Properties of a nematic spin vortex in an antiferromagnetic spin-1 Bose-Einstein condensate
Abstract
A spin-1 condensate with antiferromagnetic interactions supports nematic spin vortices in the easy-plane polar phase. These vortices have a 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 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
(1) |
with the three components representing the condensate amplitude in the spin levels , respectively, where is the quantum number associated with the -component of spin. In weak fields the short-ranged contact interactions between atoms are rotationally invariant with a Hamiltonian density
(2) |
Here the first term, with coupling constant , describes the density dependent interactions, where is the total density. The second term describes the spin-dependent interactions, where is the spin-dependent coupling constant, is the spin density, and are the spin-1 matrices. In addition, we consider the presence of a (uniform) quadratic Zeeman shift. Taking the field to be along this is described by
(3) |
In practice the coefficient 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) -magnetization of the system. Here we consider only .
The case of antiferromagnetic interactions, where , 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 and spin-density the spinor is in a polar state
(10) |
where the real unit vector is the nematic director and is the global phase. Noting that is invariant under and , we see that defines a preferred axis in spin space, but not a preferred direction along that axis. The ground state orientation of is determined by the quadratic Zeeman energy, which is given by . Thus for the system maximizes , by being in the EAP phase, i.e. . The case of interest in this paper is the EPP phase for where . 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 , i.e. , where is the angle the director takes with respect to the -axis.
Note that in the EPP phase we have 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 may be expressed in the binary form
(11) |
with intra-species coupling constant (identical for both components) and interspecies coupling constant . For the antiferromagnetic case 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 -plane with the vortex core taken at the origin. Sufficiently far from the core the general vortex state is of the form
(15) |
where is the azimuthal angle in the -plane [i.e. from , taking and to vary as and in space, respectively]. Here and 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 components we see that must both be integers for the field to be singled valued. There are 8 non-trivial cases with which define the elementary vortices of interest.
II.2.1 HQV
The HQVs come in 4 types with and , thus exhibiting both spin and mass currents. In these vortices the director completes a 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 or can mutually annihilate.
II.2.2 Mass vortex
The mass vortex has , 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.
(16) |
as anticipated by theoretical studies Ji et al. (2008); Eto et al. (2011); Lovegrove et al. (2012).

II.2.3 NSV
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 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. ). 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 changes.
III.1 Spin-1 Gross-Pitaevskii equation
The evolution of a spin-1 condensate is described by the Gross-Pitaevskii equation (GPE)
(18) |
where the is the nonlinear GPE operator
(19) |
with . Here denotes the identity matrix in spin space. The nonlinear terms and are determined using . Here we focus our attention on a quasi-2D system with spatial coordinates , where and are the quasi-2D coupling constants.
Stationary solutions of the form satisfy the time-independent Gross-Pitaevskii equation
(20) |
with nonlinear terms in now evaluated with . Here and are the chemical and magnetic potentials introduced as a Lagrange multipliers to conserve norm and -magnetization , respectively111These are defined by the thermodynamic relations , and . For the bulk EPP phase at (with entropy ) [from (3) and (2)], with and for the appropriately dimensioned volume .. For a uniform EPP phase condensate of bulk density and magnetization density we have
(21) |
and (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)]:
(22) |
where we have used the radial coordinate . Here we take as real, and have explicitly imposed the circulation on each component using . Substituting (22) into Eq. (20), gives a radial equation for the :
(23) |
with
(24) |
Here
(25) | ||||
(26) | ||||
(27) | ||||
(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 off the single particle energy [see Eq. (24)] for convenience, so that the adjusted chemical potential appearing in Eq. (26) is given by
(29) |
and is independent of [cf. Eq.(21)]. Here we take as a useful characteristic energy scale of the system, with associated healing length
(30) |

III.3 Phases of a NSV
In this work we consider a NSV with . From our choice of being fixed [Eq. (29)], far away from the vortex core the system will approach the bulk value for number density , i.e. . Additionally, we focus on the case (and thus ), for which the bulk spin density is . In this case the NSV is completely unmagnetized222The discrete symmetry associated with the operation is explicitly broken for the “biased” case with , where a magnetization () is generally preferred. In this work, we consider the “unbiased” case with . Even for , a spontaneous magnetization can occur locally, e.g. in the core of a HQV. In this case, the spinor field is no longer represented by the real vector like Eq. (10). with , and the boundary conditions on the are:
(31) | ||||
(32) |
where the prime denotes a derivative with respect to . 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 points with . We choose the point spacing to be much smaller than and typically use a maximum radius of , with . We choose to implement the outer boundary conditions in Eqs. (31) and (32) as .
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
(33) |
[or , see Eq. (29)], which separates the normal-core and polar-core forms of the vortex.
III.3.1 Normal-core NSV
For the stationary state has and is otherwise is independent of [e.g., see Fig. 2(a)]. In this regime the vortex profile is , where is the radial profile333Defining the scalar vortex state as . of a single component (scalar) vortex in uniform system satisfying
(34) |
with chemical potential used to ensure that goes to as .
III.3.2 Polar-core NSV
When the component is non-zero, and constitutes a polar core of the vortex [e.g., see Fig. 2(b)]. As further increases the density of the component and the core radius of the NSV both increase, with and the core radius diverging as [see Fig. 2(c) and (d)].
We can qualitatively understand this behaviour by noting that the effective quadratic Zeeman energy for the system is spatially dependent due to the vortex kinetic energy. Within a local density approximation the sign of determines local nematic spin order: the EPP phase occurs where and EAP phase occurs where (see Sec. II.1). We see that the effect of the vortex kinetic energy is to transition a central region of radius into the EAP phase, where is defined by . In Fig. 2(d) we observe that provides a good estimate of the core size for , noting that for 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 is small (i.e. ) and a single-particle treatment of this component can be employed. In this regime the component of the GPE (23) reduces to (neglecting the nonlinear terms in ):
(35) |
where we have set (i.e. the normal-core NSV density, see Sec. III.3.1) in the interaction term. The lowest energy eigenstate of the linear operator is a core localized state of energy444The lowest energy eigenstate of is with a numerically determined energy of . This state is bound within the vortex core [see Fig. 2(e)]. Note that the harmonic approximation to , i.e. , with Bradley and Anderson (2012), is inaccurate and predicts , which exceeds the core well-depth of . . The system will “condense" into this state when the chemical potential, , exceeds . Thus we take to define the critical value of the quadratic Zeeman energy, yielding a value of 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 on the vortex stationary state [see Eq. (22)]:
(36) |
where are arbitrary (small) linearization amplitudes, and
(37) | |||
(38) |
Here we have introduced the radial quasiparticle amplitudes and eigenvalues , with being the quantum number associated with the -component of angular momentum (relative to the condensate) and representing the remaining quantum numbers. Linearizing the time-dependent GPE (18) we obtain that the quasiparticle amplitudes satisfy the Bogoliubov-de Gennes (BdG) equation
(39) |
where nonlinear terms in
(40) |
are evaluated with , and
(41) | ||||
(42) |
Because of the symmetry of the radial BdG equation for a solution (i) there are potentially three additional solutions which relate to the first as: (ii) , (iii) , and (iv) . If the eigenvalue is real, then and 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 as
(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 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 (36) we see modes (i) and (iv) are identical perturbations, so here we can choose to focus on solution (i).

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 the NSV only exhibits dynamic instabilities which occur in excitations with , arising from modes that are localized in the vortex core555The centrifugal term in the () component of the operator () vanishes for , allowing the corresponding component of the () amplitude to develop amplitude in the NSV core. [see Fig. 3(d)]. The growth of these unstable modes causes the 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 and vary. In general the imaginary part is always relatively small (), so we expect the instability to manifest slowly in the system dynamics (see Sec. V). The dynamic instability is seen to depend on both and . It is independent of for the normal-core NSV (i.e. ), but reduces with increasing in the polar-core regime [Fig. 3(b)]. The instability also decreases with increasing [Fig. 3(c)].
IV.2.1 -dependence of instability: Pinning effect of polar core
For the polar-core NSV the magnitude of the dynamic instability decreases with increasing . 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 . From this we can project out the component to arrive at an effective normal-core NSV that is a stationary solution for the same value if we add the scalar potential to the GPE666I.e. add a term to to compensate for the reduction of by the removal of the component.. The unstable modes in the BdG analysis of (including the pinning potential) have larger imaginary parts than those obtained for . 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 , with ) then eventually the dynamical instability is suppressed.
IV.2.2 -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 . For a uniform spinor condensate in the EPP phase the critical relative velocity for the onset of the instability is 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 components and varies radially as . Approximating the NSV density by the background value we identify the critical radius , from the condition . For the relative velocity exceeds 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 and hence the critical velocity increases, 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 , and the critical radius is equal to this when
(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.

IV.3 Zero energy modes from broken translational symmetry
For 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 zero energy mode, which can be written in terms of the condensate mode as and is associated with the breaking of the gauge symmetry. The new zero energy modes with 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 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 -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 , which has the form [see Eq. (22)] . Considering the vortex to be displaced by a small amount such that the change in the condensate wavefunction is , we obtain for the change in the nontrivial components
(45) |
where
(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 . For definiteness we consider a zero energy mode of the BdG solution with , which we denote as , with amplitude , such that
(47) |
In comparison to Eq. (45) we observe , and , with the complex amplitude determining the vortex displacement. In Fig. 4 we show the numerically calculated BdG zero energy mode result and the functions (obtained from the vortex solution), showing they coincide.
IV.4 Numerical considerations
It is challenging to numerically calculate the unstable and zero modes accurately. One issue is that our grids are of finite spatial extent . This is problematic for the zero energy mode which decays rather slowly [noting that , see Eq. (46)]. Additionally, the 2D radial Laplacian is difficult to evaluate accurately with finite differences, particularly near Arsoski et al. (2015); Laliena and Campo (2018). In the BdG equations for 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

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
(48) |
where is the trap depth, and is the radius. The initial state is a NSV centered at the origin, obtained by solving the radial GPE [Eq. (23) including ] in the flat-bottomed trap potential using the approach described in Sec. III. This state is interpolated onto a uniform -point 2D grid with spacing . 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 , and finally spatially filtered to the region within the trap. Typically adding this noise to the initial state causes a increase in the wavefunction norm and a 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 -spin density to illustrate the evolution of a normal-core NSV in Fig. 5. Initially is zero (to the level of the noise) but as the component vortices separate, clear structure develops. The vortex core in the component is filled by the component, and thus appears as a negative peak. Similarly the vortex core in the component appears with a positive 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 and [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 ).

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 () 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 as the time when first exceeds . In Fig. 6(b) we show obtained from simulations conducted over a range of values. Here we see that increases with for , and appears to diverge as approaches . These results are consistent with the BdG analysis [see Fig. 3] if we identify as scaling with , where 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
- Ho (1998) T.-L. Ho, Phys. Rev. Lett. 81, 742 (1998).
- Ohmi and Machida (1998) T. Ohmi and K. Machida, J. Phys. Soc. Jpn 67, 1822 (1998).
- Stamper-Kurn et al. (1998) D. M. Stamper-Kurn, M. R. Andrews, A. P. Chikkatur, S. Inouye, H.-J. Miesner, J. Stenger, and W. Ketterle, Phys. Rev. Lett. 80, 2027 (1998).
- Kawaguchi and Ueda (2012) Y. Kawaguchi and M. Ueda, Phys. Rep. 520, 253 (2012).
- Stamper-Kurn and Ueda (2013) D. M. Stamper-Kurn and M. Ueda, Rev. Mod. Phys. 85, 1191 (2013).
- Bray (1994) A. Bray, Adv. Phys. 43, 357 (1994).
- Sadler et al. (2006) L. E. Sadler, J. M. Higbie, S. R. Leslie, M. Vengalattore, and D. M. Stamper-Kurn, Nature 443, 312 (2006).
- Kudo and Kawaguchi (2015) K. Kudo and Y. Kawaguchi, Phys. Rev. A 91, 053609 (2015).
- Williamson and Blakie (2016a) L. A. Williamson and P. B. Blakie, Phys. Rev. Lett. 116, 025301 (2016a).
- Seo et al. (2015) S. W. Seo, S. Kang, W. J. Kwon, and Y.-i. Shin, Phys. Rev. Lett. 115, 015301 (2015).
- Seo et al. (2016) S. W. Seo, W. J. Kwon, S. Kang, and Y. Shin, Phys. Rev. Lett. 116, 185301 (2016).
- Chen et al. (2018) H.-R. Chen, K.-Y. Lin, P.-K. Chen, N.-C. Chiu, J.-B. Wang, C.-A. Chen, P.-P. Huang, S.-K. Yip, Y. Kawaguchi, and Y.-J. Lin, Phys. Rev. Lett. 121, 113204 (2018).
- Turner (2009) A. M. Turner, Phys. Rev. Lett. 103, 080603 (2009).
- Eto et al. (2011) M. Eto, K. Kasamatsu, M. Nitta, H. Takeuchi, and M. Tsubota, Phys. Rev. A 83, 063603 (2011).
- Williamson and Blakie (2016b) L. A. Williamson and P. B. Blakie, Phys. Rev. A 94, 063615 (2016b).
- Kasamatsu et al. (2016) K. Kasamatsu, M. Eto, and M. Nitta, Phys. Rev. A 93, 013615 (2016).
- Zibold et al. (2016) T. Zibold, V. Corre, C. Frapolli, A. Invernizzi, J. Dalibard, and F. Gerbier, Phys. Rev. A 93, 023614 (2016).
- Symes and Blakie (2017) L. M. Symes and P. B. Blakie, Phys. Rev. A 96, 013602 (2017).
- Isoshima (2002) T. Isoshima, J. Low Temp. Phys. 126, 431 (2002).
- Lovegrove et al. (2016) J. Lovegrove, M. O. Borgh, and J. Ruostekoski, Physical Review A 93, 033633 (2016).
- Leonhardt and Volovik (2000) U. Leonhardt and G. E. Volovik, JETP Letters 72, 46 (2000).
- Kang et al. (2017) S. Kang, S. W. Seo, J. H. Kim, and Y. Shin, Phys. Rev. A 95, 053638 (2017).
- Kang et al. (2019) S. Kang, S. W. Seo, H. Takeuchi, and Y. Shin, Phys. Rev. Lett. 122, 095301 (2019).
- Symes et al. (2018) L. M. Symes, D. Baillie, and P. B. Blakie, Phys. Rev. A 98, 063618 (2018).
- Lovegrove et al. (2014) J. Lovegrove, M. O. Borgh, and J. Ruostekoski, Phys. Rev. Lett. 112, 075301 (2014).
- Choi et al. (2012a) J.-Y. Choi, W. J. Kwon, and Y.-I. Shin, Phys. Rev. Lett. 108, 035301 (2012a).
- Choi et al. (2012b) J.-y. Choi, W. J. Kwon, M. Lee, H. Jeong, K. An, and Y.-i. Shin, New J. Phys. 14, 053013 (2012b).
- Gerbier et al. (2006) F. Gerbier, A. Widera, S. Fölling, O. Mandel, and I. Bloch, Phys. Rev. A 73, 041602(R) (2006).
- Leslie et al. (2009) S. R. Leslie, J. Guzman, M. Vengalattore, J. D. Sau, M. L. Cohen, and D. M. Stamper-Kurn, Phys. Rev. A 79, 043631 (2009).
- Mineev (1974) V. Mineev, Zh. Eksp. Teor. Fiz. 67, 263 (1974).
- Ji et al. (2008) A.-C. Ji, W. M. Liu, J. L. Song, and F. Zhou, Phys. Rev. Lett. 101, 010402 (2008).
- Lovegrove et al. (2012) J. Lovegrove, M. O. Borgh, and J. Ruostekoski, Phys. Rev. A 86, 013613 (2012).
- Ishino et al. (2013) S. Ishino, M. Tsubota, and H. Takeuchi, Phys. Rev. A 88, 063617 (2013).
- Bao and Cai (2013) W. Bao and Y. Cai, Kinet. Relat. Mod. 6, 1 (2013).
- Bradley and Anderson (2012) A. S. Bradley and B. P. Anderson, Phys. Rev. X 2, 041001 (2012).
- Takahashi et al. (2009) M. Takahashi, V. Pietilä, M. Möttönen, T. Mizushima, and K. Machida, Phys. Rev. A 79, 023618 (2009).
- Isoshima and Machida (1999) T. Isoshima and K. Machida, Phys. Rev. A 59, 2203 (1999).
- Simula et al. (2002) T. P. Simula, S. M. M. Virtanen, and M. M. Salomaa, Phys. Rev. A 65, 033614 (2002).
- Hayashi et al. (2013) S. Hayashi, M. Tsubota, and H. Takeuchi, Phys. Rev. A 87, 063628 (2013).
- Law et al. (2001) C. K. Law, C. M. Chan, P. T. Leung, and M.-C. Chu, Phys. Rev. A 63, 063612 (2001).
- Yukalov and Yukalova (2004) V. I. Yukalov and E. P. Yukalova, Laser Physics Letters 1, 50 (2004).
- Takeuchi et al. (2010) H. Takeuchi, S. Ishino, and M. Tsubota, Phys. Rev. Lett. 105, 205301 (2010).
- Suzuki et al. (2010) N. Suzuki, H. Takeuchi, K. Kasamatsu, M. Tsubota, and H. Saito, Phys. Rev. A 82, 063604 (2010).
- Abad et al. (2015) M. Abad, A. Recati, and S. Stringari, Eur. Phys. J. D 69, 126 (2015).
- Hoefer et al. (2011) M. A. Hoefer, J. J. Chang, C. Hamner, and P. Engels, Phys. Rev. A 84, 041605(R) (2011).
- Hamner et al. (2011) C. Hamner, J. J. Chang, P. Engels, and M. A. Hoefer, Phys. Rev. Lett. 106, 065302 (2011).
- Fujimoto and Tsubota (2012) K. Fujimoto and M. Tsubota, Phys. Rev. A 85, 033642 (2012).
- Zhu et al. (2015) Q. Zhu, Q.-f. Sun, and B. Wu, Phys. Rev. A 91, 023633 (2015).
- Kim et al. (2017) J. H. Kim, S. W. Seo, and Y. Shin, Phys. Rev. Lett. 119, 185302 (2017).
- Arsoski et al. (2015) V. Arsoski, N. Čukarić, M. Tadić, and F. Peeters, Comput. Phys. Commun. 197, 17 (2015).
- Laliena and Campo (2018) V. Laliena and J. Campo, J. Phys. A 51, 325203 (2018).
- Symes et al. (2016) L. M. Symes, R. I. McLachlan, and P. B. Blakie, Phys. Rev. E 93, 053309 (2016).
- Shin et al. (2004) Y. Shin, M. Saba, M. Vengalattore, T. A. Pasquini, C. Sanner, A. E. Leanhardt, M. Prentiss, D. E. Pritchard, and W. Ketterle, Phys. Rev. Lett. 93, 160406 (2004).
- Schmied et al. (2019) C.-M. Schmied, T. Gasenzer, and P. B. Blakie, Phys. Rev. A 100, 033603 (2019).