<<

Stationary states, dynamical stability, and vorticity of Bose-Einstein condensates in tilted rotating harmonic traps

Srivatsa B. Prasad,1, ∗ Brendan C. Mulkerin,2 and Andrew M. Martin1 1School of , University of Melbourne, Melbourne, 3010, Australia 2Centre for Quantum and Optical Science, Swinburne University of Technology, Melbourne, 3122, Australia (Dated: May 13, 2020) We theoretically investigate a Bose-Einstein condensate confined by a rotating harmonic trap whose rotation axis is not aligned with any of its principal axes. The principal axes of the Thomas- Fermi density profiles of the resulting stationary solutions are found to be tilted with respect to those of the rotating trap, representing an extra degree of freedom that is associated with the existence of additional branches of stationary solutions for any given rotation axis alignment. By linearizing the time-dependent theory about the stationary states, we obtain a semi-analytical prediction of their dynamical instability at high rotation frequencies against collective modes arising from environmen- tal perturbations. Comparing the stationary states to direct simulations of the Gross-Pitaevskii equation, we predict the nucleation of quantum vortices in the dynamically unstable rotational regime. These vortex lines are aligned along the rotation axis despite the tilting of the rotating trap although the background density profile is tilted with respect to the trapping and rotation axes.

I. INTRODUCTION in the condensate [11]. Alternate methods for produc- ing vortices in BECs also exist, such as stirring with a For more than 70 years the behaviour of rotating su- Gaussian beam [12], dragging a laser configuration perfluids have been of considerable experimental and the- through a trapped condensate (or, equivalently, a con- oretical interest to several generations of . Ini- densate through a laser configuration) [13, 14], applying tially, it was recognized by Onsager [1] and Feynman [2] oscillatory perturbations to the trapping [15, 16], con- that superfluid flow is characterized by a quantized cir- densing a rotating thermal (non-condensed) atomic va- culation, which was subsequently experimentally veri- por [17], and utilizing the Kibble-Zurek mechanism by fied by Hall and Vinen [3,4]. Since then, it has been quenching a thermal vapor across the BEC critical tem- recognized that the presence of a nonzero angular mo- perature [18]. The theoretical and analytical study of the mentum in a superfluid leads to complex, nontrivial be- resulting vortices have uncovered phenomena rich in vari- haviour, spurring discoveries such as the recent realiza- ety; some examples that are relevant to scalar, nondipo- tion of negative-temperature Onsager vortex clusters in lar, single-component condensates at zero temperature two-dimensional Bose gases [5,6]. In particular, Bose- includes Kelvin waves [19–21], Abrikosov vortex lattices Einstein condensates (BECs) offer a uniquely flexible and their Tkachenko modes [22–24], quantum Hall-like platform for the study of superfluid rotation and have physics [25–28], vortex reconnections [29, 30], quantum been the focus of intense study for several years [7–9]. analogs of classical fluid instabilities [14, 31, 32], and hys- Consider a superfluid in a rotating bucket. This su- teresis [33, 34]. perfluid cannot support rigid-body rotation if the bucket By contrast, previous studies relating to tilting effects is symmetric about the rotation axis, due to the absence in rotating BECs have mainly focussed on the collec- of shear forces, and so its angular momentum manifests tive modes of vortices in response to tilting perturba- in the form of quantum vortices above a certain critical tions of the trap [35–38], while the literature concerning rotation frequency [10]. However, when the symmetry of the steady rotation of the external confinement about a the bucket about the rotation axis is broken, it is able to non-principal axis is chiefly limited to the stability of the transfer angular momentum to the superfluid even in the centre-of-mass oscillations in the rotating frame [39, 40]. absence of shear forces, and the condensate exhibits solid- Given that such tilted rotating traps may be experimen- body rotation at slow rotation frequencies and quantized tally generated in a similar manner to the excitation of vortices above a critical rotation frequency. For Bose- the tilting modes, and that roughly analogous systems Einstein condensates, in which the ‘bucket’ is replaced such as dipolar BECs with tilted rotating dipole moments by an atomic trap generated by electromagnetic fields, have been realized experimentally [41, 42], a systematic arXiv:2004.02568v3 [cond-mat.quant-gas] 12 May 2020 angular momentum may be transferred from the trap to study of BECs confined by a tilted rotating trap is war- the condensate by modulating the applied fields such that ranted. In this paper, we analytically obtain stationary the condensate is confined by a rotating potential that is solutions of the Gross-Pitaevskii equation in the Thomas- asymmetrical about the rotation axis [7]. At sufficiently Fermi limit for a condensate subject to a range of differ- high rotation frequencies, this method induces vorticity ent tilting angles and harmonic trapping regimes. For all but the most trivial cases, the stationary solution den- sities are found to be tilted about the rotation axis by a different angle than the trap itself. One of the conse- ∗ [email protected] quences of this additional degree of freedom is the exis- 2 tence of two previously unknown branches of stationary (GPE) [7–9, 52]: solutions. These exist even when the trap is not tilted ∂ψ 1 2 2 away from the rotation axis, a result that is analogous to i ψ VT r, t ψ g˜ ψ ψ iΩ r ψ. (2) the tilted triaxial ellipsoids that are rotating-frame sta- ∂t 2 tionary solutions for self-gravitating irrotational classical Here we= − define∇ g ˜+ 4πNa( s) l⊥+asS theS effective+ ⋅ ( strength× ∇) of a fluids [43]. Focusing on the stationary solution branch mean-field, two-body interaction, with a corresponding s- existing in the nonrotating limit, we semi-analytically lin- wave scattering length= given~ by as, and denote the time- earize the condensate’s fluctuations in response to small dependent harmonic trapping potential by VT. perturbations and thus predict a dynamical instability at Previously, theoretical and experimental studies of the higher rotation frequencies, where the amplitude of one angular momentum of trapped BECs have tended to as- or more collective modes is expected to amplify expo- sume that the rotation axis of the confinement coincides nentially in time. In the regions of dynamical instabil- with one of its symmetry axes. The rotation of a har- ity we show via numerical Gross-Pitaevskii simulations monic trap about an arbitrary axis can effectively be that untilted vortices are nucleated from a tilted conden- modeled by fixing Ω Ωˆz, without loss of generality, sate, despite the background condensate density still be- and assuming that the trapping potential is not symmet- ing tilted. The theoretical formalism utilized represents a ric under the transformation= z z. In the co-rotating generalization of existing theoretical methods for study- reference frame, this potential can be specified by ing BECs in asymmetric, untilted rotating traps [44–49] → − and are readily conducive to experimental investigation 1 2 2 VT r 1 ε x cos θ z sin θ 1 ε y along the lines of previous studies that have probed the 2 untilted regime [11, 50, 51]. 1 2 ( ) = γ(2 x−sin)(θ z cos+θ , ) + ( + )  (3) This paper is structured as follows. SectionII defines 2 the concept of a tilted rotating trap and introduces the where ε + 1, 1 (and γ − R. This) external potential is relevant coordinate reference frames, while Sec.III dis- equivalent to cusses the methodology for solving for the vortex-free sta- ∈ (− ) ∈ tionary solutions in the Thomas-Fermi limit. In Sec. IV, 1 V R 1 ε X2 1 ε Y 2 γ2Z2 , (4) we examine the features of these stationary solutions for T 2 two distinct trapping regimes, and in Sec.V, the time- via a rotation( ) = of the( co-rotating− ) + ( coordinates+ ) + as given by dependent theory is linearized in order to characterize the dynamical stability of the vortex-free stationary so- X cos θ 0 sin θ x lutions during a quasi-adiabatic rampup of the trapping Y 0 1 0 y . (5) rotation frequency. Finally, Sec.VI contains a discussion ⎛Z ⎞ ⎛ sin θ 0 cos θ⎞ ⎛z⎞ of the outcomes of a series of numerical simulations of ⎜ ⎟ = ⎜ ⎟ ⎜ ⎟ such rampups, where the dynamical route to vortex nu- By inspection,⎝ Eq.⎠ (3⎝)− is equivalent to⎠ ⎝ Eq.⎠ (4) when, for cleation in a vorticity-free condensate in a tilted rotating integer n, the tilting angle obeys θ nπ. A similar equiv- trap is demonstrated. alence, albeit with modified values of γ and ε, holds when n takes on half-integer values. = To simulate the stationary state and dynamics of a II. THE TILTED, ROTATING HARMONIC BEC in this trap via numerical methods, it is sufficient TRAP to use Eqs. (2) and (3). However, the vorticity-free sta- tionary solutions of Eq. (3), and their linear response to environmental perturbations, are well-described in the In order to describe a dilute, scalar BEC, at zero tem- θ 0 limit by purely semi-analytical methods [44, 45]. perature in a rotating, tilted harmonic trap, we utilize To utilise these methods for an arbitrary value of θ it is the Gross-Pitaevskii equation for the condensate order necessary= to set up a hydrodynamic formalism. This in- parameter, ψ. We assume that N condensed bosons, volves the definition of the condensate’s density, n, phase, each with a mass m, are confined in the trap and that S, and superfluid velocity, v, via the relations [7]: the root mean squared harmonic trapping frequency in the x-y plane is given by ω⊥. This may be used to rescale ψ neiS, (6) t as t ω⊥t and r as r r l⊥, where l⊥ h mω⊥ is v √S. (7) the in-plane harmonic oscillator length. We» also rescale = 3 ̵ ψ as ψ→ l⊥ Nψ, such→ that~ it is normalized= as~( ) Substituting Eqs. (6) and= (∇7) into Eq. (2) yields a pair of » hydrodynamic equations given by [8,9]: → ~ 3 2 d r Ψ r, t 1. (1) ∂n n v Ω r , (8) ∂t S S ( )S = 2 Subsequently, in a reference frame rotating with respect ∂v = −∇ ⋅ [v2( − × )] n to the inertial laboratory frame with the angular velocity VT gn˜ v Ω r . (9) ∂t ⎧ 2 2 √n ⎫ Ω, ψ obeys the dimensionless Gross-Pitaevskii equation ⎪ ∇ ‰ Ž⎪ = −∇ ⎨ + + − ⋅ ( × ) − √ ⎬ ⎪ ⎪ ⎩⎪ ⎭⎪ 3

When Nas l⊥ the quantum pressure term in Eq. (9), Ω 2 n n , is negligible due to the minimal effects of zero-point≫ kinetic energy fluctuations in the conden- z √ √ z˜ yˆ sate∇ ∇ [7(, 53)~, 54]. In the Thomas-Fermi (TF) limit, where π X 2 θ this term may be neglected, Eq. (9) is approximated by Z the simplified form − x˜ ξ θ ξ ∂v v2 x V gn˜ v Ω r . (10) ∂t 2 T − = −∇ œ + + − ⋅ ( × )¡ Rz We also note that the vector Ω r lies in the x-y plane Rx whereas the principal axes of the trap are given by Xˆ, Yˆ , and Zˆ, withx ˆ and Xˆ not coinciding× with each other unless the trap is not tilted. The resulting competition between the trapping and rotating-frame transformation terms necessitates the introduction of a second angle, ξ, and a third co-rotating coordinate frame, ˜r, in order to FIG. 1. Shaded cross-section, at y = Y = y˜ = 0, of the el- find the axes of symmetry of the solutions of Eqs. (8) and lipsoidal surface of constant density for a Thomas-Fermi sta- (10). Let us define ξ and ˜r via the transformation tionary state with its semi-axes along thex ˜- andz ˜-axes, Rx and Rz, respectively, illustrated for reference. The Cartesian x˜ cos ξ 0 sin ξ X axes corresponding to the coordinate frames r, R, and ˜r are y˜ 0 1 0 Y overlaid on the cross-section, and Ω ∥ zˆ. ⎛z˜⎞ ⎛sin ξ 0− cos ξ ⎞ ⎛Z ⎞ ⎜ ⎟ = ⎜ cos θ ξ 0⎟ sin⎜ θ⎟ ξ x ⎝ ⎠ ⎝ ⎠ ⎝ ⎠ Let us impose the following Ans¨atze for nTF and vTF: 0 1 0 y . (11) ⎛ sin( θ− ξ) 0 cos(θ − ξ)⎞ ⎛z⎞ 2 2 r˜i r˜i = ⎜ ⎟ ⎜ ⎟ nTF ˜r n0 1 Θ 1 , (17) R2 R2 In this new reference⎝− ( frame,− ) the trapping( − )⎠ is⎝ given⎠ by ⎛ i∈x,y,z i ⎞ ⎛ i∈x,y,z i ⎞ ( ) = − − 1 vTF ˜r ⎝αxyx˜y˜Qαyzy˜z˜⎠ α⎝zxz˜x˜ Q. ⎠ (18) V ˜r 1 ε x˜ cos ξ z˜sin ξ 2 1 ε y˜2 T 2 Here, n(0 ) 15= ∇ [8πRxR+yRz is a+ normalization] parameter 1 2 2 ( ) = γ( x˜−sin)(ξ z˜cos+ξ , ) + ( + ) (12) that ensures that nTF obeys Eq. (1)[53, 54]. The form of 2 Eq. (17)= shows~( that the angle) ξ in the coordinate trans- while the rotating-frame+ ( − term, Ω) r transforms to formation given by Eq. (11) is fixed by the requirement that the principal axes of the TF stationary state den- Ω ˜r Ω cos θ ξ y˜x˜ˆ x˜y˜ˆ ×sin θ ξ y˜z˜ˆ z˜y˜ˆ . sity coincide with the Cartesian axes of the r coordinate (13) frame. The parameters Ri thus denote the semi-axes of To clarify× = the relationship( − ) ‰− between+ Ž + the( co-rotating− ) ‰ − refer-Ž the paraboloid TF profile along ther ˜i-axis. We illustrate ence frames, we overlay the coordinate axes of r, R, and these features in the TF{ density} cross-section in Fig.1 ˆ ˆ ˜r at constant y Y y˜ 0 on a typical cross-section of by labeling the ellipsoid’s semi-axes along x˜ and z˜ as Rx an TF surface of constant density in Fig.1. and Rz, respectively. = = = Equation (18) is consistent with the quadrupolar flow of a TF stationary state in an untilted harmonic trap III. THOMAS-FERMI STATIONARY (θ ξ 0) rotating about the z-axis, vTF α xy [44]. SOLUTIONS An inspection of Eq. (15) shows that the kth component of v=, =j≠k αjkr˜j, is nonzero only if ijkΩir˜=j ∇0,( which) in turn shows that α 0 only if  Ω 0. This suggests The stationary solutions of the GPE are specified ij ijk k that for the problem at hand, we have α≠ 0 since through the condensate’s chemical potential, µ, via [7] ∑ zx Ωy 0. By substituting≠ Eq. (17) and≠ (18) into Eq. (15) ψ ˜r, t ψ ˜r, t 0 exp iµt . (14) and equating the coefficients of the spatial coordinates,= we can= verify the property that αzx is null and also derive the relations Therefore, the stationary( ) = ( state= density,) (− nTF) , and velocity, vTF, obey κ2 κ2 α α x y Ω cos θ ξ , (19) xy 2 2 0 n v Ω ˜r , (15) κx κy − 2 ≡ = Œκ2 1 ‘ ( − ) vTF y µ =V∇ ⋅ [ gn˜( − v× )] Ω ˜r . (16) δ αzx + Ω sin θ ξ , (20) 2 T TF TF κ2 1 y − ∇ = ∇ œ + + − ⋅ ( × )¡ ≡ = Œ ‘ ( − ) + 4 where κx Rx Rz and κy Ry Rz. Thus the trial solu- a)˜ωx 0 andω ˜y, ω˜z 0, tion employed for the velocity field is = ~ = ~ b)˜ωy → 0 andω ˜x, ω˜z ≠ 0, vTF ˜r α x˜y˜ δ y˜z˜ . (21) c)˜ωx,→ω˜y 0 andω ˜z ≠ 0, This quadrupolar( profile) = ∇( for) the+ ∇ velocity( ) field, and d)˜ω , ω˜ 0 andω ˜ 0. thereby the spatial dependence of the condensate’s phase, y z → x ≠ may be considered as the quantum analog of the classical For the remainder of this→ paper, the≠ subscripts xc, yc, velocity potential for an inviscid fluid inside an ellipsoid xyc, and yzc are used to denote the values of quantities container rotating about a non-principal axis of the el- such as ξ in the limits a), b), c) and d), respectively. A lipsoid [55, 56]. In both systems, solid-body rotation is detailed description of the self-consistency relations sat- possible only when the density is asymmetric about the isfied by Ω, α, δ and ξ at each of these limits is provided rotation axis. We also note that Eqs. (19) and (20) are in AppendixB. We also provide a description of how the formally similar to the equations of motion appearing in shape of the TF distribution can be understood via in- the context of the rotational energy bands in the tilted- spection of the signs of α, δ and θ ξ can be found in axis cranked shell model of rotating triaxial nuclei [57]. AppendixA. The problem of determining the stationary solutions − of Eqs. (15) and (16) may now be reduced to solving a set of five self-consistency relations for κx, κy, α, δ, ξ . IV. STATIONARY SOLUTION BRANCHES These are obtained by substituting Eqs. (17) and (21) into Eq. (16) and subsequently reading off{ the coefficients} of like terms. Firstly, from the coefficients ofx ˜2,y ˜2 and Keeping in mind the possible limits of the stationary z˜2, we find that the TF semi-axes are given by solution branches, Ω Ωxc, Ωyc, Ωxyc, Ωyzc , we pro- ceed to solve Eqs. (26)–(29) and plot the resulting val- 2˜gn ues of α, δ, and ξ as functions→ { of Ω for fixed values} of θ, γ R2 0 . (22) i ω˜2 and ε. To provide a representative sample of the variety i of trapping regimes, we analyze the following cases: = In Eq. (22), we make use of generalized harmonic trap- 2 1. γ 3 4, ε 0, θ 0, π 8, π 4, 3π 8 , ping frequencies,ω ˜i , that are defined as

2 2 2 2 2 2. γ = 4~3, ε = 0.05,∈θ{ 0~, π 8~, π 4,~3π} 8 . ω˜x 1 ε cos ξ γ sin ξ α 2Ωα cos θ ξ , (23) ω˜2 1 ε α2 δ2 2Ω α cos θ ξ δ sin θ ξ , When Ω= ~0, stationary= ∈ states{ ~ in the~ harmonic~ } trap de- y = ( − ) + + − ( − ) (24) scribed by case 1 are prolate and are axially symmetric aboutz ˆ, while= those for case 2 are oblate and do not ex- 2 = 2+ +2 + + [ 2 ( 2− ) − ( − )] ω˜z γ cos ξ 1 ε sin ξ δ 2Ωδ sin θ ξ . (25) hibit this axial symmetry. We do not analyze the trap tilting angle θ π 2 as this limit is easily transformed to This= implies that+ ( the− quantities) + κx+ and κy (obey− ) an untilted trap with a different set of trapping frequen-

2 cies by rotating= the~ coordinate frame abouty ˆ by π 2. 2 ω˜z κi 2 i x, y. (26) ω˜i ~ = ∶ = A. Prolate, Symmetric Trapping By recognizing that there is nox ˜z˜ term in Eq. (17), we also obtain the condition that We initially focus on the prolate, symmetric trap, 1 ε γ2 sin ξ cos ξ αδ Ω α sin θ ξ δ cos θ ξ 0. where γ 3 4 and ε 0. For this trap, we specify the ro- (27) tational frequencies as defined by cases a) – d) in Sec.III The( − two− ) final self-consistency+ + [ relations( − )− are obtained( − )] = via in TableI=: ~ = substituting Eq. (22) into Eqs. (19) and (20), which yields = ~ = 2 2 TABLE I. Endpoints of the branches - γ 3 4, ε 0 α Ω cos θ ξ ω˜x α Ω cos θ ξ ω˜y 0, (28) Ωxc~ω⊥ Ωyc~ω⊥ Ωxyc~ω⊥ Ωyzc~ω⊥ δ Ω sin θ ξ ω˜2 δ Ω sin θ ξ ω˜2 0. (29) [ + ( − )] y + [ − ( − )] z = θ = 0 1 1 1.75 1.75 θ = π~8 0.9475 1 1.8958 1.6495 Equations[ + ( 26−)–()] 29)+ describe[ − branches( − )] of= stationary θ = π~4 0.8485 1 1.9899 1.6578 solutions as functions of Ω that terminate when one or θ = 3π~8 0.7752 1 1.9833 1.7563 more ofω ˜x, ω˜y, ω˜z equal zero. The locations of these end- points determine the number of real stationary solutions for a given value of Ω. We identify four such limits which In Fig.2(a), we plot α as a function of Ω for the val- are of use to us, noting that a rotation of the rotating ues of θ listed in TableI. Here we see that there exist frame by π 2 about they ˜-axis transformsx ˜ toz ˜: five distinct stationary solution branches, four of which

~ 5 exhibit the endpoints defined by cases a) – d). Ini- is the only solution that exists for Ω max Ωxyc, Ωyzc . tially we describe the limit θ 0, as explored in pre- vious theoretical studies. For a trap with axial symme- > { } try about the rotation axis, i.e.= ε 0, an α 0 sta- tionary solution exists for all Ω 0, while two further solutions emerge at the rotational bifurcation= frequency= Ω Ωb1 ω⊥ 2 [44, 48, 49].≥ We note that the po- sition of this bifurcation√ is attributable to an energetic instability= ≡ of the~ l 2, m 2 quadrupolar surface mode, which has a frequency ω l 2, m 2 2ω⊥ 2Ω and is thus energetically favourable= = for Ω Ωb1√[44]. These ad- ditional stationary solutions( = are symmetric= ) = about− the Ω axis and terminate in the limit Ω Ω≥xc Ωyc ω⊥, where α ω⊥ as well. Furthermore, we find evidence for the ex- istence of a second bifurcation where→ two= more= stationary solutions→ emerge from the stationary solution defined by α 0 and terminate when Ω Ωxyc Ωyzc 1.75ω⊥. We attribute the existence of this bifurcation to the energetic instability= of the l 2, m →1 quadrupole= = mode, which 2 boasts the frequency ω l 2, m 1 ω⊥ 1 γ Ω[7] and is therefore associated= with= the bifurcation» frequency Ωb2 5ω⊥ 4 when γ 3( 4.= These= new) = branches+ are− not symmetric about the Ω axis unlike those emerging from the m= 1~ bifurcation,= and~ their existence had not pre- viously been predicted in the context of rotating BECs due to= the omission of the additional degrees of freedom given by δ and ξ. However, we note that the energetic instability of an l 2, m 1 surface mode causes similar bifurcations in other systems. For instance, in a rotating reference frame, the= equilibrium= density of a irrotational gravitationally-bound fluid can undergo just such a bi- furcation from a Maclaurin spheroid to a tilted Riemann ellipsoid [43]. The new class of stationary solutions described by θ 0 behaves markedly differently to those for the untilted trap. When Ω 0 we have a solution defined by α ≠ 0 and this solution, which we denote as Branch I, persists for Ω min Ωxc=, Ωyc . From TableI, this rotation fre-= quency is given by Ω Ωxc for all of the values of θ that we consider< { in this case.} Branch I is, in general, the solution that the condensate= will follow in response to a quasi-adiabatic acceleration of the trap’s rotation fre- quency from zero. Two additional, connected, branches emerge at a bifurcation frequency, denoted as Ωb1, and initially have values of α with opposite sign to the first solution. One of these solutions, denoted here as Branch II, terminates at Ω max Ωxc, Ωyc ω⊥. The other so- lution, denoted here as Branch III, persists until the end- point defined by Ω =min Ω{xyc, Ωyzc} ,≡ which is equivalent to Ω Ω for this trap. The behavior of Branch III con- yzc − trasts with that of= the solutions{ for θ} 0, where it is pos- FIG. 2. Stationary solutions as a function of Ω for α (a), θ ξ (b), and δ (c), when γ = 3~4, ε = 0, θ ∈ {0, π~8, π~4, 3π~8}. sible= for a condensate to follow the same solution branch from Ω Ωb1 till Ω . A second bifurcation= frequency, Ωb2, heralds the emergence of an additional pair of con- We also present the corresponding solutions of θ ξ, nected= branches that→ ∞ exhibit the same sign of α. One as a function of Ω, in Fig.2(b) where we observe that of these, denoted here as Branch IV, terminates when both of the bifurcations are clearly evident in the− be- Ω max Ωxyc, Ωyzc Ωxyc, while the other solution, havior of θ ξ as well as that of α. Furthermore, for denoted here as Branch V, exists for Ω Ωb2, and θ 0, the solutions that emerge at Ω Ωb2 and termi- → { } ≡ − ∈ [ +∞) = = 6 nate at Ω Ωxyc Ωyzc 1.75ω⊥ are closely related to Fig.3(a), where we have plotted α as a function of Ω for each other; they correspond to density profiles with the the angles θ 0, π 8, π 4, 3π 8 . identical TF= semi-axes= but= with opposite tilting angles about the rotation axis. As such, their respective values ∈ { ~ ~ ~ } of ξ are symmetric about the value ξ π 4. In Fig.2(c), where δ is plotted as a function of Ω, we find that for the θ 0 branches emerging when Ω Ω=b2−,~ the values of α for one branch are equivalent to those of δ for the other branch.= We also note that unlike= the corresponding be- havior of α and ξ, a qualitative discrepancy− in δ along Branch I for θ π 8 is evident when compared to the angles θ π 4 and θ 3π 8. Specifically, δ is a monoton- ically increasing= function~ of Ω when θ π 4 and θ 3π 8 but exhibits= ~ a maximum= ~ at Ω 0.78ω⊥ when θ π 8. However, such qualitative differences with= ~ respect= to the~ trap tilting angle are not exhibited≈ by Branches II= - V.~

B. Oblate, Asymmetric Trapping

We proceed to discuss the condensate’s behavior in the oblate, asymmetric trap where γ 4 3 and ε 0.05. Here, the lack of axial symmetry of the trapping along any axis results in the features of= the~ stationary= solu- tions being qualitatively different to those described in Sec.IVA. In TableII, we specify the rotation frequen- cies that correspond to the termination cases a) – d):

TABLE II. Endpoints of the branches - case 2

Ωxc~ω⊥ Ωyc~ω⊥ Ωxyc~ω⊥ Ωyzc~ω⊥ θ = 0 0.9747 1.0247 2.3334 2.3334 θ = π~8 1.0097 1.0247 2.1311 2.4914 θ = π~4 1.1128 1.0247 1.9924 2.5176 θ = 3π~8 1.2556 1.0247 2.0012 2.3892

When the rotating trap is untilted, i.e. θ 0, the sta- tionary solutions corresponding to Branches I, II and III are also untilted, i.e. θ ξ 0. We find that= Branch I, for which α 0, terminates when Ω Ωxc ω⊥ 1 ε. Branches II and III, which= = both exhibit α 0, are√ con- nected at the≥ bifurcation frequency Ω= Ωb1 =but are dis-− connected from Branch I. While Branch II< terminates when Ω Ωyc ω⊥ 1 ε, Branch III is= characterized by α monotonically tending√ to zero as Ω [44, 48, 49]. The extra= degrees= of freedom+ that are represented by δ and ξ manifest themselves when θ 0 through→ ∞ the pres- ence of the additional, previously unknown, branches IV and V, which are connected at Ω = Ωb2 and terminate at the same rotation frequency, Ω Ωxyc Ωyzc. How- ever, when the rotating trap is tilted= the stationary solu- FIG. 3. Stationary solutions as a function of Ω for α (a), θ −ξ tions more closely resemble those in= Sec.IVA= except that (b), and δ (c), when γ = 4~3, ε = 0.05, θ ∈ {0, π~8, π~4, 3π~8}. Ωxc Ωyc ω⊥ 1 ε when θ 0, π 8 and Ωxc ω⊥ 1 ε when θ π 4, 3√π 8; the crossover, where Ωxc Ωyc√, oc- We have also plotted θ ξ as a function of Ω in curs< when=θ 0.4693+ 26.89○=. This~ results in> Branches+ Fig.3(b) for this oblate, axially asymmetric trap, and I – III possessing∈ ~ ~ the opposite signs for Ωxc =Ωyc when thereby find that θ ξ similarly− behaves differently for θ π 4, 3π 8≈ to the solutions≡ when Ωxc Ωyc, a feature Branch I when θ π 8 as compared to the angles θ π 4 not seen in Sec.IVA. This behavior is demonstrated> in and θ 3π 8. For− instance, the behavior of Branch I ∈ ~ ~ < = ~ = ~ = ~ 7 for θ π 8 is not monotonic but has a maximum at time evolution of the fluctuations that is given by [45, 48] Ω 0.65ω⊥. This contrasts sharply with the monotonic behavior= of~ θ ξ as a function of Ω for θ π 4 and ∂ δS δS , (32) θ ≈3π 8. However, Branches II - V exhibit merely quan- ∂t δn δn titative differences− with respect to the tilting= angle.~ As ‹ vc= M ‹ g˜ in= Sec.~IVA, the values of ξ for the branches that emerge , (33) nTF vc when Ω Ωb2 are symmetric about the value ξ π 4, ⋅ ∇ M =v−c‹ STF Ω ˜r.  (34) suggesting that the density profiles for these two branches ∇ ⋅ ( ∇) ⋅ ∇ are physically= equivalent with the same TF semi-axes= − but~ Hence we can express each collective mode, indexed by ν, exhibit opposite tilting angles about the rotation axis. = ∇ − × as a combination of a density fluctuation, δn ˜r eλν t, and Thus the values of α for one branch is equivalent to those ν a phase fluctuation, δS ˜r eλν t, that satisfies Eq. (32) if of δ for the other branch, which may be inferred from ν the constant λ is an eigenvalue of the operator( ) . the corresponding plots of δ, as a function of Ω, that ν Since the time-dependence( ) of the collective modes is are− provided in Fig.3(c). Interestingly, the maximum exponential it is evident that, to linear order,M the dy- of θ ξ for θ π 8 along Branch I when Ω 0.65ω is ⊥ namical stability of a stationary state is determined by reflected in a similar maximum in δ, which eventually the set of all eigenvalues of , λ . If a given eigen- attains− negative= values~ as Ω Ω . ≈ ν xc value has a positive real component, the amplitude of the corresponding collective modeM will{ grow} exponentially in → time and will overwhelm the stationary state, rendering the stationary state dynamically unstable. Conversely we V. LINEARIZED TIME-DEPENDENT have dynamical stability only if all of the eigenvalues of HYDRODYNAMICS have a negative real component, while purely imagi- nary eigenvalues are characteristic of excitations with an infiniteM lifetime. To diagonalise , we expand δn and Via the hydrodynamic formalism elucidated in Secs.III δS as polynomials in 3 [45, 48]. Since it is not possible andIV, we have shown in Sec.IV that the tilting of a R to consider all possible collective modes,M we truncate the rotating harmonic trap induces a nontrivial tilting an- polynomial) expansion of the fluctuations such that the gle of the condensate’s TF stationary state density. The maximum allowed order of the polynomials is N 10. hydrodynamic formalism may also be used to determine max This proves to be a sufficiently high order to explore the the parametric domain of dynamical stability against en- dynamical stability of the stationary states in the= lin- vironmental perturbations, a procedure that has been earized regime. However, we note that even if no unstable achieved in the θ 0 limit [45]. Let us specifically address modes are found from this procedure for a given station- the scenario where the rotation frequency, Ω, is quasi- ary state, it is not a guarantee of dynamical stability as a adiabatically accelerated= from zero for a fixed choice of higher value of N may admit a collective mode whose ε, γ, and θ. In the TF limit, the condensate will follow max eigenvalue has a positive real component. Furthermore, the stationary solution Branch I and therefore we solely by limiting our analysis to the linearized regime, we ne- investigate the dynamical stability of Branch I. glect nonlinear effects that could destabilize modes that In general, the perturbating of a trapped BEC in a are stable at linear order in the fluctuations. stationary state can excite one or more of its collective We now proceed to describe the eigenvalues of the col- modes. For perturbations of sufficiently small magni- lective modes for Branch I of the stationary solutions tude the condensate’s response may be assumed to be presented in Sec.IV. From an inspection of , ev- linear and the collective excitations may be obtained by ery possible collective mode features the same maximum linearizing Eqs. (8) and (10) about the TF stationary polynomial order for both δn ˜r and δS ˜r , exceptM for state. In this formalism the collective modes are ex- a spatially uniform phase fluctuation without a corre- pressed as time-dependent fluctuations of the density and sponding density fluctuation( that) is associated( ) with a phase that are equivalent to linear combinations of the so- null eigenvalue. This is a manifestation of the Goldstone lutions of the Bogoliubov–de Gennes equations [7, 58, 59]. mode and is a consequence of the broken U 1 symme- To determine the spectrum of collective modes, we write: try that characterizes Bose-Einstein condensation [7, 58]. Fixing Nmax 10, we diagonalize Eq. (33()) over the discretely binned parameter space specified by the do- n ˜r, t nTF ˜r δn ˜r, t , (30) main of Branch= I of the stationary solutions described in S ˜r, t STF ˜r, t δS ˜r, t . (31) ○ ○ ( ) = ( ) + ( ) Sec.IVA( ε 0, γ 3 4 and θ 0 , 90 ). In Fig.4, we have shaded the bins where the respective Branch I so- ( ) = ( ) + ( ) Here, STF ˜r, t µt αx˜y˜ δy˜z˜. The subsequent lin- lutions are associated= = ~ with at least∈ [ one eigenvalue] of earization of Eqs. (8) and (10) is equivalent to neglecting with a real positive component. To linear order, these contributions( from) = − terms+ that+ are quadratic in the den- points in parameter space comprise a domain of guaran-M sity and phase fluctuations, δn and δS, respectively. This teed dynamical instability. A similar diagonalization of results in a coupled set of first-order equations for the with respect to the stationary solutions in Sec.IVB,

M 8 i.e. ε 0.05, γ 4 3 and θ 0○, 90○ , yields the stability trap tilt angles, which we attribute to the effective el- diagram depicted in Fig.5. lipticity of the trapping in the upright co-rotating frame = = ~ ∈ [ ] becoming larger as θ π 2. Due to the high order of polynomial perturbations that is required to realize un- stable collective modes→ in~ the limit Ω min Ωxc, Ωyc , Figs.4 and5 erroneously predict a region of dynamical stability. This limit is represented in Figs.→ 4 and{ 5 by the} red lines that plot min Ωxc, Ωyc as a function of θ; for a sufficiently large value of Nmax, these red lines would be the boundary of the{ domain of} dynamical instability.

VI. GROSS-PITAEVSKII EQUATION SIMULATIONS

In the preceding two sections we have found that the rotation of a tilted harmonic trap induces a nontrivial tilted angle of the condensate’s density profile and that the stationary solutions become dynamically unstable as Ω min Ωxc, Ωyc . However, the Thomas-Fermi ap- proximation does not provide information about the be- FIG. 4. Phase diagram of the dynamical stability of Branch havior→ of{ the condensate} after the dynamical instability I for ε = 0, γ = 3~4, with Nmax = 10; Branch I is dynamically unstable at the shaded points of parameter space. The red has manifested itself, nor does it predict whether or not dashed lines and markers denote the trajectories of the GPE the narrow regions of instability that extend to lower simulations and the corresponding instability frequency, re- rotation frequencies in Figs.4 and5 are negligible dur- spectively. The red unbroken curve denotes the endpoints of ing a quasi-adiabatic rampup of Ω. In order to attempt Branch I, Ω = min{Ωxc, Ωyc}. to answer these questions, we have also directly explored this system via numerically solving the GPE and thereby simulating a quasi-adiabatic rampup of a harmonic trap’s rotation frequency from zero. In this section, we employ the same set of trapping parameters that were specified in the discussion of the TF stationary states and their dynamical stability, i.e. θ π 8, π 4, 3π 8 with either γ 3 4, ε 0 or γ 4 3, ε 0.05 , and discuss the results of the GPE simulations.∈ { ~ ~ ~ } { Our= ~ procedure= } for{ solving= ~ Eq.= (2) in} the upright, co- rotating coordinate frame (denoted by r) is as follows. We set the rescaled two-body interaction strength as g˜ 104 and specify a 200 200 200 spatial grid with the intervals ∆x ∆y ∆z 0.25l⊥; these parameters are= sufficient for the ground× state× at Ω 0 to be well- described by the= TF stationary= = solution. Initially, the backward Euler method is utilized to simulate= Eq. (2) in imaginary time, with a suitable trial state as the ini- tial condition, and the converged solution is taken as the ground state solution at zero rotation [60]. Before propagating this resulting solution in real time, the local FIG. 5. Phase diagram of the dynamical stability of Branch I value of the condensate density is randomly perturbed by for ε = 0.05, γ = 4~3, with Nmax = 10; Branch I is dynamically unstable at the shaded points of parameter space. The red up to 5% of the original value in order to represent the dashed lines and markers denote the trajectories of the GPE environmental noise or experimental imperfections that simulations and the corresponding instability frequency, re- would seed any potentially unstable collective modes. spectively. The red unbroken curve denotes the endpoints of This perturbed state is used as the initial condition for Branch I, Ω = min{Ωxc, Ωyc}. the real-time evolution of the GPE, which is achieved us- ing the Alternate Direction Implicit-Time Splitting pseu- From Figs.4 and5, we can see that Branch I is stable doSPectral (ADI-TSSP) Strang scheme [61]. We employ −1 for small rotation frequencies and becomes dynamically a timestep of ∆t 0.004ω⊥ and an angular acceleration ∆Ω 2 unstable as Ω ω⊥. In both cases, we find that the ∆t 0.0005ω⊥, where the resulting increase in Ω at each first rotation frequency of instability is lower for larger timestep, ∆Ω 2=10−6, is sufficiently small that the con- → = = × 9 dition of adiabaticity holds. Therefore, the condensate is expected to smoothly follow Branch I of the TF station- ary solutions during the rampup procedure. During the real-time evolution of the GPE, we extract the observ- ables Rx, Ry, Rz, and ξ by fitting the density at x z 0 2 2 to the 1D TF density profile, n y n0 1 y Ry , and similarly the density at y 0 to the 2D TF density= pro-= 2 2 2 2 2 file, n x, z n0 1 x cos θ (ξ )R=x (sin− θ ~ ξ )Rz 2 2 2 2 2 2 2 z sin θ ξ Rx cos θ ξ= Rz 1 Rx 1 Rz sin 2 θ ξ xz(. Note) = that{ the− form[ of( these− )~ density+ cross-sections( − )~ ] − can[ be( found− )~ by+ applying( − the)~ transformation]−( ~ − ~ in) Eq.[ ((11−) to)] Eq.} (17). Subsequently, we may determine α and δ via Eqs. (19) and (20).

A. Prolate, Symmetric Trapping

In Fig.6 we compare α and ξ as obtained from the FIG. 6. Comparison of the TF stationary solutions along GPE simulations of a quasi-adiabatic rampup of Ω, when Branch I to GPE simulations, for ε = 0 and γ = 3~4, during a ε 0 and γ 3 4, to the TF results in Fig.2(a) and (b). quasi-adiabatic rampup of Ω. The first column (a, c, e) plots Here, the first (a, c, e) and second (b, d, f) columns α as a function of Ω and the column (b, d, f) plots θ − ξ as correspond= = to ~α and θ ξ, respectively, as functions of a function of Ω, where θ is equal to π~8 (a, b), π~4 (c, d), or ~ Ω while the rows correspond to distinct tilting angles: 3π 8 (e, f). θ π 8 in the first row− (a, b), θ π 4 in the second row (c, d), and θ 3π 8 in the third row (e, f). Figure 6 demonstrates= ~ that the condensate= initially~ follows the θ 3π 8 simulation when Ω 0.72ω⊥, they are very close TF stationary state= closely~ during the quasi-adiabatic ac- to the continuous domain of dynamical instability and celeration of the rotation frequency, which confirms the thus= their~ effect is relatively≈ minimal. Furthermore, the prediction in Fig.4 that the TF stationary states are dy- GPE simulations also capture nonlinear effects that are namically stable for low rotation frequencies. However, ignored in the the linearized hydrodynamic formalism. as Ω ω⊥, each of the trajectories from the numeri- The deviation of the simulations from the respective cal simulations diverge dramatically from the TF-based TF stationary states at higher rotation frequencies can predictions.→ This indicates the onset of a dynamical in- be better understood by examining the cross-sections of stability, as predicted in Fig.4, where the condensate the condensate densities at y 0 and z 0, which provide has been forced away from the TF stationary state due us information about the density profiles; tilting angles, to the uncontrolled growth of collective modes. Similar ξ, and the semi-axes, Rx,R= y,Rz ,= during the accel- behavior is seen in the analogous comparison of the TF- eration of the rotation frequency. These cross-sections and GPE-derived values of δ, which we have included in are presented in Fig.7 {for the trap} tilt θ π 4 when AppendixC for the reader’s reference. Ω equals 0.25ω⊥ (first row) and 0.575ω⊥ (second row). The rotation frequencies at which the condensate den- When Ω 0.575ω⊥, we halt the rampup of= Ω and~ then sities in each of the three simulations diverge from the evolve the GPE at constant rotation frequency for a du- −1 corresponding TF stationary state densities are depicted ration of 500= ω⊥ ; the cross-sections at the end of this pro- as red circular markers in Fig.4. When θ π 8 or 3π 8, cedure are given in the third row of Fig.7. We include the onset of dynamical instability agrees well with the the results for the analogous procedures performed with predictions of the linearized hydrodynamical= ~ formalism.~ the same trapping geometry but with θ π 8, 3π 8 However, when θ π 4 the critical rotation frequency is in AppendixC. In order to aid the reader’s visualiza- approximately 0.55ω⊥, whereas Fig.4 predicts that the tion of how the density’s principal axes do= not{ generally~ ~ } stationary solution= is~ always unstable when Ω 0.7ω⊥. coincide with those of either the trapping frame or the We attribute this discrepancy to the existence of the rotation axis, the X-Z Cartesian axes, i.e. the princi- small fringes of dynamical instability that intersect≳ the pal axes of the trapping, are overlaid in white upon the trajectory of the θ π 4 simulation at Ω 0.50ω⊥ and cross-sections at y 0. 0.51ω⊥, which are sufficient to destabilize the station- In Fig.7 we can see that the density profile is smooth ary solution. The fringes= ~ at Ω 0.57ω⊥ and≈ Ω 0.44ω⊥ when the condensate= is dynamically stable against the that are crossed by the trajectories of the simulations for initially seeded perturbation and, as predicted by the TF θ π 8 and θ 3π 8, respectively,≈ seem to be too≈ narrow theory, its symmetry axes in the x z plane are slightly to sufficiently destabilize the stationary states. While an- tilted away from those of the trap. However, when the other= ~ such fringe= is~ crossed by the simulation for θ π 8 condensate initially enters the regime− of dynamic insta- when Ω 0.51ω⊥, and a set of fringes is crossed by the bility, the density develops surface ripples and a sur- = ~ ≈ 10

48].

B. Oblate, Asymmetric Trapping

We now describe the results of the analogous GPE simulations for a trap with the parameters γ 4 3 and ε 0.05. In Fig.8 we compare α and θ ξ from these GPE simulations to the TF results in Fig.3(a)= and~ (b). Here,= the first (a, c, e) and second (b, d, f)− columns corre- spond to α and θ ξ, respectively, as functions of Ω while the rows correspond to distinct tilting angles: θ π 8 in the first row (a, b),− θ π 4 in the second row (c, d), and θ 3π 8 in the third row (e, f). Just as in the= simula-~ tions described in Sec.=VIA~ , the condensate is seen to be unstable= ~ at higher rotation frequencies against collective modes seeded by the random perturbation at t 0. This agrees with the behavior seen in a comparison of the semi- analytically and numerically obtained values of=δ, which we have included in AppendixC. A comparison may also be made with the prediction of dynamical instability in Fig.5, where we have indicated the rotation frequencies at which the GPE states diverge considerably from the TF states via red circular markers. When θ π 8, π 4 , these rotation frequencies are greater than the respective threshold frequencies above which the stationary∈ { ~ states~ } are always dynamically unstable. However, for θ 3π 8, the rotation frequency where the GPE solution diverges FIG. 7. Cross-sections of the condensate density in the co- wildly from the TF prediction occurs at Ω 0=.69ω~⊥, rotating x-y (first column) and x-z (second column) planes for which is considerably lower than the prediction of Fig.5 = = ~ = ~ ε 0, γ 3 4, and θ π 4, during a quasi-adiabatic rampup that the stationary state is dynamically unstable≈ when of Ω at Ω = 0.25ω⊥ (first row) and Ω = 0.575ω⊥ (second row), −1 Ω 0.80ω⊥. This may be attributed to the fact that the and after 500ω⊥ at constant Ω = 0.575ω⊥ (third row). The white lines represent the co-rotating X-Z axes. trajectory of the quasi-adiabatic rampup crosses a fringe of≳ dynamical instability when Ω 0.63ω⊥. We note that a similar fringe is crossed when θ π 4 and Ω 0.65ω⊥, but this fringe is narrower than the≈ one that destabilizes rounding cloud as some of the are ejected from the θ 3π 8 stationary state. While= the~ quasi-adiabatic≈ the centre of the condensate, as seen in the second col- trajectory for θ π 8 crosses several narrow fringes when umn of Fig.7. Moreover, we see that after evolution Ω 0=.75ω~⊥, 0.78ω⊥ , their effect is relatively minimal as −1 over a period of 500ω⊥ at constant rotation frequency, they are closely= followed~ by the threshold for dynamical Ω 0.575ω⊥, the condensate does not resemble a smooth instability∈ ( at Ω 0.)8ω⊥. TF distribution but has been subject to quantum vortex We can also visualize the GPE solutions for θ π 4 by nucleation= after further atoms have been ejected from plotting the cross-sections≈ of the density in the x-y and the centre of the condensate. This behaviour is a well- x-z planes for θ π 4 in Fig.9, with the corresponding= ~ known phenomenon that occurs in the rotation of an up- plots for θ π 8, 3π 8 included for reference in Ap- right, anisotropic harmonic trap containing a BEC [45– pendixC. Just as= in~ Sec.VIA, the density cross-sections 50, 62, 63] and thus it is not surprising that it occurs in are smooth= and{ ~ ellipsoidal~ } at low rotation frequencies this system. Crucially, an inspection of Fig.7(f) shows during a quasi-adiabatic rampup of Ω. This is evident in that the vortex lines coincident upon the x z plane are the first row of Fig.9, where Ω 0.4ω⊥, which also shows almost completely aligned along the rotation axis. This is that the condensate density’s principal axes are slightly in contrast to the background condensate density− whose tilted away from those of the= trapping. Similarly, we symmetry axes are tilted with respect to both x,ˆ zˆ and again observe that the onset of the dynamical instabil- X,ˆ Zˆ . While the vortices that are seen in Fig.7 are not ity is marked by the presence of a high-density core with ordered in a lattice, we expect that after a considerably{ } surface rippling, surrounded by a low-density halo-like {longer} period of evolution of the GPE at a constant rota- cloud, in the second row of Fig.9 where Ω 0.85ω⊥. tion frequency, the final state of the system is a triangular Upon halting the acceleration of the rotation frequency Abrikosov vortex lattice, as is seen in BECs subject to when Ω 0.85ω⊥ and then evolving the GPE at= constant −1 rotation about a principal axis of the trapping [22, 46– rotation frequency for the duration 500ω⊥ , the conden- = 11

FIG. 8. Comparison of the TF stationary solutions along Branch I to GPE simulations, for ε = 0.05 and γ = 4~3, during a quasi-adiabatic rampup of Ω. The first column (a, c, e) plots α as a function of Ω and the column (b, d, f) plots θ − ξ as a function of Ω, where θ is equal to π~8 (a, b), π~4 (c, d), or 3π~8 (e, f). sate is subject to the nucleation of a large number of vortices as seen in the third row of Fig.9. More vor- tices are found in Fig.9 than in Fig.7, which is likely due to the higher rotation frequency at which the quasi- FIG. 9. Cross-sections of the condensate density in the co- adiabatic rampup was halted. In both cases, however, we rotating x-y (first column) and x-z (second column) planes find that the vortex lines coincident upon the x z plane for ε = 0.05, γ = 4~3, and θ = π~4, during a quasi-adiabatic are almost completely aligned along the z-axis and that rampup of Ω at Ω = 0.4ω⊥ (first row) and Ω = 0.85ω⊥ (second −1 the background condensate density profile is tilted− with row), and after 500ω⊥ at constant Ω = 0.85ω⊥ (third row). respect to both the rotating trap and the rotation axis. The white lines represent the co-rotating X-Z axes.

VII. CONCLUSION has been proposed for accessing the branch defined for Ω ωb1, is to start from the α 0 stationary solution In this work, we have extended the Thomas-Fermi when ε 0 and then quasi-adiabatically tune ε to the de- theory for slowly rotating Bose-Einstein condensates in sired∈ [ final∞ value) whilst keeping Ω fixed= [44]. In principle a anisotropic harmonic traps to account for rotations of similar method= could be utilized to explore the stationary the trap about an axis that is not one of its three prin- solution along Branch III for Ω max Ωxc, Ωyc , Ωb2 cipal axes. In traps subject to tilted rotation, the sta- and θ 0, in an anisotropic harmonic trap, by starting tionary state density profile’s principal axes are gener- from an isotropic trap rotating at∈ a[ fixed{ frequency} and] ally tilted with respect to those of both the confinement adiabatically≠ tuning its anisotropy as desired. and the rotation. The quadrupolar irrotational velocity Our work also suggests that vortices are nucleated in profile describing the vorticity-free flow of the conden- response to a tilted rotating trap and are aligned along sate is also modified as a consequence of the tilting of the rotation axis, and not along one of the tilted principal the rotating harmonic trap. Our analysis of the resulting axes of the trap. Although we expect that the conden- stationary solutions demonstrate the existence of previ- sate’s final state in the dynamically unstable domain to ously unknown, tilted, solution branches (Branches III be a triangular vortex lattice, further work in this direc- and IV) that exist when Ω ω⊥. Although we have only tion is needed to resolve this, as well as the tilting angle conducted a systematic study of the dynamical stability of the background condensate density and the response of one of the five stationary> solution branches, Branch I, of the vortices to perturbations [19–21, 23, 24, 35–38]. it is nonetheless interesting to consider whether Branch The formalism outlined here for finding rotating frame III, in particular, becomes dynamically unstable immedi- stationary solutions with a tilting of the trap’s symme- ately upon reaching Ω ω⊥ or if its stability persists for try axes can be extended to more exotic condensates than a larger window. When θ 0 and ε 0, a method that the scalar one we have considered. Notably, in the field = = ≠ 12 of dipolar quantum gases, we expect that dipolar Bose- Equations (A1) and (A2) admit the solution pair Einstein condensates in the TF limit can be described in a similar manner, based on previous work on rotating ei- γ2 1 ε ω2 Ω2 ⊥ , (A3) ther the trapping or the dipole polarization about a prin- xc γ2 cos2 θ 1 ε sin2 θ cipal axis of the trapping [64–68]. Similarly we would ex- ( − ) = γ4 cos2 θ pect that spin-orbit-coupled BECs subject to an artificial cos2 ξ , (A4) xc 4 2 + ( − )2 2 gauge field that induces a synthetic rotation about a non- γ cos θ 1 ε sin θ principal axis would be described analogously, in the TF = which together yield the solution for α via α limit, to the formalism we have introduced here [69, 70]. + ( − ) xc xc Ω cos θ ξxc . We may also solve for δxc by substitut- ing these roots into Eq. (29). Crucially, when θ 0,= Eqs.( (A1− ) and) (A2) imply that ξxc δxc 0 and ACKNOWLEDGMENTS Ωxc αxc ω⊥ 1 ε, as expected [44]. The same= lim- iting forms are√ also valid when the trapping= is= axially S. B. P. is supported by an Australian Government symmetric= = about the− y-axis, i.e. γ 1 ε. For case b), Research Training Program Scholarship and by the Uni- we find that we have √ versity of Melbourne. The numerical simulations were = − conducted on the Spartan HPC cluster [71], and we thank αyc Ωyc cos θ ξyc , (A5) Research Computing Services at the University of Mel- δyc Ωyc sin θ ξyc . (A6) bourne for access to this resource. We also thank Nick = − ( − ) Parker and Thomas Bland for several stimulating discus- Substitution of these= into Eq.( (24− ) yields) sions that motivated this work. Ωyc ω⊥ 1 ε, (A7) √ Appendix A: Endpoints of the Stationary Solution and thus = + Branches 2 3 1 ε sin 2 θ ξyc 1 ε γ sin 2ξyc , (A8) The endpoints of the branches, save for Branch V, are which( admits+ ) [ the( solution− )] = ( − − ) ( ) defined by the limits given by: 3 1 ε sin 2θ tan 2ξyc . (A9) a)˜ωx 0 andω ˜y, ω˜z 0, 3 1 ε cos 2θ γ2 1 ε ( + ) ( ) ( ) = b)˜ωy → 0 andω ˜x, ω˜z ≠ 0, Via Eqs. (A5)–(A7( ),+ and) (A9( ),) we− may+ − obtain the so- lutions of αyc and δyc in this limit. For the special cases c)˜ωx,→ω˜y 0 andω ˜z ≠ 0, that the trap is not tilted, i.e. θ 0, and/or is axially symmetric about the y-axis, i.e. γ 1 ε, we have d)˜ωy, ω˜z → 0 andω ˜x ≠ 0. αyc Ωyc ω⊥ 1 ε and δyc ξ=yc 0√ [44]. The limits c) and d) are somewhat= more− involved. In Let us denote the critical→ values of quantities≠ such as Ω, √ case= c)− we= have− δ + Ω sin =θ ξ= , but the limit α, δ, and ξ at the limiting cases a), b), c) and d) by the xyc xyc xyc of α is not as obvious and must be found by solving subscripts xc, yc, xyc, and yzc respectively. xyc Eq. (27). This gives us= ( − ) For case a), Eq. (28) implies that we have αxc Ωxc cos θ ξxc . Substituting this into Eqs. (23) and 2 2 Ωxyc sin 2 θ ξxyc γ 1 ε sin 2ξxyc (27) yields = αxyc . ( − ) 4Ωxyc sin θ ξxyc 2 2 2 2 2 [ ( − )] + ( − + ) ( (A10)) Ω cos θ ξxc 1 ε cos ξxc γ sin ξxc, (A1) = xc Substituting these relations into( − Eqs. ()23) and (24) re- Ω2 sin 2 θ ξ γ2 1 ε sin 2ξ . (A2) xc ( − xc ) = ( − ) + xc sults in the system of equations given by: [ ( − )] = ( − + ) ( )

4Ω2 γ2 1 ε sin 2ξ 12Ω4 cos2 θ ξ 8 γ2 1 ε Ω2 γ2 1 ε 8Ω2 cos 2ξ sin 2ξ 0, tan θ ξ sin2 θ ξ ( − + ) ( ) ( − ) − ( + − ) + ( − + ) œ ( ) + ( )  − ¡ = (A11) ( − ) γ2 (1 −ε )sin 2ξ 2Ω2 8 1 ε Ω2 1 9 cos 2 θ ξ sin θ ξ γ2 1 ε sin 2ξ 12Ω2 cos θ ξ 0. sin θ ξ ( − + ) ( ) { ( + ) + [ + ( ( − ))]} ( − ) + ( − + ) ( )  ( − ) + =(A12) ( − ) 13

Solving these simultaneously for Ω and ξ yields the lim- The substitution of these relations into Eqs. (24) and (25) iting values, Ωxyc and ξxyc, which subsequently allows results in the following system of equations: for the solution of αxyc and δxyc from Eq. (A10) and the relation δxyc Ωxyc sin θ ξxyc respectively. In case d), we have α Ω cos θ ξ and from solving Eq. (27) we also= find that( − ) = − ( − ) 2 2 1 ε γ sin 2ξyzc Ωyzc sin 2 θ ξyzc δyzc . 4Ωyzc cos θ ξyzc ( − − ) ( ) − [ ( − )] = (A13) ( − )

γ2 1 ε sin 2ξ 2Ω2 8 1 ε Ω2 1 9 cos 2 θ ξ cos θ ξ γ2 1 ε sin 2ξ 12Ω2 sin θ ξ 0, cos θ ξ ( − + ) ( ) { ( + ) + [ − ( ( − ))]} ( − ) + ( − + ) ( )  ( − ) + =(A14) ( − ) 2 γ2 1 ε sin 2ξ tan θ ξ Ω sin θ ξ γ2 1 ε sin 2ξ γ2 cos2 ξ 1 ε sin2 ξ Ω2 sin2 θ ξ 0. 2 2 4Ω cos θ ξ ( − + ) ( ) ( − ) ( − ) ( − + ) ( ) + ( − ) − ( − ) − +  + (A15)= ( − )

As in case c), solving these equations for Ω and ξ yields δ in terms of the TF semi-axes: Ωyzc and ξyzc, and thus also αyzc and δyzc via the relation R2 R2 αyzc Ωyzc cos θ ξyzc and Eq. (A13) respectively. α x y Ω cos θ ξ , (B1) For both limits c) and d), the limits for the special case R2 R2 x − y where= θ− 0 evaluate( − to simple) closed forms given by: = ŒR2 R2 ‘ ( − ) δ y + z Ω sin θ ξ . (B2) R2 R2 = 2 2 2 y z Ωxyc Ωyzc ω⊥ 1 γ 4γ ε , (A16) − ¼ = Œ ‘ ( − ) »2 2 Since Eq. (17) exhibits+ a twofold rotation symmetry 2 2 ε 4γ ε ε cos= ξxyc = + + + about the y-axis, ξ has a period of π and so we assume 2 γ2 1 »ε γ2 2 2 ε ( + )( + − ) that θ ξ π 2, π 2 without loss of generality. This = γ2 4γ2 ε2 1 ε γ2 choice of the principal branch fixes cos θ ξ 0, whereas ( − + )[ − ( + )], (A17) γ2» 1 ε γ2 2 2 ε sin θ −ξ ∈ 0(− when~ ξ~ ] θ and sin θ ξ 0 when ξ θ. ( + + + − ) 2 2 2 Therefore, by inspection of Eqs. (B1) and( − ()B2≥), we have: 2 − γ 4γ ε 4 cos ξyzc ( − + )[ − ( + )] ( − ) > < ( − ) < > γ2 1 »ε γ2 2 2 ε i) Rx Ry Rz when α 0, δ 0, θ ξ 0 or when ( + − ) α 0, δ 0, θ ξ 0, = 2 ε 4γ2 ε2 4 3ε ( − + )[ − ( + )] , (A18) > > > > − > 2 γ2 1» ε γ2 2 2 ε ii) Rx> Ry > > > − < the respective limits. It is noted that Eqs. (A17) and iii) Ry Rx>and R α 0, δ 0, θ ξ 0 or (A18) formally exhibit a removable singularity when the when α 0, δ 0, θ ξ 0, trap is axially symmetric abouty ˆ, i.e. γ2 1 , and > > < > − > iv) R R R when α 0, δ 0, θ ξ 0 or when in this limit we have cos2 ξ 2 1  3 2  and z y< x < − < xyc α 0, δ 0, θ ξ 0. cos2 ξ 4  3 2  . = − yzc > > < > − < = ( − )~[ ( − )] Note that< the< signs− of α>, δ, and θ ξ cannot conclusively = ( − )~[ ( − )] determine an inequality or equality relating Rx and Rz Appendix B: Visualizing the TF Density Profiles in the scenarios ii) and iii). − In Fig. 10 we illustrate these relations by providing In this section, we provide the reader with a description examples of the typical cross-sections of the TF density of how the signs of the velocity amplitudes, α and δ, and in the upright co-rotating x-y (first and third columns) the angle θ ξ provide us with a considerable amount and x-z (second and fourth columns) planes. Each row of of qualitative information of the shape of the Thomas- Fig. 10 corresponds to a different combination of positive Fermi density− profile corresponding to a given solution of or negative values of α and δ, with α 0, δ 0 presented Eqs. (26)–(29). Let us restate the definitions of α and in the first row, α 0, δ 0 in the second row, α 0, δ 0 > > > < < > 14 in the third row, and α 0, δ 0 in the fourth row. the values of Ω where the GPE cross-section snapshots In addition, the first and second columns correspond to are taken have chosen in order to illustrate the three main θ ξ 0 and the third and< fourth< columns correspond to stages of the evolution of the BEC from TF-like, via the θ ξ 0. intermediate stage with a halo-like cloud surrounding the − > deformed core, to a state containing many vortices. − < Appendix C: Additional GPE Data

For the sake of completeness, we present a comparison of the values of δ, between those pertaining to the TF stationary solutions and those obtained from the GPE simulations in Fig. 11. In this figure, the first column (a, c, e) pertains to the trapping parameters ε 0 and γ 3 4, and in the second column (b, d, f), ε 0.05 and γ 4 3. The trap tilt angles represented in the Fig.= 11 are θ = π~ 8 (first row), θ π 4 (second row), and= θ 3π 8 (third= ~ row); the analogous comparisons of α and ε are found= ~ in Figs.6 and8 for= the~ parameters represented= in~ the first and second columns, respectively. The deviation of the GPE-derived values of δ from the corresponding stationary state values illustrates the transition from the TF state to that with vortices, due to the dynamical instability of the TF states as Ω min Ωxc, Ωyc , that is discussed in the main text. → { }

FIG. 11. Comparison of δ, as a function of Ω, between the TF stationary solution values along Branch I and those derived from GPE simulations of a quasi-adiabatic rampup of Ω, for ε = 0 and γ = 3~4 (a, c, e) and ε = 0.05 and γ = 4~3 (b, d, f). θ equals π~8 in the first row (a, b), π~4 in the second row (c, d), or 3π~8 in the third row (e, f).

We also present the cross-sections of the density, in the x-y and x-z planes, for the angles θ π 8, 3π 8 that were not discussed in the main text. Specifically, for θ π 8, these GPE-derived density cross-sections= { ~ are~ plotted} FIG. 12. Cross-sections of the condensate density in the co- in Fig. 12 for the trap with ε 0 and γ 3 4 and in Fig. 13= rotating x-y (first column) and x-z (second column) planes for for~ the parameters ε 0.05 and γ 4 3. Similarly, for θ ε = 0, γ = 3~4, and θ = π~8, during a quasi-adiabatic rampup 3π 8, the density cross-sections= are presented= ~ in Fig. 14 of Ω at Ω = 0.5ω⊥ (first row) and Ω = 0.8ω⊥ (second row), and −1 for the trapping parameters= ε =0 and~ γ 3 4 and in= after 500ω⊥ at constant Ω = 0.8ω⊥ (third row). The white Fig.~ 15 for the parameters ε 0.05 and γ 4 3. Note that lines represent the co-rotating X-Z axes. = = ~ = = ~ 15

y z y z

x x x x α > 0, δ

> Rx Rx Rz Ry Rz 0 Ry α > 0,

δ R < Rx z Rx Ry 0 Rz Ry α < 0,

δ Rx Rx >

0 Rz Ry Rz Ry α < 0,

δ Rx Rx < Rz 0 Ry Rz Ry

Θ ξ 0 Θ ξ 0

FIG. 10. Shaded examples of cross-sections− > at constant z = 0 (first and third columns) and y−= 0< (second and fourth columns), where the TF semi-axes are related to each other in the ratio 5 ∶ 7 ∶ 10. Here, θ − ξ is positive (first and second columns) or negative (third and fourth columns), α is positive (first and second rows) or negative (third and fourth rows) and δ is positive (first and third rows) or negative (second and fourth rows). 16

FIG. 13. Cross-sections of the condensate density in the co- FIG. 14. Cross-sections of the condensate density in the co- rotating x-y (first column) and x-z (second column) planes rotating x-y (first column) and x-z (second column) planes for for ε = 0.05, γ = 4~3, and θ = π~8, during a quasi-adiabatic ε = 0, γ = 3~4, and θ = 3π~8, during a quasi-adiabatic rampup rampup of Ω at Ω = 0.5ω⊥ (first row) and Ω = 0.85ω⊥ (second of Ω at Ω = 0.4ω⊥ (first row) and Ω = 0.6ω⊥ (second row), and −1 −1 row), and after 500ω⊥ at constant Ω = 0.85ω⊥ (third row). after 500ω⊥ at constant Ω = 0.6ω⊥ (third row). The white The white lines represent the co-rotating X-Z axes. lines represent the co-rotating X-Z axes. 17

FIG. 15. Cross-sections of the condensate density in the co- rotating x-y (first column) and x-z (second column) planes for ε = 0.05, γ = 4~3, and θ = 3π~8, during a quasi-adiabatic rampup of Ω at Ω = 0.4ω⊥ (first row) and Ω = 0.7ω⊥ (second −1 row), and after 500ω⊥ at constant Ω = 0.7ω⊥ (third row). The white lines represent the co-rotating X-Z axes.

[1] Lars Onsager, “Statistical Hydrodynamics,” Il Nuovo Ci- [6] Shaun P. Johnstone, Andrew J. Groszek, Philip T. mento 6, 279–287 (1949). Starkey, Christopher J. Billington, Tapio P. Simula, and [2] Richard Phillips Feynman, “Chapter II Application of Kristian Helmerson, “Evolution of Large-Scale Flow from Quantum Mechanics to Liquid Helium,” (Elsevier, 1955) Turbulence in a Two-Dimensional Superfluid,” Science pp. 17–53. 364, 1267–1271 (2019). [3] Henry Edgar Hall and William Frank Vinen, “The Ro- [7] Lev Pitaevskii and Sandro Stringari, Bose-Einstein Con- tation of Liquid Helium II I. Experiments on the Prop- densation and Superfluidity, 1st ed., International Series agation of Second Sound in Uniformly Rotating Helium of Monographs on Physics No. 164 (Clarendon Press, II,” Proceedings of the Royal Society A: Mathematical, 2016). Physical and Engineering Sciences 238, 204–214 (1956). [8] Nigel R. Cooper, “Rapidly Rotating Atomic Gases,” Ad- [4] Henry Edgar Hall and William Frank Vinen, “The Rota- vances in Physics 57, 539 (2008). tion of Liquid Helium II II. The Theory of Mutual Fric- [9] Alexander L. Fetter, “Rotating Trapped Bose-Einstein tion in Uniformly Rotating Helium II,” Proceedings of Condensates,” Reviews of Modern Physics 81, 647–691 the Royal Society A: Mathematical, Physical and Engi- (2009). neering Sciences 238, 215–234 (1956). [10] , Quantum Liquids: Bose- [5] Guillaume Gauthier, Matthew T. Reeves, Xiaoquan Yu, Condensation and Cooper Pairing in Condensed Matter Ashton S. Bradley, Mark Baker, Thomas A. Bell, Halina Systems (Oxford University Press, 2006). Rubinsztein-Dunlop, Matthew J. Davis, and Tyler W. [11] Eleanor Hodby, Gerald Hechenblaikner, Stephen A. Hop- Neely, “Giant Vortex Clusters in a Two-Dimensional kins, Onofrio M. Marag`o, and Christopher J. Foot, Quantum Fluid,” Science 364, 1264–1267 (2019). “Vortex Nucleation in Bose-Einstein Condensates in an 18

Oblate, Purely Magnetic Potential,” Physical Review [27] Nicolas Regnault and Thierry Jolicoeur, “Quantum Hall Letters 88, 010405 (2001). Fractions in Rotating Bose-Einstein Condensates,” Phys- [12] Kirk W. Madison, Fr´ed´ericChevy, Wendel Wohlleben, ical Review Letters 91, 030402 (2003). and Jean Dalibard, “Vortex Formation in a Stirred Bose- [28] Volker Schweikhard, Ian Coddington, Peter Engels, Einstein Condensate,” Physical Review Letters 84, 806– Veronique P. Mogendorff, and , 809 (2000). “Rapidly Rotating Bose-Einstein Condensates in and [13] Tyler W. Neely, Edward Carlo Samson, Ashton S. near the Lowest Landau Level,” Physical Review Letters Bradley, Matthew J. Davis, and Brian P. Anderson, 92, 040404 (2004). “Observation of Vortex Dipoles in an Oblate Bose- [29] Simone Zuccher, Marco Caliari, Andrew W. Baggaley, Einstein Condensate,” Physical Review Letters 104, and Carlo F. Barenghi, “Quantum Vortex Reconnec- 160401 (2010). tions,” Physics of Fluids 24, 125108 (2012). [14] Woo Jin Kwon, Joon Hyun Kim, Sang Won Seo, and [30] Simone Serafini, Luca Galantucci, Elena Iseni, Tom Bien- Y. Shin, “Observation of von K´arm´anVortex Street in aim´e,Russell N. Bisset, Carlo F. Barenghi, Franco Dal- an Atomic Superfluid Gas,” Physical Review Lett. 117, fovo, Giacomo Lamporesi, and Gabriele Ferrari, “Vortex 245301 (2016). Reconnections and Rebounds in Trapped Atomic Bose- [15] Emanuel Alves Lima Henn, Jorge Amin Seman, Edmir Einstein Condensates,” Physical Review X 7, 021031 R. F. Ramos, M. Caracanhas, Patricia Castilho, Ed- (2017). uardo P. Ol´ımpio, Giacomo Roati, Daniel Varelas Ma- [31] Kazuki Sasaki, Naoya Suzuki, and Hiroki Saito, galh˜aes,Kilvia Mayre Farias Magalh˜aes, and Vander- “B´enardvon K´arm´anVortex Street in a Bose-Einstein lei Salvador Bagnato, “Observation of Vortex Formation Condensate,” Physical Review Letters 104, 150404 in an Oscillating Trapped Bose-Einstein Condensate,” (2010). Physical Review A 79, 043618 (2009). [32] Andrew W. Baggaley and Nick G. Parker, “Kelvin- [16] Emanuel Alves Lima Henn, Jorge Amin Seman, Gia- Helmholtz Instability in a Single-Component Atomic Su- como Roati, Kilvia Mayre Farias Magalh˜aes, and Van- perfluid,” Physical Review A 97, 053608 (2018). derlei Salvador Bagnato, “Emergence of Turbulence in an [33] Juan J. Garc´ıa-Ripoll and V´ıctorM. P´erez-Garc´ıa,“Vor- Oscillating Bose-Einstein Condensate,” Physical Review tex Nucleation and Hysteresis Phenomena in Rotat- Letters 103, 045301 (2009). ing Bose-Einstein Condensates,” Physical Review A 63, [17] P. C. Haljan, I. Coddington, P. Engels, and Eric Allin 041603(R) (2001). Cornell, “Driving Bose-Einstein-Condensate Vorticity [34] Brian Jackson and Carlo F. Barenghi, “Hysteresis Ef- with a Rotating Normal Cloud,” Physical Review Let- fects in Rotating Bose-Einstein Condensates,” Physical ters 87, 210403 (2001). Review A 74, 043618 (2006). [18] Chad N. Weller, Tyler W. Neely, David R. Scherer, Ash- [35] Sandro Stringari, “Superfluid Gyroscope with ton S. Bradley, Matthew J. Davis, and Brian P. An- Atomic Gases,” Physical Review Letters 86, 4725–4728 derson, “Spontaneous Vortices in the Formation of Bose- (2001). Einstein Condensates,” Nature 455, 948–951 (2008). [36] Eleanor Hodby, Stephen A. Hopkins, Gerald Hechen- [19] Anatoly A. Svidzinsky and Alexander L. Fetter, “Dy- blaikner, Nathan L. Smith, and Christopher J. Foot, namics of a Vortex in a Trapped Bose-Einstein Conden- “Experimental Observation of a Superfluid Gyroscope sate,” Physical Review A 62, 063617 (2000). in a Dilute Bose-Einstein Condensate,” Physical Review [20] Vincent Bretin, Peter Rosenbusch, Fr´ed´eric Chevy, Letters 91, 090403 (2003). Georgy V. Shlyapnikov, and Jean Dalibard, [37] Nathan L. Smith, William H. Heathcote, Jan Max “Quadrupole Oscillation of a Single-Vortex Bose- Krueger, and Christopher J. Foot, “Experimental Ob- Einstein Condensate: Evidence for Kelvin Modes,” servation of the Tilting Mode of an Array of Vortices Physical Review Letters 90, 100403 (2003). in a Dilute Bose-Einstein Condensate,” Physical Review [21] Tapio P. Simula, Takeshi Mizushima, and Kazushige Letters 93, 080406 (2004). Machida, “Kelvin Waves of Quantized Vortex Lines in [38] Andrew Tasman Powis, Steven J. Sammut, and Tapio P. Trapped Bose-Einstein Condensates,” Physical Review Simula, “Vortex Gyroscopic Imaging of Planar Superflu- Letters 101, 020402 (2008). ids,” Physical Review Letters 113, 165303 (2014). [22] Jamil R. Abo-Shaeer, Chandra Raman, Johnny M. Vo- [39] Iwo Bialynicki-Birula and Zofia Bialynicka-Birula, gels, and , “Observation of Vortex Lat- “Center-of-Mass Motion in the Many-Body Theory of tices in Bose-Einstein Condensates,” Science 292, 476– Bose-Einstein Condensates,” Physical Review A 65, 479 (2001). 063606 (2002). [23] Gordon Baym, “Tkachenko Modes of Vortex Lattices in [40] Iwo Bialynicki-Birula and Tomasz Sowinski, “Gravity- Rapidly Rotating Bose-Einstein Condensates,” Physical Induced Resonances in a Rotating Trap,” Physical Re- Review Letters 91, 110402 (2003). view A 71, 043610 (2005). [24] Ian Coddington, Peter Engels, Volker Schweikhard, and [41] Stefano Giovanazzi, Axel G¨orlitz, and Tilman Pfau, Eric Allin Cornell, “Observation of Tkachenko Oscilla- “Tuning the Dipolar Interaction in Quantum Gases,” tions in Rapidly Rotating Bose-Einstein Condensates,” Physical Review Letters 89, 130401 (2002). Physical Review Letters 91, 100402 (2003). [42] Yijun Tang, Wil Kao, Kuan-Yu Li, and Benjamin L. Lev, [25] Tin-Lun Ho, “Bose-Einstein Condensates with Large “Tuning the Dipole-Dipole Interaction in a Quantum Gas Number of Vortices,” Physical Review Letters 87, 060403 with a Rotating Magnetic Field,” Physical Review Let- (2001). ters 120, 230401 (2018). [26] Nigel R. Cooper, Nicola K. Wilkin, and J. Mike F. Gunn, [43] Eugene R. Marshalek, “An Overlooked Figure of Equilib- “Bose-Einstein Condensates with Large Number of Vor- rium of a Rotating Ellipsoidal Self-Gravitating Fluid and tices,” Physical Review Letters 87, 120405 (2001). the Riemann Theorem,” Physics of Fluids 8, 3414–3422 19

(1996). [59] Anatoly A. Svidzinsky and Alexander L. Fetter, “Normal [44] Alessio Recati, Francesca Zambelli, and Sandro Modes of a Vortex in a Trapped Bose-Einstein Conden- Stringari, “Overcritical Rotation of a Trapped Bose- sate,” Physical Review A 58, 3168–3179 (1998). Einstein Condensate,” Physical Review Letters 86, 377– [60] Weizhu Bao and Yongyong Cai, “Mathematical The- 380 (2001). ory and Numerical Methods for Bose-Einstein Conden- [45] Subhasis Sinha and Yvan Castin, “Dynamical Instability sation,” Kinetic and Related Models 6, 1–135 (2013). of a Rotating Bose-Einstein Condensate,” Physical Re- [61] Weizhu Bao and Hanquan Wang, “An Efficient and Spec- view Letters 87, 190402 (2001). trally Accurate Numerical Method for Computing Dy- [46] Carlos Lobo, Alice Sinatra, and Yvan Castin, “Vortex namics of Rotating Bose-Einstein Condensates,” Journal Lattice Formation in Bose-Einstein Condensates,” Phys- of Computational Physics 217, 612–626 (2006). ical Review Letters 92, 020403 (2004). [62] Makoto Tsubota, Kenichi Kasamatsu, and Masahito [47] Nick G. Parker and Charles S. Adams, “Emergence Ueda, “Vortex Lattice Formation in a Rotating Bose- and Decay of Turbulence in Stirred Atomic Bose- Einstein Condensate,” Physical Review A 65, 023603 Einstein Condensates,” Physical Review Letters 95, (2002). 145301 (2005). [63] Kenichi Kasamatsu, Makoto Tsubota, and Masahito [48] Nick G. Parker, Rick M. W. van Bijnen, and Andrew Mc- Ueda, “Nonlinear Dynamics of Vortex Formation in a Callum Martin, “Instabilities Leading to Vortex Forma- Rotating Bose-Einstein Condensate,” Physical Review A tion in Rotating Bose-Einstein Condensates,” Physical 67, 033610 (2003). Review A 73, 061603(R) (2006). [64] Rick M. W. van Bijnen, Duncan H. J. O’Dell, Nick G. [49] Ivan Corro, Nick G. Parker, and Andrew McCallum Parker, and Andrew McCallum Martin, “Dynamical In- Martin, “Rotation of an Atomic BoseEinstein Conden- stability of a Rotating Dipolar Bose-Einstein Conden- sate With and Without a Quantized Vortex,” Journal of sate,” Physical Review Letters 98, 150401 (2007). Physics B: Atomic, Molecular and Optical Physics 40, [65] Rick M. W. van Bijnen, Anita Jane Dow, Duncan H. J. 3615–3628 (2007). O’Dell, Nick G. Parker, and Andrew McCallum Mar- [50] Kirk W. Madison, Fr´ed´ericChevy, Vincent Bretin, and tin, “Exact Solutions and Stability of Rotating Dipolar Jean Dalibard, “Stationary States of a Rotating Bose- Bose-Einstein Condensates in the Thomas-Fermi Limit,” Einstein Condensate: Routes to Vortex Nucleation,” Physical Review A 80, 033617 (2009). Physical Review Letters 86, 4443–4446 (2001). [66] Andrew McCallum Martin, Neil G. Marchant, Duncan [51] Gerald Hechenblaikner, Eleanor Hodby, Stephen A. Hop- H. J. O’Dell, and Nick G. Parker, “Vortices and Vor- kins, Onofrio M. Marag`o, and Christopher J. Foot, “Di- tex Lattices in Quantum Ferrofluids,” Journal of Physics: rect Observation of Irrotational Flow and Evidence of Condensed Matter 29, 103004 (2017). Superfluidity in a Rotating Bose-Einstein Condensate,” [67] Srivatsa Badari Prasad, Thomas Bland, Brendan Craig Physical Review Letters 88, 070406 (2002). Mulkerin, Nick G. Parker, and Andrew McCallum Mar- [52] Christopher J. Pethick and Henrik Smith, Bose-Einstein tin, “Instability of Rotationally Tuned Dipolar Bose- Condensation in Dilute Gases, 2nd ed. (Cambridge Uni- Einstein Condensates,” Physical Review Letters 122, versity Press, 2008). 050401 (2019). [53] Mark Edwards and Keith Burnett, “Numerical Solution [68] Srivatsa Badari Prasad, Thomas Bland, Brendan Craig of the Nonlinear Schr¨odingerEquation for Small Samples Mulkerin, Nicholas G. Parker, and Andrew McCal- of Trapped Atoms,” Physical Review A 51, 1382–1386 lum Martin, “Vortex Lattice Formation in Dipolar Bose- (1995). Einstein Condensates via Rotation of the Polarization,” [54] Gordon Baym and Christopher J. Pethick, “Ground- Physical Review A 100, 023625 (2019). State Properties of Magnetically Trapped Bose- [69] Lachlan B. Taylor, Rick M. W. van Bijnen, Duncan H. J. Condensed Gas,” Physical Review Letters 76, O’Dell, Nick G. Parker, Servaas J. J. M. F. Kokkelmans, 6–9 (1996). and Andrew McCallum Martin, “Synthetic Magnetohy- [55] Horace Lamb, Hydrodynamics, 6th ed. (Dover Publica- drodynamics in Bose-Einstein Condensates and Routes tions, 1945) Chap. 5, pp. 146–147. to Vortex Nucleation,” Physical Review A 84, 021604(R) [56] Lev Davidovich Landau and Evgeny Mikhailovich Lif- (2011). shitz, Fluid Mechanics, 2nd ed., Course of Theoretical [70] Chunlei Qu and Sandro Stringari, “Angular Momentum Physics No. 6 (Butterworth-Heinemann, 1987). of a Bose-Einstein Condensate in a Synthetic Rotational [57] W. D. Heiss and R. G. Nazmitdinov, “Self-Consistent Field,” Physical Review Letters 120, 183202 (2018). Harmonic Oscillator Model and Tilted Rotation,” Phys- [71] Bernard Meade, Lev Lafayette, Greg Sauter, and Daniel ical Review C 65, 054304 (2002). Tosello, “Spartan HPC-Cloud Hybrid: Delivering Perfor- [58] Wen-Chiu Wu and Allan Griffin, “Quantized Hydrody- mance and Flexibility,” University of Melbourne (2017), namic Model and the Dynamic Structure Factor for a 10.4225/49/58ead90dceaaa. Trapped Bose Gas,” Physical Review A 54, 4204–4212 (1996).