Global Crustal Dynamics of Magnetars in Relation to Their Bright X-Ray Outbursts

, , and

Published 2017 May 23 © 2017. The American Astronomical Society. All rights reserved.
, , Citation Christopher Thompson et al 2017 ApJ 841 54 DOI 10.3847/1538-4357/aa6c30

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/841/1/54

Abstract

This paper considers the yielding response of a neutron star crust to smooth, unbalanced Maxwell stresses imposed at the core–crust boundary, and the coupling of the dynamic crust to the external magnetic field. Stress buildup and yielding in a magnetar crust are global phenomena: an elastic distortion radiating from one plastically deforming zone is shown to dramatically increase the creep rate in distant zones. Runaway creep to dynamical rates is shown to be possible, being enhanced by in situ heating and suppressed by thermal conduction and shearing of an embedded magnetic field. A global and time-dependent model of elastic, plastic, magnetic, and thermal evolution is developed. Fault-like structures develop naturally, and a range of outburst timescales is observed. Transient events with time profiles similar to giant magnetar flares (millisecond rise, ∼0.1 s duration, and decaying power-law tails) result from runaway creep that starts in localized sub-kilometer-sized patches and spreads across the crust. A one-dimensional model of stress relaxation in the vertically stratified crust shows that a modest increase in applied stress allows embedded magnetic shear to escape the star over ∼3–10 ms, dissipating greater energy if the exterior field is already sheared. Several such zones coupled to each other naturally yield a burst of duration ∼0.1 s, as is observed over a wide range of burst energies. The collective interaction of many plastic zones forces an overstability of global elastic modes of the crust, consistent with quasi-periodic oscillation (QPO) activity extending over ∼100 s. Giant flares probably involve sudden meltdown in localized zones, with high-frequency (≫100 Hz) QPOs corresponding to standing Alfvén waves within these zones.

Export citation and abstract BibTeX RIS

1. Introduction

Magnetars are neutron stars that emit intense and broadband electromagnetic radiation by dissipating very strong (≳1015 G) magnetic fields (Woods & Thompson 2006; Mereghetti et al. 2015; Turolla et al. 2015). Our interest here is in the extraordinary bursts of X-rays and gamma rays that are detected sporadically from some magnetars. We consider how stress imbalances build up in the liquid core and crust of the neutron star, global effects of yielding in the crust, and the coupling between crust and magnetosphere.

A role for the crust in mediating magnetar outbursts is suggested by (i) the detection of quasi-periodic oscillations (QPOs) during giant X-ray flares, many of which resemble global elastic modes of the crust (Israel et al. 2005; Strohmayer & Watts 2005; Watts & Strohmayer 2006); and (ii) the similarity between the time for an elastic wave to propagate around the star (∼0.1 s) and the duration of the most common soft gamma repeater (SGR) bursts, as measured over a wide range of burst energies (Göǧüş et al. 2000). SGR outbursts therefore appear to be a global phenomenon.

Some magnetars—the anomalous X-ray pulsars (AXPs)—appear to dissipate their magnetic fields gradually, consistent with plastic creep within the active parts of their crusts (Thompson & Duncan 1996; Jones 2003). The SGRs behave similarly in between their brief episodes of intense burst activity. Fluctuations in the X-ray output and spindown torque of magnetars show a broad range of timescales and energies. A central question is how changes in stress balance and current flow can emerge over timescales much shorter than the age of the star but only rarely result in subsecond transient X-ray emission, and even then only in some sources.

A promising approach starts with the extreme sensitivity of deformation rate in a solid to applied stress and temperature. This is a principal focus of research in geodynamics (Turcotte & Schubert 2002), and the extension to the extreme densities and pressures of neutron stars is supported by recent ab initio calculations using molecular dynamics methods (Chugunov & Horowitz 2010; Hoffman & Heyl 2012). Magnetar outbursts give evidence of very localized dissipation (Ibrahim et al. 2001; Lenters et al. 2003; Woods et al. 2004; Esposito et al. 2007). We show that narrow concentrations of rapid plastic flow are the natural response of the crustal solid to relatively smooth, large-scale magnetic stresses acting on it from below. We also investigate how magnetic stresses are communicated from crust to magnetosphere on short timescales, resulting in intense X-ray emission. Finally, we argue that the global redistribution of stresses by elastic forces in the crust is tied to QPO activity during large flares, in other words, that global elastic modes may become overstable during large flares.

Burst-active and burst-quiet magnetars both persistently emit nonthermal X-rays with luminosity ∼1035–1036 erg s−1, up to ∼103 times what could be supplied by the spindown of the star. This is consistent with the presence of strong external electric currents (magnetospheric twist) in both bursting and nonbursting sources (Thompson et al. 2002; Beloborodov & Thompson 2007). The nondetection of energetically significant burst emission from several magnetars with active magnetospheres suggests to us that an external current-driven instability is not the primary driver of subsecond burst activity, although such an instability may play a supporting role during episodes of fast and localized crustal shear deformation (Gill & Heyl 2010; Elenbaas et al. 2016). This conclusion is supported by detailed timing, spectral, and X-ray flux measurements before and after the 2004 giant flare from SGR 1806–20, the most energetic magnetar flare ever detected (Woods et al. 2007). For example, the spindown torque, which is sensitive to the external twist, increased substantially in the year before the flare, but then it decreased in the final 4–5 months and did not show any obvious anomaly after the flare.

1.1. Overview of Results

We begin by listing several open questions relating to magnetar activity, describing our approach to them, and offering the conclusions we have drawn.

The basic picture developed here views the magnetar crust as being composed of many interacting elastic units, which experience intermittent plastic creep over a wide range of rates. Maxwell stresses imparted to the crust at its lower boundary source nonlocal solid stresses. The creep rate in a given patch is determined in part by the collective stress imparted by its neighbors, and it is regulated by conductive cooling, neutrino emission, and the shearing of an embedded magnetic field.

Can the crust experience runaway deformation to dynamical rates, comparable to (shear wave speed)/(crustal scale height)? Or are its deformations restricted to relatively slow plastic creep? We show that a slow increase in applied stress, sourced by distant Maxwell stresses, can drive rapid creep within a small part of the crust. The localized rapid creep does not feed back immediately on the imposed global stress. The size of this zone, in which inertial forces start to counterbalance the Lorentz force and elastic force, depends on the growth rate of the applied stress. Locally the dense solid crust of a neutron star can momentarily behave like a fluid.

What triggers a growth in creep rate? Molecular dynamics simulations of dense Coulomb solids (Chugunov & Horowitz 2010; Hoffman & Heyl 2012) indicate that the creep rate $\dot{\varepsilon }$ is much more sensitive to changes in stress σ than it is to changes in temperature T (Figure 1). This opens up the possibility of a chain reaction within the solid strain pattern, in which creep in one part of the magnetar crust changes the stress in another part, thereby triggering even faster creep.

Figure 1.

Figure 1. Sensitivity of plastic creep rate in a Coulomb solid to changes in applied stress and temperature. Creep rate is determined by Equation (1), taken from Chugunov & Horowitz (2010). The mass density is 1014 g cm−3, with nuclear composition taken from Negele & Vautherin (1973). The range of stress plotted corresponds to temperature 0.99Tmelt (on the left) decreasing to 0.02 Tmelt (on the right). Creep rate is much more sensitive to changes in stress than temperature, except when T is close to the melt temperature Tmelt.

Standard image High-resolution image

What is the time profile of the dissipation? The net plastic dissipation rate in the crust peaks at a wide range of values, depending on the details of how stresses are configured. In some cases, we find energy release profiles similar to those observed in giant magnetar flares: very fast (millisecond) rise, ∼0.1 s duration, followed by an extended power-law decay of plastic creep.

How does fast creep couple to the magnetosphere? Magnetic twist or shear can be directly ejected on a vertical shear wave timescale, consistent with some early theoretical ideas (Thompson & Duncan 1995). This happens preferentially in zones where plastic flow has previously deposited a horizontal magnetic field. The efficiency of this process, as measured by the escaping Poynting flux, is greatly enhanced by the presence of a background magnetic twist extending into the magnetosphere. The energy that powers an SGR burst is not primarily stored in an elastic wave, which only couples relatively slowly to magnetospheric modes (Blaes et al. 1989; Link 2014). The ejected magnetic displacement may or may not oscillate in the magnetosphere, depending on the length of the excited field lines.

What is the origin of the QPO activity seen in large magnetar flares? Stresses are redistributed throughout the magnetar crust by elastic forces, modified by an embedded magnetic field (Duncan 1998; Piro 2005). Changes in the stress balance in one part of the crust will have a larger influence on a neighboring part than changes in temperature, both because of the greater sensitivity of creep rate and because elastic waves propagate much more rapidly than temperature fluctuations. We find that modulation of the local creep rate by a global elastic mode can feed back positively on the global mode.

How does a temperature increase influence subsequent yielding? As has been explored in models of terrestrial faults (Turcotte & Schubert 2002), in situ plastic heating raises the rate of plastic creep; it also enhances the duration and extent of the creep, by reducing the equilibrium stress. These effects have been quantified by Chugunov & Horowitz (2010) for neutron star crusts. They are especially important in layers with a strongly temperature-dependent specific heat, e.g., where the temperature is well below the Debye temperature and the dripped neutrons are strongly paired.

Rapid temperature feedback in a magnetar crust does not automatically follow from transport of the magnetic field. Transport in the crust (by electron drift) or in the core (by electron–proton drift) is generally slow compared with the conduction of heat in the crust. As a result, heat deposited by plastic flow may be conducted away faster than creep can grow; growth of the Maxwell stress may be compensated by the solid response.

Beloborodov & Levin (2014) posit that creep has already reached high enough rates for the heat generated to be trapped in a part of the magnetar crust. In a one-dimensional calculation, they consider the conduction of heat away from the yielding zone. This triggers the creation of a propagating front and a continuing release of both elastic and Maxwell stresses. Because heat transfer is needed to keep the process going, energy is released on a timescale intermediate between the local dynamical time and the conduction time. A one-dimensional simulation of Hall drift, in which the vertical electric current is assumed to vanish, produces small intermittent episodes of thermal feedback in the upper crust (Li et al. 2016).

On the other hand, creep triggered by a growing global stress does not depend on sharp temperature gradients. It is also worth emphasizing that a thermal runaway feeds off the local elastic stress, not the local Maxwell stress. The yielding behavior uncovered here cannot be understood in terms of a local "burning" process: the horizontal (xy) component of the embedded Maxwell stress typically increases in response to rapid plastic creep: in other words, the embedded field can act as a "fire retardant." We also find that fast dynamical creep develops more easily if the initial temperature lies below ∼5 × 108 K. This means that the crust of a neutron star has a much "jumpier" response to magnetic stresses at temperatures well below the melt temperature, supporting the conjecture of Ruderman (1991).

What is the role of electron thermal conduction in regulating plastic creep? We find that thermal conduction generally reduces the creep rate, and in some circumstances it can prevent runaway creep. Equilibrium solutions for a plastic deformation "spot" (zero-dimensional) and "fault" (one-dimensional) can be constructed in which the heat generated by plastic deformation is conducted away and/or radiated to neutrinos.

What is the relative importance of core and crustal magnetic fields to "breaking" the crust? Having fixed the magnitude of the magnetic field, one obtains a larger solid stress from an imbalance in the xz and yz components of the Maxwell stress at the crust–core boundary, as compared with the stress locally imparted by magnetic irregularities within the crust (Thompson & Duncan 2001). The ratio of the two is R/δRc, where R is the stellar radius and δRc the crustal thickness. This Maxwell stress imbalance has a nonlocal effect, increasing the solid stress at distant points in the crust (Figure 2).

Figure 2.

Figure 2. An imbalance in Maxwell stress BBz/4π between the bottom and top of the magnetar crust drives horizontal stresses that peak at distant locations. The symmetric case shown here contains a narrow zone of plastic creep, which bisects two zones of peak Maxwell stress and horizontal elastic displacement. The problem of plastic creep in the crust is formulated here in terms of a displacement field ${{\boldsymbol{\xi }}}_{B}$, which imparts tilt to an initially vertical (radial) magnetic field. A horizontal magnetic field embedded in the crust and oriented transverse to the plastic flow plays an important role in compensating growth of the applied stress and limiting creep.

Standard image High-resolution image

How important is the role of hydromagnetic instability in driving the largest magnetar flares? Solid stresses in the crust can resist only a limited imbalance between hydrostatic, gravitational, and elastic stresses. We find that giant flare energies larger than ∼1045 erg depend on the onset of a hydromagnetic instability in the magnetar core. We also show that large stress imbalances may develop spontaneously in the core on timescales as short as a day. This is fast enough for the heat deposited in crustal fault zones to remain trapped. A similar effect could be produced by current-driven instabilities in the magnetosphere.

Can fault-like features form in the magnetar crust? We find that if the crust is stressed from below in a slightly anisotropic manner (corresponding, e.g., to a large-scale winding of the core magnetic field), then concentrated zones of strong shear develop naturally (Figure 3). Adding a horizontal magnetic field to the crust does not suppress this effect, but it can modify the width of the shear zone and the peak strain rate within it.

Figure 3.

Figure 3. Map of plastic dissipation in a planar solid with periodic boundary conditions. The dissipation has been integrated in time over a major outburst of total energy ≃1 × 1045 erg and peak duration ∼0.1 s. The black circle marks the first onset of rapid creep, and the black line follows the shifting location of peak creep rate. Red, gold, green, cyan, and blue pixels have energy release >1043 erg, >1042 erg, ..., >1039 erg. The forcing Maxwell stress at the crust–core boundary has power spectrum ${{dB}}^{2}/d\mathrm{ln}k\propto {({k}_{\parallel }^{2}+0.3{k}_{\perp }^{2})}^{-1/2},$ where k/k are the components of the two-dimensional wavevector perpendicular/parallel to the axis along which power is concentrated. This axis is chosen as a 45° diagonal in order to avoid grid-alignment effects.

Standard image High-resolution image

The formation of open cracks is inhibited by the extreme ratio of hydrostatic pressure to shear modulus in the crust of a neutron star (Jones 2003). An earlier analysis by Levin & Lyutikov (2012) started by positing the spontaneous formation of a crack near a surface of maximal stress, which was assumed to run perpendicular to the magnetic field, and followed the acceleration of the magnetized solid near the crack. Yielding was restricted to a narrow layer whose thickness is set by ohmic diffusion, and the energy release was found to be minimal.

The fault-like features presented here are driven by continued stress growth, which, being nonlocally sourced, does not stop as creep begins to pick up. Much of the expansion of creep is associated with a nonlocal redistribution of stress in the neighborhood of plastically deforming zones, which is an inherently two- or three-dimensional effect. As a result, plastic flow develops over a much wider band than the ohmic diffusion length. These processes clearly have a strong influence on the spectral signature of localized heating.

In a thin-shell geometry, the shell thickness defines a characteristic scale for extended linear bands of peak stress, below which the strain field is no longer essentially two-dimensional. We evolve the global solid stress pattern by dividing the crust into $\sim 4\pi {({R}_{\star }/\delta {R}_{c})}^{2}\sim {10}^{4}$ elastic units.

Why are all magnetars not bursting SGRs, and why are SGRs not bursting all the time, given the extreme nonlinearity and temperature sensitivity of the creep rate? We find that shearing of an embedded magnetic field feeds back more strongly on peak creep rate than on the net energy dissipated. The Coulomb solid attains a state of equilibrium creep (at rates intermediate between the dynamical rate and the inverse age of the star) if the magnetic field transverse to the direction of shear reaches a substantial strength, greater than 1015 G at the base of the crust (corresponding to a Maxwell stress ≳0.1 the shear modulus). As this embedded field weakens, the crust becomes more sensitive to slow changes in applied stress. We also find that a tangled (as opposed to laminar) field is most effective at limiting runaway large-scale creep. Hall drift is a plausible source of small-scale magnetic irregularities (Goldreich & Reisenegger 1992), and in this way it may actually limit SGR burst activity.

1.2. Where is the Magnetic Field Anchored?

In contrast with the approach taken here, a number of calculations of magnetic field evolution assume that the magnetic field is anchored entirely in the magnetar crust, setting aside any effects of a poloidal field threading crust/core boundary (e.g., Perna & Pons 2011; Beloborodov & Levin 2014; Gourgouliatos et al. 2016, and references therein). This is done, in part, for reasons of computational facility: one can then neglect the response of the liquid core to changing Maxwell stresses.

It is therefore worth asking whether a purely crustal magnetic field is a realistic configuration in a neutron star. Considerations of the origin of this field suggest that it is not. The progenitor star experiences a series of hydrodynamic instabilities that successively modify the magnetic field, starting with main-sequence core convection and growing in power to the most extreme core collapse phase (Thompson & Duncan 1993). Post-collapse accretion following the brief stalling of a bounce shock deposits some 3–10 times the mass of the eventual solid neutron star crust. This allows a transient activation of the magnetorotational instability in a rarefied mantle (Akiyama et al. 2003), which is followed by more persistent linear winding of the magnetic field as the mantle collapses. Rapid neutrino heating drives buoyant motions of the toroidal magnetic field if the seed field exceeds a critical value (Thompson & Murray 2001), thereby potentially distinguishing magnetars from neutron stars with a combination of slower rotation and weaker seed fields.

The dilution of the core magnetic field by entrainment with outward-drifting superfluid vortices (Ruderman et al. 1998) is impeded in comparison with ordinary radio pulsars by a relatively low ratio of rotational to magnetic energy. Given that the core superfluid transition takes place at age ∼300 yr in a more weakly magnetized neutron star such as the Cas A remnant (Ho et al. 2015), one sees that a magnetar must reach a spin period >1 s before superfluid vortices appear in its core. Heating of the cores of the most active magnetars can delay this transition further; indeed, an active lifetime of ∼103–104 yr was used to infer a core superfluid transition temperature of ≲6 × 108 K (Arras et al. 2004).

We therefore adopt the following simplified magnetic field configuration: the field is strongly tilted or twisted in the core but becomes essentially vertical in the crust. Of course, a hybrid configuration is plausibly attained in magnetars, with the axis of the magnetic field winding tilted with respect to the local direction of gravity. But the chosen configuration is adequate to demonstrate the effects we are interested in.

Empirically there appears to be a strong correlation between the detection of dense episodes of short burst activity from magnetars and their ability to generate a giant flare. This suggests that the same structures responsible for the giant flares also give rise to the shorter bursts. The giant flares, in turn, appear to depend on large-scale unbalanced stresses in the neutron star, which partly motivates our approach that starts with a core magnetic field making a transition out of magnetostatic equilibrium.

1.3. Plan of the Paper

The plan of this paper is as follows. Section 2 analyzes the plastic response of a small patch of Coulomb solid to an externally imposed stress.

Section 3 sets up the problem of plastic flow and elastic response in two dimensions, in response to an inhomogeneous Maxwell stress that is applied at the base of the crust.

Section 4 analyzes the results of global time-dependent calculations of such a stressed model crust. Section 5 presents outburst light curves and shows how the energy dissipation is distributed between magnetic, plastic, and elastic components and how the peak of dissipation circulates around the crust.

In Section 6 we consider the vertical redistribution of stresses in a solid with a strong density stratification, appropriate to a neutron star crust. Fast ejection of magnetic twist is encountered when a solid crust experiences runaway creep in response to an externally applied stress. This provides, we believe, a promising foundation for understanding the short (∼0.1 s) SGR bursts.

The slow damping of free vertical Alfvénic oscillations in liquified patches of the crust is described in Section 7. This suggests that the high-frequency QPOs observed in magnetar outbursts (Strohmayer & Watts 2006) may involve the vertical excitation of the magnetic field in melted parts of the crust (e.g., within heated fault zones).

The interaction between the local Maxwell stresses and global elastic oscillations is addressed in Section 8: we show that the global mode is overstable if the localized stresses have a "patch"-like geometry.

The concluding Section 9 makes a comparison with independent theoretical approaches and summarizes the implications of our work for the physics of magnetars. The two Appendices describe resolution tests of the two-dimensional crustal model, and present a simple example of a plastic spot.

2. Stability of Plastic Flow in Reduced Dimensions

We begin by considering the simplest problem of local plastic flow in response to an externally imposed stress. The nonlocal origin of this stress within a magnetically deformed crust is described in Section 3.

We include the effects of in situ heating, thermal conduction, neutrino cooling, and the shearing out of a magnetic field flowing perpendicular to the axis of creep. This is done first in a zero-dimensional model, followed by planar (one-dimensional) geometry (e.g., a fault).

We take the plastic deformation locally to have Cartesian symmetry, e.g., to be described by a scalar function ${\dot{\varepsilon }}_{\mathrm{pl}}={\partial }_{x}{v}_{y}$. The dependence of ${\dot{\varepsilon }}_{\mathrm{pl}}$ on solid stress σ and temperature T has been calibrated by molecular dynamics calculations. We adopt the following relation from Chugunov & Horowitz (2010):

Equation (1)

Here ${\omega }_{p}={(4\pi {n}_{i}{Z}^{2}{e}^{2}/{m}_{i})}^{1/2}$ is the ion plasma frequency, ni is the ion number density, +Ze and mi are the average charge and mass per ion, a is the mean separation between ions, Γ = Z2e2/(aT) is the melting parameter, $\bar{\sigma }=\sigma /({n}_{i}{Z}^{2}{e}^{2}/a)$, and

Equation (2)

A very similar expression relating creep rate and yield stress is used by Preston et al. (2003) to fit the yielding behavior of metals in the regime of extreme loading (see their Equation (4), which is easily inverted when T ≪ Tmelt).

The melting temperature Tmelt of the crystal corresponds to Γ ≃ 176 (Stringfellow et al. 1990). Throughout this paper, we adopt a low-temperature approximation for the shear modulus $\mu \approx 0.183\,{n}_{i}{Z}^{2}{e}^{2}/a$ (Strohmayer et al. 1991). Finite-temperature corrections to μ are generally less important than the temperature dependence of the creep rate. We also neglect the effects of nonspherical nuclei, which can weaken the crust (Sotani 2011): when considering vertically averaged stresses, we focus on a density of 1014 g cm−3. Then Equation (1) can be rewritten as

Equation (3)

Expression (1) contains the familiar Boltzmann factor ${e}^{-\bar{U}{\rm{\Gamma }}}$ representing thermal activation of dislocation drift within the solid. The second term in the exponential is more important for the initiation of fast creep, in the sense that the creep rate is more sensitive to changes in stress than to changes in temperature (Figure 1). We typically consider a starting temperature well below the melt temperature at the base of the crust. Then Γ starts as a large number where most of the magnetic shear energy is concentrated, and the first term on the right-hand side of Equation (3) dominates.

It should be emphasized that expression (1) is deployed in the regime of extremely slow creep. The prefactor ωp is some 20 orders of magnitude larger than a dynamical rate (∼103 s−1) in a neutron star crust. This means that the crust behaves approximately like a magnetofluid when ${\dot{\varepsilon }}_{\mathrm{pl}}\gg {10}^{3}$ s−1, with the inertial force balancing the Lorentz force. It also should be kept in mind that the molecular dynamics simulations do not currently capture the effects of large-scale networks of defects on the stress–strain relation, e.g., effects such as strain hardening; but in the absence of direct measurements they permit calculations without free adjustable parameters.

One sometimes describes a yielding solid as a very viscous fluid with a temperature-dependent yield stress—see Beloborodov & Levin (2014) and Li & Beloborodov (2015) for considerations of magnetar crusts. In such an approach the strong exponential dependence of creep rate on applied stress is only roughly captured.

2.1. Plastic Spot with Imposed Global Stress and Compensating Local Maxwell Stress

We consider the solid stress to be the sum of local and global contributions,

Equation (4)

The local stress σl compensates a local Maxwell stress that grows in magnitude in response to plastic shear. The applied global stress σg is taken to grow on a timescale tgrowth. It is sourced by Maxwell stresses at a distance Lx ≫ δRc from the plastic zone and varies slowly within the plastically deforming patch.

We work in a frame where the plastic flow velocity v vanishes at a given point in the solid, so that

Equation (5)

Here $\perp \text{and}\,\parallel $ are label coordinates perpendicular and parallel to the direction of plastic flow. Here there is an additional contribution to the solid displacement from the elastic response to an imbalance between σl and ${B}_{\perp }{B}_{\parallel }/4\pi $, which develops as the result of plastic creep. The system moves through a series of equilibrium states in which ${\partial }_{\perp }({B}_{\perp }{B}_{\parallel }/4\pi +{\sigma }_{l}+{\sigma }_{g})\,={\partial }_{\perp }({B}_{\perp }{B}_{\parallel }/4\pi +{\sigma }_{l})=0.$ Hence,

Equation (6)

where

Equation (7)

Combining these equations, one sees the local stress changes in a negative manner with respect to the global stress, so as to partly compensate its growth,

Equation (8)

The equation for σg includes a second term representing the decrease in large-scale elastic stress caused by the net displacement ${v}_{\parallel }\sim \,({\dot{\varepsilon }}_{\mathrm{pl}}+{\dot{\varepsilon }}_{\mathrm{el},l}){\rm{\Delta }}$ $=\,(1-{f}_{\mathrm{local}}){\dot{\varepsilon }}_{\mathrm{pl}}{\rm{\Delta }}$ across the plastic patch of width Δ.

The energy equation is evolved at density 1014 g cm−3, where the global stress is concentrated (Section 3),

Equation (9)

In Equation (9), CV is the specific heat, which is approximated by the contribution of the ion lattice (Chabrier 1993), with the effective ion mass raised by the entrainment of a fraction 0.3 of the superfluid neutrons (Onsi et al. 2008). Then CV ∼ T3 at temperatures below the Debye temperature, meaning that the creep rate is more sensitive to heat input from in situ dissipation at T ∼ 108 K (appropriate for transient magnetars in their low states) as opposed to (5–10) × 108 K (appropriate for persistently bright magnetars). The contribution of dripped neutrons to the specific heat varies rapidly with density: it is strongly suppressed by superfluidity where the pairing temperature is high compared with T. Runaway creep is therefore concentrated within the superfluid part of the lower crust.

Two terms are added to the right-hand side of Equation (9), representing thermal conduction out of the "spot" over a characteristic size (temperature gradient scale) , and neutrino cooling. The thermal conductivity varies weakly with temperature near 1014 g cm−3, Kcond ≃ 1 × 1020 erg (cm s K)−1 (Potekhin et al. 2015). Neutrino emission is dominated by thermal bremsstrahlung and synchrotron radiation by relativistic electrons and calculated using formulae given in Yakovlev et al. (2001). The conductive cooling length is taken to be  = 0.3 km, approximately the vertical density scale height at the base of the crust.

Here ohmic effects can be neglected: we are studying smooth plastic flow on macroscopic scales, driven by a continued growth in applied stress. Sharper gradients in the strain field and magnetic field are encountered if a zero-stress discontinuity is introduced near the stress maximum, and the solid evolves passively thereafter (Levin & Lyutikov 2012). If instead the global stress is allowed to grow, then strengthening the transverse magnetic field does not limit the energy deposition in a shear band, but instead pushes it to larger scales (Section 2.2). The regulating effect of horizontal thermal conduction perpendicular to the fault is addressed in the one-dimensional model of Section 2.3.

A simple numerical experiment evolves the "spot" starting from an initial temperature Tbase, with the initial total stress ${\sigma }_{0}={\sigma }_{g,0}+{\sigma }_{l,0}$ corresponding to a very long creep time (>1 Myr) at this temperature. The transverse magnetic field is normalized by the parameter flocal in Equation (8), and we consider a range of tgrowth for the global stress ranging from about a day up to 105 yr (longer than a typical magnetar lifetime).

Figure 4 shows the evolution of creep rate, total stress, and temperature in the spot. As the applied stress ramps up more quickly, one finds, not surprisingly, a faster and sharper response of the plastic flow in the spot. The initial response, over a timescale (0.01–0.1)tgrowth, is driven by the nonlinear stress dependence. Thereafter, there is a brief interval where thermal feedback takes over, corresponding to the sharpest rise in T. The creep rate ${\dot{\varepsilon }}_{\mathrm{pl}}$ peaks and then subsides as the transverse magnetic field is sheared out, which allows the local Maxwell stress to cancel out further increase in the global stress. Most of the drop in total stress occurs after peak creep is reached, over a timescale (0.1–0.3)tgrowth. Smaller initial temperatures correspond to higher peak creep rates, a consequence of the T3 scaling of the specific heat.

Figure 4.

Figure 4. Top panels: time dependence of creep rate in a small patch of magnetar crust that is exposed to an external stress with growth time 10−2–105 yr, uniformly spaced in ${\mathrm{log}}_{10}{t}_{\mathrm{growth}}$. Curves with faster stress growth lie further to the left. Density is 1014 g cm−3, and background temperature Tbase ranges from 2 × 108 K to 6 × 109 K (corresponding to 0.02–0.6 of the melt temperature, blue to red curves). Oscillations appear when the growth time is comparable to the conductive cooling time. Plateaus in creep rate represent a balance between growth of the global stress and a compensating local Maxwell stress; when an initial spike is present (absent), plastic heating is balanced by neutrino cooling (conductive cooling). Bottom left panel: corresponding change in the total stress. Bottom right panel: corresponding change in temperature. There is a significant reduction in stress as temperature builds up.

Standard image High-resolution image

The stress, temperature, and creep rate all eventually reach a plateau. Dynamically this can arise in two distinct ways. There are two negative terms on the right-hand sides of Equation (8) arising from (i) field line stretching and (ii) direct elastic feedback of creep on σg. The first dominates when

Equation (10)

and one has

Equation (11)

The key feature here is that ${\dot{\varepsilon }}_{\mathrm{pl}}\propto {B}_{\perp }^{-2}$ unless the transverse magnetic field is extremely strong. This gives ${\dot{\varepsilon }}_{\mathrm{pl}}{t}_{\mathrm{growth}}\gg 1$, and there is significant amplification of the creep rate compared with the "seed" rate σ/μtgrowth.

The plateaus also represent a balance between in situ plastic heating and neutrino cooling for all except the longest tgrowth and the lower values of Tbase. One observes some bimodality in final temperature and stress in the latter cases, representing a switch between a warmer, neutrino-cooled state and a cooler state in which plastic heating is balanced by conductive cooling. Large swings in creep rate are observed in a few cases where tgrowth is comparable to 1 yr, the conductive cooling time across  ∼ 0.3 km.

It is also instructive to compare the behavior of this zero-dimensional model with the one-dimensional thermal Hall front proposed by Beloborodov & Levin (2014); this gives a clearer sense of how the different physical effects interact with each other. As is noted by these authors, runaway damping of the solid stress can readily be found by initializing the stress to a high enough value that creep is faster than thermal conduction. But this immediately raises the question of how the minimal creep is established in the first place. Instead of introducing a nonlocal source term for the solid stress (as done here), Beloborodov & Levin (2014) consider an increase in temperature driven by thermal conduction from a neighboring yielding layer. The front that forms is much slower than an elastic wave and involves sharp gradients in temperature and Lorentz force. Another distinction is that the Maxwell stress decreases in magnitude behind such a thermal-magnetic front, instead of increasing so as to cancel growth in the global stress.

Returning to the present model, one infers from Figure 4 that there is a critical growth rate ${t}_{\mathrm{growth}}^{-1}$ for the applied stress above which ${\dot{\varepsilon }}_{\mathrm{pl}}$ approaches the dynamical value Vμ/, where inertial forces must be taken into account and the flow rate must saturate. Here Vμ = (μ/ρ)1/2 is the shear wave speed in an unmagnetized Coulomb solid. Figure 5 shows that the crust is most susceptible to runaway creep at lower temperatures: slower growth of the applied stress (higher tgrowth) is required. Alternatively, strengthening the transverse magnetic field makes runaway creep more difficult. Indeed, at fixed Tbase one can consider there to be a threshold value of flocal leading to explosive creep growth. For example, at Tbase = 5 × 108 K the critical tgrowth drops from 1011 s (comparable to the magnetar lifetime) down to ∼1 s as flocal rises from 0.05 to 0.1.

Figure 5.

Figure 5. Critical value of the growth time tgrowth of the global stress that is applied to a small plastic spot in a neutron star crust, below which the spot experiences runaway plastic creep reaching a peak rate of $\sim 0.1{V}_{\mu }/\delta {R}_{c}\sim 300$ s−1. This is plotted vs. the baseline temperature of the spot at the onset of heating, for a range of values of the magnetic field strength transverse to the direction of plastic creep. The curves are uniformly spaced in ${\mathrm{log}}_{10}{f}_{\mathrm{local}}$ between −2 and 0 (see Equation (8)).

Standard image High-resolution image

2.2. Transition from Zero to One Dimension

The preceding zero-dimensional creep model did not specify the width Δ of the plastic layer. Consider now the case where the applied global stress has a broad maximum at x = 0 and falls away slowly on some length scale Lx ≫ δRc to $| x| \gt 0$ (Figure 2). As the amplitude of the global stress σg rises, the width of the layer that achieves a certain stress grows.

Here we must distinguish between the width Δeq of the layer that achieves the equilibrium creep rate (11), where growth in the applied stress is canceled off, and the (instantaneous) width ΔT of the layer experiencing thermal runaway. The second is much narrower than the first, except when the applied stress grows on subsecond timescales. This means that Δeq provides a first estimate of the width of the dissipative zones that persist for months to decades (or longer) in transient magnetars and the AXPs. On the other hand, we observe rapid and extensive growth of fault-like features during giant outbursts in the global two-dimensional yielding calculations described in Section 4. In this situation tgrowth is shorter than ∼0.1 s, implying that Δeq and ΔT are comparable, as we now show.

If the transverse magnetic field B is very small, then Equation (10) is only satisfied as long as ${\rm{\Delta }}\lesssim {{\rm{\Delta }}}_{\mathrm{eq}}\,={L}_{x}({B}_{\perp }^{2}/4\pi \bar{\mu })\ll \delta {R}_{c}$. Beyond that stage of growth, the long-term, equilibrium creep rate becomes large enough to compensate any further growth of σg, stress growth stalls outside the fault, and Δeq saturates.

On the other hand, if ${B}_{\perp }^{2}/4\pi \gtrsim (\delta {R}_{c}/{L}_{x})\mu $, then magnetic shearing sufficiently limits ${\dot{\varepsilon }}_{\mathrm{pl}}$ that Δeq can reach (or exceed) δRc as the stress builds up. In a neutron star crust with δRc ∼ 0.03R this corresponds to B ≳ 0.6 × 1015 G at a density of ∼1014 g cm−3, well within the magnetar range. Even weaker fields embedded in the lower-density parts of the crust will facilitate faults of width Δeq ∼ δRc.

Now consider ΔT, which we define as the width of the layer experiencing peak strain rate ${\dot{\varepsilon }}_{\mathrm{pl},\max }$. The latter quantity depends in a highly nonlinear manner on initial temperature, tgrowth, and B, so we simply normalize it to the dynamical rate Vμ/Lx ∼ 102 s−1. As is seen in the top left panel of Figure 4, peak creep involves a rapid increase $\delta ({B}_{\perp }{B}_{\parallel }/4\pi )\sim {10}^{-2}\mu $ in the local Maxwell stress (in the case of low tgrowth and low initial temperature Tbase). Taking flocal ≪ 1, the duration of peak creep is seen from Equation (8) to be

Equation (12)

where B = B⊥,15 × 1015 G. The width of the layer experiencing peak creep is

Equation (13)

which is similar to the crustal thickness δRc.

Widening of a fault is also caused by the redistribution of stress into the surrounding solid. This is a two-dimensional effect, which is associated with curvature of the fault and also with the expansion of stress around the tip of a growing fault. It cannot be captured by the (effectively one-dimensional) Cartesian geometry adopted here. A nice example of nonlocal dissipation around a curved fault is shown in Figure 22.

2.3. Plastic Response in One Dimension: Conductive Fault with Imposed Global Stress

We now turn to a steady, one-dimensional model of plastic creep in a fault-like geometry. The maximum of temperature and stress sits in a vertical plane situated (say) at horizontal coordinate x = 0. The injection of heat by plastic creep is balanced by volumetric neutrino cooling and thermal conduction away from the stress maximum. Similar solutions were constructed some time ago by Schubert & Yuen (1978), with a goal of explaining the formation of faults in Earth's lithosphere.

The vertical scale height δRc provides a characteristic scale below which the temperature gradient is mainly horizontal, and vertical conductive losses can be neglected, so that T = T(x). We first consider a weakly magnetized fault (i.e., neglect the calming effect of an embedded magnetic field within the creep layer).

We construct steady profiles of temperature, strain rate, and heat flux, with a range of background temperature T(δRc) = Tbase, from the coupled equations

Equation (14)

Each value of Tbase defines a sequence of solutions with different peak temperatures Tmid ≡ T(0) at the middle of the fault and different total stress σ = σg = const. The conductive energy flux Fcond vanishes at x = 0, corresponding to a solution that is symmetric about the temperature maximum: T(−x) = T(x) and Fcond(−x) = −Fcond(x).

The global stress relaxes on a timescale

Equation (15)

where

Equation (16)

is the net differential creep speed across the fault. Figure 6 shows the sequence of curves relating tcreep to the magnitude of the stress. Each curve is labeled by Tbase, with lower stresses corresponding to higher midplane temperatures. One also sees a tight relation between Tmid and σ, reflecting the extreme sensitivity of ${\dot{\varepsilon }}_{\mathrm{pl}}$ to σ and T.

Figure 6.

Figure 6. Top panel: global creep time (15) in the neutron star crust resulting from plastic flow at a narrow fault vs. the magnitude of the global stress that is applied to the fault. Here the magnetic field transverse to the fault direction is assumed to vanish. Curves correspond to a range of background temperatures outside the fault. Bottom panel: peak temperature in the fault, again for a range of background temperatures.

Standard image High-resolution image

3. Elastic–Plastic–Thermal Evolution of a Neutron Star Crust: Two-dimensional Model

We now construct a global model of yielding in neutron star crusts. Stresses are redistributed horizontally in response to an imbalance in the Maxwell stress at the crust–core boundary relative to the magnetar surface. An effectively two-dimensional description is made possible by the small scale height δRc ∼ 0.03 R at the base of the crust. Our focus is on the feedback of the global two-dimensional stress field on localized yielding.

We first consider departures from hydromagnetic stability in the magnetar core, and then describe the method. Section 4 presents results of time-dependent calculations, and Section 5 provides the implications for magnetar giant flares.

3.1. Slow Evolution of the Core Magnetic Field away from Magnetohydrostatic Equilibrium

The magnetic field within the magnetar core is slowly reconfigured by ambipolar drift and beta reactions (Haensel et al. 1990; Goldreich & Reisenegger 1992; Pethick 1992) over its active lifetime of tNS ∼ 1011 s. The feedback of heating on the rate of modified URCA reactions allows the diffusion time to scale with the lifetime of the star while the core neutrons remain in a normal state (Thompson & Duncan 1996). Slow transport of the magnetic field through a fluid star allows the field to evolve into a nonaxisymmetric configuration (Braithwaite & Spruit 2006) that becomes susceptible to a global hydromagnetic instability.

Many details of how this happens cannot yet be captured by global hydromagnetic simulations, in which numerical diffusion is typically rapid in the outer parts of the model star. Nonetheless, simple considerations of continuity suggest that modes of a very low growth rate must develop during a gradual transition away from stability. We make the simplest approximation of treating the core field as a one-dimensional spring, with a spring constant $\propto {\,\omega }_{0}^{2}$. Stable equilibrium corresponds to ω0 ∼ VA/R, where VA is the Alfvén speed within the core.

Now we let the spring constant evolve linearly through zero over the diffusion time ∼tinst (which initially may be comparable to tNS),

Equation (17)

Near t = tinst, one finds that $| {\omega }_{0}| $ is much smaller than the spin frequency Ω (typically ∼1 s−1 for an X-ray-bright magnetar), meaning that the growth of a hydromagnetic instability is limited by the Coriolis force. Balancing this force against internal magnetic stresses gives ${\rm{\Omega }}| v| \sim | {\omega }_{0}^{2}| {R}_{\star }$. The hydromagnetic displacement at speed v can then be converted to a growth rate

Equation (18)

Instability occurs where ${\omega }_{0}^{2}\lt 0$, but in a fluid star it can develop only when ΓeffΔt > 1, where ${\rm{\Delta }}t=t-{t}_{\mathrm{inst}}$. This gives a minimum growth time

Equation (19)

which is much shorter than the neutron star lifetime, but still longer than the Alfvén crossing time R/VA ∼ 0.1 s. Taking tinst ∼ tNS ∼ 1011 s gives ${{\rm{\Gamma }}}_{\mathrm{eff}}^{-1}\sim 3\times {10}^{4}\,{\rm{s}}$.

These considerations apply to a fluid star. The next question that arises is whether the crustal elastic stress should force the transition from a decaying to a growing mode away from the vanishing core spring constant (ω0 = 0). That would not happen if, at each instant in the evolution of the star, the core could be placed in magnetohydrostatic equilibrium, with the vanishing shear component of the Maxwell stress at the base of the crust. But the magnetar will start at some early time t with the shear stress above the yield stress. When the core is otherwise close to magnetohydrostatic equilibrium, the shear Maxwell stress can only relax to the point where the plastic creep rate has dropped to $\dot{\varepsilon }\sim {\varepsilon }_{b}/t$. Here ${\varepsilon }_{b}\equiv \sigma ({\dot{\varepsilon }}_{\mathrm{pl}})/\mu \sim 0.02\mbox{--}0.1$ is the yield strain at creep rate ${\dot{\varepsilon }}_{\mathrm{pl}}$ as described by Equations (1) and (3).

3.2. Response of the Crust to a Hydromagnetically Imbalanced Core Magnetic Field

This suggests a numerical experiment in which the creep rate starts at a very low value in the plastic parts of the crust. In the initial state, the lattice strain exceeds the yield strain only in small patches. We allow the Maxwell stress to grow very slowly and then see whether the crust is capable of a much faster response. Only then do we vary the sign of the core spring constant.

We idealize the crust as a planar solid layer that is threaded by a uniform vertical magnetic field Bz, which is tilted by some angle at the lower boundary (Figure 2). The horizontal magnetic field at the lower boundary can be expressed in terms of a solenoidal displacement field ${{\boldsymbol{\xi }}}_{B}$ at the base of the core (depth hcore ∼ R)

Equation (20)

with ${\partial }_{x}{B}_{x}^{-}+{\partial }_{y}{B}_{y}^{-}={\partial }_{x}{\xi }_{B,x}+{\partial }_{y}{\xi }_{B,y}=0$. We define ${{\boldsymbol{\xi }}}_{B}$ as the net displacement before any compensating horizontal motion of the crust. The applied Maxwell stress drives both elastic and plastic displacements in the crust,

Equation (21)

in response to which

Equation (22)

Since the crustal shear modulus is only ∼10−3 of the hydrostatic pressure, the displacement field (21) is nearly two-dimensional and incompressible. Therefore, both ${{\boldsymbol{\xi }}}_{B}$ and ${\boldsymbol{\xi }}$ can be expressed using stream functions ΨB, ${\rm{\Psi }}={{\rm{\Psi }}}_{\mathrm{el}}+{{\rm{\Psi }}}_{\mathrm{pl}}$, with, e.g., ${\xi }_{i}={\epsilon }_{{ij}}{\partial }_{j}{\rm{\Psi }}$. Here ${\epsilon }_{{ij}}=\pm 1$ is the antisymmetric symbol and i = x, y.

In this approximation, the wavelength of horizontal variations in ΨB and Ψ is constrained to be larger than or comparable to δRc. We neglect any vertical structure in ${\boldsymbol{\xi }}$, that is, magnetic tilt interior to the crust, whose evolution will be considered in Section 6.

In some calculations, we also include a magnetic field localized to the crust. The seed horizontal magnetic field ${{\boldsymbol{B}}}_{\perp ,0}^{c}$ is uniform and configured independently of the core field. Elastic and plastic displacements of the crust perturb this field and also change the tilt of the core field,

Equation (23)

A tangled component of the magnetic field is easy to implement in Fourier space, as discussed below.

The equation of magnetoelastic equilibrium in the horizontal direction reads

Equation (24)

where the solid stress is

Equation (25)

the horizontal Maxwell stress is

Equation (26)

and

Equation (27)

Only the elastic displacement field enters into the solid stress (25). We neglect forces arising from background gradients in ${{\boldsymbol{B}}}_{\perp ,0}^{c}$, in which case equality holds in the second line after Equation (27).

We integrate Equation (24) in vertical coordinate z from the bottom to the top of the crust, neglecting the exterior4 Maxwell stress ${B}_{i}^{+}{B}_{z}/4\pi $ and defining vertically averaged shear modulus and pressure via $\int \mu {dz}=\bar{\mu }\delta {R}_{c}$ and $\int {Pdz}=\bar{P}\delta {R}_{c}$. This gives

Equation (28)

One observes that the pressure perturbation must be included only when ${{\boldsymbol{B}}}_{\perp ,0}^{c}$ is nonvanishing.

One solves for $\bar{P}$ by taking x- and y- derivatives of the first and second lines in Equation (28) and subtracting:

Equation (29)

Substituting this back into Equation (28) gives

Equation (30)

where

Equation (31)

Uniqueness of the solution implies that Φ vanishes.

The elastic displacement stream function Ψel is easily solved for in Fourier space (kx, ky) in terms of ΨB and the cumulative plastic stream function Ψpl:

Equation (32)

To introduce a tangled magnetic field Bt with no preferred direction, we replace ${({\boldsymbol{k}}\cdot {{\boldsymbol{B}}}_{0}^{c})}^{2}\to {({\boldsymbol{k}}\cdot {{\boldsymbol{B}}}_{0}^{c})}^{2}+{k}^{2}{B}_{t}^{2}$ in Equation (32).

3.2.1. Numerical Procedure

Our procedure therefore is as follows:

1. The computational domain is a square in the xy plane with periodic boundary conditions, and $4\pi {R}_{\star }^{2}/\delta {R}_{c}^{2}\sim {({2}^{6})}^{2}$ pixels. Runs with (25)2 or (27)2 pixels have been performed and show qualitatively similar results. The unit of length L is normalized to give a total area L2 = 4π × (10 km)2. The two-dimensional balance between boundary Maxwell stress and horizontal elastic stress is calculated using δRc = 0.3 km, with density 1 × 1014 g cm−3 and shear modulus corresponding to the Negele & Vautherin (1973) equation of state and gravity g = 2 × 1014 cm s−2. The specific heat is approximated in the same way as in the zero-dimensional model of Section 2, with the contribution from non-superfluid neutrons neglected. The strength and energy of the core magnetic field are calculated using an effective core depth5 hcore = L/(2π).

2. A Maxwell stress is imposed at the lower crust boundary with a power-law spectrum

Equation (33)

and random phases.6 This corresponds to a stress ${B}_{x}{B}_{z}/4\pi \propto {k}^{-2\alpha }$ per logarithm of wavenumber. The parameter η can be chosen to be less than unity to represent an anisotropic Maxwell stress, e.g., a nearly axially symmetric core field with predominantly latitudinal gradients. The basis (k, k) can also be rotated with respect to (kx, ky): we typically choose a 45° rotation when η < 1, so as to eliminate the possibility of grid-alignment effects.

3. The magnitude Az0 of the applied Maxwell stress is gradually raised until the peak creep rate is a small value (∼10−4 yr−1). From this point on, the amplitude is raised linearly in time, dAz/dt = Az0/tgrowth. Here tgrowth is a free parameter, describing the speed of the transition to hydromagnetic instability in the core. We adaptively test the time step, raising or reducing it so as to maintain a maximum fractional change of ∼10−2 in the creep rate in any pixel. As a result, the time step can vary from ∼103 yr or longer down to a fraction of a millisecond. The creep rate, as determined by Equation (1), is capped at 0.1Vμ/δRc, representing the feedback of inertial forces on unbalanced stresses.

4. In each time step δt, the plastic stream function is updated according to the method described in Section 3.3: ${{\rm{\Psi }}}_{\mathrm{pl},{\boldsymbol{k}}}\to {{\rm{\Psi }}}_{\mathrm{pl},{\boldsymbol{k}}}+{\dot{{\rm{\Psi }}}}_{\mathrm{pl},{\boldsymbol{k}}}\delta t$, where ${\dot{{\rm{\Psi }}}}_{\mathrm{pl},{\boldsymbol{k}}}$ is given by Equation (48). In the linear approximation, the update in the deformation field due to plastic flow can be separated from the back-reaction on the applied Maxwell stress and the self-consistent elastic response of the crust. The core magnetic stream function is updated according to ${{\rm{\Psi }}}_{B}\to {{\rm{\Psi }}}_{B}\mp {\dot{{\rm{\Psi }}}}_{\mathrm{pl},{\boldsymbol{k}}}\delta t$. Then the elastic response is recalculated according to Equation (32). The temperature is updated according to Equation (9).

5. The upper or lower sign in the update for ΨB corresponds to a positive or negative core spring constant. The change in core field due to creep supplements the imposed linear rise in the Maxwell stress and mediates a runaway global instability when the core spring constant is negative.

6. We separately test the presence or absence of a horizontal crustal magnetic field. In particular, a tangled crustal field has a strong effect on suppressing runaway yielding (Section 4).

3.3. Formulation of Plastic Flow in a Compact Zone

Here we make use of the ab initio relation ${\sigma }_{b}({\dot{\varepsilon }}_{\mathrm{pl}})$ (Equation (1)) between breaking strain and strain rate to describe inhomogeneous plastic creep in one and two dimensions. The generalization to fully three-dimensional deformations is not addressed.

The molecular dynamics simulations of three-dimensional solids on which this is based take the imposed shear to have planar symmetry and grow at a uniform rate, e.g., $\varepsilon ={\dot{\varepsilon }}_{\mathrm{pl}}t$ with ${\dot{\varepsilon }}_{\mathrm{pl}}={\partial }_{x}{v}_{y}=\mathrm{const}$ (Chugunov & Horowitz 2010; Hoffman & Heyl 2012).

The plastic strain rate ${\dot{\varepsilon }}_{\mathrm{pl}}$ is defined as a positive scalar function of the scalar stress σ. In the planar model described here, it is given by

Equation (34)

where

Equation (35)

The first step is therefore to write the creep rate as a function of scalar strain, ${\dot{\varepsilon }}_{\mathrm{pl}}={\dot{\varepsilon }}_{\mathrm{pl}}(\varepsilon ),$ at each position in the solid.

The full evolution of the strain tensor by plastic creep is described by a single scalar function ${\dot{{\rm{\Psi }}}}_{\mathrm{pl}}$. Here we treat all parts of the crust as effectively plastic, but with exponentially varying creep rate. As long as the deformation rate is low enough that inertial forces can be neglected, the crust moves through a series of quasi-equilibrium states. Then the new elastic equilibrium can be obtained from Equation (32) after the plastic stream function is updated to ${{\rm{\Psi }}}_{\mathrm{pl}}\to {{\rm{\Psi }}}_{\mathrm{pl}}+{\dot{{\rm{\Psi }}}}_{\mathrm{pl}}\delta t.$

The procedure in outline is therefore

Equation (36)

Here the label ${\boldsymbol{x}}$ or ${\boldsymbol{k}}$ refers to whether the step takes place in real space or Fourier space. As a final step, we can derive a scalar strain rate $\dot{\varepsilon }^{\prime} $ directly from ${\dot{\varepsilon }}_{{ij}}$ using Equation (34). Our approach to two-dimensional creep starts by equating $\dot{\varepsilon }^{\prime} =\dot{\varepsilon }$ and working backward.

3.3.1. Inhomogeneous One-dimensional Creep and Crack Formation

Some subtleties are involved in obtaining a self-consistent time evolution equation for Ψpl in a compact domain. Consider first the case of linear creep in cylindrical symmetry with one periodic coordinate,

Equation (37)

At first it might appear straightforward to apply the continuous creep law (1) to the stress profile $\sigma =\mu | {\xi }_{y,\mathrm{el}}^{\prime }| $:

Equation (38)

where $\dot{\varepsilon }={\dot{\varepsilon }}_{\mathrm{pl}}+{\dot{\varepsilon }}_{\mathrm{el}}$ and $\varepsilon \equiv | {\xi }_{y,\mathrm{el}}^{\prime }| $. But an obstruction arises from the fact that the net displacement

Equation (39)

must be invariant under continuous changes in ξy, whereas the integral

Equation (40)

does not vanish in general.

Equivalently, Equation (38) does not self-consistently prescribe the time evolution of the zero-wavenumber mode of Ψ. To see this, we can work in Fourier space,

Equation (41)

We first use the identity ${\varepsilon }_{{xy}}={\xi }_{y,\mathrm{el}}^{\prime }={\partial }_{x}^{2}{{\rm{\Psi }}}_{\mathrm{el}}$ and write the plastic term on the right-hand side of Equation (38) as ${\partial }_{x}^{2}{{\rm{\Psi }}}_{\mathrm{el}}\times {\dot{\varepsilon }}_{\mathrm{pl}}(\varepsilon )/{\varepsilon }_{}$ before Fourier transforming, giving7

Equation (42)

This equation is not satisfied for r = 0. This conundrum exists only if the deformation is continuous. Introducing a dislocation near the location of peak strain would remove Equation (39) as a conserved quantity and restore consistency with the local creep law (38).

3.3.2. Self-consistent Two-dimensional Creep

A different constraint on the creep law is present in a two-dimensional solid with inhomogeneous deformation. Here it is natural to consider aligning the plastic deformation tensor with the stress tensor (Hill 1998)

Equation (43)

This solves the expression

Equation (44)

which is obtained by taking the time derivative of Equation (34). However, the strain tensor is derived from an incompressible displacement field and must therefore satisfy consistency rules. Equation (35) implies that

Equation (45)

The update $\delta {\varepsilon }_{{ij},\mathrm{pl}}={\dot{\varepsilon }}_{{ij},\mathrm{pl}}\delta t$ obtained from Equation (43) over a small time interval δt will satisfy Equation (45) when ${\dot{\varepsilon }}_{\mathrm{pl}}$ is constant, and also when the strain field has a planar or rotational symmetry, but not more generally.

Expressing the ansatz (43) in terms of the time derivative ${\dot{{\rm{\Psi }}}}_{\mathrm{pl}}$ gives two separate equations. Although these cannot be individually correct, because they do not satisfy the consistency rule, we will combine them by calculating the quantity ${\partial }_{x}{\partial }_{y}{\dot{\varepsilon }}_{{xx},\mathrm{pl}}+\tfrac{1}{2}({\partial }_{y}^{2}-{\partial }_{x}^{2}){\dot{\varepsilon }}_{{xy},\mathrm{pl}}=\tfrac{1}{4}{\partial }^{4}{\dot{{\rm{\Psi }}}}_{\mathrm{pl}}$. This gives

Equation (46)

One can check that this ansatz reduces to Equation (42) when the strain field has a planar symmetry, e.g., when only εxy or εxx = −εyy, or some constant linear combination of these components, is nonvanishing throughout the solid.8

We then convert to Fourier space,

Equation (47)

giving

Equation (48)

Similarly to the one-dimensional example discussed above, this equation does not evolve the (r = 0, s = 0) Fourier mode.

The two-dimensional description of the strain pattern must break down on length scales shorter than the crust scale height δRc, and so we impose a cutoff N ∼ R/δRc in Fourier space.

3.4. Plastic and Magnetic Dissipation

Dissipation in the crust and the associated change in internal magnetic energy are calculated as follows. The change in the plastic deformation stream function (48) is converted back to coordinate space and then used to calculate the time derivative of the strain tensor, ${\dot{\varepsilon }}_{\mathrm{pl}{xy}}=\tfrac{1}{2}({\partial }_{y}^{2}-{\partial }_{x}^{2}){\dot{{\rm{\Psi }}}}_{\mathrm{pl}}$, ${\dot{\varepsilon }}_{\mathrm{pl}{xx}}={\partial }_{x}{\partial }_{y}{\dot{{\rm{\Psi }}}}_{\mathrm{pl}}$. The dissipated energy over a time interval δt is $\int {dz}\,2({\sigma }_{{xy}}{\dot{\varepsilon }}_{\mathrm{pl}{xy}}+{\sigma }_{{xx}}{\dot{\varepsilon }}_{\mathrm{pl}{xx}})\delta t\,=$ $4\bar{\mu }({\varepsilon }_{{xx}}\cdot {\dot{\varepsilon }}_{\mathrm{pl}{xx}}+{\varepsilon }_{{xy}}\cdot {\dot{\varepsilon }}_{\mathrm{pl}{xy}})\delta t$ per unit area of crust. The change in core magnetic energy per unit area of crust is ${h}_{\mathrm{core}}\cdot {B}_{i}^{-}\delta {B}_{i}^{-}/4\pi $, where $\delta {{\boldsymbol{B}}}^{-}$ is obtained from Equation (23) by substituting $\delta {\rm{\Psi }}-\delta {{\rm{\Psi }}}_{B}={\dot{{\rm{\Psi }}}}_{\mathrm{pl}}\delta t\,+\delta {{\rm{\Psi }}}_{\mathrm{el}}-{\dot{{\rm{\Psi }}}}_{B}\delta t$. Here ${\dot{{\rm{\Psi }}}}_{B}$ describes the imposed secular growth in the Maxwell stress at the crust–core boundary, and δΨel is the self-consistent elastic response obtained from the sum of $\delta {{\rm{\Psi }}}_{\mathrm{pl}}={\dot{{\rm{\Psi }}}}_{\mathrm{pl}}\delta t$ and $\delta {{\rm{\Psi }}}_{B}={\dot{{\rm{\Psi }}}}_{B}\delta t$ via Equation (32).

4. Results for the Two-dimensional Model

We now examine how the rate and extent of plastic flow depend on the applied core Maxwell stress, as well as on compensating stresses arising from the shearing of a crustal magnetic field. What circumstances lead to outbursts similar to giant magnetar flares?

A typical applied Maxwell stress pattern as adopted in our calculations is depicted in Figure 7. Here the core "spring constant" is taken to be negative, as defined by the sign of ${\dot{{\rm{\Psi }}}}_{B}$ (Section 3.2.1), so that the stress rapidly runs away once yielding begins. A sequence of snapshots reveals the growing area of the zones where the creep rate exceeds a critical value of 10−5 s−1.

Figure 7.

Figure 7. Two-dimensional pattern of yielding in a planar Coulomb solid of characteristic density 1014 g cm−3 that is threaded with a vertical magnetic field Bz = 1015 G and subjected to Maxwell stress BBz/4π from below. Imposed stress is mildly anisotropic (parameter η = 0.3 in Equation (33)) with a 45° tilt from the grid. The grid is 26 cells on a side. In this realization, the core magnetic spring constant is taken to be negative, so that the applied Maxwell stress grows exponentially in time in response to plastic motion in the crust. Mauve, black, blue, cyan, green, gold, and red points show the growth of the zones with creep rate exceeding 10−5 s−1. Figure 3 shows for comparison the result of long-term evolution with a hydromagnetically stable (but imbalanced) core magnetic field. In that case a much more prominent fault-like structure emerges during the long-term plastic-elastic-thermal response of the crust.

Standard image High-resolution image

Much more elaborate evolution is possible when the core spring constant is positive and the Maxwell stress is slowly raised at the core–crust interface. The results of the lower-dimensional yielding models presented in Section 2 suggest that runaway creep will result in the absence of a tangled crustal magnetic field.

Figure 8 shows that the crust passes through a series of rapid creep events, with energies approaching those of magnetar giant flares; their properties are discussed in more detail in Section 5. Here the applied Maxwell stress grows on a 2000 yr timescale and has the same power spectrum as assumed in Figures 3 and 7. The dissipation pattern shown in Figure 3 corresponds to the sixth outburst in Figure 8. The separation between outbursts is several times larger than the rough estimate of 40 yr−1 that is obtained from the detection of three giant flares from the three most active Galactic magnetars over the past ∼40 yr.

Figure 8.

Figure 8. "Light curve" of the integrated plastic dissipation rate in a two-dimensional model crust, for the same power spectrum of applied Maxwell stress that gives rise to the yielding pattern shown in Figures 3 and 7. Core–crust Maxwell stress has a growth time of 2000 yr. Here the embedded crustal magnetic field is set to zero, so the global solid stress is not compensated by local stretching of the magnetic field. As a result, the first pixel to reach the yielding threshold experiences runaway creep, with the stress change being redistributed throughout the crust and leading to rapid yielding at distant sites. The yielding pattern shown in Figure 3 corresponds to the sixth peak in this light curve.

Standard image High-resolution image

There is a direct connection in this model between the rate of giant flares and the quiescent plastic dissipation rate Lpl. Both are inversely proportional to the growth time tgrowth of the applied Maxwell stress. Figure 8 shows Lpl ≃ 1 × 1035 erg s−1 in quiescence, which is comparable to (or in some cases several times smaller than) the observed X-ray output of quiescent magnetars. The normalization of this quantity is determined by the crustal shear modulus (which is obtained using a realistic equation of state and crustal volume; see Section 3.2.1). We find that Lpl is somewhat larger than the rate of change of the energy in the external magnetic field when Bz ∼ 1015 G (see e.g., Figure 11). Hence, closer agreement with the observed burst interval and dissipation rate could be obtained by lowering tgrowth by a factor of ∼0.1.

Including an embedded magnetic field allows the magnetar crust to sustain larger zones with moderately fast creep, over intervals much longer than a giant flare. The time evolution shown in Figure 8 corresponds to a vanishing horizontal field, but similar behavior is encountered if the seed horizontal field ${{\boldsymbol{B}}}_{0}^{c}$ is uniform (e.g., toroidal). Figure 9 compares the yielding pattern following a major outburst for (i) vanishing ${{\boldsymbol{B}}}_{0}^{c}$, (ii) ${({B}_{0}^{c})}^{2}/4\pi =0.1\,\mu $ and the direction of ${{\boldsymbol{B}}}_{0}^{c}$ oriented with the symmetry direction of the applied Maxwell stress (e.g., the direction ∥ in Equation (33), and (iii) the same strength of ${{\boldsymbol{B}}}_{0}^{c}$ but now oriented 45° with respect to the ∥ direction. Fault-like features form in all three cases, with increasing linearity when the crustal field is present and aligned with the core field.

Figure 9.

Figure 9. Comparison of yielding pattern with different strengths and orientation of a uniform (e.g., toroidal) crustal magnetic field Bc. Here Tbase = 2 × 108 K and applied Maxwell stress is moderately anisotropic: η = 0.1 in Equation (33), corresponding to a kinked core toroidal magnetic field. Top left panel: Bc = 0. Top right panel ${B}_{c}^{2}/4\pi =0.1\,\mu $ and both crustal field and dominant symmetry direction of core field oriented 45°. Bottom panel: same Bc but now pointing in the x-direction, at 45° with respect to the core field. Colors label the plastic energy release per pixel in the same way as Figure 3.

Standard image High-resolution image

Introducing a tangled (statistically isotropic) magnetic field Bt to the two-dimensional model, according to the prescription listed following Equation (32), has a more dramatic effect, as is shown in Figure 10. The strength of the nonlinear response was found to depend on background temperature Tbase, before the onset of plastic heating. Here Tbase = 3 × 108 K, and we choose embedded field strengths ranging from ${B}_{t}^{2}/4\pi =0.2\,\mu $ up to 0.6 μ. A giant flare-like outburst results when ${B}_{t}^{2}/4\pi \sim (0.2\mbox{--}0.25)\,\mu $, whereas slower outbursts more similar to those of AXPs are observed for the strongest chosen values of the embedded field.

Figure 10.

Figure 10. Outburst behavior in two-dimensional yielding model with embedded tangled magnetic field of varying strengths and Tbase = 3 × 108 K.

Standard image High-resolution image

These results, along with those of Section 2, suggest that a wide range of outburst behavior can appear in different magnetars with only slightly different magnetic fields and temperatures. For example, lowering Tbase to 2 × 108 K in the preceding calculation broadens the transition from rapid to slow outburst behavior, as measured by the range of Bt over which this transition occurs. Increasing Tbase to 6 × 108 K makes the transition very sharp.

5. Giant Magnetar Flares

A large energy release (∼1045 erg) is observed in our two-dimensional model when runaway creep starts within a single pixel and spreads to other locations in the crust. This energy is partitioned between internal plastic heating and magnetic deformation outside the star, with a compensating decrease in internal magnetic energy. The net deformation in a small subset of 212 pixels can approach δε ∼ 1, corresponding to a plastic energy release >3 × 1043 erg. The total outburst energy represents a sum over many pixels. Indeed, the peak of dissipation repeatedly moves around, as is seen in Figure 3. Some (but not all) of this movement represents repeat bursts, which are produced by the model with a range of energies during the ∼106 s following the main dissipation peak.

Figure 11 shows the various components of the energy shift for several giant-flare-like events that are drawn from the same simulation as Figure 3. The change in external magnetic field is calculated after the fact, but it is not included in the time evolution equation for Ψpl, due to the complications of implementing a realistic external magnetic geometry. We approximate

Equation (49)

where Lmag = 10 km is an effective length of the magnetospheric cavity. On this basis, we expect that 20%–30% of the change in internal magnetic energy is communicated to the exterior in the form of magnetic curl or twist (compare the black to the cyan curve in Figure 11). Slightly more energy could be transferred to the magnetosphere if it were already twisted (see Section 6 for local calculations demonstrating this).

Figure 11.

Figure 11. Relative size of different components of the energy during a sequence of outbursts drawn from the same two-dimensional simulation as the burst in Figure 3 (burst 6). Curves show (i) change ${\rm{\Delta }}{E}_{\mathrm{mag},\mathrm{int}}\lt 0$ in internal magnetic energy (magenta), (ii) energy ΔEplast dissipated plastically in the crust (gold), (iii) energy ${\rm{\Delta }}{E}_{\mathrm{mag},\mathrm{ex}}$ injected into the external magnetic field (black), and (iv) change in crustal elastic energy ΔEelas (blue). Parameters Bz = 1015 G and η = 0.3 in Equation (33), with growth time 2000 yr for the applied Maxwell stress at the lower crust boundary.

Standard image High-resolution image

The two-dimensional pattern of magnetic energy injection outside the star and the corresponding reduction in the core are shown in Figure 12 for the same outburst whose plastic dissipation profile appears in Figure 3. The current flowing through the magnetar surface after the outburst is highly inhomogeneous (see the top right panel of Figure 12). These zones of strong magnetic curl are a promising source of nonthermal X-ray emission.

Figure 12.

Figure 12. Change in external magnetic energy (top left panel), vertical current density (top right panel), and internal magnetic energy (bottom panel) during the same outburst whose plastic dissipation profile appears in Figure 3. The energy scale per pixel is the same (ranging from >1039 up to >1043 erg per pixel).

Standard image High-resolution image

A fundamental point about energy conservation during magnetic field decay is raised by Figure 11. One observes that the magnitude of the change in internal magnetic energy exceeds the sum of the plastically dissipated energy and the change in elastic energy. Some of the energy difference must be deposited in hydromagnetic motions in the fluid core (Thompson & Duncan 2001; Levin 2006), which are implicitly assumed to have damped between successive time steps. Core oscillations are an interesting driver of low-frequency QPO behavior, although the spectrum has not been convincingly established.

To see this, we refer back to the elastic equilibrium defined by Equation (32). Defining L as the horizontal gradient scale for the solid stress σxy, one has

Equation (50)

and the ratio of elastic energy in the crust to available magnetic energy in the core is

Equation (51)

This ratio is always less than unity, and the plastic dissipation therefore does not fully compensate the change in core magnetic energy.

One sees from Equations (50) and (51) that energy injection into the magnetosphere by the yielding magnetar crust is maximized at ${{ \mathcal R }}_{B}=1$, corresponding to Bz ∼ 1 × 1015 G, in the case where the Maxwell stress at the crust–core boundary decreases in response to crustal yielding. Larger energy release is possible when the core experiences a large-scale hydromagnetic instability.

5.1. Time Profile

Figure 13 shows the dissipation profiles of two outbursts (the first and 12th) from the same calculation as in Figure 3. The duration of the main pulse (∼0.1 s) is comparable to the time for unbalanced stresses to redistribute around the star at the limiting creep speed of ∼0.1 Vμ. There is a decaying power-law tail of emission, scaling as t−1 and with total energy ∼10−2 of the main pulse energy, which is powered by continuing plastic creep. This extended dissipation is associated with the growth of a fault (by an increase in length and in width). Figure 14 shows an example of fault growth.

Figure 13.

Figure 13. Top panel: plastic dissipation rate vs. time in the first (black line) and 12th (dotted red line) major outbursts seen in the two-dimensional simulation shown in Figure 3. Bottom panel: rise and decay time of the plastic dissipation.

Standard image High-resolution image
Figure 14.

Figure 14. Lengthening of a fault during a giant yielding event. Here η = 0.1 and Bc = 0, and we use higher resolution (27)2. Top panel: plastic energy dissipated after the first 30 s. Bottom panel: cumulative distribution of dissipated energy, after the dissipation rate drops below ∼1036 erg s−1.

Standard image High-resolution image

The outburst light curve as shown does not include emission from a trapped fireball (Thompson & Duncan 1995) or from current dissipation in the magnetosphere (Thompson et al. 2002; Beloborodov & Thompson 2007). Excess decaying persistent X-ray emission with a similar power-law index (−0.7) and relative energy was observed following the 1998 giant flare of SGR 1900+14 (Woods et al. 2001). Similar power-law behavior is seen following shorter SGR bursts (e.g., Ibrahim et al. 2001; Lenters et al. 2003), but our two-dimensional simulations do not have the resolution to adequately address the time profiles of these lower-energy events.

The outburst rise time is ∼10−3 s, which is comparable to that observed in giant flares. There are also repeated injections of energy near the peak of the dissipation, similar to the actual burst light curves recorded by Terasawa et al. (2005) and Tanaka et al. (2007). Our calculation represents the accumulation and release of internal stresses. Although magnetospheric currents will slightly lag the internal motions, one can expect current-driven instabilities eventually to be triggered, with slightly faster rise times over a scale of ∼10 km (Thompson & Duncan 1995; Lyutikov 2003; Gill & Heyl 2010). The internal crustal dissipation would experience a faster rise if the limiting creep speed were taken to be a larger multiple of Vμ (than 0.1).

Figure 15 provides an explicit demonstration of how the energy release (both internal [black points] and external [red points]) is distributed in a nonlocal manner. (The black lines in Figure 3 only show the shifting position of peak creep.) In fact, the peak of energy release does not necessarily coincide with the initial trigger point.

Figure 15.

Figure 15. Pixel-by-pixel energy release vs. horizontal distance from the trigger site, in burst 6 shown in Figure 3. The threshold for burst onset is set at a creep rate >30 times the recent average in some pixel. Black points: internal plastic dissipation; red points: external change in magnetic energy.

Standard image High-resolution image

The total flare energy repeatably reaches ∼1045 erg (including changes in external magnetic energy) but does not approach the 6 × 1046 erg that was observed from the 2004 giant flare of SGR 1900+14 (assuming isotropic emission; Hurley et al. 2005; Palmer et al. 2005). Larger energies are possible if the magnetar core passes through a transition to hydromagnetic instability. Based on our model, we infer that this was the case for the 2004 flare.

Repeat bursts of intermediate energy appear in these simulations and are also observed following giant flares (e.g., the ∼4 s burst of energy ∼1043 erg detected the day after the 1998 giant flare; Ibrahim et al. 2001). An analysis of the energy distribution of bursts is beyond the scope of this paper; a proper analysis would require introducing physical scales smaller than δRc ∼ 0.3 km. Sizable repeat bursts appear to be too common, in comparison with the limited sample of two well-resolved giant flares. In this regard, it should be kept in mind that the moving locus of peak activity allows the burst to probe crustal zones with differing internal magnetizations and differing susceptibility to rapid creep.

6. Crust–Magnetosphere Coupling

We now consider vertical magnetic gradients within the crust. We calculate their evolution, starting from a state of near magnetoelastic balance and then developing runaway creep in response to an applied global stress. The stable stratification of the stellar crust (Lattimer & Mazurek 1981; Goldreich & Reisenegger 1992) impedes vertical buoyant motion of the magnetic field, as is observed in the early stages of solar flares, but still allows a coupling to the magnetosphere via sideways shear motions. We find that magnetic curl or twist can be ejected into the magnetosphere from deep in the crust in a few milliseconds, much shorter than the durations of most SGR bursts, if a background global stress is itself growing on a ∼0.1 s timescale.

We have described the crust–core Maxwell stress by a two-dimensional horizontal shuffling of vertical field lines. The rigidity of the crust prevents the displacement of the external field lines by plastic creep (Equation (49)) from instantly mirroring the core displacement. We can compare the gradient energy that accumulates in the crustal magnetic field with the energy that would be stored in the magnetosphere if the shuffling of the external field lines were smooth and instantaneous (that is, if effectively the crust were absent). The answer is sensitive to the presence of a background magnetic curl $\delta {B}_{\perp ,0}\sim ({\xi }_{\perp ,0}/{R}_{\star }){B}_{z}$ associated with a radial current flowing across the magnetar surface (Thompson et al. 2002). The additional crustal gradient energy associated with a horizontal displacement ξ is

Equation (52)

per unit area, where ${B}_{\perp }^{c}\sim ({\xi }_{\perp }/\delta {R}_{c}){B}_{z}$. If the same displacement field were immediately to spread to the magnetosphere, then the factor of R/δRc in the first term on the right-hand side of Equation (52) would be absent.

The net result is that trapping of vertical magnetic gradients in the crust will enhance the energy in the horizontal magnetic field by a factor of ∼R/δRc if background curl is absent, but it only provides an order-unity enhancement if the background shear is stronger, $| {\xi }_{\perp }| \lesssim | {\xi }_{\perp ,0}| $ × δRc/R. In this section, we consider how the magnetic field relaxes to a smoother, lower-energy state.

In contrast with the zero-dimensional model described in Section 2, there is no interference between the local stress and the applied global stress, because they are orthogonal. The vertical gradient of BiBz/4π within the crust (here i = x, y) is compensated by a gradient in σiz, and the total stress is

Equation (53)

Here we write explicitly the dominant component of the global stress.

This situation is also distinguished from the spot and fault models of Section 2 by the strong vertical gradients now present in the shear modulus and the yield stress. The question therefore arises as to whether temperature feedback is needed to trigger creep runaway. Here we show that it is not: an upward shift in the envelope of the horizontal magnetic field (initiated by growth of σg) triggers a secondary nonlinear effect. The net horizontal displacement of the vertical field is conserved by plastic creep within the crust, at least until the disturbance reaches the stellar surface. As a result, the ratio of Maxwell stress to yield stress increases as the peak of the horizontal field shifts to shallower layers with lower shear modulus.

It may appear at first sight that the two Maxwell stress terms are subdominant on the right-hand side of Equation (53) when σxy approaches the yield stress, because ${\sigma }_{{xy}}\sim {({R}_{\star }/\delta {R}_{c})({B}_{\perp }{B}_{z}/4\pi )}_{-}$. A stronger embedded horizontal magnetic field is expected to develop in plastic fault zones that have a net displacement >δRc along the fault. This suggests that vertical energy leakage is localized in the yielding zones, which, as we show in Section 8, has interesting implications for magnetar QPO activity.

6.1. One-dimensional Vertical Evolution

We consider a vertical one-dimensional test problem that illustrates the dynamical ejection of a horizontal magnetic field. The crust is threaded by a uniform vertical magnetic field Bz = 1015 G that is bent by a variable amount in the horizontal x-direction, ${B}_{x}(z)={\partial }_{z}{\xi }_{B,x}{B}_{z}$. Our method assumes that the crust is in a state of slow plastic creep at the onset of the calculation: the strain rate ${\dot{\varepsilon }}_{\mathrm{pl}}$ is taken as a dynamical variable, and the solid stress is computed as a function of ${\dot{\varepsilon }}_{\mathrm{pl}}$. As a result, the calculation applies to parts of the crust, such as the shear bands described in Sections 2 and 4, in which the Maxwell stress BxBz/4π has been raised close to the solid yield stress.

We show that the timescale for the ejection of magnetic shear is at least 30 times shorter than the growth time of the global stress that triggers it. The zero-dimensional creep model of Section 2 suggests that the result will be qualitatively similar if the local Maxwell stress lies initially below the yield stress and runaway creep is preceded by a brief interval of growth in the total stress.

The creep rate is initialized to a uniform value at all depths z below the magnetar surface. (The results turn out to be similar when creep is relatively suppressed in the upper parts [lower densities] of the computational domain.) Therefore, the Maxwell stress BxBz/4π saturates the yield stress σb (Equation (3)). The corresponding transverse field, shown in Figure 16, equals Bz only at a particular depth. Everywhere else ${B}^{2}/4\pi =[({B}_{x}/{B}_{z})+({B}_{z}/{B}_{x})]{\sigma }_{b}\gt 2{\sigma }_{b}$, meaning that the field needed to induce plastic creep is generally stronger than (8πσb)1/2 in the absence of an addition global stress.

Figure 16.

Figure 16. Vertical profile of a neutron star crust that starts in a slowly deforming magneto-elastic state: creep rate $\dot{{\varepsilon }_{\mathrm{pl}}}=0.003$ s−1 (independent of depth) and Bz = 1015 G. Right panel: ratio of transverse to vertical magnetic field Bx/Bz obtained from the steady solution to Equation (56) using stress formula (3). Left panel: available magnetic energy compared with elastic energy, as a function of depth z. The crust model is described in the text.

Standard image High-resolution image

The vertical dependence of crustal density, melting temperature, and shear modulus are obtained by fitting the data from nuclear equation-of-state calculations, both below (Negele & Vautherin 1973) and above (Haensel et al. 1989) the neutron drip point, for a surface gravity g = 2 × 1014 cm s−2. Smooth fitting functions are applied to the density and shear modulus profiles ρ(z), μ(z), which allows us to avoid complications associated with discrete jumps in these variables.

Figure 16 also shows the ratio between ${B}_{x}^{2}/(8\pi )$ and the initial elastic energy density $\tfrac{1}{2}{\sigma }_{b}^{2}/\mu $, as a function of depth in the crust. The magnetic field is the dominant energy reservoir near the base of the crust, where most of the potential energy is concentrated.

The evolution away from the initial slow creep state can be encapsulated in two variables

Equation (54)

Freezing of magnetic flux in the plastically deforming crust implies that

Equation (55)

so that v is the combined transverse velocity of magnetic field and crustal material. The transverse component of the relativistic Euler equation reads

Equation (56)

The magnetic field dominates the effective inertia in the upper crust, ${\rho }_{B}=\rho +{B}_{z}^{2}/4\pi {c}^{2}$, so that the Alfvén speed ${V}_{{\rm{A}}}\,={B}_{z}/{(4\pi {\rho }_{B})}^{1/2}$ is limited to the speed of light c. Figure 17 shows the vertical dependence of VA and transverse shear wave speed ${V}_{\mu }=\sqrt{\mu /{\rho }_{B}}$. The density field is extended to low values ($\rho \ll {B}_{z}^{2}/4\pi {c}^{2}$) where electromagnetic inertia dominates. The initial equilibrium state is

Equation (57)

where ${\chi }_{0}\sim { \mathcal O }(0.1)$ is the constant horizontal magnetic field (representing a shear or twist) that is imposed by the magnetosphere.

Figure 17.

Figure 17. Vertical profile of Alfvén speed VA (red thick line) and its fitting curve (black dashed line) and of transverse shear wave speed Vμ (blue thick line) and its fitting curve (orange dotted line). In the upper part of the crust Vμ ≪ VA ≃ c.

Standard image High-resolution image

Because we are studying the transition between magneto-elastic equilibrium and a dynamical state of the crust, we cannot assume that at each time step the crust relaxes to magneto-elastic balance. Therefore, we approximate ${\partial }_{z}v={\dot{\varepsilon }}_{\mathrm{pl}}$ and use Equation (3) to write

Equation (58)

The yield stress is insensitive to small changes in ${\dot{\varepsilon }}_{\mathrm{pl}}$ following Equation (3).

Near the top of the crust, μ is relatively small and VA ∼ c ≫ Vμ. This means that translations of the upper crust follow passively a magnetic perturbation evolving below. This is especially clear in the case of horizontal shear deformations, e.g., a translation of the upper crust in the y-direction that is differential in x. These involve an approximate balance between the inertial force and the Lorentz force; during the first stages of the deformation, the plastic flow simply follows the magnetic field. In some circumstances the upper parts of the magnetar crust may actually be in a fluid state, but the melt depth of a quiescent magnetar is typically shallow enough that VA ≃ c, and the solid-to-fluid transition has a small influence on the dynamics.

We impose an outgoing wave boundary condition at the upper boundary,

Equation (59)

This choice represents an external magnetic cavity large enough to sustain a full oscillation and is done in part to avoid dealing with complications associated with magnetospheric dissipation. In the deeper part of the crust VA < Vμ. We keep the evolution short enough that the initial magnetoelastic equilibrium is only slightly disturbed near the lower boundary, and therefore we simply assume that the magnetic field is pinned there with constant Maxwell stress,

Equation (60)

To avoid numerical instabilities, the uniform creep rate ${\dot{\varepsilon }}_{\mathrm{pl}}$ is smoothly set to zero in a thin layer near the lower boundary. It is also essential to smoothly interpolate the sgn function in Equation (58), which otherwise is discontinuous when ∂zv changes sign:

Equation (61)

Here ${\dot{\varepsilon }}_{0}$ is a small parameter characterizing the thickness of the transition. With this modification, Equation (56) takes a form similar to a heat diffusion equation with a spatially variable thermal conduction coefficient.

Next, an external stress σxy is smoothly added to the system, representing, e.g., a horizontally propagating elastic wave. This means that, at a fixed vertical creep rate, the vertical stress is reduced to

Equation (62)

Strictly speaking, ${\dot{\varepsilon }}_{\mathrm{pl}}^{2}={({\partial }_{z}v)}^{2}+{\dot{\varepsilon }}_{{xy}}^{2}$, where the second term is the horizontal strain rate. As the timescale for horizontal stress growth is longer by a factor of ∼R/δRc, we approximate ${\dot{\varepsilon }}_{\mathrm{pl}}\approx | {\partial }_{z}v| $ during the one-dimensional evolution.

Collecting Equations (55)–(62), we obtain a complete system to model the growing stress imbalance and vertical hydromagnetic wave ejection.

6.2. Numerical Results

Figure 18 shows the result of introducing an external "driver" wave stress

Equation (63)

The radial eigenfunction f(z) reproduces the ${}_{2}{t}_{0}$ global elastic mode (McDermott et al. 1988), which has a frequency $\omega \simeq {2}^{1/2}{\bar{V}}_{\mu }/{R}_{\star }$, where ${\bar{V}}_{\mu }$ is the vertically averaged shear wave speed. To obtain numerically clean results, we take a driver amplitude εxy = 0.06. The background creep rate is $\dot{{\varepsilon }_{\mathrm{pl}}}$ = 1 s−1, which is significantly slower than the ensuing evolution. The smoothing parameter ${\dot{\varepsilon }}_{0}$ in the stress formulae (61) and (62) is necessarily lower than this, and it is taken to be ∼0.1 s−1, as limited by numerical precision.

Figure 18.

Figure 18. Top panels: evolution of $\chi -\chi {| }_{t=0}$ (left) and v (right) at time 0 ms (blue), 1 ms (orange), 2 ms (magenta), 3 ms (red), and 4 ms (black). Bottom panels: evolution of vertical Poynting flux measured at the star surface (left), and integrated surface energy density (kinetic + magnetic) during the first 15 ms (right). The maximum in kinetic energy roughly coincides with the maximum of outgoing energy flux, both being driven by unwinding of the initial magnetic curl.

Standard image High-resolution image

The net result is a burst of upward Poynting flux SP that builds up significantly over a few milliseconds, tapping the horizontal magnetic field χ. The Poynting flux is proportional to χ,

Equation (64)

and is significantly enhanced in the upper crust if the shear has a baseline value χ0 that is imposed by external magnetic twist or shear. The net electromagnetic energy ejected to the magnetosphere is

Equation (65)

where Δx is the net horizontal displacement of the field lines. Taking χ0 = 0.1, this works out to about 7% of the total energy loss over the first 10 ms of the evolution. The majority of the change in magnetic gradient energy is dissipated within the crust, at a rate ${\sigma }_{b}| \dot{{\varepsilon }_{\mathrm{pl}}}| $ per unit volume. Numerical tests show that this efficiency is insensitive to the amplitude of the driver elastic wave.

The magnetic tilt fluctuation is small near the stellar surface, χ ≃ v/VA ∼ v/c, and consequently the plastic creep rate is also suppressed, $\dot{{\varepsilon }_{\mathrm{pl}}}\sim \chi /(10\,\mathrm{ms})\sim v/(c\,\cdot \,10\,\mathrm{ms})$. Combining this with the relatively small shear modulus, we find negligible plastic heating at shallow depths. The crust may behave like a fluid even while it is still in the solid phase.

What happens to the magnetic pulse after it is ejected from the crust depends on details such as the length of the magnetospheric flux bundle and the horizontal twist structure, which is not captured by this one-dimensional model. Effectively we have taken an infinitely long magnetospheric cavity, so that reflection of the outgoing "wave" can be neglected. This corresponds in a dipole geometry to a field line extending to a maximum radius ${R}_{\max }\sim \tfrac{1}{3}c\cdot 3\,\mathrm{ms}\sim 30\,{R}_{\star }$. Here we make the implicit assumption that twist ejection is accompanied by enough pair creation or particle ejection to sustain the MHD approximation in the magnetosphere. Then the twist propagates with a speed VA ∼ c over a cavity of length ∼3Rmax. In more compact parts of the magnetosphere, a quasi-static but strongly localized twist is established, which probably is subject to secondary current-driven instabilities. Details of the subsequent damping are not investigated here.

7. Trapped Alfvén Waves in Liquified Patches

Narrow zones of strong plastic flow experience melting in our two-dimensional simulations. Figure 19 shows the temperature profile just after the peak of the outburst shown in Figure 3. We now consider the implications of this result for the damping of vertical magnetic gradients, following the one-dimensional approach of Section 6.

Figure 19.

Figure 19. Temperature profile right after the peak of the outburst whose dissipation profile is shown in Figure 3. Pixel colors red, gold, green, cyan, and white correspond to T > 0.5Tmelt, (0.3–0.5)Tmelt, (0.2–0.3)Tmelt, (0.1–0.2)Tmelt, and (0–0.1)Tmelt, respectively, at density 1014 g cm−3.

Standard image High-resolution image

The energy needed per unit volume to melt the crust started at a temperature ≪Tmelt is

Equation (66)

(Here we include the specific heat of the ions as calculated by Chabrier 1993 and assume that the specific heat of dripped neutrons is suppressed by superfluidity.) This compares with an elastic energy at the yielding point

Equation (67)

Thus, when the crust starts relatively cold, it must continue to be deformed through ∼5–10 times the yield strain in order to melt, increasing to ∼30–102 times when it starts close to melting. Such a large relative displacement is possible if the solid stress is sourced by distant Maxwell stresses, and it occurs in the pixels in Figure 15 with dissipated energy ≳1043 erg.

Let us now consider the evolution of an inhomogeneous horizontal magnetic field in such a melted layer. The corresponding time evolution equations are Equations (55) and (56) with σxz set to zero. Now the magnetic field provides the only restoring force.

The boundary condition Equation (59) is still valid, and at z = δRc we continue to impose v = 0. To study the ejection of a trapped Alfvén wave, we prepare initial data such that v(z, t = 0) = 0 and

Equation (68)

Here $d\simeq 1.1$ km is the depth of the crust. Any constant background twist χ0 can be added to χ to obtain a new solution. In Figure 20, we show the numerical evolution using the above initial data. Within the first 0.12 s, the Alfvén wave packet bounces within the crustal cavity for N ∼ 30 cycles. The quality factor extracted from the evolution of the total energy is

Equation (69)

The energy loss shown in the bottom right panel of Figure 20 is consistent with integrating the flux in the bottom left panel. Such a high quality factor originates in the strong stratification of fluid density. The density scale height lρ decreases toward the magnetar surface (z = 0), providing a turning point for an upward-propagating Alfvén wave where ${{kl}}_{\rho }\,=\omega {l}_{\rho }/{V}_{{\rm{A}}}\propto \rho {l}_{\rho }\sim 1.$ A similar effect is seen in the tunneling of an ideal elastic wave out of a magnetized neutron star crust (Blaes et al. 1989; Link 2014).

Figure 20.

Figure 20. Evolution of twist and kinetic energy in a melted crust, showing decaying periodic behavior. Top panels: evolution of χ (left) and v (right) at time 0 ms (blue), 0.5 ms (orange), 1 ms (magenta), 1.5 ms (red), and 2 ms (black). The initial profile of χ is given by Equation (68) and initially v = 0. Bottom panels: evolution of vertical Poynting flux measured at the star surface, and integrated surface energy density (kinetic + magnetic) during the first 0.12 s. The time-integrated Poynting flux balances the loss of horizontal magnetic energy from the resonant cavity.

Standard image High-resolution image

This trapped Alfvén wave, being slowly damped and strongly periodic, is a good candidate for the mode underlying high-frequency QPO behavior. The frequency of such a standing wave will exceed ∼1 kHz if it has radial nodes, or if the lower part of the crust remains solid, thereby reducing the Alfvén crossing time of the resonant cavity. Coherence of the excited mode probably depends on strong two-dimensional localization in the crust, which may be possible if the local dissipation rate is very high.

A background twist χ0 outside the star modifies the outgoing Poynting flux in the following way. The first half oscillation inside the star supplies an additional component of the time integral SPdt. Thereafter, the contribution from the term in SP proportional to 0 averages nearly to zero.

8. Overstability of Global Elastic Modes: Interaction with Plastic Patches

We now consider how a global elastic mode interacts with plastically deforming zones in the magnetar crust, explaining how it may gain energy from them. We previously showed that a static global stress enhances the creep rate in localized plastic "spots" and "faults" (Section 2). A growing global stress triggers a localized flow of Poynting flux into the magnetosphere (Section 6). An oscillating elastic mode will produce a periodic modulation of this dissipation; at the same time, it can feed off the collective shear or vortical motion of the plastic zones.

To start, we idealize each plastic patch as a circular spot of characteristic radius rp, which we may define as the radius of maximal stress and creep rate. Each spot is much smaller than the wavelength of an imposed elastic wave, e.g., rp ≪ R (Figure 21). Creep within a spot is driven at a rate ${\dot{\varepsilon }}_{\mathrm{pl},0}$ by a Maxwell stress ${({B}_{\phi }{B}_{z}/4\pi )}_{-}$ at the lower crust boundary. Outside the spot the strain field is source free and satisfies the wave equation

Equation (70)

We continue to work with a vertically averaged strain field, and here we adopt cylindrical coordinates centered on the spot. Then ${\sigma }_{\phi \varpi }=\bar{\mu }\varpi {\partial }_{\varpi }({\xi }_{\phi }/\varpi ).$

Figure 21.

Figure 21. Global elastic response of the crust to enhanced yielding in localized plastic zones. Secondary waves are superposed on the original wave, and their cumulative effect is measured at a displaced position "O." This introduces a damping/growth term into the elastic wave dispersion relation, Equations (79) and (84).

Standard image High-resolution image

The plastic flow within a spot is vortical, and at constant stress the elastic deformation outside the spot is independent of time, satisfying ${\varpi }^{2}{\sigma }_{\phi \varpi }=\mathrm{const}$ and ${\xi }_{\phi }\propto 1/\varpi $. But when the stress is increasing or decreasing, the elastic solid experiences a net circulation

Equation (71)

An explicit construction showing this is given in Appendix B. Here $\langle X\rangle $ denotes the average ${(2\pi )}^{-1}\int X(\phi )d\phi $ over the azimuthal angle ϕ centered on the spot.

The feedback of many spots on the global strain field now follows easily. Given a spot surface number density np and filling factor ${f}_{p}\sim {n}_{p}\pi {r}_{p}^{2}$, the vorticity equation gives

Equation (72)

For simplicity, each spot is assumed to have an identical internal stress pattern, except for a relative phase shift between spots at different positions.

It is essential here to include the response of the Maxwell stress to accelerated creep associated with the imposed wave stress. Consider a planar wave strain field ${\xi }^{y}(x-{\bar{V}}_{\mu }t)$ interacting with a single plastic spot. Here x is a Cartesian coordinate and the spot is centered at coordinate x1. Then we write

Equation (73)

where the coefficient Kσ is determined below. Equation (72) becomes

Equation (74)

This represents negative damping of the strain field, with growth rate

Equation (75)

Here ${\dot{\varepsilon }}_{\mathrm{pl},0}$ is the peak background strain rate in the spot.

To evaluate the coefficient Kσ, we need to consider the response of the plastic flow to the additional stress ${\sigma }_{{xy}}^{w}$ of the elastic wave. Here our experience with the vertical ejection of magnetic shear or twist (Section 6) is useful. Writing the total stress $\sigma ={({\sigma }_{{ij}}{\sigma }_{{ij}})}^{1/2}$ within a spot in Cartesian coordinates and linearizing, the change in stress is $\delta \sigma ={\sigma }_{{xy},0}{\sigma }_{{xy}}^{w}/{\sigma }_{0}$. The peak creep rate, as given by Equation (1), rises to

Equation (76)

The change in stress in the plastic spot depends crucially on the sign of the derivative $\partial \sigma /\partial {\varepsilon }_{\mathrm{pl}}$, where εpl is the cumulative plastic strain:

Equation (77)

A positive sign corresponds to a locally unstable magnetoelastic equilibrium. Then, in the absence of the wave, the elastic stress surrounding the plastic spot experiences persistent growth, which accelerates in response to the additional wave stress. Working in the linear regime, but neglecting the background creep, we recover

Equation (78)

The angular average does not vanish as long as the spot has some nonaxisymmetric structure. A rapidly propagating crack has the same property: it depends on a concentration of stress around the crack tip, and as such it can only appear if the yielding pattern lacks rotational symmetry.

The same result is obtained by a direct construction. We introduce a damping term into the wave equation for the global elastic mode,

Equation (79)

which has the solution ${v}_{y}={v}_{y}^{+}(x-{\bar{V}}_{\mu }t)+{v}_{y}^{-}(x+{\bar{V}}_{\mu }t)$ in the absence of damping. Treating the damping term as a perturbation, we consider the change δvy in the velocity field at some position x2 due to spot rotation within a zone of thickness δx near the position x1. Then δvy satisfies

Equation (80)

where x± = x ± ${\bar{V}}_{\mu }t$. Integrating once with respect to x+ and noting that δx = 2δx at constant x+, we obtain

Equation (81)

The velocity perturbation is obtained by linear superposition. The imposed elastic wave is assumed to be periodic with frequency ω and wavenumber $k=\omega /{\bar{V}}_{\mu }$. The velocity response around each spot has the same frequency. Hence, the solution to the wave Equation (70) outside a single spot is

Equation (82)

The Hankel function ${{iH}}_{1}^{(1)}(x)\simeq 2/\pi x$ at small argument; hence, v0 ≃ kΓ/4.

Following Figure 21, the collective contribution to δvy at x2 from all the spots between x1 and ${x}_{1}+\delta x$ (denote ${\rm{\Delta }}x\equiv {x}_{2}-{x}_{1}$) is

Equation (83)

Identifying this with Equation (81) and substituting for v0, we find growth at a rate

Equation (84)

We conclude that a global elastic mode can experience "super-radiant" amplification by scattering off localized plastic spots in the magnetar crust, if a localized Maxwell stress driving creep in these spots increases with displacement from the initial equilibrium. In other words, parts of the magnetar crust must be in a hydromagnetically unstable state.

The dissipation rate in these plastic zones is modulated by the global mode, which feeds off part of the magnetic energy liberated. Some of this energy is transported into the magnetosphere (Section 6). The power output from the boundary of a single spot is

Equation (85)

and the power output per unit area is (with ${E}_{w}\sim \delta {R}_{c}{\sigma }_{w}^{2}/\bar{\mu}$ and ${\bar{\mu}}^{-1}\partial {\sigma }_{\phi \varpi }/\partial {\varepsilon }_{\mathrm{pl}}={\rm{O}}(1))$

Equation (86)

The corresponding energy growth per cycle r ∼ γP = γ2π/ω is

Equation (87)

Temperature growth may impose a negative feedback on super-radiant amplification of elastic waves, because it implies a decrease in the equilibrium Maxwell stress that can be sustained inside a plastic spot.

We have found that the geometry of the plastic zones can have a strong impact on the feedback process. One can show that an extended and connected (e.g., equatorial) fault is capable of damping a shear wave significantly within one wave cycle.

9. Discussion and Comparison with Other Theoretical Approaches

We have described a quantitative approach to magnetar activity whose physical ingredients are (i) departures from magnetohydrostatic equilibrium in the core, combined with (ii) the nonlinear and temperature-dependent elastic and plastic response of the crust. A global treatment of the system is essential: without it, one cannot begin to understand the temporal and spatial distrbutions of magnetar activity.

This framework allows us to make constrained statements about properties as diverse as the rise times and durations of super-Eddington X-ray bursts, the energies of giant flares and waiting times between them, the origin of QPO activity and its connection with the triggering mechanism of SGR bursts, the mechanism by which magnetic energy in the magnetar crust and core is transmitted to the magnetosphere, the degree of melting and heating of the magnetar crust during a burst, the surface covering fraction and pattern of the dissipation, repeat burst activity, delayed energy release that is manifested as afterglow emission with a power-law time dependence, and the distribution of electric currents that power the persistent nonthermal X-ray emission of magnetars. We are also able to shed some light on why super-Eddington bursts are observed only rarely in all magnetars, and at an insignificant level in many of them, even though all magnetars appear to be strong nonthermal sources with active magnetospheres and strong torque noise.

The yielding crust of a magnetar reveals itself to be a very nonlinear system in the sense that relatively slow variations in one location can trigger much faster variations in another. Variability emerges over a wide range of timescales, which turns out to be computationally challenging in the case where runaway creep is limited by the stretching of an embedded magnetic field. The first onset of strong dissipation in a single pixel that is found in our two-dimensional simulations does not realistically represent the underlying processes that first trigger dissipation on a macroscopic scale (e.g., over an area $\sim \delta {R}_{c}^{2}$). However, once rapid creep is established in a small but macroscopic part of the magnetar crust, these simulations show how intricate dissipative structures can develop and extend far from the initial yielding site.

In the case of a hydromagnetically unstable core, it was not possible to follow the post-outburst response (e.g., to measure the time profile of the extended dissipation) since we lacked a plausible prescription for restoring hydromagnetic stability after the first ∼0.1 s. This underlines the need for a more fundamental understanding of the normal modes of the magnetar core near the threshold for hydromagnetic instability.

We close by comparing our results and conclusions with other theoretical approaches and outlining some open problems.

9.1. Nonlocal Triggering of Rapid Creep

Stress buildup and release in the magnetar crust is a nonlocal phenomenon, in good part due to the small aspect ratio of crustal thickness to stellar radius. A yielding criterion involving a local balance between Maxwell stress and solid stress (Thompson & Duncan 1996; Perna & Pons 2011; Beloborodov & Levin 2014; Lander et al. 2015) does not offer an adequate description of runaway creep at dynamical rates. On the observational side, spectral analysis suggests that some slower transient behavior seen in the AXPs involves global response of the crust to local dissipation (Woods et al. 2004). The nonlocal yielding process described here also offers a framework for investigating small precursor events that have been detected shortly before giant magnetar flares (Hurley et al. 2005).

9.2. Overstability of Global Elastic Modes

A consequence of the preceding item is that a standing elastic wave will modulate the creep rate at discrete plastic spots within the crust. The elastic wave becomes overstable if the covering fraction of the plastic spots is large enough, and if the crust is hydromagnetically unstable within these zones. The oscillation is not excited at a discrete time, thereafter passively decaying, but is continually reinforced.

Magnetar QPOs are therefore identified with periodically forced yielding in concentrated shear zones. This allows the emission zone of the modulated X-rays to be fairly compact, smaller than or comparable to R. Indeed, the strong rotational modulation detected in the QPO power (Strohmayer & Watts 2006) implies an occultation of the emission zone by the star and requires a compact size. A persistent oscillation of the magnetar crust also couples to extended dipolar magnetic field lines outside the star (Timokhin et al. 2008), but only relatively weakly (Thompson & Duncan 2001), and in a way that should be continuously visible over a full rotation.

This mechanism provides a promising explanation for the observed persistence of discrete QPOs in magnetar flare light curves (Israel et al. 2005; Strohmayer & Watts 2005; Watts & Strohmayer 2006). The objection of Levin (2006) to identifying magnetar QPOs with modes confined to the magnetar crust is weakened: we find that growth can be faster than the damping caused by coupling to a continuous spectrum of modes in the magnetized core. In other words, the relative concentration of the excited mode between the crust and the magnetized core (e.g., Levin 2007) may shift back toward the crust in the presence of strong forcing at a natural frequency of the crust. Active forcing also permits QPO activity to extend well beyond the core coupling time. We do find continuing plastic dissipation well after the impulsive energy release (Figure 13), but the hydromagnetic instability of the core is only treated schematically.

9.3. Intermittency of Super-Eddington Outbursts

We have uncovered a plausible reason why super-Eddington outbursts are relatively rare events in magnetars (in comparison with slower but still dramatic changes in persistent X-ray output and spindown torque, both of which require strong magnetospheric currents). The extreme sensitivity of creep rate to applied stress and temperature can be compensated by the buildup of a reverse Maxwell stress during plastic flow. In a two-dimensional simulation, the seed magnetic field must exceed a substantial minimum strength. This need not suppress the formation of narrow zones of plastic creep, but it can help to cap the rate of creep within these zones. As the strength of this embedded field is increased, one sees a transition from a giant-flare-like phenomenon to a slower and less energetic outburst more typical of the transient magnetars. Super-Eddington outbursts may depend on a favorable magnetic geometry.

9.4. Core Hydromagnetic Instability

A transition from hydromagnetic stability to instability in the core is needed to explain the most energetic magnetar X-ray flares (>1045 erg). A significant structural change in the core magnetic field must come with longer-term side effects; indeed, our simulations produce repeat bursts of lower energy. It should be taken seriously as the underlying driver of lower-energy SGR bursts as well. We show that such a transition can, even in the absence of a crustal response, produce transient effects over timescales of a day, significantly shorter than the thermal conduction time across the crust.

9.5. Characteristic Burst Timescale

The ∼0.1 s duration seen in a wide range of magnetar X-ray bursts (e.g., Göǧüş et al. 2001) is a signature of stress redistribution within the crust. We demonstrate this for large (∼1045 erg) energy releases using a global two-dimensional crust model that includes magnetic, elastic, plastic, and thermal effects. In major outbursts, the peak of plastic dissipation jumps by kilometers, often repeatedly. Models that postulate that most of the available energy is stored in a twisted magnetosphere at the onset of a burst (Lyutikov 2003) have not yet offered a quantitative explanation for the ∼0.1 s timescale.

9.6. Rapid (Millisecond) Dissipation Growth

Millisecond growth times in the dissipation rate arise from the nonlinear and temperature-dependent plastic response of the crust: they correspond to stress rebalancing over a distance ∼ δRc ∼ 0.3 km. Rapid horizontal shear motions in the crust are accompanied by strong vertical magnetospheric currents and plausibly drive secondary current-driven instabilities (Thompson & Duncan 2001; Elenbaas et al. 2016). The shortest (submillisecond) growth seen in giant flares may point to such a magnetospheric phenomenon (Thompson & Duncan 1995; Lyutikov 2003).

9.7. Outbursts from Ejection of Crustal Magnetic Shear

Horizontal shuffling of poloidal magnetic field lines in the core is not fully mirrored by shuffling in the magnetosphere and therefore leaves behind vertical magnetic field gradients in the crust. We show that such embedded shear is rapidly (over several milliseconds) ejected from the crust when the nonlocal component of the stress builds up to a critical value. This is in line with early theoretical ideas suggesting the ejection of magnetic shear as the source of low-energy magnetar bursts (Thompson & Duncan 1995). The efficiency of this process is significantly enhanced if the background magnetic field is already sheared outside the neutron star. It will be further enhanced by secondary current-driven instabilities. The coupling of an elastic wave to the magnetosphere is slower by comparison (Blaes et al. 1989; Thompson & Duncan 2001; Link 2014).

This process starts with a decorrelation between solid and magnetic stresses throughout the crust, which is not present in a one-dimensional formulation. For example, restricting gradients to the vertical (z) direction and setting ${\partial }_{z}{T}_{{iz}}={\partial }_{z}({B}_{i}{B}_{z}/4\pi +{\sigma }_{{iz}})\,=0$, one finds a local relation between the maxima/minima of the solid stress and the Maxwell stress, ${\sigma }_{{iz}}=-{B}_{i}{B}_{z}/4\pi +\mathrm{const}.$ The constant boundary term does not change this result: it only biases up or down the local magnetic stress. By restricting the dynamics to one dimension, it is possible to consider a local extraction of magnetic energy through the deformation of the solid in which it is embedded. An example is the thermal-magnetic front studied by Beloborodov & Levin (2014) and Li et al. (2016), which operates on an intermediate timescale. Adding a growing global stress produces faster growth and larger energy release without the need for sharp temperature gradients.

9.8. Narrow Shear Bands

Narrow fault-like structures easily form in the magnetar crust in response to unbalanced Maxwell stresses at the lower crust boundary. These structures have dissipative envelopes extending to scales much larger than δRc and covering more than 10 percent of the surface area of the magnetar. As we show in Appendix A, they are well resolved, in the sense that the distribution of dissipation over the magnetar surface does not depend on resolution, and only a fraction of the total energy is dissipated in the highest-temperature core of the fault. We find that rapid development of extensive faults is associated with subsecond giant outbursts and is driven by the nonlocal response of the crust to rapid plastic flow in localized zones.

Fault-like structures provide a promising explanation for the spectral inference of hot spots on the surfaces of active magnetars (Ibrahim et al. 2001; Lenters et al. 2003; Woods et al. 2004; Esposito et al. 2007). They are especially prominent when the applied Maxwell stress has a rough rotational symmetry (one example being a kinked toroidal field in the outer core). The core magnetic field can otherwise be fairly smooth and need not contain current sheets.

In contrast with the analysis of Levin & Lyutikov (2012), we find that magnetic shearing within a fault need not suppress the release of energy. Instead, the resistance of the embedded magnetic field to the growing stress forces the shear layer to spread out as the global stress grows. In the presence of magnetar-strength (∼1015 G) fields, the fault width can approach the crustal thickness. This is energetically feasible because plastic creep with a narrow fault is driven by a stress accumulated over a much wider area of crust. Curvature of the fault in two dimensions is associated with a redistribution of stress outside the fault, creating wider dissipative structures.

9.9. Melting

Our two-dimensional model shows that the crust is nearly fully melted within active faults during major outbursts. We show that a standing Alfvén wave in a vertically magnetized liquid crust damps relatively slowly and oscillates with a roughly kilohertz frequency. Melted faults are promising locations for the high-frequency QPO activity seen in giant magnetar flares (Strohmayer & Watts 2006).

9.10. X-Ray Afterglow from Continuing Crustal Creep

The same two-dimensional model shows extended plastic dissipation in the aftermath of a giant energy release, declining as a power law ∼t−1 and releasing a net energy ∼10−2 of the outburst energy. These properties are remarkably similar to the X-ray afterglow profile detected following the 1998 giant flare of SGR 1900+14 (Woods et al. 2001).

One extracts from this example a broader lesson that may apply to other instances of transient X-ray emission (e.g., to the slower outbursts of transient magnetars; Ibrahim et al. 2004; Mori et al. 2013; Alford & Halpern 2016, and references therein). The magnetospheric twist associated with transient emission need not be injected all at once, as in the model of Beloborodov (2009), and is probably localized closer to the star than is implied by the dipolar "j-bundle" construction.

9.11. Some Open Questions

Implications for the magnetar eigenvalue problem. The potential of magnetar QPOs as a diagnostic of the internal structure has inspired detailed approaches to the eigenvalue problem by several groups (Piro 2005; Glampedakis et al. 2006; Levin 2007; Samuelsson & Andersson 2007; Sotani et al. 2008; Cerdá-Durán et al. 2009; Colaiuda & Kokkotas 2011; Gabler et al. 2011; van Hoven & Levin 2011, 2012; Passamonti & Pons 2016). The eigensolutions are not only sensitive to magnetic and compositional effects in the elastic parts of the crust but also subtly modified by plastic flow. In particular, creep in localized faults may contribute to the observed QPO frequency drifts by introducing phase shifts into shear wave propagation.

Hard and persistent X-ray emission. Magnetars are copious sources of hard X-rays in quiescence, with a power greatly exceeding the spindown power (Murakami et al. 1994; Kouveliotou et al. 1999; Kuiper et al. 2006). This emission is probably a signature of very strong electric currents flowing through the closed magnetosphere. The source location has variously been ascribed to compact and mildly relativistic plasma near the magnetar surface (Thompson & Beloborodov 2005), or to outward relativistic flows of pairs along a thin bundle of closed field lines surrounding the magnetic dipole axis (Beloborodov 2013). Our two-dimensional simulations naturally produce compact currents associated with strong gradients in creep rate and distributed broadly across the crust. They therefore favor the first, more localized type of X-ray emission process. The pulsed radio emission detected from some transient magnetars (Camilo et al. 2006, 2007) also gives evidence for a dynamic magnetosphere, but it transmits much less energy than the X-rays and is much more rapidly variable, and it is consistent with emission from open (or nearly open) magnetic field lines (Thompson 2008).

Thermalization of magnetic shear energy ejected from magnetar crusts. We do not address how a time-dependent magnetic displacement will damp after injection into the magnetosphere. Emission of quasi-thermal X-rays during super-Eddington magnetar bursts depends on a transfer of energy from relatively large scale (≳δRc) magnetic gradients to sub-relativistic particles with gyrational radii some 10−17 times smaller. Some type of cascade process must be involved (Thompson & Blaes 1998).

Plastic dissipation in the upper crust is found to be negligible, in agreement with Li & Beloborodov (2015). We nonetheless note that any cascade process that operates outside the star should also operate within the upper crust (where B2/8π ≫ ρc2 and VA ∼ c). The profile of heat deposition may therefore be flat compared with the pressure profile (Lyubarsky et al. 2002). Prompt burst afterglow emission is a probe of the thermalization process.

Surface Maxwell stresses. Persistent electric currents flowing through the surfaces of magnetars would cut the vertical imbalance in BBz/4π. The outer core and inner crust generally can sustain a stronger magnetic twist than the magnetosphere, but this effect could be important in zones with modest internal twist. A reduction in magnetospheric twist by current-driven instabilities would increase the net Maxwell stress applied to the crust, thereby facilitating the growth of rapid creep. The magnetosphere could also have a nonlocal effect on crustal yielding by communicating twist along closed magnetic field lines: injected twist would be focused at the opposing magnetic footpoint, although at the cost of diluting the twist over the length of the flux bundle.

Do lower-energy SGR bursts represent a "bottom-up" or "top-down" phenomenon? Earthquakes are ultimately the consequence of large-scale convective motions and compositional divisions within Earth (Turcotte & Schubert 2002). We have shown that fault-like structures emerge naturally in magnetar crusts, but our global model does not resolve features smaller than ∼0.3 km. Therefore, we cannot offer a quantitative model for the origin of low-energy SGR bursts.

The ∼0.1 s characteristic duration of low-energy SGR bursts does highlight an important difference with earthquakes: this duration is comparable to the shear wave propagation time across the magnetar crust, pointing to strongly nonlocal energy release, whereas the terrestrial events are relatively short by the analogous measure. In spite of this distinction, the energy distributions of SGR X-ray bursts and earthquakes follow similar power laws, with smaller bursts supplying a smaller cumulative energy release than the largest flares (Cheng et al. 1996; Göǧüş et al. 2001).

Finally, we note that a role for a large-scale hydromagnetic instability in facilitating low-energy burst activity is suggested by the relative activity of the same sources that (so far) have produced giant flares.

What role does Hall drift play in driving transient magnetar phenomena? Our approach here is partly motivated by the fact that the rate of crustal Hall drift may be exceeded in young and hot magnetars by the rate by magnetic drift in the core (Haensel et al. 1990; Goldreich & Reisenegger 1992), given the strong feedback of magnetic dissipation on the temperature and therefore on the ability of the magnetic field to overcome the back-pressure induced by the radial electron-fraction gradient (Thompson & Duncan 1996).

The sign of the effect of Hall drift on magnetar activity is not clear. The time-dependent yielding calculations presented here raise the possibility that Hall drift has a calming effect, opposite to the one usually considered. Sourcing small-scale magnetic irregularities (Goldreich & Reisenegger 1992; Hollerbach & Rüdiger 2002; Wareing & Hollerbach 2009; Pons & Geppert 2010; Takahashi et al. 2011; Li et al. 2016) helps to create components of the magnetic field transverse to the local direction of shear, which we find are key to suppressing runaway creep. Alternatively, small zones of enhanced magnetic flux density driven by Hall drift could become preferred dissipation sites in the presence of time-dependent global stresses. The Lorentz force density increases with wavenumber in an electron-MHD cascade (Cho & Lazarian 2004), and localized zones of magnetic flux enhancement are seen in three-dimensional simulations of Hall transport (e.g., Gourgouliatos et al. 2016).

The relevance of Hall drift for large-scale reorganizations of the magnetic field depends on the location of the currents that support the stellar magnetic field. Large-scale simulations (Perna & Pons 2011; Gourgouliatos et al. 2016) achieve a large redistribution of the dipole magnetic flux within the active lifetimes of a magnetar (∼103–104 yr) only if the external field is anchored in the crust, so that the crustal toroidal field approaches 1016 G. Otherwise, the latitudinal electron drift is too slow to transport the poloidal magnetic field a significant distance. In the alternative approach advanced here, global twists, which can strongly modulate the spindown of the magnetar, are sourced by global imbalances in the core.

Comments from the anonymous referee helped to sharpen the presentation of our results. H.Y. thanks Jonah Miller for discussion of the numerical implementation of wave equations. This research was supported in part by NSERC and in part by the Perimeter Institute for Theoretical Physics. Research at Perimeter Institute is supported by the Government of Canada through the Department of Innovation, Science and Economic Development Canada and by the Province of Ontario through the Ministry of Research Innovation, and Science.

Appendix A: Convergence of the Two-dimensional Plastic/Elastic/Thermal Crust Model

A comparison of the results for simulations with resolution (25)2, (26)2, and (27)2 shows that the energy dissipated by plastic creep and the energy injected by surface shearing motions into the magnetosphere are both reasonably converged. This conclusion holds for both the total energy and the cumulative distribution of energy with respect to surface energy density. At the highest resolution attainable for reasonable running time, the pixel scale was ∼0.3 km.

The test is performed by using the same power spectrum of Maxwell stress at the lower crust boundary as in the other simulations (α = 1/2 in Equation (33)). In order to align the mode amplitudes and phases between the different resolutions, the spectrum is cut off at wavenumber ${r}_{\max }={s}_{\max }={2}^{4}-1$. The maps of the plastic dissipation and magnetospheric energy injection are shown for the case Bc = 0 and anisotropy parameter η = 0.3 in Figures 22 and 23. The total plastic energy released in the first two bursts is (7.0, 6.4, 6.8) ×1044 erg and (1.1, 1.3, 1.4) × 1045 erg, as recorded at the three successively higher resolutions. The total increase in external magnetic energy is (4.0, 3.9, 3.8) × 1044 erg in burst 1 and (6.8, 8.0, 7.5) × 1044 erg in burst 2. The variation is therefore at the several percent level during the first burst, being compounded in the second burst to the 10%–15% level.

Figure 22.

Figure 22. Map of energy dissipated by plastic creep, comparing three different levels of resolution ((25)2, top left; (26)2, top right; and (27)2, bottom) for the second burst of a simulation with initial core magnetic spectrum α = 1/2, anisotropy parameter η = 0.3, and ${\boldsymbol{B}}$c = 0. The initial stress spectrum is cut off at wavenumber ${r}_{\max }={s}_{\max }={2}^{4}-1$ so as to align the mode amplitudes and phases. The color scheme for energy dissipated per pixel is the same as in Figure 3 for resolution (26)2, with energy per pixel scaled up (down) by a factor of 22 (2−2) at lower (higher) resolution.

Standard image High-resolution image
Figure 23.

Figure 23. Same as Figure 22, but now showing change in external magnetic energy.

Standard image High-resolution image

The cumulative distribution of plastic/magnetic energy with respect to energy dissipated/injected per unit area is shown in Figure 24. There is excellent agreement between the three levels of resolution.

Figure 24.

Figure 24. Black curves (burst 1) and blue curves (burst 2): cumulative distribution of energy dissipated by plastic creep, with respect to energy dissipated per unit area. Solid curve: resolution (26)2; long-dashed curve: resolution (25)2; dotted curve: resolution (27)2. Red curves (burst 1) and orange curves (burst 2): cumulative distribution of energy injected into the external magnetic field.

Standard image High-resolution image

Now we consider yielding models with a horizontal magnetic field embedded in the crust, which is taken to lie in the y-direction (at ${45}^{^\circ }$ with respect to the symmetry axis of the applied stress). The fault-like structure that emerges is not quite as extended as in the unmagnetized case, but there is no evidence for a continuing shrinkage between a magnetization ${B}_{c}^{2}/4\pi \bar{\mu }=0.03$ (Bc ≃ 1015 G) and ${B}_{c}^{2}/4\pi \bar{\mu }=0.1$ (Figure 25). The distributions of the dissipated plastic and magnetic energy still show a strong overlap (Figure 26).

Figure 25.

Figure 25. Plastic dissipation patterns for the simulations with highest resolution, anisotropy parameter η = 0.1, and ${B}_{c}^{2}/4\pi \bar{\mu }=0.03$ (top) and 0.1 (bottom). Magnetic field is oriented at 45° to the dominant fault direction. Faults appear to be somewhat broadened at the highest flow displacements compared with the unmagnetized simulations. The distributions of plastic and magnetic energy release in these simulations and in corresponding lower-resolution simulations with identical initial conditions are shown in Figure 26.

Standard image High-resolution image
Figure 26.

Figure 26. Black (${B}_{y}^{2}/4\pi \bar{\mu }=0.03$) and blue (${B}_{y}^{2}/4\pi \bar{\mu }=0.1$) curves show the cumulative distribution of energy dissipated by plastic creep. Dotted curves correspond to the sequence of simulations shown in Figure 25 (resolution (27)2). Long-dashed curves: resolution (25)2. Solid curves: resolution (26)2. Red (${B}_{y}^{2}/4\pi \bar{\mu }=0.03$) and orange (${B}_{y}^{2}/4\pi \bar{\mu }=0.1$) curves show the cumulative distribution of energy injected into the external magnetic field.

Standard image High-resolution image

The convergence properties are not quantitatively quite as tight as in the unmagnetized simulations. The total plastic energies at successive resolutions are (1.6, 0.54, 0.46) × 1045 for the ${B}_{y}^{2}/4\bar{\mu}=0.03$ simulations and (0.6, 0.35, 0.43) × 1045 erg for the ${B}_{y}^{2}/4\bar{\mu}=0.1$ simulations. Here the plastic and magnetic energies dissipated agreed to about 15%–20% at the two highest resolutions.

The results obtained are consistent with the picture of fault smoothing outlined in Section 2.2: for the highest value of Bc, the effective fault width is Δ ∼ 0.1Lx ∼ 1 km (a few pixels), as is seen in Figure 25. Continuing dissipation is associated with the growth of a fault (Figure 14). A nonlocal redistribution of stress around the end of the fault pushes neighboring patches of the crust above the threshold for runaway yielding. We conclude that a horizontal magnetic field of strength ${B}_{c}^{2}/4\pi \bar{\mu }\sim 0.03\mbox{--}0.1$ does not suppress the development of narrow dissipative structures.

Appendix B: Localized Plastic Spot

Here we give a concrete example of a plastically deforming circular spot in a planar shell that is subject to a Maxwell stress BϕBz/4π at its lower boundary. The horizontal gradient scale of the applied stress is large everywhere compared with the shell thickness δRc, so that magnetoelastic equilibrium can be described in terms of a planar stress ${\sigma }_{\phi \varpi }=\bar{\mu }\varpi {\partial }_{\varpi }({\xi }_{\phi }/\varpi ).$ The toroidal displacement field is assumed to be independent of height, and $\bar{\mu }$ is the vertically mass-averaged shear modulus.

The vertical magnetic field Bz is everywhere uniform, and we take

Equation (88)

so that the Maxwell stress is strongly localized at ϖ ≲ r0. The solution to the equation of magnetoelastic equilibrium

Equation (89)

is

Equation (90)

This peaks at $\varpi ={r}_{p}=\sqrt{2}{r}_{0}$. The displacement field

Equation (91)

scales as ϖ−1 at large radius.

Time dependence of the applied stress corresponds to ${\partial }_{t}{B}_{0}\ne 0$. The net circulation is given by

Equation (92)

Footnotes

  • The magnetosphere generally supports only a weaker magnetic twist than the magnetar interior.

  • This Cartesian approximation therefore overestimates the core volume by a factor of ∼1.7.

  • Strictly the vector potential Az is first calculated, and then Bx,y are derived from it.

  • Our numerical approach has been checked by restricting the power spectrum of the seed Maxwell stress to one dimension and converting ${\dot{{\rm{\Psi }}}}_{\mathrm{pl},r}$ back to coordinate space. The plastic displacement agrees with the direct evaluation of Equation (38), once its spatial average is subtracted.

  • Imagine that all variables are functions $f(x,y)=f(\cos \alpha x+\sin \alpha y)$, with α = constant. Substituting into Equation (46), one ends up with the one-dimensional relation (38).

Please wait… references are loading.
10.3847/1538-4357/aa6c30