A publishing partnership

Letters

MAGNETIC ENERGY PRODUCTION BY TURBULENCE IN BINARY NEUTRON STAR MERGERS

and

Published 2013 May 15 © 2013. The American Astronomical Society. All rights reserved.
, , Citation Jonathan Zrake and Andrew I. MacFadyen 2013 ApJL 769 L29 DOI 10.1088/2041-8205/769/2/L29

2041-8205/769/2/L29

ABSTRACT

The simultaneous detection of electromagnetic and gravitational wave emission from merging neutron star binaries would greatly aid in their discovery and interpretation. By studying turbulent amplification of magnetic fields in local high-resolution simulations of neutron star merger conditions, we demonstrate that magnetar-level (≳ 1016 G) fields are present throughout the merger duration. We find that the small-scale turbulent dynamo converts 60% of the randomized kinetic energy into magnetic fields on a merger timescale. Since turbulent magnetic energy dissipates through reconnection events that accelerate relativistic electrons, turbulence may facilitate the conversion of orbital kinetic energy into radiation. If 10−4 of the ∼1053 erg of orbital kinetic available gets processed through reconnection and creates radiation in the 15–150 keV band, then the fluence at 200 Mpc would be 10−7 erg cm−2, potentially rendering most merging neutron stars in the advanced LIGO and Virgo detection volumes detectable by Swift BAT.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

The inspiral and coalescence of binary neutron star (BNS) systems is a topic of increasingly intensive research in observational and theoretical astrophysics. It is anticipated that the first direct detections of gravitational wave (GW) will be from compact binary mergers. BNS mergers are also thought to produce short-hard gamma-ray bursts (SGRBs; Eichler et al. 1989; Narayan et al. 1992; Belczynski et al. 2006; Metzger et al. 2008). Simultaneous detections of a prompt GW signal with a spatially coincident electromagnetic (EM) counterpart dramatically increase the potential science return of the discovery. For this reason, there has been considerable interest as to which, if any, detectable EM signature may result from the merger (Metzger & Berger 2012; Piran et al. 2013). Other than SGRBs and their afterglows, including those viewed off-axis (Rhoads 1997; van Eerten & MacFadyen 2011), suggestions include optical afterglows associated with the radioactive decay of tidally expelled r-process material (Li & Paczyński 1998; Metzger et al. 2010; though detailed calculations indicate that they are faint (Barnes & Kasen 2013)), radio afterglows following the interaction of a mildly relativistic shell with the interstellar medium (Nakar & Piran 2011), and high-energy pre-merger emission from resistive magnetosphere interactions (Hansen & Lyutikov 2001; Piro 2012).

Merging neutron stars possess abundant orbital kinetic energy (∼1053 erg). A fraction of this energy is certain to be channeled through a turbulent cascade triggered by hydrodynamical instabilities during the merger. Turbulence is known to amplify magnetic fields by stretching and folding embedded field lines in a process known as the small-scale turbulent dynamo (Vaĭnshteĭn & Zel'dovich 1972; Chertkov et al. 1999; Tobias et al. 2011). Amplification stops when the magnetic energy grows to equipartition with the energy containing turbulent eddies. An order of magnitude estimate of the magnetic energy available at saturation of the dynamo can be informed by global merger simulations. These studies indicate the presence of turbulence following the nonlinear saturation of the Kelvin–Helmholtz instability activated by shearing at the NS surface layers (Price & Rosswog 2006; Liu et al. 2008; Anderson et al. 2008; Rezzolla et al. 2011; Giacomazzo et al. 2011). The largest eddies produced are on the ∼1 km scale and rotate at ∼0.1c, setting the cascade time teddy and kinetic energy injection rate εKEK/teddy at 0.1 ms and 5 × 1050 erg s−1, respectively. When kinetic equipartition is reached, each turbulent eddy contains 5 × 1046 erg of magnetic energy, and a mean magnetic field strength

Equation (1)

Whether such conditions are realized in merging neutron star systems depends upon the dynamo saturation time tsat and equipartition level Esat/EK. In particular, if tsattmerge then turbulent volumes of neutron star material will contain magnetar-level fields throughout the early merger phase. Once saturation is reached, a substantial fraction of the injected kinetic energy, 0.7εK, is resistively dissipated (Haugen et al. 2004) at small scales. Magnetic energy dissipated by reconnection in optically thin surface layers will accelerate relativistic electrons (Blandford & Eichler 1987; Lyutikov & Uzdensky 2003), potentially yielding an observable EM counterpart, independently of whether the merger eventually forms a relativistic outflow capable of powering a short gamma-ray burst (GRB).

In this Letter, we demonstrate that the small-scale turbulent dynamo saturates quickly, on a time tsattmerge, and that ≳ 1016 G magnetic fields are present throughout the early merger phase. This implies that the magnetic energy budget of merging BNSs is controlled by the rate with which hydrodynamical instabilities randomize the orbital kinetic energy. Our results are derived from simulations of the small-scale turbulent dynamo operating in the high-density, trans-relativistic, and highly conductive material present in merging neutron stars. We have carefully examined the approach to numerical convergence and report grid resolution criteria sufficient to resolve aspects of the small-scale dynamo. Our Letter is organized as follows. The numerical setup is briefly described in Section 2. Section 3 reports the resolution criterion for numerical convergence of the dynamo completion time and the saturated field strength. In Section 4, we assess the possibility that magnetic reconnection events may convert a sufficiently large fraction of the magnetic energy into high-energy photons to yield a prompt EM counterpart detectable by high-energy observatories including Swift and Fermi.

2. NUMERICAL SCHEME AND PHYSICAL SETUP

The equations of ideal relativistic magnetohydrodynamics (MHD) have been solved on the periodic unit cube with resolutions between 163 and 10243:

Equation (2a)

Equation (2b)

Equation (2c)

Here, $T^{\mu \nu} = ph^{*} u^{\mu } u^{\nu } + p^{*} g^{\mu \nu } - b^{\mu } b^{\nu }, b^{\mu } = F^{\mu }_{\nu } u^{\nu }$ is the magnetic field four-vector, and h* = 1 + e* + p*/ρ is the total specific enthalpy, where p* = pg + b2/2 is the total pressure, pg is the gas pressure, and e* = e + b2/2ρ is the specific internal energy. The source term Sμ = ρaμ − ρ(T/T0)4uμ includes injection of energy and momentum at the large scales and the subtraction of internal energy (with parameter T0 = 40 MeV) to permit stationary evolution. Vortical modes at k/2π ⩽ 3 are forced by the four-acceleration field aμ = duμ/dτ which smoothly decorrelates over a large-eddy turnover time, as described in Zrake & MacFadyen (2012, 2013).

We have employed a realistic micro-physical equation of state (EOS) appropriate for the conditions of merging neutron stars. It includes contributions from high-density nucleons according to a relativistic mean field model (Shen et al. 1998, 2010), a relativistic and degenerate electron–positron component, neutrino and anti-neutrino pairs in beta equilibrium with the nucleons, and radiation pressure. For our conditions, all the components make comparable contributions to the pressure. We have also employed a simpler gamma-law EOS and found close agreement for the conditions explored in this Letter, indicating that the MHD effects are insensitive to the EOS for transonic conditions. The models presented in our resolution study use the far less expensive gamma-law EOS.

All of the simulations presented in this study use the HLLD approximate Riemann solver (Mignone et al. 2009), which has been demonstrated as crucial in providing the correct spatial distribution of magnetic energy in MHD turbulence (Beckwith & Stone 2011). The solution is advanced with an unsplit, second-order MUSCL-Hancock scheme. Spatial reconstruction is accomplished with the piecewise linear method configured to yield the smallest possible degree of numerical dissipation. The divergence constraint on the magnetic field is maintained to machine precision at cell corners using the finite volume CT method of Tóth (2002). Full details of the numerical scheme may be found in Zrake & MacFadyen (2012).

3. RESULTS

3.1. Growth and Saturation of the Magnetic Energy

Magnetic fields are amplified in our simulations by the small-scale turbulent dynamo. Turbulent fluid motions stretch and fold the magnetic field lines, causing exponential growth of magnetic energy (e.g., Moffatt 1978). This growth is attributed to the advection and diffusion of B through the MHD induction equation (Equation (2c)). When the magnetic field is weak (EM ≪ EK), B evolves passively and the turbulence is hydrodynamical. This limit is referred to as small-scale kinematic turbulent dynamo, and is well described by Kazantzev's model (Kazantsev 1968). This model predicts that the power spectrum of magnetic energy peaks at the resistive scale ℓη and obeys a power law k3/2 at longer wavelengths. The kinematic phase ends when the magnetic field acquires sufficient tension to modify the hydrodynamic motions, after which a dynamical balance between kinetic and magnetic energy is established.

Numerical simulations of MHD turbulence are typically limited to magnetic Prandtl numbers Pm ≡ ν/η ≈ 1. However, neutron star material is characterized by Pm ≫ 1, with the viscous cutoff due to neutrino diffusion occurring at around 10 cm, while the resistive scale is significantly smaller (Thompson & Duncan 1993). However, the disparity between true and simulated magnetic Prandtl number does not influence our conclusions, since dynamos are generically easier to establish in the high Pm regime than the small (Haugen et al. 2004; Ponty & Plunian 2011).

We use an initially uniform, pulsar-level (1011 G) seed magnetic field. This field is sub-dominant to the kinetic energy by 10 orders of magnitude, so that the initial field amplification is expected to be well described by Kazantsev's model. Indeed, we find that during this phase the power spectrum of magnetic energy follows k3/2 (Figure 4), peaking at around 10 grid zones, which we identify as the effective scale of resistivity. The saturation process begins at ever-earlier times with increasing numerical resolution. This reflects the fact that during the kinematic phase, magnetic energy exponentiates on a timescale controlled by shearing at the smallest scales. In numerically converged runs, full saturation occurs with EM ≈ 0.6 EK and is characterized by scale-by-scale super-equipartition, with EM ≈ 4 EK at all but the largest scale.

3.2. Numerical Convergence

The same driven turbulence model was run through magnetic saturation at resolutions 163, 323, 643, 1283, 2563, and 5123. Another model at 10243 was run through the end of the kinematic phase, but further evolution was computationally prohibitive. Figure 1 shows the development of BRMS, EM, and EK as a function of time at each resolution.

Figure 1.

Figure 1. Time development of volume-averaged kinetic and magnetic energies at resolutions between 163 and 10243. Lower resolutions are shown in red and graduate to black with higher resolution. Top: the root-mean-square magnetic field strength in units of 1016 G. When a turbulent volume is resolved by 163 zones, the small-scale dynamo proceeds so slowly that almost no amplification is observed in the first 1 ms. Middle: the magnetic energy in units of the rest mass ρoc2 shown on logarithmic axes. It is clear that the linear growth rate increases at each resolution. Bottom: the kinetic energy (upper curves) shown against the magnetic energy (lower curves) again in units of ρoc2. For all resolutions, the kinetic energy saturates in less than 1 teddy.

Standard image High-resolution image

We find that sufficiently resolved runs (⩾5123) attain mean magnetic field strengths of 1016 G within two large-eddy rotations. All models with resolutions ⩾323 eventually attain mean fields of ≳ 3 × 1016 G. The saturated field strength increases until resolution 2563. We find that the kinematic growth rate is higher at each higher resolution, while the timescale for the nonlinear saturation converges at 2563 to roughly five large-eddy rotation times, or about 0.5 ms for the physical parameters of BNS mergers.

In order to quantitatively describe the time development of magnetic energy EM(t), we describe it with an empirical model,

Equation (3)

where the six parameters (summarized in Table 1) are obtained by a least-squares optimization. Figure 2 shows the empirical model given in Equation (3) applied to a representative run at 1283. The first phase, t < t1, is a startup transient and lasts until the hydrodynamic cascade is fully developed at t1teddy. The kinematic dynamo phase lasts between t1 and tnl, during which the magnetic energy exponentiates on the timescale τ1. At tnl, the smallest scales reach kinetic equipartition and the growth rate slows. In the final stage, EM asymptotically approaches Esat on the timescale τ2. We define the dynamo completion time tsat as tnl + τ2. Figure 3 shows the best-fit Esat, tsat, and τ1 as a function of the resolution.

Figure 2.

Figure 2. Time history of the magnetic energy for a representative run at 1283, together with the empirical model (Equation (3)) with best-fit parameters. The horizontal dashed line indicates the magnetic energy, Esat, at the dynamo completion. From left to right, the vertical dashed lines mark the end of the startup, kinematic, and saturation phases.

Standard image High-resolution image
Figure 3.

Figure 3. Top: convergence study of the kinematic dynamo growth time τ1 (blue) and the dynamo completion time tsat (green) defined as tnl + τ2. Bottom: convergence study of the best-fit model parameter Esat expressed as the ratio of magnetic to kinetic energy EM/EK. The converged value of the volume-averaged EM/EK ≈ 0.6. Nevertheless, at intermediate wavelengths EM(k)/EK(k) ≈ 4. As shown in Figure 2, the largest scales remain kinetically dominated. This indicates the suppression of coherent magnetic structure formation near the integral scale of turbulence.

Standard image High-resolution image

Table 1. Empirical Model for Magnetic Energy Growth

Fit Parameter Description Numerically Converged Value
τ0 Startup timescale None, artifact
t1 Hydrodynamic cascade fully developed teddy
τ1 Kinematic growth timescale None, ∝N−2/3
tnl End of kinematic phase 2teddy
τ2 Nonlinear saturation timescale tnl + τ2 ≈ 5teddy, 2563
Esat Saturated magnetic energy 0.6 EK, 5123

Notes. Model parameters characterizing the growth of magnetic energy before and during the dynamo saturation. Numerical convergence is attained for the fully saturated magnetic energy Esat and the dynamo completion time tsat defined as tnl + τ2.

Download table as:  ASCIITypeset image

The magnetic energy Esat at dynamo completion shows signs of converging to a value of 0.6 EK by resolution 5123. The timescale τ2 on which the magnetic energy asymptotically approaches Esat is consistently ≈3teddy at different resolutions. The dynamo completion time tsat is numerically converged at ≈5teddy by 2563. The best-fit kinematic growth time follows a power law in the resolution, τ1N−0.6. This is consistent with the value of −2/3 expected if the dynamo time is controlled by shearing at the smallest scale, the cascade is Kolmogorov (i.e., u∝ℓ1/3), and the viscous cutoff ℓν occurs at a fixed number of grid zones. In that case, $\tau _1 \sim t_\nu \propto \ell _\nu ^{2/3} \sim N^{-2/3}$.

3.3. Power Spectrum of Kinetic and Magnetic Energy

The time development of kinetic and magnetic energy power spectra has been studied for a single run with resolution 5123. We present three-dimensional, spherically integrated power spectra with the dimensions of erg cm−3 m defined as

Equation (4a)

Equation (4b)

where the Newtonian versions of kinetic and magnetic energy are appropriate since the conditions are only mildly relativistic. The definitions in Equations (4) satisfy ∫P(k)dk = 〈E〉 for PK and PM. Figure 4 shows the power spectrum of kinetic and magnetic energy at various times throughout the growth and saturation of magnetic field. During the kinematic phase, the kinetic energy has a power spectrum PK(k)∝k−5/3 consistent with the Kolmogorov theory for incompressible hydrodynamical turbulence, while $P_M(k,t) \propto e^{t/\tau _1} k^{3/2}$ consistent with Kazanstev's model. PM(k) maintains the same shape, but exponentiates in amplitude at the timescale τ1 which is controlled by shearing at the resistive scale. According to Kazantsev's model, PM(k) should peak at the resistive scale. This is consistent with the observed peak in the magnetic energy at roughly 10 grid zones, the same scale at which we observe the viscous cutoff. This is also consistent with Pm = 1 expected from the numerical scheme employed.

Figure 4.

Figure 4. Time development of the spectrum of kinetic (left) and magnetic (right) energy, given in erg cm−3 m for resolution 5123. During the startup phase the hydrodynamic cascade is not yet fully developed; this phase lasts for 1teddy. The kinematic phase refers to the time when the hydrodynamic cascade is fully developed, but the magnetic energy is still energetically sub-dominant. During the kinematic phase, the kinetic energy power spectrum is consistent with a Kolmogorov cascade, having a slope k−5/3 over intermediate wavelengths. Meanwhile, the magnetic energy spectrum is consistent with the Kazantsev description, having a positive slope k3/2 over the same intermediate wavelengths. During this phase the magnetic energy exponentiates rapidly, with an e-folding time controlled by shearing at the smallest available scales. After saturation, the magnetic energy spectrum conforms to the kinetic energy spectrum, such that EM(k) ≈ 4 EK(k) over the intermediate wavelengths. The kinetic energy spectrum after the dynamo completion has a slightly shallower slope than during the kinematic phase, but is still consistent with a 5/3 spectral index.

Standard image High-resolution image

When the magnetic energy at the resistive scale surpasses the level of the kinetic energy at that scale, PM(k) changes shape. The equipartition scale ℓK, M moves into the inertial range and migrates to larger scale until full saturation occurs with ℓK, ML/4. The movement of ℓK, M to larger scale is associated with the formation of coherent and dynamically substantial magnetic structures of increasing size. For Kolmogorov turbulence, ℓK, Mt3/2 (Beresnyak et al. 2009; Schleicher et al. 2013). In the fully saturated state, the magnetic field is in scale-by-scale super-kinetic equipartition throughout the inertial range, with PM(k) ≈ 4PK(k). The largest scales remain kinetically dominated so that the numerically converged saturation level is EM ≈ 0.6 EK.

4. DISCUSSION

In this Letter we have determined the timescale and saturation level of the small-scale turbulent dynamo operating in the conditions of BNS mergers. We have presented numerically converged simulations showing that magnetic fields are amplified to the ∼1016 G level within a small fraction of the merger dynamical time (tsattmerge), independently of the seed field strength. If hydrodynamical instabilities create fluctuating velocities on the order of 0.1c as indicated by global simulations, then each 1 km3 turbulent volume dissipates ≳ 1046 erg of magnetic energy per 0.1 ms. If fturb represents the fraction of the merger remnant that contains such turbulence, the magnetic energy dissipated during the merger is at least

Equation (5)

A fraction of that dissipation will occur through magnetic reconnection in optically thin surface layers, supplying relativistic electrons which synchrotron radiate in the merger remnant magnetosphere. If 5% of that magnetic energy dissipation creates radiation in the 15–150 keV band, then the fluence at 200 Mpc would be 10−7 erg cm−2, potentially rendering most merging neutron stars in the advanced LIGO and Virgo detection volumes detectable by Swift BAT. If so, then merging neutron stars are accompanied by a prompt EM counterpart, independently of whether a later merger phase yields a collimated outflow capable of powering a short GRB.

We suggest that merger flares may be present in the current sample of short GRBs and may be roughly isotropic on the sky since they are seen to distances where the cosmological matter distribution becomes homogeneous (e.g., Tanvir et al. 2005). Searches for merger flares should seek to identify short flares, not unlike soft-gamma repeaters, among the short burst population. If mergers also produce short GRBs, then the flaring activity could be detectable as a separate component of the emission, which may appear as a precursor (e.g., Troja et al. 2010) if there is a considerable delay between merger onset and the formation of short GRB engine (e.g., Baiotti et al. 2008). The possibility of prompt flares from BNS mergers has also been suggested to follow from a shock breakout scenario (Kyutoku et al. 2012) or tidal interaction in the late inspiral (Dall'Osso & Rossi 2013).

The presence of strong magnetic fields may also aid in the ejection of neutron-rich material from surface layers of the merger remnant, possibly enhancing the enrichment of interstellar medium by r-process nuclei (Freiburghaus et al. 1999; Rosswog et al. 1999; Hotokezaka et al. 2013). Enhanced production of r-process nuclei also increases the likelihood of EM detection by radioactive decay powered afterglows, or "kilonovae" (Li & Paczyński 1998; Metzger et al. 2010).

Finally, it has been shown that magnetic fields will directly influence the GW signature and remnant disk mass if they exist at the 1017 G level (Etienne et al. 2012). Such strong fields are unlikely on evolutionary grounds in merging BNSs, but our results suggest they may be revived up to the scale of turbulent cells during the merger. To exert influence, such turbulent cells would have to fill a considerable fraction of the merger volume. As we have shown in this Letter, the overall magnetic energy budget is controlled by the prevalence (fturb) and vigor (εK) of the turbulent volumes. This fact motivates the use of higher resolution global simulations aimed at measuring fturb and εK.

This research was supported in part by the NSF through grant AST-1009863 and by NASA through grant NNX10AF62G issued through the Astrophysics Theory Program. Resources supporting this work were provided by the NASA High-End Computing (HEC) Program through the NASA Advanced Supercomputing (NAS) Division at Ames Research Center.

Please wait… references are loading.
10.1088/2041-8205/769/2/L29