A publishing partnership

UNTWISTING MAGNETOSPHERES OF NEUTRON STARS

Published 2009 September 4 © 2009. The American Astronomical Society. All rights reserved.
, , Citation Andrei M. Beloborodov 2009 ApJ 703 1044 DOI 10.1088/0004-637X/703/1/1044

0004-637X/703/1/1044

ABSTRACT

Magnetospheres of neutron stars are anchored in the rigid crust and can be twisted by sudden crustal motions ("starquakes"). The twisted magnetosphere does not remain static and gradually untwists, dissipating magnetic energy and producing radiation. The equation describing this evolution is derived, and its solutions are presented. Two distinct regions coexist in untwisting magnetospheres: a potential region where ∇ ×B = 0 ("cavity") and a current-carrying bundle of field lines with ∇ ×B ≠ 0 ("j-bundle"). The cavity has a sharp boundary, which expands with time and eventually erases all of the twist. In this process, the electric current of the j-bundle is sucked into the star. Observational appearance of the untwisting process is discussed. A hot spot forms at the footprints of the j-bundle. The spot shrinks with time toward the magnetic dipole axis, and its luminosity and temperature gradually decrease. As the j-bundle shrinks, the amplitude of its twist ψ can grow to the maximum possible value ψmax ∼ 1. The strong twist near the dipole axis increases the spindown rate of the star and can generate a broad beam of radio emission. The model explains the puzzling behavior of magnetar XTE J1810−197, a canonical example of magnetospheric evolution following a starquake. We also discuss implications for other magnetars. The untwisting theory suggests that the nonthermal radiation of magnetars is preferentially generated on a bundle of extended closed field lines near the dipole axis.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Neutron stars are highly conducting and strongly magnetized. Their extended magnetospheres are anchored deep in the rigid crust and corotate with the star. The magnetosphere is usually assumed to be static in the corotating frame or evolving very slowly as the star ages. Electric currents are confined to a narrow bundle of open field lines that connect the star to its light cylinder (Goldreich & Julian 1969). The main, closed, part of the magnetosphere is usually assumed to be current-free and potential, ∇ ×B = 0.

The standard picture of a static potential magnetosphere is reasonable for ordinary pulsars, yet apparently it does not describe all neutron stars. In particular, neutron stars with ultrastrong fields B ≳ 1014 G (magnetars) are inferred to have dynamic magnetospheres (see, e.g., reviews by Woods & Thompson 2006; Kaspi 2007; Mereghetti 2008). Their activity is believed to be caused by crustal motions relieving internal stresses.2 In contrast to ordinary pulsars, these objects show huge temporal variations in luminosity, spectrum, and spindown rate. Theoretically, stresses are expected to build up in the deep crust as the star ages (e.g., Ruderman 1991; Thompson & Duncan 1995). In particular, the large Ampere forces  j ×B/c inside magnetars can break the crust and shear it in a catastrophic way.3 Such a starquake twists the magnetic field anchored in the crust, creating ∇ ×B ≠ 0 and inducing electric currents in the closed magnetosphere (Thompson et al. 2000). The currents are approximately force-free and flow along the magnetic field lines,  j ×B = 0. They emerge from the deep crust sheared in the starquake.

Twisted force-free magnetospheric configurations were studied extensively in the context of the solar corona, and these models can be applied to neutron stars. A simple example is the self-similarly twisted dipole. It was constructed by Wolfson (1995) and applied to neutron stars by Thompson et al. (2002). This and similar force-free configurations are magnetostatic solutions. A sequence of such configurations may be constructed by changing their boundary conditions, i.e., displacing the footpoints of the magnetic field lines. If the footpoints freeze, the configuration freezes as well. At a first glance, this seems to suggest that the implanted twist must freeze when the starquake ends, and wait for another starquake.

In fact, the magnetosphere must evolve after the starquake, even though it remains anchored in the motionless deep crust. Indeed, energy is continually dissipated in the twisted magnetosphere, because the twist current  j = (c/4π)∇ ×B is maintained by a voltage Φe ≠ 0 established along the magnetic field lines. Thompson et al. (2000) estimated voltage Φe assuming that the currents are carried by electrons and ions lifted from the star's surface against gravity. Beloborodov & Thompson (2007; hereafter BT07) found that Φe is regulated by an e± discharge.4 The voltage is significant—comparable to 1 GeV—and implies a modest lifetime of the twist, comparable to one year.

The untwisting dynamics of the magnetosphere remained, however, unknown. Usually, resistivity in a plasma leads to diffusion of currents across the magnetic field. Voltage Φe ≠ 0 implies an effective resistivity, and one could expect the decaying twist to spread diffusively across the magnetosphere. This expectation is incorrect, as will be shown below.

The goal of this paper is to develop an electrodynamic theory of twisted magnetospheres that describes their evolution. We focus on axially symmetric configurations. In this case, the twist is created through a latitude-dependent azimuthal rotation of the crust. An introductory description of twisted magnetic configurations is given in Section 2, and their untwisting dynamics is qualitatively discussed in Section 3. In Section 4 we derive the electrodynamic equation for axisymmetric magnetospheres.

Section 5 presents solutions to the evolution equation and explores the mechanism of untwisting. Observational effects of this process are described in Section 6. Section 7 compares the theory with the recent observations of a starquake in the anomalous X-ray pulsar (AXP) XTE J1810−197. Section 8 summarizes the results of the paper and discusses implications for magnetars.

2. TWISTED MAGNETOSPHERE

In spherical coordinates r, θ, ϕ, the magnetic field can be written as the sum of poloidal and toroidal components,

Equation (1)

where $\hat{{\mathbf e}}_r$, $\hat{{\mathbf e}}_\theta$, $\hat{{\mathbf e}}_\phi$ are unit vectors pointing in the r, θ, ϕ directions. We assume that the magnetic field is symmetric about the polar axis, i.e.,  B does not depend on ϕ. The axisymmetric field can be viewed as a foliation of magnetic flux surfaces. (Each surface may be obtained by rotating a field line around the axis of symmetry.)

Let f be the magnetic flux through a circular contour of fixed r = const and θ = const. The function f(r, θ) is constant on a flux surface. As is usual in plasma physics, we will use f to label the flux surfaces. Note that flux surfaces extending farther from the star have smaller f, and f = 0 corresponds to the polar axis θ = 0. We focus on the closed magnetosphere in this paper and neglect the narrow bundles of open magnetic field lines; effectively, rotation of the star is neglected.

Let R be the radius of the star. The magnetic field is force-free outside the star:5j ×B ≈ 0 at r > R. If we follow the magnetic field lines into the star, significant deviations from the force-free condition appear. The ϕ-component of  j ×B remains, however, small as the crust is relatively fragile to axisymmetric azimuthal displacements (which involve no compression). This component can be written as  jp ×Bp where  jp is the poloidal component of the current density. Let rc be the radius of the lower crust that is strong enough to sustain significant azimuthal Ampere forces  jp ×Bp/c ≠ 0. Outside this radius we assume

Equation (2)

A rough estimate rc ∼ 0.9R is sufficient for the purposes of this paper (the exact rc depends on the strength of the magnetic field). The effective footpoints of the magnetospheric field lines sit at r = rc. We focus in this paper on the region r > rc and call it "force-free" (in the restricted sense  jp ×Bp = 0).

The spheres r = R and r = rc define two special flux surfaces (Figure 1).

  • 1.  
    Flux surface fR touches the surface of the star.6 Flux surfaces f > fR are confined to the star, and flux surfaces f < fR extend outside the star and form the magnetosphere. fR represents the total magnetic flux emerging from the star in the region of positive polarity (where Br > 0).
  • 2.  
    Flux surface fc touches the boundary of the inner crust r = rc. Flux surfaces f > fc are confined to the inner crust and the core of the star.
Figure 1.

Figure 1. Poloidal cross section of an axisymmetric magnetic configuration. The star is the sphere of radius R (shaded). The lower crust is inside radius rc (dashed circle). The figure shows the magnetic axis f = 0, flux surface fR that touches the surface of the star, and flux surface fc that touches the sphere rc. The magnetosphere is composed of nested closed flux surfaces f < fR; one such flux surface is shown in the figure.

Standard image High-resolution image

A twist is pumped into the force-free region r > rc when the footpoints of field lines at r = rc are displaced by a starquake. The crust is practically incompressible, and hence any axisymmetric starquake produces a pure azimuthal displacement (an axisymmetric displacement in the θ-direction would imply compression).

Let us call the footpoints of positive polarity (Br > 0) northern. Consider an initially pure poloidal magnetosphere and suppose a starquake shifts the northern footpoints of magnetic field lines through angle Δϕn(f) and southern footpoints through Δϕs(f). The created twist is described by the relative angular displacement ψ(f) = Δϕs − Δϕn = ϕs − ϕn. This angle can be expressed as an integral along the closed field line: ψ = ϕs − ϕn is accumulated as we move along the field line from its northern footpoint at r = rc to the southern footpoint. An infinitesimal displacement dl along the field line corresponds to azimuthal displacement hϕdϕ = (Bϕ/B) dl where hϕ = rsin θ. Therefore, the twist angle is given by

Equation (3)

where the integral is taken along the field line outside rc. Creation of ψ ≠ 0 implies the appearance of a toroidal magnetic field Bϕ in the magnetosphere.

The toroidal field Bϕ(r, θ) determines the circulation of  B along the circular contour r = const, θ = const. By Stokes' theorem, it is related to the electric current I flowing through the contour,7

Equation (4)

I is determined by the poloidal component of the current  jp, and f is determined by the poloidal component of the magnetic field  Bp. At r > rc, the condition  jp ×Bp = 0 implies that the poloidal currents flow along the poloidal flux surfaces. Therefore, I is a function of f. Note also that the definition of I is similar to that of f except that  B is replaced by  j, which implies

Equation (5)

Any axisymmetric force-free field outside the star satisfies the Grad–Shafranov equation that expresses the condition  B × (∇ ×B) = 0 in terms of I and f. Its exact solutions (matching an interior non-force-free solution) are needed to describe twists with large angles ψ. Note that configurations with ψ ≫ 1 are not expected as they are unstable (Uzdensky 2002 and references therein). If the twist grows beyond the instability threshold $\psi _{\rm max}={\cal O}(1)$, the magnetosphere becomes kink-unstable and ejects a closed plasmoid, which prevents the twist growth above ψmax.

For moderate twists ψ < 1 one does not have to solve the Grad–Shafranov equation. Instead, a simple linear approximation may be sufficient: the configuration can be thought of as a linear superposition of an initial non-twisted poloidal field  B0 and a toroidal field Bϕ < B0 that was created by the footpoint displacement ψ. The appearance of Bϕ does not affect the poloidal field in the linear order—the poloidal correction to  B is quadratic in Bϕ/B. The linear approximation may break at large distances from the star (Low 1986), however it describes well most of the magnetosphere. Wolfson & Low (1992) found that the relation between I and f obtained in the linear approximation works well even for twists ψ ∼ 1.

During the starquake, the twisting motion of the footpoints at r = rc pumps energy into the magnetosphere, which may be released later. The free energy of twisted force-free configurations was extensively studied (e.g., Aly 1984). For linear twists, Bϕ < B, the free energy simply equals the energy of the toroidal field component Bϕ,

Equation (6)

3. FATE OF THE EJECTED CURRENT

The current I through the magnetosphere is maintained by electric field E ≠ 0 (parallel to  B), which implies Ohmic dissipation of the twist energy,  E ·  j ≠ 0. Thus, the magnetic field must be gradually untwisted (even though it remains anchored in the static deep crust), and eventually the magnetospheric current must vanish. On the other hand, the ejected current cannot disappear because it emerges from a static and almost ideal conductor—the deep crust r < rc, where the magnetic field and electric currents remain unchanged. We conclude that Ohmic dissipation must redirect the ejected poloidal current so that it closes below the surface of the star. The current must be redirected across the magnetic flux surfaces, which can happen only in the transition layer rrc between the heavy static conductor and the force-free region. Then the current does not penetrate the force-free region on flux surfaces f < fR and avoids the magnetospheric dissipation.

It is instructive to consider an idealized problem where the entire star is a perfect conductor, so that E ≠ 0 only outside the star. When the magnetospheric dissipation is completed,  jp = 0 on flux surfaces f < fR in the region r > rc. The initially ejected current now flows in a current sheet inside the star (Figure 2). The current sheet serves as a screen between the twisted field inside the star and the untwisted (potential) field in the magnetosphere. When the finite resistivity of the crust is taken into account, the state shown in Figure 2 is not final. The current sheet in the non-ideal conductor will acquire a non-zero thickness and gradually spread to the inner flux surfaces until the currents reach deeper crust with so high conductivity that it can be treated as an ideal conductor on timescales equal to the age of the star. The currents that initially emerged during the starquake will eventually close deep under the surface of the star. The tendency of currents to diffuse toward regions of higher conductivity was observed in numerical simulations of Ohmic dissipation in neutron star crusts (Sang & Chanmugam 1987).

Figure 2.

Figure 2. Pattern of electric currents after a global starquake that has twisted the entire magnetosphere. Poloidal electric currents are shown by arrows; they flow along the magnetic flux surfaces outside rc (dashed circle). Left: initial twisted state. Right: final untwisted state. When the magnetospheric dissipation is completed, all currents are sucked into the star and close below its surface. The thick curve shows the poloidal cross section of the current sheet formed inside the star. It extends from the axis along the sphere rrc, across the flux surfaces f < fR. The current sheet turns where it reaches the flux surface fR, and continues along this flux surface.

Standard image High-resolution image

In summary, two stages are expected in the evolution of a magnetic twist created by a starquake. (1) The current ejected by the starquake into the magnetosphere is gradually drawn into the star. Most of the twist energy is released at this stage. (2) The current spreads into deeper layers of the star and eventually collects near the highly conducting inner crust. The second, subsurface untwisting stage is much slower because the resistivity inside the star is smaller than the effective resistivity of the magnetosphere. We shall focus below on the faster first stage and treat the star as an ideal conductor.

4. TWIST EVOLUTION EQUATION

4.1. Resistive Evolution

The evolution of magnetic field is related to electric field  E according to the induction equation,

Equation (7)

We will express ∇ ×E in curvilinear coordinates qi defined as follows. Let us label magnetic field lines on a flux surface f by the azimuthal angle ϕ0 of their northern footpoints at r = rc (northern footpoints have Br > 0). Thus, the set of all field lines is parameterized by two coordinates f and ϕ0. Let s be a parameter running along the field line (increasing in the direction of  B). The parameter s may be chosen, e.g., equal to length l measured along the field line from its northern footpoint (then |es| = 1). The coordinate system qi = (s, f, ϕ0) covers the entire magnetic field that passes through the sphere r = rc and emerges in the force-free region.

The electric field can be written in components in the new coordinate basis,

Equation (8)

The length of the basis vectors will be denoted by hi = |ei|. Note that

Equation (9)

We will need the determinant g of metric gik = ei · ek, which is given by $\sqrt{g}={{\mathbf e}}_s\cdot ({\mathbf e}_{f}\times {\mathbf e}_\phi)$. Consider an infinitesimal axisymmetric ring perpendicular to the poloidal component of  B. Its surface element is ef × eϕdfdϕ, and the area of the ring is 2π |ef × eϕ|df. The magnetic flux through the ring, df, is

Equation (10)

which implies the identity 2π B · (ef × eϕ) = 1. We substitute  B = Bes/hs and find

Equation (11)

The general expression for ∇ ×E in curvilinear coordinates is

Equation (12)

where epsilonijk is the Levi-Civita symbol and Ei = ei ·  E are the covariant components of the electric field in the coordinate system qi. Using ${\mathbf e}_i=h_i\hat{{\mathbf e}}_i$, we obtain the azimuthal component of ∇ ×E in the normalized basis $\hat{{\mathbf e}}_i$ and find

Equation (13)

Let us divide both sides of this equation by 2πBhϕ/hs and integrate it over s along an entire closed field line (including its part at r < rc). Then the second term on the right-hand side disappears, and we get

Equation (14)

The integral on the right-hand side is the net voltage induced along the magnetic field line,8

Equation (15)

where $E_\parallel ={\,{\mathbf E}}\cdot \hat{{\mathbf e}}_s$ and dl = hsds is the length element along the field line. Equation (14) shows that the twist evolution is controlled by the longitudinal voltage, as expected. Using Bϕ = 2I(f, t)/chϕ (Equation (4)), we rewrite Equation (14) as

Equation (16)

Note that ∂Bϕ/∂t = 0 and ∂I/∂t = 0 at r < rc since the magnetic field remains static inside the static ideally conducting inner region. ∂I/∂t jumps from 0 to its value in the force-free region at rrc, in a transition layer of thickness Δ ≪ rc. The contribution from this layer to the integral on the left-hand side of Equation (16) is small and we neglect it. Then, effectively, the integral is taken only along the force-free part of the field line at r > rc. The current I and its time derivative ∂I/∂t are constant along this part of the field line. Then, using the expression for the twist angle (Equation (3)), we find

Equation (17)

Equation (17) describes the evolution of I(f, t) and the corresponding ψ(f, t). The twist is changing with time as the magnetic field lines gradually slip in the resistive magnetosphere and connect new footpoints with different ϕs − ϕn. Thus, ψ is changing despite the fact that the magnetosphere remains anchored in the static deep crust.

4.2. Linear Twists

For small twists ψ < 1, the magnetic field can be written as  B ≈  B0 +  Bϕ(t) (Section 2), where poloidal field  Bp =  B0 remains static and Bϕ < B0. Then $B=B_0[1+{\cal O}(B_\phi ^2/B^2)]\approx {\rm const}$ and the twist evolution equation (16) simplifies to

Equation (18)

This approximate equation quickly becomes accurate for BϕB: its error is decreasing as (Bϕ/B)2. In the following sections we will use Equation (18) to study the evolution of the twist amplitude ψ. The exact evolution equation (17) would have to be used when the effect of the twist on  Bp is of interest (e.g., for calculations of the spindown rate of the star).

The linearized description of twisted configurations is useful for the first calculations of the resistive untwisting. However, the limitations of this approximation should be kept in mind. The linearized description may fail at large distances from the star (Low 1986). Besides, the force-free configuration may be unable to smoothly adjust to the growing twist. Sudden relaxation to a new topological configuration is possible, with partial opening of the field lines and the loss of connectivity between the footpoints. The twist evolution equation derived above (linear or nonlinear) does not describe such transitions.

4.3. Energy Conservation Law

The energy of a linear twist Bϕ < B0 is given by

Equation (19)

Differentiating with respect to time and using Equations (17) and (18), we get

Equation (20)

Integrating by parts and taking into account that I(0) = 0 and Φe(fc) = 0 (E = 0 on flux surfaces confined to the perfect conductor), we find

Equation (21)

The right-hand side represents the net Ohmic losses. Etw is the free energy of the twist that is gradually dissipated as the magnetic field evolves according to Equation (18).

4.4. Magnetosphere with Moving Footpoints

The evolution equation derived above assumes that the magnetosphere is anchored in the deep static crust following a starquake. It is straightforward to generalize Equation (18) for magnetospheres with moving footpoints,

Equation (22)

where ω = dϕs/dtdϕn/dt is the differential angular velocity of the northern and southern footpoints of the magnetic field lines. A crust that remains motionless at all times except a sudden starquake at t = 0 is described by

Equation (23)

where δ(t) is the Dirac function and ψ0 is the amplitude of the twist imparted by the starquake. The impulsive twisting is a good approximation for recurring starquakes if the time between subsequent starquakes is longer than the timescale of untwisting. In the opposite limit, the crust is frequently deformed my mini-starquakes or moves plastically. Then the footpoint motion can be described by a continuous function ω(f, t). In this paper, we focus on the case of impulsive twisting described by Equation (23).

4.5. Twisted Dipole

For the study of the untwisting mechanism in the next section, it is useful to consider a concrete simple magnetic configuration. We will consider a dipole field with the symmetry axis passing through the center of the star. Let $\vec{\mu }$ be the dipole moment. The poloidal flux function for the dipole is given by (Appendix A),

Equation (24)

where Rmax = r/sin2θ is the maximum radius reached by the flux surface passing through given r, θ. Hereafter, instead of f or Rmax we will label flux surfaces by the dimensionless coordinate

Equation (25)

The last flux surface in the force-free region (marginally emerging from the inner crust: Rmax = rc) has u = uc = R/rc ≈ 1.1. The region 0 < u < 1 corresponds to the magnetospheric flux surfaces. In this region, u = sin2θ1, where θ1 is the polar angle of the northern footprint of the field line on the star surface r = R.

Suppose now that the field has been twisted by a starquake that was symmetric about the dipole axis and resulted in a differential rotation of the crust through angle ψ(u) (Section 2). The poloidal current I and twist angle ψ will be viewed below as functions of u and time t. The following relation holds between ψ and I,

Equation (26)

(see Appendix A). The twist evolution equation (18) becomes

Equation (27)

or

Equation (28)

The current density in the twisted dipole magnetosphere is given by (cf. Equation (5)),

Equation (29)

5. MECHANISM OF UNTWISTING

This section will explore the mechanism of untwisting for twists created by axisymmetric starquakes in the dipole magnetosphere (Section 4.5). Their evolution is described by Equation (27). Before this equation can be solved, the voltage Φe must be specified.

5.1. Twist Evolution in a Medium with Fixed Conductivity

Consider first what would happen if the magnetosphere was filled by a medium with a fixed conductivity σ; we will assume in this toy model σ(r) = const. Then the current density is related to E by Ohm's law j = σE. The star will be modeled as a perfect conductor, so E = 0 is assumed inside the star. Then the voltage Φe along a magnetospheric field line is given by

Equation (30)

(The integral has been calculated using Equations (29) and (A10).) Substitution of this result to the twist evolution equation (28) yields the equation for I(u, t),

Equation (31)

Given an initial twist with current function I(u, 0), one can calculate the evolution of I(u, t) by solving this differential equation with two boundary conditions: I(0) = 0 and I(1) = const. The latter condition is valid as long as the perfect-conductor approximation is used for the star.

Equation (31) is of diffusion type. It has a special feature: the effective diffusion coefficient is proportional to $\sqrt{1-u}$ and vanishes at u = 1. This fact allows one to qualitatively understand the evolution of I(u, t) before solving the equation numerically. It is instructive to consider the analogous problem of particle diffusion with a position-dependent diffusion coefficient D(u) that vanishes at the boundary u = 1. The vanishing of D means that particles "stick" to the boundary. With time, more and more particles get stuck, and eventually the particle density vanishes everywhere except at the boundary. The magnetospheric current behaves in a similar way. Far from the boundary, it simply spreads diffusively: ∂I/∂t = const u2 ∂2I/∂u2 at u ≪ 1. At the same time, near the boundary u = 1, the current is sucked toward u = 1. The current keeps accumulating at u = 1 until I = 0 at all u < 1.

For illustration, we solved numerically Equation (31) for a twist that initially has a uniform amplitude ψ0(u) = 0.2. The corresponding initial current function I(u, 0) and ∂I/∂uj/B are shown in Figure 3. The evolution of this twist is shown in Figure 4. The characteristic diffusion timescale is tσ = R2c2/4πσ, and we express time in units of tσ. As expected, the current tends to spread to smaller u and, at the same time, it is quickly drawn into the current sheet at u = 1. The sum of the currents flowing in the magnetopshere and in the current sheet, Itot, remains constant. Our numerical model assumes Φe = 0 at u > 1 (inside the star), which allows the current sheet to persist at u = 1. A real star has a finite conductivity and the current sheet will slowly spread into the star (toward larger u > 1).

Figure 3.

Figure 3. Uniform twist created by a global starquake involving the entire magnetosphere 0 < u < 1. The twist amplitude in this example is ψ0(u) = const = 0.2. The figure shows the corresponding current function I0(u) = I(u, 0) and its derivative dI0/du, which is proportional to current density j (see Equation (29)). Current I is measured in units of $\hat{I}=c\mu /4R^2$, where μ and R are the magnetic dipole moment and the radius of the star. The coordinate u = f/fR = R/Rmax labels flux surfaces (see Equation (25) and Figure 1); Rmax is the maximum radius reached by the flux surface. u = 0 on the magnetic axis (Rmax on the axis). For magnetospheric flux surfaces u < 1 and u = sin2θ1 where θ1 is the polar angle of the northern footprints of the flux surface on the star. Flux surfaces with u > 1 close inside the star; this region is shaded in the figure.

Standard image High-resolution image
Figure 4.

Figure 4. Evolution of the twist (with the initial configuration shown in Figure 3), calculated under the assumption that the magnetosphere is filled with a medium of a fixed conductivity σ, and the star is an ideal conductor. Three panels show the twist at times t/tσ = 0.01, 0.1, and 0.7, where tσ = R2c2/4πσ. The solid curve shows the current distribution ∂I(u, t)/∂u; I in the figure is measured in units of $\hat{I}=c\mu /4R^2$. The initial dI0(u)/du is shown by the dotted curve (from Figure 3). The red shaded area represents the magnetospheric current I and the current sheet I1 at u = 1. The total current Itot = I + I1 is conserved. I1 grows with time, and the magnetospheric current I decreases and spreads toward the magnetic axis u = 0. The dashed-dotted curve shows the twist amplitude ψ. It grows near the axis.

Standard image High-resolution image

Note that the twist amplitude ψ grows near u = 0, because ∂Φe/∂u > 0 in this region. A strongly twisted bundle develops near the magnetic axis; however, its thickness shrinks with time, so its net current decreases. At late times, the twist growth near the axis enters a self-similar regime. The amplitude peak ψpeak would grow indefinitely in the limit t, if the twist remained stable. In fact, the growth must be stopped by the MHD instability expected at ψ ≳ 1. An upper limit on ψ is set by the field-line opening, which may give a partially opened configuration with a lower energy (Wolfson & Low 1992). Overtwisted configurations are prone to kink instability, which is likely to reconnect part of the twisted region away from the star. These dynamic processes are not described by our model. We shall assume that they keep ψ below $\psi _{\rm max}={\cal O}(1)$.

5.2. Threshold Voltage

The toy model discussed in Section 5.1 is deficient: the real magnetospheres of neutron stars are not filled by a medium of a fixed conductivity. Instead, the magnetospheric currents are maintained through a discharge with a threshold voltage that is approximately the same for any current density j ≠ 0 (or, more precisely, for any j exceeding a small value j). The threshold nature of the magnetospheric voltage changes the evolution of the twist. Remarkably, it simplifies the solution: the partial differential equation (27) will be reduced to an ordinary differential equation.

We denote the threshold voltage by ${\cal V}$. Once Φe reaches ${\cal V}$, copious particle supply is available to carry any large current. The particle supply and voltage regulation in magnetars was studied in BT07. In principle, there are two sources of particles: (1) the surface of the star and (2) e± creation in the magnetosphere. The current is carried by charges of both signs, since the net charge density must be nearly zero to avoid huge voltages. If no e± are created, electrons and ions must be lifted from the star's surface (if ions are available in an atmospheric layer atop the solid crust). Maintaining a flow of ions along a magnetic loop requires a minimum voltage,

Equation (32)

where Rmax is the maximum radius reached by the loop and mi is the ion mass. This voltage is ∼0.2mic2 for loops with RmaxR. The plasma is lifted into the loop by the self-induction electric field (BT07). The exact solution was obtained for this process in the one-dimensional circuit model. It shows that voltage keeps growing even after reaching ${\cal V}_{ei}$, and the circuit evolves toward a global double-layer configuration. This result may not hold in the complete three-dimensional model that includes the excitation of transverse waves in the magnetosphere. Nevertheless, it suggests that lifting plasma from the surface is not the ultimate regulator of the voltage.

A robust mechanism for limiting the voltage is the e± discharge. If Φe exceeds a certain threshold ${\cal V}_{\pm }$, the exponential runaway of pair creation occurs (cf. Figure 5 in BT07), and the e± pairs screen E. In a magnetar magnetosphere, the e± avalanche is triggered when the accelerated electrons resonantly scatter X-rays streaming from the star, and the scattered photons convert to e± off the magnetic field. The threshold for the discharge is given by

Equation (33)

where ωX is the typical frequency of target X-rays and γres ∼ (B/BQ)(mec2/ℏωX) is the electron Lorentz factor at which the electron begins to scatter >1 X-rays as it travels along the magnetic loop. Here BQ = m2ec3/eℏ ≈ 4.4 × 1013 G. Numerical experiments in BT07 show that the e± discharge is intermittent on a timescale ∼r/c and proceeds in the regime of self-organized criticality. The time-averaged current equals the current imposed by the magnetospheric twist, (c/4π)∇ ×B, and the time-averaged voltage Φe is close to ${\cal V}_{\pm }$.

The discharge voltage remains almost independent of the imposed current j unless j is reduced below j. The value of j is unknown but small. It may be comparable to cρGJ, where ρGJ is the corotation charge density (Goldreich & Julian 1969) that should be maintained in the magnetosphere in the absence of electric currents. For the purposes of the present paper, the following description for Φe(j) will be sufficient:

Equation (34)

where j is much smaller than the characteristic currents induced by the starquake. As will be shown below, neither the value of j nor the behavior of Φe(j) in the transition region jj matters. In essence, $\Phi _e={\cal V}(u)\Theta (j)$, where Θ is the Heaviside step function. In computer simulations, we use a smoothed step function,

Equation (35)

where Δ ≪ 1. Note that the discharge voltage ${\cal V}$ can be different for different flux surfaces, i.e., ${\cal V}$ in general depends on u. There is a sharp drop in ${\cal V}(u)$ at u = 1 (voltage is small inside the highly conducting star). In computer simulations, we model this drop by introducing the factor exp{[epsilon/(1 − u)]10} with epsilon ≪ 1. The exact form of this factor plays no role for the twist dynamics.

5.3. Expanding Cavity

Consider again the twist shown in Figure 3, and let us calculate its evolution with the new threshold relation between j and Φe (Equation (34)). The numerical solution of Equations (28), (29), and (35) is shown in Figure 5 for the simplest model that assumes ${\cal V}(u)={\rm const}$ in the magnetosphere and ${\cal V}(u)=0$ inside the star.

Figure 5.

Figure 5. Evolution of the magnetospheric twist with initial uniform amplitude ψ0 = 0.2 (Figure 3). The evolution is caused by the discharge voltage Φe that is described by Equation (34), with ${\cal V}(u)={\rm const}$ at u < 1 and ${\cal V}(u)=0$ at u > 1. The dashed curve shows Φe(u); it vanishes in the shaded region u > 1 (inside the highly conducting star). The dotted curve shows the initial current distribution dI0/du. The solid curve shows the current distribution ∂I/∂u at time $t=0.01t_{\cal V}$ (upper panel) and $t=0.04t_{\cal V}$ (lower panel), where $t_{\cal V}=\mu /cR{\cal V}$; current is plotted in units of $\hat{I}=c\mu /4R^2$. Immediately following the starquake (t = 0), a cavity with jj ≈ 0 forms at u = 1 and grows with time. Its boundary—the current front—moves to smaller u, i.e., larger Rmax = R/u, erasing the magnetospheric currents. The red shaded area shows the magnetospheric current I and the current sheet I1 at u = 1. The total current Itot = I + I1 is conserved: the erased magnetospheric current flows in the current sheet.

Standard image High-resolution image

Initially, the drop in Φe near u = 1 is very steep, i.e., ∂Φe/∂u is large and negative, and hence the current density here is quickly reduced (cf. Equation (27)). The reduction continues until jj, which permits $\Phi _e<{\cal V}$. Then a smooth profile of $0<\Phi _e(u)<{\cal V}$ is established in a region u(t) < u < 1. The threshold nature of the discharge leads to the formation of two distinct regions in the magnetosphere:

  • 1.  
    "Cavity" u < u < 1 where $0<\Phi _e<{\cal V}$ and jj (essentially zero current). The current originally injected in this region is sucked into the star and flows in the current sheet at u = 1.
  • 2.  
    Region u < u where $\Phi _e={\cal V}$. Here the current remains equal to its initial value at t = 0, i.e., the twist remains static.9

The voltage profile in the cavity has ∂Φe/∂u < 0 which implies ∂I/∂t < 0 (Equation (28)). Hence the cavity must grow, as indeed seen in the simulation. The structure of the untwisting magnetosphere is shown in Figure 6.

Figure 6.

Figure 6. Structure of untwisting dipole magnetosphere (poloidal cross section). The magnetospheric currents are confined to the j-bundle u < u (the outer region shaded in yellow). Field lines shown in green are potential (∇ ×B = 0) and form the inner cavity with j ≈ 0. The cavity is bounded by the current front (located at the magnetic flux surface u = u, shown by the red curve). The front expands with time, moving to flux surfaces closer to the magnetic axis, and the j-bundle shrinks. The erased currents (that initially flowed in the cavity) are closed inside the star and flow in the current sheet (thick blue curve). The final state shown in Figure 2 is achieved when the current front u reaches the magnetic axis u = 0.

Standard image High-resolution image

The boundary of the cavity forms a sharp front, which resembles a shock wave. The front starts as a tiny arc emerging from the star at the magnetic equator, and propagates toward smaller u (i.e., outward and poleward, to flux surfaces with larger Rmax). We denote its instantaneous position by u(t). The profile of the front—the shape of j(u) near u—is controlled by the behavior of Φe at jj.10 However, we are not interested in the exact profile of the front. It is sufficient to know that it is steep and can be treated as a step function. The quantity of interest is the speed of the front propagation du/dt. In the limit of small j (steep front), one can use

Equation (36)

and derive an explicit expression for du/dt (see Appendix B),

Equation (37)

where ${\cal V}^\prime =d{\cal V}/du$ and I0(u) ≡ I(u, 0) is the initial current function. This ordinary differential equation can be solved for u(t). If ${\cal V}^\prime =0$ (as in the model in Figure 5) the front equation simplifies to

Equation (38)

The history of the front propagation is shown in Figure 7. The front starts with extremely high speed near u = 1, then decelerates and approaches the axis u = 0 with $du_{\star }/dt=-cR{\cal V}/2\mu \psi _0$. It reaches u = 0 (and erases all of the twist) at $t_{\rm end}=\mu \psi _0/cR{\cal V}$.

Figure 7.

Figure 7. Lower panel: propagation of the current front u(t) in the magnetosphere with the initial twist ψ0(u) = 0.2 and a discharge voltage ${\cal V}(u)={\rm const}$. Time is expressed in units of $t_{\cal V}=\mu /cR{\cal V}$. Immediately after the starquake, the front emerges from the star at the magnetic equator and begins to expand outward and poleward. The front reaches u = 0 (the magnetic axis) and erases all of the magnetospheric current in a finite time $t_{\rm end}=\psi _0t_{\cal V}$. The dashed line shows the slope of u(t) when u approaches 0. This slope equals −(2ψ0)−1. Upper panel: evolution of the dissipation power L (magnetospheric luminosity). L is expressed in units of $L_{\cal V}={\cal V}\hat{I}=c\mu {\cal V}/4R^2$. The dashed curve corresponds to the dashed line in the bottom panel. Both panels show the model with ψ0 = 0.2. Similar plots for twists with different initial ψ0 = const are obtained by simple stretching of the time coordinate tt0/0.2); the evolution slows down by the factor ψ0/0.2 for stronger twists.

Standard image High-resolution image

The current function I of the evolving twist is given by Equation (B8) in Appendix B. The corresponding twist amplitude is given by

Equation (39)

where I(t) ≡ I[u(t), t] is the net current that flows through the magnetosphere at time t. Equation (39) together with u(t) gives a complete analytical description for untwisting magnetospheres (in the linear-twist approximation; see Section 4.2).

Note that the numerical simulation shown in Figure 5 gives $\Phi _e/{\cal V}<1$ for u ≪ 1, i.e., for flux surfaces with footpoints near the magnetic axis. This is not predicted by the analytical model, which assumes j = 0. The drop appears in the numerical simulation at a finite u ≪ 1 because the numerical model assumes a finite j. The twist with ψ0(u) = const has the current density near the axis j(u) ∝ u2 and hence, in a small polar region where jj, Φe must drop. In the limit j → 0, this effect disappears in the sense that the polar region with jj shrinks to one point u = 0.

The twist behavior on the axis becomes important when the spindown of the star is of interest. The model in Figure 5 assumes a finite j; however, it does not take into account rotation of the star; therefore j(0) = 0 and Φe(0) = 0. Rotation with angular velocity Ω implies the additional twisting and opening of the field lines that extend to the light cylinder Rlc = c/Ω. This persistent "external" twisting induces small but finite currents on the magnetic dipole axis, along the bundle of open field lines. The open bundle has the parameter ulc = R/Rlc ≪ 1. In particular, observed magnetars have ulc ≲ 10−4. As long as the behavior of the closed magnetosphere u > ulc is concerned, the rotation can be neglected, and j → 0 is a good approximation.

5.4. j-bundle with Growing Twist

As the cavity expands, the current-carrying region u < u shrinks. We will call the current-carrying bundle of magnetic field lines "j-bundle," for brevity. The numerical model of Section 5.3 (Figures 5 and 7) assumed that the discharge voltage is the same for all magnetospheric flux surfaces, ${\cal V}(u)={\rm const}$. Then the twist remains static inside the j-bundle. It freezes and waits while it is eaten by the expanding front u(t).

In contrast, if ${\cal V}(u)\ne {\rm const}$, the twist in the j-bundle will change linearly with time as it waits for the front to come. This change is described by Equation (27) (note that $\Phi _e={\cal V}(u)$ inside the bundle u < u) or Equation (39). The twist amplitude ψ decreases if ${\cal V}^\prime <0$ and grows if ${\cal V}^\prime >0$. Observational data (discussed below) suggest ${\cal V}^\prime >0$ and the growth of ψ near the axis u = 0. Despite the twist growth at small u, its total energy Etw is decreasing with time as the j-bundle shrinks. This evolution is consistent with the energy conservation law (Equation (21)).

For illustration, we calculated the same model as in Figure 5 but with new threshold voltage ${\cal V}(u)=(0.04+2u)^{1/2}\bar{{\cal V}}$, where $\bar{{\cal V}}$ approximately equals the average of ${\cal V}(u)$ in the magnetosphere 0 < u < 1. Figure 8 shows the evolution of the twist amplitude ψ in this case and compares it with the case of ${\cal V}(u)={\rm const}$.

Figure 8.

Figure 8. Evolution of the twist profile ψ(u). (a) Model with ${\cal V}(u)={\rm const}$ (same model as in Figure 5). (b) Similar model but with ${\cal V}(u)=(0.02+2u)^{1/2}\bar{{\cal V}}$. Solid curves show ψ(u, t) at three different moments of time t. Time is expressed in units of $t_{\cal V}=\mu /cR{\cal V}$ in panel (a) and $t_{\cal V}=\mu /cR\bar{{\cal V}}$ in panel (b). The dashed curve shows ψ = ψ(u)—the twist amplitude at the boundary of the cavity throughout the entire history of its expansion from u = 1 to u = 0.

Standard image High-resolution image

The twist evolution is described by simple analytical formulae when the j-bundle is narrow, u ≪ 1. In the leading order of u ≪ 1 the front equation becomes

Equation (40)

where prime denotes the differentiation with respect to u. The relation between I and ψ (Equation (26)) gives $I_0(u)=(c\mu \psi _0/4R^2)u^2+{\cal O}(u^3)$ where ψ0(u) = ψ(u, 0) is the initial amplitude of the twist. Then we get,

Equation (41)

If ${\cal V}^\prime (0)\ne 0$, integration of this equation yields

Equation (42)

where tend is the time when the front reaches u = 0; it is finite in all cases.

The twist amplitude ψ(u, t) inside the j-bundle grows according to Equation (39) (unless ψ reaches ψmax ≳ 1),

Equation (43)

It grows by a large factor if tendt0.

For example, consider a model with linear ${\cal V}(u)={\cal V}(0)+{\cal V}^\prime (0)u$ at $u<\hat{u}$ and constant ${\cal V}(u)={\cal V}(\hat{u})$ at $u>\hat{u}$. Suppose ${\cal V}(\hat{u})\gg {\cal V}(0)$. Then we find,

Equation (44)

The drop in ${\cal V}(u_{\star })$ at small u delays the arrival of the front to u = 0 and gives an exponentially longer time for the twist growth at u = 0. Thus, even a small twist with ψ0 ≪ 1 can grow to ψmax inside the j-bundle. Further growth is impeded by the MHD instability.

5.5. Localized Starquakes

Starquakes may rotate part of the crust and leave the rest of it untouched. Suppose that a ring u2 < u < u1 has been rotated. Then the twist ψ ≠ 0 is created only in the region u2 < u < u1. This implies I(u > u1) = 0, i.e. the net ejected current is zero, and hence the current density must change sign at some um in the region u2 < u < u1. The current function I0(u) reaches a maximum at um. An example of such a localized twist and its evolution are shown in Figure 9. Voltage ${\cal V}(u)=(\frac{1}{2}+u)\bar{{\cal V}}\ne {\rm const}$ is assumed in the model.

Figure 9.

Figure 9. Evolution of the magnetospheric twist that is created by rotation of a crustal ring. The initial twist amplitude ψ0 ≈ 0.2 in the region 0.3 < u < 0.7 and close to zero outside this region. The corresponding initial current distribution dI0/du is shown by dotted curve. Current is plotted in units of $\hat{I}=c\mu /4R^2$. The discharge voltage ${\cal V}(u)=(\frac{1}{2}+u)\bar{{\cal V}}$ is assumed in this model ($\bar{{\cal V}}$ is the average of ${\cal V}(u)$ in the magnetosphere 0 < u < 1). The solid curve shows the current distribution ∂I/∂u at time $t=0.003t_{\cal V}$ (upper panel), $t=0.007t_{\cal V}$ (middle panel) and $t=0.05t_{\cal V}$ (lower panel), where $t_{\cal V}=\mu /cR\bar{{\cal V}}$.

Standard image High-resolution image

Three stages may be noted in the evolution of the ring twist.

  • 1.  
    One current front is launched from u = 1. It immediately jumps to the twist boundary u1, and continues to propagate toward smaller u. In addition, two divergent current fronts are immediately launched from u = um. Thus, two cavities form in the magnetosphere. Both cavities grow until they merge: the two fronts erasing the spike of negative current (Figure 9) eventually meet and "annihilate." Stage 1 ends at this point (at time $t\approx 0.005t_{\cal V}$ where $t_{\cal V}=\mu /cR\bar{{\cal V}}$).
  • 2.  
    The merged cavity continues to expand poleward and erase the remaining spike of positive current at smaller u. All of the initially injected current is erased (and stage 2 ends) at $t\approx 0.05t_{\cal V}$.
  • 3.  
    The current front proceeds toward the axis, erasing the currents that have grown there (from zero) since the beginning of the twist evolution.

Voltage ${\cal V}(u)\ne {\rm const}$ was chosen in the model to allow the twist growth near the axis. The simulation shows, however, that the growth is slow compared to the expansion of the cavity. Only at the last stage (stage 3) does the grown current create a significant twist ψ near the axis and somewhat decelerate the expansion of the cavity.

Figure 10 shows the evolution of the twist luminosity L(t). Its decrease is quickest during stage 1, as the spike of negative current is quickly erased. Figure 10 also shows the propagation of the current front u(t) that starts at um and moves to smaller u, erasing the ejected positive current.

Figure 10.

Figure 10. Upper panel: evolution of the dissipated power L in the model shown in Figure 9. L is expressed in units of $L_{\cal V}=\hat{I}\bar{{\cal V}}=c\mu \bar{{\cal V}}/4R^2$. Time is expressed in units of $t_{\cal V}=\mu /cR\bar{{\cal V}}$. Lower panel: outward-propagating front that erases the ejected positive current (cf. Figure 9). This front starts at um ≈ 0.63 where the ejected current density jdI0/du changes sign. Vertical dotted lines mark stages 1–3 in the twist evolution (see the text).

Standard image High-resolution image

Instead of a ring u2 < u < u1, an axisymmetric localized starquake may rotate a cap u < u1. The cap-twist and ring-twist evolve in a similar way. In particular, stages I and II are similar. However, stage 3 is absent for the cap-twist. It has u2 = 0, i.e., the initially ejected positive currents occupy the entire region around the polar axis. Erasing these currents takes a longer time. As a result, the twist grows to a large amplitude at u ≪ 1 before the cavity reaches the axis (Section 5.4). Then a maximally twisted narrow bundle with ψ = ψmax forms.

6. OBSERVATIONAL EFFECTS

6.1. Luminosity

Energy dissipation in the untwisting magnetosphere is confined to the bundle of current-carrying field lines u < u (j-bundle). It can be a bright source of radiation, with luminosity equal to the rate of Ohmic dissipation L. The luminosity generally decreases as the magnetosphere untwists. Examples of the evolution of L(t) are shown in Figures 7 and 10.

A large fraction of the dissipated power may be radiated quasi-thermally at the footprints of the j-bundle as the accelerated magnetospheric particles run into the star (BT07), creating a hot spot on the surface θ < θ. The area of this spot is given by

Equation (45)

As the cavity expands in an untwisting magnetosphere (Figure 6), the spot shrinks.

The evolution of a narrow j-bundle (u ≪ 1) with a uniform twist ψ (e.g., ψ ≈ ψmax ∼ 1) is described by simple formulae. The free energy of the twist (Equation (19)) is then given by11

Equation (46)

where R6R/106 cm, B14Bpole/1014 G, and Bpole ≡ 2μ/R3. The luminosity of the j-bundle with ψ ≈ const can be immediately calculated for a given voltage ${\cal V}(u)$,

Equation (47)

The simplest model with ${\cal V}(u)={\rm const}$ gives

Equation (48)

where ${\cal V}_9\equiv {\cal V}/10^9$ V. The j-bundle with ${\cal V}(u)={\rm const}$ shrinks according to Equation (38), which yields at u ≪ 1

Equation (49)

This equation is also easy to derive from dEtw/dt = −L, using Equations (46) and (48). The evolution timescale of the luminosity is given by

Equation (50)

Approximate formulae can also be derived for ${\cal V}(u)\ne {\rm const}$. For example, for ${\cal V}(u)={\cal V}(0)+{\cal V}^\prime (0)u$ we find

Equation (51)

This equation again assumes ψ(u < u) ≈ const. It may approximately describe e.g., the maximally twisted j-bundle, ψ ≈ ψmax.

The decay of the twist luminosity L(t) was previously estimated assuming that the current is decaying uniformly in the twisted region, which gave a linear L(t) ∝ ttend (BT07). The electrodynamic theory developed in this paper shows that the untwisting is strongly non-uniform: the twist is erased by the propagating front that resembles a shock wave. The speed of this front depends on the initial twist configuration ψ0(u). Similar to the simple estimate L(t) ∝ ttend, we find that the twist is erased in a finite time and L(t) vanishes at tend (unless new starquakes occur). However, no universal linear shape of L(t) is predicted. In some cases L(t) may resemble a linear decay (e.g., segment II in Figure 10 is almost linear in a linear plot). In observed sources, L(t) was close to linear in AXP 1E 1048.1−5937 (Dib et al. 2009) and nonlinear in XTE J1810−197 (Gotthelf & Halpern 2007).

6.2. Nonthermal Radiation

The energy released in the j-bundle can power nonthermal magnetospheric emission. Such emission is observed in most magnetars. Two distinct nonthermal components are detected in their spectra: (1) soft X-ray tail that extends from 1 keV to ∼10–20 keV with a photon index Γ ∼ 2–4 (e.g., Woods & Thompson 2006 and references therein) and (2) hard X-ray component that extends to ∼300 keV with Γ ∼ 0.8–1.5 (e.g., Kuiper et al. 2008 and references therein).

The 1–20 keV tail is usually explained by resonant scattering of thermal radiation by the magnetospheric plasma (Thompson et al. 2002; Lyutikov & Gavriil 2006; Fernández & Thompson 2007; Nobili et al. 2008; Rea et al. 2008). Ions resonantly scatter thermal photons near the star where ℏeB/mic∼ keV, and e± scatter at radii r ∼ 10R where ℏeB/mec∼ keV. The growth of cavity in the untwisting magnetosphere implies that the scattering in the inner magnetosphere is suppressed, because the dense plasma is confined to the narrow j-bundle. At larger radii rR/u, the j-bundle broadens and forms an outer corona that subtends a large solid angle as viewed from the star. The cyclotron energy in this region is

Equation (52)

and resonant scattering by e± can give 1–20 keV photons. The luminosity expected from a strongly twisted j-bundle with u ∼ 0.1 is consistent with the typical nonthermal luminosity of magnetars, L ∼ 1035 erg s−1 (see Equation (48) and substitute the typical μ ∼ 3 × 1032 G cm3 and ψ = 1–2). If no new starquakes occur, the j-bundle must shrink toward the magnetic axis with time, and u will be reduced below 0.1. Then the resonant scattering must be suppressed. This suppression is caused by two reasons: ℏeB/mec in the outer corona rR/u decreases below keV (it is proportional to u3), and the power dissipated in the j-bundle becomes small as Lu2.

The j-bundle with u ∼ 0.1–0.2 dissipates sufficient energy to explain also the hard X-ray component. This component was detected in three AXPs and two soft gamma-ray repeaters (SGRs; see Kuiper et al. 2008 for a recent review). In AXPs, the hard X-ray emission has a huge pulsed fraction, approaching 100% at high energies (Kuiper et al. 2006; den Hartog et al. 2008a, 2008b). This may be explained if the emission is produced in the narrow j-bundle near the star.

Observations indicate that the nonthermal emission can be stable on timescales as long as a decade (den Hartog et al. 2008a). There may be two reasons for this stability. (1) The discharge voltage ${\cal V}$ is relatively low (below 1 GeV) and the j-bundle is relatively thick, u ∼ 0.2. Then the untwisting timescale becomes long (see Equation (50)). (2) The j-bundle is kept in a quasi-steady, maximally twisted state ψ ∼ ψmax by frequently repeating (possibly continual) shearing motion of the crust in a fixed region u < u.

The plasma filling the j-bundle may also produce optical and infrared radiation by mechanisms discussed in BT07. Besides, it can be a bright source of radio waves, as suggested in Section 7.2.

6.3. Outer Magnetosphere and Spindown Rate of the Star

The magnetospheric twist is expected to impact the spindown rate when the twist amplitude is large, ψ ≳ 1. This impact may occur in two ways.

  • 1.  
    The strong twist inflates the poloidal field lines and increases the magnetic field at the light cylinder, which leads to stronger spindown torque acting on the star (Thompson et al. 2002). The poloidal inflation is common for twisted configurations. It is seen, e.g., in the self-similarly twisted dipole (Wolfson 1995). A similar inflation must occur when the currents are confined to the j-bundle. Its calculation will require a full nonlinear model. Here we limit our consideration to simple estimates.One can think of poloidal inflation as an increase of magnetic dipole moment with radius. This effect is easiest to evaluate for moderate twists ψ < 1. The current dI flowing along a bundle of twisted field lines du creates a toroidal current dIϕdI ψ(u)/2π. The field lines carrying this current extend to Rmax = R/u, and the dipole moment created by dIϕ is dμ ∼ dIϕR2max/c. Integrating over u, we obtain the net change of dipole moment of the star due to the twist
    Equation (53)
    The j-bundle ulc < u < u with a uniform twist ψ increases the dipole moment of the star by
    Equation (54)
    This effect is quadratic in ψ and quickly becomes small for ψ < 1. For strong twists ψ ∼ ψmax, the estimate (54) must be replaced by a full nonlinear calculation. Qualitatively, it suggests that Δμ/μ is reduced as the j-bundle shrinks (u decreases). Therefore, the spindown torque is expected to be reduced with time and gradually come back to the standard dipole torque as uulc.
  • 2.  
    When the twist has grown to ψmax, it will begin to "boil over" through a repeated instability. The energy that would be stored in the magnetosphere if ψ kept growing above ψmax is then carried away by an intermittent magnetic outflow. The outflow may also carry away a significant angular momentum.The outflow can be generated where the overtwisted field lines open up. This opening occurs because the twist is constantly pumped near the axis by ${\cal V}^\prime >0$. A possible structure of untwisting magnetospheres with ${\cal V}^\prime >0$ is schematically shown in Figure 11. It resembles the picture of a rotationally powered pulsar (see, e.g., Arons 2008 for a review); however, the opening is caused by the internal resistive dynamics of the magnetosphere and depends on the profile of ${\cal V}(u)$. By contrast, in ordinary pulsars the twist is pumped by the star rotation.
Figure 11.

Figure 11. Possible structure of untwisting magnetospheres (poloidal cross section). The growing twist near the axis inflates the outer magnetosphere until its field lines open. Then the j-bundle (shaded in yellow) becomes confined between the last closed flux surface (blue curve) and the inner cavity (red curve), which expands with time. An equatorial outflow is expected to form just outside the last closed flux surface. The outflow is driven by $d{\cal V}/du>0$, which forces the twist to grow until part of the overtwisted field lines reconnect away from the star. The small circle in the center (shaded in cyan) shows the neutron star.

Standard image High-resolution image

Mechanisms (1) and (2) start immediately after the starquake if it implants a strong initial twist ψ0 > 1 near the magnetic axis. If ψ0 < 1, the spindown torque may not be affected until ψ grows to ∼1, which takes time (cf. Equation (27) or Equation (39))

Equation (55)

where ${\cal V}^\prime =d{\cal V}/du$ is evaluated near the axis u ≈ 0; ${\cal V}^\prime$ may be large, because u is small. For example, a change in ${\cal V}$ from ${\cal V}=10^9$ V at u = 0 to 2 × 109 V at u = 10−2 corresponds to ${\cal V}^\prime =10^{11}$ V. The delay is observed in some objects, with a characteristic tdelay ∼ 107 s (e.g., Gavriil & Kaspi 2004). This requires ${\cal V}^\prime \sim 10^{11}$ V.

In contrast to luminosity L(t) (which always tends to decrease after the starquake), the torque behavior is generally non-monotonic. If ψ0 < 1, the torque is expected to grow as ψ grows in the shrinking j-bundle. (Such an anti-correlation between the torque and the X-ray luminosity was observed in 1E 1048.1−5937; see Gavriil & Kaspi 2004). Once ψ has reached ψmax ∼ 1 the torque should start to decrease, because the shrinking of the j-bundle at constant ψ ≈ ψmax leads to the reduction of Δμ. The power of the outflow from the magnetosphere is also expected to decrease. The simultaneous decrease in torque and luminosity was observed, e.g., in XTE J1810−197 (Camilo et al. 2007).

7. UNTWISTING MAGNETOSPHERE IN XTE J1810−197

XTE J1810−197 is an AXP with period P = 5.54 s and estimated dipole magnetic moment μ ∼ 1.5 × 1032 G cm3, which corresponds to the surface field at the polar cap B ∼ 3 × 1014 G (Gotthelf & Halpern 2007). An X-ray outburst was detected from this object in January 2003 (Ibrahim et al. 2004). Its luminosity approximately followed an exponential decay on a timescale of 233 days for three years (Gotthelf & Halpern 2007). During the X-ray decay, the source became radio-bright (Halpern et al. 2005) and switched on as a powerful radio pulsar with unusual spectrum and pulse-profile variations (Camilo et al. 2007). The spindown rate of the star dramatically increased following the outburst. In the subsequent years, the object gradually evolved toward its quiescent (pre-outburst) state.

These observations clearly indicate that the magnetosphere of XTE J1810−197 changed in the outburst, i.e., the footpoints of field lines must have moved, imparting a twist to the magnetosphere. The observational data give significant hints about the twist geometry and evolution.

  • 1.  
    The change in spindown rate suggests that the open field-line bundle was affected by a strong twist ψ ≳ 1 in the closed magnetosphere near the magnetic dipole axis. On the other hand, ψ is limited to $\psi _{\rm max}={\cal O}(1)$ by the MHD instability. Therefore, we infer $\psi ={\cal O}(1)$ near the axis.
  • 2.  
    Already one year after the outburst, the object luminosity was below 1035 erg s−1. The theoretically expected luminosity from a global twist with ψ ∼ 1 (Equation (48)) would be much higher: $L\sim 3\times 10^{36}{\cal V}_9$ erg s−1, and it would decay much slower than observed (Equation (50)). Therefore, we conclude that the twisted region was small: the current-carrying field lines formed a narrow bundle emerging from a small spot on the star surface.
  • 3.  
    Remarkably, the spot was discovered: a hot blackbody component with a small emission area was found in the X-ray spectrum following the outburst (Gotthelf & Halpern 2007; Perna & Gotthelf 2008). Its emission area shrank with time until the spot became barely detectable (Figure 12).
  • 4.  
    The X-ray and radio pulse profiles had almost simultaneous peaks, consistent with the X-ray hot spot being near the magnetic dipole axis (Camilo et al. 2007).
Figure 12.

Figure 12. Comparison of the model with the observed evolution of the area A and luminosity L of the hot spot formed after the outburst in XTE J1810−197. The data (open squares) are from Gotthelf & Halpern (2007). The dashed line shows the object luminosity in quiescence. The solid curve shows the theoretical model (see the text after Equation (56)).

Standard image High-resolution image

These observations are consistent with the theory developed in the present paper. The hot spot is explained as the footprint of the j-bundle on the star (Section 6.1). It has a sharp boundary and is shrinking with time as the j-bundle shrinks toward the magnetic dipole axis. The large amplitude of the twist ψ ∼ 1 may have been created initially by the starquake, but not necessarily: ψ can naturally grow to ≳1 following the starquake (Section 5.4). A quantitative comparison of the model with the observational data is given below.

7.1. X-ray Emitting Spot

Useful preliminary estimates can be made if we assume a uniform voltage ${\cal V}(u)\approx {\rm const}$ and uniform twist ψ(u) ≈ const across the j-bundle u < u (Section 6.1). Then the produced luminosity L(t) is given by Equation (48). Suppose a large fraction of L is emitted thermally at the footprint of the j-bundle. The estimates for L (Equation (48)) and its evolution timescale tev (Equation (50)) for a given spot area A (Equation (45)) may be compared with observations. For example, in the fall of 2004, the spot area was A ≈ 1011 cm (and hence u ≈ 0.03). One then finds that a strongly twisted j-bundle (ψ ∼ 1–1.5) with voltage ${\cal V}_9\sim 3$ explains both observed L ≈ 2 × 1034 erg s−1 and tev ≈ 0.6 yr.12

The model ${\cal V}(u)\approx {\rm const}$ gives good estimates for L, A, and tev; however, it has a drawback: it is unable to describe the possible growth of ψ from a smaller ψ0 before the j-bundle became maximally twisted. The growth occurs if ${\cal V}^\prime >0$ (Section 5.4). Therefore, we adopt a slightly more general model that includes the next (linear) term in the expansion of ${\cal V}(u)$ near u = 0,

Equation (56)

${\cal V}^\prime$ is unknown and probably large near the axis (see the text after Equation (55)).

A simplest model of a twisted magnetosphere has four parameters: magnetic dipole moment of the star μ, radius of the star R, the initial size of the j-bundle created by the starquake u0 = sin2θ0, and the initial amplitude of the twist ψ0 (it equals the angular displacement of the crustal cap rotated by the starquake; in our fiducial model, the starquake imparts a uniform twist ψ0(u < u0) = const). The twist evolution after the starquake is controlled by two parameters ${\cal V}_0$ and ${\cal V}^\prime$ that specify voltage ${\cal V}(u)$ in Equation (56). The evolution is described by Equations (39) and (37) that we solve numerically. If ψ reaches $\psi _{\rm max}={\cal O}(1)$, it is assumed to stall at ψmax.13 The luminosity of a maximally twisted j-bundle, L(u), is approximately given by Equation (51).

Figure 12 compares the model with observations of XTE J1810−197. The star is assumed to have μ = 1.5 × 1032 G cm3 and R = 9 km. The parameters of the starquake are u0 = 0.15 and ψ0 = 0.5. The discharge voltage is assumed to drop linearly from 5.5 GeV at u = 0.15 to 1 GeV at u = 0 (which corresponds to ${\cal V}_0=10^9$ V and ${\cal V}^\prime =3\times 10^{10}$ V). Although we did not attempt a formal fitting of the data, the figure suggests that the model is successful in explaining the evolution of L(t) and A(t). The data firmly constrain the voltage to be in 1–6 GeV range, which is close to the theoretical estimate (Equation (33)).

The accuracy of the model is limited by several idealizing assumptions: uniform initial twist in the starquake region, the linear form of ${\cal V}(u)$, and axial symmetry. The actual value of μ is probably smaller than inferred from spindown measurements (Section 6.3). The linear twist-evolution equation becomes approximate when ψ ∼ 1. Note also that our simplest fiducial model assumes that all energy dissipated in the j-bundle is emitted thermally at one (e.g., anode) footprint. A more realistic model can predict a more complicated spectrum with both footprints emitting. Finally, the hot-spot emission may be significantly anisotropic (Perna & Gotthelf 2008), which may increase or reduce its apparent luminosity depending of the average inclination of the spot to the line of sight.

Note that the observed spectrum was well fitted by a two-temperature blackbody. The second blackbody component was cooler and had emission area comparable to that of the star. No nonthermal component from magnetospheric scattering was required by the data. The weakness of the scattered component may be explained by the small size u of the j-bundle: the star's radiation does not scatter in the cavity around the j-bundle, and the probability of scattering in the narrow bundle is low (Section 6.2).

7.2. Radio Pulsations

In ordinary radio pulsars, radio emission is believed to be produced by the open field lines passing through the light cylinder, because they are the only part of the magnetosphere that carries electric currents. In magnetars, these field lines have a tiny ulc = R/Rlc ≲ 10−4 where Rlc = c/Ω is the light-cylinder radius for a star rotating with angular velocity Ω. Radio emission from the open bundles of magnetars may be undetectable for two reasons. (1) e± discharge has a low threshold in magnetars, ${\cal V}\sim 10^9$ V. When this voltage is multiplied with the current in the bundle, $I_{\rm lc}\approx \hat{I}u_{\rm lc}^2=c\mu /4R_{\rm lc}^2$, one obtains the dissipated power $L_{\rm lc}={\cal V}I_{\rm lc}\sim 10^{28}{\cal V}_9$ erg s−1. This power appears to be too small to feed the observed radio luminosity of XTE J1810−197, Lradio ∼ 1030 erg s−1. (2) The radio beam may be narrow (because of small ulc). Then the probability of its passing through our line of sight is small. This suggests that radio pulsations are hardly detectable when the magnetar is in quiescence, i.e., when its magnetosphere is untwisted.

In contrast, after the starquake, the j-bundle forms. It is much thicker and more energetic than the bundle passing through the light cylinder and can produce much brighter radio emission with a much broader pulse. The untwisting magnetosphere in XTE J1810−197 had u/ulc ≳ 3 × 102. The net current flowing in the j-bundle is ∼105 times larger than Ilc,

Equation (57)

This gives L/Llc ∼ 105 (assuming a comparable discharge voltage at u and ulc). A small fraction epsilonradio of this luminosity may escape as radio waves; epsilonradio ∼ 10−3 is consistent with observations.

Radio waves are efficiently absorbed by the magnetospheric plasma. Only waves produced close to the magnetic axis, u < uesc, are likely to escape, while waves emitted at u > uesc are trapped in the magnetosphere. The value of uesc depends on the frequency of the wave and the state of the plasma (particle density and velocity distribution); uesc might be inferred from the observed opening angle of the radio beam. The radio luminosity should quickly decrease when the j-bundle shrinks to u < uesc.

Radio pulsar XTE J1810−197 is distinguished from ordinary radio pulsars by its hard spectrum, very strong linear polarization, and variable pulse profile. If its emission is produced on the bundle of closed field lines with uulc, it may be expected to be different from ordinary radio pulsars. The spectrum of radio waves may form as the sum of emissions from different (frequency-dependent) uesc inside the j-bundle. The sporadic changes in the pulse profile may be caused by the instabilities in the maximally twisted outer magnetosphere (Section 6.3).

Radio observations provided accurate measurements of the spindown torque acting on the star (Camilo et al. 2007). The torque was decreasing together with the X-ray luminosity 2–3 years after the starquake. Its history at earlier times was not observed; the torque is believed to have increased after the starquake (Camilo et al. 2007). Such a non-monotonic evolution of the torque would be consistent with the theoretical expectations (Section 6.3).

8. DISCUSSION

Electrodynamics of untwisting may be summarized as follows.

  • 1.  
    The twist evolution following a starquake is not diffusive spreading that was pictured previously. Instead, the twist current is sucked into the star, a cavity immediately forms in the inner magnetosphere and grows until it erases all of the twist (Figure 6). A sharp, step-like drop in current density j is maintained at the boundary of the cavity. It is caused by the threshold nature of the discharge that conducts magnetospheric currents.
  • 2.  
    As the cavity expands and the j-bundle shrinks toward the magnetic axis, the twist amplitude ψ in the j-bundle can grow. The growth occurs if $d{\cal V}/du>0$, i.e., the discharge voltage is smaller on field lines extending farther from the star. This is plausible if the current is conducted through e± discharge.14 The j-bundle with the growing twist is shrinking with time so that the total twist energy Etw decreases consistently with the Ohmic dissipation rate.
  • 3.  
    The growing twist that has reached the threshold for instability $\psi _{\rm max}={\cal O}(1)$ is expected to "boil over" and drive an intermittent outflow of magnetic energy from the star. The twist in the j-bundle then remains near ψmax for the rest of its lifetime. It may be regulated by the limit-cycle instability—the repeated growth of ψ to ψmax followed by a sudden reduction of ψ below ψmax. The value of ψmax and the nonlinear behavior of the outer magnetosphere at ψ ∼ ψmax needs to be studied further using numerical simulations.
  • 4.  
    In addition to rare large starquakes, the magnetosphere can be gradually twisted by the continued motion of its footpoints, either plastic or through a sequence of small starquakes. The continued footpoint motion leads to either a very strong twist or a tiny (negligible) twist that has no observational effects. A quasi-steady state with a non-negligible twist amplitude is possible only with ψ ∼ ψmax.

Recent observations of XTE J1810−197 are particularly useful for testing the untwisting theory, for a few reasons. (1) XTE J1810−197 displayed a clean post-starquake evolution, which was observed for years uninterrupted by new starquakes. (2) The low level of quiescent luminosity from this object allows one to see clearly the shrinking hot spot on the star, and its evolving luminosity was tracked from 1035 erg s−1 down to 1033 erg s−1. (3) The detection of radio pulsations and detailed measurements of spindown torque make this object yet more interesting for testing theoretical models. For these reasons, this paper focused mainly on XTE J1810−197 (Section 7). Recently, an outburst was detected in the similar radio magnetar 1E 1547.0−5408 (Camilo et al. 2008; Halpern et al. 2008). Its behavior appears to be more complicated than that of XTE J1810−197; apparently, episodes of repeated (and overlapping in time) activity occurred a few months after the outburst. The analysis of this object is deferred to a future work.

Other, more active, AXPs and SGRs display a diverse and complicated behavior of X-ray luminosity, pulse profile, and spindown rate (see Woods & Thompson 2006; Kaspi 2007; Mereghetti 2008 for reviews). Repeated starquakes of various amplitudes and possible plastic deformations of the crust make these objects more difficult to analyze. The continuing injection of a magnetospheric twist can slow down the decay of its luminosity. It can also affect the star's spindown in a more complicated way than described in Section 6.3 for an isolated starquake. In general, twist injection should lead to higher X-ray activity and faster spindown (Thompson et al. 2002). Such a general correlation exists in the magnetar population (Marsden & White 2001) but not always observed in individual objects. Note that spindown is controlled by the X-ray-dim bundle of open field lines. The impact of a starquake on this narrow bundle may be immediate, occur with a delay, or never occur, depending on the starquake geometry and amplitude (cf. the case of a "ring" starquake in Section 5.5). Repeated starquakes may lead to a non-trivial relation between spindown and X-ray emission.

Despite the diverse behavior of active magnetars, some general features may be inferred. It is clear that the observed sources do not have strong global twists, for two reasons. First, the evolution timescale of such twists would be too long, $t_{\rm ev}\sim (10\hbox{--}10^2){\cal V}_9^{-1}$ yr, where ${\cal V}_9$ is the discharge voltage in units of 109 V (see Equation (50)). In contrast, the magnetospheres of observed magnetars usually evolve on timescales ∼1 yr or even shorter.15 Second, the luminosity produced by a strong global twist would be too high, $L\sim 10^{37}{\cal V}_9$ erg s−1 (see Equation (48)). It is 2 orders of magnitude higher than the typical observed luminosities of magnetars, L ∼ 1035 erg s−1 (e.g., Durant & van Kerkwijk 2006). This leaves two possibilities: the twist is weak (ψ ≪ 1) or localized to a narrow bundle of field lines. The changing spindown rate suggests a strong twist, at least near the magnetic dipole axis. Thus, observations appear to support the picture of a strongly twisted j-bundle near the dipole axis. It is certainly supported by the observations of XTE J1810−197 (Section 7).

The localized strong twist may be created by starquakes localized to the polar region. It also tends to form dynamically from a weak global twist as the j-bundle shrinks to the axis. The luminosity and evolution timescale of an untwisting magnetosphere with u = sin2θ ∼ 0.1 and ψ ∼ 1 is consistent with typical L ∼ 1035 erg s−1 and tev ∼ 1 yr of active magnetars (see Equations (48) and (50)). Both nonthermal X-ray components in the magnetar spectra, 1–20 keV and 20–300 keV, can be produced by the j-bundle. They are likely emitted at different radii. If the hard component is produced near the star, where the j-bundle is narrow, a relatively narrow 20–300 keV pulse is expected.

The resonant-scattering model for 1–20 keV radiation is consistent with the picture of a narrow j-bundle near the dipole axis. The cyclotron resonance for e± with keV photons takes place at radii r ∼ 10R. This means that the resonant scattering is confined to the field-line bundle with u = R/Rmax ∼ 0.1 (Section 6.2). The luminosity of upscattered radiation Lsc is supplied by Ohmic dissipation of electric currents in this bundle.

The footprints of a bundle with u ∼ 0.1 form ∼3 km spot on the star. A significant fraction of energy released in the j-bundle may be transported to its footprints and emitted there quasi-thermally. If the spot radiates a significant part of the bundle luminosity, L ∼ 1035 erg s−1, then it must have a temperature kT ≈ 1 keV. Such hot spots were observed in XTE J1810−197 and 1E 1547.0−5408, but were not reported for other, brighter magnetars. The typical reported temperatures of blackbody components are 0.3–0.6 keV (e.g., Perna et al. 2001). The disappearance of hot spots in the presence of large Lsc may be caused by the outward drag created by resonant scattering, which suppresses the transport of released energy to the footprints of the j-bundle.

This work was supported by NASA grant NNG-06-G107G.

APPENDIX A: TWISTED DIPOLE MAGNETOSPHERE

Consider a weakly twisted dipole field  B =  B0 +  Bϕ. Its poloidal component  B0 is that of untwisted dipole,

Equation (A1)

The poloidal flux function is given by

Equation (A2)

where hϕ = rsin θ and hθ = r. Since f = const on any flux surface,

Equation (A3)

which is the radius where the flux surface crosses the equatorial plane (the maximum radius reached by the flux surface).

The magnitude of the twisted magnetic field is given by (neglecting terms ${\cal O}[B_\phi ^2/B^2$]),

Equation (A4)

The twist angle ψ (Equation (3)) is easy to calculate using θ as the parameter along the field line and substituting dl = (B/Bθ)dθ,

Equation (A5)

Here θ0 is the polar angle of the northern footpoint of the field line, and π − θ0 is the polar angle of the southern footpoint. Substituting Bϕ = 2I/chϕ (Equation (4)), Bθ from Equation (A1), and using Equation (A3), one finds

Equation (A6)

where

Equation (A7)

When the poloidal current I and twist angle ψ are viewed as functions of u = f/fR and time t, the relation between ψ and I (Equation (A6)) becomes Equation (26). The free energy of a twist with given I(u) is found using Equations (19) and (26),

Equation (A8)

We also calculate here the integral that is needed in Section 5.1,

Equation (A9)

It is taken along the field line outside the star. Consider the contour that closes the path of integration between the two footpoints along the surface of the star (and across the magnetic field). The line integral of  B0 along this closed contour vanishes (as follows from Stokes' theorem and ∇ ×B0 = 0). This gives

Equation (A10)

Here θ1 and π − θ1 are the polar angles of the field-line footprints on the star surface, and $\cos \theta _1=\sqrt{1-u}$.

APPENDIX B: DERIVATION OF THE FRONT EQUATION

The front is located on the flux surface u where Φe(u) deviates from ${\cal V}(u)$ and current j jumps to ∼j (Figure 5). We consider below the limit j → 0. Then the jump of j is described by the Heaviside step function Θ(uu) and j = 0 in the region u < u < 1.

This implies that I(u) = const for u < u < 1, and this fact can be used to derive the dynamical equation for u. Let us differentiate the twist evolution equation (28) with respect to u. Then the left-hand side vanishes in the region u < u < 1 and we get

Equation (B1)

which implies

Equation (B2)

where K is constant in the region u < u < 1. Integrating this equation and using the boundary condition Φe(1) = 0 we find the shape of Φe(u),

Equation (B3)

The explicit expression for K can be found from the condition $\Phi _e(u_{\star })={\cal V}(u_{\star })$,

Equation (B4)

Equation (B5)

The numerical factor ξ(u) is well approximated by a simpler formula,

Equation (B6)

The accuracy of this approximation is better than 2.2% for 0 < u < 0.65. The approximation is worst (30% accuracy) when u = 1 and quickly becomes excellent as the front propagates away from the boundary.

The twist evolution equation (28) can now be written as

Equation (B7)

where ${\cal V}^\prime =d{\cal V}/du$. Note that ∂I/∂t does not vary with time at u < u, and hence I(u, t) can be obtained by simple integration. Thus we find the current function of the twist,

Equation (B8)

where I0(u) ≡ I(u, 0) is the initial current function, and I(t) ≡ I[u(t), t] is the net current that flows through the magnetosphere at time t. Using the relation between I and ψ (Equation (26)), we obtain Equation (39). It gives a complete solution to the problem of twist evolution if we find u(t).

The equation for u(t) can be obtained if one considers the evolution of I. Its time derivative is given by

Equation (B9)

which can be evaluated separately on each side of the front,

Equation (B10)

Equation (B11)

The two results must match because they describe the same I(t). The jump in ∂I/∂u across the front (which corresponds to the jump in j) is compensated by the jump in ∂I/∂t. Equating the two expressions for dI/dt, we find

Equation (B12)

This ordinary differential equation describes the propagation of the front. Substitution of Equation (B4) for K(u) gives Equation (37).

Footnotes

  • Well studied ordinary pulsars (Crab, Vela) also show glitches in their spindown rates, which are associated with sudden crustal deformations.

  • In some cases, the crust may move plastically.

  • The discharge on closed field lines differs from that on open field lines (see Arons 2008 and Beloborodov 2008 for a recent discussion of the polar-cap discharge in ordinary pulsars). In both cases, however, the discharge is ultimately driven by the magnetic twist ∇ ×B ≠ 0 that imposes an electric current.

  • Electric fields maintaining the currents in a twisted magnetosphere are relatively weak and do not spoil this approximation.

  • Figure 1 assumes for simplicity that there is only one such flux surface. This is the case for a bipolar magnetosphere, with two regions on the star's surface with opposite polarities of the magnetic field (i.e., opposite signs of Br). One can imagine more complicated axisymmetric magnetic configurations with many rings of opposite polarities on the star's surface. The development of electrodynamic theory would be similar in those cases. When calculating the evolution of currents on a given flux surface f we would need to know the magnetic field only between the two nearest flux surfaces that touch the surface of the star.

  • The current is maintained through a magnetospheric discharge which fluctuates on a very short (light-crossing) timescale r/c (BT07). We consider the time-average I and treat it as a quasi-steady current. Its evolution caused by resistivity is slow (year timescale) and can be viewed as a slow progression through a sequence of steady states. The displacement current vanishes in a steady state, so ∇ ×B = (4π/c) j.

  • This voltage is not electrostatic and does not vanish for a closed contour. It is the self-induction voltage of the gradually decaying twist; see Section 2 in BT07.

  • This is a consequence of ${\cal V}(u)={\rm const}$. The twist at u < u will not remain static if ${\cal V}(u)\ne {\rm const}$, as discussed below.

  • 10 

    The profile of the front is controlled by the form of function W(j/j) in Equation (35); in the simulation shown in Figure 5 we chose j = 10−2cB0/8πR and Δ = 0.2.

  • 11 

    Only 1/4 of Etw resides in the region u < u; 3/4 of the twist energy is contained in the potential region u > u. The fact that ∇ ×B = 0 in the potential region does not imply that Bϕ = 0; Bϕ is determined by Equation (4).

  • 12 

    We use here B14 ≈ 3 (Gotthelf & Halpern 2007).

  • 13 

    In the simulations, we set ψmax = 1.5. The exact value of the stability threshold needs to be calculated in the full nonlinear twist model.

  • 14 

    In contrast, if the current-carrying charges were lifted from the star by voltage (32), the voltage would be larger for field lines that extend to larger altitudes, i.e., dV/du < 0. Then the twist on the j-bundle would diminish with time rather than grow.

  • 15 

    The theoretical tev would be reduced for a larger discharge voltage ${\cal V}\gg 10^9$ V, however it appears impossible to sustain such a voltage as it leads to runaway e± creation (BT07). Even if ${\cal V}\gg 10^9$ V were theoretically possible, strong global twists would still be ruled out by observations because they are overluminous.

Please wait… references are loading.
10.1088/0004-637X/703/1/1044