Articles

THE INFLUENCE OF THERMAL PRESSURE ON EQUILIBRIUM MODELS OF HYPERMASSIVE NEUTRON STAR MERGER REMNANTS

, , , , , and

Published 2014 June 27 © 2014. The American Astronomical Society. All rights reserved.
, , Citation J. D. Kaplan et al 2014 ApJ 790 19 DOI 10.1088/0004-637X/790/1/19

0004-637X/790/1/19

ABSTRACT

The merger of two neutron stars leaves behind a rapidly spinning hypermassive object whose survival is believed to depend on the maximum mass supported by the nuclear equation of state (EOS), angular momentum redistribution by (magneto-)rotational instabilities, and spindown by gravitational waves. The high temperatures (∼5–40 MeV) prevailing in the merger remnant may provide thermal pressure support that could increase its maximum mass and, thus, its life on a neutrino-cooling timescale. We investigate the role of thermal pressure support in hypermassive merger remnants by computing sequences of spherically symmetric and axisymmetric uniformly and differentially rotating equilibrium solutions to the general-relativistic stellar structure equations. Using a set of finite-temperature nuclear EOS, we find that hot maximum-mass critically spinning configurations generally do not support larger baryonic masses than their cold counterparts. However, subcritically spinning configurations with mean density of less than a few times nuclear saturation density yield a significantly thermally enhanced mass. Even without decreasing the maximum mass, cooling and other forms of energy loss can drive the remnant to an unstable state. We infer secular instability by identifying approximate energy turning points in equilibrium sequences of constant baryonic mass parameterized by maximum density. Energy loss carries the remnant along the direction of decreasing gravitational mass and higher density until instability triggers collapse. Since configurations with more thermal pressure support are less compact and thus begin their evolution at a lower maximum density, they remain stable for longer periods after merger.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Coalescing double neutron stars (NSs) are prime candidate progenitors of short-hard gamma-ray bursts (GRBs; e.g., Nakar 2007 and references therein). The strong gravitational wave emission driving the coalescence makes NSNS systems the primary targets of the network of second-generation gravitational-wave interferometers currently under construction (Advanced LIGO: Harry et al. 2010; Advanced Virgo: Accadia et al. 2011; and KAGRA: Somiya et al. 2012).

Until the last moments of inspiral, the constituent NSs may essentially be treated as cold NSs. Tidal heating is mild and the NS crust may not fail until the NSs touch (Penner et al. 2012, but see Tsang et al. 2012; Weinberg et al. 2013). Merger results in the formation of a shocked, extremely rapidly differentially spinning central object, commonly referred to as a hypermassive NS (HMNS), since it comprises the vast majority of the baryonic mass of the two premerger NSs and is thus expected to be more massive than the maximum mass supported by the nuclear equation of state (EOS) in the spherical and uniformly rotating limits limit (see, e.g., Faber & Rasio 2012 for a review of NSNS mergers). The subsequent evolution of the HMNS has important ramifications for gravitational wave emission and the possible transition to a short-hard GRB. If the HMNS survives for an extended period, nonaxisymmetric rotational instability may enhance the high-frequency gravitational-wave emission, possibly allowing gravitational-wave observers to constrain the nuclear EOS (e.g., Bauswein et al. 2012). On the other hand, the neutrino-driven wind blown off a surviving HMNS, producing mass loss at a rate of order 10−4M s−1, will lead to strong baryon loading in polar regions (Dessart et al. 2009), making the formation of the relativistic outflows needed for a GRB more difficult, even if a black hole with an accretion disk forms eventually. If the HMNS collapses to a black hole within milliseconds of merger, baryon loading will not hamper a GRB, but strong gravitational-wave and neutrino emission would be shut off rapidly.

The long-term survival of the HMNS depends sensitively on the maximum mass of a nonrotating cold NS supported by the nuclear EOS, which most certainly is above ∼2 M (Demorest et al. 2010; Antoniadis et al. 2013) and very likely below ∼3.2 M (Lattimer & Prakash 2007). At its formation, the HMNS is rapidly and strongly differentially rotating. Extreme differential rotation alone may increase the maximum HMNS mass by more than 100% (e.g., Baumgarte et al. 2000). Angular momentum redistribution by (magneto-)rotational instabilities and spindown by gravitational wave emission are expected to remove this additional support. This will ultimately lead to black hole formation if the HMNS mass is above the maximum mass that can be supported by the nuclear EOS and uniform rotation (≲20% greater than the maximum in the nonrotating limit; Baumgarte et al. 2000).

Recently, Sekiguchi et al. (2011), Paschalidis et al. (2012), Bauswein et al. (2010), and, in earlier work, Baiotti et al. (2008), have argued that thermal pressure support at moderately high temperatures of ∼5–40 MeV (Oechslin et al. 2007; Sekiguchi et al. 2011) may significantly influence the structure and evolution of the postmerger HMNS and prolong its lifetime until collapse to a black hole. If true, the HMNS may survive on the neutrino cooling timescale provided that the combined premerger mass of the NSs is sufficiently close to the thermally enhanced maximum HMNS mass. These authors estimate the neutrino cooling timescale to be comparable to or longer than the timescale for angular momentum redistribution and spindown by gravitational waves.

The focus of this paper is on the role of thermal pressure support in HMNS merger remnants. Postmerger HMNS configurations that survive for multiple dynamical times quickly assume dynamical equilibrium and, after the extremely dynamic merger phase, show only mild deviation from axisymmetry (e.g., Sekiguchi et al. 2011; Shibata et al. 2005). Hence, instead of performing computationally expensive full merger simulations, we investigate the role of thermal effects by approximating HMNS configurations as sequences of rotational equilibrium solutions, which we compute with the relativistic self-consistent field method (Komatsu et al. 1989a, 1989b; Cook et al. 1992). We consider the spherical limit (Tolman–Oppenheimer–Volkoff (TOV) solutions), uniform, and differential rotation. We employ multiple finite-temperature microphysical nuclear EOS and, since the equilibrium solver requires a barotropic EOS, a range of temperature and composition parameterizations that are motivated by the merger simulations of Sekiguchi et al. (2011). An overall similar approach, though only considering isothermal and isentropic configurations, has been used in the past to study thermal effects on uniformly and differentially rotating proto-NSs (Goussard et al. 1997, 1998).

The key quantity relevant in the secular evolution of HMNSs is the baryonic mass (Mb; also called "rest mass") that can be supported by a given combination of EOS, thermal/compositional structure, and rotational setup. The gravitational mass (Mg) is not conserved and is reduced by cooling and angular momentum loss. Our results show that the maximum baryonic mass of TOV, uniformly rotating, and differentially rotating configurations is essentially unaffected by thermal pressure support. Thermal pressure support is negligible at supranuclear densities and becomes significant only at densities below nuclear saturation density. Since maximum-mass configurations always have maximum and mean densities above nuclear, thermal pressure support is minimal. The thermal contribution to the stress-energy tensor (which sources curvature) may, depending on the EOS, even lead to a net decrease of the $M_\mathrm{b}^\mathrm{max}$ with increasing temperature.

We find thermal enhancement of Mb for configurations with mean densities less than a few times nuclear saturation density that are nonrotating or rotating subcritically (i.e., below the mass-shedding limit). A hot configuration in this regime will support the same baryonic mass at a lower mean (and maximum) density. However, hot rotating configurations are spatially more extended than their cold counterparts, and thus reach mass shedding at lower angular velocities. This counteracts the thermal enhancement and results in $M_\mathrm{b}^\mathrm{max}$ that are within a few percent of cold configurations.

The secular evolution of an HMNS toward collapse is driven by energy losses to gravitational waves and neutrinos, and, potentially, by loss of angular momentum transported to the surface by processes such as the magnetorotational instability (MRI). It proceeds along trajectories of constant (or nearly constant) baryonic mass and in the direction of decreasing total energy (i.e., gravitational mass Mg) and increasing maximum baryon density ρb, max (i.e., more compact configurations). We conjecture, based on established results of the theory of rotating relativistic stars (Friedman & Stergioulas 2013), that instability to collapse occurs when the configuration reaches an unstable part of the parameter space and not necessarily because the maximum supportable baryonic mass $M_\mathrm{b}^\mathrm{max}$ drops below Mb. We formalize this via an approximate variant of the turning-point theorem (e.g., Sorkin 1982; Friedman & Stergioulas 2013): The turning-point theorem states that for uniformly rotating NSs, a local extremum in Mg at fixed angular momentum, entropy, and baryonic mass constitutes a point at which secular instability to collapse must set in. We argue that the turning point theorem carries over to differentially rotating hot HMNSs. The precise turning points become approximate and are distributed over a narrow range of ρb, max and Mg for all degrees of differential rotation and temperature prescriptions that we consider here. The regime of instability is thus largely independent of HMNS temperature. However, a hotter configuration will be less compact initially and, hence, will begin its secular evolution to its turning point at a lower ρb, max than a colder one. It will thus have to evolve further until it reaches its turning point and, at a fixed rate of energy loss, will survive for longer.

This paper is structured as follows. In Section 2, we introduce the set of EOS we employ and discuss the relative importance of thermal pressure as a function of density. We also introduce the temperature and composition parameterizations and the methods used for constructing equilibrium models without and with rotation. In Section 3, we lay out our results for nonrotating NSs and then discuss uniformly and differentially rotating configurations in Sections 4.1 and 4.2, respectively. We consider evolutionary sequences of HMNSs at constant baryonic mass in the context of an approximate turning point theorem and compare with results from recent merger simulations in Section 5. Finally, in Section 6, we summarize our results and conclude.

2. METHODS AND EQUATIONS OF STATE

2.1. Equations of State

We use a set of eight EOS in this study. All EOS produce cold NSs in β-equilibrium that can have gravitational masses Mg above 2 M. These include two EOS from Lattimer & Swesty (1991), the K0 = 220 MeV and K0 = 375 MeV variants (where K0 is the nuclear compressibility modulus), denoted LS220 and LS375; the relativistic mean field (RMF) model EOS from Shen et al. (2011c), denoted HShen; two RMF models based on the NL3 and the FSUGold parameter set (Shen et al. 2011a, 2011b) denoted GShen-NL3 and GShen-FSU2.1; an unpublished7 RMF model based on the DD2 interaction denoted HSDD2; and two recent RMF model EOS fit to astrophysical measurements of NS masses and radii (Steiner et al. 2013), denoted SFHo and SFHx. All of these EOS are available in a common format for download from http://www.stellarcollapse.org.

The EOS of finite-temperature nuclear matter in nuclear statistical equilibrium (NSE) has contributions from a baryonic component (nucleons and nuclei), a relativistic electron/positron Fermi gas, a photon gas, and, if neutrinos are trapped, a neutrino gas. The Helmholtz free energies of these components add linearly, and the pressure is then the sum of the partial pressures and a function of baryon density ρ, temperature T, and electron fraction Ye,

Equation (1)

While Pbaryon varies between the employed EOS, we add Pe and Pγ using the Timmes EOS (Timmes & Arnett 1999) available from http://cococubed.asu.edu. In hot HMNSs, like in protoNSs, neutrinos are trapped and in equilibrium with matter. We include their pressure contribution to the EOS by treating them as a non-interacting relativistic Fermi gas with chemical potential $\mu _{\nu _i}$. For a single species of neutrinos and antineutrinos, the neutrino pressure in equilibrium is

Equation (2)

where $\eta _{\nu _i} = \mu _{\nu _i} / (k_B T)$ is the neutrino degeneracy parameter. For HMNS conditions, all neutrino species are present, but νμ and ντ have $\mu _{\nu _i}=0$, since they appear only in particle–anti-particle pairs that have equal and opposite chemical potentials. For electron neutrinos we use $\mu _{\nu _e} = \mu _e + \mu _p - \mu _n$, for electron antineutrinos we use $\mu _{\bar{\nu }_e} = -\mu _{\nu _e}$. We include an attenuation factor exp (− ρtrap/ρ) to account for the fact that neutrinos decouple from matter at low densities. We set ρtrap = 1012.5 g cm−3, which is a fiducial trapping density for protoNSs (e.g., Liebendörfer 2005). Taking the exact expression for the difference of the Fermi integrals from Bludman & van Riper (1978), we have the total neutrino pressure summed over all three species,

Equation (3)

We note that due to the neutrino statistical weight g = 1, for a single species of relativistic non-degenerate $\nu \hbox{{--}}\bar{\nu }$ pairs, the pressure is a factor of two lower than for ee+ pairs, since e and e+ have statistical weight (spin degeneracy) 2.

Figure 1 illustrates the contributions of the partial pressures to the total pressure as a function of baryon density ρb for neutron-rich HMNS matter at two temperatures, 0.5 MeV (a representative "cold" temperature) and 20 MeV (a representative "hot" temperature for HMNSs). For the 0.5 MeV EOS, we set the electron fraction Ye by solving for ν-less β-equilibrium ($\mu _{\nu _e} = 0$). The resulting EOS describes ordinary cold NSs (at 0.5 MeV any thermal effects are negligible). For the 20 MeV case, we solve for Ye by assuming ν-full β-equilibrium. We do so by making the assumption that any neutrinos produced during the merger are immediately trapped in the HMNS core, but stream away from regions below trapping density. The procedure is discussed in the next Section 2.2 and detailed in Appendix B.

Figure 1.

Figure 1. Individual pressure contributions of baryons, electrons/positrons, photons, and trapped neutrinos and the total pressure as a function of baryon density in the LS220 EOS for ν-full β-equilibrium as described in the text and T = 0.5 MeV (dashed lines) and T = 20 MeV (solid lines). The qualitative and quantitative behavior of the LS220 EOS with increasing temperature is representative for all EOS considered in this study. Note that the baryon pressure becomes negative at ρb ≲ 1012 g cm−3, and dips around 1013.5 g cm−3 due to Coulomb effects at low temperatures (Lattimer & Swesty 1991).

Standard image High-resolution image

Near and above nuclear saturation density, ρnuc ≃ 2.6 × 1014 g cm−3 for the LS220 EOS, the baryon pressure is due to the repulsive core of the nuclear force and dominates in both cold and hot regimes. The thermal enhancement above ρnuc remains small even at 20 MeV. In the cold case, relativistically degenerate electrons (Γ = (dln P)(dln ρ)−1 = 4/3) dominate below ρnuc. At 20 MeV, relativistic non-degenerate electron/positron pairs and photons (for both, PT4, independent of ρb; see, e.g., van Riper & Bludman 1977) are the primary contributors at low densities, while the baryon pressure is significantly thermally enhanced below nuclear saturation density and dominates above ∼1012 g cm−3. The neutrino pressure is comparable to the degenerate electron pressure between ∼1012.5 and 1014 g cm−3, but still subdominant to the nuclear component. The contribution of pairs and photons gradually becomes more important at all densities as the temperature increases. We note that for T = 0.5 MeV, the neutrino chemical potentials are all zero and the pressure of trapped neutrinos is 3 × (7/8) × Pγ, thermodynamically insignificant at T = 0.5 MeV.

2.2. Temperature and Composition Parameterizations

The hydrostatic and rotational equilibrium equations that we solve in this study assume a barotropic EOS (P = P(ρ)) and do not provide constraints on thermal structure and composition (Ye is the only relevant compositional variable in NSE). We must make some assumptions to be able to proceed and obtain P = P(ρ, T(ρ), Ye(ρ, T(ρ))) for our general finite-temperature microphysical EOS. Old NSs in isolation are nearly isothermal (e.g., Prakash et al. 2001) and so are coalescing NSs until tidal heating becomes significant (e.g., Kochanek 1992; Lai 1994). During merger, the NS matter is shock-heated to tens of MeV and results of the few merger simulations that have been carried out with temperature-dependent EOS (e.g., Sekiguchi et al. 2011; Bauswein et al. 2010; Oechslin et al. 2007; Rosswog & Liebendörfer 2003; Ruffert & Janka 2001) indicate that the HMNS is far from being isothermal or isentropic. It has a very hot dense core with T ∼ 20–40 MeV surrounded by a lower-density cooler envelope/torus of 5–20 MeV, which may also be almost Keplerian and, hence, centrifugally supported. This result appears to be robust for equal-mass or near equal-mass NSNS systems (which may dominate the population; e.g., Lattimer 2012 and references therein). Mergers of non-equal mass systems in which the lower-mass NS is tidally wrapped around its more massive companion reach similar temperatures, but generally tend to have more mass at lower densities in the disk/torus (Oechslin et al. 2007).

There is no unique model/EOS independent mapping T = T(ρ), thus we must explore a variety of possibilities. In Figure 2, we contrast our set of temperature parameterizations with a T(ρ) profile obtained from a 1.35–1.35 M simulation using the HShen EOS by Sekiguchi et al. (2011) at ∼12 ms after merger. We consider very hot cores at 20, 30, and 40 MeV with cold envelopes (parameterizations c20p0, c30p0, and c40p0) and two parameterizations with very hot cores at 30 MeV and cool envelopes at 10 MeV and 5 MeV, c30p10 and c30p5, respectively. Since low-density regions have shorter neutrino cooling times, the c30p10 and c30p5 parameterization may represent early HMNSs, while the cold-envelope parameterizations c20p0, c30p0, and c40p0 may correspond to late-time HMNSs. Note that the c30p10 parameterization fits the temperature profile from the Sekiguchi et al. (2011) simulation quite well. Details on the functional forms of our parameterizations can be found in Appendix A. For the TOV case we also consider isothermal configurations as a limiting case.

Figure 2.

Figure 2. Temperature (T, left panel) and electron fraction (Ye, right panel) as a function of baryon density for the T and Ye prescriptions we explore in this work compared to three-dimensional NSNS simulation data of Sekiguchi et al. (2011; dashed brown lines). The profiles are created by taking T, Ye, and ρb data along the +x-axis from their low-mass (two 1.35 M progenitor NSs) simulation at 12.1 ms after merger. In the right panel, the dashed brown graph denotes the Ye obtained from the simulation, while the solid and the dash-dotted graphs are Ye obtained from the simulation temperature profile for ν-full and ν-less β-equilibrium, respectively.

Standard image High-resolution image

The choice of Ye(ρ, T(ρ)) is equally difficult. Before merger, the NSs are in ν-less β-equilibrium (μν = μe + μp − μn = 0). After merger, neutrinos are present. They are trapped in hot dense matter (μν ≠ 0) and are streaming away from low-density regions. The equilibrium Ye will shift and mixing due to non-linear hydrodynamics in the HMNS phase will distort any initial Ye(ρ, T(ρ)) profile.

We deem the following prescription for Ye to be the physically most sensible. We assume that the NSNS merger occurs so rapidly that the electron fraction Ye of the ν-less β-equilibrium in the NSs becomes the trapped postmerger lepton fraction $Y_\mathrm{lep} = Y_e + Y_{\nu _e} - Y_{\bar{\nu }_e}$ above ρtrap. Using the β-equilibrium condition with nonzero μν, we solve for Ye. At densities below ρtrap we transition to Ye given by ν-less β-equilibrium. Details of this procedure are given in Appendix B. We refer to this parameterization of Ye as ν-full β-equilibrium. In addition and for comparison, we consider choices of constant Ye = 0.1 and Ye set according to ν-less β-equilibrium. We note that our parameterization of Ye is ad hoc and cannot account for mixing and neutrino transport effects in the merger process. The right panel of Figure 2 depicts Ye(ρ, T(ρ)) as obtained from the simulation of Sekiguchi et al. (2011) contrasted with Ye profiles computed under the assumption of ν-less and ν-full β-equilibrium for various temperature parameterizations and for the T(ρ) as given by the simulation. None of the prescriptions fit the simulation-Ye particularly well, which indicates that mixing and neutrino transport effects are important (but cannot be included here). The Ye obtained using the temperature data from the simulation naturally fits best, in particular at low densities where neutrinos have decoupled from the matter and ν-less β-equilibrium holds.

In the top panels of Figure 3, we show the fractional pressure increase due to thermal effects as a function of baryon density for our set of temperature parameterizations for the LS220 EOS (left panel) and the HShen EOS (right panel) as two representative example EOS. We also distinguish between the choices of Ye parameterization. For the parameterizations with cold "mantles" (cXp0), thermal effects are most important at densities near ∼ρnuc and quickly lose significance at lower and higher densities in both EOS. The thermal pressure enhancement is at most a factor of three (for the HShen) to five (for the LS220 EOS) for these parameterizations. The situation is different for the cases with hot plateaus, c30p10 and c30p5. For these, the thermal pressure is up to 20 times larger at low densities than predicted by the cold EOS. The Ye parameterizations corresponding to ν-full and ν-less β-equilibrium yield qualitatively and quantitatively very similar results for both EOS.

Figure 3.

Figure 3. Effects of temperature and Ye parameterizations on the pressure and relevance of the neutrino pressure component. Top panels: fractional increase of the pressure over the cold ν-less β-equilibrium pressure for the LS220 EOS (left panel) and the HShen EOS (right panel). The different line styles correspond to Ye(ρ) obtained in ν-full β-equilibrium (solid), ν-less β-equilibrium (dash-dotted), and constant Ye = 0.1 (dashed). Bottom panels: relative contribution of the neutrinos to the total pressure (see Equation (3)) in the five temperature and three Ye parameterizations and the LS220 EOS (left panel) and the HShen EOS (right panel).

Standard image High-resolution image

At low densities, the ν-full and ν-less β-equilibrium cases both lead to Ye > 0.1 (see Figure 2). As a consequence, the pressure in the unrealistic Ye = const. = 0.1, cXp0 parameterizations is lower than in the cold ν-less case at ρb ≲1012.2 g cm−3. Due to the logarithmic scale of Figure 3, the graphs of cXp0 with Ye = 0.1 start only there and the predicted pressure enhancement is higher than in the β-equilibrium cases, which lead to lower Ye above ∼1012.2 g cm−3 and below ∼ρnuc. In the cases with hot plateau (c30p10 and c30p5), thermal effects dominate over differences in Ye at low densities. Finally, at ρ > ρnuc, where temperature effects are smaller, differences in Ye become important. Since the nuclear component dominates there, lower Ye corresponds to higher pressure (e.g., Lattimer & Prakash 2001) and both β-equilibrium cases yield Ye > 0.1.

The lower panels of Figure 3 depict the relative contribution of the neutrinos to the total (hot) pressure in the HMNS temperature and Ye parameterizations considered in this study. While there are clear temperature (see Equation (3)) and Ye (through $\mu _{\nu _e}$) dependences, neutrino pressure plays only a minor role, making up at most ∼2% of the total pressure of the LS220 EOS. This is true also for the HShen EOS with the exception of the unrealistic Ye = 0.1 case in which the neutrino pressure contribution grows to ≳10% of the total pressure at supranuclear densities.

Finally, we note that the temperature and Ye prescriptions discussed here lead to regions that may be unstable to convection if not stabilized by a positive specific angular momentum gradient (e.g., Tassoul 1978). The spherically and axially symmetric equilibrium models that we construct in this study cannot account for convection and we leave an analysis of convective instability to future work.

2.3. Spherically Symmetric Equilibrium Models

We solve the TOV equation (e.g., Shapiro & Teukolsky 1983),

Equation (4)

where r is the areal (circumferential) radius, ρb is the baryon density, epsilon is the specific internal energy, and Mg(r) is the gravitational mass enclosed by radius r, determined via

Equation (5)

The baryonic mass is larger and given by

Equation (6)

We construct the TOV solutions using a standard fourth-order Runge–Kutta integrator on an equidistant grid with δR = 102 cm zones. After each integration sub-step, the EOS P = Pb) is inverted to obtain ρb. We use a variety of Pb) parameterizations: (1) T = const. (isothermal) with ν-full β-equilibrium above ρtrap and ν-less β-equilibrium below, (2) T = const. with ν-less β-equilibrium, (3) T = const. with constant Ye = 0.1, and (4) the phenomenological cXpX temperature parameterizations with ν-full β-equilibrium above ρtrap and ν-less equilibrium below. We compute TOV solutions for all EOS and define the surface of the NS as the areal radius at which one of the following two conditions is true: (1) the pressure equals 10−10 of the central pressure; (2) the pressure predicted by the integration of Equation (4) drops below the lowest pressure value available in the equation of state table. The latter is not a limitation, because the high-density TOV configurations considered here have steep density and pressure profiles near their surfaces. The pressure dropping to very small values thus indicates that the surface has been reached.

Besides the EOS, temperature, and Ye prescription, the central baryon density ρb, c is the only other free parameter. Since we are interested in the maximum mass that can be supported, we compute sequences with varying ρb, c for each EOS, but limit ourselves to ρnuc < ρb, c ⩽ ρmax, EOS, where the latter is just the maximum density entry in the respective EOS table. HMNSs with central densities below ρnuc are not realistic (see Sekiguchi et al. 2011).

We make our TOV solver, all P = Pb) tables, and the Python scripts used to create the results in this paper available on http://www.stellarcollapse.org.

2.4. Axisymmetric Equilibrium Models

We generate axisymmetric equilibrium models using the code originally presented in (Cook et al. 1992, hereafter CST; see also Cook et al. 1994a, 1994b), which is based on the relativistic self-consistent field method of Komatsu et al. (1989a). The axisymmetric equilibrium equations are solved iteratively on a grid in (s, μ), where s is a compactified radial coordinate and μ = cos θ, where θ is the usual spherical polar angle. Additionally, metric functions are solved using Green's functions integrals expanded in terms of Nl Legendre polynomials. Consequently, the total numerical resolution is specified via a tuple of (Ns, Nμ, Nl), which we set to (500, 300, 16). The resolution is chosen so that the resulting integral quantities of the equilibrium solution (e.g., its gravitational mass) are precise to about one part in 103. The surface of the star is defined by an enthalpy contour which is specified in the code by setting a surface energy density. This energy density has a default value of 7.9 g cm−3, and we have checked that increasing its value by a factor of 106 leaves the physical quantities of the solution unchanged to our stated general error level of 10−3.

An axisymmetric HMNS equilibrium configuration is constructed by the CST code based on choices of (1) a barotropic EOS, (2) a rotation law, (3) the rotation rate, and, (4) the maximum mass-energy density Emax = [ρb(1 + epsilon/c2)]max of the configuration.

In order to keep the size of the parameter space manageable, we restrict rotating configurations to the LS220 and HShen EOS and set up barotropic versions using the temperature and composition parameterizations described in Section 2.2. Since the EOS obtained with ν-full and ν-less β-equilibrium differ only very mildly (see Figure 3), we construct rotating configurations under the simple assumption of ν-less β-equilibrium.

We employ the "j − const." rotation law (see, e.g., CST), which is commonly used in the literature for HMNS models (e.g., Baumgarte et al. 2000). The degree of differential rotation is parameterized by $\tilde{A}$.8 In the Newtonian limit, this rotation law becomes $\Omega = \Omega _c / (1 + \tilde{A}^2 \varpi ^2/r_{\rm e}^2)$, where ϖ is the cylindrical radius, re is the radius of the star at its equator, and Ωc is the central angular velocity. For $\tilde{A} = 0$, one recovers uniform rotation, while for large $\tilde{A}$, the specific angular momentum becomes constant (i.e., Ω∝ϖ−2 in the Newtonian limit). We explore values of $\tilde{A}$ between 0 and 1. The latter value of $\tilde{A}$ corresponds to roughly a factor of two decrease of the angular velocity from the center to the HMNS surface, which is in the ball park of what is found in merger simulations (e.g., Shibata et al. 2005). Once the rotation law is fixed, the rotation rate is determined by specifying the axis ratiorp/e, defined as the ratio of the HMNS radius along the pole rp divided by the radius at the equator re.

The final parameter to be chosen is the maximum energy density of the configuration. For simplicity and consistency with the choice of variables for the TOV solutions discussed in Section 2.3, we set Emax by choosing a maximum baryon density ρb, max and obtain Eb, max) from the EOS.

For each choice of EOS, $\rho _\mathrm{b,\mathrm{max}}$, and $\tilde{A}$, we compute a sequence of models with increasing rotation rate, stepping down from rp/e = 1 (the nonrotating TOV case) until we reach mass shedding or until the code fails to converge to an equilibrium solution. In the case of uniform rotation ($\tilde{A} = 0$) the sequence always ends at mass shedding, the resulting rotating NS has spheroidal shape, and the maximum and central density coincide (ρb, max = ρc). Differentially rotating sequences, on the other hand, can bifurcate into two branches: one with ρb, max = ρc and spheroidal geometry and one with an off-center location of ρb, max and quasitoroidal shape. For differentially rotating models, the CST solver generally fails to converge to a solution at rp/e before mass shedding and, therefore, possibly before the maximum mass for a given configuration is reached. This limitation means that the maximum masses we state for differentially rotating models are to be interpreted as lower bounds on the true maximum masses. The code developed by Ansorg et al. (2003) is far more robust than CST for such extreme configurations and these authors have argued that with increasing degree of differential rotation, arbitrarily large masses could be supported in extremely extended tori, but such configurations are unlikely to be astrophysically relevant.

3. RESULTS: SPHERICALLY SYMMETRIC MODELS

Our main interest is in how temperatures in the range encountered in HMNS of NSNS postmerger simulations change the maximum mass that can be supported. Since baryonic mass is a conserved quantity and can be related to the number of baryons present in the individual NSNS before merger (modulo a small amount of potential ejecta), we treat it as a the most important variable and define the maximum gravitational masses $M_\mathrm{g}^\mathrm{max}$ as the gravitational mass at which $M_\mathrm{b}^\mathrm{max}$ is maximal. We consider the isothermal TOV solution as a limiting case of maximal thermal support but note that such configurations with T ≳ 5–8 MeV develop very large, non-degenerate envelopes at the low end of the central baryon densities ρb, c considered here. With increasing temperature, degeneracy is more and more lifted at those densities and the TOV model approaches an isothermal sphere whose pressure is dominated by relativistic non-degenerate pairs and whose mass and radius become infinite. We discard such solutions.

The results of our TOV calculations are summarized by Figure 4 for all considered EOS. We provide numerical results in Table 1 for fiducial isothermal cold (T = 0.5 MeV) and parameterized temperature choices.

Figure 4.

Figure 4. Effect of temperature T on the maximum masses of TOV solutions. Top panel: gravitational mass $M_\mathrm{g}^\mathrm{max}$ at the maximum baryonic mass for T = const. configurations (lines) and parameterized cXpX profiles (symbols). The cXpX solutions are computed only for ν-full β-equilibrium. With increasing T, $M_\mathrm{g}^\mathrm{max}$ increases. This trend is independent of Ye prescription, but the sensitivity to Ye is highly EOS dependent. Center panel: relative increase of $M_\mathrm{g}^\mathrm{max}$ with T for solutions in ν-full β-equilibrium. The increase is modest and below ∼10% even in the T = const. case. Bottom panel: maximum baryonic mass $M_\mathrm{b}^\mathrm{max}$ that can be supported as a function of temperature. For most EOS, there is little variation in $M_\mathrm{b}^\mathrm{max}$ at low T, but the increasing thermal contribution to the TOV energy density (see Equation (4)) leads to a decrease of $M_\mathrm{b}^\mathrm{max}$ for high-T solutions. A linear vertical shift of −0.30 (−0.50) M has been applied to the LS375 (GShen-NL3) curves to enhance the vertical dynamic range of the plot.

Standard image High-resolution image

Table 1. Summary of TOV Results for All EOS

EOS T(ρ) $M^\mathrm{max}_\mathrm{b}$ $M^\mathrm{max}_\mathrm{g}$ R ρc
(M) (M) (km) (1015 g cm−3)
 
LS220, ν-less 0.5 MeV 2.406 2.042 10.63 1.863
LS220, ν-full c20p0 2.434 2.068 10.69 1.873
  c30p0 2.433 2.078 10.89 1.840
  c30p10 2.433 2.079 11.86 1.840
  c30p5 2.433 2.078 11.23 1.840
  c40p0 2.428 2.087 11.07 1.808
 
LS375, ν-less 0.5 MeV 3.349 2.715 12.34 1.243
LS375, ν-full c20p0 3.322 2.717 12.59 1.232
  c30p0 3.294 2.717 12.68 1.221
  c30p10 3.293 2.718 13.49 1.221
  c30p5 3.293 2.717 12.95 1.221
  c40p0 3.264 2.714 12.75 1.210
 
HShen, ν-less 0.5 MeV 2.560 2.214 12.59 1.357
HShen, ν-full c20p0 2.584 2.246 13.17 1.321
  c30p0 2.601 2.273 13.48 1.276
  c30p10 2.604 2.277 15.08 1.276
  c30p5 2.603 2.275 14.01 1.276
  c40p0 2.613 2.295 13.69 1.243
 
GShen-NL3, ν-less 0.5 MeV 3.353 2.765 13.34 1.115
GShen-NL3, ν-full c20p0 3.354 2.781 13.51 1.098
  c30p0 3.344 2.791 13.70 1.081
  c30p10 3.346 2.793 15.04 1.081
  c30p5 3.345 2.792 14.30 1.081
  c40p0 3.330 2.796 13.86 1.070
 
GShen-FSU2.1, ν-less 0.5 MeV 2.468 2.114 11.67 1.505
GShen-FSU2.1, ν-full c20p0 2.488 2.140 12.15 1.474
  c30p0 2.497 2.159 12.40 1.428
  c30p10 2.502 2.164 14.30 1.420
  c30p5 2.497 2.160 12.44 1.428
  c40p0 2.504 2.176 12.56 1.398
 
HSDD2, ν-less 0.5 MeV 2.896 2.419 11.92 1.395
HSDD2, ν-full c20p0 2.891 2.429 12.28 1.381
  c30p0 2.883 2.436 12.43 1.367
  c30p10 2.884 2.437 13.47 1.367
  c30p5 2.883 2.436 12.79 1.367
  c40p0 2.871 2.440 12.55 1.353
 
SFHo, ν-less 0.5 MeV 2.433 2.057 10.31 1.906
SFHo, ν-full c20p0 2.434 2.068 10.67 1.884
  c30p0 2.433 2.078 10.86 1.862
  c30p10 2.434 2.079 11.81 1.862
  c30p5 2.433 2.078 11.21 1.851
  c40p0 2.428 2.087 11.03 1.829
 
SFHx, ν-less 0.5 MeV 2.529 2.127 10.79 1.722
SFHx, ν-full c20p0 2.531 2.139 11.18 1.705
  c30p0 2.530 2.150 11.37 1.688
  c30p10 2.531 2.151 12.39 1.688
  c30p5 2.531 2.150 11.72 1.688
  c40p0 2.527 2.160 11.51 1.671

Notes. "ν-less" indicates neutrino-less β-equilibrium, which we use only for the "cold" configurations. "ν-full" indicates neutrino-full β-equilibrium with neutrino pressure. T(ρ) is the temperature parameterization, $M_\mathrm{b}^\mathrm{max}$ is the maximum baryonic mass, $M_\mathrm{g}^\mathrm{max}$ is the gravitational mass at the maximum baryonic mass, R is the radius of the $M_\mathrm{b}^\mathrm{max}$ configuration, and ρc is the central baryon density at which $M_\mathrm{b}^\mathrm{max}$ obtains.

Download table as:  ASCIITypeset image

In the top panel of Figure 4, we show the maximum gravitational mass (defined as Mg at $M_\mathrm{b}^\mathrm{max}$) as a function of isothermal temperature for our three Ye prescriptions. The considered EOS show a great degree of variation in their sensitivity to Ye prescriptions, but the overall trend is clear: increasing temperature generally leads to increasing $M_\mathrm{g}^\mathrm{max}$. The fractional increase over the cold value, however, is not large, as shown by the center panel. The HShen and GShen-FSU2.1 RMF TOV stars are the most sensitive to temperature variations,9 but even their maximum gravitational TOV mass increases only by ∼12%–15% at isothermal T = 50 MeV. The cXpX temperature parameterizations, shown as symbols in Figure 4 located at their respective central temperatures, generally follow the trend of the isothermal sequences for each EOS, but their $M_\mathrm{g}^\mathrm{max}$ enhancement is systematically lower, since they are only centrally hot.

The lower panel of Figure 4 depicts the change of the maximum baryonic TOV mass $M_\mathrm{b}^\mathrm{max}$ with increasing temperature. For most EOS, $M_\mathrm{b}^\mathrm{max}$ stays roughly constant at low temperatures, but decreases at high temperatures. This shows that the increase in $M_\mathrm{g}^\mathrm{max}$ in the TOV solutions is primarily due to thermal contributions to the total mass-energy density. Since it is the mass-energy density, and not just the baryonic mass, which sources curvature (the relativistic gravitational field), the thermal effects lead to a decrease in $M_\mathrm{b}^\mathrm{max}$ with temperature even if $M_\mathrm{g}^\mathrm{max}$ is still increasing. The HShen and GShen-FSU2.1 are the only two EOS that exhibit an increase of $M_\mathrm{b}^\mathrm{max}$ at intermediate to high temperatures, but they too reverse this trend at isothermal T ≳ 50 MeV. The LS375 EOS, on the other hand, has monotonically decreasing $M_\mathrm{b}^\mathrm{max}$ with T, which was seen before by O'Connor & Ott (2011). The more realistic cXpX temperature parameterizations show a similar trend as their isothermal counterparts, but for the HShen and GShen-FSU2.1 EOS, the increase in $M_\mathrm{b}^\mathrm{max}$ at intermediate T is smaller in these only centrally hot parameterized models.

It is interesting to compare our findings with the results of O'Connor & Ott (2011), who studied black hole formation through protoNS collapse in failing core-collapse supernovae. These authors found much larger maximum baryonic and gravitational masses of their protoNSs at the onset of collapse than reported here. The collapsing protoNSs in their study have moderately high central temperatures T ≲ 40 MeV. However, at ρ ≈ 4 × 1014–1015 g cm−3, a region of extremely hot material with T ≳ 80–100 MeV is present due to compression of multiple M of accreted shock-heated material. O'Connor & Ott (2011) demonstrated that this extremely hot region is responsible for the observed thermal enhancement of the maximum protoNS mass. In NSNS mergers the situation is quite different and fully dynamical NSNS merger simulations have not found such extremely hot high-density regions (e.g., Sekiguchi et al. 2011; Oechslin et al. 2007). It is thus unlikely that the findings of O'Connor & Ott (2011) apply to the merger HMNS case. Table 1 summarizes key parameters of the computed TOV solutions.

4. RESULTS: AXISYMMETRIC MODELS IN ROTATIONAL EQUILIBRIUM

4.1. Uniformly Rotating Configurations

It has been widely recognized that uniform rotation can support a supramassive NS against gravitational collapse (see, e.g., Friedman et al. 1986; Friedman & Ipser 1987). A supramassive NS is defined as a stable NS with a mass greater than the maximum mass of a TOV star with the same EOS (CST). At a given central density, the mass that may be supported rises with increasing angular velocity until the material on the NS's equator becomes unbound (the mass-shedding limit). This leads to the supramassive limit, a well defined maximum mass for uniformly rotating NSs with a specified EOS.

In Figure 5, we plot the baryonic mass Mb as a function of maximum baryon density for TOV and uniformly rotating mass-shedding sequences obtained with the LS220 EOS (left panel) and the HShen EOS (right panel). Focusing first on the TOV sequences, one notes that at low central densities (ρb ≲ few × ρnuc), Mb is significantly increased by thermal effects. This is because the mean density $\bar{\rho }_{{\rm _b}}$ of such configurations is in the regime in which thermal pressure is of greatest relevance (see Figure 3) and can alter the structure of the bulk of the NS. This carries over to the uniformly rotating case. The extended hot configurations reach mass shedding at lower angular velocities than their cold counterparts, but the extended, low $\bar{\rho }_{{\rm b}}$ cores of hot configurations receive sufficient rotational support to yield a higher Mb. This, however, is the case only for centrally hot cXp0 configurations. Models with hot envelopes (with parameterizations c30p5 and c30p10) benefit less from rotational support.

Figure 5.

Figure 5. Baryonic mass Mb as a function of maximum baryon density ρb, max of uniformly rotating ($\tilde{A}=0$) equilibrium models at the mass-shedding limit for different temperature prescriptions (solid lines). We also plot the corresponding TOV sequences (dashed lines) and show results for the LS220 EOS (left panel) and HShen EOS (right panel). There is a large thermal enhancement of Mb at low densities, but the sequences converge toward the cold supramassive limit as the maximum density increases and the configurations become more compact.

Standard image High-resolution image

With increasing maximum density, the baryonic masses of the TOV models for different temperature parameterizations converge for a given EOS. Near the density at which the maximum mass is reached, the increase in Mb in hot configurations has turned into a slight decrease for models computed with the LS220 EOS and has dropped to ≲5% for the HShen EOS (see also Figure 4). The mass-shedding sequences show a more complex behavior with increasing maximum density. As in the TOV case, the mean density $\bar{\rho }_b$ of the NSs increases and less material is experiencing enhanced pressure support due to high temperatures in the cXp0 models. Hence, these models move toward the $M_\mathrm{b}^\mathrm{max}$ of the cold supramassive limit (see the inset plots in Figure 5). For both EOS, the $M_\mathrm{b}^\mathrm{max}$ of hot configurations are all lower than the cold value. The cXp0 models reach supramassive limits that are within less than 2% of the cold supramassive limit for both EOS. The c30p10 and c30p5 models, on the other hand, have $M_\mathrm{b}^\mathrm{max}$ that are ∼5%–10% lower than the cold supramassive limit for both EOS. Table 2 summarizes key parameters of the hot and cold configurations at the supramassive limit.

Table 2. Uniformly Rotating Neutron Stars at the Supramassive Limit

Model ρb, max $M_\mathrm{b}^\mathrm{max}$ $M_\mathrm{g}^\mathrm{max}$ re rp/e Ω T/|W|
(1015 g cm−3) (M) (M) (km) (103 rad s−1)
LS220 cold 1.653 2.823 2.419 14.429 0.566 10.096 0.118
LS220 c20p0 1.652 2.760 2.384 14.788 0.574 9.647 0.106
LS220 c30p0 1.652 2.737 2.382 15.000 0.576 9.441 0.103
LS220 c30p5 1.710 2.671 2.322 15.300 0.587 9.031 0.088
LS220 c30p10 1.769 2.587 2.247 16.130 0.599 8.215 0.066
LS220 c40p0 1.625 2.717 2.383 15.201 0.577 9.262 0.101
 
HShen cold 1.220 3.046 2.649 17.101 0.564 8.233 0.117
HShen c20p0 1.196 3.006 2.629 17.760 0.573 7.745 0.105
HShen c30p0 1.171 3.009 2.648 18.173 0.574 7.511 0.103
HShen c30p5 1.228 2.916 2.564 18.665 0.588 7.086 0.084
HShen c30p10 1.261 2.808 2.467 20.070 0.604 6.238 0.060
HShen c40p0 1.139 3.012 2.664 18.474 0.574 7.355 0.101

Notes. Summary of mass-shedding uniformly rotating supramassive neutron star configurations at the maximum mass for each EOS and temperature prescription. These models are in ν-less β-equilibrium (see Section 2.2). ρb, max is the central density of the model with the maximum baryonic mass $M_\mathrm{b}^\mathrm{max}$. $M_\mathrm{g}^\mathrm{max}$ is the gravitational mass at the ρb, max at which $M_\mathrm{b}^\mathrm{max}$ occurs. re is the equatorial radius, rp/e is the axis ratio, Ω is the angular velocity, and T/|W| is the ratio of rotating kinetic energy T to gravitational energy |W|.

Download table as:  ASCIITypeset image

The systematics of the supramassive limit with temperature prescription becomes clear when considering Figure 6. This figure shows the baryonic mass Mb and gravitational mass Mg for uniformly rotating NSs as a function of angular velocity Ω for the LS220 and HShen EOS at fixed densities near the maximum of Mbb, max) (see Table 2). At fixed angular velocity below mass shedding, hotter configurations always yield higher Mg than their colder counterparts. For the LS220 EOS, as in the TOV case discussed in the previous section 3, hotter configurations have lower Mb. In the case of the HShen EOS, which generally yields less compact equilibrium models, the opposite is true, but the increase in Mb caused by thermal support is smaller than the increase in Mg.

Figure 6.

Figure 6. Gravitational mass (Mg, solid lines, right ordinate) and baryonic mass (Mb, dashed lines, left ordinate) as a function of angular velocity Ω for uniformly spinning models at a fixed density near the density that yields the maximum Mb for the LS220 (left panel) and the HShen EOS (right panel). The sequences terminate at the mass-shedding limit, which is the point with the maximum angular velocity for a specific temperature prescription. Configurations with higher temperatures, in particular, the c30p5 and c30p10 models with high-temperature plateaus at low densities, have larger radii than colder models and thus reach the mass-shedding limit at lower angular velocities. Hence, such models have lower maximum masses at the supramassive limit than colder models. Note that hotter models with the LS220 have lower baryonic masses than colder models.

Standard image High-resolution image

With increasing Ω, the mass-shedding limit is approached and hotter configurations systematically reach the mass shedding limit at lower angular velocities. The reason for this is best illustrated by comparing c30p0 models with c30p10 and c30p5 models, which have a high-temperature plateau at low densities of 10 MeV and 5 MeV, respectively. At low angular velocities, all c30pX models show the same thermal increase in Mg. However, the high pressure at low densities in the c30p10 and c30p5 models leads to significantly larger radii compared to the model without temperature plateau. Consequently, as Ω is increased, the configurations with plateau reach the mass-shedding limit at lower Ω. For the LS220 EOS, the c30p10 sequence terminates at ∼8200 rad s−1, the c30p5 sequence terminates at ∼9200 rad s−1, while the c30p0 sequence does not terminate before ∼9800 rad s−1. The HShen model sequences show the same qualitative trends.

4.2. Differentially Rotating Configurations

Differential rotation can provide centrifugal support at small radii while allowing a NS configuration to stay below the mass-shedding limit at its equatorial surface. Differentially rotating equilibrium configurations have been shown to support masses well in excess of the supramassive limit (e.g., Ostriker et al. 1966; Baumgarte et al. 2000; Morrison et al. 2004). Such configurations are referred to as "hypermassive." However, since there is (mathematically speaking) an infinite number of possible differential rotation laws, it is impossible to define a formal "hypermassive limit" for the maximum mass of HMNSs in the way it is possible for uniformly rotating supramassive NSs. Nevertheless, we can study the systematics of the supported baryonic (and gravitational) masses with variations in the HMNS temperature profile, maximum baryon density, and degree and rate of differential rotation for the rotation law considered in this study, which is not drastically different from what is found in merger simulations (e.g., Shibata et al. 2005).

In Figure 7, we show the supported baryonic mass Mb as a function of maximum baryon density ρb, max for cold, c20p0, and c40p0 temperature prescriptions, both EOS, and for different choices of $\tilde{A}$ (see Table 3 for a summary of quantitative results). The curves represent configurations with the minimum rp/e at which an equilibrium solution is found by the CST solver (i.e., the most rapidly spinning setup). Note that the peaks of these curves represent only lower limits on the maximum HMNS mass. In addition, we plot only solutions with ratios T/|W| of rotational kinetic energy T to gravitational energy |W| below 25%, since more rapidly spinning models would be dynamically nonaxisymmetrically unstable (Chandrasekhar 1969; Baiotti et al. 2007). It is this limit which defines the rising branch of the Mbb, max) curve at the lowest densities in Figure 7 for $\tilde{A} = 1.0$. Note that many of these configurations may still be unstable to secular rotational instabilities or rotational shear instabilities (e.g., Watts et al. 2005; Ott et al. 2007; Corvino et al. 2010).

Figure 7.

Figure 7. Maximum baryonic mass configurations for sequences of uniformly rotating ($\tilde{A}=0$) and differentially rotating ($\tilde{A} = \lbrace 0.4,0.5,1.0\rbrace$) models with cold, c20p0, and c40p0 temperature parameterizations and the LS220 EOS (left panel) and HShen EOS (right panel). We note that for differentially rotating models these curves represent lower limits on the maximum baryonic mass (i.e., the solver fails to converge at lower axis ratios without reaching the true mass shedding limit). We limit the sequences to models with T/|W| ≲ 0.25 and this limit defines the rising part of the graphs for $\tilde{A}=1$ at low densities. We show the TOV case (thinnest and shortest dash-dotted lines) for comparison. The raggedness of the curves with $\tilde{A} \gtrsim 0.4$ is a consequence of finite resolution in the parameter rp/e that is varied to find the maximum mass at a given ρb, max. Thermal effects are most pronounced at low densities and for high $\tilde{A}$. For uniform and moderate differential rotation, hotter models have lower global maximum Mb than colder models.

Standard image High-resolution image

Table 3. Differentially Rotating Hypermassive Neutron Stars

Model ρb, max $M_\mathrm{b}^\mathrm{max}$ $M_\mathrm{g}^\mathrm{max}$ re rp/e $\tilde{A}$ Ωc T/|W|
(1015 g cm−3) (M) (M) (km) (103 rad s−1)
LS220 cold 0.993 3.648 3.140 17.258 0.376 0.5 15.476 0.244
LS220 c20p0 0.852 3.573 3.124 18.538 0.364 0.6 15.047 0.243
LS220 c30p0 0.706 3.568 3.167 19.611 0.344 0.7 14.888 0.249
LS220 c30p5 0.600 3.413 3.064 21.870 0.320 0.9 14.461 0.250
LS220 c30p10 0.990 3.090 2.723 19.208 0.421 0.9 16.330 0.187
LS220 c40p0 0.692 3.597 3.211 19.931 0.344 0.7 14.677 0.249
 
HShen cold 0.766 4.101 3.562 19.800 0.372 0.5 13.450 0.245
HShen c20p0 0.641 4.076 3.585 21.352 0.360 0.6 13.042 0.245
HShen c30p0 0.532 4.099 3.650 22.305 0.344 0.7 13.131 0.249
HShen c30p5 0.517 3.942 3.527 24.371 0.340 0.8 12.426 0.243
HShen c30p10 0.646 3.529 3.141 23.521 0.400 1.0 13.934 0.196
HShen c40p0 0.514 4.148 3.708 22.701 0.344 0.7 12.888 0.249

Notes. Summary of the differentially rotating HMNS configurations with the largest baryonic masses for each EOS and temperature prescription. These configurations are obtained in a sequence from $\tilde{A} = 0$ to $\tilde{A} = 1$ with spacing $\delta \tilde{A} = 0.1$ and are to be seen as lower bounds on the maximum achievable masses. The sequences considered here exclude dynamically nonaxisymmetrically unstable models with ratio of rotational kinetic energy to gravitational energy T/|W| > 0.25. The quantities listed in the table are the following: ρb, max is the baryon density at which the maximum baryonic mass $M_\mathrm{b}^\mathrm{max}$ occurs, $M_\mathrm{g}^\mathrm{max}$ is the gravitational mass at that density, re is the equatorial radius of the configuration, rp/e is its axis ratio, $\tilde{A}$ is the differential rotation parameter at which $M_\mathrm{b}^\mathrm{max}$ obtains. Ωc is the central angular velocity of the configuration and T/|W| is its ratio of rotational kinetic energy to gravitational energy. We note that the accuracy of the results listed in this table is set by the step size in rp/e, which we set to δrp/e = 0.004.

Download table as:  ASCIITypeset image

The overall shape of the Mbb, max) curves in Figure 7 is qualitatively similar to what is shown in Figure 1 of Baumgarte et al. (2000) for Γ = 2 polytropes and Figure 2 of Morrison et al. (2004) for the cold Friedman–Pandharipande EOS (Friedman & Pandharipande 1981). The LS220 and HShen EOS yield qualitatively very similar results, but the supported HMNS masses found by the CST solver are, as expected, systematically higher for models with the HShen EOS than for those using the LS220 EOS. One notes, however, interesting variations with temperature prescription. At low ρb, max, thermal pressure support leads to increased Mb and more differentially rotating configurations have higher Mb. Sequences with $\tilde{A} \lesssim 0.5$ show similar systematics with density and temperature prescription as the uniformly spinning ones discussed in Section 4.1: As the density increases, hot configurations converge toward the cold sequence and reach their maximum Mb near and below the maximum of the cold sequence. Models with $\tilde{A} \gtrsim 0.5$, on the other hand, have more steeply rising curves with ρb, max and are discontinuous (i.e., exhibit a "kink") at their global maxima. At these points quasitoroidal solutions appear. Furthermore, the slope of the curve describing (as a function of ρb, max) the axis ratios rp/e at which the solver stops converging discontinuously changes sign. We attribute this behavior, which was also observed by Morrison et al. 2004, to a bifurcation of the sequence between models, which continue shrinking in axis ratio until they become completely toroidal (rp/e = 0), and less extreme models that stay quasitoroidal or spheroidal. Beyond the "kink" in $\tilde{A}\gtrsim 0.5$ sequences, thermal effects play little role.

The lower bounds of the range of ρb, max shown in the two panels of Figure 7 (and also Figure 8) are chosen for the following reason: fully dynamical merger simulations by, e.g., Sekiguchi et al. (2011), Baiotti et al. (2008), Shibata et al. (2005), Kiuchi et al. (2009), Bauswein et al. (2012), and Thierfelder et al. (2011), all suggest a rule of thumb that the postmerger maximum baryon density of the HMNS is typically not less than ∼80% of the central density of the progenitor NSs. We can derive a rather solid EOS-dependent lower limit on ρb, max for HMNS remnants from (equal mass) NSNS mergers in the following way. In order to form an HMNS, constituent equal-mass NSs must at the very least have a mass that is 50% of the maximum mass in the cold TOV limit. Hence, the premerger central density must at least be that of a TOV solution with $M_\mathrm{b} = 0.5 M_\mathrm{b}^\mathrm{max,TOV}$. Using the aforementioned empirical result from merger simulations, we arrive at

Equation (7)

For the LS220 EOS, ρb, TOV(Mb = Mb, max/2) ∼ 5.8 × 1014 g cm−3 and occurs at Mb (Mg) of 1.19 M (1.10 M). For the HShen EOS, ρb, TOV(Mb = Mb, max/2) ∼ 4.4 × 1014 g cm−3 and occurs at Mb (Mg) of 1.28 M (1.20 M). Applying the density cut given by Equation (7) excludes most dynamically nonaxisymmetrically unstable configurations.

Figure 8.

Figure 8. Same as Figure 7, but comparing cold configurations with models with the c30p5 and c30p10 temperature prescriptions, which have a hot plateau at low densities. The overall systematics are the same for the LS220 EOS (left panel) and the HShen EOS (right panel). In the TOV case, Mb is thermally enhanced at low densities, but the global maximum of Mb of hot configurations is near that of the cold TOV solution. Uniformly and moderately differentially rotating sequences of c30p10 and c30p5 models have systematically smaller maximum masses than cold models throughout the considered density range. Only very differentially rotating models ($\tilde{A}\gtrsim 0.7$; $\tilde{A}=1.0$ shown here) exhibit a thermal enhancement of the maximum mass at low to intermediate densities. The c30p10 sequence for $\tilde{A}=1.0$ exhibits a discontinuous jump, which occurs when the sequence transitions from spheroidal to quasitoroidal shape. See text for discussion.

Standard image High-resolution image

Figure 8, like Figure 7, shows baryonic mass as a function of maximum baryon density for both EOS and a variety of $\tilde{A}$, but contrasts models c30p5 and c30p10, which have hot plateaus at low densities, with cold models. The qualitative features discussed in the following are identical for both EOS. In the TOV case and at low densities, Mb is enhanced primarily by the hot core, since nonrotating solutions are compact and dominated by ρb ≳ 1014 g cm−3, where the high-temperature plateaus play no role. At higher densities, the Mb curves of hot models converge to near or below the cold TOV maximum Mb. The situation is different for uniformly and moderately differentially rotating models ($\tilde{A} \lesssim 0.5$). Rotation shifts these configurations to lower mean densities and the hot plateaus lead to equatorially bloated solutions. These reach their minimum rp/e for which a solution can be found at lower angular velocities. Hence, centrifugal support is weaker and the configuration with the hottest plateau has the lowest Mb, max. The behavior is different at high degrees of differential rotation ($\tilde{A} = 1$). The cold and the c30p5 models are HMNSs and quasitoroidal already at the lowest densities shown in Figure 8. The c30p5 sequence has slightly larger Mb than the cold sequence. The c30p10 sequence, however, is spheroidal at low ρb, max and then discontinuously transitions to the quasitoroidal branch, which is marked by a large jump in Mb.

In order to illustrate this discontinuous behavior further, we plot in Figure 9 the equatorial radius of equilibrium solutions as a function of central angular velocity at $\tilde{A} = 1$ and for three different fixed ρb, max. We show curves obtained with the LS220 EOS for the cold, c30p5, and c30p10 temperature prescriptions. The curves are parameterized by decreasing rp/e and terminate at the smallest value at which the solver converges. The three densities are chosen so that the first two are below and the third is above the jump of the c30p10 curve in Figure 8. At all ρb, max, the hot configurations have significantly larger radii than the cold models, but decreasing rp/e leads to increasing Ωc and only modest radius changes for cold and c30p5 models. This is very different for the c30p10 sequence. At ρb, max = 7.11 × 1014 g cm−3 these models do not become quasitoroidal and the re − Ωc mapping becomes double-valued as the decrease in rp/e turns from a decrease of rp at nearly fixed re and increasing Ωc into a steep increase of re and a decrease of Ωc. As ρb, max increases, less material is at low densities where thermal pressure support is strong in the c30p10 models. Consequently, the solutions are more compact and stay so to smaller rp/e. ρb, max = 8.16 × 1014 g cm−3 is the critical density at which the very last point in the sequence of decreasing rp/e (the one shown in Figure 8) jumps discontinuously to large re. At ρb, max = 9.21 × 1014 g cm−3, which is above the critical density for c30p10 in Figure 8, the c30p10 models become quasitoroidal as rp/e decreases and Ωc increases. They exhibit the same systematics as the c30p5 and cold models. We note that what we have described for the c30p10 models also occurs for the c30p5 models, although at significantly lower densities ρb, max ≲ 5 × 1014 g cm−3 and even the cold models show similar trends at low densities.

Figure 9.

Figure 9. Equatorial radii re vs. central angular velocity Ωc in sequences parameterized by the axis ratio rp/e for models using the LS220 EOS, differential rotation parameter $\tilde{A}=1.0$, and cold, c30p5, and c30p10 temperature parameterizations. We show curves for three densities, two below the discontinuous jump of the c30p10 curve in Figure 8 and one above. At the same density, hotter configurations have larger radii and transition to quasitoroidal shape (marked by dots) at higher Ωc. The transition between spheroidal and quasitoroidal shape is discontinuous in ρb, max for critical models at the minimum rp/e that can be found (shown in Figures 7 and 8), but smooth in rp/e at fixed ρb, max. The low-density sequences with the c30p10 temperature prescription (10 MeV plateau at low densities; see Section 2.2) become double valued in Ωc with increasing rp/e, stay spheroidal and have very large re.

Standard image High-resolution image

The sequences shown in Figures 7 and 8 are extreme configurations in the sense that models with smaller rp/e cannot be found by the CST solver and may not exist for the rotation law that we consider here. Real HMNS may not by such critical rotators. In Figure 10, we plot Mb and Mg for the LS220 EOS as a function of central angular velocity Ωc and temperature prescription. We fix the degree of differential rotation to $\tilde{A}=1$ and show sequences in Ωc for a fixed maximum density ρb, max = 9.21 × 1014 g cm−3, which is the highest density shown in Figure 9. The transition to quasitoroidal shape is smooth and quasitoroidal configurations are marked with symbols. The end points of the Mb curves shown in Figure 10 and in Figure 9 correspond to the Mb values of the $\tilde{A}=1$ curves in Figures 7 and 8 at 9.21 × 1014 g cm−3.

Figure 10.

Figure 10. Baryonic mass Mb and gravitational mass Mg vs. central angular velocity Ωc parameterized by the axis ratio rp/e at fixed degree of differential rotation $\tilde{A}=1$, and fixed maximum density of ρb, max = 9.21 × 1014 g cm−3. Curves for all temperature parameterizations are shown for the LS220 EOS. Quasitoroidal configurations are marked by symbols and the transitions between spheroidal and quasitoroidal solutions are smooth. The end points of all graphs correspond to the values plotted in Figures 7 and 8 for the various temperature prescriptions at $\tilde{A}=1.0$ and the ρb, max chosen here. Sequences with hot plateaus (using temperature prescriptions c30p5 and c30p10) exhibit significant thermal enhancements of Mb and Mg at rapid rotation rates, but have lower maximum rotation rates due to their larger radii.

Standard image High-resolution image

Figure 10 shows that, as in the case of uniform rotation (see Figure 6), hotter subcritically differentially spinning configurations have higher Mg. At the density chosen for this plot, they also have higher Mb, but at the higher densities at which the masses of uniformly spinning models peak, the Mb of hotter configurations are smaller than those of colder ones. It is particularly remarkable that the models with the hot plateau at low densities show the greatest thermal enhancement. They also transition to a quasitoroidal shape last but terminate the earliest in Ωc. Nevertheless, for the ρb, max chosen here, they can support slightly more mass at critical rotation than their counterparts without low-density temperature plateau.

5. DISCUSSION AND COMPARISON WITH THREE-DIMENSIONAL NSNS SIMULATIONS

5.1. The Stability of HMNS Equilibrium Sequences

The existence of a maximum mass for equilibrium sequences of nonrotating (TOV) NSs is one of the most important astrophysical consequences of general relativity and, hence, is well known in the study of compact objects. The parameter space of hot differentially rotating HMNS models studied here is vast and complex. In the following, we briefly review the classical results on the stability of stationary NSs and formulate how one may reason regarding the stability of HMNS equilibrium models.

A particular useful approach to the stability problem is the turning-point method of Sorkin (1982). The turning-point method allows one to reason about the stability of sequences of equilibrium solutions solely by examining the parameter space of equilibrium models without dynamical simulations or linear perturbation analysis. The turning-point method has been used extensively in previous work on the stability of cold and uniformly rotating NSs (e.g., CST; Friedman et al. 1988; Stergioulas & Friedman 1995; Read et al. 2009).

An equilibrium sequence is a one-dimensional slice from the space of equilibrium models indexed by some parameter. Here we use ρb, max as our sequence parameter. A model in the space of equilibrium models may be defined by the following conserved quantities: the gravitational mass Mg, baryonic mass Mb, total angular momentum J, and total entropy S. Generally, as one changes the sequence parameter, ρb, max, the quantities (Mg, Mb, J, S) will vary. A turning point in the sequence occurs when 3 out of 4 of the derivatives d/dρb, max of (Mg, Mb, J, S) vanish. For this point in ρb, max, the turning point theorem shows (1) that the derivative of the fourth quantity in the tuple also vanishes, and (2) that the sequence must have transitioned from stable to unstable (Sorkin 1982; Kaplan 2014). This characterization of the space of equilibrium models relies on the assumption that the change in Mg depends to first order only on the total changes in baryonic mass Mb, angular momentum J, and entropy S, and not on changes to their higher moments. That is, changes in the distribution of entropy, baryonic mass and angular momentum. In nature, this will generally not be the case, since cooling and angular momentum redistribution will change the entropy and angular momentum distributions, respectively. However, these changes will be slow and not drastic so that changes to the total energy due to changes in these higher order moments will be small. We account for such changes approximately by considering different degrees of differential rotation and a range of temperature prescriptions in the following.

If we are considering the special case of zero-temperature configurations, then the entropy S is no longer relevant to the equilibrium's stability, since the change to the configuration's energy due to a change in entropy is also zero. In this case, a turning point may be identified when two out of three of the set d/dρb, max(Mg, Mb, J) are zero. Zero temperature is a very good approximation for our cold equilibrium models. In Figure 11, we plot Mg along constant Mb sequences with Mb = 2.9 M for the HShen EOS (Mb = 2.9 M corresponds to Mb of an HMNS formed from two NSs of Mg = 1.35 M, assuming no mass loss). All of these curves have a minimum located at ρb, max ≳ 1 × 1015g cm−3.

Figure 11.

Figure 11. Gravitational mass Mg as a function of maximum baryon density ρb, max for models with Mb = 2.9 M. Each curve is for a fixed degree of differential rotation $\tilde{A}$, with the axis ratios rp/e chosen such that Mb = 2.9 M. Symbols mark equilibrium solutions at the minimum rp/e for which a solution can be found for Mb = 2.9 M and a given $\tilde{A}$ (i.e., the solver fails to converge when searching for a Mb = 2.9 M mass model at densities outside the bounds of the symbols). The local minima of these curves are approximate turning points of the sequences. For the cold (c40p0) models, we have noted the range in Mg and ρb, max across models with different amounts of differential rotation with dashed (solid) blue lines. Consequently, ρb, max = 1.30 × 1015 g cm−3 represents the upper limit for the baryon density of a stable HMNS with the HShen EOS. Note also that the difference in Mg of the approximate turning points between sequences with the same temperature prescription is only ∼0.005 M.

Standard image High-resolution image

For the cold sequences, these minima are turning points because dMg/dρb, max and dMb/dρb, max are both zero. Any models along those curves at densities in excess of ρb, max at the minima are secularly unstable to collapse. For the hot temperature parameterizations,10 the minima are only approximations to the turning point (which we shall call approximate turning points) because only two out of four (dMg/dρb, max and dMb/dρb, max) of the derivatives of (Mg, Mb, J, S) are zero. We argue that these approximate turning points are good indicators of the onset of instability for the equilibrium sequences for several reasons. (1) We find that the approximate turning points for all considered temperature parameterizations and measures of differential rotation ($\tilde{A} = 0$ to $\tilde{A} = 1.1$ with spacing $\delta \tilde{A} = 0.1$) lie within the same ∼25% range in ρb, max indicated by the blue lines in Figure 11 (similarly within a ∼25% range in ρb, max for the LS220 EOS). (2) In cold uniformly rotating NS models, approximate turning points occur where one out of three of d/dρb, max(Mg, Mb, J) vanish. The study of such models shows that the actual turning point density is within only ∼1% of the approximate turning point density (where dMg/dρb, max = 0 along the mass-shed sequence; see Figure 10 of Stergioulas & Friedman 1995). (3) The turning-point condition is a sufficient, but not necessary, criterion for secular instability. Thus instability must set in at ρb, max greater than the turning-point ρb, max, but may set in already at lower densities (see, e.g., Takami et al. 2011 for an example). It is thus conservative to use the approximate turning point located at the highest ρb, max over all sequences for a given EOS as an upper bound for the maximum stable ρb, max of HMNS models for that EOS.

Further to the above, we have verified (see Section 9.1 of Kaplan 2014) that the same density ranges contain approximate turning points when examining alternate pairs of conserved variables: both J and Mb, and J and Mg (in contrast to Figure 11, where we examine Mb and Mg). This gives us confidence that the method of approximate turning points is self consistent with respect to choice of the vanishing derivatives. Unfortunately, since the CST code employs only barytropic EOSs, we lack the infrastructure necessary to study the total entropy of the configurations, and note that an examination of total entropy of these models is an important goal for future work.

5.2. The Secular Evolution of HMNS from Mergers

An HMNS remnant resulting from the merger of two NSs that does not promptly collapse into a black hole will settle into a quasiequilibrium state. More precisely, this is a state in which the HMNS is no longer in dynamical evolution, measured, for example, by oscillations in the HMNS maximum density. This should occur several dynamical times after merger. From this point on, the HMNS will evolve secularly along some sequence of equilibrium models. A secular evolution is, by definition, a dissipative process that may involve energy loss11 from the system. Consequently, we may parameterize the secular evolution of the HMNS toward a turning point via the change in its total mass-energy, which, in our case, is the change in gravitational mass of the equilibrium model. This occurs in HMNSs via neutrino cooling and the emission of gravitational radiation. In addition, the rotational energy of the HMNS may be reduced by angular momentum redistribution via the MRI, provided this occurs sufficiently slowly to be characterized as a secular process. This can lead to a build up of magnetic field, or dissipation of the free energy of differential rotation as heat (see, e.g., Thompson et al. 2005 for a detailed discussion), which may lead to increased neutrino cooling. Furthermore, specific angular momentum transported to the HMNS surface may unbind surface material, leading to a decrease in J and Mb. These changes of Mb and J may be significant, but cannot be taken into account by the approximate description of the HMNS's evolution we are considering here. Our results should thus be interpreted with these limitations in mind.

A secularly evolving HMNS will, in general, evolve in the direction of decreasing gravitational mass Mg while (at least approximately) conserving its total baryonic mass Mb. This results in an increasing density and compactness of the star. Figure 11 shows, for a fixed temperature prescription and differential rotation parameter, that the gravitational mass Mg of a sequence with fixed baryonic mass Mb = 2.9 M (using the HShen EOS; we find qualitatively the same for the LS220) is decreasing with increasing density. This continues until, Mg reaches a minimum at an approximate turning point for ρb, max ≳ 1 × 1015 g cm−3. Here, δMg = 0, and δMb vanishes by our choice of a constant Mb sequence.

The curves in Figure 11 are shown for constant differential rotation parameter $\tilde{A}$. However, an HMNS of Mb = 2.9 M is not necessarily constrained to a specific curve. One would expect the HMNS to evolve to neighboring curves of less extreme differential rotation (decreasing $\tilde{A}$), in accordance with its loss of angular momentum due to gravitational waves and its redistribution of angular momentum due to other secular processes. Nevertheless, consider the limit in which the HMNS is constrained to a curve of constant $\tilde{A}$. Then it would evolve secularly until reaching the curve's minimum. At this point, any further energy loss implies that the HMNS must either (1) secularly evolve to a nearby equilibrium sequence with lower temperature or lower degree of differential rotation and higher density (another curve on the plot) or (2) undergo collapse to a black hole. Note that the densities at which the minimum occurs for different $\tilde{A}$ and temperatures are remarkably close to each other. For the sequences using the HShen EOS shown in Figure 11, the approximate turning points lie in the range 1.05 × 1015 g cm−3 < ρb, max < 1.30 × 1015 g cm−3 for all considered $\tilde{A}$ and both shown temperature prescriptions. The constant-Mb curves for other temperature parameterizations (c20p0, c30p0, c30p5, c30p10) are all located in-between the curves for the c40p0 and cold cases shown. Thus, we expect that the point of collapse for an HMNS will be marked by its evolution to this density regime regardless of the temperature distribution of the model.

From the above findings, we conclude that thermal effects have little influence on the stability of HMNSs in rotational equilibrium against gravitational collapse. However, our results do imply that thermal support will affect at what density the HMNS first settles to its quasiequilibrium state. The discussion in Section 4.2 and, in particular, Figure 10, illustrates that at subcritical rotation rates and densities significantly below those of the approximate turning points, models with hot temperature profiles have a larger Mb compared to models with cooler temperatures at the same ρb, max. Thus an HMNS with greater thermal support will reach a quasiequilibrium at a lower ρb, max, and thus have more energy to lose before it can evolve to the critical density regime for collapse.

While thermal effects may be important in setting the initial conditions for the secular evolution of an HMNS, they appear to be of little consequence to the stability of an HMNS in quasiequilibrium. Once in a quasiequilibrium state, the energy lost by an HMNS during its secular evolution is the most robust indicator for its progress toward instability and collapse. Figure 11 shows that this is true regardless of the degree of differential rotation of the HMNS. For a fixed temperature parameterization, the difference in Mg between different degrees of differential rotation is at most ∼0.005 M, corresponding to ≲ 10% of the total energy lost during the HMNS's secular evolution.

5.3. Comparison with NSNS Merger Simulations

Sekiguchi et al. (2011) conducted simulations of NSNS mergers using the HShen EOS and included neutrino cooling via an approximate leakage scheme. They considered three equal-mass binaries with component NS gravitational (baryonic) masses of 1.35 M (1.45 M), 1.50 M (1.64 M), 1.60 M (1.77 M) denoted as L, M, and H, respectively. The HMNS formed from their high-mass binary collapses to a black hole within ≲ 9 ms of merger. The low-mass and the intermediate-mass binaries, however, form hot (T ∼ 5–30 MeV) spheroidal quasiequilibrium HMNSs that remain stable for at least 25 ms, the duration of their postmerger simulations.

Sekiguchi et al. (2011) argue that thermal pressure support could increase the maximum mass of HMNSs with T ≳ 20 MeV by 20%–30%. The results that we lay out in Sections 3 and 4 of our study suggest that it is not straightforward to disentangle centrifugal and thermal effects for differentially rotating HMNS. Our findings show that critically spinning configurations (i.e., configurations at which the maximum Mb is obtained for a given $\tilde{A}$) of hot models do not lead to an increase in the maximum supported baryonic mass by more than a few percent and in most cases predict a lower maximum mass than in the cold case. We find it more useful to consider the results of Sekiguchi et al. (2011) in the context of the evolutionary scenario outlined in Section 5.2.

In Figure 12, we plot Mb as a function of ρb, max for select sequences of uniformly and differentially rotating models obtained with the HShen EOS with the cold and c40p0 temperature prescriptions. We also mark the immediate postmerger densities of the L, M, and H models of Sekiguchi et al. (2011) and their evolutionary tracks (in ρb, max). The high-mass model H never settles into a quasiequilibrium and collapses to a black hole during the dynamical early postmerger phase. Its ρb, max evolves within ∼9 ms from 0.58 × 1015 g cm−3 to values beyond the range of the plot. Our secular-evolution approach cannot be applied to this model since it never reaches a quasiequilibrium state. The lower-mass M and L models enter Figure 12 at successively lower densities. Their "ring-down" oscillations are damped by ∼9 ms after which the HMNSs evolve secularly with ρb, max that increase roughly at the same rate in both models, suggesting that their rate of energy loss is comparable. At such early times, gravitational waves are most likely dominating energy loss (see the discussion of timescales in Paschalidis et al. 2012), and, indeed, model M and L exhibit similar gravitational wave amplitudes and frequencies (Sekiguchi et al. 2011, Figure 4). Focusing on model L, we now consider Figure 11, which shows sequences of constant Mb (for model L with Mb ∼ 2.9 M). As the HMNS loses energy, Mg decreases and the HMNS evolves to the right (toward higher ρb, max). Model L enters its secular evolution at a central density of ∼0.56 × 1015 g cm−3 and evolves secularly to ∼0.68 × 1015 g cm−3 within ∼16 ms. Largely independent of its specific angular momentum distribution and thermal structure, Figure 11 suggests that this model will reach its global minimum Mg and, thus, instability in a small density range of ∼1.05–1.30 × 1015 g cm−3.

Figure 12.

Figure 12. Similar to Figure 7 but for the HShen EOS and showing the approximate evolution of HMNSs from Sekiguchi et al. (2011). We show the evolution of maximum density of the HMNS for the low, medium and high mass configurations (thick lines L, M and H) starting from the premerger density (noted by circles), and ending at the simulation termination densities (squares, or, in the H configuration, an arrow indicating collapse to a black hole). After ∼9 ms (noted with diamonds), the L and M models show negligible dynamical oscillations and have settled to a quasiequilibrium state. From there until the end of the simulation, the L and M HMNS are evolving secularly (indicated by thick dotted lines). We note that given the limitations of our approach discussed in the main text, the evolutionary tracks of constant baryonic mass shown in this figure should not be considered quantitatively reliable.

Standard image High-resolution image

Using our approximate secular evolution model for HMNSs discussed in Section 5.2, we linearly extrapolate the density evolution of model L in Sekiguchi et al. (2011). We expect a possible onset of collapse at t ≳ 58 ms after merger (and ≳49 ms after the start of the secular evolution). These numbers should be regarded as very rough estimates, given the limitations and rather qualitative nature of our model. Depending on its angular momentum when entering its secular evolution, its cooling rate, angular momentum redistribution and loss, model L may alternatively evolve into a long-term stable supramassive NS, since a baryonic mass of ∼2.9 M can in principle be supported by the HShen EOS at the supramassive limit (see Table 2). Furthermore, we have also checked that model L of Sekiguchi et al. (2011) contains sufficient angular momentum to be represented by the sequences identified in Figures 11 and 12. At a time of ∼10–15 ms after merger, model L has an angular momentum of 6 × 1049 g cm2 s−1(6.8 in $c = G = M_{_\odot }$ units). Plots of similar sequences can be found in Kaplan (2014). They are consistent with this value.

The role of thermal pressure effects in all of the above is relatively minor (see the very similar ρb, max locations of the Mg minima in hot and cold configurations shown in Figure 11). However, when first entering the secular regime as a subcritical HMNS, a configuration with higher temperature and stronger thermal pressure support will be less compact and will have a lower ρb, max at a fixed Mb than a colder one. Hence, in the picture of secular HMNS evolution discussed in Section 5.2, such a configuration would have to evolve "farther" in ρb, max to reach criticality and, thus, can survive longer at fixed energy loss rates.

Paschalidis et al. (2012) performed NSNS merger simulations of Γ = 2 polytropes in which they approximated a thermal pressure component with a Γ = 2 Γ-law. Their postmerger HMNS enters its secular evolution in a quasitoroidal configuration with two high-density, low-entropy cores, a central, lower-density, hot region and a high-entropy low-density envelope. The total mass of their model can be arbitrarily rescaled, but in order to estimate temperatures and thermal pressure contributions, the authors scaled their HMNS remnant to a gravitational mass of 2.69 M. With this, they estimated in their quasitoroidal HMNS peak and rms temperatures of ∼20 MeV and ∼5 MeV, respectively. Paschalidis et al. (2012) studied the effect of neutrino cooling on the HMNS evolution by introducing an ad-hoc cooling function that removes energy proportional to the thermal internal energy (neglecting the stiff temperature dependence of neutrino cooling). In order to capture effects of cooling during the limited simulated physical postmerger time, they drained energy from their HMNS at rates ∼100–200 times higher than realistic cooling by neutrinos.

The authors considered cases without cooling and with two different accelerated cooling timescales. As cooling is turned on in their simulations, the slope of the maximum baryon density ρb, max(t) of the HMNS increases discontinuously and the higher the cooling rate, the faster the evolution to higher ρb, max(t). The HMNSs in both cases with cooling become unstable at different times, but roughly at the same ρb, max. This is consistent with the secular HMNS evolution picture laid out in Section 5.2. Cooling reduces the total energy of the system (Mg) and drives the HMNS to higher ρb, max at fixed Mb until the (approximate) turning point is reached and collapse ensues. However, losses due to gravitational wave emission and angular momentum redistribution and shedding will have the same effect and may dominate in nature, since they are likely to operate more rapidly than neutrino cooling (see the discussion of timescales by Paschalidis et al. 2012).

Bauswein et al. (2010) carried out smoothed-particle hydrodynamics simulations of HMNSs in the conformal-flatness approximation to general relativity. They compared simulations using the full temperature dependence of the HShen and LS180 EOS12 with an approximate treatment of thermal pressure via a Γ-law, Pth = (Γth − 1)epsilonthρb. Although Bauswein et al. (2010) do not provide a figure showing the evolution of maximum baryon density, they show (in their Figure 5) graphs of cumulative mass as a function of distance from the center of the LS180-EOS HMNS at 8 ms after merger, roughly the time when the dynamical early postmerger phase is over and the secular HMNS evolution begins. From this, it may be observed that the HMNS with the lower thermal Γ (Γth = 1.5) is more compact than the model with Γth = 2. The HMNS evolved with the fully temperature-dependent LS180 EOS is in between the two, but closer to the Γth = 2 model. Bauswein et al. (2010) found that the more compact HMNS with Γth = 1.5 collapses after 10 ms, while the less compact Γth = 2.0 and full-LS180 cases collapse after ∼20 ms. This is consistent with the picture of secular HMNS evolution drawn in Section 5.2. Given a fixed number of baryons, a less compact configuration has a lower maximum baryon density after merger and, therefore, begins its secular evolution (in the sense of Figures 11 and 12) at a lower density than a more compact configuration. Consequently, it must lose more energy before reaching the critical density for collapse.

The above illustrates how thermal pressure effects may increase the lifetime of an HMNS by affecting the initial conditions for its secular evolution. From Section 4 one notes that hot configurations, at densities below ≲ 1015 g cm−3 (the exact value being EOS dependent), may support significantly larger masses than their cold counterparts at the same ρb, max. Thus, during the dynamical settle-down of two merging NSs to a secularly evolving HMNS remnant, a configuration with lower thermal pressure will need to evolve to higher ρb, max to reach an equilibrium configuration.

6. SUMMARY AND CONCLUSIONS

The merger of double NSs with component masses in the most commonly observed mass range (∼1.3–1.4 M; Lattimer 2012) is most likely to result in a hot, differentially spinning HMNS remnant that is stable against collapse on a dynamical timescale, but likely secularly evolving toward instability, driven by energy loss. While a number of merger simulations in approximate or full general-relativity with the necessary microphysics are now available, the role of thermal pressure support on the postmerger HMNS and its stability is not well understood.

In this study, we have attempted to gain insight into the role of thermal pressure support by constructing nonrotating, uniformly rotating and differentially rotating axisymmetric equilibrium solutions with multiple microphysical, fully temperature and composition dependent EOS and parameterized temperature distributions motivated by results from full merger simulations. Such axisymmetric equilibrium models may be acceptable approximations to merger remnants that have survived the initial highly dynamical and strongly nonaxisymmetric postmerger evolution and have settled down into longer-term stable quasiequilibrium. How far away the equilibrium configurations really are from real HMNSs, and the reliability of our results, will ultimately have to be established by more detailed comparisons with merger simulations in future work.

In the secular postmerger phase, the baryonic mass Mb of the hypermassive merger remnant is approximately conserved. Thus the dependence of the maximum of Mb on temperature is the most interesting quantity to study. In spherical symmetry (the TOV case), we find that at densities significantly lower than the density at which the maximum mass configuration occurs, thermal enhancement of the NS mass can be strong. Generally, hotter configurations yield the same Mb at lower central densities than their colder counterparts. However, when considering compact maximum-Mb configurations, thermal effects are small. For reasonable temperature prescriptions, hot temperatures lead to a small (≲ 1%) decrease of $M_\mathrm{b}^\mathrm{max}$ for five out of the seven EOS that we consider. The two other EOS, the HShen EOS and the GShen-FSU2.1 EOS, show up to ∼2% thermal enhancement of Mb. As expected, none of the considered EOS could support a remnant of the merger of a canonical double NS system with typical masses.

Rapidly uniformly spinning configurations can support supramassive NSs. We have studied uniformly spinning sequences generated with the LS220 and HShen EOS. As in the TOV case, we find significant thermal enhancement of Mb at low central densities and rotation rates up to mass shedding. At high densities, however, thermal pressure is much less important for the support of the inner NS core, but bloats the envelope. This results in hotter configurations reaching mass shedding at lower angular velocities than colder configurations. Hence, at the mass-shedding supramassive limit, Mb and Mg decrease with increasing temperature for uniformly spinning NSs. For the LS220 EOS (HShen EOS), the cold supramassive Mb limit is ∼2.823 M (∼3.046 M). Under the plausible assumption that the HMNS merger remnant evolves toward a uniformly rotating configuration, assuming no mass loss during or after merger, the cold supramassive limit corresponds to component gravitational masses in an equal-mass progenitor binary of Mg ∼ 1.287 M (Mg ∼ 1.403 M). On the other hand, a supramassive LS220 (HShen) NS with a 30 MeV core and a 10 MeV envelope has a supramassive limit Mb ∼ 2.587 M (Mb ∼ 2.808 M), which corresponds to binary component Mg ∼ 1.185 M (Mb ∼ 1.300 M). Hence, cold maximally uniformly rotating configurations of LS220 and HShen NSs may barely support the merger remnant of canonical double NS binaries, but hot ones might not.

Differential rotation adds yet another layer of complexity, but is the most interesting scenario, since hypermassive merger remnants are born with differential rotation. The notion of a maximum mass of a differentially rotating HMNS is somewhat misleading, since different rotation laws will give different masses and different solvers may converge to different branches in the solution space. Hence, all "maximum" masses quoted are lower limits. For the commonly used j − const. rotation-law, parameterized by the dimensionless parameter $\tilde{A}$, we find Mb up to ∼3.65 M and ∼4.10 M, for the LS220 EOS and the HShen EOS, respectively. These high-mass configurations generally occur at densities that are up to a factor of two lower than those of maximum-Mb TOV and uniformly rotating models. Even higher masses could be found, but such configurations would be dynamically nonaxisymmetrically unstable.

Our results indicate that the role of thermal effects depends very much on the degree of differential rotation in addition to maximum density and (central) angular velocity. All qualitative findings are identical for the LS220 EOS and the HShen EOS. For critically rotating models (with minimum axis ratio rp/e for which a solution is found) the dependence on differential rotation is as follows. (1) For a low degree of differential rotation ($\tilde{A} \lesssim 0.4$), the same systematics as found for the uniformly rotating case hold. (2) In models with intermediate degree of differential rotation ($\tilde{A} \sim 0.5\hbox{--}0.7$), hot configurations have systematically lower "maximum" Mb than colder ones. (3) Models with high degree of differential rotation ($\tilde{A} \gtrsim 0.7$) are mostly quasitoroidal and the "maximum" Mb occurs at low densities (≲ 5 × 1014 g cm−3) and is mildly enhanced by thermal pressure support for models with hot cores, but cold envelopes. Models with high-temperature envelopes remain spheroidal until higher densities and have lower "maximum" Mb. The situation is yet different for differentially rotating configurations that are rotating rapidly, but subcritically. For example, for LS220 EOS configurations with $\tilde{A} = 1$, models with thermally supported envelopes have the highest Mb at subcritical rotation, but their sequences terminate at lower angular velocities (higher rp/e) than the cold configuration, which ultimately catches up in Mb at critical rotation.

To summarize all of the above: the forecast is mixed—the role of thermal effects on the baryonic mass that is supported by a given configuration depends sensitively and in a complicated way on its details, that is, central/mean baryon density, temperature distribution, degree of differential rotation and rotation rate, to name the most important parameters. Configurations that yield "maximum" Mb are essentially unaffected by thermal effects. Beyond that, no simple general statements can be made.

A more useful way to reason about the role of thermal pressure support is to consider evolutionary sequences of equilibrium models representing the secular quasiequilibrium evolution of an HMNS. This evolution occurs along tracks of approximately constant baryonic mass Mb parameterized by maximum baryon density ρb, max. Since energy is lost by gravitational wave and neutrino emission, a configuration always evolves into the direction of decreasing total energy (i.e., decreasing gravitational mass Mg and increasing ρb, max). The turning point theorem (Sorkin 1982; Friedman & Stergioulas 2013) says that an extremum in Mg may mark the point at which the sequence becomes secularly unstable to collapse. While this can be proven rigorously only for uniformly rotating (or nonrotating) configurations, we conjecture that it also holds at least approximately for the much more complex HMNS case. Provided this is true, we can define approximate turning points using constant-Mb sequences with different degrees of differential rotation and temperature parameterizations. With this, we find that the approximate turning points for a given Mb always lie in narrow ranges of ρb, max and Mg, which define the Mg–ρb, max space in which collapse to a black hole occurs. Furthermore, the approximate turning point density at which collapse must set in depends only very weakly on temperature. Finally, we note that all approximate turning points found in this work are at baryon densities below the critical value for stable TOV stars. This may suggest that HMNS with maximum densities at or higher than the critical TOV central density could always be unstable to collapse. This possibility should be investigated further in future work.

Under the assumptions of the model laid out in this paper, the secular evolution of an HMNS can then be described by the progressive decrease of its gravitational mass Mg and increase of its maximum density ρb, max. Our results show that an HMNS with more thermal pressure support will enter its secular evolution at a higher Mg and lower ρb, max than a colder one (with the same rotational setup). Hence, the hot HMNS will have to evolve further in ρb, max until reaching its approximate turning point. This explains the effects of thermal pressure observed in merger simulations (e.g., Bauswein et al. 2010; Sekiguchi et al. 2011). We note that the same argument may also be applied to differences in HMNS spin: a more rapidly spinning HMNS will enter its secular evolution at lower ρb, max and higher total energy and, hence, will have to evolve further in ρb, max to reach its approximate turning point.

The goal of the work presented in this paper was to elucidate the role of thermal pressure support in hypermassive NSNS merger remnants on the basis of stationary spherically symmetric and axisymmetric equilibrium solutions of the Einstein–Euler equations. While yielding new insights, our present approach is limited in multiple ways. (1) Even in the secular quasiequilibrium evolution phase, HMNS are not exactly axisymmetric. The CST solver used in this study does not support nonaxisymmetric configurations, which makes it impossible for us to test how sensitive our results are to symmetry assumptions. (2) The equilibrium sequences considered here rely on an ad-hoc rotation law and ad-hoc temperature and composition parameterizations motivated by the simulations of Sekiguchi et al. (2011). In general, the angular velocity distribution will be more complex (see, e.g., Galeazzi et al. 2012) and the temperature and composition of an HMNS will not be single-parameter functions of density. (3) The CST solver has difficulties converging for configurations with a high degree of differential rotation and it is not clear if the terminating axis ratio rp/e is set by the formulation and implementation of the equations by the CST solver or if the termination occurs for physical reasons. This could be checked only by a comparison study with a more robust solver, e.g., the one of Ansorg et al. (2003). (4) The approximate turning point theorem that we have used to reason about the evolution and stability of HMNSs is heuristic and lacks rigorous foundation. Fully reliable statements about the stability of differentially rotating HMNSs with complex temperature and compositional distributions will require at least perturbative stability analysis or direct non-linear simulation.

Future work should address the above limitations (1)–(4) and should also consider rotating configurations constructed with a broader set of finite-temperature microphysical EOS.

We thank Eliot Quataert for inspiration and acknowledge helpful discussions with Lars Bildsten, Ursula C. T. Gamma, Jim Lattimer, Lee Lindblom, Sterl Phinney, Jocelyn Read, Yuichiro Sekiguchi, Masaru Shibata, Saul Teukolsky, Kip Thorne, and especially Aaron Zimmerman. Furthermore, we thank the anonymous referee for suggestions that improved this paper. This work was initiated at a Palomar Transient Factory Theory Network meeting at the Sky House, Los Osos, CA. C.D.O. thanks the Yukawa Institute for Theoretical Physics for hospitality during the long-term workshop Gravitational Waves and Numerical Relativity 2013 when this work was completed. This research is supported in part by NASA under the Astrophysics Theory grant No. NNX11AC37G, by the National Science Foundation under grant Nos. AST-1205732, PHY-1151197, AST-1212170, PHY-1068881, and PHY-1068243, by the Alfred P. Sloan Foundation and by the Sherman Fairchild Foundation. The calculations underlying the results presented in this paper were performed on the Caltech compute cluster "Zwicky" (NSF MRI award No. PHY-0960291). The EOS tables, driver, and TOV solver routines used in this work are available for download at http://www.stellarcollapse.org. The solver for axisymmetric equilibrium configurations of rotating HMNSs is not open source, but a similar solver may be obtained from http://www.lorene.obspm.fr/. The figures in this paper were generated with the matplotlib library for Python (Hunter 2007). This work was supported by Grant-in-Aid for Scientific Research (25103510, 25105508, 24740163), by HPCI Strategic Program of Japanese MEXT. The simulations were performed on XC30 at CfCA of NAOJ and SR16000 at YITP of Kyoto University.

APPENDIX A: TEMPERATURE PARAMETERIZATIONS

We consider temperature prescriptions with only a hot core at and above nuclear density and with a hot core and a more extended high-density plateau at lower densities. We emphasize that these prescriptions are rather ad-hoc and motivated primarily by the data from the simulations of Sekiguchi et al. (2011). All high-temperature regions are smoothly tapered-off ("rolled-off") using tanh functions.

The prescriptions with only a hot core (i.e., prescriptions cXp0) are given by

Equation (A1)

where m is the roll-off midpoint (in log10b(g cm−3)) and s is the roll-off e-folding scale (also in log10b(g cm−3)). For prescriptions that only have hot cores, T1 is set to the peak temperature Tmax and T2 is set to Tmin = 0.01 MeV. The prescriptions with a high-temperature plateau at lower densities, i.e., c30p5 and c30p10, are constructed as the sum of two of the above functions as follows:

Equation (A2)

where m' is the roll-off midpoint, s' is the roll-off scale, and Tp is the plateau temperature. Writing this out more explicitly, we have

Equation (A3)

Table 4 summarizes the parameters for generating the temperature prescriptions used in this study.

Table 4. Temperature Prescription Parameters

Model Tmax Midpoint m Scale s Plateau
  (MeV) log10b(g cm−3)) log10b(g cm−3)) Temperature Tp
(MeV)
cold  ⋅⋅⋅   ⋅⋅⋅   ⋅⋅⋅   ⋅⋅⋅ 
c20p0 20 14.0–0.07 $0.25\phantom{25}$ 0
c30p0 30 14.125–0.07 $0.375\phantom{2}$ 0
c30p5 30 14.1875–0.07 0.3125 5
c30p10 30 14.25–0.07 $0.25\phantom{25}$ 10
c40p0 40 14.25–0.07 $0.5\phantom{752}$ 0

Notes. Parameters used for the temperature parameterizations used in this study. The notation is c〈core temperature〉p〈plateau temperature〉. All low-density temperature plateaus are tapered off at densities below ∼1012 g cm−3 with a tanh function with a midpoint at log10b(g cm−3)) = 11.5 and an e-folding width of log10b(g cm−3)) = 0.25. All minimum temperatures are 0.01 MeV. See Figure 2 for a comparison of the various temperature prescriptions. The functional form of the prescriptions is given by Equations (A1) and (A3).

Download table as:  ASCIITypeset image

APPENDIX B: SOLVING FOR THE ELECTRON FRACTION

For a given EOS and temperature prescription, we find the electron fraction Ye by first solving for Ye assuming neutrino-less β-equilibrium for the cold case (T = 0.01 MeV or the lowest temperature point available in the EOS table), using the condition

Equation (B1)

for the chemical potentials. In the absence of neutrinos, the lepton fraction Ylep = Ye. In the hot case, neutrinos are trapped in the HMNS matter above ρ = ρtrap ≈ 1012.5 g cm−3 and Ylep = Ye + Yν, where $Y_\nu = Y_{\nu _e} - Y_{\bar{\nu }_e}$.

We then take Ylep and solve for Ye in the hot case with neutrinos by treating the latter as a relativistic Fermi gas in equilibrium for which Yν can be calculated from the neutrino number density $n_\nu = n_{\nu _e} - n_{\bar{\nu }_e}$ via

Equation (B2)

The neutrino number density is

Equation (B3)

where ην = μν/(kBT) is the neutrino degeneracy parameter (Bludman & van Riper 1978). Note that in equilibrium, νe and $\bar{\nu }_e$ have equal and opposite chemical potentials. F2 is a Fermi integral given by

Equation (B4)

In practice, we use

Equation (B5)

which is given in Bludman & van Riper (1978) and is exact for any degeneracy parameter η.

We find Ye by finding the root

Equation (B6)

Ylep is a fixed input. We set Ye = Ylep as an initial guess and Yν is calculated using Equations (B2), (B3), and (B5), with μν = μn + μp − μe obtained from the EOS. Ye is then adjusted and we iterate until convergence.

Since neutrinos begin to stream freely below ρtrap, we also compute Ye using the ν-less β-equilibrium condition (Equation (B1)). We then compute a final effective Ye using

Equation (B7)

Footnotes

  • Available from http://phys-merger.physik.unibas.ch/~hempel/eos.html, based on Hempel et al. (2012) and Hempel & Schaffner-Bielich (2010).

  • Note that $\tilde{A} = 1/\hat{A}$, where $\hat{A}$ is the same $\hat{A}$ as used in Baumgarte et al. (2000).

  • See, e.g., Hempel et al. (2012) for a discussion of EOS physics and temperature dependence of various EOS models.

  • 10 

    We show only the c40p0 and cold temperature parameterizations in Figure 11, because we find them to be the limiting cases. All other parameterizations have minima at intermediate locations in the (Mg, ρb, max) plane.

  • 11 

    The trapped lepton number is, of course, also changing, since the fluxes of νe and $\bar{\nu }_e$ will at least initially not be symmetric. However the effect of the trapped lepton fraction on stability is minimal, since electron degeneracy pressure is present only at high densities where it is much smaller than the baryon pressure in hot HMNSs that lose energy to neutrino emission (see Figure 1).

  • 12 

    The LS180 is the variant of the Lattimer & Swesty (1991) EOS with nuclear compressibility modulus K0 = 180 MeV.

Please wait… references are loading.
10.1088/0004-637X/790/1/19