Unveiling Supersolid Order via Vortex Trajectory Correlations

Subrata Das [email protected]    Vito W. Scarola [email protected] Department of Physics, Virginia Tech, Blacksburg, VA 24061, USA
Abstract

The task of experimentally investigating the inherently dual properties of a supersolid, a simultaneous superfluid and solid, has become more critical following the recent experimental evidence for supersolids in dipolar Bose-Einstein condensates (BECs) of Dy164superscriptDy164{}^{164}\text{Dy}start_FLOATSUPERSCRIPT 164 end_FLOATSUPERSCRIPT Dy. We introduce a supersolid order parameter that uses vortex-vortex trajectory correlations to simultaneously reveal the periodic density of the underlying solid and superfluidity in a single measure. We propose experiments using existing technology to optically create and image trajectories of vortex dipoles in dipolar BECs. We numerically test our observable and find that vortex-vortex correlations reveal the supersolid lattice structure while distinguishing it from superfluidity. Our method sets the stage for experiments to use vortex trajectory correlations to investigate fundamental properties of supersolids arising from their dynamics and phase transitions.

The coexistence of a superfluid and solid at the same time and place defines a supersolid [1, 2, 3, 4]. The superfluid component spontaneously breaks the U(1)𝑈1U(1)italic_U ( 1 ) gauge symmetry of the many-body wavefunction with signatures such as quantized vortices and irrotational flow. Whereas the solid component spontaneously breaks translational symmetry with signatures that include persistent spatial oscillations in particle-particle pair correlations. Combined observation of such overlap** order parameters signals a supersolid, Fig.  1(a).

Definitive signatures of supersolids obtained from globally averaged observables are challenging to interpret. For example, torsional oscillation experiments were proposed to reveal irrotational flow of supersolid He4superscriptHe4{}^{4}\text{He}start_FLOATSUPERSCRIPT 4 end_FLOATSUPERSCRIPT He [5]. Experiments observing expected behavior [6] were brought into question [7, 8] by the prospect of solid dislocations and defects [9, 4] containing superfluid components that mimic irrotational flow expected in a supersolid.

The search for informative experimental probes of supersolids has become more pressing with the advent of engineered supersolidity [10] of ultracold bosonic atoms [11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27], particularly in highly magnetic atoms such as Dy164superscriptDy164{}^{164}\text{Dy}start_FLOATSUPERSCRIPT 164 end_FLOATSUPERSCRIPT Dy. The tunable dipole-dipole interaction between Dy164superscriptDy164{}^{164}\text{Dy}start_FLOATSUPERSCRIPT 164 end_FLOATSUPERSCRIPT Dy atoms allows strong interaction to drive the formation of periodic high-density crystalline structures (quantum droplets) [28, 29, 30, 31] that spontaneously break translational invariance [27]. This solid gives way to a supersolid as the strength of dipolar interaction is lowered to become comparable to kinetic energy. Experiments identifying and probing dysprosium supersolids [15, 19, 17, 16, 18, 21, 20, 22, 23, 24, 25, 26] rely on close support from theory [32, 33, 34, 35, 36, 37, 38, 39, 40].

Refer to caption
Figure 1: (a) Schematic phase diagram where weak particle-particle interaction (compared to kinetic energy) favors a superfluid and strong interaction favors a solid. Overlap** order parameters define a supersolid. (b) The same as (a) but here a non-zero vortex-vortex trajectory correlation function offers a distinct supersolid order parameter. (c) Proposal to create a vortex dipole amid quantum droplets defining a quasi-two dimensional supersolid in a dipolar BEC using a pair of Laguerre-Gaussian beams.

Quantized vortices offer probes of superfluid order and therefore supersolids [41, 42, 43, 44, 45, 46]. Several methods have been used to create vortices in atomic BECs. Rotation by stirring [47, 48, 49] was recently used to observe vortices in rotating supersolid Dy164superscriptDy164{}^{164}\text{Dy}start_FLOATSUPERSCRIPT 164 end_FLOATSUPERSCRIPT Dy [45]. Other methods include quenching [50, 51, 52, 53, 54], optical phase-engineering and holographic methods [55, 56, 57, 58], and Laguerre-Gaussian (LG) beams [59, 60, 61, 62, 63, 64, 65, 66]. Ref. 57 holographically created a vortex dipole (a pair of vortices with opposite angular momentum) [67] in a BEC with no net angular momentum thus avoiding net rotation. Furthermore, vortex dipoles move as quasiparticles making them excellent candidates for recent experimental advances observing vortex trajectories in BECs [68, 69, 70, 71, 72, 73], including the use of non-destructive in situ trajectory imaging [74]. The creation of vortex dipoles and the direct imaging of their trajectories is therefore within reach of near-term experiments in supersolid Dy164superscriptDy164{}^{164}\text{Dy}start_FLOATSUPERSCRIPT 164 end_FLOATSUPERSCRIPT Dy.

We propose vortex-vortex correlations in trajectory imaging as quantitative probes of the fundamental duality inherent to supersolids, Fig.  1(b). Our central observation is that quantized vortices necessitate a superfluid component but also move along density gradient contours [75, 76, 77] to reveal the underlying lattice structure of a supersolid (akin to pair-correlations of the constituent bosons). Fig.  1(c) depicts a proposed experiment wherein a pair of LG beams creates vortex dipoles in a supersolid. We numerically test the efficacy of our setup as a route to extract vortex-vortex trajectory correlations experimentally. We consider uniform [78] and harmonic traps. We find that our correlation function distinguishes the supersolid from the superfluid and, in the uniform trap, reveals the supersolid lattice structure.

Our proposal for a unique supersolid order parameter and its experimental implementation establishes a method for identifying supersolids, measuring lattice properties, tracking phase transitions, and monitoring supersolid lattice dynamics.

Model.— We consider N𝑁Nitalic_N 164Dy atoms trapped in an external potential Vext(𝐫)subscript𝑉ext𝐫V_{\text{ext}}(\mathbf{r})italic_V start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT ( bold_r ) while interacting via contact and dipolar interaction potentials [79, 80, 81]:

Vc(𝐫)=gδ(𝐫) and Vdd(𝐫)=μ0μ24π13cos2θr3,subscript𝑉c𝐫𝑔𝛿𝐫 and subscript𝑉dd𝐫subscript𝜇0superscript𝜇24𝜋13superscript2𝜃superscript𝑟3\displaystyle V_{\rm c}(\mathbf{r})=g\delta(\mathbf{r})\mbox{\quad and\quad}V_% {\rm dd}(\mathbf{r})=\frac{\mu_{0}\mu^{2}}{4\pi}\frac{1-3\cos^{2}\theta}{r^{3}},italic_V start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT ( bold_r ) = italic_g italic_δ ( bold_r ) and italic_V start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT ( bold_r ) = divide start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_μ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_π end_ARG divide start_ARG 1 - 3 roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ end_ARG start_ARG italic_r start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG , (1)

respectively. The contact interaction strength g𝑔gitalic_g is related to the s𝑠sitalic_s-wave scattering length assubscript𝑎𝑠a_{s}italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT by g=4π2as/m𝑔4𝜋superscriptPlanck-constant-over-2-pi2subscript𝑎𝑠𝑚g=4\pi\hbar^{2}a_{s}/mitalic_g = 4 italic_π roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_m where m𝑚mitalic_m is the mass of the atom, μ0subscript𝜇0\mu_{0}italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the vacuum permeability, and μ𝜇\muitalic_μ is the magnetic moment of the atom. The dipolar interaction strength varies with the angle θ𝜃\thetaitalic_θ between the relative position vector 𝐫𝐫\mathbf{r}bold_r and the polarization direction (z𝑧zitalic_z-axis) of the dipole.

We model the dynamics of the entire system with the extended Gross-Pitaevskii equation (eGPE) [28, 82]:

itψ(𝐫)=[\displaystyle\mathrm{i}\hbar\partialderivative{t}\psi(\mathbf{r})=\Big{[}-roman_i roman_ℏ start_DIFFOP divide start_ARG ∂ end_ARG start_ARG ∂ start_ARG italic_t end_ARG end_ARG end_DIFFOP italic_ψ ( bold_r ) = [ - 22m2+Vext(𝐫)+g|ψ(𝐫)|2+γ|ψ(𝐫)|3superscriptPlanck-constant-over-2-pi22𝑚subscript𝑉ext𝐫𝑔superscript𝜓𝐫2𝛾superscript𝜓𝐫3\displaystyle\frac{\hbar^{2}}{2m}\laplacian+V_{\text{ext}}(\mathbf{r})+g% \absolutevalue{\psi(\mathbf{r})}^{2}+\gamma\absolutevalue{\psi(\mathbf{r})}^{3}divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG start_OPERATOR ∇ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_OPERATOR + italic_V start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT ( bold_r ) + italic_g | start_ARG italic_ψ ( bold_r ) end_ARG | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_γ | start_ARG italic_ψ ( bold_r ) end_ARG | start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT
+d𝐫Vdd(𝐫𝐫)|ψ(𝐫)|2]ψ(𝐫).\displaystyle+\int\differential{\mathbf{r^{\prime}}}V_{\rm dd}(\mathbf{r}-% \mathbf{r^{\prime}})\absolutevalue{\psi(\mathbf{r^{\prime}})}^{2}\Big{]}\psi(% \mathbf{r}).+ ∫ roman_d start_ARG bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG italic_V start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT ( bold_r - start_ID bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ID ) | start_ARG italic_ψ ( start_ID bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ID ) end_ARG | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_ψ ( bold_r ) . (2)

Here, ψ(𝐫)𝜓𝐫\psi(\mathbf{r})italic_ψ ( bold_r ) is the system wavefunction normalized with d𝐫|ψ|2=N𝐫superscript𝜓2𝑁\int\differential{\mathbf{r}}\absolutevalue{\psi}^{2}=N∫ roman_d start_ARG bold_r end_ARG | start_ARG italic_ψ end_ARG | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = italic_N. The Lee-Huang-Yang [83, 84, 85] correction term contains γ(ϵdd)=(128π2as5/2/3m)(1+32ϵdd2)𝛾subscriptitalic-ϵdd128𝜋superscriptPlanck-constant-over-2-pi2superscriptsubscript𝑎𝑠523𝑚132superscriptsubscriptitalic-ϵdd2\gamma(\epsilon_{\rm dd})=(128\sqrt{\pi}\hbar^{2}a_{s}^{5/2}/3m)(1+\frac{3}{2}% \epsilon_{\rm dd}^{2})italic_γ ( italic_ϵ start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT ) = ( 128 square-root start_ARG italic_π end_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 5 / 2 end_POSTSUPERSCRIPT / 3 italic_m ) ( 1 + divide start_ARG 3 end_ARG start_ARG 2 end_ARG italic_ϵ start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) [86, 87], which depends on the ratio of the dipolar length scale to assubscript𝑎𝑠a_{s}italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, ϵdd=add/assubscriptitalic-ϵddsubscript𝑎ddsubscript𝑎𝑠\epsilon_{\rm dd}=a_{\rm dd}/a_{s}italic_ϵ start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT = italic_a start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT / italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, where add=μ0μ2m/(12π2)subscript𝑎ddsubscript𝜇0superscript𝜇2𝑚12𝜋superscriptPlanck-constant-over-2-pi2a_{\rm dd}=\mu_{0}\mu^{2}m/(12\pi\hbar^{2})italic_a start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT = italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_μ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_m / ( 12 italic_π roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ). ϵddsubscriptitalic-ϵdd\epsilon_{\rm dd}italic_ϵ start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT can be tuned using a Feshbach resonance in 164Dy to vary assubscript𝑎𝑠a_{s}italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT [88, 27] thus allowing exploration of phase transitions [82]. For 164Dy, the dipole moment μ=9.93μB𝜇9.93subscript𝜇𝐵\mu{=}9.93\mu_{B}italic_μ = 9.93 italic_μ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT, sets a dipolar length scale add=131a0subscript𝑎dd131subscript𝑎0a_{\rm dd}{=}131a_{0}italic_a start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT = 131 italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT where μBsubscript𝜇𝐵\mu_{B}italic_μ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT and a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT are the Bohr magneton and Bohr radius, respectively. We choose as=94a0(ϵdd=1.39)subscript𝑎𝑠94subscript𝑎0subscriptitalic-ϵdd1.39a_{s}=94a_{0}(\epsilon_{\rm dd}=1.39)italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 94 italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_ϵ start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT = 1.39 ) for and 110a0(ϵdd=1.19)110subscript𝑎0subscriptitalic-ϵdd1.19110a_{0}(\epsilon_{\rm dd}=1.19)110 italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_ϵ start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT = 1.19 ) to achieve the supersolid and superfluid phases, respectively [16].

To study vortex dynamics in superfluid and supersolid phases, we assume two different external trap** potentials: (i) uniform in the xy𝑥𝑦xyitalic_x italic_y plane and harmonic along z𝑧zitalic_z-axis, Vext=mωz2z2/2subscript𝑉ext𝑚superscriptsubscript𝜔𝑧2superscript𝑧22V_{\rm ext}{=}m\omega_{z}^{2}z^{2}/2italic_V start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT = italic_m italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 and (ii) an oblate 3D harmonic trap, Vext=mω2(x2+y2+λ2z2)/2subscript𝑉ext𝑚superscript𝜔2superscript𝑥2superscript𝑦2superscript𝜆2superscript𝑧22V_{\rm ext}{=}m\omega^{2}(x^{2}+y^{2}+\lambda^{2}z^{2})/2italic_V start_POSTSUBSCRIPT roman_ext end_POSTSUBSCRIPT = italic_m italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) / 2 where trap** anisotropy λ𝜆\lambdaitalic_λ is the ratio of trap** frequency along the z𝑧zitalic_z-axis to that in the xy𝑥𝑦xyitalic_x italic_y plane. In both cases, we induce vortex dipoles using the phase-imprinting method [89, 90] by multiplying the wavefunction ψ(𝐫)𝜓𝐫\psi(\mathbf{r})italic_ψ ( bold_r ) with a phase factor of exp[il(yy0)/(xx0)]i𝑙𝑦subscript𝑦0𝑥subscript𝑥0\exp[\mathrm{i}l(y-y_{0})/(x-x_{0})]roman_exp [ roman_i italic_l ( italic_y - italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / ( italic_x - italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ]. Here l𝑙Planck-constant-over-2-pil\hbaritalic_l roman_ℏ, with l=±1𝑙plus-or-minus1l=\pm 1italic_l = ± 1, is the vortex angular momentum and (x0,y0)subscript𝑥0subscript𝑦0(x_{0},y_{0})( italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) is the vortex initial position.

Vortex motion in a uniform trap.— We first consider a uniform planar trap to ignore the effect of density variation due to trap** confinement. We solve the eGPE  \tagform@2 using the split-step Crank-Nicolson method [91] and study the vortex dynamics in the system by tracking the vortex trajectories over time. In the uniform trap, N=1.3×106𝑁1.3superscript106N{=}1.3\times 10^{6}italic_N = 1.3 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT atoms are trapped along the z𝑧zitalic_z-axis with ωz=2π×150subscript𝜔𝑧2𝜋150\omega_{z}{=}2\pi\times 150italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 2 italic_π × 150 Hz, and hard walls in the xy𝑥𝑦xyitalic_x italic_y plane. The simulation box size is Lx=Ly=2Lz=32.8μsubscript𝐿𝑥subscript𝐿𝑦2subscript𝐿𝑧32.8𝜇L_{x}{=}L_{y}{=}2L_{z}{=}32.8\muitalic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT = italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT = 2 italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 32.8 italic_μm and the grid size is (256×256×128)256256128(256\times 256\times 128)( 256 × 256 × 128 ).

Refer to caption
Figure 2: Groundstate density profile in the z=0𝑧0z=0italic_z = 0 plane and vortex motion in a uniform trap in (a) supersolid and (b) superfluid phases. vi+superscriptsubscriptv𝑖\rm v_{\it{i}}^{+}roman_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT and visuperscriptsubscriptv𝑖\rm v_{\it{i}}^{-}roman_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT are the initial position of the l=+1𝑙1l=+1italic_l = + 1 and l=1𝑙1l=-1italic_l = - 1 vortex of i𝑖iitalic_i-th vortex dipole. The red (green) line represents the trajectory of a l=+1𝑙1l=+1italic_l = + 1 (l=1𝑙1l=-1italic_l = - 1) vortex. The black arrows in (b) represent the direction of motion of vortex dipoles in the superfluid phase. In (b), the simulations of vortex dipoles (v1+,v1)superscriptsubscriptv1superscriptsubscriptv1(\rm v_{1}^{+},v_{1}^{-})( roman_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT , roman_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) and (v2+,v2)superscriptsubscriptv2superscriptsubscriptv2(\rm v_{2}^{+},v_{2}^{-})( roman_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT , roman_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) are done separately and the background density is the product of two groundstates to identify the vortices.

By solving the eGPE, we obtain the superfluid and supersolid ground state for different ϵddsubscriptitalic-ϵdd\epsilon_{\rm dd}italic_ϵ start_POSTSUBSCRIPT roman_dd end_POSTSUBSCRIPT. In a uniform trap the superfluid phase has a flat density profile whereas the supersolid phase shows a density modulation. The supersolid crystalline structure of quantum droplets contains a low density background, Fig.  1(c).

Vortices follow qualitatively distinct trajectories in the supersolid and superfluid phases. To see this, we dynamically imprint a vortex dipole. The vortex dipole will have a linear momentum in the superfluid phase which is inversely proportional to the separation d𝑑ditalic_d of l=+1𝑙1l=+1italic_l = + 1 and l=1𝑙1l=-1italic_l = - 1 vortices [92, 93]. Both vortices therefore move in straight lines parallel to each other with a constant velocity in the superfluid. In contrast, in the supersolid phase, the vortex trajectories depend on non-uniform and periodic density patterns. In the vortex’s frame of reference, density changes with time and space leading to a non-uniform force on the vortices. Due to this non-uniform force, vortices move in non-linear contours outlining the supersolid density gradient. In contrast to the superfluid phase, the vortex dipole is not stable in the supersolid phase and eventually annihilates when the l=+1𝑙1l=+1italic_l = + 1 and l=1𝑙1l=-1italic_l = - 1 vortices approach each other. We test these qualitative expectations quantitatively using eGPE.

Figure  2 shows the vortex trajectories in the superfluid and supersolid phases. We choose a vortex dipole (v1+,v1)superscriptsubscriptv1superscriptsubscriptv1(\rm v_{1}^{+},v_{1}^{-})( roman_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT , roman_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) in both phases as shown in Figs.  2(a) and  2(b). The l=+1𝑙1l=+1italic_l = + 1 and l=1𝑙1l=-1italic_l = - 1 vortices follow non-uniform trajectories and eventually annihilate in the supersolid phase. While in the superfluid phase, they move in straight lines parallel to each other. The vortex dipole motion in the superfluid phase is stable, and they do not annihilate.

We verify that the velocity of the vortices is inversely proportional to the separation. In the superfluid phase we prepare a well separated l=±1𝑙plus-or-minus1l=\pm 1italic_l = ± 1 vortex pair (v1+,v1)superscriptsubscriptv1superscriptsubscriptv1(\rm v_{1}^{+},v_{1}^{-})( roman_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT , roman_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ). Fig.  2(b) shows that the distances traveled by this vortex dipole are small. In Fig.  2(b), we also show motion of a closely spaced vortex dipole (v2+,v2)superscriptsubscriptv2superscriptsubscriptv2(\rm v_{2}^{+},v_{2}^{-})( roman_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT , roman_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) in the superfluid phase (this is studied separately from (v1+,v1)superscriptsubscriptv1superscriptsubscriptv1(\rm v_{1}^{+},v_{1}^{-})( roman_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT , roman_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ), not shown in Fig.  2(a) for clarity of the figure). These vortices traveled a longer distance (full paths are not shown) compared to the vortex dipole (v1+,v1)superscriptsubscriptv1superscriptsubscriptv1(\rm v_{1}^{+},v_{1}^{-})( roman_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT , roman_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ).

Correlation of vortex trajectories.— We now turn to our central proposal. Given distinct behavior of vortex motion in each phase we seek to extract quantitative information from the vortex trajectories. We define a vortex-vortex trajectory correlation function:

C(𝐫)=s(𝐫0)s(𝐫0+𝐫),𝐶𝐫expectation-value𝑠subscript𝐫0𝑠subscript𝐫0𝐫\displaystyle C(\mathbf{r})=\expectationvalue{s(\mathbf{r}_{0})s(\mathbf{r}_{0% }+\mathbf{r})},italic_C ( bold_r ) = ⟨ start_ARG italic_s ( bold_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) italic_s ( bold_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + bold_r ) end_ARG ⟩ , (3)

where s(𝐫)𝑠𝐫s(\mathbf{r})italic_s ( bold_r ) is 1 if a vortex of any l𝑙litalic_l passes through 𝐫𝐫\mathbf{r}bold_r and 0 otherwise. In the following calculations we choose to compute C(𝐫)𝐶𝐫C(\mathbf{r})italic_C ( bold_r ) using trajectories of opposite l𝑙litalic_l with no loss of generality. Note that C(𝐫)𝐶𝐫C(\mathbf{r})italic_C ( bold_r ) vanishes in the quantum droplet phase (because there is no phase coherence for vortices to form) and should show only trivial structure in the superfluid phase because the vortices move in straight lines. But in the supersolid phase, measurements of C(𝐫)𝐶𝐫C(\mathbf{r})italic_C ( bold_r ) will probe the underlying crystal structure of the supersolid. C(𝐫)𝐶𝐫C(\mathbf{r})italic_C ( bold_r ) can therefore be used as a supersolid order parameter. Importantly, the initial configuration of vortices must be chosen to break the underlying translational symmetry of the supersolid phase, otherwise straight-line trajectories will arise [94]. Straight trajectories will not allow measurements to distinguish between superfluid and supersolid states. In the following, we choose the initial positions of the vortex dipoles on the opposite arcs of a circle in the simulations to avoid biased choices in probing the quantum droplet square lattice in the supersolid phase.

We test the utility of C(𝐫)𝐶𝐫C(\mathbf{r})italic_C ( bold_r ) using eGPE simulations designed to replicate repeated measurements in a uniform trap. Figs.  3(a) and  3(b) show nine separate simulations of the resulting motion of the vortex dipoles [(v1+,v1),,(v9+,v9)]superscriptsubscriptv1superscriptsubscriptv1superscriptsubscriptv9superscriptsubscriptv9[(\rm v_{1}^{+},v_{1}^{-}),\cdots,(\rm v_{9}^{+},v_{9}^{-})][ ( roman_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT , roman_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) , ⋯ , ( roman_v start_POSTSUBSCRIPT 9 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT , roman_v start_POSTSUBSCRIPT 9 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) ] in both superfluid and supersolid phases. For all the vortex dipoles, the l=+1𝑙1l=+1italic_l = + 1 and l=1𝑙1l=-1italic_l = - 1 vortex separation d13μm𝑑13𝜇md\approx 13{\mu\rm m}italic_d ≈ 13 italic_μ roman_m is the same as shown by the initial position of the l=+1𝑙1l=+1italic_l = + 1 (l=1𝑙1l=-1italic_l = - 1) vortex by red (green) dots in Figs.  3(a) and  3(b). We locate the vortex position at different instants of time and find the trajectories of the vortices as time evolves. We find different trajectories for different sets of initial positions of the vortices. By merging all of the trajectories we obtain a combined 2D trajectory on the xy𝑥𝑦xyitalic_x italic_y plane shown in Figs.  3(a) and  3(b).

Refer to caption
Figure 3: Combined vortex trajectories of nine vortex dipoles [(v1+,v1),,(v9+,v9)]superscriptsubscriptv1superscriptsubscriptv1superscriptsubscriptv9superscriptsubscriptv9[(\rm v_{1}^{+},v_{1}^{-}),\cdots,(\rm v_{9}^{+},v_{9}^{-})][ ( roman_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT , roman_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) , ⋯ , ( roman_v start_POSTSUBSCRIPT 9 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT , roman_v start_POSTSUBSCRIPT 9 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) ] from separate simulations in (a) supersolid and (b) superfluid phases. The red (green) dots represent the initial position of the l=+1𝑙1l=+1italic_l = + 1 (l=1𝑙1l=-1italic_l = - 1) vortex of the vortex dipoles. The grey lines represent the vortex trajectories. The correlation function C(𝐫)𝐶𝐫C(\mathbf{r})italic_C ( bold_r ) of the vortex trajectories in (c) superfluid and (d) supersolid phases. (c) and (d) show the correlation function C(𝐫)𝐶𝐫C(\mathbf{r})italic_C ( bold_r ) of the vortex trajectories in the supersolid and superfluid phases, respectively. In (c), apart from the primary peak at 𝐫=0𝐫0\mathbf{r}=0bold_r = 0, there are the secondary peaks at 𝐫0𝐫0\mathbf{r}\neq 0bold_r ≠ 0 denoted by yellow arrows for C(0,y)𝐶0𝑦C(0,y)italic_C ( 0 , italic_y ). In contrast, in (d), the correlation function C(𝐫)𝐶𝐫C(\mathbf{r})italic_C ( bold_r ) shows only the primary peak at 𝐫=0𝐫0\mathbf{r}=0bold_r = 0. All simulation parameters are the same as in Fig.  2.

We find that even with a limited number of trajectories, C(𝐫)𝐶𝐫C(\mathbf{r})italic_C ( bold_r ) reveals aspects of the underlying lattice structure of the supersolid. Fig.  3(c) shows that the correlation function C(𝐫)𝐶𝐫C(\mathbf{r})italic_C ( bold_r ), apart from the primary trivial peak at 𝐫=0𝐫0\mathbf{r}=0bold_r = 0, shows secondary peaks at 𝐫0𝐫0\mathbf{r}\neq 0bold_r ≠ 0 due to the non-uniform and periodic density patterns in the supersolid phase. Whereas in the superfluid phase, C(𝐫)𝐶𝐫C(\mathbf{r})italic_C ( bold_r ) shows only the primary peak at 𝐫=0𝐫0\mathbf{r}=0bold_r = 0 due to the straight-line motion of the vortices. Thus, the correlation function C(𝐫)𝐶𝐫C(\mathbf{r})italic_C ( bold_r ) of vortex trajectories is a useful tool to distinguish the superfluid and supersolid phases.

Use of C(𝐫)𝐶𝐫C(\mathbf{r})italic_C ( bold_r ) to probe a supersolid requires the following conditions to be met: (i) As mentioned above, initial positions must break the supersolid crystal symmetry. If initial vortex positions are chosen, for example, to be random, a sufficient number of trajectories must be imaged to resolve auxiliary peaks. The limited number of runs in Fig.  3 were chosen near this threshold. (ii) Vortices are, ideally, to be used as non-invasive probes that leave the supersolid intact. We assume an LG beam spot size (1μmsimilar-toabsent1𝜇m\sim 1\mu\text{m}∼ 1 italic_μ m) below the supersolid lattice spacing (3μmsimilar-toabsent3𝜇m\sim 3\mu\text{m}∼ 3 italic_μ m in our simulations) and with initial inter-vortex spacing high enough to avoid high vortex speeds. (iii) We assume the use of a uniform trap (or a sufficiently large number of atoms) to allow resolution of translational symmetry breaking. The supplementary material [94] shows data relaxing condition (i) in order to map the supersolid lattice whereas the following simulations relax conditions (ii) and (iii).

Vortex motion in a harmonic trap.— We now turn to small system sizes and harmonic trap** relevant for ongoing experiments. Tight trap** will limit the range of the solid and can even lead to spurious supersolid signatures [95]. We choose N=80000𝑁80000N{=}80000italic_N = 80000 atoms to be trapped in an oblate-shaped 3D harmonic trap with trap** frequencies (ω,ωz)=2π×(45,133)𝜔subscript𝜔𝑧2𝜋45133(\omega,\omega_{z}){=}2\pi\times(45,133)( italic_ω , italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) = 2 italic_π × ( 45 , 133 ) Hz [34], where ω𝜔\omegaitalic_ω is the transverse trap** frequency. The simulation box size is Lx=Ly=Lz=30μsubscript𝐿𝑥subscript𝐿𝑦subscript𝐿𝑧30𝜇L_{x}{=}L_{y}{=}L_{z}{=}30\muitalic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT = italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT = italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 30 italic_μm.

We studied the motion of vortex dipoles in both superfluid (as=110a0subscript𝑎𝑠110subscript𝑎0a_{s}=110a_{0}italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 110 italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) and supersolid (as=94a0subscript𝑎𝑠94subscript𝑎0a_{s}=94a_{0}italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 94 italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) regimes. Without any vortices, the density profile of the superfluid is Gaussian, whereas the supersolid regime shows a single quantum droplet in the center of the trap with a low-density superfluid background [16, 18]. In both phases, we prepare ground states with a vortex dipole (v1+,v1)superscriptsubscriptv1superscriptsubscriptv1(\rm v_{1}^{+},v_{1}^{-})( roman_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT , roman_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) separated by a distance d6μm𝑑6𝜇md\approx 6\mu\rm mitalic_d ≈ 6 italic_μ roman_m, and we observe the motion of the vortices by tracking the vortex trajectories over time.

Refer to caption
Figure 4: (a) Snapshots of density profiles in the z=0𝑧0z=0italic_z = 0 plane at different times in the harmonic trap after the vortex dipole is imprinted in the supersolid phase. The red (green) circled “+”(“-”) marker represents the l=+1𝑙1l=+1italic_l = + 1 (l=1𝑙1l=-1italic_l = - 1) vortex. (b) and (c) show the l=+1𝑙1l=+1italic_l = + 1 (red dots) and l=1𝑙1l=-1italic_l = - 1 (green dots) vortex trajectories in the supersolid and superfluid regimes, respectively.

Figure  4(a) shows density profiles in the supersolid regime after the vortex dipole is imprinted. At t=0𝑡0t=0italic_t = 0 the creation of the vortices causes the single droplet at the center to split into multiple droplets. Over time, these droplets continuously deform and form at different locations due to the motion of the vortices. Here the dynamics explores a variety of metastable small supersolid configurations. Once the l=±1𝑙plus-or-minus1l=\pm 1italic_l = ± 1 vortices annihilate each other, the small supersolid reaches an equilibrium state with three droplets, as we see at t=700ms𝑡700mst=700\,\rm msitalic_t = 700 roman_ms in  4(a). The trajectories of the vortices are affected by the density modulation in the supersolid regime, moving in non-linear paths before annihilation [Fig.  4(b)]. In contrast, the vortices in the superfluid regime do not annihilate. They keep moving in an elliptical path with a constant speed, as shown in Fig.  4(c) [96, 97]. We therefore see that vortex dipole trajectories in the supersolid and superfluid regimes show qualitatively distinct behavior even in small systems with tight trap**.

Discussion.— We introduced and numerically tested a supersolid order parameter defined by correlations in vortex trajectories. Our tests used a limited number of trajectories to replicate sampling for near term experiments. Larger numbers of samples allow resolution of complete supersolid lattice structures [94]. Our correlation function can also be used to study important properties of supersolids, including lattice dynamics due to supersolid phonons and phase transitions. Possible applications to different systems include exploration of broken translational symmetry in thin films of superfluid He4superscriptHe4{}^{4}\text{He}start_FLOATSUPERSCRIPT 4 end_FLOATSUPERSCRIPT He [98, 99, 100] and Fulde-Ferrell-Larkin-Ovchinnikov [101, 102] superconductors.

Acknowledgements.
S.D. thanks Soumyadeep Halder and Bankim C. Das for the useful discussions. We acknowledge support from AFOSR-FA9550-23-1-0034, FA9550-19-1-0272 and ARO-W911NF2210247.

References

  • Gross [1957] E. P. Gross, Unified Theory of Interacting Bosons, Phys. Rev. 106, 161 (1957).
  • Andreev and Lifshitz [1969] A. F. Andreev and I. M. Lifshitz, Quantum theory of defects in crystals, Soviet Physics JETP 29, 1107 (1969).
  • Chester [1970] G. V. Chester, Speculations on Bose-Einstein Condensation and Quantum Crystals, Phys. Rev. A 2, 256 (1970).
  • Boninsegni and Prokof’ev [2012] M. Boninsegni and N. V. Prokof’ev, Colloquium: Supersolids: What and where are they?, Rev. Mod. Phys. 84, 759 (2012).
  • Leggett [1970] A. J. Leggett, Can a Solid Be “Superfluid”?, Phys. Rev. Lett. 25, 1543 (1970).
  • Kim and Chan [2004] E. Kim and M. H. W. Chan, Probable observation of a supersolid helium phase, Nature 427, 225 (2004).
  • Kim and Chan [2012] D. Y. Kim and M. H. W. Chan, Absence of supersolidity in solid helium in porous vycor glass, Phys. Rev. Lett. 109, 155301 (2012).
  • Chan et al. [2013] M. H. W. Chan, R. B. Hallock, and L. Reatto, Overview on solid 4he and the issue of supersolidity, Journal of Low Temperature Physics 172, 317 (2013).
  • Pollet et al. [2007] L. Pollet, M. Boninsegni, A. B. Kuklov, N. V. Prokof’ev, B. V. Svistunov, and M. Troyer, Superfluidity of grain boundaries in solid 4He, Phys. Rev. Lett. 98, 135301 (2007).
  • Scarola and Das Sarma [2005] V. W. Scarola and S. Das Sarma, Quantum phases of the extended bose-hubbard hamiltonian: Possibility of a supersolid state of cold atoms in optical lattices, Phys. Rev. Lett. 95, 033003 (2005).
  • Lu et al. [2011] M. Lu, N. Q. Burdick, S. H. Youn, and B. L. Lev, Strongly Dipolar Bose-Einstein Condensate of Dysprosium, Phys. Rev. Lett. 107, 190401 (2011).
  • Léonard et al. [2017a] J. Léonard, A. Morales, P. Zupancic, T. Esslinger, and T. Donner, Supersolid formation in a quantum gas breaking a continuous translational symmetry, Nature 543, 87 (2017a).
  • Léonard et al. [2017b] J. Léonard, A. Morales, P. Zupancic, T. Donner, and T. Esslinger, Monitoring and manipulating Higgs and Goldstone modes in a supersolid quantum gas, Science 358, 1415 (2017b).
  • Li et al. [2017] J.-R. Li, J. Lee, W. Huang, S. Burchesky, B. Shteynas, F. Ç. Top, A. O. Jamison, and W. Ketterle, A stripe phase with supersolid properties in spin–orbit-coupled Bose–Einstein condensates, Nature 543, 91 (2017).
  • Tanzi et al. [2019a] L. Tanzi, E. Lucioni, F. Famà, J. Catani, A. Fioretti, C. Gabbanini, R. N. Bisset, L. Santos, and G. Modugno, Observation of a Dipolar Quantum Gas with Metastable Supersolid Properties, Phys. Rev. Lett. 122, 130405 (2019a).
  • Böttcher et al. [2019] F. Böttcher, J.-N. Schmidt, M. Wenzel, J. Hertkorn, M. Guo, T. Langen, and T. Pfau, Transient Supersolid Properties in an Array of Dipolar Quantum Droplets, Phys. Rev. X 9, 011051 (2019).
  • Tanzi et al. [2019b] L. Tanzi, S. M. Roccuzzo, E. Lucioni, F. Famà, A. Fioretti, C. Gabbanini, G. Modugno, A. Recati, and S. Stringari, Supersolid symmetry breaking from compressional oscillations in a dipolar quantum gas, Nature 574, 382 (2019b).
  • Chomaz et al. [2019] L. Chomaz, D. Petter, P. Ilzhöfer, G. Natale, A. Trautmann, C. Politi, G. Durastante, R. M. W. van Bijnen, A. Patscheider, M. Sohmen, M. J. Mark, and F. Ferlaino, Long-Lived and Transient Supersolid Behaviors in Dipolar Quantum Gases, Phys. Rev. X 9, 021012 (2019).
  • Guo et al. [2019] M. Guo, F. Böttcher, J. Hertkorn, J.-N. Schmidt, M. Wenzel, H. P. Büchler, T. Langen, and T. Pfau, The low-energy Goldstone mode in a trapped dipolar supersolid, Nature 574, 386 (2019).
  • Norcia et al. [2021] M. A. Norcia, C. Politi, L. Klaus, E. Poli, M. Sohmen, M. J. Mark, R. N. Bisset, L. Santos, and F. Ferlaino, Two-dimensional supersolidity in a dipolar quantum gas, Nature 596, 357 (2021).
  • Ilzhöfer et al. [2021] P. Ilzhöfer, M. Sohmen, G. Durastante, C. Politi, A. Trautmann, G. Natale, G. Morpurgo, T. Giamarchi, L. Chomaz, M. J. Mark, and F. Ferlaino, Phase coherence in out-of-equilibrium supersolid states of ultracold dipolar atoms, Nat. Phys. 17, 356 (2021).
  • Sohmen et al. [2021] M. Sohmen, C. Politi, L. Klaus, L. Chomaz, M. J. Mark, M. A. Norcia, and F. Ferlaino, Birth, Life, and Death of a Dipolar Supersolid, Phys. Rev. Lett. 126, 233401 (2021).
  • Tanzi et al. [2021] L. Tanzi, J. G. Maloberti, G. Biagioni, A. Fioretti, C. Gabbanini, and G. Modugno, Evidence of superfluidity in a dipolar supersolid from nonclassical rotational inertia, Science 371, 1162 (2021).
  • Norcia et al. [2022] M. A. Norcia, E. Poli, C. Politi, L. Klaus, T. Bland, M. J. Mark, L. Santos, R. N. Bisset, and F. Ferlaino, Can Angular Oscillations Probe Superfluidity in Dipolar Supersolids?, Phys. Rev. Lett. 129, 040403 (2022).
  • Bland et al. [2022a] T. Bland, E. Poli, C. Politi, L. Klaus, M. A. Norcia, F. Ferlaino, L. Santos, and R. N. Bisset, Two-Dimensional Supersolid Formation in Dipolar Condensates, Phys. Rev. Lett. 128, 195302 (2022a).
  • Sánchez-Baena et al. [2023] J. Sánchez-Baena, C. Politi, F. Maucher, F. Ferlaino, and T. Pohl, Heating a dipolar quantum fluid into a solid, Nat Commun 14, 1868 (2023).
  • Chomaz et al. [2023] L. Chomaz, I. Ferrier-Barbut, F. Ferlaino, B. Laburthe-Tolra, B. L. Lev, and T. Pfau, Dipolar physics: A review of experiments with magnetic quantum gases, Rep. Prog. Phys. 86, 026401 (2023).
  • Wächtler and Santos [2016a] F. Wächtler and L. Santos, Quantum filaments in dipolar Bose-Einstein condensates, Phys. Rev. A 93, 061603 (2016a).
  • Wächtler and Santos [2016b] F. Wächtler and L. Santos, Ground-state properties and elementary excitations of quantum droplets in dipolar Bose-Einstein condensates, Phys. Rev. A 94, 043618 (2016b).
  • Baillie and Blakie [2018] D. Baillie and P. B. Blakie, Droplet Crystal Ground States of a Dipolar Bose Gas, Physical Review Letters 121, 195301 (2018).
  • Mishra et al. [2020] C. Mishra, L. Santos, and R. Nath, Self-Bound Doubly Dipolar Bose-Einstein Condensates, Physical Review Letters 124, 073402 (2020).
  • Roccuzzo and Ancilotto [2019] S. M. Roccuzzo and F. Ancilotto, Supersolid behavior of a dipolar Bose-Einstein condensate confined in a tube, Phys. Rev. A 99, 041601 (2019).
  • Blakie et al. [2020] P. B. Blakie, D. Baillie, L. Chomaz, and F. Ferlaino, Supersolidity in an elongated dipolar condensate, Phys. Rev. Res. 2, 043318 (2020).
  • Halder et al. [2022] S. Halder, K. Mukherjee, S. I. Mistakidis, S. Das, P. G. Kevrekidis, P. K. Panigrahi, S. Majumder, and H. R. Sadeghpour, Control of Dy164superscriptDy164{}^{164}\mathrm{Dy}start_FLOATSUPERSCRIPT 164 end_FLOATSUPERSCRIPT roman_Dy Bose-Einstein condensate phases and dynamics with dipolar anisotropy, Phys. Rev. Res. 4, 043124 (2022).
  • Halder et al. [2023] S. Halder, S. Das, and S. Majumder, Two-dimensional miscible-immiscible supersolid and droplet crystal state in a homonuclear dipolar bosonic mixture, Phys. Rev. A 107, 063303 (2023).
  • Smith et al. [2023] J. C. Smith, D. Baillie, and P. B. Blakie, Supersolidity and crystallization of a dipolar Bose gas in an infinite tube, Phys. Rev. A 107, 033301 (2023).
  • Ripley et al. [2023] B. T. E. Ripley, D. Baillie, and P. B. Blakie, Two-dimensional supersolidity in a planar dipolar Bose gas, Phys. Rev. A 108, 053321 (2023).
  • Mukherjee et al. [2024] K. Mukherjee, T. A. Cardinale, and S. M. Reimann, Selective Rotation and Attractive Persistent Currents in Anti-Dipolar Ring Supersolids (2024), arXiv:2402.19126 [cond-mat] .
  • Bland et al. [2022b] T. Bland, E. Poli, L. A. P. Ardila, L. Santos, F. Ferlaino, and R. N. Bisset, Alternating-domain supersolids in binary dipolar condensates, Phys. Rev. A 106, 053322 (2022b).
  • Halder et al. [2024] S. Halder, S. Das, and S. Majumder, Induced supersolidity in a Dy-Er mixture, Phys. Rev. A 109, 063321 (2024).
  • Gallemí et al. [2020] A. Gallemí, S. M. Roccuzzo, S. Stringari, and A. Recati, Quantized vortices in dipolar supersolid Bose-Einstein-condensed gases, Phys. Rev. A 102, 023322 (2020).
  • Roccuzzo et al. [2020] S. M. Roccuzzo, A. Gallemí, A. Recati, and S. Stringari, Rotating a Supersolid Dipolar Gas, Phys. Rev. Lett. 124, 045702 (2020).
  • Ancilotto et al. [2021] F. Ancilotto, M. Barranco, M. Pi, and L. Reatto, Vortex properties in the extended supersolid phase of dipolar Bose-Einstein condensates, Phys. Rev. A 103, 033314 (2021).
  • Bland et al. [2023] T. Bland, G. Lamporesi, M. J. Mark, and F. Ferlaino, Vortices in dipolar Bose–Einstein condensates, Comptes Rendus. Physique 24, 133 (2023).
  • Casotti et al. [2024] E. Casotti, E. Poli, L. Klaus, A. Litvinov, C. Ulm, C. Politi, M. J. Mark, T. Bland, and F. Ferlaino, Observation of vortices in a dipolar supersolid (2024), arXiv:2403.18510 [cond-mat, physics:quant-ph] .
  • Ghosh et al. [2024] H. S. Ghosh, S. Halder, S. Das, and S. Majumder, Induced supersolidity and hypersonic flow of a dipolar Bose-Einstein Condensate in a rotating bubble trap (2024), arXiv:2402.13422 [cond-mat] .
  • Madison et al. [2000] K. W. Madison, F. Chevy, W. Wohlleben, and J. Dalibard, Vortex Formation in a Stirred Bose-Einstein Condensate, Phys. Rev. Lett. 84, 806 (2000).
  • Abo-Shaeer et al. [2001] J. R. Abo-Shaeer, C. Raman, J. M. Vogels, and W. Ketterle, Observation of Vortex Lattices in Bose-Einstein Condensates, Science 292, 476 (2001).
  • Klaus et al. [2022] L. Klaus, T. Bland, E. Poli, C. Politi, G. Lamporesi, E. Casotti, R. N. Bisset, M. J. Mark, and F. Ferlaino, Observation of vortices and vortex stripes in a dipolar condensate, Nat. Phys. 18, 1453 (2022).
  • Sadler et al. [2006] L. E. Sadler, J. M. Higbie, S. R. Leslie, M. Vengalattore, and D. M. Stamper-Kurn, Spontaneous symmetry breaking in a quenched ferromagnetic spinor Bose–Einstein condensate, Nature 443, 312 (2006).
  • Weiler et al. [2008] C. N. Weiler, T. W. Neely, D. R. Scherer, A. S. Bradley, M. J. Davis, and B. P. Anderson, Spontaneous vortices in the formation of Bose–Einstein condensates, Nature 455, 948 (2008).
  • Lamporesi et al. [2013] G. Lamporesi, S. Donadello, S. Serafini, F. Dalfovo, and G. Ferrari, Spontaneous creation of Kibble–Zurek solitons in a Bose–Einstein condensate, Nature Phys 9, 656 (2013).
  • Donadello et al. [2016] S. Donadello, S. Serafini, T. Bienaimé, F. Dalfovo, G. Lamporesi, and G. Ferrari, Creation and counting of defects in a temperature-quenched Bose-Einstein condensate, Phys. Rev. A 94, 023628 (2016).
  • Goo et al. [2021] J. Goo, Y. Lim, and Y. Shin, Defect Saturation in a Rapidly Quenched Bose Gas, Phys. Rev. Lett. 127, 115701 (2021).
  • Matthews et al. [1999] M. R. Matthews, B. P. Anderson, P. C. Haljan, D. S. Hall, C. E. Wieman, and E. A. Cornell, Vortices in a Bose-Einstein Condensate, Phys. Rev. Lett. 83, 2498 (1999).
  • Denschlag et al. [2000] J. Denschlag, J. E. Simsarian, D. L. Feder, C. W. Clark, L. A. Collins, J. Cubizolles, L. Deng, E. W. Hagley, K. Helmerson, W. P. Reinhardt, S. L. Rolston, B. I. Schneider, and W. D. Phillips, Generating Solitons by Phase Engineering of a Bose-Einstein Condensate, Science 287, 97 (2000).
  • Brachmann et al. [2011] J. F. S. Brachmann, W. S. Bakr, J. Gillen, A. Peng, and M. Greiner, Inducing vortices in a Bose-Einstein condensate using holographically produced light beams, Opt. Express, OE 19, 12984 (2011).
  • Kumar et al. [2018] A. Kumar, R. Dubessy, T. Badr, C. De Rossi, M. de Goër de Herve, L. Longchambon, and H. Perrin, Producing superfluid circulation states using phase imprinting, Phys. Rev. A 97, 043615 (2018).
  • Allen et al. [1992] L. Allen, MW. Beijersbergen, RJC. Spreeuw, and JP. Woerdman, Orbital angular momentum of light and the transformation of Laguerre-Gaussian laser modes, Physical Review A 45, 8185 (1992).
  • Marzlin et al. [1997] K.-P. Marzlin, W. Zhang, and E. M. Wright, Vortex Coupler for Atomic Bose-Einstein Condensates, Phys. Rev. Lett. 79, 4728 (1997).
  • Andersen et al. [2006] MF. Andersen, C. Ryu, P. Cladé, V. Natarajan, A. Vaziri, K. Helmerson, and WD. Phillips, Quantized Rotation of Atoms from Photons with Orbital Angular Momentum, Physical Review Letters 97, 170406 (2006).
  • Beattie et al. [2013] S. Beattie, S. Moulder, R. J. Fletcher, and Z. Hadzibabic, Persistent Currents in Spinor Condensates, Phys. Rev. Lett. 110, 025301 (2013).
  • Mondal et al. [2014] P. K. Mondal, B. Deb, and S. Majumder, Angular momentum transfer in interaction of Laguerre-Gaussian beams with atoms and molecules, Phys. Rev. A 89, 063418 (2014).
  • Bhowmik et al. [2016] A. Bhowmik, P. K. Mondal, S. Majumder, and B. Deb, Interaction of atom with nonparaxial Laguerre-Gaussian beam: Forming superposition of vortex states in Bose-Einstein condensates, Phys. Rev. A 93, 063852 (2016).
  • Das et al. [2020] S. Das, A. Bhowmik, K. Mukherjee, and S. Majumder, Transfer of orbital angular momentum superposition from asymmetric Laguerre–Gaussian beam to Bose–Einstein Condensate, J. Phys. B: At. Mol. Opt. Phys. 53, 025302 (2020).
  • Ghosh Dastidar et al. [2022] M. Ghosh Dastidar, S. Das, K. Mukherjee, and S. Majumder, Pattern formation and evidence of quantum turbulence in binary Bose-Einstein condensates interacting with a pair of Laguerre-Gaussian laser beams, Physics Letters A 421, 127776 (2022).
  • Neely et al. [2010] T. W. Neely, E. C. Samson, A. S. Bradley, M. J. Davis, and B. P. Anderson, Observation of Vortex Dipoles in an Oblate Bose-Einstein Condensate, Physical Review Letters 104, 160401 (2010).
  • Haljan et al. [2001] P. C. Haljan, B. P. Anderson, I. Coddington, and E. A. Cornell, Use of Surface-Wave Spectroscopy to Characterize Tilt Modes of a Vortex in a Bose-Einstein Condensate, Phys. Rev. Lett. 86, 2922 (2001).
  • Anderson et al. [2000] B. P. Anderson, P. C. Haljan, C. E. Wieman, and E. A. Cornell, Vortex Precession in Bose-Einstein Condensates: Observations with Filled and Empty Cores, Phys. Rev. Lett. 85, 2857 (2000).
  • Engels et al. [2002] P. Engels, I. Coddington, P. C. Haljan, and E. A. Cornell, Nonequilibrium Effects of Anisotropic Compression Applied to Vortex Lattices in Bose-Einstein Condensates, Phys. Rev. Lett. 89, 100403 (2002).
  • Freilich et al. [2010] D. V. Freilich, D. M. Bianchi, A. M. Kaufman, T. K. Langin, and D. S. Hall, Real-Time Dynamics of Single Vortex Lines and Vortex Dipoles in a Bose-Einstein Condensate, Science 329, 1182 (2010).
  • Serafini et al. [2017] S. Serafini, L. Galantucci, E. Iseni, T. Bienaimé, R. N. Bisset, C. F. Barenghi, F. Dalfovo, G. Lamporesi, and G. Ferrari, Vortex Reconnections and Rebounds in Trapped Atomic Bose-Einstein Condensates, Phys. Rev. X 7, 021031 (2017).
  • Seo et al. [2017] S. W. Seo, B. Ko, J. H. Kim, and Y. Shin, Observation of vortex-antivortex pairing in decaying 2D turbulence of a superfluid gas, Sci Rep 7, 4587 (2017).
  • Wilson et al. [2015] K. E. Wilson, Z. L. Newman, J. D. Lowney, and B. P. Anderson, In situ imaging of vortices in Bose-Einstein condensates, Phys. Rev. A 91, 023621 (2015).
  • Mason and Berloff [2008] P. Mason and N. G. Berloff, Motion of quantum vortices on inhomogeneous backgrounds, Phys. Rev. A 77, 032107 (2008).
  • Simula [2018] T. Simula, Vortex mass in a superfluid, Phys. Rev. A 97, 023609 (2018).
  • Groszek et al. [2018] A. J. Groszek, D. M. Paganin, K. Helmerson, and T. P. Simula, Motion of vortices in inhomogeneous Bose-Einstein condensates, Phys. Rev. A 97, 023617 (2018).
  • Gaunt et al. [2013] A. L. Gaunt, T. F. Schmidutz, I. Gotlibovych, R. P. Smith, and Z. Hadzibabic, Bose-einstein condensation of atoms in a uniform potential, Phys. Rev. Lett. 110, 200406 (2013).
  • Pethick and Smith [2008] C. J. Pethick and H. Smith, Bose-Einstein Condensation in Dilute Gases, 2nd ed. (Cambridge University Press, Cambridge, 2008).
  • Pitaevskii and Stringari [2016] L. Pitaevskii and S. Stringari, Bose-Einstein Condensation and Superfluidity, International Series of Monographs on Physics (Oxford University Press, Oxford, 2016).
  • Lahaye et al. [2009] T. Lahaye, C. Menotti, L. Santos, M. Lewenstein, and T. Pfau, The physics of dipolar bosonic quantum gases, Rep. Prog. Phys. 72, 126401 (2009).
  • Baillie et al. [2016] D. Baillie, R. M. Wilson, R. N. Bisset, and P. B. Blakie, Self-bound dipolar droplet: A localized matter wave in free space, Phys. Rev. A 94, 021602 (2016).
  • Lee et al. [1957] T. D. Lee, K. Huang, and C. N. Yang, Eigenvalues and Eigenfunctions of a Bose System of Hard Spheres and Its Low-Temperature Properties, Phys. Rev. 106, 1135 (1957).
  • Lima and Pelster [2011] A. R. P. Lima and A. Pelster, Quantum fluctuations in dipolar Bose gases, Phys. Rev. A 84, 041604 (2011).
  • Petrov [2015] D. S. Petrov, Quantum Mechanical Stabilization of a Collapsing Bose-Bose Mixture, Phys. Rev. Lett. 115, 155302 (2015).
  • Schützhold et al. [2006] R. Schützhold, M. Uhlmann, Y. Xu, and U. R. Fischer, Mean-field expansion in bose–einstein condensates with finite-range interactions, Int. J. Mod. Phys. B 20, 3555 (2006).
  • Lima and Pelster [2012] A. R. P. Lima and A. Pelster, Beyond mean-field low-lying excitations of dipolar Bose gases, Phys. Rev. A 86, 063609 (2012).
  • Tang et al. [2018] Y. Tang, W. Kao, K.-Y. Li, and B. L. Lev, Tuning the dipole-dipole interaction in a quantum gas with a rotating magnetic field, Phys. Rev. Lett. 120, 230401 (2018).
  • Leanhardt et al. [2002] A. E. Leanhardt, A. Görlitz, A. P. Chikkatur, D. Kielpinski, Y. Shin, D. E. Pritchard, and W. Ketterle, Imprinting Vortices in a Bose-Einstein Condensate using Topological Phases, Phys. Rev. Lett. 89, 190403 (2002).
  • Bandyopadhyay et al. [2017] S. Bandyopadhyay, A. Roy, and D. Angom, Dynamics of phase separation in two-species Bose-Einstein condensates with vortices, Physical Review A 96, 043603 (2017).
  • Muruganandam and Adhikari [2009] P. Muruganandam and S. Adhikari, Fortran programs for the time-dependent Gross–Pitaevskii equation in a fully anisotropic trap, Computer Physics Communications 180, 1888 (2009).
  • Sasaki et al. [2010] K. Sasaki, N. Suzuki, and H. Saito, Bénard–von Kármán Vortex Street in a Bose-Einstein Condensate, Phys. Rev. Lett. 104, 150404 (2010).
  • Aioi et al. [2011] T. Aioi, T. Kadokura, T. Kishimoto, and H. Saito, Controlled Generation and Manipulation of Vortex Dipoles in a Bose-Einstein Condensate, Phys. Rev. X 1, 021003 (2011).
  • [94] See Supplemental Material.
  • Scarola et al. [2006] V. W. Scarola, E. Demler, and S. Das Sarma, Searching for a supersolid in cold-atom optical lattices, Phys. Rev. A 73, 051601 (2006).
  • Li et al. [2008] W. Li, M. Haque, and S. Komineas, Vortex dipole in a trapped two-dimensional Bose-Einstein condensate, Phys. Rev. A 77, 053610 (2008).
  • Middelkamp et al. [2011] S. Middelkamp, P. J. Torres, P. G. Kevrekidis, D. J. Frantzeskakis, R. Carretero-González, P. Schmelcher, D. V. Freilich, and D. S. Hall, Guiding-center dynamics of vortex dipoles in Bose-Einstein condensates, Phys. Rev. A 84, 011605 (2011).
  • Nakamura et al. [2016] S. Nakamura, K. Matsui, T. Matsui, and H. Fukuyama, Possible quantum liquid crystal phases of helium monolayers, Phys. Rev. B 94, 180501 (2016).
  • Nyéki et al. [2017] J. Nyéki, A. Phillis, A. Ho, D. Lee, P. Coleman, J. Parpia, B. Cowan, and J. Saunders, Intertwined superfluid and density wave order in two-dimensional 4He, Nature Phys 13, 455 (2017).
  • Choi et al. [2021] J. Choi, A. A. Zadorozhko, J. Choi, and E. Kim, Spatially Modulated Superfluid State in Two-Dimensional He 4 Films, Phys. Rev. Lett. 127, 135301 (2021).
  • Fulde and Ferrell [1964] P. Fulde and R. A. Ferrell, Superconductivity in a strong spin-exchange field, Phys. Rev. 135, A550 (1964).
  • Larkin and Ovchinnikov [1964] A. I. Larkin and Y. N. Ovchinnikov, Nonuniform state of superconductors, Zh. Eksp. Teor. Fiz. 47, 1136 (1964), [Sov.Phys.JETP 20,762 (1965)].