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

Properties of a nematic spin vortex in an antiferromagnetic spin-1 Bose-Einstein condensate

Andrew P. C. Underwood Dodd-Walls Centre for Photonic and Quantum Technologies, New Zealand Department of Physics, University of Otago, Dunedin 9016, New Zealand    D. Baillie Dodd-Walls Centre for Photonic and Quantum Technologies, New Zealand Department of Physics, University of Otago, Dunedin 9016, New Zealand    P. Blair Blakie Dodd-Walls Centre for Photonic and Quantum Technologies, New Zealand Department of Physics, University of Otago, Dunedin 9016, New Zealand    H. Takeuchi Department of Physics and Nambu Yoichiro Institute of Theoretical and Experimental Physics (NITEP), Osaka City University, Osaka 558-8585, Japan
Abstract

A spin-1 condensate with antiferromagnetic interactions supports nematic spin vortices in the easy-plane polar phase. These vortices have a 2π2\pi winding of the nematic director, with a core structure that depends on the quadratic Zeeman energy. We characterize the properties of the nematic spin vortex in a uniform quasi-two-dimensional system. We also obtain the vortex excitation spectrum and use it to quantify its stability against dissociating into two half-quantum vortices, finding a parameter regime where the nematic spin vortex is dynamically stable. These results are supported by full dynamical simulations.

I Introduction

Spinor Bose-Einstein condensates are superfluid quantum gases with spin degrees of freedom. These can exist in various spin-ordered phases depending on the nature of the inter-particle interactions and quadratic Zeeman shift Ho (1998); Ohmi and Machida (1998); Stamper-Kurn et al. (1998); Kawaguchi and Ueda (2012); Stamper-Kurn and Ueda (2013). Quantized vortices are regarded a hallmark of superfluidity and often play a unifying role in nonequilibrium dynamics Bray (1994); Sadler et al. (2006); Kudo and Kawaguchi (2015); Williamson and Blakie (2016a). The rich order parameter symmetries of spinor condensates give rise to an array of different spin vortices. Emerging experimental capabilities to produce and monitor the dynamics of spin vortices Seo et al. (2015, 2016); Chen et al. (2018) motivate the need for a better understanding of their properties and interactions (e.g. see Turner (2009); Eto et al. (2011); Williamson and Blakie (2016b); Kasamatsu et al. (2016)).

In this paper we consider a spin-1 antiferromagnetic condensate with polar (nematic spin) order which can be characterized by a director, i.e. a preferred unoriented axis in spin-space Zibold et al. (2016); Symes and Blakie (2017). When the quadratic Zeeman energy is negative, the condensate is in the easy-plane polar (EPP) phase where the director lies in the plane transverse to the magnetic field. In this case, the director breaks the continuous rotational symmetry and the order-parameter manifold supports various spin vortices as topological defects (e.g. see Isoshima (2002); Lovegrove et al. (2016, 2016)). A significant amount of attention has been given to half-quantum vortices (HQVs) Leonhardt and Volovik (2000), which consist of mass and nematic spin current circulation and have recently been prepared in experiments Seo et al. (2015, 2016). The role of HQVs in post-quench dynamics Kang et al. (2017); Symes and Blakie (2017); Kang et al. (2019); Symes et al. (2018) and the interactions between HQVs Eto et al. (2011); Kasamatsu et al. (2016); Seo et al. (2016) have been studied. Here we focus on a second type of vortex in the EPP phase: a pure spin-vortex (i.e. without mass current), which we refer to as the nematic spin vortex (NSV) (also see Lovegrove et al. (2014, 2016); Choi et al. (2012a, b)). These NSVs can decay by splitting into a pair of HQVs, and thus an important consideration is their stability. We note previous work has considered NSVs in a harmonically trapped system, and explored their energetic stability under external rotation and varying magnetization Lovegrove et al. (2014, 2016).

In this paper we develop theory for a NSV in an infinite uniform quasi-two-dimensional (quasi-2D) EPP condensate. This allows us to describe the core structure and excitation spectrum using two parameters: the quadratic Zeeman energy scaled by the chemical potential, and the ratio of the spin-dependent to spin-independent interactions. We determine a critical value qcq_{c} of the quadratic Zeeman energy where the NSV undergoes a continuous transition from having a normal (unfilled) core to having a core filled by an easy-axis polar (EAP) component. We quantify the dissociation instability of the NSV by solving the Bogoliubov-de Gennes (BdG) equations, and by performing dynamical simulations. Importantly we find that at small negative values of the quadratic Zeeman energy, the NSV is dynamically stable.

The structure of the paper is as follows. In Sec. II we introduce the background theory for the spin-1 system and overview the vortices in the EPP phase. In Sec. III we specialize the theory to NSV stationary states in a quasi-2D system, and present numerical results for the vortex properties. In Sec. IV we formulate the BdG equations for the NSV and present a phase diagram characterizing the strength of dynamic instabilities. Using dynamical simulations of the spin-1 system in a flat-bottomed trap we verify the splitting instability emerging from the dynamic instability in Sec. V. We conclude in Sec. VI.

II Background theory

II.1 General formalism for spin-1 BECs

A spin-1 condensate is described by the spinor field

𝚿[Ψ1,Ψ0,Ψ1]T,\displaystyle\bm{\Psi}\equiv[\Psi_{1},\Psi_{0},\Psi_{-1}]^{\mathrm{T}}, (1)

with the three components representing the condensate amplitude in the spin levels m=1,0,1m=1,0,-1, respectively, where mm is the quantum number associated with the zz-component of spin. In weak fields the short-ranged contact interactions between atoms are rotationally invariant with a Hamiltonian density

int=c02n2+c12|F|2.\displaystyle\mathcal{H}_{\mathrm{int}}=\frac{c_{0}}{2}n^{2}+\frac{c_{1}}{2}|\vec{F}|^{2}. (2)

Here the first term, with coupling constant c0c_{0}, describes the density dependent interactions, where n𝚿𝚿n\equiv\bm{\Psi}^{\dagger}\bm{\Psi} is the total density. The second term describes the spin-dependent interactions, where c1c_{1} is the spin-dependent coupling constant, F𝚿f𝚿\vec{F}\equiv\bm{\Psi}^{\dagger}\vec{f}\bm{\Psi} is the spin density, and f(fˇx,fˇy,fˇz)\vec{f}\equiv(\check{f}_{x},\check{f}_{y},\check{f}_{z}) are the spin-1 matrices. In addition, we consider the presence of a (uniform) quadratic Zeeman shift. Taking the field to be along zz this is described by

QZ=q𝚿fˇz2𝚿=q(|Ψ1|2+|Ψ1|2).\displaystyle\mathcal{H}_{\mathrm{QZ}}=q\bm{\Psi}^{\dagger}\check{f}_{z}^{2}\bm{\Psi}=q\left(|\Psi_{1}|^{2}+|\Psi_{-1}|^{2}\right). (3)

In practice the coefficient qq is readily changed in experiments using microwave dressing (e.g. see Gerbier et al. (2006); Leslie et al. (2009)). We note that the uniform linear Zeeman term can be removed using a gauge rotation, and can be neglected. The spin properties also depend on the (conserved) zz-magnetization Mz=dVFzM_{z}=\int\differential{V}F_{z} of the system. Here we consider only Mz=0M_{z}=0.

The case of antiferromagnetic interactions, where c1>0c_{1}>0, is realized with 23Na atoms in their lowest hyperfine manifold. Here the condensate prefers to minimize the spin-density to reduce the spin-dependent interaction energy. For a condensate of uniform density nbn_{b} and spin-density F=0\vec{F}=\vec{0} the spinor is in a polar state

𝚿P=[Ψ1Ψ0Ψ1]=nbeiθ[dx+idy2dzdx+idy2],\displaystyle\bm{\Psi}_{\text{P}}=\left[\begin{array}[]{c}\Psi_{1}\\ \Psi_{0}\\ \Psi_{-1}\end{array}\right]=\sqrt{n_{b}}e^{i\theta}\left[\begin{array}[]{c}\frac{-d_{x}+id_{y}}{\sqrt{2}}\\ d_{z}\\ \frac{d_{x}+id_{y}}{\sqrt{2}}\end{array}\right], (10)

where the real unit vector d=(dx,dy,dz)\vec{d}=(d_{x},d_{y},d_{z}) is the nematic director and θ\theta is the global phase. Noting that 𝚿P\bm{\Psi}_{\text{P}} is invariant under θθ+π\theta\to\theta+\pi and dd\vec{d}\to-\vec{d}, we see that d\vec{d} defines a preferred axis in spin space, but not a preferred direction along that axis. The ground state orientation of d\vec{d} is determined by the quadratic Zeeman energy, which is given by QZ=qnb(1dz2)\mathcal{H}_{\mathrm{QZ}}=qn_{b}(1-d_{z}^{2}). Thus for q>0q>0 the system maximizes dz2d_{z}^{2}, by being in the EAP phase, i.e. 𝚿EAP=nbeiθ(0,1,0)T\bm{\Psi}_{\text{EAP}}=\sqrt{n_{b}}e^{i\theta}(0,1,0)^{T}. The case of interest in this paper is the EPP phase for q<0q<0 where d=(dx,dy,0)\vec{d}=(d_{x},d_{y},0). In addition to the global phase, the EPP ground state also breaks a U(1) symmetry in spin space. This can be seen from the director, which can be written as d=(cosφ,sinφ,0)\vec{d}=(\cos\varphi,\sin\varphi,0), i.e. 𝚿EPP=nb2eiθ(eiφ,0,eiφ)T\bm{\Psi}_{\text{EPP}}=\sqrt{\frac{n_{b}}{2}}e^{i\theta}(-e^{-i\varphi},0,e^{i\varphi})^{T}, where φ\varphi is the angle the director takes with respect to the xx-axis.

Note that in the EPP phase we have Ψ0=0\Psi_{0}=0 and the system is effectively a two-component condensate. Indeed, several studies of the relevant EPP vortices we present in the next subsection have been performed in a two-component (or binary) condensate. For completeness we briefly mention the mapping of the spinor parameters onto an equivalent binary system. To do this we note that interaction Hamiltonian density int\mathcal{H}_{\mathrm{int}} may be expressed in the binary form

int=12i,j=1,1gij|Ψi|2|Ψj|2,\displaystyle\mathcal{H}_{\mathrm{int}}=\frac{1}{2}\sum_{i,j=-1,1}g_{ij}|\Psi_{i}|^{2}|\Psi_{j}|^{2}, (11)

with intra-species coupling constant gii=c0+c1g_{ii}=c_{0}+c_{1} (identical for both components) and interspecies coupling constant g1,1=c0c1g_{1,-1}=c_{0}-c_{1}. For the antiferromagnetic case g1,1<giig_{1,-1}<g_{ii} and the components are miscible Mineev (1974).

II.2 EPP phase vortex classification by winding numbers

Here we consider a quasi-2D spinor gas where the order parameter manifold of the EPP phase permits vortices as point defects. To give the basic structure of such vortex states we write the wave function on the xyxy-plane with the vortex core taken at the origin. Sufficiently far from the core the general vortex state is of the form

𝚿V=nb2eiσMϕ[eiσSϕ0eiσSϕ]\displaystyle\bm{\Psi}_{\text{V}}=\sqrt{\frac{n_{b}}{2}}e^{i\sigma_{\text{M}}\phi}\left[\begin{array}[]{c}-e^{-i\sigma_{\text{S}}\phi}\\ 0\\ e^{i\sigma_{\text{S}}\phi}\end{array}\right] (15)

where ϕ\phi is the azimuthal angle in the xyxy-plane [i.e. from 𝚿EPP\bm{\Psi}_{\text{EPP}}, taking θ\theta and φ\varphi to vary as σMϕ\sigma_{\text{M}}\phi and σSϕ\sigma_{\text{S}}\phi in space, respectively]. Here σM\sigma_{\text{M}} and σS\sigma_{\text{S}} are the winding numbers associated with the mass and spin current around the vortex, respectively (e.g. see Kudo and Kawaguchi (2015)). Combining the circulations for the m=±1m=\pm 1 components we see that σM±σS\sigma_{\text{M}}\pm\sigma_{\text{S}} must both be integers for the field to be singled valued. There are 8 non-trivial cases with |σM±σS|1|\sigma_{\text{M}}\pm\sigma_{\text{S}}|\leq 1 which define the elementary vortices of interest.

II.2.1 HQV

The HQVs come in 4 types with (σM,σS)=(±12,±12)(\sigma_{\text{M}},\sigma_{\text{S}})=(\pm\frac{1}{2},\pm\frac{1}{2}) and (±12,12)(\pm\frac{1}{2},\mp\frac{1}{2}), thus exhibiting both spin and mass currents. In these vortices the director completes a π\pi winding around the vortex core. These vortices have recently been studied in experiments and observed to be long lived topological defects Seo et al. (2015, 2016). Indeed, HQVs are expected to be stable topological defects of the EPP phase and have been the focus of studies of nonequilibrium dynamics in this phase (e.g. see Kang et al. (2017); Symes and Blakie (2017); Symes et al. (2018)). Experiments have also observed annihilation events between suitable pairs of HQVs Seo et al. (2016), e.g. the pair {(12,12),(12,12)}\{(\frac{1}{2},\frac{1}{2}),(-\frac{1}{2},-\frac{1}{2})\} or {(12,12),(12,12)}\{(\frac{1}{2},-\frac{1}{2}),(-\frac{1}{2},\frac{1}{2})\} can mutually annihilate.

II.2.2 Mass vortex

The mass vortex has (σM,σS)=(±1,0)(\sigma_{\text{M}},\sigma_{\text{S}})=(\pm 1,0), with mass current but no spin current. These vortices have been prepared in the EPP phase Seo et al. (2015), but were observed to rapidly decay by dissociation into two HQVs, i.e.

(±1,0)(±12,±12)+(±12,12),\displaystyle(\pm 1,0)\to(\pm\tfrac{1}{2},\pm\tfrac{1}{2})+(\pm\tfrac{1}{2},\mp\tfrac{1}{2}), (16)

as anticipated by theoretical studies Ji et al. (2008); Eto et al. (2011); Lovegrove et al. (2012).

Refer to caption
Figure 1: The basic structure of the NSV at the origin in a quasi-2D EPP phase condensate. (a) Normal-core NSV and (b) a polar-core NSV. Cylindrical rods indicate the orientation of the nematic director d\vec{d}. Far from the vortex core the director lies in the plane and undergoes a 2π2\pi rotation as we complete a closed loop around the core. The background shaded colours indicate the total density.

II.2.3 NSV

The NSV has (σM,σS)=(0,±1)(\sigma_{\text{M}},\sigma_{\text{S}})=(0,\pm 1), thus exhibits a spin current but no mass current. Similar to the mass vortex the NSV can potentially dissociate into two HQVs Ishino et al. (2013) as

(0,±1)(±12,±12)+(12,±12).\displaystyle(0,\pm 1)\to(\pm\tfrac{1}{2},\pm\tfrac{1}{2})+(\mp\tfrac{1}{2},\pm\tfrac{1}{2}). (17)

We consider the stability to this dissociation process in Sec. IV.

III Stationary NSV Solutions

Here, we investigate the structure of the NSV core. Examples of the two types of NSV are shown in Fig. 1. We see that completing a loop around the core (at a radius sufficiently far from the core) the director remains in the easy-plane manifold and completes a 2π2\pi winding. The two examples differ in their structure near the vortex core. The case in Fig. 1(a), which we refer to as the normal-core NSV, has the total density vanish at the vortex core, with the director always remaining in the plane (i.e. dz=0d_{z}=0). The case in Fig. 1(b), which we refer to as the polar-core NSV, instead has a finite density at the vortex core. Here the director moves out of the EPP order parameter manifold (i.e. tilts out of the plane near the core) showing the emergence of the EAP state within the core. We adopt the name polar core for this case, with polar being a conventional name for the EAP phase. Here we show that there is a continuous transition between these two types of NSV as qq changes.

III.1 Spin-1 Gross-Pitaevskii equation

The evolution of a spin-1 condensate is described by the Gross-Pitaevskii equation (GPE)

i𝚿t=ˇ𝚿,\displaystyle i\hbar\frac{\partial{\bm{\Psi}}}{\partial t}=\check{\mathcal{L}}\bm{\Psi}, (18)

where the ˇ\check{\mathcal{L}} is the nonlinear GPE operator

ˇ=222M𝟙+qfˇz2+c0n𝟙+c1αFαfˇα,\displaystyle\check{\mathcal{L}}=-\frac{\hbar^{2}\nabla^{2}}{2M}\mathds{1}+q\check{f}_{z}^{2}+c_{0}n\mathds{1}+c_{1}\sum_{\alpha}F_{\alpha}\check{f}_{\alpha}, (19)

with α{x,y,z}\alpha\in\{x,y,z\}. Here 𝟙\mathds{1} denotes the identity matrix in spin space. The nonlinear terms nn and F\vec{F} are determined using 𝚿\bm{\Psi}. Here we focus our attention on a quasi-2D system with spatial coordinates ρ=(x,y)\vec{\rho}=(x,y), where c0c_{0} and c1c_{1} are the quasi-2D coupling constants.

Stationary solutions of the form Ψm(ρ,t)=ψm(ρ)ei(μ+mλ)t/\Psi_{m}(\vec{\rho},t)=\psi_{m}(\vec{\rho}\,)e^{-i(\mu+m\lambda)t/\hbar} satisfy the time-independent Gross-Pitaevskii equation

(μ𝟙+λfˇz)𝝍=ˇ𝝍,\displaystyle(\mu\mathds{1}+\lambda\check{f}_{z})\bm{\psi}=\check{\mathcal{L}}\bm{\psi}, (20)

with nonlinear terms in ˇ\check{\mathcal{L}} now evaluated with 𝝍\bm{\psi}. Here μ\mu and λ\lambda are the chemical and magnetic potentials introduced as a Lagrange multipliers to conserve norm NN and zz-magnetization MzM_{z}, respectively111These are defined by the thermodynamic relations μ=(EN)S,V\mu=\left(\frac{\partial E}{\partial N}\right)_{\!S,V}, and λ=(EMz)S,V\lambda=\left(\frac{\partial E}{\partial M_{z}}\right)_{\!S,V}. For the bulk EPP phase at T=0T=0 (with entropy S=0S=0) E=N(q+12c0nb)+12c1Fz,bMzE=N(q+\frac{1}{2}c_{0}n_{b})+\frac{1}{2}c_{1}F_{z,b}M_{z} [from (3) and (2)], with N=VnbN=Vn_{b} and Mz=VFz,bM_{z}=VF_{z,b} for the appropriately dimensioned volume VV.. For a uniform EPP phase condensate of bulk density nbn_{b} and magnetization density Fz,bF_{z,b} we have

μ=c0nb+q,\displaystyle\mu=c_{0}n_{b}+q, (21)

and λ=c1Fz,b\lambda=c_{1}F_{z,b} (e.g. see Kawaguchi and Ueda (2012)).

III.2 Radial Spin-1 GPE

An EPP phase vortex stationary state takes the form [generalizing Eq. (15)]:

𝝍V(ρ,ϕ)=eiσMϕ[eiσSϕχ1(ρ)χ0(ρ)eiσSϕχ1(ρ)]=Cˇ(ϕ)𝝌(ρ),\displaystyle\bm{\psi}_{\text{V}}(\rho,\phi)=e^{i\sigma_{\text{M}}\phi}\begin{bmatrix}e^{-i\sigma_{\text{S}}\phi}\chi_{1}(\rho)\\ \chi_{0}(\rho)\\ e^{i\sigma_{\text{S}}\phi}\chi_{-1}(\rho)\end{bmatrix}=\check{C}(\phi)\bm{\chi}(\rho), (22)

where we have used the radial coordinate ρ=x2+y2\rho=\sqrt{x^{2}+y^{2}}. Here we take 𝝌=[χ1,χ0,χ1]T\bm{\chi}=[\chi_{1},\chi_{0},\chi_{-1}]^{\mathrm{T}} as real, and have explicitly imposed the circulation on each component using Cˇ=diag{ei(σMσS)ϕ,eiσMϕ,ei(σM+σS)ϕ}\check{C}=\text{diag}\{e^{i(\sigma_{\text{M}}-\sigma_{\text{S}})\phi},e^{i\sigma_{\text{M}}\phi},e^{i(\sigma_{\text{M}}+\sigma_{\text{S}})\phi}\}. Substituting (22) into Eq. (20), gives a radial equation for the 𝝌\bm{\chi}:

μ~𝝌=𝒦ˇ𝝌,\displaystyle\tilde{\mu}\bm{\chi}=\check{\mathcal{K}}\bm{\chi}, (23)

with

𝒦ˇ=Tρ𝟙+p~fˇz+q~(fˇz2𝟙)+c0n𝟙+c1α(𝝌Tfˇα𝝌)fˇα.\displaystyle\check{\mathcal{K}}=T_{\rho}\mathds{1}+\tilde{p}\check{f}_{z}+\tilde{q}(\check{f}_{z}^{2}-\mathds{1})+c_{0}n\mathds{1}+c_{1}\sum_{\alpha}\left(\bm{\chi}^{\mathrm{T}}\check{f}_{\alpha}\bm{\chi}\right)\check{f}_{\alpha}. (24)

Here

Tρ\displaystyle T_{\rho} 22M1ρddρ(ρddρ),\displaystyle\equiv-\frac{\hbar^{2}}{2M}\frac{1}{\rho}\derivative{\rho}\left(\rho\derivative{\rho}\right), (25)
μ~(ρ)\displaystyle\tilde{\mu}(\rho) μb2(σM2+σS2)2Mρ2,\displaystyle\equiv\mu_{b}-\frac{\hbar^{2}(\sigma_{\text{M}}^{2}+\sigma_{\text{S}}^{2})}{2M\rho^{2}}, (26)
p~(ρ)\displaystyle\tilde{p}(\rho) λ2σMσSMρ2,\displaystyle\equiv-\lambda-\frac{\hbar^{2}\sigma_{\text{M}}\sigma_{\text{S}}}{M\rho^{2}}, (27)
q~(ρ)\displaystyle\tilde{q}(\rho) q+2σS22Mρ2,\displaystyle\equiv q+\frac{\hbar^{2}\sigma_{\text{S}}^{2}}{2M\rho^{2}}, (28)

are the radial part of the kinetic energy operator, the effective chemical potential, and the effective linear and quadratic Zeeman shifts, respectively. We have also subtracted qq off the single particle energy [see Eq. (24)] for convenience, so that the adjusted chemical potential appearing in Eq. (26) is given by

μbμq=c0nb,\displaystyle\mu_{b}\equiv\mu-q=c_{0}n_{b}, (29)

and is independent of qq [cf. Eq.(21)]. Here we take μb=c0nb\mu_{b}=c_{0}n_{b} as a useful characteristic energy scale of the system, with associated healing length

ξb=Mμb.\displaystyle\xi_{b}=\frac{\hbar}{\sqrt{M\mu_{b}}}. (30)
Refer to caption
Figure 2: Stationary state component densities |χm(ρ)|2|\chi_{m}(\rho)|^{2} of a NSV with (a) a normal core and (b) a polar core. (c) The central (peak) density of the m=0m=0 component. (d) The core radius ρcore\rho_{\text{core}} defined as the radius where |χ±1(ρcore)|2=14nb|\chi_{\pm 1}(\rho_{\text{core}})|^{2}=\frac{1}{4}n_{b} [see arrow in subplot(a)]. The red dotted line shows 12ξq\frac{1}{\sqrt{2}}\xi_{q}. (e) The effective potential c0|χv|2c_{0}|\chi_{\mathrm{v}}|^{2} for the core localized state ψcore\psi_{\mathrm{core}} used to identify the critical point (see text).

III.3 Phases of a NSV

In this work we consider a NSV with (σM,σS)=(0,1)(\sigma_{\text{M}},\sigma_{\text{S}})=(0,1). From our choice of μb\mu_{b} being fixed [Eq. (29)], far away from the vortex core the system will approach the bulk value for number density , i.e. nbn_{b}. Additionally, we focus on the case λ=0\lambda=0 (and thus p~=0\tilde{p}=0), for which the bulk spin density is Fz,b=0F_{z,b}=0. In this case the NSV is completely unmagnetized222The discrete symmetry associated with the operation fˇzfˇz\check{f}_{z}\to-\check{f}_{z} is explicitly broken for the “biased” case with λ0\lambda\neq 0, where a magnetization (Fz0F_{z}\neq 0) is generally preferred. In this work, we consider the “unbiased” case with λ=0\lambda=0. Even for λ=0\lambda=0, a spontaneous magnetization can occur locally, e.g. Fz0F_{z}\neq 0 in the core of a HQV. In this case, the spinor field is no longer represented by the real vector d\vec{d} like Eq. (10). with χ1=χ1\chi_{1}=-\chi_{-1}, and the boundary conditions on the χm\chi_{m} are:

χ0(0)=0,\displaystyle\chi_{0}^{\prime}(0)=0,\qquad χ0(ρ)=0,\displaystyle\chi_{0}(\rho\to\infty)=0, (31)
χ±1(0)=0,\displaystyle\chi_{\pm 1}(0)=0,\qquad χ±1(ρ)=nb2,\displaystyle\chi_{\pm 1}(\rho\to\infty)=\mp\sqrt{\frac{n_{b}}{2}}, (32)

where the prime denotes a derivative with respect to ρ\rho. We solve for the stationary state solutions numerically using a gradient flow technique (e.g. see Bao and Cai (2013)) with a finite difference implementation of the derivative operators and boundary conditions. We use an equally spaced radial grid of NρN_{\rho} points ρj=(j12)Δρ\rho_{j}=(j-\frac{1}{2})\Delta\rho with 1jNρ1\leq j\leq N_{\rho}. We choose the point spacing Δρ\Delta\rho to be much smaller than ξb\xi_{b} and typically use a maximum radius of ρmax410ξb\rho_{\mathrm{max}}\approx 410\xi_{b}, with Nρ=8192N_{\rho}=8192. We choose to implement the outer boundary conditions in Eqs. (31) and (32) as χm(ρmax)=0\chi^{\prime}_{m}(\rho_{\mathrm{max}})=0.

Because the NSV stationary solutions are completely unmagnetized they are independent of the strength of the spin-dependent interaction. However, they depend on the quadratic Zeeman energy, and we find that there is a critical value

qc0.2545μb,\displaystyle q_{c}\approx-0.2545\mu_{b}, (33)

[or qc0.3414μq_{c}\approx-0.3414\mu, see Eq. (29)], which separates the normal-core and polar-core forms of the vortex.

III.3.1 Normal-core NSV

For q<qcq<q_{c} the stationary state has χ0=0\chi_{0}=0 and is otherwise is independent of qq [e.g., see Fig. 2(a)]. In this regime the vortex profile is χ1(ρ)=12χv(ρ)\chi_{-1}(\rho)=\tfrac{1}{\sqrt{2}}\chi_{\text{v}}(\rho), where χv\chi_{\text{v}} is the radial profile333Defining the scalar vortex state as ψv(ρ)=eiϕχv(ρ)\psi_{\text{v}}(\vec{\rho}\,)=e^{i\phi}\chi_{\text{v}}(\rho). of a single component (scalar) vortex in uniform system satisfying

μbχv=(Tρ+22Mρ2+c0|χv|2)χv,\displaystyle\mu_{b}\chi_{\text{v}}=\left(T_{\rho}+\frac{\hbar^{2}}{2M\rho^{2}}+c_{0}|\chi_{\text{v}}|^{2}\right)\chi_{\text{v}}, (34)

with chemical potential μb\mu_{b} used to ensure that |χv|2|\chi_{\text{v}}|^{2} goes to nbn_{b} as ρ\rho\to\infty.

III.3.2 Polar-core NSV

When q>qcq>q_{c} the χ0\chi_{0} component is non-zero, and constitutes a polar core of the vortex [e.g., see Fig. 2(b)]. As qq further increases the density of the χ0\chi_{0} component and the core radius of the NSV both increase, with |χ0(0)|2nb|\chi_{0}(0)|^{2}\to n_{b} and the core radius diverging as q0q\to 0 [see Fig. 2(c) and (d)].

We can qualitatively understand this behaviour by noting that the effective quadratic Zeeman energy for the system q~(ρ)\tilde{q}(\rho) is spatially dependent due to the vortex kinetic energy. Within a local density approximation the sign of q~\tilde{q} determines local nematic spin order: the EPP phase occurs where q~<0\tilde{q}<0 and EAP phase occurs where q~>0\tilde{q}>0 (see Sec. II.1). We see that the effect of the vortex kinetic energy is to transition a central region of radius ρcoreξq=/2M|q|\rho_{\text{core}}\sim\xi_{q}=\hbar/\sqrt{2M|q|} into the EAP phase, where ξq\xi_{q} is defined by q~(ξq)=0\tilde{q}(\xi_{q})=0. In Fig. 2(d) we observe that ξq\xi_{q} provides a good estimate of the core size for q>qcq>q_{c}, noting that for q<qcq<q_{c} the local density approximation breaks because of finite-size effects in the vortex core.

III.3.3 Identifying the critical point

Near the critical point χ0\chi_{0} is small (i.e. |χ0|2nb|\chi_{0}|^{2}\ll n_{b}) and a single-particle treatment of this component can be employed. In this regime the m=0m=0 component of the GPE (23) reduces to (neglecting the nonlinear terms in χ0\chi_{0}):

μbχ0=(Tρ+c0|χv|2q)χ0,|χ0|2nb,\displaystyle\mu_{b}\chi_{0}=(T_{\rho}+c_{0}|\chi_{\text{v}}|^{2}-q)\chi_{0},\qquad|\chi_{0}|^{2}\ll n_{b}, (35)

where we have set n=|χv|2n=|\chi_{\text{v}}|^{2} (i.e. the normal-core NSV density, see Sec. III.3.1) in the interaction term. The lowest energy eigenstate of the linear operator Tρ+c0|χv|2qT_{\rho}+c_{0}|\chi_{\text{v}}|^{2}-q is a core localized state of energy444The lowest energy eigenstate of Tρ+c0|χv|2T_{\rho}+c_{0}|\chi_{\text{v}}|^{2} is ψcore\psi_{\text{core}} with a numerically determined energy of ϵ0=0.7455μb\epsilon_{0}=0.7455\mu_{b}. This state is bound within the vortex core [see Fig. 2(e)]. Note that the harmonic approximation to c0|χv|2c_{0}|\chi_{\text{v}}|^{2}, i.e. Uharm=μb(Λρ/ξb)2U_{\text{harm}}=\mu_{b}(\Lambda\rho/\xi_{b})^{2}, with Λ=0.8249\Lambda=0.8249 Bradley and Anderson (2012), is inaccurate and predicts ϵ0=1.14μb\epsilon_{0}=1.14\mu_{b}, which exceeds the core well-depth of μb\mu_{b}. ϵcore=0.7455μbq\epsilon_{\text{core}}=0.7455\mu_{b}-q. The system will “condense" into this state when the chemical potential, μb\mu_{b}, exceeds ϵcore\epsilon_{\text{core}}. Thus we take μb=ϵcore\mu_{b}=\epsilon_{\text{core}} to define the critical value of the quadratic Zeeman energy, yielding a value of qcq_{c} in agreement with the value identified from the GPE calculations for the NSV [Eq. (33)].

IV Linear stability of the NSV

We now examine the excitations of the NSV by directly solving the BdG equations. We begin by deriving the BdG equations for an EPP phase vortex. We then use numerical solutions of these equations to quantify NSV stability.

IV.1 Bogoliubov-de Gennes equations

To derive the BdG equations we introduce a time-dependent fluctuation δ𝝍(ρ,t)=𝝍(ρ,t)𝝍V(ρ)\delta\bm{\psi}(\vec{\rho},t)=\bm{\psi}(\vec{\rho},t)-\bm{\psi}_{\text{V}}(\vec{\rho}\,) on the vortex stationary state 𝝍V\bm{\psi}_{\text{V}} [see Eq. (22)]:

δ𝝍(ρ,t)=Cˇ(ϕ)ν,η[βν,η𝒖ν,ηβν,η𝒗ν,η],\displaystyle\delta\bm{\psi}(\vec{\rho},t)=\check{C}(\phi)\sum_{\nu,\eta}\left[\beta_{\nu,\eta}\bm{u}_{\nu,\eta}-\beta_{\nu,\eta}^{*}\bm{v}_{\nu,\eta}^{*}\right], (36)

where βν,η\beta_{\nu,\eta} are arbitrary (small) linearization amplitudes, and

𝒖ν,η(ρ,t)=eiEν,ηt/+iηϕ𝒖~ν,η(ρ),\displaystyle\bm{u}_{\nu,\eta}(\vec{\rho},t)=e^{-iE_{\nu,\eta}t/\hbar+i\eta\phi}\tilde{\bm{u}}_{\nu,\eta}(\rho), (37)
𝒗ν,η(ρ,t)=eiEν,ηt/+iηϕ𝒗~ν,η(ρ).\displaystyle\bm{v}_{\nu,\eta}(\vec{\rho},t)=e^{-iE_{\nu,\eta}t/\hbar+i\eta\phi}\tilde{\bm{v}}_{\nu,\eta}(\rho). (38)

Here we have introduced the radial quasiparticle amplitudes {𝒖~ν,η(ρ),𝒗~ν,η(ρ)}\{\tilde{\bm{u}}_{\nu,\eta}(\rho),\tilde{\bm{v}}_{\nu,\eta}(\rho)\} and eigenvalues Eν,ηE_{\nu,\eta}, with η\eta being the quantum number associated with the zz-component of angular momentum (relative to the condensate) and ν\nu representing the remaining quantum numbers. Linearizing the time-dependent GPE (18) we obtain that the quasiparticle amplitudes satisfy the Bogoliubov-de Gennes (BdG) equation

[𝒦ˇη++Xˇ1Xˇ2Xˇ2(𝒦ˇη+Xˇ1)][𝒖~ν,η𝒗~ν,η]=Eν,η[𝒖~ν,η𝒗~ν,η],\displaystyle\begin{bmatrix}\check{\mathcal{K}}^{+}_{\eta}+\check{X}_{1}&-\check{X}_{2}\\ \check{X}_{2}^{*}&-(\check{\mathcal{K}}^{-}_{\eta}+\check{X}_{1})^{*}\end{bmatrix}\begin{bmatrix}\tilde{\bm{u}}_{\nu,\eta}\\ \tilde{\bm{v}}_{\nu,\eta}\end{bmatrix}=E_{\nu,\eta}\begin{bmatrix}\tilde{\bm{u}}_{\nu,\eta}\\ \tilde{\bm{v}}_{\nu,\eta}\end{bmatrix}, (39)

where nonlinear terms in

𝒦ˇη±\displaystyle\check{\mathcal{K}}^{\pm}_{\eta} =𝒦ˇ+22Mρ2[(η2±2σMη)𝟙2σSηfˇz]μ~𝟙,\displaystyle=\check{\mathcal{K}}+\frac{\hbar^{2}}{2M\rho^{2}}\left[\left(\eta^{2}\pm 2\sigma_{\text{M}}\eta\right)\!\mathds{1}\mp 2\sigma_{\text{S}}\eta\check{f}_{z}\right]-\tilde{\mu}\mathds{1}, (40)

are evaluated with 𝝍V\bm{\psi}_{\text{V}}, and

Xˇ1\displaystyle\check{X}_{1} =c0𝝌𝝌T+c1αfˇα𝝌𝝌Tfˇα,\displaystyle=c_{0}\bm{\chi}\bm{\chi}^{\mathrm{T}}+c_{1}\sum_{\alpha}\check{f}_{\alpha}\bm{\chi}\bm{\chi}^{\mathrm{T}}\check{f}_{\alpha}, (41)
Xˇ2\displaystyle\check{X}_{2} =c0𝝌𝝌T+c1αfˇα𝝌𝝌Tfˇα.\displaystyle=c_{0}\bm{\chi}\bm{\chi}^{\mathrm{T}}+c_{1}\sum_{\alpha}\check{f}_{\alpha}\bm{\chi}\bm{\chi}^{\mathrm{T}}\check{f}_{\alpha}^{*}. (42)

Because of the symmetry of the radial BdG equation for a solution (i) (E,η,𝐮~,𝐯~)(E,\eta,\tilde{\mathbf{u}},\tilde{\mathbf{v}}) there are potentially three additional solutions which relate to the first as: (ii) (E,η,𝐯~,𝐮~)(-E,-\eta,\tilde{\mathbf{v}},\tilde{\mathbf{u}}), (iii) (E,η,𝐮~,𝐯~)(E^{*},\eta,\tilde{\mathbf{u}}^{*},\tilde{\mathbf{v}}^{*}), and (iv) (E,η,𝐯~,𝐮~)(-E^{*},-\eta,\tilde{\mathbf{v}}^{*},\tilde{\mathbf{u}}^{*}) . If the eigenvalue is real, then 𝐮~\tilde{\mathbf{u}} and 𝐯~\tilde{\mathbf{v}} can also be taken real (see Ref. Takahashi et al. (2009)) and only (i) and (ii) are unique solutions. Furthermore quasiparticle amplitudes with real non-zero eigenvalues can be normalized to ±1\pm 1 as

02πρdρ(𝒖ν,η𝒖ν,η𝒗ν,η𝒗ν,η)=±δννδη,η.\displaystyle\int_{0}^{\infty}2\pi\rho\,\differential{\rho}\,(\bm{u}_{\nu,\eta}^{\dagger}\bm{u}_{\nu^{\prime},\eta^{\prime}}-\bm{v}_{\nu,\eta}^{\dagger}\bm{v}_{\nu^{\prime},\eta^{\prime}})=\pm\delta_{\nu\nu^{\prime}}\delta_{\eta,\eta^{\prime}}. (43)

In the description of equilibrium condensates only positively normalized quasiparticles are considered to be physical. We note that the partner (ii) to a positively normed quasiparticle (i) has negative norm.

Here we are particularly interested in quasiparticles with complex eigenvalues where all the symmetries (i)-(iv) furnish unique solutions. If Im(E)>0\text{Im}(E)>0 then the solution (i) is dynamically unstable (exponentially growing in time), and so is the partner solution (iv), while (ii) and (iii) are exponentially decaying solutions. Examining the effect of the unstable mode perturbation δ𝝍\delta\bm{\psi} (36) we see modes (i) and (iv) are identical perturbations, so here we can choose to focus on solution (i).

Refer to caption
Figure 3: Unstable (splitting) mode of the NSV. (a) Phase diagram showing the imaginary part of the eigenvalue of the unstable mode with η=±1\eta=\pm 1. Dotted line indicates the simple model for instability boundary based on counter-superflow instability. Several slices through the phase diagram showing the strength of instability as (b) qq varies, and (c) as c1c_{1} varies. In (b) the dashed line shows the zero mode for c1=0.15c0c_{1}=0.15c_{0}. The inset shows the behaviour of the c1=0.15c0c_{1}=0.15c_{0} modes near where the instability vanishes. (d) The non-zero uu-components of the η=1\eta=-1 unstable mode for q=0.3μbq=-0.3\mu_{b} and c1=0.05c0c_{1}=0.05c_{0}. (e) The zz-component of the spin density of the NSV after the unstable mode shown in (d) is added to the stationary vortex state, revealing that the vortices of the two components spatially separate.

IV.2 Dynamic instability phase diagram

The BdG results can reveal two types of instabilities for the NSV: (i) A dynamic instability revealed by a solution with a complex eigenvalue indicating that the respective eigenmode will exponentially grow with time. (ii) A Landau instability marked by a (positively normed) solution with a negative real eigenvalue, such that the system could reduce its energy if some dissipative mechanism allowed transfer of population into this state.

We have numerically calculated the BdG spectrum of the NSV over a wide parameter regime. We find that for q<0q<0 the NSV only exhibits dynamic instabilities which occur in excitations with η=±1\eta=\pm 1, arising from modes that are localized in the vortex core555The centrifugal term in the m=ηm=-\eta (m=ηm=\eta) component of the operator 𝒦ˇη+\check{\mathcal{K}}_{\eta}^{+} (𝒦ˇη\check{\mathcal{K}}_{\eta}^{-}) vanishes for η=±1\eta=\pm 1, allowing the corresponding component of the 𝒖ν,η\bm{u}_{\nu,\eta} (𝒗ν,η\bm{v}_{\nu,\eta}) amplitude to develop amplitude in the NSV core. [see Fig. 3(d)]. The growth of these unstable modes causes the m=±1m=\pm 1 cores of the NSV to separate [see Fig. 3(e)], thus initiating the dissociation of the NSV into two HQVs (see Sec. II.2.3).

In Fig. 3(a) we present a stability phase diagram quantifying the dynamic instability of the NSV (i.e. showing the imaginary part of the eigenvalue of the dynamically unstable mode) as qq and c1c_{1} vary. In general the imaginary part is always relatively small (101μb\lesssim 10^{-1}\mu_{b}), so we expect the instability to manifest slowly in the system dynamics (see Sec. V). The dynamic instability is seen to depend on both qq and c1c_{1}. It is independent of qq for the normal-core NSV (i.e. q<qcq<q_{c}), but reduces with increasing qq in the polar-core regime [Fig. 3(b)]. The instability also decreases with increasing c1c_{1} [Fig. 3(c)].

IV.2.1 qq-dependence of instability: Pinning effect of polar core

For the polar-core NSV the magnitude of the dynamic instability decreases with increasing qq. We interpret this as a pinning effect of the polar core that helps bind the two component vortices together, and thus stabilize the NSV. The effects of pinning (e.g. due to an external potential, the other superfluid component, or thermal component) has previously been considered as a mechanism for stabilizing vortices (e.g. see Isoshima and Machida (1999); Isoshima (2002); Simula et al. (2002); Hayashi et al. (2013)).

To quantify the pinning effect we consider a polar-core NSV solution 𝝌PC=[χ1,χ0,χ1]T\bm{\chi}_{\text{PC}}=[-\chi_{-1},\chi_{0},\chi_{-1}]^{\mathrm{T}}. From this we can project out the m=0m=0 component to arrive at an effective normal-core NSV 𝝌ENC=[χ1,0,χ1]T\bm{\chi}_{\text{ENC}}=[-\chi_{-1},0,\chi_{-1}]^{\mathrm{T}} that is a stationary solution for the same qq value if we add the scalar potential Upin=c0|χ0|2U_{\text{pin}}=c_{0}|\chi_{0}|^{2} to the GPE666I.e. add a term Upin𝟙U_{\text{pin}}\mathds{1} to ρ\mathcal{L}_{\rho} to compensate for the reduction of c0nc_{0}n by the removal of the χ0\chi_{0} component.. The unstable modes in the BdG analysis of 𝝌ENC\bm{\chi}_{\text{ENC}} (including the pinning potential) have larger imaginary parts than those obtained for 𝝌PC\bm{\chi}_{\text{PC}}. This demonstrates that there is an intrinsic spin-dependent aspect to the pinning stabilization. If we artificially increase the strength of the pinning potential (i.e. set UpinγUpinU_{\text{pin}}\to\gamma U_{\text{pin}}, with γ>1\gamma>1) then eventually the dynamical instability is suppressed.

IV.2.2 c1c_{1}-dependence of instability: Counter-superflow instability

Counter-superflow instability involves the breakdown of spin-superfluidity when the relative velocity of two miscible superfluids exceeds a critical value Law et al. (2001); Yukalov and Yukalova (2004); Takeuchi et al. (2010); Suzuki et al. (2010); Abad et al. (2015); Hoefer et al. (2011); Hamner et al. (2011); Fujimoto and Tsubota (2012); Zhu et al. (2015); Kim et al. (2017), and affords a qualitative understanding of the dependence of the NSV instability on c1c_{1}. For a uniform spinor condensate in the EPP phase the critical relative velocity for the onset of the instability is vcrit=2c1n/Mv_{\text{crit}}=2\sqrt{c_{1}n/M} Zhu et al. (2015). We can apply this criteria to NSV using the local density approximation (similar to the treatment presented in Ref. Ishino et al. (2013)). The relative velocity arises from the counter-rotating vortices in the m=±1m=\pm 1 components and varies radially as vrel(ρ)=2Mρv_{\text{rel}}(\rho)=\frac{2\hbar}{M\rho}. Approximating the NSV density by the background value nbn_{b} we identify the critical radius ρcrit=c0c1ξb\rho_{\text{crit}}=\sqrt{\frac{c_{0}}{c_{1}}}\xi_{b}, from the condition vrel(ρcrit)=vcritv_{\text{rel}}(\rho_{\text{crit}})=v_{\text{crit}}. For ρ<ρcrit\rho<\rho_{\text{crit}} the relative velocity exceeds vcritv_{\text{crit}} and counter-superflow instability is activated. This analysis suggests that the instability will be stronger closer to the core, consistent with unstable modes being localized near the core [see Fig. 3(d)]. Also, as c1/c0c_{1}/c_{0} and hence the critical velocity increases, ρcrit\rho_{\text{crit}} decreases, suggesting that the instability should be weaker.

This analysis does not apply to the core region as here the density varies rapidly so that the local density approximation is inapplicable. If we assume that the instability is suppressed once the critical radius is comparable to the vortex core size we can quantify a stability boundary for the system. In the polar-core regime we estimate the core size as ξq\xi_{q}, and the critical radius is equal to this when

c1c0=qμb=qμq,(stability boundary).\displaystyle\frac{c_{1}}{c_{0}}=-\frac{q}{\mu_{b}}=-\frac{q}{\mu-q},\qquad\mbox{(stability boundary)}. (44)

This is shown as a dotted line in Fig. 3(a), and is seen to reasonably characterize the boundary of instability in the polar-core regime.

Refer to caption
Figure 4: The non-trivial uu-components of a η=1\eta=-1 zero energy mode and the functions Λ±\Lambda_{\pm}. The same scaling factor is applied to the uu-amplitudes. Note the nontrivial vv-components for the zero energy mode are given by v~±=u~\tilde{v}_{\pm}=\tilde{u}_{\mp}. Parameters are as in Fig. 3(d).

IV.3 Zero energy modes from broken translational symmetry

For η=±1\eta=\pm 1 we find that in addition to the unstable modes there are also zero-energy modes. These modes often reveal a broken symmetry in the system. These new zero modes are in addition to the usual η=0\eta=0 zero energy mode, which can be written in terms of the condensate mode as {𝒖0,0,𝒗0,0}={𝝌,𝝌}\{\bm{u}_{0,0},\bm{v}_{0,0}\}=\{\bm{\chi},-\bm{\chi}\} and is associated with the breaking of the gauge symmetry. The new zero energy modes with η=±1\eta=\pm 1 are associated with the breaking of translational symmetry in our uniform system by the presence of the NSV. These modes are localized in the core, similar to the unstable modes, but have a flat phase profile [see Fig. 4 and cf. Fig. 3(d)]. Whereas the unstable mode causes the m=±1m=\pm 1 component vortices to separate, the zero energy mode causes both to translate together. For a scalar condensate a similar zero mode emerges and in the case of a vortex line (i.e. finite zz-extent), it is associated with the Kelvin-wave spectrum of helical modes that propagate along the vortex line.

We can develop an analytic expression for these zero energy modes that we compare to the numerical solution of the BdG equations. We restrict our attention to a normal-core NSV (σM,σS)=(0,1)(\sigma_{\text{M}},\sigma_{\text{S}})=(0,1), which has the form [see Eq. (22)] 𝝍V=[χ1(ρ)eiϕ,0,χ1(ρ)eiϕ]T\bm{\psi}_{V}=[-\chi_{-1}(\rho)e^{-i\phi},0,\chi_{-1}(\rho)e^{i\phi}]^{\mathrm{T}}. Considering the vortex to be displaced by a small amount dρ\text{d}\vec{\rho} such that the change in the condensate wavefunction is δ𝝍=dρ𝝍V\delta\bm{\psi}=\text{d}\vec{\rho}\cdot\vec{\nabla}\bm{\psi}_{V}, we obtain for the change in the nontrivial m=±1m=\pm 1 components

δψ±1=eiϕ[(dxidy)Λ±eiϕ+(dx+idy)Λeiϕ]\displaystyle\delta{\psi}_{\pm 1}=e^{\mp i\phi}\left[-\left(\differential{x}-i\differential{y}\right)\Lambda_{\pm}e^{i\phi}+\left(\differential{x}+i\differential{y}\right)\Lambda_{\mp}e^{-i\phi}\right] (45)

where

Λ±(ρ)=12(χ1ρ±dχ1dρ).\displaystyle\Lambda_{\pm}(\rho)=\frac{1}{2}\left(\frac{\chi_{-1}}{\rho}\pm\frac{\text{d}\chi_{-1}}{\text{d}{\rho}}\right). (46)

By inspecting the form of the BdG linearization [Eqs. (36)-(38)] we see that the vortex shift can be mapped to quasiparticle amplitudes with η=±1\eta=\pm 1. For definiteness we consider a zero energy mode of the BdG solution with η=1\eta=-1, which we denote as {𝒖~,𝒗~}\{\tilde{\bm{u}},\tilde{\bm{v}}\}, with amplitude β\beta, such that

δψ±1=eiϕ[βu~±1eiϕβv~±1eiϕ].\displaystyle\delta{\psi}_{\pm 1}=e^{\mp i\phi}[{\beta}\tilde{u}_{\pm 1}e^{-i\phi}-{\beta}^{*}\tilde{v}_{\pm 1}^{*}e^{i\phi}]. (47)

In comparison to Eq. (45) we observe u~±1Λ\tilde{u}_{\pm 1}\sim\Lambda_{\mp}, and v~±1Λ±\tilde{v}_{\pm 1}\sim\Lambda_{\pm}, with the complex amplitude β\beta determining the vortex displacement. In Fig. 4 we show the numerically calculated BdG zero energy mode result and the functions Λ±\Lambda_{\pm} (obtained from the vortex solution), showing they coincide.

IV.4 Numerical considerations

It is challenging to numerically calculate the η=±1\eta=\pm 1 unstable and zero modes accurately. One issue is that our grids are of finite spatial extent ρmax=410ξb\rho_{\mathrm{max}}=410\xi_{b}. This is problematic for the zero energy mode which decays rather slowly [noting that Λ±ρ1\Lambda_{\pm}\sim\rho^{-1}, see Eq. (46)]. Additionally, the 2D radial Laplacian is difficult to evaluate accurately with finite differences, particularly near ρ=0\rho=0 Arsoski et al. (2015); Laliena and Campo (2018). In the BdG equations for η=±1\eta=\pm 1 the unstable and zero energy modes are particularly sensitive to these issues. We have found that second order finite difference schemes (including the improved schemes presented in Refs. Arsoski et al. (2015); Laliena and Campo (2018)) require an impractically large number of grid points to obtain accurate results. This motivated us to implement an 8th order finite difference scheme, which is the basis of the results we present. With this scheme small artefacts are still apparent such as the zero mode having a small imaginary part [e.g. see Fig. 3(b) and inset], causing it to couple to the unstable mode. These artefacts reduce as the range and point density of the numerical grid increase.

V Dynamical simulations

Refer to caption
Figure 5: (a)-(f) Time-evolution of the zz-component of the spin-density of a normal-core NSV in a circular flat-bottomed trap. The vortices in the m=1m=1 (white cross) and m=1m=-1 (black cross) components are indicated. Subplots (a) and (b) zoom in to reveal the vortex dynamics in the central region, whereas (c)-(f) show the full simulation domain and indicate the trap boundary (black dashed line). In (f) the vortex trajectories are also shown for the vortices in the m=1m=1 (white line) and m=1m=-1 (black line) components. Other simulation parameters are q=0.3μbq=-0.3\mu_{b}, c1=0.05c0c_{1}=0.05c_{0}, U0=100μbU_{0}=100\mu_{b}, and R=95ξbR=95\xi_{b} is the radius.

We further investigate the stability of NSVs by simulating their dynamics using the time-dependent GPE (18). The simulations are performed with a flat-bottomed circular trap (i.e. scalar potential added to the GPE) of the form

U(ρ)=12U0[tanh(ρRξb)+1],\displaystyle U(\rho)=\frac{1}{2}U_{0}\left[\tanh\left(\frac{\rho-R}{\xi_{b}}\right)+1\right], (48)

where U0U_{0} is the trap depth, and RR is the radius. The initial state is a NSV centered at the origin, obtained by solving the radial GPE [Eq. (23) including U(ρ)U(\rho)] in the flat-bottomed trap potential using the approach described in Sec. III. This state is interpolated onto a uniform 2048×20482048\times 2048-point 2D grid with spacing 0.1ξb0.1\xi_{b}. A small amount of complex Gaussian noise is added to seed any instabilities in the dynamics. This is first prepared as white noise (on the position space grid), then restricted in reciprocal space to have maximum wave-number 8ξb18\xi_{b}^{-1}, and finally spatially filtered to the region within the trap. Typically adding this noise to the initial state causes a 0.005%0.005\% increase in the wavefunction norm and a 0.05%0.05\% increase in the system energy. We time-evolve the resulting state using the second-order symplectic method described in Ref. Symes et al. (2016).

We use the zz-spin density to illustrate the evolution of a normal-core NSV in Fig. 5. Initially FzF_{z} is zero (to the level of the noise) but as the component vortices separate, clear structure develops. The vortex core in the m=1m=1 component is filled by the m=1m=-1 component, and thus appears as a negative FzF_{z} peak. Similarly the vortex core in the m=1m=-1 component appears with a positive FzF_{z} peak. As time progresses the component vortex separation tends to increase and they move away from trap centre. Eventually the vortices approach the boundary where they undergo a sudden change in their motion causing the appreciable emission of spinwaves [see Fig. 5(f)]. In contrast for a polar-core NSV with q>0.05μbq>-0.05\mu_{b} and c1nb>0.05μbc_{1}n_{b}>0.05\mu_{b} [which is stable according to the BdG analysis, see Fig. 3(a) and (b)] we observed the component vortices to remain together at the origin for the entire evolution (i.e. up to tfinal=1200/μbt_{\mathrm{final}}=1200\hbar/\mu_{b}).

Refer to caption
Figure 6: (a) Vortex separation Δr\Delta r of a normal-core NSV, obtained with parameters q=0.3μbq=-0.3\mu_{b}, c1=0.05c0c_{1}=0.05c_{0}. (b) Quadratic Zeeman dependence of the separation time tsept_{\text{sep}} for c1=0.05c0c_{1}=0.05c_{0}. Black dots indicate numerical results. The red line is a fit to the BdG results, given by 4.07/Eun4.07/E_{\text{un}}, where EunE_{\text{un}} is shown in Fig. 3(b).

The above results motivate us to quantify the instability of the NSV in terms of the rate that the component vortices separate. In Fig. 6(a) we show the evolution of the distance between the component vortices (Δr\Delta r) for the case examined in Fig. 5 over the initial period of its dynamics (i.e. well before the boundary collision occurs). We identify the separation time tsept_{\text{sep}} as the time when Δr\Delta r first exceeds ξb\xi_{b}. In Fig. 6(b) we show tsept_{\text{sep}} obtained from simulations conducted over a range of qq values. Here we see that tsept_{\text{sep}} increases with qq for q>qcq>q_{c}, and appears to diverge as qq approaches 0.1μb-0.1\mu_{b}. These results are consistent with the BdG analysis [see Fig. 3] if we identify tsept_{\text{sep}} as scaling with /|Eun|\hbar/|E_{\text{un}}|, where EunE_{\text{un}} is the (imaginary) eigenvalue of the dynamically unstable mode. A comparison to the BdG results is presented in Fig. 6(b) and is seen to have good quantitative agreement.

VI Discussion and conclusions

Here we have presented a description of the NSV, outlining the stationary state properties and a transition between a normal-core and polar-core form occurring at a critical value of the quadratic Zeeman energy. The NSV generally is unstable to dissociating into two HQVs. Using a BdG analysis we quantify this instability, and find that it can be reduced by increasing the strength of the spin-dependent interactions. For the polar-core NSV the instability also decreases by increasing the quadratic Zeeman energy.

It should be possible to controllably produce NSVs using established experimental schemes involving magnetic and optical fields Shin et al. (2004); Chen et al. (2018), which would allow the properties of individual NSVs to be studied. It is also interesting to ask if NSVs could play a role in the non-equilibrium dynamics of an antiferromagnetic spinor condensate quenched into the EPP phase. To date studies have considered the role of HQVs (e.g. Kang et al. (2017); Symes and Blakie (2017)), however we have identified regimes where NSVs are stable (or quasi-stable) defects. In such regimes they may be important in the description of phenomena such as phase ordering and quantum turbulence. We note that the easy-plane ferromagnetic spinor condensate similarly supports two types of vortices, with the dominant vortex type determining the universal ordering dynamics Kudo and Kawaguchi (2015); Williamson and Blakie (2016a); Schmied et al. (2019).

Acknowledgements.
APCU, DB, and PBB acknowledge support from the Marsden Fund of the Royal Society of New Zealand. HT was supported by JSPS KAKENHI Grant Numbers JP17K05549, JP18KK0391, JP20H01842, and in part by the OCU “Think globally, act locally” Research Grant for Young Scientists 2019 through the hometown donation fund of Osaka City.

References