This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Accumulation of elastic strain toward crustal fracture in magnetized neutron stars

Yasufumi Kojima Department of Physics, Hiroshima University
Higashi-Hiroshima,739-8526, Japan
Abstract

This study investigates elastic deformation driven by the Hall drift in a magnetized neutron-star crust. Although the dynamic equilibrium initially holds without elastic displacement, the magnetic-field evolution changes the Lorentz force over a secular timescale, which inevitably causes the elastic deformation to settle in a new force balance. Accordingly, elastic energy is accumulated, and the crust is eventually fractured beyond a particular threshold. We assume that the magnetic field is axially symmetric, and we explicitly calculate the breakup time, maximum elastic energy stored in the crust, and spatial shear-stress distribution. For the barotropic equilibrium of a poloidal dipole field expelled from the interior core without a toroidal field, the breakup time corresponds to a few years for the magnetars with a magnetic field strength of 1015\sim 10^{15} G; however, it exceeds 1 Myr for normal radio pulsars. The elastic energy stored in the crust before the fracture ranges from 104110^{41} to 104510^{45} erg, depending on the spatial-energy distribution. Generally, a large amount of energy is deposited in a deep crust. The energy released at fracture is typically 1041\sim 10^{41} erg when the rearrangement of elastic displacements occurs only in the fragile shallow crust. The amount of energy is comparable to the outburst energy on the magnetars.

Neutron stars; Compact objects; Magnetars; High-energy astrophysics

1 Introduction

Neutron-star crust is considered a key aspect for understanding several astrophysical phenomena. The solid layer near the stellar surface can support non-spherical deformations, called mountains, with a height of less than 1 cm. Such asymmetries on a spinning star cause the continuous emission of gravitational waves. Therefore, calculation of the maximum size of these mountains is key to the detection of gravitational waves and has been discussed in several theoretical studies (Ushomirsky et al., 2000; Payne & Melatos, 2004; Haskell et al., 2006; Gittins et al., 2021). Gravitational-wave observation provides valuable information (Abbott et al., 2021a, b, 2022, for recent upper limit); thus, by continuously improving the sensitivity of the LIGO-Virgo-KAGRA detectors, the physics relevant to the phenomenon may be explored.

Pulsar glitches are sudden spin-up events that are observed in radio pulsars (Espinoza et al., 2011; Basu et al., 2022, for glitch catalogue). Similar spin-up and peculiar spin-down events are observed in anomalous X-ray pulsars (Dib et al., 2008; Kaspi & Beloborodov, 2017). A sudden spin-up in a radio pulsar is produced by the transfer of angular momentum from the superfluid components of the core to the normal crust (Anderson & Itoh, 1975; Alpar et al., 1984). Crust quakes were also discussed in other models (Franco et al., 2000; Giliberti et al., 2020; Rencoret et al., 2021, for recent studies). The elastic deformation is caused by a decrease in centrifugal force, owing to a secular spin-down, and the crust eventually fractures when the strain exceeds a critical threshold. However, this simple model does not explain the observation; the loading of the solid crust between glitches is too insignificant to trigger a quake.

Giant flares in magnetars are rare, albeit highly energetic. They are typically 10441046\sim 10^{44}-10^{46} erg released within a second (Turolla et al., 2015; Kaspi & Beloborodov, 2017; Esposito et al., 2021, for a review). Quasi-periodic oscillations (QPOs) with some discrete frequencies in the range of 20 Hz – 2 kHz were observed in the tails of these flares. Per the order-of-magnitude estimate, these frequencies correspond to the torsional shear or Alfvén modes with magnetic field strength 1015\sim 10^{15} G. Outbursts, which are less energetic, are also observed in magnetars. These activities are considered to be powered by internal strong magnetic fields of 1015\sim 10^{15} G. The crustal fracture of the magnetar is proposed as a model for fast radio bursts (FRBs) (Suvorov & Kokkotas, 2019; Wadiasingh & Chirenti, 2020), and it may be supported by QPOs (Li et al., 2022) in the radio burst from SGR J1935+2154 in the galaxy (Mereghetti et al., 2020; CHIME/FRB Collaboration et al., 2020; Bochenek et al., 2020). Most FRBs are located at a cosmological distance, and further observation sheds light on whether FRBs originate from magnetars or a subclass. A recent observation of the magnetar SGR 1830-0645 revealed pulse-peak migration during the first 37 days of outburst decay (Younes et al., 2022). This provides important information concerning the crust motion coupled with the exterior magnetosphere.

Most theoretical studies have been focused on the crustal-deformation limit. Elastic stresses gradually accumulate until a particular threshold. Beyond this threshold, the elastic behavior of the lattice abruptly ceases, and the transition is exhibited as a star-quake or burst. An evolutionary calculation of the deformation is necessary to understand the related astrophysical phenomena.

In this study, we consider the crust in a magnetized neutron star. The static magneto-elastic force balance was studied for various magnetic-field configurations Kojima et al. (2021a, 2022). A variety of magneto-elastic equilibria was demonstrated, and is considerably different from the barotropic equilibrium without a solid crust. Herein, we explore the accumulation of shear stress induced by the Hall evolution, which is an important process in the strong field-strength regime. Suppose that the magneto-hydro-dynamical (MHD) equilibrium in the crust holds at a particular time without the elastic force. The equilibrium is not that for electrons (Gourgouliatos et al., 2013; Gourgouliatos & Cumming, 2014); thus, the magnetic field tends to the Hall equilibrium in a secular timescale. According to the magnetic-field evolution, the Lorentz force also changes. The deviation is assumed to be balanced with the elastic force. Thus, the shear stress in the crust gradually accumulates and reaches a critical limit. We cannot follow the post-failure evolution because some uncertainties are involved in the discontinuous transition. Therefore, our study provides the recurrent time and magnitude of the bursts.

The models and equations used in the study are discussed in Section 2. For MHD equilibrium in a barotropic star, the evolution of the magnetic field is driven by a spatial gradient of electron density. In Section 3, the critical configuration at the elastic limit is evaluated, and the accumulating elastic energy is calculated. In Section 4, we also consider non-barotropic effects using simple models. The non-barotropicity results in another driving process of the magnetic-field evolution, and consequently elastic deformation. The numerical results of these models are given. Finally, our conclusions are presented in Section 5.

2 Formulation and Model

2.1 Magnetic Equilibrium

We consider the dynamical force balance between pressure, gravity, and the Lorentz force. The MHD equilibrium is described as follows:

1ρPΦg+1cρ𝐣×𝐁=0,-\frac{1}{\rho}{{\bf{\nabla}}}P-{{\bf{\nabla}}}\Phi_{\rm g}+\frac{1}{c\rho}{{\bf{j}}}\times{{\bf{B}}}=0, (1)

where Φg\Phi_{\rm g} is the gravitational potential including the centrifugal terms. The third term has a magnitude 107(B/1014G)2\sim 10^{-7}(B/10^{14}{\rm G})^{2} times smaller than those of the first and second terms. Deviation owing to the Lorentz force is small enough to be treated as a perturbation on a background equilibrium.

We limit our consideration to an axially symmetric configuration for the magnetic-field configuration. The poloidal and toroidal components of the magnetic field are expressed by two functions Ψ\Psi and SS, respectively, as follows:

𝐁=×(Ψϖ𝐞ϕ)+Sϖ𝐞ϕ,{{\bf{B}}}={{\bf{\nabla}}}\times\left(\frac{\Psi}{\varpi}{{\bf{e}}}_{\phi}\right)+\frac{S}{\varpi}{{\bf{e}}}_{\phi}, (2)

where ϖ=rsinθ\varpi=r\sin\theta is the cylindrical radius, and 𝐞ϕ{\bf{e}}_{\phi} is the azimuthal unit-vector in (r,θ,ϕ)(r,\theta,\phi) coordinates. When the equilibrium is barotropic, i.e., the constant surfaces of ρ\rho and PP are parallel, the azimuthal current jϕj_{\phi} is described in the form

4πjϕc=ρϖdKdΨ+SϖdSdΨ,\frac{4\pi j_{\phi}}{c}=\rho\varpi\frac{dK}{d\Psi}+\frac{S}{\varpi}\frac{dS}{d\Psi}, (3)

where the current function SS should be a function of Ψ\Psi, and KK is another function of Ψ\Psi. For the axially symmetric barotropic equilibrium, acceleration due to the Lorentz force, which is abbreviated as 𝐚(cρ)1𝐣×𝐁{{\bf{a}}}\equiv(c\rho)^{-1}{{\bf{j}}}\times{{\bf{B}}}, is given by

𝐚=14πK.{{\bf{a}}}=\frac{1}{4\pi}{{\bf{\nabla}}}K. (4)

The force balance (1) is described by gradient terms of scalar functions.

The magnetic function Ψ\Psi is obtained by solving the Ampére–Biot–Savart law with the source term (Equation (3)) after the functional forms S(Ψ)S(\Psi) and K(Ψ)K(\Psi) are specified. For simplicity, we assume that KK is a linear function of Ψ\Psi, K=K0ΨK=K_{0}\Psi, and S=0S=0 in Equation (3). The poloidal magnetic field is purely dipolar and is given by Ψ=g(r)sin2θ\Psi=g(r)\sin^{2}\theta. The radial function gg is solved with appropriate boundary conditions: in a vacuum at the surface RR, and g=0g=0 at the core-crust interface rcr_{c}. The latter refers to the magnetic field expelled from the core. For the case where the field penetrates into the core, gg smoothly connects to the interior solution at rcr_{c}. We normalize the radial function by the dipole field strength B0B_{0} at the surface, g(R)=B0R2/2g(R)=B_{0}R^{2}/2.

The magnetic geometry discussed above is a simple initial model to examine the magnetic field evolution. However, purely poloidal magnetic field configurations are unstable according to an energy principle (Tayler, 1973; Markey & Tayler, 1973; Wright, 1973) and numerical MHD simulation (Braithwaite & Nordlund, 2006; Braithwaite, 2009; Lander & Jones, 2011a, b; Mitchell et al., 2015). Dynamical simulations revealed that the final state after a few Alfvén-wave crossing times is a twisted-torus configuration, in which the poloidal and toroidal components of comparable field strengths are tangled. Moreover, recent three-dimensional simulation shows very asymmetric equilibrium (Becerra et al., 2022). These studies are concerned with the configuration in an entire star. The information relevant to our study corresponds to the magnetic field in a thin outer layer; therefore, the present understanding is quite incomplete. For example, the ratio of toroidal to poloidal components decreases near the surface, because the toroidal field should vanish outside the exterior. However, the ratio in the crust located near the surface is uncertain, almost zero or in the order of unity, although both components are comparable in magnitude inside the star. Initially, a simple magnetic field configuration is used in this paper, but it is necessary to improve the configuration.

We discuss the non-barotropic equilibrium of the magnetic field. From Equation (1), the acceleration owing to the Lorentz force satisfies

×𝐚=×(1ρP)0.{{\bf{\nabla}}}\times{{\bf{a}}}={{\bf{\nabla}}}\times\left(\frac{1}{\rho}{{\bf{\nabla}}}P\right)\neq 0. (5)

The acceleration 𝐚{{\bf{a}}} is no longer described by the gradient of a scalar, but may be generalized as

𝐚=14πK+αβ,{{\bf{a}}}=\frac{1}{4\pi}{{\bf{\nabla}}}K+\alpha{{\bf{\nabla}}}\beta, (6)

where α\alpha and β\beta are functions of rr and θ\theta, and ×𝐚=α×β0{{\bf{\nabla}}}\times{{\bf{a}}}={{\bf{\nabla}}}\alpha\times{{\bf{\nabla}}}\beta\neq 0 is assumed. Owing to almost arbitrary functions α\alpha and β\beta, the constraint on the electric current, and hence to the magnetic-field configuration, is relaxed in the non-barotropic case. Non-barotropicity has been studied in magnetic deformation (Mastrano et al., 2011, 2013, 2015). Barotropic (Haskell et al., 2008; Kojima et al., 2021b) and non-barotropic models are significantly different. (Mastrano et al., 2011, 2013, 2015) up to approximately one order of magnitude in the resulting ellipticity. The effect is important; however, the treatment remains unclear. Therefore, we introduce the models of ×𝐚{{\bf{\nabla}}}\times{{\bf{a}}} to study non-barotropicity in Section 4.

2.2 Magnetic-field Evolution

Our consideration is limited to the inner crust of a neutron star, where the mass density ranges from ρc=1.4×1014\rho_{c}=1.4\times 10^{14} g cm-3 at the core–crust boundary rcr_{c} to the neutron-drip density ρ1=4×1011\rho_{1}=4\times 10^{11} g cm-3 at R=12R=12 km. We ignore the outer crust compared to “ocean” and treat the exterior region of r>Rr>R as the vacuum. The crust thickness ΔrRrc\Delta r\equiv R-r_{c} is assumed to be Δr/R=0.05\Delta r/R=0.05, Δr=0.6\Delta r=0.6 km.

The Lorentz force 𝐣×𝐁{{\bf{j}}}\times{{\bf{B}}} due to the magnetic-field evolution is not fixed in a secular timescale. The evolution in the crust is governed by the induction equation

t𝐁=×[1ene𝐣×𝐁+cσ𝐣],\displaystyle\frac{\partial}{\partial t}{{\bf{B}}}=-{{\bf{\nabla}}}\times\left[\frac{1}{en_{\rm e}}{{\bf{j}}}\times{{\bf{B}}}+\frac{c}{\sigma}{{\bf{j}}}\right], (7)

where nen_{\rm e} is the electron-number density, and σ\sigma is the electric conductivity. The first term in Equation (7) represents the Hall drift, and the Hall timescale τH\tau_{\rm H} is estimated as follows:

τH=4πenec(Δr)2cB0=7.9×105yr(B01014G)1,\displaystyle\tau_{\rm H}=\frac{4\pi en_{\rm ec}(\Delta r)^{2}}{cB_{0}}=7.9\times 10^{5}~{}{\rm yr}\left(\frac{B_{0}}{10^{14}{\rm G}}\right)^{-1}, (8)

where the electron-number density necn_{\rm ec} at the core–crust boundary, crust thickness Δr\Delta r, and dipole field strength B0B_{0} at the surface are used. This timescale is shorter than that of the Ohmic decay, which is the second term in Equation (7) in the strong-magnetic-field regime.

The Hall-Ohmic evolution was numerically simulated in the Hall timescale Pons & Geppert (2007); Kojima & Kisaka (2012); Viganò et al. (2013); Gourgouliatos et al. (2013); Gourgouliatos & Cumming (2014); Viganò et al. (2021) for axially symmetric models. Recently, the calculation has been extended to 3D models Wood & Hollerbach (2015); Viganò et al. (2019); De Grandis et al. (2020); Gourgouliatos et al. (2020); Igoshev et al. (2021), revealing some of the effects ignored in the 2D models. Here, our calculation is limited to the early phase of the evolution in a simpler axial-symmetric model. We consider only the Hall drift term in Equation (7) and rewrite the equation as follows:

t𝐁=×[1ene𝐣×𝐁]=χ×𝐚χ×𝐚,\frac{\partial}{\partial t}{{\bf{B}}}=-{{\bf{\nabla}}}\times\left[\frac{1}{en_{\rm e}}{{\bf{j}}}\times{{\bf{B}}}\right]=-{{\bf{\nabla}}}\chi\times{{\bf{a}}}-\chi{{\bf{\nabla}}}\times{{\bf{a}}}, (9)

where

χcρene=4πρc(Δr)2τHB0χ^.\chi\equiv\frac{c\rho}{en_{e}}=\frac{4\pi\rho_{c}(\Delta r)^{2}}{\tau_{\rm H}B_{0}}{\hat{\chi}}. (10)

In Equation (10), χ^{\hat{\chi}} is a dimensionless function, which represents an inverse of the electron fraction. The electron-number density is obtained from the proton fraction of the equilibrium nucleus in ”cold catalyzed matter,” i.e., it is determined in the ground state at T=0T=0 K. The data given by Douchin & Haensel (2001) is approximately fitted by a smooth function:

χ^=ρ^1/20.32+0.66ρ^,{\hat{\chi}}=\frac{{\hat{\rho}}^{1/2}}{0.32+0.66{\hat{\rho}}}, (11)

where ρ^ρ/ρc.{\hat{\rho}}\equiv\rho/\rho_{c}. The spatial-density profile of a neutron star in rcrRr_{c}\leq r\leq R is approximated as follows (Lander & Gourgouliatos, 2019):

ρ^=ρρc=[1(1(ρ1ρc)1/2)(rrcΔr)]2.{\hat{\rho}}=\frac{\rho}{\rho_{c}}=\left[1-\left(1-\left(\frac{\rho_{1}}{\rho_{c}}\right)^{1/2}\right)\left(\frac{r-r_{c}}{\Delta r}\right)\right]^{2}. (12)

The radial derivative dχ^/dr=(dχ^/dρ)(dρ/dr)d{\hat{\chi}}/dr=(d{\hat{\chi}}/d\rho)(d\rho/dr) sharply changes, owing to an abrupt decrease in the density near the surface; however, χ^{\hat{\chi}} is a smoothly varying function of 𝒪(1)\mathcal{O}(1). This functional behavior, which originates from the stellar-density profile that is inherent in neutron stars, is crucial in our numerical calculation. Different fitting formulae are discussed for different equation of state in Pearson et al. (2018); however, the difference in χ^{\hat{\chi}} is not significant in our analysis.

We consider the early phase of the evolution in the axially symmetric equilibrium model, in which aϕ=0a_{\phi}=0. From Equation (9) at t=0t=0, we obtain that Bϕ/t0{\partial B_{\phi}}/{\partial t}\neq 0, but Br/t=Bθ/t=0{\partial B_{r}}/{\partial t}={\partial B_{\theta}}/{\partial t}=0. The azimuthal component BϕB_{\phi} changes linearly with time tt, whereas the poloidal components change with t2t^{2}. We limit our consideration to the lowest order of tt only and ignore the change in the poloidal magnetic field. The early phase of the toroidal magnetic field may be approximated as

δBϕ=δSϖ(tτH),\delta B_{\phi}=\frac{\delta S}{\varpi}\left(\frac{t}{\tau_{\rm H}}\right), (13)

where δS\delta S is a function of rr and θ\theta. Because it is associated with δBϕ\delta B_{\phi}, the poloidal current changes; thus, the Lorentz force δ𝐟=c1(δ𝐣×𝐁+𝐣×δ𝐁)\delta{{\bf{f}}}=c^{-1}(\delta{{\bf{j}}}\times{{\bf{B}}}+{{\bf{j}}}\times\delta{{\bf{B}}}) also changes. We observe that the non-zero component is δfϕ\delta f_{\phi} because 𝐣p=Bϕ=0{{\bf{j}}}_{p}=B_{\phi}=0 at t=0t=0, and we explicitly write

δfϕ=14πϖ2[δS×Ψ]ϕ(tτH).\delta f_{\phi}=\frac{1}{4\pi\varpi^{2}}[{{\bf{\nabla}}}\delta S\times{{\bf{\nabla}}}\Psi]_{\phi}\left(\frac{t}{\tau_{\rm H}}\right). (14)

2.3 Quasi-stationary Elastic Response

We assume that the solid crust elastically acts against the force δfϕ\delta f_{\phi}. The change is so slow that the elastic evolution is quasi-stationary. The acceleration 2ξi/t2\partial^{2}\xi_{i}/\partial t^{2} of the elastic displacement vector ξi\xi_{i} is dismissed; thus, the elastic force is balanced with the change in the Lorentz force, i.e., when the gravity and pressure in Equation (1) is assumed to be fixed. The elastic force is expressed by the trace-free strain tensor σij\sigma_{ij} and a shear modulus μ\mu; therefore, the force balance is

i(2μσϕi)+δfϕ=0,\displaystyle{\nabla}_{i}\left(2\mu\sigma^{i}_{\phi}\right)+\delta f_{\phi}=0, (15)
σij=12(iξj+jξi),\displaystyle\sigma_{ij}=\frac{1}{2}({\nabla}_{i}\xi_{j}+{\nabla}_{j}\xi_{i}), (16)

where incompressible motion kξk=0{\nabla}_{k}\xi^{k}=0 is assumed. Alternatively, the equivalent form is

i[2μσiϕ+14πBiδBϕ]=0.\nabla_{i}\left[2\mu\sigma^{i\phi}+\frac{1}{4\pi}B^{i}\delta B^{\phi}\right]=0. (17)

The relevant component induced by δfϕ\delta f_{\phi} is the azimuthal displacement ξϕ\xi_{\phi} only, and the shear tensors that are determined by it are

σrϕ=r2(ξϕr),r,σθϕ=sinθ2r(ξϕsinθ),θ.\displaystyle\sigma_{r\phi}=\frac{r}{2}\left(\frac{\xi_{\phi}}{r}\right)_{,r},~{}~{}~{}~{}\sigma_{\theta\phi}=\frac{\sin\theta}{2r}\left(\frac{\xi_{\phi}}{\sin\theta}\right)_{,\theta}. (18)

The shear modulus μ\mu increases with density, and it may be approximated as a linear function of ρ\rho, which is overall fitted to the results of a detailed calculation reported in a previous study (see Figure 43 in Chamel & Haensel, 2008).

μ=μcρρc,{\mu}=\frac{\mu_{c}\rho}{\rho_{c}}, (19)

where μc=1030ergcm3\mu_{c}=10^{30}~{}{\rm erg~{}cm}^{-3} at the core–crust interface. The shear speed vsv_{s} in Equation (19) is constant through the crust:

vs=(μρ)1/2=8.5×107cms1.v_{s}=\left(\frac{\mu}{\rho}\right)^{1/2}=8.5\times 10^{7}{\rm{cm~{}s}^{-1}}. (20)

To solve Equation (15), we use an expansion method with the Legendre polynomials Pl(cosθ)P_{l}(\cos\theta) and radial functions kl(r)k_{l}(r), al(r)a_{l}(r) as follows:

ξϕ=l1rklPl,θ(tτH),\displaystyle\xi_{\phi}=-\sum_{l\geq 1}rk_{l}P_{l,\theta}\left(\frac{t}{\tau_{\rm H}}\right), (21)
δfϕ=l1r3alPl,θ(tτH).\displaystyle\delta f_{\phi}=-\sum_{l\geq 1}r^{-3}a_{l}P_{l,\theta}\left(\frac{t}{\tau_{\rm H}}\right). (22)

The displacement ξϕ\xi_{\phi} is decoupled with respect to the index ll, owing to spherical symmetry, i.e., μ(r)\mu(r). Equation (15) is reduced to a set of ordinary differential equations for klk_{l} (Kojima et al., 2022):

(μr4kl)(l1)(l+2)μr2kl=al,(\mu r^{4}k_{l}^{\prime})^{\prime}-(l-1)(l+2)\mu r^{2}k_{l}=-a_{l}, (23)

where a prime denotes a derivative with respect to rr. The boundary conditions for the radial functions klk_{l} are given by the force-balance across the surfaces at rcr_{c} and RR. That is, the shear-stress tensor σrϕ\sigma_{r\phi} vanishes because other stresses for the fluid and magnetic field are assumed to be continuous. Explicitly, we have kl=0k_{l}^{\prime}=0 both at rcr_{c} and RR. Note that a mode with l=1l=1 is special in Equation (23), and k1k_{1} is simply obtained by integrating a1a_{1} with respect to rr.

3 Elastic Deformation in Barotropic Model

The magnetic-field evolution in Equation (9) is driven by two terms, which are separately examined. We first consider the barotropic case, in which ×𝐚=0{{\bf{\nabla}}}\times{{\bf{a}}}=0. The evolution is driven by the first term in Equation (9), i.e., the distribution of the electron fraction. The linear growth term δS\delta S in Equation (13) is obtained using Equation (4) as

δS=2K0ρc(Δr)23B0χ^gsinθP2,θ.\delta S=\frac{2K_{0}\rho_{c}({\Delta}r)^{2}}{3B_{0}}{\hat{\chi}}^{\prime}g\sin\theta P_{2,\theta}. (24)

In the case where the poloidal magnetic field (Ψ=gsin2θ\Psi=g\sin^{2}\theta) is confined in the crust, the constant K0K_{0} is numerically obtained as K0=6.0×B0/(ρc(Δr)2)K_{0}=6.0\times B_{0}/(\rho_{c}({\Delta}r)^{2}). Alternatively, we express K0=8.6×101va2/(B0R2)K_{0}=8.6\times 10^{1}v_{a}^{2}/(B_{0}R^{2}), where vav_{a} is the Alfvén speed in terms of B0B_{0} and ρ1\rho_{1}:

va=B04πρ1=4.5×107(B01014G)cms1.v_{a}=\frac{B_{0}}{\sqrt{4\pi\rho_{1}}}=4.5\times 10^{7}\left(\frac{B_{0}}{10^{14}{\rm G}}\right){\rm{cm~{}s}^{-1}}. (25)

The Lorentz force δfϕ\delta f_{\phi} in Equation (14) is calculated using Equation (24). For later convenience, we consider a general form δS=yl(r)sinθPl,θ\delta S=y_{l}(r)\sin\theta P_{l,\theta}. By using an identity for the Legendre polynomials, we reduce Equation (14) to

δfϕ=\displaystyle\delta f_{\phi}= 14πr3[(lgyl+2gyl)l+12l+1Pl1,θ+((l+1)gyl+2gyl)l2l+1Pl+1,θ](tτH).\displaystyle\frac{1}{4\pi r^{3}}\bigg{[}(lg^{\prime}y_{l}+2gy_{l}^{\prime})\frac{l+1}{2l+1}P_{l-1,\theta}+(-(l+1)g^{\prime}y_{l}+2gy_{l}^{\prime})\frac{l}{2l+1}P_{l+1,\theta}\bigg{]}\left(\frac{t}{\tau_{\rm H}}\right). (26)

Thus, the radial functions al1a_{l-1} and al+1a_{l+1} in Equation (22) are induced by yly_{l}. By numerically solving Equation (23) for klk_{l}’s, we obtain ξϕ\xi_{\phi} in Equation (21) and the shear stress tensors, σrϕ\sigma_{r\phi} and σθϕ\sigma_{\theta\phi}, in Equation (18). For Equation (24), the results are expressed using a combination of k1k_{1} and k3k_{3}.

3.1 Results

The shear stress increases homologously with time, i.e., the spatial profile of the shear force is unchanged, but the magnitude increases with time. The numerical calculation provides the maximum shear stress σmax\sigma_{\rm{max}} with respect to (r,θ)(r,\theta) in the crust as follows:

σmax=6.5×101(vavs)2(tτH).\sigma_{\rm{max}}=6.5\times 10^{1}\left(\frac{v_{a}}{v_{s}}\right)^{2}\left(\frac{t}{\tau_{\rm{H}}}\right). (27)

The maximum is determined using a ratio of the shear speed vsv_{s} in Equation (20) to the Alfvén speed vav_{a} in Equation (25). Elastic equilibrium is possible, only when the shear strain satisfies a particular criterion. We adopt the following (the Mises criterion) to determine the elastic limit:

12σijσij(σmax)2(σc)2,\frac{1}{2}\sigma_{ij}\sigma^{ij}\leq(\sigma_{\rm{max}})^{2}\leq(\sigma_{c})^{2}, (28)

where σc\sigma_{c} is a number σc102101\sigma_{c}\approx 10^{-2}-10^{-1} (Horowitz & Kadau, 2009; Caplan et al., 2018; Baiko & Chugunov, 2018).

Thus, the period of the elastic response is limited by the constraint σc\sigma_{c}:

tt\displaystyle t\leq t_{*} 1.5×103(σc0.1)(vsva)2τH,\displaystyle\equiv 1.5\times 10^{-3}\left(\frac{\sigma_{c}}{0.1}\right)\left(\frac{v_{s}}{v_{a}}\right)^{2}\tau_{{\rm H}}, (29)
=4.3×103(σc0.1)(B01014G)3yr.\displaystyle=4.3\times 10^{3}\left(\frac{\sigma_{c}}{0.1}\right)\left(\frac{B_{0}}{10^{14}{\rm G}}\right)^{-3}{\rm{yr}}. (30)

The breakup time tt_{*} becomes short, i.e., a few years for magnetars with B0=1015B_{0}=10^{15}G. This is in agreement with the recurrent time of the activity in magnetars. However, the timescale exceeds 1 Myr for most neutron stars with B<1013GB<10^{13}{\rm G}. Moreover, other evolution effects are important, and present results are not applicable.

3.2 Energy

The stored elastic energy is obtained by numerically integrating over the entire crust:

ΔEela\displaystyle\Delta E_{\rm ela} =2πrcRr2𝑑r0πsinθdθμσijσij\displaystyle=2\pi\int_{r_{c}}^{R}r^{2}dr\int_{0}^{\pi}\sin\theta d\theta~{}\mu\sigma_{ij}\sigma^{ij}
=5.8×107μcR3(σc0.1)2(tt)2,\displaystyle=5.8\times 10^{-7}~{}\mu_{c}R^{3}\left(\frac{\sigma_{c}}{0.1}\right)^{2}\left(\frac{t}{t_{*}}\right)^{2}, (31)

where we have used R3R^{3} instead of R2ΔrR^{2}{\Delta{r}} to normalize the crustal volume because Δr/R=0.05{\Delta{r}}/R=0.05 is fixed. The elastic energy EelaE_{\rm ela} increases up to 1041\approx 10^{41} erg at the breakup time tt_{*}.

The magnetic energy ΔEmag\Delta E_{\rm mag} is also obtained by

ΔEmag\displaystyle\Delta E_{\rm mag} =2πrcRr2𝑑r0πsinθdθ(δBϕ)28π\displaystyle=2\pi\int_{r_{c}}^{R}r^{2}dr\int_{0}^{\pi}\sin\theta d\theta~{}\frac{(\delta B_{\phi})^{2}}{8\pi}
=2.0×104B02R3(σc0.1)2(vsva)4(tt)2.\displaystyle=2.0\times 10^{-4}~{}B_{0}^{2}R^{3}\left(\frac{\sigma_{c}}{0.1}\right)^{2}\left(\frac{v_{s}}{v_{a}}\right)^{4}\left(\frac{t}{t_{*}}\right)^{2}. (32)

Here, the shear speed appears in ΔEmag\Delta E_{\rm mag} because the breakup time tt_{*} in Equation (29) is used instead of τH\tau_{\rm H}. The magnetic energy ΔEmag\Delta E_{\rm mag} at tt_{*} is ΔEmag2×1043(B0/1014G)2\Delta E_{\rm mag}\approx 2\times 10^{43}(B_{0}/10^{14}{\rm G})^{-2} erg. However, it is considerably smaller than the poloidal magnetic energy Emag,pE_{\rm mag,p}, which is numerically calculated as Emag,p=3.8B02R34×1046(B0/1014G)2E_{\rm mag,p}=3.8B_{0}^{2}R^{3}\approx 4\times 10^{46}(B_{0}/10^{14}{\rm G})^{2} erg. Note that total magnetic energy is conserved by the Hall evolution. Therefore, the same amount of polar magnetic energy decreases. However, we ignored the change in the poloidal component and its energy, which are proportional to t2t^{2}.

The ratio of Equations (31) and (32) is

dΔEela/dtdΔEmag/dt=ΔEelaΔEmag=8.1×102(vavs)2=2.3×102(B01014G)2,\frac{d{\Delta}E_{\rm ela}/dt}{d{\Delta}E_{\rm mag}/dt}=\frac{\Delta E_{\rm ela}}{\Delta E_{\rm mag}}=8.1\times 10^{-2}\left(\frac{v_{a}}{v_{s}}\right)^{2}=2.3\times 10^{-2}\left(\frac{B_{0}}{10^{14}{\rm G}}\right)^{2}, (33)

where μc\mu_{c} and B02B_{0}^{2} are eliminated using vsv_{s} and vav_{a}. The ratio is proportional to B02B_{0}^{2}; thus, ΔEmag\Delta E_{\rm mag} decreases in more strongly magnetized neutron stars. From the viewpoint of the energy flow from the poloidal component, the breakup energy ΔEela1041\Delta E_{\rm ela}\approx 10^{41} at the terminal is fixed, but the ΔEmag\Delta E_{\rm mag} stored in the middle depends on the Hall drift speed. The elastic energy is efficiently accumulated through toroidal magnetic energy with an increase in B0B_{0}; ΔEela>ΔEmag\Delta E_{\rm ela}>\Delta E_{\rm mag} for B0>6.7×1014B_{0}>6.7\times 10^{14} G.

3.3 Spatial Distribution

Figure 1 shows spatial-energy densities εela(r)\varepsilon_{\rm ela}(r) and εmag(r)\varepsilon_{\rm mag}(r), with regard to ΔEela\Delta E_{\rm ela} and ΔEmag\Delta E_{\rm mag} in the crust, respectively. They are normalized as εela𝑑r=εela𝑑r=1\int\varepsilon_{\rm ela}dr=\int\varepsilon_{\rm ela}dr=1. Evidently, both energies are highly concentrated near the surface rRr\approx R. This property comes from the radial derivative of χ^{\hat{\chi}} in Equation (10). dχ^/dr=(dχ^/dρ)(dρ/dr)d{\hat{\chi}}/dr=(d{\hat{\chi}}/d\rho)(d\rho/dr) is steep there even though χ^{\hat{\chi}} is 𝒪(1)\mathcal{O}(1). The large value comes from |dρ/dr||d\rho/dr|, i.e., a sharp decrease in density near the stellar surface, and it results in a smaller evolution timescale τH\ll\tau_{\rm{H}} in Equation (29).

Refer to caption
Figure 1: Energy distribution in crust as a function of the radius. Normalized energy density ε(r)\varepsilon(r) is plotted for magnetic energy, which is denoted by a dotted curve, and elastic energy, denoted by a solid curve.
Refer to caption
Figure 2: Contour map of stress tensor in crust. The left panel shows 2μσθϕ2\mu\sigma_{\theta\phi}, the middle 2μσrϕ2\mu\sigma_{r\phi}, and the right the magnitude σ\sigma. They are normalized according to the maximum. Negative values are plotted using a dotted curve in the left panel. The magnitude σ\sigma in the right panel is shown in the outer small region near the surface, R0.1ΔrrRR-0.1{\Delta}r\leq r\leq R.

Shear stresses σθϕ\sigma_{\theta\phi} and σrϕ\sigma_{r\phi} are induced by the axial displacement ξϕ\xi_{\phi}. A numerical calculation shows that the component σrϕ\sigma_{r\phi} is considerably larger than σθϕ\sigma_{\theta\phi}; (σrϕ)max200(σθϕ)max(\sigma_{r\phi})_{\rm max}\sim 200(\sigma_{\theta\phi})_{\rm max}. Figure 2 shows their spatial distribution using a contour map of 2μσθϕ2\mu\sigma_{\theta\phi} and 2μσrϕ2\mu\sigma_{r\phi} in the rθr-\theta plain. The angular dependence of σθϕ\sigma_{\theta\phi} is σθϕsin2θcosθ\sigma_{\theta\phi}\propto\sin^{2}\theta\cos\theta, and it is anti-symmetric with respect to the equator (θ=π/2\theta=\pi/2). Moreover, σrϕ\sigma_{r\phi} is the sum of P1,θP_{1,\theta} and P3,θP_{3,\theta}, and it is symmetric with respect to θ=π/2\theta=\pi/2. The magnitude σ=(σijσij/2)1/2\sigma=(\sigma_{ij}\sigma^{ij}/2)^{1/2} is also shown in the right panel, and σ\sigma is sharp near the surface, as expected from the sharp energy-density-distribution in Figure 1

We discuss the modification of the poloidal magnetic field at the core–crust boundary. Thus far, the magnetic field is expelled there. When the field is penetrated to the core, the function gg near the boundary and the constant K0K_{0} in Equation (24) are changed. The former is unimportant because the function χ^{\hat{\chi}}^{\prime} is sharp near the surface, and this fact determines the result, as shown in Figures 1 and 2. The constant K0K_{0} for the penetrated field is 4.1×1024.1\times 10^{-2} times smaller than that for the expelled one. Consequently, the profile is almost unchanged, but the break time tt_{*} increases by a factor of 24 with the same dipole field strength.

4 Elastic Deformation in Non-Barotropic Model

We consider the evolution driven by the second term, ×𝐚0{{\bf{\nabla}}}\times{{\bf{a}}}\neq 0 in Equation (9), which originates from the non-barotropic material distribution. However, ×𝐚{{\bf{\nabla}}}\times{{\bf{a}}} and the corresponding magnetic field cannot be easily estimated, unless the non-barotropic property is specified. A large freedom of choice hinders our analysis. Therefore, we simply model the term ×𝐚{{\bf{\nabla}}}\times{{\bf{a}}} and understand the non-barotropic effect in its magnitude and property. For this purpose, we assume aϕ=0a_{\phi}=0 and

(×𝐚)ϕ=N2(Δr)2Fn(r)P2,θ,({{\bf{\nabla}}}\times{{\bf{a}}})_{\phi}=\frac{N^{2}}{(\Delta r)^{2}}F_{n}(r)P_{2,\theta}, (34)

where NN is a constant that characterizes the non-barotropic strength, and it has the dimension of velocity. Additionally, FnF_{n} is a non-dimensional radial function. We consider a small deviation from the barotropic case, for which the second term in Equation (6) is smaller than the first term. Therefore, the magnetic field is approximated using the barotropic case, i.e., the poloidal magnetic function Ψ\Psi and S=0S=0. This treatment constrains the normalization NN in Equation (34) with respect to the magnitude. By the dimensional argument, we have |N|R/tff|N|\ll R/t_{\rm{ff}} 109cms1\sim 10^{9}~{}{\rm{cm~{}s}}^{-1}, where tfft_{\rm{ff}} is a free-fall timescale. Moreover, |N|<va|N|<v_{a} and |N|<vs|N|<v_{s} 108cms1\sim 10^{8}~{}{\rm{cm~{}s}}^{-1} are also likely because the crust is in magneto-elastic equilibrium.

The angular dependence of Equation (34) is chosen for δS\delta S to be the same as in Equation (24). The radial function FnF_{n} has a maximum that is normalized as unity, and it vanishes at inner and outer boundaries; the function is modeled as follows:

Fn=25627(rrcΔr)n(RrΔr)4n,F_{n}=\frac{256}{27}\left(\frac{r-r_{c}}{\Delta r}\right)^{n}\left(\frac{R-r}{\Delta r}\right)^{4-n}, (35)

where n=1n=1 or 3. The model with n=1n=1 is referred to as the ”in” model because the maximum is located at r=rc+Δr/4r=r_{c}+\Delta r/4, whereas that with n=3n=3 is referred to as the ”out” model because the maximum is located at r=RΔr/4r=R-\Delta r/4.

4.1 Results of a Simple Model

We neglect the first term in Equation (9), and consider the magnetic-field evolution driven by the term ×𝐚{{\bf{\nabla}}}\times{{\bf{a}}} (Equation (34)) only. Similar to the calculations in Section 3, a linearly growing shear–stress is obtained, owing to ξϕ\xi_{\phi}. The period of the elastic response is limited by

ttn1×(σc0.1)(vsN)2τH,t\leq t_{*}\equiv n_{1}\times\left(\frac{\sigma_{c}}{0.1}\right)\left(\frac{v_{s}}{N}\right)^{2}\tau_{{\rm H}}, (36)

where n1n_{1} is a number of the order of 10210^{-2}, depending on the models, as listed in Table 1. Owing to our simple modeling, the comparison between the barotropic and non-barotropic models is uncomplicated; the Alfvén speed vav_{a} in Equation (29) is formally substituted by NN in Equation (36).

The elastic energy ΔEela\Delta E_{\rm ela} and toroidal magnetic energy ΔEmag\Delta E_{\rm mag} stored inside the crust are also summarized as follows:

ΔEela=n2×μcR3(σc0.1)2(tt)2,\displaystyle\Delta E_{\rm ela}=n_{2}\times\mu_{c}R^{3}\left(\frac{\sigma_{c}}{0.1}\right)^{2}\left(\frac{t}{t_{*}}\right)^{2}, (37)
ΔEmag=n3×B02R3(σc0.1)2(vsN)4(tt)2,\displaystyle\Delta E_{\rm mag}=n_{3}\times B_{0}^{2}R^{3}\left(\frac{\sigma_{c}}{0.1}\right)^{2}\left(\frac{v_{s}}{N}\right)^{4}\left(\frac{t}{t_{*}}\right)^{2}, (38)

where n26×104n_{2}\approx 6\times 10^{-4}, and n3106n_{3}\approx 10^{-6}, as listed in Table 1. The elastic energy ΔEela\Delta E_{\rm ela} does not depend on NN; however, the timescale (36) does. The amount of elastic energy is unrelated to the detailed process, which affects the accumulation speed in the crust. At the breakup time, the elastic energy is ΔEela=2×10442×1045\Delta E_{\rm ela}=2\times 10^{44}-2\times 10^{45} erg. This energy is more than three orders of magnitude larger than the energy (31) considered in the previous section. The difference is made clear when considering the energy–density distribution.

model n1(t)n_{1}~{}~{}(t_{*}) n2(ΔEela)n_{2}~{}~{}(\Delta E_{\rm ela}) n3(ΔEmag)n_{3}~{}~{}(\Delta E_{\rm mag})
in 8.0×1038.0\times 10^{-3} 2.1×1032.1\times 10^{-3} 7.2×1077.2\times 10^{-7}
out 1.2×1021.2\times 10^{-2} 2.4×1042.4\times 10^{-4} 1.0×1061.0\times 10^{-6}
ave 3.6×1033.6\times 10^{-3} 2.9×1042.9\times 10^{-4} 2.9×1072.9\times 10^{-7}
min 1.9×1031.9\times 10^{-3} 0.49×1040.49\times 10^{-4} 0.73×1070.73\times 10^{-7}
max 5.5×1035.5\times 10^{-3} 7.0×1047.0\times 10^{-4} 6.4×1076.4\times 10^{-7}
Table 1: Numerical values in eqs. (36), (37), and (38).

Figure 3 shows the energy–density distribution in the crust. The difference in the toroidal magnetic energy clearly originates in the model choice; the energy density spreads more towards the interior for the ”in” model, whereas it spreads more towards the exterior for the ”out” model. Most of the elastic energy is localized near the inner core–crust boundary; however, the distribution in the ”out” model is shifted outwardly with the second peak (rc+0.8Δr\sim r_{c}+0.8{\Delta}r) produced by the input model. The integrated elastic energy in the ”out” model becomes one order of magnitude smaller than that in the ”in” model.

The amount of elastic energy at the breakup clearly depends on the spatial distribution of the energy density because the shear modulus μ\mu is a strongly decreasing function toward the surface. The elastic limit of the entire crust is typically determined using a condition to the shear σij\sigma_{ij} near the surface. The total elastic energy μσijσijd3x\sim\int\mu\sigma_{ij}\sigma^{ij}d^{3}x thereby decreases as σijσij\sigma_{ij}\sigma^{ij} is localized towards the exterior. The breakup elastic energy ΔEela1041\Delta E_{\rm ela}\sim 10^{41} erg at tt_{*} in the previous section is an extreme case because the energy density is concentrated near the surface.

Refer to caption
Figure 3: Spatial distribution of magnetic energy (dotted curve) and elastic energy (solid curve) in the crust. Normalized energy densities ε(r)\varepsilon(r) are plotted for two models.
Refer to caption
Figure 4: Crust contour map of magnitude of stress tensor normalized using the maximum. The left panel for the ”in” model shows σσθϕ\sigma\approx\sigma_{\theta\phi}, whereas the right panel for the ”out” model shows σσrϕ\sigma\approx\sigma_{r\phi}.

Figure 4 shows the magnitude of shear stress σ\sigma inside the crust. The contours of σ\sigma in the two models are different. We identified that the dominant component in the ”in” model (left panel) is σθϕ\sigma_{\theta\phi}, which has an angular dependence that is described by σθϕsin2θcosθ\sigma_{\theta\phi}\propto\sin^{2}\theta\cos\theta. The maximum of σ\sigma is given along a line cosθ=±1/3(θ55,125)\cos\theta=\pm 1/\sqrt{3}~{}(\theta\sim 55^{\circ},125^{\circ}). The component σrϕ\sigma_{r\phi} is dominant in the ”out” model (right panel). Sharp peaks are localized near the surface, similar to the right panel in Figure 2; however, the localization is not as pronounced as in Figure 2. The magnitude σ\sigma, which is important to determine the critical limit, is large near the surface.

4.2 Results of a Model Including Higher Multipoles

In a more realistic situation, the solenoidal acceleration may fluctuate spatially. We consider the sum of multi-pole components Pl,θP_{l,\theta}:

(×𝐚)ϕ=N2(Δr)2l=2lmax(l2)4/3λlFn(r)Pl,θ,({{\bf{\nabla}}}\times{{\bf{a}}})_{\phi}=\frac{N^{2}}{(\Delta r)^{2}}\sum_{l=2}^{l_{\rm max}}\left(\frac{l}{2}\right)^{-4/3}\lambda_{l}F_{n}(r)P_{l,\theta}, (39)

where λlFn\lambda_{l}F_{n} is a radial function that is randomly selected from ±F1\pm F_{1} or ±F3\pm F_{3} depending on ll. As discussed for Equation (26), the radial function klk_{l} in the azimuthal displacement ξϕ\xi_{\phi} is solved for the source term that originates from λl1Fn\lambda_{l-1}F_{n} and λl+1Fn\lambda_{l+1}F_{n}; thus, the amplitude |kl||k_{l}| fluctuates according to the randomness. We fix the overall constant NN. Equation (39) is reduced to Equation (34) when lmax=2l_{\rm max}=2. Moreover, higher ll-modes, up to lmax=30l_{\rm max}=30 with the power-law weight, are included. The power-law index is considerably steep; therefore, the dominant component is still described by l=2l=2.

We calculated 20 models by randomly mixing λlFn\lambda_{l}F_{n}. The numerical results are summarized in the same forms as eqs. (36)–(38), and the numerical values ni(i=1,2,3)n_{i}~{}(i=1,2,3) in the breakup time and energies are listed in Table 1 according to the average, minimum, and maximum for the 20 models. These numerical values are of the same order as those by a single mode with l=2l=2 because we include higher ll-modes as the correction. Interestingly, the breakup time tt_{*} generally becomes shorter than that for a single mode with l=2l=2 because the higher modes l>3l>3 are cooperative.

Refer to caption
Figure 5: Color contour map of crust for the magnitude of the stress tensor σ\sigma normalized using the maximum; two models are compared.

Figure 5 demonstrates the spatial distribution of the shear stress tensor. Two models are shown using contours of the magnitude of σ\sigma. In the left panel, the sub-critical regions are on a constant θ\theta line with a sharp peak near the surface. In the other model (right panel), a peak is observed at θ=0\theta=0 near the surface. The angular position of the peak is different between the two models. As shown in Figure 4, the spatial pattern along a constant θ\theta comes from the component σθϕ\sigma_{\theta\phi}, whereas the sharp peak near the surface is due to σrϕ\sigma_{r\phi}. The mixing of the two types of radial functions, F1F_{1} and F3F_{3}, and angular functions PlP_{l} with random phases and weights only complicates the spatial distribution of σ\sigma. A sharp peak is likely to be located near the surface. The outer part of the crust is always fragile; thus, the breakup time becomes shorter.

5 Summary and Discussion

We have considered the evolution of elastic deformation over a secular timescale (>1>1 yr) starting from zero displacement. The initial state is related to the dynamic force balance that is determined within a second. When a neutron star cools below the melting temperature T109T\sim 10^{9} K, its crust is crystallized. Meanwhile, the pressure, gravity, and Lorentz force are balanced without the elastic force. In another situation, the elastic energy settles to the ground state, and zero displacement occurs after the energy is completely released at crustal fracture. Therefore, the initial condition is simple and natural.

When the MHD equilibrium is axisymmetric, the azimuthal component of the magnetic field increases linearly according to the Hall evolution. Consequently, the elastic deformation in the azimuthal direction is induced to cancel the change in the Lorentz force, and the shear strain gradually increases. We estimate the range of the elastic response. Beyond the critical limit, the crust responds plastically or fractures. Our calculations provide the breakup time and shear distribution at the threshold.

For the barotropic case, the breakup time until fracture is proportional to the cube of the magnetic-field strength. The time becomes a few years for a magnetar with a surface dipole of B01015B_{0}\sim 10^{15}G, when the field is located outside the core. However, it exceeds 1 Myr for most radio pulsars (B0<1013B_{0}<10^{13}G), and the process is irrelevant to them. In addition to the field strength, the timescale is typically shortened by a factor of 10310^{-3} smaller than the Hall timescale because the elastic displacement is highly concentrated near the surface. The driven mechanism is related to the instability associated with electron-density gradients (Wood et al., 2014). The distribution in any realistic model of neutron-star crusts is considerably sharp; therefore, the evolution is general. Another type of Hall-drift instability occurs in the presence of a non-uniform magnetic field (Rheinhardt & Geppert, 2002), which is not considered here, and its energy would be smaller, owing to the size of the irregularity. In our calculation, we do not follow the instability; instead, we estimate the energy transferred to the elastic deformation.

The elastic energy at the critical limit in the model driven by the electron-number-density gradient is 1041\sim 10^{41} erg. The amount of energy is of the same order as that of short bursts in magnetars. The breaktime of 10\sim 10 years also reconciles with the observed recurrent-time of the bursts. However, the energy 1041\sim 10^{41} erg is smaller than that of giant flares 10441046\sim 10^{44}-10^{46} erg (Turolla et al., 2015; Kaspi & Beloborodov, 2017; Esposito et al., 2021, for a review). The total elastic energy derived in Section 3 is based on the electron-number density in cold catalyzed matter, i.e., the ground state at T=0T=0 K. If the assumption was relaxed, the non-barotropic effects might increase the total elastic energy, as considered in Section 4.

When the pressure distribution is no longer expressed solely by the density ρ\rho, the magnetic evolution is affected by the solenoidal acceleration 𝐚=×(ρ1P)0{{\bf{a}}}={{\bf{\nabla}}}\times(\rho^{-1}{{\bf{\nabla}}}P)\neq 0. We have also considered this effect by creating the model in terms of a spatial function and an overall strength parameter, which are assumed to be constant in time in our non-barotropic model. Using the simplified model, we calculated the breakup time of the crustal failure and the energies stored in the crust. The results were comparable to those for the barotropic case. The strength parameter significantly affects the breakup time; the larger the magnitude, the shorter the breakup time. However, the amount of elastic energy at the breakup does not depend on the strength parameter, but only on the spatial function. The maximum elastic energy considerably increases up to 1045\sim 10^{45} erg. However, the model is still primitive, and thermal evolution should also be incorporated to investigate a more realistic situation.

The maximum energy has been explored thus far; however, a natural question arises. What is the fraction of energy that is released at the crustal fracture when the strain exceeds the threshold? This question is important, but at present unclear, owing to our lack of understanding of the fracture dynamics. Therefore, we present the following discussion. As depicted in Figure 5, in the realistic mixture model, a peak of shear strain σ\sigma is probably located near the surface, where the crust is fragile. Therefore, the fracture should not include the whole crust, but only the shallow crust. For this case, the released energy is not the whole elastic energy 1045\sim 10^{45} erg, but the energy stored in the restricted region, i.e., a small fraction of the total, probably 1041\sim 10^{41} erg.

The elastic deformation driven by the Hall evolution is simulated for the first time. The critical structure at the breakup time is crucial for subsequent evolution, irrespective of plastic evolution or fracturing. The transition may appear as a burst on a magnetar. The magnetic-field rearrangement due to a mimic burst was incorporated in the Hall evolution (Pons & Perna, 2011; Viganò et al., 2013; Dehman et al., 2020), without solving elastic deformation. These studies estimated the critical state based on the magnetic stress ij{\mathcal{M}}_{ij}. In numerical simulations, ij{\mathcal{M}}_{ij} changed, and the critical state was assumed when a condition among ij{\mathcal{M}}_{ij} reached the threshold value. Similar approximations for the elastic limit, which were derived solely from ij{\mathcal{M}}_{ij}, were used in a previous study (Lander et al., 2015; Lander & Gourgouliatos, 2019; Suvorov & Kokkotas, 2019). Mathematically, the shear stress σij\sigma_{ij} cannot be derived from ij{\mathcal{M}}_{ij} without solving the appropriate differential equation j(2μσij+ij)=0\nabla_{j}(2\mu\sigma_{i}^{j}+{\mathcal{M}}_{i}^{j})=0 (see Equation (17)). Therefore, previous results with the criterion ij{\mathcal{M}}_{ij} are questionable.

Our calculation shows that the period of elastic evolution is typically 10310^{-3} times the Hall timescale; however, this value depends on the strength and geometry of the magnetic field. The timescale is shorter than the Ohmic timescale for B1013B\geq 10^{13} G. The magnetic-field evolution beyond the period may be described by including the viscous bulk flow when the crust responds plastically. The effect of the plastic flow on the Hall–Ohmic evolution was considered by assuming a plastic flow everywhere in the crust (Kojima & Suzuki, 2020) or by using an approximated criterion (Lander & Gourgouliatos, 2019; Gourgouliatos & Lander, 2021; Gourgouliatos, 2022). The effect may be regarded as additional energy lost to the Ohmic decay. However, the post-failure evolution significantly depends on the modeling in the numerical simulation (Gourgouliatos & Lander, 2021; Gourgouliatos, 2022). That is, the region of plastic flow is either local or global when the failure criterion is satisfied. Therefore, the manner of incorporation of crust failure in the numerical simulation must be explored.

Finally, further investigation is required before the elastic deformation toward the crust failure can be considered a viable model. The effect of magnetic field configuration should be considered because there are many degrees of freedom concerning it. Moreover, the outer boundary, i.e., inner–outer crust boundary or exterior magnetosphere is crucial as the crust becomes more fragile with increasing radius. Meanwhile, the electric conductivity decreases and the Ohmic loss becomes more important. By considering coupling to an exterior magnetosphere, twisting of the magnetosphere as well as crustal motion will be calculated in a secular timescale to match astrophysical observations, e.g., to describe the pre-stage of outbursts, like SGR 1830-0645 (Younes et al., 2022).

Acknowledgements

This work was supported by JSPS KAKENHI Grant Number JP17H06361,JP19K03850.

References

  • Abbott et al. (2021a) Abbott, R., Abe, H., Acernese, F., et al. 2021a, ApJ, 921, 80, doi: 10.3847/1538-4357/ac17ea
  • Abbott et al. (2021b) —. 2021b, ApJ, 913, L27, doi: 10.3847/2041-8213/abffcd
  • Abbott et al. (2022) —. 2022, arXiv e-prints, arXiv:2204.04523. https://arxiv.org/abs/2204.04523
  • Alpar et al. (1984) Alpar, M. A., Pines, D., Anderson, P. W., & Shaham, J. 1984, ApJ, 276, 325, doi: 10.1086/161616
  • Anderson & Itoh (1975) Anderson, P. W., & Itoh, N. 1975, Nature, 256, 25, doi: 10.1038/256025a0
  • Baiko & Chugunov (2018) Baiko, D. A., & Chugunov, A. I. 2018, MNRAS, 480, 5511, doi: 10.1093/mnras/sty2259
  • Basu et al. (2022) Basu, A., Shaw, B., Antonopoulou, D., et al. 2022, MNRAS, 510, 4049, doi: 10.1093/mnras/stab3336
  • Becerra et al. (2022) Becerra, L., Reisenegger, A., Valdivia, J. A., & Gusakov, M. E. 2022, MNRAS, 511, 732, doi: 10.1093/mnras/stac102
  • Bochenek et al. (2020) Bochenek, C. D., Ravi, V., Belov, K. V., et al. 2020, Nature, 587, 59, doi: 10.1038/s41586-020-2872-x
  • Braithwaite (2009) Braithwaite, J. 2009, MNRAS, 397, 763, doi: 10.1111/j.1365-2966.2008.14034.x
  • Braithwaite & Nordlund (2006) Braithwaite, J., & Nordlund, Å. 2006, A&A, 450, 1077, doi: 10.1051/0004-6361:20041980
  • Caplan et al. (2018) Caplan, M. E., Schneider, A. S., & Horowitz, C. J. 2018, Phys. Rev. Lett., 121, 132701, doi: 10.1103/PhysRevLett.121.132701
  • Chamel & Haensel (2008) Chamel, N., & Haensel, P. 2008, Living Reviews in Relativity, 11, 10, doi: 10.12942/lrr-2008-10
  • CHIME/FRB Collaboration et al. (2020) CHIME/FRB Collaboration, Andersen, B. C., Bandura, K. M., et al. 2020, Nature, 587, 54, doi: 10.1038/s41586-020-2863-y
  • De Grandis et al. (2020) De Grandis, D., Turolla, R., Wood, T. S., et al. 2020, ApJ, 903, 40, doi: 10.3847/1538-4357/abb6f9
  • Dehman et al. (2020) Dehman, C., Viganò, D., Rea, N., et al. 2020, ApJ, 902, L32, doi: 10.3847/2041-8213/abbda9
  • Dib et al. (2008) Dib, R., Kaspi, V. M., & Gavriil, F. P. 2008, ApJ, 673, 1044, doi: 10.1086/524653
  • Douchin & Haensel (2001) Douchin, F., & Haensel, P. 2001, A&A, 380, 151, doi: 10.1051/0004-6361:20011402
  • Espinoza et al. (2011) Espinoza, C. M., Lyne, A. G., Stappers, B. W., & Kramer, M. 2011, MNRAS, 414, 1679, doi: 10.1111/j.1365-2966.2011.18503.x
  • Esposito et al. (2021) Esposito, P., Rea, N., & Israel, G. L. 2021, in Astrophysics and Space Science Library, Vol. 461, Astrophysics and Space Science Library, ed. T. M. Belloni, M. Méndez, & C. Zhang, 97–142, doi: 10.1007/978-3-662-62110-3_3
  • Franco et al. (2000) Franco, L. M., Link, B., & Epstein, R. I. 2000, ApJ, 543, 987, doi: 10.1086/317121
  • Giliberti et al. (2020) Giliberti, E., Cambiotti, G., Antonelli, M., & Pizzochero, P. M. 2020, MNRAS, 491, 1064, doi: 10.1093/mnras/stz3099
  • Gittins et al. (2021) Gittins, F., Andersson, N., & Jones, D. I. 2021, MNRAS, 500, 5570, doi: 10.1093/mnras/staa3635
  • Gourgouliatos (2022) Gourgouliatos, K. N. 2022, arXiv e-prints, arXiv:2202.06662. https://arxiv.org/abs/2202.06662
  • Gourgouliatos & Cumming (2014) Gourgouliatos, K. N., & Cumming, A. 2014, MNRAS, 438, 1618, doi: 10.1093/mnras/stt2300
  • Gourgouliatos et al. (2013) Gourgouliatos, K. N., Cumming, A., Reisenegger, A., et al. 2013, MNRAS, 434, 2480, doi: 10.1093/mnras/stt1195
  • Gourgouliatos et al. (2020) Gourgouliatos, K. N., Hollerbach, R., & Igoshev, A. P. 2020, MNRAS, 495, 1692, doi: 10.1093/mnras/staa1295
  • Gourgouliatos & Lander (2021) Gourgouliatos, K. N., & Lander, S. K. 2021, MNRAS, 506, 3578, doi: 10.1093/mnras/stab1869
  • Haskell et al. (2006) Haskell, B., Jones, D. I., & Andersson, N. 2006, MNRAS, 373, 1423, doi: 10.1111/j.1365-2966.2006.10998.x
  • Haskell et al. (2008) Haskell, B., Samuelsson, L., Glampedakis, K., & Andersson, N. 2008, MNRAS, 385, 531, doi: 10.1111/j.1365-2966.2008.12861.x
  • Horowitz & Kadau (2009) Horowitz, C. J., & Kadau, K. 2009, Phys. Rev. Lett., 102, 191102, doi: 10.1103/PhysRevLett.102.191102
  • Igoshev et al. (2021) Igoshev, A. P., Gourgouliatos, K. N., Hollerbach, R., & Wood, T. S. 2021, ApJ, 909, 101, doi: 10.3847/1538-4357/abde3e
  • Kaspi & Beloborodov (2017) Kaspi, V. M., & Beloborodov, A. M. 2017, ARA&A, 55, 261, doi: 10.1146/annurev-astro-081915-023329
  • Kojima & Kisaka (2012) Kojima, Y., & Kisaka, S. 2012, MNRAS, 421, 2722, doi: 10.1111/j.1365-2966.2012.20509.x
  • Kojima et al. (2021a) Kojima, Y., Kisaka, S., & Fujisawa, K. 2021a, MNRAS, 506, 3936, doi: 10.1093/mnras/stab1848
  • Kojima et al. (2021b) —. 2021b, MNRAS, 502, 2097, doi: 10.1093/mnras/staa3489
  • Kojima et al. (2022) —. 2022, MNRAS, 511, 480, doi: 10.1093/mnras/stac036
  • Kojima & Suzuki (2020) Kojima, Y., & Suzuki, K. 2020, MNRAS, 494, 3790, doi: 10.1093/mnras/staa1045
  • Lander et al. (2015) Lander, S. K., Andersson, N., Antonopoulou, D., & Watts, A. L. 2015, MNRAS, 449, 2047, doi: 10.1093/mnras/stv432
  • Lander & Gourgouliatos (2019) Lander, S. K., & Gourgouliatos, K. N. 2019, MNRAS, 486, 4130, doi: 10.1093/mnras/stz1042
  • Lander & Jones (2011a) Lander, S. K., & Jones, D. I. 2011a, MNRAS, 412, 1394, doi: 10.1111/j.1365-2966.2010.17998.x
  • Lander & Jones (2011b) —. 2011b, MNRAS, 412, 1730, doi: 10.1111/j.1365-2966.2010.18009.x
  • Li et al. (2022) Li, X., Ge, M., Lin, L., et al. 2022, ApJ, 931, 56, doi: 10.3847/1538-4357/ac6587
  • Markey & Tayler (1973) Markey, P., & Tayler, R. J. 1973, MNRAS, 163, 77, doi: 10.1093/mnras/163.1.77
  • Mastrano et al. (2013) Mastrano, A., Lasky, P. D., & Melatos, A. 2013, MNRAS, 434, 1658, doi: 10.1093/mnras/stt1131
  • Mastrano et al. (2011) Mastrano, A., Melatos, A., Reisenegger, A., & Akgün, T. 2011, MNRAS, 417, 2288, doi: 10.1111/j.1365-2966.2011.19410.x
  • Mastrano et al. (2015) Mastrano, A., Suvorov, A. G., & Melatos, A. 2015, MNRAS, 447, 3475, doi: 10.1093/mnras/stu2671
  • Mereghetti et al. (2020) Mereghetti, S., Savchenko, V., Ferrigno, C., et al. 2020, ApJ, 898, L29, doi: 10.3847/2041-8213/aba2cf
  • Mitchell et al. (2015) Mitchell, J. P., Braithwaite, J., Reisenegger, A., et al. 2015, MNRAS, 447, 1213, doi: 10.1093/mnras/stu2514
  • Payne & Melatos (2004) Payne, D. J. B., & Melatos, A. 2004, MNRAS, 351, 569, doi: 10.1111/j.1365-2966.2004.07798.x
  • Pearson et al. (2018) Pearson, J. M., Chamel, N., Potekhin, A. Y., et al. 2018, MNRAS, 481, 2994, doi: 10.1093/mnras/sty2413
  • Pons & Geppert (2007) Pons, J. A., & Geppert, U. 2007, A&A, 470, 303, doi: 10.1051/0004-6361:20077456
  • Pons & Perna (2011) Pons, J. A., & Perna, R. 2011, ApJ, 741, 123, doi: 10.1088/0004-637X/741/2/123
  • Rencoret et al. (2021) Rencoret, J. A., Aguilera-Gómez, C., & Reisenegger, A. 2021, A&A, 654, A47, doi: 10.1051/0004-6361/202141499
  • Rheinhardt & Geppert (2002) Rheinhardt, M., & Geppert, U. 2002, Phys. Rev. Lett., 88, 101103, doi: 10.1103/PhysRevLett.88.101103
  • Suvorov & Kokkotas (2019) Suvorov, A. G., & Kokkotas, K. D. 2019, MNRAS, 488, 5887, doi: 10.1093/mnras/stz2052
  • Tayler (1973) Tayler, R. J. 1973, MNRAS, 161, 365, doi: 10.1093/mnras/161.4.365
  • Turolla et al. (2015) Turolla, R., Zane, S., & Watts, A. L. 2015, Reports on Progress in Physics, 78, 116901, doi: 10.1088/0034-4885/78/11/116901
  • Ushomirsky et al. (2000) Ushomirsky, G., Cutler, C., & Bildsten, L. 2000, MNRAS, 319, 902, doi: 10.1046/j.1365-8711.2000.03938.x
  • Viganò et al. (2021) Viganò, D., Garcia-Garcia, A., Pons, J. A., Dehman, C., & Graber, V. 2021, Computer Physics Communications, 265, 108001, doi: 10.1016/j.cpc.2021.108001
  • Viganò et al. (2013) Viganò, D., Rea, N., Pons, J. A., et al. 2013, MNRAS, 434, 123, doi: 10.1093/mnras/stt1008
  • Viganò et al. (2019) Viganò, D., Martínez-Gómez, D., Pons, J. A., et al. 2019, Computer Physics Communications, 237, 168, doi: 10.1016/j.cpc.2018.11.022
  • Wadiasingh & Chirenti (2020) Wadiasingh, Z., & Chirenti, C. 2020, ApJ, 903, L38, doi: 10.3847/2041-8213/abc562
  • Wood & Hollerbach (2015) Wood, T. S., & Hollerbach, R. 2015, Phys. Rev. Lett., 114, 191101, doi: 10.1103/PhysRevLett.114.191101
  • Wood et al. (2014) Wood, T. S., Hollerbach, R., & Lyutikov, M. 2014, Physics of Plasmas, 21, 052110, doi: 10.1063/1.4879810
  • Wright (1973) Wright, G. A. E. 1973, MNRAS, 162, 339, doi: 10.1093/mnras/162.4.339
  • Younes et al. (2022) Younes, G., Lander, S. K., Baring, M. G., et al. 2022, ApJ, 924, L27, doi: 10.3847/2041-8213/ac4700