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

Vortex nucleation in rotating Bose-Einstein condensates with density-dependent gauge potential

Ishfaq Ahmad Bhat1, Thudiyangal Mithun2, Bishwajyoti Dey1 1 Department of Physics, Savitribai Phule Pune University, Pune, Maharashtra, India, 411007
2 Department of Mathematics and Statistics, University of Massachusetts, Amherst Massachusetts 01003-4515, USA
Abstract

We study numerically the vortex dynamics and vortex-lattice formation in a rotating density-dependent Bose-Einstein condensate (BEC), characterized by the presence of nonlinear rotation. By varying the strength of nonlinear rotation in density-dependent BECs, we calculate the critical frequency, Ωcr\Omega_{\text{cr}}, for vortex nucleation both in adiabatic and sudden external trap rotations. The nonlinear rotation modifies the extent of deformation experienced by the BEC due to the trap and shifts the Ωcr\Omega_{\text{cr}} values for vortex nucleation. The critical frequencies and thereby, the transition to vortex-lattices in an adiabatic rotation ramp, depend on conventional s-wave scattering lengths through the strength of nonlinear rotation, C\mathit{C}, such that Ωcr(C>0)<Ωcr(C=0)<Ωcr(C<0)\Omega_{\text{cr}}(\mathit{C}>0)<\Omega_{\text{cr}}(\mathit{C}=0)<\Omega_{\text{cr}}(\mathit{C}<0). In an analogous manner, the critical ellipticity (ϵcr\epsilon_{\text{cr}}) for vortex nucleation during an adiabatic introduction of trap ellipticity (ϵ\epsilon) depends on the nature of nonlinear rotation besides trap rotation frequency. The nonlinear rotation additionally affects the vortex-vortex interactions and the motion of the vortices through the condensate by altering the strength of Magnus force on them. The combined result of these nonlinear effects is the formation of the non-Abrikosov vortex-lattices and ring-vortex arrangements in the density-dependent BECs.

I Introduction:

Soon after the successful realization of Bose-Einstein condensates (BECs) in alkali metal atoms, the study of quantized vortices and their dynamics in them gained an active interest Matthews et al. (1999); Madison et al. (2000, 2001); Hodby et al. (2001); Abo-Shaeer et al. (2001); Recati et al. (2001); Sinha and Castin (2001). Quantum vortices in BECs and superfluid He4{}^{4}\text{He} Donnelly (1991) are the topological defects that arise in response to the external rotation Madison et al. (2000, 2001); Hodby et al. (2001); Abo-Shaeer et al. (2001), shaking a BEC Yukalov et al. (2016); Middleton-Spencer et al. (2022) or even moving an obstacle through the condensate Sasaki et al. (2010); Kwon et al. (2015). The rotation in a BEC is induced by stirring the atomic cloud and can be either adiabatic or sudden (nonadiabatic). In either case the nucleation of vortices within the condensate occurs only when the trap rotation exceeds a certain minimum value called critical frequency, Ωcr\Omega_{\text{cr}}. At this stage the vortices in BEC become energetically favorable. For rotational frequencies greater than Ωcr\Omega_{\text{cr}}, multiple vortices enter the condensate and eventually form a vortex-lattice. In nonadiabatic rotations, the critical rotation frequency, Ωcr\Omega_{\text{cr}}, at which a single vortex exists stably at the minimum of the trap, decreases with the increasing strength of two- Fetter (2009); Dalfovo et al. (1999); Ghosh (2004) and three-body Mithun et al. (2013) interaction strengths. The change in Ωcr\Omega_{\text{cr}} is, however, marginal at large interaction strengths and Ωcr0.65ω\Omega_{\text{cr}}\simeq 0.65\omega_{\perp}, where ω\omega_{\perp} is the transverse trapping frequency, for the creation of single vortex in the ENS experiment Chevy et al. (2000). On the other hand, in an adiabatic rotation ramp for a fixed trap ellipticity, conventional BECs with only short range s-wave interactions form vortex lattices when the trap rotation frequencies are ω/2\approx\omega_{\perp}/\sqrt{2} Madison et al. (2000, 2001); Hodby et al. (2001). This transition from the vortex free state to the vortex-lattice structure occurs via dynamical instability Madison et al. (2001); Sinha and Castin (2001) and is independent of the s-wave interaction strength and longitudinal confinement. Likewise, vortex-lattices are also achieved by varying ϵ\epsilon in a linear fashion at fixed trap rotation Madison et al. (2001); Parker et al. (2006). Here the transition to the vortex-lattice structure is characterized by the critical ellipticity, ϵcr\epsilon_{\text{cr}}, depending only on the trap rotation frequency. Long range dipole-dipole interactions and trap geometry within the dipolar BECs are, however, found to alter these critical frequencies and ellipticities Cai et al. (2018); O’Dell and Eberlein (2007); van Bijnen et al. (2007, 2009); Prasad et al. (2019).

In addition to providing a flexible test-bed for numerous quantum phenomena, BECs are even noted for the experimental realization of artificial electromagnetism in them Dalibard et al. (2011); Goldman et al. (2014). This has been made possible through the engineering of artificial gauge potentials in these ultracold condensates. Induced electromagnetism in BECs has lead to the emergence of novel phenomena like density-dependent magnetism Edmonds et al. (2013), exotic spin-orbit Lin et al. (2011) and spin angular momentum couplings Chen et al. (2018); Zhang et al. (2019). However, the nature of gauge potentials in BECs is typically static with no feedback between light and matter waves. In dynamical gauge fields, on the other hand, nonlinear feedback between matter and gauge fields plays a role. This back-action is expected to enrich the physics of atomic and nonlinear systems and has been realized in experiments recently Clark et al. (2018); Görg et al. (2019).

Scalar and nonrotating BECs with density-dependent gauge potentials, known as Chiral condensates, have been already studied in one-dimension for anyonic structures Keilmann et al. (2011), chiral solitons Aglietti et al. (1996); Bhat et al. (2021); Dingwall et al. (2018); Dingwall and Öhberg (2019) and collective excitations Edmonds et al. (2015) in them. More recently, strong density-dependent gauge potentials are even shown to realize chaotic collective dynamics in two-dimensional BECs Chen and Zhu (2022). The response of density-dependent BECs against the trap rotation in two-dimensions has been studied in Refs. Edmonds and Nitta (2020); Edmonds (2021) where vortex ground states have been simulated numerically. The density-dependent gauge potential realizes an effective nonlinear rotation in BECs, thereby leading to non-Abrikosov vortex-lattices and ring vortex arrangements. In contrast to the Abrikosov vortex-lattices in scalar BECs Campbell and Ziff (1979); Sato et al. (2007); Mithun et al. (2016), the non-Abrikosov vortex-lattices in BECs with density-dependent gauge potentials lack the hexagonal (or triangular) symmetry in the vortex arrangements Edmonds and Nitta (2020); Edmonds (2021).

The present paper studies the effects of nonlinear rotation on the dynamics of vortex lattice formation in these BECs within the scope of time-dependent numerical simulations. The studies of vortex nucleation in conventional Madison et al. (2000, 2001); Sinha and Castin (2001); Tsubota et al. (2002); Kasamatsu et al. (2003) and dipolar van Bijnen et al. (2007, 2009); Prasad et al. (2019) BECs reveal that the transition to the vortex-lattices in an adiabatic rotation ramp depends only on the rotation frequency and trap ellipticity but is independent of the s-wave interaction strength. Here we show that unlike in conventional and dipolar condensates, this transition depends on the short range s-wave interactions in density-dependent condensates. The density-dependent BECs show similar responses against the adiabatic and sudden rotations such that the number and arrangement of the nucleated vortices varies with the strength of the nonlinear rotations. In addition to the energy considerations, the non-Abrikosov vortex-lattices and ring-vortex arrangements in these condensates are a result of the modifications in the Magnus force experienced by the vortices and the repulsive interactions between them.

The subsequent material is structured as follows. In Sec. II we introduce the model in the form of Gross-Pitaevskii equation including the density-dependent gauge potential. This is followed by Sec. III where we present the possible routes to vortex nucleation through numerical investigations based on the Crank Nicholson method. The work is finally summarized in Sec. IV.

II Model

We consider a weakly interacting BEC of NN two-level atoms coupled by a coherent light-matter interaction due to an applied laser field. Within the rotating-wave approximation, such a BEC is described by the following mean-field Hamiltonian Dalibard et al. (2011); Edmonds and Nitta (2020); Butera et al. (2016a):

^=(𝐩^22m+V(𝐫))𝟙+^int+𝒰^MF\hat{\mathcal{H}}=\left(\frac{\mathbf{\hat{p}}^{2}}{2m}+V(\mathbf{r})\right)\otimes\mathbb{1}+\hat{\mathcal{H}}_{\text{int}}+\hat{\mathcal{U}}_{\text{MF}} (1)

where in the first term 𝐩^\mathbf{\hat{p}} is the momentum operator, V(𝐫)=m(ωx2x2+ωy2y2+ωz2z2)/2V(\mathbf{r})=m(\omega^{2}_{x}x^{2}+\omega^{2}_{y}y^{2}+\omega^{2}_{z}z^{2})/2 is the trapping potential and 𝟙\mathbb{1} symbolizes the 2×22\times 2 unity matrix. The mean-field interactions, ^int=(1/2)diag[Δ1,Δ2]\hat{\mathcal{H}}_{\text{int}}=(1/2)\text{diag}[\Delta_{1},\Delta_{2}] with Δi=gii|Ψi|2+gij|Ψj|2\Delta_{i}=g_{ii}|\Psi_{i}|^{2}+g_{ij}|\Psi_{j}|^{2} and gij=4π2aij/mg_{ij}=4\pi\hbar^{2}a_{ij}/m where aija_{ij} are the respective scattering lengths of the collisions between atoms in internal states ii and jj (i,j=1,2i,j=1,2). Further, the light-matter interactions accounted by Refs. Goldman et al. (2014); Butera et al. (2016a),

𝒰^MF=Ωr2(cosθ(𝐫)eiϕ(𝐫)sinθ(𝐫)eiϕ(𝐫)sinθ(𝐫)cosθ(𝐫))\hat{\mathcal{U}}_{\text{MF}}=\frac{\hbar\Omega_{r}}{2}\begin{pmatrix}\text{cos}\theta\left(\mathbf{r}\right)&e^{-i\phi\left(\mathbf{r}\right)}\text{sin}\theta\left(\mathbf{r}\right)\\ e^{i\phi\left(\mathbf{r}\right)}\text{sin}\theta\left(\mathbf{r}\right)&-\text{cos}\theta\left(\mathbf{r}\right)\end{pmatrix} (2)

are parameterized in terms of Rabi frequency, Ωr\Omega_{r}, mixing angle, θ(𝐫)\theta\left(\mathbf{r}\right) and phase of the incident laser beam, ϕ(𝐫)\phi\left(\mathbf{r}\right). We use perturbation theory by defining the perturbed states in terms of the unperturbed eigen-states |±\ket{\pm} of U^MF\hat{U}_{\text{MF}} Butera et al. (2016a); Edmonds and Nitta (2020); Edmonds (2021):

|Ψ±=|±±ΔdΩr|\ket{\Psi_{\pm}}=\ket{\pm}\pm\frac{\Delta_{d}}{\hbar\Omega_{r}}\ket{\mp} (3)

where Δd=sin(θ/2)cos(θ/2)(Δ1Δ2)/2\Delta_{d}=\text{sin}\left(\theta/2\right)\text{cos}\left(\theta/2\right)\left(\Delta_{1}-\Delta_{2}\right)/2 is the mean-field detuning. Since the qualitative details do not depend on which of the two perturbed states is chosen, therefore, we chose the Ψ+\Psi_{+}-state and write the effective Hamiltonian as Butera et al. (2016a); Edmonds and Nitta (2020); Edmonds (2021):

H^+=(𝐩^𝐀+)22m+W++Ωr2+Δ++V(𝐫),\mathit{\hat{H}}_{+}=\frac{\left(\mathbf{\hat{p}}-\mathbf{A}_{+}\right)^{2}}{2m}+W_{+}+\frac{\hbar\Omega_{r}}{2}+\Delta_{+}+V(\mathbf{r}), (4)

where the vector gauge potential is defined as, 𝐀+=iΨ+|Ψ+\mathbf{A}_{+}=i\hbar\braket{\Psi_{+}}{\Psi_{+}}, while the scalar potential, W+=2|Ψ+|Ψ|2/2mW_{+}=\hbar^{2}|\braket{\Psi_{+}}{\nabla\Psi_{-}}|^{2}/2m and Δ+=(Δ1cos2(θ/2)+Δ2sin2(θ/2))/2\Delta_{+}=(\Delta_{1}~{}\text{cos}^{2}(\theta/2)+\Delta_{2}~{}\text{sin}^{2}(\theta/2))/2 is the dressed mean-field interaction. The generalized mean-field Gross-Pitaevskii (GP) equation for Ψ+\Psi_{+} is finally obtained by extremizing the energy functional, =Ψ+|H^+|Ψ+\mathcal{E}=\braket{\Psi_{+}}{\hat{H}_{+}}{\Psi_{+}} as Edmonds and Nitta (2020); Butera et al. (2016a, b); Butera (2017):

iΨ+t=[(𝐩^𝐀+)22m+𝐚1𝐉]Ψ++[V(𝐫)+Ωr2+2Δ++W+]Ψ++[𝔫+(W+Ψ+W+Ψ+)W+Ψ+𝔫+]\begin{split}i\hbar\frac{\partial\Psi_{+}}{\partial t}&=\left[\frac{\left(\mathbf{\hat{p}}-\mathbf{A}_{+}\right)^{2}}{2m}+\mathbf{a}_{1}\cdot\mathbf{J}\right]\Psi_{+}+\\ &\left[V(\mathbf{r})+\frac{\hbar\Omega_{r}}{2}+2\Delta_{+}+W_{+}\right]\Psi_{+}+\\ &\left[\mathfrak{n}_{+}\left(\frac{\partial W_{+}}{\partial\Psi^{*}_{+}}-\boldsymbol{\nabla}\cdot\frac{\partial W_{+}}{\partial\boldsymbol{\nabla}\Psi^{*}_{+}}\right)-\frac{\partial W_{+}}{\partial\boldsymbol{\nabla}\Psi^{*}_{+}}\cdot\boldsymbol{\nabla}\mathfrak{n}_{+}\right]\end{split} (5)

Here 𝔫+=|Ψ+(𝐫,t)|2\mathfrak{n}_{+}=|\Psi_{+}(\mathbf{r},t)|^{2} is the atomic density of the BEC in the |Ψ+\ket{\Psi_{+}} state, 𝐚1=ϕΔdsinθ/𝔫+Ωr\mathbf{a}_{1}=\boldsymbol{\nabla}\phi\Delta_{d}\text{sin}\theta/\mathfrak{n}_{+}\Omega_{r} is the strength of the coupling to the gauge field and

𝐉=2mi[Ψ+(+i𝐀+)Ψ+Ψ+(i𝐀+)Ψ+]\mathbf{J}=\frac{\hbar}{2mi}\left[\Psi_{+}\left(\boldsymbol{\nabla}+\frac{i}{\hbar}\mathbf{A}_{+}\right)\Psi^{*}_{+}-\Psi^{*}_{+}\left(\boldsymbol{\nabla}-\frac{i}{\hbar}\mathbf{A}_{+}\right)\Psi_{+}\right] (6)

is the nonlinear current. We now expand the gauge potentials 𝐀+\mathbf{A}_{+} and W+W_{+} to first order in small parameters —θ=Ωr/Δ\theta=\Omega_{r}/\Delta representing the ratio of Rabi frequency to laser detuning, and ϵ=𝔫+(𝐫)(g11g12)/4Δ\epsilon=\mathfrak{n}_{+}(\mathbf{r})(g_{11}-g_{12})/4\hbar\Delta which takes into account the collisional and coherent interactions, to obtain the simplified relations for the potentials Edmonds and Nitta (2020); Edmonds (2021):

𝐀+=θ24(14ϵ)ϕ\mathbf{A}_{+}=-\frac{\hbar\theta^{2}}{4}\left(1-4\epsilon\right)\boldsymbol{\nabla}\phi (7)
W+=22((θ)2(14ϵ)+θ2(1+4ϵ)(ϕ)24mθ2ϵ)\begin{split}\mathit{W}_{+}=&\frac{\hbar^{2}}{2}\bigg{(}\frac{\left(\boldsymbol{\nabla}\theta\right)^{2}\left(1-4\epsilon\right)+\theta^{2}\left(1+4\epsilon\right)\left(\boldsymbol{\nabla}\phi\right)^{2}}{4m}-\\ &\boldsymbol{\nabla}\theta^{2}\cdot\boldsymbol{\nabla}\epsilon\bigg{)}\end{split} (8)

We further invoke the spatial dependence in Rabi coupling through Ωr=κ0r\Omega_{\text{r}}=\kappa_{0}r, where rr is the radial distance and κ0\kappa_{0} is a constant, and define the phase, ϕ=lφ\phi=l\varphi where ll and φ\varphi are respectively the angular momenta and polar angle of the incident laser beam Juzeliūnas et al. (2005). The density-dependent gauge potentials (Eqs. (7) and (8)) when substituted in Eq. (5) in a rotating frame gives the simplified three-dimensional GP equation as Edmonds and Nitta (2020); Edmonds (2021):

iΨt=[22m2+V(𝐫)Ω𝔫(𝐫,t)L^z+geff|Ψ|2]Ψi\hbar\frac{\partial\Psi}{\partial t}=\bigg{[}-\frac{\hbar^{2}}{2m}\nabla^{2}+V(\mathbf{r})-\Omega_{\mathfrak{n}}(\mathbf{r},t)\hat{L}_{z}+\text{g}_{\text{eff}}|\Psi|^{2}\bigg{]}\Psi (9)

In Eq. (9), the subscript ++ is omitted, the operator 2=2x2+2y2+2z2\nabla^{2}=\frac{\partial^{2}}{\partial x^{2}}+\frac{\partial^{2}}{\partial y^{2}}+\frac{\partial^{2}}{\partial z^{2}}, and L^z=i(xyyx)\hat{L}_{z}=-i\hbar\bigg{(}x\frac{\partial}{\partial y}-y\frac{\partial}{\partial x}\bigg{)} is the z component of the angular momentum operator. The density-dependent rotation experienced by the BEC, Ω𝔫(𝐫,t)\Omega_{\mathfrak{n}}(\mathbf{r},t) has the form,

Ω𝔫(𝐫,t)=Ω0+C𝔫(𝐫,t),\Omega_{\mathfrak{n}}(\mathbf{r},t)=\Omega_{0}+\mathit{C}\mathfrak{n}(\mathbf{r},t), (10)

where Ω0\Omega_{0} represents the trap rotation and 𝔫(𝐫,t)=|Ψ(𝐫,t)|2\mathfrak{n}(\mathbf{r},t)=|\Psi(\mathbf{r},t)|^{2} is the number density of the BEC. The nonlinear rotation strength, C=lθ02(g11g12)/(mΔ)\mathit{C}=l\theta^{2}_{0}{(g_{11}-g_{12})}/(m\Delta) where θ0=κ0/Δ\theta_{0}=\kappa_{0}/\Delta while the mean-field interaction, geff=g11+(2l21)C/4l\text{g}_{\text{eff}}=g_{11}+\hbar(2l^{2}-1)\mathit{C}/{4l}. It is worth mentioning that the allowed values of C\mathit{C} in a density-dependent BEC with geff>0\text{g}_{\text{eff}}>0 are bounded by ±Cmax\pm\mathit{C}_{max}, where Cmax=8π(ln2)geff/3Nm\mathit{C}_{max}=\sqrt{8\pi(\text{ln}2)\text{g}_{\text{eff}}/3Nm} Edmonds (2021). Otherwise the nonlinear rotation overcomes the trap and the BEC ceases to exist. As a consequence of the density-dependent gauge potentials, the nonlinear rotation, C𝔫(𝐫,t)\mathit{C}\mathfrak{n}(\mathbf{r},t) gives rise to the density modulated angular velocity within the BEC. The different density regions in a BEC with density-dependent gauge potentials, therefore, experience different rotations depending on the nature of nonlinear rotation. Such effective nonlinear rotations in the BEC can be induced by employing laser beams with spatially varying intensities and carrying ll units of orbital angular momenta. From an experimental point of view, one can employ Laguerre-Gaussian laser beams with cylindrically varying intensity profiles and l=1l=1, recently used in realizing spin-angular-momentum-coupled BECs Chen et al. (2018). Furthermore, the density-dependent rotation in Eq. (10) is the combined (rigid-body and nonlinear) rotation, and the second gauged rotation still exists even when the externally driven rotation is absent (Ω0=0\Omega_{0}=0). The larger the number of atoms in a BEC, the easier it is to induce the nonlinear rotation and, therefore, the nucleation of the vortices in the BEC.

By expressing Ψ(𝐫,t)=ψ(𝝆,t)exp(z2/2σz2)/πσz24\Psi(\mathbf{r},t)=\psi(\boldsymbol{\rho},t)\text{exp}(-z^{2}/2\sigma^{2}_{z})/\sqrt[4]{\pi\sigma^{2}_{z}} for a BEC of thickness σz\sigma_{z} confined in a highly anisotropic trap with ωz>>ωx,y\omega_{z}>>\omega_{\text{x,y}}, and following the usual procedure of dimensional reduction Yunyi (2005), Eq. (9) results in the following two-dimensional (2D) GP equation:

iψt=[22m𝝆2+V(𝝆)Ωn(𝝆,t)L^z+g|ψ|2]ψi\hbar\frac{\partial\psi}{\partial t}=\bigg{[}-\frac{\hbar^{2}}{2m}\nabla^{2}_{\boldsymbol{\rho}}+V(\boldsymbol{\rho})-\Omega_{n}(\boldsymbol{\rho},t)\hat{L}_{z}+\text{g}|\psi|^{2}\bigg{]}\psi (11)

The BEC, because of the high anisotropy and the absence of axial modes in case of sufficiently weak interactions Petrov et al. (2000), assumes a pancake geometry such that 𝝆=(x,y)\boldsymbol{\rho}=(x,y) represents the position vector in xy plane. In quasi-2D settings, the operator 𝝆2=2x2+2y2\nabla^{2}_{\boldsymbol{\rho}}=\frac{\partial^{2}}{\partial x^{2}}+\frac{\partial^{2}}{\partial y^{2}}, V(𝝆)=12m(ωx2x2+ωy2y2)V(\boldsymbol{\rho})=\frac{1}{2}m(\omega^{2}_{x}x^{2}+\omega^{2}_{y}y^{2}) is the 2D harmonic trapping potential, and the nonlinear coefficient, g=geff/2πσz\text{g}=\text{g}_{\text{eff}}/\sqrt{2\pi}\sigma_{z}. The density-dependent trap rotation Ωn(𝝆,t)\Omega_{n}(\boldsymbol{\rho},t) has the same form as that in Eq. (10) except that C\mathit{C} is now scaled by a factor of 1/2πσz1/\sqrt{2\pi}\sigma_{z} and n(𝝆,t)=|ψ|2n(\boldsymbol{\rho},t)=|\psi|^{2} is the 2D density of the BEC. In an elliptical trap with ω2=(ωx2+ωy2)/2\omega^{2}_{\perp}=(\omega^{2}_{x}+\omega^{2}_{y})/2 as the transverse trapping frequency, we assume ωx=ω1ϵ\omega_{x}=\omega_{\perp}\sqrt{1-\epsilon} and ωy=ω1+ϵ\omega_{y}=\omega_{\perp}\sqrt{1+\epsilon} such that ϵ=(ωx2ωy2)/(ωx2+ωy2)\epsilon=(\omega^{2}_{x}-\omega^{2}_{y})/(\omega^{2}_{x}+\omega^{2}_{y}) determines the trap ellipticity. This ellipticity in the trapping potential breaks the rotational symmetry which is crucial for the nucleation of vortices in BECs. By means of spatiotemporal scaling Kasamatsu et al. (2003); Pérez-García et al. (1998), t=ω1t,(x,y)=a(x,y),andψ=Naψt=\omega^{-1}_{\perp}t^{\prime},(x,y)=a_{\perp}(x^{\prime},y^{\prime}),~{}\text{and}~{}\psi=\frac{\sqrt{N}}{a_{\perp}}\psi^{\prime}, Eq. (11) can be recast in the following normalized form:

(iγ)ψt=[𝝆22+V(𝝆)Ω~n(𝝆,t)L^z+g~|ψ|2]ψ.(i-\gamma)\frac{\partial\psi}{\partial t}=\bigg{[}-\frac{\nabla^{2}_{\boldsymbol{\rho}}}{2}+V(\boldsymbol{\rho})-\tilde{\Omega}_{n}(\boldsymbol{\rho},t)\hat{L}_{z}+\tilde{\text{g}}|\psi|^{2}\bigg{]}\psi. (12)

In Eq. (12), the scaled parameters, g~=gNm/2\tilde{\text{g}}=\text{g}Nm/\hbar^{2} and Ω~n(𝝆,t)=Ω0/ω+C~n(𝝆,t)\tilde{\Omega}_{n}(\boldsymbol{\rho},t)=\Omega_{0}/\omega_{\perp}+\tilde{\mathit{C}}n(\boldsymbol{\rho},t), where C~=CNm/(2πσz)\tilde{\mathit{C}}=\mathit{C}Nm/(\hbar\sqrt{2\pi}\sigma_{z}). In this dimensionless form the 2D wavefunction ψ(𝝆,t)\psi(\boldsymbol{\rho},t) is now normalized to unity, the elliptical trapping potential, V(𝝆)=12[(1ϵ)x2+(1+ϵ)y2]V(\boldsymbol{\rho})=\frac{1}{2}[(1-\epsilon)x^{2}+(1+\epsilon)y^{2}] and L^z=i(xyyx)\hat{L}_{z}=-i\big{(}x\frac{\partial}{\partial y}-y\frac{\partial}{\partial x}\big{)} is the angular-momentum operator. The term γ\gamma is introduced to account for any dissipation in the system Choi et al. (1998). By setting γ=0.03\gamma=0.03 Tsubota et al. (2002); Kasamatsu et al. (2003); Mithun et al. (2016, 2014) and ω=2π×200\omega_{\perp}=2\pi\times 200 Hz in the following, we study the effects of density-dependent gauge potentials on the dynamics of vortex nucleation in a harmonically confined BEC by directly solving the generalised GP equation (12) numerically. The dissipation term fastens the convergence to the equilibrium state characterized by a vortex-lattice and does not affect the dynamics of the BEC.

III Numerical Analysis

The introduction of trap rotation, Ω0\Omega_{0} and anisotropy, ϵ\epsilon have been employed both experimentally Madison et al. (2000, 2001) and numerically Tsubota et al. (2002); Kasamatsu et al. (2003); Parker et al. (2006) to evolve the harmonically confined BECs for the nucleation of vortices. In either case, the BEC is first deformed to an elliptical cloud, undergoes shape oscillations and is subsequently nucleated with vortices. The deformation of the condensate due to the trap rotation is parameterized by the factor α0=Ω0(x2y2/x2+y2)\alpha_{0}=\Omega_{0}(\braket{x^{2}-y^{2}}/\braket{x^{2}+y^{2}}) with \braket{\cdots} representing the expectation values and was first obtained theoretically by Recati et al. Recati et al. (2001) by employing a hydrodynamic formulation Recati et al. (2001); Sinha and Castin (2001). In this formalism the value of α0\alpha_{0} in units of ω\omega_{\perp} satisfies the cubic equation, α03+(1+2Ω0/ω)α0ϵΩ0/ω=0\alpha^{3}_{0}+\left(1+2\Omega_{0}/\omega_{\perp}\right)\alpha_{0}-\epsilon\Omega_{0}/\omega_{\perp}=0, whose solutions for a given free parameter (Ω0\Omega_{0} or ϵ\epsilon), form branches that mark the routes toward the instability in the respective parameter space. The free parameter (Ω0\Omega_{0} or ϵ\epsilon) is varied adiabatically until either a dynamical instability is reached or the α\alpha branch back-bends Madison et al. (2001); Recati et al. (2001); van Bijnen et al. (2009). For α0>0\alpha_{0}>0, the BEC system is always dynamically unstable; otherwise the instability is caused by the branch back-bending. In the current scenario, the spatial dependence of the rotation complicates the hydrodynamic formulation. The phase profile of a rotating density-dependent BEC now also depends locally on the density distribution within the BEC. We, therefore, study the effects of density-dependent gauge potentials on the BEC deformation (α\alpha) and vortex nucleation in a harmonically trapped 2D BEC by employing the split-step Crank Nicholson method Muruganandam and Adhikari (2009). The modifications in the effective mean-field coupling constant due to the gauge potential are neglected.

An initial state for g~=420\tilde{\text{g}}=420 and C~=0,±5\tilde{\mathit{C}}=0,\pm 5 is simulated via imaginary time propagation such that Δx=0.05\Delta x=0.05 and Δt=0.001\Delta t=0.001 are the space and time steps respectively. The initially simulated ground state is then followed in real time by ramping up either Ω0\Omega_{0} or ϵ\epsilon linearly in time, thereby marking different routes for nucleation of vortices. The experimental tuning of interaction parameters, g~\tilde{\text{g}} and C~\tilde{\mathit{C}} through Feshbach resonances Inouye et al. (1998); Blatt et al. (2011) also paves way for controlling the dynamics of the respective condensate in the α\alpha-Ω0\Omega_{0} and α\alpha-ϵ\epsilon parameter spaces. In any case, the number of atoms in the BEC is conserved by fixing the chemical potential after every time step. Moreover, the results remain unchanged for a ground state simulated with C~=0\tilde{\mathit{C}}=0, followed by introducing Ω0\Omega_{0} or ϵ\epsilon ramp and the nonlinear rotation simultaneously in real time.

We first consider the adiabatic increase of trap rotation frequency by making it time-dependent, Ω0(t)=ft\Omega_{0}(t)=ft for f1.6×103ω2f\approx 1.6\times 10^{-3}\omega^{2}_{\perp} and ϵ=0.025\epsilon=0.025.

Refer to caption
Refer to caption
Figure 1: (a) Time development of |α|/ω|\alpha|/\omega_{\perp} for an ascending trap frequency, Ω0(t)\Omega_{0}(t) and ϵ=(ωx2ωy2)/(ωx2+ωy2)=0.025\epsilon=(\omega^{2}_{x}-\omega^{2}_{y})/(\omega^{2}_{x}+\omega^{2}_{y})=0.025. The color coded curves describe the numerical result for C~=CNm/(2πσz)=5,0and5\tilde{\mathit{C}}=\mathit{C}Nm/(\hbar\sqrt{2\pi}\sigma_{z})=5,0~{}\text{and}~{}-5 respectively while the black dashed curves describe the analytical result for C=0\mathit{C}=0 Recati et al. (2001); Sinha and Castin (2001). The discrepancy between the analytical and numerical results for C~=0\tilde{\mathit{C}}=0 correspond to the deviation of the density profile from the Thomas-Fermi limit. Columns I, II and III represent the time evolution of condensate density with C~=5,0,and5\tilde{\mathit{C}}=5,0,~{}\text{and}~{}-5 respectively.

The effect of the density-dependent gauge potentials on the deformation of the BEC is taken into account by defining,

α=Ω~n(𝝆,t)(x2y2)x2+y2\alpha=\frac{{\braket{\tilde{\Omega}_{n}(\boldsymbol{\rho},t)(x^{2}-y^{2})}}}{{\braket{x^{2}+y^{2}}}} (13)

for Ω~n(𝝆,t)=Ω0/ω+C~|ψ(𝝆,t)|2\tilde{\Omega}_{n}(\boldsymbol{\rho},t)=\Omega_{0}/\omega_{\perp}+\tilde{\mathit{C}}|\psi(\boldsymbol{\rho},t)|^{2} and keeping the different parameter values same as mentioned in Ref. Madison et al. (2000). The expressions \braket{\cdots} in Eq. (13) represent the expectation values and hence α\alpha represents the net average deformation experienced by the density-dependent BEC. As the rotation frequency, Ω0\Omega_{0} is ramped up from zero, and the condensate follows the left branch for α\alpha (See Fig 1(a)), except for the small discrepancy due to its non-Thomas-Fermi nature, and becomes increasingly elongated as it goes through the higher values of α\alpha. Once Ωcr\Omega_{\text{cr}} is reached, the elongation in the BEC is maximum and the nucleated vortices at the surface start entering the condensate because of dynamical instability Sinha and Castin (2001); Parker et al. (2006). The nucleated condensate finally returns to an axis-symmetric state specified by the drop in the α\alpha values. In Fig. 1 it is clear that the critical frequency values shift with the nature and strength of nonlinear rotation, C~\tilde{\mathit{C}}. For C~=±5\tilde{\mathit{C}}=\pm 5, the critical frequencies are found to change by ΔΩcr0.02ω\Delta\Omega_{\text{cr}}\approx\mp 0.02\omega_{\perp}respectively against the value of Ωcr0.74ω\Omega_{\text{cr}}\simeq 0.74\omega_{\perp} for C~=0\tilde{\mathit{C}}=0. In addition to shifting the critical frequency, the nonlinear rotation is seen to alter the values of α\alpha such that the condensates with larger C~\tilde{\mathit{C}} values are more elongated while following the left branch. This is directly linked with the number of the nucleated vortices within the condensate and hence, angular momentum. Consequently, the BECs with C~>0\tilde{\mathit{C}}>0 are densely populated with the vortices and have large angular momentum values compared to the ones with C~<0\tilde{\mathit{C}}<0 where the vortices mostly lie toward the periphery. This occurs due to the different rotations and hence energies (E=E0Ω~n(𝝆)LzE=E_{0}-\tilde{\Omega}_{n}(\boldsymbol{\rho})L_{z}) associated with the varying condensate densities. The BECs with C~>0\tilde{\mathit{C}}>0 have their lowest energy regions at the center and hence get populated with vortices profusely. On the other hand, density-dependent BECs with C~<0\tilde{\mathit{C}}<0 have their lowest energy zones at the periphery, thereby populating the vortices around the condensate outskirts. The values of critical frequency, Ωcr\Omega_{\text{cr}} and the maximum condensate deformity, |αmax||\alpha_{max}| in units of ω\omega_{\perp} for the corresponding values of C~\tilde{\mathit{C}} are shown in Fig. 2.

Refer to caption
Figure 2: Variation of the critical frequency, Ωcr\Omega_{\text{cr}}, and maximum deformation, |αmax||\alpha_{max}| with the strength of nonlinear rotation, C~=CNm/(2πσz)\tilde{\mathit{C}}=\mathit{C}Nm/(\hbar\sqrt{2\pi}\sigma_{z}) in a harmonically confined BEC with g~=420\tilde{g}=420.

It is clear that the shifts in critical frequency vary linearly with C~\tilde{\mathit{C}} for C~<0\tilde{\mathit{C}}<0 and small C~>0\tilde{\mathit{C}}>0 values. For large C~>0\tilde{\mathit{C}}>0, the deviation from the linear behavior can be attributed to the non-Thomas-Fermi profiles of the corresponding BECs. In an analogous manner, the maximum condensate deformations, |αmax||\alpha_{max}| corresponding to the Ωcr\Omega_{\text{cr}} values increase with increasing C~\tilde{\mathit{C}} and show anomaly for large values of C~>0\tilde{\mathit{C}}>0.

Next, we consider the adiabatic increase in trap anisotropy by allowing ϵ\epsilon to ramp up linearly from 0 to 0.025 at a rate dϵ/dt6.2×105ωd\epsilon/dt\approx 6.2\times 10^{-5}\omega_{\perp} for a fixed trap rotation frequency, Ω0\Omega_{0}. Like in the case of Ω0\Omega_{0} ramp, the BEC follows the respective α\alpha branch until the vortices start nucleating within the BEC at ϵ=ϵcr\epsilon=\epsilon_{\text{cr}}. In the ENS experiment involving an ϵ\epsilon ramp Madison et al. (2001), the nucleation of vortices in BECs with C~=0\tilde{\mathit{C}}=0 was observed only if Ω0>ω/2\Omega_{0}>\omega_{\perp}/\sqrt{2} whence the α\alpha branch shows back-bending. However, for Ω0<ω/2\Omega_{0}<\omega_{\perp}/\sqrt{2}, α\alpha is a monotonic function of ϵ\epsilon and the numerical simulations Parker et al. (2006) suggest that vortex nucleation can occur at a larger ϵ\epsilon due to ripple instability. Since the net rotation and deformation experienced by the BEC is altered by the presence of density-dependent gauge potentials. Therefore, for a given Ω0\Omega_{0}, the nature of nonlinear rotation determines the branch followed by the BEC in α\alpha-ϵ\epsilon space such that for large values of C~>0\tilde{\mathit{C}}>0, the α\alpha branches exhibit back-bending even if Ω0<ω/2\Omega_{0}<\omega_{\perp}/\sqrt{2}. Similarly with Ω0ω/2\Omega_{0}\geq\omega_{\perp}/\sqrt{2}, the α\alpha branches show no back-bending for large magnitudes of C~<0\tilde{\mathit{C}}<0. Consequently, this regulates the values of ϵcr\epsilon_{\text{cr}} for vortex nucleation.

Refer to caption
Refer to caption
Figure 3: (a) Time development of |α|/ω|\alpha|/\omega_{\perp} for an ascending trap anisotropy, ϵ(t)\epsilon(t) at fixed Ω0\Omega_{0} values. For Ω0/ω=0.65,0.7,0.75\Omega_{0}/\omega_{\perp}=0.65,0.7,0.75, columns I (b)-(d), II (b)-(d) and III (b)-(d) respectively represent the density profiles at mentioned ϵcr\epsilon_{\text{cr}} and C~\tilde{\mathit{C}} values.

Figures 3 and 4 show that for a BEC with given value of C~\tilde{\mathit{C}} and following an α\alpha branch without (with) back-bending, ϵcr\epsilon_{\text{cr}} decreases (increases) with the increase of the trap frequency. A similar effect due by the nonlinear rotation is inferred such that ϵcr(C>0)<ϵcr(C=0)<ϵcr(C<0)\epsilon_{\text{cr}}(\mathit{C}>0)<\epsilon_{\text{cr}}(\mathit{C}=0)<\epsilon_{\text{cr}}(\mathit{C}<0) for a given value of Ω0\Omega_{0} and the BEC follows an α\alpha branch without back-bending. The numerically obtained ϵcr\epsilon_{\text{cr}} values as a function of Ω0\Omega_{0} and C~\tilde{\mathit{C}} as well as the nature of the α\alpha branch followed by the BEC during an ϵ\epsilon ramp are presented in Fig. 4 whereby it is evident that for an α\alpha branch without back-bending, the ϵcr\epsilon_{\text{cr}} values decrease with C~\tilde{\mathit{C}} values for fixed Ω0\Omega_{0}. However, for a BEC following an α\alpha branch with back-bending, ϵcr\epsilon_{\text{cr}} increases with C~\tilde{\mathit{C}} at a given value of Ω0\Omega_{0}.

Refer to caption
Figure 4: Variation of ϵcr\epsilon_{\text{cr}} for corresponding values of Ω0\Omega_{0} and C~=CNm/(2πσz)\tilde{\mathit{C}}=\mathit{C}Nm/(\hbar\sqrt{2\pi}\sigma_{z}) in a BEC with g~=420\tilde{g}=420 and following an α\alpha branch with (BB) or without (NB) back-bending.

This is due to the drastic change in the BEC deformation by the presence of density-dependent gauge potentials and it depends on which side of ω/2\omega_{\perp}/\sqrt{2}, Ω0\Omega_{0} lies. For Ω0<ω/2\Omega_{0}<\omega_{\perp}/\sqrt{2}, the condensate deformation, α\alpha, decreases together with C~\tilde{\mathit{C}} while the situation is reversed with Ω0>ω/2\Omega_{0}>\omega_{\perp}/\sqrt{2}, where the condensate deformation increases with decreasing C~\tilde{\mathit{C}} values. This is nonetheless true for a narrow range of C~\tilde{\mathit{C}} values and anomalies occur at larger magnitudes. In an ϵ\epsilon ramp, the vortex nucleation in BECs with C=0\mathit{C}=0 is shown to occur either by ripple, interbranch or catastrophic instability mechanisms, depending on the value of Ω0\Omega_{0} with respect to ω/2\omega_{\perp}/\sqrt{2} Parker et al. (2006). In case of interbranch and catastrophic instability mechanisms, the respective α\alpha branches show back-bending in the αϵ\alpha-\epsilon parameter space. However, the back-bending of the α\alpha branch in catastrophic instability occurs at a larger ϵ\epsilon compared to that of the interbranch instability. Though the nature of the α\alpha branch followed by the BEC during an ϵ\epsilon ramp is modified by nonlinear rotation. However, in our numerical simulations, we did not observe the interbranch and catastrophic instability mechanisms and hence the role of C~\tilde{\mathit{C}} in tuning these mechanisms for vortex nucleation. As shown in Fig. 3, the vortex nucleation in all cases is due to the generation of ripples at the condensate surface rather than due to the shedding of low-density areas and contortion in interbranch and catastrophic instability mechanisms respectively.

So far we have focused on the response of the BEC with density-dependent gauge potentials to the slow turn on of the rotation of the trapping potential. We now consider the response of the BEC with the above set parameters against the sudden turn on of the rotation of the trap. When a BEC is rotated suddenly, it undergoes damped shape oscillations and ultimately relaxes into a state with vortex-lattice Madison et al. (2000); Tsubota et al. (2002). As shown in Fig. 5, the nonlinear rotation has a prominent effect on these shape oscillations and thereby on the vortex nucleation, vortex dynamics and vortex-lattice formation. The shape oscillations become aperiodic and their amplitude increases for increasing C~>0\tilde{\mathit{C}}>0. On the other hand, the amplitude and time period of the shape oscillations decreases with decreasing C~<0\tilde{\mathit{C}}<0 as shown in Fig. 5(a).

Refer to caption
Refer to caption
Figure 5: (a) Time development of α/ω\alpha/\omega_{\perp} for different strengths of nonlinear rotation, C~\tilde{\mathit{C}} in a BEC suddenly rotated with Ω0=ω/2\Omega_{0}=\omega_{\perp}/\sqrt{2}. In the same scenario with g~=420\tilde{g}=420, columns I-IV for C~=5,0,5,45\tilde{\mathit{C}}=5,0,-5,-45 respectively represent the time development of the BEC density.

The nature of the vortex ground states and their achieving times are controlled by the nature and strength of nonlinear rotation, C~\tilde{\mathit{C}}. For C~=5\tilde{\mathit{C}}=5, the vortex ground state is achieved at short times and the corresponding density profiles show that the BEC relaxes from the point of maximum elongation by shedding low-density material in a spiral pattern —fragmentation Parker and Adams (2005). Consequently, surface waves and ghost vortices are generated which induce vortex nucleation. The vortices in the ground state, as shown in Fig. 5(I)(e), crystallize into a non-Abrikosov lattice which lack the hexagonal symmetry in the vortex arrangements. The BEC becomes disordered for large values of C~>0\tilde{\mathit{C}}>0. In case C~=0\tilde{\mathit{C}}=0, perfect triangular lattices, as shown in Fig. 5(II)(e), are obtained while the non-Abrikosov vortex pattern again occurs for small values of C~<0\tilde{\mathit{C}}<0. The ring-vortex arrangements, shown in Fig. 5(IV)(e), are obtained for large values of C~<0\tilde{\mathit{C}}<0. Moreover, the vortex nucleation in BECs with C~0\tilde{\mathit{C}}\leq 0 occurs through ripples rather than shedding of low density regions in BECs with C~>0\tilde{\mathit{C}}>0. In deformed BECs, the excitations occur on the surfaces with less curvature. Since the BEC deformation decreases together with C~\tilde{\mathit{C}}, the condensates with large values of C~<0\tilde{\mathit{C}}<0 are much symmetric and excitations occur uniformly throughout the surface as shown in Figs. 5(IV)(b) and 5(IV)(c). With regard to the effect on the number of vortices in the equilibrium states due by the nonlinear rotation, Fig. 6 displays the variation of the number of vortices with the strength of nonlinear rotation, C~\tilde{\mathit{C}} for given values of Ω0\Omega_{0}. In the vicinity of Ω0=ω/2\Omega_{0}=\omega_{\perp}/\sqrt{2}, the number of vortices do not change for a larger domain of C~\tilde{\mathit{C}} values, suggesting that the number of vortices in a suddenly rotated density-dependent BEC is fixed by the interaction strength, g, trap ellipticity, ϵ\epsilon, and rotation frequency, Ω0\Omega_{0}, while nonlinear rotation, C~n(𝝆,t)\tilde{\mathit{C}}n(\boldsymbol{\rho},t) only manipulates the position of the vortices within the condensate. However, this is not true for all trap rotation frequencies and the number of vortices also change on either side of C~=0\tilde{\mathit{C}}=0. Except at high trap rotations, the number of vortices change in multiples of two and is attributed to the twofold symmetry of the trapping potential.

Refer to caption
Figure 6: Variation of the number of vortices with C~=CNm/(2πσz)\tilde{\mathit{C}}=\mathit{C}Nm/(\hbar\sqrt{2\pi}\sigma_{z}) for fixed values of Ω0\Omega_{0} in a BEC with g~=420\tilde{g}=420 and confined in an elliptical potential, V(𝐫)=[(1ϵ)x2+(1+ϵ)y2]/2V(\mathbf{r})=[(1-\epsilon)x^{2}+(1+\epsilon)y^{2}]/2 with ϵ=(ωx2ωy2)/(ωx2+ωy2)=0.025\epsilon=(\omega^{2}_{x}-\omega^{2}_{y})/(\omega^{2}_{x}+\omega^{2}_{y})=0.025

.

Further, it is worth mentioning that the vortex nucleation due to nonlinear rotation alone (Ω0=0\Omega_{0}=0) is anticipated for a BEC with large number of atoms Butera et al. (2016a); Butera (2017). Our numerical experiments along this line shown in Fig. 7 displays that the average angular momentum, Lz=ψ|i(y/xx/y)|ψ\braket{L_{z}}=\braket{\psi}{i(y\partial/\partial x-x\partial/\partial y)}{\psi} values are still very small to ensure the formation of vortices within the BEC even for the mean-field interactions as large as, g~=10 000\tilde{g}=10\,000. Neverthless, Lz\braket{L_{z}} increases with C~\tilde{\mathit{C}} and g~\tilde{g}.

Refer to caption
Figure 7: Variation of Lz\braket{L_{z}} with C~\tilde{\mathit{C}} in the equilibrium states of the BEC for different strengths of g~\tilde{g} and no external trap rotation, Ω0=0\Omega_{0}=0. The topmost data point corresponds to a value closer to C~max\tilde{\mathit{C}}_{\text{max}} for the respective case.

We further examined the critical frequencies and the related single vortex dynamics by suddenly rotating the density-dependent BEC with g~=420\tilde{g}=420 in a harmonic trap displaced by x0=0.3ax_{0}=0.3a_{\perp} Kevrekidis et al. (2008):

V(𝝆)=12((xx0)2+y2)V(\boldsymbol{\rho})=\frac{1}{2}\bigg{(}(x-x_{0})^{2}+y^{2}\bigg{)} (14)

The critical frequencies in units of ω\omega_{\perp} are found to be 0.443,0.461,and0.470.443,0.461,\text{and}~{}0.47 for C~=5,0,and5\tilde{C}=5,0,\text{and}-5 respectively. These results are in accordance with the already mentioned results, though the BEC in this case deviates much from its stationary state because of sudden rotation.

Refer to caption
Refer to caption
Figure 8: (a) Single vortex trajectories in BECs with g~=420\tilde{g}=420, C~=0,±5\tilde{\mathit{C}}=0,\pm 5 and rotated with Ω0=0.475ω\Omega_{0}=0.475\omega_{\perp} in xyxy-coordinate space. (b) Time evolution of vortex position in terms of its radial distance |𝐑𝐯|\mathbf{|R_{v}}| from the trap center. The vortex finally sits around the trap minimum at (x0,y0)=(0.3,0.0)a(x_{0},y_{0})=(0.3,0.0)a_{\perp}

The single vortex nucleated at the boundary of the BEC traverses a spiral path before settling down at the minimum of the trap. Because of the density-dependent gauge potentials, the single vortex for a fixed trap rotation follows different trajectories depending on the nature of nonlinear rotation. The effect of nonlinear rotation strengths on the trajectory of a single vortex is presented in Fig. 8, where vortex positions and its distance from the trap center, |𝐑𝐯||\mathbf{R_{v}}| are traced at different instants of time. For fixed Ω0=0.475ω\Omega_{0}=0.475\omega_{\perp}, the single vortex first enters in BEC with C~>0\tilde{\mathit{C}}>0, followed by the ones with C~=0,and5\tilde{\mathit{C}}=0,\text{and}-5 respectively. The first appearances of the vortex in the BECs with C~=5,0,5\tilde{\mathit{C}}=5,0,-5 were at |𝐑𝐯|(a)4.6,4.2,3.9|\mathbf{R_{v}}|(a_{\perp})\approx 4.6,4.2,3.9. This is due to the larger net rotation and the associated size in case of BECs with C~>0\tilde{\mathit{C}}>0. Moreover, in density-dependent BECs, the Magnus force, 𝐟=n(𝝆,t)𝐊×𝐯\mathbf{f}=n(\boldsymbol{\rho},t)\mathbf{K}\times\mathbf{v} for a given density distribution, n(𝝆,t)n(\boldsymbol{\rho},t), circulation, 𝐊\mathbf{K} and vortex velocity, 𝐯=yΩ~n(𝝆,t)e^x+xΩ~n(𝝆,t)e^y\mathbf{v}=-y\tilde{\Omega}_{n}(\boldsymbol{\rho},t)\hat{e}_{x}+x\tilde{\Omega}_{n}(\boldsymbol{\rho},t)\hat{e}_{y} is modified by the nonlinear rotation. In fact, the vorticity, ω=[2Ω~n(𝝆,t)+C~(xn(𝝆,t)/x+yn(𝝆,t)/y)]e^z\mathbf{\omega}=[2\tilde{\Omega}_{n}(\boldsymbol{\rho},t)+\tilde{\mathit{C}}(x\partial n(\boldsymbol{\rho},t)/\partial x+y\partial n(\boldsymbol{\rho},t)/\partial y)]\hat{e}_{z} is now a function of both space and time. The result is that in case of harmonically confined density-dependent BECs with C~>0\tilde{\mathit{C}}>0, the Magnus force is enhanced radially inwards and thereby the vortex moves swiftly toward the trap center, as confirmed in inset of Fig. 8(b). This is contrary to the case when C~<0\tilde{\mathit{C}}<0 where the Magnus force is reduced by the nonlinear rotation and hence the vortex moves slowly toward the trap center, as shown by the corresponding inset in Fig. 8(b). The vortex stays at the boundary for long before getting within the bulk of the condensate. The time taken in units of ω\omega_{\perp} by the single vortex to reach the minimum of the trap is found to be 115,186and221\approx 115,186~{}\text{and}~{}221 respectively for C~=5,0and5\tilde{\mathit{C}}=5,0~{}\text{and}-5. For large negative values of C~\tilde{\mathit{C}}, the Magnus force is outbalanced and the vortices reside at the fast rotating periphery, thereby favoring ring-vortex arrangements.

The formation of ring-vortex arrangements is further supported by the presence of modified repulsive interactions between the vortices within the BEC. In this direction we simulated the equilibrium states with two and three vortices respectively. It is found that the separation between the vortices increases with the strength of C~<0\tilde{\mathit{C}}<0 as shown in Fig. 9.

Refer to caption
Figure 9: Contour-plots showing vortices in density- dependent BEC with C~=±5\tilde{\mathit{C}}=\pm 5 and rotated with (a) Ω0=0.51ω\Omega_{0}=0.51\omega_{\perp} and (b) Ω0=0.525ω\Omega_{0}=0.525\omega_{\perp} respectively. The respective results show same number of vortices with different positions for the same trap rotation frequency.

For C~>0\tilde{\mathit{C}}>0, the vortices come closer, making their separation smaller than C~=0\tilde{\mathit{C}}=0. Consequently, the strength of vortex-vortex repulsion decreases with the increase in the strength of nonlinear rotation and vice-versa. A condensate with the above set of values for the s-wave interactions when rotated at Ω0=0.51ω\Omega_{0}=0.51\omega_{\perp} is nucleated with two vortices as shown in Fig. 9(a). The separation in units of aa_{\perp} between the two vortices for C~=5and5\tilde{\mathit{C}}=5~{}\text{and}-5 are respectively found to be 2.05 and 2.41. These values correspond to an \sim 8% change in mean vortex separation with C~=0\tilde{\mathit{C}}=0. The same results are inferred from the simulations of the ground states with three vortices shown in Fig. 9(b), where the area changes by \sim 15% from the mean value. The change in separation between the vortices can be qualitatively explained by considering the method of images Verhelst and Tempere (2017). This method estimates that the equilibrium separation between two neighboring vortices in a BEC is d=12Ω0/ωd=\frac{1}{\sqrt{2\Omega_{0}/\omega_{\perp}}} as Ω0ω\Omega_{0}\rightarrow\omega_{\perp}. Along similar lines, for a density-dependent BEC d1Ω~n(𝝆,t)d\propto\frac{1}{\sqrt{\tilde{\Omega}_{n}(\boldsymbol{\rho},t)}} where Ω~n(𝝆,t)=Ω0/ω+C~n(𝝆,t)\tilde{\Omega}_{n}(\boldsymbol{\rho},t)=\Omega_{0}/\omega_{\perp}+\tilde{\mathit{C}}n(\boldsymbol{\rho},t). This relation shows that the distance between vortices decreases (increases) with the increasing (decreasing) strength of nonlinear rotation as shown in Fig. 9. Moreover, for a fixed value of C~\tilde{\mathit{C}} the separation between the vortices will also depend on the density distribution of the BEC. For a given value of Ω0\Omega_{0} and C~>0\tilde{\mathit{C}}>0 the vortices in the center will be closer to each other than the ones in the periphery of the BEC. However if C~<0\tilde{\mathit{C}}<0, then the vortices in the center will be distant from each other and for large negative values of C~\tilde{\mathit{C}}, the vortices reside only in the periphery of the BEC. The modifications in the vortex-vortex separation by the density-dependent gauge potentials thus result in the formation of non-Abrikosov vortex lattices and ring-vortex arrangements. Consequently, the areal density of vortices within the stationary states of density-dependent BECs is nonuniform and is regulated by the density distribution of the BEC through nonlinear rotation. These modifications due by the nonlinear rotation have an additional effect on the number of vortices within the BEC. Figure 10 shows the variation of the number of vortices in the BEC trapped in a harmonic potential (14) with C~\tilde{\mathit{C}} for given values of Ω0\Omega_{0}. The results are consistent with those presented in Fig. 6, except that the change in the vortex number now can be even or odd.

Refer to caption
Figure 10: Variation of the number of vortices with C~=CNm/(2πσz)\tilde{\mathit{C}}=\mathit{C}Nm/(\hbar\sqrt{2\pi}\sigma_{z}) for fixed values of Ω0\Omega_{0} in a BEC with g~=420\tilde{g}=420 and V(𝐫)=((xx0)2+y2)/2V(\mathbf{r})=((x-x_{0})^{2}+y^{2})/2.

IV Conclusions

In this work we consider a harmonically trapped 2D BEC subjected to a density-dependent gauge potential and theoretically studied the vortex nucleation in it by employing direct numerical simulations. The presence of density-dependent gauge potential realises nonlinear rotation in BECs. The nonlinear rotation alters the net rotation and deformation experienced by the BEC. This consequently shifts the critical frequencies, Ωcr\Omega_{\text{cr}} for the vortex nucleation. The shifts in the frequency values depend on the strength and nature of the nonlinear rotation. Consequently, the BECs with C>0\mathit{C}>0 experience an increased net rotation and their critical frequencies are lowered. However, in BECs with C<0\mathit{C}<0, the net rotation of the condensate is lowered while the critical frequencies soar higher. In general, for any BEC, the critical frequencies follow the order Ωcr(C>0)<Ωcr(C=0)<Ωcr(C<0)\Omega_{\text{cr}}(\mathit{C}>0)<\Omega_{\text{cr}}(\mathit{C}=0)<\Omega_{\text{cr}}(\mathit{C}<0). The critical frequencies for the transition to vortex-latices in BECs with density-dependent gauge potentials, therefore, depend on the strength of the short range s-wave interactions within them, contrary to the conventional and dipolar BECs with C=0\mathit{C}=0. The nonlinear rotation due to the density-dependent gauge potential also determines the number and arrangement of vortices within the rotating BECs. Moreover, in nonuniform BECs subjected to density-dependent gauge potentials, different density regions experience different rotations. This has a direct impact on the motion of the vortices within the condensate. In comparison to the BECs with C0\mathit{C}\leq 0, it is found that a single vortex in BECs with C>0\mathit{C}>0 moves swiftly and takes less time to reach the trap center when the system attains equilibrium. The Magnus force on the vortices is expected to get modified for the cause. Moreover, the mutual repulsive interactions between the vortices in a multiple vortex system are also modified by the presence of density-dependent gauge potentials. The distance between the vortices and their areal density now depends on the density distribution of the BEC and the strength and sign of nonlinear rotation. The combined effects of these modified forces and the interactions along with energy considerations result in non-Abrikosov vortex-lattices and ring-vortex arrangements.

Like critical frequency, nonlinear rotation due to the density-dependent gauge potentials also regulate the critical ellipticity, ϵcr\epsilon_{\text{cr}} for vortex nucleation. At a fixed Ω0\Omega_{0}, the critical ellipticities follow the trend of critical frequencies i.e, ϵcr(C>0)<ϵcr(C=0)<ϵcr(C<0)\epsilon_{\text{cr}}(\mathit{C}>0)<\epsilon_{\text{cr}}(\mathit{C}=0)<\epsilon_{\text{cr}}(\mathit{C}<0) in case the BEC follows an α\alpha branch without back-bending. However, the trend is reversed for a BEC following an α\alpha branch with back-bending.

It would be interesting to extend the present work by considering density-dependent BECs in optical lattices on top of the harmonic traps. In addition to the surface instability, vortex nucleation in conventional BECs with optical lattices even occurs through the generation of vortex-antivortex pairs in the bulk Kato et al. (2011). The pair generation occurs only if the lattice spacings are greater than the critical values. It is, therefore, natural to investigate the effects of density-dependent gauge potentials on the dynamics of vortex nucleation in such BEC systems. Moreover, the findings can be realized in experiments since the density-dependent gauge potentials have already been demonstrated in two-dimensional optical lattices Clark et al. (2018); Görg et al. (2019).

V Acknowledgements

The authors thank P. G. Kevrekidis for fruitful discussions. B.D. thanks Science and Engineering Research Board, Government of India for funding through research project CRG/2020/003787. I.A.B. acknowledges CSIR, Government of India, for funding via CSIR Research Associateship (09/137(0627)/2020 EMR-I).

References

  • Matthews et al. (1999) M. R. Matthews, B. P. Anderson, P. C. Haljan, D. S. Hall, C. E. Wieman, and E. A. Cornell, Phys. Rev. Lett. 83, 2498 (1999).
  • Madison et al. (2000) K. W. Madison, F. Chevy, W. Wohlleben, and J. Dalibard, Phys. Rev. Lett. 84, 806 (2000).
  • Madison et al. (2001) K. W. Madison, F. Chevy, V. Bretin, and J. Dalibard, Phys. Rev. Lett. 86, 4443 (2001).
  • Hodby et al. (2001) E. Hodby, G. Hechenblaikner, S. A. Hopkins, O. M. Maragò, and C. J. Foot, Phys. Rev. Lett. 88, 010405 (2001).
  • Abo-Shaeer et al. (2001) J. R. Abo-Shaeer, C. Raman, J. M. Vogels, and W. Ketterle, Science 292, 476 (2001).
  • Recati et al. (2001) A. Recati, F. Zambelli, and S. Stringari, Phys. Rev. Lett. 86, 377 (2001).
  • Sinha and Castin (2001) S. Sinha and Y. Castin, Phys. Rev. Lett. 87, 190402 (2001).
  • Donnelly (1991) R. J. Donnelly, Quantized Vortices in Helium II, vol. 2 (Cambridge University Press, Cambridge, UK, 1991).
  • Yukalov et al. (2016) V. Yukalov, A. Novikov, E. Yukalova, and V. S. Bagnato, in Journal of Physics: Conference Series (IOP Publishing, Bristol, 2016), vol. 691, p. 012019.
  • Middleton-Spencer et al. (2022) H. Middleton-Spencer, A. Orozco, L. Galantucci, M. Moreno, N. Parker, L. Machado, V. Bagnato, and C. Barenghi, arXiv:2204.08544 (2022).
  • Sasaki et al. (2010) K. Sasaki, N. Suzuki, and H. Saito, Physical review letters 104, 150404 (2010).
  • Kwon et al. (2015) W. J. Kwon, G. Moon, S. W. Seo, and Y.-i. Shin, Physical Review A 91, 053615 (2015).
  • Fetter (2009) A. L. Fetter, Rev. Mod. Phys. 81, 647 (2009).
  • Dalfovo et al. (1999) F. Dalfovo, S. Giorgini, L. P. Pitaevskii, and S. Stringari, Rev. Mod. Phys. 71, 463 (1999).
  • Ghosh (2004) T. K. Ghosh, Phys. Rev. A 69, 043606 (2004).
  • Mithun et al. (2013) T. Mithun, K. Porsezian, and B. Dey, Phys. Rev. E 88, 012904 (2013).
  • Chevy et al. (2000) F. Chevy, K. W. Madison, and J. Dalibard, Phys. Rev. Lett. 85, 2223 (2000).
  • Parker et al. (2006) N. G. Parker, R. M. W. van Bijnen, and A. M. Martin, Phys. Rev. A 73, 061603 (2006).
  • Cai et al. (2018) Y. Cai, Y. Yuan, M. Rosenkranz, H. Pu, and W. Bao, Phys. Rev. A 98, 023610 (2018).
  • O’Dell and Eberlein (2007) D. H. J. O’Dell and C. Eberlein, Phys. Rev. A 75, 013604 (2007).
  • van Bijnen et al. (2007) R. M. W. van Bijnen, D. H. J. O’Dell, N. G. Parker, and A. M. Martin, Phys. Rev. Lett. 98, 150401 (2007).
  • van Bijnen et al. (2009) R. M. W. van Bijnen, A. J. Dow, D. H. J. O’Dell, N. G. Parker, and A. M. Martin, Phys. Rev. A 80, 033617 (2009).
  • Prasad et al. (2019) S. B. Prasad, T. Bland, B. C. Mulkerin, N. G. Parker, and A. M. Martin, Phys. Rev. Lett. 122, 050401 (2019).
  • Dalibard et al. (2011) J. Dalibard, F. Gerbier, G. Juzeliūnas, and P. Öhberg, Reviews of Modern Physics 83, 1523 (2011).
  • Goldman et al. (2014) N. Goldman, G. Juzeliūnas, P. Öhberg, and I. B. Spielman, Reports on Progress in Physics 77, 126401 (2014).
  • Edmonds et al. (2013) M. J. Edmonds, M. Valiente, G. Juzeliūnas, L. Santos, and P. Öhberg, Phys. Rev. Lett. 110, 085301 (2013).
  • Lin et al. (2011) Y.-J. Lin, K. Jiménez-García, and I. B. Spielman, Nature 471, 83 (2011).
  • Chen et al. (2018) H.-R. Chen, K.-Y. Lin, P.-K. Chen, N.-C. Chiu, J.-B. Wang, C.-A. Chen, P. Huang, S.-K. Yip, Y. Kawaguchi, and Y.-J. Lin, Phys. Rev. Lett. 121, 113204 (2018).
  • Zhang et al. (2019) D. Zhang, T. Gao, P. Zou, L. Kong, R. Li, X. Shen, X.-L. Chen, S.-G. Peng, M. Zhan, H. Pu, et al., Phys. Rev. Lett. 122, 110402 (2019).
  • Clark et al. (2018) L. W. Clark, B. M. Anderson, L. Feng, A. Gaj, K. Levin, and C. Chin, Phys. Rev. Lett. 121, 030402 (2018).
  • Görg et al. (2019) F. Görg, K. Sandholzer, J. Minguzzi, R. Desbuquois, M. Messer, and T. Esslinger, Nat. Phys. 15, 1161 (2019).
  • Keilmann et al. (2011) T. Keilmann, S. Lanzmich, I. McCulloch, and M. Roncaglia, Nat. commun. 2, 1 (2011).
  • Aglietti et al. (1996) U. Aglietti, L. Griguolo, R. Jackiw, S.-Y. Pi, and D. Seminara, Phys. Rev. Lett. 77, 4406 (1996).
  • Bhat et al. (2021) I. A. Bhat, S. Sivaprakasam, and B. A. Malomed, Phys. Rev. E 103, 032206 (2021).
  • Dingwall et al. (2018) R. Dingwall, M. Edmonds, J. Helm, B. Malomed, and P. Öhberg, New Journal of Physics 20, 043004 (2018).
  • Dingwall and Öhberg (2019) R. Dingwall and P. Öhberg, Physical Review A 99, 023609 (2019).
  • Edmonds et al. (2015) M. J. Edmonds, M. Valiente, and P. Öhberg, EPL (Europhysics Letters) 110, 36004 (2015).
  • Chen and Zhu (2022) L. Chen and Q. Zhu, New Journal of Physics 24, 053044 (2022).
  • Edmonds and Nitta (2020) M. Edmonds and M. Nitta, Phys. Rev. A 102, 011303 (2020).
  • Edmonds (2021) M. Edmonds, Phys. Rev. A 104, 043310 (2021).
  • Campbell and Ziff (1979) L. J. Campbell and R. M. Ziff, Physical Review B 20, 1886 (1979).
  • Sato et al. (2007) T. Sato, T. Ishiyama, and T. Nikuni, Physical Review A 76, 053628 (2007).
  • Mithun et al. (2016) T. Mithun, K. Porsezian, and B. Dey, Physical Review A 93, 013620 (2016).
  • Tsubota et al. (2002) M. Tsubota, K. Kasamatsu, and M. Ueda, Phys. Rev. A 65, 023603 (2002).
  • Kasamatsu et al. (2003) K. Kasamatsu, M. Tsubota, and M. Ueda, Phys. Rev. A 67, 033610 (2003).
  • Butera et al. (2016a) S. Butera, M. Valiente, and P. Öhberg, Journal of Physics B: Atomic, Molecular and Optical Physics 49, 015304 (2016a).
  • Butera et al. (2016b) S. Butera, M. Valiente, and P. Öhberg, New Journal of Physics 18, 085001 (2016b).
  • Butera (2017) S. Butera, Ph.D. thesis, Heriot-Watt University (2017).
  • Juzeliūnas et al. (2005) G. Juzeliūnas, P. Öhberg, J. Ruseckas, and A. Klein, Physical Review A 71, 053614 (2005).
  • Yunyi (2005) G. Yunyi, Master’s thesis, National University of Singapore (2005).
  • Petrov et al. (2000) D. S. Petrov, M. Holzmann, and G. V. Shlyapnikov, Phys. Rev. Lett. 84, 2551 (2000).
  • Pérez-García et al. (1998) V. M. Pérez-García, H. Michinel, and H. Herrero, Phys. Rev. A 57, 3837 (1998).
  • Choi et al. (1998) S. Choi, S. Morgan, and K. Burnett, Phys. Rev. A 57, 4057 (1998).
  • Mithun et al. (2014) T. Mithun, K. Porsezian, and B. Dey, Physical Review A 89, 053625 (2014).
  • Muruganandam and Adhikari (2009) P. Muruganandam and S. K. Adhikari, Comput. Phys. Commun. 180, 1888 (2009).
  • Inouye et al. (1998) S. Inouye, M. Andrews, J. Stenger, H.-J. Miesner, D. M. Stamper-Kurn, and W. Ketterle, Nature 392, 151 (1998).
  • Blatt et al. (2011) S. Blatt, T. Nicholson, B. Bloom, J. Williams, J. Thomsen, P. Julienne, and J. Ye, Phys. Rev. Lett. 107, 073202 (2011).
  • Parker and Adams (2005) N. G. Parker and C. S. Adams, Phys. Rev. Lett. 95, 145301 (2005).
  • Kevrekidis et al. (2008) P. G. Kevrekidis, D. J. Frantzeskakis, and R. Carretero-González, Emergent Nonlinear Phenomena in Bose-Einstein Condensates: Theory and Experiment, vol. 45 (Springer, Berlin, 2008).
  • Verhelst and Tempere (2017) N. Verhelst and J. Tempere, in Vortex Dynamics and Optical Vortices, edited by H. Perez-De-Tejada (Vortex Dynamics. Intech, 2017), chap. 1.
  • Kato et al. (2011) A. Kato, Y. Nakano, K. Kasamatsu, and T. Matsui, Phys. Rev. A 84, 053623 (2011).