Vortex Avalanches and Collective Motion in Neutron Stars111Released on March, 1st, 2024

I-Kang Liu ({CJK}UTF8bkai劉翼綱) School of Mathematics, Statistics and Physics, Newcastle University,
Newcastle upon Tyne, NE1 7RU, United Kingdom
Andrew W. Baggaley School of Mathematics, Statistics and Physics, Newcastle University,
Newcastle upon Tyne, NE1 7RU, United Kingdom
Carlo F. Barenghi School of Mathematics, Statistics and Physics, Newcastle University,
Newcastle upon Tyne, NE1 7RU, United Kingdom
Toby S. Wood School of Mathematics, Statistics and Physics, Newcastle University,
Newcastle upon Tyne, NE1 7RU, United Kingdom
Abstract

We simulate the dynamics of about 600 quantum vortices in a spinning-down cylindrical container using a Gross–Pitaevskii model. For the first time, we find convincing spatial-temporal evidence of avalanching behaviour resulting from vortex depinning and collective motion. During a typical avalanche, about 10 to 20 vortices exit the container in a short period, producing a glitch in the superfluid angular momentum and a localised void in the vorticity. After the glitch, vortices continue to depin and circulate around the vorticity void in a similar manner to that seen in previous point-vortex simulations. We present evidence of collective vortex motion throughout this avalanche process. We also show that the effective Magnus force can be used to predict when and where avalanches will occur. Lastly, we comment on the challenge of extrapolating these results to conditions in real neutron stars, which contain many orders of magnitude more vortices.

Neutron Stars (1108) – Pulsars (1306) – Hydrodynamical simulations (767)

1 Introduction

Rotational glitches are sudden, spasmodic changes in the rotation rate of a neutron star, which result in changes to the (otherwise regular) pulses of radiation detected from pulsars. Glitches are believed to arise from the spontaneous transfer of angular momentum to the star’s solid crust from neutron superfluid in its interior, which generally rotates more rapidly. The inner part of the crust comprises a lattice of nuclei immersed in a sea of superfluid neutrons and degenerate electrons (Baym et al., 1971). The superfluid component cannot rotate in the manner of a classical fluid, and instead contains a multitude (typically 1018superscript101810^{18}10 start_POSTSUPERSCRIPT 18 end_POSTSUPERSCRIPT1020superscript102010^{20}10 start_POSTSUPERSCRIPT 20 end_POSTSUPERSCRIPT) of superfluid vortices, each carrying a quantum of circulation 2π/m2𝜋Planck-constant-over-2-pi𝑚2\pi\hbar/m2 italic_π roman_ℏ / italic_m, where Planck-constant-over-2-pi\hbarroman_ℏ is Planck’s reduced constant and m=2mn𝑚2subscript𝑚𝑛m=2m_{n}italic_m = 2 italic_m start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is the mass of a neutron Cooper pair. These vortices pin to the crustal nuclei (e.g. Avogadro et al., 2008; Chamel & Haensel, 2008) preventing the superfluid from spinning down at the same rate as the crust, which is constantly losing angular momentum through electromagnetic braking. It is thought that, once the rotational lag between the curst and the superfluid exceeds some threshold, many vortices spontaneously depin and a fraction of the superfluid’s angular momentum is suddenly transferred to the crust, producing a glitch. The leading paradigm of this process is the avalanche model (Melatos et al., 2008), which is motivated by the close analogy with magnetic flux avalanches in Type-II superconductors (Anderson & Itoh, 1975; Field et al., 1995). This model predicts self-organised criticality (Bak et al., 1987) with a power-law distribution of glitch sizes and an exponential distribution of waiting times between glitches, consistent with the majority of pulsar observations (Melatos et al., 2008; Fuentes et al., 2019). The observed range of glitch sizes implies that the number of vortices involved ranges from 107superscript10710^{7}10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT to 1015superscript101510^{15}10 start_POSTSUPERSCRIPT 15 end_POSTSUPERSCRIPT, and that the amount of superfluid involved in the glitch is comparable to that residing in the crust (although part of the core may also be involved, e.g. Andersson et al., 2012; Gügercinoğlu & Alpar, 2014; Newton et al., 2015; Haskell et al., 2018).

The simplest superfluid model that self-consistently describes vortex pinning is the Gross–Pitaevskii (GP) model, and this has previously been used to study rotational glitches in the neutron star crust (Warszawski & Melatos, 2011; Warszawski et al., 2012; Melatos et al., 2015; Lönnborn et al., 2019). Given the huge disparity between the length scales of individual vortex cores (100similar-toabsent100\sim 100\,∼ 100fm) and the star itself (10similar-toabsent10\sim 10∼ 10 km), such models can only resolve a much smaller system, and previous studies have been limited to 200less-than-or-similar-toabsent200\lesssim 200≲ 200 vortices. Although these models have reproduced some features of pulsar glitches, it is unclear whether the results can be scaled up to the true parameter regime. Indeed, Warszawski & Melatos (2011) found that glitch size decreased as they increased the number of vortices, and so their results focussed on simulations with <100absent100<100< 100 vortices.

An alternative approach is the point-vortex (or vortex-filament) model (Howitt et al., 2020; Cheunchitra et al., 2024), which tracks only the position of vortices rather than the density and velocity of the superfluid. This is computationally more efficient and such models have produced glitches in systems of up to 5,000 vortices. However, the interactions between vortices, and their interactions with pinning sites, cannot be treated fully in such a model, and must instead be parameterised. Moreover, the dynamics depend qualitatively on the degree of dissipation in the model, which must be incorporated in an ad hoc manner, and greatly exceeds the dissipation expected in a neutron star.

We study the dynamics of up to 600600600600 vortices in a two-dimensional GP model, in the presence of a spinning down crust. We show that, if the spin down is sufficiently slow, then rotational glitches do occur and are associated with vortex avalanches. We also demonstrate that the results are essentially independent of the degree of dissipation, provided that this is made sufficiently small.

2 Gross–Pitaevskii Modeling

2.1 Numerical Model

In the GP framework, the superfluid is characterised by a mean-field wavefunction, ψ(𝐫,t)𝜓𝐫𝑡\psi(\mathbf{r},t)italic_ψ ( bold_r , italic_t ), which satisfies the (damped) GP equation:

iψt=(1iγ)(H^GPμ)ψ.iPlanck-constant-over-2-pi𝜓𝑡1i𝛾subscript^𝐻GP𝜇𝜓\mathrm{i}\hbar\frac{\partial\psi}{\partial t}=(1-\mathrm{i}\gamma)\left(\hat{% H}_{\mathrm{GP}}-\mu\right)\psi.roman_i roman_ℏ divide start_ARG ∂ italic_ψ end_ARG start_ARG ∂ italic_t end_ARG = ( 1 - roman_i italic_γ ) ( over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT roman_GP end_POSTSUBSCRIPT - italic_μ ) italic_ψ . (1)

Here H^GPsubscript^𝐻GP\hat{H}_{\mathrm{GP}}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT roman_GP end_POSTSUBSCRIPT is the Gross–Pitaevskii Hamiltonian and μ𝜇\muitalic_μ is the effective chemical potential, i.e. the energetic cost of increasing the superfluid density. We emphasize that μ𝜇\muitalic_μ should not be equated to the chemical potential of the neutrons themselves; as we explain below, its physical role is to set the superfluid density, and hence the neutron coherence length. In the rotating frame, the Hamiltonian can be expressed as

H^GP=22m2+V(𝐫,t)+g|ψ(𝐫,t)|2𝛀(t)𝐋^,subscript^𝐻GPsuperscriptPlanck-constant-over-2-pi22𝑚superscript2𝑉𝐫𝑡𝑔superscript𝜓𝐫𝑡2𝛀𝑡^𝐋\hat{H}_{\mathrm{GP}}=-\frac{\hbar^{2}}{2m}\nabla^{2}+V(\mathbf{r},t)+g\left|% \psi(\mathbf{r},t)\right|^{2}-\boldsymbol{\Omega}(t)\cdot\hat{\mathbf{L}},over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT roman_GP end_POSTSUBSCRIPT = - divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG ∇ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_V ( bold_r , italic_t ) + italic_g | italic_ψ ( bold_r , italic_t ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - bold_Ω ( italic_t ) ⋅ over^ start_ARG bold_L end_ARG , (2)

where m=2mn𝑚2subscript𝑚nm=2m_{\mathrm{n}}italic_m = 2 italic_m start_POSTSUBSCRIPT roman_n end_POSTSUBSCRIPT is the mass of a superconducting particle (i.e. a neutron Cooper pair), V(𝐫,t)𝑉𝐫𝑡V(\mathbf{r},t)italic_V ( bold_r , italic_t ) is an imposed potential, g𝑔gitalic_g measures the self-repulsion of the superfluid, 𝛀(t)𝛀𝑡\boldsymbol{\Omega}(t)bold_Ω ( italic_t ) is the angular velocity of the reference frame, and 𝐋^^𝐋\hat{\mathbf{L}}over^ start_ARG bold_L end_ARG is the angular momentum operator. We will adopt Cartesian coordinates (x,y,z)𝑥𝑦𝑧(x,y,z)( italic_x , italic_y , italic_z ) such that 𝛀=(0,0,Ω(t))𝛀00Ω𝑡\boldsymbol{\Omega}=\bigl{(}0,0,\Omega(t)\bigr{)}bold_Ω = ( 0 , 0 , roman_Ω ( italic_t ) ), and so 𝛀𝐋^=iΩ/ϕ𝛀^𝐋iPlanck-constant-over-2-piΩitalic-ϕ\boldsymbol{\Omega}\cdot\hat{\mathbf{L}}=-\mathrm{i}\hbar\,\Omega\,\partial/\partial\phibold_Ω ⋅ over^ start_ARG bold_L end_ARG = - roman_i roman_ℏ roman_Ω ∂ / ∂ italic_ϕ, where ϕitalic-ϕ\phiitalic_ϕ is the angular coordinate around the z𝑧zitalic_z-axis. We solve Eq. (1) numerically in the xy𝑥𝑦xyitalic_x italic_y-plane, assuming that ψ𝜓\psiitalic_ψ has no z𝑧zitalic_z-dependence; for details see Sec. 2.2.

The number density and superfluid velocity of the superfluid (in the rotating frame) are defined as

n=|ψ|2and𝐯=mIm{lnψ}𝛀×𝐫.formulae-sequence𝑛superscript𝜓2and𝐯Planck-constant-over-2-pi𝑚bold-∇Im𝜓𝛀𝐫n=|\psi|^{2}\qquad\mbox{and}\qquad\mathbf{v}=\frac{\hbar}{m}\boldsymbol{\nabla% }\text{Im}\{\ln\psi\}-\boldsymbol{\Omega}\times\mathbf{r}.italic_n = | italic_ψ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and bold_v = divide start_ARG roman_ℏ end_ARG start_ARG italic_m end_ARG bold_∇ Im { roman_ln italic_ψ } - bold_Ω × bold_r . (3)

In the absence of rotation or any imposed potential, the ground state for this system would have zero velocity and a uniform density nbμ/gsubscript𝑛𝑏𝜇𝑔n_{b}\equiv\mu/gitalic_n start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ≡ italic_μ / italic_g, thus we take μ𝜇\muitalic_μ and nbsubscript𝑛𝑏n_{b}italic_n start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT as characteristic units for energy and density, respectively. The characteristic length scale is the coherence length, ξ/mμ𝜉Planck-constant-over-2-pi𝑚𝜇\xi\equiv\hbar/\sqrt{m\mu}italic_ξ ≡ roman_ℏ / square-root start_ARG italic_m italic_μ end_ARG, which sets the vortex core size, and the characteristic time scale is τ/μ𝜏Planck-constant-over-2-pi𝜇\tau\equiv\hbar/\muitalic_τ ≡ roman_ℏ / italic_μ.

As mentioned earlier, the radius of a neutron star exceeds ξ𝜉\xiitalic_ξ by many orders of magnitude, and so it is impracticable to model the entire crust (and all of its vortices) in a single numerical GP model. Previous GP models have therefore resorted to modelling, in effect, a tiny neutron star, containing of order 100 vortices only. In this work, we take a slightly different approach, and aim to model a small but representative piece of the crust. In contrast to previous models (Warszawski & Melatos, 2011; Melatos et al., 2015; Drummond & Melatos, 2017; Lönnborn et al., 2019) the dynamics within our numerical domain are therefore assumed to play a negligible role in the overall balance of angular momentum within the star. For this reason, we simply impose the rotation rate of the crust, Ω(t)Ω𝑡\Omega(t)roman_Ω ( italic_t ), without taking account of any transfer of angular momentum between the superfluid in our domain and the crust. In common with previous models, we assume that the crust is spun down by electromagnetic radiation at a constant rate, and so

Ω(t)=Ω0Ω˙tΩ𝑡subscriptΩ0˙Ω𝑡\Omega(t)=\Omega_{0}-\dot{\Omega}troman_Ω ( italic_t ) = roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - over˙ start_ARG roman_Ω end_ARG italic_t (4)

where the initial value Ω0subscriptΩ0\Omega_{0}roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and spin-down rate Ω˙˙Ω\dot{\Omega}over˙ start_ARG roman_Ω end_ARG are (positive) constants.

The imposed potential V(𝐫,t)𝑉𝐫𝑡V(\mathbf{r},t)italic_V ( bold_r , italic_t ) is a sum of three contributions:

V(𝐫,t)=Vcon(𝐫)+Vpin(𝐫)+Vcen(𝐫,t)𝑉𝐫𝑡subscript𝑉con𝐫subscript𝑉pin𝐫subscript𝑉cen𝐫𝑡V(\mathbf{r},t)=V_{\mathrm{con}}(\mathbf{r})+V_{\mathrm{pin}}(\mathbf{r})+V_{% \mathrm{cen}}(\mathbf{r},t)italic_V ( bold_r , italic_t ) = italic_V start_POSTSUBSCRIPT roman_con end_POSTSUBSCRIPT ( bold_r ) + italic_V start_POSTSUBSCRIPT roman_pin end_POSTSUBSCRIPT ( bold_r ) + italic_V start_POSTSUBSCRIPT roman_cen end_POSTSUBSCRIPT ( bold_r , italic_t ) (5)

where Vconsubscript𝑉conV_{\mathrm{con}}italic_V start_POSTSUBSCRIPT roman_con end_POSTSUBSCRIPT is a confining potential, Vpinsubscript𝑉pinV_{\mathrm{pin}}italic_V start_POSTSUBSCRIPT roman_pin end_POSTSUBSCRIPT is a pinning potential, and Vcensubscript𝑉cenV_{\mathrm{cen}}italic_V start_POSTSUBSCRIPT roman_cen end_POSTSUBSCRIPT is a centripetal potential that balances the centrifugal force. The confining potential is taken to be a hard-wall potential at cylindrical radius Rconsubscript𝑅conR_{\mathrm{con}}italic_R start_POSTSUBSCRIPT roman_con end_POSTSUBSCRIPT, i.e. Vcon(𝐫)=V0,conΘ(rRcon)subscript𝑉con𝐫subscript𝑉0conΘ𝑟subscript𝑅conV_{\mathrm{con}}(\mathbf{r})=V_{0,\mathrm{con}}\Theta(r-R_{\mathrm{con}})italic_V start_POSTSUBSCRIPT roman_con end_POSTSUBSCRIPT ( bold_r ) = italic_V start_POSTSUBSCRIPT 0 , roman_con end_POSTSUBSCRIPT roman_Θ ( italic_r - italic_R start_POSTSUBSCRIPT roman_con end_POSTSUBSCRIPT ), where Θ(x)Θ𝑥\Theta(x)roman_Θ ( italic_x ) is the Heaviside step function and r=(x2+y2)1/2𝑟superscriptsuperscript𝑥2superscript𝑦212r=(x^{2}+y^{2})^{1/2}italic_r = ( italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT. The pinning potential consists of Npinsubscript𝑁pinN_{\mathrm{pin}}italic_N start_POSTSUBSCRIPT roman_pin end_POSTSUBSCRIPT identical circular Gaussian barriers of height V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and width w𝑤witalic_w:

Vpin(𝐫)=j=1NpinV0e[(xxj)2+(yyj)2]/w2.subscript𝑉pin𝐫superscriptsubscript𝑗1subscript𝑁pinsubscript𝑉0superscriptedelimited-[]superscript𝑥subscript𝑥𝑗2superscript𝑦subscript𝑦𝑗2superscript𝑤2V_{\mathrm{pin}}(\mathbf{r})=\sum_{j=1}^{N_{\mathrm{pin}}}V_{0}\mathrm{e}^{-[(% x-x_{j})^{2}+(y-y_{j})^{2}]/w^{2}}.italic_V start_POSTSUBSCRIPT roman_pin end_POSTSUBSCRIPT ( bold_r ) = ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_pin end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_e start_POSTSUPERSCRIPT - [ ( italic_x - italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ( italic_y - italic_y start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] / italic_w start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT . (6)

In all of the results presented later, the pinning site locations, (xj,yj)subscript𝑥𝑗subscript𝑦𝑗(x_{j},y_{j})( italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ), are arranged in a square lattice with separation dpsubscript𝑑𝑝d_{p}italic_d start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT. However, taking randomly distributed pinning sites does not qualitatively affect the dynamics. Finally, the centripetal potential is taken to be Vcen(𝐫,t)=12mΩ2(t)r2subscript𝑉cen𝐫𝑡12𝑚superscriptΩ2𝑡superscript𝑟2V_{\mathrm{cen}}(\mathbf{r},t)=\tfrac{1}{2}m\Omega^{2}(t)r^{2}italic_V start_POSTSUBSCRIPT roman_cen end_POSTSUBSCRIPT ( bold_r , italic_t ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_m roman_Ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_t ) italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, so that the superfluid density remains roughly uniform, with |ψ|2nbsimilar-to-or-equalssuperscript𝜓2subscript𝑛𝑏|\psi|^{2}\simeq n_{b}| italic_ψ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≃ italic_n start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT, despite the overall rotation. This approach is consistent with regarding the system as only a small piece of the star’s crust, across which the superfluid density should not vary significantly. This is in contrast to previous GP models, which have generally used a harmonic confining potential (Warszawski & Melatos, 2011; Warszawski et al., 2012; Melatos et al., 2015; Lönnborn et al., 2019), resulting in a density that varies significantly across the domain. Such density variations also affect the vortex core size and dynamical time scales, which we prefer to avoid.

The dimensionless coefficient γ𝛾\gammaitalic_γ in Eq. (1) introduces dissipation into the superfluid dynamics. It acts to reduce the total free energy, defined as

d2𝐫ψ(H^GPμ)ψ,superscriptd2𝐫superscript𝜓subscript^𝐻GP𝜇𝜓\int\mathrm{d}^{2}\mathbf{r}\,\psi^{\ast}\left(\hat{H}_{\mathrm{GP}}-\mu\right% )\psi,∫ roman_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_r italic_ψ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT roman_GP end_POSTSUBSCRIPT - italic_μ ) italic_ψ , (7)

decreases monotonically with time. In the context of ultracold quantum gases, γ𝛾\gammaitalic_γ arises from the interaction between superfluid and normal components at nonzero temperatures, and typically has a value 103less-than-or-similar-toabsentsuperscript103\lesssim 10^{-3}≲ 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT (e.g. Bradley et al., 2008; Blakie et al., 2008; Rooney et al., 2012). In the inner crust of a neutron star, the true dissipation mechanisms are more complicated, and γ𝛾\gammaitalic_γ serves only as a crude parameterisation. Nonetheless, previous studies have typically adopted values γ0.02𝛾0.02\gamma\geqslant 0.02italic_γ ⩾ 0.02 (Warszawski & Melatos, 2011; Warszawski et al., 2012; Melatos et al., 2015; Lönnborn et al., 2019). We aim to determine how small γ𝛾\gammaitalic_γ must be such that it plays little (if any) role in the qualitative behaviour of the system.

2.2 Numerical Procedure

We numerically solve the two-dimensional damped GP equation in dimensionless form by scaling the energy, density, length and time by μ𝜇\muitalic_μ, nbsubscript𝑛𝑏n_{b}italic_n start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT, ξ𝜉\xiitalic_ξ and τ𝜏\tauitalic_τ, respectively. We use 4th-order Runge–Kutta method with a timestep of 103τsuperscript103𝜏10^{-3}\tau10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT italic_τ; the wavefunction is discretized on a uniform Cartesian grid and spatial derivatives are evaluated spectrally. The domain is a square of size (512ξ)2superscript512𝜉2(512\xi)^{2}( 512 italic_ξ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, with 10242superscript102421024^{2}1024 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT grid points. The confining potential has height V0,con=1000μsubscript𝑉0con1000𝜇V_{0,\mathrm{con}}=1000\muitalic_V start_POSTSUBSCRIPT 0 , roman_con end_POSTSUBSCRIPT = 1000 italic_μ and radius Rcon=230ξsubscript𝑅con230𝜉R_{\mathrm{con}}=230\xiitalic_R start_POSTSUBSCRIPT roman_con end_POSTSUBSCRIPT = 230 italic_ξ. The pinning potential height and width are V0=2μsubscript𝑉02𝜇V_{0}=2\muitalic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 2 italic_μ and w=ξ𝑤𝜉w=\xiitalic_w = italic_ξ, and the separation between pinning sites is dp=10ξsubscript𝑑𝑝10𝜉d_{p}=10\xiitalic_d start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = 10 italic_ξ, creating similar-to\sim1,600 pinning sites.

Simulations are prepared by first evolving in imaginary time, i.e. by setting γ=0𝛾0\gamma=0italic_γ = 0, Ω=Ω0ΩsubscriptΩ0\Omega=\Omega_{0}roman_Ω = roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, and replacing tit𝑡i𝑡t\rightarrow\mathrm{i}titalic_t → roman_i italic_t in Eq. (1(e.g. Modugno et al., 2003), beginning with a random phase at each grid point. After evolving in imaginary time for a sufficiently long period (typically >>>7,500 τ𝜏\tauitalic_τ), the wavefunction achieves a quasi-equilibrium where the superfluid density and angular momentum are essentially steady. This is taken as the initial condition for the damped GP equation.

Each vortex is associated with a quantum of circulation 2π/m2𝜋Planck-constant-over-2-pi𝑚2\pi\hbar/m2 italic_π roman_ℏ / italic_m. In the absence of pinning sites we would therefore expect the initial number of vortices to be roughly NvΩ0Rcon2m/=(Ω0τ)(Rcon/ξ)2similar-to-or-equalssubscript𝑁𝑣subscriptΩ0superscriptsubscript𝑅con2𝑚Planck-constant-over-2-pisubscriptΩ0𝜏superscriptsubscript𝑅con𝜉2N_{v}\simeq\Omega_{0}R_{\mathrm{con}}^{2}m/\hbar=(\Omega_{0}\tau)(R_{\mathrm{% con}}/\xi)^{2}italic_N start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ≃ roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT roman_con end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_m / roman_ℏ = ( roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_τ ) ( italic_R start_POSTSUBSCRIPT roman_con end_POSTSUBSCRIPT / italic_ξ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, so that the average rotation rate of the superfluid is roughly Ω0subscriptΩ0\Omega_{0}roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. For example, with Rcon=230ξsubscript𝑅con230𝜉R_{\mathrm{con}}=230\xiitalic_R start_POSTSUBSCRIPT roman_con end_POSTSUBSCRIPT = 230 italic_ξ and Ω0=2π×103τ1subscriptΩ02𝜋superscript103superscript𝜏1\Omega_{0}=2\pi\times 10^{-3}\tau^{-1}roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 2 italic_π × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT italic_τ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT we would expect Nv332similar-to-or-equalssubscript𝑁𝑣332N_{v}\simeq 332italic_N start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ≃ 332. To study significantly more vortices than previous GP studies we take Ω0=2π×103τ1subscriptΩ02𝜋superscript103superscript𝜏1\Omega_{0}=2\pi\times 10^{-3}\tau^{-1}roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 2 italic_π × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT italic_τ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT or Ω0=4π×103τ1subscriptΩ04𝜋superscript103superscript𝜏1\Omega_{0}=4\pi\times 10^{-3}\tau^{-1}roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 4 italic_π × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT italic_τ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. In practice, the presence of pinning sites means that there are many different quasi-equilibrium states that can be achieved during imaginary time propagation, and the actual number of vortices can differ from this prediction by up to 20%. Even though the number of pinning sites greatly exceeds the number of vortices, there are usually a small number (5%less-than-or-similar-toabsentpercent5\lesssim 5\%≲ 5 %) of vortices that are unpinned in the initial state (see the bottom-left inset of Fig. 1 (a), for example). Following (Liu et al., 2024), we define a vortex to be pinned if it is located within 1.25w1.25𝑤1.25w1.25 italic_w of a pinning site.

Adopting a characteristic value of ξ=200𝜉200\xi=200italic_ξ = 200 fm for the coherence length in the crust (Graber et al., 2017), we find that μ2/mξ2518eV𝜇superscriptPlanck-constant-over-2-pi2𝑚superscript𝜉2similar-to-or-equals518eV\mu\equiv\hbar^{2}/m\xi^{2}\simeq 518\,\text{eV}italic_μ ≡ roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_m italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≃ 518 eV and τ/μ1.3×1018𝜏Planck-constant-over-2-pi𝜇similar-to-or-equals1.3superscript1018\tau\equiv\hbar/\mu\simeq 1.3\times 10^{-18}italic_τ ≡ roman_ℏ / italic_μ ≃ 1.3 × 10 start_POSTSUPERSCRIPT - 18 end_POSTSUPERSCRIPT s. For computational feasibility, the domain size and angular velocity used are far from the typical values for a neutron star. A comparison between the estimated physical parameters for real neutron stars and the parameters used in our simulations is presented in Appendix A. Our objective in the present work is to include as many vortices as feasible to study any resulting collective dynamics, which requires compromise with regard to the other physical parameters.

3 Spin-down dynamics

3.1 Macroscopic Observables

Refer to caption
Figure 1: Spin-down dynamics for Ω0=4π×103τ1subscriptΩ04𝜋superscript103superscript𝜏1\Omega_{0}=4\pi\times 10^{-3}\tau^{-1}roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 4 italic_π × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT italic_τ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, Ω˙=2.5π×108τ2˙Ω2.5𝜋superscript108superscript𝜏2\dot{\Omega}=2.5\pi\times 10^{-8}\tau^{-2}over˙ start_ARG roman_Ω end_ARG = 2.5 italic_π × 10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT italic_τ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT and γ=5×103𝛾5superscript103\gamma=5\times 10^{-3}italic_γ = 5 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT over a time span of 4.45×104τ4.45superscript104𝜏4.45\times 10^{4}\tau4.45 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT italic_τ. (a) Time series of Nvsubscript𝑁𝑣N_{v}italic_N start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT (black solid line), Lzdelimited-⟨⟩subscript𝐿𝑧\langle L_{z}\rangle⟨ italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩ (green dashed line), and Lzvsubscriptdelimited-⟨⟩subscript𝐿𝑧𝑣\langle L_{z}\rangle_{v}⟨ italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT (green dotted line). Glitches, i.e. sudden jumps in Lzdelimited-⟨⟩subscript𝐿𝑧\langle L_{z}\rangle⟨ italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩, are labelled as G.1 to G.9 and color coded by time. The bottom-left inset shows the initial vortex locations, whether pinned (blue filled circles) or unpinned (red hollow circles). The pinning sites are at the vertices of the square grid shown. The top-right inset shows a close-up of the first glitch (G.1). (b) Vortex trajectories in the rotating frame, colour-coded by time. The black dashed circle indicates the region r210ξ𝑟210𝜉r\leqslant 210\xiitalic_r ⩽ 210 italic_ξ within which analysis is performed, i.e. we take R=210ξ𝑅210𝜉R=210\xiitalic_R = 210 italic_ξ in Eqs. (8) and (9).

We now examine how the initial quasi-equilibrium vortex configuration responds to the linear spin-down imposed by Eq. (4). As a representative example, we focus on the case with γ=5×103𝛾5superscript103\gamma=5\times 10^{-3}italic_γ = 5 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, Ω0=4π×103τ1subscriptΩ04𝜋superscript103superscript𝜏1\Omega_{0}=4\pi\times 10^{-3}\tau^{-1}roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 4 italic_π × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT italic_τ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and Ω˙=2.5π×108τ2˙Ω2.5𝜋superscript108superscript𝜏2\dot{\Omega}=2.5\pi\times 10^{-8}\tau^{-2}over˙ start_ARG roman_Ω end_ARG = 2.5 italic_π × 10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT italic_τ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT, which is illustrated in Fig. 1. In what follows, we will pay particular attention to the number of vortices, Nvsubscript𝑁𝑣N_{v}italic_N start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT, and to the mean angular momentum,

Lz=|𝐫|Rd2𝐫ψL^zψ/|𝐫|Rd2𝐫|ψ|2.\langle L_{z}\rangle=\left.{\displaystyle\int_{|\mathbf{r}|\leqslant R}\mathrm% {d}^{2}\mathbf{r}\,\psi^{\ast}\hat{L}_{z}\psi}\middle/{\displaystyle\int_{|% \mathbf{r}|\leqslant R}\mathrm{d}^{2}\mathbf{r}\,|\psi|^{2}}\right..⟨ italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩ = ∫ start_POSTSUBSCRIPT | bold_r | ⩽ italic_R end_POSTSUBSCRIPT roman_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_r italic_ψ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT over^ start_ARG italic_L end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT italic_ψ / ∫ start_POSTSUBSCRIPT | bold_r | ⩽ italic_R end_POSTSUBSCRIPT roman_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_r | italic_ψ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (8)

To count the number of vortices we first determine their locations, at a given time, by interpolating the superfluid velocity to a sub-grid scale and identifying points of singularity. Identifying vortices is problematic close to the edge of the container, r=Rcon𝑟subscript𝑅conr=R_{\mathrm{con}}italic_r = italic_R start_POSTSUBSCRIPT roman_con end_POSTSUBSCRIPT, where the density vanishes and “ghost vortices” often arise. Hence, our subsequent analysis is performed within the subdomain |𝐫|R=210ξ0.91Rcon𝐫𝑅210𝜉similar-to-or-equals0.91subscript𝑅con|\mathbf{r}|\leqslant R=210\xi\simeq 0.91R_{\mathrm{con}}| bold_r | ⩽ italic_R = 210 italic_ξ ≃ 0.91 italic_R start_POSTSUBSCRIPT roman_con end_POSTSUBSCRIPT.

Fig. 1 (a) presents time-series of Nvsubscript𝑁𝑣N_{v}italic_N start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT and Lzdelimited-⟨⟩subscript𝐿𝑧\langle L_{z}\rangle⟨ italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩. The number of vortices exhibits clear step-like drops, each representing a loss of 1010101025252525 vortices from the subdomain within a period of <3000τabsent3000𝜏<3000\tau< 3000 italic_τ. Each drop is color coded in Fig. 1 (a), with dashed and dotted vertical lines indicating its beginning and end time, respectively. The mean angular momentum exhibits simultaneous step-like behavior, representing a sequence of rotational glitches.

Between glitches, Nvsubscript𝑁𝑣N_{v}italic_N start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT remains essentially constant, but Lzdelimited-⟨⟩subscript𝐿𝑧\langle L_{z}\rangle⟨ italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩ decreases smoothly (at a rate much smaller than Ω˙˙Ω\dot{\Omega}over˙ start_ARG roman_Ω end_ARG). These periods are associated with a spatial redistribution of vortices, essentially filling in the gaps left by vortices that have left the domain.

The close correlation between Nvsubscript𝑁𝑣N_{v}italic_N start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT and Lzdelimited-⟨⟩subscript𝐿𝑧\langle L_{z}\rangle⟨ italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩ is not surprising; if we neglect variations in the superfluid density, then we can obtain the following estimate for the mean angular momentum within |𝐫|R𝐫𝑅|\mathbf{r}|\leqslant R| bold_r | ⩽ italic_R (Fetter, 1965):

Lzv=j[1|rj|2/R2].subscriptdelimited-⟨⟩subscript𝐿𝑧𝑣Planck-constant-over-2-pisubscript𝑗delimited-[]1superscriptsubscript𝑟𝑗2superscript𝑅2\langle L_{z}\rangle_{v}=\hbar\sum_{j}\left[1-|\vec{r}_{j}|^{2}/R^{2}\right].⟨ italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT = roman_ℏ ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT [ 1 - | over→ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] . (9)

Here rj(t)subscript𝑟𝑗𝑡\vec{r}_{j}(t)over→ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) denotes the position of vortex j𝑗jitalic_j; for clarity we will use bold typeface for vector fields, and right arrow accents for vector properties of vortices. As shown in Fig. 1 (a), Eq. (9) provides an excellent approximation to the true angular momentum, Eq. (8), with a relative error of 0.5%similar-to-or-equalsabsentpercent0.5\simeq 0.5\%≃ 0.5 %. From Eq. (9) we see why the step-like behavior of Nvsubscript𝑁𝑣N_{v}italic_N start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT is also reflected in Lzdelimited-⟨⟩subscript𝐿𝑧\langle L_{z}\rangle⟨ italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩. Indeed, as long as the vortices remain roughly uniformly distributed throughout the domain we would expect Lz12Nvsimilar-to-or-equalsdelimited-⟨⟩subscript𝐿𝑧12Planck-constant-over-2-pisubscript𝑁𝑣\langle L_{z}\rangle\simeq\tfrac{1}{2}\hbar N_{v}⟨ italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩ ≃ divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_ℏ italic_N start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT. The slow decreases in Lzdelimited-⟨⟩subscript𝐿𝑧\langle L_{z}\rangle⟨ italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩ that occurs between glitches must then correspond to a slow outward migration of vortices.

The motion of vortices is illustrated in Fig. 1 (b), which plots vortex trajectories (in the spinning down frame) color coded according to time as in Fig. 1 (a). Each glitch is associated with the unpinning and outward migration of multiple vortices, localized in both time and space, which we interpret as a vortex avalanche. Each avalanche occurs within a narrow channel that is aligned roughly in the radial direction. However, individual vortex trajectories are not purely radial, and most follow roughly circular arcs in a clockwise direction, as shown in Fig. 2. This glitching behaviour occurs only if the spin-down rate, Ω˙˙Ω\dot{\Omega}over˙ start_ARG roman_Ω end_ARG, and dissipation, γ𝛾\gammaitalic_γ, are sufficiently small (see Appendix B).

Refer to caption
Figure 2: Vortex trajectories for the same simulation shown in Fig. 1 for four time windows (a) during and (b) after the first glitch (G.1). The locations of vortices at the end of each time window are shown with filled circles, which are color coded to show the magnitude of the Magnus force. The trajectories are color coded by time. The coarsened vorticity, w~~𝑤\tilde{w}over~ start_ARG italic_w end_ARG, is shown in grayscale.

3.2 Vortex Avalanches

Several mechanisms can cause a pinned vortex to depin (Warszawski et al., 2012), but the most important for neutron stars is the Magnus force, due to the relative velocity between a pinned vortex and the ambient superfluid flow. In the case of a point-vortex model, the Magnus force on the j𝑗jitalic_j-th vortex, with circulation κjsubscript𝜅𝑗\vec{\kappa}_{j}over→ start_ARG italic_κ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, is given by

Fj=κj×(vs,jdrj/dt),subscript𝐹𝑗subscript𝜅𝑗subscript𝑣𝑠𝑗dsubscript𝑟𝑗d𝑡\vec{F}_{j}=\vec{\kappa}_{j}\times(\vec{v}_{s,j}-\mathrm{d}\vec{r}_{j}/\mathrm% {d}t),over→ start_ARG italic_F end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = over→ start_ARG italic_κ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT × ( over→ start_ARG italic_v end_ARG start_POSTSUBSCRIPT italic_s , italic_j end_POSTSUBSCRIPT - roman_d over→ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT / roman_d italic_t ) , (10)

where vs,jsubscript𝑣𝑠𝑗\vec{v}_{s,j}over→ start_ARG italic_v end_ARG start_POSTSUBSCRIPT italic_s , italic_j end_POSTSUBSCRIPT represents the superfluid velocity at the point rj(t)subscript𝑟𝑗𝑡\vec{r}_{j}(t)over→ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ). This superflow is induced by all of the other vortices (plus any image vortices resulting from boundaries):

vs,j=kjκk×(rjrk)|rjrk|2𝛀×rj.subscript𝑣𝑠𝑗subscript𝑘𝑗subscript𝜅𝑘subscript𝑟𝑗subscript𝑟𝑘superscriptsubscript𝑟𝑗subscript𝑟𝑘2𝛀subscript𝑟𝑗\displaystyle\vec{v}_{s,j}=\sum_{k\neq j}\frac{\vec{\kappa}_{k}\times(\vec{r}_% {j}-\vec{r}_{k})}{\left|\vec{r}_{j}-\vec{r}_{k}\right|^{2}}-\boldsymbol{\Omega% }\times\vec{r}_{j}.over→ start_ARG italic_v end_ARG start_POSTSUBSCRIPT italic_s , italic_j end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_k ≠ italic_j end_POSTSUBSCRIPT divide start_ARG over→ start_ARG italic_κ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT × ( over→ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - over→ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) end_ARG start_ARG | over→ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - over→ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - bold_Ω × over→ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT . (11)

A vortex is expected to depin when the Magnus force exceeds a critical value.

In practice, the depinning of a vortex in the GP model is more complicated than in this simplified description, and the critical value can vary by about 20%percent2020\%20 % (e.g. Liu et al., 2024). For the values of V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and w𝑤witalic_w used here, the critical velocity is vc0.2ξ/τsimilar-to-or-equalssubscript𝑣𝑐0.2𝜉𝜏v_{c}\simeq 0.2\xi/\tauitalic_v start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≃ 0.2 italic_ξ / italic_τ, this can help anticipate depinning events. Fig. 2 presents snapshots of the vortex locations in the vicinity of the first glitch (G.1) from Fig. 1. Vortices are coloured according to |vs,j|subscript𝑣𝑠𝑗|\vec{v}_{s,j}|| over→ start_ARG italic_v end_ARG start_POSTSUBSCRIPT italic_s , italic_j end_POSTSUBSCRIPT | ( Eq. (11)), where the sum is taken over all vortices in the system. By t=8,500τ𝑡8,500𝜏t=\text{8,500}\,\tauitalic_t = 8,500 italic_τ a few vortices have reached the threshold, |vs,j|>vcsubscript𝑣𝑠𝑗subscript𝑣𝑐|\vec{v}_{s,j}|>v_{c}| over→ start_ARG italic_v end_ARG start_POSTSUBSCRIPT italic_s , italic_j end_POSTSUBSCRIPT | > italic_v start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, at which depinning is expected; the first vortex to depin is labelled with a black arrow. This vortex has a “knock-on” collision with another pinned vortex, causing it to depin (see supplementary movie). As vortices migrate outward, the residual Magnus force on other pinned vortices increases, as shown in Fig. 2(a)(iii), causing further depinning.

Fig. 2 also shows the coarsened vorticity of the superfluid, which we define as (Baggaley et al., 2012a, b)

ω¯(𝐫)=jNvκjW(|rj𝐫|,h),¯𝜔𝐫superscriptsubscript𝑗subscript𝑁𝑣subscript𝜅𝑗𝑊subscript𝑟𝑗𝐫\bar{\omega}(\mathbf{r})=\sum_{j}^{N_{v}}\kappa_{j}W(\left|\vec{r}_{j}-\mathbf% {r}\right|,h),over¯ start_ARG italic_ω end_ARG ( bold_r ) = ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_κ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_W ( | over→ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - bold_r | , italic_h ) , (12)

where W𝑊Witalic_W is a smoothing kernel (Monaghan, 1992) of width hhitalic_h. We choose the value of hhitalic_h based on the average distance between vortices, with h(t)=2.45πRcon2/Nv(t)𝑡2.45𝜋superscriptsubscript𝑅con2subscript𝑁𝑣𝑡h(t)=2.45\sqrt{\pi R_{\mathrm{con}}^{2}/N_{v}(t)}italic_h ( italic_t ) = 2.45 square-root start_ARG italic_π italic_R start_POSTSUBSCRIPT roman_con end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_N start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_t ) end_ARG. Figure  2(b) shows the subsequent dynamics, including the post-glitch period. As a result of vortices exiting the domain, a void appears in the (coarsened) vorticity. Multiple vortices depin and orbit clockwise around this void, causing the void to propagate inward and gradually disperse. By t=13,000τ𝑡13,000𝜏t=\text{13,000}\,\tauitalic_t = 13,000 italic_τ, as shown in Fig. 2(b)(iv), all vortices have Magnus forces below the threshold for depinning. We observe similar dynamics in each of the subsequent glitches.

The vortex trajectories shown in Fig. 2 suggest that vortices behave collectively both during and after the glitch. In Appendix C, we confirm this by introducing an order parameter that quantifies the local correlation of the vortex motions.

4 Conclusions

We simulated a rotating superfluid with 600similar-toabsent600\sim 600∼ 600 vortices, coupled to a spinning-down lattice of pinning sites, using a GP model. For sufficiently slow spin-down, and sufficiently small dissipation, the vortices undergo avalanches that produce glitches in the superfluid angular momentum. Each avalanche is triggered when the effective Magnus force on a few neighboring vortices exceeds a critical value, causing depinning. The movement of these vortices results in stronger Magnus forces on other pinned vortices, producing a cascade of depinning and creating a localized void in the vorticity. Depinned vortices circulate anti-cyclonically around this void, which propagates inward and gradually dissipates, until a new quasi-equilibrium state is achieved. Throughout this process, the vortex motions are locally correlated, i.e. they behave collectively.

In a real neutron star the number of vortices, and their mean separation, is many orders of magnitude larger than can be achieved computationally. However, we have shown that avalanching persists as the mean separation between vortices increases, provided that the spin-down rate and dissipation are kept sufficiently small. For more rapid spin-downs, avalanches become so frequent that they overlap in time, and so the superfluid angular momentum evolves stochastically but without sporadic changes that could be identified as glitches. A future study may test whether this scenario arises in pulsars, by examining the power series of their spin-down rates.

Our results bear the hallmarks of collective motion, as expected in the standard picture of vortex avalanches and self-organized criticality (Jensen, 1998; Melatos et al., 2008). This collective motion begins during the glitch and often continues into the post-glitch relaxation dynamics. To determine whether the glitch sizes and waiting times are consistent with the predictions of self-organized criticality theory will require many further simulations, and will be studied in later work.

We thank Brynmor Haskell and Marco Antonelli for fruitful discussions and Vanessa Graber for useful comments. This work was supported by the Science and Technology Facilities Council grant ST/W001020/1.
Table 1: The dimensionless parameters used in the model and typical orders of magnitude in a neutron star (e.g. Warszawski & Melatos, 2011; Harding, 2013). Rconsubscript𝑅conR_{\mathrm{con}}italic_R start_POSTSUBSCRIPT roman_con end_POSTSUBSCRIPT is the domain radius; w𝑤witalic_w and V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT are the width and height of the pinning potential; dpsubscript𝑑𝑝d_{p}italic_d start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is the separation between pinning sites; Ω0subscriptΩ0\Omega_{0}roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the initial angular velocity; Ω˙=dΩ/dt˙ΩdΩd𝑡\dot{\Omega}=-\mathrm{d}\Omega/\mathrm{d}tover˙ start_ARG roman_Ω end_ARG = - roman_d roman_Ω / roman_d italic_t is the spin-down rate.
Simulation Neutron Star Neutron Star
(dimensionless) (dimensional) (dimensionless)
Rconsubscript𝑅conR_{\mathrm{con}}italic_R start_POSTSUBSCRIPT roman_con end_POSTSUBSCRIPT 230 10101010 km 1013superscript101310^{13}10 start_POSTSUPERSCRIPT 13 end_POSTSUPERSCRIPT
w𝑤witalic_w 1 10 fm 0.050.050.050.05
V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT 2 1111 MeV 2×1032superscript1032\times 10^{3}2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT
dpsubscript𝑑𝑝d_{p}italic_d start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT 10 10101010 fm 0.050.050.050.05
Ω0subscriptΩ0\Omega_{0}roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (2,4)π×10324𝜋superscript103(2,4)\pi\times 10^{-3}( 2 , 4 ) italic_π × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT (100,103)superscript100superscript103(10^{0},10^{3})( 10 start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ) s-1 (1018,1015)superscript1018superscript1015(10^{-18},10^{-15})( 10 start_POSTSUPERSCRIPT - 18 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT - 15 end_POSTSUPERSCRIPT )
Ω˙˙Ω\dot{\Omega}over˙ start_ARG roman_Ω end_ARG 2.5π×(108,107)2.5𝜋superscript108superscript1072.5\pi\times(10^{-8},10^{-7})2.5 italic_π × ( 10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT ) (1024,105)superscript1024superscript105(10^{-24},10^{-5})( 10 start_POSTSUPERSCRIPT - 24 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT ) s-2 (1060,1041)superscript1060superscript1041(10^{-60},10^{-41})( 10 start_POSTSUPERSCRIPT - 60 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT - 41 end_POSTSUPERSCRIPT )

Appendix A Numerical vs. Neutron Star Parameters

In the GP model, the characteristic scales of length, time and energy are related. Taking the coherence length in the crust as a guide, ξ=200𝜉200\xi=200italic_ξ = 200 fm  (Graber et al., 2017), one can find that the characteristic energy scale is

μ2mξ2518eV𝜇superscriptPlanck-constant-over-2-pi2𝑚superscript𝜉2similar-to-or-equals518eV\mu\equiv\frac{\hbar^{2}}{m\xi^{2}}\simeq 518\,\text{eV}italic_μ ≡ divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_m italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ≃ 518 eV (A1)

and

τμ1.3×1018s.𝜏Planck-constant-over-2-pi𝜇similar-to-or-equals1.3superscript1018𝑠\tau\equiv\frac{\hbar}{\mu}\simeq 1.3\times 10^{-18}\,s.italic_τ ≡ divide start_ARG roman_ℏ end_ARG start_ARG italic_μ end_ARG ≃ 1.3 × 10 start_POSTSUPERSCRIPT - 18 end_POSTSUPERSCRIPT italic_s . (A2)

Then one can compare the relevant parameters of neutron stars in the dimensionless scale in Table 1.

Appendix B Parameter dependence in GP modeling

The results of Sec. 3 demonstrate the same kind of vortex avalanche behavior believed to occur in neutron stars. However, given that it is not possible to replicate the true parameter conditions of a neutron star in the computational model, it is important to determine the extent to which the results depend on the key parameters: the initial rotation rate, Ω0subscriptΩ0\Omega_{0}roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, the spin-down rate, Ω˙˙Ω\dot{\Omega}over˙ start_ARG roman_Ω end_ARG, and the dissipation, γ𝛾\gammaitalic_γ.

Role of initial vortex numbers

In the simulation presented in Sec. 3, the number of vortices in the region |𝐫|210ξ𝐫210𝜉|\mathbf{r}|\leqslant 210\xi| bold_r | ⩽ 210 italic_ξ decreases from Nv550similar-to-or-equalssubscript𝑁𝑣550N_{v}\simeq 550italic_N start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ≃ 550 to Nv420similar-to-or-equalssubscript𝑁𝑣420N_{v}\simeq 420italic_N start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ≃ 420, corresponding to only a modest increase in the average distance between vortices, which remains far smaller than that expected in a neutron star. To determine whether the same avalanche dynamics persists as the density of vortices decreases, we halve the initial rotation rate to Ω0=2π×103τ1subscriptΩ02𝜋superscript103superscript𝜏1\Omega_{0}=2\pi\times 10^{-3}\tau^{-1}roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 2 italic_π × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT italic_τ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. This simulation initially has around 300 vortices. As illustrated in Fig. 3, we observe similar glitching behavior in both Nvsubscript𝑁𝑣N_{v}italic_N start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT and Lzdelimited-⟨⟩subscript𝐿𝑧\langle L_{z}\rangle⟨ italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩; in fact, on average these glitches are larger than those seen in the previous results. The top-right inset in Fig. 3 shows the vortex positions at the end of the first glitch in this simulation; the bottom-left inset presents a similar plot for the previous simulation illustrated in Fig. 1 and 2. We see that, as the average density of vortices decreases, the vorticity voids produced by each avalanche become larger and more pronounced.

Role of spin-down rate

Next we investigate the role of spin-down rate by increasing Ω˙˙Ω\dot{\Omega}over˙ start_ARG roman_Ω end_ARG by a factor of 10 to 2.5π×107τ22.5𝜋superscript107superscript𝜏22.5\pi\times 10^{-7}\tau^{-2}2.5 italic_π × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT italic_τ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT. As shown in Fig. 3, the time series of Nvsubscript𝑁𝑣N_{v}italic_N start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT and Lzdelimited-⟨⟩subscript𝐿𝑧\langle L_{z}\rangle⟨ italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩ for this simulation show signs of stochasticity, but in contrast to the previous results they do not exhibit clear glitching behavior. Despite this, the vortex trajectories (not shown) display similar patterns to those presented in Figs. 1 and 2. We interpret these results as evidence of multiple vortex avalanches that overlap in time, producing time series that are smoother on average.

Role of dissipation

Finally, we investigate the role of dissipation by decreasing or increasing γ𝛾\gammaitalic_γ by a factor of 10. In the simulation with dissipation decreased to γ=5×104𝛾5superscript104\gamma=5\times 10^{-4}italic_γ = 5 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT we observe similar glitching behavior, and the typical magnitude of the glitches is not significantly affected. By contrast, in the simulation with dissipation increased to γ=5×102𝛾5superscript102\gamma=5\times 10^{-2}italic_γ = 5 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT we do not observe glitches. We conclude from this that glitches can only occur when the level of dissipation is sufficiently small, and that once it is sufficiently small it does not play a significant role in the glitch dynamics.

Refer to caption
Figure 3: Time series of Nvsubscript𝑁𝑣N_{v}italic_N start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT for different simulations. The solid black line is the representative simulation, as shown in Figs. 12. The other lines are for faster spin down (solid gray), slower rotation (solid red), stronger dissipation (dotted purple), and weaker dissipation (dashed blue). The light grey dashed line indicates the expected value of Nvsubscript𝑁𝑣N_{v}italic_N start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT in the absence of pinning sites. The insets show the vortex trajectories, Magnus forces and coarsened vorticity immediately after the first glitch in two of the simulations, using the same color scheme as Fig. 2.

Appendix C Vortex Collective Motion

Refer to caption
Figure 4: The order parameter, φ(r,t)𝜑𝑟𝑡\varphi(r,t)italic_φ ( italic_r , italic_t ), (heatmap) averaged over three consecutive measurements, and the number of moving vortices, Nv,moving(t)subscript𝑁𝑣moving𝑡N_{v,\mathrm{moving}}(t)italic_N start_POSTSUBSCRIPT italic_v , roman_moving end_POSTSUBSCRIPT ( italic_t ) (green lines). The light/dark green lines are smoothed/unsmoothed Nv,movingsubscript𝑁𝑣movingN_{v,\mathrm{moving}}italic_N start_POSTSUBSCRIPT italic_v , roman_moving end_POSTSUBSCRIPT. The glitches identified in Fig. 1 are indicated by vertical dashed and dotted lines. The order parameter vanishes for r10ξless-than-or-similar-to𝑟10𝜉r\lesssim 10\xiitalic_r ≲ 10 italic_ξ, because there are rarely any vortices within this distance, and saturates for r50ξgreater-than-or-equivalent-to𝑟50𝜉r\gtrsim 50\xiitalic_r ≳ 50 italic_ξ, which is the typical size for a cluster of moving vortices.

In order to quantify whether the vortices are exhibiting collective motion, we introduce an order parameter,

φ(r,t)=1Nv,movingj=1Nv,moving|uj|,𝜑𝑟𝑡1subscript𝑁𝑣movingsuperscriptsubscript𝑗1subscript𝑁𝑣movingsubscript𝑢𝑗\varphi(r,t)=\displaystyle\frac{1}{N_{v,\mathrm{moving}}}\sum_{j=1}^{N_{v,% \mathrm{moving}}}\left|\vec{u}_{j}\right|,italic_φ ( italic_r , italic_t ) = divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUBSCRIPT italic_v , roman_moving end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_v , roman_moving end_POSTSUBSCRIPT end_POSTSUPERSCRIPT | over→ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | , (C1)

where the sum is taken over vortices that are not pinned and are moving, and where uj(r,t)subscript𝑢𝑗𝑟𝑡\vec{u}_{j}(r,t)over→ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_r , italic_t ) measures the correlation of all moving vortices within a distance r𝑟ritalic_r of vortex j𝑗jitalic_j, i.e.

uj(r,t)=1𝒩jkCjΔrj/Δt|Δrj/Δt|,subscript𝑢𝑗𝑟𝑡1subscript𝒩𝑗subscript𝑘subscript𝐶𝑗Δsubscript𝑟𝑗Δ𝑡Δsubscript𝑟𝑗Δ𝑡\vec{u}_{j}(r,t)=\frac{1}{\mathcal{N}_{j}}\sum_{k\in C_{j}}\frac{\Delta\vec{r}% _{j}/\Delta t}{|\Delta\vec{r}_{j}/\Delta t|},over→ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_r , italic_t ) = divide start_ARG 1 end_ARG start_ARG caligraphic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_k ∈ italic_C start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT divide start_ARG roman_Δ over→ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT / roman_Δ italic_t end_ARG start_ARG | roman_Δ over→ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT / roman_Δ italic_t | end_ARG , (C2)

where Cj={kj:|rj(t)rk(t)|r}subscript𝐶𝑗conditional-set𝑘𝑗subscript𝑟𝑗𝑡subscript𝑟𝑘𝑡𝑟C_{j}=\{k\neq j:|\vec{r}_{j}(t)-\vec{r}_{k}(t)|\leqslant r\}italic_C start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = { italic_k ≠ italic_j : | over→ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) - over→ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_t ) | ⩽ italic_r }, and 𝒩jsubscript𝒩𝑗\mathcal{N}_{j}caligraphic_N start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT is the number of vortices in Cjsubscript𝐶𝑗C_{j}italic_C start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT. The locations of vortices are tracked every one unit of time, namely, Δt=τΔ𝑡𝜏\Delta t=\tauroman_Δ italic_t = italic_τ, allowing us to identify the vortex motion with sufficiently fine time resolution.

Fig. 4 presents a time series of φ𝜑\varphiitalic_φ for the same simulation shown in Fig. 1. As in Fig. 1, the start and end times of each glitch are indicated by vertical dashed and dotted lines, respectively. We see that periods of collective motion, indicated by values of φ(r,t)0.5greater-than-or-equivalent-to𝜑𝑟𝑡0.5\varphi(r,t)\gtrsim 0.5italic_φ ( italic_r , italic_t ) ≳ 0.5 over a range of r𝑟ritalic_r, typically occur throughout most of the glitch, and often continue significantly after the glitch.

References

  • Anderson & Itoh (1975) Anderson, P. W., & Itoh, N. 1975, Nature, 256, 25, doi: 10.1038/256025a0
  • Andersson et al. (2012) Andersson, N., Glampedakis, K., Ho, W. C. G., & Espinoza, C. M. 2012, Phys. Rev. Lett., 109, 241103, doi: 10.1103/PhysRevLett.109.241103
  • Avogadro et al. (2008) Avogadro, P., Barranco, F., Broglia, R. A., & Vigezzi, E. 2008, Nucl. Phys. A, 811, 378, doi: 10.1016/j.nuclphysa.2008.07.010
  • Baggaley et al. (2012a) Baggaley, A. W., Barenghi, C. F., Shukurov, A., & Sergeev, Y. A. 2012a, EPL (Europhysics Letters), 98, 26002, doi: 10.1209/0295-5075/98/26002
  • Baggaley et al. (2012b) Baggaley, A. W., Laurie, J., & Barenghi, C. F. 2012b, Phys. Rev. Lett., 109, 205304, doi: 10.1103/PhysRevLett.109.205304
  • Bak et al. (1987) Bak, P., Tang, C., & Wiesenfeld, K. 1987, Phys. Rev. Lett., 59, 381, doi: 10.1103/PhysRevLett.59.381
  • Baym et al. (1971) Baym, G., Bethe, H. A., & Pethick, C. J. 1971, Nucl. Phys. A, 175, 225, doi: 10.1016/0375-9474(71)90281-8
  • Blakie et al. (2008) Blakie, P. B., Bradley, A. S., Davis, M. J., Ballagh, R. J., & Gardiner, C. W. 2008, Advances in Physics, 57, 363, doi: 10.1080/00018730802564254
  • Bradley et al. (2008) Bradley, A. S., Gardiner, C. W., & Davis, M. J. 2008, Phys. Rev. A, 77, 033616, doi: 10.1103/PhysRevA.77.033616
  • Chamel & Haensel (2008) Chamel, N., & Haensel, P. 2008, Living Reviews in Relativity, 11, 10, doi: 10.12942/lrr-2008-10
  • Cheunchitra et al. (2024) Cheunchitra, T., Melatos, A., Carlin, J. B., & Howitt, G. 2024, Mon. Not. R. Astron. Soc., 528, 1360, doi: 10.1093/mnras/stae130
  • Drummond & Melatos (2017) Drummond, L. V., & Melatos, A. 2017, Mon. Not. R. Astron. Soc., 472, 4851, doi: 10.1093/mnras/stx2301
  • Fetter (1965) Fetter, A. L. 1965, Physical Review, 138, 429, doi: 10.1103/PhysRev.138.A429
  • Field et al. (1995) Field, S., Witt, J., Nori, F., & Ling, X. 1995, Phys. Rev. Lett., 74, 1206, doi: 10.1103/PhysRevLett.74.1206
  • Fuentes et al. (2019) Fuentes, J. R., Espinoza, C. M., & Reisenegger, A. 2019, Astron. Astrophys., 630, A115, doi: 10.1051/0004-6361/201935939
  • Graber et al. (2017) Graber, V., Andersson, N., & Hogg, M. 2017, International Journal of Modern Physics D, 26, 1730015, doi: 10.1142/S0218271817300154
  • Gügercinoğlu & Alpar (2014) Gügercinoğlu, E., & Alpar, M. A. 2014, Astrophys. J. Lett., 788, L11, doi: 10.1088/2041-8205/788/1/L11
  • Harding (2013) Harding, A. K. 2013, Frontiers of Physics, 8, 679, doi: 10.1007/s11467-013-0285-0
  • Haskell et al. (2018) Haskell, B., Khomenko, V., Antonelli, M., & Antonopoulou, D. 2018, Mon. Not. R. Astron. Soc., 481, L146, doi: 10.1093/mnrasl/sly175
  • Howitt et al. (2020) Howitt, G., Melatos, A., & Haskell, B. 2020, Mon. Not. R. Astron. Soc., 498, 320, doi: 10.1093/mnras/staa2314
  • Jensen (1998) Jensen, H. J. 1998, Self-Organized Criticality: Emergent Complex Behavior in Physical and Biological Systems (Cambridge University Press), doi: 10.1017/cbo9780511622717
  • Liu et al. (2024) Liu, I. K., Prasad, S. B., Baggaley, A. W., Barenghi, C. F., & Wood, T. S. 2024, Journal of Low Temperature Physics, doi: 10.1007/s10909-024-03064-7
  • Lönnborn et al. (2019) Lönnborn, J. R., Melatos, A., & Haskell, B. 2019, Mon. Not. R. Astron. Soc., 487, 702, doi: 10.1093/mnras/stz1302
  • Melatos et al. (2015) Melatos, A., Douglass, J. A., & Simula, T. P. 2015, Astrophys. J., 807, 132, doi: 10.1088/0004-637X/807/2/132
  • Melatos et al. (2008) Melatos, A., Peralta, C., & Wyithe, J. S. B. 2008, Astrophys. J., 672, 1103, doi: 10.1086/523349
  • Modugno et al. (2003) Modugno, M., Pricoupenko, L., & Castin, Y. 2003, European Physical Journal D, 22, 235, doi: 10.1140/epjd/e2003-00015-y
  • Monaghan (1992) Monaghan, J. J. 1992, Annu. Rev. Astron. Astrophys., 30, 543, doi: 10.1146/annurev.aa.30.090192.002551
  • Newton et al. (2015) Newton, W. G., Berger, S., & Haskell, B. 2015, Mon. Not. R. Astron. Soc., 454, 4400, doi: 10.1093/mnras/stv2285
  • Rooney et al. (2012) Rooney, S. J., Blakie, P. B., & Bradley, A. S. 2012, Phys. Rev. A, 86, 053634, doi: 10.1103/PhysRevA.86.053634
  • Warszawski & Melatos (2011) Warszawski, L., & Melatos, A. 2011, Mon. Not. R. Astron. Soc., 415, 1611, doi: 10.1111/j.1365-2966.2011.18803.x
  • Warszawski et al. (2012) Warszawski, L., Melatos, A., & Berloff, N. G. 2012, Phys. Rev. B, 85, doi: 10.1103/PhysRevB.85.104503