show the strain at nominal distance of 100 Mpc. The lower panels show the instantaneous frequency. (From Ref. [216])

of the EoS may be thermodynamically non-convex12 at some of the juncture points even if the physical EoS is convex [253, 249, 252] and this may also lead to spurious deformations of the emitted GWs [252]. Furthermore, it was found that, in principle, in some cases a piecewise-polytropic parameterization does not admit an invertible mapping between the phenomenological EoS parameters and quantities like gravitational mass, equatorial radius, and of the star [238, 239]. However, this may not be a serious problem, as in practice Bayesian analysis of observations needs to be done [254], and it has been shown that it is feasible to use gravitational signals to solve the relativistic inverse stellar problem for piecewise-polytropic parameterizations, i.e., to reconstruct the parameters of the EoS from measurements of the stellar masses and tidal deformability (or tidal Love number; see Sect. 3.1) with a few observations made by advanced interferometers [255]. Another parameterization is the spectral parameterization [256, 257, 258], which expresses the log- arithm of the adiabatic index of the EoS, Γ(p, γi), as a polynomial of the pressure p; γi are the free EoS parameters. This can match a wide variety of EoSs, usually better than piecewise-polytropic models with the same number of parameters (however, in cases in which the candidate EoSs contain phase tran- sitions the residuals are similar) [256] and without the above-mentioned problems at the joining points. The spectral parameterization has recently seen further developments, including the self-imposition of the causality constraint [241, 259], which would lead to improved computational efficiency. However,

12The convexity of any EoS is mathematically defined in terms of the value of the fundamental derivative, which measures the convexity of the isentropes in the pressure-density plane. If the fundamental derivative is positive, isentropes in the pressure-density plane are convex, and thus rarefaction waves are expansive and shocks are compressive [101, 249]. This is the usual regime in which many astrophysical scenarios develop. However, some EoSs may display regimes in which the fundamental derivative is negative, namely, the EoS is non-convex. In this case rarefaction waves are compressive and shocks are expansive. These non-classical or exotic phenomena have been observed experimentally [250, 251]. Also, it has been shown that any first-order transition (see Sect. 7.3) would lead to a non-convex thermodynamics [252].

9 the improved version has not yet been used in applications to real data. It was also shown that the spectral parameterization is free from the invertibility problems that may affect the piecewise-polytropic representation (as long as the number of measured data points is greater than or equal to the number of EoS parameters) [259]. Both the piecewise-polytropic and the spectral parameterizations can be stitched to an EoS known to be realistic at lower densities (e.g., below half the nuclear saturation density), like the SLy EoS [260]. When they are used with random parameters to survey a large portion of the space of EoSs, rather than used to mimic specific tabulated EoSs, this connection at low density is actually the only detailed input that these representations get from known nuclear physics. Another limitation of these param- eterizations is related to the fact that they automatically include the assumption of β-equilibrium13. After they form, NSs cool rapidly via neutrino emission and quickly reach β-equilibrium, therefore thermal and dissipative corrections to the EoS are negligible, and NS matter is typically modelled as a perfect fluid. However, during tidal disruption and merger of a BNS system, matter heats up and is not in such an equilibrium [262]. For some time interval, then, the piecewise-polytropic and the spectral parameterizations may not offer a good description of matter. The common workaround used for this in numerical simulations is the addition by hand of a thermal component of the EoS modelled as an ideal fluid, in which the pressure p is related to the internal  and the rest-mass density ρ through the adiabatic index Γ as p = (Γ 1)ρ. − The two representations introduced above do not allow for a simple connection to knowledge on nu- clear physics and they cannot bring information about matter composition, such as the fraction. A third and complementary way to obtain a parameterization of the EoS is to Taylor expand the en- ergy per E of isospin asymmetric near the saturation density ρ0 (see, e.g., Refs. [263, 264, 265, 266, 267, 268, 269]). The expansion is carried out in the isospin symmetry parameter δ (ρn ρp)/ρ (with ρp and ρn representing the proton and neutron number densities respectively and ≡ − ρ ρp + ρn) around the symmetric nuclear matter case δ = 0 [270]: ≡ 2 4 E(ρ, δ) = E(ρ, 0) + Esym(ρ)δ + (δ ), (1) O where E(ρ, 0) corresponds to the energy of symmetric nuclear matter. E(ρ, 0) and Esym(ρ) are then expanded themselves around the saturation density

K0 2 Q0 3 4 E(ρ, 0) = E0 + y + y + (y ), (2) 2 6 O Ksym 2 Qsym 3 4 Esym(ρ) = Esym,0 + Ly + y + y + (y ), (3) 2 6 O where y (ρ ρ0)/3ρ0. These lowest-order parameters are known as the energy per particle E0, ≡ − the incompressibility coefficient K0, the third derivative of the energy of symmetric matter Q0, the symmetry energy Esym,0, its slope L, its curvature Ksym, and its skewness Qsym at saturation density. Some of these parameters (E0, K0, and Esym,0) are either strongly constrained by terrestrial experiments or have no noticeable impact on the bulk properties of NSs (see, e.g., Refs. [265, 29, 268, 271] and Ref.

13β-equilibrium is the condition of equilibrium between , and neutrons with respect to the β-decay and inverse β-decay processes:

n p + e +ν ¯ , → e− + p n + ν , → respectively. Neutrinos escape from the system. Inverse β-decay can proceed whenever the has enough energy to balance the mass difference between the proton and the neutron. β-decay is blocked if the density is high enough that all electron energy levels in the Fermi sea are occupied up to the one that the emitted electron would fill. See, e.g., Ref. [261].

10 [272] for a recent review.). The slope of the incompressibility

M0 = Q0 + 12K0 (4) is also addressed in some works [273, 274, 267, 268]. For some results on constraints on the above parameters from GW observations see Sect. 7.1.1. A different way of parameterizing the nuclear-matter EoS in terms of nuclear parameters consists in expanding in the proton fraction xp = ρp/(ρn + ρp) from pure neutron matter to symmetric neutron matter [275]. The EoS representations discussed above are used only or mainly to describe smooth, nucleonic EoSs and may not work very well if, at higher densities, hadrons undergo deconfinement and NSs have a quark-matter core [276, 277, 278]. In the literature, three main effective models have been proposed and used to address this phase (see, e.g., Refs. [279, 280] for reviews): (i) the relatively simple MIT bag model [281, 282] (and its many variations), where quarks are treated as massless particles confined in a bag of finite dimension through a bag pressure B (called the bag constant), a phenomenological quantity introduced to take into account the nonperturbative effects of QCD; (ii) the much more complicated Nambu–Jona-Lasinio model [283, 284], and (iii) a simpler, phenomenological description that can mimic the above two sophisticated models by assuming a density-independent speed of sound, known as constant-speed-of-sound parameterization [285, 286, 287], and that is much better suited for numerical computations [288, 289]. The speed of sound may be a better parameter also because it is physically constrained by both causality requirements and an asymptotic limit at 2 ultrahigh density, cs 1/3, suggested by theoretical insights [290] and perturbative QCD calculations [291, 292]14. Alternative→ parameterizations of the speed of sound have been proposed and used recently in other works [297, 298, 299, 240, 275, 296]; see Sect. 7.3. Other models to parameterize EoSs in general have been proposed, but are currently rarely or not used in simulations and data analysis. One is a model in which the EoS is divided into four density regimes: a fixed crust below the nuclear saturation point, one pressure-energy relation depending on nuclear-physics parameters (as symmetry energy and proton/electron fraction) for densities around the saturation point, and two polytropic relations at larger densities to fit the inner core [242]15. This parameterization was devised from considering data from photospheric-radius-expansion bursts and quiescent thermal emission from x-ray transients [242] and cannot include EoS that allow for twin stars [302, 303, 304] (see Sect. 7.3 for more on twin stars). Another method combines the better-known EoS of nuclear matter at lower densities (as done in the representations described above) with the requirement that the pressure approaches that of deconfined quark matter at high densities [305]. This places significant further restrictions on the possible EoSs, if one accepts the assumptions made. Up to this point I have summarized parametric models that describe the EoS. However, as mentioned above and as recently pointed out in Ref. [306], parametric EoS inference has limitations, due to un- avoidable modelling errors [256, 245, 247], especially for sharp features like first-order phase transitions (see Sect. 7.3 for more on phase transitions and hybrid stars). Given the extremely varied phenomenol- ogy of (hybrid) EoSs, it is unlikely that any single parametric model will be able to faithfully represent the full range of EoS variability with only a few parameters. In order to remedy these drawbacks, non-parametric representations of the EoS have been proposed [306, 264, 265]. Non-parametric repre- sentations do effectively involve some parameters (called hyperparameters) that control allowed types

14Note that not all agree on this limit [293, 294]. Furthermore, such an asymptotic value would probably be reached only at densities orders of magnitude higher than the maximum densities attained in NSs, even after the merger (see, e.g., Refs. [31, 295]). It is likely that the speed of sound reaches a maximum at some density before tending to the asymptotic limit [31]. Ref. [296] has recently discussed the constraints that astrophysical observations place on the speed of sound of quark matter in compact stars . 15This representation turned out to be a generalization of that of Refs. [300, 301].

11 of functional behaviour, but their coverage of the space of EoSs is much larger than that provided by a parametric model. The key difference between parametric and non-parametric models is that the faithfulness of the representation increases with the number of parameters in the former case, while in the latter case it scales not with the number of hyperparameters but with how representative the input knowledge (in the specific case, the set of EoSs) used to constrain the functional behaviour is. It can then be expected that systematic errors are much smaller in non-parametric representations, if the a priori knowledge is realistic and accurately input. One example of non-parametric representations is the metamodel for the nucleonic EoS of Refs. [264, 265]. A metamodel, or surrogate model, is a model of models, i.e. a simplified model of actual models, EoS representations in this case. Conceptually, a metamodel is like building a hypersurface from a limited amount of known data from the underlying models and approximating the output over a much wider parameter space. Metamodels have to be evaluated with respect to their goodness (how well they fit a set of data) and there is no proof of existence or of uniqueness in general. Some of the advantages of this metamodel of Refs. [264, 265] are that (i) it covers a wider space of EoSs, which may not be covered by the underlying models or by specific elements like specific EoSs (as opposed to parameterized EoSs) or experimental data; (ii) it allows the incorporation of knowledge on nuclear physics acquired from laboratory experiments or the results of complex ab initio models; (iii) it can easily be used in the Bayesian framework, facilitating the estimation of the experimental and theoretical error bars into confidence levels for the astrophysical observables. Refs. [264, 265] even argue that predictions done with their metamodel are without assumptions on the functional form of the EoS and only require that the EoS is nucleonic and satisfies basic physical constraints, like thermodynamical stability and causality, and therefore can be qualified as model independent. As a final addition to this Section, note that all recent works on NS EoS, especially those reviewed in Sect. 7, impose physical limits on the parameterizations used. In particular, the following is required for viable EoSs. (1) Causality: The speed of sound must be less than the everywhere, in particular up to the central pressure of the heaviest NS supported by the EoS16. (2) Observational consistency: The EoS allows for a maximum equal to or larger than the largest observed mass of a NS. Among the observations and inferences therefrom that are well- understood and therefore reliable, it has been recently announced in April 2019 that a NS of estimated 0.11 mass 2.17 M was observed in PSR J0740+6620 [310]. The other highest NS masses reliably ±0.10 observed up to now are 1.908 0.016M in PSR J16142230 [311, 63] and 2.01 0.04M in PSR J0348+0432 [312]. Different works± published before April 2019 use different, more or± less conservative, values for such lower limit to the maximum mass to be reached by NSs, spanning from 1.93 to 2.01M . Note that in Bayesian analyses also observations of record masses should be treated with a likelihood- based procedure, which includes multiple observations and the error on their measurements, rather than strict bounds [254], as already done in Ref. [313]. (3) All NSs are described by the same EoS17.

2.5 Other compact objects possibly similar to neutron stars Up to this point, I have focused only on compact stars made of ordinary matter, usually referred to in general as NSs. In this subsection I will briefly introduce other possible exotic objects that could mimic NSs (and some of them also black holes, actually), to the point that current or near-future observations may not be able to make a distinction. Such exotic objects are of course interesting on their own and

16General limits imposed by causality have been known for a long time [307] and applied to the tidal deformability [308, 309] also before the GW170817 event. 17This last requirement was actually not imposed in the detection article for GW170817 [1], but only later, in analyses by the LIGO-Virgo Collaborations [220, 314] and by others (see Sect. 7.1).

12 not only in comparison to NSs. Research on these astrophysical objects may, for example, give hints on the of dark matter, help identify quantum effects that could halt collapse to black holes and thus prevent the formation of spacetime singularities, or contribute to solve the information problem in black-hole physics (loss of unitarity in Hawking evaporation). See Refs. [315, 86] for reviews on all these topics and more. Perhaps the most popular alternative proposal is that of boson stars [316, 317], which are compact, stationary configurations of a scalar field bound by gravity. There have been several proposals about what such a scalar field could be, like the [318, 319] or particles [320]. Boson stars could either be composed of stable fundamental bosonic particles bound by gravity, or of unstable particles for which a process inverse to their decay is enhanced by gravitational binding and becomes efficient enough to reach equilibrium, similarly to what happens with β-decay in NSs. Boson stars may have formed through during the primordial stages of the [321, 322, 323]. Boson stars made of are also called axion stars [324, 325]. While yet unobserved, axions are supported and motivated by theory, since they solve the strong CP problem of QCD [326], arise naturally in string theory compactifications (see, e.g., Ref. [327]), and are candidates for dark matter [328]. Another example of boson stars mentioned later in this review is Proca stars [329], which are the vector analogues of the scalar boson stars. Other hypothesized compact objects are (short for gravitational vacuum condensate star) [330], which are made of a thin shell of matter with radius very close to its . Inside it, a originated by the collapse of the original massive star would form a core described by a metric (de Sitter metric) that has a repulsive effect that stops the collapse, thus preventing the birth of both a singularity and an (see again Ref. [315] for a review and, e.g., Ref. [331] for a shorter summary of work on gravastars). Very recently, another type of compact objects that may mimic black holes and compact NSs has been proposed in Ref. [332] and named nonlocal stars. See Sect. 6 for current results on the possibility to distinguish observationally these exotic binaries from binaries composed of stars made of ordinary matter. Some comments on the possibility and the effects of accumulation of dark matter in NSs will be made in Sects. 4.1, 6 and 7.4.

2.6 Universal relations for neutron stars In works about the connection between GW observations from BNS systems and the EoS at supranuclear density, several empirical relations (found through numerical-relativity simulations) between physical quantities of NSs and observed quantities are used. Often the term employed to address them is universal relations. This subsection contains an introduction to such relations. Universal here means approximately independent of the EoS. There are two types of such relations: relations that connect different physical quantities of a NS isolated or in a (like radius, mass, moment of inertia, tidal deformability; see Sect. 3.1 for definitions) among themselves and relations that connect physical quantities of BNS systems (like the compactness C M/R of the component stars) with quantities that can be measured directly (like the main frequency≡ of the post- merger GW spectrum) in GW detectors. Relations of the former type often have interesting names, like the I-Love-Q relations [333, 334, 335, 336, 337], which connect the moment of inertia I, the Love number [which is related to the tidal deformability; see Eq. (7)] and the quadrupole moment Q of a NS. Such universality may originate from the fact that these relations depend most sensitively on the internal structure far from the core, where all realistic EoSs are rather similar. The universality is found to hold at the 1% level, except in exotic cases, like bare solid quark stars [338, 339, 340] (but even in this example they hold at the 20% level), but it may be affected by strong magnetic fields [341, 335] and large differential rotation [335]. Also, theories of gravity different from general relativity predict different relations18. 18This fact can actually be taken advantage of for testing general relativity versus other theories of gravity [333, 334,

13 Other relations of this type are the C-Love relations [333, 334, 348, 335, 349] between the compactness and the Love number of a NS. They hold at the few percent level for all EoSs on which they were tried, including those with phase transitions [350]. Still another is the I C (moment-of-inertia–compactness) relation [351, 352, 346, 353], which was actually the first to be found,− but possesses a lesser degree of EoS insensitiveness (accurate at the 10% level) [354, 335], unless improved (to the several percent level) with a different normalization [355]. More general relations between the lowest few multipole moments of NSs have also been found [336, 356, 357]. A relation, holding at the level of a few percent, between the total mass of the binary and the angular momentum in the remnant has been presented in Ref. [358]. Then, the binary Love relations, often used in the analysis of the data of GW170817, connect the tidal 19 deformabilities of the two stars in a BNS system ΛA and ΛB [361, 362, 363, 364, 365, 359, 360, 350]. They hold at about the 20% level for nucleonic EoSs, but it has been shown that, as intuition may suggest, they do not hold for EoSs with phase transitions when one star in the binary is a NS and the other a hybrid star [289, 366, 367, 350] (see also Sect. 7.3). Making use of combinations of the above-mentioned relations, R-Love (NS-radius–Love-number) relations have been proposed [350]. Belonging to the other type of universal relations is, for example, the relation between the tidal deformability and the frequency of the merger of a BNS system, defined as the instantaneous GW frequency at the time when the amplitude reaches its first peak. This relation was first found in Ref. [71] and later confirmed by more advanced and comprehensive works [368, 369, 370, 371, 372, 373]. It had been shown to hold at around the 1% level, but a work posted to arXiv.org just before this review was accepted for publication revised the relative error between data and (an improved) fitting formula to 3%, through resolution studies that included also a more varied set of binaries, in particular with more varied mass ratios [373]. Furthermore, the relation is found to hold only for equal-mass or very nearly equal-mass binaries [371, 373] and it has not been tested yet for magnetized and/or highly spinning binaries. A similar relation [71, 374, 373] was found between the tidal deformability and the GW amplitude at its first peak, which defines the merger of a BNS system, as said above. It has been shown to hold at the 4% level [373]. Another set of universal relations, with different degrees of reliability, has been found to connect the frequencies of the main peaks of the power spectral density of the post-merger GW signal with properties (radius at a fiducial mass, compactness, etc.) of a spherical star in equilibrium [375, 376, 377, 378, 379, 370, 369, 380, 120, 122, 381, 371, 382, 383, 373] (see Sect. 4). The spin of the NSs in the inspiral has been found to affect these relations [384, 369, 146, 385], that otherwise hold at the 10% level for EoSs without phase transitions. ≈ Other relations have been found between the threshold mass for prompt collapse, the maximum T mass for a nonrotating NS and its radius [386, 358, 387, 388] (see Sect. 4), between the quantity κ2 [defined in Eq. (14)], which parameterizes the late-inspiral of tidally interacting binaries, and the main peak of the post-merger GW spectrum [369] (again, see Sect. 4), and between the total gravitational radiation emitted in a merger event and the angular momentum of the remnant [160, 373]. Still other universal relations have been found between the frequencies of the fundamental modes (f-modes) of oscillation of stars and certain combinations of the stellar mass and radius [389, 390, 391, 392, 393, 394, 395, 396, 397]. The best of these relations hold at the 1% level [394, 395, 396]. Combining these with the I-Love relation, it is possible to find a universal relation between the f-mode oscillation frequency and the tidal deformability [395, 335, 398].

342, 343, 344, 345], if one can obtain for a given NS independent measurements of two of the quantities involved in the relations, e.g., the Love number through GW observations and the moment of inertia or the quadrupole moment through observations of binary pulsars [346, 347] or x-ray binaries [49, 62, 63, 64, 65]. 19Similar relations have been proposed that are valid for any two NSs (that have the same EoS), independently of whether they are in a binary system or not [359, 360].

14 As mentioned above, these universal relations are in fact not literally universal, since they have some (small) levels of dependence on the EoS. Hence, if, on one side, using these relations in data analysis allows to perform estimates impossible otherwise (see the rest of this review), on the other side they do include further uncertainties in the analyses. In some cases, the issue has been addressed by marginalizing over the EoS variability, for example when inferring the NS radii of GW170817 employing binary Love and C-Love relations [399]. Since in current detectors statistical uncertainties in parameter estimation are much larger than systematic errors added by using these universal relations, such a marginalization procedure is not a noticeable handicap, but, in order to be useful with the higher sensitivity of (near) future detectors and the large number of expected observations, more accurate universal relations will be necessary [350]. On the other hand, the observations themselves will impose more and more stringent constraints on the EoS, and the reduced allowed space of EoSs will allow to decrease the number of EoSs used for computing the uncertainties in the universal relations and thus decrease their variability [350]. For example, by imposing the 90% credible region constructed from the posterior probability distributions on the pressure-density plane [314, 247] obtained from GW170817, it has been found that the EoS insensitivity increases by a factor of 60% in the binary Love relations, by a factor of 70% in the C-Love relations, and by factors of 50%∼ in the I-Love-Q relations [350]. ∼ ∼

3 Extracting information on the equation of state from grav- itational waves emitted before the merger

The dynamics of BNS systems around merger depend on the EoS of ultrahigh-density matter. Essen- tially two methods to link the observed GWs to the NS EoS have been studied. One method uses tidal deformations during the last before merger [68, 69, 70, 400, 401, 140, 71, 334, 67, 402, 245, 403, 404, 405, 406, 407, 408, 409, 410]; tidal deformations have been measured for GW17081720 [1, 220, 314]. The other method uses the spectra of the gravitational radiation of the post-merger object (if it does not collapse to a black hole too soon) [375, 376, 377, 378, 379, 370, 369, 380, 371, 382]. More energy may be emitted in GWs in this phase than in the inspiral, but, because of their higher frequency, their signal-to-noise ratio in current and projected detectors is smaller than in the inspiral. In this Section I will focus on the first method.

3.1 The basic idea Stars in a binary system undergo tidal deformations that become larger as they get closer. These deformations affect the orbital trajectory of the binary and thus the emitted GWs [412, 413], encoding in the latter the NS EoS. Tidal deformations are described through the tidal deformability coefficient defined as the proportionality constant λ between the external tidal field ij (the field generated by the E companion star) and the quadrupole moment of the star Qij [68, 69, 70, 71, 334]:

Qij = λ ij . (5) − E However, tidal deformations are more usefully described through the dimensionless tidal deformability21

λ Λ , (6) ≡ M 5

20Note that one analysis suggested that the noise in the high-frequency region of the data from the Livingston LIGO interferometer may have corrupted information about the tidal deformability estimated for GW170817 [411]. 21Here and in the other equations below I use geometric units, in which G = c = 1, unless otherwise stated.

15 where M is the stellar mass. Equivalently, this can be written as

2  R 5 Λ κ2 , (7) ≡ 3 M where κ2 is the quadrupole Love number and R the stellar radius. This gives the quadrupole component of Λ, which can be calculated via the following expression [69, 414, 334] 16 Λ = (1 2C)2[2 + 2C(y(R) 1) y(R)] 15 − − − · n 2C[6 3y(R) + 3C(5y(R) 8)] + 4C3[13 11y(R) + C(3y(R) 2) + 2C2(1 + y(R))] + − − − − o−1 3(1 2C)2[2 y(R) + 2C(y(R) 1)] ln (1 2C) , (8) − − − − where C M/R is the stellar compactness and y(r) satisfies the differential equation ≡ dy 4(m + 4πr3p)2 6 y2 r + 4πr3(p ρ) 4πr2  ρ + p  = + − y 5ρ + 9p + , (9) dr r(r 2m)2 r 2m − r − r(r 2m) − r 2m (dp/dρ) − − − − where p and ρ represent pressure and mass density, respectively, and

1 e−γ(r) r m(r) − , (10) ≡ 2 where γ(r) is in the definition of the metric coefficients [415]

ds2 = eν(r)dt2 + eγ(r)dr2 + r2(dθ2 + sin2 θdϕ2) . (11) 0 − Eq. (9) can be solved for a given EoS together with the Tolman-Oppenheimer-Volkoff (TOV) equa- tions [416, 417], which describe spherically symmetric stars in static equilibrium in general relativity. Appropriate boundary conditions need to be imposed [69]. In general, the extraction of higher-order GW parameters (like tidal deformabilities) from the GW signal is difficult because these parameters can be efficiently extracted only in the late part of the inspiral, which may not be very long, and because there exist degeneracies between different higher- order parameters, like the individual NS spins and the tidal polarizability22. Moreover, because of the strong correlation between the two tidal deformabilities, ΛA and ΛB, associated with each star in a binary, it is challenging to extract them separately from the gravitational waveform, unless one assumes some empirical universal relations connecting ΛA and ΛB [364, 361, 362, 365, 350], which may add uncertainties to the estimates, as said in Sect. 2.6. Also, as mentioned earlier, it has been shown that such universal relations do not necessarily hold for EoSs with phase transitions [289, 367, 350]. What can be more easily directly measured from BNS waveforms is the dominant23 tidal parameter in the waveform, corresponding to the mass-weighted average tidal deformability (also called effective tidal deformability) given by [418]

4 4 ˜ 16 (mA + 12mB)mAΛA + (12mA + mB)mBΛB Λ = 5 (12) 13 (mA + mB)

22This is why the tidal deformabilities estimated by the LIGO-Virgo Collaborations were separated in a low-spin and a high-spin scenario [1, 220, 314]. 23 More in detail, the lowest-order post-Newtonian correction (see Sect. 3.2) of ΛA and ΛB can be written as Λ˜ and another parameter δΛ,˜ whose contribution, however, is very small (δΛ˜/Λ˜ . 0.01) and can be ignored, also because it cannot be extracted from data of current detectors [67, 402].

16 or, equivalently, 16 (1 + 12q)Λ + (12 + q)q4Λ Λ˜ = A B , (13) 13 (1 + q)5 where q mB/mA( 1) is the ratio of the masses of the two stars in the binary. ≡ ≤ Another useful related quantity that is part of some universal relations (see Sects. 2.6 and 4.1) is

A B T  1 κ2 1 κ2  κ2 2 4 5 5 + 5 5 , (14) ≡ q (1 + 1/q) CA q(1 + 1/q) CB

A,B where κ2 are the quadrupole Love numbers of the two stars of the binary, respectively, and CA,B are their compactnesses [369]. It has been shown that the dimensionless GW frequency depends on the T stellar EoS, binary mass and mass ratio, only through this tidal coupling constant κ2 [369]. An alternative way to gain information on the tidal deformability from detector data would be to extract the (mass-independent) coefficients of a Taylor expansion of the tidal deformabilities about some fiducial mass [419, 401, 361, 362]. This parameterization, however, can be efficiently applied only to systems whose NS masses are close to the fiducial mass, otherwise, the systematic error on the leading tidal coefficient due to mismodelling the tidal deformability can dominate the statistical one. This drawback, on the other hand, may be compensated and overcome by combining the information from multiple events (also with different masses), which can be done easily with this method. However, a sensitivity higher than that of current detectors would be necessary to accurately measure any of these coefficients. Since the Taylor expansion of the tidal deformabilities just mentioned has not been used yet in practice, I will focus on the use of the mass-weighted average tidal deformability. Reducing the tidal parameters to one (Λ)˜ allows for better statistical estimations, but of course some physical information about the two stars is lost. As mentioned in Sect. 2.6, a way around this problem was proposed in Ref. [361], that showed the existence of an EoS-insensitive relation (with variations of at most 20%) between symmetric and antisymmetric combinations of the tidal deformabilities

ΛA + ΛB ΛA ΛB Λs , Λa − . (15) ≡ 2 ≡ 2

These binary Love relations [e.g. Λa(Λs)] allow to compute the individual tidal deformabilities from the mass-weighted average tidal deformability. A simple Fisher analysis has shown that the binary Love relations improve parameter estimation of the individual tidal deformabilities by up to an order of magnitude with respect to estimations done by extracting ΛA and ΛB from the data directly [361, 362]. Modified binary Love relations for specific purposes have also been proposed [363, 359, 360]. As an important note on this introductory material, it may be useful to notice that the relationship between the tidal deformability Λ and the radius R cannot be simply derived from Eq. (7), namely 5 Λ is not necessarily proportional to R , because the quadrupole Love number κ2 also depends (and differently for each EoS) on the radius in a complicated manner, determined by differential equation (9) coupled to the Tolman-Oppenheimer-Volkov equations [416, 417]. A few empirical model-dependent estimations for such a relation [420, 421, 422, 30, 267, 364, 31] found exponents between 5.28 and 7.5. Even if the correlation between radii and tidal polarizabilities for different EoSs is tight, these two quantities provide complementary information, since for a given tidal polarizability, different EoSs may lead to somewhat different radii [31]. Also, it has been shown that these relations have errors of 5% or more for hybrid EoSs [367]. If, on one hand, the mass-weighted tidal deformability depends strongly on stellar radius, on the other hand it was recently found that, if the chirp mass24 is specified, the mass-weighted tidal deformability

24As mentioned earlier, the chirp mass can be extracted very accurately.

17 Figure 4: Mass-weighted average tidal deformability [Eq. (12)] of a binary system as a func- tion of the radius of the primary NS. The tidal deformability is calculated for various primary masses (corresponding to the different symbols) using several proposed EoSs (corresponding to the different clusters of radii). The mass of the secondary NS is computed assuming the chirp mass. The horizontal dotted line indicates the observed 90% confidence upper limit on the effective tidal deformability (from the original LIGO-Virgo analysis [1]). The narrow solid band (which is indistinguishable from a single curve) is some quasi-Newtonian expression for the binary tidal deformability [423] for 0.7 < q < 1.0. (From Ref. [423])

is approximately independent of the component masses for a BNS merger [423]. This empirical result was obtained by calculating the tidal deformabilities for several EoSs after choosing various values for one of the component masses in such a way that they lie within the mass range inferred for GW170817. The corresponding values for the mass of the other component star were calculated from the chirp mass measured for GW170817. The results of Ref. [423] are summarized in Fig. 4, where one sees, for example, that the upper limit25 of Λ˜ < 800 [1] immediately excludes radii above 13 km at the 90% ≈ confidence level, without requiring detailed knowledge of the component masses (m1 in Fig. 4). This idea of Ref. [423] came from using the I-Love-Q relations between stellar compactness and tidal deformability [333, 334, 335] (see Sect. 2.6) to express the mass-weighted tidal deformability of the binary as a function of component masses and stellar radii. Let me conclude this introductory Subsection by mentioning that, in addition to measurements based on the tidal deformability, other ways to gain information on the interiors of NSs from GW observations have been proposed, even if they require higher sensitivities and thus may be applicable only when third- generation detectors [234, 235, 233]) become operational. Some of these studies involve tidal excitations of resonant modes [424, 425, 426, 427, 428, 429, 430, 431, 432, 433, 434], gravitomagnetic excitations of resonant modes [435], resonant shattering of the NS crust by tides [436, 437] and non-linear tidal effects

25Note that the value Λ˜ < 800 was incorrectly reported in the detection article of GW170817 [1]: The corrected value in the case of the low-spin prior was Λ˜ 900, but the analysis of Ref. [423] and others made before the publication of the correction [220] could not but use≤ the mistaken value. See also Sect. 7.1. Ref. [423] and others also used the +0.004 original value given for the chirp mass Mchirp = 1.188−0.002M in Ref. [1], which was later revised in Ref. [220] to Mchirp = 1.186 0.001M . ± 18 [438].

3.2 Applications to data analysis One needs to treat carefully many aspects of the basic idea delineated in the previous Subsection when applying it concretely to GW data. Estimating the parameters of BNS systems during the inspiral phase is based on matched filtering: the GW data stream is cross-correlated with theoretically predicted template waveforms (approximants) for different possible physical parameters. These trial waveforms need to be accurate to allow for correct estimates of the stellar masses and spins, and of the internal structure of the stars. This is especially true for the very last orbits before the merger, where instead approximants become increasingly inaccurate (see, e.g., Refs. [400, 67, 402, 405, 408, 409, 439, 440, 441]). As it can be easily imagined, approximants that do not consider tidal effects are not sufficient, especially for spinning BNS systems or stiff EoSs. For example, it was estimated that for searches in Advanced LIGO at design sensitivity neglecting tidal effects would cause roughly a 5% additional loss of signals 2 [442]. Furthermore, for spinning systems (spin parameter χ J/M & 0.1, where J is the NS angular momentum and M its mass), in order to reduce mismatches≡ to an acceptable level, it is crucial to include spin-induced and EoS-dependent higher order terms [443, 444] in the waveform approximants [410, 445, 439, 440, 441, 446, 447]. Approximants including nonprecessing (namely aligned to the orbital angular momentum) stellar spins and self-spin effects have become available recently [439, 440, 441]. Figure 5 offers an illustration, based on computations, of which frequency ranges are the most important for extracting information about intrinsic binary parameters. Considering that the maxi- mum sensitivities of interferometric detectors are for frequencies . 1 kHz, one immediately sees that information about tidal parameters accumulates where the detectors are not at their best sensitivity. Another issue is the fact that the computation of trial waveforms needs to be efficient and fast, because source properties are generally inferred via a coherent Bayesian analysis that involves repeated cross-correlation of the measured GW strain with predicted waveforms [448]. Computational efficiency is crucial also because BNS systems are visible by GW detectors for a long time, several seconds or even minutes before the merger. Another development that would help data analysis would be producing approximants that include also the post-merger regime consistently, but no such model exists yet (see the end of Sect. 4.2 and Sect. 5 for related work). A few types of waveform models have become the standard in GW analysis: post-Newtonian, effective-one-body (EOB) [449, 450] and the Phenom models [451, 452]. Post-Newtonian expansions are approximate solutions, valid for weaker fields, of the Einstein field equations. The expansion is made in parameters that are small when the approximation is valid, like the velocity v of the objects (for the binary systems considered in this review it would be the relative velocity of binary constituents) with respect to the speed of light or deviations from a background metric. A post-Newtonian term of order n is proportional to v2n relative to the leading-order term in the expression (see Ref. [453] for an introduction). In post-Newtonian expansions, the GW signal can be approximated by imposing that the power radiated by a binary system in GWs is equal to the change in the energy of the binary. Even if physically tidal effects are present even in a Newtonian description, in the post-Newtonian framework, the lowest-order tidal effects appear as terms of the fifth post-Newtonian order. However, point-mass terms for binaries are currently only known to fourth order in the dynamics [454, 455] and only to order 3.5 in the GW phasing [453], and this lack of complete information has raised concerns about systematic errors in GW measurements of tidal effects [402, 67, 357]. In order to solve this problem, two classes of effective or phenomenological models for point-mass binaries that to some extent include all post-Newtonian orders in an approximate way have been devel- oped: the EOB model [449, 450] and the so-called Phenom models [451, 452, 456, 457]. Both approaches combine analytical results with information from numerical relativity simulations and, while address- ing the problems of the systematic errors of the post-Newtonian expansion, they may introduce other,

19 1.0 ) signal-to-noise ratio f symmetric mass ratio 0.8 chirp mass spin-spin parameter spin- parameter 0.6 tidal parameter

0.4

0.2

Normalized integrand per log(0.0 101 102 Frequency (Hz)

Figure 5: Illustration of where in frequency the information about intrinsic binary parameters predominantly comes from. The quantity shown on the y-axis is a normalized quantity characterizing the accumulation of information about the binary parameters (see Ref. [410] for details) per logarithmic frequency interval. The zero-detuned high-power configuration of Advanced LIGO is used and each curve is normalized to its maximum value. (From Ref. [410]) smaller, systematic errors (which can be assessed by comparison with numerical relativity simulations). The Phenom models are phenomenological waveforms that approximate a set of hybrid waveforms constructed by matching numerical-relativity waveforms with analytical post-Newtonian waveforms26 [451, 452, 456, 457, 406]. The fundamental idea of the effective-one-body method consists in representing the two-body dy- namics by those of a single effective particle in an effective potential. In practice, the dynamics and GWs from a binary are computed by solving the coupled system of ordinary differential equations for the orbital motion, GW generation, and radiation backreaction in the time-domain. Tidal effects have been fully incorporated in the effective-one-body model [458, 459, 460, 461, 462, 463, 430, 429], including for (nonprecessing) spinning binaries [439, 440, 441]. Refinements and calibrations of the models are performed by comparison with numerical-relativity simulations [464, 465, 466, 467], which suggest that further improvements of the tidal effective-one-body models are still necessary for a satisfactory descrip- tion of the signal [468, 382, 409, 445], at least in some regions of the parameter space of BNS systems. For example, a full Bayesian analysis determined that different tidal and point-particle/binary–black- hole descriptions for the waveform approximant yield estimated tidal parameters that can differ by more than a factor of two [409] (see also Sect. 3.3). Furthermore, the effective-one-body model still has a rather high computational cost per waveform [445]. Solutions to the latter problem may come from reduced-order-modelling techniques [469], which, however, also add further complexity, from additional inclusion of numerical-relativity results, or from other modelling techniques complementary to the effective-one-body model, like those in Refs. [470,

26In general, hybrid waveforms are constructed by matching numerical waveforms with some approximant of the general- relativistic equations.

20 471, 472, 473]. In the future numerical-relativity–based waveform models will likely be the ones that allow more precise and stringent extraction of the source properties [445].

3.3 Estimates about inspiral waveforms Starting from the work27 of Ref. [71], which quantified data-analysis estimates of the measurability of matter effects in gravitational waveforms from NS binaries with different EoSs (approximated as piecewise polytropes) by analysing numerical waveforms produced with codes of different groups, with different numerical setups, and combined with detector noise curves, it was made clear that it is actually possible to measure the tidal deformability (and the radius) of compact stars from BNS inspirals with current detectors. By using hybrid waveforms constructed as a match between the numerical waveforms at higher fre- quencies and some approximant waveform (see previous Section) at lower frequency, Ref. [71] found that for Advanced LIGO the radius of a NS can be estimated with an error of δR 0.5 km (Deff /100 Mpc), ' × or δR/R 5% (Deff /100 Mpc), where Deff is the effective distance to the source. Hybrid waveforms have been' later× noticeably improved by using better numerical-relativity simulations (higher resolutions, smaller initial orbital eccentricity28), better approximants (resummed post-Newtonian expressions [406], tidal effective one body [407]), either in the time domain [406] or directly in the frequency domain [407]. Through comparison with numerical-relativity simulations, these works obtained tidal corrections to the GW phase and amplitude that can be efficiently used in data analysis. The studies of Ref. [71] relied on the Fisher matrix approximation, which holds for loud signals. However, making strong statements about estimating source parameters requires a full Bayesian anal- ysis. With more sophisticated Bayesian statistical analyses and/or increasing amounts of (physical and detector) details taken into account, other works later confirmed that for NS binaries with individual masses around 1.4M , the dimensionless tidal deformability Λ could be realistically determined with about 10% accuracy by combining information from about 20 100 sources, depending on assumptions about the BNS population parameters (for example, if one considers− also nonzero spins for the initial NSs the necessary number of sources is higher) [401, 140, 402, 67, 245, 403, 405]. In particular, the first Bayesian study for the estimation of tidal deformability parameters was carried out in Ref. [401]. Using post-Newtonian approximants that included tidal corrections up to the sixth post-Newtonian order, they concluded that tens of detections are required to constrain the tidal deformability parameter to an accuracy of 10%. If one expresses the tidal deformability as a linear expansion in mass, this would allow to distinguish≈ between different types of EoSs at that accuracy level; however a caveat is necessary here, since EoSs with phase transitions do not necessarily allow for such a linear expansion (see Sect. 7.3 for further discussion). Following Ref. [401], an analytical approach to study the systematic and statistical uncertainties arising from neglecting physical effects in the estimation of the Love number of NSs (and so of their tidal deformability) was made by Refs. [67, 402]. It was found that relevant estimation biases that exceed statistical ones are introduced (i) if post-Newtonian terms of order 4 and higher are neglected, (ii) if even relatively small spins (χ 0.003) are not taken into account, and (iii) if eccentricities larger than about 10−3 are neglected. Other≥ biases may originate from inaccurate numerical simulations if they are used to tune the approximants, even if in the most advanced numerical-relativity codes these are relatively under control [486, 374, 373].

27Previous works [68, 474, 460] had already suggested that information on EoSs from the inspiral waveforms could be obtained with future detectors, but these were based solely on theoretical considerations and did not employ numerical- relativity simulations and a detailed treatment of detector noise. 28In BNS systems born bound, GW emission reduces the eccentricity [475, 476, 477] to smaller values, but compact binaries formed through dynamical scattering and dynamical capture in dense stellar environments could have non- negligible eccentricity [478, 479, 480, 481, 482, 483, 484] in the LIGO band (see, e.g., Ref. [477, 485]).

21 In further developments, Ref. [245], provided a method to estimate the EoS parameters for piecewise polytropes by stacking tidal-deformability measurements from multiple detections and concluded, in accordance with previous estimates, that a few bright sources would allow to constrain the NS EoS. Ref. [404], however, showed that stacking multiple detections with moderately low signal-to-noise ratio should be carried out with caution as the procedure may fail when the prior information dominates over new information from the data. Ref. [403] then found similar results after extending the previous analysis [245] by including a larger number of simulated BNS signals and taking into account more physical ingredients, such as spins, the quadrupole-monopole interaction (this did not affect parameter estimation for the considered configurations), and tidal effects to the highest known order. In addition, the tidal deformability parameter was expanded to include up to a quadratic function of mass. Using Bayesian model selection analysis, Ref. [404] found that detectors like Advanced LIGO and Virgo can heavily constrain EoSs containing only quark matter (with a signal-to-noise ratio of 20), but hybrid stars (see Section 7.3) would be more difficult to distinguish from hadronic stars, because≈ they differ only at higher densities, that contribute less to the tidal deformability. They considered , hyperon, and hybrid EoSs with exotic matter parameters within the range allowed from experiments and theoretical calculations. They concluded that the presence of kaon and hyperons cannot be easily confirmed (but easily excluded). Many of the works mentioned above (which date to before the detection event GW170817) were carried out by using piecewise-polytropic parameterizations of the EoS [103]. As already mentioned in Sect. 2.4, however, it had been shown in Ref. [245] that near the fixed joining densities of the piecewise- polytropic representation systematic error arises that may exceed statistical error. Refs. [220, 247] extended the work of Ref. [245] by using also a spectral EoS parameterization29 [256] (see Section 2.4), which does not have such problems, for analysing the data of GW170817 (for more on estimates from GW170817 see Sect. 7). In particular, in Ref. [247] it was found that both the piecewise-polytropic and spectral EoS parameterizations allow to recover consistent tidal information from the simulated signals, but, as expected, the spectral model allows for smaller errors. In Ref. [220], in addition to the adoption of the spectral representation of the EoSs, by probing different mass-ratios for non-spinning signals with a tidal effective-one-body–based model [429], it was concluded in an exhaustive Bayesian analysis that for GW170817 the systematic uncertainties due to the modelling of matter effects are smaller than the statistical errors in the measurement. Also through Bayesian inference, Ref. [408] investigated numerical-relativity–based tidal waveform models and showed the importance of the inclusion of tidal effects for the extraction of the NS masses and spins from the GW signal for high signal-to-noise ratios. They found that inaccurate modelling of tidal effects in the analysis of the inspiral GW signal for stiff EoSs leads to a large bias in the measurement of the masses and spins. They also studied whether omitting the post-merger waveforms from the global analysis of the signal leads to significant loss of information, or possibly to biases in the estimation of the source properties, and concluded that, as expected but not rigorously shown before, the post- merger part has no impact on GW measurements with current detectors for the signal-to-noise ratios considered. Ref. [408] stated that systematic errors in the analysis are under control, but Ref. [409] pointed out that in some cases there may be potential systematic biases during the extraction of parameters, from non-spinning sources as well. By experimenting with many different approximants, they found that for unequal-mass binaries30, while the mass and spin recovery shows almost no systematic bias with respect to the chosen waveform model, the extracted tidal effects can be significantly biased, up to a point where the injected value is not contained within the 90% credible interval. In particular,

29However, Refs. [220, 247] have not adopted the latest developments in such a spectral representation, namely the self-imposition of the causality constraint [241, 259]. 30For equal-mass binaries sufficiently similar predictions with any modelling were found.

22 they concluded that, generally, post-Newtonian approximants predict NSs with larger deformability and radii than models tuned by using numerical-relativity simulations. Furthermore, it was highlighted that the use of higher post-Newtonian orders in the tidal phasing does not lead to a monotonic change in the estimated properties [409]. Their remarkable note of caution states that for a signal with strength similar to GW170817, but observed with design sensitivity by the Advanced LIGO and Virgo detectors, different tidal descriptions of the waveform approximant yield estimated tidal parameters that can differ by more than a factor of two (see also Ref. [410]). This could lead to misclassification of the observed system (NS binaries, black-holes binaries, or NS–black-hole binaries). Thus, further improved waveform models with improved tidal descriptions are imperative for characterizing unequal-mass mergers. Other notes of caution have been rung, among others, by Ref. [487], which found that the posterior distribution of the tidal deformability is strongly affected by the choice of the maximum frequency considered in the analysis, and by Ref. [240], which found that the choice of parameterization of the EoS (see Sect. 2.4) can have a significant effect on posterior EoS constraints inferred from GW data. Also, the effect of precession in the late inspiral has been found, through numerical simulations, to be visible in the GW signal [146], even if it may be of secondary importance in practical cases, since it is appreciable only for edge-on systems, which are harder to detect because of the smaller observable GW strain for such inclinations. More comprehensive Bayesian analyses have been performed that include not only gravitational emission from GW170817, but also the related electromagnetic emission [488, 387, 489, 490, 491, 492, 493], but these are not treated in detail in this review (see also Sect. 7). In the previous paragraphs, I reported the main advances in estimations from the inspiral waveforms, especially as far as statistical treatments (Bayesian techniques) are concerned. In addition to these, the work of Ref. [71], described at the beginning of this Subsection, has been improved in other ways as well. Ref. [405] has quantitatively improved its computations and estimations in two principal directions. First, they employed new numerical-relativity simulations of irrotational binaries with longer inspirals (i.e. 14 16 orbits) and higher accuracy both in the initial-data setup (i.e. residual eccentricity of the order− of 10−3). Second, they included in the analysis lower frequencies, down to 30 Hz , to which ground-based detectors like Advanced LIGO and Virgo are reasonably sensitive. They also adopted additional EoSs. Results were very similar to those of Ref. [71], namely that deformability Λ and radius can be determined to about 10% accuracy for sources at 200 Mpc. More recent works [406, 407] noticeably improved procedures for efficient data analysis of the pre- merger signal in relation to tidal deformations. In particular, Ref. [407] found that the statistical error for the measurement of the mass-weighted tidal deformability is more than six times larger than the systematic error for a signal-to-noise ratio of 50. They also showed that the statistical error for the measurement of the mass-weighted tidal deformability is larger than the variation of the mass-weighted tidal deformability with respect to the mass ratio even for signal-to-noise ratio 100 (see also the last paragraphs of Sect. 3.1 and Fig. 4). This suggests that even for events with signal-to-noise ratio 100, the systematic error in current waveform models is unlikely to cause serious problems in the parameter estimation [407].

3.4 Eccentric mergers As briefly mentioned in Sect. 3.3, BNS systems that have evolved without close interaction with other stars are thought to have lost, by the time their GW signals enter the detectability range, any significant initial eccentricity they may have had through the emission of gravitational radiation [475, 476] and −3 therefore to have very small orbital eccentricities, i.e. . 10 [477]. Instead, compact binaries formed in dense stellar environments (globular clusters and galactic nuclei) through dynamical scattering, dy- namical capture and, in general, multibody interactions could still have non-negligible eccentricity by the time they merge [478, 479, 480, 481, 482, 483, 484]. These events are probably rarer, even though

23 the estimates of their event rates are very uncertain (see, e.g., Refs. [481, 437, 484]), and more difficult to detect with detectors like Advanced LIGO and Virgo, mostly because the signal power at frequencies around 100Hz, where the detectors perform best, is smaller (see, e.g., Refs. [485, 494, 495]). ∼ However, if detected in GWs, they would provide, possibly more easily, measurements of several stellar physical quantities, from which then information on the EoS may be gained. This comes from the fact that tidal perturbations during pericentre passage can excite stellar fundamental modes (the previously mentioned f-modes) of oscillation that have a time-varying quadrupole moment and that therefore act as sources of gravitational radiation, on top of the inspiral waveform31. This f-mode signal depends on the EoS (in general, stiff EoSs store more energy in the oscillations compared to soft EoSs [153]) and can significantly affect the phase of the gravitational radiation by enhancing the loss of orbital energy by up to tens of percent over that radiated away by GWs during an orbit. Part of the orbital angular momentum may also be transferred to the stars. Measurements of the frequency, damping time, and amplitude of the tidally excited f-modes could yield simultaneous measurements of their masses, moments of inertia, and tidal Love numbers and thus present a prime opportunity to test the I-Love-Q relations (see Sect. 2.6) observationally. Mergers of eccentric BNS systems may also produce brighter electromagnetic emission than quasicircular mergers (see, e.g., Refs. [150, 151, 111, 144, 106, 155]) and contribute to the overall r-process [497, 184, 72, 73, 498, 74] element abundances [150, 202, 106, 155]. Actually, if accurate enough templates for eccentric BNS systems are used, these binaries can be de- tected from farther away, and parameters such as the chirp mass and sky localization can be estimated more accurately [494, 495], mostly because of the their richer structure that breaks parameter degenera- cies. Detection of a few highly eccentric BNS mergers per year might be possible with third-generation detectors [153, 155] or even with the LIGO-Virgo-KAGRA detector network at design sensitivity [495]. Several numerical studies have been performed on eccentric binaries [499, 500, 150, 151, 111, 144, 106, 152, 501, 153, 154, 155]. Ref. [154] proposed a basic model that, using a post-Newtonian description, can predict the timing of different pericentre passages within a factor of two with respect to results of numerical simulations. The error is probably dominated by systematic effects in measuring orbital properties from the numerical-relativity results [154]. A refined version of this model (that takes into account post-Newtonian corrections to the tidal coupling and the oscillations of the stars) may serve as a template for detection and analysis of gravitational radiation from eccentric systems [154]. Extensive studies on the subject have found that, on one side, the same qualitative relation between the merger frequency and the stiffness of the EoS that is known for quasicircular binaries is valid for eccentric binaries [152, 153], while, on the other side, there is no direct correlation between initial eccentricity and post-merger frequencies [153] (see Sect. 4), and post-merger peak frequencies do not follow the approximate universal relations with stellar properties like compactness [155]. Actually for eccentric orbits the position of the main peak in the power spectrum has been found to vary with the eccentricity of the orbit, as a result of the different angular momentum of the merger remnant [155].

4 Extracting information on the equation of state from grav- itational waves emitted after the merger

Also GW signals from the merger and post-merger phases of GW170817 have of course been searched in the data from Advanced LIGO and Advanced Virgo, but no signal was found [220, 502]. In fact, the strain upper limits set by the detector were found to be about one order of magnitude above the numerical-relativity expectations for post-merger emission from a hypermassive NS at the distance of GW170817.

31The excitation of f-modes in low-eccentricity inspirals can also be measured, but only with a network of third- generation detectors [496].

24 The first detection of a BNS post-merger signal is still to come, but simulations and investigation on their possible connections with observations of GWs from the post-merger phase of BNS mergers has been very active. The strong interest springs also from the fact that such observations would probe densities higher (several times the nuclear density) than typical densities in inspiralling stars and would also probe effects of temperature, which becomes much higher (up to 50MeV) after the merger. However, even though the energy emitted in GWs after the merger may∼ be higher than that emitted during the inspiral (if there is not a prompt collapse), since post-merger GW frequencies are higher (from 1 to several kHz [375, 379, 383, 205]), their signal-to-noise ratio in current and projected detectors is smaller than that of the inspiral and they are probably only marginally measurable by detectors like Advanced LIGO or Advanced Virgo. Third-generation detectors, such as Einstein Telescope [234, 235] and Cosmic Explorer [233], are needed. Also on the theoretical side, the post-merger phase presents more difficulties. Numerical simulations of the merger and post-merger dynamics are more difficult than for the inspiral part, because of strong shocks, turbulence, large magnetic fields, various physical instabilities, neutrino cooling, viscosity and other microphysical effects32. Therefore, their accuracy is not as good as for the inspiral. For example, currently there exist no reliable determinations of the phase of post-merger gravitational radiation, but only of its spectrum. Furthermore, the complicated morphology of post-merger signals makes constructing accurate templates challenging. The basic idea for obtaining information on the supranuclear EoS from GWs emitted after the merger is that the main peak frequencies of the post-merger power spectrum (see Fig. 6) strongly correlate with properties (radius at a fiducial mass, compactness, etc.) of a zero-temperature spherical equilibrium star in a rather EoS-insensitive way. These are also called universal relations, as illustrated in Sect 2.6, even though they depend on the spin of the stars in the inspiral and hold only approximately even for irrotational binaries. These relations have been employed to estimate constraints on the NS radius from the post-merger signal of future observations [507, 226, 228]. Some details of the post-merger spectrum are still debated, but there is widespread consensus that (i) the post-merger GW signal possesses spectral features that are robust (present in all simulations, irrespective of the numerical methods, codes used and numerical settings) and that emerge irrespective of the EoS or the mass ratio of the original binary; (ii) for the frequencies of the most salient peaks, analytic fitting functions can be obtained in terms of the stellar tidal deformability or compactness.

4.1 Basic ideas A first attempt to study systematically the post-merger waveforms was put forward by Ref. [213], that tried to codify the whole spectrum, instead of singling out individual peaks. They decomposed the merger and post-merger GW amplitude into three parts: (i) a peak in frequency and amplitude that appears soon after the merger starts; (ii) a decrease in amplitude during the merger and a new increase when the compact star forms; (iii) a final decrease in the amplitude during the hypermassive NS phase, which is either monotonical or with modulations. They also identified a damping of the oscillations of the frequency during the phase (spanning several oscillation periods) in which the merged object is a compact star; the frequency eventually settles to an approximately constant value (although a long-

32It has been hinted at that viscosity and neutrino cooling are probably subdominant with respect to the emission of gravitational radiation in the dynamics of the post-merger phase, at least for the first 20 ms after merger [159, 503, 160, 504, 505], but this is not true in all cases. In particular, it was found that thermal conduction and shear viscosity are not subdominant if both neutrino trapping and short-distance gradients (for example due to turbulence on a scale of 10 m) are present [503]. Moreover, a more recent analysis, assuming that nuclear matter remains neutrino transparent∼ up to temperatures of a few MeV, found that bulk viscosity can damp oscillations on timescales of 10 ms, comparable to those of a BNS merger, for nuclear matter at temperatures of 2 4MeV and for densities between∼ 0.5 2 times the saturation density [506]. This means that bulk viscous damping should− probably be considered seriously for− inclusion in future simulations.

25 Figure 6: Time-frequency analysis for the waveform of an equal-mass BNS merger for an optimally-oriented source at 50Mpc. The top and right panels show the time-domain waveform-component h+ and its Fourier magnitude spectrum, respectively. The time- frequency map is constructed from the magnitudes of the coefficients of a continuous wavelet transform (see Ref. [224] for details). Horizontal straight lines emphasize the locations of the main peak frequency fmain peak and a secondary peak which the authors of Ref. [224] call fspiral (a peak with similar frequency and amplitude has been called f1 in other works; see main text for discussion). The vertical lines have no meaning out of the context of the original article. (From Ref. [224])

term secular change associated with the change of the state of the hypermassive NS is always present). Based on this, they found an optimal 13-parameter fitting function, using which it may be possible to constrain the NS radius with errors of about 1 km. The first to propose relations between single peak frequencies and stellar properties (mass, radius, compactness) were Refs. [375, 376, 377, 378], using simulations with approximate treatment of general relativity. Subsequent analyses were performed by a few groups with general-relativistic codes [379, 370, 369, 380, 120, 381, 371, 508, 382, 509, 383, 373], which confirmed that the conformally flat approximation employed by other authors provided a rather accurate estimate of the frequencies of the largest peak in the power spectral densities. In more detail, Refs. [375, 376] pointed out that the largest peak in the power spectral density (whose frequency is dubbed there as fpeak and as f2 in other works, while in this review I will call it fmain peak for the sake of the reader) correlates with the radius Rmax of the maximum-mass nonrotating star for a given EoS and with other similar quantities, like the radius of a 1.6M NS with a given EoS. It was found later that the scatter in the relations of fmain peak with these stellar properties was such that it does not allow to determine the radius with an accuracy below 1km, even if fmain peak were measured exactly [213, 373]. It was also found later that the relation between fmain peak and the dimensionless tidal deformability is more universal and holds at the 10% level, but only for equal-mass or very nearly equal-mass binaries [371, 373]. ≈ Refs. [379, 370] then presented another method based on results form a large number of numerical- relativity simulations of equal- and unequal-mass irrotational binaries with different EoSs and different masses. They identified two distinct and robust main spectral features in the post-merger power spectral density: the largest peak fmain peak mentioned above and a smaller peak at lower frequencies (see Fig. 6), dubbed f1.

26 Other works then pointed out that what was called f1 in Refs. [379, 370] may be composed of two peaks, either or both present depending on the total mass and EoS [510, 384, 371, 383, 511, 373]. One of these subdominant lower-frequency peaks is thought to originate from the quasi-linear interaction between the dominant quadrupolar oscillation and the quasi-radial mode of the remnant [512] (even if the quasi-radial mode by itself may not be visible in the power spectrum). Its frequency, named f2−0, is equal to the difference of the frequencies of these two modes. The f2−0 feature is particularly pronounced for relatively high total masses of the binary system and soft EoSs, while it is often not visible for less compact stars in the binary, either low-mass stars or stars described with a stiff EoS [510, 383, 511, 373]. The other peak, called fspiral, is thought to be produced by the orbital motion of antipodal bulges that form during the merger and persist for a few milliseconds [510] (see Fig. 6). Different theoretical toy models were proposed to explain the origin of the f1 or fspiral peak [370, 510], but one thing on which everyone agrees is that it originates (or they originate) only from the dynamics immediately following the merger (within 3 or 4 ms after the merger). Also, in practical uses of the relations involving f1 or fspiral, the different origin of these subdominant features or whether the lower frequency post-merger peak has composite substructure have not been taken into account. Ref. [228], for example, defines what they call fsub as the frequency of the second highest peak with a frequency at least 400 Hz below the main peak. Coming back to the findings of Refs. [379, 370], it was shown that, at least for the sample of EoSs tested, a single function relates the f1 frequency to the average compactness M/¯ R¯ (where M¯ (MA + ≡ MB)/2 and R¯ (RA+RB)/2, where MA,B and RA,B are the masses and the radii of the nonrotating stars associated to those≡ of the binary). Knowing the masses from measurements of the inspiral waveform and the f1 frequency from measurements of the post-merger waveform would then allow to compute the radius of the NSs. Also, since the f1 or fspiral peak is produced soon after the merger, it should not be affected significantly by magnetic fields and radiative effects, whose modifications emerge on much larger timescales. However, a work (already mentioned in Sect. 2.6) posted to arXiv.org just before this review was accepted for publication pointed out that the existence and accurate location of this peak depends significantly on the numerical resolution of the simulations [373]. Also, it has to be noted that, especially in the first milliseconds after the merger, when also the subdominant modes are active and when the merger remnant is rapidly evolving towards a more stable configuration, post-merger frequencies evolve in time (see, e.g., spectrograms in Refs. [371, 382, 383] and Fig. 6), albeit only slightly. In particular, the frequency of the main peak increases up to the formation of the black hole, while its amplitude decreases, as physically expected from the increase of rotational velocity and compactness of the merged object as it loses angular momentum [382]. Hence, the spectral properties of the GW signal can only be asserted reliably when the signal-to-noise ratio is sufficiently strong so that even these changes in time can be measured in the evolution of the power spectral densities. In light of these considerations, as mentioned earlier, the prospects for high-frequency searches for the post-merger signal are limited to rare nearby events. An interesting extension of the works described above has been done in Ref. [369]. They found T that the coupling constant κ2 [defined in Eq. (14)] efficiently parametrizes the late-inspiral of tidally interacting binaries and observed that it can also be used to determine the main features of the post- T merger GW spectrum, instead of the tidal deformability parameter Λ. The relation fmain peak(κ2 ) depends very weakly on the total mass of the binary, mass-ratio, and EoS. However, there is dependence on stellar spins. The proposed physical explanation is that at fixed separation, the tidal interaction T T is more attractive for larger values of κ2 . Binaries with larger κ2 merge at lower frequencies (larger T separations). As a consequence, the remnants of binaries with larger κ2 are less bound and have larger angular-momentum support at formation. The fmain peak frequency seems mainly determined by these initial conditions, other physical effects having negligible influence on the value of the frequency. Dependence of the main peak frequency fmain peak itself on the initial state of rotation, especially for very rapidly rotating NSs, has been pointed out in several works [140, 384, 369, 382, 146, 385]. In

27 particular, Ref. [146] emphasized that their spectra are qualitatively similar, but quantitatively different, from the non-spinning cases. More in detail, the frequencies of all peaks are about 200 Hz higher and the f1 peak is significantly more pronounced for systems in which the stellar spins are aligned with the orbital angular momentum and less pronounced for anti-aligned system. This means that actually the quasi-universal relations found for irrotational binaries [375, 376, 377, 378, 379, 370, 369] might contain systematic biases for spinning systems. Furthermore, Ref. [385] found that the secondary peaks are less significant in mergers of NSs with higher spin magnitude. Additional checks on the validity of the relations discussed above and estimations of their error were done independently through simulations with codes different from the ones used in the first place to obtain the fitting parameters in those relations [383, 373]. Furthermore, it was found in the same works that the post-merger frequencies of unequal-mass binaries differ at most of 10% from those of equal-mass binaries, as long as the differences in the masses of the stars in the binary≈ are within 10% [122, 371, 382, 383, 373]. For lower mass ratios the spectra become more complicated: for fixed EoS and total mass, spectra for smaller mass ratios show less power in the main peak and more peaks at frequencies below it. Incidentally, there is actually a third, higher-frequency peak, named f3, often identifiable in com- puted spectra. It may be explained, together with f2−0 and perhaps f1 (as mentioned above), as the result of the combinations of the quasiradial oscillation mode and the fundamental mode fmain peak [512]. In fact, in most cases, the lower- and higher-frequency subdominant peaks are almost equidistant from the dominant one and the frequency predicted for the mode combination is a very good approximation for these subdominant peaks. As said above, however, this is not observed in binaries composed of less compact stars [510, 383, 511]. Also, the dependency of f1 and f3 on the binary mass ratio seems to be similar (and much stronger than that of the fmain peak mode): The more unequal a system is, the lower both frequencies seem to be. In any case, the f3 peak is too weak and at too high frequencies to foreseeably help in parameter estimation from GW observations. Another peak in the spectrum may be visible in some cases because of an m = 1 (see note on page 7) deformation of the merged object [199, 200, 176, 140, 141, 201, 202, 144, 203, 204, 152, 205], especially in mergers of eccentric binaries [152]. This deformation, due to the so-called one-armed spiral instability, is found to be present generically in BNS simulations and to carry information about the EoS33, but its peak in the spectrum, located at about half the frequency of fmain peak, has a much smaller amplitude and the prospects of observations in GWs appear unlikely in current detectors [203]. Third-generation detectors may be able to target this signals. Talking about additional peaks, it has been proposed that the presence of dark-matter cores inside NSs may produce one or two supplementary peaks in the post-merger GW spectrum of NS mergers, with frequencies between fmain peak and f3 [514]. This result was obtained by adding the effects of dark matter (in amounts up to 10% of the stellar mass34) into the mechanical model proposed in Ref. [370], and mentioned above, that captures the essential features of the dynamics of the hyper-massive NS formed in the first instants after the merger. Again, because of their higher frequency and lower power, it will be difficult to distinguish such peaks observationally in the near future. See Sects. 6 and 7.4 for other results involving the hypothetical presence of dark matter in NSs. Still further additional peaks may appear if the lifetime of the merged stellar object before collapse is long enough that convective excitation of inertial modes occurs, as pointed out in Ref. [518]. These are related to the thermal properties of the EoS but a systematic study of these modes and of how they are related to the properties of the stellar EoS is still missing. A different way to gain knowledge on the mass-radius relation of NSs was proposed in Refs. [386, 358, 387]. It consists in using some empirical relations obtained from numerical simulations of BNS

33See Sect. 6 for a possible relation of the m = 1 mode to dark matter [513]. 34Note that dark matter cannot condensate inside stars in quantities as large as this, therefore non-standard properties or mechanism involving dark matter have to be thought [515, 516, 517].

28 mergers between threshold mass for prompt collapse, maximum mass for a nonrotating NS, and its radius [386, 358]. Considering that observations of the electromagnetic emission associated with GW170817 may imply (but see Ref. [519] for alternative interpretations) that GW170817 was not a prompt collapse35 [130, 387] and that therefore the total mass of GW170817 is a lower bound on the threshold mass described above, an estimate was found for the maximum mass for a nonrotating NS and its radius [387] (see Table 1 for results on radii). The authors remark that the constraints they set are particularly robust because they only require a measurement of the chirp mass and a distinction between prompt and delayed collapse of the merger remnant. Ref. [388] later used the same arguments to update the obtained results in a fully general-relativistic framework.

4.2 Applications to data analysis For data-analysis standards and in comparison to the inspiral, numerical simulations of the post-merger phase are still sparse and not accurate enough (as explained above) and this contributes to the lack of analytical, physically parameterized waveform templates. This reduces the feasibility of matched- filtering. Generic analyses that target signals of unknown morphology might be less efficient than matched-filtering, but they have been shown to be able to extract the main features of post-merger signals, such as its main frequency components [223, 224, 520, 521, 226, 522, 372]. Below, more details on these works are given. Ref. [223] was the first to carry out a systematic study of the detectability of the high-frequency content of the merger and post-merger parts of BNS post-merger GW signals with methodologies used to search for unmodelled GW transients. Using a morphology-independent algorithm based on constrained likelihood statistics that identifies whether a signal is significant with respect to the noise, they focused on the problem of discriminating among different post-merger scenarios (prompt collapse or not) and on measuring the dominant oscillation frequency in the post-merger signal, in case of non-prompt collapse (see Sect. 4.3 for a summary of their results too). With the goal of reducing the complexity of the problem from a high-dimensional physical parameter space, where the waveforms are modelled directly through numerical simulations, to a lower-dimensional problem that includes only the dominant features of the waveform, the same group then proposed a low-dimensionality frequency-domain template based on principal component analysis [224]. They constructed a catalogue of numerical waveforms and decomposed the magnitude and phase spectra into orthogonal bases, as per the principal component analysis. By excluding each element of the catalogue one at a time and checking how well just the first principal component from the principal component analysis could reproduce that element, a match of 0.93 was found. The usually accepted desired threshold is 0.97, but Ref. [224] suggested that the first principal component is robust enough to model the high-frequency GW spectrum for BNS mergers for data-analysis purposes. Ref. [226], then, also proposed a method that relies less on numerical-relativity simulations: recon- structing the post-merger GW signal in observed data through a sum of wavelets without assumptions on the morphology of the system. The algorithm they use is BAYESWAVE [523, 524] and it was shown to be capable of reconstructing the dominant features of the injected signal with an overlap of above 90% for post-merger signal-to-noise ratios above 5. This would provide a measurement of the ≈ dominant post-merger frequency fmain peak to within a few tens of Hz. Additional information about the signal, such as its broadband structure or the finite extent of the post-merger peak, could increase the sensitivity of the analysis, making it easier to detect and characterize the post-merger signal. After the reconstruction an empirical relation based on numerical-relativity calculations to relate radius and fmain peak was used [226]. The final results (namely that the radius can be measured to a few hundred metres for signals with a signal-to-noise ratio 5) are comparable to the methods that are based on ≈ 35In general, of course, GW observations as well can inform us about the fate of the merged object (prompt or delayed collapse, or maybe no collapse).

29 a description of the post-merger power spectral density obtained from numerical-relativity simulations [375, 376, 377, 378, 379, 370, 369, 373]. It is to be noted that their error on the radius is dominated by the scatter in the universal relation used, rather than the statistical error of the reconstruction itself [226]. Very recently, Ref. [372] has done a similar work based on a Bayesian analysis that employs simplified Lorentzian model functions and found that the main emission frequency of the post-merger remnant, for signal-to-noise ratios of 8 and above, can be extracted within a 1σ uncertainty of about 100Hz for Advanced LIGO and Advanced Virgo at design sensitivities. The methodology described in Ref. [226] was applied to the data of GW170817 and results were reported by the LIGO-Virgo Collaborations [220, 502]. That is how it was found that upper limits on energy are at least an order of magnitude larger than expectations based on simulations performed with the EoSs they chose [502]. Recently, the first draft suggestion of a way of generating reliable post-merger spectra rapidly enough to be useful for templated data analysis was proposed in Ref. [522]. This is done through a hierarchical model built to represent the amplitude spectra and trained on the numerical-relativity simulations of Ref. [371]. They report a mean of 0.95 for the noise-weighted fitting factors across all tested spectra, for sources simulated at a distance of 50 Mpc. However, they do not test their results in a Bayesian framework to find its actual efficiency in parameter estimation of detected events. Actually, in order to be useful in search and parameter-estimation studies, the model of Ref. [522] should be extended to include more numerical simulations for training.

4.3 Estimates about future observations Realistic estimates of what information can be obtained with the methods described in the previous subsections require Bayesian analyses, like those performed in Refs. [222, 223, 224, 225, 226, 228]. As mentioned at the beginning, GW measurements at the expected frequencies and amplitudes for f1 and fmain peak are very difficult; in practice they are limited to sources within 30 Mpc for second generation detectors, as shown by Refs. [223, 224, 225] via large-scale Monte Carlo∼ studies. As mentioned above, Refs. [223, 224] provided the first systematic studies of the detectability of the high-frequency content of the merger and post-merger parts of GW signals from BNS systems. These investigations did not rely on waveform models and optimal filtering, but exploited methodologies used to search for unmodelled GW transients. They concluded that second generation detectors could measure post-merger signals and constrain the NS EoS for sources up to a distance of 10 25 Mpc (assuming optimal orientation). Ref. [224] also showed that the error in the estimate of the− NS radius would be of the order of 400 m in Advanced LIGO. Ref. [225] proposed methods that stack the post-merger signal from multiple BNS observations to boost the post-merger detection probability. They found that, after one year of operation of Cosmic Explorer [233], the dominant-peak frequency can be measured to a statistical error of 4 20 Hz for certain EoSs, corresponding to a radius measurement to within 15 56 m, a fractional≈ − error of about 4%. They showed that errors in the universal relations between≈ post-merger− oscillation frequency and total mass of the binary, and in the template construction dominate over the statistical error. Detectability of individual events could potentially improve if one considers all components/peaks that arise in the post-merger waveform, and not only the dominant peak. As said in Sect. 4.2, waveform templates that span the pre- and post-merger signals are currently not available, but observations of the pre-merger signal from GW170817 has been used to inform expectations about the properties of the undetected post-merger signal. Ref. [228] used the posterior samples for the masses, radii and tidal parameters of the inspiral of GW170817 to estimate its expected post-merger signal to be approximately 2.5 kHz fmain peak 4 kHz. ≤ ≤ It was also pointed out that, the full LIGO-Virgo network operating at design sensitivity may provide reasonable estimates of the dominant post-merger oscillation frequency and corresponding constraints on

30 the NS EoS [220] and this will be certainly measured if the sensitivity in the kHz regime of the detectors are improved by a factor 2 or 3 over their current design sensitivity [228]. With further improvements and next-generation detectors, it is reasonable to believe that we will also be able to extract subdominant frequencies [228], as long as the numerical waveforms currently available approximately reflect the true strength of the emitted signal (which could be weaker if physical viscosity is very high; see also Ref. [373]) and the expected calibration of the detectors at high frequencies are realized.

4.4 Investigating phase transitions with post-merger waveforms Hadron-quark (and other strong) phase transitions or phase transition to hyperonic matter may occur at some high-density threshold. Since the densities reached after the merger are larger than those in the original stars in the binary, it is possible that phase transitions occur only after the merger. In this case, which is the focus of this Section, a measurement of the tidal deformabilities, of course, cannot contain information on phase transitions. The case in which such phase transitions occur in both or one of the stars in the binary already before the merger will be treated in Sect. 7.3. Several simulations of BNS mergers with EoSs containing a phase transition to hyperonic matter (see the end of Sect 7.1 for more comments on stars containing hyperonic matter) have been carried out, starting with the work of Refs. [525, 526], where it was found that the post-merger GW frequency significantly changes during the time between merger and collapse. If it were so, the above-mentioned relations between post-merger spectral properties and stellar physical quantities may have to be taken with care. However, more recent works [527, 109] found that also in case hyperons appear the post- merger main GW frequency remains rather constant over time and that its frequency is very similar to (and, in the foreseeable future, indistinguishable from) that given from BNS mergers with the cor- responding purely nucleonic EoS. What differentiates hyperonic EoSs in the post-merger GW signal is the amplitude and phase modulation and the total [527], but these quantities are both more difficult to measure36 and more difficult to obtain accurate estimates of from numerical simulations. In a strong phase transition, like one from nucleonic to quark matter, the main feature of the power spectral density of the post-merger phase, fmain peak, may instead change rapidly, because of the abrupt speed up of the rotation of the differentially rotating core of the remnant [109]. The lifetime of the merged object before collapse and the black-hole ringdown waveforms may also be rather different from the ones expected from purely nucleonic EoSs [108]. In addition to GWs, the rearrangement of the angular momentum in the remnant as a result of the formation of a quark core could be accompanied by a prompt burst of neutrinos followed by a gamma-ray burst [528, 529, 530]. However, a preliminary study found that there would be no significant qualitative differences in the electromagnetic counterpart of NS mergers between a system undergoing a phase transition to quark matter and purely hadronic mergers [531]. These indications come from simulations of merging NSs described by EoSs that include a phase transition [108, 109, 531]. As mentioned in Sect. 2.1, such simulations have been performed for the first time very recently and published at the same time by two groups, one that used a smoothed-particle- hydrodynamics code with approximate treatment of general relativity [109, 531] and the other that carried out fully general-relativistic simulations [108]. Neither work includes magnetic fields. The latter simulations included quarks at finite temperatures (within a temperature-dependent chiral mean field model37 [533]) and were used to present for the first time a phase diagram (see Fig. 7) illustrating the properties of the phase transition, as often done in studies about heavy-ion collisions. This work then showed that a quark-hadron phase transition (occurring after the merger, in this case)

36Ref. [527] estimated that Advanced LIGO could distinguish between the two EoSs they employed with a single merger at a distance of up to 20Mpc, depending on the total mass of the binary. 37The EoS employed∼ does not produce gravitationally stable hybrid stars, therefore the effects of such a hadron-quark phase transition can only be observed in a BNS merger.

31 log10 Yquark

6 5 4 3 2 1 0 60 − − − − − −

14 14 50 crossover transition

40 12 12

30 10 [ms] [MeV] 10 t T

20 8 8 n /n 10 max sat st Tmax 1 order 6 6 phase transition 0 1.0 2.0 3.0 4.0 5.0 10.0 nb/nsat

Figure 7: Phase diagram illustrating the properties of the phase transition by showing the evolution of the temperature and density in the merger remnant. Diamonds refer to the maximum normalized baryon-number density and circles to the temperature after the merger of a BNS system described with an EoS that includes a strong first-order phase transition to quark matter. Different times of the evolution are represented with a colour code, together with the quark fraction Yquark . The grey shaded area shows the first-order phase-transition region. See Ref.[532] for details. (From the the version of Ref. [532] on arXiv.org)

would indeed leave its strong imprint in the GW signal (collapse time, ringdown), but has only a small influence on the post-merger GW frequencies. This is because the EoS used in Ref. [108] produces a small (see below for a comparison with another EoS with a strong phase transition, the DD2F-SF EoS [534]) quark-matter fraction during the early phases of the post-merger. When, at later times, the quark matter fraction increases, it immediately induces the gravitational collapse of the remnant. Ref. [109] more systematically studied the effects of phase transitions on the post-merger phase by adopting purely hadronic EoSs, EoSs with a second-order phase transition to hyperonic matter [535, 536, 537] and EoSs with a first-order phase transition to quark matter (the DD2F-SF EoS [534]), which produce a non-negligible amount of quark matter soon after the merger. It was shown that the phase transition increases the dominant post-merger GW frequency fmain peak if and only if a strong first- order-like phase transition leads to the formation of a gravitationally stable extended quark-matter core in the post-merger remnant [109]. This happens because the formation of a quark-matter core makes the remnant more compact. Also, comparing the outcomes of all their simulations on the Λ-fmain peak plane, they found that only the DD2F-SF models with a phase transition to deconfined quark matter lie clearly outside the relation visible for other EoSs38, including those with a second-order phase transition (see Fig. 8). The maximum deviation from this relation is about 100Hz for EoS models without strong first-order phase transitions and about 500Hz for those with a strong phase transition to deconfined

38Note, however, that error bars on the points in the graph have not been computed, but similar computations by the same authors had given uncertainties a few 10Hz [375]. See also Ref. [373] for more on computations of error bars in these kinds of relations.

32 3.75 ∆n = 0.121 ∆n = 0.094 3.50 ∆n = 0.030 3.25 [kHz]

peak 3.00 f

2.75

200 400 600 800 Λ1.35

Figure 8: Dominant post-merger GW frequency fmain peak as a function of tidal deformability Λ for mergers of two 1.35M stars. The DD2F-SF [534] models with a phase transition to deconfined quark matter appear as clear outliers. The solid curve displays the least square fit for all purely hadronic EoSs (including three models with hyperons, marked by asterisks). Arrows mark DD2F-SF [534] models with roughly the same onset density and stiffness of quark matter, but with different strength of the density jumps. (From Ref. [109])

quark matter. Namely, it was found that only a sufficiently strong first-order phase transition (or, actually, any transition, even not formally first-order, which causes a strong softening of the EoS like the one seen for DD2F-SF models) has a noticeable impact on the and thus alters the post-merger GW signal in a measurable way. What would be determinant for the discovery of such a phase transition is a shift of the fmain peak frequency as compared to that expected from the tidal deformability of the inspiralling NSs. Such a shift of the dominant post-merger GW frequency might be revealed by future GW observations using second- and third-generation GW detectors [109] and would also allow to constrain the density at which the phase transition occurs.

5 Combining analyses of waveforms emitted before and after the merger

Even if (as mentioned in Sect. 3.3), with current detectors, omitting the post-merger waveforms from the global analysis of the signal does not lead to significant loss of information or biased estimation of the source properties [408], it would be ideal to have waveform templates that consistently and exhaustively cover the inspiral, merger, and post-merger phases, so that one could perform matched-filter searches, as mentioned in Sect. 4.2. This is true especially for the shorter-duration signals of the post-merger phase. However, complete numerical-relativity waveforms are too sparse and semianalytical models for the whole signal are in their infancy. One work has carried out a global analysis that combines the two types of estimates from the

33 pre-merger and post-merger waveforms [507]. Through Monte-Carlo simulations that combine mea- surements of the total mass from the inspiral phase with those of the compactness from the post-merger oscillation frequencies, improved estimates were found for the mean population radius of NSs: With about 50 observations, the error on the average radius was computed to be 2 5% for stiff EoSs and 7 12% for soft EoSs. − −

6 The merger of binaries of stars not made of ordinary matter

As mentioned in Sect. 2.5, there may exist compact objects similar in mass and size to NSs, but not made of ordinary matter, like boson stars or gravastars. A few recent works have proposed preliminary studies about whether current and future observations can distinguish between inspirals and mergers of NSs and boson stars [538, 95, 539, 540, 97, 541] or gravastars [542, 331, 315]. GW detectors may also be able to probe the structure of these different compact objects (if they exist) through their tidal interactions in the inspiral of binary systems and through the phenomenology of their merger, post-merger and ringdown phases. The first step of such studies is the computation of the tidal deformability of these objects [538, 539, 543, 544, 545]. Boson stars, actually, have no surface, so defining their radius (and hence compactness) may be ambiguous. One common convention is to define the radius as that of a shell containing a fixed fraction of the total mass of the star. Another option consists in using only quantities that can be extracted from the asymptotic geometry of the boson stars, like the total mass and the dimensionless tidal deformability [539]. The boson-star tidal deformability can be obtained similarly to how it is done with NSs (see Sect. 3.1). Different varieties of boson stars have been considered in the literature (see Refs. [316, 317, 315] for full reviews and, e.g., Ref. [539] for a short summary), but in general, the compactness and tidal deformability of boson stars is comparable to that of NSs, with most types of boson stars having tidal deformability as much as 25% smaller than that of a NS of similar compactness. Solitonic boson stars [546] and Proca stars≈ [329] can have larger compactness and smaller tidal deformability than any realistic NS. Numerical simulations of the merger of boson stars have been performed for some years, with the first works focussing on head-on collisions [547, 548, 549], which are computationally less difficult. More recent simulations have investigated collisions of very compact boson stars and compared them to collisions of black holes [550, 551, 552]. In one work, numerical simulations of the head-on collision of a NS with an axion star [541] were performed. In this particular example, it was found that the GW emission after the merger extends for a much longer duration with respect to a NS head-on collision, thus releasing more energy in GWs. Assuming that a binary merger of a NS and an axion star would show similar dynamics, it has been estimated [541] that it would be observable up to about 100 Mpc with current GW detectors and up to 1 Gpc with the Einstein Telescope [234, 235]. Simulations of binary mergers of a NS and an axion star have not been performed yet, but those of the merger of two inspiralling boson stars have. The first such simulations were done with the main goal to study the properties of the merged object [94]. One interesting finding was that, if collapse to black hole does not occur, a nonrotating boson star forms from the merger, having lost all the original angular momentum by emitting scalar-field and gravitational waves. This is obviously a very different outcome from that of BNS mergers. Other works then studied the frequency of the GW emission that carries away the angular momentum [95], finding that it depends on the compactness of the initial boson stars and that it is in most cases higher than the main post-merger frequency in BNS mergers (that do not end in a prompt collapse). As for BNS mergers (see Sect. 2.6), Ref. [95] found that the main post-merger frequency can be related to the frequency of the merger, defined as the instantaneous GW frequency at the time when the amplitude reaches its first peak after the inspiral phase [71]. It was suggested that third-generation detectors like the Einstein Telescope [234, 235] or Cosmic Explorer

34 [233] may be able to distinguish post-merger gravitational signals of boson-star mergers from those of NS mergers [95]. Ref. [97] considered simulations of binaries of dark boson stars, namely binaries composed of two boson stars each described by a different complex scalar field and thus not interacting with each other except through gravity. It was found that their merger produces a gravitational signature in principle distinguishable from that of binaries composed of black holes, NSs and also interacting boson stars. Similar results were found for mergers of BNS with dark matter particles trapped on their interior [513]. The distinctive feature of this scenario is the generic development of a strong m = 1 one-armed instability, in principle distinguishable from that of mergers of NSs (see Sect. 4.1). I have already reviewed in Sect. 4.1 some works that found other distinctive features in the post-merger spectrum due to the possible presence of dark matter [514]. Finally, Refs. [538, 539] investigated more systematically for different types of boson stars the extent to which tidal effects in the inspiral GW signal can be used to discriminate between standard sources and boson stars. Only some types of boson stars and only nonrotating boson stars were studied. It was found that Advanced LIGO could differentiate between massive boson stars and NSs or black holes, but only in some cases, namely only for some types of boson stars and for systems with large mass asymmetry. For example, it was found that the lower limit for the tidal deformability for boson stars with a quartic self-interaction is 280, while that for boson stars with a solitonic interaction is 1.3 [539], implying that Advanced LIGO≈ is not sensitive enough to discriminate between solitonic boson≈ stars and black holes. Third-generation detectors like the Einstein Telescope [234, 235] or Cosmic Explorer [233] should be able to distinguish between them. A Fisher-matrix analysis confirmed these first estimates [538].

7 Constraints on stellar radius and equation-of-state param- eters deduced from GW170817

In this Section, I discuss the numerous works and results that appeared after the GW observation of GW170817 [1] and tried to estimate physical quantities from it and, ultimately, the EoS. Before starting, however, let me note that some articles [29, 30, 31] claimed that actually GW170817 has not added new insights about the EoS, because the constraints it imposes are less stringent than those obtained from current knowledge in nuclear physics, with the possible exceptions of the estimate of the lower limit for the stellar radius [488, 387, 489, 388], which however is said to be possibly affected by systematic errors [31]. In particular, Ref. [29] pointed out that limits on the high-density EoS parameters from the constraints set by GW170817 on the tidal deformability are weaker than the constraints extracted from analysing other astrophysical observations unrelated to GWs. Ref. [31] then stated that the EoSs that are said to be ruled out by the upper limit on the tidal polarizability derived from GW170817 were actually already ruled out on the basis of state-of-the-art nuclear theory describing microscopic EoSs at densities where error estimates are still credible [31]. More details on this view are presented below. In addition to the GW measurement, additional information has been obtained through the electo- magnetic emission (in the gamma-ray, x-ray, , optical, infrared, and radio bands) related to GW170817, named GRB 170817A (for the gamma-ray burst) and AT2017gfo for the rest of the astro- nomical transient [4, 32, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28], alone or in combination with GW constraints on the tidal deformability. However, electromagnetic radiation from binary mergers is out of the scope of this review, so I limit myself to refer the reader to some of the relevant references [488, 387, 489, 490, 491, 492, 493] and to reporting estimates from some of these works for the radius of a 1.4M star in Table 1. Note also that extracting accurate information from electromagnetic radiation from GW170817 is difficult because the underlying models are still based on assumptions that may significantly introduce systematic errors [519].

35 Refs. [360, 553, 275] combined the GW measurement with electromagnetic observations of other NSs in accreting low-mass x-ray binaries (from Ref. [554]) to give a more precise estimate of the tidal deformability and Ref. [254] proposed an affordable Bayesian approach that can include diverse evidence, such as nuclear data and the inferred masses, radii, tidal deformabilities, moments of inertia, and gravitational binding of NSs obtained through different types of observations [254]. The method allows any parameterization of the EoS. Estimates and constraints on EoSs from measurements of NSs have been derived on the basis of (i) bounds on the tidal deformability in GW170817 [220, 314, 420, 532, 30, 364, 365, 487, 298, 29, 399, 306, 555, 264, 265, 266, 267, 268, 269, 556, 422, 557, 340, 532, 558, 559, 560, 561, 298, 562, 563, 564, 565, 289, 288, 566, 367, 567, 568, 31], (ii) the upper bound for the maximum mass of a cold non-rotating compact star deduced in various ways from GW170817 [130, 569, 363, 570, 306, 288, 556, 571], (iii) the lower bound for the maximum mass for a non-rotating compact star from observations39, (iv) the lower bound on the radius of a non-rotating compact star deduced from GW170817 [488, 387, 489, 388]. Condition (iii) requires stiff enough EoS, while conditions (i) and (ii) require soft enough EoS. According to Refs. [298, 31], condition (ii) is powerful for smooth EoS models, because stars without phase transitions exhibit a strong correlation between their maximum mass and the radii of a NS of typical mass40, but it is not very constraining for general EoS models that may contain strong first-order phase transitions. Also note that there is an open debate on what may be the (systematic) error bar of the upper bound for the maximum mass of a cold non-rotating compact star, since some assumptions need to be made in its calculation. Some of the the articles cited at point (ii) above do report an error bar on the values they give, but a recently posted work explained that all previous results were obtained with oversimplified assumptions and that presently all one can reliably say is that the upper bound for the maximum mass of a cold non-rotating compact star is only weakly constrained as . 2.3M [571]. Since studies performed for hadronic EoSs, quark EoSs, and EoSs with phase transitions lead to rather different results, below I will separate them in different subsections.

7.1 Hadronic equations of state As mentioned earlier, the first published analysis of GW170817 by the LIGO-Virgo Collaborations [1] was very conservative: the EoSs of the two stars were assumed to be completely independent, which resulted in a not-much-constraining upper limit for the tidal deformability and no estimate of radius. Later, the LIGO-Virgo Collaborations published a revised analysis with improved estimates [220, 314] (see below) and the correction of an error in their reporting of the 90% confidence upper limit of the tidal deformability41, but in the meanwhile some notable works were published that used the original very conservative (and affected by the bookkeeping error) observational results [420, 532, 30, 566]. Some of these works studied how parameterized smooth EoSs fit with the constraints from GW170817. The two most prominent articles to carry out the first systematic studies of the statistical properties of the tidal deformability on this line [532, 30] generalized efforts of previous works that had used a more limited set of EoSs to check whether the observation of GW170817 allowed them or not [420, 573, 566]42. Both Refs. [532, 30] parameterized a very large range of physically plausible EoSs for compact stars with a piecewise-polytropic representation and computed equilibrium stellar solutions for up to one million different EoS by numerically solving the Tolman-Oppenheimer-Volkoff [416, 417] equations. By

39 As mentioned earlier, works published before the discovery of a 2.17M NS [310] use for this limit different numerical values in the range 1.97 0.04M . 40 ± For example, considering smooth EoSs, Ref. [572] showed that the discovery of a 2.17M NS [310] significantly reduced the allowed range for the skewness Q0 (see Sect. 2.4). 41A mentioned previously, the corrected value in the case of the low-spin prior is Λ˜ 900 instead of 800; see the caption of Table IV in Ref. [220]. ≤ ≤ 42Ref. [555] later updated the results of [420] with the new LIGO-Virgo Collaborations estimates for GW170817.

36 imposing only one condition, that the sound speed is subluminal, Ref. [30] focussed on smooth EoSs and concluded that the 95% credibility range of predicted tidal deformabilities for a 1.4M NS is consistent with the upper bound deduced from GW170817 [1]. Ref. [532] considered also EoSs with phase transitions and imposed some additional constraints, namely that the tidal deformability for a 1.4M NS is Λ1.4 < 800 (from the first official analysis of GW170817 [1]), that the lower bound on the maximum mass for a non-rotating NS is Mmax > 2.01 and that an upper bound on the maximum mass for a non-rotating NS is Mmax < 2.16 [363]. Ref. [532] proposed for the first time a lower limit for the tidal deformability Λ1.4 > 400 at a 2σ confidence level (see Table 1 for their radius estimates). They pointed out that the distributions are very robust upon changes in the upper limit of the maximum mass and of the tidal deformability and that different prescriptions on the treatment of the nuclear matter in the outer core (0.08 < ρ/fm−3 < 0.21) [532, 420] may have a large impact on the statistical properties of NS radii. Progress in theoretical knowledge of the outer core is then important for a good description of the NS structure. Also the NS crust may have a large impact on the statistical properties of NS radii [30, 574], even if it does not strongly impact tidal deformability [574, 575, 576] and thus it is not very important in relation to current GW measurements from BNS systems. Another work that was published before the updated analysis by the LIGO-Virgo Collaborations [220, 314] compared the tidal deformability observation from GW170817 (the first announcement [1]) with data from the PREX experiment [45, 46] (elastic scattering of polarized electrons) on the neutron- skin thickness (the difference between the root-mean-square radii of the distribution of neutrons and protons) of 208Pb [421] (see also Refs. [577, 263]). In order to connect the observed tidal polarizability to nuclear observables, they assumed EoSs described with a relativistic mean-field approach and found that the central value of the neutron-skin thickness of 208Pb measured in PREX is rather far from the one derived from GW170817, even if there is no tension because the error bar of PREX is very large. If this experimental datum is confirmed by more accurate experiments (like PREX-II) in the future, then this might be in tension with GW170817 and be indicative of a phase transition in the interior of NSs [421]. An alternative view on this possible tension has been expressed in Ref. [31], where it was stated that actually smooth EoSs (without phase transitions) different from those employed in Ref. [421] can be found that are compatible with both GW170817 and future more accurate nuclear experimental data that may confirm the value of the neutron-skin thickness of 208Pb found in PREX [31]. Applying the calculation of Ref. [421] to the revised analysis of GW170817 [220, 314] would probably not change their conclusions. As mentioned above, several months after the announcement of the observation of GW170817, the LIGO-Virgo Collaborations published a revised analysis [220, 314] that provided improved estimates on the tidal deformability and on the radius by incorporating new theoretical insights, improved analysis tools, re-calibrated data (also starting at a lower frequency threshold), and some likely assumptions that had not been assumed in the first article [1] in order to maintain the widest generality and conser- vativeness. These assumptions are the source location (known mostly by electromagnetic observations [4, 25, 578]), the fact the two stars have the same EoS, the fact that the more massive star in the binary has smaller radius and lower tidal deformability, the fact that, because of causality, the speed of sound in the NS must be less than the speed of light (plus 10% to allow for imperfect parameterization) up to the central pressure of the heaviest star supported≈ by the EoS, and the fact that the EoS must allow for non-rotating stars of mass > 1.97M . The improved analysis tools consisted in more accurate post-Newtonian waveform models, the use of approximately EoS-insensitive relations to relate observed quantities to radii, and sampling the EoS also as a spectral parameterization [256], which has the advantages mentioned in Sect. 2.4. Such an analysis is more model dependent than the original one, but apparently the different models used in Refs. [220, 314] give the same results. However, Ref. [306] warned that, according to their own Bayesian analysis, constraints on the EoS from GW170817 alone may be relatively prior-dominated and thus should be interpreted with care (see also the end of Sect. 7.3).

37 ˜ +420 The refined estimate for the mass-weighted average dimensionless tidal deformability is Λ = 300−230 +390 and that for the dimensionless tidal deformability of a NS of mass 1.4M is Λ1.4 = 190−120 [314]. For comparison, the widely used candidate EoS Sly [260], one of the softest theoretical models, has Λ1.4 290, while the ms1b EoS [579], one of the stiffest, has Λ1.4 1220. Overall, these constraints favour≈ a relatively soft NS EoS, and are believed to rule out several≈ candidate EoSs at 90% confidence level [220, 314] (but see below for ideas on how to revive the excluded EoSs and Refs. [29, 30, 31] for an alternative view). See Table 1 for the radius estimates. Various other works re-analysed the publicly available data of GW170817 in similar but distinct ways and drew similar and consistent conclusions [364, 399, 365, 487, 298]. Ref. [399] revisited the early analysis of the LIGO-Virgo Collaborations [1] by incorporating some physical assumptions that had been left out for conservativeness. First of all, they assumed that the two stars in GW170817 had the same EoS (this has then been assumed in all subsequent studies). Then, they made use of the binary Love universal relation [361, 362] (see Sect. 2.6) between the dimensionless tidal deformabilities of the two stars and the ratio of their masses. The use of such a universal relation implies that, for a given realistic EoS, the dimensionless tidal deformability is a decreasing function of the mass of the NS (for stable NS configurations) and so that the most massive component in a BNS has the smallest tidal deformability. The use of this relation to link the tidal deformabilities of the two stars leads to a reduction in the extension of the credible region by factors of 2 to 10 and to an improvement in the measurement of the individual tidal parameters by up to an order of magnitude, depending on the EoS and the mass ratio. Also Ref. [364] was published before the updated analysis of the LIGO-Virgo Collaborations [220, 314] and assumed that the two stars in GW170817 had the same EoS. This is the first published work to place also lower bounds on the deformability and radii of NSs on the basis of a revised statistical analysis of GW170817. Their statistical errors are comparable to the error reported later by the LIGO-Virgo Collaborations, but total errors were estimated to be larger than in Refs. [220, 314] because systematic errors (from unknown physics related to the EoS) of 0.2 km were added as conservative bounds. In order to carry out their estimations, Ref. [364] justified and used the approximate relation 6 ΛA = q ΛB, where q is the mass ratio and ΛA,B the dimensionless tidal deformabilities of the two stars of the binary. Later, the treatment was improved by considering instead analytic limits on the dependence of the dimensionless tidal deformability on the stellar mass, Λ(M) [365]. This method is alternative to that used in Refs. [399, 220, 314], where the binary Love relations between the dimensionless tidal deformabilities of the two stars and the ratio of their masses [361, 362] were used. It is claimed that the latter method is more suitable to describe modifications to deformability correlations because it also contemplates the existence of a strong first-order phase transition (see also Sect. 7.3). In fact, if one of the stars in the binary has undergone a strong phase transition and the other has not, the correlation found in Ref. [364] would be weakened [220, 314]. In a different line of research, Refs. [580, 271, 581] studied whether the constraints imposed by GW170817, together with those from other astrophysical observations (in particular the lower limit on the maximum NS mass) and from nuclear theory and experiments, can allow for hyperons to appear in NSs. It is well known that, while, on one side, beyond nuclear density the conversion of into hyperons is energetically favourable (see, e.g., Ref. [582] for a review), on the other side it is not easy to explain how NSs containing hyperons can have masses around 2M , since hyperonic EoSs are usually too soft at the required densities. A few ideas to solve this puzzle have been proposed (see, e.g., Ref. [3] for a review). One of these consists in tuning the interactions in the hyperonic sector to allow for high-mass NSs [583, 584, 585, 586, 587, 588, 589, 590]. However, most of the hyperonic stars proposed have tidal deformabilities and radii that are larger than what indicated by GW170817 (and other observations). A hyperonic EoS should be sufficiently soft below 2-3 times the saturation density, in order to have a tidal deformability compatible with GW170817, and sufficiently stiff at higher densities to sustain NSs with masses larger than 2M . This is exactly what was found to be the effect of

38 considering ∆-isobars in the EoS, in addition to hyperons [591, 580, 271, 581, 592, 593]. Namely, EoSs including both hyperons and ∆-isobars are found to satisfy also all current astrophysical constraints [580, 271, 581].

7.1.1 Correlations of nuclear-matter parameters of the equation of state A few works focused on estimating physical parameters of the expansion of the EoS, as introduced in Sect. 2.4, for the data of GW170817. The first estimates, based on the LIGO-Virgo first analysis release 43 [1], concluded that the limits on the high-density EoS parameters from the Λ1.4 < 800 constraint alone are weaker than those imposed by measurements of the NS radius performed through x-ray observations [29]. Other works, based on ideas of previous studies [594, 595, 274] proposed before the observa- tion of GW170817, pointed out that actually the (mass-averaged) tidal deformability is found to be weakly or only moderately correlated with the individual nuclear-matter parameters of the EoS [264, 265, 266, 267, 268, 269], while stronger correlation is found for specific combinations of the EoS parameters, in particular for a linear combination of the slopes of the incompressibility and symmetry- energy coefficients [267, 268]. The correlation coefficients found are only around 50%, but the authors set conservative bounds on some nuclear parameters (the incompressibility K0, its slope M0, and the curvature of symmetry energy Ksym,0 at nuclear saturation density) to be 81 MeV K0 362 MeV, ≤ ≤ 1556 MeV M0 4971 MeV, and 259 MeV Ksym,0 32 MeV at 90% confidence level from GW170817≤ [268]. ≤ − ≤ ≤ Also Ref. [269] found that the individual measurements of the tidal deformability and stellar radius can only set limits on combinations of nuclear parameters; in their case, stringent constraints on the correlation between the slope L of the symmetry energy and its curvature Ksym were found. The same authors, by combining several physical and astrophysical constraints, also claimed that the symmetry energy at twice the saturation density is constrained to 46.9 10.1 MeV [556]. ± It was also pointed out in a Bayesian study [254] (already mentioned above), which included nuclear data, NS masses, radii, tidal deformabilities, moments of inertia, and gravitational binding energies obtained through different types of observations, that more precise measurements of the slope L of the symmetry energy would strongly constrain the EOS only below nuclear saturation density and would have little implications for densities above saturation density. An important caveat needs to be highlighted here: Most correlations of the kind mentioned above that are described in the literature originate or may originate from the lack of generality of existing phenomenological functionals describing EoSs. Instead, in these estimates it is important to account for the widest possible range of valid EoSs and EoS variation [264, 265, 267, 306, 268]. Furthermore, bounds like those mentioned above have been derived only for NSs and may not be valid for hybrid stars with a quark core and a nuclear matter envelope, or for compact stars containing heavy baryons (hyperons and/or ∆-isobars) [580, 271] (see Sect. 7.3).

7.2 Quark stars Some works have checked whether the data from GW170817 (the first announcement [1]) are compatible with stars made purely of free quarks or of strange-quark clusters (which have been called strangeons [596]). It is important to note that, because of the finite surface density of quark stars and of strangeon stars, the standard relations involving the tidal deformability for nucleonic EoSs need to be modified; therefore constraints on NSs according to the analysis of GW170817 performed by the LIGO-Virgo Collaborations [1, 220, 314] and similar studies cannot be simply applied in the scenario of binary– quark-star mergers [422, 557, 340] (see also Sect. 7.3).

43 The corrected upper limit Λ1.4 < 900 (see note on page 36) would have been even less constraining.

39 With this caveat in mind, Ref. [422], making use of the MIT bag model [281, 282], and considering the limits on the tidal deformability from the GW170817 discovery article [1], inferred that the constraints on the tidal deformability from GW170817 are compatible with a binary–quark-star merger. This remains true also when considering the updated constraints from the LIGO-Virgo Collaborations [220, 314], but very possibly only for stars made of superfluid quarks (where quarks form Cooper pairs) [422]. Also Ref. [340] showed that the tidal deformability of quark stars may be reduced to values compatible with current GW observations by the effect of elasticity generated if a crystalline colour-superconducting phase occurs, like in the proposed solid quark stars [339], solid quark stars covered with a thin layer of nuclear matter [339], or solid quark-cluster stars [597]. It must be said, anyway, that even if their tidal deformabilities were compatible with GW170817, it is difficult to explain how quark stars can provide the ejecta necessary to produce the observed macronova emission [598]. Even if no detailed studies exist, it has been also claimed that the tidal deformability of a strangeon star is compatible with GW170817 [557, 599], also when considering the updated constraints from the LIGO-Virgo Collaborations [220, 314]. Also for strangeon stars there are issues related to the ejecta: As for free quarks, the ejecta composed of strangeon nuggets would not lead to r-process . However, it has been claimed that different components of the observed macronova AT2017gfo may be explained in the strangeon scenario in the following way: The blue component could be powered by the decay of ejected strangeon nuggets, while the late red component could be powered by the spin-down of the remnant strangeon star after merging [557]. Finally, the so-called two family scenario [600, 565, 601, 561, 602] involves also quark stars, but I will describe it in the next Sect. 7.3, for convenience.

7.3 Hybrid stars Several works have investigated whether GW170817 could have originated by the coalescence of a NS with a hybrid star or that of two hybrid stars [532, 558, 559, 560, 561, 298, 562, 563, 564, 565, 289, 288, 603, 604, 566, 367, 605, 567, 568, 606]. All found that GW170817 is also consistent with the coalescence of a binary system containing at least one hybrid star. As mentioned earlier and widely known, the neutron matter EoS is expected to be reliable up to densities as high as twice the nuclear saturation density of 0.16 fm3. Beyond that not much is known for sure and matter could also undergo a phase transition to quark matter [276, 277, 278], possibly giving rise to a hybrid hadron-, in which the EoS is nucleonic for lower densities and contains quark matter at higher densities. The possibility of more than one phase transition is also envisaged [607, 608, 289]. Such phase transitions would change drastically the properties of the star, in particular because they would produce drastic softening or stiffening of the EoS (according to the type of the discontinuity, namely whether it is sharp or smoothed, the Love number of strange quark stars can be smaller or larger than that for hadronic stars [609]). Multiple detections of GWs from different BNS mergers with the same value for the tidal deformability of one star ΛA but considerably different values of the tidal deformability of the second star ΛB would probably indicate the existence of a strong first-order phase transition in NS matter [567, 288, 610]. The first numerical-relativity simulations of BNS mergers in which matter undergoes a hadron-quark phase transition have been performed recently [108, 109] (see Sect. 4.4). Some hadronic EoSs that have been found not to satisfy the GW170817 constraints on the tidal deformability [1, 420, 563, 220, 314, 555] and/or the lower limit for the maximum NS mass [312, 311, 310] can be salvaged from the closed files of theories44, by proposing that they are the lower-density part

44In the next paragraphs I will write about some ways to salvage EoSs, but other physical effects have been found that may make the tidal deformability obtained with a given EoS higher, and therefore more easily excluded by GW observations of BNS systems. In particular, Ref. [611] found that, at least for the EoSs tried, considering superfluidity may increase the tidal deformability of 5 10%. ∼ − 40 of an EoS with such a hadron-quark phase transition, since hybrid stars built in that way have been shown to be compatible with the constraints imposed by GW170817 [558, 559, 560]. Stiff EoSs excluded at a first investigation may also be salvaged by considering the additional effects that vector-meson interactions can have on EoS models that do not consider them, since these interactions were found to decrease the tidal deformability of a 1.4M star by 30% [612], or by assuming that dark matter interacts with nucleonic matter inside the NS, since this∼ softens the EoS and lowers the value of tidal deformability [514, 613, 614, 615, 616] (see Sect. 7.4). Before continuing the topic of hybrid stars, I mention here for completeness that a third way to salvage EoSs that seem ruled out at first sight by recent observations consists in considering anisotropic compact stars [349], which have been studied in many works since the seminal one of Ref. [617] (see Refs. [618, 619, 620, 361, 621], for some of the most recent references). Pressure in stars in equilibrium is usually assumed to be isotropic, but the possibility that the radial pressure and the transverse pressure are not equal has been proposed in several works [622, 623, 624, 625, 626]. Basically, pressure anisotropy can arise whenever the velocity distribution of particles in a fluid is anisotropic, which in turn can be caused by magnetic fields, turbulence, convection [622], phase transitions [623], superfluidity or solid cores [622, 623]. Additionally, some braneworld models of gravity, with extra spatial dimensions, predict an effective four-dimensional stress-energy tensor with anisotropic pressure45 [625, 626]. Using the formula for pressure anisotropy proposed (for facility of use, rather than for physical reasons) in Ref. [617], Ref. [349] computed the tidal deformability of anisoptropic stars and found that, for a wide range of values for the parameters of the anisotropy ansatz and for stars with mass in the range of interest [1, 2]M , the tidal Love number for an anisotropic star is smaller than that for a spherically symmetric star with the same mass. This leads to the conclusion that EoSs that are apparently ruled out by GW170817 because of their stiffness might still be viable if the stars in that BNS system were supported by a sufficiently anisotropic pressure. Additionally, it has been noted that there is then a degeneracy between EoS and anisotropy in the interpretation of the tidal deformability measured from BNS inspirals. This degeneracy may be solved if the stellar radius is measured independently, since anisotropic stars usually have a larger radius for a given mass [349]. After digressing about the effects of anisotropy, I now resume the main topic of this Section, saying that the interest in hybrid EoSs goes even further than what written above, since some choices of phase transition predict the existence of twin stars [302, 303, 304]: NSs and hybrid stars having the same mass but different radii. The appearance of a second stable branch in the mass-radius relation of compact stars (see e.g. Fig. 9) is thought to arise from the occurrence of a phase transition. The lower-mass star in the twin is made solely of nucleons and has larger radius and larger tidal deformability, while the more massive star is a hybrid star with a quark content and smaller radius and tidal deformability46. See Refs. [287, 299] for studies of classifications of twin-star scenarios, Ref. [288] for a recent extensive analysis of the features of the phase transitions that lead to twin-star configurations and Refs. [628, 629, 630] for recent works on viable EoSs that allow for twin stars. It has been recently pointed out that strong magnetic fields can influence noticeably twin-star scenarios [568]. Even if, given an EoS, non-magnetized or weakly magnetized twin stars are found to exist, it is possible that strongly magnetized twin stars with that same EoS do not exist. Conversely, even if, given an EoS, non-magnetized or weakly magnetized twin stars are found not to exist, it is possible that strongly magnetized twin stars with that same EoS do exist [568]. As a somewhat different idea, the two-family scenario [565] has been proposed, in which hadronic stars exist up to 1.5 1.6M (their radius and tidal deformability decreasing with increasing mass) and more massive stars− are strange quark stars, with larger radii and tidal deformabilities (note that the

45Conversely, with accurate enough measurements of stellar properties, one could constrain the brane tension of those alternative models of gravity [625, 626]. 46There are exceptions: The so-called rising twins are twin star combinations where the more massive one has a larger radius [627, 367].

41 relations between mass and radii or tidal deformabilities are qualitatively different from those of hybrid stars in the twin-star scenario). A complete transition may occur in finite time during the merger of the binary (or the collapse of a star, e.g. in core-collapse supernovae) [565] and the process of conversion may have two different stages: a turbulent combustion, which, in a timescale of the order of a few ms, converts most of the star, and a diffusive combustion, which converts the unburnt hadronic layer in a timescale of the order of 10 s [631]. Strictly speaking, then, in the two-family scenario hybrid stars exist only briefly during the conversion of hadronic matter into quark matter. In addition to influencing the computation of tidal deformabilities, the two-family scenario predicts different rates for prompt collapses after the merger with respect to other EoSs, because the threshold mass [358] against prompt collapse to black hole after the merger is different. Related to this, it has also been found that GW170817 cannot be a binary of hadronic stars with the hadronic EoS adopted in the two-family scenario (containing hyperons and ∆ resonances), because it would have undergone prompt collapse [601], while this is not believed to be the case [20, 130, 632, 633, 634], but it could be a binary of one hadronic and one quark star: The prompt collapse would be avoided by the formation of a hypermassive hybrid configuration, whose ejecta may give rise to the macronova [601]. The first numerical-relativity simulations of BNS mergers with the hadronic EoS of the two-family scenario have been recently performed [602]. Hybrid stars that give rise to the twin-stars or the two-family scenarios allow for both stars that have a tidal deformability large enough to fit with the lower-bound from GW170817 and stars that have radii small enough47 to fit with data from x-ray observations of isolated NSs. If both such observational constraints are confirmed and become tighter in the future, the twin-stars or the two-family scenarios may become the only viable ones [561]. One way to show the qualitative difference between scenarios that allow for twin stars and those that do not is to plot the dimensionless tidal deformabilities of the stars in the binary against each other in a ΛA ΛB plane. This is often done when examining the constraints imposed by GW observations of BNS on− the EoS. Since the tidal deformabilities are related to each other through the total mass (or, more practically, the chirp mass) of the system, in the absence of twin stars, such a plot results in a broad band, whose position and shape are governed by the EoS and whose width by the error of the measurement of the total mass. Since a twin-star pair has two significantly different values of the tidal deformability, more distinct bands may appear in the ΛA ΛB plot, one for the NS–NS case, one for the NS-hybrid-star case and one for the hybrid-star–hybrid-star− case [567, 288]. Before going into some details about work on hybrid stars, I would like to stress again two general points. The first is that, as already noted in Sect. 7.2, the constraints on the tidal deformability set by the analysis of GW170817 performed by the LIGO-Virgo Collaborations [1, 220, 314] and by similar works were obtained by expanding the tidal deformability as a function of mass Λ(M) linearly about M = 1.4 M . However, such a linear expansion is valid only for EoSs without phase transitions, as can be seen, e.g., in Fig. 9. In fact, using this approximation to estimate the tidal deformability of a 1.4M star should be avoided because it excludes the possibility of testing for hybrid stars. In particular, it has been shown that if one does not make that linearity assumption the upper bound on Λ1.4 could be smaller [532, 558, 298, 561, 289, 288]. The second general note on hybrid stars is that up to the time of the writing of this review there are actually no numerical-relativity simulations in full general relativity of the merger of binaries composed of one hybrid star and one NS or of two hybrid stars. Such simulations are necessary for further comparison with mergers of stars described by purely nucleonic EoSs.

47By using a set of relativistic polytropes connected to a quark bag model for the phase transition with a Maxwell construction (see note on page 43) and with a realistic description of the NS crust (SLy4 [635, 260] up to the nuclear saturation density), Ref. [367] studied the radii and tidal deformabilities of compact stars compatible with current observational constraints and found that the minimum radius that can be produced on a twin branch lies between 9.5 and 10.5 km, while the minimum radius for lighter stars that do not undergo the phase transition is about 12km.

42 For the rest of this Section, I will give more details on specific works involving hybrid stars. Ref. [558] was the first to investigate how GW170817 can constrain the properties of hybrid hadron-quark compact stars. Ref. [559] later did a very similar work. They constructed (with the Maxwell method48 [558] or the Gibbs method [559]) parameterized hybrid hadron-quark EoSs that (i) consist of zero- temperature nuclear matter in β-equilibrium with a low-density phase of nucleonic matter and a high- density single-phase core of quark matter (parameterized with piecewise polytropes and other methods), with a first-order phase transition at their interface, (ii) are supplemented with a low-density EoSs for crustal matter, and (iii) are consistent with the existence of 2M pulsars. Stars built with such EoSs ≈ were found to have the twin-star property at 1.5M . ≈ These articles [558, 559] were then the first to stress that the tidal deformation observed from GW170817 is consistent with the coalescence of a binary composed of one hybrid star and one NS, that certain hadronic EoSs that do not satisfy the GW170817 constraints on the tidal deformation become compatible with GW170817 if a first-order phase transition occurs in one of the stars, and that, while for purely hadronic EoSs the dimensionless tidal deformability can be approximated as a linear function of the gravitational mass for masses in the vicinity of 1.4M , this is not true for hybrid hadron-quark EoSs with low-mass twins [558]. This issue was later elaborated in detail in Ref. [367]. Similarly, Ref. [563], adopting an EoS that predicts stars possibly composed of a core of two-flavour quark matter and a shell of hadronic matter and that allows for twin stars, interpreted the GW170817 event as the merger of either such a hybrid star and a NS or a merger of two such hybrid stars, while the BNS scenario was deemed disfavoured mostly because the stiffness of the hadronic EoS employed made a BNS merger incompatible with the compactness expected from GW170817 (see also Fig. 9). Also Refs. [567, 288] interpreted the GW170817 event as the merger scenario of either a NS-NS, a NS–hybrid-star, or a hybrid-star–hybrid-star binary. In particular, Ref. [288] performed an extensive analysis of the merger scenarios in which the parameters characterizing the phase transition have also been varied and both Maxwell and Gibbs constructions for the phase transition were considered. For the lower-density region of the inner core the hadronic EoSs of Refs. [587, 639], for the inner and outer crust the EoS of Ref. [640], and for the quark phase a constant-speed-of-sound parameterization [285, 286, 287] were used. They also predicted that a phase transition would be revealed in the inspiral waveform if GW detectors measured values of the chirp mass smaller than 1.2M together with tidal deformabilities of Λ1.4 . 400. Ref. [610] performed a Bayesian analysis that included some models for the hadronic and quark matter phases that were said to be realistic and that give rise to twin stars and found that their analysis favours models with a strong mixed phase. Ref. [289] has explored the sensitivity of the tidal deformability to the properties of a phase transi- tions or of two sequential phase transitions (the first from nuclear matter to quark matter, the second, at a higher density, to a different quark-matter phase49), not necessarily producing twin stars, using the constant-speed-of-sound parameterization [285, 286, 287] [641, 608, 607]. In their calculations, they found that the tidal deformability is smaller than that of purely hadronic stars and small enough to be distinguishable in future GW observations of BNS merger events. Refs. [642, 636], on the basis of an idea of Ref. [643], investigated the effects of the surface

48The Maxwell method, or Maxwell construction, describes a constant-pressure phase transition in which density is discontinuous at the interface between hadronic and quark matter, without a phase in which the two are mixed; the Gibbs method (or construction), instead, allows for a mixed phase in which pressure and density vary monotonically. More in general, a hadron-quark phase transition is described by the surface tension between the two phases. Infinite surface tension corresponds to the Maxwell construction and a zero surface tension to a Gibbs construction. It is largely unknown what may be realistic values for the surface tension and different values for it (i.e. different constructions) lead to stars with different properties, especially tidal deformability and radius [636]. Particular care should be therefore taken when using results obtained from such constructions. 49for example, from the two-flavour colour-superconducting to the colour-flavor-locked phase [608]; see, e.g., Ref. [607] for a review.

43 2.2 GW170817 PSR J0348+0432 2

1.8 PSR J1614-2230

1.6 GW170817 ] . O M 1.4 1 PSR J0437-4715 GW170817

M[M M 1.2 2

1 APR 0.8 set 1 set 2 0.6 set 3 9 10 11 12 13 14 15 16 17 R [km]

Figure 9: Mass vs radius for sequences of compact stars with the hybrid EoS of Ref. [563]. Sets 1, 2 and 3 correspond to different onset masses for the deconfinement transition. The dotted lines denote unstable configurations and connect two stable branches. For compari- son, the mass-radius sequence for the APR EoS [637] (a standard EoS for nuclear applications) is shown with a dashed line. The horizontal bands denote: the mass measure- ment for PSR J0348+432 [312], PSR J1614-2230 [63], and PSR J0437-4715 [638]; the mass ranges for the compact stars in GW170817 (labelled M1 and M2); some exclusion regions from the observation of GW170817 (in particular, Ref. [387] excluded radii of smaller than 10.68 km for stars of 1.6M , Ref. [420] excluded radii exceeding 13.6 km for stars of 1.4M , and Ref. [363] excluded masses higher than 2.16M ). (From Ref. [563])

tension50 between the hadron and quark phases (its value is largely unknown) and found that assuming different values of the surface tensions leads to variations in the maximum mass of the hybrid star of 0.02M (which is not particularly significant), in its radius of 0.6km and in its tidal deformability of ≈∆Λ/Λ 50% (which are both very significant). Conversely, with≈ future, more accurate measurements of tidal≈ deformabilities and radii of hybrid stars it may be possible to constrain the surface tension. The current measurements from GW170817 were found to be unconstraining [636]. Ref. [603] pointed out that GW measurements from BNS mergers may offer for the first time access to the structure of objects that might have a non-negligible contribution from vacuum energy to their total mass. The presence of such vacuum energy in the inner cores of NSs would occur in new QCD phases at large densities [644, 645]. If such phase transitions, which are different from those mentioned above, occur, this would lead to a change in the internal structure of NSs and thus possibly influence their tidal deformabilities. In their computations, performed with a piecewise-polytropic representation of some commonly used EoSs augmented with a phase transition involving vacuum energy, they found that the effects of vacuum energy could be measurable (in future detectors and with enough events) for mergers with high chirp masses ( 1.9M ), while for smaller chirp masses like that of GW170817 ≈ 50See also the footnote on the Maxwell and Gibbs constructions, on page 43.

44 the deviations of the tidal deformability are below detection capabilities [603]. Measurements of this sort may be the only possibility for testing the gravitational properties of vacuum energy independently from the acceleration of the Universe. Ref. [298] extended the work of the LIGO-Virgo Collaborations [1, 220, 314] by using a different general representation of EoSs that allows for a phase transition at high densities through a parame- terization of the speed of sound [297, 298], which was developed starting from ideas of Refs. [287, 294]. This is different from the constant-speed-of-sound parameterization [285, 286, 287] mentioned in Sect. 2.4, but analogously offers the advantage that the speed of sound is continuous everywhere except when first-order phase transitions are explicitly inserted. Six values of the density are randomly selected in the interval between the saturation density and twelve times the saturation density and the values of the speed of sound at those densities are treated as reference points, while all other points are computed by connecting the reference points through linear segments. It is claimed that this model parameterizes the widest possible domain of EoSs [298, 31]. With this setup, the authors computed probability contours smaller than those from the LIGO-Virgo Collaborations [220, 314] and obtained 80 < Λ˜ < 570. In a follow-up work [31], it was estimated that, in order to test the predictions of nuclear theory (in the form of chiral effective field theory) in the range between one and two times the saturation density, the uncertainty in the dimensionless tidal deformability needs to be less than about 300, while in order to test the existence of phase transitions in denser matter it needs to be below 100. As discussed in Sects. 2.4 and 3.3, the choice of parameterization can have a significant effect on the global mass-radius relation and the EoS constraints inferred from GW observations of BNS sys- tems. This was shown also for the case of EoSs allowing for first-order phase transitions in Ref. [240], that compared a piecewise-polytropic parameterization to their own constant-speed-of-sound parame- terization, similar but different from that of Ref. [298]. It was also mentioned above that, in order to overcome the drawbacks of parameterized models, non-parametric models have been proposed. Ref. [306] was the first to introduce a non-parametric method for directly inferring the NS EoS from GW data. It used Gaussian process regression (see, e.g., Ref. [646]) to generate a large number of possible dependencies of density on pressure, which span a very wide range of stiffnesses and core , while being consistent with thermodynamic stability, causality, observed astrophysical data, and candi- date models from nuclear theory. In their work, they train what they call their synthetic EoSs on seven (of which five purely hadronic and two containing quarks or hyperons) fiducial tabulated candidate EoSs selected from well-established nuclear-theory models that span a wide range of stiffnesses and support a 1.93M star. Two models were considered: one built with a very uninformative EoS prior (tabulated EoSs provide only weak, general guidance to the form of the synthetic EoSs) and the other more conformed to the chosen fiducial EoSs (in practice the synthetic EoSs depart relatively little from the average of the tabulated EoSs) [306]. Applying the method to GW170817 gave tidal deformabilities and their error bands that for both models agree with the results of the LIGO-Virgo Collaborations [220, 314]. Perhaps the most important claim of Ref. [306] is that, since a Bayesian analysis of the two models (the one weakly influenced by the tabulated EoSs and the one strongly influenced by them) finds that neither is strongly preferred by the data (Bayes factor of about 1.12), constraints on the EoS from GW170817 alone may be relatively prior-dominated and thus should be interpreted with care.

7.4 Possible influence of dark matter A few works have been published on more exotic possibilities of what could also be inside compact stars and how this would influence the interpretation of GW data, starting from those of GW170817. Some works have studied the influence that dark matter may have on NSs (e.g., their maximum possible mass and the mass-radius relation [647, 613]) and their tidal deformability [648, 649, 650], in particular in view of the data of GW170817 [514, 613, 614, 615, 616]. In general, these works have determined that the tidal deformability and the maximum possible mass for NSs are smaller in the presence of a

45

3

DM

k (0 GeV)

F

DM

k (0.02 GeV)

F

DM

k (0.03 GeV)

F

DM

k (0.04 GeV)

F

DM

k (0.05 GeV)

F

2 DM

k (0.06 GeV) )

F

mass(M

1

8 12 16 20

R (km)

Figure 10: Mass-radius relation of NSs for EoSs with different dark-matter Fermi momentum. (From Ref. [615])

dark-matter core [613, 615, 616] (see Fig. 10). As already mentioned in Sects. 4 and 6, it has been found that the presence of sizeable dark matter in merging NSs may cause the appearance of one or two additional peaks in the post-merger frequency spectrum [514, 513] and that it can affect the one-armed deformations of the stellar object formed after the merger [513]. However, these effects of dark matter are non-negligible only if enough dark matter accumulates in (or in and around) NSs and this does not happen during what is considered the usual evolution of NSs. Namely, via gravitational accretion a NS cannot accumulate dark matter in quantities sufficient to form substantial dark-matter cores or halos. The formation of substantial dark-matter structures would require additional processes to occur, like dark conversion of neutrons to scalar dark matter [514, 613], production of dark-matter particles via bremsstrahlung in neutron-neutron scatterings [614], or a nonstandard dark-matter sector with a subdominant dissipative component [515, 647, 648, 651, 516, 517]. Ref. [613] considered eleven representative nucleonic EoSs from the literature and added to NSs built with them self-interacting bosonic dark matter (they also argued that asymmetric dark matter [652] with or without self-interactions would yield similar results). It was found that, if enough dark matter accumulates in NSs, the presence of a dark-matter core with a mass of about 5% of the NS mass could make the NS more compact (and so decrease its tidal deformability and the maximum mass attainable) to the point that some EoS currently compatible with the existence of a 2M NS would instead become incompatible and that some EoSs that lie outside the confidence region≈ of tidal deformability determined from GW170817 [220, 314] could instead no longer be excluded with the current observation. These results are nearly independent from the strength of the dark-matter self-interaction. Ref. [615], then, arrived at similar conclusions, considering fermionic dark matter interacting with nucleonic matter (described with the Walecka model [653, 654, 647]) inside the NS. Refs. [613, 614] considered not only cores of dark matter, but also halos extending outside the NS. In this case, NSs with dark matter would have a larger tidal deformability. In conclusion to this Section, note that the dark-matter mass fraction in a NS may not be the same in all stars, since it may depend on the NS age, initial temperature, or the environment in which it was formed. Such possibilities would further complicate the interpretations of measurements of NS

46 properties.

7.5 Summary of radius measurements A summary of radius estimates obtained from GW170817 (and some other observations) from various works is presented in Table 1 (see also Ref. [288] for a similar table). The Table is not meant to be comprehensive of all published results, but only to give the reader an image of the radius intervals and error bars. In addition to the estimates obtained from GW170817, radii of NSs have been and are being measured with several other approaches, mostly based on x-ray, optical and radio observations. Comments on these observations and results are outside the scope of this review. Several other recent reviews are available on the topic [50, 49, 61, 64]. However, let me note that one of the most fruitful methods is the analysis of thermal emissions from low-mass x-ray binaries in their quiescent stage and photospheric- radius expansion of type I x-ray bursters [242, 50, 655, 656, 657, 658]. Furthermore, it needs to be said that the systematic errors of these methods are not completely known (some of the data are more controversial than others; see, e.g., discussions in Refs. [659, 61, 64]) and different methods yield different values for the estimates of the radius R1.4 of a 1.4M NS: with the so-called touchdown method R1.4 is in the range 10 11km [50] and with the so-called cooling tail method it is in the range 11 13km [657]. − − Future observations by the NICER (Neutron star Interior Composition Explorer) mission [660, 62], expected to provide first results within this year, and the planned eXTP (enhanced x-ray Timing and Polarimetry) Mission [65], LOFT (Large Observatory For x-ray Timing) satellite [661], and ATHENA (Advanced Telescope for High Energy Astrophysics) [662] are seen to have completely different sys- tematic errors, and thus, by combining many measurements with differing systematics one can hope to significantly narrow the range for R1.4. These endeavours are based on the idea of measuring pulsations in the x-rays emitted from the hot cap of a pulsar due to its rotation. These x-ray waveforms depend on the relativistic gravitational properties of the star because they initially propagate through the highly curved NS space-time. Ray-tracing techniques then can yield the radius of the NS or be used directly in Bayesian analyses of EoS parameters (see, e.g., Refs. [663, 664] and references therein for the latest results). In the future, accurate estimates of the NS radius may also come from simultaneous measurements of the mass and the moment of inertia of NSs. These can be obtained mainly in two ways: (i) exploiting geodetic precession, namely the precession of the rotation and orbital axes (if these are not aligned) around the direction of the total angular momentum; (ii) exploiting the advance of the pericentre of the system caused by spin-orbit coupling. Refs. [346, 665] estimated that the moment of inertia will eventually be determined to a precision of about 10%. As described in the previous sections and in Table 1, various works extracted estimates of the NS radius from data obtained from GW170817 and the overall picture that appears from such estimates is relatively robust. Nuclear theory (see, e.g., Refs. [667, 668, 669, 670]) and experiments currently predict approximately the same radius intervals (see, e.g., Ref. [48] for a review) as estimated from GW170817. Results from GW170817 are also consistent with the above-mentioned measurements from electromagnetic observations.

8 Conclusion

Even if not all agree, most people in the community are saying that, just from the one detection of GWs from BNS mergers that has been analysed, it has already been possible to set new constraints on the EoS of very high-density matter. It may be not so useful to debate on whether such a statement is

47 Table 1: Constraints on the radius of NSs from GW170817 (and some other observations) from works that report in their text estimates for the radius R1.4 of a 1.4M star or the radius of the NSs in GW170817, considering EoSs without and with a phase transition and from multimessenger analyses. In each category, the entries are in order of publication date. See also Ref. [288].

Reference Ri [km] Notes Without a phase transition +0.15 ◦ Bauswein et al. [387] 10.68 R1.6 −0.03 ≤ Fattoyev et al. [421] R1.4 13.76 ≤ × Most et al. [532] 12.00 R1.4 13.45 ≤ ≤ × Lim Holt [30] 10.36 R1.4 12.87 ≤ ≤ ÷ De et al. [364] 8.9 R1.4 13.2 ≤ ≤ + Malik et al. [267] 11.82 R1.4 13.72 ≤ ≤ ∗ LIGO/Virgo [314] 10.5 RGW170817 13.3 ≤ ≤ ? Tews et al. [298] 11.3 R1.4 12.1 ≤ ≤ K¨oppel et al. [388] 10.92 R1.4 ≤ ∗ Raithel [666] 9.8 RGW170817 13.2 With a phase transition ≤ ≤ Annala et al. [420] 9.9 R1.4 13.6 ≤ ≤ × Most et al. [532] 8.53 R1.4 13.74 ≤ ≤ ? Tews et al. [298] 9.2 R1.4 12.5 ≤ ≤ × Montana et al. [288] 10.1 R1.4 13.11 From multimessanger analyses ≤ ≤ ∗ † Radice Dai [489] 11.4 R1.4 13.2 ≤ ≤ ∗ † Coughlin et al. [491] 11.1 R1.4 13.4 ≤ ≤ ∗ ‡ Kumar Landry [360] 9.4 R1.4 12.8 ≤ ≤ ◦ This work gives only an estimate for the radius R1.6 of an 1.6M star. However, I still include it in this table because it was probably the first work to give an estimate of NS radii based on the observation of GW170817. × At 2 σ confidence level. ÷ At 90% confidence level and including an estimate of systematic error. + The lower limit is based on results from Ref. [488], related to macronova emission. ∗ At 90% confidence level. ? Total envelopes with GW input at 90% confidence level. † These authors suggest an additional systematic uncertainty of 0.2km on these values. ‡ This work combined the GW data with electromagnetic observations of other NSs in accreting low-mass x-ray binaries (from Ref. [554]). true or not, since the new observation run of the LIGO-Virgo Collaborations has already started to give us new datasets on BNS mergers, which will definitively tell us more about the EoS of compact stars, including, perhaps, about possible phase transitions at supranuclear density. While the GW170817 event was very fortunate because of its proximity to us, it can be expected that future observations will not all have such a high signal-to-noise ratio, but the increasing sensitivity of advanced GW detectors over the next years will surely lead to a large number of detections of merging BNS systems (if we are lucky, we may also observe the merger of an eccentric NS binary) [671, 227] and, together with accurate electromagnetic observations related and unrelated to GW events, will allow for an even more precise measurement of the supranuclear EoSs. It has been estimated that a number of detections (at 100Mpc) of order 10 will set percent limits on NS radii and 10% limits on the EoS [493, 254, 31]. The techniques reviewed in this article will continue to be used in the data analyses and will of course undergo further developments. Numerical simulations, in addition to delivering waveforms with higher and higher accuracy for constructing improved approximants, may also at some point finally provide us with more realistic hints on the evolution of magnetic fields and radiation after the merger. Semianalytic techniques that provide the actual templates for the data analyses are also continuously

48 improving. This is all really exciting.

9 Acknowledgements

I would like to thank all the colleagues who gave me valuable suggestions on how to improve this review. Partial support has come from JSPS Grant-in-Aid for Scientific Research (C) No. T18K036220.

References

[1] The LIGO Scientific Collaboration, The Virgo Collaboration, Gw170817: Observation of grav- itational waves from a binary neutron star inspiral, Phys. Rev. Lett. 119 (16) (2017) 161101. arXiv:1710.05832, doi:10.1103/PhysRevLett.119.161101.

[2] T. Hinderer, S. Nissanke, F. Foucart, K. Hotokezaka, T. Vincent, M. Kasliwal, P. Schmidt, A. R. Williamson, D. Nichols, M. Duez, L. E. Kidder, H. P. Pfeiffer, M. A. Scheel, Discerning the binary neutron star or neutron star-black hole nature of GW170817 with and Electromagnetic Measurements, arXiv e-prints (2018) arXiv:1808.03836arXiv:1808.03836.

[3] G. Baym, T. Hatsuda, T. Kojo, P. D. Powell, Y. Song, T. Takatsuka, From hadrons to quarks in neutron stars: a review, Reports on Progress in Physics 81 (5) (2018) 056902. arXiv:1707.04966, doi:10.1088/1361-6633/aaae14.

[4] The LIGO Scientific Collaboration, the Virgo Collaboration, B. P. Abbott, R. Abbott, T. D. Abbott, F. Acernese, K. Ackley, C. aAdams, T. Adams, P. Addesso, R. X. Adhikari, V. B. Adya, et al., Multi-messenger observations of a binary , Astrophys. J. Lett. 848 (2) (2017) L12. arXiv:1710.05833, doi:10.3847/2041-8213/aa91c9.

[5] B. P. Abbott, R. Abbott, T. D. Abbott, F. Acernese, K. Ackley, C. Adams, T. Adams, P. Addesso, R. X. Adhikari, V. B. Adya, et al., Estimating the Contribution of Dynamical Ejecta in the Kilonova Associated with GW170817, Astrophys. J. Lett. 850 (2) (2017) L39. arXiv:1710.05836, doi:10.3847/2041-8213/aa9478.

[6] K. D. Alexander, E. Berger, W. Fong, P. K. G. Williams, C. Guidorzi, R. Margutti, B. D. Metzger, J. Annis, P. K. Blanchard, D. Brout, D. A. Brown, H.-Y. Chen, R. Chornock, P. S. Cowperthwaite, M. Drout, T. Eftekhari, J. Frieman, D. E. Holz, M. Nicholl, A. Rest, M. Sako, M. Soares-Santos, V. A. Villar, The Electromagnetic Counterpart of the Binary Neutron Star Merger LIGO/Virgo GW170817. VI. Radio Constraints on a Relativistic Jet and Predictions for Late-time Emission from the Kilonova Ejecta, Astrophys. J. Letters 848 (2) (2017) L21. arXiv:1710.05457, doi: 10.3847/2041-8213/aa905d.

[7] I. Arcavi, G. Hosseinzadeh, D. A. Howell, C. McCully, D. Poznanski, D. Kasen, J. Barnes, M. Zaltzman, S. Vasylyev, D. Maoz, S. Valenti, Optical emission from a kilonova following a gravitational-wave-detected neutron-star merger, Nature 551 (2017) 64–66. arXiv:1710.05843, doi:10.1038/nature24291.

[8] I. Arcavi, C. McCully, G. Hosseinzadeh, D. A. Howell, S. Vasylyev, D. Poznanski, M. Zaltzman, D. Maoz, L. Singer, S. Valenti, D. Kasen, J. Barnes, T. Piran, W.-f. Fong, Optical Follow-up of Gravitational-wave Events with Las Cumbres Observatory, Astrophys. J. Lett. 848 (2) (2017) L33. arXiv:1710.05842, doi:10.3847/2041-8213/aa910f.

49 [9] I. Arcavi, The First Hours of the GW170817 Kilonova and the Importance of Early Optical and Ultraviolet Observations for Constraining Emission Models, Astrophys. J. Lett. 855 (2) (2018) L23. arXiv:1802.02164, doi:10.3847/2041-8213/aab267. [10] R. Chornock, E. Berger, D. Kasen, P. S. Cowperthwaite, M. Nicholl, V. A. Villar, K. D. Alexander, P. K. Blanchard, T. Eftekhari, W. Fong, R. Margutti, P. K. G. Williams, J. Annis, D. Brout, D. A. Brown, H.-Y. Chen, M. R. Drout, B. Farr, R. J. Foley, J. A. Frieman, C. L. Fryer, K. Herner, D. E. Holz, R. Kessler, T. Matheson, B. D. Metzger, E. Quataert, A. Rest, M. Sako, D. M. Scolnic, N. Smith, M. Soares-Santos, The Electromagnetic Counterpart of the Binary Neutron Star Merger LIGO/Virgo GW170817. IV. Detection of Near-infrared Signatures of r-process Nucleosynthesis with -South, Astrophys. J. Letters 848 (2) (2017) L19. arXiv:1710.05454, doi:10.3847/ 2041-8213/aa905c. [11] D. A. Coulter, R. J. Foley, C. D. Kilpatrick, M. R. Drout, A. L. Piro, B. J. Shappee, M. R. Siebert, J. D. Simon, N. Ulloa, D. Kasen, B. F. Madore, A. Murguia-Berthier, Y.-C. Pan, J. X. Prochaska, E. Ramirez-Ruiz, A. Rest, C. Rojas-Bravo, Swope Survey 2017a (SSS17a), the optical counterpart to a gravitational wave source, Science 358 (2017) 1556–1558. arXiv:1710.05452, doi:10.1126/science.aap9811. [12] S. Covino, K. Wiersema, Y. Z. Fan, K. Toma, A. B. Higgins, A. Melandri, P. D’Avanzo, C. G. Mundell, R. A. M. J. Wijers, The unpolarized macronova associated with the gravitational wave event GW 170817, Nature 1 (2017) 791–794. arXiv:1710.05849, doi:10.1038/ s41550-017-0285-z. [13] P. S. Cowperthwaite, E. Berger, V. A. Villar, B. D. Metzger, M. Nicholl, R. Chornock, P. K. Blan- chard, W. Fong, R. Margutti, M. Soares-Santos, K. D. Alexander, S. Allam, J. Annis, D. Brout, D. A. Brown, R. E. Butler, H.-Y. Chen, H. T. Diehl, Z. Doctor, M. R. Drout, T. Eftekhari, B. Farr, D. A. Finley, R. J. Foley, J. A. Frieman, C. L. Fryer, J. Garc´ıa-Bellido,M. S. S. Gill, J. Guil- lochon, K. Herner, D. E. Holz, D. Kasen, R. Kessler, J. Marriner, T. Matheson, E. H. Neilsen, Jr., E. Quataert, A. Palmese, A. Rest, M. Sako, D. M. Scolnic, N. Smith, D. L. Tucker, P. K. G. Williams, E. Balbinot, J. L. Carlin, E. R. Cook, F. Durret, T. S. Li, P. A. A. Lopes, A. C. C. Louren¸co,J. L. Marshall, G. E. Medina, J. Muir, R. R. Mu˜noz,M. Sauseda, D. J. Schlegel, L. F. Secco, A. K. Vivas, W. Wester, A. Zenteno, Y. Zhang, T. M. C. Abbott, M. Banerji, K. Bech- tol, A. Benoit-L´evy,E. Bertin, E. Buckley-Geer, D. L. Burke, D. Capozzi, A. Carnero Rosell, M. Carrasco Kind, F. J. Castander, M. Crocce, C. E. Cunha, C. B. D’Andrea, L. N. da Costa, C. Davis, D. L. DePoy, S. Desai, J. P. Dietrich, A. Drlica-Wagner, T. F. Eifler, A. E. Evrard, E. Fernandez, B. Flaugher, P. Fosalba, E. Gaztanaga, D. W. Gerdes, T. Giannantonio, D. A. Goldstein, D. Gruen, R. A. Gruendl, G. Gutierrez, K. Honscheid, B. Jain, D. J. James, T. Jel- tema, M. W. G. Johnson, M. D. Johnson, S. Kent, E. Krause, R. Kron, K. Kuehn, N. Nuropatkin, O. Lahav, M. Lima, H. Lin, M. A. G. Maia, M. March, P. Martini, R. G. McMahon, F. Menan- teau, C. J. Miller, R. Miquel, J. J. Mohr, E. Neilsen, R. C. Nichol, R. L. C. Ogando, A. A. Plazas, N. Roe, A. K. Romer, A. Roodman, E. S. Rykoff, E. Sanchez, V. Scarpine, R. Schindler, M. Schubnell, I. Sevilla-Noarbe, M. Smith, R. C. Smith, F. Sobreira, E. Suchyta, M. E. C. Swanson, G. Tarle, D. Thomas, R. C. Thomas, M. A. Troxel, V. Vikram, A. R. Walker, R. H. Wechsler, J. Weller, B. Yanny, J. Zuntz, The Electromagnetic Counterpart of the Binary Neu- tron Star Merger LIGO/Virgo GW170817. II. UV, Optical, and Near-infrared Light Curves and Comparison to Kilonova Models, Astrophys. J. Lett. 848 (2) (2017) L17. arXiv:1710.05840, doi:10.3847/2041-8213/aa8fc7. [14] M. R. Drout, A. L. Piro, B. J. Shappee, C. D. Kilpatrick, J. D. Simon, C. Contreras, D. A. Coulter, R. J. Foley, M. R. Siebert, N. Morrell, K. Boutsia, Z. Wan, D. D. Whitten, Light curves

50 of the neutron star merger GW170817/SSS17a: Implications for r-process nucleosynthesis, Science 358 (2017) 1570–1574. arXiv:1710.05443, doi:10.1126/science.aaq0049.

[15] P. A. Evans, S. B. Cenko, J. A. Kennea, S. W. K. Emery, N. P. M. Kuin, O. Korobkin, R. T. Wollaeger, G. Tagliaferri, N. R. Tanvir, A. Tohuvavohu, Swift and NuSTAR observations of GW170817: Detection of a blue kilonova, Science 358 (2017) 1565–1570. arXiv:1710.05437, doi:10.1126/science.aap9580.

[16] A. Goldstein, P. Veres, E. Burns, M. S. Briggs, R. Hamburg, D. Kocevski, C. A. Wilson-Hodge, R. D. Preece, S. Poolakkil, O. J. Roberts, C. M. Hui, V. Connaughton, J. Racusin, A. von Kienlin, T. Dal Canton, N. Christensen, T. Littenberg, K. Siellez, L. Blackburn, J. Broida, E. Bissaldi, W. H. Cleveland, M. H. Gibby, M. M. Giles, R. M. Kippen, S. McBreen, J. McEnery, C. A. Meegan, W. S. Paciesas, M. Stanbro, An Ordinary Short Gamma-Ray Burst with Extraordinary Implications: Fermi-GBM Detection of GRB 170817A, Astrophys. J. Letters 848 (2) (2017) L14. arXiv:1710.05446, doi:10.3847/2041-8213/aa8f41.

[17] G. Hallinan, A. Corsi, K. P. Mooley, K. Hotokezaka, E. Nakar, M. M. Kasliwal, D. L. Kaplan, D. A. Frail, S. T. Myers, T. Murphy, K. De, D. Dobie, J. R. Allison, K. W. Bannister, V. Bhalerao, P. Chandra, T. E. Clarke, S. Giacintucci, A. Y. Q. Ho, A. Horesh, N. E. Kassim, S. R. Kulkarni, E. Lenc, F. J. Lockman, C. Lynch, D. Nichols, S. Nissanke, N. Palliyaguru, W. M. Peters, T. Piran, J. Rana, E. M. Sadler, L. P. Singer, A radio counterpart to a neutron star merger, Science 358 (2017) 1579–1583. arXiv:1710.05435, doi:10.1126/science.aap9855.

[18] M. M. Kasliwal, E. Nakar, L. P. Singer, D. L. Kaplan, D. O. Cook, A. Van Sistine, R. M. Lau, C. Fremling, O. Gottlieb, J. E. Jencson, S. M. Adams, U. Feindt, K. Hotokezaka, S. Ghosh, D. A. Perley, P.-C. Yu, T. Piran, J. R. Allison, G. C. Anupama, A. Balasubramanian, K. W. Bannister, J. Bally, J. Barnes, S. a. Barway, Illuminating gravitational waves: A concordant picture of from a neutron star merger, Science 358 (2017) 1559–1565. arXiv:1710.05436, doi:10.1126/science.aap9455.

[19] R. Margutti, E. Berger, W. Fong, C. Guidorzi, K. D. Alexander, B. D. Metzger, P. K. Blanchard, P. S. Cowperthwaite, R. Chornock, T. Eftekhari, M. Nicholl, V. A. Villar, P. K. G. Williams, J. Annis, D. A. Brown, H. Chen, Z. Doctor, J. A. Frieman, D. E. Holz, M. Sako, M. Soares-Santos, The Electromagnetic Counterpart of the Binary Neutron Star Merger LIGO/Virgo GW170817. V. Rising X-Ray Emission from an Off-axis Jet, Astrophys. J. Letters 848 (2) (2017) L20. arXiv: 1710.05431, doi:10.3847/2041-8213/aa9057.

[20] A. Murguia-Berthier, E. Ramirez-Ruiz, C. D. Kilpatrick, R. J. Foley, D. Kasen, W. H. Lee, A. L. Piro, D. A. Coulter, M. R. Drout, B. F. Madore, B. J. Shappee, Y.-C. Pan, J. X. Prochaska, A. Rest, C. Rojas-Bravo, M. R. Siebert, J. D. Simon, A Neutron Star Binary Merger Model for GW170817/GRB 170817A/SSS17a, Astrophys. J. Lett. 848 (2) (2017) L34. arXiv:1710.05453, doi:10.3847/2041-8213/aa91b3.

[21] M. Nicholl, E. Berger, D. Kasen, B. D. Metzger, J. Elias, C. Brice˜no, K. D. Alexander, P. K. Blanchard, R. Chornock, P. S. Cowperthwaite, T. Eftekhari, W. Fong, R. Margutti, V. A. Villar, P. K. G. Williams, W. Brown, J. Annis, A. Bahramian, D. Brout, D. A. Brown, H.-Y. Chen, J. C. Clemens, E. Dennihy, B. Dunlap, D. E. Holz, E. Marchesini, F. Massaro, N. Moskowitz, I. Pelisoli, A. Rest, F. Ricci, M. Sako, M. Soares-Santos, J. Strader, The Electromagnetic Counterpart of the Binary Neutron Star Merger LIGO/Virgo GW170817. III. Optical and UV Spectra of a Blue Kilonova from Fast Polar Ejecta, Astrophys. J. Letters 848 (2) (2017) L18. arXiv:1710.05456, doi:10.3847/2041-8213/aa9029.

51 [22] E. Pian, P. D’Avanzo, S. Benetti, M. Branchesi, E. Brocato, S. Campana, E. Cappellaro, S. Covino, V. D’Elia, J. P. U. Fynbo, F. Getman, G. Ghirlanda, G. Ghisellini, A. Grado, G. Greco, J. Hjorth, C. Kouveliotou, A. Levan, L. Limatola, D. Malesani, P. A. Mazzali, A. Melandri, P. Møller, S. D. Vergani, D. Vergani, Spectroscopic identification of r-process nu- cleosynthesis in a double neutron-star merger, Nature 551 (2017) 67–70. arXiv:1710.05858, doi:10.1038/nature24298. [23] V. Savchenko, C. Ferrigno, E. Kuulkers, A. Bazzano, E. Bozzo, S. Brandt, J. Chenevez, T. J.-L. Courvoisier, R. Diehl, A. Domingo, L. Hanlon, E. Jourdain, A. von Kienlin, P. Laurent, F. Lebrun, A. Lutovinov, A. Martin-Carrillo, S. Mereghetti, L. Natalucci, J. Rodi, J.-P. Roques, R. Sunyaev, P. Ubertini, INTEGRAL Detection of the First Prompt Gamma-Ray Signal Coincident with the Gravitational-wave Event GW170817, Astrophys. J. Letters 848 (2017) L15. arXiv:1710.05449, doi:10.3847/2041-8213/aa8f94. [24] S. Smartt, T. e. a. Chen, A kilonova as the electromagnetic counterpart to a gravitational-wave source, Nature 551 (2017) 75–79. arXiv:1710.05841, doi:10.1038/nature24303. [25] M. Soares-Santos, D. E. Holz, J. Annis, R. Chornock, K. Herner, E. Berger, D. Brout, H.-Y. Chen, R. Kessler, M. Sako, S. Allam, D. L. Tucker, R. E. Butler, A. Palmese, Z. Doctor, H. T. Diehl, J. Frieman, B. Yanny, H. Lin, D. Scolnic, P. Cowperthwaite, E. Neilsen, J. Marriner, Survey, Dark Energy Camera GW-EM Collaboration, The Electromagnetic Counterpart of the Binary Neutron Star Merger LIGO/Virgo GW170817. I. Discovery of the Optical Counterpart Using the Dark Energy Camera, Astrophys. J. Letters 848 (2) (2017) L16. arXiv:1710.05459, doi:10.3847/2041-8213/aa9059. [26] M. Tanaka, Y. Utsumi, P. A. Mazzali, N. Tominaga, M. Yoshida, Y. Sekiguchi, T. Morokuma, K. Motohara, K. Ohta, K. S. Kawabata, F. Abe, K. Aoki, Y. Asakura, S. Baar, S. Barway, I. A. Bond, M. Doi, T. Fujiyoshi, H. Furusawa, S. Honda, Y. Itoh, M. Kawabata, N. Kawai, J. H. Kim, C.-H. Lee, S. Miyazaki, K. Morihana, H. Nagashima, T. Nagayama, T. Nakaoka, F. Nakata, R. Ohsawa, T. Ohshima, H. Okita, T. Saito, T. Sumi, A. Tajitsu, J. Takahashi, M. Takayama, Y. Tamura, I. Tanaka, T. Terai, P. J. Tristram, N. Yasuda, T. Zenko, Kilonova from post-merger ejecta as an optical and near-Infrared counterpart of GW170817, Public. Astron. Soc. of Japan 69 (2017) 102. arXiv:1710.05850, doi:10.1093/pasj/psx121. [27] N. R. Tanvir, A. J. Levan, C. Gonz´alez-Fern´andez,O. Korobkin, I. Mandel, S. Rosswog, J. Hjorth, P. D’Avanzo, A. S. Fruchter, C. L. Fryer, T. Kangas, B. Milvang-Jensen, S. Rosetti, D. Steeghs, R. T. Wollaeger, Z. Cano, C. M. Copperwheat, R. A. M. J. Wijers, The Emergence of a Lanthanide-rich Kilonova Following the Merger of Two Neutron Stars, Astrophys. J. Letters 848 (2) (2017) L27. arXiv:1710.05455, doi:10.3847/2041-8213/aa90b6. [28] E. Troja, L. Piro, H. van Eerten, R. T. Wollaeger, M. Im, O. D. Fox, N. R. Butler, S. B. Cenko, T. Sakamoto, C. L. Fryer, R. Ricci, A. Lien, S. Veilleux, M. H. Wieringa, Y. Yoon, The X-ray counterpart to the gravitational-wave event GW170817, Nature 551 (2017) 71–74. arXiv:1710.05433, doi:10.1038/nature24290. [29] N.-B. Zhang, B.-A. Li, J. Xu, Combined Constraints on the Equation of State of Dense Neutron- rich Matter from Terrestrial Nuclear Experiments and Observations of Neutron Stars, Astrophys. J. 859 (2) (2018) 90. arXiv:1801.06855, doi:10.3847/1538-4357/aac027. [30] Y. Lim, J. W. Holt, Neutron Star Tidal Deformabilities Constrained by Nuclear Theory and Experiment, Phys. Rev. Lett. 121 (6) (2018) 062701. arXiv:1803.02803, doi:10.1103/ PhysRevLett.121.062701.

52 [31] I. Tews, J. Margueron, S. Reddy, Confronting gravitational-wave observations with modern nu- clear physics constraints, arXiv e-printsarXiv:1901.09874.

[32] LIGO Scientific Collaboration, Virgo Collaboration, Gamma-Ray Burst Monitor, INTEGRAL, B. P. Abbott, R. Abbott, T. D. Abbott, F. Acernese, K. Ackley, C. Adams, T. Adams, P. Addesso, R. X. Adhikari, V. B. Adya, et al., Gravitational waves and gamma-rays from a binary neutron star merger: Gw170817 and grb 170817a, Astrophys. J. Lett. 848 (2) (2017) L13. arXiv:1710.05834, doi:10.3847/2041-8213/aa920c.

[33] I. I. Shapiro, Fourth Test of General Relativity, Letters 13 (26) (1964) 789–791. doi:10.1103/PhysRevLett.13.789.

[34] T. Baker, E. Bellini, P. G. Ferreira, M. Lagos, J. Noller, I. Sawicki, Strong Constraints on Cos- mological Gravity from GW170817 and GRB 170817A, Phys. Rev. Lett. 119 (25) (2017) 251301. arXiv:1710.06394, doi:10.1103/PhysRevLett.119.251301.

[35] P. Creminelli, F. Vernizzi, Dark Energy after GW170817 and GRB170817A, Phys. Rev. Lett. 119 (25) (2017) 251302. arXiv:1710.05877, doi:10.1103/PhysRevLett.119.251302.

[36] J. M. Ezquiaga, M. Zumalac´arregui, Dark Energy After GW170817: Dead Ends and the Road Ahead, Phys. Rev. Lett. 119 (25) (2017) 251304. arXiv:1710.05901, doi:10.1103/ PhysRevLett.119.251304.

[37] J. Sakstein, B. Jain, Implications of the Neutron Star Merger GW170817 for Cosmological Scalar- Tensor Theories, Phys. Rev. Lett. 119 (25) (2017) 251303. arXiv:1710.05893, doi:10.1103/ PhysRevLett.119.251303.

[38] C. de Rham, S. Melville, Gravitational Rainbows: LIGO and Dark Energy at its Cutoff, Phys. Rev. Lett. 121 (22) (2018) 221101. arXiv:1806.09417, doi:10.1103/PhysRevLett.121.221101.

[39] P. Creminelli, M. Lewandowski, G. Tambalo, F. Vernizzi, Gravitational wave decay into dark energy, Journal of Cosmology and Astroparticle Physics 2018 (12) (2018) 025. arXiv:1809.03484, doi:10.1088/1475-7516/2018/12/025.

[40] P. Creminelli, G. Tambalo, F. Vernizzi, V. Yingcharoenrat, Resonant Decay of Gravitational Waves into Dark Energy, arXiv e-prints (2019) arXiv:1906.07015arXiv:1906.07015.

[41] J. M. Lattimer, The Nuclear Equation of State and Neutron Star Masses, Annual Review of Nuclear and Particle Science 62 (2012) 485–515. doi:10.1146/annurev-nucl-102711-095018.

[42] P. Danielewicz, R. Lacey, W. G. Lynch, Determination of the Equation of State of Dense Matter, Science 298 (2002) 1592–1596. arXiv:nucl-th/0208016, doi:10.1126/science.1078070.

[43] A. W. Steiner, M. Prakash, J. M. Lattimer, P. J. Ellis, Isospin asymmetry in nuclei and neutron stars [review article], Physics Reports 411 (2005) 325–375. arXiv:nucl-th/0410066, doi:10. 1016/j.physrep.2005.02.004.

[44] B.-A. Li, L.-W. Chen, C. M. Ko, Recent progress and new challenges in isospin physics with heavy-ion reactions, Physics Reports 464 (2008) 113–281. arXiv:0804.3580, doi:10.1016/j. physrep.2008.04.005.

53 [45] S. Abrahamyan, Z. Ahmed, H. Albataineh, K. Aniol, D. S. Armstrong, W. Armstrong, T. Averett, B. Babineau, A. Barbieri, V. Bellini, R. Beminiwattha, J. Benesch, F. Benmokhtar, T. Bielarski, W. Boeglin, A. Camsonne, M. Canan, P. Carter, G. D. Cates, C. Chen, J.-P. Chen, O. Hen, F. Cu- sanno, M. M. Dalton, R. de Leo, K. de Jager, W. Deconinck, P. Decowski, X. Deng, A. Deur, D. Dutta, A. Etile, D. Flay, G. B. Franklin, M. Friend, S. Frullani, E. Fuchey, F. Garibaldi, E. Gasser, R. Gilman, A. Giusa, A. Glamazdin, J. Gomez, J. Grames, C. Gu, O. Hansen, J. Hansknecht, D. W. Higinbotham, R. S. Holmes, T. Holmstrom, C. J. Horowitz, J. Hoskins, J. Huang, C. E. Hyde, F. Itard, C.-M. Jen, E. Jensen, G. Jin, S. Johnston, A. Kelleher, K. Kli- akhandler, P. M. King, S. Kowalski, K. S. Kumar, J. Leacock, J. Leckey, IV, J. H. Lee, J. J. Lerose, R. Lindgren, N. Liyanage, N. Lubinsky, J. Mammei, F. Mammoliti, D. J. Margaziotis, P. Markowitz, A. McCreary, D. McNulty, L. Mercado, Z.-E. Meziani, R. W. Michaels, M. Mi- hovilovic, N. Muangma, C. Mu˜noz-Camacho, S. Nanda, V. Nelyubin, N. Nuruzzaman, Y. Oh, A. Palmer, D. Parno, K. D. Paschke, S. K. Phillips, B. Poelker, R. Pomatsalyuk, M. Posik, A. J. R. Puckett, B. Quinn, A. Rakhman, P. E. Reimer, S. Riordan, P. Rogan, G. Ron, G. Russo, K. Saenboonruang, A. Saha, B. Sawatzky, A. Shahinyan, R. Silwal, S. Sirca, K. Slifer, P. Solvi- gnon, P. A. Souder, M. L. Sperduto, R. Subedi, R. Suleiman, V. Sulkosky, C. M. Sutera, W. A. Tobias, W. Troth, G. M. Urciuoli, B. Waidyawansa, D. Wang, J. Wexler, R. Wilson, B. Wo- jtsekhowski, X. Yan, H. Yao, Y. Ye, Z. Ye, V. Yim, L. Zana, X. Zhan, J. Zhang, Y. Zhang, X. Zheng, P. Zhu, Measurement of the Neutron Radius of Pb208 through Parity Violation in Electron Scattering, Physical Review Letters 108 (11) (2012) 112502. arXiv:1201.2568, doi:10.1103/PhysRevLett.108.112502.

[46] C. J. Horowitz, Z. Ahmed, C.-M. Jen, A. Rakhman, P. A. Souder, M. M. Dalton, N. Liyanage, K. D. Paschke, K. Saenboonruang, R. Silwal, G. B. Franklin, M. Friend, B. Quinn, K. S. Kumar, D. McNulty, L. Mercado, S. Riordan, J. Wexler, R. W. Michaels, G. M. Urciuoli, Weak charge form factor and radius of 208Pb through parity violation in electron scattering, Physical Review C 85 (3) (2012) 032501. arXiv:1202.1468, doi:10.1103/PhysRevC.85.032501.

[47] M. B. Tsang, J. R. Stone, F. Camera, P. Danielewicz, S. Gandolfi, K. Hebeler, C. J. Horowitz, J. Lee, W. G. Lynch, Z. Kohley, R. Lemmon, P. M¨oller,T. Murakami, S. Riordan, X. Roca- Maza, F. Sammarruca, A. W. Steiner, I. Vida˜na,S. J. Yennello, Constraints on the symmetry energy and neutron skins from experiments and theory, Physical Review C 86 (1) (2012) 015803. arXiv:1204.0466, doi:10.1103/PhysRevC.86.015803.

[48] J. M. Lattimer, M. Prakash, The equation of state of hot, dense matter and neutron stars, Physics Reports 621 (2016) 127–164. arXiv:1512.07820, doi:10.1016/j.physrep.2015.12.005.

[49] A. L. Watts, N. Andersson, D. Chakrabarty, M. Feroci, K. Hebeler, G. Israel, F. K. Lamb, M. C. Miller, S. Morsink, F. Ozel,¨ A. Patruno, J. Poutanen, D. Psaltis, A. Schwenk, A. W. Steiner, L. Stella, L. Tolos, M. van der Klis, Colloquium: Measuring the neutron star equation of state using x-ray timing, Reviews of Modern Physics 88 (2) (2016) 021001. arXiv:1602.01081, doi:10.1103/RevModPhys.88.021001.

[50] F. Ozel,¨ P. Freire, Masses, Radii, and the Equation of State of Neutron Stars, Annual Re- view of Astronomy and Astrophysics 54 (2016) 401–440. arXiv:1603.02698, doi:10.1146/ annurev-astro-081915-023322.

[51] M. Oertel, M. Hempel, T. Kl¨ahn,S. Typel, Equations of state for supernovae and compact stars, Reviews of Modern Physics 89 (1) (2017) 015007. arXiv:1610.03361, doi:10.1103/RevModPhys. 89.015007.

54 [52] M. B. Tsang, T. X. Liu, L. Shi, P. Danielewicz, C. K. Gelbke, X. D. Liu, W. G. Lynch, W. P. Tan, G. Verde, A. Wagner, H. S. Xu, W. A. Friedman, L. Beaulieu, B. Davin, R. T. de Souza, Y. Larochelle, T. Lefort, R. Yanez, V. E. Viola, R. J. Charity, L. G. Sobotka, Isospin Diffusion and the Nuclear Symmetry Energy in Heavy Ion Reactions, Physical Review Letters 92 (6) (2004) 062701. doi:10.1103/PhysRevLett.92.062701.

[53] C. Fuchs, pp i Kaon production in heavy ion reactions at intermediate energies [review article], Progress in Particle and Nuclear Physics 56 (2006) 1–103. arXiv:nucl-th/0507017, doi:10. 1016/j.ppnp.2005.07.004.

[54] W. G. Lynch, M. B. Tsang, Y. Zhang, P. Danielewicz, M. Famiano, Z. Li, A. W. Steiner, Probing the symmetry energy with heavy ions, Progress in Particle and Nuclear Physics 62 (2009) 427–432. arXiv:0901.0412, doi:10.1016/j.ppnp.2009.01.001.

[55] M. B. Tsang, Y. Zhang, P. Danielewicz, M. Famiano, Z. Li, W. G. Lynch, A. W. Steiner, Con- straints on the Density Dependence of the Symmetry Energy, Physical Review Letters 102 (12) (2009) 122701. arXiv:0811.3107, doi:10.1103/PhysRevLett.102.122701.

[56] M. Kortelainen, T. Lesinski, J. Mor´e,W. Nazarewicz, J. Sarich, N. Schunck, M. V. Stoitsov, S. Wild, Nuclear energy density optimization, Physical Review C 82 (2) (2010) 024313. arXiv: 1005.5145, doi:10.1103/PhysRevC.82.024313.

[57] A. Tamii, I. Poltoratska, P. von Neumann-Cosel, Y. Fujita, T. Adachi, C. A. Bertulani, J. Carter, M. Dozono, H. Fujita, K. Fujita, K. Hatanaka, D. Ishikawa, M. Itoh, T. Kawa- bata, Y. Kalmykov, A. M. Krumbholz, E. Litvinova, H. Matsubara, K. Nakanishi, R. Neveling, H. Okamura, H. J. Ong, B. Ozel-Tashenov,¨ V. Y. Ponomarev, A. Richter, B. Rubio, H. Sak- aguchi, Y. Sakemi, Y. Sasamoto, Y. Shimbara, Y. Shimizu, F. D. Smit, T. Suzuki, Y. Tameshige, J. Wambach, R. Yamada, M. Yosoi, J. Zenihiro, Complete Electric Dipole Response and the Neutron Skin in Pb208, Physical Review Letters 107 (6) (2011) 062502. arXiv:1104.5431, doi:10.1103/PhysRevLett.107.062502.

[58] B. A. Brown, Constraints on the Skyrme Equations of State from Properties of Doubly Magic Nuclei, Physical Review Letters 111 (23) (2013) 232502. arXiv:1308.3664, doi:10.1103/ PhysRevLett.111.232502.

[59] A. Le F`evre,Y. Leifels, W. Reisdorf, J. Aichelin, C. Hartnack, Constraining the nuclear matter equation of state around twice saturation density, Nuclear Physics A 945 (2016) 112–133. arXiv: 1501.05246, doi:10.1016/j.nuclphysa.2015.09.015.

[60] P. Russotto, S. Gannon, S. Kupny, P. Lasko, L. Acosta, M. Adamczyk, A. Al-Ajlan, M. Al- Garawi, S. Al-Homaidhi, F. Amorini, L. Auditore, T. Aumann, Y. Ayyad, Z. Basrak, J. Benlliure, M. Boisjoli, K. Boretzky, J. Brzychczyk, A. Budzanowski, C. Caesar, G. Cardella, P. Cammarata, Z. Chajecki, M. Chartier, A. Chbihi, M. Colonna, M. D. Cozma, B. Czech, E. De Filippo, M. Di Toro, M. Famiano, I. Gaˇspari´c,L. Grassi, C. Guazzoni, P. Guazzoni, M. Heil, L. Heilborn, R. In- trozzi, T. Isobe, K. Kezzar, M. Kiˇs,A. Krasznahorkay, N. Kurz, E. La Guidara, G. Lanzalone, A. Le F`evre,Y. Leifels, R. C. Lemmon, Q. F. Li, I. Lombardo, J.Lukasik, W. G. Lynch, P. Marini, Z. Matthews, L. May, T. Minniti, M. Mostazo, A. Pagano, E. V. Pagano, M. Papa, P. Paw lowski, S. Pirrone, G. Politi, F. Porto, W. Reviol, F. Riccio, F. Rizzo, E. Rosato, D. Rossi, S. San- toro, D. G. Sarantites, H. Simon, I. Skwirczynska, Z. Sosin, L. Stuhl, W. Trautmann, A. Trifir`o, M. Trimarchi, M. B. Tsang, G. Verde, M. Veselsky, M. Vigilante, Y. Wang, A. Wieloch, P. Wigg, J. Winkelbauer, H. H. Wolter, P. Wu, S. Yennello, P. Zambon, L. Zetta, M. Zoric, Results of the

55 ASY-EOS experiment at GSI: The symmetry energy at suprasaturation density, Physical Review C 94 (3) (2016) 034608. arXiv:1608.04332, doi:10.1103/PhysRevC.94.034608. [61] M. C. Miller, F. K. Lamb, Observational constraints on neutron star masses and radii, European Physical Journal A 52 (2016) 63. arXiv:1604.03894, doi:10.1140/epja/i2016-16063-8. [62] K. C. Gendreau, Z. Arzoumanian, P. W. Adkins, C. L. Albert, J. F. Anders, A. T. Aylward, C. L. Baker, E. R. Balsamo, W. A. Bamford, S. S. Benegalrao, D. L. Berry, S. Bhalwani, J. K. Black, C. Blaurock, G. M. Bronke, G. L. Brown, J. G. Budinoff, J. D. Cantwell, T. Cazeau, P. T. Chen, T. G. Clement, A. T. Colangelo, J. S. Coleman, J. D. Coopersmith, W. E. Dehaven, J. P. Doty, M. D. Egan, T. Enoto, T. W. Fan, D. M. Ferro, R. Foster, N. M. Galassi, L. D. Gallo, C. M. Green, D. Grosh, K. Q. Ha, M. A. Hasouneh, K. B. Heefner, P. Hestnes, L. J. Hoge, T. M. Jacobs, J. L. Jørgensen, M. A. Kaiser, J. W. Kellogg, S. J. Kenyon, R. G. Koenecke, R. P. Kozon, B. LaMarr, M. D. Lambertson, A. M. Larson, S. Lentine, J. H. Lewis, M. G. Lilly, K. A. Liu, A. Malonis, S. S. Manthripragada, C. B. Markwardt, B. D. Matonak, I. E. Mcginnis, R. L. Miller, A. L. Mitchell, J. W. Mitchell, J. S. Mohammed, C. A. Monroe, K. M. Montt de Garcia, P. D. Mul´e,L. T. Nagao, S. N. Ngo, E. D. Norris, D. A. Norwood, J. Novotka, T. Okajima, L. G. Olsen, C. O. Onyeachu, H. Y. Orosco, J. R. Peterson, K. N. Pevear, K. K. Pham, S. E. Pollard, J. S. Pope, D. F. Powers, C. E. Powers, S. R. Price, G. Y. Prigozhin, J. B. Ramirez, W. J. Reid, R. A. Remillard, E. M. Rogstad, G. P. Rosecrans, J. N. Rowe, J. A. Sager, C. A. Sanders, B. Savadkin, M. R. Saylor, A. F. Schaeffer, N. S. Schweiss, S. R. Semper, P. J. Serlemitsos, L. V. Shackelford, Y. Soong, J. Struebel, M. L. Vezie, J. S. Villasenor, L. B. Winternitz, G. I. Wofford, M. R. Wright, M. Y. Yang, W. H. Yu, The Neutron star Interior Composition Explorer (NICER): design and development, in: Space Telescopes and Instrumentation 2016: Ultraviolet to , Vol. 9905 of Proceedings of the SPIE, 2016, p. 99051H. doi:10.1117/12.2231304. [63] Z. Arzoumanian, A. Brazier, S. Burke-Spolaor, S. Chamberlin, S. Chatterjee, B. Christy, J. M. Cordes, N. J. Cornish, F. Crawford, H. Thankful Cromartie, K. Crowter, M. E. DeCesar, P. B. Demorest, T. Dolch, J. A. Ellis, R. D. Ferdman, E. C. Ferrara, E. Fonseca, N. Garver-Daniels, P. A. Gentile, D. Halmrast, E. A. Huerta, F. A. Jenet, C. Jessup, G. Jones, M. L. Jones, D. L. Kaplan, M. T. Lam, T. J. W. Lazio, L. Levin, A. Lommen, D. R. Lorimer, J. Luo, R. S. Lynch, D. Madison, A. M. Matthews, M. A. McLaughlin, S. T. McWilliams, C. Mingarelli, C. Ng, D. J. Nice, T. T. Pennucci, S. M. Ransom, P. S. Ray, X. Siemens, J. Simon, R. Spiewak, I. H. Stairs, D. R. Stinebring, K. Stovall, J. K. Swiggum, S. R. Taylor, M. Vallisneri, R. van Haasteren, S. J. Vigeland, W. Zhu, NANOGrav Collaboration, The NANOGrav 11-year Data Set: High-precision Timing of 45 Millisecond Pulsars, Astrophys. J., Supp. 235 (2018) 37. arXiv:1801.01837, doi: 10.3847/1538-4365/aab5b0. [64] N. Degenaar, V. F. Suleimanov, Testing the equation of state of neutron stars with electromagnetic observations, in: L. Rezzolla, P. Pizzochero, D. I. Jones, N. Rea, I. Vida˜na(Eds.), The Physics and Astrophysics of Neutron Stars, Vol. 457 of Astrophysics and Space Science Library, 2018, pp. 185–253. arXiv:1806.02833, doi:10.1007/978-3-319-97616-7\_5. [65] A. L. Watts, W. Yu, J. Poutanen, S. Zhang, S. Bhattacharyya, S. Bogdanov, L. Ji, A. Pa- truno, T. E. Riley, P. Bakala, A. Baykal, F. Bernardini, I. Bombaci, E. Brown, Y. Cavec- chi, D. Chakrabarty, J. Chenevez, N. Degenaar, M. Del Santo, T. Di Salvo, V. Doroshenko, M. Falanga, R. D. Ferdman, M. Feroci, A. F. Gambino, M. Ge, S. K. Greif, S. Guillot, C. Gun- gor, D. H. Hartmann, K. Hebeler, A. Heger, J. Homan, R. Iaria, J. i. Zand, O. Kargaltsev, A. Kurkela, X. Lai, A. Li, X. Li, Z. Li, M. Linares, F. Lu, S. Mahmoodifar, M. M´endez, M. Coleman Miller, S. Morsink, J. N¨attil¨a,A. Possenti, C. Prescod-Weinstein, J. Qu, A. Rig- gio, T. Salmi, A. Sanna, A. Santangelo, H. Schatz, A. Schwenk, L. Song, E. Sr´amkov´a,B.ˇ Stap-

56 pers, H. Stiele, T. Strohmayer, I. Tews, L. Tolos, G. T¨or¨ok,D. Tsang, M. Urbanec, A. Vacchi, R. Xu, Y. Xu, S. Zane, G. Zhang, S. Zhang, W. Zhang, S. Zheng, X. Zhou, Dense matter with eXTP, Science China Physics, Mechanics, and Astronomy 62 (2019) 29503. arXiv:1812.04021, doi:10.1007/s11433-017-9188-4.

[66] B. D. Metzger, E. Berger, What is the Most Promising Electromagnetic Counterpart of a Neutron Star Binary Merger?, Astrophys. J. 746 (2012) 48. arXiv:1108.6056, doi:10.1088/0004-637X/ 746/1/48.

[67] M. Favata, Systematic Parameter Errors in Inspiraling Neutron Star Binaries, Phys. Rev. Lett. 112 (10) (2014) 101101. arXiv:1310.8288, doi:10.1103/PhysRevLett.112.101101.

[68] E.´ E.´ Flanagan, T. Hinderer, Constraining neutron-star tidal Love numbers with gravitational- wave detectors, Phys. Rev. D 77 (2) (2008) 021502. arXiv:0709.1915, doi:10.1103/PhysRevD. 77.021502.

[69] T. Hinderer, Tidal Love Numbers of Neutron Stars, Astrophys. J. 677 (2008) 1216–1220. arXiv: 0711.2420, doi:10.1086/533487.

[70] J. S. Read, C. Markakis, M. Shibata, K. Ury¯u,J. D. E. Creighton, J. L. Friedman, Measuring the neutron star equation of state with gravitational wave observations, Phys. Rev. D 79 (12) (2009) 124033. arXiv:0901.3258, doi:10.1103/PhysRevD.79.124033.

[71] J. S. Read, L. Baiotti, J. D. E. Creighton, J. L. Friedman, B. Giacomazzo, K. Kyutoku, C. Markakis, L. Rezzolla, M. Shibata, K. Taniguchi, Matter effects on binary neutron star waveforms, Phys. Rev. D 88 (4) (2013) 044042. arXiv:1306.4065, doi:10.1103/PhysRevD. 88.044042.

[72] L.-X. Li, B. Paczy´nski,Transient Events from Neutron Star Mergers, Astrophys. J. 507 (1998) L59–L62. arXiv:astro-ph/9807272, doi:10.1086/311680.

[73] S. R. Kulkarni, Modeling Supernova-like Explosions Associated with Gamma-ray Bursts with Short Durations, astro-ph/0510256arXiv:astro-ph/0510256.

[74] B. D. Metzger, G. Mart´ınez-Pinedo,S. Darbha, E. Quataert, A. Arcones, D. Kasen, R. Thomas, P. Nugent, I. V. Panov, N. T. Zinner, Electromagnetic counterparts of compact object mergers powered by the radioactive decay of r-process nuclei, Mon. Not. R. Astron. Soc. 406 (2010) 2650– 2662. arXiv:1001.5029, doi:10.1111/j.1365-2966.2010.16864.x.

[75] S.-Z. Li, L.-D. Liu, Y.-W. Yu, B. Zhang, What Powered the Optical Transient AT2017gfo Associated with GW170817?, Astrophys. J. Lett. 861 (2018) L12. arXiv:1804.06597, doi: 10.3847/2041-8213/aace61.

[76] S. Kisaka, K. Ioka, T. Nakamura, Isotropic Detectable X-Ray Counterparts to Gravitational Waves from Neutron Star Binary Mergers, Astrophys. J. Lett. 809 (2015) L8. arXiv:1506.02030, doi:10.1088/2041-8205/809/1/L8.

[77] The Open Kilonova Catagolue. [link]. URL https://kilonova.space/

[78] J. Guillochon, J. Parrent, L. Z. Kelley, R. Margutti, An Open Catalog for Supernova Data, Astrophys. J. 835 (2017) 64. arXiv:1605.01054, doi:10.3847/1538-4357/835/1/64.

57 [79] D. A. Perley, B. D. Metzger, J. Granot, N. R. Butler, T. Sakamoto, E. Ramirez-Ruiz, A. J. Levan, J. S. Bloom, A. A. Miller, A. Bunker, H.-W. Chen, A. V. Filippenko, N. Gehrels, K. Glazebrook, P. B. Hall, K. C. Hurley, D. Kocevski, W. Li, S. Lopez, J. Norris, A. L. Piro, D. Poznanski, J. X. Prochaska, E. Quataert, N. Tanvir, GRB 080503: Implications of a Naked Short Gamma-Ray Burst Dominated by Extended Emission, Astrophys. J. 696 (2009) 1871–1885. arXiv:0811.1044, doi:10.1088/0004-637X/696/2/1871. [80] E. Berger, W. Fong, R. Chornock, An r-process Kilonova Associated with the Short-hard GRB 130603B, Astrophys. J. 774 (2013) L23. arXiv:1306.3960, doi:10.1088/2041-8205/774/2/L23. [81] N. R. Tanvir, A. J. Levan, A. S. Fruchter, J. Hjorth, R. A. Hounsell, K. Wiersema, R. L. Tun- nicliffe, A ‘kilonova’ associated with the short-duration γ-ray burst GRB130603B, Nature 500 (2013) 547–549. arXiv:1306.4971, doi:10.1038/nature12505. [82] B. Yang, Z.-P. Jin, X. Li, S. Covino, X.-Z. Zheng, K. Hotokezaka, Y.-Z. Fan, T. Piran, D.-M. Wei, A possible macronova in the late afterglow of the long-short burst GRB 060614, Nature Communications 6 (2015) 7323. arXiv:1503.07761, doi:10.1038/ncomms8323. [83] Z.-P. Jin, X. Li, Z. Cano, S. Covino, Y.-Z. Fan, D.-M. Wei, The Light Curve of the Macronova Associated with the Long-Short Burst GRB 060614, Astrophys. J. Lett. 811 (2015) L22. arXiv: 1507.07206, doi:10.1088/2041-8205/811/2/L22. [84] Z.-P. Jin, K. Hotokezaka, X. Li, M. Tanaka, P. D’Avanzo, Y.-Z. Fan, S. Covino, D.-M. Wei, T. Piran, The Macronova in GRB 050709 and the GRB-macronova connection, Nature Commu- nications 7 (2016) 12898. arXiv:1603.07869, doi:10.1038/ncomms12898. [85] E. Berti, E. Barausse, V. Cardoso, L. Gualtieri, P. Pani, U. Sperhake, L. C. Stein, N. Wex, K. Yagi, T. Baker, C. P. Burgess, F. S. Coelho, D. Doneva, A. De Felice, P. G. Ferreira, P. C. C. Freire, J. Healy, C. Herdeiro, M. Horbatsch, B. Kleihaus, A. Klein, K. Kokko- tas, J. Kunz, P. Laguna, R. N. Lang, T. G. F. Li, T. Littenberg, A. Matas, S. Mirshekari, H. Okawa, E. Radu, R. O’Shaughnessy, B. S. Sathyaprakash, C. Van Den Broeck, H. A. Winther, H. Witek, M. Emad Aghili, J. Alsing, B. Bolen, L. Bombelli, S. Caudill, L. Chen, J. C. De- gollado, R. Fujita, C. Gao, D. Gerosa, S. Kamali, H. O. Silva, J. G. Rosa, L. Sadeghian, M. Sampaio, H. Sotani, M. Zilhao, Testing general relativity with present and future astrophys- ical observations, Classical and 32 (24) (2015) 243001. arXiv:1501.07274, doi:10.1088/0264-9381/32/24/243001. [86] L. Barack, V. Cardoso, S. Nissanke, T. P. Sotiriou, A. Askar, C. Belczynski, G. Bertone, E. Bon, D. Blas, R. Brito, et al., Black holes, gravitational waves and fundamental physics: a roadmap, Classical and Quantum Gravity 36 (14) (2019) 143001. arXiv:1806.05195, doi: 10.1088/1361-6382/ab0587. [87] N. Yunes, K. Yagi, F. Pretorius, Theoretical physics implications of the binary black-hole mergers GW150914 and GW151226, Phys. Rev. D 94 (8) (2016) 084002. arXiv:1603.08955, doi:10. 1103/PhysRevD.94.084002. [88] E. Barausse, C. Palenzuela, M. Ponce, L. Lehner, Neutron-star mergers in scalar-tensor theories of gravity, Phys. Rev. D 87 (8) (2013) 081506. arXiv:1212.5053, doi:10.1103/PhysRevD.87. 081506. [89] M. Shibata, K. Taniguchi, H. Okawa, A. Buonanno, Coalescence of binary neutron stars in a scalar-tensor theory of gravity, Phys. Rev. D 89 (8) (2014) 084005. arXiv:1310.0627, doi: 10.1103/PhysRevD.89.084005.

58 [90] C. Palenzuela, E. Barausse, M. Ponce, L. Lehner, Dynamical scalarization of neutron stars in scalar-tensor gravity theories, Phys. Rev. D 89 (4) (2014) 044024. arXiv:1310.4481, doi:10. 1103/PhysRevD.89.044024.

[91] L. Sampson, N. Yunes, N. Cornish, M. Ponce, E. Barausse, A. Klein, C. Palenzuela, L. Lehner, Projected constraints on scalarization with gravitational waves from neutron star binaries, Phys. Rev. D 90 (12) (2014) 124091. arXiv:1407.7038, doi:10.1103/PhysRevD.90.124091.

[92] K. Taniguchi, M. Shibata, A. Buonanno, Quasiequilibrium sequences of binary neutron stars undergoing dynamical scalarization, Phys. Rev. D 91 (2) (2015) 024033. arXiv:1410.0738, doi: 10.1103/PhysRevD.91.024033.

[93] M. Ponce, C. Palenzuela, E. Barausse, L. Lehner, Electromagnetic outflows in a class of scalar- tensor theories: Binary neutron star coalescence, Phys. Rev. D 91 (8) (2015) 084038. arXiv: 1410.0638, doi:10.1103/PhysRevD.91.084038.

[94] M. Bezares, C. Palenzuela, C. Bona, Final fate of compact boson star mergers, Phys. Rev. D 95 (12) (2017) 124005. arXiv:1705.01071, doi:10.1103/PhysRevD.95.124005.

[95] C. Palenzuela, P. Pani, M. Bezares, V. Cardoso, L. Lehner, S. Liebling, Gravitational wave signatures of highly compact boson star binaries, Phys. Rev. D 96 (10) (2017) 104058. arXiv: 1710.09432, doi:10.1103/PhysRevD.96.104058.

[96] L. Sagunski, J. Zhang, M. C. Johnson, L. Lehner, M. Sakellariadou, S. L. Liebling, C. Palenzuela, D. Neilsen, Neutron star mergers as a probe of modifications of general relativity with finite-range scalar forces, Phys. Rev. D 97 (6) (2018) 064016. arXiv:1709.06634, doi:10.1103/PhysRevD. 97.064016.

[97] M. Bezares, C. Palenzuela, Gravitational waves from dark boson star binary mergers, Classical and Quantum Gravity 35 (23) (2018) 234002. arXiv:1808.10732, doi:10.1088/1361-6382/aae87c.

[98] M. Alcubierre, Introduction to 3 + 1 Numerical Relativity, Oxford University Press, Oxford, UK, 2008. doi:10.1093/acprof:oso/9780199205677.001.0001.

[99] T. W. Baumgarte, S. L. Shapiro, Numerical Relativity: Solving Einstein’s Equations on the Computer, Cambridge University Press, Cambridge, UK, 2010. doi:10.1017/cbo9781139193344.

[100] E. Gourgoulhon, 3+1 Formalism in General Relativity, Vol. 846 of Lecture Notes in Physics, Berlin Springer Verlag, Springer, 2012. doi:10.1007/978-3-642-24525-1.

[101] L. Rezzolla, O. Zanotti, Relativistic Hydrodynamics, Oxford University Press, Oxford, UK, 2013. doi:10.1093/acprof:oso/9780198528906.001.0001.

[102] M. Shibata, Numerical Relativity, World Scientific, Singapore, 2016. doi:10.1142/9692.

[103] J. S. Read, B. D. Lackey, B. J. Owen, J. L. Friedman, Constraints on a phenomenologically parametrized neutron-star equation of state, Phys. Rev. D 79 (12) (2009) 124032. arXiv:0812. 2163, doi:10.1103/PhysRevD.79.124032.

[104] D. W. Neilsen, S. L. Liebling, M. Anderson, L. Lehner, E. O’Connor, C. Palenzuela, Magnetized neutron stars with realistic equations of state and neutrino cooling, Phys. Rev. D 89 (10) (2014) 104029. arXiv:1403.3680, doi:10.1103/PhysRevD.89.104029.

59 [105] C. Palenzuela, S. L. Liebling, D. Neilsen, L. Lehner, O. L. Caballero, E. O’Connor, M. An- derson, Effects of the microphysical equation of state in the mergers of magnetized neutron stars with neutrino cooling, Phys. Rev. D 92 (4) (2015) 044045. arXiv:1505.01607, doi: 10.1103/PhysRevD.92.044045.

[106] D. Radice, F. Galeazzi, J. Lippuner, L. F. Roberts, C. D. Ott, L. Rezzolla, Dynamical Mass Ejection from Binary Neutron Star Mergers, Mon. Not. R. Astron. Soc. 460 (2016) 3255–3271. arXiv:1601.02426, doi:10.1093/mnras/stw1227.

[107] A. Endrizzi, D. Logoteta, B. Giacomazzo, I. Bombaci, W. Kastaun, R. Ciolfi, Effects of chiral effective field theory equation of state on binary neutron star mergers, Phys. Rev. D 98 (4) (2018) 043015. arXiv:1806.09832, doi:10.1103/PhysRevD.98.043015.

[108] E. R. Most, L. J. Papenfort, V. Dexheimer, M. Hanauske, S. Schramm, H. St¨ocker, L. Rezzolla, Signatures of Quark-Hadron Phase Transitions in General-Relativistic Neutron-Star Mergers, Physical Review Letters 122 (6) (2019) 061101. arXiv:1807.03684, doi:10.1103/PhysRevLett. 122.061101.

[109] A. Bauswein, N.-U. F. Bastian, D. B. Blaschke, K. Chatziioannou, J. A. Clark, T. Fischer, M. Oertel, Identifying a First-Order Phase Transition in Neutron-Star Mergers through Grav- itational Waves, Physical Review Letters 122 (6) (2019) 061102. arXiv:1809.01116, doi: 10.1103/PhysRevLett.122.061102.

[110] A. Bauswein, S. Goriely, H.-T. Janka, Systematics of Dynamical Mass Ejection, Nucleosynthesis, and Radioactively Powered Electromagnetic Signals from Neutron-star Mergers, Astrophys. J. 773 (2013) 78. arXiv:1302.6530, doi:10.1088/0004-637X/773/1/78.

[111] S. Rosswog, T. Piran, E. Nakar, The multimessenger picture of compact object encounters: binary mergers versus dynamical collisions, Mon. Not. R. Astron. Soc. 430 (2013) 2585–2604. arXiv: 1204.6240, doi:10.1093/mnras/sts708.

[112] T. Piran, E. Nakar, S. Rosswog, The electromagnetic signals of compact binary mergers, Mon. Not. R. Astron. Soc. 430 (2013) 2121–2136. arXiv:1204.6242, doi:10.1093/mnras/stt037.

[113] S. Rosswog, O. Korobkin, A. Arcones, F.-K. Thielemann, T. Piran, The long-term evolution of neutron star merger remnants - I. The impact of r-process nucleosynthesis, Mon. Not. R. Astron. Soc. 439 (2014) 744–756. arXiv:1307.2939, doi:10.1093/mnras/stt2502.

[114] D. Grossman, O. Korobkin, S. Rosswog, T. Piran, The long-term evolution of neutron star merger remnants - II. Radioactively powered transients, Mon. Not. R. Astron. Soc. 439 (2014) 757–770. arXiv:1307.2943, doi:10.1093/mnras/stt2503.

[115] A. Perego, S. Rosswog, R. M. Cabez´on,O. Korobkin, R. K¨appeli, A. Arcones, M. Liebend¨orfer, Neutrino-driven winds from neutron star merger remnants, Mon. Not. R. Astron. Soc. 443 (2014) 3134–3156. arXiv:1405.6730, doi:10.1093/mnras/stu1352.

[116] S. Wanajo, Y. Sekiguchi, N. Nishimura, K. Kiuchi, K. Kyutoku, M. Shibata, Production of All the r-process Nuclides in the Dynamical Ejecta of Neutron Star Mergers, Astrophys. J. 789 (2014) L39. arXiv:1402.7317, doi:10.1088/2041-8205/789/2/L39.

[117] B. D. Metzger, A. Bauswein, S. Goriely, D. Kasen, Neutron-powered precursors of kilonovae, Mon. Not. R. Astron. Soc. 446 (2015) 1115–1120. arXiv:1409.0544, doi:10.1093/mnras/stu2225.

60 [118] O. Just, A. Bauswein, R. A. Pulpillo, S. Goriely, H.-T. Janka, Comprehensive nucleosynthesis analysis for ejecta of compact binary mergers, Mon. Not. R. Astron. Soc. 448 (2015) 541–567. arXiv:1406.2687, doi:10.1093/mnras/stv009.

[119] Y. Sekiguchi, K. Kiuchi, K. Kyutoku, M. Shibata, Dynamical mass ejection from binary neutron star mergers: Radiation-hydrodynamics study in general relativity, Phys. Rev. D 91 (6) (2015) 064059. arXiv:1502.06660, doi:10.1103/PhysRevD.91.064059.

[120] F. Foucart, R. Haas, M. D. Duez, E. O’Connor, C. D. Ott, L. Roberts, L. E. Kidder, J. Lippuner, H. P. Pfeiffer, M. A. Scheel, Low mass binary neutron star mergers: Gravita- tional waves and neutrino emission, Phys. Rev. D 93 (4) (2016) 044019. arXiv:1510.06398, doi:10.1103/PhysRevD.93.044019.

[121] O. Just, M. Obergaulinger, H.-T. Janka, A. Bauswein, N. Schwarz, Neutron-star Merger Ejecta as Obstacles to Neutrino-powered Jets of Gamma-Ray Bursts, Astrophys. J. Lett. 816 (2016) L30. arXiv:1510.04288, doi:10.3847/2041-8205/816/2/L30.

[122] L. Lehner, S. L. Liebling, C. Palenzuela, O. L. Caballero, E. O’Connor, M. Anderson, D. Neilsen, Unequal mass binary neutron star mergers and multimessenger signals, Classical and Quantum Gravity 33 (18) (2016) 184002. arXiv:1603.00501, doi:10.1088/0264-9381/33/18/184002.

[123] Y. Sekiguchi, K. Kiuchi, K. Kyutoku, M. Shibata, K. Taniguchi, Dynamical mass ejection from the merger of asymmetric binary neutron stars: Radiation-hydrodynamics study in general relativity, Phys. Rev. D 93 (12) (2016) 124046. arXiv:1603.01918, doi:10.1103/PhysRevD.93.124046.

[124] K. Hotokezaka, S. Wanajo, M. Tanaka, A. Bamba, Y. Terada, T. Piran, Radioactive decay prod- ucts in neutron star merger ejecta: heating efficiency and γ-ray emission, Mon. Not. R. Astron. Soc. 459 (2016) 35–43. arXiv:1511.05580, doi:10.1093/mnras/stw404.

[125] J. Barnes, D. Kasen, M.-R. Wu, G. Mart´ınez-Pinedo, Radioactivity and Thermalization in the Ejecta of Compact Object Mergers and Their Impact on Kilonova Light Curves, Astrophys. J. 829 (2016) 110. arXiv:1605.07218, doi:10.3847/0004-637X/829/2/110.

[126] L. Bovard, D. Martin, F. Guercilena, A. Arcones, L. Rezzolla, O. Korobkin, On r-process nucle- osynthesis from matter ejected in binary neutron star mergers, Phys. Rev. D 96 (2017) 124005. arXiv:1709.09630.

[127] M. Tanaka, K. Hotokezaka, Radiative Transfer Simulations of Neutron Star Merger Ejecta, As- trophys. J. 775 (2013) 113. arXiv:1306.3742, doi:10.1088/0004-637X/775/2/113.

[128] B. D. Metzger, R. Fern´andez,Red or blue? A potential kilonova imprint of the delay until black hole formation following a neutron star merger, Mon. Not. R. Astron. Soc. 441 (2014) 3444–3453. arXiv:1402.4803, doi:10.1093/mnras/stu802.

[129] A. Perego, D. Radice, S. Bernuzzi, AT 2017gfo: An Anisotropic and Three-component Kilonova Counterpart of GW170817, Astrophys. J. Lett. 850 (2017) L37. arXiv:1711.03982, doi:10. 3847/2041-8213/aa9ab9.

[130] M. Shibata, S. Fujibayashi, K. Hotokezaka, K. Kiuchi, K. Kyutoku, Y. Sekiguchi, M. Tanaka, Modeling GW170817 based on numerical relativity and its implications, Phys. Rev. D 96 (12) (2017) 123012. arXiv:1710.07579, doi:10.1103/PhysRevD.96.123012.

61 [131] K. Hotokezaka, K. Kiuchi, M. Shibata, E. Nakar, T. Piran, from the Fast Tail of Dynamical Ejecta of Neutron Star Mergers, Astrophys. J. 867 (2018) 95. arXiv: 1803.00599, doi:10.3847/1538-4357/aadf92. [132] M. Tanaka, D. Kato, G. Gaigalas, P. Rynkun, L. Radˇzi¯ut˙e,S. Wanajo, Y. Sekiguchi, N. Nakamura, H. Tanuma, I. Murakami, H. A. Sakaue, Properties of Kilonovae from Dynamical and Post- merger Ejecta of Neutron Star Mergers, Astrophys. J. 852 (2018) 109. arXiv:1708.09101, doi: 10.3847/1538-4357/aaa0cb. [133] P. Marronetti, S. L. Shapiro, Relativistic models for binary neutron stars with arbitrary spins, Phys. Rev. D 68 (10) (2003) 104024. arXiv:gr-qc/0306075, doi:10.1103/PhysRevD.68.104024. [134] T. W. Baumgarte, S. L. Shapiro, Formalism for the construction of binary neutron stars with arbitrary circulation, Phys. Rev. D 80 (6) (2009) 064009. arXiv:0909.0952, doi:10.1103/ PhysRevD.80.064009. [135] W. Tichy, Initial data for binary neutron stars with arbitrary spins, Phys. Rev. D 84 (2) (2011) 024041. arXiv:1107.1440, doi:10.1103/PhysRevD.84.024041. [136] W. Tichy, Constructing quasi-equilibrium initial data for binary neutron stars with arbitrary spins, Phys. Rev. D 86 (6) (2012) 064024. arXiv:1209.5336, doi:10.1103/PhysRevD.86.064024. [137] W. E. East, F. M. Ramazanoˇglu,F. Pretorius, Conformal thin-sandwich solver for generic ini- tial data, Phys. Rev. D 86 (10) (2012) 104053. arXiv:1208.3473, doi:10.1103/PhysRevD.86. 104053. [138] W. Kastaun, F. Galeazzi, D. Alic, L. Rezzolla, J. A. Font, Black hole from merging binary neutron stars: How fast can it spin?, Phys. Rev. D 88 (2) (2013) 021501. arXiv:1301.7348, doi:10.1103/PhysRevD.88.021501. [139] P. Tsatsin, P. Marronetti, Initial data for neutron star binaries with arbitrary spins, Phys. Rev. D 88 (6) (2013) 064060. arXiv:1303.6692, doi:10.1103/PhysRevD.88.064060. [140] S. Bernuzzi, T. Dietrich, W. Tichy, B. Br¨ugmann,Mergers of binary neutron stars with realis- tic spin, Phys. Rev. D 89 (10) (2014) 104021. arXiv:1311.4443, doi:10.1103/PhysRevD.89. 104021. [141] W. Kastaun, F. Galeazzi, Properties of hypermassive neutron stars formed in mergers of spinning binaries, Phys. Rev. D 91 (6) (2015) 064027. arXiv:1411.7975, doi:10.1103/PhysRevD.91. 064027. [142] A. Tsokaros, K. Ury¯u,L. Rezzolla, New code for quasiequilibrium initial data of binary neutron stars: Corotating, irrotational, and slowly spinning systems, Phys. Rev. D 91 (10) (2015) 104030. arXiv:1502.05674, doi:10.1103/PhysRevD.91.104030. [143] N. Tacik, F. Foucart, H. P. Pfeiffer, R. Haas, S. Ossokine, J. Kaplan, C. Muhlberger, M. D. Duez, L. E. Kidder, M. A. Scheel, B. Szil´agyi,Binary neutron stars with arbitrary spins in numerical relativity, Phys. Rev. D 92 (12) (2015) 124012. arXiv:1508.06986, doi:10.1103/PhysRevD.92. 124012. [144] T. Dietrich, N. Moldenhauer, N. K. Johnson-McDaniel, S. Bernuzzi, C. M. Markakis, B. Br¨ugmann,W. Tichy, Binary neutron stars with generic spin, eccentricity, mass ratio, and compactness: Quasi-equilibrium sequences and first , Phys. Rev. D 92 (12) (2015) 124007. arXiv:1507.07100, doi:10.1103/PhysRevD.92.124007.

62 [145] W. Kastaun, R. Ciolfi, A. Endrizzi, B. Giacomazzo, Structure of stable binary neutron star merger remnants: Role of initial spin, Phys. Rev. D 96 (4) (2017) 043019. arXiv:1612.03671, doi:10.1103/PhysRevD.96.043019.

[146] T. Dietrich, S. Bernuzzi, B. Br¨ugmann, M. Ujevic, W. Tichy, Numerical relativity simulations of precessing binary neutron star mergers, Phys. Rev. D 97 (6) (2018) 064002. arXiv:1712.02992, doi:10.1103/PhysRevD.97.064002.

[147] A. Tsokaros, K. Ury¯u,M. Ruiz, S. L. Shapiro, Constant circulation sequences of binary neutron stars and their spin characterization, Phys. Rev. D 98 (12) (2018) 124019. arXiv:1809.08237, doi:10.1103/PhysRevD.98.124019.

[148] M. Ruiz, S. L. Shapiro, A. Tsokaros, Jet launching from -neutron star mergers: Dependence on black hole spin, binary mass ratio, and magnetic field orientation, Phys. Rev. D 98 (12) (2018) 123017. arXiv:1810.08618, doi:10.1103/PhysRevD.98.123017.

[149] M. Ruiz, A. Tsokaros, V. Paschalidis, S. L. Shapiro, Effects of spin on magnetized binary neutron star mergers and jet launching, Phys. Rev. D 99 (8) (2019) 084032. arXiv:1902.08636, doi: 10.1103/PhysRevD.99.084032.

[150] W. E. East, F. Pretorius, Dynamical Capture Binary Neutron Star Mergers, Astrophys. J. 760 (2012) L4. arXiv:1208.5279, doi:10.1088/2041-8205/760/1/L4.

[151] R. , S. Bernuzzi, M. Thierfelder, B. Br¨ugmann,F. Pretorius, Eccentric binary neutron star mergers, Phys. Rev. D 86 (12) (2012) 121501. arXiv:1109.5128, doi:10.1103/PhysRevD.86. 121501.

[152] W. E. East, V. Paschalidis, F. Pretorius, Equation of state effects and one-arm spiral instability in hypermassive neutron stars formed in eccentric neutron star mergers, Classical and Quantum Gravity 33 (24) (2016) 244004. arXiv:1609.00725, doi:10.1088/0264-9381/33/24/244004.

[153] S. V. Chaurasia, T. Dietrich, N. K. Johnson-McDaniel, M. Ujevic, W. Tichy, B. Br¨ugmann,Grav- itational waves and mass ejecta from binary neutron star mergers: Effect of large eccentricities, Phys. Rev. D 98 (10) (2018) 104005. arXiv:1807.06857, doi:10.1103/PhysRevD.98.104005.

[154] H. Yang, W. E. East, V. Paschalidis, F. Pretorius, R. F. P. Mendes, Evolution of highly eccentric binary neutron stars including tidal effects, Phys. Rev. D 98 (4) (2018) 044007. arXiv:1806. 00158, doi:10.1103/PhysRevD.98.044007.

[155] L. J. Papenfort, R. Gold, L. Rezzolla, Dynamical ejecta and nucleosynthetic yields from eccentric binary neutron-star mergers, Phys. Rev. D 98 (2018) 104028. arXiv:1807.03795, doi:10.1103/ PhysRevD.98.104028.

[156] D. Radice, General-relativistic Large-eddy Simulations of Binary Neutron Star Mergers, Astro- phys. J. Lett. 838 (2017) L2. arXiv:1703.02046, doi:10.3847/2041-8213/aa6483.

[157] M. G. Alford, L. Bovard, M. Hanauske, L. Rezzolla, K. Schwenzer, Viscous Dissipation and Heat Conduction in Binary Neutron-Star Mergers, Phys. Rev. Lett. 120 (4) (2018) 041101. arXiv: 1707.09475, doi:10.1103/PhysRevLett.120.041101.

[158] M. Shibata, K. Kiuchi, Gravitational waves from remnant massive neutron stars of binary neutron star merger: Viscous hydrodynamics effects, Phys. Rev. D 95 (12) (2017) 123003. arXiv:1705. 06142, doi:10.1103/PhysRevD.95.123003.

63 [159] K. Kiuchi, K. Kyutoku, Y. Sekiguchi, M. Shibata, Global simulations of strongly magnetized remnant massive neutron stars formed in binary neutron star mergers, Phys. Rev. D 97 (12) (2018) 124039. arXiv:1710.01311, doi:10.1103/PhysRevD.97.124039.

[160] F. Zappa, S. Bernuzzi, D. Radice, A. Perego, T. Dietrich, Gravitational-Wave Luminosity of Binary Neutron Stars Mergers, Physical Review Letters 120 (11) (2018) 111101. arXiv:1712. 04267, doi:10.1103/PhysRevLett.120.111101.

[161] C. Palenzuela, L. Lehner, S. L. Liebling, M. Ponce, M. Anderson, D. Neilsen, P. Motl, Linking electromagnetic and gravitational radiation in coalescing binary neutron stars, Phys. Rev. D 88 (4) (2013) 043011. arXiv:1307.7372, doi:10.1103/PhysRevD.88.043011.

[162] M. Ponce, C. Palenzuela, L. Lehner, S. L. Liebling, Interaction of misaligned in the coalescence of binary neutron stars, Phys. Rev. D 90 (4) (2014) 044007. arXiv:1404.0692, doi:10.1103/PhysRevD.90.044007.

[163] K. Dionysopoulou, D. Alic, L. Rezzolla, General-relativistic resistive-magnetohydrodynamic sim- ulations of binary neutron stars, Phys. Rev. D 92 (8) (2015) 084064. arXiv:1502.02021, doi:10.1103/PhysRevD.92.084064.

[164] B. Giacomazzo, L. Rezzolla, L. Baiotti, Can magnetic fields be detected during the inspiral of binary neutron stars?, Mon. Not. R. Astron. Soc. 399 (2009) L164–L168. arXiv:0901.2722, doi:10.1111/j.1745-3933.2009.00745.x.

[165] F. Galeazzi, W. Kastaun, L. Rezzolla, J. A. Font, Implementation of a simplified approach to radiative transfer in general relativity, Phys. Rev. D 88 (6) (2013) 064009. arXiv:1306.4953, doi:10.1103/PhysRevD.88.064009.

[166] K. Kyutoku, K. Kiuchi, Y. Sekiguchi, M. Shibata, K. Taniguchi, Neutrino transport in black hole-neutron star binaries: Neutrino emission and dynamical mass ejection, Phys. Rev. D 97 (2) (2018) 023009. arXiv:1710.00827, doi:10.1103/PhysRevD.97.023009.

[167] L. Baiotti, B. Giacomazzo, L. Rezzolla, Accurate evolutions of inspiralling neutron-star binaries: Prompt and delayed collapse to a black hole, Phys. Rev. D 78 (8) (2008) 084033. arXiv:0804. 0594, doi:10.1103/PhysRevD.78.084033.

[168] L. Baiotti, L. Rezzolla, Binary neutron-star mergers: a review of Einstein’s richest laboratory, Rept. Prog. Phys. 80 (9) (2017) 096901. arXiv:1607.03540, doi:10.1088/1361-6633/aa67bb.

[169] V. Paschalidis, General relativistic simulations of compact binary mergers as engines for short gamma-ray bursts, Classical and Quantum Gravity 34 (8) (2017) 084002. arXiv:1611.01519, doi:10.1088/1361-6382/aa61ce.

[170] M. D. Duez, Y. Zlochower, Numerical relativity of compact binaries in the 21st century, Reports on Progress in Physics 82 (1) (2019) 016902. arXiv:1808.06011, doi:10.1088/1361-6633/aadb16.

[171] S. Chandrasekhar, Hydrodynamic and hydromagnetic stability, Dover Edition, New York, USA, 1981.

[172] J. Braithwaite, A.˚ Nordlund, Stable magnetic fields in stellar interiors, Astron. Astrophys. 450 (2006) 1077–1095. arXiv:arXiv:astro-ph/0510316, doi:10.1051/0004-6361:20041980.

64 [173] R. Ciolfi, V. Ferrari, L. Gualtieri, J. A. Pons, Relativistic models of magnetars: the twisted torus magnetic field configuration, Mon. Not. R. Astron. Soc. 397 (2009) 913–924. arXiv:0903.0556, doi:10.1111/j.1365-2966.2009.14990.x.

[174] S. K. Lander, D. I. Jones, Are there any stable magnetic fields in barotropic stars?, Mon. Not. R. Astron. Soc. 424 (2012) 482–494. arXiv:1202.2339, doi:10.1111/j.1365-2966.2012.21213.x.

[175] D. J. Price, S. Rosswog, Producing Ultrastrong Magnetic Fields in Neutron Star Mergers, Science 312 (2006) 719–722. arXiv:astro-ph/0603845, doi:10.1126/science.1125201.

[176] M. Anderson, E. W. Hirschmann, L. Lehner, S. L. Liebling, P. M. Motl, D. Neilsen, C. Palenzuela, J. E. Tohline, Magnetized Neutron-Star Mergers and Gravitational-Wave Signals, Phys. Rev. Lett. 100 (19) (2008) 191101. arXiv:0801.4387, doi:10.1103/PhysRevLett.100.191101.

[177] M. Obergaulinger, M. A. Aloy, E. M¨uller, Local simulations of the magnetized Kelvin-Helmholtz instability in neutron-star mergers, Astron. Astrophys. 515 (2010) A30. arXiv:1003.6031, doi: 10.1051/0004-6361/200913386.

[178] B. Giacomazzo, L. Rezzolla, L. Baiotti, Accurate evolutions of inspiralling and magnetized neutron stars: Equal-mass binaries, Phys. Rev. D 83 (4) (2011) 044014. arXiv:1009.2468, doi:10.1103/ PhysRevD.83.044014.

[179] L. Rezzolla, B. Giacomazzo, L. Baiotti, J. Granot, C. Kouveliotou, M. A. Aloy, The Missing Link: Merging Neutron Stars Naturally Produce Jet-like Structures and Can Power Short Gamma-ray Bursts, Astrophys. J. Letters 732 (2011) L6. arXiv:1101.4298, doi:10.1088/2041-8205/732/ 1/L6.

[180] K. Kiuchi, K. Kyutoku, Y. Sekiguchi, M. Shibata, T. Wada, High resolution numerical relativity simulations for the merger of binary magnetized neutron stars, Phys. Rev. D 90 (4) (2014) 041502. arXiv:1407.2660, doi:10.1103/PhysRevD.90.041502.

[181] K. Kiuchi, P. Cerd´a-Dur´an,K. Kyutoku, Y. Sekiguchi, M. Shibata, Efficient magnetic-field am- plification due to the Kelvin-Helmholtz instability in binary neutron star mergers, Phys. Rev. D 92 (12) (2015) 124034. arXiv:1509.09205, doi:10.1103/PhysRevD.92.124034.

[182] R. C. Duncan, C. Thompson, Formation of very strongly magnetized neutron stars - Implications for gamma-ray bursts, Astrophys. J. 392 (1992) L9–L13. doi:10.1086/186413.

[183] V. M. Kaspi, A. M. Beloborodov, Magnetars, Annual Review of Astronomy and Astrophysics 55 (1) (2017) 261–301. arXiv:https://doi.org/10.1146/annurev-astro-081915-023329, doi:10.1146/annurev-astro-081915-023329.

[184] D. Eichler, M. Livio, T. Piran, D. N. Schramm, Nucleosynthesis, neutrino bursts and gamma-rays from coalescing neutron stars, Nature 340 (1989) 126–128. doi:10.1038/340126a0.

[185] R. Narayan, B. Paczynski, T. Piran, Gamma-ray bursts as the death throes of massive binary stars, Astrophys. J. Lett. 395 (1992) L83–L86. arXiv:astro-ph/9204001, doi:10.1086/186493.

[186] E. Nakar, Short-hard gamma-ray bursts, Phys. Rep. 442 (2007) 166–236. arXiv:astro-ph/ 0701748, doi:10.1016/j.physrep.2007.02.005.

[187] P. D’Avanzo, Short gamma-ray bursts: A review, Journal of High Energy Astrophysics 7 (2015) 73–80. doi:10.1016/j.jheap.2015.07.002.

65 [188] T. W. Baumgarte, S. L. Shapiro, M. Shibata, On the Maximum Mass of Differentially Rotating Neutron Stars, Astrophys. J. Lett. 528 (2000) L29–L32. arXiv:astro-ph/9910565, doi:10. 1086/312425. [189] N. Stergioulas, Rotating Stars in Relativity, Living Reviews in Relativity 6 (1) (2003) 3. arXiv: gr-qc/0302034, doi:10.12942/lrr-2003-3. [190] M. Shibata, T. W. Baumgarte, S. L. Shapiro, The Bar-Mode Instability in Differentially Rotating Neutron Stars: Simulations in Full General Relativity, Astrophys. J. 542 (2000) 453–463. arXiv: astro-ph/0005378, doi:10.1086/309525. [191] M. Shibata, S. Karino, Y. Eriguchi, Dynamical bar-mode instability of differentially rotating stars: effects of equations of state and velocity profiles, Mon. Not. R. Astron. Soc. 343 (2) (2003) 619–626. arXiv:astro-ph/0304298, doi:10.1046/j.1365-8711.2003.06699.x. [192] L. Baiotti, R. De Pietri, G. M. Manca, L. Rezzolla, Accurate simulations of the dynamical bar- mode instability in full general relativity, Phys. Rev. D 75 (4) (2007) 044023. arXiv:astro-ph/ 0609473, doi:10.1103/PhysRevD.75.044023. [193] R. de Pietri, L. Baiotti, G. M. Manca, L. Rezzolla, Accurate simulations of the barmode instability in General Relativity, in: L. Mornas, J. Diaz Alonso (Eds.), A Century of Relativity Physics: ERE 2005, Vol. 841 of American Institute of Physics Conference Series, 2006, pp. 416–419. doi: 10.1063/1.2218203. [194] L. Franci, R. De Pietri, K. Dionysopoulou, L. Rezzolla, Bar-mode instability suppression in mag- netized relativistic stars, Journal of Physics Conference Series 470 (1) (2013) 012008. arXiv: 1309.6549, doi:10.1088/1742-6596/470/1/012008. [195] L. Franci, R. De Pietri, K. Dionysopoulou, L. Rezzolla, Dynamical bar-mode instability in rotating and magnetized relativistic stars, Phys. Rev. D 88 (10) (2013) 104028. arXiv:1308.3989, doi: 10.1103/PhysRevD.88.104028. [196] D. M. Siegel, R. Ciolfi, A. I. Harte, L. Rezzolla, Magnetorotational instability in relativistic hypermassive neutron stars, Phys. Rev. D R 87 (12) (2013) 121302. arXiv:1302.4368, doi: 10.1103/PhysRevD.87.121302. [197] R. De Pietri, A. Feo, L. Franci, F. L¨offler,Neutron star instabilities in full general relativity using a Γ =2.75 ideal fluid, Phys. Rev. D 90 (2) (2014) 024034. arXiv:1403.8066, doi:10.1103/ PhysRevD.90.024034. [198] F. L¨offler, R. De Pietri, A. Feo, F. Maione, L. Franci, Stiffness effects on the dynamics of the bar-mode instability of neutron stars in full general relativity, Phys. Rev. D 91 (6) (2015) 064057. arXiv:1411.1963, doi:10.1103/PhysRevD.91.064057. [199] S. Ou, J. E. Tohline, Unexpected Dynamical Instabilities in Differentially Rotating Neutron Stars, Astrophys. J. 651 (2006) 1068. arXiv:astro-ph/0604099, doi:10.1086/507597. [200] G. Corvino, L. Rezzolla, S. Bernuzzi, R. De Pietri, B. Giacomazzo, On the shear instability in relativistic neutron stars, Class. Quantum Grav. 27 (11) (2010) 114104. arXiv:1001.5281, doi:10.1088/0264-9381/27/11/114104. [201] V. Paschalidis, W. E. East, F. Pretorius, S. L. Shapiro, One-arm spiral instability in hypermassive neutron stars formed by dynamical-capture binary neutron star mergers, Phys. Rev. D 92 (12) (2015) 121502. arXiv:1510.03432, doi:10.1103/PhysRevD.92.121502.

66 [202] W. E. East, V. Paschalidis, F. Pretorius, S. L. Shapiro, Relativistic simulations of eccentric binary neutron star mergers: One-arm spiral instability and effects of neutron star spin, Phys. Rev. D 93 (2) (2016) 024011. arXiv:1511.01093, doi:10.1103/PhysRevD.93.024011.

[203] D. Radice, S. Bernuzzi, C. D. Ott, One-armed spiral instability in neutron star mergers and its detectability in gravitational waves, Phys. Rev. D 94 (6) (2016) 064011. arXiv:1603.05726, doi:10.1103/PhysRevD.94.064011.

[204] L. Lehner, S. L. Liebling, C. Palenzuela, P. M. Motl, m =1 instability and gravitational wave signal in binary neutron star mergers, Phys. Rev. D 94 (4) (2016) 043003. arXiv:1605.02369, doi:10.1103/PhysRevD.94.043003.

[205] V. Paschalidis, N. Stergioulas, Rotating stars in relativity, Living Reviews in Relativity 20 (2017) 7. arXiv:1612.03050, doi:10.1007/s41114-017-0008-x.

[206] W. Kastaun, R. Ciolfi, B. Giacomazzo, Structure of stable binary neutron star merger remnants: A case study, Phys. Rev. D 94 (4) (2016) 044060. arXiv:1607.02186, doi:10.1103/PhysRevD. 94.044060.

[207] P. Diener, A new general purpose event horizon finder for 3D numerical , Class. Quan- tum Grav. 20 (22) (2003) 4901–4917. arXiv:gr-qc/0305039, doi:10.1088/0264-9381/20/22/ 014.

[208] J. Thornburg, Finding apparent horizons in numerical relativity, Phys. Rev. D 54 (1996) 4899– 4918. arXiv:gr-qc/9508014, doi:10.1103/PhysRevD.54.4899.

[209] M. Shibata, K. Ury¯u,Apparent horizon finder for general three-dimensional spaces, Phys. Rev. D 62 (8) (2000) 087501. doi:10.1103/PhysRevD.62.087501.

[210] J. Thornburg, A fast apparent-horizon finder for 3-dimensional Cartesian grids in numerical rel- ativity, Class. Quantum Grav. 21 (2) (2004) 743–766. arXiv:gr-qc/0306056.

[211] J. Thornburg, Event and Apparent Horizon Finders for 3 + 1 Numerical Relativity, Living Reviews in Relativity 10 (1) (2007) 3. doi:10.12942/lrr-2007-3.

[212] S. L. Shapiro, Differential Rotation in Neutron Stars: Magnetic Braking and Viscous Damping, Astrophys. J. 544 (1) (2000) 397–408. arXiv:astro-ph/0010493, doi:10.1086/317209.

[213] K. Hotokezaka, K. Kiuchi, K. Kyutoku, T. Muranushi, Y.-i. Sekiguchi, M. Shibata, K. Taniguchi, Remnant massive neutron stars of binary neutron star mergers: Evolution process and gravita- tional waveform, Phys. Rev. D 88 (4) (2013) 044026. arXiv:1307.5888, doi:10.1103/PhysRevD. 88.044026.

[214] G. B. Cook, S. L. Shapiro, S. A. Teukolsky, Spin-up of a Rapidly Rotating Star by Angular Momentum Loss: Effects of General Relativity, Astrophys. J. 398 (1992) 203. doi:10.1086/ 171849.

[215] V. Ravi, P. D. Lasky, The birth of black holes: neutron star collapse times, gamma-ray bursts and fast radio bursts, Mon. Not. R. Astron. Soc. 441 (2014) 2433–2439. doi:10.1093/mnras/stu720.

[216] T. Kawamura, B. Giacomazzo, W. Kastaun, R. Ciolfi, A. Endrizzi, L. Baiotti, R. Perna, Binary neutron star mergers and short gamma-ray bursts: Effects of magnetic field orientation, equation of state, and mass ratio, Phys. Rev. D 94 (6) (2016) 064012. arXiv:1607.01791, doi:10.1103/ PhysRevD.94.064012.

67 [217] C. Cutler, E.´ E. Flanagan, Gravitational waves from merging compact binaries: How accurately can one extract the binary’s parameters from the inspiral waveform ?, Phys. Rev. D 49 (1994) 2658–2697. arXiv:gr-qc/9402014, doi:10.1103/PhysRevD.49.2658\.

[218] L. Blanchet, T. Damour, B. R. Iyer, C. M. Will, A. G. Wiseman, Gravitational-Radiation Damping of Compact Binary Systems to Second Post-Newtonian Order, Phys. Rev. Lett. 74 (18) (1995) 3515–3518. arXiv:gr-qc/9501027, doi:10.1103/PhysRevLett.74.3515.

[219] L. Blanchet, B. R. Iyer, C. M. Will, A. G. Wiseman, Gravitational waveforms from inspiralling compact binaries to second-post-Newtonian order, Classical and Quantum Gravity 13 (4) (1996) 575–584. arXiv:gr-qc/9602024, doi:10.1088/0264-9381/13/4/002.

[220] The LIGO Scientific Collaboration, the Virgo Collaboration, B. P. Abbott, R. Abbott, T. D. Abbott, F. Acernese, K. Ackley, C. Adams, T. Adams, P. Addesso, R. X. Adhikari, V. B. Adya, et al., Properties of the Binary Neutron Star Merger GW170817, Physical Review X 9 (1) (2019) 011001. arXiv:1805.11579, doi:10.1103/PhysRevX.9.011001.

[221] K. Kokkotas, B. Schmidt, Quasi-Normal Modes of Stars and Black Holes, Living Rev. Relativity 2 (1999) 2. arXiv:gr-qc/9909058, doi:10.12942/lrr-1999-2.

[222] C. Messenger, K. Takami, S. Gossan, L. Rezzolla, B. S. Sathyaprakash, Source from Gravitational-Wave Observations of Binary Neutron Star Mergers, Physical Review X 4 (2014) 041004. arXiv:1312.1862, doi:10.1103/PhysRevX.4.041004.

[223] J. Clark, A. Bauswein, L. Cadonati, H.-T. Janka, C. Pankow, N. Stergioulas, Prospects for high frequency burst searches following binary neutron star coalescence with advanced gravitational wave detectors, Phys. Rev. D 90 (6) (2014) 062004. arXiv:1406.5444, doi:10.1103/PhysRevD. 90.062004.

[224] J. A. Clark, A. Bauswein, N. Stergioulas, D. Shoemaker, Observing gravitational waves from the post-merger phase of binary neutron star coalescence, Class. Quantum Grav. 33 (8) (2016) 085003. arXiv:1509.08522, doi:10.1088/0264-9381/33/8/085003.

[225] H. Yang, V. Paschalidis, K. Yagi, L. Lehner, F. Pretorius, N. Yunes, Gravitational wave spec- troscopy of binary neutron star merger remnants with mode stacking, Phys. Rev. D 97 (2) (2018) 024049. arXiv:1707.00207, doi:10.1103/PhysRevD.97.024049.

[226] K. Chatziioannou, J. A. Clark, A. Bauswein, M. Millhouse, T. B. Littenberg, N. Cornish, Inferring the post-merger gravitational wave emission from binary neutron star coalescences, Phys. Rev. D 96 (12) (2017) 124035. arXiv:1711.00040, doi:10.1103/PhysRevD.96.124035.

[227] B. P. Abbott, R. Abbott, T. D. Abbott, M. R. Abernathy, F. Acernese, K. Ackley, C. Adams, T. Adams, P. Addesso, R. X. Adhikari, et al., Prospects for observing and localizing gravitational- wave transients with Advanced LIGO, Advanced Virgo and KAGRA, Living Reviews in Relativity 21 (2018) 3. arXiv:1304.0670, doi:10.1007/s41114-018-0012-9.

[228] A. Torres-Rivas, K. Chatziioannou, A. Bauswein, J. A. Clark, Observing the post-merger signal of GW170817-like events with improved gravitational-wave detectors, Phys. Rev. D 99 (4) (2019) 044014. arXiv:1811.08931, doi:10.1103/PhysRevD.99.044014.

[229] J. Aasi, et al., Advanced LIGO, Class. Quant. Grav. 32 (2015) 074001. arXiv:1411.4547, doi: 10.1088/0264-9381/32/7/074001.

68 [230] F. Acernese, et al., Advanced Virgo: a second-generation interferometric gravitational wave de- tector, Class. Quant. Grav. 32 (2) (2015) 024001. arXiv:1408.3978, doi:10.1088/0264-9381/ 32/2/024001.

[231] Y. Aso, Y. Michimura, K. Somiya, M. Ando, O. Miyakawa, T. Sekiguchi, D. Tatsumi, H. Ya- mamoto, Interferometer design of the KAGRA gravitational wave detector, Phys. Rev. D 88 (4) (2013) 043007. arXiv:1306.6747, doi:10.1103/PhysRevD.88.043007.

[232] S. Fairhurst, Improved source localization with LIGO-India, Journal of Physics Conference Series 484 (1) (2014) 012007. arXiv:1205.6611, doi:10.1088/1742-6596/484/1/012007.

[233] D. McClelland, M. Evans, B. Lantz, I. Martin, V. Quetschke, R. Schnabel, Instrument science white paper, Tech. Rep. LIGO Document T1500290-v2, LIGO Scientific Collaboration (2015).

[234] M. Punturo, et al., The Einstein Telescope: a third-generation gravitational wave observatory, Class. Quantum Grav. 27 (2010) 194002. doi:10.1088/0264-9381/27/19/194002.

[235] B. S. Sathyaprakash, B. F. Schutz, Physics, Astrophysics and Cosmology with Gravitational Waves, Living Reviews in Relativity 12 (1) (2009) 2. arXiv:0903.0338, doi:10.12942/ lrr-2009-2.

[236] D. Martynov, H. Miao, H. Yang, F. H. Vivanco, E. Thrane, R. Smith, P. Lasky, W. E. East, R. Adhikari, A. Bauswein, A. Brooks, Y. Chen, T. Corbitt, A. Freise, H. Grote, Y. Levin, C. Zhao, A. Vecchio, Exploring the sensitivity of gravitational wave detectors to neutron star physics, Phys. Rev. D 99 (10) (2019) 102004. arXiv:1901.03885, doi:10.1103/PhysRevD.99.102004.

[237] N. K. Glendenning, S. A. Moszkowski, Reconciliation of neutron-star masses and binding of the lambda in hypernuclei, Phys. Rev. Lett. 67 (18) (1991) 2414–2417. doi:10.1103/PhysRevLett. 67.2414.

[238] T. E. Riley, G. Raaijmakers, A. L. Watts, On parametrized cold dense matter equation-of-state inference, Mon. Not. R. Astron. Soc. 478 (2018) 1093–1131. arXiv:1804.09085, doi:10.1093/ mnras/sty1051.

[239] G. Raaijmakers, T. E. Riley, A. L. Watts, A pitfall of piecewise-polytropic equation of state inference, Mon. Not. R. Astron. Soc. 478 (2018) 2177–2192. arXiv:1804.09087, doi:10.1093/ mnras/sty1052.

[240] S. K. Greif, G. Raaijmakers, K. Hebeler, A. Schwenk, A. L. Watts, Equation of state sensitivities when inferring neutron star and dense matter properties, Mon. Not. R. Astron. Soc. 485 (4) (2019) 5363–5376. arXiv:1812.08188, doi:10.1093/mnras/stz654.

[241] L. Lindblom, Causal representations of neutron-star equations of state, Phys. Rev. D 97 (12) (2018) 123019. arXiv:1804.04072, doi:10.1103/PhysRevD.97.123019.

[242] A. W. Steiner, J. M. Lattimer, E. F. Brown, The Equation of State from Observed Masses and Radii of Neutron Stars, Astrophys. Journ. 722 (2010) 33–54. arXiv:1005.0811, doi:10.1088/ 0004-637X/722/1/33.

[243] A. W. Steiner, J. M. Lattimer, E. F. Brown, The Neutron Star Mass-Radius Relation and the Equation of State of Dense Matter, Astrophys. J. Lett. 765 (1) (2013) L5. arXiv:1205.6871, doi:10.1088/2041-8205/765/1/L5.

69 [244] K. Hebeler, J. M. Lattimer, C. J. Pethick, A. Schwenk, Equation of State and Neutron Star Properties Constrained by Nuclear Physics and Observation, Astrophys. J. 773 (2013) 11. arXiv: 1303.4662, doi:10.1088/0004-637X/773/1/11.

[245] B. D. Lackey, L. Wade, Reconstructing the neutron-star equation of state with gravitational- wave detectors from a realistic population of inspiralling binary neutron stars, Phys. Rev. D 91 (4) (2015) 043002. arXiv:1410.8866, doi:10.1103/PhysRevD.91.043002.

[246] C. A. Raithel, F. Ozel,¨ D. Psaltis, From Neutron Star Observables to the Equation of State. I. An Optimal Parametrization, Astrophys. J. 831 (2016) 44. arXiv:1605.03591, doi:10.3847/ 0004-637X/831/1/44.

[247] M. F. Carney, L. E. Wade, B. S. Irwin, Comparing two models for measuring the neutron star equation of state from gravitational-wave signals, Phys. Rev. D 98 (6) (2018) 063004. arXiv: 1805.11217, doi:10.1103/PhysRevD.98.063004.

[248] C. A. Raithel, F. Ozel,¨ D. Psaltis, From Neutron Star Observables to the Equation of State. II. Bayesian Inference of Equation of State Pressures, Astrophys. J. 844 (2017) 156. arXiv: 1704.00737, doi:10.3847/1538-4357/aa7a5a.

[249] J. M. Ib´a˜nez,A. Marquina, S. Serna, M. A. Aloy, Anomalous dynamics triggered by a non- convex equation of state in relativistic flows, Mon. Not. R. Astron. Soc. 476 (1) (2018) 1100–1110. arXiv:1712.03248, doi:10.1093/mnras/sty137.

[250] P. Cinnella, P. M. Congedo, Inviscid and viscous aerodynamics of dense gases, Journal of Fluid Mechanics 580 (2007) 179. doi:10.1017/S0022112007005290.

[251] P. Cinnella, P. M. Congedo, V. Pediroda, L. Parussini, Sensitivity analysis of dense gas flow simulations to thermodynamic uncertainties, Physics of Fluids 23 (2011) 116101. doi:10.1063/ 1.3657080.

[252] M. A. Aloy, J. M. Ib´a˜nez,N. Sanchis-Gual, M. Obergaulinger, J. A. Font, S. Serna, A. Marquina, Neutron star collapse and gravitational waves with a non-convex equation of state, Mon. Not. R. Astron. Soc. 484 (4) (2019) 4980–5008. arXiv:1806.03314, doi:10.1093/mnras/stz293.

[253] J. M. Ib´a˜nez,I. Cordero-Carri´on,J. M. Mart´ı,J. A. Miralles, On the convexity of relativistic hydrodynamics, Classical and Quantum Gravity 30 (5) (2013) 057002. doi:10.1088/0264-9381/ 30/5/057002.

[254] M. C. Miller, C. Chirenti, F. K. Lamb, Constraining the equation of state of high-density cold mat- ter using nuclear and astronomical measurements, arXiv e-prints (2019) arXiv:1904.08907arXiv: 1904.08907.

[255] T. Abdelsalhin, A. Maselli, V. Ferrari, Solving the relativistic inverse stellar problem through gravitational waves observation of binary neutron stars, Phys. Rev. D 97 (8) (2018) 084014. arXiv:1712.01303, doi:10.1103/PhysRevD.97.084014.

[256] L. Lindblom, Spectral representations of neutron-star equations of state, Phys. Rev. D 82 (10) (2010) 103011. arXiv:1009.0738, doi:10.1103/PhysRevD.82.103011.

[257] L. Lindblom, N. M. Indik, Spectral approach to the relativistic inverse stellar structure problem, Phys. Rev. D 86 (8) (2012) 084003. arXiv:1207.3744, doi:10.1103/PhysRevD.86.084003.

70 [258] L. Lindblom, N. M. Indik, Spectral approach to the relativistic inverse stellar structure problem II, Phys. Rev. D 89 (6) (2014) 064003. arXiv:1310.0803, doi:10.1103/PhysRevD.89.064003.

[259] L. Lindblom, Inverse structure problem for neutron-star binaries, Phys. Rev. D 98 (4) (2018) 043012. arXiv:1807.02538, doi:10.1103/PhysRevD.98.043012.

[260] F. Douchin, P. Haensel, A unified equation of state of dense matter and neutron star struc- ture, Astron. Astrophys. 380 (2001) 151–167. arXiv:arXiv:astro-ph/0111092, doi:10.1051/ 0004-6361:20011402.

[261] S. L. Shapiro, S. A. Teukolsky, Black Holes, White Dwarfs, and Neutron Stars, John Wiley & Sons, New York, 1983.

[262] M. G. Alford, S. P. Harris, β equilibrium in neutron-star mergers, Phys. Rev. C 98 (6) (2018) 065806. arXiv:1803.00662, doi:10.1103/PhysRevC.98.065806.

[263] F. J. Fattoyev, W. G. Newton, B.-A. Li, Probing the high-density behavior of symmetry energy with gravitational waves, European Physical Journal A 50 (2014) 45. arXiv:1309.5153, doi: 10.1140/epja/i2014-14045-6.

[264] J. Margueron, R. Hoffmann Casali, F. Gulminelli, Equation of state for dense nucleonic matter from metamodeling. I. Foundational aspects, Physical Review C 97 (2) (2018) 025805. arXiv: 1708.06894, doi:10.1103/PhysRevC.97.025805.

[265] J. Margueron, R. Hoffmann Casali, F. Gulminelli, Equation of state for dense nucleonic matter from metamodeling. II. Predictions for neutron star properties, Physical Review C 97 (2) (2018) 025806. arXiv:1708.06895, doi:10.1103/PhysRevC.97.025806.

[266] Z.-Y. Zhu, E.-P. Zhou, A. Li, Neutron Star Equation of State from the Quark Level in Light of GW170817, Astrophys. J. 862 (2018) 98. arXiv:1802.05510, doi:10.3847/1538-4357/aacc28.

[267] T. Malik, N. Alam, M. Fortin, C. Providˆencia,B. K. Agrawal, T. K. Jha, B. Kumar, S. K. Patra, GW170817: Constraining the nuclear matter equation of state from the neutron star tidal deformability, Physical Review C 98 (3) (2018) 035804. arXiv:1805.11963, doi:10.1103/ PhysRevC.98.035804.

[268] Z. Carson, A. W. Steiner, K. Yagi, Constraining nuclear matter parameters with GW170817, Phys. Rev. D 99 (4) (2019) 043010. arXiv:1812.08910, doi:10.1103/PhysRevD.99.043010.

[269] N.-B. Zhang, B.-A. Li, Delineating effects of nuclear symmetry energy on the radii and tidal polarizabilities of neutron stars, Journal of Physics G Nuclear Physics 46 (1) (2019) 014002. arXiv:1808.07955, doi:10.1088/1361-6471/aaef54.

[270] I. Vida˜na,C. Providˆencia,A. Polls, A. Rios, Density dependence of the nuclear symmetry energy: A microscopic perspective, Physical Review C 80 (4) (2009) 045806. arXiv:0907.1165, doi: 10.1103/PhysRevC.80.045806.

[271] J. J. Li, A. Sedrakian, Constraining compact star properties with nuclear saturation parameters, Phys. Rev. C 100 (1) (2019) 015809. arXiv:1903.06057, doi:10.1103/PhysRevC.100.015809.

[272] B.-A. Li, P. G. Krastev, D.-H. Wen, N.-B. Zhang, Towards Understanding Astrophysical Effects of Nuclear Symmetry Energy, arXiv e-prints (2019) arXiv:1905.13175arXiv:1905.13175.

71 [273] N. Alam, B. K. Agrawal, J. N. De, S. K. Samaddar, G. Col`o,Equation of state of nuclear matter from empirical constraints, Physical Review C 90 (5) (2014) 054317. arXiv:1411.0404, doi:10.1103/PhysRevC.90.054317.

[274] N. Alam, B. K. Agrawal, M. Fortin, H. Pais, C. Providˆencia,A. R. Raduta, A. Sulaksono, Strong correlations of neutron star radii with the slopes of nuclear matter incompressibility and symmetry energy at saturation, Physical Review C 94 (5) (2016) 052801. arXiv:1610.06344, doi:10.1103/ PhysRevC.94.052801.

[275] M. McNeil Forbes, S. Bose, S. Reddy, D. Zhou, A. Mukherjee, S. De, Constraining the neutron- matter equation of state with gravitational waves, arXiv e-prints (2019) arXiv:1904.04233arXiv: 1904.04233.

[276] D. D. Ivanenko, D. F. Kurdgelaidze, Hypothesis concerning quark stars, Astrophysics 1 (1965) 251–252. doi:10.1007/BF01042830.

[277] N. Itoh, Hydrostatic Equilibrium of Hypothetical Quark Stars, Progress of Theoretical Physics 44 (1970) 291–292.

[278] G. Baym, S. A. Chin, Can a neutron star be a giant MIT bag?, Physics Letters B 62 (1976) 241–244. doi:10.1016/0370-2693(76)90517-7.

[279] M. Buballa, V. Dexheimer, A. Drago, E. Fraga, P. Haensel, I. Mishustin, G. Pagliara, J. Schaffner- Bielich, S. Schramm, A. Sedrakian, F. Weber, EMMI rapid reaction task force meeting on quark matter in compact stars, Journal of Physics G Nuclear Physics 41 (12) (2014) 123001. arXiv: 1402.6911, doi:10.1088/0954-3899/41/12/123001.

[280] M. G. Alford, S. Han, Characteristics of hybrid compact stars with a sharp hadron-quark in- terface, European Physical Journal A 52 (2016) 62. arXiv:1508.01261, doi:10.1140/epja/ i2016-16062-9.

[281] C. Alcock, E. Farhi, A. Olinto, Strange stars, Astrophys. J. 310 (1986) 261–272. doi:10.1086/ 164679.

[282] P. Haensel, J. L. Zdunik, R. Schaefer, Strange quark stars, Astron. Astrophys. 160 (1986) 121–128.

[283] Y. Nambu, G. Jona-Lasinio, Dynamical Model of Elementary Particles Based on an Analogy with . I, Physical Review 122 (1961) 345–358. doi:10.1103/PhysRev.122.345.

[284] Y. Nambu, G. Jona-Lasinio, Dynamical Model of Elementary Particles Based on an Analogy with Superconductivity. II, Physical Review 124 (1961) 246–254. doi:10.1103/PhysRev.124.246.

[285] N. Chamel, A. F. Fantina, J. M. Pearson, S. Goriely, Phase transitions in dense matter and the maximum mass of neutron stars, Astron. Astrophys. 553 (2013) A22. arXiv:1205.0983, doi:10.1051/0004-6361/201220986.

[286] J. L. Zdunik, P. Haensel, Maximum mass of neutron stars and strange neutron-star cores, Astron. Astrophys. 551 (2013) A61. arXiv:1211.1231, doi:10.1051/0004-6361/201220697.

[287] M. G. Alford, S. Han, M. Prakash, Generic conditions for stable hybrid stars, Phys. Rev. D 88 (8) (2013) 083013. arXiv:1302.4732, doi:10.1103/PhysRevD.88.083013.

[288] G. Monta˜na,L. Tol´os,M. Hanauske, L. Rezzolla, Constraining twin stars with GW170817, Phys. Rev. D 99 (10) (2019) 103009. arXiv:1811.10929, doi:10.1103/PhysRevD.99.103009.

72 [289] S. Han, A. W. Steiner, Tidal deformability with sharp phase transitions in binary neutron stars, Phys. Rev. D 99 (8) (2019) 083014. arXiv:1810.10967, doi:10.1103/PhysRevD.99.083014.

[290] S. Weinberg, Gravitation and Cosmology: Principles and Applications of the General , John Wiley and Sons, New York, 1972.

[291] A. Kurkela, P. Romatschke, A. Vuorinen, Cold quark matter, Phys. Rev. D 81 (10) (2010) 105021. arXiv:0912.1856, doi:10.1103/PhysRevD.81.105021.

[292] E. S. Fraga, A. Kurkela, A. Vuorinen, Interacting Quark Matter Equation of State for Compact Stars, Astrophys. J. Lett. 781 (2014) L25. arXiv:1311.5154, doi:10.1088/2041-8205/781/2/ L25.

[293] T. S. Olson, Maximally incompressible neutron star matter, Physical Review C 63 (1) (2001) 015802. arXiv:astro-ph/0011107, doi:10.1103/PhysRevC.63.015802.

[294] P. Bedaque, A. W. Steiner, Sound Velocity Bound and Neutron Stars, Physical Review Letters 114 (3) (2015) 031103. arXiv:1408.5116, doi:10.1103/PhysRevLett.114.031103.

[295] Y. Fujimoto, K. Fukushima, K. Murase, Mapping neutron star data to the equation of state of the densest matter using the deep neural network, arXiv e-prints (2019) arXiv:1903.03400arXiv: 1903.03400.

[296] E. Annala, T. Gorda, A. Kurkela, J. N¨attil¨a,A. Vuorinen, Quark-matter cores in neutron stars, arXiv e-prints (2019) arXiv:1903.09121arXiv:1903.09121.

[297] I. Tews, J. Carlson, S. Gandolfi, S. Reddy, Constraining the Speed of Sound inside Neutron Stars with Chiral Effective Field Theory Interactions and Observations, Astrophys. J. 860 (2018) 149. arXiv:1801.01923, doi:10.3847/1538-4357/aac267.

[298] I. Tews, J. Margueron, S. Reddy, Critical examination of constraints on the equation of state of dense matter obtained from GW170817, Physical Review C 98 (4) (2018) 045804. arXiv: 1804.02783, doi:10.1103/PhysRevC.98.045804.

[299] J.-E. Christian, A. Zacchi, J. Schaffner-Bielich, Classifications of twin star solutions for a constant speed of sound parameterized equation of state, European Physical Journal A 54 (2018) 28. arXiv:1707.07524, doi:10.1140/epja/i2018-12472-y.

[300] F. Ozel,¨ D. Psaltis, Reconstructing the neutron-star equation of state from astrophysical mea- surements, Phys. Rev. D 80 (10) (2009) 103003. arXiv:0905.1959, doi:10.1103/PhysRevD.80. 103003.

[301] F. Ozel,¨ G. Baym, T. G¨uver, Astrophysical measurement of the equation of state of neutron star matter, Phys. Rev. D 82 (10) (2010) 101301. arXiv:1002.3153, doi:10.1103/PhysRevD.82. 101301.

[302] N. K. Glendenning, C. Kettner, Possible third family of compact stars more dense than neutron stars, Astron. Astrophys. 353 (2000) L9–L12. arXiv:astro-ph/9807155.

[303] J. Schaffner-Bielich, M. Hanauske, H. St¨ocker, W. Greiner, Phase Transition to Hyperon Matter in Neutron Stars, Physical Review Letters 89 (17) (2002) 171101. arXiv:astro-ph/0005490, doi:10.1103/PhysRevLett.89.171101.

73 [304] E. S. Fraga, R. D. Pisarski, J. Schaffner-Bielich, New class of compact stars at high density, Nuclear Physics A 702 (2002) 217–223. arXiv:nucl-th/0110077, doi:10.1016/S0375-9474(02) 00709-1. [305] A. Kurkela, E. S. Fraga, J. Schaffner-Bielich, A. Vuorinen, Constraining Neutron Star Matter with , Astrophys. J. 789 (2014) 127. doi:10.1088/0004-637X/789/2/127. [306] P. Landry, R. Essick, Nonparametric inference of the neutron star equation of state from grav- itational wave observations, Phys. Rev. D 99 (8) (2019) 084049. arXiv:1811.12529, doi: 10.1103/PhysRevD.99.084049. [307] C. E. Rhoades, R. Ruffini, Maximum Mass of a Neutron Star, Physical Review Letters 32 (1974) 324–327. doi:10.1103/PhysRevLett.32.324. [308] C. C. Moustakidis, T. Gaitanos, C. Margaritis, G. A. Lalazissis, Bounds on the speed of sound in dense matter, and neutron star structure, Physical Review C 95 (4) (2017) 045801. arXiv: 1608.00344, doi:10.1103/PhysRevC.95.045801. [309] E. D. Van Oeveren, J. L. Friedman, Upper limit set by causality on the tidal deformability of a neutron star, Phys. Rev. D 95 (8) (2017) 083014. arXiv:1701.03797, doi:10.1103/PhysRevD. 95.083014. [310] T. H. Cromartie, E. Fonseca, S. M. Ransom, P. B. Demorest, Z. Arzoumanian, H. Blumer, P. R. Brook, M. E. DeCesar, T. Dolch, J. A. Ellis, R. D. Ferdman, E. C. Ferrara, N. Garver-Daniels, P. A. Gentile, M. L. Jones, M. T. Lam, D. R. Lorimer, R. S. Lynch, M. A. McLaughlin, C. Ng, D. J. Nice, T. T. Pennucci, R. Spiewak, I. H. Stairs, K. Stovall, J. K. Swiggum, W. Zhu, A very massive neutron star: relativistic Shapiro delay measurements of PSR J0740+6620, arXiv e-prints (2019) arXiv:1904.06759arXiv:1904.06759. [311] P. B. Demorest, T. Pennucci, S. M. Ransom, M. S. E. Roberts, J. W. T. Hessels, A two-solar-mass neutron star measured using Shapiro delay, Nature 467 (2010) 1081–1083. arXiv:1010.5788, doi:10.1038/nature09466. [312] J. Antoniadis, P. C. C. Freire, N. Wex, T. M. Tauris, R. S. Lynch, M. H. van Kerkwijk, M. Kramer, C. Bassa, V. S. Dhillon, T. Driebe, J. W. T. Hessels, V. M. Kaspi, V. I. Kondratiev, N. Langer, T. R. Marsh, M. A. McLaughlin, T. T. Pennucci, S. M. Ransom, I. H. Stairs, J. van Leeuwen, J. P. W. Verbiest, D. G. Whelan, A Massive Pulsar in a Compact Relativistic Binary, Science 340 (2013) 448. arXiv:1304.6875, doi:10.1126/science.1233232. [313] D. Alvarez-Castillo, A. Ayriyan, S. Benic, D. Blaschke, H. Grigorian, S. Typel, New class of hybrid EoS and Bayesian M - R data analysis, European Physical Journal A 52 (2016) 69. arXiv: 1603.03457, doi:10.1140/epja/i2016-16069-2. [314] B. P. Abbott, R. Abbott, T. D. Abbott, F. Acernese, K. Ackley, C. Adams, T. Adams, P. Addesso, R. X. Adhikari, V. B. Adya, et al., GW170817: Measurements of Neutron Star Radii and Equation of State, Physical Review Letters 121 (16) (2018) 161101. arXiv:1805.11581, doi:10.1103/ PhysRevLett.121.161101. [315] V. Cardoso, P. Pani, Testing the nature of dark compact objects: a status report, arXiv e-prints (2019) arXiv:1904.05363arXiv:1904.05363. [316] F. E. Schunck, E. W. Mielke, TOPICAL REVIEW: General relativistic boson stars, Classical and Quantum Gravity 20 (2003) R301–R356. arXiv:0801.0307, doi:10.1088/0264-9381/20/20/ 201.

74 [317] S. L. Liebling, C. Palenzuela, Dynamical Boson Stars, Living Reviews in Relativity 15 (1) (2012) 6. arXiv:1202.5809, doi:10.12942/lrr-2012-6.

[318] Y. Brihaye, Y. Verbin, Spherical non-Abelian solutions in conformal gravity, Phys. Rev. D 81 (4) (2010) 044041. arXiv:0910.0973, doi:10.1103/PhysRevD.81.044041.

[319] F. H. Vincent, Z. Meliani, P. Grandclement, E. Gourgoulhon, O. Straub, Imaging a boson star at the Galactic center, Class. Quant. Grav. 33 (10) (2016) 105015. arXiv:1510.04170, doi: 10.1088/0264-9381/33/10/105015.

[320] J. Barranco, A. Bernal, Self-gravitating system made of axions, Phys. Rev. D 83 (4) (2011) 043525. arXiv:1001.1769, doi:10.1103/PhysRevD.83.043525.

[321] C. J. Hogan, M. J. Rees, Axion miniclusters, Physics Letters B 205 (1988) 228–230. doi:10. 1016/0370-2693(88)91655-3.

[322] M. S. Madsen, A. R. Liddle, The cosmological formation of boson stars, Physics Letters B 251 (1990) 507–510. doi:10.1016/0370-2693(90)90788-8.

[323] E. W. Kolb, I. I. Tkachev, Axion miniclusters and Bose stars, Phys. Rev. Lett. 71 (1993) 3051– 3054. arXiv:hep-ph/9303313, doi:10.1103/PhysRevLett.71.3051.

[324] A. Iwazaki, A possible origin of gamma ray bursts and axionic boson stars, Physics Letters B 455 (1999) 192–196. arXiv:astro-ph/9903251, doi:10.1016/S0370-2693(99)00423-2.

[325] J. Eby, M. Leembruggen, J. Leeney, P. Suranyi, L. C. R. Wijewardhana, Collisions of dark matter axion stars with astrophysical sources, Journal of High Energy Physics 4 (2017) 99. arXiv: 1701.01476, doi:10.1007/JHEP04(2017)099.

[326] R. D. Peccei, H. R. Quinn, CP conservation in the presence of pseudoparticles, Physical Review Letters 38 (1977) 1440–1443. doi:10.1103/PhysRevLett.38.1440.

[327] A. Arvanitaki, S. Dimopoulos, S. Dubovsky, N. Kaloper, J. March-Russell, String axiverse, Phys. Rev. D 81 (12) (2010) 123530. arXiv:0905.4720, doi:10.1103/PhysRevD.81.123530.

[328] D. J. E. Marsh, Axion cosmology, Physics Reports 643 (2016) 1–79. arXiv:1510.07633, doi: 10.1016/j.physrep.2016.06.005.

[329] R. Brito, V. Cardoso, C. A. R. Herdeiro, E. Radu, Proca stars: Gravitating Bose-Einstein con- densates of massive spin 1 particles, Physics Letters B 752 (2016) 291–295. arXiv:1508.05395, doi:10.1016/j.physletb.2015.11.051.

[330] P. O. Mazur, E. Mottola, Gravitational vacuum condensate stars, Proceedings of the Na- tional Academy of Science 101 (2004) 9545–9550. arXiv:gr-qc/0407075, doi:10.1073/pnas. 0402717101.

[331] C. Chirenti, L. Rezzolla, Did GW150914 produce a rotating ?, Phys. Rev. D 94 (8) (2016) 084016. arXiv:1602.08759, doi:10.1103/PhysRevD.94.084016.

[332] L. Buoninfante, A. Mazumdar, Nonlocal star as a blackhole mimicker, Physical Review D 100 (2) (2019) 024031. arXiv:1903.01542, doi:10.1103/PhysRevD.100.024031.

[333] K. Yagi, N. Yunes, I-Love-Q: Unexpected Universal Relations for Neutron Stars and Quark Stars, Science 341 (6144) (2013) 365–368. arXiv:1302.4499, doi:10.1126/science.1236462.

75 [334] K. Yagi, N. Yunes, I-Love-Q relations in neutron stars and their applications to astrophysics, gravitational waves, and fundamental physics, Phys. Rev. D 88 (2) (2013) 023009. arXiv:1303. 1528, doi:10.1103/PhysRevD.88.023009.

[335] K. Yagi, N. Yunes, Approximate universal relations for neutron stars and quark stars, Phys. Rep. 681 (2017) 1–72. arXiv:1608.02582, doi:10.1016/j.physrep.2017.03.002.

[336] G. Pappas, T. A. Apostolatos, Effectively Universal Behavior of Rotating Neutron Stars in General Relativity Makes Them Even Simpler than Their Newtonian Counterparts, Phys. Rev. Lett. 112 (12) (2014) 121101. arXiv:1311.5508, doi:10.1103/PhysRevLett.112.121101.

[337] S. Chakrabarti, T. Delsate, N. G¨urlebeck, J. Steinhoff, I-Q Relation for Rapidly Rotating Neutron Stars, Phys. Rev. Lett. 112 (20) (2014) 201102. arXiv:1311.6509, doi:10.1103/PhysRevLett. 112.201102.

[338] R. Anglani, R. Casalbuoni, M. Ciminale, N. Ippolito, R. Gatto, M. Mannarelli, M. Ruggieri, Crystalline color superconductors, Reviews of Modern Physics 86 (2014) 509–561. arXiv:1302. 4264, doi:10.1103/RevModPhys.86.509.

[339] S. Y. Lau, P. T. Leung, L.-M. Lin, Tidal deformations of compact stars with crystalline quark matter, Phys. Rev. D 95 (10) (2017) 101302. arXiv:1705.01710, doi:10.1103/PhysRevD.95. 101302.

[340] S. Y. Lau, P. T. Leung, L.-M. Lin, Two-layer compact stars with crystalline quark mat- ter: Screening effect on the tidal deformability, Phys. Rev. D 99 (2) (2019) 023018. doi: 10.1103/PhysRevD.99.023018.

[341] B. Haskell, R. Ciolfi, F. Pannarale, L. Rezzolla, On the universality of I-Love-Q relations in magnetized neutron stars, Mon. Not. R. Astron. Soc. 438 (2014) L71–L75. arXiv:1309.3885, doi:10.1093/mnrasl/slt161.

[342] Y.-H. Sham, L.-M. Lin, P. T. Leung, Testing Universal Relations of Neutron Stars with a Nonlinear Matter-Gravity Coupling Theory, Astrophys. J. 781 (2014) 66. arXiv:1312.1011, doi:10.1088/ 0004-637X/781/2/66.

[343] P. Pani, E. Berti, Slowly rotating neutron stars in scalar-tensor theories, Phys. Rev. D 90 (2) (2014) 024025. arXiv:1405.4547, doi:10.1103/PhysRevD.90.024025.

[344] T. Gupta, B. Majumder, K. Yagi, N. Yunes, I-Love-Q relations for neutron stars in dynamical Chern Simons gravity, Classical and Quantum Gravity 35 (2) (2018) 025009. arXiv:1710.07862, doi:10.1088/1361-6382/aa9c68.

[345] D. D. Doneva, G. Pappas, Universal Relations and Alternative Gravity Theories, in: L. Rezzolla, P. Pizzochero, D. I. Jones, N. Rea, I. Vida˜na (Eds.), Astrophysics and Space Science Library, Vol. 457 of Astrophysics and Space Science Library, 2018, p. 737. arXiv:1709.08046, doi: 10.1007/978-3-319-97616-7\_13.

[346] J. M. Lattimer, B. F. Schutz, Constraining the Equation of State with Moment of Inertia Mea- surements, Astrophys. J. 629 (2005) 979–984. arXiv:astro-ph/0411470, doi:10.1086/431543.

[347] F. Ozel,¨ D. Psaltis, Z. Arzoumanian, S. Morsink, M. Baub¨ock, Measuring Neutron Star Radii via Pulse Profile Modeling with NICER, Astrophys. J. 832 (2016) 92. arXiv:1512.03067, doi: 10.3847/0004-637X/832/1/92.

76 [348] A. Maselli, V. Cardoso, V. Ferrari, L. Gualtieri, P. Pani, Equation-of-state-independent relations in neutron stars, Phys. Rev. D 88 (2) (2013) 023007. arXiv:1304.2052, doi:10.1103/PhysRevD. 88.023007.

[349] B. Biswas, S. Bose, Tidal deformability of an anisotropic compact star: Implications of GW170817, Phys. Rev. D 99 (10) (2019) 104002. arXiv:1903.04956, doi:10.1103/PhysRevD.99.104002.

[350] Z. Carson, K. Chatziioannou, C.-J. Haster, K. Yagi, N. Yunes, Equation-of-state insensitive relations after GW170817, Phys. Rev. D 99 (8) (2019) 083016. arXiv:1903.03909, doi: 10.1103/PhysRevD.99.083016.

[351] D. G. Ravenhall, C. J. Pethick, Neutron Star Moments of Inertia, Astrophys. J. 424 (1994) 846. doi:10.1086/173935.

[352] M. Bejger, P. Haensel, Moments of inertia for neutron and strange stars: Limits derived for the , Astron. Astrophys. 396 (2002) 917–921. arXiv:arXiv:astro-ph/0209151, doi: 10.1051/0004-6361:20021241.

[353] M. Baub¨ock, E. Berti, D. Psaltis, F. Ozel,¨ Relations between Neutron-star Parameters in the Hartle-Thorne Approximation, Astrophys. J. 777 (2013) 68. arXiv:1306.0569, doi:10.1088/ 0004-637X/777/1/68.

[354] T. K. Chan, A. P. O. Chan, P. T. Leung, Universality and stationarity of the I-Love relation for self-bound stars, Phys. Rev. D 93 (2) (2016) 024033. arXiv:1511.08566, doi:10.1103/ PhysRevD.93.024033.

[355] C. Breu, L. Rezzolla, Maximum mass, moment of inertia and compactness of relativistic stars, Mon. Not. R. Astron. Soc. 459 (2016) 646–656. arXiv:1601.06083, doi:10.1093/mnras/stw575.

[356] L. C. Stein, K. Yagi, N. Yunes, Three-hair Relations for Rotating Stars: Nonrelativistic Limit, Astrophys. J. 788 (2014) 15. arXiv:1312.4532, doi:10.1088/0004-637X/788/1/15.

[357] K. Yagi, K. Kyutoku, G. Pappas, N. Yunes, T. A. Apostolatos, Effective no-hair relations for neutron stars and quark stars: Relativistic results, Phys. Rev. D 89 (12) (2014) 124013. arXiv: 1403.6243, doi:10.1103/PhysRevD.89.124013.

[358] A. Bauswein, N. Stergioulas, Semi-analytic derivation of the threshold mass for prompt collapse in binary neutron-star mergers, Mon. Not. R. Astron. Soc. 471 (2017) 4956–4965. arXiv:1702. 02567, doi:10.1093/mnras/stx1983.

[359] P. Landry, B. Kumar, Constraints on the Moment of Inertia of PSR J0737-3039A from GW170817, Astrophys. J. Lett. 868 (2018) L22. arXiv:1807.04727, doi:10.3847/2041-8213/aaee76.

[360] B. Kumar, P. Landry, Inferring neutron star properties from GW170817 with universal relations, Phys. Rev. D 99 (12) (2019) 123026. arXiv:1902.04557, doi:10.1103/PhysRevD.99.123026.

[361] K. Yagi, N. Yunes, Binary Love relations, Classical and Quantum Gravity 33 (13) (2016) 13LT01. arXiv:1512.02639, doi:10.1088/0264-9381/33/13/13LT01.

[362] K. Yagi, N. Yunes, Approximate universal relations among tidal parameters for neutron star binaries, Classical and Quantum Gravity 34 (1) (2017) 015006. arXiv:1608.06187, doi:10. 1088/1361-6382/34/1/015006.

77 [363] L. Rezzolla, E. R. Most, L. R. Weih, Using Gravitational-wave Observations and Quasi-universal Relations to Constrain the Maximum Mass of Neutron Stars, Astrophys. J. Lett. 852 (2018) L25. arXiv:1711.00314, doi:10.3847/2041-8213/aaa401.

[364] S. De, D. Finstad, J. M. Lattimer, D. A. Brown, E. Berger, C. M. Biwer, Tidal Deformabilities and Radii of Neutron Stars from the Observation of GW170817, Physical Review Letters 121 (9) (2018) 091102. arXiv:1804.08583, doi:10.1103/PhysRevLett.121.091102.

[365] T. Zhao, J. M. Lattimer, Tidal deformabilities and neutron star mergers, Phys. Rev. D 98 (6) (2018) 063020. arXiv:1808.02858, doi:10.1103/PhysRevD.98.063020.

[366] S. A. Bhat, D. Bandyopadhyay, Neutron star equation of state and GW170817, Journal of Physics G Nuclear Physics 46 (1) (2019) 014003. arXiv:1807.06437, doi:10.1088/1361-6471/aaef45.

[367] M. Sieniawska, W. Turcza´nski,M. Bejger, J. L. Zdunik, Tidal deformability and other global parameters of compact stars with strong phase transitions, Astron. Astrophys. 622 (2019) A174. arXiv:1807.11581, doi:10.1051/0004-6361/201833969.

[368] S. Bernuzzi, A. Nagar, S. Balmelli, T. Dietrich, M. Ujevic, Quasiuniversal Properties of Neu- tron Star Mergers, Phys. Rev. Lett. 112 (20) (2014) 201101. arXiv:1402.6244, doi:10.1103/ PhysRevLett.112.201101.

[369] S. Bernuzzi, T. Dietrich, A. Nagar, Modeling the Complete Gravitational Wave Spectrum of Neutron Star Mergers, Phys. Rev. Lett. 115 (9) (2015) 091101. arXiv:1504.01764, doi:10. 1103/PhysRevLett.115.091101.

[370] K. Takami, L. Rezzolla, L. Baiotti, Spectral properties of the post-merger gravitational-wave signal from binary neutron stars, Phys. Rev. D 91 (6) (2015) 064001. arXiv:1412.3240, doi: 10.1103/PhysRevD.91.064001.

[371] L. Rezzolla, K. Takami, Gravitational-wave signal from binary neutron stars: A systematic analysis of the spectral properties, Phys. Rev. D 93 (12) (2016) 124051. arXiv:1604.00246, doi:10.1103/PhysRevD.93.124051.

[372] K. W. Tsang, T. Dietrich, C. Van Den Broeck, Modeling the postmerger gravitational wave signal and extracting binary properties from future binary neutron star detections, arXiv e-prints (2019) arXiv:1907.02424arXiv:1907.02424.

[373] K. Kiuchi, K. Kawaguchi, K. Kyutoku, Y. Sekiguchi, M. Shibata, Sub-radian-accuracy gravita- tional waves from coalescing binary neutron stars II: Systematic study on the equation of state, binary mass, and mass ratio, arXiv e-prints (2019) arXiv:1907.03790arXiv:1907.03790.

[374] K. Kiuchi, K. Kawaguchi, K. Kyutoku, Y. Sekiguchi, M. Shibata, K. Taniguchi, Sub-radian- accuracy gravitational waveforms of coalescing binary neutron stars in numerical relativity, Phys. Rev. D 96 (8) (2017) 084060. arXiv:1708.08926, doi:10.1103/PhysRevD.96.084060.

[375] A. Bauswein, H.-T. Janka, Measuring Neutron-Star Properties via Gravitational Waves from Neutron-Star Mergers, Phys. Rev. Lett. 108 (1) (2012) 011101. arXiv:1106.1616, doi:10.1103/ PhysRevLett.108.011101.

[376] A. Bauswein, H.-T. Janka, K. Hebeler, A. Schwenk, Equation-of-state dependence of the gravitational-wave signal from the ring-down phase of neutron-star mergers, Phys. Rev. D 86 (6) (2012) 063001. arXiv:1204.1888, doi:10.1103/PhysRevD.86.063001.

78 [377] A. Bauswein, H.-T. Janka, Measuring Neutron-Star Properties via Gravitational Waves from Neutron-Star Mergers, Phys. Rev. Lett. 108 (1) (2012) 011101. arXiv:1106.1616, doi:10.1103/ PhysRevLett.108.011101. [378] A. Bauswein, N. Stergioulas, H.-T. Janka, Revealing the high-density equation of state through binary neutron star mergers, Phys. Rev. D 90 (2) (2014) 023002. arXiv:1403.5301, doi:10. 1103/PhysRevD.90.023002. [379] K. Takami, L. Rezzolla, L. Baiotti, Constraining the Equation of State of Neutron Stars from Binary Mergers, Phys. Rev. Lett. 113 (9) (2014) 091104. arXiv:1403.5672, doi:10.1103/ PhysRevLett.113.091104. [380] T. Dietrich, S. Bernuzzi, M. Ujevic, B. Br¨ugmann,Numerical relativity simulations of neutron star merger remnants using conservative mesh refinement, Phys. Rev. D 91 (12) (2015) 124041. arXiv:1504.01266, doi:10.1103/PhysRevD.91.124041. [381] R. De Pietri, A. Feo, F. Maione, F. L¨offler,Modeling equal and unequal mass binary neutron star mergers using public codes, Phys. Rev. D 93 (6) (2016) 064047. arXiv:1509.08804, doi: 10.1103/PhysRevD.93.064047. [382] T. Dietrich, M. Ujevic, W. Tichy, S. Bernuzzi, B. Br¨ugmann,Gravitational waves and mass ejecta from binary neutron star mergers: Effect of the mass ratio, Phys. Rev. D 95 (2) (2017) 024029. arXiv:1607.06636, doi:10.1103/PhysRevD.95.024029. [383] F. Maione, R. De Pietri, A. Feo, F. L¨offler,Spectral analysis of gravitational waves from binary neutron star merger remnants, Phys. Rev. D 96 (6) (2017) 063011. arXiv:1707.03368, doi: 10.1103/PhysRevD.96.063011. [384] A. Bauswein, N. Stergioulas, H.-T. Janka, Exploring properties of high-density matter through remnants of neutron-star mergers, European Physical Journal A 52 (2016) 56. arXiv:1508.05493, doi:10.1140/epja/i2016-16056-7. [385] W. E. East, V. Paschalidis, F. Pretorius, A. Tsokaros, Binary neutron star mergers: effects of spin and post-merger dynamics, arXiv e-prints (2019) arXiv:1906.05288arXiv:1906.05288. [386] A. Bauswein, T. W. Baumgarte, H.-T. Janka, Prompt Merger Collapse and the Maximum Mass of Neutron Stars, Phys. Rev. Lett. 111 (13) (2013) 131101. arXiv:1307.5191, doi:10.1103/ PhysRevLett.111.131101. [387] A. Bauswein, O. Just, H.-T. Janka, N. Stergioulas, Neutron-star Radius Constraints from GW170817 and Future Detections, Astrophys. J. Lett. 850 (2017) L34. arXiv:1710.06843, doi:10.3847/2041-8213/aa9994. [388] S. K¨oppel, L. Bovard, L. Rezzolla, A General-relativistic Determination of the Threshold Mass to Prompt Collapse in Binary Neutron Star Mergers, Astrophys. J. Lett. 872 (1) (2019) L16. arXiv:1901.09977, doi:10.3847/2041-8213/ab0210. [389] N. Andersson, K. D. Kokkotas, Gravitational Waves and Pulsating Stars: What Can We Learn from Future Observations?, Phys. Rev. Lett. 77 (20) (1996) 4134–4137. arXiv:gr-qc/9610035, doi:10.1103/PhysRevLett.77.4134. [390] N. Andersson, K. D. Kokkotas, Towards gravitational wave , Mon. Not. R. As- tron. Soc. 299 (1998) 1059–1068. arXiv:arXiv:gr-qc/9711088, doi:10.1046/j.1365-8711. 1998.01840.x.

79 [391] O. Benhar, E. Berti, V. Ferrari, The imprint of the equation of state on the axial w-modes of oscillating neutron stars, Mon. Not. R. Astron. Soc. 310 (3) (1999) 797–803. arXiv:gr-qc/ 9901037, doi:10.1046/j.1365-8711.1999.02983.x.

[392] O. Benhar, V. Ferrari, L. Gualtieri, Gravitational wave asteroseismology revisited, Phys. Rev. D 70 (2004) 124015. arXiv:astro-ph/0407529, doi:10.1103/PhysRevD.70.124015.

[393] J. M. Lattimer, M. Prakash, The Physics of Neutron Stars, Science 304 (2004) 536–542. arXiv: arXiv:astro-ph/0405262, doi:10.1126/science.1090720.

[394] H. K. Lau, P. T. Leung, L. M. Lin, Inferring Physical Parameters of Compact Stars from their f-mode Gravitational Wave Signals, Astrophys. J. 714 (2010) 1234–1238. arXiv:0911.0131, doi: 10.1088/0004-637X/714/2/1234.

[395] T. K. Chan, Y. H. Sham, P. T. Leung, L. M. Lin, Multipolar universal relations between f - mode frequency and tidal deformability of compact stars, Phys. Rev. D 90 (12) (2014) 124023. arXiv:1408.3789, doi:10.1103/PhysRevD.90.124023.

[396] C. Chirenti, G. H. de Souza, W. Kastaun, Fundamental oscillation modes of neutron stars: Validity of universal relations, Phys. Rev. D 91 (4) (2015) 044034. arXiv:1501.02970, doi:10.1103/ PhysRevD.91.044034.

[397] D. D. Doneva, K. D. Kokkotas, Asteroseismology of rapidly rotating neutron stars: An alternative approach, Phys. Rev. D 92 (12) (2015) 124004. arXiv:1507.06606, doi:10.1103/PhysRevD.92. 124004.

[398] D.-H. Wen, B.-A. Li, H.-Y. Chen, N.-B. Zhang, GW170817 implications on the frequency and damping time of f -mode oscillations of neutron stars, Physical Review C 99 (4) (2019) 045806. arXiv:1901.03779, doi:10.1103/PhysRevC.99.045806.

[399] K. Chatziioannou, C.-J. Haster, A. Zimmerman, Measuring the neutron star tidal deformability with equation-of-state-independent relations and gravitational waves, Phys. Rev. D 97 (10) (2018) 104036. arXiv:1804.03221, doi:10.1103/PhysRevD.97.104036.

[400] S. Bernuzzi, A. Nagar, M. Thierfelder, B. Br¨ugmann,Tidal effects in binary neutron star coa- lescence, Phys. Rev. D 86 (4) (2012) 044030. arXiv:1205.3403, doi:10.1103/PhysRevD.86. 044030.

[401] W. Del Pozzo, T. G. F. Li, M. Agathos, C. Van Den Broeck, S. Vitale, Demonstrating the Fea- sibility of Probing the Neutron-Star Equation of State with Second-Generation Gravitational- Wave Detectors, Phys. Rev. Lett. 111 (7) (2013) 071101. arXiv:1307.8338, doi:10.1103/ PhysRevLett.111.071101.

[402] L. Wade, J. D. E. Creighton, E. Ochsner, B. D. Lackey, B. F. Farr, T. B. Littenberg, V. Raymond, Systematic and statistical errors in a Bayesian approach to the estimation of the neutron-star equation of state using advanced gravitational wave detectors, Phys. Rev. D 89 (10) (2014) 103012. arXiv:1402.5156, doi:10.1103/PhysRevD.89.103012.

[403] M. Agathos, J. Meidam, W. Del Pozzo, T. G. F. Li, M. Tompitak, J. Veitch, S. Vitale, C. Van Den Broeck, Constraining the neutron star equation of state with gravitational wave signals from coalescing binary neutron stars, Phys. Rev. D 92 (2) (2015) 023012. arXiv:1503.05405, doi:10.1103/PhysRevD.92.023012.

80 [404] K. Chatziioannou, K. Yagi, A. Klein, N. Cornish, N. Yunes, Probing the internal composition of neutron stars with gravitational waves, Phys. Rev. D 92 (10) (2015) 104008. arXiv:1508.02062, doi:10.1103/PhysRevD.92.104008.

[405] K. Hotokezaka, K. Kyutoku, Y.-i. Sekiguchi, M. Shibata, Measurability of the tidal deformability by gravitational waves from coalescing binary neutron stars, Phys. Rev. D 93 (6) (2016) 064082. arXiv:1603.01286, doi:10.1103/PhysRevD.93.064082.

[406] T. Dietrich, S. Bernuzzi, W. Tichy, Closed-form tidal approximants for binary neutron star grav- itational waveforms constructed from high-resolution numerical relativity simulations, Phys. Rev. D 96 (12) (2017) 121501. arXiv:1706.02969, doi:10.1103/PhysRevD.96.121501.

[407] K. Kawaguchi, K. Kiuchi, K. Kyutoku, Y. Sekiguchi, M. Shibata, K. Taniguchi, Frequency-domain gravitational waveform models for inspiraling binary neutron stars, Phys. Rev. D 97 (4) (2018) 044044. arXiv:1802.06518, doi:10.1103/PhysRevD.97.044044.

[408] R. Dudi, F. Pannarale, T. Dietrich, M. Hannam, S. Bernuzzi, F. Ohme, B. Br¨ugmann,Relevance of tidal effects and post-merger dynamics for binary neutron star parameter estimation, Phys. Rev. D 98 (8) (2018) 084061. arXiv:1808.09749, doi:10.1103/PhysRevD.98.084061.

[409] A. Samajdar, T. Dietrich, Waveform systematics for binary neutron star gravitational wave sig- nals: Effects of the point-particle baseline and tidal descriptions, Phys. Rev. D 98 (12) (2018) 124030. arXiv:1810.03936, doi:10.1103/PhysRevD.98.124030.

[410] I. Harry, T. Hinderer, Observing and measuring the neutron-star equation-of-state in spinning binary neutron star systems, Classical and Quantum Gravity 35 (14) (2018) 145010. arXiv: 1801.09972, doi:10.1088/1361-6382/aac7e3.

[411] T. Narikawa, N. Uchikata, K. Kawaguchi, K. Kiuchi, K. Kyutoku, M. Shibata, H. Tagoshi, Discrepancy in tidal deformability of GW170817 between the Advanced LIGO twins, arXiv e- printsarXiv:1812.06100.

[412] C. S. Kochanek, Coalescing binary neutron stars, Astrophys. J. 398 (1992) 234–247. doi:10. 1086/171851.

[413] D. Lai, A. G. Wiseman, Innermost stable circular orbit of inspiraling neutron-star binaries: Tidal effects, post-Newtonian effects, and the neutron-star equation of state, Phys. Rev. D 54 (1996) 3958–3964. arXiv:gr-qc/9609014, doi:10.1103/PhysRevD.54.3958.

[414] T. Damour, A. Nagar, Relativistic tidal properties of neutron stars, Phys. Rev. D 80 (8) (2009) 084035. arXiv:0906.0096, doi:10.1103/PhysRevD.80.084035.

[415] J. B. Hartle, Slowly Rotating Relativistic Stars. I. Equations of Structure, Astrophys. J. 150 (1967) 1005. doi:10.1086/149400.

[416] R. C. Tolman, Static Solutions of Einstein’s Field Equations for Spheres of Fluid, Physical Review 55 (4) (1939) 364–373. doi:10.1103/PhysRev.55.364.

[417] J. R. Oppenheimer, G. M. Volkoff, On Massive Neutron Cores, Physical Review 55 (4) (1939) 374–381. doi:10.1103/PhysRev.55.374.

[418] E.´ E.´ Flanagan, T. Hinderer, Constraining neutron-star tidal Love numbers with gravitational- wave detectors, Physical Review D 77 (2) (2008) 021502. arXiv:0709.1915, doi:10.1103/ PhysRevD.77.021502.

81 [419] C. Messenger, J. Read, Measuring a Cosmological Distance- Relationship Using Only Gravitational Wave Observations of Binary Neutron Star Coalescences, Phys. Rev. Lett. 108 (9) (2012) 091101. arXiv:1107.5725, doi:10.1103/PhysRevLett.108.091101. [420] E. Annala, T. Gorda, A. Kurkela, A. Vuorinen, Gravitational-Wave Constraints on the Neutron- Star-Matter Equation of State, Phys. Rev. Lett. 120 (17) (2018) 172703. arXiv:1711.02644, doi:10.1103/PhysRevLett.120.172703. [421] F. J. Fattoyev, J. Piekarewicz, C. J. Horowitz, Neutron Skins and Neutron Stars in the Mul- timessenger Era, Physical Review Letters 120 (17) (2018) 172702. arXiv:1711.06615, doi: 10.1103/PhysRevLett.120.172702. [422] E.-P. Zhou, X. Zhou, A. Li, Constraints on interquark interaction parameters with GW170817 in a binary scenario, Phys. Rev. D 97 (8) (2018) 083015. arXiv:1711.04312, doi: 10.1103/PhysRevD.97.083015. [423] C. Raithel, F. Ozel,¨ D. Psaltis, Tidal Deformability from GW170817 as a Direct Probe of the Neu- tron Star Radius, Astrophys. J. 857 (2018) L23. arXiv:1803.07687, doi:10.3847/2041-8213/ aabcbf. [424] M. Shibata, Effects of Tidal Resonances in Coalescing Compact Binary Systems, Progress of Theoretical Physics 91 (5) (1994) 871–883. doi:10.1143/ptp/91.5.871. [425] A. Reisenegger, P. Goldreich, Excitation of Neutron Star Normal Modes during Binary Inspiral, Astrophys. J. 426 (1994) 688. doi:10.1086/174105. [426] D. Lai, Resonant Oscillations and Tidal Heating in Coalescing Binary Neutron Stars, Mon. Not. R. Astron. Soc. 270 (1994) 611. arXiv:astro-ph/9404062, doi:10.1093/mnras/270.3.611. [427] K. D. Kokkotas, G. Schafer, Tidal and tidal-resonant effects in coalescing binaries, Mon.Not.Roy.Astron.Soc. 275 (2) (1995) 301–308. arXiv:gr-qc/9502034, doi:10.1093/mnras/ 275.2.301. [428] W. C. G. Ho, D. Lai, Resonant tidal excitations of rotating neutron stars in coalescing binaries, Mon. Not. R. Astron. Soc. 308 (1) (1999) 153–166. arXiv:astro-ph/9812116, doi:10.1046/j. 1365-8711.1999.02703.x. [429] T. Hinderer, A. Taracchini, F. Foucart, A. Buonanno, J. Steinhoff, M. Duez, L. E. Kidder, H. P. Pfeiffer, M. A. Scheel, B. Szilagyi, K. Hotokezaka, K. Kyutoku, M. Shibata, C. W. Carpenter, Effects of Neutron-Star Dynamic Tides on Gravitational Waveforms within the Effective-One- Body Approach, Phys. Rev. Lett. 116 (18) (2016) 181101. arXiv:1602.00599, doi:10.1103/ PhysRevLett.116.181101. [430] J. Steinhoff, T. Hinderer, A. Buonanno, A. Taracchini, Dynamical Tides in General Relativ- ity: Effective Action and Effective-One-Body Hamiltonian, Phys. Rev. D 94 (10) (2016) 104028. arXiv:1608.01907, doi:10.1103/PhysRevD.94.104028. [431] H. Yu, N. N. Weinberg, Resonant tidal excitation of superfluid neutron stars in coalescing binaries, Mon. Not. R. Astron. Soc. 464 (3) (2017) 2622–2637. arXiv:1610.00745, doi:10.1093/mnras/ stw2552. [432] H. Yu, N. N. Weinberg, Dynamical tides in coalescing superfluid neutron star binaries with hy- peron cores and their detectability with third-generation gravitational-wave detectors, Mon. Not. R. Astron. Soc. 470 (1) (2017) 350–360. arXiv:1705.04700, doi:10.1093/mnras/stx1188.

82 [433] W. Xu, D. Lai, Resonant tidal excitation of oscillation modes in merging binary neutron stars: Inertial-gravity modes, Physical Review D 96 (8) (2017) 083005. arXiv:1708.01839, doi:10. 1103/PhysRevD.96.083005.

[434] P. Schmidt, T. Hinderer, A Frequency Domain Model of f-Mode Dynamic Tides in Gravitational Waveforms from Compact Binaries, arXiv e-prints (2019) arXiv:1905.00818arXiv:1905.00818.

[435] E.´ E.´ Flanagan, E.´ Racine, Gravitomagnetic resonant excitation of Rossby modes in coalescing neutron star binaries, Physical Review D 75 (4) (2007) 044001. arXiv:gr-qc/0601029, doi: 10.1103/PhysRevD.75.044001.

[436] D. Tsang, J. S. Read, T. Hinderer, A. L. Piro, R. Bondarescu, Resonant Shattering of Neutron Star Crusts, Phys. Rev. Lett. 108 (1) (2012) 011102. arXiv:1110.0467, doi:10.1103/PhysRevLett. 108.011102.

[437] D. Tsang, Shattering Flares During Close Encounters of Neutron Stars, Astrophys. J. 777 (2013) 103. arXiv:1307.3554, doi:10.1088/0004-637X/777/2/103.

[438] R. Essick, S. Vitale, N. N. Weinberg, Impact of the tidal p -g instability on the gravitational wave signal from coalescing binary neutron stars, Physical Review D 94 (10) (2016) 103012. arXiv:1609.06362, doi:10.1103/PhysRevD.94.103012.

[439] A. Nagar, S. Bernuzzi, W. Del Pozzo, G. Riemenschneider, S. Akcay, G. Carullo, P. Fleig, S. Babak, K. W. Tsang, M. Colleoni, F. Messina, G. Pratten, D. Radice, P. Rettegno, M. Agathos, E. Fauchon-Jones, M. Hannam, S. Husa, T. Dietrich, P. Cerd´a-Duran,J. A. Font, F. Pannarale, P. Schmidt, T. Damour, Time-domain effective-one-body gravitational waveforms for coalescing compact binaries with nonprecessing spins, tides, and self-spin effects, Phys. Rev. D 98 (10) (2018) 104052. arXiv:1806.01772, doi:10.1103/PhysRevD.98.104052.

[440] T. Abdelsalhin, L. Gualtieri, P. Pani, Post-Newtonian spin-tidal couplings for compact binaries, Phys. Rev. D 98 (10) (2018) 104046. arXiv:1805.01487, doi:10.1103/PhysRevD.98.104046.

[441] X. Jim´enezForteza, T. Abdelsalhin, P. Pani, L. Gualtieri, Impact of high-order tidal terms on binary neutron-star waveforms, Phys. Rev. D 98 (12) (2018) 124014. arXiv:1807.08016, doi: 10.1103/PhysRevD.98.124014.

[442] T. Cullen, I. Harry, J. Read, E. Flynn, Matter effects on LIGO/Virgo searches for gravitational waves from merging neutron stars, Classical and Quantum Gravity 34 (24) (2017) 245003. arXiv: 1708.04359, doi:10.1088/1361-6382/aa9424.

[443] E. Poisson, Gravitational waves from inspiraling compact binaries: The quadrupole-moment term, Phys. Rev. D 57 (1998) 5287–5290. arXiv:gr-qc/9709032, doi:10.1103/PhysRevD.57.5287.

[444] A. Boh´e,G. Faye, S. Marsat, E. K. Porter, Quadratic-in-spin effects in the orbital dynamics and gravitational-wave energy flux of compact binaries at the 3PN order, Classical and Quantum Gravity 32 (19) (2015) 195010. arXiv:1501.01529, doi:10.1088/0264-9381/32/19/195010.

[445] T. Dietrich, S. Khan, R. Dudi, S. J. Kapadia, P. Kumar, A. Nagar, F. Ohme, F. Pannarale, A. Samajdar, S. Bernuzzi, G. Carullo, W. Del Pozzo, M. Haney, C. Markakis, M. P¨urrer, G. Riemenschneider, Y. E. Setyawati, K. W. Tsang, C. Van Den Broeck, Matter imprints in waveform models for neutron star binaries: Tidal and self-spin effects, Phys. Rev. D 99 (2) (2019) 024029. arXiv:1804.02235, doi:10.1103/PhysRevD.99.024029.

83 [446] A. Samajdar, T. Dietrich, Waveform systematics for binary neutron star gravitational wave sig- nals: Effects of spin, precession, and the observation of electromagnetic counterparts, Phys. Rev. D 100 (2) (2019) 024046. arXiv:1905.03118, doi:10.1103/PhysRevD.100.024046.

[447] A. Tsokaros, M. Ruiz, V. Paschalidis, S. L. Shapiro, K. Ury¯u,Effect of spin on the inspiral of binary neutron stars, Phys. Rev. D 100 (2) (2019) 024061. arXiv:1906.00011, doi:10.1103/ PhysRevD.100.024061.

[448] J. Veitch, V. Raymond, B. Farr, W. Farr, P. Graff, S. Vitale, B. Aylott, K. Blackburn, N. Chris- tensen, M. Coughlin, W. Del Pozzo, F. Feroz, J. Gair, C.-J. Haster, V. Kalogera, T. Littenberg, I. Mandel, R. O’Shaughnessy, M. Pitkin, C. Rodriguez, C. R¨over, T. Sidery, R. Smith, M. Van Der Sluys, A. Vecchio, W. Vousden, L. Wade, Parameter estimation for compact binaries with ground-based gravitational-wave observations using the LALInference software library, Phys. Rev. D 91 (4) (2015) 042003. arXiv:1409.7215, doi:10.1103/PhysRevD.91.042003.

[449] A. Buonanno, T. Damour, Effective one-body approach to general relativistic two-body dynamics, Phys. Rev. D 59 (8) (1999) 084006. arXiv:gr-qc/9811091, doi:10.1103/PhysRevD.59.084006.

[450] A. Buonanno, T. Damour, Transition from inspiral to plunge in binary black hole coalescences, Phys. Rev. D 62 (2000) 064015. arXiv:gr-qc/0001013, doi:10.1103/PhysRevD.62.064015.

[451] P. Ajith, S. Babak, Y. Chen, M. Hewitson, B. Krishnan, J. T. Whelan, B. Br¨ugmann,P. Diener, J. Gonzalez, M. Hannam, S. Husa, M. Koppitz, D. Pollney, L. Rezzolla, L. Santamar´ıa,A. M. Sintes, U. Sperhake, J. Thornburg, A phenomenological template family for black-hole coalescence waveforms, Classical and Quantum Gravity 24 (19) (2007) S689–S699. arXiv:0704.3764, doi: 10.1088/0264-9381/24/19/S31.

[452] P. Ajith, S. Babak, Y. Chen, M. Hewitson, B. Krishnan, A. M. Sintes, J. T. Whelan, B. Br¨ugmann, P. Diener, N. Dorband, J. Gonzalez, M. Hannam, S. Husa, D. Pollney, L. Rezzolla, L. Santamar´ıa, U. Sperhake, J. Thornburg, Template bank for gravitational waveforms from coalescing binary black holes: Nonspinning binaries, Phys. Rev. D 77 (10) (2008) 104017. arXiv:0710.2335, doi:10.1103/PhysRevD.77.104017.

[453] L. Blanchet, Gravitational Radiation from Post-Newtonian Sources and Inspiralling Compact Bi- naries, Living Reviews in Relativity 17 (2014) 2. arXiv:1310.1528, doi:10.12942/lrr-2014-2.

[454] T. Damour, P. Jaranowski, G. Sch¨afer,Conservative dynamics of two-body systems at the fourth post-Newtonian approximation of general relativity, Phys. Rev. D 93 (8) (2016) 084014. arXiv: 1601.01283, doi:10.1103/PhysRevD.93.084014.

[455] T. Marchand, L. Bernard, L. Blanchet, G. Faye, Ambiguity-free completion of the equations of motion of compact binary systems at the fourth post-Newtonian order, Phys. Rev. D 97 (4) (2018) 044023. arXiv:1707.09289, doi:10.1103/PhysRevD.97.044023.

[456] P. Schmidt, F. Ohme, M. Hannam, Towards models of gravitational waveforms from generic binaries: II. Modelling precession effects with a single effective precession parameter, Phys. Rev. D 91 (2) (2015) 024043. arXiv:1408.1810, doi:10.1103/PhysRevD.91.024043.

[457] S. Khan, S. Husa, M. Hannam, F. Ohme, M. P¨urrer,X. J. Forteza, A. Boh´e,Frequency-domain gravitational waves from nonprecessing black-hole binaries. II. A phenomenological model for the advanced detector era, Phys. Rev. D 93 (4) (2016) 044007. arXiv:1508.07253, doi:10.1103/ PhysRevD.93.044007.

84 [458] T. Damour, A. Nagar, Effective one body description of tidal effects in inspiralling compact binaries, Phys. Rev. D 81 (8) (2010) 084016. arXiv:0911.5041, doi:10.1103/PhysRevD.81. 084016.

[459] J. Vines, E. E. Flanagan, T. Hinderer, Post-1-Newtonian tidal effects in the gravitational waveform from binary inspirals, Phys. Rev. D 83 (2011) 084051. arXiv:1101.1673, doi: 10.1103/PhysRevD.83.084051.

[460] T. Damour, A. Nagar, L. Villain, Measurability of the tidal polarizability of neutron stars in late-inspiral gravitational-wave signals, Phys. Rev. D 85 (12) (2012) 123007. arXiv:1203.4352, doi:10.1103/PhysRevD.85.123007.

[461] D. Bini, T. Damour, G. Faye, Effective action approach to higher-order relativistic tidal interac- tions in binary systems and their effective one body description, Phys. Rev. D 85 (2012) 124034. arXiv:1202.3565, doi:10.1103/PhysRevD.85.124034.

[462] D. Bini, T. Damour, Gravitational self-force corrections to two-body tidal interactions and the effective one-body formalism, Phys. Rev. D 90 (12) (2014) 124037. arXiv:1409.6933, doi: 10.1103/PhysRevD.90.124037.

[463] S. Bernuzzi, A. Nagar, T. Dietrich, T. Damour, Modeling the Dynamics of Tidally Interacting Binary Neutron Stars up to the Merger, Phys. Rev. Lett. 114 (16) (2015) 161103. arXiv:1412. 4553, doi:10.1103/PhysRevLett.114.161103.

[464] A. Boh´e, L. Shao, A. Taracchini, A. Buonanno, S. Babak, I. W. Harry, I. Hinder, S. Os- sokine, M. P¨urrer,V. Raymond, T. Chu, H. Fong, P. Kumar, H. P. Pfeiffer, M. Boyle, D. A. Hemberger, L. E. Kidder, G. Lovelace, M. A. Scheel, B. Szil´agyi,Improved effective-one-body model of spinning, nonprecessing binary black holes for the era of gravitational-wave astro- physics with advanced detectors, Phys. Rev. D 95 (4) (2017) 044028. arXiv:1611.03703, doi:10.1103/PhysRevD.95.044028.

[465] S. Babak, A. Taracchini, A. Buonanno, Validating the effective-one-body model of spinning, precessing binary black holes against numerical relativity, Phys. Rev. D 95 (2) (2017) 024010. arXiv:1607.05661, doi:10.1103/PhysRevD.95.024010.

[466] A. Nagar, G. Riemenschneider, G. Pratten, Impact of Numerical Relativity information on effective-one-body waveform models, Phys. Rev. D96 (8) (2017) 084045. arXiv:1703.06814, doi:10.1103/PhysRevD.96.084045.

[467] R. Cotesta, A. Buonanno, A. Boh´e,A. Taracchini, I. Hinder, S. Ossokine, Enriching the symphony of gravitational waves from binary black holes by tuning higher harmonics, Phys. Rev. D 98 (8) (2018) 084028. arXiv:1803.10701, doi:10.1103/PhysRevD.98.084028.

[468] K. Hotokezaka, K. Kyutoku, H. Okawa, M. Shibata, Exploring tidal effects of coalescing binary neutron stars in numerical relativity. II. Long-term simulations, Phys. Rev. D 91 (6) (2015) 064060. arXiv:1502.03457, doi:10.1103/PhysRevD.91.064060.

[469] B. D. Lackey, S. Bernuzzi, C. R. Galley, J. Meidam, C. Van Den Broeck, Effective-one-body waveforms for binary neutron stars using surrogate models, Phys. Rev. D 95 (10) (2017) 104036. arXiv:1610.04742, doi:10.1103/PhysRevD.95.104036.

[470] F. Pannarale, The Black Hole Remnant of Black Hole-Neutron Star Coalescing Binaries, Phys. Rev. D 88 (2013) 104025. arXiv:1208.5869, doi:10.1103/PhysRevD.88.104025.

85 [471] B. D. Lackey, K. Kyutoku, M. Shibata, P. R. Brady, J. L. Friedman, Extracting equation of state parameters from black hole-neutron star mergers: Nonspinning black holes, Phys. Rev. D 85 (4) (2012) 044061. arXiv:1109.3402, doi:10.1103/PhysRevD.85.044061.

[472] B. D. Lackey, K. Kyutoku, M. Shibata, P. R. Brady, J. L. Friedman, Extracting equation of state parameters from black hole-neutron star mergers: Aligned-spin black holes and a preliminary waveform model, Phys. Rev. D 89 (4) (2014) 043009. arXiv:1303.6298, doi:10.1103/PhysRevD. 89.043009.

[473] K. Barkett, M. A. Scheel, R. Haas, C. D. Ott, S. Bernuzzi, D. A. Brown, B. Szil´agyi,J. D. Kaplan, J. Lippuner, C. D. Muhlberger, F. Foucart, M. D. Duez, Gravitational waveforms for neutron star binaries from binary black hole simulations, Phys. Rev. D 93 (4) (2016) 044064. arXiv:1509.05782, doi:10.1103/PhysRevD.93.044064.

[474] T. Hinderer, B. D. Lackey, R. N. Lang, J. S. Read, Tidal deformability of neutron stars with realistic equations of state and their gravitational wave signatures in binary inspiral, Phys. Rev. D 81 (2010) 123016. arXiv:0911.3535, doi:10.1103/PhysRevD.81.123016.

[475] P. C. Peters, J. Mathews, Gravitational Radiation from Point Masses in a Keplerian Orbit, Phys- ical Review 131 (1) (1963) 435–440. doi:10.1103/PhysRev.131.435.

[476] P. C. Peters, Gravitational Radiation and the Motion of Two Point Masses, Physical Review 136 (4B) (1964) 1224–1232. doi:10.1103/PhysRev.136.B1224.

[477] I. Kowalska, T. Bulik, K. Belczynski, M. Dominik, D. Gondek-Rosinska, The eccentricity dis- tribution of compact binaries, Astron. Astrophys. 527 (2011) A70. arXiv:1010.0511, doi: 10.1051/0004-6361/201015777.

[478] A. C. Fabian, J. E. Pringle, M. J. Rees, Tidal capture formation of binary systems and X-ray sources in globular clusters, Mon. Not. R. Astron. Soc. 172 (1975) 15p–18p. doi:10.1093/mnras/ 172.1.15P.

[479] D. Pooley, W. H. G. Lewin, S. F. Anderson, H. Baumgardt, A. V. Filippenko, B. M. Gaensler, L. Homer, P. Hut, V. M. Kaspi, J. Makino, B. Margon, S. McMillan, S. Portegies Zwart, M. van der Klis, F. Verbunt, Dynamical Formation of Close Binary Systems in Globular Clusters, Astrophys. J. Lett. 591 (2003) L131–L134. arXiv:astro-ph/0305003, doi:10.1086/377074.

[480] R. M. O’Leary, B. Kocsis, A. Loeb, Gravitational waves from scattering of stellar-mass black holes in galactic nuclei, Mon. Not. R. Astron. Soc. 395 (2009) 2127–2146. arXiv:0807.2638, doi:10.1111/j.1365-2966.2009.14653.x.

[481] W. H. Lee, E. Ramirez-Ruiz, G. van de Ven, Short Gamma-ray Bursts from Dynamically Assem- bled Compact Binaries in Globular Clusters: Pathways, Rates, Hydrodynamics, and Cosmological Setting, Astrophys. J. 720 (2010) 953–975. arXiv:0909.2884, doi:10.1088/0004-637X/720/1/ 953.

[482] B. Kocsis, J. Levin, Repeated bursts from relativistic scattering of compact objects in galactic nuclei, Phys. Rev. D 85 (12) (2012) 123005. arXiv:1109.4170, doi:10.1103/PhysRevD.85. 123005.

[483] J. Samsing, M. MacLeod, E. Ramirez-Ruiz, The Formation of Eccentric Compact Binary Inspirals and the Role of Gravitational Wave Emission in Binary-Single Stellar Encounters, Astrophys. J. 784 (2014) 71. arXiv:1308.2964, doi:10.1088/0004-637X/784/1/71.

86 [484] G. Fragione, E. Grishin, N. W. C. Leigh, H. B. Perets, R. Perna, Black Hole and Neutron Star Mergers in Galactic Nuclei, Mon. Not. R. Astron. Soc. (2019) 1560arXiv:1811.10627, doi: 10.1093/mnras/stz1651.

[485] N. Seto, Highly Eccentric Kozai Mechanism and Gravitational-Wave Observation for Neutron- Star Binaries, Physical Review Letters 111 (6) (2013) 061106. arXiv:1304.5151, doi:10.1103/ PhysRevLett.111.061106.

[486] D. Radice, L. Rezzolla, F. Galeazzi, Beyond second-order convergence in simulations of binary neutron stars in full general-relativity, Mon. Not. R. Astron. Soc. L. 437 (2014) L46–L50. arXiv: 1306.6052, doi:10.1093/mnrasl/slt137.

[487] L. Dai, T. Venumadhav, B. Zackay, Parameter Estimation for GW170817 using Relative Binning, arXiv e-printsarXiv:1806.08793.

[488] D. Radice, A. Perego, F. Zappa, S. Bernuzzi, GW170817: Joint Constraint on the Neutron Star Equation of State from Multimessenger Observations, Astrophys. J. Lett. 852 (2018) L29. arXiv:1711.03647, doi:10.3847/2041-8213/aaa402.

[489] D. Radice, L. Dai, Multimessenger parameter estimation of GW170817, European Physical Jour- nal A 55 (4) (2019) 50. arXiv:1810.12917, doi:10.1140/epja/i2019-12716-4.

[490] M. W. Coughlin, T. Dietrich, Z. Doctor, D. Kasen, S. Coughlin, A. Jerkstrand, G. Leloudas, O. McBrien, B. D. Metzger, R. O’Shaughnessy, S. J. Smartt, Constraints on the neutron star equation of state from AT2017gfo using radiative transfer simulations, Mon. Not. R. Astron. Soc. 480 (2018) 3871–3878. arXiv:1805.09371, doi:10.1093/mnras/sty2174.

[491] M. W. Coughlin, T. Dietrich, B. Margalit, B. D. Metzger, Multi-messenger Bayesian parameter inference of a binary neutron-star merger, arXiv e-printsarXiv:1812.04803.

[492] Y.-Z. Wang, D.-S. Shao, J.-L. Jiang, S.-P. Tang, X.-X. Ren, F.-W. Zhang, Z.-P. Jin, Y.-Z. Fan, D.-M. Wei, GW170817: The Energy Extraction Process of the Off-axis Relativistic Outflow and the Constraint on the Equation of State of Neutron Stars, Astrophys. J. 877 (1) (2019) 2. arXiv: 1811.02558, doi:10.3847/1538-4357/ab1914.

[493] B. Margalit, B. D. Metzger, The Multi-messenger Matrix: The Future of Neutron Star Merger Constraints on the Nuclear Equation of State, Astrophys. J. Lett. 880 (1) (2019) L15. arXiv: 1904.11995, doi:10.3847/2041-8213/ab2ae2.

[494] L. Gond´an,B. Kocsis, P. Raffai, Z. Frei, Accuracy of Estimating Highly Eccentric Binary Black Hole Parameters with Gravitational-wave Detections, Astrophys. J. 855 (2018) 34. arXiv:1705. 10781, doi:10.3847/1538-4357/aaad0e.

[495] L. Gond´an,B. Kocsis, Measurement Accuracy of Inspiraling Eccentric Neutron Star and Black Hole Binaries Using Gravitational Waves, Astrophys. J. 871 (2019) 178. arXiv:1809.00672, doi:10.3847/1538-4357/aaf893.

[496] G. Pratten, P. Schmidt, T. Hinderer, Gravitational-Wave Asteroseismology with Fundamen- tal Modes from Compact Binary Inspirals, arXiv e-prints (2019) arXiv:1905.00817arXiv:1905. 00817.

[497] J. M. Lattimer, F. Mackie, D. G. Ravenhall, D. N. Schramm, The decompression of cold neutron star matter., Astrophys. J. 213 (1977) 225–233. doi:10.1086/155148.

87 [498] M. Arnould, S. Goriely, K. Takahashi, The r-process of : Astrophysics and nuclear physics achievements and mysteries, Physics Reports 450 (2007) 97–213. arXiv: 0705.4512, doi:10.1016/j.physrep.2007.06.002.

[499] M. Turner, Tidal generation of gravitational waves from orbiting Newtonian stars. I - General formalism, Astrophys. J. 216 (1977) 914–929. doi:10.1086/155536.

[500] M. Turner, Gravitational radiation from point-masses in unbound orbits - Newtonian results, Astrophys. J. 216 (1977) 610–619. doi:10.1086/155501.

[501] C. Chirenti, R. Gold, M. C. Miller, Gravitational waves from f-modes excited by the inspiral of highly eccentric neutron star binaries, Astrophys. J. 837 (1) (2017) 67. arXiv:1612.07097, doi:10.3847/1538-4357/aa5ebb.

[502] The LIGO Scientific Collaboration, the Virgo Collaboration, B. P. Abbott, R. Abbott, T. D. Abbott, F. Acernese, K. Ackley, C. Adams, T. Adams, P. Addesso, R. X. Adhikari, V. B. Adya, et al., Search for Gravitational Waves from a Long-lived Remnant of the Binary Neu- tron Star Merger GW170817, Astrophys. J. 875 (2) (2019) 160. arXiv:1810.02581, doi: 10.3847/1538-4357/ab0f3d.

[503] M. G. Alford, L. Bovard, M. Hanauske, L. Rezzolla, K. Schwenzer, Viscous dissipation and heat conduction in binary neutron-star mergers, Phys. Rev. Lett. 120 (2018) 041101. arXiv:1707. 09475, doi:10.1103/PhysRevLett.120.041101.

[504] D. Radice, A. Perego, S. Bernuzzi, B. Zhang, Long-lived remnants from binary neutron star mergers, Mon. Not. R. Astron. Soc. 481 (2018) 3670–3682. arXiv:1803.10865, doi:10.1093/ mnras/sty2531.

[505] M. Alford, A. Harutyunyan, A. Sedrakian, Bulk viscosity of baryonic matter with trapped neu- trinos, arXiv e-prints (2019) arXiv:1907.04192arXiv:1907.04192.

[506] M. G. Alford, S. P. Harris, Damping of density oscillations in neutrino-transparent nuclear matter, arXiv e-prints (2019) arXiv:1907.03795arXiv:1907.03795.

[507] S. Bose, K. Chakravarti, L. Rezzolla, B. S. Sathyaprakash, K. Takami, Neutron-Star Radius from a Population of Binary Neutron Star Mergers, Phys. Rev. Lett. 120 (3) (2018) 031102. arXiv:1705.10850, doi:10.1103/PhysRevLett.120.031102.

[508] F. Maione, R. De Pietri, A. Feo, F. L¨offler,Binary neutron star merger simulations with different initial orbital frequency and equation of state, Classical and Quantum Gravity 33 (17) (2016) 175009. arXiv:1605.03424, doi:10.1088/0264-9381/33/17/175009.

[509] A. Feo, R. De Pietri, F. Maione, F. L¨offler,Modeling mergers of known galactic systems of binary neutron stars, Class. Quantum Grav. 34 (3) (2017) 034001. arXiv:1608.02810, doi: 10.1088/1361-6382/aa51fa.

[510] A. Bauswein, N. Stergioulas, Unified picture of the post-merger dynamics and gravitational wave emission in neutron star mergers, Phys. Rev. D 91 (12) (2015) 124056. arXiv:1502.03176, doi:10.1103/PhysRevD.91.124056.

[511] A. Bauswein, N. Stergioulas, Spectral classification of gravitational-wave emission and equation of state constraints in binary neutron star mergers, arXiv e-prints (2019) arXiv:1901.06969arXiv: 1901.06969.

88 [512] N. Stergioulas, A. Bauswein, K. Zagkouris, H.-T. Janka, Gravitational waves and non- axisymmetric oscillation modes in mergers of compact object binaries, Mon. Not. R. Astron. Soc. 418 (2011) 427–436. arXiv:1105.0368, doi:10.1111/j.1365-2966.2011.19493.x. [513] M. Bezares, D. Vigano, C. Palenzuela, Signatures of dark matter cores in binary neutron star mergers, arXiv e-prints (2019) arXiv:1905.08551arXiv:1905.08551. [514] J. Ellis, A. Hektor, G. H¨utsi,K. Kannike, L. Marzola, M. Raidal, V. Vaskonen, Search for dark matter effects on gravitational signals from neutron star mergers, Physics Letters B 781 (2018) 607–610. arXiv:1710.05540, doi:10.1016/j.physletb.2018.04.048. [515] R. Foot, Mirror Matter-Type Dark Matter, International Journal of Modern Physics D 13 (2004) 2161–2192. arXiv:astro-ph/0407623, doi:10.1142/S0218271804006449. [516] J. Fan, A. Katz, L. Randall, M. Reece, Double-Disk Dark Matter, Physics of the Dark Universe 2 (2013) 139–156. arXiv:1303.1521, doi:10.1016/j.dark.2013.07.001. [517] J. Pollack, D. N. Spergel, P. J. Steinhardt, Supermassive Black Holes from Ultra-strongly Self- interacting Dark Matter, Astrophys. J. 804 (2015) 131. arXiv:1501.00017, doi:10.1088/ 0004-637X/804/2/131. [518] R. De Pietri, A. Feo, J. A. Font, F. L¨offler,F. Maione, M. Pasquali, N. Stergioulas, Convective Excitation of Inertial Modes in Binary Neutron Star Mergers, Phys. Rev. Lett. 120 (22) (2018) 221101. arXiv:1802.03288, doi:10.1103/PhysRevLett.120.221101. [519] K. Kiuchi, K. Kyutoku, M. Shibata, K. Taniguchi, Revisiting the Lower Bound on Tidal De- formability Derived by AT 2017gfo, Astrophys. J. 876 (2) (2019) L31. arXiv:1903.01466, doi:10.3847/2041-8213/ab1e45. [520] S. Klimenko, G. Vedovato, M. Drago, F. Salemi, V. Tiwari, G. A. Prodi, C. Lazzaro, K. Ackley, S. Tiwari, C. F. Da Silva, G. Mitselmakher, Method for detection and reconstruction of gravita- tional wave transients with networks of advanced detectors, Phys. Rev. D 93 (4) (2016) 042004. arXiv:1511.05999, doi:10.1103/PhysRevD.93.042004. [521] S. Vinciguerra, M. Drago, G. A. Prodi, S. Klimenko, C. Lazzaro, V. Necula, F. Salemi, V. Tiwari, M. C. Tringali, G. Vedovato, Enhancing the significance of gravitational wave bursts through signal classification, Classical and Quantum Gravity 34 (9) (2017) 094003. arXiv:1702.03208, doi:10.1088/1361-6382/aa6654. [522] P. J. Easter, P. D. Lasky, A. R. Casey, L. Rezzolla, K. Takami, Computing fast and reliable gravitational waveforms of binary neutron star merger remnants, Phys. Rev. D 100 (4) (2019) 043005. arXiv:1811.11183, doi:10.1103/PhysRevD.100.043005. [523] N. J. Cornish, T. B. Littenberg, Bayeswave: Bayesian inference for gravitational wave bursts and instrument glitches, Classical and Quantum Gravity 32 (13) (2015) 135012. arXiv:1410.3835, doi:10.1088/0264-9381/32/13/135012. [524] T. B. Littenberg, N. J. Cornish, Bayesian inference for spectral estimation of gravitational wave detector noise, Phys. Rev. D 91 (8) (2015) 084034. arXiv:1410.3852, doi:10.1103/PhysRevD. 91.084034. [525] Y. Sekiguchi, K. Kiuchi, K. Kyutoku, M. Shibata, Gravitational Waves and Neutrino Emission from the Merger of Binary Neutron Stars, Phys. Rev. Lett. 107 (5) (2011) 051102. arXiv: 1105.2125, doi:10.1103/PhysRevLett.107.051102.

89 [526] K. Kiuchi, Y. Sekiguchi, K. Kyutoku, M. Shibata, Gravitational waves, neutrino emissions and effects of hyperons in binary neutron star mergers, Class. Quantum Grav. 29 (12) (2012) 124003. arXiv:1206.0509, doi:10.1088/0264-9381/29/12/124003.

[527] D. Radice, S. Bernuzzi, W. Del Pozzo, L. F. Roberts, C. D. Ott, Probing Extreme-density Matter with Gravitational-wave Observations of Binary Neutron Star Merger Remnants, Astrophys. J. Lett. 842 (2017) L10. arXiv:1612.06429, doi:10.3847/2041-8213/aa775f.

[528] K. S. Cheng, Z. G. Dai, Conversion of Neutron Stars to Strange Stars as a Possible Origin of γ-Ray Bursts, Physical Review Letters 77 (1996) 1210–1213. arXiv:astro-ph/9510073, doi: 10.1103/PhysRevLett.77.1210.

[529] I. Bombaci, B. Datta, Conversion of Neutron Stars to Strange Stars as the Central Engine of Gamma-Ray Bursts, Astrophys. J. Lett. 530 (2000) L69–L72. arXiv:astro-ph/0001478, doi: 10.1086/312497.

[530] I. N. Mishustin, M. Hanauske, A. Bhattacharyya, L. M. Satarov, H. St¨ocker, W. Greiner, Catas- trophic rearrangement of a compact star due to the quark core formation, Physics Letters B 552 (2003) 1–8. arXiv:hep-ph/0210422, doi:10.1016/S0370-2693(02)03108-8.

[531] A. Bauswein, N.-U. F. Bastian, D. Blaschke, K. Chatziioannou, J. A. e. Clark, T. Fischer, H.-T. Janka, O. Just, M. Oertel, N. Stergioulas, Equation-of-state Constraints and the QCD Phase Tran- sition in the Era of Gravitational-Wave Astronomy, arXiv e-prints (2019) arXiv:1904.01306arXiv: 1904.01306.

[532] E. R. Most, L. R. Weih, L. Rezzolla, J. Schaffner-Bielich, New Constraints on Radii and Tidal Deformabilities of Neutron Stars from GW170817, Phys. Rev. Lett. 120 (26) (2018) 261103. arXiv:1803.00549, doi:10.1103/PhysRevLett.120.261103.

[533] V. A. Dexheimer, S. Schramm, A Novel Approach to Model Hybrid Stars, Phys. Rev. C81 (2010) 045201. arXiv:0901.1748, doi:10.1103/PhysRevC.81.045201.

[534] T. Fischer, N.-U. F. Bastian, M.-R. Wu, P. Baklanov, E. Sorokina, S. Blinnikov, S. Typel, T. Kl¨ahn, D. B. Blaschke, Quark deconfinement as a supernova explosion engine for mas- sive blue supergiant stars, 2 (2018) 980–986. arXiv:1712.08788, doi: 10.1038/s41550-018-0583-0.

[535] S. Banik, M. Hempel, D. Bandyopadhyay, New Hyperon Equations of State for Supernovae and Neutron Stars in Density-dependent Hadron Field Theory, Astrohys. J. Suppl. 214 (2014) 22. arXiv:1404.6173, doi:10.1088/0067-0049/214/2/22.

[536] M. Fortin, M. Oertel, C. Providˆencia,Hyperons in hot dense matter: what do the constraints tell us for equation of state?, Pub. Astron. Soc. Aust. 35 (2018) E044. arXiv:1711.09427, doi:10.1017/pasa.2018.32.

[537] M. Marques, M. Oertel, M. Hempel, J. Novak, New temperature dependent hyperonic equation of state: Application to rotating neutron star models and I -Q relations, Phys. Rev. C 96 (4) (2017) 045806. arXiv:1706.02913, doi:10.1103/PhysRevC.96.045806.

[538] V. Cardoso, E. Franzin, A. Maselli, P. Pani, G. Raposo, Testing strong-field gravity with tidal Love numbers, Phys. Rev. D 95 (8) (2017) 084014. arXiv:1701.01116, doi:10.1103/PhysRevD. 95.084014.

90 [539] N. Sennett, T. Hinderer, J. Steinhoff, A. Buonanno, S. Ossokine, Distinguishing boson stars from black holes and neutron stars from tidal interactions in inspiraling binary systems, Phys. Rev. D 96 (2) (2017) 024002. arXiv:1704.08651, doi:10.1103/PhysRevD.96.024002.

[540] K. Clough, T. Dietrich, J. C. Niemeyer, Axion star collisions with black holes and neutron stars in full 3D numerical relativity, Phys. Rev. D 98 (8) (2018) 083020. arXiv:1808.04668, doi: 10.1103/PhysRevD.98.083020.

[541] T. Dietrich, F. Day, K. Clough, M. Coughlin, J. Niemeyer, Neutron star-axion star collisions in the light of multimessenger astronomy, Mon. Not. R. Astron. Soc. 483 (2019) 908–914. arXiv: 1808.04746, doi:10.1093/mnras/sty3158.

[542] C. B. M. H. Chirenti, L. Rezzolla, How to tell a gravastar from a black hole, Class. Quantum Grav. 24 (2007) 4191–4206. arXiv:0706.1513, doi:10.1088/0264-9381/24/16/013.

[543] R. F. P. Mendes, H. Yang, Tidal deformability of boson stars and dark matter clumps, Classical and Quantum Gravity 34 (18) (2017) 185001. arXiv:1606.03035, doi:10.1088/1361-6382/ aa842d.

[544] P. Pani, I-Love-Q relations for gravastars and the approach to the black-hole limit, Phys. Rev. D 92 (12) (2015) 124030. arXiv:1506.06050, doi:10.1103/PhysRevD.92.124030.

[545] N. Uchikata, S. Yoshida, P. Pani, Tidal deformability and I-Love-Q relations for gravastars with polytropic thin shells, Phys. Rev. D 94 (6) (2016) 064015. arXiv:1607.03593, doi:10.1103/ PhysRevD.94.064015.

[546] R. Friedberg, T. D. Lee, Y. Pang, Mini-soliton stars, Phys. Rev. D 35 (1987) 3640–3657. doi: 10.1103/PhysRevD.35.3640.

[547] C. W. K. Lai, A numerical study of boson stars, Ph.D. thesis, The University of British Columbia (Canada), Canada (Jan 2005).

[548] C. Palenzuela, I. Olabarrieta, L. Lehner, S. L. Liebling, Head-on collisions of boson stars, Phys. Rev. D 75 (6) (2007) 064005. arXiv:arXiv:gr-qc/0612067, doi:10.1103/PhysRevD.75.064005.

[549] C. Palenzuela, L. Lehner, S. L. Liebling, Orbital dynamics of binary boson star systems, Phys. Rev. D 77 (4) (2008) 044036. arXiv:0706.2435, doi:10.1103/PhysRevD.77.044036.

[550] V. Cardoso, S. Hopper, C. F. B. Macedo, C. Palenzuela, P. Pani, Gravitational-wave signatures of exotic compact objects and of quantum corrections at the horizon scale, Physical Review D 94 (8) (2016) 084031. arXiv:1608.08637, doi:10.1103/PhysRevD.94.084031.

[551] T. Helfer, E. A. Lim, M. A. G. Garcia, M. A. Amin, Gravitational wave emission from collisions of compact scalar solitons, Phys. Rev. D 99 (4) (2019) 044046. arXiv:1802.06733, doi:10.1103/ PhysRevD.99.044046.

[552] N. Sanchis-Gual, C. Herdeiro, J. A. Font, E. Radu, F. Di Giovanni, Head-on collisions and orbital mergers of Proca stars, Phys. Rev. D 99 (2) (2019) 024017. arXiv:1806.07779, doi:10.1103/ PhysRevD.99.024017.

[553] M. Fasano, T. Abdelsalhin, A. Maselli, V. Ferrari, Constraining the neutron star equation of state using multi-band independent measurements of radii and tidal deformabilities, arXiv e- printsarXiv:1902.05078.

91 [554] F. Ozel,¨ D. Psaltis, T. G¨uver, G. Baym, C. Heinke, S. Guillot, The Dense Matter Equation of State from Neutron Star Radius and Mass Measurements, Astrophys. J. 820 (2016) 28. arXiv: 1505.05155, doi:10.3847/0004-637X/820/1/28.

[555] A. Vuorinen, Neutron stars and stellar mergers as a laboratory for dense QCD matter, Nuclear Physics A 982 (2019) 36–42. arXiv:1807.04480, doi:10.1016/j.nuclphysa.2018.10.011.

[556] N.-B. Zhang, B.-A. Li, Extracting nuclear symmetry energies at high densities from observations of neutron stars and gravitational waves, European Physical Journal A 55 (2019) 39. arXiv: 1807.07698, doi:10.1140/epja/i2019-12700-0.

[557] X.-Y. Lai, Y.-W. Yu, E.-P. Zhou, Y.-Y. Li, R.-X. Xu, Merging strangeon stars, Research in Astronomy and Astrophysics 18 (2018) 024. arXiv:1710.04964, doi:10.1088/1674-4527/18/ 2/24.

[558] V. Paschalidis, K. Yagi, D. Alvarez-Castillo, D. B. Blaschke, A. Sedrakian, Implications from GW170817 and I-Love-Q relations for relativistic hybrid stars, Phys. Rev. D 97 (8) (2018) 084038. arXiv:1712.00451, doi:10.1103/PhysRevD.97.084038.

[559] R. Nandi, P. Char, Hybrid Stars in the Light of GW170817, Astrophys. J. 857 (2018) 12. arXiv: 1712.08094, doi:10.3847/1538-4357/aab78c.

[560] R. Nandi, P. Char, S. Pal, Constraining the relativistic mean-field model equations of state with gravitational wave observations, Phys. Rev. C 99 (5) (2019) 052802. arXiv:1809.07108, doi: 10.1103/PhysRevC.99.052802.

[561] G. F. Burgio, A. Drago, G. Pagliara, H.-J. Schulze, J.-B. Wei, Are Small Radii of Compact Stars Ruled out by GW170817/AT2017gfo?, Astrophys. J. 860 (2018) 139. doi:10.3847/1538-4357/ aac6ee.

[562] R. O. Gomes, P. Char, S. Schramm, Constraining Strangeness in Dense Matter with GW170817, Astrophys. J. 877 (2) (2019) 139. arXiv:1806.04763, doi:10.3847/1538-4357/ab1751.

[563] D. E. Alvarez-Castillo, D. B. Blaschke, A. G. Grunfeld, V. P. Pagura, Third family of compact stars within a nonlocal chiral quark model equation of state, Phys. Rev. D 99 (6) (2019) 063010. arXiv:1805.04105, doi:10.1103/PhysRevD.99.063010.

[564] C.-M. Li, Y. Yan, J.-J. Geng, Y.-F. Huang, H.-S. Zong, Constraints on the hybrid equation of state with a crossover hadron-quark phase transition in the light of GW170817, Phys. Rev. D 98 (8) (2018) 083013. arXiv:1808.02601, doi:10.1103/PhysRevD.98.083013.

[565] A. Drago, G. Pagliara, Merger of Two Neutron Stars: Predictions from the Two-families Scenario, Astrophys. J. Lett. 852 (2018) L32. arXiv:1710.02003, doi:10.3847/2041-8213/aaa40a.

[566] J.-B. Wei, A. Figura, G. F. Burgio, H. Chen, H.-J. Schulze, Neutron star universal relations with microscopic equations of state, Journal of Physics G Nuclear Physics 46 (3) (2019) 034001. arXiv:1809.04315, doi:10.1088/1361-6471/aaf95c.

[567] J.-E. Christian, A. Zacchi, J. Schaffner-Bielich, Signals in the tidal deformability for phase tran- sitions in compact stars with constraints from GW170817, Phys. Rev. D 99 (2) (2019) 023009. arXiv:1809.03333, doi:10.1103/PhysRevD.99.023009.

92 [568] R. O. Gomes, V. Dexheimer, S. Han, S. Schramm, Can magnetic fields (de)stabilize twin stars?, Mon. Not. R. Astron. Soc. 485 (4) (2019) 4873–4877. arXiv:1810.07046, doi:10.1093/mnras/ stz542.

[569] M. Ruiz, S. L. Shapiro, A. Tsokaros, GW170817, general relativistic magnetohydrodynamic sim- ulations, and the neutron star maximum mass, Phys. Rev. D 97 (2) (2018) 021501. arXiv: 1711.00473, doi:10.1103/PhysRevD.97.021501.

[570] B. Margalit, B. D. Metzger, Constraining the Maximum Mass of Neutron Stars from Multi- messenger Observations of GW170817, Astrophys. J. Lett. 850 (2017) L19. arXiv:1710.05938, doi:10.3847/2041-8213/aa991c.

[571] M. Shibata, E. Zhou, K. Kiuchi, S. Fujibayashi, Constraint on the maximum mass of neutron stars using GW170817 event, Phys. Rev. D 100 (2) (2019) 023015. arXiv:1905.03656, doi: 10.1103/PhysRevD.100.023015.

[572] N.-B. Zhang, B.-A. Li, Implications of the Mass M= 2.17 -0.10 ˆ+0.11 M of PSR J0740+6620 on the Equation of State of Super-dense Neutron-rich{ Nuclear} { Matter,}{ Astrophys.} J. 879 (2) (2019) 99. arXiv:1904.10998, doi:10.3847/1538-4357/ab24cb.

[573] P. G. Krastev, B.-A. Li, Imprints of the nuclear symmetry energy on the tidal deformability of neutron stars, J. Phys. G 46 (7) (2019) 074001. arXiv:1801.04620, doi:10.1088/1361-6471/ ab1a7a.

[574] R. Gamba, J. S. Read, L. E. Wade, The impact of the crust equation of state on the analysis of GW170817, arXiv e-printsarXiv:1902.04616.

[575] B. Biswas, R. Nandi, P. Char, S. Bose, Role of crustal physics in the tidal deformation of a neutron star, arXiv e-prints (2019) arXiv:1905.00678arXiv:1905.00678.

[576] A. M. Kalaitzis, T. F. Motta, A. W. Thomas, Roles of crust and core in the tidal deformability of neutron stars, arXiv e-prints (2019) arXiv:1905.05907arXiv:1905.05907.

[577] F. J. Fattoyev, J. Carvajal, W. G. Newton, B.-A. Li, Constraining the high-density behavior of the nuclear symmetry energy with the tidal polarizability of neutron stars, Phys. Rev. C 87 (1) (2013) 015806. arXiv:1210.3402, doi:10.1103/PhysRevC.87.015806.

[578] M. Cantiello, J. B. Jensen, J. P. Blakeslee, E. Berger, A. J. Levan, N. R. Tanvir, G. Raimondo, E. Brocato, K. D. Alexander, P. K. Blanchard, M. Branchesi, Z. Cano, R. Chornock, S. Covino, P. S. Cowperthwaite, P. D’Avanzo, T. Eftekhari, W. Fong, A. S. Fruchter, A. Grado, J. Hjorth, D. E. Holz, J. D. Lyman, I. Mandel, R. Margutti, M. Nicholl, V. A. Villar, P. K. G. Williams, A Precise Distance to the Host of the Binary Neutron Star Merger GW170817 Using Surface Brightness Fluctuations, Astrophys. J. Lett. 854 (2018) L31. arXiv:1801.06080, doi: 10.3847/2041-8213/aaad64.

[579] H. M¨uller,B. D. Serot, Relativistic mean-field theory and the high-density nuclear equation of state, Nuclear Physics A 606 (1996) 508–537. arXiv:arXiv:nucl-th/9603037, doi:10.1016/ 0375-9474(96)00187-X.

[580] J. J. Li, A. Sedrakian, Implications from GW170817 for ∆-isobar Admixed Hypernuclear Compact Stars, Astrophys. J. 874 (2) (2019) L22. arXiv:1904.02006, doi:10.3847/2041-8213/ab1090.

93 [581] T.-T. , S.-S. Zhang, Q.-L. Zhang, C.-J. Xia, Strangeness and ∆ resonance in compact stars with relativistic-mean-field models, Phys. Rev. D 99 (2) (2019) 023004. arXiv:1808.02207, doi:10.1103/PhysRevD.99.023004.

[582] F. Weber, Strange quark matter and compact stars, Progress in Particle and Nuclear Physics 54 (1) (2005) 193–288. arXiv:astro-ph/0407155, doi:10.1016/j.ppnp.2004.07.001.

[583] S. Weissenborn, D. Chatterjee, J. Schaffner-Bielich, Hyperons and massive neutron stars: Vector repulsion and SU(3) symmetry, Physical Review C 85 (6) (2012) 065802. arXiv:1112.0234, doi:10.1103/PhysRevC.85.065802.

[584] S. Weissenborn, D. Chatterjee, J. Schaffner-Bielich, Erratum: Hyperons and massive neutron stars: Vector repulsion and SU(3) symmetry [Phys. Rev. C 85, 065802 (2012)], Physical Review C 90 (1) (2014) 019904. doi:10.1103/PhysRevC.90.019904.

[585] E. N. E. van Dalen, G. Colucci, A. Sedrakian, Constraining hypernuclear density functional with Λ-hypernuclei and compact stars, Physics Letters B 734 (2014) 383–387. arXiv:1406.0744, doi:10.1016/j.physletb.2014.06.002.

[586] M. Oertel, C. Providˆencia,F. Gulminelli, A. R. Raduta, Hyperons in neutron star matter within relativistic mean-field models, Journal of Physics G Nuclear Physics 42 (7) (2015) 075202. arXiv: 1412.4545, doi:10.1088/0954-3899/42/7/075202.

[587] L. Tolos, M. Centelles, A. Ramos, Equationn of State for Nucleonic and Hyperonic Neutron Stars with Mass and Radius Constraints, Astrophys. J. 834 (2017) 3. arXiv:1610.00919, doi: 10.3847/1538-4357/834/1/3.

[588] M. Fortin, S. S. Avancini, C. Providˆencia,I. Vida˜na,Hypernuclei and massive neutron stars, Physical Review C 95 (6) (2017) 065803. doi:10.1103/PhysRevC.95.065803.

[589] J. J. Li, W. H. Long, A. Sedrakian, Hypernuclear stars from relativistic Hartree-Fock density functional theory, European Physical Journal A 54 (8) (2018) 133. arXiv:1801.07084, doi: 10.1140/epja/i2018-12566-6.

[590] H. S. Sahoo, R. Mishra, D. K. Mohanty, P. K. Panda, N. Barik, Neutron star matter with strange interactions within constraints by GW170817 in a relativistic quark model, Phys. Rev. C 99 (5) (2019) 055803. doi:10.1103/PhysRevC.99.055803.

[591] J. J. Li, A. Sedrakian, F. Weber, Competition between delta isobars and hyperons and properties of compact stars, Physics Letters B 783 (2018) 234–240. arXiv:1803.03661, doi:10.1016/j. physletb.2018.06.051.

[592] A. Drago, A. Lavagno, G. Pagliara, D. Pigato, Early appearance of ∆ isobars in neutron stars, Physical Review C 90 (6) (2014) 065809. doi:10.1103/PhysRevC.90.065809.

[593] Z.-Y. Zhu, A. Li, J.-N. Hu, H. Sagawa, ∆ (1232 ) effects in density-dependent relativistic Hartree- Fock theory and neutron stars, Phys. Rev. C 94 (4) (2016) 045803. arXiv:1607.04007, doi: 10.1103/PhysRevC.94.045803.

[594] H. Sotani, K. Iida, K. Oyamatsu, A. Ohnishi, Mass and radius formulas for low-mass neutron stars, Progress of Theoretical and Experimental Physics 2014 (5) (2014) 051E01. arXiv:1401.0161, doi:10.1093/ptep/ptu052.

94 [595] H. O. Silva, H. Sotani, E. Berti, Low-mass neutron stars: universal relations, the nuclear symmetry energy and gravitational radiation, Mon. Not. R. Astron. Soc. 459 (4) (2016) 4378–4388. arXiv: 1601.03407, doi:10.1093/mnras/stw969.

[596] X. Y. Lai, R. X. Xu, Lennard-Jones quark matter and massive quark stars, Mon. Not. R. Astron. Soc. 398 (2009) L31–L35. arXiv:0905.2839, doi:10.1111/j.1745-3933.2009.00701.x.

[597] R. X. Xu, Solid Quark Stars?, Astrophys. J. Lett. 596 (2003) L59–L62. arXiv:astro-ph/0302165, doi:10.1086/379209.

[598] H. Wang, F.-W. Zhang, Y.-Z. Wang, Z.-Q. Shen, Y.-F. Liang, X. Li, N.-H. Liao, Z.-P. Jin, Q. Yuan, Y.-C. Zou, Y.-Z. Fan, D.-M. Wei, The GW170817/GRB 170817A/AT 2017gfo Association: Some Implications for Physics and Astrophysics, Astrophys. J. Lett. 851 (2017) L18. arXiv:1710.05805, doi:10.3847/2041-8213/aa9e08.

[599] X. Lai, E. Zhou, R. Xu, Strangeons constitute bulk strong matter: Test using GW 170817, European Physical Journal A 55 (4) (2019) 60. arXiv:1811.00193, doi:10.1140/epja/ i2019-12720-8.

[600] A. Drago, A. Lavagno, G. Pagliara, Can very compact and very massive neutron stars both exist?, Phys. Rev. D 89 (4) (2014) 043014. arXiv:1309.7263, doi:10.1103/PhysRevD.89.043014.

[601] A. Drago, G. Pagliara, S. Popov, S. Traversi, G. Wiktorowicz, The Merger of Two Compact Stars: A Tool for Dense Matter Nuclear Physics, Universe 4 (2018) 50. arXiv:1802.02495, doi:10.3390/universe4030050.

[602] R. De Pietri, A. Drago, A. Feo, G. Pagliara, M. Pasquali, S. Traversi, G. Wiktorowicz, Merger of compact stars in the two-families scenario, arXiv e-prints (2019) arXiv:1904.01545arXiv:1904. 01545.

[603] C. Cs´aki,C. Er¨oncel,J. Hubisz, G. Rigo, J. Terning, Neutron star mergers chirp about vac- uum energy, Journal of High Energy Physics 9 (2018) 87. arXiv:1802.04813, doi:10.1007/ JHEP09(2018)087.

[604] E. Annala, C. Ecker, C. Hoyos, N. Jokela, D. R. Fern´andez, A. Vuorinen, Holographic compact stars meet gravitational wave constraints, Journal of High Energy Physics 2018 (12) (2018) 78. arXiv:1711.06244, doi:10.1007/JHEP12(2018)078.

[605] D. Sen, T. K. Jha, Effects of hadron-quark phase transition on properties of neutron stars, Journal of Physics G Nuclear Physics 46 (1) (2019) 015202. arXiv:1811.07434, doi:10.1088/ 1361-6471/aaf0b0.

[606] G. Bozzola, P. L. Espino, C. Davis Lewin, V. Paschalidis, Maximum mass and universal relations of rotating relativistic hybrid hadron-quark stars, arXiv e-prints (2019) arXiv:1905.00028arXiv: 1905.00028.

[607] M. G. Alford, A. Schmitt, K. Rajagopal, T. Sch¨afer, in dense quark matter, Reviews of Modern Physics 80 (2008) 1455–1515. arXiv:0709.4635, doi:10.1103/ RevModPhys.80.1455.

[608] M. Alford, A. Sedrakian, Compact Stars with Sequential QCD Phase Transitions, Physical Review Letters 119 (16) (2017) 161104. arXiv:1706.01592, doi:10.1103/PhysRevLett.119.161104.

95 [609] S. Postnikov, M. Prakash, J. M. Lattimer, Tidal Love numbers of neutron and self-bound quark stars, Phys. Rev. D 82 (2) (2010) 024016. arXiv:1004.5098, doi:10.1103/PhysRevD.82.024016.

[610] A. Ayriyan, D. Alvarez-Castillo, D. Blaschke, H. Grigorian, Bayesian Analysis for Extract- ing Properties of the Nuclear Equation of State from Observational Data Including Tidal De- formability from GW170817, Universe 5 (2) (2019) 61. arXiv:1812.10796, doi:10.3390/ universe5020061.

[611] P. Char, S. Datta, Relativistic tidal properties of superfluid neutron stars, Phys. Rev. D 98 (8) (2018) 084010. arXiv:1806.10986, doi:10.1103/PhysRevD.98.084010.

[612] V. Dexheimer, R. de Oliveira Gomes, S. Schramm, H. Pais, What do we learn about vector interactions from GW170817?, Journal of Physics G Nuclear Physics 46 (3) (2019) 034002. arXiv: 1810.06109, doi:10.1088/1361-6471/ab01f0.

[613] J. Ellis, G. H¨utsi,K. Kannike, L. Marzola, M. Raidal, V. Vaskonen, Dark matter effects on neutron star properties, Phys. Rev. D 97 (12) (2018) 123007. arXiv:1804.01418, doi:10.1103/ PhysRevD.97.123007.

[614] A. E. Nelson, S. Reddy, D. Zhou, Dark halos around neutron stars and gravitational waves, Journal of Cosmology and Astroparticle Physics 2019 (7) (2019) 012. arXiv:1803.03266, doi: 10.1088/1475-7516/2019/07/012.

[615] A. Das, T. Malik, A. C. Nayak, Confronting nuclear equation of state in the presence of dark matter using GW170817 observation in relativistic mean field theory approach, Phys. Rev. D 99 (4) (2019) 043016. arXiv:1807.10013, doi:10.1103/PhysRevD.99.043016.

[616] A. Quddus, G. Panotopoulos, B. Kumar, S. Ahmad, S. K. Patra, GW170817 constraints on the properties of a neutron star in the presence of WIMP dark matter, arXiv e-printsarXiv: 1902.00929.

[617] R. L. Bowers, E. P. T. Liang, Anisotropic Spheres in General Relativity, Astrophys. J. 188 (1974) 657. doi:10.1086/152760.

[618] D. D. Doneva, S. S. Yazadjiev, Nonradial oscillations of anisotropic neutron stars in the Cowling approximation, Phys. Rev. D 85 (12) (2012) 124023. arXiv:1203.3963, doi:10.1103/PhysRevD. 85.124023.

[619] H. O. Silva, C. F. B. Macedo, E. Berti, L. C. B. Crispino, Slowly rotating anisotropic neutron stars in general relativity and scalar-tensor theory, Classical and Quantum Gravity 32 (14) (2015) 145008. arXiv:1411.6286, doi:10.1088/0264-9381/32/14/145008.

[620] K. Yagi, N. Yunes, I-Love-Q anisotropically: Universal relations for compact stars with scalar pressure anisotropy, Phys. Rev. D 91 (12) (2015) 123008. arXiv:1503.02726, doi:10.1103/ PhysRevD.91.123008.

[621] G. Raposo, P. Pani, M. Bezares, C. Palenzuela, V. Cardoso, Anisotropic stars as ultracompact objects in general relativity, Phys. Rev. D 99 (10) (2019) 104072. arXiv:1811.07917, doi: 10.1103/PhysRevD.99.104072.

[622] L. Herrera, N. O. Santos, Jeans Mass for Anisotropic Matter, Astrophys. J. 438 (1995) 308. doi:10.1086/175075.

96 [623] R. Kippenhahn, A. Weigert, A. Weiss, Stellar Structure and Evolution, Springer, Berlin, Heidel- berg, 2012. doi:10.1007/978-3-642-30304-3.

[624] A. Sulaksono, Anisotropic pressure and hyperons in neutron stars, International Journal of Modern Physics E 24 (2015) 1550007. arXiv:1412.7247, doi:10.1142/S021830131550007X.

[625] K. Chakravarti, S. Chakraborty, S. Bose, S. SenGupta, Tidal Love numbers of black holes and neutron stars in the presence of higher dimensions: Implications of GW170817, Phys. Rev. D 99 (2) (2019) 024036. arXiv:1811.11364, doi:10.1103/PhysRevD.99.024036.

[626] K. Chakravarti, S. Chakraborty, K. S. Phukon, S. Bose, S. SenGupta, Constraining extra- spatial dimensions with multi-messenger observations of GW170817, arXiv e-prints (2019) arXiv:1903.10159arXiv:1903.10159.

[627] K. Schertler, C. Greiner, J. Schaffner-Bielich, M. H. Thoma2, Quark phases in neutron stars and a third family of compact stars as signature for phase transitions1, Nuclear Physics A 677 (2000) 463–490. arXiv:astro-ph/0001467, doi:10.1016/S0375-9474(00)00305-5.

[628] D. E. Alvarez-Castillo, D. B. Blaschke, High-mass twin stars with a multipolytrope equation of state, Physical Review C 96 (4) (2017) 045809. arXiv:1703.02681, doi:10.1103/PhysRevC.96. 045809.

[629] M. A. R. Kaltenborn, N.-U. F. Bastian, D. B. Blaschke, Quark-nuclear hybrid star equation of state with excluded volume effects, Phys. Rev. D 96 (5) (2017) 056024. arXiv:1701.04400, doi:10.1103/PhysRevD.96.056024.

[630] A. Ayriyan, N.-U. Bastian, D. Blaschke, H. Grigorian, K. Maslov, D. N. Voskresensky, Robustness of third family solutions for hybrid stars against mixed phase effects, Physical Review C 97 (4) (2018) 045802. arXiv:1711.03926, doi:10.1103/PhysRevC.97.045802.

[631] A. Drago, G. Pagliara, Combustion of a hadronic star into a quark star: The turbulent and the diffusive regimes, Phys. Rev. C 92 (4) (2015) 045801. arXiv:1506.08337, doi:10.1103/ PhysRevC.92.045801.

[632] P.-X. Ma, J.-L. Jiang, H. Wang, Z.-P. Jin, Y.-Z. Fan, D.-M. Wei, GW170817 and the Prospect of Forming Supramassive Remnants in Neutron Star Mergers, Astrophys. J. 858 (2018) 74. arXiv: 1711.05565, doi:10.3847/1538-4357/aabafe.

[633] D. Pooley, P. Kumar, J. C. Wheeler, B. Grossan, GW170817 Most Likely Made a Black Hole, Astrophys. J. Letters 859 (2018) L23. arXiv:1712.03240, doi:10.3847/2041-8213/aac3d6.

[634] S. Ai, H. Gao, Z.-G. Dai, X.-F. Wu, A. Li, B. Zhang, M.-Z. Li, The Allowed Parameter Space of a Long-lived Neutron Star as the Merger Remnant of GW170817, Astrophys. J. 860 (2018) 57. arXiv:1802.00571, doi:10.3847/1538-4357/aac2b7.

[635] P. Haensel, B. Pichon, Experimental nuclear masses and the ground state of cold dense matter, Astron. Astrophys. 283 (1) (1994) 313–318. arXiv:nucl-th/9310003.

[636] C.-J. Xia, T. Maruyama, N. Yasutake, T. Tatsumi, Constraining quark-hadron interface tension in the multimessenger era, Phys. Rev. D 99 (10) (2019) 103017. arXiv:1902.08766, doi:10. 1103/PhysRevD.99.103017.

97 [637] A. Akmal, V. R. Pandharipande, D. G. Ravenhall, Equation of state of nucleon matter and neutron star structure, Phys. Rev. C 58 (3) (1998) 1804–1828. arXiv:nucl-th/9804027, doi: 10.1103/PhysRevC.58.1804.

[638] S. Johnston, D. R. Lorimer, P. A. Harrison, M. Bailes, A. G. Lynet, J. F. Bell, V. M. Kaspi, R. N. Manchester, N. D’Amico, L. Nleastrol, J. Shengzhen, Discovery of a very bright, nearby binary , Nature 361 (1993) 613–615. doi:10.1038/361613a0.

[639] L. Tolos, M. Centelles, A. Ramos, The Equation of State for the Nucleonic and Hyperonic Core of Neutron Stars, Publications of the Astronomical Society of Australia 34 (2017) e065. arXiv: 1708.08681, doi:10.1017/pasa.2017.60.

[640] B. K. Sharma, M. Centelles, X. Vi˜nas,M. Baldo, G. F. Burgio, Unified equation of state for neutron stars on a microscopic basis, Astron. Astrophys. 584 (2015) A103. arXiv:1506.00375, doi:10.1051/0004-6361/201526642.

[641] D. Blaschke, F. Sandin, T. Kl¨ahn,J. Berdermann, Sequential deconfinement of quark flavors in neutron stars, Physical Review C 80 (6) (2009) 065807. arXiv:0807.0414, doi:10.1103/ PhysRevC.80.065807.

[642] K. Maslov, N. Yasutake, A. Ayriyan, D. Blaschke, H. Grigorian, T. Maruyama, T. Tatsumi, D. N. Voskresensky, Hybrid equation of state with pasta phases and third family of compact stars I: Pasta phases and effective mixed phase model, arXiv e-prints (2018) arXiv:1812.11889arXiv: 1812.11889.

[643] D. N. Voskresensky, M. Yasuhira, T. Tatsumi, Charge screening in hadronquark mixed phase, Phys. Lett. B 541 (2002) 93–100. doi:10.1016/S0370-2693(02)02186-X.

[644] F. Pacini, High-energy Astrophysics and a Possible Sub-nuclear Energy Source, Nature 209 (1966) 389–390. doi:10.1038/209389a0.

[645] J. R. Ellis, J. I. Kapusta, K. A. Olive, Phase transition in dense nuclear matter with quark and gluon condensates, Phys. Lett. B273 (1991) 123–127. doi:10.1016/0370-2693(91)90564-7.

[646] C. E. Rasmussen, C. K. I. Williams, Gaussian Processes for Machine Learning, MIT Press, 2006.

[647] G. Panotopoulos, I. Lopes, Radial oscillations of strange quark stars admixed with condensed dark matter, Phys. Rev. D 96 (8) (2017) 083013. arXiv:1709.06643, doi:10.1103/PhysRevD. 96.083013.

[648] F. Sandin, P. Ciarcelluti, Effects of mirror dark matter on neutron stars, Astroparticle Physics 32 (2009) 278–284. arXiv:0809.2942, doi:10.1016/j.astropartphys.2009.09.005.

[649] P. Ciarcelluti, F. Sandin, Have neutron stars a dark matter core?, Physics Letters B 695 (2011) 19–21. arXiv:1005.0857, doi:10.1016/j.physletb.2010.11.021.

[650] L. Tolos, J. Schaffner-Bielich, Dark compact , Phys. Rev. D 92 (12) (2015) 123002. arXiv: 1507.08197, doi:10.1103/PhysRevD.92.123002.

[651] S.-C. Leung, M.-C. Chu, L.-M. Lin, Dark-matter admixed neutron stars, Phys. Rev. D 84 (10) (2011) 107301. arXiv:1111.1787, doi:10.1103/PhysRevD.84.107301.

[652] K. Petraki, R. R. Volkas, Review of Asymmetric Dark Matter, International Journal of Modern Physics A 28 (19) (2013) 1330028. arXiv:1305.4939, doi:10.1142/S0217751X13300287.

98 [653] J. D. Walecka, A theory of highly condensed matter., Annals of Physics 83 (1974) 491–529. doi:10.1016/0003-4916(74)90208-5. [654] B. D. Serot, J. D. Walecka, Recent Progress in Quantum Hadrodynamics, International Journal of Modern Physics E 6 (1997) 515–631. arXiv:nucl-th/9701058, doi:10.1142/ S0218301397000299. [655] J. N¨attil¨a,A. W. Steiner, J. J. E. Kajava, V. F. Suleimanov, J. Poutanen, Equation of state constraints for the cold dense matter inside neutron stars using the cooling tail method, Astron. Astrophys. 591 (2016) A25. arXiv:1509.06561, doi:10.1051/0004-6361/201527416. [656] V. F. Suleimanov, J. J. E. Kajava, S. V. Molkov, J. N¨attil¨a,A. A. Lutovinov, K. Werner, J. Pouta- nen, Basic parameters of the -accreting X-ray bursting neutron star in 4U 1820-30, Mon. Not. R. Astron. Soc. 472 (2017) 3905–3913. arXiv:1708.09168, doi:10.1093/mnras/stx2234. [657] J. N¨attil¨a,M. C. Miller, A. W. Steiner, J. J. E. Kajava, V. F. Suleimanov, J. Poutanen, Neutron star mass and radius measurements from atmospheric model fits to X-ray burst cooling tail spectra, Astron. Astrophys. 608 (2017) A31. arXiv:1709.09120, doi:10.1051/0004-6361/201731082. [658] A. W. Steiner, C. O. Heinke, S. Bogdanov, C. K. Li, W. C. G. Ho, A. Bahramian, S. Han, Constraining the mass and radius of neutron stars in globular clusters, Mon. Not. R. Astron. Soc. 476 (2018) 421–435. arXiv:1709.05013, doi:10.1093/mnras/sty215. [659] J. J. E. Kajava, J. N¨attil¨a,O.-M. Latvala, M. Pursiainen, J. Poutanen, V. F. Suleimanov, M. G. Revnivtsev, E. Kuulkers, D. K. Galloway, The influence of accretion geometry on the spectral evolution during thermonuclear (type I) X-ray bursts, Mon. Not. R. Astron. Soc. 445 (2014) 4218–4234. arXiv:1406.0322, doi:10.1093/mnras/stu2073. [660] Z. Arzoumanian, K. C. Gendreau, C. L. Baker, T. Cazeau, P. Hestnes, J. W. Kellogg, S. J. Kenyon, R. P. Kozon, K.-C. Liu, S. S. Manthripragada, C. B. Markwardt, A. L. Mitchell, J. W. Mitchell, C. A. Monroe, T. Okajima, S. E. Pollard, D. F. Powers, B. J. Savadkin, L. B. Winternitz, P. T. Chen, M. R. Wright, R. Foster, G. Prigozhin, R. Remillard, J. Doty, The neutron star interior composition explorer (NICER): mission definition, in: Space Telescopes and Instrumentation 2014: Ultraviolet to Gamma Ray, Vol. 9144 of Proceedings of the SPIE, 2014, p. 914420. doi: 10.1117/12.2056811. [661] C. A. Wilson-Hodge, P. S. Ray, D. Chakrabarty, M. Feroci, L. Alvarez, M. Baysinger, C. Becker, E. Bozzo, S. Brandt, B. Carson, J. Chapman, A. Dominguez, L. Fabisinski, B. Gangl, J. Garcia, C. Griffith, M. Hernanz, R. Hickman, R. Hopkins, M. Hui, L. Ingram, P. Jenke, S. Korpela, T. Maccarone, M. Michalska, M. Pohl, A. Santangelo, S. Schanne, A. Schnell, L. Stella, M. van der Klis, A. Watts, B. Winter, S. Zane, Large Observatory for x-ray Timing (LOFT-P): a Probe-class mission concept study, in: Space Telescopes and Instrumentation 2016: Ultraviolet to Gamma Ray, Vol. 9905 of Proceedings of the SPIE, 2016, p. 99054Y. arXiv:1608.06258, doi:10.1117/ 12.2232944. [662] C. Motch, J. Wilms, D. Barret, W. Becker, S. Bogdanov, L. Boirin, S. Corbel, E. Cackett, S. Campana, D. de Martino, F. Haberl, J. in’t Zand, M. M´endez,R. Mignani, J. Miller, M. Orio, D. Psaltis, N. Rea, J. Rodriguez, A. Rozanska, A. Schwope, A. Steiner, N. Webb, L. Zampieri, S. Zane, The Hot and Energetic Universe: End points of , arXiv e-printsarXiv: 1306.2334. [663] J. N¨attil¨a,P. Pihajoki, Radiation from rapidly rotating oblate neutron stars, Astron. Astrophys. 615 (2018) A50. arXiv:1709.07292, doi:10.1051/0004-6361/201630261.

99 [664] T. Salmi, J. N¨attil¨a,J. Poutanen, Bayesian parameter constraints for neutron star masses and radii using X-ray timing observations of accretion-powered millisecond pulsars, Astron. Astrophys. 618 (2018) A161. arXiv:1805.01149, doi:10.1051/0004-6361/201833348.

[665] M. Kramer, N. Wex, TOPICAL REVIEW: The double pulsar system: a unique laboratory for gravity, Class. Quantum Grav. 26 (7) (2009) 073001. doi:10.1088/0264-9381/26/7/073001.

[666] C. A. Raithel, Constraints on the neutron star equation of state from GW170817, European Physical Journal A 55 (5) (2019) 80. arXiv:1904.10002, doi:10.1140/epja/i2019-12759-5.

[667] S. Gandolfi, J. Carlson, S. Reddy, Maximum mass and radius of neutron stars, and the nuclear symmetry energy, Physical Review C 85 (3) (2012) 032801. arXiv:1101.1921, doi:10.1103/ PhysRevC.85.032801.

[668] J. E. Lynn, I. Tews, J. Carlson, S. Gandolfi, A. Gezerlis, K. E. Schmidt, A. Schwenk, Chiral Three- Nucleon Interactions in Light Nuclei, Neutron-α Scattering, and Neutron Matter, Physical Review Letters 116 (6) (2016) 062501. arXiv:1509.03470, doi:10.1103/PhysRevLett.116.062501.

[669] C. Drischler, K. Hebeler, A. Schwenk, Asymmetric nuclear matter based on chiral two- and three-nucleon interactions, Physical Review C 93 (5) (2016) 054314. arXiv:1510.06728, doi: 10.1103/PhysRevC.93.054314.

[670] I. Tews, J. M. Lattimer, A. Ohnishi, E. E. Kolomeitsev, Symmetry Parameter Constraints from a Lower Bound on Neutron-matter Energy, Astrophys. J. 848 (2017) 105. arXiv:1611.07133, doi:10.3847/1538-4357/aa8db9.

[671] The LIGO Scientific Collaboration, The Virgo Collaboration, B. P. Abbott, R. Abbott, T. D. Abbott, M. R. Abernathy, F. Acernese, K. Ackley, C. Adams, T. Adams, P. Addesso, R. X. Adhikari, et al., Upper Limits on the Rates of Binary Neutron Star and Neutron Star-Black Hole Mergers from Advanced LIGOs First Observing Run, Astrophys. J. Lett. 832 (2) (2016) L21. arXiv:1607.07456, doi:10.3847/2041-8205/832/2/L21.

100