Crow instability of vortex lines in dipolar superfluids

Srivatsa B. Prasad srivatsa.badariprasad@newcastle.ac.uk Joint Quantum Centre Durham-Newcastle, School of Mathematics, Statistics and Physics, Newcastle University, Newcastle upon Tyne, NE1 7RU, United Kingdom    Nick G. Parker nick.parker@newcastle.ac.uk Joint Quantum Centre Durham-Newcastle, School of Mathematics, Statistics and Physics, Newcastle University, Newcastle upon Tyne, NE1 7RU, United Kingdom    Andrew W. Baggaley andrew.baggaley@newcastle.ac.uk Joint Quantum Centre Durham-Newcastle, School of Mathematics, Statistics and Physics, Newcastle University, Newcastle upon Tyne, NE1 7RU, United Kingdom
(July 4, 2024)
Abstract

In classical inviscid fluids, parallel vortices perturbed by Kelvin waves may exhibit the Crow instability, where the mutual interaction of the Kelvin modes renders them dynamically unstable. This results in the mutual approach and reconnection of the vortices, leading to a cascaded decay into smaller and smaller vortex loops. We study the Crow instability of quantum vortex lines in a superfluid enjoying the anisotropic, long-ranged dipole-dipole interaction through mean-field simulations of the dynamics of the superfluid order parameter. We observe that both the strength and direction of the dipole-dipole interaction play a crucial role in determining the Kelvin modes that are dynamically favoured. This is shown to be linked to effective enhancements or suppressions of the vortices’ curvature and can be explained by the effective dipole-dipole interaction between the vortex lines themselves. This paves the way to a deeper understanding of vortex reconnection dynamics, vortex loop cascading and turbulence in dipolar superfluids.

I Introduction

Classical inviscid fluids play host to a variety of instabilities arising from the interaction of waves and vortices, and understanding their properties is a central tenet of fluid dynamics. One such example is the celebrated Crow instability, where a pair of vortices with mutually antiparallel circulations is unstable against transverse symmetric perturbations that induce helical Kelvin waves upon the vortex tubes. These Kelvin waves grow in amplitude until the vortices reconnect, thereby forming a series of vortex loops. The subsequent cascade of reconnections to ever-smaller vortex loops eventuates in dissipation of the loops, as seen in the mitigation of wingtip vortices of aircraft by the interaction with the corresponding contrails [1]. The existence and properties of analogues of such hydrodynamic instabilities has proved to be a fruitful source of inspiration for studying quantum fluids, such as Bose-Einstein condensates, that exhibit superfluidity. Unlike classical fluids described theoretically by the Navier-Stokes equations, superfluids can flow without viscous dissipation and are characterised by a locally coherent order parameter [2]. The phase coherence of this order parameter ensures that the vorticity of a superfluid can only be nonzero along discrete, infinitesimally thin lines, each a topological defect, such that each vortex line boasts a quantised circulation and a total depletion of the superfluid density at its core [3]. As such, dynamical processes involving quantum vortices, such as reconnections and annihilations, [4, 5] cannot change their total net circulation as it is a topologically conserved quantity [3]. Whereas the ground state of an ensemble of quantum vortices – typically, a triangular vortex lattice – is well-understood [6, 7, 8, 9], the interwoven phenomena of instabilities, turbulence and disorder of incompressible (vortices) and compressible (phonon) excitations in superfluids have been of great interest for some time [10, 11, 12, 13, 14]. While certain superfluid instabilities such as the snake instability [15] have no classical counterpart it has been shown that superfluids, or mixtures thereof, can become turbulent through quantum analogues of classical hydrodynamic instabilities such as the Kelvin-Helmholtz [16, 17, 18], Rayleigh-Taylor [19, 20] and Richtmyer-Meshkov [21] instabilities. The Crow instability, too, manifests itself in the dynamics of pairs of quantum vortices of opposite circulations when subjected to either an environmental perturbation or if there are pre-existing Kelvin waves on the vortex lines [15, 22, 23, 24, 25]. An example of the latter scenario is depicted in Fig. 1, showing how the magnitude of the transverse excitations of the vortex lines grows in time till the vortices’ cores overlap at a point in space to the extent that they are able to reconnect into a pair of loops 111The method by which this figure is obtained is presented later in this article..

Refer to caption
Figure 1: The superfluid Crow instability of an antiparallel quantum vortex pair initially separated along the y𝑦yitalic_y-axis. With trsubscript𝑡rt_{\mathrm{r}}italic_t start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT the time till the first reconnection, the vortices are depicted at t=tr/3𝑡subscript𝑡r3t=t_{\mathrm{r}}/3italic_t = italic_t start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT / 3, t=2tr/3𝑡2subscript𝑡r3t=2t_{\mathrm{r}}/3italic_t = 2 italic_t start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT / 3 and t=tctr0.5𝑡subscript𝑡csubscript𝑡r0.5t=t_{\mathrm{c}}\equiv t_{\mathrm{r}}-0.5italic_t = italic_t start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT ≡ italic_t start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT - 0.5.

The vast majority of investigations of hydrodynamic instabilities in superfluids have focussed on systems where the interactions between its constituents can be approximated as being effectively isotropic and short-ranged. Dipolar Bose-Einstein condensates (dBECs), comprised of certain lanthanide atoms with large, permanent magnetic dipole moments, represent a counterpoint to this paradigm. When its atomic dipole moments are uniformly polarized by an applied magnetic field, a dBEC exhibits a wealth of exotic phenomena that arise from a delicate interplay between the long-ranged and anisotropic dipole-dipole interaction (DDI) and the residual short-ranged, effectively isotropic van der Waals repulsion [26, 27, 28]. The foremost consequence of this competition is the tendency of a dBEC to exhibit anisotropy, whether it be its density profile under external confinement [29, 30], the speed of sound and superfluid critical velocity [31, 32], or the core of an embedded quantum vortex [33, 34, 35, 36]. Another consequence is a tendency towards stratification or short-ranged order, often associated with the presence of intermediate-wavelength roton excitations, resulting in the manifestation of novel features such as non-triangular vortex lattices with density striping [37, 38, 27, 39, 35], density ripples about the cores of quasi-two-dimensional point vortices [34], stratification during turbulence [40], pattern formation [41, 42], and the hitherto elusive supersolid phase [43, 44, 45]. Theoretical studies of quantum vortex dynamics in dBECs, which are now of greater relevance than ever after the experimental realization of dipolar quantum vortices [46, 47], have also demonstrated evidence for considerable deviations away from the nondipolar limit of properties such as the velocities and trajectories of vortex pairs [34, 48, 49, 36, 50] and the Kelvin wave spectra of single vortices [51, 52]. Together, these results would suggest that the onset of a hydrodynamic instability such as the Crow instability is sensitive to the magnitude and direction of the dipole moment polarization.

Thus, in this article we present the results of a systematic analysis of the Crow instability manifesting itself in a dBEC with a uniform background density and elucidate the ways in which the various properties of the instability depend on the parameters of the DDI. In Section II we provide an overview of the dipolar Gross-Pitaevskii equation (dGPE) as employed to model perturbed quantum vortices in a uniform, three-dimensional dBEC. We also discuss the numerical methods used for propagating the dGPE and identifying vortex structures in the solutions. In Sec. III we introduce a scenario where each of a pair of antiparallel vortices is subjected to a symmetric, random perturbation and provide a qualitative discussion of the results of one such simulation, thereby illustrating the influence that the DDI has upon the vortex profiles that emerge when the Crow instability is triggered. Section IV presents a spectral study of these vortex profiles, where tracking the populations of each Kelvin wave mode allows us to identify the modes that are excited most strongly by the Crow instability, and their corresponding rates of amplitude growth, as well as the line-averaged curvature of the vortices. These findings are summarised in Sec. V along with some implications and possible generalisations of this work.

II Formalism

In this article, we investigate the dynamics of perturbed pairs of quantum vortices in a dBEC through propagating the dipolar Gross-Pitaevskii equation (dGPE) governing the superfluid order parameter, ψ(𝐫,t)𝜓𝐫𝑡\psi(\mathbf{r},t)italic_ψ ( bold_r , italic_t ). For a system composed of a single atomic species of mass m𝑚mitalic_m and magnetic dipole moment μdsubscript𝜇d\mu_{\mathrm{d}}italic_μ start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT, polarised uniformly by a magnetic field parallel to d𝑑ditalic_d, the dGPE is given by [26, 27, 28],

iψt={22m2+4π2asm[δ(𝐫)+3εddVdd]nμ}ψ.𝑖Planck-constant-over-2-pi𝜓𝑡superscriptPlanck-constant-over-2-pi22𝑚superscript24𝜋superscriptPlanck-constant-over-2-pi2subscript𝑎s𝑚delimited-[]𝛿𝐫3subscript𝜀ddsubscript𝑉dd𝑛𝜇𝜓i\hbar\frac{\partial\psi}{\partial t}=\left\{-\frac{\hbar^{2}}{2m}\nabla^{2}+% \frac{4\pi\hbar^{2}a_{\mathrm{s}}}{m}\left[\delta(\mathbf{r})+3\varepsilon_{% \mathrm{dd}}V_{\mathrm{dd}}\right]\circledast n-\mu\right\}\psi.italic_i roman_ℏ divide start_ARG ∂ italic_ψ end_ARG start_ARG ∂ italic_t end_ARG = { - divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG ∇ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + divide start_ARG 4 italic_π roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT end_ARG start_ARG italic_m end_ARG [ italic_δ ( bold_r ) + 3 italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT ] ⊛ italic_n - italic_μ } italic_ψ . (1)

The dGPE encapsulates the effects of both the short-ranged two-body interaction, mediated by assubscript𝑎sa_{\mathrm{s}}italic_a start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT, the scattering length of the atom-atom scattering potential, and the dipole-dipole interaction (DDI), which is defined in real and reciprocal space as [26]

Vdd(𝐫)subscript𝑉dd𝐫\displaystyle V_{\mathrm{dd}}(\mathbf{r})italic_V start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT ( bold_r ) =14π[13(𝐁^𝐫^)2r3],absent14𝜋delimited-[]13superscript^𝐁^𝐫2superscript𝑟3\displaystyle=\frac{1}{4\pi}\left[\frac{1-3(\hat{\mathbf{B}}\cdot\hat{\mathbf{% r}})^{2}}{r^{3}}\right],= divide start_ARG 1 end_ARG start_ARG 4 italic_π end_ARG [ divide start_ARG 1 - 3 ( over^ start_ARG bold_B end_ARG ⋅ over^ start_ARG bold_r end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_r start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG ] , (2)
V~dd(𝐪)subscript~𝑉dd𝐪\displaystyle\widetilde{V}_{\mathrm{dd}}(\mathbf{q})over~ start_ARG italic_V end_ARG start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT ( bold_q ) =(𝐁^𝐪^)213,absentsuperscript^𝐁^𝐪213\displaystyle=(\hat{\mathbf{B}}\cdot\hat{\mathbf{q}})^{2}-\frac{1}{3},= ( over^ start_ARG bold_B end_ARG ⋅ over^ start_ARG bold_q end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 3 end_ARG , (3)

respectively. We represent the ratio between the interaction strengths of the DDI and the short-ranged interaction through the parameter εdd=mμ0μd2/(12π2as)subscript𝜀dd𝑚subscript𝜇0superscriptsubscript𝜇d212𝜋superscriptPlanck-constant-over-2-pi2subscript𝑎s\varepsilon_{\mathrm{dd}}=m\mu_{0}\mu_{\mathrm{d}}^{2}/(12\pi\hbar^{2}a_{% \mathrm{s}})italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT = italic_m italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 12 italic_π roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT ), where μ0subscript𝜇0\mu_{0}italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the permeability of free space. Additionally, in Eq. (1) the superfluid’s density and chemical potential are represented as n=|ψ|2𝑛superscript𝜓2n=|\psi|^{2}italic_n = | italic_ψ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and μ𝜇\muitalic_μ, respectively.

As encapsulated by Eq. (2), the DDI between any two atoms in the dBEC is repulsive (attractive) when the angle between them, arccos(𝐁^𝐫^)^𝐁^𝐫\arccos(\hat{\mathbf{B}}\cdot\hat{\mathbf{r}})roman_arccos ( over^ start_ARG bold_B end_ARG ⋅ over^ start_ARG bold_r end_ARG ), is less (more) than the critical angle θc=arccos(1/3)54.7degsubscript𝜃c1354.7deg\theta_{\mathrm{c}}=\arccos(1/\sqrt{3})\approx 54.7\,\mathrm{deg}italic_θ start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT = roman_arccos ( 1 / square-root start_ARG 3 end_ARG ) ≈ 54.7 roman_deg. The resulting tendency of the superfluid atoms to realign themselves to minimise their dipolar energy is responsible for magnetostrictive effects where the condensate density exhibits anisotropy where a nondipolar condensate would not. When εdd>1subscript𝜀dd1\varepsilon_{\mathrm{dd}}>1italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT > 1, the solutions of Eq. (1) are unstable to collapse as the anisotropic mutual dipolar attraction of the atoms overwhelms the short-ranged repulsion, necessitating the inclusion of a beyond-mean-field energy correction [53, 54, 55] in such regimes to account for the existence and stability of exotic states such as quantum droplets [56, 57] and supersolids [43, 44, 45]. For the sake of simplicity, we focus on the regime εdd<1subscript𝜀dd1\varepsilon_{\mathrm{dd}}<1italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT < 1, where no such correction is necessary [28]. This renders it convenient to work in natural units where energies are scaled by the ground state value of the chemical potential, μg=4π2asn0/msubscript𝜇g4𝜋superscriptPlanck-constant-over-2-pi2subscript𝑎ssubscript𝑛0𝑚\mu_{\mathrm{g}}=4\pi\hbar^{2}a_{\mathrm{s}}n_{0}/mitalic_μ start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT = 4 italic_π roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_m, for a uniform background density n0subscript𝑛0n_{0}italic_n start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Simultaneously, space and time are scaled in units of the healing length ξ=/mμg𝜉Planck-constant-over-2-pi𝑚subscript𝜇g\xi=\hbar/\sqrt{m\mu_{\mathrm{g}}}italic_ξ = roman_ℏ / square-root start_ARG italic_m italic_μ start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT end_ARG, and /μgPlanck-constant-over-2-pisubscript𝜇g\hbar/\mu_{\mathrm{g}}roman_ℏ / italic_μ start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT, respectively, while the order parameter is scaled by n0subscript𝑛0\sqrt{n_{0}}square-root start_ARG italic_n start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG [58].

Our results are based on numerical solutions of the dGPE obtained by propagation via the split-step pseudospectral method [59]. Throughout this article, we work in a domain with dimensions {Lx,Ly,Lz}={100ξ, 100ξ, 200ξ}subscript𝐿𝑥subscript𝐿𝑦subscript𝐿𝑧100𝜉100𝜉200𝜉\{L_{x},\,L_{y},\,L_{z}\}=\{100\xi,\,100\xi,\,200\xi\}{ italic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT , italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT } = { 100 italic_ξ , 100 italic_ξ , 200 italic_ξ } along the x𝑥xitalic_x, y𝑦yitalic_y- and z𝑧zitalic_z-axes, respectively, corresponding to a spatial grid of 256× 256× 512256256512256\,\times\,256\,\times\,512256 × 256 × 512 points. The use of discrete Fourier transforms in the pseudospectral scheme to calculate spatial derivatives results in periodic boundary conditions being imposed upon the solutions of Eq. (1). The dGPE admits quantum vortex solutions, characterised by a quantized circulation [3], viz.

Γ=d𝐬𝐯=2πq:q,:Γcontour-integraldifferential-d𝐬𝐯2𝜋𝑞𝑞\Gamma=\oint\,\mathrm{d}\mathbf{s}\cdot\mathbf{v}=2\pi q\,:\,q\in\mathbb{Z},roman_Γ = ∮ roman_d bold_s ⋅ bold_v = 2 italic_π italic_q : italic_q ∈ blackboard_Z , (4)

of the superfluid velocity, 𝐯=[Im(logψ)]𝐯Im𝜓\mathbf{v}=\nabla\left[\mathrm{Im}(\log\psi)\right]bold_v = ∇ [ roman_Im ( roman_log italic_ψ ) ], around the vortex core. As ψ𝜓\psiitalic_ψ is singly-valued, this implies that the superfluid density vanishes at the vortex core. To efficiently identify regions of the spatial domain where the superfluid is characterized by both a quantized circulation and a vanishing superfluid density and a quantized circulation, we compute the superfluid’s pseudovorticity [60],

𝝎ps=12×(n𝐯)Re[ψ]×Im[ψ].subscript𝝎ps12𝑛𝐯Redelimited-[]𝜓Imdelimited-[]𝜓\boldsymbol{\omega}_{\mathrm{ps}}=\frac{1}{2}\nabla\times\left(n\mathbf{v}% \right)\equiv\nabla\mathrm{Re}[\psi]\times\nabla\mathrm{Im}[\psi].bold_italic_ω start_POSTSUBSCRIPT roman_ps end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∇ × ( italic_n bold_v ) ≡ ∇ roman_Re [ italic_ψ ] × ∇ roman_Im [ italic_ψ ] . (5)

This quantity is only nonzero in the vicinity of a quantum vortex core and has thus been used extensively in recent studies of superfluid vortex dynamics to locate and track vortices [25, 61, 62, 63]. Subsequently, we employ the Newton-Raphson method to obtain subgrid interpolations of the space curve defining each vortex line in the domain.

III Evolution of the Vortex Lines

In a realistic experimental scenario, the Crow instability of two antiparallel vortices would arise from an initial perturbation by a source of compressible energy, where vortex-sound interactions seed Kelvin waves oscillations along the vortices [1]. In a trapped Bose-Einstein condensate, one such scenario would occur when a pair of vortices is located in a region where the trapping potential landscape is not flat, thereby resulting in background density gradients that perturb the vortices [23, 24]. In this article we have no such external perturbation and, instead, the perturbation is implicit in the initial conditions of the vortices [15, 22, 25]. Thus, we initialize our simulations with the superfluid phase of an unperturbed pair of straight, parallel vortex lines with mutually opposed circulations and apply a random perturbation to the loci of the vortex lines along their respective lengths. Let us assume without loss of generality that the two respective vorticities of the unperturbed vortices are (anti-)parallel to the z𝑧zitalic_z-axis. Noting that the phase of ψ𝜓\psiitalic_ψ, S𝑆Sitalic_S, is effectively a superfluid velocity potential, i.e. 𝐯=S𝐯𝑆\mathbf{v}=\nabla Sbold_v = ∇ italic_S, let us recall the classical expression for the velocity potential generated by a pair of straight, antiparallel vortices in an incompressible fluid as defined in the periodic domain x[0,Lx),y[0,Ly)formulae-sequence𝑥0subscript𝐿𝑥𝑦0subscript𝐿𝑦x\in[0,L_{x}),\,y\in[0,L_{y})italic_x ∈ [ 0 , italic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ) , italic_y ∈ [ 0 , italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) [64, 65, 66]:

S(x,y)=p={\displaystyle S(x,y)=\sum_{p=-\infty}^{\infty}\biggl{\{}italic_S ( italic_x , italic_y ) = ∑ start_POSTSUBSCRIPT italic_p = - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT { atan[tanh(πY±Ly+pπ)tan(πX±Lyπ2)]minus-or-plusatan𝜋subscript𝑌plus-or-minussubscript𝐿𝑦𝑝𝜋𝜋subscript𝑋plus-or-minussubscript𝐿𝑦𝜋2\displaystyle\mp\operatorname{atan}\left[\tanh\left(\frac{\pi Y_{\pm}}{L_{y}}+% p\pi\right)\tan\left(\frac{\pi X_{\pm}}{L_{y}}-\frac{\pi}{2}\right)\right]∓ roman_atan [ roman_tanh ( divide start_ARG italic_π italic_Y start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT end_ARG start_ARG italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT end_ARG + italic_p italic_π ) roman_tan ( divide start_ARG italic_π italic_X start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT end_ARG start_ARG italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT end_ARG - divide start_ARG italic_π end_ARG start_ARG 2 end_ARG ) ]
+π[Θ(X+)Θ(X)]}2π(x+x)yLxLy,\displaystyle+\pi\left[\Theta(X_{+})-\Theta(X_{-})\right]\biggr{\}}-\frac{2\pi% (x_{+}-x_{-})y}{L_{x}L_{y}},+ italic_π [ roman_Θ ( italic_X start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ) - roman_Θ ( italic_X start_POSTSUBSCRIPT - end_POSTSUBSCRIPT ) ] } - divide start_ARG 2 italic_π ( italic_x start_POSTSUBSCRIPT + end_POSTSUBSCRIPT - italic_x start_POSTSUBSCRIPT - end_POSTSUBSCRIPT ) italic_y end_ARG start_ARG italic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT end_ARG , (6)

Here, (X±,Y±)=(xx±,yy±)subscript𝑋plus-or-minussubscript𝑌plus-or-minus𝑥subscript𝑥plus-or-minus𝑦subscript𝑦plus-or-minus(X_{\pm},Y_{\pm})=(x-x_{\pm},y-y_{\pm})( italic_X start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT , italic_Y start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ) = ( italic_x - italic_x start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT , italic_y - italic_y start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ) where the vortex with circulation ±2πplus-or-minus2𝜋\pm 2\pi± 2 italic_π is located at (x±,y±)subscript𝑥plus-or-minussubscript𝑦plus-or-minus(x_{\pm},y_{\pm})( italic_x start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ) 222In practice, replacing the infinite sum over p𝑝pitalic_p with a partial sum from p=P𝑝𝑃p=-Pitalic_p = - italic_P to p=P𝑝𝑃p=Pitalic_p = italic_P results in a rapid convergence of S(x,y)���𝑥𝑦S(x,y)italic_S ( italic_x , italic_y ) for small positive P𝑃Pitalic_P; in this investigation we have fixed P=11𝑃11P=11italic_P = 11 in line with our previous work.. Let us define the following random perturbation profile,

δw(z)=q=2020exp[iπ(2ηq1)+2πiqzLz]exp[iπ(2η01)],𝛿𝑤𝑧superscriptsubscript𝑞2020𝑖𝜋2subscript𝜂𝑞12𝜋𝑖𝑞𝑧subscript𝐿𝑧𝑖𝜋2subscript𝜂01\delta w(z)=\sum_{q=-20}^{20}\exp\left[i\pi(2\eta_{q}-1)+\frac{2\pi iqz}{L_{z}% }\right]-\exp\left[i\pi(2\eta_{0}-1)\right],italic_δ italic_w ( italic_z ) = ∑ start_POSTSUBSCRIPT italic_q = - 20 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 20 end_POSTSUPERSCRIPT roman_exp [ italic_i italic_π ( 2 italic_η start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT - 1 ) + divide start_ARG 2 italic_π italic_i italic_q italic_z end_ARG start_ARG italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG ] - roman_exp [ italic_i italic_π ( 2 italic_η start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 1 ) ] , (7)

representing the lowest 40404040 excited modes along the z𝑧zitalic_z-axis equally populated with a random phase ηqsubscript𝜂𝑞\eta_{q}italic_η start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT drawn from the uniform distribution in the interval [0,1)01[0,1)[ 0 , 1 ). The initial phase of ψ𝜓\psiitalic_ψ in our simulations of the Crow instability is then given by Eq. (6) with (x±,y±)[x±Re(δw),y±Im(δw)]maps-tosubscript𝑥plus-or-minussubscript𝑦plus-or-minussubscript𝑥plus-or-minusRe𝛿𝑤subscript𝑦plus-or-minusIm𝛿𝑤(x_{\pm},y_{\pm})\mapsto[x_{\pm}-\mathrm{Re}(\delta w),y_{\pm}-\mathrm{Im}(% \delta w)]( italic_x start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ) ↦ [ italic_x start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT - roman_Re ( italic_δ italic_w ) , italic_y start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT - roman_Im ( italic_δ italic_w ) ]. Starting from a spatially uniform profile as an initial ansatz for the superfluid density, the dGPE is evolved in imaginary time, viz. under the substitution titmaps-to𝑡𝑖𝑡t\mapsto ititalic_t ↦ italic_i italic_t, with our phase ansatz held fixed until convergence of ψ𝜓\psiitalic_ψ is achieved.

In the dimensionless dGPE theory there exist two independent parametric degrees of freedom: 𝐁𝐁\mathbf{B}bold_B and εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT. In our work we consider εdd{0,0.1,0.2,0.8,0.9}subscript𝜀dd00.10.20.80.9\varepsilon_{\mathrm{dd}}\in\{0,0.1,0.2,\mathellipsis 0.8,0.9\}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT ∈ { 0 , 0.1 , 0.2 , … 0.8 , 0.9 }. This allows us to investigate the emergence of the Crow instability across a range of regimes from the nondipolar limit to just below the upper limit of the range of validity of the mean-field theory. For each value of εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT, the initial unperturbed separation of the vortices in the x𝑥xitalic_x-y𝑦yitalic_y plane is taken to be d=6.25ξ𝑑6.25𝜉d=6.25\xiitalic_d = 6.25 italic_ξ along the y𝑦yitalic_y-axis and 𝐁𝐁\mathbf{B}bold_B is allowed to be (anti)parallel to each vortex line (𝐁z^conditional𝐁^𝑧\mathbf{B}\parallel\hat{z}bold_B ∥ over^ start_ARG italic_z end_ARG), parallel to the mean vortex separation (𝐁y^conditional𝐁^𝑦\mathbf{B}\parallel\hat{y}bold_B ∥ over^ start_ARG italic_y end_ARG) or parallel to the net velocity of the vortex pair (𝐁x^conditional𝐁^𝑥\mathbf{B}\parallel\hat{x}bold_B ∥ over^ start_ARG italic_x end_ARG). For each choice of 𝐁𝐁\mathbf{B}bold_B and εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT, we have conducted simulations of the Crow instability over an ensemble consisting of 4444 randomly selected sets of the initial Kelvin wave phase profile, {ηq}subscript𝜂𝑞\{\eta_{q}\}{ italic_η start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT }.

Refer to caption
Figure 2: The Crow instability depicted for εdd=0.9subscript𝜀dd0.9\varepsilon_{\mathrm{dd}}=0.9italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT = 0.9 and 𝐁𝐁\mathbf{B}bold_B parallel to x^^𝑥\hat{x}over^ start_ARG italic_x end_ARG (a), y^^𝑦\hat{y}over^ start_ARG italic_y end_ARG (b) and z^^𝑧\hat{z}over^ start_ARG italic_z end_ARG (c).

First, let us inspect the profiles of the vortex filaments during the evolution of the dGPE. In Fig. 1 we presented snapshots of two vortex lines exhibiting the Crow instability in the nondipolar limit. Below, in Fig. 2, we depict the corresponding snapshots of the vortex pair with the same initial conditions but with εdd=0.9subscript𝜀dd0.9\varepsilon_{\mathrm{dd}}=0.9italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT = 0.9 and 𝐁𝐁\mathbf{B}bold_B being parallel to x^^𝑥\hat{x}over^ start_ARG italic_x end_ARG (first row), y^^𝑦\hat{y}over^ start_ARG italic_y end_ARG (second row), or z^^𝑧\hat{z}over^ start_ARG italic_z end_ARG (third row). In each of these subplots, the vortices are plotted at the times t=tr/3𝑡subscript𝑡r3t=t_{\mathrm{r}}/3italic_t = italic_t start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT / 3, t=2tr/3𝑡2subscript𝑡r3t=2t_{\mathrm{r}}/3italic_t = 2 italic_t start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT / 3 and 0.50.50.50.5 units of dimensionless time before the first reconnection occurs at t=tr𝑡subscript𝑡rt=t_{\mathrm{r}}italic_t = italic_t start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT; for the sake of brevity we denote this value of t𝑡titalic_t as tc=tr1/2subscript𝑡csubscript𝑡r12t_{\mathrm{c}}=t_{\mathrm{r}}-1/2italic_t start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT = italic_t start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT - 1 / 2. Note that trsubscript𝑡rt_{\mathrm{r}}italic_t start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT, and thus tcsubscript𝑡ct_{\mathrm{c}}italic_t start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT, is a nontrivial function of 𝐁𝐁\mathbf{B}bold_B and εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT as well as initial conditions, and that the translation speed of the mean orientation of the vortex pair is also a function of 𝐁𝐁\mathbf{B}bold_B and εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT [36], such that in each row trsubscript𝑡rt_{\mathrm{r}}italic_t start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT attains a different value and the vortices are displaced by different distances along the y𝑦yitalic_y-axes from their initial positions.

Figure 2 highlights that the final vortex profiles shortly before the reconnection induced by the Crow instability are strongly dependent on the direction of the dipolar polarisation relative to the orientation of the vortices. Compared to the corresponding vortex lines in the nondipolar limit in Fig. 1, the vortex lines when 𝐁z^conditional𝐁^𝑧\mathbf{B}\parallel\hat{z}bold_B ∥ over^ start_ARG italic_z end_ARG are noticeably straighter for each of the three times as depicted in Fig. 1 (c); whereas the nondipolar limit is characterised by a highly agitated pair of vortex lines when t=tc𝑡subscript𝑡ct=t_{\mathrm{c}}italic_t = italic_t start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT with multiple points along each vortex line where the vortices are curved towards each other and appear to heading for reconnection, only one such point is evident in the z𝑧zitalic_z-polarized case. The opposite case occurs for a dipole polarization along the y𝑦yitalic_y-axis, which is parallel to the mean separation of the vortices and is depicted in Fig. 2 (b). In this case, while the first reconnection occurs at approximately the same location along the z𝑧zitalic_z-axis as in the nondipolar limit, the spatial separation of the vortices at the other two separation minima is appreciably smaller, suggesting that the vortices would undergo their second and third reconnections sooner than when εdd=0subscript𝜀dd0\varepsilon_{\mathrm{dd}}=0italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT = 0. The third case, where 𝐁z^conditional𝐁^𝑧\mathbf{B}\parallel\hat{z}bold_B ∥ over^ start_ARG italic_z end_ARG, is distinguishable from both of the other polarizations as the points where the vortex separation attains local minima are characterised by larger displacements of each vortex from its line-averaged mean position along the x𝑥xitalic_x-axis than for any of the other polarizations; this is particularly evident when t=tc𝑡subscript𝑡ct=t_{\mathrm{c}}italic_t = italic_t start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT.

IV Properties of the Vortex Lines

In Sec. III we have shown that the profiles of antiparallel vortex lines subjected to a small initial perturbation depend considerably on the strength and orientation of the DDI experienced by the background dipolar superfluid. This would suggest that the populations of the Kelvin waves that are excited through the mechanism underlying the Crow instability are dependent in some manner on the specifics of the DDI. In this section we investigate the spectral properties of the vortex line profiles and characterise the Kelvin wave populations during the simulations of the dGPE presented in the previous section.

Let us proceed by representing the coordinates of the n𝑛nitalic_nth vortex line as wn(z)=xn(z)+iyn(z)subscript𝑤𝑛𝑧subscript𝑥𝑛𝑧𝑖subscript𝑦𝑛𝑧w_{n}(z)=x_{n}(z)+iy_{n}(z)italic_w start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_z ) = italic_x start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_z ) + italic_i italic_y start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_z ). Shifting the average of these coordinates over the z𝑧zitalic_z-axis to the origin of the x𝑥xitalic_x-y𝑦yitalic_y plane yields the quantity w~n(z,t)=wn(z,t)wn(t)subscript~𝑤𝑛𝑧𝑡subscript𝑤𝑛𝑧𝑡delimited-⟨⟩subscript𝑤𝑛𝑡\tilde{w}_{n}(z,t)=w_{n}(z,t)-\langle w_{n}\rangle(t)over~ start_ARG italic_w end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_z , italic_t ) = italic_w start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_z , italic_t ) - ⟨ italic_w start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ ( italic_t ), where w~n(z,t=0)=(1)nδw(z)subscript~𝑤𝑛𝑧𝑡0superscript1𝑛𝛿𝑤𝑧\tilde{w}_{n}(z,t=0)=(-1)^{n}\delta w(z)over~ start_ARG italic_w end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_z , italic_t = 0 ) = ( - 1 ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_δ italic_w ( italic_z ) is the initial Kelvin wave perturbation given by Eq. (7). The corresponding mode amplitudes, Wn(kz,t)subscript𝑊𝑛subscript𝑘𝑧𝑡W_{n}(k_{z},t)italic_W start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_t ), are the discrete Fourier transforms of w~n(z,t)subscript~𝑤𝑛𝑧𝑡\tilde{w}_{n}(z,t)over~ start_ARG italic_w end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_z , italic_t ) and the periodicity of the domain ensures a restriction of the mode wavenumber, kzsubscript𝑘𝑧k_{z}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, such that kz=2πq/Lz,qformulae-sequencesubscript𝑘𝑧2𝜋𝑞subscript𝐿𝑧𝑞k_{z}=2\pi q/L_{z},\,q\in\mathbb{Z}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 2 italic_π italic_q / italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_q ∈ blackboard_Z. We also note that, by construction, Wn(kz=0,t)=0subscript𝑊𝑛subscript𝑘𝑧0𝑡0W_{n}(k_{z}=0,t)=0italic_W start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0 , italic_t ) = 0. Initially, we study the temporal evolution of these mode amplitudes, focussing on those initially occupied at t=0𝑡0t=0italic_t = 0. Let us define the proportional amplitude of a mode, W(kz,t)superscript𝑊subscript𝑘𝑧𝑡W^{\prime}(k_{z},t)italic_W start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_t ), as

W(kz,t)=|W1(kz,t)|+|W2(kz,t)|kz>0[|W1(kz,t)|+|W2(kz,t)|].superscript𝑊subscript𝑘𝑧𝑡subscript𝑊1subscript𝑘𝑧𝑡subscript𝑊2subscript𝑘𝑧𝑡subscriptsubscript𝑘𝑧0delimited-[]subscript𝑊1subscript𝑘𝑧𝑡subscript𝑊2subscript𝑘𝑧𝑡W^{\prime}(k_{z},t)=\frac{|W_{1}(k_{z},t)|+|W_{2}(-k_{z},t)|}{\sum_{k_{z}>0}% \left[|W_{1}(k_{z},t)|+|W_{2}(-k_{z},t)|\right]}.italic_W start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_t ) = divide start_ARG | italic_W start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_t ) | + | italic_W start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( - italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_t ) | end_ARG start_ARG ∑ start_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT > 0 end_POSTSUBSCRIPT [ | italic_W start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_t ) | + | italic_W start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( - italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_t ) | ] end_ARG . (8)

Figure 3 depicts W(kz,t)superscript𝑊subscript𝑘𝑧𝑡W^{\prime}(k_{z},t)italic_W start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_t ), with kzsubscript𝑘𝑧k_{z}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT restricted to the 20202020 lowest-lying modes with positive wavenumber, for the nondipolar system (a) as well as the maximally dipolar regime, εdd=0.9subscript𝜀dd0.9\varepsilon_{\mathrm{dd}}=0.9italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT = 0.9, with 𝐁𝐁\mathbf{B}bold_B parallel to x^^𝑥\hat{x}over^ start_ARG italic_x end_ARG (b), y^^𝑦\hat{y}over^ start_ARG italic_y end_ARG (c), or z^^𝑧\hat{z}over^ start_ARG italic_z end_ARG (d). For each distinct regime, we average the results over the 4444 sets of initial conditions over the duration of time in which all 4444 simulations have not yet undergone a vortex reconnection.

Refer to caption
Figure 3: The relative occupation of the first 20 Kelvin wave modes with positive kzsubscript𝑘𝑧k_{z}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT during time evolution, averaged over the ensemble of initial state Kelvin wave profiles. In (a), (top left), εdd=0subscript𝜀dd0\varepsilon_{\mathrm{dd}}=0italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT = 0 and in the remaining subfigures, εdd=0.9subscript𝜀dd0.9\varepsilon_{\mathrm{dd}}=0.9italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT = 0.9; 𝐁𝐁\mathbf{B}bold_B is parallel to x^^𝑥\hat{x}over^ start_ARG italic_x end_ARG in (b), y^^𝑦\hat{y}over^ start_ARG italic_y end_ARG (c) and z^^𝑧\hat{z}over^ start_ARG italic_z end_ARG (d).

An inspection of Fig. 3 affords us the following insights into the evolution of the Kelvin modes on the vortices. Regardless of the dipolar regime, we can see that at early times the interaction of the modes results in fluctuations of the relative mode populations, accompanied by the suppression of strongly energetic high order modes with p10greater-than-or-equivalent-to𝑝10p\gtrsim 10italic_p ≳ 10. It is not until the vortices enter their final stage of evolution – that is, with reference to the snapshots of the vortex states in Figs. 1 and  2, when t>2tr/3𝑡2subscript𝑡r3t>2t_{\mathrm{r}}/3italic_t > 2 italic_t start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT / 3 – that unambiguous signatures of a dependence of the Kelvin mode population on 𝐁𝐁\mathbf{B}bold_B and εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT become apparent. For large values of εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT with the dipole moments polarised (anti-)parallel to the vortex lines, as is the case in Fig. 3(d), the Kelvin mode populations become monotonically decreasing as a function of kzsubscript𝑘𝑧k_{z}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT as ttr𝑡subscript𝑡rt\rightarrow t_{\mathrm{r}}italic_t → italic_t start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT. A comparison with Fig. 2 (c) shows that these spectral features are consistent with the qualitative appearance of the vortex lines at late times since, for both vortices at t=tc𝑡subscript𝑡ct=t_{\mathrm{c}}italic_t = italic_t start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT, only one pair of antinodes of the vortex displacement from its mean position is apparent. By contrast, it is clear that polarising the dipole moments orthogonal to the vorticity of either vortex stimulates the population of higher Kelvin modes. The averaged spectrum for polarisations along the translational axis of the vortex pair in Fig. 3(b) illustrates clearly that the q=3𝑞3q=3italic_q = 3 mode becomes overwhelmingly preferred as the Crow instability manifests itself and the vortices approach their first reconnection. In line with our observations for the 𝐁z^conditional𝐁^𝑧\mathbf{B}\parallel\hat{z}bold_B ∥ over^ start_ARG italic_z end_ARG case, this aspect of the Kelvin wave spectrum for the 𝐁x^conditional𝐁^𝑥\mathbf{B}\parallel\hat{x}bold_B ∥ over^ start_ARG italic_x end_ARG case is consistent with the plot of the vortices’ state at t=tc𝑡subscript𝑡ct=t_{\mathrm{c}}italic_t = italic_t start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT in Fig. 2 (a) where 3333 distinct antinode pairs are distinguishable on each vortex. Intriguingly, while Fig. 3 (c) shows that odd values of the Kelvin mode index q𝑞qitalic_q appear to be preferred as ttc𝑡subscript𝑡ct\rightarrow t_{\mathrm{c}}italic_t → italic_t start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT when the dipole polarisation is parallel to the vortices’ mean separation, it is evident that no single mode is overwhelmingly dominant in this regime and, similarly to the nondipolar limit in Fig. 3 (a), the first 6666 modes contribute to the majority of the mode population when 𝐁y^conditional𝐁^𝑦\mathbf{B}\parallel\hat{y}bold_B ∥ over^ start_ARG italic_y end_ARG. Indeed, while the two antinodes of the vortices at z20𝑧20z\approx-20italic_z ≈ - 20 and z50𝑧50z\approx-50italic_z ≈ - 50 when t=tc𝑡subscript𝑡ct=t_{\mathrm{c}}italic_t = italic_t start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT appear to be closer to instigating reconnections in Fig. 2 (b) than the corresponding antinodes are in Fig. 1, the nondipolar vortices are otherwise much more similar to the dipolar vortices in Fig. 2 (b) than those in (a) and (c) when t=tc𝑡subscript𝑡ct=t_{\mathrm{c}}italic_t = italic_t start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT.

Refer to caption
Figure 4: The wavenumber, kzsubscript𝑘𝑧k_{z}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, of the Kelvin mode that is maximally occupied by the vortex pair immediately prior to their mutual reconnection as a function of εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT where the dipole polarisation is parallel to x^^𝑥\hat{x}over^ start_ARG italic_x end_ARG (a), y^^𝑦\hat{y}over^ start_ARG italic_y end_ARG (b), and z^^𝑧\hat{z}over^ start_ARG italic_z end_ARG (c). In each of these subplots, distinct curves correspond to distinct Kelvin wave initial conditions indexed by η{1, 2, 3, 4}𝜂1234\eta\in\{1,\,2,\,3,\,4\}italic_η ∈ { 1 , 2 , 3 , 4 }.

With these insights on the evolution of W(kz,t)superscript𝑊subscript𝑘𝑧𝑡W^{\prime}(k_{z},t)italic_W start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_t ) over all of the low-lying modes for the maximally dipolar regime, εdd=0.9subscript𝜀dd0.9\varepsilon_{\mathrm{dd}}=0.9italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT = 0.9, we proceed to study specific features of the mode spectrum for the full range of εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT from 00 to 0.90.90.90.9. Initially, let us consider the maximally unstable Kelvin mode for a given choice of {𝐁,εdd}𝐁subscript𝜀dd\{\mathbf{B},\varepsilon_{\mathrm{dd}}\}{ bold_B , italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT }, viz. the value of kzsubscript𝑘𝑧k_{z}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, denoted hereafter as kcsubscript𝑘ck_{\mathrm{c}}italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT, that maximises W(kz,tc)superscript𝑊subscript𝑘𝑧subscript𝑡cW^{\prime}(k_{z},t_{\mathrm{c}})italic_W start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT ). Figure 4 presents plots of this wavenumber as a function of εdd[0.0.9]subscript𝜀dddelimited-[]0.0.9\varepsilon_{\mathrm{dd}}\in[0.0.9]italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT ∈ [ 0.0.9 ], with 𝐁𝐁\mathbf{B}bold_B parallel to x^^𝑥\hat{x}over^ start_ARG italic_x end_ARG (a), y^^𝑦\hat{y}over^ start_ARG italic_y end_ARG (b), and z^^𝑧\hat{z}over^ start_ARG italic_z end_ARG (c), where each choice of initial condition has been plotted as a distinct curve in each subplot. In the nondipolar limit we find that, depending on the initial conditions, either the 3rd or 5th Kelvin modes are most strongly excited through the Crow instability. Taking the higher of the two modes and comparing its wavenumber, kcξ=10π/2000.157subscript𝑘c𝜉10𝜋2000.157k_{\mathrm{c}}\xi=10\pi/200\approx 0.157italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT italic_ξ = 10 italic_π / 200 ≈ 0.157, to earlier studies of the nondipolar Crow instability shows that this is consistent with predictions that dkc1similar-to𝑑subscript𝑘𝑐1dk_{c}\sim 1italic_d italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ∼ 1 [22].

When the dipolar interaction strength is nonzero the behaviour of kcsubscript𝑘ck_{\mathrm{c}}italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT is dependent to the dipole polarisation relative to the vortex configuration. When 𝐁y^conditional𝐁^𝑦\mathbf{B}\parallel\hat{y}bold_B ∥ over^ start_ARG italic_y end_ARG, kcsubscript𝑘ck_{\mathrm{c}}italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT is particularly insensitive to εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT and the specific mode that is most strongly excited as a consequence of the Crow instability appears to be determined only by our choice of initial conditions. This, we note, is consistent with the ensemble-averaged relative mode populations in Fig. 3 being qualitatively similar for both the nondipolar limit (a) and for the regime εdd=0.9,𝐁y^subscript𝜀dd0.9conditional𝐁^𝑦\varepsilon_{\mathrm{dd}}=0.9,\,\mathbf{B}\parallel\hat{y}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT = 0.9 , bold_B ∥ over^ start_ARG italic_y end_ARG (c) as ttc𝑡subscript𝑡ct\rightarrow t_{\mathrm{c}}italic_t → italic_t start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT. However, such robustness against changes in the relative dipolar interactions strength is not evident when the dipole moments are polarised along either 𝐱𝐱\mathbf{x}bold_x or 𝐳𝐳\mathbf{z}bold_z. In Fig. 2 (d), for the regime εdd=0.9,𝐁z^subscript𝜀dd0.9conditional𝐁^𝑧\varepsilon_{\mathrm{dd}}=0.9,\,\mathbf{B}\parallel\hat{z}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT = 0.9 , bold_B ∥ over^ start_ARG italic_z end_ARG it was apparent that the q=1𝑞1q=1italic_q = 1 Kelvin mode becomes overwhelmingly preferred as ttc𝑡subscript𝑡ct\rightarrow t_{\mathrm{c}}italic_t → italic_t start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT. This is supported by the behaviour of kcsubscript𝑘ck_{\mathrm{c}}italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT in Fig. 4 (c) for this regime where we see that only one of the 4444 initial conditions results in the q=2𝑞2q=2italic_q = 2 mode being most strongly excited with the remainder preferentially occupying the q=1𝑞1q=1italic_q = 1 mode. We also find that, regardless of the initial conditions applied, kcsubscript𝑘ck_{\mathrm{c}}italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT is a decreasing function of εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT when 𝐁z^conditional𝐁^𝑧\mathbf{B}\parallel\hat{z}bold_B ∥ over^ start_ARG italic_z end_ARG, such that the vortex loops formed as a consequence of the Crow instability are unambiguously larger for more strongly dipolar superfluids when the dipole polarisation is (anti-)parallel to the initial vorticity of each vortex. The case when 𝐁x^conditional𝐁^𝑥\mathbf{B}\parallel\hat{x}bold_B ∥ over^ start_ARG italic_x end_ARG is a little more ambiguous, though, as the evident drift to lower kcsubscript𝑘ck_{\mathrm{c}}italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT for higher εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT is not obeyed for all of the choices of perturbation initial conditions. We note that each of the initial conditions imply the same mode, q=3𝑞3q=3italic_q = 3 becoming preferential for the largest value of εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT considered in our work, εdd=0.9subscript𝜀dd0.9\varepsilon_{\mathrm{dd}}=0.9italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT = 0.9, but an explanation for the volatility of this quantity with respect to the initial conditions for this particular dipole alignment is not readily apparent.

Having identified the maximally unstable Kelvin mode, kcsubscript𝑘ck_{\mathrm{c}}italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT, in a given dipolar regime, we have also studied the evolution of the corresponding mode amplitude, W(kc,t)𝑊subscript𝑘c𝑡W(k_{\mathrm{c}},t)italic_W ( italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT , italic_t ), from the start of each dGPE simulation till the reconnection induced by the Crow instability. After an early transient arising from nonlinear interactions with the other Kelvin modes, we find that W(kc,t)𝑊subscript𝑘c𝑡W(k_{\mathrm{c}},t)italic_W ( italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT , italic_t ) grows approximately exponentially in time, i.e. W(kc,t)exp(σt)similar-to𝑊subscript𝑘c𝑡𝜎𝑡W(k_{\mathrm{c}},t)\sim\exp(\sigma t)italic_W ( italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT , italic_t ) ∼ roman_exp ( italic_σ italic_t ) with σ𝜎\sigmaitalic_σ the growth rate. This is consistent with a linear dynamical instability of this mode arising from the Crow mechanism, which is evident in Fig. 5 (a) where a plot of log10W(kc,t)subscript10𝑊subscript𝑘c𝑡\log_{10}\,W(k_{\mathrm{c}},t)roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT italic_W ( italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT , italic_t ) is approximately linear as a function of t𝑡titalic_t after sufficient time has passed. Here, εdd=0.9subscript𝜀dd0.9\varepsilon_{\mathrm{dd}}=0.9italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT = 0.9 and 𝐁{x^,y^,z^}conditional𝐁^𝑥^𝑦^𝑧\mathbf{B}\parallel\{\hat{x},\hat{y},\hat{z}\}bold_B ∥ { over^ start_ARG italic_x end_ARG , over^ start_ARG italic_y end_ARG , over^ start_ARG italic_z end_ARG }. Figure 5 (a) also illustrates that the time taken till the first reconnection varies considerably depending on the dipole polarisation, though it is known from prior investigations that the velocities of vortex pairs are themselves dipole-dependent [36].

Refer to caption
Refer to caption
Figure 5: (a) log10W(kc,t)subscript10𝑊subscript𝑘c𝑡\log_{10}\,W(k_{\mathrm{c}},t)roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT italic_W ( italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT , italic_t ) as a function of t𝑡titalic_t, with kcsubscript𝑘ck_{\mathrm{c}}italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT the maximally unstable mode wavenumber and εdd=0.9subscript𝜀dd0.9\varepsilon_{\mathrm{dd}}=0.9italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT = 0.9. The initial conditions correspond to those in Figs. 1 and 2. (b) The ensemble-average of σ𝜎\sigmaitalic_σ, the exponential growth rate of W(kc,t)𝑊subscript𝑘c𝑡W(k_{\mathrm{c}},t)italic_W ( italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT , italic_t ), as a function of εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT. In both plots, 𝐁{x^,y^,z^}conditional𝐁^𝑥^𝑦^𝑧\mathbf{B}\parallel\{\hat{x},\hat{y},\hat{z}\}bold_B ∥ { over^ start_ARG italic_x end_ARG , over^ start_ARG italic_y end_ARG , over^ start_ARG italic_z end_ARG }.

To excise the transient, the exponential growth rate σ𝜎\sigmaitalic_σ has been computed through linear regression fitting of logW(kc,t)𝑊subscript𝑘c𝑡\log\,W(k_{\mathrm{c}},t)roman_log italic_W ( italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT , italic_t ) to t𝑡titalic_t for times after the its growth is exponential, which for the sake of consistency we take to be the interval t[tr/2,tc]𝑡subscript𝑡r2subscript𝑡ct\in[t_{\mathrm{r}}/2,t_{\mathrm{c}}]italic_t ∈ [ italic_t start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT / 2 , italic_t start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT ]. The ensemble average of σ𝜎\sigmaitalic_σ over the set of four Kelvin wave initial conditions is plotted as a function of εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT as distinct curves for 𝐁{x^,y^,z^}conditional𝐁^𝑥^𝑦^𝑧\mathbf{B}\parallel\{\hat{x},\hat{y},\hat{z}\}bold_B ∥ { over^ start_ARG italic_x end_ARG , over^ start_ARG italic_y end_ARG , over^ start_ARG italic_z end_ARG } in Fig. 5 (b). In the nondipolar limit, we can compare our computed value of the growth rate, σ0.0118𝜎0.0118\sigma\approx 0.0118italic_σ ≈ 0.0118, to predictions of σ𝜎\sigmaitalic_σ obtained in the literature through linear stability analysis of the corresponding Bogoliubov-de Gennes (BdG) equations [15, 22]:

σ(kc)2kc2d2[ln(2d)+0.38].similar-to𝜎superscriptsubscript𝑘c2superscriptsubscript𝑘c2superscript𝑑2delimited-[]2𝑑0.38\sigma(k_{\mathrm{c}})^{2}\sim\frac{k_{\mathrm{c}}^{2}}{d^{2}}\left[\ln(\sqrt{% 2}d)+0.38\right].italic_σ ( italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∼ divide start_ARG italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG [ roman_ln ( square-root start_ARG 2 end_ARG italic_d ) + 0.38 ] . (9)

This yields the estimation σ0.04𝜎0.04\sigma\approx 0.04italic_σ ≈ 0.04; while the two predictions are of the same order of magnitude, the discrepancy between the two may be due to the nonlinear growth of W(kc,t)𝑊subscript𝑘c𝑡W(k_{\mathrm{c}},t)italic_W ( italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT , italic_t ) evident in Fig. 5 (a). For finite values of εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT, our results show that σ𝜎\sigmaitalic_σ is smaller when 𝐁x^conditional𝐁^𝑥\mathbf{B}\parallel\hat{x}bold_B ∥ over^ start_ARG italic_x end_ARG than for the other two polarisations regardless of εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT and that as εdd1subscript𝜀dd1\varepsilon_{\mathrm{dd}}\rightarrow 1italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT → 1, a hierarchy such that σ(x^)<σ(z^)<σ(y^)𝜎^𝑥𝜎^𝑧𝜎^𝑦\sigma(\hat{x})<\sigma(\hat{z})<\sigma(\hat{y})italic_σ ( over^ start_ARG italic_x end_ARG ) < italic_σ ( over^ start_ARG italic_z end_ARG ) < italic_σ ( over^ start_ARG italic_y end_ARG ) manifests itself. Now, given that the initial mean separation of the vortices is identical for all of the simulations we have conducted, one would expect that longer reconnection times are consistent with smaller values of σ𝜎\sigmaitalic_σ and the hierarchy of σ𝜎\sigmaitalic_σ in Fig. 5 (b) is consistent with that of the ensemble-averaged reconnection times, tr(x^)>tr(z^)>tr(y^)subscript𝑡r^𝑥subscript𝑡r^𝑧subscript𝑡r^𝑦t_{\mathrm{r}}(\hat{x})>t_{\mathrm{r}}(\hat{z})>t_{\mathrm{r}}(\hat{y})italic_t start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT ( over^ start_ARG italic_x end_ARG ) > italic_t start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT ( over^ start_ARG italic_z end_ARG ) > italic_t start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT ( over^ start_ARG italic_y end_ARG ), seen in Fig. 2 in the regime εdd=0.9subscript𝜀dd0.9\varepsilon_{\mathrm{dd}}=0.9italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT = 0.9.

From our spectral analysis of the vortex lines during their evolution, it is clear that there is a significant dependence on the polarisation of the dipole moments relative to the initial vortex alignment and the DDI interaction strength. The dependence of the maximally unstable mode kcsubscript𝑘ck_{\mathrm{c}}italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT on the characteristics of the DDI, in particular, suggests that local geometric properties of the vortex line are sensitive to these parameters. One such quantity is the curvature of the vortices, which we now proceed to as a function of time for each of the simulations conducted. Parametrising a given vortex line as 𝜸(s)=γi(s)e^i𝜸𝑠subscript𝛾𝑖𝑠subscript^𝑒𝑖\boldsymbol{\gamma}(s)=\gamma_{i}(s)\hat{e}_{i}bold_italic_γ ( italic_s ) = italic_γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_s ) over^ start_ARG italic_e end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, with γz(s)ssubscript𝛾𝑧𝑠𝑠\gamma_{z}(s)\equiv sitalic_γ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ( italic_s ) ≡ italic_s, its curvature is defined through this parametrisation as [67]

κ(s)=|𝜸(s)×𝜸′′(s)||𝜸(s)|3.𝜅𝑠superscript𝜸𝑠superscript𝜸′′𝑠superscriptsuperscript𝜸𝑠3\kappa(s)=\frac{|\boldsymbol{\gamma}^{\prime}(s)\times\boldsymbol{\gamma}^{% \prime\prime}(s)|}{|\boldsymbol{\gamma}^{\prime}(s)|^{3}}.italic_κ ( italic_s ) = divide start_ARG | bold_italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_s ) × bold_italic_γ start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_s ) | end_ARG start_ARG | bold_italic_γ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_s ) | start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG . (10)

In Eq. (10), we compute the derivatives directly from the mode amplitudes {Wn(kz,t)}subscript𝑊𝑛subscript𝑘𝑧𝑡\{W_{n}(k_{z},t)\}{ italic_W start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_t ) } by utilising the standard properties of the discrete Fourier transform and by applying a simple low-pass filter such that only the modes Lz|kz|/(2π)100subscript𝐿𝑧subscript𝑘𝑧2𝜋100L_{z}|k_{z}|/(2\pi)\leq 100italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT | italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT | / ( 2 italic_π ) ≤ 100 contribute to the curvature 333Through trial and error we have found that such a low-pass filter is necessary to avoid aliasing effects, as well as noise from the vortex detection procedure, that adversely affect the computation of higher-order derivatives..

Refer to caption
Figure 6: The mean curvature over the length of the two vortices ensemble-averaged over the set of vortex initial conditions, κdelimited-⟨⟩𝜅\langle\kappa\rangle⟨ italic_κ ⟩, at the time t=tc𝑡subscript𝑡ct=t_{\mathrm{c}}italic_t = italic_t start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT. The magnetic dipole orientation is parallel to x^^𝑥\hat{x}over^ start_ARG italic_x end_ARG in (a), y^^𝑦\hat{y}over^ start_ARG italic_y end_ARG in (b), and z^^𝑧\hat{z}over^ start_ARG italic_z end_ARG in (c).

Let us focus on the vortex profiles when t=tc𝑡subscript𝑡ct=t_{\mathrm{c}}italic_t = italic_t start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT and compare the curvature of the two vortices across the various regimes of εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT and 𝐁𝐁\mathbf{B}bold_B. Finding the mean curvatures of both vortices over their respective line lengths and averaging the two resulting quantities yields the quantity κ(tc)delimited-⟨⟩𝜅subscript𝑡c\langle\kappa(t_{\mathrm{c}})\rangle⟨ italic_κ ( italic_t start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT ) ⟩, which is plotted in Fig. 6 as a function of εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT. Here, 𝐁𝐁\mathbf{B}bold_B is assumed to be parallel to the x𝑥xitalic_x-, y𝑦yitalic_y- and z𝑧zitalic_z-axes in (a), (b) and (c), respectively, and for each choice of {𝐁,εdd}𝐁subscript𝜀dd\{\mathbf{B},\,\varepsilon_{\mathrm{dd}}\}{ bold_B , italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT } we have performed an ensemble average over the 4444 initial vortex perturbation profiles. Figure 6 illustrates clearly the following salient characteristics of the vortex line profiles arising from the Crow instability. While fluctuations in the overall behaviour of the vortex lines occur due to the specific choice of initial conditions, it is evident in Fig. 6 (a) that κ(tc)delimited-⟨⟩𝜅subscript𝑡c\langle\kappa(t_{\mathrm{c}})\rangle⟨ italic_κ ( italic_t start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT ) ⟩ is largely independent of εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT when the dipole moments are polarised along the x𝑥xitalic_x-axis, the translation axis of the vortices, when compared to the y𝑦yitalic_y- and z𝑧zitalic_z-axes. As for dipole polarisations parallel to the y𝑦yitalic_y-axis, Fig. 6 (b) suggests that κ(tc)delimited-⟨⟩𝜅subscript𝑡c\langle\kappa(t_{\mathrm{c}})\rangle⟨ italic_κ ( italic_t start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT ) ⟩ generally increases with εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT and that the values it attains are comparable in magnitude to the corresponding values for polarisations along the x𝑥xitalic_x-axis.

These observations are in sharp contrast to the results presented in Fig. 6 (c), where the mean vortex curvature at t=tc𝑡subscript𝑡ct=t_{\mathrm{c}}italic_t = italic_t start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT is seen to be strongly dependent on εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT when the dipole moments are polarised along the z𝑧zitalic_z-axis, which is (anti-)parallel to the mean vorticity of each axis. As εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT is increased, the curvature of the vortex is suppressed rather than enhanced. This has been attributed in earlier studies of Kelvin waves perturbing single, isolated vortices to the effects of magnetostriction, where there exists an effective interaction between the virtual dipole moments, induced in the vortex line, that are polarised antiparallel to the real dipole moments in the bulk superfluid [51, 68, 52]. When 𝐁z^conditional𝐁^𝑧\mathbf{B}\parallel\hat{z}bold_B ∥ over^ start_ARG italic_z end_ARG, viz. (anti-)parallel to the mean orientation of the vortex line, the effective dipolar interaction energy arising from the mutual interaction of the virtual dipole moments is minimised when the dipole moments are aligned along this mean orientation axis. Thus, a suppression of the curvature of the vortex arises since a larger curvature would effectively cause the configuration of the vortex’s virtual dipole moments to become misaligned with the magnetic field and thus increase the dipolar interaction energy. This also results in the wavenumber of the preferentially excited Kelvin mode, kcsubscript𝑘ck_{\mathrm{c}}italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT, being smaller when 𝐁z^conditional𝐁^𝑧\mathbf{B}\parallel\hat{z}bold_B ∥ over^ start_ARG italic_z end_ARG than for polarisations along the other principal axes as seen in Fig. 4. These effects are naturally more pronounced for larger values of εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT, which leads to smaller curvatures and thus also the lower values of kcsubscript𝑘ck_{\mathrm{c}}italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT seen in Figs. 4 (c).

We also note that magnetostriction arising from the virtual dipole moments in the vortex cores serves to explain the salient features of the vortices when 𝐁x^,y^conditional𝐁^𝑥^𝑦\mathbf{B}\parallel\hat{x},\,\hat{y}bold_B ∥ over^ start_ARG italic_x end_ARG , over^ start_ARG italic_y end_ARG as well. When the dipole polarisation are polarised along the y𝑦yitalic_y-axis, the energetic preference of the vortices’ virtual dipole moments is to align along this axis and such vortex curvature is stimulated along this axis. While this results in the monotonic behaviour of κcdelimited-⟨⟩subscript𝜅c\langle\kappa_{\mathrm{c}}\rangle⟨ italic_κ start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT ⟩ in Fig. 6 (b), it is also responsible for the reconnections occurring on a faster timescale at fixed εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT than for the other polarisation orientations, as observed in Fig. 2 and, as we have argued, also explains why the exponential mode growth factor σ𝜎\sigmaitalic_σ is larger when 𝐁y^conditional𝐁^𝑦\mathbf{B}\parallel\hat{y}bold_B ∥ over^ start_ARG italic_y end_ARG than the other polarisations. Similarly, when the dipole polarisation is parallel to the x𝑥xitalic_x-axis the contributions to the overall curvature arising from curvature orthogonal to this axis are suppressed, such that the majority of the curvature arises from curvature along the propagation axis of the vortex pair. While this results in the strongly transversely perturbed vortex profiles in the x𝑥xitalic_x-z𝑧zitalic_z plane as is seen in Fig. 2 (a), curvature along the y𝑦yitalic_y-axis is greatly inhibited and the reconnection is suppressed. Thus, the exponential growth of the dominant Kelvin mode is suppressed by the DDI when 𝐁x^conditional𝐁^𝑥\mathbf{B}\parallel\hat{x}bold_B ∥ over^ start_ARG italic_x end_ARG, resulting in smaller values of the mode growth factor σ𝜎\sigmaitalic_σ and correspondingly longer reconnection times.

V Conclusion

Through studying the evolution of a pair of quantum vortices in a dipolar superfluid with an initial transverse, symmetric perturbation applied to each vortex, we have found that the quantum analogue of the Crow instability behaves in vastly different ways depending on the magnetic dipole polarisation and the relative strength of the DDI. When the dipole polarisation, 𝐁𝐁\mathbf{B}bold_B, is (anti-)parallel to the initial mean vorticity of each vortex, the vortices’ curvature is strongly suppressed due to the influence of the interaction between virtual dipole moments inside the vortex core. This results in Kelvin waves of decreasing wavenumber contributing most strongly to the Crow instability as the relative dipolar interaction strength, εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT, increases. This magnetostrictive effect has the converse effect when the dipole polarisation is orthogonal to the initial vorticities, with the curvature of each vortex as well as the growth of Kelvin modes of higher wavenumber being enhanced by the influence of the DDI. While the Kelvin mode occupation spectrum for a magnetic dipole polarisation parallel to the separation between the two vortices appears rather similar to its nondipolar analogue, these modes induce a greater curvature in the vortices as εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT increases and, correspondingly, a Crow instability that develops more rapidly. However, for dipole polarisations parallel to the binormal axis, the Kelvin modes become concentrated at one specific higher wavenumber mode, p=3𝑝3p=3italic_p = 3, in the large-εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT limit and the magnetostrictive elongation of the vortex lines along the binormal axis suppresses the Crow instability when compared to the other two dipole polarisations considered here.

Our studies of the dependence of the Crow instability on the orientation of the dipole moments suggest that other hydrodynamical instabilities might exhibit directional dependence in the presence of a DDI. Notably, the snake instability of dark solitons in three-dimensional dipolar condensates has been shown to be suppressed when the dipole alignment is parallel to the solitonic phase slip [69]. The corresponding phase slip due to the vortex line pair in our system is parallel to the x𝑥xitalic_x-axis and our results demonstrate a suppression of the Crow instability for dipole polarisations parallel to this axis. Given the intrinsic correspondence of the snake instability of exhibited by dark solitons and the Crow instability of antiparallel vortex lines [15], this is likely not a coincidence. This suppression of the snake instability was demonstrated through linear stability analysis applied to the BdG spectra of a three-dimensional dark soliton [69]; a similar procedure was conducted in Ref. [22] to characterise the linear stability of antiparallel vortex lines in a nondipolar BEC against the transverse perturbations that induce the Crow instability. An application of linear stability analysis to the corresponding dipolar problem as studied in this article may thus prove to be illuminating. We also note that Ref. [69] studied only dipole orientations parallel to the phase slip for dark solitons in three-dimensional regimes, and that our analysis of the Crow instability for distinct dipole orientations suggest that it is prudent to also examine the snake instability for dipole orientations in the soliton plane.

We also note that our results bear implications for aspects of quantum vortex dynamics in dipolar superfluids that, to the best of our knowledge, have not yet been investigated thoroughly. For instance, in Sec. IV we have posited a relationship between the exponential growth factor of the dominant Kelvin mode, σ𝜎\sigmaitalic_σ, and the time taken for the vortices to undergo their first reconnection, trsubscript𝑡rt_{\mathrm{r}}italic_t start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT. While the values of σ𝜎\sigmaitalic_σ, ensemble-averaged over the set of initial conditions for the vortex perturbation profiles, exhibited a clear dependence on 𝐁𝐁\mathbf{B}bold_B and εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT, it would be desirable to investigate such scaling behaviour in a setting that is independent of the randomness of the initial conditions present in our system. In a dipolar superfluid, it would be of considerable interest to elucidate the relationship between the parameters of the DDI and the temporal scaling of quantities such as the inter-vortex separation, the angle between the normal axes of the vortex lines, and global properties such as the vortices’ curvature and torsion, and establish whether or not they exhibit deviations from the familiar nondipolar paradigm [67, 70]. Furthermore,in the nondipolar limit a considerable degree of insight may be gleaned into the dynamics of quantum vortex lines by assuming that the superfluid flow is purely compressible, which allows the use of the Biot-Savart law and the local induction approximation to model vortices before reconnective processes occur [71, 72]. In this article, a strong influence has been demonstrated of the magnetic dipole polarisation on the bending of vortex line pairs, while the dynamics of quasi-two-dimensional point-vortices and three-dimensional straight vortex lines have been shown to exhibit dipole-mediated deviations from the predictions of the point-vortex model, the two-dimensional limit of the Biot-Savart law. We also note that the Kelvin wave dispersion relation for a single vortex line is modified in the presence of the DDI [51, 52]. Given these pre-existing and new examples, an extension of the Biot-Savart law [71, 72] to account for the dipolar interaction between the virtual dipole moments in the vortex line cores would provide an intuitive understanding of how these disparate results can be generalised for a larger number of dipolar vortices. Furthermore, we also note that while this investigation assumed a value of εddsubscript𝜀dd\varepsilon_{\mathrm{dd}}italic_ε start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT less than 1111, such that the dBEC does not exhibit the supersolid phase [43, 44, 45], the interaction of supersolid quantum droplets and vortices produces complicated dynamics that are inaccessible in the superfluid phase. It is feasible to suggest that this might result in Kelvin waves and subsequently the Crow instability in pairs of antiparallel vortices [73].

Acknowledgements.
This work was funded by Grant No. RPG-2021-108 from the Leverhulme Trust. S. B. P. thanks Luca Galantucci for providing the pseudovorticity-based vortex detection program used herein to extract the vortex line profiles. This research made use of the Rocket High Performance Computing service at Newcastle University.

References

  • Crow [1970] S. C. Crow, Stability Theory for a Pair of Trailing Vortices, AIAA Journal 8, 2172 (1970).
  • Pitaevskii and Stringari [2016] L. Pitaevskii and S. Stringari, Bose-Einstein Condensation and Superfluidity, 1st ed., International Series of Monographs on Physics No. 164 (Clarendon Press, 2016).
  • Fetter [2009] A. L. Fetter, Rotating Trapped Bose-Einstein Condensates, Reviews of Modern Physics 81, 647 (2009).
  • Koplik and Levine [1993] J. Koplik and H. Levine, Vortex Reconnection in Superfluid Helium, Physical Review Letters 71, 1375 (1993).
  • Zuccher et al. [2012] S. Zuccher, M. Caliari, A. W. Baggaley, and C. F. Barenghi, Quantum Vortex Reconnections, Physics of Fluids 24, 125108 (2012).
  • Ho [2001] T.-L. Ho, Bose-Einstein Condensates with Large Number of Vortices, Physical Review Letters 87, 060403 (2001).
  • Cooper et al. [2001] N. R. Cooper, N. K. Wilkin, and J. M. F. Gunn, Quantum Phases of Vortices in Rotating Bose-Einstein Condensates, Physical Review Letters 87, 120405 (2001).
  • Baym [2003] G. Baym, Tkachenko Modes of Vortex Lattices in Rapidly Rotating Bose-Einstein Condensates, Physical Review Letters 91, 110402 (2003).
  • Watanabe et al. [2004] G. Watanabe, G. Baym, and C. J. Pethick, Landau Levels and the Thomas-Fermi Structure of Rapidly Rotating Bose-Einstein Condensates, Physical Review Letters 93, 190401 (2004).
  • Henn et al. [2009] E. A. L. Henn, J. A. Seman, G. Roati, K. M. F. Magalhães, and V. S. Bagnato, Emergence of Turbulence in an Oscillating Bose-Einstein Condensate, Physical Review Letters 103, 045301 (2009).
  • Tsatos et al. [2016] M. C. Tsatos, P. E. S. Tavares, A. Cidrim, A. R. Fritsch, M. A. Caracanhas, F. E. A. dos Santos, C. F. Barenghi, and V. S. Bagnato, Quantum Turbulence in Trapped Atomic Bose-Einstein Condensates, Physics Reports 622, 1 (2016).
  • Navon et al. [2016] N. Navon, A. L. Gaunt, R. P. Smith, and Z. Hadzibabic, Emergence of a Turbulent Cascade in a Quantum Gas, Nature 539, 72 (2016).
  • Madeira et al. [2020] L. Madeira, M. A. Caracanhas, F. E. A. dos Santos, and V. S. Bagnato, Quantum Turbulence in Quantum Gases, Annual Review of Condensed Matter Physics 11, 37 (2020).
  • Barenghi et al. [2023] C. F. Barenghi, H. A. J. Middleton-Spencer, L. Galantucci, and N. G. Parker, Types of Quantum Turbulence, AVS Quantum Science 5, 025601 (2023).
  • Kuznetsov and Rasmussen [1995] E. A. Kuznetsov and J. J. Rasmussen, Instability of Two-Dimensional Solitons and Vortices in Defocusing Media, Physical Review E 51, 4479 (1995).
  • Baggaley and Parker [2018] A. W. Baggaley and N. G. Parker, Kelvin-Helmholtz Instability in a Single-Component Atomic Superfluid, Physical Review A 97, 053608 (2018).
  • Giacomelli and Carusotto [2023] L. Giacomelli and I. Carusotto, Interplay of Kelvin-Helmholtz and Superradiant Instabilities of an Array of Quantized Vortices in a Two-Dimensional Bose-Einstein Condensate, Physical Review A 14, 025 (2023).
  • Hernández-Rajkov et al. [2024] D. Hernández-Rajkov, N. Grani, F. Scazza, G. Del Pace, W. J. Kwon, M. Inguscio, K. Xhani, C. Fort, M. Modugno, F. Marino, and G. Roati, Connecting Shear Flow and Vortex Array Instabilities in Annular Atomic Superfluids, Nature Physics 20, 939 (2024).
  • Sasaki et al. [2009] K. Sasaki, N. Suzuki, D. Akamatsu, and H. Saito, Rayleigh-Taylor Instability and Mushroom-Pattern Formation in a Two-Component Bose-Einstein Condensate, Physical Review A 80, 063611 (2009).
  • Gautam and Angom [2010] S. Gautam and D. Angom, Rayleigh-Taylor Instability in Binary Condensates, Physical Review A 81, 053601 (2010).
  • Bezett et al. [2010] A. Bezett, V. Bychkov, E. Lundh, D. Kobyakov, and M. Marklundr, Magnetic Richtmyer-Meshkov Instability in a Two-Component Bose-Einstein Condensate, Physical Review A 82, 043608 (2010).
  • Berloff and Roberts [2001] N. G. Berloff and P. H. Roberts, Motion in a Bose Condensate: IX. Crow Instability of Antiparallel Vortex Pairs, Journal of Physics A: Mathematical and Theoretical 34, 10057 (2001).
  • Simula [2011] T. P. Simula, Crow Instability in Trapped Bose-Einstein Condensates, Physical Review A 84, 021603 (2011).
  • Gautam [2013] S. Gautam, Crow Instability in Unitary Fermi Gas, Modern Physics Letters B 27, 1350097 (2013).
  • Villois et al. [2017] A. Villois, D. Proment, and G. Krstulovic, Universal and Nonuniversal Aspects of Vortex Reconnections in Superfluids, Physical Review Fluids 2, 044701 (2017).
  • Lahaye et al. [2009] T. Lahaye, C. Menotti, L. Santos, M. Lewenstein, and T. Pfau, The Physics of Dipolar Bosonic Quantum Gases, Reports on Progress in Physics 72, 126401 (2009).
  • Martin et al. [2017] A. M. Martin, N. G. Marchant, D. H. J. O’Dell, and N. G. Parker, Vortices and Vortex Lattices in Quantum Ferrofluids, Journal of Physics: Condensed Matter 29, 103004 (2017).
  • Chomaz et al. [2023] L. Chomaz, I. Ferrier-Barbut, F. Ferlaino, B. Laburthe-Tolra, B. L. Lev, and T. Pfau, Dipolar Physics: A Review of Experiments with Magnetic Quantum Gases, Reports on Progress in Physics 86, 026401 (2023).
  • Eberlein et al. [2005] C. Eberlein, S. Giovanazzi, and D. H. J. O’Dell, Exact Solution of the Thomas-Fermi Equation for a Trapped Bose-Einstein Condensate with Dipole-Dipole Interactions, Physical Review A 71, 033618 (2005).
  • Stuhler et al. [2005] J. Stuhler, A. Griesmaier, T. Koch, M. Fattori, T. Pfau, S. Giovanazzi, P. Pedri, and L. Santos, Observation of Dipole-Dipole Interaction in a Degenerate Quantum Gas, Physical Review Letters 95, 150406 (2005).
  • Ticknor et al. [2011] C. Ticknor, R. M. Wilson, and J. L. Bohn, Anisotropic Superfluidity in a Dipolar Bose Gas, Physical Review Letters 106, 065301 (2011).
  • Wenzel et al. [2018] M. Wenzel, F. Böttcher, J.-N. Schmidt, M. Eisenmann, T. Langen, T. Pfau, and I. Ferrier-Barbut, Anisotropic Superfluid Behaviour of a Dipolar Bose-Einstein Condensate, Physical Review Letters 121, 030401 (2018).
  • Abad et al. [2009] M. Abad, M. Guilleumas, R. Mayol, M. Pi, and D. M. Jezek, Vortices in Bose-Einstein Condensates with Dominant Dipole-Dipole Interactions, Physical Review A 79, 063622 (2009).
  • Mulkerin et al. [2013] B. C. Mulkerin, R. M. W. van Bijnen, D. H. J. O’Dell, A. M. Martin, and N. G. Parker, Anisotropic and Long-Range Vortex Interactions in Two-Dimensional Dipolar Bose Gases, Physical Review Letters 111, 170402 (2013).
  • Bland et al. [2023] T. Bland, G. Lamporesi, M. J. Mark, and F. Ferlaino, Vortices in Dipolar Bose-Einstein Condensates, Comptes Rendus Physique 24, 1 (2023).
  • Prasad et al. [2024] S. B. Prasad, N. G. Parker, and A. W. Baggaley, Vortex-Pair Dynamics in Three-Dimensional Homogeneous Dipolar Superfluids, Physical Review A 109, 063323 (2024).
  • Cooper et al. [2005] N. R. Cooper, E. H. Rezayi, and S. H. Simon, Vortex Lattices in Rotating Atomic Bose Gases with Dipolar Interactions, Physical Review Letters 95, 200402 (2005).
  • Zhang and Zhai [2005] J. Zhang and H. Zhai, Vortex Lattices in Planar Bose-Einstein Condensates with Dipolar Interactions, Physical Review Letters 95, 200403 (2005).
  • Cai et al. [2018] Y. Cai, Y. Yuan, M. Rosenkranz, H. Pu, and W. Bao, Vortex Patterns and the Critical Rotation Frequency in Rotating Dipolar Bose-Einstein Condensates, Physical Review A 98, 023610 (2018).
  • Bland et al. [2018] T. Bland, G. W. Stagg, L. Galantucci, A. W. Baggaley, and N. G. Parker, Quantum Ferrofluid Turbulence, Physical Review Letters 121, 174501 (2018).
  • yan M.Wilson et al. [2012] yan M.Wilson, C. Ticknor, J. L. Bohn, and E. Timmermans, Roton Immiscibility in a Two-Component Dipolar Bose Gas, Physical Review A 86, 033606 (2012).
  • Hertkorn et al. [2021] J. Hertkorn, J.-N. Schmidt, M. Guo, F. Böttcher, K. S. H. Ng, S. D. Graham, P. Uerlings, T. Langen, M. Zwierlein, and T. Pfau, Pattern Formation in Quantum Ferrofluids: From Supersolids to Superglasses, Physical Review Research 3, 033125 (2021).
  • Böttcher et al. [2019] F. Böttcher, J.-N. Schmidt, M. Wenzel, J. Hertkorn, M. Guo, T. Langen, and T. Pfau, Transient Supersolid Properties in an Array of Dipolar Quantum Droplets, Physical Review X 9, 011051 (2019).
  • Chomaz et al. [2019] L. Chomaz, D. Petter, P. Ilzhöfer, G. Natale, A. Trautmann, C. Politi, G. Durastante, R. M. W. van Bijnen, A. Patscheider, M. Sohmen, M. J. Mark, and F. Ferlaino, Long-Lived and Transient Supersolid Behaviors in Dipolar Quantum Gases, Physical Review X 9, 021012 (2019).
  • Tanzi et al. [2019] L. Tanzi, E. Lucioni, F. Famà, J. Catani, A. Fioretti, C. Gabbanini, R. N. Bisset, L. Santos, and G. Modugno, Observation of a Dipolar Quantum Gas with Metastable Supersolid Properties, Physical Review Letters 122, 130405 (2019).
  • Klaus et al. [2022] L. Klaus, T. Bland, E. Poli, C. Politi, G. Lamporesi, E. Casotti, R. N. Bisset, M. J. Mark, and F. Ferlaino, Observation of Vortices and Vortex Stripes in a Dipolar Condensate, Nature Physics 18, 1453 (2022).
  • Casotti et al. [2024] E. Casotti, E. Poli, L. Klaus, A. Litvinov, C. Ulm, C. Politi, M. J. Mark, T. Bland, and F. Ferlaino, Observation of vortices in a dipolar supersolid,   (2024), arXiv:2403.18510 [cond-mat.quant-gas] .
  • Gautam [2014] S. Gautam, Dynamics of the Corotating Vortices in Dipolar Bose–Einstein Condensates in the Presence of Dissipation, Journal of Physics B: Atomic, Molecular and Optical Physics 47, 165301 (2014).
  • Zhao [2021] Q. Zhao, Effects of Dipole–Dipole Interaction on Vortex Motion in Bose–Einstein Condensates, Journal of Low Temperature Physics 204, 1 (2021).
  • Sabari et al. [2024] S. Sabari, R. K. Kumar, and L. Tomio, Vortex Dynamics and Turbulence in Dipolar Bose-Einstein Condensates, Physical Review A 109, 023313 (2024).
  • Klawunn et al. [2008] M. Klawunn, R. Nath, P. Pedri, and L. Santos, Transverse Instability of Straight Vortex Lines in Dipolar Bose-Einstein Condensates, Physical Review Letters 100, 240403 (2008).
  • Lee et al. [2018] A.-C. Lee, D. Baillie, R. N. Bisset, and P. B. Blakie, Excitations of a Vortex Line in an Elongated Dipolar Condensate, Physical Review A 98, 063620 (2018).
  • Schützhold et al. [2006] R. Schützhold, M. Uhlmann, Y. Xu, and U. R. Fischer, Mean-Field Expansion in Bose–Einstein Condensates with Finite-Range Interactions, International Journal of Modern Physics B 20, 3555 (2006).
  • Lima and Pelster [2012] A. R. P. Lima and A. Pelster, Beyond Mean-Field Low-Lying Excitations of Dipolar Gases, Physical Review A 86, 063609 (2012).
  • Bisset et al. [2016] R. N. Bisset, R. M. Wilson, D. Baillie, and P. B. Blakie, Ground-State Phase Diagram of a Dipolar Condensate with Quantum Fluctuations, Physical Review A 94, 033619 (2016).
  • Schmitt et al. [2016] M. Schmitt, M. Wenzel, F. Böttcher, I. Ferrier-Barbut, and T. Pfau, Self-Bound Droplets of a Dilute Magnetic Quantum Liquid, Nature 539, 259 (2016).
  • Chomaz et al. [2016] L. Chomaz, S. Baier, D. Petter, M. J. Mark, F. Wächtler, L. Santos, and F. Ferlaino, Quantum-Fluctuation-Driven Crossover from a Dilute Bose-Einstein Condensate to a Macrodroplet in a Dipolar Quantum Fluid, Physical Review X 6, 041039 (2016).
  • Barenghi and Parker [2016] C. F. Barenghi and N. G. Parker, A Primer on Quantum Fluids, Springer Briefs in Physics (Springer, 2016).
  • Bao and Cai [2013] W. Bao and Y. Cai, Mathematical Theory and Numerical Methods for Bose-Einstein Condensation, Kinetic and Related Models 6, 1 (2013).
  • Villois et al. [2016] A. Villois, G. Krstulovic, D. Proment, and H. Salman, A Vortex Filament Tracking Method for the Gross–Pitaevskii Model of a Superfluid, Journal of Physics A: Mathematical and Theoretical 49, 415502 (2016).
  • Serafini et al. [2017] S. Serafini, L. Galantucci, E. Iseni, T. Bienaimé, R. N. Bisset, C. F. Barenghi, F. Dalfovo, G. Lamporesi, and G. Ferrari, Vortex Reconnections and Rebounds in Trapped Atomic Bose-Einstein Condensates, Physical Review X 7, 021031 (2017).
  • Hossain et al. [2022] K. Hossain, K. Kobuszewski, M. M. Forbes, P. Magierski, K. Sekizawa, and G. Wlazłowski, Rotating Quantum Turbulence in the Unitary Fermi Gas, Physical Review A 105, 013304 (2022).
  • Middleton-Spencer et al. [2023] H. A. J. Middleton-Spencer, A. D. G. Orozco, L. Galantucci, M. Moreno, N. G. Parker, L. A. Machado, V. S. Bagnato, and C. F. Barenghi, Strong Quantum Turbulence in Bose-Einstein Condensates, Physical Review Research 5, 043081 (2023).
  • Weiss and McWilliams [1991] J. B. Weiss and J. C. McWilliams, Nonergodicity of Point Vortices, Physics of Fluids A: Fluid Dynamics 3, 835 (1991).
  • Billam et al. [2014] T. P. Billam, M. T. Reeves, B. P. Anderson, and A. S. Bradley, Onsager-Kraichnan Condensation in Decaying Two-Dimensional Quantum Turbulence, Physical Review Letters 112, 145301 (2014).
  • Griffin et al. [2020] A. Griffin, V. Shukla, M.-E. Brachet, and S. Nazarenko, Magnus-Force model for Active Particles Trapped on Superfluid Vortices, Physical Review A 101, 053601 (2020).
  • Rorai et al. [2016] C. Rorai, J. Skipper, R. M. Kerr, and K. R. Sreenivasan, Approach and Separation of Quantised Vortices with Balanced Cores, Journal of Fluid Mechanics 808, 641 (2016).
  • Klawunn and Santos [2009] M. Klawunn and L. Santos, Phase Transition from Straight into Twisted Vortex Lines in Dipolar Bose–Einstein Condensates, New Journal of Physics 11, 055612 (2009).
  • Nath et al. [2008] R. Nath, P. Pedri, and L. Santos, Stability of Dark Solitons in Three Dimensional Dipolar Bose-Einstein Condensates, Physical Review Letters 101, 210402 (2008).
  • Galantucci et al. [2019] L. Galantucci, A. W. Baggaley, N. G. Parker, and C. F. Barenghi, Crossover from Interaction to Driven Regimes in Quantum Vortex Reconnections, Proceedings of the National Academy of Sciences 116, 12204 (2019).
  • Schwarz [1985] K. W. Schwarz, Three-Dimensional Vortex Dynamics in Superfluid 4He: Line-Line and Line-Boundary Interactions, Physical Review B 31, 5782 (1985).
  • Schwarz [1988] K. W. Schwarz, Three-Dimensional Vortex Dynamics in Superfluid 4He: Homogeneous Superfluid Turbulence, Physical Review B 38, 2398 (1988).
  • Poli et al. [2023] E. Poli, T. Bland, S. J. M. White, M. J. Mark, F. Ferlaino, S. Trabucco, and M. Mannarelli, Glitches in Rotating Supersolids, Physical Review Letters 131, 223401 (2023).