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

Synthetic half-integer magnetic monopole and single-vortex dynamics in spherical Bose-Einstein condensates

Xi-Yu Chen These authors contribute equally. Institute of Modern Physics and school of Physics, Northwest University, Xi’an 710127, China    Lijia Jiang These authors contribute equally. Institute of Modern Physics and school of Physics, Northwest University, Xi’an 710127, China Shaanxi Key Laboratory for Theoretical Physics Frontiers, Xi’an 710127, China Peng Huanwu Center for Fundamental Theory, Xi’an 710127, China    Wen-Kai Bai Institute of Modern Physics and school of Physics, Northwest University, Xi’an 710127, China Shaanxi Key Laboratory for Theoretical Physics Frontiers, Xi’an 710127, China Peng Huanwu Center for Fundamental Theory, Xi’an 710127, China    Tao Yang yangt@nwu.edu.cn Institute of Modern Physics and school of Physics, Northwest University, Xi’an 710127, China Shaanxi Key Laboratory for Theoretical Physics Frontiers, Xi’an 710127, China Peng Huanwu Center for Fundamental Theory, Xi’an 710127, China    Jun-Hui Zheng junhui.zheng@nwu.edu.cn Institute of Modern Physics and school of Physics, Northwest University, Xi’an 710127, China Shaanxi Key Laboratory for Theoretical Physics Frontiers, Xi’an 710127, China Peng Huanwu Center for Fundamental Theory, Xi’an 710127, China
Abstract

Magnetic monopoles are crucial in explaining the quantization of electric charges and quantum hall effects, while artificially creating a minimal magnetic monopole in experiments remains a challenge. Here, we come up with a flexible way to simulate a half-integer-type monopole in Bose gases and investigate the induced vortex dynamics on a sphere. We list the possible experiment parameter settings for different isotopes and discuss their experimental feasibility. With the assumption of a rigid monopole-vortex structure, we analytically predict the vortex trajectory in an external magnetic field. We then confirm the result by numerically solving the Gross-Pitaevskii equation, which employs two gauges simultaneously (the Wu-Yang approach) to prevent singularity in the one-gauge method when a monopole is present. The study offers significant insight into the characteristics of monopoles and vortices, facilitating avenues for experimental validation.

I Introduction

Magnetic monopoles, which are particles with isolated magnetic charges, have been a subject of speculation in theoretical physics. Although the existence of elementary magnetic monopoles remains a mystery [1, 2, 3, 4, 5, 6, 7], analogues emerge in condensed matter systems, for instance, on the interfaces of rotating superfluid mixtures [8], at the ends of spin chains in spin ices [9, 10, 11, 12] and that of skyrmion-line excitations in chiral magnets [13], or in the momentum space of quantum Hall systems [14]. Monopoles can also be mimicked by creating a thin magnetic needle [15], the Berry curvature of a varying superconducting qubit state [16], or through magneto-optical effects [17].

The existence of magnetic monopoles implies that the magnetic field is no longer sourceless, i.e., 𝑩0\nabla\cdot\bm{B}\neq 0. To retain the electromagnetic vector potential 𝑨(𝒓)\bm{A}(\bm{r}) in this scenario, it is necessary to introduce a singular line in 𝑨(𝒓)\bm{A}(\bm{r}), known as the Dirac string. The Dirac string shows its physical effects, when a charged particle (with electric charge qeq_{e}) moves around the string closely. The particle acquires an Aharonov-Bohm phase eiqeΦe^{iq_{e}\Phi}, where Φ=4πgm\Phi=4\pi g_{m} is the magnetic flux (gmg_{m} is the magnetic charge) passing through a surface (which does not intersect the Dirac string) bounded by the particle’s trajectory, as illustrated in Fig.1. Crucially, since the Dirac string is hypothetical and unobservable, it must yield no detectable consequences. This requires the phase factor to satisfy qeΦ=2nπq_{e}\Phi=2n\pi with nn\in\mathbb{Z} (in natural units), leading directly to the Dirac charge quantization condition: 2gmqe2g_{m}q_{e}\in\mathbb{Z} [1, 2]. This duality explains why electric charges are quantized in the universe if a magnetic monopole exists, i.e. qe=n/2gmq_{e}=n/2g_{m}. Conversely, it also restricts magnetic charges to be integer multiples of 1/2qe1/2q_{e}.

Refer to caption
Figure 1: Dirac string and the Wu-Yang approach. When the particle moves around the string closely, the Aharonov-Bohm phase is eiqeΦe^{iq_{e}\Phi}, where Φ=4πgm\Phi=4\pi g_{m} is the magnetic flux. In the Wu-Yang approach, two different gauges are needed to avoid the Dirac string.

In experiments of cold atom simulation with Bose-Einstein condensates (BEC), the effective magnetic field originates from the Berry curvature, which stems from the adiabatic evolution of the atomic spin state |χ(𝒓)|\chi(\bm{r})\rangle in space. The corresponding Berry connection, 𝑨(𝒓)=χ(𝒓)|i|χ(𝒓)\bm{A}(\bm{r})=\langle\chi(\bm{r})|i\nabla|\chi(\bm{r})\rangle, serves as an effective gauge field, which introduces a geometric phase ei𝑨(𝒓)𝑑𝒓e^{i\int\bm{A}(\bm{r})\cdot d\bm{r}} into the atomic spatial wavefunction, depending on the path. By analogy with the form of the Aharonov-Bohm effect, the effective charge in the geometric phase is identified to qe=1q_{e}=1. Synthetic magnetic monopoles can be constructed by aligning the spin of the atoms (with the quantum number FF) with a special configuration in space so that moving atoms experience a Berry curvature corresponding to a monopole with gm=Fg_{m}=F [18]. From the quantization condition, the allowed magnetic charge is either an integer or a half-integer. However, magnetic monopole realized in the BEC simulations so far, contains an integer magnetic charge [18, 19, 20, 21] since the atomic spin quantum number FF must be an integer for bosons. Whether and how half-integer-type magnetic monopoles, especially the minimal monopole, can be realized in boson gases remains an open question.

With a magnetic monopole situated in the center of a sphere, a uniform magnetic field perpendicular to the spherical surface is created. For the noninteracting case, the single-particle wave functions on the sphere surface, the so-called monopole harmonics, have been analytically obtained by Wu and Yang [3]. The eigenstate has vortices with a total vortex charge of 2gm2g_{m}. Especially for the gm=1g_{m}=1 bosonic case, the ground state contains two vortices [3] and these vortices can be manipulated to desired position [22]. In the interacting case, for a BEC system with an integer gmg_{m}, total 2gm2g_{m} vortices exist in the ground state and form stable lattice patterns, like Thomson plum pudding model [23]. So far, the vortex dynamics on the spherical BEC in the literature mainly focus on cases in the absence of magnetic monopoles, such as vortex-antivortex dynamics, BKT phase transition, and the formation of vortex clusters [24, 29, 26, 31, 32, 30, 25, 27, 28, 33, 34, 35]. Yet, it is unclear how a monopole influences the dynamics of vortex.

On the other hand, the spherical system with a monopole serves as an ideal framework for investigating fractional quantum hall effects [36] and vortex dynamics without boundary effects in a finite size [31]. It is imperative to examine the realization of the minimal monopole and the characteristics of the minimal vortex-monopole unit in the spherical system, which underpins the study of intricate vortex structures in the presence of monopoles with high magnetic charges. In this article, we explore the realization of a magnetic monopole with a half-integer magnetic charge in a cold atom system, as well as the dynamics of a single vortex (in the presence of a minimal monopole) driven by an external field. The rest of the article is organized as follows:

In Sec. II, we propose a flexible scheme to realize a magnetic monopole with a half-integer magnetic charge in the alkali- or hydrogen-atom system. The key steps are to freeze the electron’s angular momentum (𝑱)(\bm{J}) and polarize the nuclear spin (with a half-integer quantum number II) aligning to the direction of the atom’s position vector, leading to an effective radial magnetic field corresponding to a monopole field of gm=Ig_{m}=I.

In Sec. III, we develop the ground state of spherical BEC in the presence of a monopole with a minimal charge gm=1/2g_{m}=1/2. For strong interaction, we take the trial wave function ψ(θ,ϕ)tanh(bθ)eiϕ\psi(\theta,\phi)\propto\tanh(b\theta)e^{i\phi} for the ground state with a vortex locating at the north pole (in spherical coordinates). The fitting function agrees very well with the results obtained from the numerical calculation. We generalize the solution by rotating the vortex core to an arbitrary position. We demonstrate that each atom in this vortex-monopole system possesses angular momentum of magnitude /2{\hbar}/{2} orientating to the vortex position.

In Sec. IV, we study the vortex dynamics of spherical BEC with a monopole gm=1/2g_{m}=1/2 and an additional external electromagnetic field. Assuming a rigid vortex-monopole configuration, we demonstrate that the vortex-monopole structure performs a precession motion and develop an equation of motion to describe the trajectory of the vortex center.

In Sec. V, we employ the Gross-Pitaevskii equation (GPE) to simulate the vortex dynamics in a 3D shell-shaped system with a magnetic monopole, where a ‘bi-gauge evolution’ approach is developed to avoid singularities of the gauge fields in the numerical calculation. The numerical results are consistent with our theoretical prediction.

Our work offers a scheme for the realization of magnetic monopoles with half-integer charges, validates the point-vortex picture for the strong interaction in the presence of monopoles, and advances the study of the dynamics of the simplest vortex-monople structure.

II Realizing monopoles of half-integer magnetic charges

In this section, we present the scheme of realizing monopoles with half-integer magnetic charges. Considering an atom in the presence of a magnetic field 𝑩\bm{B}, the Hamiltonian describing the energy level splitting for the internal degree of freedom of atoms contains the hyperfine interaction and Zeeman energies,

H^spin=A𝑰^𝑱^+gμB𝑩𝑱^(μI/I)𝑩𝑰^,\hat{H}_{\text{spin}}=A\bm{\hat{I}}\cdot\bm{\hat{J}}+g\mu_{B}\bm{B}\cdot\bm{\hat{J}}-(\mu_{I}/I)\bm{B}\cdot\bm{\hat{I}}, (1)

where AA is a constant, 𝑰^\bm{\hat{I}} is the nuclear spin, 𝑱^\bm{\hat{J}} denotes the total angular momentum of electron in the atom, g2g\simeq 2 is the Landé gg-factor, μB\mu_{B} is the Bohr magneton, and μI𝑰^/I\mu_{I}\bm{\hat{I}}/I is the nuclear magnetic moment [37]. The total angular momentum of electron is 𝑱^=𝑳^+𝑺^\bm{\hat{J}}=\bm{\hat{L}}+\bm{\hat{S}}, where 𝑳^\bm{\hat{L}} is the electron’s orbital angular momentum and 𝑺^\bm{\hat{S}} is the electron spin. The atomic spin is defined as 𝑭^=𝑰^+𝑱^\bm{\hat{F}}=\bm{\hat{I}}+\bm{\hat{J}}. For a ground-state alkali or hydrogen atom, the electrons have no orbital angular momentum (L=0L=0), so the electron’s total angular momentum quantum number is equal to its spin quantum number, that is J=S=1/2J=S=1/2 [37]. In this scenario, if the atom is a boson, the nuclear spin must be half-integer, resulting in an integer atomic spin quantum number FF. To realize the half-integer type of magnetic monopoles in these bosonic systems, we adopt the following two steps:

First, a strong magnetic field 𝑩0=B0𝒆z\bm{B}_{0}=B_{0}{\bm{e}}_{z} with a strength of B0AI/gμBB_{0}\gg AI/g\mu_{B} is applied to the system, so that the second term in Eq.(1) is much larger than the hyperfine interaction. Note that the nuclear magnetic moment is three orders of magnitude smaller than the electron magnetic moment. Thus, the second term of H^spin\hat{H}_{\text{spin}} is dominant, and the angular momentum 𝑱\bm{J} is locked to the state |J,Jz|J,-J\rangle_{{z}}. In this manifold, the second term becomes gμBB0J-g\mu_{B}B_{0}J, and AI^zJ^zA{\hat{I}_{z}}{\hat{J}_{z}} from the first term becomes AI^z/2A\hat{I}_{z}/2 (here, J=1/2J=1/2 has been used). The other terms AI^xJ^x+AI^yJ^y=A(I^+J^+I^J^+)/2A\hat{I}_{x}\hat{J}_{x}+A\hat{I}_{y}\hat{J}_{y}=A(\hat{I}_{+}\hat{J}_{-}+\hat{I}_{-}\hat{J}_{+})/2 couples the high energy manifold |J,Jz|J,J\rangle_{{z}}, which will contribute a second order correction. Together with the last term, (μI/I)B0I^z-(\mu_{I}/I)B_{0}{\hat{I}_{z}}, the low-energy effective Hamiltonian in the low-energy manifold |J,Jz|J,-J\rangle_{{z}}, up to the second order approximation, becomes

H^spin\displaystyle\hat{H}_{\text{spin}} \displaystyle\simeq ωLI^zA24gμBB0J,J|J^I^+I^J^+|J,Jzz\displaystyle-\omega_{L}\hat{I}_{z}-\frac{A^{2}}{4g\mu_{B}B_{0}}{{}_{z}\langle J,-J|\hat{J}_{-}\hat{I}_{+}\hat{I}_{-}\hat{J}_{+}|J,-J\rangle_{{z}}} (2)
gμBB0J,\displaystyle-g\mu_{B}B_{0}J,

where ωL=A/2+μIB0/I\omega_{L}=A/2+\mu_{I}B_{0}/I. Using the fact that J=1/2J=1/2, we have

J,J|J^I^+I^J^+|J,Jzz=I2I^z2+I^z,{{}_{z}\langle J,-J|\hat{J}_{-}\hat{I}_{+}\hat{I}_{-}\hat{J}_{+}|J,-J\rangle_{{z}}}=I^{2}-\hat{I}_{z}^{2}+\hat{I}_{z}, (3)

Considering the condition B0AI/gμBB_{0}\gg AI/g\mu_{B} and after discarding a constant term, we approximately have

H^spinωLI^z,\hat{H}_{\text{spin}}\simeq-\omega_{L}\hat{I}_{z}, (4)

which represents the energy splitting for the nuclear spin. This is different from the weak magnetic field case, where the low-energy physics is captured by the Zeeman splitting of the total spin 𝑭\bm{F} [19]. Note that we have neglected the second term (the second order correction) of Eq. (2) for simplicity, which contributes a linear I^z\propto\hat{I}_{z} and a quadratic term I^z2\propto\hat{I}_{z}^{2}. The linear term slightly corrects the magnitude of ωL\omega_{L} and does not change the form (4). The quadratic term can be compensated by the Floquet mechanics as shown below. For 𝑰=1/2\bm{I}=1/2, which is our main focus for generating minimal monopoles, it is always a two-level system and the quadratic term becomes a constant term, so the second term of Eq. (2) just contributes a correction to the splitting ωL\omega_{L}.

Second, similar to the proposal in Ref. [23], a periodic oscillating quadrupolar field

𝑩1=B1(14λ0cosωt)(x𝒆x+y𝒆y2z𝒆z),\bm{B}_{1}=B^{\prime}_{1}(1-4\lambda_{0}\cos{\omega t})(x\bm{e}_{x}+y\bm{e}_{y}-2z\bm{e}_{z}), (5)

is applied. The quadratic term mentioned in above can be compensated by a proper choice of λ0\lambda_{0} [23]. Here we focus on λ0=1\lambda_{0}=1. We require ω=ωL\omega=\omega_{L} to couple different nuclear spin states resonantly. In the rotating frame of

U^=exp(iωLtI^z),\hat{U}=\exp(i\omega_{L}t\hat{I}_{z}), (6)

the Hamiltonian becomes

H^I=U^[ωLI^z(μI/I)𝑩1𝑰^]U^iU^tU^.\hat{H}_{I}=\hat{U}^{\dagger}\big{[}-\omega_{L}\hat{I}_{z}-(\mu_{I}/I)\bm{B}_{1}\cdot\bm{\hat{I}}\big{]}\hat{U}-i\hat{U}^{\dagger}\partial_{t}\hat{U}. (7)

The low-energy effective Hamiltonian in the rotating wave approximation [23], where ωLμIB1L/I\omega_{L}\gg\mu_{I}B^{\prime}_{1}L/I being much larger than all other energy scales (LL is the typical size of the BEC and B1LB^{\prime}_{1}L charactorizes the magnitude of the quadrupolar magnetic field), becomes

H^I=2μIIB1𝑰^𝒓.\hat{H}_{I}=2\frac{\mu_{I}}{I}B^{\prime}_{1}\bm{\hat{I}}\cdot\bm{{r}}. (8)

When the two magnetic fields 𝑩0\bm{B}_{0} and 𝑩1\bm{B}_{1} are applied, the lowest spin state reads

|χ(𝒓)±=|J,Jz|I,Ir±,|\chi(\bm{r})\rangle_{\pm}=|J,-J\rangle_{z}\otimes|I,-I\rangle_{{r}}^{\pm}, (9)

where

|I,Ir±=eiIϕexp(iI^zϕ)exp(iI^yθ)|I,Iz|I,-I\rangle_{{r}}^{\pm}=e^{\mp iI\phi}\exp(-i\hat{I}_{z}\phi)\exp(-i\hat{I}_{y}\theta)|I,-I\rangle_{z} (10)

is the common eigenstate of 𝑰^2\bm{\hat{I}}^{2} and 𝑰^𝒆r\bm{\hat{I}}\cdot\bm{e}_{r}. Here, (θ,ϕ)(\theta,\phi) are the polar and azimuthal angles, respectively. In contrast to the integer II case, for a half-integer II, the phase e±iIϕe^{\pm iI\phi} is necessary to ensure |I,Ir±|I,-I\rangle_{{r}}^{\pm} remains the same value after the transformation ϕϕ+2π\phi\rightarrow\phi+2\pi, where ±\pm labels different gauge choices [23].

Now we consider the full single-particle Hamiltonian, which includes HspinH_{\text{spin}}, the external trap potential VH(r){V}_{H}(r), and the kinetic energy. It reads

H^=12m2+VH(r)+H^spin,\hat{H}=-\frac{1}{2m}\nabla^{2}+{V}_{H}(r)+\hat{H}_{\text{spin}}, (11)

where the external trap potential is VH(r)=mωH2𝒓2/2{V}_{H}(r)=m\omega^{2}_{H}\bm{{r}}^{2}/2. Both magnetic fields, 𝑩0\bm{B}_{0} and 𝑩1\bm{B}_{1}, are applied. In the adiabatic approximation of the lowest spin manifold (9), the effective Hamiltonian governing the dynamics of atoms in space is

H^±=(i𝑨±)2/2m+V(r),\hat{H}_{\pm}=(-i\nabla-\bm{{A}}_{\pm})^{2}/2m+{V}(r), (12)

where the vector potential is

𝑨±(𝒓)=χ(𝒓)|i|χ(𝒓)±±=(±1cosθ)Irsinθ𝒆ϕ,\bm{A}_{\pm}(\bm{r})={{}_{\pm}\langle}\chi(\bm{r})|{i\nabla}|\chi(\bm{r})\rangle_{\pm}=\frac{(\pm 1-\cos{\theta})I}{r\sin{\theta}}\bm{e}_{\phi}, (13)

and the scalar potential V(r)=VH(r)2μIB1r+I/2mr2{V}(r)={V}_{H}(r)-2\mu_{I}B^{\prime}_{1}r+I/2mr^{2} makes the BEC shell-shaped [23]. The reorientation of the spin state during the motion of the atom induces a gauge field and corrects the scalar potential. The effective magnetic field is 𝑩=×𝑨=I𝒆r/r2\bm{B}=\nabla\times\bm{A}=I{\bm{e}}_{r}/r^{2}. This field exactly corresponds to an effective magnetic charge of gm=Ig_{m}=I, which is half-integer in our systems.

In summary, to freeze the degree of freedom of the angular momentum of electrons, it requires a strong field B0BcAI/gμBB_{0}\gg B_{c}\equiv AI/g\mu_{B}, and to employ the rotating wave approximation, it requires the drive frequency to be much larger than other energy scales of the quadrupolar field, ωL=A/2+μIB0/IωcμIB1L/I\omega_{L}=A/2+\mu_{I}B_{0}/I\gg\omega_{c}\equiv\mu_{I}B^{\prime}_{1}L/I. On the other hand, to use the adiabatic approximation for the lowest spin manifold, it is necessary to require the system’s temperature is much smaller than the energy splitting ωc\hbar\omega_{c}. These three conditions are the key to realize a half-integer type of magnetic monopoles. In Table 1, we list the value of BcB_{c} using the known experimental data, estimate the value of ωc\omega_{c} by assuming that the typical size of the BEC is L=100μL=100~{}\mum and the magnetic field gradient is B1=0.5T/mB^{\prime}_{1}=0.5~{}\text{T/m}, and evaluate ωL\omega_{L} by setting B0=1TB_{0}=1~{}\text{T}, for different isotopes with a nuclear spin I=1/2I=1/2, 3/23/2, 5/25/2, and 7/27/2, respectively. The table shows that in this set, both conditions B0BcB_{0}\gg B_{c} and ωLωc\omega_{L}\gg\omega_{c} are satisfied. Indeed, ωL\omega_{L} is of the order of A/2A/2, which is always much larger than ωc\omega_{c}. So B0BcB_{0}\gg B_{c} is the only factor that should be taken into account. For hydrogen, ωc100nK\hbar\omega_{c}\simeq 100~{}\text{nK}, a uniform field with B0=0.25TB_{0}=0.25~{}\text{T} (about ten times of BcB_{c}) is sufficient to induce a minimal monopole of gm=1/2g_{m}=1/2. For K39{}^{39}\text{K}, K41{}^{41}\text{K},Rb85{}^{85}\text{Rb},Cs133{}^{133}\text{Cs}, ωc\hbar\omega_{c} is very small. Therefore, the required temperature is too low to be achievable in the current setting. A strategy to overcome this issue is to increase the radius of the shell-shaped BEC so that the energy splitting ωc\hbar\omega_{c} can be enhanced. Another concern is about gravitational effects, which may distort the shape of the BEC shell. A strategy to weaken this effect involves using optical trapping to counteract the gravitational potential or employing a microgravity environment.

Isotope II A{A}[MHz] μI/μN{\mu_{I}}/{\mu_{N}} BcB_{c}[mT] ωc\omega_{c}[kHz] ωL\omega_{L}[MHz]
1H 1/2 8925 2.793 25.4 13.38 4730
7Li 3/2 2524 3.256 21.5 5.20 1366
23Na 3/2 5566 2.218 47.5 3.54 2854
39K 3/2 1451 0.391 12.4 0.62 738
41K 3/2 798 0.215 6.8 0.34 406
87Rb 3/2 21472 2.751 183.1 4.39 10824
85Rb 5/2 6358 1.353 90.4 1.30 3205
133Cs 7/2 14440 2.579 287.4 1.76 7255
Table 1: The parameter estimation of realizing half-integer magnetic monopoles for different isotopes. The data for II, AA, and μI/μN\mu_{I}/\mu_{N} are selected from Refs.[38, 39, 37], where μN\mu_{N} is the nuclear magneton. BcAI/gμBB_{c}\equiv AI/g\mu_{B} and ωcμIB1L/I\omega_{c}\equiv\mu_{I}B^{\prime}_{1}L/I are evaluated accordingly, assuming the typical size of BEC to be L=100μL=100~{}\mum and B1=0.5T/mB^{\prime}_{1}=0.5~{}\text{T/m}[19, 20]. ωL=A/2+μIB0/I\omega_{L}=A/2+\mu_{I}B_{0}/I is obtained by setting B0=1B_{0}=1 T. To realize monopoles, it requires B0BcB_{0}\gg B_{c}, ωLωc\omega_{L}\gg\omega_{c}, and ωckBT\hbar\omega_{c}\gg k_{B}T.

III Single vortex on a sphere

In this section, we calculate the ground state of the spherical BEC in the presence of a minimal monopole with I=1/2I=1/2. The dynamical behaviors of the BEC are described by the GPE,

itψ±=H^0ψ±,i\partial_{t}{\psi_{\pm}}=\hat{H}_{0}\psi_{\pm}, (14)

where H^0=H^±+λ3D|ψ±|2\hat{H}_{0}=\hat{H}_{\pm}+\lambda_{3D}|\psi_{\pm}|^{2} contains the mean-field contribution from the contact atom-atom interaction, and the wavefunction is normalized to one in the whole space. To avoid Dirac strings—the singularities—from 𝑨±\bm{A}_{\pm}, the two gauges are utilized for ψ±\psi_{\pm} in the regions

R+\displaystyle R_{+} \displaystyle\equiv {(θ,ϕ)|0θπ/2+δ,0ϕ<2π},\displaystyle\{(\theta,\phi)|0\leq\theta\leq\pi/2+\delta,0\leq\phi<2\pi\}, (15)
R\displaystyle R_{-} \displaystyle\equiv {(θ,ϕ)|π/2δθπ,0ϕ<2π},\displaystyle\{(\theta,\phi)|\pi/2-\delta\leq\theta\leq\pi,0\leq\phi<2\pi\}, (16)

respectively, where 0<δ<π20<\delta<\frac{\pi}{2} [3]. We sketch this Wu-Yang approach in Fig. 1. In the overlap region, R+RR_{+}\cap R_{-}, the wave functions are connected by the U(1)U(1) gauge transformation

S=e2iIϕ:ψ+=Sψ,S=e^{2iI\phi}:~{}~{}\psi_{+}=S\psi_{-}, (17)

and accordingly, 𝑨+=𝑨ilnS{\bm{A}}_{+}={\bm{A}}_{-}-i\nabla\ln{S} [2]. Note the density distribution is gauge invariant, i.e., |ψ+|2=|ψ|2|\psi_{+}|^{2}=|\psi_{-}|^{2}.

We calculate the ground-state wave function when l0μIB1/mωH21/mωHl_{0}\equiv\mu_{I}B^{\prime}_{1}/m\omega_{H}^{2}\gg 1/\sqrt{m\omega_{H}}, in which case the BEC is narrow shell-shaped with radius r0=2l0r_{0}=2l_{0} and thickness 1/mωH\sim 1/\sqrt{m\omega_{H}} [23]. For λ3D=0\lambda_{3D}=0, the ground state of I=1/2I=1/2 can be constructed from the monopole harmonics [3], i.e.,

ψ+(θ,ϕ;θv,ϕv)=1π[ei(ϕϕv)sinθ2cosθv2cosθ2sinθv2],\psi_{+}(\theta,\phi;\theta_{v},\phi_{v})=\frac{1}{\sqrt{\pi}}\big{[}e^{i(\phi-\phi_{v})}\sin\frac{\theta}{2}\cos\frac{\theta_{v}}{2}-\cos\frac{\theta}{2}\sin\frac{\theta_{v}}{2}\big{]}, (18)

which vanishes at the vortex’s core position (θv,ϕv)(\theta_{v},\phi_{v}) and its eigen-energy is 1/4mr021/4mr_{0}^{2}. On the other sector, ψ\psi_{-} is obtained via the gauge transformation: ψ=S1ψ+\psi_{-}=S^{-1}\psi_{+}.

In a closed manifold with monopoles inside, a 2π2\pi phase change of the wave function along a loop on the manifold does not always indicate a vortex, since it may be eliminated by a gauge transformation. Note that the velocity field is gauge invariant. Specifically for the case θv=ϕv=0\theta_{v}=\phi_{v}=0, we have

𝒗=1m[argψ𝑨]=1m(1+cosθ)2r0sinθ𝒆ϕ.\bm{v}=\frac{1}{m}[\nabla{\arg\psi}-\bm{A}]=\frac{1}{m}\frac{(1+\cos{\theta})}{2r_{0}\sin{\theta}}\bm{e}_{\phi}. (19)

The singularity of ×𝒗\nabla\times\bm{v} indicates the position of the vortex. The winding number (i.e., the vortex charge) is given by the limit of the loop integral surrounding the vortex: liml0(m/2π)l𝒗𝑑𝒓=1\lim_{l\rightarrow 0}({m}/{2\pi\hbar})\oint_{l}\bm{v}\cdot d\bm{r}=1. Beyond the vortex center, the vorticity ×𝒗=𝑩/m\nabla\times\bm{v}=-\bm{B}/m is nonzero everywhere due to the presence of a monopole. The velocity field 𝒗\bm{v} cannot be expressed as the gradient of a scalar function with singular points, unlike in a system without monopoles [24].

Note that the system is symmetric under rotations with generators 𝑳^=𝒓×𝒑^I𝒓/r\bm{\hat{L}}=\bm{r}\times\bm{\hat{p}}-I{\bm{r}}/r, where 𝒑^=i𝑨±\bm{\hat{p}}=-i\nabla-\bm{{A}}_{\pm} [3]. The appearance of vortices breaks the rotational symmetry spontaneously.

For interacting cases, the ground state with a vortex locating at θv=ϕv=0\theta_{v}=\phi_{v}=0, in general, has the form

ψ+(θ,ϕ;θv=0,ϕv=0)=1r0f(θ)eiϕeiμt/mr02\psi_{+}(\theta,\phi;\theta_{v}=0,\phi_{v}=0)=\frac{1}{r_{0}}f(\theta)e^{i\phi}e^{-i\mu t/mr_{0}^{2}} (20)

and ψ=S1ψ+\psi_{-}=S^{-1}\psi_{+}. We will also use Dirac notation ψ±(θ,ϕ;θv,ϕv)θ,ϕ|ψ±(θv,ϕv)\psi_{\pm}(\theta,\phi;\theta_{v},\phi_{v})\equiv\langle\theta,\phi|\psi_{\pm}(\theta_{v},\phi_{v})\rangle. The velocity field is the same as Eq. (19), which is completely independent of the interaction strength. From GPE, we find that the amplitude f(θ)f(\theta) satisfies the gauge independent equation

μf=12[θ2cotθθ+1+cosθ4(1cosθ)]f+λ|f|2f,\mu f=\frac{1}{2}\Big{[}-\partial_{\theta}^{2}-\cot\theta\partial_{\theta}+\frac{1+\cos\theta}{4(1-\cos\theta)}\Big{]}f+\lambda|f|^{2}f, (21)

where λ=mλ2D\lambda=m\lambda_{2D} is dimensionless, λ2D\lambda_{2D} is the effective two-dimensional interaction strength on the sphere (see Appendix A), and the normalization condition is 2π|f|2sinθdθ=12\pi\int|f|^{2}\sin\theta d\theta=1.

Refer to caption
Figure 2: (a) The ground-state wave functions for different inter-atom interaction strengths (λ\lambda) [symbols for the numerical results of Eq. (21) and solid lines for the fitting results of the trial function]; (b) the chemical potential μ\mu and (c) the fitting parameters as functions of λ\lambda. For large λ\lambda, we have b20.051λ+0.76b^{2}\simeq 0.051\lambda+0.76.

Approximately, we employ a trial solution f(θ)=atanh(bθ)f(\theta)=a\tanh(b\theta) with fitting parameters aa and bb. The fitting results agree very well with the numerical results of Eq. (21) especially for strong interaction cases, as shown in Fig. 2. For a small θ\theta, the square of the slope of f(θ)f(\theta), i.e., a2b2a^{2}b^{2}, linearly increases with λ\lambda, like the vortex solution in the plane case [37]. For large interactions, the density away from the vortex center becomes uniform and the vortex size is small. Consequently, a21/4πa^{2}\simeq 1/4\pi and μλ/4π\mu\simeq\lambda/4\pi (in the Thomas-Fermi approximation). The numerical result of μ\mu [see Fig. 2(b)] is a bit larger because the velocity fields are finite. For large λ\lambda, from Fig. 2(c), we have b20.051λ+0.76b^{2}\simeq 0.051\lambda+0.76 from the numerical fitting by replacing a2a^{2} with 1/4π1/4\pi. Therefore, the healing length θc=1/b\theta_{c}=1/b decreases with the interaction strength. Correspondingly, the total particle number missing in the vortex decreases as well, and for θc1\theta_{c}\ll 1, we have nmiss2π(a2[f(θ)]2)sinθdθ0.347θc2n_{\text{miss}}\equiv 2\pi\int(a^{2}-[f(\theta)]^{2})\sin\theta d\theta\simeq 0.347\theta_{c}^{2}.

The solution with the vortex core at an arbitrary position (θv,ϕv)(\theta_{v},\phi_{v}) is obtained by the rotation transformation,

ψ+(θ,ϕ;θv,ϕv)=exp(iL^zϕv)exp(iL^yθv)ψ+(θ,ϕ;0,0).\psi_{+}(\theta,\phi;\theta_{v},\phi_{v})=\exp(-i\hat{L}_{z}\phi_{v})\exp(-i\hat{L}_{y}\theta_{v})\psi_{+}(\theta,\phi;0,0). (22)

To engineer vortex at the target position in experiments, a small positive Gaussian potential with the center at 𝒓v=r0𝒏v\bm{r}_{v}=r_{0}{\bm{n}}_{v} can be employed, where 𝒏v=(sinθvcosϕv,sinθvsinϕv,cosθv){\bm{n}}_{v}=(\sin\theta_{v}\cos\phi_{v},\sin\theta_{v}\sin\phi_{v},\cos\theta_{v}) is a unit vector. The potential breaks the rotation symmetry of the system explicitly, and the system has minimal total energy when the vortex center coincides with 𝒓v\bm{r}_{v}.

In the following, we calculate the expectation of the angular momentum, which will be employed in the study of the dynamics of the vortex. For the special case with the vortex center locating at the north pole, the expectation reads

𝒍=ψ+(0,0)|𝑳^|ψ+(0,0),\bm{l}=\langle\psi_{+}(0,0)|\hat{\bm{L}}|\psi_{+}(0,0)\rangle, (23)

where the angular momentum operator in the presence of a monopole is

𝑳^\displaystyle\hat{\bm{L}} =\displaystyle= 𝒓×(i𝑨+)I𝒓/r\displaystyle\bm{r}\times(-i\nabla-\bm{A}_{+})-I\bm{r}/r (24)
=\displaystyle= i𝒆ϕθ+i𝒆θsinθϕ+𝒆θ1cosθ2sinθ𝒆r2.\displaystyle-i\bm{e}_{\phi}\frac{\partial}{\partial\theta}+i\bm{e}_{\theta}\frac{\partial}{\sin\theta\partial\phi}+\bm{e}_{\theta}\frac{1-\cos\theta}{2\sin\theta}-\frac{\bm{e}_{r}}{2}.

Using the wave function (20), we have

𝒍=sinθdθdϕf(θ)[i𝒆ϕθ𝒆θ1+cosθ2sinθ𝒆r12]f(θ).\bm{l}=\int\sin\theta d\theta d\phi f^{*}(\theta)\Big{[}-i\bm{e}_{\phi}\frac{\partial}{\partial\theta}-\bm{e}_{\theta}\frac{1+\cos\theta}{2\sin\theta}-\bm{e}_{r}\frac{1}{2}\Big{]}f(\theta). (25)

Now we consider the three components li=𝒆i𝒍l_{i}=\bm{e}_{i}\cdot\bm{l} for i=x,y,zi=x,y,z. Note that 𝒆x𝒆ϕ=sinϕ\bm{e}_{x}\cdot\bm{e}_{\phi}=-\sin\phi, 𝒆x𝒆θ=cosθcosϕ\bm{e}_{x}\cdot\bm{e}_{\theta}=\cos\theta\cos\phi, and 𝒆x𝒆r=sinθcosϕ\bm{e}_{x}\cdot\bm{e}_{r}=\sin\theta\cos\phi. Therefore, lxl_{x} vanishes since the integral over ϕ\phi is null. For the yy component, we have 𝒆y𝒆ϕ=cosϕ\bm{e}_{y}\cdot\bm{e}_{\phi}=\cos\phi, 𝒆y𝒆θ=cosθsinϕ\bm{e}_{y}\cdot\bm{e}_{\theta}=\cos\theta\sin\phi, and 𝒆y𝒆r=sinθsinϕ\bm{e}_{y}\cdot\bm{e}_{r}=\sin\theta\sin\phi. Similarly, we obtain ly=0l_{y}=0. For the zz component, we have 𝒆z𝒆ϕ=0\bm{e}_{z}\cdot\bm{e}_{\phi}=0, 𝒆z𝒆θ=sinθ\bm{e}_{z}\cdot\bm{e}_{\theta}=-\sin\theta , 𝒆z𝒆r=cosθ\bm{e}_{z}\cdot\bm{e}_{r}=\cos\theta. As a result,

lz=sinθdθdϕf(θ)[1+cosθ2cosθ2]f(θ)=12.l_{z}=\int\sin\theta d\theta d\phi f^{*}(\theta)\Big{[}\frac{1+\cos\theta}{2}-\frac{\cos\theta}{2}\Big{]}f(\theta)=\frac{1}{2}. (26)

Therefore, 𝒍=(0,0,1/2)\bm{l}=(0,0,1/2), which is the angular momentum for each atom. It is also easy to check that |ψ+(0,0)|\psi_{+}(0,0)\rangle is indeed the eigenstate of L^z\hat{L}_{z} with eigenvalue 1/21/2. We can generalize the expectation value of the angular momentum to the case with an arbitrary vortex position (θv,ϕv)(\theta_{v},\phi_{v}) by a rotation transformation (22), which gives 𝒍𝑳^=𝒏v/2{\bm{l}}\equiv\langle\hat{\bm{L}}\rangle=\bm{n}_{v}/2. Note that the angular momentum is interaction-independent and its direction coincides with the vortex position vector.

IV Single vortex-monopole dynamics

In this section, we study the dynamics of the single vortex induced by the minimal monopole in an additional uniform magnetic field. Without loss of generality, we suppose that the magnetic field is along the zz direction, 𝑩2=B2𝒆z{\bm{B}}_{2}=B_{2}{\bm{e}}_{z}, and the corresponding vector potential is δ𝑨=B2(y/2,x/2,0)\delta\bm{A}=B_{2}(-y/2,x/2,0). We assume that the monopole-vortex configuration is rigid. This is inspired by the Thiele equation for topological particles [40], such as magnetic skyrmions or Hopfions [41, 42]. These particles are typically treated as undeformed, particle-like entities that maintain their structure during motion, thereby simplifying their dynamics. Our assumption is also motivated by the analogy to vortex dynamics under strong interactions. When the healing length is small, a point vortex model is commonly employed to describe vortex motion [24], as phonon excitations are strongly suppressed in this regime. Our theoretical predictions, derived under this assumption, show excellent agreement with numerical results, thereby confirming its validity in the strong interaction region. For more general cases, where phonon excitations may deform the vortex profile and influence its trajectory, further research will be conducted to explore these effects.

Using this assumption, the evolving wavefunction becomes |ψ(t)|ψ±[θv(t),ϕv(t)]|\psi(t)\rangle\propto|\psi_{\pm}[\theta_{v}(t),\phi_{v}(t)]\rangle with a varying vortex center. With this ansatz, the system is described by the dynamics of the vortex center, or equivalently, the dynamics of the angular momentum 𝒍(t)=𝒏v(t)/2{\bm{l}}(t)={\bm{n}}_{v}(t)/2. From the GPE (14) and the definition 𝒍=ψ(t)|𝑳^|ψ(t)\bm{l}=\langle\psi(t)|\hat{\bm{L}}|\psi(t)\rangle, we obtain

𝒍˙=iψ(t)|[𝑳^,H^eff]|ψ(t),\dot{\bm{l}}=-i\langle\psi(t)|[\hat{\bm{L}},\hat{H}_{\text{eff}}]|\psi(t)\rangle, (27)

where the effective Hamiltonian

H^eff(δ𝑨)=(𝒑^δ𝑨)22m+V(r)+λ3D|ψ|2\hat{H}_{\text{eff}}(\delta\bm{A})=\frac{(\hat{\bm{p}}-\delta\bm{A})^{2}}{2m}+V(r)+\lambda_{3D}|\psi|^{2} (28)

contains the gauge fields from both the monopole and the external magnetic field. We can split the effective Hamiltonian into two parts,

H^eff(δ𝑨)=H^eff(δ𝑨=0)+ΔH^,\displaystyle\hat{H}_{\text{eff}}(\delta\bm{A})=\hat{H}_{\text{eff}}(\delta\bm{A}=0)+\Delta\hat{H}, (29)

where

ΔH^=(iδ𝑨)2δ𝑨𝒑^+δ𝑨δ𝑨2m\Delta\hat{H}=\frac{(i\nabla\cdot\delta\bm{A})-2\delta\bm{A}\cdot\hat{\bm{p}}+\delta\bm{A}\cdot\delta\bm{A}}{2m} (30)

is the correction contributed by the external field. Using the fact that the instantaneous |ψ(t)|\psi(t)\rangle is the eigen-wavefunction of H^eff(δ𝑨=0)\hat{H}_{\text{eff}}(\delta\bm{A}=0), we have

𝒍˙=iψ(t)|[𝑳^,ΔH^]|ψ(t).\dot{\bm{l}}=-i\langle\psi(t)|[\hat{\bm{L}},\Delta\hat{H}]|\psi(t)\rangle. (31)

Note that δ𝑨=0\nabla\cdot\delta\bm{A}=0, 2δ𝑨𝒑^=B2[L^z+(1/2)cosθ]2\delta\bm{A}\cdot\hat{\bm{p}}=B_{2}[\hat{L}_{z}+(1/2)\cos\theta], and δ𝑨δ𝑨=(1/4)B22r2sin2θ\delta\bm{A}\cdot\delta\bm{A}=(1/4)B_{2}^{2}r^{2}\sin^{2}\theta, the Hamiltonian becomes

ΔH^=αL^z+g(r,θ),\Delta\hat{H}=-\alpha\hat{L}_{z}+g(r,\theta), (32)

where α=B2/(2m)\alpha=B_{2}/(2m) is the Larmor precession frequency and the scalar potential is

g(r,θ)=B2cosθ4m+B22r2sin2θ8m.g(r,\theta)=-\frac{B_{2}\cos\theta}{4m}+\frac{B_{2}^{2}r^{2}\sin^{2}\theta}{8m}. (33)

Detailed calculations give

iψ(t)|[𝑳^,αL^z]|ψ(t)=αlx𝒆y+αly𝒆x=𝝎×𝒍,-i\langle\psi(t)|[\hat{\bm{L}},-\alpha\hat{L}_{z}]|\psi(t)\rangle=-\alpha l_{x}\bm{e}_{y}+\alpha l_{y}\bm{e}_{x}=\bm{\omega}\times\bm{l}, (34)

where 𝝎=α𝒆z\bm{\omega}=-\alpha\bm{e}_{z}, and

iψ(t)|[𝑳^,g(r,θ)]|ψ(t)\displaystyle-i\langle\psi(t)|[\hat{\bm{L}},g(r,\theta)]|\psi(t)\rangle
=i[𝒓×(i𝑨+)I𝒓/r,g(r,θ)]\displaystyle~{}~{}~{}~{}~{}~{}~{}=-i\langle[\bm{r}\times(-i\nabla-\bm{A}_{+})-I\bm{r}/r,g(r,\theta)]\rangle
=[𝒓×,g(r,θ)]=𝒆ϕθg(r,θ).\displaystyle~{}~{}~{}~{}~{}~{}~{}=-\langle[\bm{r}\times\nabla,g(r,\theta)]\rangle=-\langle\bm{e}_{\phi}\partial_{\theta}g(r,\theta)\rangle. (35)

If the corresponding density of the wave function |ψ(t)|\psi(t)\rangle is exactly uniform on the whole sphere, then we have 𝒆ϕθg(r,θ)=0\langle\bm{e}_{\phi}\partial_{\theta}g(r,\theta)\rangle=0 because 𝒆ϕ\bm{e}_{\phi} changes its direction along the circle of latitude. However, since the system has a small size of vacancy at the vortex center, and considering the shell-shaped case, rr0r\simeq r_{0}, we have

𝒆ϕθg(r,θ)=𝒆ϕvθg(r0,θ)|θ=θvnmiss.\langle\bm{e}_{\phi}\partial_{\theta}g(r,\theta)\rangle=-\bm{e}_{\phi_{v}}\partial_{\theta}g(r_{0},\theta)|_{\theta=\theta_{v}}n_{\text{miss}}. (36)

Therefore,

𝒍˙=𝝎×𝒍+𝝉,\dot{\bm{l}}=\bm{\omega}\times\bm{l}+\bm{\tau}, (37)

where 𝝉=θg(r0,θ)|θ=θvnmiss𝒆ϕv\bm{\tau}=\partial_{\theta}g(r_{0},\theta)|_{\theta=\theta_{v}}n_{\text{miss}}\bm{e}_{\phi_{v}} is a torque contributed by the potential g(r,θ)g(r,\theta). In the following, we provide an interpretation of this result from a classical perspective. The effective energy for a vortex in the potential g(r0,θ)g(r_{0},\theta) is E(θv)=g(r0,θv)nmissE(\theta_{v})=-g(r_{0},\theta_{v})n_{miss} due to the particle missing, then, the torque is 𝝉=𝒍E(θv)×𝒍\bm{\tau}=\nabla_{\bm{l}}E(\theta_{v})\times\bm{l}, which is exactly the same as the result given above. As a result, lzl_{z} is conserved during the evolution. Therefore, θv\theta_{v} remains a constant value, while the precession angular frequency is

ωmϕv˙(t)=α+θg(r0,θ)|θ=θvnmiss|𝒍|sinθv.\omega_{m}\equiv\dot{\phi_{v}}(t)=-\alpha+\frac{\partial_{\theta}g(r_{0},\theta)|_{\theta=\theta_{v}}n_{\text{miss}}}{|\bm{l}|\sin\theta_{v}}. (38)

where ωm=(1nmiss)B2/2m+nmissB22r02cosθv/2m\omega_{m}=-(1-n_{\text{miss}}){B_{2}}/{2m}+n_{\text{miss}}{B_{2}^{2}r_{0}^{2}\cos\theta_{v}}/{2m} arise from the magnetic field B2𝒆^zB_{2}\hat{\bm{e}}_{z}. The external magnetic field drives the precession of the vortex. ωm\omega_{m} is slightly different from the Larmor precession frequency due to the particle missing in the vortex. When considering the influence of a microgravity or an electric field along the zz direction, which gives an additional contribution in the scalar potential (cosθ\propto\cos\theta) in (33), the precession frequency will be further corrected.

V Numerical simulations

In this section, we numerically simulate the vortex dynamics in the presence of 𝑩2\bm{B}_{2} and compare it with the analytical results. For this purpose, we first develop the numerical approach. To avoid the singularities (Dirac strings) of the gauge field, we use the bi-gauge-evolution method to cover the whole sphere.

V.1 Bi-gauge evolution of the GPE

The dynamics of BEC in the presence of a monopole are described by GPE (14), where the regions for ψ+\psi_{+} and ψ\psi_{-} are R±R_{\pm}, respectively. The two wave functions in the overlapped region are linked by the gauge transformation ψ(r)=S1ψ+(r)\psi_{-}(\textbf{r})=S^{-1}\psi_{+}(\textbf{r}). We use the split-step Crank-Nicolson algorithm to solve the GPE. This approach has been introduced in details in Refs. [43, 44]. In the current system, we split the Hamiltonian into four parts: H±=H1±+H2±+H3±+H4±H_{\pm}=H_{1}^{\pm}+H_{2}^{\pm}+H_{3}^{\pm}+H_{4}^{\pm}, where

H1±\displaystyle H_{1}^{\pm} =\displaystyle= V+|𝑨±|22m+λ3D|ψ±|2,\displaystyle V+\frac{\left|\bm{A}_{\pm}\right|^{2}}{2m}+\lambda_{3D}\left|\psi_{\pm}\right|^{2}, (39)
H2±\displaystyle H_{2}^{\pm} =\displaystyle= 12mx2+i2Ax±x+i2xAx±,\displaystyle-\frac{1}{2m}\frac{\partial}{\partial x^{2}}+\frac{i}{2}A_{x}^{\pm}\frac{\partial}{\partial x}+\frac{i}{2}\frac{\partial}{\partial x}A_{x}^{\pm}, (40)
H3±\displaystyle H_{3}^{\pm} =\displaystyle= 12my2+i2mAy±y+i2myAy±,\displaystyle-\frac{1}{2m}\frac{\partial}{\partial y^{2}}+\frac{i}{2m}A_{y}^{\pm}\frac{\partial}{\partial y}+\frac{i}{2m}\frac{\partial}{\partial y}A_{y}^{\pm}, (41)
H4±\displaystyle H_{4}^{\pm} =\displaystyle= 12mz2+i2mAz±z+i2mzAz±.\displaystyle-\frac{1}{2m}\frac{\partial}{\partial z^{2}}+\frac{i}{2m}A_{z}^{\pm}\frac{\partial}{\partial z}+\frac{i}{2m}\frac{\partial}{\partial z}A_{z}^{\pm}. (42)

In the 3D numerical simulations, we use 150×150×150150\times 150\times 150 grids for space and discretize time into {,tn1,tn,}\{\cdots,t_{n-1},t_{n},\cdots\} with spacing Δt\Delta t. For each gauge choice, the evolution operator is split into four steps:

eiH±Δt=eiH4±ΔteiH3±ΔteiH2±ΔteiH1±Δt,e^{-iH_{\pm}\Delta t}=e^{-iH_{4}^{\pm}\Delta t}e^{-iH_{3}^{\pm}\Delta t}e^{-iH_{2}^{\pm}\Delta t}e^{-iH_{1}^{\pm}\Delta t}, (43)

and ψ¯±n+1=eiH±Δtψ±n\underline{\psi}_{\pm}^{n+1}=e^{-iH_{\pm}\Delta t}{\psi}_{\pm}^{n}. More details can be seen in Appendix B.

Refer to caption
Figure 3: Schematic illustration of the iteration

To avoid Dirac strings, we use two gauges in the two regions: R+R_{+} and RR_{-}. Each region has an open boundary. The boundary would induce non-physical effects on the profile of the wavefunction in numerical simulations. In order to get the correct behavior of the boson gas on the shell with a monopole, we bridge the two wave functions at each iteration:

ψ+n+1={ψ¯+n+1forz0Sψ¯n+1forz<0,{\psi}_{+}^{n+1}=\left\{\begin{array}[]{ccc}\underline{\psi}_{+}^{n+1}&&\text{for}~{}~{}z\geq 0\\ S\underline{\psi}_{-}^{n+1}&&\text{for}~{}~{}z<0,\end{array}\right. (44)

and

ψn+1={S1ψ¯+n+1forz>0ψ¯n+1forz0.{\psi}_{-}^{n+1}=\left\{\begin{array}[]{ccc}S^{-1}\underline{\psi}_{+}^{n+1}&&\text{for}~{}~{}z>0\\ \underline{\psi}_{-}^{n+1}&&\text{for}~{}~{}z\leq 0.\end{array}\right. (45)

These finish one iteration. We sketch the iteration loop in Fig. 3. The bi-gauge evolution eliminates the non-physical effects from boundary and overcomes the singularity in the one-gauge method, so we can get the correct wavefunction.

In the imaginary-time evolution simulation, we rigorously monitor the convergence of both the chemical potential and the wavefunction. We successfully reproduce the stable vortex lattice configuration in a spherical BEC with integer-type magnetic monopoles [23], thereby validating the numerical method.

Refer to caption
Figure 4: (a) The single-vortex core trajectory under a magnetic field: ϕv(t)\phi_{v}(t) (left) and θv(t)\theta_{v}(t) (right). All lines are the analytical results (38) of the evolving vortex center, and all symbols are 3D numerical simulation results for different magnetic field strengths B2B_{2} (unit [mωH][m\omega_{H}]), initial vortex core positions θv\theta_{v}, and radii r0r_{0} (unit /mωH\ell\equiv\sqrt{\hbar/m\omega_{H}}). Here, λ3D=13375ωH3\lambda_{3D}=13375~{}\hbar\omega_{H}\ell^{3}, tH=1/ωHt_{H}=1/\omega_{H}, and the initial ϕv\phi_{v} equals to π\pi. (b) The scalar shell potential V(r)V(r) for the case of r0=16.72r_{0}=16.72\ell. (c) The deviation in the normalization of the wavefunction during dynamical evolution, ΔN=𝑑𝒓|ψ(𝒓)|21\Delta_{N}=\int d\bm{r}|\psi(\bm{r})|^{2}-1. (d) The velocity distribution and the density profile of the BEC on the shell with a single vortex at (θv=π/3,ϕv=3π/2)(\theta_{v}=\pi/3,\phi_{v}=3\pi/2) and (e) the corresponding density profile on the x=0x=0 plane.

V.2 Simulation results

In this subsection, we present our 3D numerical simulations of vortex trajectory on a narrow shell. For strong interaction with a thin shell, we have bridged the 3D interaction strength and the effective 2D interaction strength in Appendix A, by using the Thomas-Fermi approximation along the radial direction. Consequently, we can compare the 3D numerical simulation with the 2D analytical results shown in Sec.IV. The initial state in the simulation is prepared by the imaginary-time evolution of the GPE [43, 44]. The initial position of the vortex can be controlled by adding a small positive Gaussian potential to pin it, as illustrated in Sec. III. The real-time dynamical simulation is implemented after switching on δ𝑨\delta\bm{A}. To prevent singularities in the simulation, we utilize the bi-gauge numerical method that we developed in the previous subsection.

In Fig. 4(a), we present the numerical results of the vortex trajectory. From the result of θv(t)\theta_{v}(t), we observe that θv(t)\theta_{v}(t) almost remains constant during the volution, while ϕv(t)\phi_{v}(t) changes linearly with time. This shows that the vortex experiences a precession around the zz-axis.

With a fixed sphere radius of r0=16.72r_{0}=16.72\ell (length unit =/mωH\ell=\sqrt{\hbar/m\omega_{H}}), we further confirm that as the external magnetic field B2B_{2} gets stronger, the vortex’s precession frequency gets stronger. For a vortex initially posed at θv=π/6\theta_{v}=\pi/6, the simulated results agree very well with our prediction for different magnitudes of B2=0.067{B_{2}}=0.067, 0.090.09, and 0.1120.112 (unit mωHm\omega_{H}).

To see how the initial position of the vortex influences the precession, for fixed B2=0.09mωH{B_{2}}=0.09~{}m\omega_{H}, we select different initial vortex positions at θv=π/6\theta_{v}=\pi/6, π/2\pi/2, and 5π/65\pi/6, respectively. The value of initial θv\theta_{v} slightly changes the precession frequency, and the results qualitatively agree with our analytic prediction.

We also try to check the effect of the shell radius. For B2=0.112mωH{B_{2}}=0.112~{}m\omega_{H}, we simulate the motion for the vortex (with initial position θv=π/6\theta_{v}=\pi/6) at the shell with radius r0=16.72r_{0}=16.72\ell and 8.368.36\ell, respectively. The results again qualitatively agree with our prediction in (38).

In Fig. 4(b-e), we present additional details of the simulation. Fig. 4(b) shows the scalar potential V(r)V(r) in Eq. (12) for the case of r0=16.72r_{0}=16.72\ell. In the real-time evolution, we monitor the normalization of the wavefunction, which should remain constant throughout the evolution. Fig. 4(c) illustrates the deviation of the normalization during dynamical evolution, defined as ΔN=𝑑𝒓|ψ(𝒓)|21\Delta_{N}=\int d\bm{r}|\psi(\bm{r})|^{2}-1. We confirm that the deviation is kept below 0.2%. Fig. 4(d) visualizes the velocity distribution and the density profile of the BEC on the shell with a single vortex at (θv=π/3,ϕv=3π/2)(\theta_{v}=\pi/3,\phi_{v}=3\pi/2), while Fig. 4(e) shows the corresponding density profile on the x=0x=0 plane.

VI Conclusion

In conclusion, we have developed a pioneering method that enables the realization of magnetic monopoles with half-integer charges in the BEC systems. We discussed the suggested parameters for the experiment setup and explored the experimental flexibility. We have systemically studied ground-state vortex solutions and single-vortex dynamics on a sphere, which includes analytical calculations under the assumption of a rigid vortex-monopole structure and numerical simulations using a novel method to prevent singularities of the gauge fields. Larmor precession with a modified frequency captures the vortex dynamics on the sphere in the presence of a uniform electromagnetic field.

We note that the spherical BEC itself has garnered intense attention due to its nontrivial geometric properties. The bubble system realized in microgravity environments is also a platform to simulate cosmic inflation [45, 47, 46, 51, 50, 48, 49]. The nontrivial topology and the finite size endow phase transitions in spherical systems with novel characters [45, 29]. On a sphere, the vortex and antivortex always appear in pairs when there are no monopoles [52]. The existence of monopoles enables the investigation of vortex dynamics in a wider range of configurations.

Our research substantially advances the realization of magnetic monopoles, enhances our understanding of vortex dynamics and monopole properties, lays the foundation for experimental verification, and is helpful for further study of vortex dynamics in closed manifolds.

Acknowledgements.
We acknowledge the discussion from Congjun Wu, Lichen Zhao, Shanchao Zhang, and Biao Wu. This work is supported by the National Natural Science Foundation of China under grants No. 12105223, No. 12175180, No. 11934015, No. 12305029, and No. 12247103, the Major Basic Research Program of Natural Science of Shaanxi Province under grants No. 2017KCT-12 and No. 2017ZDJC-32, Shaanxi Fundamental Science Research Project for Mathematics and Physics under grant No. 22JSZ005 and No. 22JSQ041, and Natural Science Basic Research Program of Shaanxi under Grant No. 2024JC-YBMS-022 and No. 2023-JC-QN-0054. This research is also supported by the Youth Innovation Team of Shannxi Universities and The Double First-class University Construction Project of Northwest University.

Appendix A 2D effective interaction

Here, we derive the 2D effective interaction strength λ2D\lambda_{2D} for a shell-shaped BEC with large r0r_{0}. The Gross-Pitaevskii equation (GPE) is

itψ±=(i𝑨±)22mψ±+Vψ±+λ3D|ψ±|2ψ±,\displaystyle i\hbar{\partial_{t}{\psi_{\pm}}}=\frac{(-i\nabla-\bm{A}_{\pm})^{2}}{2m}\psi_{\pm}+V\psi_{\pm}+\lambda_{3D}\left|\psi_{\pm}\right|^{2}\psi_{\pm}, (46)

where V(r)=VH(r)2μIB1r+I/2mr2{V}(r)={V}_{H}(r)-2\mu_{I}B^{\prime}_{1}r+I/2mr^{2}. For a large shell geometry, we approximately have V(r)=12mωH2(rr0)2{V}(r)=\frac{1}{2}m\omega_{H}^{2}(r-r_{0})^{2}, after discarding a constant. We write the wave function as ψ=1rg(r)f(θ,ϕ,t)\psi=\frac{1}{r}g(r)f(\theta,\phi,t). The normalization condition is

1\displaystyle 1 =\displaystyle= g2(r)𝑑r,\displaystyle\int g^{2}(r)dr, (47)
1\displaystyle 1 =\displaystyle= |f(θ,ϕ,t)|2sinθdθdϕ.\displaystyle\int\int|f(\theta,\phi,t)|^{2}\sin\theta d\theta d\phi. (48)

To get λ2D\lambda_{2D}, We left-multiply the GPE (46) by rg(r)rg(r) and then perform a radial integration 𝑑r\int dr. Then, we obtain

itf(θ,ϕ,t)=+λ3D|f|2f1r2g4(r)𝑑r.i\hbar{\partial_{t}{f}}(\theta,\phi,t)=\cdots+\lambda_{3D}\left|f\right|^{2}f\int\frac{1}{r^{2}}g^{4}(r)dr. (49)

We denote

λ2Dr02λ3D1r2g4(r)𝑑r.\frac{\lambda_{2D}}{r_{0}^{2}}\equiv\lambda_{3D}\int\frac{1}{r^{2}}g^{4}(r)dr. (50)

For r0r_{0}\gg\ell, we have

λ2Dλ3Dg4(r)𝑑r.{\lambda_{2D}}\simeq\lambda_{3D}\int g^{4}(r)dr. (51)

In the static case, we use the Thomas-Fermi approximation,

μFψVψ+λ3D|ψ±|2ψ.\mu_{F}{\psi}\simeq V\psi+\lambda_{3D}\left|\psi_{\pm}\right|^{2}\psi. (52)

Assuming that the density is uniform for different θ\theta and ϕ\phi, i.e., |f|21/4π|f|^{2}\simeq 1/4\pi, we obtain

μF=12mωH2(rr0)2+λ3Dg2(r)/4πr2.\mu_{F}=\frac{1}{2}m\omega_{H}^{2}(r-r_{0})^{2}+\lambda_{3D}g^{2}(r)/4\pi r^{2}. (53)

Thus, for large r0r_{0},

g2(r)\displaystyle g^{2}(r) =\displaystyle= 4πr2[μFmωH2(rr0)2/2]/λ3D\displaystyle 4\pi r^{2}[\mu_{F}-m\omega_{H}^{2}(r-r_{0})^{2}/2]/\lambda_{3D} (54)
\displaystyle\simeq 4πr02[μFmωH2(rr0)2/2]/λ3D\displaystyle 4\pi r_{0}^{2}[\mu_{F}-m\omega_{H}^{2}(r-r_{0})^{2}/2]/\lambda_{3D}

We denote μF=mωH2RTF2/2\mu_{F}=m\omega_{H}^{2}R^{2}_{TF}/2. Then from the normalization of g(r)g(r), we have

1=8π3mωH2r02RTF3λ3D.1=\frac{8\pi}{3}\frac{m\omega_{H}^{2}r_{0}^{2}R^{3}_{TF}}{\lambda_{3D}}. (55)

Finally, we obtain

λ2D\displaystyle{\lambda_{2D}} =\displaystyle= λ3Dg4(r)𝑑r\displaystyle\lambda_{3D}\int g^{4}(r)dr (56)
=\displaystyle= 25(9πmωH2r02λ3D2)13.\displaystyle\frac{2}{5}(9\pi m\omega_{H}^{2}r_{0}^{2}\lambda_{3D}^{2})^{\frac{1}{3}}.

Appendix B Crank-Nicolson algorithm for gauge field GPE

Here, we illustrate the discretization of GPE. When the time spacing Δt\Delta t is sufficient small, for each gauge choice, the evolution operator is split into four steps:

eiH±Δt=eiH4±ΔteiH3±ΔteiH2±ΔteiH1±Δt.e^{-iH_{\pm}\Delta t}=e^{-iH_{4}^{\pm}\Delta t}e^{-iH_{3}^{\pm}\Delta t}e^{-iH_{2}^{\pm}\Delta t}e^{-iH_{1}^{\pm}\Delta t}. (57)

we solve the GPE step by step individually:

(1) In the first step, since H1±H_{1}^{\pm} is already diagonalized,

ψ±n+1/4=eiH1±Δtψ±n\psi_{\pm}^{n+1/4}=e^{-iH^{\pm}_{1}\Delta t}\psi_{\pm}^{n} (58)

can be implemented directly.

(2) In the second step, ψ±n+1/2=eiH2±Δtψ±n+1/4\psi_{\pm}^{n+1/2}=e^{-iH^{\pm}_{2}\Delta t}\psi_{\pm}^{n+1/4}. Linear expansion gives (m=1m=1)

iψ±,jn+1/2ψ±,jn+1/4Δt=1212Δx2{(ψ±,j+1n+1/22ψ±,jn+1/2\displaystyle i\frac{\psi_{\pm,j}^{n+1/2}-\psi_{\pm,j}^{n+1/4}}{\Delta t}=-\frac{1}{2}\frac{1}{2\Delta_{x}^{2}}\Big{\{}\big{(}\psi_{\pm,j+1}^{n+1/2}-2\psi_{\pm,j}^{n+1/2} (59)
+ψ±,j1n+1/2)+(ψ±,j+1n+1/42ψ±jn+1/4+ψ±j1n+1/4)}\displaystyle+\psi_{\pm,j-1}^{n+1/2}\big{)}+\big{(}\psi_{\pm,j+1}^{n+1/4}-2\psi_{\pm j}^{n+1/4}+\psi_{\pm j-1}^{n+1/4}\big{)}\Big{\}}
+iAx,j±8Δx{(ψ±,j+1n+1/2ψ±,j1n+1/2)+(ψ±,j+1n+1/4ψ±,j1n+1/4)}\displaystyle+\frac{iA_{x,j}^{\pm}}{8\Delta_{x}}\Big{\{}\big{(}\psi_{\pm,j+1}^{n+1/2}-\psi_{\pm,j-1}^{n+1/2}\big{)}+\big{(}\psi_{\pm,j+1}^{n+1/4}-\psi_{\pm,j-1}^{n+1/4}\big{)}\Big{\}}
+i8Δx{(Ax,j+1±ψ±,j+1n+1/2Ax,j1±ψ±,j1n+1/2)\displaystyle+\frac{i}{8\Delta_{x}}\Big{\{}\big{(}A_{x,j+1}^{\pm}\psi_{\pm,j+1}^{n+1/2}-A_{x,j-1}^{\pm}\psi_{\pm,j-1}^{n+1/2}\big{)}
+(Ax,j+1±ψ±,j+1n+1/4Ax,j1±ψ±,j1n+1/4)},\displaystyle+\big{(}A_{x,j+1}^{\pm}\psi_{\pm,j+1}^{n+1/4}-A_{x,j-1}^{\pm}\psi_{\pm,j-1}^{n+1/4}\big{)}\Big{\}},

where j=1,2,3,j=1,2,3,... is the jj-th grid point in xx direction for given yy and zz, and Δx\Delta_{x} is the spacing. Eq. (59) results in a set of equations

α±,jψ±,j1n+1/2+α±,j0ψ±,jn+1/2+α±,j+ψ±,j+1n+1/2=β±,j,\alpha_{\pm,j}^{-}\psi_{\pm,j-1}^{n+1/2}+\alpha_{\pm,j}^{0}\psi_{\pm,j}^{n+1/2}+\alpha_{\pm,j}^{+}\psi_{\pm,j+1}^{n+1/2}=\beta_{\pm,j}, (60)

where

β±,j=\displaystyle\beta_{\pm,j}= iΔt4Δx2(ψ±,j+1n+1/42ψ±,jn+1/4+ψ±,j1n+1/4)\displaystyle\frac{i\Delta t}{4\Delta_{x}^{2}}\big{(}\psi_{\pm,j+1}^{n+1/4}-2\psi_{\pm,j}^{n+1/4}+\psi_{\pm,j-1}^{n+1/4}\big{)} (61)
+Δt8Δx{Ax,j±(ψ±,j+1n+1/4ψ±,j1n+1/4)\displaystyle+\frac{\Delta t}{8\Delta_{x}}\Big{\{}A_{x,j}^{\pm}\big{(}\psi_{\pm,j+1}^{n+1/4}-\psi_{\pm,j-1}^{n+1/4}\big{)}
+(Ax,j+1±ψ±,j+1n+1/4Ax,j1±ψ±,j1n+1/4)}\displaystyle+\big{(}A_{x,j+1}^{\pm}\psi_{\pm,j+1}^{n+1/4}-A_{x,j-1}^{\pm}\psi_{\pm,j-1}^{n+1/4}\big{)}\Big{\}}
α±,j0=\displaystyle\alpha_{\pm,j}^{0}= 1+iΔt2Δx2,\displaystyle 1+\frac{i\Delta t}{2\Delta_{x}^{2}},
α±,j=\displaystyle\alpha_{\pm,j}^{-}= iΔt8Δx[2Δx+i(Ax,j1±+Ax,j±)],\displaystyle-\frac{i\Delta t}{8\Delta_{x}}\Big{[}\frac{2}{\Delta_{x}}+i(A_{x,j-1}^{\pm}+A_{x,j}^{\pm})\Big{]},
α±,j+=\displaystyle\alpha_{\pm,j}^{+}= iΔt8Δx[2Δxi(Ax,j+1±+Ax,j±)].\displaystyle-\frac{i\Delta t}{8\Delta_{x}}\Big{[}\frac{2}{\Delta_{x}}-i(A_{x,j+1}^{\pm}+A_{x,j}^{\pm})\Big{]}.

Then, the wave function ψ±n+1/2\psi_{\pm}^{n+1/2} is obtained accordingly. A similar method can be performed for the third step ψ±n+3/4=eiH3±Δtψ±n+1/2\psi_{\pm}^{n+3/4}=e^{-iH_{3}^{\pm}\Delta t}\psi_{\pm}^{n+1/2} and the fourth step ψ¯±n+1=eiH4±Δtψ±n+3/4\underline{\psi}_{\pm}^{n+1}=e^{-iH_{4}^{\pm}\Delta t}\psi_{\pm}^{n+3/4}.

References

  • [1] P. A. M. Dirac, Quantised singularities in the electromagnetic field, Proc. R. Soc. A 133, 60 (1931).
  • [2] T. T. Wu and C. N. Yang, Concept of nonintegrable phase factors and global formulation of gauge fields, Phys. Rev. D 12, 3845 (1975).
  • [3] T. T. Wu and C. N. Yang, Dirac monopole without strings: Monopole harmonics, Nucl. Phys. B 107, 365 (1976).
  • [4] M. Bonnardeau and A. K. Drukier, Creation of magnetic monopoles in pulsars, Nature 277, 543 (1979).
  • [5] H. J. Friseh, Quest for magnetic monopoles, Nature 344, 706 (1990).
  • [6] A. Vilenkin and E. P. S. Shellard, Cosmic Strings and Other Topological Defects, Cambridge Monographs on Mathematical Physics (Cambridge University Press, Cambridge, 2000).
  • [7] A. H. Guth, Inflationary universe: A possible solution to the horizon and flatness problems, Phys. Rev. D 23, 347 (1981).
  • [8] M. M. Salomaa, Monopoles in the rotating superfluid Helium-3 a-b interface, Nature 326, 367 (1987).
  • [9] C. Castelnovo, R. Moessner, and S. L. Sondhi, Magnetic monopoles in spin ice, Nature 451, 42 (2008).
  • [10] M. J. P. Gingras, Observing monopoles in a magnetic analog of ice, Science 326, 375 (2009).
  • [11] S. T. Bramwell, S. R. Giblin, S. Calder, R. Aldus, D. Prabhakaran, and T. Fennell, Measurement of the charge and current of magnetic monopoles in spin ice, Nature 461, 956 (2009).
  • [12] S. R. Giblin, S. T. Bramwell, P. C. W. Holdsworth, D. Prabhakaran, and I. Terry, Creation and measurement of long-lived magnetic monopole currents in spin ice, Nat. Phys. 7, 252 (2011).
  • [13] S.-Z. Lin and A. Saxena, Dynamics of Dirac strings and monopole like excitations in chiral magnets under a current drive, Phys. Rev. B 93, 060401 (2016).
  • [14] Z. Fang, N. Nagaosa, K. S. Takahashi, A. Asamitsu, R. Mathieu, T. Ogasawara, H. Yamada, M. Kawasaki, Y. Tokura, and K. Terakura, The anomalous Hall effect and magnetic monopoles in momentum space, Science 302, 92 (2003).
  • [15] A. Béché, R. V. Boxem, G. V. Tendeloo, and J. Verbeeck, Magnetic monopole field exposed by electrons, Nature Physics 10, 26 (2013).
  • [16] Z.-L. Zhang, M.-F. Chen, H.-Z. Wu, and Z.-B. Yang, Quantum simulation of abelian Wu-Yang monopoles in spin-1/2 systems, Laser Phys. Lett. 14, 045205 (2017).
  • [17] M. I. Marqués, S. Edelstein, P. A. Serena, B. C. L. de Larrinzar, and A. Garcia-Martín, Magneto-optical particles in isotropic spinning fields mimic magnetic monopoles, Phys. Rev. Lett. 133, 046901 (2024).
  • [18] M. W. Ray, E. Ruokokoski, S. Kandel, M. Möttönen, and D. S. Hall, Observation of Dirac monopoles in a synthetic magnetic field, Nature 505, 657 (2014).
  • [19] V. Pietilä and M. Möttönen, Creation of Dirac monopoles in spinor Bose-Einstein condensates, Phys. Rev. Lett. 103, 030401 (2009).
  • [20] M. W. Ray, E. Ruokokoski, K. Tiurev, M. Möttönen, and D. S. Hall, Observation of isolated monopoles in a quantum field, Science 348, 544 (2015).
  • [21] E. Ruokokoski, V. Pietilä, and M. Möttönen, Ground-state Dirac monopole, Phys. Rev. A 84, 063627 (2011).
  • [22] G.-S. Xu, M. Jain, X.-F. Zhou, G.-C. Guo, M. A. Amin, H. Pu, and Z.-W. Zhou, Engineering and revealing Dirac strings in spinor condensates, Phys. Rev. Research 6, 023272 (2024).
  • [23] X.-F. Zhou, C. Wu, G.-C. Guo, R. Wang, H. Pu, and Z.-W. Zhou, Synthetic Landau levels and spinor vortex matter on a Haldane spherical surface with a magnetic monopole, Phys. Rev. Lett. 120, 130402 (2018).
  • [24] A. M. Turner, V. Vitelli, and D. R. Nelson, Vortices on curved surfaces, Rev. Mod. Phys. 82, 1301 (2010).
  • [25] K. Padavić, K. Sun, C. Lannert, and S. Vishveshwara, Vortex-antivortex physics in shell-shaped Bose-Einstein condensates, Phys. Rev. A 102, 043305 (2020).
  • [26] N.-E. Guenther, P. Massignan, and A. L. Fetter, Superfluid vortex dynamics on a torus and other toroidal surfaces of revolution, Phys. Rev. A 101, 053606 (2020).
  • [27] S. J. Bereta, M. A. Caracanhas, and A. L. Fetter, Superfluid vortex dynamics on a spherical film, Phys. Rev. A 103, 053306 (2021).
  • [28] M. A. Caracanhas, P. Massignan, and A. L. Fetter, Superfluid vortex dynamics on an ellipsoid and other surfaces of revolution, Phys. Rev. A 105, 023307 (2022).
  • [29] A. Tononi, A. Pelster, and L. Salasnich, Topological superfluid transition in bubble-trapped condensates, Phys. Rev. Res. 4, 013122 (2022).
  • [30] D. G. Dritschel, M. Lucia, and Andrew C. Poje, Ergodicity and spectral cascades in point vortex flows on the sphere, Phys. Rev. E 91, 063014 (2015).
  • [31] T. Kanai and W. Guo, True mechanism of spontaneous order from turbulence in two-dimensional superfluid manifolds, Phys. Rev. Lett. 127, 095301 (2021).
  • [32] Y. Xiong and X. Yu, Hydrodynamics of quantum vortices on a closed surface, Phys. Rev. Research 6, 013133 (2024).
  • [33] G. Li and D. K. Efimkin, Equatorial waves in rotating bubble-trapped superfluids, Phys. Rev. A 107, 023319 (2023).
  • [34] A. C. White, Triangular vortex lattices and giant vortices in rotating bubble Bose-Einstein condensates, Phys. Rev. A 109, 013301 (2024).
  • [35] Y. He and C.-C. Chien, Vortex structure and spectrum of an atomic fermi superfluid in a spherical bubble trap, Phys. Rev. A 108, 053303 (2023).
  • [36] F. D. M. Haldane, Fractional quantization of the Hall effect: A hierarchy of incompressible quantum fluid states, Phys. Rev. Lett. 51, 605 (1983).
  • [37] C. J. Pethick and H. Smith, Bose-Einstein Condensation in Dilute Gases, 2nd ed. (Cambridge University Press, Cambridge, 2008).
  • [38] M. Diermaier, C. B. Jepsen, B. Kolbinger, C. Malbrunot, O. Massiczek, C. Sauerzopf, M. C. Simon, J. Zmeskal, and E. Widmann, In-beam measurement of the hydrogen hyperfine splitting and prospects for antihydrogen spectroscopy, Nat. Commun. 8, 15749 (2017).
  • [39] E. Arimondo, M. Inguscio, and P. Violino, Experimental determinations of the hyperfine structure in the alkali atoms, Rev. Mod. Phys. 49, 31 (1977).
  • [40] A. A. Thiele, Steady-State Motion of Magnetic Domains, Phys. Rev. Lett. 30, 230 (1973).
  • [41] C. Reichhardt, C. J. O. Reichhardt, and M. V. Milošević, Statics and dynamics of skyrmions interacting with disorder and nanostructures, Rev. Mod. Phys. 94, 035005 (2022).
  • [42] X. S. Wang, A. Qaiumzadeh, and A. Brataas, Current-Driven Dynamics of Magnetic Hopfions, Phys. Rev. Lett. 123, 147203 (2019).
  • [43] P. Muruganandam and S. Adhikari, Fortran programs for the time-dependent Gross-Pitaevskii equation in a fully anisotropic trap, Comput. Phys. Commun. 180, 1888 (2009).
  • [44] R. Kishor Kumar, V. Lonc̆ar, P. Muruganandam, S. K. Adhikari, and A. Balaz̆, C and Fortran OpenMP programs for rotating Bose-Einstein condensates, Comput. Phys. Commun. 240, 74 (2019).
  • [45] A. Tononi and L. Salasnich, Bose-Einstein condensation on the surface of a sphere, Phys. Rev. Lett. 123, 160403 (2019).
  • [46] A. Tononi, F. Cinti, and L. Salasnich, Quantum bubbles in microgravity, Phys. Rev. Lett. 125, 010402 (2020).
  • [47] R. A. Carollo, D. C. Aveline, B. Rhyno, S. Vishveshwara, C. Lannert, J. D. Murphree, E. R. Elliott, J. R. Williams, R. J. Thompson, and N. Lundblad, Observation of ultracold atomic bubbles in orbital microgravity, Nature 606, 281 (2022).
  • [48] B. Rhyno, N. Lundblad, D. C. Aveline, C. Lannert, and S. Vishveshwara, Thermodynamics in expanding shell-shaped Bose-Einstein condensates, Phys. Rev. A 104, 063310 (2021).
  • [49] F. Jia, Z. Huang, L. Qiu, R. Zhou, Y. Yan, and D. Wang, Expansion dynamics of a shell-shaped Bose-Einstein condensate, Phys. Rev. Lett. 129, 243402 (2022).
  • [50] Y. He, H. Guo, and C.-C. Chien, BCS-BEC crossover of atomic Fermi superfluid in a spherical bubble trap, Phys. Rev. A 105, 033324 (2022).
  • [51] M. Brooks, M. Lemeshko, D. Lundholm, and E. Yakaboylu, Molecular impurities as a realization of anyons on the two-sphere, Phys. Rev. Lett. 126, 015301 (2021).
  • [52] V. A. Bogomolov, Dynamics of vorticity at a sphere, Fluid Dyn. 12, 863(1977).