Articles

GRAVITATIONAL RADIATION FROM MAGNETICALLY FUNNELED SUPERNOVA FALLBACK ONTO A MAGNETAR

and

Published 2014 October 7 © 2014. The American Astronomical Society. All rights reserved.
, , Citation A. Melatos and M. Priymak 2014 ApJ 794 170 DOI 10.1088/0004-637X/794/2/170

0004-637X/794/2/170

ABSTRACT

Protomagnetars spun up to millisecond rotation periods by supernova fallback are predicted to radiate gravitational waves via hydrodynamic instabilities for ∼102 s before possibly collapsing to form a black hole. It is shown that magnetic funneling of the accretion flow (1) creates a magnetically confined polar mountain, which boosts the gravitational wave signal, and (2) "buries" the magnetic dipole moment, delaying the propeller phase and assisting black hole formation.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Magnetars born rotating with millisecond periods have attracted theoretical attention as the central engines powering long-soft gamma-ray bursts (Thompson et al. 2004; Dessart et al. 2008) and optically brightened core-collapse supernovae (Kasen & Bildsten 2010) and as sources of relativistic, Poynting-flux-dominated outflows (Usov 1992; Yi & Blackman 1998; Metzger et al. 2011), ultra-high-energy cosmic rays (Arons 2003), and gravitational waves (Stella et al. 2005; Dall'Osso et al. 2009; Piro & Thrane 2012). The latter signal is predicted to be detectable out to the Virgo Cluster by current-generation, long-baseline antennas like the Laser Interferometer Gravitational Wave Observatory (LIGO; Abbott et al. 2009) at a rate of about one event per year. Two gravitational radiation mechanisms have been analyzed: a permanent mass quadrupole created by magnetic stresses in the stellar interior (Bonazzola & Gourgoulhon 1996; Dall'Osso et al. 2009; Mastrano et al. 2011), and a transient quadrupole generated by hydrodynamic instabilities like bar modes, which are excited when the protomagnetar spins up during fallback (Piro & Thrane 2012). In the latter context, it is crucial to understand how the rotation and magnetization of the protomagnetar evolve as fallback proceeds. If the magnetic field is relatively high and/or the star rotates relatively rapidly, the magnetic propeller effect (Illarionov & Sunyaev 1975) shields the star from infalling material, limiting the strength of the gravitational wave signal and preventing the formation of a black hole (Piro & Ott 2011).

In this paper, we extend the magnetar fallback scenario by incorporating magnetic funneling of the accretion flow onto the magnetic poles of the star (Romanova et al. 2003). In other accreting systems, such as low-mass X-ray binaries, magnetic funneling strongly modifies the surface distributions of mass and magnetic flux and hence observable properties like the magnetic dipole moment (Brown & Bildsten 1998; Payne & Melatos 2004), equilibrium spin (Wang et al. 2011), thermonuclear X-ray burst recurrence times, energies, and harmonic content (Payne & Melatos 2006b; Misanovic et al. 2010; Cavecchi et al. 2011; Patruno 2012), cyclotron lines (Mukherjee & Bhattacharya 2012), and gravitational wave output (Melatos & Payne 2005; Vigelius & Melatos 2009b). Here we show that magnetar fallback is modified in two important ways: (1) the accreting material forms a magnetically supported polar mountain, whose gravitational radiation supplements the signal from the internal magnetic quadrupole and hydrodynamic instabilities; and (2) polar magnetic burial reduces the magnetic dipole moment and hence the effectiveness of the propeller mechanism, making it easier to form a black hole.

The paper is structured as follows. In Section 2, we introduce an idealized, general-purpose model of magnetar fallback proposed recently (Piro & Ott 2011; Piro & Thrane 2012). We investigate how magnetic funneling modifies gravitational wave emission, mass capture, and black hole formation in the context of the model in Sections 3 and 4, respectively. Conclusions and a survey of the limitations of the calculation are presented in Section 5. We emphasize at the outset that the results pertain specifically to the scenario, addressed by the authors above and many others, where a magnetar is born spinning fast, with a rotation period of a few milliseconds. At least some magnetars may not be born in this manner. X-ray observations of three supernova remnants associated with anomalous X-ray pulsars and soft gamma-ray repeaters in the Milky Way imply that the explosion energy of their progenitors is close to the canonical supernova value of 1051 erg, arguing against fast rotation at birth in three out of ∼20 known objects (Vink & Kuiper 2006).1

2. PROTOMAGNETAR EVOLUTION DURING FALLBACK

We begin by summarizing the key ingredients of magnetar fallback. In Section 2.1, we adopt the standard, parameterized prescriptions for the propeller torque, accretion rate, and hydrodynamic instability threshold favored in the literature (Piro & Ott 2011; Piro & Thrane 2012). In Section 2.2, we calculate how the mass ellipticity epsilon and magnetic dipole moment μ evolve as functions of the accreted mass Ma, applying the rigorous theory of polar magnetic burial developed originally for recycled neutron stars (Payne & Melatos 2004; Priymak et al. 2011). At every step we emphasize the idealizations in the model, which are unavoidable. The messy physics of fallback, especially its geometry, is incompletely understood even for ∼1012 G fields, let alone for magnetars (Bernal et al. 2010). A related calculation without magnetic funneling was performed by Watts & Andersson (2002) in the context of gravitational radiation from r-modes during supernova fallback. In addition to bar-mode instabilities and the propeller effect, Watts & Andersson (2002) treated viscous damping and the star's thermal response to accretion and r-mode heating, effects that are neglected here.

2.1. Accreted Mass and Angular Velocity

Accretion during fallback has been studied thoroughly in the context of collapsars, "mild" core-collapse events in which part of the stellar mantle initially explodes then stalls and implodes. Numerical simulations of collapsars that treat the hydrodynamics, shock physics, and neutrino transport in detail find accretion rates in the range 10−4–10−2M s−1 lasting for 103–104 s (MacFadyen et al. 2001; Zhang et al. 2008). Fallback passes through early and late stages, with $\dot{M}_{\rm early} \approx 10^{-3} \eta t^{1/2} \,M_{\odot } \, {\rm s^{-1}}$ and $\dot{M}_{\rm late} \approx 50\, t^{-5/3}\, M_{\odot } \, {\rm s^{-1}}$, respectively, where t is the time after core bounce (in seconds), and η depends sensitively on the explosion energy, Es, with 0.1 ⩽ η ⩽ 10 for 0.3 ⩽ Es/1051 erg ⩽ 1.2. The transition occurs at t ∼ 102 s, after the oxygen shell falls in. We adopt the convenient parameterization $\dot{M}_{\rm a} = (\dot{M}_{\rm early}^{-1} + \dot{M}_{\rm late}^{-1})^{-1}$ for the total accretion rate, $\dot{M}_{\rm a}$, introduced by Piro & Ott (2011). For η in the above range, $\dot{M}_{\rm a}$ is mostly high enough to overcome the outward ram pressure of the magnetar's neutrino- and Poynting-flux-driven wind, even when the system is spherically symmetric (Piro & Ott 2011; Piro & Thrane 2012).

The angular velocity of the magnetar, Ω(t), evolves in response to three torques, Ndip, Na, and Ngw, defined below, according to $I\dot{\Omega } = N_{\rm dip} + N_{\rm a} + N_{\rm gw}$, where I = 0.35 MR2 is the moment of inertia in terms of the stellar mass M and radius R (Lattimer & Prakash 2001). Figure 6 of Lattimer & Prakash (2001) gives 0.2 ⩽ I/MR2 ⩽ 0.5 for a range of plausible equations of state; we adopt a central value here. The magnetic dipole torque Ndip obeys the standard vacuum formula, Ndip = −μ2Ω3/(6 c3), as long as the light cylinder (radius c/Ω) lies inside the hydromagnetic lever arm, given by the nominal Alfvén radius $r_{\rm m} = \mu ^{4/7} (G\!M)^{-1/7} \dot{M}_{\rm a}^{-2/7}$. For millisecond magnetars, this is a fair approximation, setting aside extended-dipole corrections (Melatos 1997). The accretion torque Na takes either sign, depending on the ratio of its mechanical and magnetic components, and is controlled by the fastness parameter ω = (rm/rc)3/2, where rc = (GM2)1/3 is the corotation radius (Ghosh & Lamb 1979). Here we follow Eksi et al. (2005) and Piro & Ott (2011) and set $N_{\rm a} = (1-\omega) (G\!Mr_{\rm m})^{1/2} \dot{M}_{\rm a}$. A more careful treatment of the disk–magnetosphere boundary and hence Na, to include episodic accretion in a corotation-limited disk (Rappaport et al. 2004; D'Angelo & Spruit 2011) and radiation pressure (Andersson et al. 2005),2 lies outside the scope of this paper; our aim here is to explore how magnetic funneling affects magnetar fallback irrespective of the specific mode of disk accretion. The gravitational wave torque Ngw is treated phenomenologically, following Piro & Ott (2011), by setting Ngw = −Naθ(β − βc), where θ(...) is the Heaviside step function, β(M, R, Ω) is the ratio of the star's kinetic and gravitational potential energies, and βc is the threshold for radiation-reaction-driven (βc = 0.14), viscosity-driven (0.14), and bar-mode (0.27) instabilities. Qualitatively, gravitational wave friction acts to cancel the other torques, once a suitable instability is triggered, with Ngw ≈ −NaNdip and |Ndip| ≪ |Na| typically. In Section 2.2, we add to Ngw the secular contribution from the quadrupole associated with a polar magnetic mountain.

The mass of the star, M(t), evolves according to $\dot{M} = \dot{M}_{\rm a} \theta (1-\omega)$, where $\dot{M}_{\rm a}$ is the parameterized fallback rate above, and the propeller effect shuts off accretion for ω > 1. This oversimplifies things: even under propeller conditions, some of the infalling plasma penetrates to the stellar surface due to instabilities at the disk–magnetosphere boundary (Rappaport et al. 2004; Romanova et al. 2005; D'Angelo & Spruit 2011). We bundle the uncertainties surrounding propeller leakage specifically and disk dynamics generally into the parameter η, an approach justified approximately elsewhere through one-zone, Shakura–Sunyaev disk calculations (Piro & Ott 2011; Piro & Thrane 2012). To stay consistent with previous work, the initial mass is set to M(0) = 1.4 M, the distinction between gravitational and baryonic mass is ignored, and R is held constant as M increases, a fair approximation for most equations of state until just before black hole formation (Lattimer & Prakash 2001).

2.2. Polar Magnetic Burial

The strong magnetic field of a magnetar funnels the fallback accretion flow preferentially onto the magnetic poles to form a mountain. Funneling is imperfect, due to Rayleigh–Taylor mixing on the boundary of the polar flux tube and filamentation caused by obliquity, but simulations confirm that it occurs for a wide range of parameters and geometries (Romanova et al. 2003; Kulkarni & Romanova 2008). As matter splashes down onto the poles, it slides sideways under its own weight, dragging along magnetic field lines by flux freezing. Thus the magnetic field is compressed at the equator and pushes back to confine the mountain, while the radial component of the polar field and hence μ are reduced (Melatos & Phinney 2001; Payne & Melatos 2004).

To calculate accurately the magnetic field structure B(x) and density profile ρ(x) of a polar mountain is a subtle task. Order-of-magnitude pressure balance arguments underestimate the mass quadrupole ∼104 fold by implicitly imposing outflow boundary conditions at the edge of the polar cap and hence underestimating the compression of, and tension in, the equatorial field, e.g., Brown & Bildsten (1998, and references therein). To avoid this, one must solve the force balance equation (quasistatic equilibrium; Alfvén crossing timescale $\ll M_{\rm a} / \dot{M}_{\rm a}$) simultaneously with a mass–flux constraint equation enforcing flux freezing and equatorial magnetic compression. Writing B = (rsin θ)−1∇ψ(r, θ) × ∇ϕ in spherical polar coordinates (r, θ, ϕ), where ψ(r, θ) is a scalar flux function, we obtain

Equation (1)

and

Equation (2)

for the force balance equation and mass–flux constraint respectively for a polytropic equation of state $P=K\rho ^\Gamma$, where P is the pressure, ϕ(r) is the gravitational potential in the Cowling approximation, ϕ0 is the gravitational potential at the stellar surface before accretion begins, Δ2 is the Grad–Shafranov operator, the integral in Equation (2) is computed along the field line ψ = constant, and the braces {...} in Equation (2) contain the same expression as the braces in Equation (1). The reader is referred to the literature for a detailed derivation of Equations (1) and (2) and a key to the notation (Payne & Melatos 2004; Priymak et al. 2011; Mukherjee et al. 2013a). A vital point concerns the barometric function F(ψ) and mass–flux ratio dM/dψ, i.e., the mass (initial plus accreted) per unit flux enclosed within (ψ, ψ + dψ). Some authors guess F(ψ) in Equation (1) and drop Equation (2) for convenience (Brown & Bildsten 1998; Melatos & Phinney 2001; Mukherjee et al. 2013a), but in general such a guess produces the wrong amount of equatorial magnetic compression. In reality F(ψ) is determined uniquely by accretion through the relevant initial value problem and dM/dψ. In equilibrium, this path-specific information is stored in Equation (2), which maps one-to-one the initial (pre-accretion) and final states to preserve flux freezing at all intermediate steps, as required by the mass continuity and magnetic induction equations of ideal magnetohydrodynamics (MHD; Priymak et al. 2011).

Figure 1 displays the equilibrium structure of a typical magnetic mountain ∼10η−1/2 s after fallback begins, with initial magnetic dipole moment μi = 5 × 1032 G cm3, accreted mass Ma = 1.2 × 10−3M, and line-tying boundary conditions (see Sections 2.1 and 2.2 in Priymak et al. 2011 for details). From the solid contours, one observes that the magnetic field is compressed into an equatorial band, with BrBθ near the pole and BθBr on either side of the equator. The magnetic tension, directed along the field-line radius of curvature, confines the accreted matter into a mound, whose isodensity surfaces are drawn as dashed contours. The equilibrium is calculated for Ma = 1.6 Mc, where Mc = 3 × 10−3i/1033 G cm3)2M is the characteristic burial mass for a polytropic equation of state with nondegenerate neutrons (model D in Priymak et al. 2011). Other polytropes with 1 ⩽ Γ ⩽ 5/3 give qualitatively similar results (see Figure 8 in Priymak et al. 2011). By constructing a quasistatic sequence of equilibria with increasing Ma, one finds that the magnetic dipole and mass quadrupole moments scale roughly as μ = μi(1 + Ma/Mc)−1 and epsilon = (Ma/M)(1 + Ma/Mc)−1 respectively (Shibazaki et al. 1989; Payne & Melatos 2004; Zhang & Kojima 2006; Priymak et al. 2011); polar magnetic burial screens μ, even as the mountain builds up, and |B| rises at the equator. In Figure 1, for example, we have μ = 0.6 μi and epsilon = 5 × 10−4 for Ma = 1.6 Mc, within 50% of the predictions from the rule-of-thumb formulae. The formula for epsilon, which implies epsilonMc/M, is deliberately conservative, to avoid overpredicting the resulting gravitational wave signal. In reality, epsilon rises gradually above Mc/M for MaMc in time-dependent MHD simulations that grow the mountain from scratch (see Figure 5 in Vigelius & Melatos 2009b), accompanied by transient, localized, loss-of-equilibrium events (cf. instabilities), where the magnetic field pinches off at isolated points to form topologically disconnected loops without disrupting the main body of the mountain (Klimchuk & Sturrock 1989; Payne & Melatos 2004; Vigelius & Melatos 2008; Mukherjee et al. 2013a).

Figure 1.

Figure 1. Hydromagnetic structure of a polar magnetic mountain on a magnetar, showing magnetic field lines (solid blue curves) and isodensity contours (dashed black curves and grayscale in red). One quadrant of a meridional section is displayed, with altitude above the base (in centimeter) on the vertical axis and colatitude on the horizontal axis. Mountain parameters: Ma = 1.2 × 10−3M = 1.6 Mc, μi = 5 × 1032 G cm3, equation of state model D from Priymak et al. (2011), μ = 0.6 μi, and epsilon = 5 × 10−4.

Standard image High-resolution image

It is natural to wonder whether the stressed magnetic configuration in Figure 1 is stable on the Alfvén timescale τA. The answer, counterintuitively, seems to be yes. Independent MHD simulations by two groups using the solvers ZEUS (Payne & Melatos 2007; Vigelius & Melatos 2008) and PLUTO (Mukherjee et al. 2013a, 2013b) confirm stability in two and three dimensions. A self-consistent solution of Equations (1) and (2), when imported into ZEUS or PLUTO, initially experiences the undular submode of the Parker instability, but it is not disrupted; after releasing a "magnetic blister," the mountain settles down to a new equilibrium with epsilon reduced by 50%–70% and mass loss ≲ 1% (Vigelius & Melatos 2008). Magnetic line tying at the inner boundary stabilizes the undular submode and switches off the interchange submode and growing modes in the continuous spectrum completely (Vigelius & Melatos 2008, 2009a). Mukherjee et al. (2013a) found unstable, pressure-driven, filamentary, toroidal modes at the periphery of filled and hollow mountains, but (1) these modes do not disrupt the mountain overall, consistent with Vigelius & Melatos (2008); (2) they are suppressed when fixed-gradient boundary conditions at the edge of the polar cap are replaced by north–south symmetry at the equator, to allow properly for stabilization by the equatorial magnetic belt (cf. Litwin et al. 2001); and (3) they arise from equilibria satisfying Equation (1) but not Equation (2) (see above). Grid-refinement tests buttress these findings (Vigelius & Melatos 2008; Mukherjee et al. 2013a). For example, the linear growth rate of the undular submode is observed to vary ∝ (grid  scale)−1/2 and is grid-limited, as theory predicts, but the final, saturation values of μ and epsilon are found to be independent of the grid scale (Vigelius & Melatos 2008).

A magnetic mountain does not relax resistively or by sinking on protomagnetar timescales (≲ 104 s; Vigelius & Melatos 2009c; Wette et al. 2010), nor is it disrupted by resistive instabilities. The buried polar magnetic field and hence μ resurrect on the shorter of the Ohmic diffusion and Hall timescales τd and τH, with min(τd, τH) ≳ 1 yr even with enhanced accretion-driven heating (Vigelius & Melatos 2009c; Viganò et al. 2013). Likewise, subduction and meridional redistribution occur long after the magnetar spins down and stops emitting gravitational waves at a detectable level, if no black hole forms (Choudhuri & Konar 2002). The values of μ and epsilon calculated from Equations (1) and (2) are within a factor of two of ZEUS mountains grown from scratch on a soft surface, as long as the boundary at r = Rin in Figure 1 is set deep enough that the mountain is light compared to the substrate, its base does not move much laterally, and magnetic line tying stays a good approximation (Wette et al. 2010). In MHD simulations where the electrical resistivity is boosted artificially, the mountain is stable on the tearing mode timescale, (τdτA)1/2 ∼ 104 s (Vigelius & Melatos 2008).

3. GRAVITATIONAL RADIATION

A magnetic mountain supplements the gravitational radiation from instabilities. We find in Section 3.1 that the added signal is substantial. The improvement in detectability is potentially even greater, as shown in Section 3.2, because the mountain radiates longer than instabilities under certain conditions, and its waveform is cleaner (cf. nonlinear fluid motions near centrifugal break-up). We estimate the gravitational wave strain assuming that the neutron star survives then look at how the conclusions are modified by black hole formation in Section 4.

3.1. Peak Wave Strain

Figure 2 displays contours of the peak gravitational wave strain hmax = max[h0(t)] as a function of μi and the initial spin period Pi. The maximum is computed for 0 ⩽ t ⩽ 104 s and typically occurs at ttpk = 85η−6/13 s, i.e., the accretion timescale, which does not depend on μi or Pi. To clarify the physics, we examine separately the consequences of magnetic funneling and hydrodynamic instabilities in the top and bottom panels of Figure 2. To help the reader interpret Figure 2, we also plot representative examples of the star's rotational evolution and dominant torque components (Na, NgwNdip) in Figure 3 for four scenarios (top to bottom rows), defined by whether or not magnetic funneling occurs, hydrodynamic instabilities are switched on, and μi is low (solid curves) or high (dashed curves). The rotational evolution differs between the scenarios, as we now discuss.

Figure 2.

Figure 2. Contours of peak gravitational wave strain hmax = max [h0(t)] (in units of 10−23) as a function of initial spin period Pi (in units of milliseconds) and initial magnetic dipole moment μi (in units of 1033 G cm3), excluding and including magnetic dipole reduction by burial (solid and dashed curves, respectively). Top panel: radiation from magnetic mountain only. Bottom panel: radiation from magnetic mountain and hydrodynamic instabilities. Parameters: η = 1, βc = 0.14, Mc = 3 × 10−3i/1033 G cm3)2, M(0) = 1.4 M, R = 10 km, and D = 1 Mpc.

Standard image High-resolution image
Figure 3.

Figure 3. Rotational evolution of a protomagnetar experiencing fallback accretion. Left column: angular velocity Ω vs. time t. Right column: torques due to accretion (Na: black curves) and gravitational radiation (Ngw: red curves) vs. t. Four scenarios are considered, according to whether or not a magnetic mountain forms and radiates, hydrodynamic instabilities are excited and radiate, and μ is reduced by burial. Top row: mountain forms, μ constant, instabilities switched off. Second row: mountain forms, μ reduced, instabilities switched off. Third row: mountain forms, μ constant, instabilities switched on. Bottom row: mountain forms, μ reduced, instabilities switched on. In every panel, the solid and dashed curves correspond to low magnetization (μi = 0.7 × 1033 G cm3, Pi = 5 ms) and high magnetization (μi = 1.5 × 1033 G cm3, Pi = 5 ms), respectively. Other parameters as for Figure 2. The plotted quantities vary by ∼15% over the range 0.2 ⩽ I/MR2 ⩽ 0.5.

Standard image High-resolution image

In the top panel, we switch off artificially the instabilities, so that they contribute neither to h0 nor to Ngw, and focus on the radiation from the mountain, with characteristic strain h0(t) = 2 GΩ(t)2epsilon(t)I(t)/(c4D), where D is the distance to the source. The solid contours describe what happens, when the mountain quadrupole grows with t, yet μ stays constant—an unlikely scenario, given the physics of magnetic burial in Section 2.2, but still instructive as a stepping stone to the full problem. For μi ≳ 0.7 × 1033 G cm3, the star quickly attains magnetocentrifugal equilibrium [ω ≈ 1, Ω ≈ 7 × 103(μ/1033 G cm3)−6/7 rad s−1] at ttpk, getting there either by spinning up by accretion (Pi ≳ 1 ms) or spinning down by the propeller effect (Pi ≲ 1 ms). It then spins down gradually over ∼103 s mainly under the action of Ngw, with Na assisting the deceleration for t ≳ 5 × 102 s, and Ndip contributing ∼10% of the total torque (see Figure 3, top row, right panel). The instability threshold is never crossed; β peaks at ≈0.05, so it is consistent to switch off the instabilities in this regime. The peak wave strain increases gently with μi according to $\epsilon \propto M_{\rm c} \propto \mu _{\rm i}^2$ (MaMc), $\Omega \propto \mu _{\rm i}^{-6/7}$ (magnetocentrifugal equilibrium), and hence $h_{\rm max} \propto \epsilon \Omega ^2 \propto \mu _{\rm i}^{2/7}$. The hmax scaling steepens at low Pi, where the solid contours turn upward, because the propeller physics asserts itself earlier.

For μi ≲ 0.7 × 1033 G cm3 in the above scenario (Figure 3, top row, dashed curves), the star spins up toward its centrifugal limit,3 entering the propeller phase relatively late at t ≳ 5 × 102 s ≫ tpk. It then spins down, but more gradually than for μi ≳ 0.7 × 1033 G cm3 because, by the time the star enters the propeller phase, $N_{\rm a} \propto \dot{M}_{\rm a}$ is well below its peak (NdipNa here). The gravitational wave signal is also weaker, as the solid contours indicate, and as expected from $h_{\rm max} \propto \mu _{\rm i}^{2/7}$. However, the predicted hmax is unrealistic, because β promptly exceeds βc at ttpk in the low-μi regime, i.e., the switched-off instabilities would switch on rapidly in reality to emit gravitational radiation and spin down the star through Ngw. The latter behavior is discussed further below with reference to the bottom panel of Figure 2.

Now suppose that μ diminishes through magnetic burial, while the mountain grows, and the instabilities remain artificially switched off. The results are described by the dashed contours in the top panel of Figure 2; see also the second row of Figure 3. The contours in Figure 2 are nearly horizontal. Burial suppresses the propeller effect observed at low Pi and high μi in the solid contours, because at ttpk we have MaMc and hence μ ≪ μi everywhere in the plotted region. Moreover, with μ ≪ μi everywhere, Na behaves similarly for low and high μi; compare the dashed and solid black curves in the second row of Figure 3. The accretion torque rapidly spins up the star toward torque balance (Na + Ngw ≈ 0), while maintaining ω < 1 for all t except t ≲ 10 s ≪ tpk, when Ma is still less than Mc. The wave strain hmax∝Ω2 is ≈4 times greater than for the solid contours, because Ω peaks at the stall frequency (Bildsten 1998), which turns out to be about double the magnetocentrifugal equilibrium frequency (see above). The scaling for the stall frequency, $\Omega \propto \epsilon ^{-2/5} \propto \mu _{\rm i}^{-4/5}$, implies $h_{\rm max} \propto \epsilon \Omega ^2 \propto \mu _{\rm i}^{2/5}$, which is confirmed by inspecting the dashed contours. The trend with Pi is weak, as the propeller is inactive. The stall frequency is reached at ttpk, after which $\dot{M}_{\rm a}$ and hence Na drop away rapidly, and h0(t) decays on the characteristic gravitational wave spin-down timescale $I\Omega /N_{\rm gw} \propto \mu _{\rm i}^{-4/5}$, evaluated at the stall point, with IΩ/Ngw ∼ 103 s typically (noting |Ndip| ≪ |Ngw|). The decay timescale influences the detectability of the signal, as discussed in Section 3.2. For μi ≳ 1 × 1033 G cm3, β peaks below βc, so switching off artificially the instabilities does not affect the conclusions. At lower μi, the conclusions are affected, as discussed below.

Finally, let us switch on the hydrodynamic instabilities. The results including (dashed contours) and excluding (solid contours) μ reduction by magnetic burial are presented in the bottom panel of Figure 2 and the third and fourth rows of Figure 3. The instability-sourced contribution to the wave strain is given by h0(t) = [5GNa(t)]1/2/[8c3D2Ω(t)]]1/2 from torque balance (Na = Ngw; see Section 2.1 and Piro & Ott 2011).4 We set the threshold at βc = 0.14; the results are qualitatively similar for βc = 0.27. Consider first the solid contours. They are distorted at the left edge of the plot (Pi ≲ 2 ms), because the propeller effect spins down the star initially, before Ω rises to a local maximum ≳ 3 × 103 rad s−1 at ttpk. Along the lower edge of the plot, for μi ≲ 0.3 × 1033 G cm3 (where the contours display kinks), accretion spins up the star toward its centrifugal limit, triggering hydrodynamic instabilities; β stays above βc for ttpk, overshooting as far as β ≈ 0.2, and the star spins down electromagnetically on the timescale ∼103i/1033 G cm3)−2 s, with Ngw balancing $N_{\rm a}\propto \dot{M}_{\rm a}$, as the instabilities operate. For μi ≳ 0.3 × 1033 G cm3, no instabilities are triggered, β < βc decreases monotonically with t, and the rotational evolution in the first and third rows of Figure 3 is similar.

On the other hand, the dashed contours in Figure 2 describe what happens when μ is reduced by burial. Across the whole plot, we find ω < 1 for all t, i.e., there is no propeller effect. Likewise Ndip is insignificant. For μi ≳ 1 × 1033 G cm3, no instabilities are triggered. The star spins up to the stall frequency, satisfying NaNgw (Bildsten 1998), then spins down in response to the mountain component of Ngw (with |Ngw| ≈ 1.5|Na|) on the timescale ${\sim} I\Omega /N_{\rm gw} \propto \mu _{\rm i}^{-4/5}$. The latter evolution resembles the dashed contours in the top panel of Figure 2 and the dashed curves in the second row of Figure 3, as the mountain quadrupole is significant. For μi ≲ 1 × 1033 G cm3, instabilities are triggered, Na drops sharply due to the instability back-reaction (see solid curve at t ≈ 70 s in Figure 3, fourth row, right panel), and Ω ∼ 104 rad s−1 decays slowly under the action of Ngw and Ndip, which is much reduced by burial.

3.2. Signal-to-noise Ratio

An accurate estimate of the signal-to-noise ratio (S/N) for a protomagnetar undergoing fallback requires detailed calculations of the waveform and Monte Carlo simulations of the search pipeline, both of which lie outside the scope of this paper. However, it is useful to convert the wave strain predictions in Figure 2 into a rough detectability measure, in order to clarify the relative importance of the peak strain, hmax, and signal lifetime, Tobs. To give a flavor of what is possible, we consider two extremes: (1) a matched filter search, which assumes optimistically that one can track the phase of the signal coherently for its duration, and (2) an excess cross-power search targeting a well-localized electromagnetic counterpart, which does not assume any phase model at all.

The key factors governing detectability are summarized in Table 1. Four scenarios emerge from Section 3.1, classified according to whether the maximum Ω (achieved at ttpk) is set by magnetocentrifugal equilibrium or gravitational radiation stalling (C or G, first column), and whether gravitational wave emission is dominated by a magnetic mountain or hydrodynamic instabilities (M or I, second column).5 Scalings are presented for hmax and Tobs (third and fourth columns) in terms of the normalized initial magnetic dipole moment, $\tilde{\mu } = \mu _{\rm i} / (10^{33}\,{\rm G\,cm^3})$, and the fallback parameter, η. The source lifetime generally satisfies Tobstpk, i.e., h0(t) rises faster than it decays. It is set by the propeller effect (TobsIΩ/Na, with ω ≈ 1.05) or the back-reaction from mountain gravitational radiation (TobsIΩ/Ngw) in the scenarios C or G, respectively. The matched filter S/N is calculated from ${\rm (S/N)}_{\rm mf}^2 \approx 32 h_{\rm max}^2 T_{\rm obs} / [375 S_h(f=\Omega /\pi)]$ (e.g., Equation (13) in Vigelius & Melatos 2009b), where Sh(f) is the one-sided detector noise power spectral density at the observing frequency f. We thereby assume that most of the power is emitted at twice the spin frequency (Jaranowski et al. 1998)—almost certainly an oversimplification for instabilities and possibly also for a mountain, if it wobbles in response to vigorous accretion (see footnote 4 and Payne & Melatos 2006a). In all table entries the source distance is normalized to 1 Mpc, and (S/N)mf is quoted in terms of $\tilde{S}_h = S_h(f=0.2\,{\rm kHz}) / (10^{-47} \, {\rm Hz^{-1}})$, with $\tilde{S}_h = 1.4$, 0.90, and 6.7 × 10−3 for zero-detuning high-power Advanced LIGO, neutron-star-inspiral-optimized Advanced LIGO, and the conventional Einstein Telescope, respectively (Bennett et al. 2010, and references therein). Interferometer configurations optimized for f ≲ 40 Hz (e.g., black-hole-inspiral-optimized Advanced LIGO, xylophone Einstein Telescope) are not considered in this paper, where we are interested in signals with f ≳ 0.2 kHz and hence Sh(f)∝f2. The final column quantifies roughly the reduction in detection distance expected when replacing a matched filter with an excess cross-power search, based on the results in Table 1 in Piro & Thrane (2012).

Table 1. Gravitational Wave Detection: Accretion and Emission Scenarios

Spin Limita Signalb 1023DMpchmaxc Tobs (102 s)d $\tilde{S}_h^{1/2} \,D_{\rm Mpc} \,{\rm (S/N)}_{\rm mf}$e Dmf/Dcpf
C M $2.0 \tilde{\mu }^{2/7} \eta ^{60/91}$ $8.8 \tilde{\mu }^{-8/7} \eta ^{-30/91}$ $3.7 \tilde{\mu }^{4/7} \eta ^{15/91}$ ∼14
C I $6.3 \tilde{\mu }^{3/7} \eta ^{20/91}$ As above $12 \tilde{\mu }^{5/7} \eta ^{-25/91}$ As above
G M $6.7 \tilde{\mu }^{2/5} \eta ^{4/13}$ $1.2 \tilde{\mu }^{-4/5} \eta ^{-8/3}$ $3.8 \tilde{\mu }^{4/5} \eta ^{-46/39}$ ∼6
G I $7.3 \tilde{\mu }^{2/5} \eta ^{4/13}$ As above $4.2 \tilde{\mu }^{4/5} \eta ^{-46/39}$ As above

Notes. In all entries, $\tilde{\mu } = \mu _{\rm i} /(10^{33} \, {\rm G\,cm^3})$ denotes the normalized initial magnetic dipole moment, and η is the fallback accretion parameter defined in Section 2.1. aAccretion mechanism that determines the maximum value of Ω at ttpk: magnetocentrifugal equilibrium (C) or gravitational radiation stalling (G). bGravitational wave emission mechanism: magnetic mountain (M) or hydrodynamic instabilities (I). cPeak wave strain hmax at ttpk; DMpc denotes the source distance measured in Mpc. dSignal duration Tobs = IΩ/Na (C) or IΩ/Ngw (G) in units of 102 s; see Section 3.2. eSignal-to-noise ratio (S/N)mf for a matched filter search; $\tilde{S}_h = S_h(0.2\,{\rm kHz}) / (10^{-47}\,{\rm Hz^{-1}})$ denotes the normalized detector noise power spectral density at 0.2 kHz. fMatched filter (Dmf) and excess cross-power (Dcp) detection distances; ratio calibrated against Table 1 in Piro & Thrane (2012).

Download table as:  ASCIITypeset image

Table 1 demonstrates that the prospects for detecting a nearby protomagnetar remain respectable, as stated by previous authors (Piro & Ott 2011; Piro & Thrane 2012), when the physics of magnetic funneling and burial is included. For example, a Local Group object with μi = 5 × 1033 G cm3 and D = 4 Mpc reaches (S/N)mf = 42 in a matched filter search with the Einstein Telescope just from its magnetic mountain emission, even before adding instabilities.6 The S/N drops to 3.6 for Advanced LIGO, still marginally detectable, while the effective detection distance is ∼6 times smaller for a more realistic excess cross-power search (see below). Table 1 contains two main trends: (1) gravitational radiation stalling leads to higher hmax and shorter Tobs than magnetocentrifugal equilibrium, so both evolutionary pathways produce similar S/Ns; and (2) the mountain and instability emission are comparable, with the latter typically being stronger and favored by higher μi and lower η.

The excess cross-power algorithm (Thrane et al. 2011; Prestegard et al. 2012), a generalized form of the stochastic radiometer statistic (Ballmer 2006), does not assume a priori how the phase evolves. It is therefore a fairer guide to search performance. It is implemented in two steps: one computes a spectrogram S/N(t; f) of the S/N, which is proportional to the cross-correlation of the strains at two interferometers (or a larger network more generally); then one scans the spectrogram for a contiguous track of positive-valued pixels using a clustering algorithm (after excising environmental noise artifacts), in order to get a total S/N for the track. A detection threshold of S/N ≈ 23 can be achieved realistically with Advanced LIGO, assuming a false alarm rate of 0.1%, a false dismissal rate of 50%, and a 1 ks × 1.7 kHz on-source region divided into 0.5 s × 1 Hz pixels. Tripling the on-source time increases the threshold by ≈10%.

Detailed Monte Carlo simulations of the excess cross-power S/N lie outside the scope of this paper; there is considerable uncertainty surrounding the waveforms, when hydrodynamic instabilities operate, and even the magnetic mountain is unlikely to be a static quadrupole, when $\dot{M}_{\rm a}$ is so high during fallback. Instead, we calibrate against the results presented by Piro & Thrane (2012). Four of the cases presented in Table 1 in the latter reference, with η = 1 and maximum stellar mass 2.5–2.9 M, are representative of the results in Section 3.1 and Figure 2: the emission frequency ranges from 0.79 kHz to 2.3 kHz, and the duration of the burst ranges from 0.21 ks to 0.35 ks, consistent with the Ω(t) evolution underlying Figure 2. In the four cases, the ratio of detection distances for the matched filter and cross-power searches ranges from 6.2 to 14. Other examples computed by Piro & Thrane (2012) are consistent with this range; at one extreme, for η = 0.3, the duration increases to 0.64 ks and the detection distance ratio is 19. As a rough guide, we adopt these results here (final column, Table 1), by applying a 14 fold reduction to the detection distance for magnetocentrifugal equilibrium, where Tobs is longer, and a 6 fold reduction for gravitational stalling, where Tobs is shorter. When the phase evolution is unknown, and a template search is prohibitive computationally, excess cross-power performs better relative to a matched filter the longer the signal lasts, as the tf pixel tracker asserts its advantage.

4. BLACK HOLE FORMATION

The gravitational wave signals discussed in Section 3 are predicted to truncate suddenly at ttpk, once enough material falls back onto the neutron star to form a black hole. The maximum gravitational mass of a stable protomagnetar, Mmax, is given by the nonrotating Tolman–Oppenheimer–Volkoff mass, MTOV, corrected for centrifugal support from differential rotation. To a first approximation, one can write $M_{\rm max} = M_{\rm TOV}(1 + \gamma \tilde{\Omega }^2)$, where $\tilde{\Omega }$ is the angular velocity normalized by its centrifugal limit, γ is a constant of the order of unity (Lyford et al. 2003), and the nonrotating maximum mass spans the range 2 ≲ MTOV/M ≲ 3 for a selection of popular equations of state allowed by observations (Lasky et al. 2014).

Figure 4 shows how the nominal mass accreted onto the neutron star varies with initial magnetization and spin, before taking black hole formation into account. Contours are drawn for three fallback scenarios: η = 0.3 (blue), 1.0 (red), and 3.0 (green). The results are to be compared with Figures 3 and 4 in Piro & Ott (2011). We include mountain and instability contributions to the torques but not μ reduction by magnetic burial, as for the solid contours in the bottom panel of Figure 2. If μ is buried, the propeller effect is suppressed for all Pi and μi across the plotted region; all the infalling matter, $\int dt^{\prime }\, \dot{M}_a(t^{\prime })$, finds its way onto the protomagnetar, which turns into a black hole at ttpk for all η ⩾ 0.3. If μ is not buried, the formation of a magnetic mountain still modifies the results of Piro & Ott (2011) to some extent through Ngw.7

Figure 4.

Figure 4. Contours of nominal accreted mass, M(104 s) − M(0), in solar masses, as a function of initial spin period Pi (in units of milliseconds) and initial magnetic dipole moment μi (in units of 1033 G cm3), for η = 0.3 (blue), 1.0 (red), and 3.0 (green). All curves include mountain and instability contributions to Na and Ngw and exclude μ reduction by burial; if μ is buried, there is no propeller effect, and one obtains $M(10^4\,{\rm s})-M(0) = \int _0^{10^4\,{\rm s}} dt^{\prime } \dot{M}_{\rm a}(t^{\prime }) = 2.0\, M_\odot,$ 3.0 M, and 4.3 M for η = 0.3, 1.0, and 3.0, respectively. Parameters: βc = 0.14, Mc = 3 × 10−3i/1033 G cm3)2, M(0) = 1.4 M, R = 10 km.

Standard image High-resolution image

In general, the trends in Figure 4 are qualitatively similar to those observed in Figures 3 and 4 in Piro & Ott (2011). The main difference is that the nominal accreted mass is ≈20% higher at μi = 1 × 1033 G cm3, and ≈50% higher at μi = 3 × 1033 G cm3, because the mountain gravitational wave torque, $N_{\rm gw} \propto \mu _{\rm i}^4$, decelerates the star and moderates the propeller effect. When fallback is vigorous, e.g., η = 3, the nominal accreted mass exceeds 3.5 M over the plotted region, and a black hole always forms. When fallback is less vigorous, e.g., η = 0.3, a black hole still forms in most parts of the figure except the top left. Taking Mmax = 2.2 M, at the lower end of the stable range, black hole formation is prevented for μi ≳ 0.8 × 1033 G cm3 and Pi ≲ 2 ms for η = 0.3. Taking Mmax = 3.5 M, at the upper end of the stable range after correcting for differential rotation (Lasky et al. 2014), black hole formation is prevented across the whole plotted region for η = 0.3 and for μi ≳ 0.5 × 1033 G cm3 for η = 1.0.

Recent observations of a sharp truncation in the X-ray flux from some short gamma-ray bursts, ∼102 s after the initial trigger, have been interpreted as evidence that we are seeing a protomagnetar forming in a binary neutron star coalescence event and collapsing to form a black hole (Rowlinson et al. 2013). Although this phenomenon is not described by the analysis in this paper, which is motivated by core-collapse supernovae, it is related in two ways. First, Fan et al. (2013) and Lasky et al. (2014) have proposed a multimessenger experiment targeting short gamma-ray bursts, in which the X-ray light curve and an Advanced LIGO measurement of the progenitor chirp mass are combined with a simple magnetic braking model and the known mass distribution in neutron star binaries to constrain the nuclear equation of state. It is worth investigating whether a similar experiment is feasible in the supernova context, as suggested originally in Section 5 of the paper by Piro & Thrane (2012). Second, in the gamma-ray burst context, it may be worthwhile to generalize the magnetic braking model assumed by Lasky et al. (2014) to include some of the ingredients discussed in this paper, such as hydrodynamic instabilities, the fallback accretion torque (including in the propeller regime), the gravitational wave torque, magnetic funneling, magnetic mountain growth, and magnetic dipole moment reduction by burial.

5. CONCLUSION

In this paper we revisit the scenario of protomagnetar spin up by supernova fallback studied by Piro & Ott (2011) and Piro & Thrane (2012), in which hydrodynamic instabilities radiate gravitational waves, and the magnetocentrifugal propeller effect impedes black hole formation. We add one extra ingredient: magnetic funneling of the accretion flow. Following closely the analysis in the original papers, we show that magnetic funneling modifies the scenario in two ways. First, a polar magnetic mountain forms, confined by the accretion-compressed equatorial magnetic field as in low-mass X-ray binaries (Payne & Melatos 2004), whose mass quadrupole moment is substantial, with epsilon = 3 × 10−3i/1033 G cm3)2. The associated, quasimonochromatic gravitational wave signal is somewhat weaker than the instability signal at its peak but can last somewhat longer (see Table 1), contributing comparably to the S/N and boosting the likelihood of detection. Figure 2 and Table 1 present the wave strain and S/N as functions of Pi and μi. They show that the peak wave strain and burst duration are chiefly controlled by whether the initial, fallback-driven spin up of the protomagnetar stalls in response to the magnetocentrifugal or gravitational wave torque, with the latter dominating for μi ≳ 1 × 1033 G cm3. Second, as in low-mass X-ray binaries, growth of a polar magnetic mountain is accompanied by reduction of the stellar magnetic dipole moment, even as the local magnetic field at the equator intensifies (Payne & Melatos 2004). As μ diminishes, the propeller effect rapidly shuts off, and most of the supernova debris falls back onto the protomagnetar, assisting black hole formation. Figure 4 presents the nominal accreted mass as a function of Pi, μi, and η. Gravitational wave emission is dominated by the magnetic mountain for μi ≳ 1 × 1033 G cm3 and instabilities for μi ≲ 1 × 1033 G cm3.

We emphasize in closing that the results in Figures 2 and 4 and Table 1 are indicative only. The model is idealized. The equations of motion for Ω(t) and M(t) implicitly assume magnetized, thin-disk accretion, which may not hold at the high accretion rates characteristic of supernova fallback. For example, the magnetic propeller may work less efficiently in a messy, high-$\dot{M}_{\rm a}$ accretion environment, whether or not μ is reduced by burial. The degree to which hydrodynamic instabilities self-adjust to cancel Na is a subtle issue; further study is needed before the approximation made here, namely perfect cancelation for β > βc, can be accepted as valid (see footnote 4). The complicated waveforms produced at high $\dot{M}_{\rm a}$ by hydrodynamic instabilities and a wobbling magnetic mountain (Payne & Melatos 2006a) are unknown at the time of writing; they are approximated by toy waveforms (or circumvented by computing the total radiated power) in this paper and by previous authors. Precession is neglected, even though it is likely to modify both the waveform (Payne & Melatos 2006a) and the torques (especially Na), when $\dot{M}_{\rm a}$ is high. In these and other respects, therefore, the model is emphatically not a substitute for large-scale, MHD simulations. Its value rests in drawing attention to the physical implications of a single, new effect—magnetic funneling—by comparing directly with otherwise identical models computed previously (Piro & Ott 2011; Piro & Thrane 2012). In summary, when magnetic funneling is included, (1) the gravitational wave amplitude and duration are multiplied by factors of roughly 0.3–0.9 and 1–7 respectively, boosting the S/N by up to ∼3 times overall, and (2) a black hole is more likely to form, because the propeller phase is suppressed, and ≳ 50% of the infalling matter lands on the compact object.

A.M. thanks the Theoretical Astrophysics Group at the Eberhard Karls Universität Tübingen for its generous hospitality while the paper was started. This work was supported by an Australian Research Council Discovery Project grant and a Group of Eight Germany Joint Research Cooperation grant.

Footnotes

  • Strictly speaking, Vink & Kuiper (2006) argued against fast rotation lasting longer than a (brief) initial time window. Strong gravitational radiation immediately after birth can nullify the issue of overpowering the remnant (Dall'Osso et al. 2009); cf. Lai et al. (2001).

  • Radiation pressure effects are modified by the strong polarization dependence of the relevant opacities at magnetar field strengths, a problem currently under investigation (van Putten et al. 2013).

  • Following Piro & Ott (2011), we do not model centrifugal break-up. Hence there are brief time intervals in some scenarios in Figure 3, during which Ω approaches the centrifugal limit for low μi (see solid curves in Figure 3), and the model breaks down. In a real system, the instability back-reaction self-adjusts to prevent break-up, and the results do not change qualitatively.

  • By summing the mountain and instability contributions to h0(t) and Ngw(t), we assume implicitly that they are independent, which is debatable; e.g., hydrodynamic instabilities may move the magnetic footpoints at the base of the mountain and hence affect epsilon(t). Likewise, once β exceeds βc, it is unclear whether Na self-adjusts to balance all of Ngw or just its instability-sourced component. Large-scale MHD simulations outside the scope of this paper are needed to resolve these subtle issues properly.

  • In reality, the signal is the sum of the M and I components, but they are presented separately for the convenience of the reader, because certain search pipelines may be more sensitive to one component. For example, the mountain signal may be easier to detect with a matched filter, because it is arguably cleaner than the instability signal, which carries the imprint of nonlinear physics like hydrodynamic turbulence; see also footnote 4.

  • A mountain is detected with false alarm and dismissal probabilities of 1% and 10%, respectively, when h0 exceeds the threshold 11.4 [Sh(f)/Tobs]1/2 (Jaranowski et al. 1998), corresponding to (S/N)mf = 3.3.

  • This is an unlikely scenario prima facie, because the growth of epsilon goes hand in hand with μ reduction in the rigorous theory of magnetic burial developed for low-mass X-ray binaries (Payne & Melatos 2004; Vigelius & Melatos 2009b; Priymak et al. 2011). Nevertheless we mention it for completeness, in case epsilon and μ decouple for some reason during supernova fallback in the high-$\dot{M}_{\rm a}$ regime.

Please wait… references are loading.
10.1088/0004-637X/794/2/170