12.07.2015 Views

Vortices in a trapped dilute Bose–Einstein condensate

Vortices in a trapped dilute Bose–Einstein condensate

Vortices in a trapped dilute Bose–Einstein condensate

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

INSTITUTE OF PHYSICS PUBLISHINGJOURNAL OF PHYSICS: CONDENSED MATTERJ. Phys.: Condens. Matter 13 (2001) R135–R194 www.iop.org/Journals/cm PII: S0953-8984(01)08644-1TOPICAL REVIEW<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>Alexander L Fetter and Anatoly A Svidz<strong>in</strong>skyDepartment of Physics, Stanford University, Stanford, CA 94305-4060, USAReceived 29 November 2000, <strong>in</strong> f<strong>in</strong>al form 2 February 2001AbstractWe review the theory of vortices <strong>in</strong> <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>sand compare theoretical predictions with exist<strong>in</strong>g experiments. Mean-fieldtheory based on the time-dependent Gross–Pitaevskii equation describes thema<strong>in</strong> features of the vortex states, and its predictions agree well with availableexperimental results. We discuss various properties of a s<strong>in</strong>gle vortex, <strong>in</strong>clud<strong>in</strong>gits structure, energy, dynamics, normal modes, and stability, as well as vortexarrays. When the nonuniform <strong>condensate</strong> conta<strong>in</strong>s a vortex, the excitationspectrum <strong>in</strong>cludes unstable (‘anomalous’) mode(s) with negative frequency.Trap rotation shifts the normal-mode frequencies and can stabilize the vortex.We consider the effect of thermal quasiparticles on vortex normal modes aswell as possible mechanisms for vortex dissipation. Vortex states <strong>in</strong> mixturesand sp<strong>in</strong>or <strong>condensate</strong>s are also discussed.Contents1. Introduction2. The time-dependent Gross–Pitaevskii equation2.1. Unbounded <strong>condensate</strong>2.2. Quantum-hydrodynamic description of the <strong>condensate</strong>2.3. Vortex dynamics <strong>in</strong> two dimensions2.4. Trapped <strong>condensate</strong>3. Static vortex states3.1. Structure of a s<strong>in</strong>gle <strong>trapped</strong> vortex3.2. Thermodynamic critical angular velocity for vortex stability3.3. Experimental creation of a s<strong>in</strong>gle vortex3.4. Vortex arrays4. Bogoliubov equations: stability of small-amplitude perturbations4.1. General features for nonuniform <strong>condensate</strong>4.2. Uniform <strong>condensate</strong>4.3. Quantum-hydrodynamic description of small-amplitude normal modes4.4. A s<strong>in</strong>gly quantized vortex <strong>in</strong> an axisymmetric trap0953-8984/01/120135+60$30.00 © 2001 IOP Publish<strong>in</strong>g Ltd Pr<strong>in</strong>ted <strong>in</strong> the UK R135


R136A L Fetter and A A Svidz<strong>in</strong>sky5. Vortex dynamics5.1. Time-dependent variational analysis5.2. The method of matched asymptotic expansions5.3. Normal modes of a vortex <strong>in</strong> a rotat<strong>in</strong>g two-dimensional TF <strong>condensate</strong>5.4. Normal modes of a vortex <strong>in</strong> a rotat<strong>in</strong>g three-dimensional TF <strong>condensate</strong>6. The effect of thermal quasiparticles, vortex lifetime, and dissipation6.1. Bogoliubov and Hartree–Fock–Bogoliubov theories6.2. Dissipation and vortex lifetimes7. Vortex states <strong>in</strong> mixtures and sp<strong>in</strong>or <strong>condensate</strong>s7.1. Basic phenomena7.2. Stability theory8. Conclusions and outlook1. IntroductionThe recent dramatic achievement of Bose–E<strong>in</strong>ste<strong>in</strong> condensation <strong>in</strong> <strong>trapped</strong> alkali-metal gasesat ultralow temperatures [1–3] has stimulated <strong>in</strong>tense experimental and theoretical activity.The atomic Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>s (BECs) differ fundamentally from the helium BEC<strong>in</strong> several ways. First, BECs <strong>in</strong> helium are uniform. In contrast, the trapp<strong>in</strong>g potential thatconf<strong>in</strong>es an alkali-metal-atom vapour BEC yields a significantly nonuniform density. Anotherdifference is that <strong>in</strong> bulk superfluid 4 He, measurements of the momentum distribution haveshown that the low-temperature <strong>condensate</strong> fraction is ∼0.1, with the rema<strong>in</strong>der of the particles<strong>in</strong> f<strong>in</strong>ite-momentum states [4, 5], whereas the low-temperature atomic <strong>condensate</strong>s can beprepared with essentially all atoms <strong>in</strong> the Bose <strong>condensate</strong>. F<strong>in</strong>ally, the <strong>condensate</strong>s of alkalivapours are pure and <strong>dilute</strong> (with mean particle density ¯n and ¯n|a| 3 ≪ 1), so the <strong>in</strong>teractionscan be accurately parametrized <strong>in</strong> terms of a scatter<strong>in</strong>g length a (<strong>in</strong> current experiments,alkali-metal-atom BECs are much less dense than air at normal pressure). This situationdiffers from that for superfluid 4 He, where the relatively high density and strong repulsive<strong>in</strong>teractions greatly complicate the analytical treatments. As a result, a relatively simplenonl<strong>in</strong>ear Schröd<strong>in</strong>ger equation (the Gross–Pitaevskii equation) gives a precise description ofthe atomic <strong>condensate</strong>s and their dynamics (at least at low temperatures). One should mention,however, that unlike the sp<strong>in</strong>less 4 He atoms, alkali atoms have nonzero hyperf<strong>in</strong>e sp<strong>in</strong>s, andvarious forms of sp<strong>in</strong>-gauge effects can be important [6].Bulk superfluids are dist<strong>in</strong>guished from normal fluids by their ability to supportdissipationless flow. Such persistent currents are <strong>in</strong>timately related to the existence of quantizedvortices, which are localized phase s<strong>in</strong>gularities with <strong>in</strong>teger topological charge. The superfluidvortex is an example of a topological defect that is well known <strong>in</strong> liquid helium [7,8] and<strong>in</strong> superconductors [9]. The occurrence of quantized vortices <strong>in</strong> superfluids has been the focusof fundamental theoretical and experimental work [10–14]. Vortex-like excitations exist <strong>in</strong> theEarth’s atmosphere [15], <strong>in</strong> superfluid hadronic matter (neutron stars) [16], and even <strong>in</strong> rotat<strong>in</strong>gnuclei [17]. Examples of other topological defects that could exist <strong>in</strong> <strong>dilute</strong> gas <strong>condensate</strong>sare ‘textures’ found <strong>in</strong> Fermi superfluid 3 He [18], skyrmions [19,20], and sp<strong>in</strong> monopoles [21].<strong>Vortices</strong> <strong>in</strong> the A and B phases of 3 He are discussed <strong>in</strong> detail <strong>in</strong> the review articles [22, 23].In superfluid 3 He the Cooper pairs have both orbital and sp<strong>in</strong> angular momentum. These<strong>in</strong>ternal quantum numbers imply a rich phase diagram of allowed vortex structures, <strong>in</strong>clud<strong>in</strong>gnonquantized vortices with cont<strong>in</strong>uous vorticity (see also references [24, 25]).In the framework of hydrodynamics, the vortices obta<strong>in</strong>ed from the Gross–Pitaevskii(GP) equation are analogous to vortices <strong>in</strong> classical fluids [26]. Also the GP equationprovides an approximate description of some aspects of superfluid behaviour of helium,


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R137such as the annihilation of vortex r<strong>in</strong>gs [27], the nucleation of vortices [28], and vortex-l<strong>in</strong>ereconnection [29, 30].The <strong>in</strong>itial studies of <strong>trapped</strong> Bose <strong>condensate</strong>s concentrated on measur<strong>in</strong>g the energy and<strong>condensate</strong> fraction, along with the lowest-ly<strong>in</strong>g collective modes and quantum-mechanical<strong>in</strong>terference effects (see, for example, reference [31]). Although the possibility of <strong>trapped</strong>quantized vortices was quickly recognized [32], successful experimental verification has takenseveral years [33–37]. This review focuses on the behaviour of quantized vortices <strong>in</strong> <strong>trapped</strong><strong>dilute</strong> Bose <strong>condensate</strong>s, emphasiz<strong>in</strong>g the qualitative features along with the quantitativecomparison between theory and experiment.The plan of the paper is the follow<strong>in</strong>g. In section 2 we discuss the basic formalism ofmean-field theory (the time-dependent Gross–Pitaevskii equation) that describes <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>s <strong>in</strong> the low-temperature limit. We summarize properties of vortices <strong>in</strong> auniform <strong>condensate</strong> and also <strong>in</strong>troduce relevant length and energy scales of a <strong>condensate</strong> <strong>in</strong>a harmonic trap. In section 3 we discuss the structure of stationary vortex states <strong>in</strong> <strong>trapped</strong><strong>condensate</strong>s. We analyse the energy of a straight vortex as a function of displacement fromthe trap centre and consider conditions of vortex stability when the trap rotates. Also wediscuss the recent experimental creation of a s<strong>in</strong>gle vortex and vortex arrays. In section 4 we<strong>in</strong>troduce the concept of elementary excitations (the Bogoliubov equations) and analyse thelowest (unstable) mode of the vortex for different values of the <strong>in</strong>teraction parameter. We alsoconsider the splitt<strong>in</strong>g of the <strong>condensate</strong> normal modes due to the presence of a vortex l<strong>in</strong>e.In section 5 we <strong>in</strong>vestigate the general dynamical behaviour of a vortex, on the basis of atime-dependent variational analysis and on the method of matched asymptotic expansions.The latter method allows us to take <strong>in</strong>to account effects of both nonuniform <strong>condensate</strong>density and vortex curvature. We consider normal modes of a vortex <strong>in</strong> two- and threedimensional<strong>condensate</strong>s. Also we discuss the energy of a curved vortex l<strong>in</strong>e and a nonl<strong>in</strong>eartilt<strong>in</strong>g of a vortex <strong>in</strong> slightly anisotropic <strong>condensate</strong>s. In section 6 we analyse the effectof thermal quasiparticles on the vortex normal modes and discuss possible mechanisms ofvortex dissipation. Also we discuss the <strong>in</strong>fluence of vortex generation on energy dissipation <strong>in</strong>superfluids. In section 7 we consider vortices <strong>in</strong> multicomponent <strong>condensate</strong>s and analysevarious sp<strong>in</strong>-gauge effects. In particular, we focus on the successful method of vortexgeneration <strong>in</strong> a two-component system that was recently used by the JILA group to createa vortex. In section 8 we draw our conclusions and discuss perspectives <strong>in</strong> the field.2. The time-dependent Gross–Pitaevskii equationBogoliubov’s sem<strong>in</strong>al treatment [38] of a uniform Bose gas at zero temperature emphasizedthe crucial role of (repulsive) <strong>in</strong>teractions both for the structure of the ground state and forthe existence of superfluidity. Subsequently, Gross [39,40] and Pitaevskii [41] <strong>in</strong>dependentlyconsidered an <strong>in</strong>homogeneous <strong>dilute</strong> Bose gas, generaliz<strong>in</strong>g Bogoliubov’s approach to <strong>in</strong>cludethe possibility of nonuniform states, especially quantized vortices.An essential feature of a <strong>dilute</strong> Bose gas at zero temperature is the existence of amacroscopic wave function (an ‘order parameter’) that characterizes the Bose <strong>condensate</strong>.For a uniform system with N particles <strong>in</strong> a stationary box of volume V , the order parameter = √ N 0 /V reflects the presence of a macroscopic number N 0 of particles <strong>in</strong> the zeromomentumstate, with the rema<strong>in</strong><strong>in</strong>g N ′ = N − N 0 particles distributed among the variousexcited states with k ≠ 0. The s<strong>in</strong>gle-particle states for periodic boundary conditions areplane waves V −1/2 e ik·r labelled with the wave vector k, and the correspond<strong>in</strong>g creationand annihilation operators a † k and a k obey the usual Bose–E<strong>in</strong>ste<strong>in</strong> commutation relations


R138A L Fetter and A A Svidz<strong>in</strong>sky[a k ,a † k ′] = δ k,k ′. In the presence of a uniform Bose <strong>condensate</strong> with k = 0, the ground-stateexpectation value 〈a † 0 a 0〉 0 = N 0 is macroscopic, whereas the ground-state expectation valueof the commutator of these zero-mode operators 〈[a 0 ,a † 0 ]〉 0 necessarily equals 1. Hence thecommutator is of order 1/ √ N 0 relative to each separate operator, and they can be approximatedby classical numbers a 0 ≈ a † 0 ≈ √ N 0 . This ‘Bogoliubov’ approximation identifies such aclassical field as the order parameter for the stationary uniform <strong>condensate</strong>. In contrast, theground-state expectation value for all of the other normal modes 〈a † k a k〉 0 is of order unity, andthe associated operators a † k and a k require a full quantum-mechanical treatment.The existence of nonuniform states of a <strong>dilute</strong> Bose gas can be understood by consider<strong>in</strong>ga second-quantized Hamiltonian∫ [H ˆ = dV ˆψ † (T + V tr ) ˆψ + 1 2 g ˆψ † ˆψ † ˆψ ˆψ](1)expressed <strong>in</strong> terms of Bose field operators ˆψ(r) and ˆψ † (r) that obey the Bose–E<strong>in</strong>ste<strong>in</strong>commutation relations[ ˆψ(r), ˆψ † (r ′ )] = δ(r − r ′ ) [ ˆψ(r), ˆψ(r ′ )] = [ ˆψ † (r), ˆψ † (r ′ )] = 0. (2)Here T =−¯h 2 ∇ 2 /2M is the k<strong>in</strong>etic energy operator for the particles of mass M, V tr (r) isan external (trap) potential, and the <strong>in</strong>terparticle potential has been approximated by a shortrange<strong>in</strong>teraction ≈gδ(r − r ′ ), where g is a coupl<strong>in</strong>g constant with the dimensions of energy× volume. For a <strong>dilute</strong> cold gas, only b<strong>in</strong>ary collisions at low energy are relevant, and thesecollisions are characterized by a s<strong>in</strong>gle parameter, the s-wave scatter<strong>in</strong>g length a, <strong>in</strong>dependentlyof the details of the two-body potential. An analysis of the scatter<strong>in</strong>g by such a potential (see,for example [42, 43]) shows that g ≈ 4πa¯h 2 /M. Determ<strong>in</strong>ations of the scatter<strong>in</strong>g length forthe atomic species used <strong>in</strong> the experiments on Bose condensation give: a = 2.75 nm for23 Na [44], a = 5.77 nm for 87 Rb [45], and a =−1.45 nm for 7 Li [46]. In a uniform bulksystem, a must be positive to prevent an <strong>in</strong>stability lead<strong>in</strong>g to a collapse, but a Bose <strong>condensate</strong><strong>in</strong> an external conf<strong>in</strong><strong>in</strong>g trap can rema<strong>in</strong> stable for a


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R139the chemical potential µ ≈ E 0 (N)−E 0 (N −1) associated with remov<strong>in</strong>g one particle from theground state. Thus the stationary solutions take the form (r,t)= (r)e −iµt/¯h , where (r)obeys the stationary GP equation (frequently identified as a nonl<strong>in</strong>ear Schröd<strong>in</strong>ger equation,although the eigenvalue µ is not the energy per particle)(T + V tr + g|| 2 ) = µ. (5)Apart from very recent work on 85 Rb us<strong>in</strong>g a Feshbach resonance to tune a to large positivevalues [48], essentially all studies of <strong>trapped</strong> atomic gases <strong>in</strong>volve the <strong>dilute</strong> limit ( ¯n|a| 3 ≪ 1,where ¯n is the average density of the gas), so the depletion of the <strong>condensate</strong> is small withN ′ = N − N 0 ∝ √ ¯n|a| 3 N ≪ N. Typically ¯n|a| 3 is always less than 10 −3 . Hence most of theparticles rema<strong>in</strong> <strong>in</strong> the <strong>condensate</strong>, and the difference between the <strong>condensate</strong> number N 0 andthe total number N can usually be neglected. In this case, the stationary GP equation (5) forthe <strong>condensate</strong> wave function follows on m<strong>in</strong>imiz<strong>in</strong>g the Hamiltonian functional∫H = dV [ ∗ (T + V tr ) + 1 2 g||4] (6)subject to a constra<strong>in</strong>t of fixed <strong>condensate</strong> number N 0 = ∫ dV || 2 ≈ N (readily <strong>in</strong>cludedwith a Lagrange multiplier that is simply the chemical potential µ).2.1. Unbounded <strong>condensate</strong>The nonl<strong>in</strong>ear Schröd<strong>in</strong>ger equation (5) conta<strong>in</strong>s a local self-consistent Hartree potential energyV H (r) = g|(r)| 2 aris<strong>in</strong>g from the <strong>in</strong>teraction with the other particles at the same po<strong>in</strong>t. Inan unbounded <strong>condensate</strong> with V tr = 0, the left-hand side of equation (5) <strong>in</strong>volves both thek<strong>in</strong>etic energy T and this repulsive Hartree potential g|| 2 = gn for a uniform medium withbulk density n. On dimensional grounds, the balance between these two terms implies a‘correlation’ or ‘heal<strong>in</strong>g’ length¯h 1ξ = √ = √ . (7)2Mng 8πnaThis length characterizes the distance over which the <strong>condensate</strong> wave function heals backto its bulk value when perturbed locally (for example, at a vortex core, where the densityvanishes).√For a uniform system <strong>in</strong> a box of volume V , the <strong>condensate</strong> wave function is =N0 /V ≈ √ N/V, and equation (6) shows that the ground-state energy E 0 arises solelyfrom the repulsive <strong>in</strong>terparticle energy of the <strong>condensate</strong> E <strong>in</strong>t ≈ 1 2 gN2 /V. The bulk chemicalpotential is then given byµ =( ∂E0∂N)V= gn = 4πa¯h2 nM . (8)The correspond<strong>in</strong>g pressure follows from the thermodynamic relation( ) ∂E0p =− = 1 2∂Vgn2 = E <strong>in</strong>tNV . (9)F<strong>in</strong>ally, the compressibility determ<strong>in</strong>es the bulk speed of sound s:s 2 = 1 ( ) ∂p= gnM ∂n M = µ M = 4πa¯h2 n¯hor, equivalently, s = √ . (10)M 2 2MξEquations (7) and (10) both <strong>in</strong>dicate that a bulk uniform Bose <strong>condensate</strong> requires a repulsive<strong>in</strong>teraction (a > 0), s<strong>in</strong>ce otherwise the heal<strong>in</strong>g length and the speed of sound becomeimag<strong>in</strong>ary.


R140A L Fetter and A A Svidz<strong>in</strong>sky2.2. Quantum-hydrodynamic description of the <strong>condensate</strong>It is often <strong>in</strong>structive to represent the <strong>condensate</strong> wave function <strong>in</strong> an equivalent ‘quantumhydrodynamic’form:(r,t)=|(r,t)|e iS(r,t) (11)with the <strong>condensate</strong> densityn(r,t)=|(r,t)| 2 . (12)The correspond<strong>in</strong>g current density j = (¯h/2Mi)[ ∗ ∇ − (∇ ∗ )] automatically assumesa hydrodynamic formj(r,t)= n(r,t)v(r,t) (13)with an irrotational flow velocityv(r,t)= ∇(r,t) (14)expressed <strong>in</strong> terms of a velocity potential(r,t)= ¯hS(r,t)M . (15)Substitute equation (11) <strong>in</strong>to the time-dependent GP equation (4). The imag<strong>in</strong>ary partyields the familiar cont<strong>in</strong>uity equation for compressible flow∂n+ ∇·(nv) = 0. (16)∂tCorrespond<strong>in</strong>gly, the real part constitutes the analogue of the Bernoulli equation for this<strong>condensate</strong> fluid:12 Mv2 + V tr + √ 1 T √ n + gn + M ∂ = 0. (17)n ∂tTo <strong>in</strong>terpret this equation, note that the assumption of a zero-temperature <strong>condensate</strong>implies vanish<strong>in</strong>g entropy; furthermore, the conventional Bernoulli equation for irrotationalcompressible isentropic flow can be rewritten as [49, 50]12 Mv2 + U + e + p + M ∂ = 0 (18)n ∂twhere U is the external potential energy, e is the energy density, and e + p is the enthalpydensity. Comparison with equations (6) and (9) shows that equation (17) for the <strong>condensate</strong>dynamics <strong>in</strong>deed <strong>in</strong>corporates the appropriate constitutive relations for the enthalpy per particle(e + p)/n = ( √ n) −1 T √ n + gn.As a result, the hydrodynamic form of the time-dependent Gross–Pitaevskii equation <strong>in</strong>equations (16) and (17) necessarily reproduces all the standard hydrodynamic behaviour foundfor classical irrotational compressible isentropic flow. In particular, the dynamics of vortexl<strong>in</strong>es at zero temperature follows from the Kelv<strong>in</strong> circulation theorem [49, 50], namely thateach element of the vortex core moves with the local translational velocity <strong>in</strong>duced by all ofthe sources <strong>in</strong> the fluid (self-<strong>in</strong>duced motion for a curved vortex, other vortices, and net appliedflow). The only explicitly quantum-mechanical feature <strong>in</strong> equation (17) is the ‘quantum k<strong>in</strong>eticpressure’ ( √ n) −1 T √ n; as seen from equation (7), this contribution determ<strong>in</strong>es the heal<strong>in</strong>glength ξ that will fix the size and structure of the vortex core.In classical hydrodynamics, the flow can be considered <strong>in</strong>compressible when the velocity|v| is small compared to the speed of sound. More generally, classical compressible flowbecomes irreversible when the flow becomes supersonic because of the emission of sound


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R141waves (which are still part of the hydrodynamic formalism). In a <strong>dilute</strong> Bose gas, however,equations (16) and (17) neglect the normal component entirely. As discussed below <strong>in</strong>section 4.2, the system becomes unstable with respect to the emission of quasiparticles oncethe flow speed exceeds the Landau critical velocity (which here is simply the speed of sound).The normal component then plays an essential role and must be <strong>in</strong>cluded <strong>in</strong> addition to the<strong>condensate</strong>. In this sense, a <strong>dilute</strong> Bose gas is <strong>in</strong>tr<strong>in</strong>sically more complicated than a classicalcompressible fluid.2.3. Vortex dynamics <strong>in</strong> two dimensionsV<strong>in</strong>en’s experiment [12] on the dynamics of a long f<strong>in</strong>e wire <strong>in</strong> rotat<strong>in</strong>g superfluid 4 He strik<strong>in</strong>glyconfirmed Onsager’s and Feynman’s theoretical prediction of quantized circulation [10, 11].These remarkable observations stimulated the study of the nonl<strong>in</strong>ear stationary GP equation(5) <strong>in</strong> the absence of a conf<strong>in</strong><strong>in</strong>g potential, build<strong>in</strong>g on an earlier analysis by G<strong>in</strong>zburg andPitaevskii of vortex-like solutions for superfluid 4 He near T λ [51]. Gross and Pitaevskii<strong>in</strong>dependently <strong>in</strong>vestigated stationary two-dimensional solutions of the form (r) = √ nχ(r),where n is the bulk density far from the orig<strong>in</strong>. Specifically, they considered axisymmetricsolutions( )χ(r) = e iφ r⊥f(19)ξwhere (r ⊥ ,φ) are two-dimensional cyl<strong>in</strong>drical polar coord<strong>in</strong>ates, and f → 1 for r ⊥ ≫ ξ.Equations (14) and (15) immediately give the local circulat<strong>in</strong>g flow velocityv = ¯h ˆφMr ⊥(20)which represents circular streaml<strong>in</strong>es with an amplitude that becomes large as r ⊥ → 0.Comparison of equations (10) and (20) shows that the circulat<strong>in</strong>g flow becomes supersonic(v ≈ s) when r ⊥ ≈ ξ.The particular <strong>condensate</strong> wave function (19) describes an <strong>in</strong>f<strong>in</strong>ite straight vortex l<strong>in</strong>e withquantized circulation∮κ = dl · v = h M(21)precisely as suggested by Onsager and Feynman [10, 11]. Stokes’s theorem then yieldsh/M = ∫ dS ·∇× v, with the correspond<strong>in</strong>g localized vorticity∇ × v = h M δ(2) (r ⊥ )ẑ. (22)Hence the velocity field around a vortex <strong>in</strong> a <strong>dilute</strong> Bose <strong>condensate</strong> is irrotational except fora s<strong>in</strong>gularity at the orig<strong>in</strong>.The k<strong>in</strong>etic energy per unit length is given by∫d 2 r ⊥ ∗ (−¯h2 ∇ 22M) = ¯h22M∫d 2 r ⊥ |∇| 2 = ¯h2 n2M∫[ ( ) ]df2d 2 r ⊥ + f 2dr ⊥ r⊥2and the centrifugal barrier <strong>in</strong> the second term forces the amplitude to vanish l<strong>in</strong>early with<strong>in</strong> acore of radius ≈ξ (see figure 1). This core structure ensures that the particle current densityj = nv vanishes and the total k<strong>in</strong>etic energy density rema<strong>in</strong>s f<strong>in</strong>ite as r ⊥ → 0. The presenceof the vortex produces an additional energy E v per unit length, both from the k<strong>in</strong>etic energy ofcirculat<strong>in</strong>g flow and from the local compression of the fluid. Numerical analysis with the GP(23)


R142f(r / ξ)1/A L Fetter and A A Svidz<strong>in</strong>sky0.80.60.40.2r /ξ01 2 3 4 5 6Figure 1. The radial wave function f(r ⊥ /ξ) obta<strong>in</strong>ed by numerical solution of the stationary GPequation for a straight vortex l<strong>in</strong>e.equation [51] yields E v ≈ (π¯h 2 n/M) ln(1.46R/ξ), where R is an outer cut-off; apart fromthe additive numerical constant, this value is simply the <strong>in</strong>tegral of 1 2 Mv2 n.To illustrate that the time-dependent GP equation <strong>in</strong>deed <strong>in</strong>corporates the correct classicalvortex dynamics, consider a state of the form(r,t)= √ ne iq·r χ(r − r 0 )e −iµt/¯h (24)where χ is the previous stationary solution (19) of the GP equation for a quantized vortex, nowshifted to the <strong>in</strong>stantaneous position r 0 (t), and µ is now a modified chemical potential. Thetotal flow velocity is the sum of a uniform velocity v 0 = ¯hq/M and the circulat<strong>in</strong>g flow aroundthe vortex. Substitute this wave function <strong>in</strong>to the time-dependent GP equation (4). S<strong>in</strong>ce χitself obeys the stationary GP equation (5) with chemical potential µ = gn, a straightforwardanalysis shows that µ = 1 2 Mv2 0+gn, where the first term arises from the centre-of-mass motionof the <strong>condensate</strong>. The rema<strong>in</strong><strong>in</strong>g terms yieldi¯h ∂χ(r − r 0)≡−i¯h dr 0·∇χ(r − r 0 ) =−i¯hv 0 ·∇χ(r − r 0 ). (25)∂tdtThis equation shows that dr 0 (t)/dt = v 0 , so the vortex wave function moves rigidly with theapplied flow velocity v 0 , correctly reproduc<strong>in</strong>g classical irrotational hydrodynamics.A similar method applies to the self-<strong>in</strong>duced motion of two well-separated vortices at r 1and r 2 with |r 1 − r 2 |≫ξ; <strong>in</strong> this case,(r,t)= √ nχ(r − r 1 )χ(r − r 2 )e −iµt/¯h (26)represents an approximate solution with µ = ng because there is no net flow velocity at<strong>in</strong>f<strong>in</strong>ity. The density n|f(r − r 1 )| 2 |f(r − r 2 )| 2 is essentially constant except near the twovortex cores, and the phase is the sum S(r − r 1 ) + S(r − r 2 ) of the two azimuthal anglesfor the variable r measured from the local vortex cores. Substitution <strong>in</strong>to the time-dependentGP equation readily shows that each vortex moves with the velocity <strong>in</strong>duced by the other; forexample,dr 1≈ ¯h dt M ∇S(r − r 2) ∣ r=r1. (27)This method also describes the two-dimensional motion of many well-separated l<strong>in</strong>e vortices[52, 53]. The dynamics of the many-vortex case <strong>in</strong> 2D was also studied <strong>in</strong> [54–56].


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R1432.4. Trapped <strong>condensate</strong>The usual condition for a uniform <strong>dilute</strong> gas requires that the <strong>in</strong>terparticle spac<strong>in</strong>g ∼n −1/3be large compared to the scatter<strong>in</strong>g length (n −1/3 ≫ a or na 3 ≪ 1). The situation is morecomplicated <strong>in</strong> the case of a <strong>dilute</strong> <strong>trapped</strong> gas, because of the three-dimensional harmonictrapp<strong>in</strong>g potential V tr = 1 2 M(ω2 x x2 + ωy 2y2 + ωz 2z2 ). The stationary GP equation (5) provides aconvenient approach for study<strong>in</strong>g the structure of the <strong>condensate</strong> <strong>in</strong> such a harmonic conf<strong>in</strong><strong>in</strong>gpotential.For an ideal non<strong>in</strong>teract<strong>in</strong>g gas (with g = 0), the states are the familiar harmonicoscillator wave functions with the characteristic spatial scale set by the oscillator lengthsd j = √¯h/Mω j (j = x, y, and z). In particular, the ground-state wave function can beobta<strong>in</strong>ed by optimiz<strong>in</strong>g the competition between the k<strong>in</strong>etic energy E k<strong>in</strong> = 〈T 〉 and theconf<strong>in</strong><strong>in</strong>g energy E tr =〈V tr 〉, where 〈···〉 = N ∫ −1 dV ∗ ··· denotes the expectationvalue for the state with the <strong>condensate</strong> wave function . The situation is more complicatedfor an <strong>in</strong>teract<strong>in</strong>g system, however, because the additional <strong>in</strong>teraction energy E <strong>in</strong>t =〈 1 2 g||2 〉provides a new dimensionless parameter. The ratio E <strong>in</strong>t /N¯hω 0 serves to quantify the effectof the <strong>in</strong>teractions, where ω 0 = (ω x ω y ω z ) 1/3 is the mean oscillator frequency. It is notdifficult to show that this ratio is of order Na/d 0 for Na/d 0 1 where d 0 = √¯h/Mω 0 isthe mean oscillator length [31, 32, 43], and of order (Na/d 0 ) 2/5 for Na/d 0 ≫ 1. Thus thepresence of the conf<strong>in</strong><strong>in</strong>g trap significantly alters the physics of the problem, for the additionalcharacteristic length d 0 and energy ¯hω 0 now imply the existence of two dist<strong>in</strong>ct regimes of<strong>dilute</strong> <strong>trapped</strong> gases.2.4.1. The near-ideal regime. In the limit Na/d 0 ≪ 1, the <strong>condensate</strong> states are qualitativelysimilar to those of an ideal gas <strong>in</strong> a three-dimensional harmonic trap, with ground-state wavefunction (r) ∝ exp [ − 1 2 (x2 /d 2 x + y2 /d 2 y + z2 /d 2 z )] . The repulsive <strong>in</strong>teractions play only asmall role, and the <strong>condensate</strong> dimensions are comparable with the oscillator lengths d j .2.4.2. The Thomas–Fermi regime. In the opposite limit Na/d 0 ≫ 1, which is relevantto current experiments on <strong>trapped</strong> Bose <strong>condensate</strong>s, the repulsive <strong>in</strong>teractions significantlyexpand the <strong>condensate</strong>, so the k<strong>in</strong>etic energy associated with the density variation becomesnegligible compared to the trap energy and <strong>in</strong>teraction energy. As a result, the k<strong>in</strong>etic energyoperator T can be omitted <strong>in</strong> the stationary GP equation (5), which yields the Thomas–Fermi(TF) parabolic profile for the ground-state density [32]:n(r) ≈| TF (r)| 2 = 1 g [µ − V tr(r)] [µ − V tr (r)](= n(0) 1 − ∑x 2 jj=x,y,zRj2) ( 1 − ∑x 2 jj=x,y,zRj2)(28)where n(0) = µ/g is the central density and (x) denotes the unit positive step function. Theresult<strong>in</strong>g ellipsoidal three-dimensional density is characterized by two physically differenttypes of parameter: (a) the central density n(0) fixed by the chemical potential (note that n(0)plays essentially the same role as the bulk density n does for the uniform <strong>condensate</strong>, whereµ = gn), and (b) the three <strong>condensate</strong> radiiRj 2 = 2µMωj2 . (29)


R144A L Fetter and A A Svidz<strong>in</strong>skyThe normalization <strong>in</strong>tegral ∫ dVn(r) = N yields the important TF relation [32]N = 8π15 n(0)R3 0 = R5 015ad 4 0or, equivalently,R 5 0d 5 0= 15 Nad 0≫ 1 (30)where R 0 = (R x R y R z ) 1/3 is the mean <strong>condensate</strong> radius. This last equality shows that therepulsive <strong>in</strong>teractions expand the mean TF <strong>condensate</strong> radius R 0 proportionally to N 1/5 . TheTF chemical potential becomesµ = 1 2 Mω2 0 R2 0 = 2¯hω 1 R020d02(31)so µ ≫ ¯hω 0 <strong>in</strong> this limit. The correspond<strong>in</strong>g ground-state energy E 0 = 14¯hω 5 0(R0 2/d2 0 )N =57 µN follows immediately from the thermodynamic relation µ = ∂E 0/∂N.The TF limit leads to several important simplifications. For a <strong>trapped</strong> <strong>condensate</strong>, it isnatural to def<strong>in</strong>e the heal<strong>in</strong>g length (7) <strong>in</strong> terms of the central density, with ξ = [8πn(0)a] −1/2 .In the TF limit, this choice implies thatξR 0 = d02 ξor, equivalently, = d 0≪ 1. (32)d 0 R 0Thus the TF limit provides a clear separation of length scales ξ ≪ d 0 ≪ R 0 , and the (small)heal<strong>in</strong>g length ξ characterizes the small vortex core. In contrast, the heal<strong>in</strong>g length (andvortex-core radius) <strong>in</strong> the near-ideal limit are comparable with d 0 and hence with the size ofthe <strong>condensate</strong>.The quantum-hydrodynamic equations also simplify <strong>in</strong> the TF limit, because the quantumk<strong>in</strong>etic pressure <strong>in</strong> equation (17) becomes negligible. For the static TF ground-state densitygiven <strong>in</strong> equation (28), the small perturbations n ′ <strong>in</strong> the density and ′ <strong>in</strong> the velocity potentialcan be comb<strong>in</strong>ed to yield the generalized wave equation [57]M ∂2 n ′∂t 2 = ∇·[(µ − V tr ) ∇n ′] or, equivalently,∂ 2 n ′∂t 2 = ∇·[s 2 (r) ∇n ′] (33)where s 2 (r) = [µ − V tr (r)] /M def<strong>in</strong>es a spatially vary<strong>in</strong>g local speed of sound. Str<strong>in</strong>gari hasused this equation to analyse the low-ly<strong>in</strong>g normal modes of the TF <strong>condensate</strong>, and severalexperimental studies have verified these predictions <strong>in</strong> considerable detail (see, for example,reference [31]).3. Static vortex statesIn the context of rotat<strong>in</strong>g superfluid 4 He, Feynman [11] noted that solid-body rotation withv sb = Ω×r has constant vorticity ∇×v sb = 2Ω. S<strong>in</strong>ce each quantized vortex l<strong>in</strong>e <strong>in</strong> rotat<strong>in</strong>gsuperfluid 4 He has an identical localized vorticity associated with the s<strong>in</strong>gular circulat<strong>in</strong>g flow(22), he argued that a uniform array of vortices can ‘mimic’ solid-body rotation on average,even though the flow is strictly irrotational away from the cores. He then considered thecirculation Ɣ = ∮ C dl · v along a closed contour C enclos<strong>in</strong>g a large number N v of vortices.The quantization of circulation ensures that Ɣ = N v κ, where κ = h/M is the quantum ofcirculation. If the vortex array mimics solid-body rotation, however, the circulation shouldalso be Ɣ = 2A v , where A v is the area enclosed by the contour C. In this way, the arealvortex density <strong>in</strong> a rotat<strong>in</strong>g superfluid becomesn v = N vA v= 2 κ . (34)


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R145Equivalently, the area per vortex is simply 1/n v = κ/2, which decreases with <strong>in</strong>creas<strong>in</strong>grotation speed. Note that equation (34) is directly analogous to the density of vortices (fluxl<strong>in</strong>es) n v = B/ 0 <strong>in</strong> a type-II superconductor, where B is the magnetic flux density and 0 = h/2e is the quantum of magnetic flux <strong>in</strong> SI units (see, for example, reference [58]).3.1. Structure of a s<strong>in</strong>gle <strong>trapped</strong> vortex3.1.1. The axisymmetric trap. Consider an axisymmetric trap with oscillator frequencies ω zand ω ⊥ and axial asymmetry parameter λ ≡ ω z /ω ⊥ (note that λ 1 yields an elongated cigarshaped<strong>condensate</strong>, and λ 1 yields a flattened disc-shaped <strong>condensate</strong>). The conservationof angular momentum L z allows a simple classification of the states of the <strong>condensate</strong>. Forexample, the macroscopic wave function for a s<strong>in</strong>gly quantized vortex located along the z-axistakes the form(r) = e iφ |(r ⊥ ,z)|. (35)The circulat<strong>in</strong>g velocity is identical to equation (20), and the centrifugal energy (compareequation (23)) gives rise to an additional term 1 2 Mv2 = ¯h 2 /2Mr⊥ 2 <strong>in</strong> the GP equation (5). Inpr<strong>in</strong>ciple, a q-fold vortex with ∝ e iqφ also satisfies the GP equation, but the correspond<strong>in</strong>genergy <strong>in</strong>creases like q 2 (compare the discussion below equation (23)); consequently, amultiply quantized vortex is expected to be unstable with respect to the formation of q s<strong>in</strong>glyquantized vortices.For a non<strong>in</strong>teract<strong>in</strong>g gas <strong>in</strong> an axisymmetric trap, the <strong>condensate</strong> wave function for a s<strong>in</strong>glyquantized vortex on the symmetry axis <strong>in</strong>volves the first excited radial harmonic oscillator statewith the non<strong>in</strong>teract<strong>in</strong>g <strong>condensate</strong> vortex wave function[(r) ∝ e iφ r ⊥ exp − 1 ( r2)]⊥2 d⊥2 + z2(36)dz2of the anticipated form (35). The <strong>in</strong>clusion of <strong>in</strong>teractions for a s<strong>in</strong>gly quantized vortex <strong>in</strong> smallto-mediumaxisymmetric <strong>condensate</strong>s with Na/d 0 1 requires numerical analysis [47, 59].Some phases of rotat<strong>in</strong>g BEC <strong>in</strong> a spherically symmetric harmonic well <strong>in</strong> the near-ideal-gaslimit (ξ d 0 ) were considered by Wilk<strong>in</strong> and Gunn [60]. By exact calculation of wavefunctions and energies for small number of particles, they show that the ground state <strong>in</strong> arotat<strong>in</strong>g trap is rem<strong>in</strong>iscent of those found <strong>in</strong> the fractional quantum Hall effect. These states<strong>in</strong>clude ‘<strong>condensate</strong>s’ of composite bosons of the atoms attached to an <strong>in</strong>teger number of quantaof angular momenta, as well as the Laughl<strong>in</strong> and Pfaffian [61] states. In addition, low-ly<strong>in</strong>gstates with a given angular momentum L z (analogous to the ‘yrast’ states <strong>in</strong> nuclear physics)have been studied <strong>in</strong> references [62, 63].In general, the density for a central vortex vanishes along the symmetry axis, and the coreradius <strong>in</strong>creases away from the centre of the trap, yield<strong>in</strong>g a toroidal <strong>condensate</strong> density (seefigure 2). This behaviour is particularly evident for a vortex <strong>in</strong> the TF limit Na/d 0 ≫ 1, whenn(r ⊥ ,z)≈ n(0)(1 − ξ 2)− r2 ⊥− z2(1 − ξ 2)− r2 ⊥− z2. (37)r 2 ⊥R 2 ⊥Here, the density differs from equation (28) for an axisymmetric vortex-free TF <strong>condensate</strong>only because of the dimensionless centrifugal barrier ξ 2 /r⊥ 2 . This term forces the density tovanish with<strong>in</strong> a core whose characteristic radius is ξ <strong>in</strong> the equatorial region |z| ≪R z and thenflares out with <strong>in</strong>creas<strong>in</strong>g |z|. The TF separation of length scales ensures that the vortex affectsthe density only <strong>in</strong> the immediate vic<strong>in</strong>ity of the core [47, 64, 65]; this behaviour can usuallybe approximated with a short-distance cut-off. For such a quantized TF vortex, the chemicalpotential µ 1 differs from µ 0 for a vortex-free TF <strong>condensate</strong> by small fractional corrections oforder (d 0 /R 0 ) 4 ln(R 0 /d 0 ).R 2 zr 2 ⊥R 2 ⊥R 2 z


R146A L Fetter and A A Svidz<strong>in</strong>skyFigure 2. A contour plot <strong>in</strong> the xz-plane for a <strong>condensate</strong> with 10 4 87 Rb atoms conta<strong>in</strong><strong>in</strong>g avortex along the z-axis. The trap is spherical and distances are <strong>in</strong> units of the oscillator lengthd = 0.791 µm. The <strong>in</strong>teraction parameter is Na/d = 72.3. Lum<strong>in</strong>osity is proportional to density,the white area be<strong>in</strong>g the most dense. (Taken from reference [31].)3.1.2. The nonaxisymmetric trap. If a s<strong>in</strong>gly quantized vortex is oriented along the z-axis ofa nonaxisymmetric trap (R x ≠ R y ) the <strong>condensate</strong> wave function is no longer an eigenfunctionof the angular momentum operator L z . In the TF limit near the trap centre the phase S of the<strong>condensate</strong> wave function has the form [66]S ≈ φ − 1 ( 1− 1 ) ( )r 24 Rx2 Ry2 ⊥ ln r⊥s<strong>in</strong>(2φ) (38)R ⊥and the <strong>condensate</strong> velocity is{v ≈ ¯h ˆφ− 1 ( 1− 1 M r ⊥ 2 Rx2 Ry2) ( )r⊥[ ] }r ⊥ ln cos(2φ) ˆφ + s<strong>in</strong>(2φ)ˆr ⊥ (39)R ⊥where R⊥ 2 = 2R2 x R2 y /(R2 x + R2 y ). Near the vortex core the <strong>condensate</strong> wave function and the<strong>condensate</strong> velocity possess cyl<strong>in</strong>drical symmetry, while far from the vortex core the <strong>condensate</strong>velocity adjusts to the anisotropy of the trap and becomes asymmetric.3.2. Thermodynamic critical angular velocity for vortex stabilityIf the <strong>condensate</strong> is <strong>in</strong> rotational equilibrium at an angular velocity around the ẑ-axis, the<strong>in</strong>tegrand of the GP Hamiltonian (6) acquires an additional term − ∗ L z [67], whereL z = xp y − yp x =−i¯h(x ∂ y − y∂ x ) is the z-component of the angular momentum operator.Thus the Hamiltonian H ′ <strong>in</strong> the rotat<strong>in</strong>g frame becomes∫H ′ = H − L z = dV [ ∗ (T + V tr − L z ) + 1 2 g||4] (40)where the variables <strong>in</strong> the <strong>in</strong>tegrand are now those <strong>in</strong> the rotat<strong>in</strong>g frame. Similarly, the GPequations (4) and (5) acquire an additional term −L z .3.2.1. The axisymmetric trap. The situation is especially simple for an axisymmetric trap,where the states can be labelled by the eigenvalues of L z . For example, the energy of a vortexfree<strong>condensate</strong> E ′ 0 () <strong>in</strong> the rotat<strong>in</strong>g frame is numerically equal to the energy E 0 <strong>in</strong> the


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R147laboratory frame because the correspond<strong>in</strong>g angular momentum vanishes. A s<strong>in</strong>gly quantizedvortex along the trap axis has the total angular momentum N¯h, so the correspond<strong>in</strong>g energyof the system <strong>in</strong> the rotat<strong>in</strong>g frame is E1 ′ () = E 1 − N¯h. The difference between these twoenergies is the <strong>in</strong>creased energyE ′ () = E 1 ′ () − E′ 0 () = E 1 − E 0 − N¯h (41)associated with the formation of the vortex at an angular velocity . In the laboratory frame( = 0), it is clear that E 1 >E 0 because of the added k<strong>in</strong>etic energy of the circulat<strong>in</strong>g flow.If the <strong>condensate</strong> is <strong>in</strong> equilibrium <strong>in</strong> the rotat<strong>in</strong>g frame, however, E1 ′ () decreases l<strong>in</strong>earlywith <strong>in</strong>creas<strong>in</strong>g , and the relative energy of the vortex vanishes at a ‘thermodynamic’ criticalangular velocity c determ<strong>in</strong>ed by E ′ ( c ) = 0. Equation (41) immediately yields c = E 1 − E 0(42)N¯hexpressed solely <strong>in</strong> terms of the energy of a <strong>condensate</strong> with and without the vortex evaluated<strong>in</strong> the laboratory frame.For a non<strong>in</strong>teract<strong>in</strong>g <strong>trapped</strong> gas, the difference E 1 − E 0 = N¯hω ⊥ follows immediatelyfrom the excitation energy for the s<strong>in</strong>gly quantized vortex <strong>in</strong> equation (36) relative to thestationary ground state. In this non<strong>in</strong>teract<strong>in</strong>g case, equation (42) gives c = ω ⊥ ,sothe non<strong>in</strong>teract<strong>in</strong>g thermodynamic critical angular velocity is just the radial trap frequency.Indeed, the same critical angular velocity value also applies to a q-fold vortex <strong>in</strong> a non<strong>in</strong>teract<strong>in</strong>g<strong>condensate</strong>, because of the special form of the non<strong>in</strong>teract<strong>in</strong>g excitation energyE q − E 0 = Nq¯hω ⊥ and the correspond<strong>in</strong>g angular momentum Nq¯h. Thus the non<strong>in</strong>teract<strong>in</strong>g<strong>condensate</strong> becomes massively degenerate as → ω ⊥ [68, 69]. Physically, this degeneracyreflects the cancellation between the centrifugal potential − 1 2 M2 r⊥ 2 and the radial trappotential 1 2 Mω2 ⊥ r2 ⊥ as → ω ⊥.Numerical analysis [47] for small and medium values of Na/d 0 shows that c /ω ⊥decreases with <strong>in</strong>creas<strong>in</strong>g N, and a perturbation analysis [69, 70] confirms this behaviourfor a weakly <strong>in</strong>teract<strong>in</strong>g system, with the analytical result c /ω ⊥ ≈ 1 − [1/(2 √ 2π)](Na/d z )for small values of the <strong>in</strong>teraction parameter Na/d z . Figure 3 shows the behaviour of c (N) <strong>in</strong>a spherical trap, derived from numerical analysis of the GP equation with parameters relevantfor 87 Rb [47].In the strongly <strong>in</strong>teract<strong>in</strong>g (TF) limit, the chemical potential µ 1 (N) for a <strong>condensate</strong>conta<strong>in</strong><strong>in</strong>g a s<strong>in</strong>gly quantized vortex can be evaluated with equation (37), and the thermodynamicidentity µ 1 = ∂E 1 /∂N then yields E 1 (N). Use of the correspond<strong>in</strong>g expressions forthe vortex-free <strong>condensate</strong> gives the approximate expression [64, 71, 72] c ≈ 5 ¯h 2 ( ) 0.67R⊥2 MR⊥2 lnfor a TF <strong>condensate</strong>. (43)ξThis expression exceeds the usual estimate [14] c ≈ (¯h/MR⊥ 2 ) ln(1.46R ⊥/ξ) for uniformsuperfluid <strong>in</strong> a rotat<strong>in</strong>g cyl<strong>in</strong>der of radius R ⊥ because the nonuniform density <strong>in</strong> the <strong>trapped</strong>gas reduces the total angular momentum relative to that for a uniform fluid. Equation (43) hasthe equivalent form c≈ 5 d 2 ( )⊥ 0.67R⊥ln . (44)ω ⊥ 2 ξR 2 ⊥This ratio is small <strong>in</strong> the TF limit, because d 2 ⊥ /R2 ⊥ ∼ ξ/R ⊥ ≪ 1. For an axisymmetric<strong>condensate</strong> with axial asymmetry λ ≡ ω z /ω ⊥ , the TF relation d 2 ⊥ /R2 ⊥ = (d ⊥/15Naλ) 2/5shows how this ratio scales with N and λ.


R148A L Fetter and A A Svidz<strong>in</strong>sky____ Ω cω ⊥NFigure 3. Thermodynamic critical angular velocity c for the formation of a s<strong>in</strong>gly quantizedvortex <strong>in</strong> a spherical trap with d 0 = 0.791 µm and N atoms of 87 Rb. (Taken from reference [31].)In contrast to the case for repulsive <strong>in</strong>teractions, the thermodynamic critical angularvelocity c for the vortex state with attractive <strong>in</strong>teractions <strong>in</strong>creases as the number of atomsgrows [47,73]. S<strong>in</strong>ce c = ω ⊥ for a non<strong>in</strong>teract<strong>in</strong>g <strong>condensate</strong>, c for a vortex <strong>in</strong> a <strong>condensate</strong>with attractive <strong>in</strong>teractions necessarily exceeds ω ⊥ . The stability or metastability of such avortex is unclear because = ω ⊥ is also the limit of mechanical stability for a non<strong>in</strong>teract<strong>in</strong>g<strong>condensate</strong>.Approximately the same functional relationship holds between the thermodynamic criticalfrequency c and the number of atoms <strong>in</strong> the <strong>condensate</strong> N 0 [74] for nonzero temperatures.A new feature, however, is that the number of atoms <strong>in</strong> the <strong>condensate</strong> becomes temperaturedependent:( )N 0 T 3N = 1 − (45)T cwhere T c is the critical temperature of Bose condensation. If the trap rotates at an angularvelocity , the distribution function of the thermal atoms changes due to the centrifugal force.As a result the critical temperature decreases accord<strong>in</strong>g to [74]T c ()T 0 c) 1/3=(1 − 2(46)ω 2 ⊥where Tc0 is the critical temperature <strong>in</strong> the absence of rotation. Equations (44)–(46) allowone to calculate the critical temperature T v (), below which the vortex corresponds to a stableconfiguration <strong>in</strong> a trap rotat<strong>in</strong>g with frequency . In figure 4 we show the critical curves T c ()and T v (). For temperatures below T c () the gas exhibits Bose–E<strong>in</strong>ste<strong>in</strong> condensation. Onlyfor temperatures below T v () does the vortex state become thermodynamically stable. Fromfigure 4 one can see that the critical temperature for the creation of stable vortices exhibits amaximum as a function of .3.2.2. The nonaxisymmetric trap. A rotat<strong>in</strong>g nonaxisymmetric trap <strong>in</strong>troduces significantnew physics, because the mov<strong>in</strong>g walls <strong>in</strong>duce an irrotational flow velocity even <strong>in</strong> the absence


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R149Ω/ω ⊥Figure 4. The phase diagram for vortices <strong>in</strong> a harmonically <strong>trapped</strong> Bose gas, N = 10 4 ,a/d ⊥ = 7.36 × 10 −3 and λ = 1. (Taken from reference [74].)of a vortex [49,75–78]. In the simplest case of a classical uniform fluid <strong>in</strong> a rotat<strong>in</strong>g ellipticalcyl<strong>in</strong>der, the <strong>in</strong>stantaneous <strong>in</strong>duced velocity potential <strong>in</strong> the laboratory frame is [49, 75] cl = xy A2 − B 2(47)A 2 + B 2where A and B are the semi-axes of the elliptical cyl<strong>in</strong>der. The <strong>in</strong>duced angular momentumand k<strong>in</strong>etic energy are reduced from the usual solid-body values by the factor I 0 /I sb =[(A 2 − B 2 )/(A 2 + B 2 )] 2 . In the extreme case B ≪ A, the moment of <strong>in</strong>ertia can approach thesolid-body value, even though the flow is everywhere irrotational.The thermodynamic critical angular velocity c for vortex creation <strong>in</strong> the same uniformclassical fluid depends on the asymmetry ratio B/A [76], and experiments on superfluid4 He confirm the theoretical predictions <strong>in</strong> considerable detail [79]. In the limit B ≪ A,a detailed calculation shows that c ≈ (¯h/2MB 2 ) ln(B/ξ); the appearance of B here isreadily understood from Feynman’s picture of a vortex occupy<strong>in</strong>g an area ≈h/2M (compareequation (34)) and hence hav<strong>in</strong>g to fit the area πB 2 fixed by the smaller lateral dimension B.The preced<strong>in</strong>g analysis for an axisymmetric <strong>dilute</strong> <strong>trapped</strong> Bose gas can be generalizedto treat the TF limit <strong>in</strong> a totally anisotropic disc-shaped harmonic trap with ωx 2 + ω2 y ≪ ω2 z ,start<strong>in</strong>g from equation (40) for the Hamiltonian <strong>in</strong> the rotat<strong>in</strong>g frame [80]. The presence ofa vortex leaves the TF <strong>condensate</strong> density essentially unchanged, and this Hamiltonian canserve as an energy functional to determ<strong>in</strong>e the phase S and hence the superfluid motion of the<strong>condensate</strong>. S<strong>in</strong>ce R x , R y ≫ R z , the curvature of the vortex is negligible. Hence we considera s<strong>in</strong>gly quantized straight vortex displaced laterally from the centre of the rotat<strong>in</strong>g trap to atransverse position r 0 = (x 0 ,y 0 ) that serves as a new orig<strong>in</strong> of coord<strong>in</strong>ates. The <strong>condensate</strong>wave function then has the form =||e iφ+iS 0(48)where φ is the polar angle around the vortex axis and S 0 is a periodic function of φ. Vary<strong>in</strong>gthe Hamiltonian gives an Euler–Lagrange equation for S 0 , and it can be well approximated


R150A L Fetter and A A Svidz<strong>in</strong>skyby the solution for a vortex-free <strong>condensate</strong>, which is M/¯h times the classical expression (47)with A and B replaced by the TF radii R x and R y given <strong>in</strong> equation (29), and with x and yshifted to the new orig<strong>in</strong>.As <strong>in</strong> equation (41) for an axisymmetric trap, E ′ (x 0 ,y 0 ,)gives the <strong>in</strong>creased energy<strong>in</strong> the rotat<strong>in</strong>g frame associated with the presence of the straight vortex. A detailed calculationwith logarithmic accuracy yields [80]E ′ (x 0 ,y 0 ,)= 8π ( )]3 µR zξ 2 n(0)(1 − ζ0 [ln2 R⊥)3/2 − 8 µξ 5 ¯h(ωx 2 + ω2 y )(1 − ζ 0 2 ) (49)where ζ0 2 ≡ x2 0 /R2 x + y2 0 /R2 y 1 is a dimensionless displacement of the vortex from the trapcentre. Here, the mean transverse <strong>condensate</strong> radius R ⊥ is given by the arithmetic mean of the<strong>in</strong>verse squared radii:1R⊥2 = 1 ( 1+ 1 )= M(ω2 x + ω2 y ). (50)2 Rx2 Ry2 4µFigure 5 shows the behaviour of E ′ (ζ 0 ,)as a function of ζ 0 for various fixed valuesof . Curve (a) for = 0 shows that the correspond<strong>in</strong>g energy E ′ (ζ 0 , = 0) decreasesmonotonically with <strong>in</strong>creas<strong>in</strong>g ζ 0 , with negative curvature at ζ 0 = 0. In the absence ofdissipation, energy is conserved and the vortex follows an elliptical trajectory at fixed ζ 0around the centre of the trap along a l<strong>in</strong>e V tr = constant. At low but f<strong>in</strong>ite temperature,1.00.8∆E’(ζ 0 ,Ω)0.6a0.4b0.2c0.0−0.2d−0.40.0 0.2 0.4 0.6 0.8 1.0ζ 0Figure 5. Energy (49) (<strong>in</strong> units of E ′ (0, 0)) associated with a s<strong>in</strong>gly quantized straight vortex <strong>in</strong>a rotat<strong>in</strong>g asymmetric trap <strong>in</strong> the TF limit as a function of a fractional vortex displacement ζ 0 fromthe symmetry axis. Different curves represent different fixed values of the external angular velocity: (a) = 0 (unstable); (b) = m (given <strong>in</strong> equation (51)) (the onset of metastability at theorig<strong>in</strong>); (c) = c (given <strong>in</strong> equation (52)) (the onset of stability at the orig<strong>in</strong>); (d) = 3 2 c,where the th<strong>in</strong> barrier h<strong>in</strong>ders vortex tunnell<strong>in</strong>g from the surface.


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R151however, the vortex experiences weak dissipation; thus it slowly reduces its energy by mov<strong>in</strong>goutward along curve (a), execut<strong>in</strong>g a spiral trajectory <strong>in</strong> the xy-plane.With <strong>in</strong>creas<strong>in</strong>g fixed rotation speed , the function E ′ (ζ 0 ,)flattens. Curve (b) showsthe special case of zero curvature at ζ 0 = 0. It corresponds to the rotation speed m = 3 ( )¯h R⊥2 MR⊥2 lnfor a disc-shaped <strong>condensate</strong> (51)ξat which angular velocity a central vortex first becomes metastable <strong>in</strong> a large disc-shaped<strong>condensate</strong>. For < m , the negative local curvature at ζ 0 = 0 means that weak dissipationimpels the vortex to move away from the centre. For > m , however, the positive localcurvature means that weak dissipation now impels the vortex to move back toward the centre ofthe trap. In this regime, the central position is locally stable; it is not globally stable, however,because E ′ (0,)is positive for ≈ m .Curve (c) shows that E ′ (0, c ) vanishes at the thermodynamic critical angular velocity c = 5 ( )¯h R⊥2 MR⊥2 ln = 5 ξ 3 m for a disc-shaped <strong>condensate</strong>. (52)As expected, this expression (52) reduces to equation (43) <strong>in</strong> the limit of an axisymmetric discshaped<strong>condensate</strong>. For > c , the central vortex is both locally and globally stable relative tothe vortex-free state, and the energy barrier near the outer surface of the <strong>condensate</strong> becomesprogressively narrower. Curve (d) illustrates this behaviour for = 3 2 c. Eventually, thebarrier thickness becomes comparable with the thickness of the boundary layer with<strong>in</strong> whichthe TF approximation fails [81], and it has been suggested that a vortex might then nucleatespontaneously through a surface <strong>in</strong>stability [78, 82, 83]. For a two-dimensional <strong>condensate</strong>, aphase diagram for different critical velocities of trap rotation versus the system parameter an z(n z is the area density) is given <strong>in</strong> [83].3.3. Experimental creation of a s<strong>in</strong>gle vortexThe first experimental detection of a vortex <strong>in</strong>volved a nearly spherical 87 Rb TF <strong>condensate</strong>conta<strong>in</strong><strong>in</strong>g two different <strong>in</strong>ternal (hyperf<strong>in</strong>e) components [33] that tend to separate <strong>in</strong>toimmiscible phases. The JILA group <strong>in</strong> Boulder created the vortex through a somewhat <strong>in</strong>tricatecoherent process that controlled the <strong>in</strong>terconversion between the two components (discussedbelow <strong>in</strong> section 7). In essence, the coupled two-component system acts like an SU(2) sp<strong>in</strong>- 1 2system whose topology differs from the usual U(1) complex one-component order parameter familiar from superfluid 4 He (and conventional BCS superconductivity). Apart from themagnitude || that is fixed by the temperature <strong>in</strong> a uniform system, a one-component orderparameter has only a phase that varies between 0 and 2π. This topology is that of a circleand yields quantized vorticity to ensure that the order parameter is s<strong>in</strong>gle valued [10, 11].In contrast, a two-component system has two degrees of freedom <strong>in</strong> addition to the overallmagnitude; its topology is that of a sphere and does not require quantized vorticity. Thequalitative difference between the two cases can be understood as follows: the s<strong>in</strong>gle degree offreedom of the one-component order parameter is like a rubber band wrapped around a cyl<strong>in</strong>der,while the correspond<strong>in</strong>g two degrees of freedom for the two-component order parameter arelike a rubber band around the equator of a sphere. The former has a given w<strong>in</strong>d<strong>in</strong>g numberthat can be removed only be cutt<strong>in</strong>g it (ensur<strong>in</strong>g the quantization of circulation), whereas thelatter can be removed simply by pull<strong>in</strong>g it to one of the poles (so that there is no quantization).The JILA group was able to sp<strong>in</strong> up the <strong>condensate</strong> by coupl<strong>in</strong>g the two components.They then turned off the coupl<strong>in</strong>g, leav<strong>in</strong>g the system with a residual <strong>trapped</strong> quantized vortexconsist<strong>in</strong>g of one circulat<strong>in</strong>g component surround<strong>in</strong>g a nonrotat<strong>in</strong>g core of the other component,


R152A L Fetter and A A Svidz<strong>in</strong>skywhose size is determ<strong>in</strong>ed by the relative fraction of the two components. By selective tun<strong>in</strong>g,they could image either component nondestructively [37]; figure 6 shows the precession of thefilled vortex core around the trap centre. In addition, an <strong>in</strong>terference procedure allowed themto map the variation of the cos<strong>in</strong>e of the phase around the vortex, clearly show<strong>in</strong>g the expecteds<strong>in</strong>usoidal variation (figure 7).(a)(b)0 ms 50 ms 100 ms 150 ms 200 ms 250 ms 300 ms(b) 45(c)0-45-90-1350 100 200 300time (ms)θ t (degrees)Figure 6. (a) Successive images of a <strong>condensate</strong> with a vortex. The recorded profile of each <strong>trapped</strong><strong>condensate</strong> is fitted with a smooth TF distribution (b). The vortex core is the dark region with<strong>in</strong>the bright <strong>condensate</strong> image. (c) The azimuthal angle of the core is determ<strong>in</strong>ed for each image,and plotted versus time held <strong>in</strong> the trap. A l<strong>in</strong>ear fit to the data gives a precession frequency of1.3(1) Hz. (Taken from reference [37].)Figure 7. The cos<strong>in</strong>e of the phase around the vortex, show<strong>in</strong>g the s<strong>in</strong>usoidal variation expected forthe azimuthal angle. (Taken from reference [33].)The JILA group were also able to remove the component fill<strong>in</strong>g the core, <strong>in</strong> which casethey obta<strong>in</strong>ed a s<strong>in</strong>gle-component vortex [37]. This one-component vortex has a small core sizeand can only be imaged by expand<strong>in</strong>g both the <strong>condensate</strong> and the core, which becomes visiblethrough its reduced density [72, 84]. They first make an image of the two-component vortex,next remove the component fill<strong>in</strong>g the core, and then make an image of the one-componentvortex after a variable time delay. In this way, they can measure the precession rate of the onecomponentempty-core vortex and compare it with theoretical predictions [85]. The data show


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R153no tendency for the core to spiral outward, suggest<strong>in</strong>g that the thermal damp<strong>in</strong>g is negligibleon the timescale of ∼1s.Separately, the ENS group <strong>in</strong> Paris observed the formation of one and more vortices <strong>in</strong> as<strong>in</strong>gle-component 87 Rb elongated cigar-shaped TF <strong>condensate</strong> with a weak nonaxisymmetricdeformation that rotates about its long axis [34–36]. In essence, a static cyl<strong>in</strong>drically symmetricmagnetic trap is augmented by a nonaxisymmetric attractive dipole potential createdby a stirr<strong>in</strong>g laser beam. The comb<strong>in</strong>ed potential produces a cigar-shaped harmonic trap witha slightly anisotropic transverse profile. The transverse anisotropy rotates slowly at a rate 200 Hz. In the first experiments [34], the trap was rotated <strong>in</strong> the normal state andthen cooled, with the clear signal of the vortex shown <strong>in</strong> figure 8 (the trap was turned off,allow<strong>in</strong>g the atomic cloud to expand so that the vortex core becomes visible). This order wasreversed (cool first, then rotate) <strong>in</strong> a later series of runs [36]. In both cases, the observed criticalangular velocity ∼0.7ω ⊥ for creat<strong>in</strong>g the first (central) vortex was roughly 70% higher than thepredicted thermodynamic value c <strong>in</strong> equation (43). These observations agree qualitativelywith the suggestion that a surface <strong>in</strong>stability might nucleate a vortex [78, 82, 83]. Alternativeexplanations of this discrepancy <strong>in</strong>volve the bend<strong>in</strong>g modes of the vortex (discussed below <strong>in</strong>sections 4.4.4 and 5.4.2).Figure 8. The optical thickness of the expanded clouds <strong>in</strong> the transverse direction show<strong>in</strong>g thedifference between the states (a) without and (b) with a vortex. (Taken from reference [34].)3.4. Vortex arraysUnder appropriate stabilization conditions, such as steady applied rotation, vortices can forma regular array. In a rotat<strong>in</strong>g uniform superfluid, the quantized vortex l<strong>in</strong>es parallel to theaxis of rotation form a lattice. This lattice rotates as a whole around the axis of rotation, thussimulat<strong>in</strong>g rigid rotation [86]. At nonzero temperature, dissipative mutual friction from thenormal component ensures that the array rotates with the same angular velocity as the conta<strong>in</strong>er.Early experiments on rotat<strong>in</strong>g superfluid 4 He [13, 87, 88] provided memorable ‘photographs’of vortex l<strong>in</strong>es and arrays with relatively small numbers of vortices, <strong>in</strong> qualitative agreementwith analytical [89, 90] and numerical [91, 92] predictions. A triangular array is favoured forvortices near the rotation axis of rapidly rotat<strong>in</strong>g vessels of superfluid helium [89]. Vortexlattices also occur <strong>in</strong> the neutron superfluid <strong>in</strong> rotat<strong>in</strong>g neutron stars [16].Even before the recent observation of vortex arrays <strong>in</strong> an elongated rotat<strong>in</strong>g <strong>trapped</strong><strong>condensate</strong> [34,35], several theoretical groups had analysed many of the expected properties. Ina weakly <strong>in</strong>teract<strong>in</strong>g (near-ideal) axisymmetric <strong>condensate</strong>, the thermodynamic critical angularvelocity c for the appearance of the first vortex is already close to the radial trap frequency ω ⊥ ,so the creation of additional vortices <strong>in</strong>volves many states φ m (r ⊥ ) ∝ e imφ r m ⊥ exp(− 1 2 r2 ⊥ /d2 ⊥ )


R154A L Fetter and A A Svidz<strong>in</strong>skywith low energy m¯h(ω ⊥ − ) per particle <strong>in</strong> the rotat<strong>in</strong>g frame. Butts and Rokhsar [69]used a l<strong>in</strong>ear comb<strong>in</strong>ation of these nearly degenerate states as a variational <strong>condensate</strong> wavefunction, m<strong>in</strong>imiz<strong>in</strong>g the total energy <strong>in</strong> the laboratory frame E lab subject to the condition offixed number N of particles and fixed angular momentum l per particle. As expected fromthe theoretical and experimental results for liquid helium, the system undergoes a sequenceof transitions between states that break rotational symmetry. Several of these have p-foldsymmetry where p is a small <strong>in</strong>teger. Each vortex represents a node <strong>in</strong> the <strong>condensate</strong> wavefunction, and their positions can vary with the specified angular momentum. Indeed, as l<strong>in</strong>creases from 0 to 1, the first vortex moves cont<strong>in</strong>uously from the edge of the <strong>condensate</strong> tothe centre. For larger number of vortices, the centrifugal forces tend to flatten and expand the<strong>condensate</strong> <strong>in</strong> the radial direction. In this approach of keep<strong>in</strong>g l fixed, the angular velocityfollows from the relation ¯h = ∂E lab /∂l. Figure 9 shows the angular momentum versus theangular velocity for the first several states. Reference [93] has carried out more detailed studiesof the states for relatively small values of the angular momentum per particle l 2.Figure 9. Dimensionless angular momentum l per particle versus dimensionless angular velocity/ω ⊥ . In the figure, γ = (2/π) 1/2 aN/d z . Black l<strong>in</strong>es show stable states and grey l<strong>in</strong>es showmetastable states. There are no stable or metastable states <strong>in</strong> the forbidden ranges l = 0–1 andl = 1–1.70. The rotational symmetry of each branch is <strong>in</strong>dicated. The total angular momentumdiverges as approaches the maximum angular velocity ω ⊥ . Three-dimensional plots of constantdensity show states with twofold and sixfold symmetry. Repr<strong>in</strong>ted by permission from Nature1999 397 327 ©1999 Macmillan Magaz<strong>in</strong>es Ltd.These analyses work at fixed angular momentum Nl, <strong>in</strong> which case the angular velocity must be determ<strong>in</strong>ed from the result<strong>in</strong>g E lab (l). In contrast, the ENS experiments fix (as doexperiments on superfluid helium) and then measure L z from the splitt<strong>in</strong>g of the quadrupolemodes [36] (see section 4.4.3). The JILA group [94] also uses this technique to detect thepresence of a vortex <strong>in</strong> a nonrotat<strong>in</strong>g <strong>condensate</strong>. The transition from fixed L z to fixed canbe considered a Legendre transformation to the Hamiltonian (40) <strong>in</strong> the rotat<strong>in</strong>g frame. Eventhough it is easier to work at fixed (because there is no constra<strong>in</strong>t of fixed L z /N = l), no


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R155such analysis has yet been carried out <strong>in</strong> the weak-coupl<strong>in</strong>g limit.In the strong-coupl<strong>in</strong>g (TF) limit, Cast<strong>in</strong> and Dum [72] have performed extensivenumerical studies of equilibrium vortex arrays <strong>in</strong> two and three dimensions, based onthe Hamiltonian <strong>in</strong> the rotat<strong>in</strong>g frame (thus work<strong>in</strong>g at fixed ). They also propose an<strong>in</strong>tuitive variational calculation based on a factorization approximation that is very similarto equation (26), apart from a different analytic form of the radial function [52, 53].The nucleation of vortices and the result<strong>in</strong>g structures of vortex arrays <strong>in</strong> zero-temperatureBECs were also <strong>in</strong>vestigated numerically by Feder et al [78]. In their simulations, vorticesare generated by rotat<strong>in</strong>g a three-dimensional, nonaxisymmetric harmonic trap. <strong>Vortices</strong>first appear at a rotation frequency significantly larger than the critical frequency for vortexstabilization. At higher frequencies, the trap geometry strongly <strong>in</strong>fluences the structure of thevortex arrays, but the lattices approach triangular arrays at large vortex densities.The ENS experiments [34, 35] have produced remarkable images of vortex arrays.Figure 10 shows three different arrays with up to 11 vortices (obta<strong>in</strong>ed after an expansionof 27 ms). The <strong>in</strong>itial <strong>condensate</strong> is very elongated (along with the vortices), so the radialexpansion predom<strong>in</strong>ates once the trap is turned off. As a result, the expanded <strong>condensate</strong>acquires a pancake shape similar to that <strong>in</strong> figure 9.Figure 10. Arrays of vortices <strong>in</strong> a Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong> stirred with a laser beam. (Takenfrom reference [35].)4. Bogoliubov equations: stability of small-amplitude perturbationsThis section considers only the behaviour of a <strong>dilute</strong> one-component Bose gas, for whichthe analysis of the eigenfrequencies is particularly direct. In the more general case of two<strong>in</strong>terpenetrat<strong>in</strong>g species, even a uniform system can have imag<strong>in</strong>ary frequencies for sufficientlystrong <strong>in</strong>terspecies repulsion [95, 96]; this dynamical <strong>in</strong>stability signals the onset of phaseseparation.4.1. General features for nonuniform <strong>condensate</strong>The special character of an elementary excitation <strong>in</strong> a <strong>dilute</strong> Bose gas largely arises fromthe role of the Bose <strong>condensate</strong> that acts as a particle reservoir. This situation is especiallyfamiliar <strong>in</strong> the uniform system, where an elementary excitation with wave vector k can arisefrom the <strong>in</strong>teract<strong>in</strong>g ground state 0 either through the creation operator a † kor through theannihilation operator a −k (<strong>in</strong> the thermodynamic limit, these two states a † k 0 and a −k 0 differonly by a normalization factor). The true excited eigenstates are l<strong>in</strong>ear comb<strong>in</strong>ations of the twostates, and the correspond<strong>in</strong>g operator for the Bogoliubov quasiparticle is a weighted l<strong>in</strong>earcomb<strong>in</strong>ation [38, 42, 43]α † k = u ka † k + v ka −k (53)


R156A L Fetter and A A Svidz<strong>in</strong>skywhere u k and v k are the (real) Bogoliubov coherence factors. This l<strong>in</strong>ear transformation (53) iscanonical if the quasiparticle operators also obey Bose–E<strong>in</strong>ste<strong>in</strong> commutation relations, whichreadily yields the conditionu 2 k − v2 k = 1 for all k ≠ 0. (54)More generally, the second-quantized Bose field operator ˆψ <strong>in</strong> equation (2) can be writtenas ˆψ(r) ≈ (r)+ ˆφ(r), where ˆφ is an operator giv<strong>in</strong>g the small deviation from the macroscopic<strong>condensate</strong> wave function . These deviation operators obey the approximate Bose–E<strong>in</strong>ste<strong>in</strong>commutation relations[ [ [ˆφ(r), ˆφ † (r )]′ ≈ δ(r − r ′ ) ˆφ(r), ˆφ(r )]′ = ˆφ † (r), ˆφ † (r )]′ ≈ 0. (55)S<strong>in</strong>ce ˆψ(r) does not conserve particle number, it is convenient to use a grand canonicalensemble, with the new Hamiltonian operator ˆK = Hˆ−µ ˆN <strong>in</strong>stead of the Hamiltonian (1). Tolead<strong>in</strong>g (second) order <strong>in</strong> the small deviations, the perturbation <strong>in</strong> ˆK conta<strong>in</strong>s not only the usual‘diagonal’ terms <strong>in</strong>volv<strong>in</strong>g ˆφ † ˆφ, but also ‘off-diagonal’ terms proportional to ˆφ ˆφ and ˆφ † ˆφ † .Consequently, the result<strong>in</strong>g Heisenberg operators ˆφ and ˆφ † obey coupled l<strong>in</strong>ear equations ofmotion (it is here that the role of the <strong>condensate</strong> is evident, for this coupl<strong>in</strong>g vanishes if vanishes). Pitaevskii [41] developed this approach for the particular case of a vortex l<strong>in</strong>e <strong>in</strong>unbounded <strong>condensate</strong>, and the formalism was subsequently extended to <strong>in</strong>clude a generalnonuniform <strong>condensate</strong> [97, 98].In direct analogy to the Bogoliubov transformation for the uniform system, assume theexistence of a l<strong>in</strong>ear transformation to quasiparticle operators α j and α † jfor a set of normalmodes labelled by j:ˆφ(r,t)= ∑ [ ]′u j (r)α j (t) − vj ∗ (r)α† j (t) (56a)jˆφ † (r,t)= ∑ [ ]′u ∗ j (r)α† j (t) − v j (r)α j (t)(56b)jwhere the primed sum means to omit the <strong>condensate</strong> mode. Here, the quasiparticle operatorsα j and α † k obey Bose–E<strong>in</strong>ste<strong>in</strong> commutation relations [α j ,α † k ] = δ jk and have simple harmonictime dependences α j (t) = α j exp(−iE j t/¯h) and α † j (t) = α† j exp(iE j t/¯h). Comparison withthe equations of motion for ˆφ and ˆφ † shows that the correspond<strong>in</strong>g spatial amplitudes obey aset of coupled l<strong>in</strong>ear ‘Bogoliubov equations’Lu j − g() 2 v j = E j u j(57a)Lv j − g( ∗ ) 2 u j =−E j v j(57b)whereL = T + V tr − µ +2g|| 2 (58)is a Hermitian operator.∫Straightforward manipulations with the Bogoliubov equations show that E j dV(|uj | 2 −|v j | 2 ) is real. If the <strong>in</strong>tegral ∫ dV (|u j | 2 −|v j | 2 ) is nonzero, then E j itself is real. Likeequation (54) for a uniform <strong>condensate</strong>, the Bose–E<strong>in</strong>ste<strong>in</strong> commutation relations (55) forthe deviations from the nonuniform <strong>condensate</strong> can be shown to imply the follow<strong>in</strong>g positivenormalization [97]:∫dV (|u j | 2 −|v j | 2 ) = 1. (59)For each solution u j ,v j with eigenvalue E j and positive normalization, the Bogoliubov equationsalways have a second solution v ∗ j ,u∗ j with eigenvalue −E j and negative normalization.


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R157The only exception to the requirement of real eigenvalues arises for zero-norm solutions with∫dV (|uj | 2 −|v j | 2 ) = 0. In this case the character of the eigenvalue requires additionalanalysis. Numerical <strong>in</strong>vestigations [99] of vortices <strong>in</strong> nonuniform <strong>trapped</strong> <strong>condensate</strong>s havereported imag<strong>in</strong>ary and/or complex eigenfrequencies for doubly quantized vortices but onlyreal eigenfrequencies for s<strong>in</strong>gly quantized vortices. Specifically, for a repulsive <strong>in</strong>terparticle<strong>in</strong>teraction, Pu et al [99] found that s<strong>in</strong>gly quantized vortices are always <strong>in</strong>tr<strong>in</strong>sically stable; <strong>in</strong>contrast, multiply quantized vortices have alternat<strong>in</strong>g stable and unstable regions with complexexcitation energy as the <strong>in</strong>teraction parameter Na/d <strong>in</strong>creases. The most unstable vortexstate decays after several periods of the harmonic trapp<strong>in</strong>g potential. In the case of multiplyquantized vortices (q > 1), the vortex core conta<strong>in</strong>s localized quasiparticle bound stateswith small exponential tails; these modes have complex frequencies and are responsible forsplitt<strong>in</strong>g the multicharged core [100]. For an attractive <strong>in</strong>teraction, stable vortices exist onlyfor the s<strong>in</strong>gly quantized case <strong>in</strong> the weak-<strong>in</strong>teraction regime; a multiply quantized vortexstate is always unstable. Similar imag<strong>in</strong>ary and complex solutions have been found for darksolitons [101–103]. For additional results on complex eigenfrequencies, see reference [104]and the appendix of reference [105].In terms of the quasiparticle operators, the approximate perturbation Hamiltonian operatortakes the simple <strong>in</strong>tuitive formˆK ′ ≈ ∑ j′Ej α † j α j (60)apart from a constant ground-state contribution of all of the normal modes. Here, the sumis over all of the states with positive normalization, and it is clear that the sign of the energyeigenvalues E j is crucial for the stability. If one or more of the eigenvalues is negative, theHamiltonian is no longer positive def<strong>in</strong>ite, and the system can lower its energy by creat<strong>in</strong>gquasiparticles <strong>in</strong> the unstable modes.The present derivation of the Bogoliubov equations and their properties emphasizes thequantum-mechanical basis for the positive normalization condition (59) and the sign of theeigenvalues. It is worth not<strong>in</strong>g an alternative purely ‘classical’ treatment [31, 106] baseddirectly on small perturbations of the time-dependent GP equation (4) around the static<strong>condensate</strong> (r). The solution is assumed to have the form(r,t)= e −iµt/¯h [ (r) + u(r)e −iωt − v ∗ (r)e iωt] (61)and the appropriate eigenvalue equations then reproduce equations (57).4.2. Uniform <strong>condensate</strong>For a uniform <strong>condensate</strong>, the solutions of equations (57) are plane waves, and the correspond<strong>in</strong>genergy is the celebrated Bogoliubov spectrum [38]√E k = gn¯h 2 k 2 /M + (¯h 2 k 2 /2M) 2 (62)where k is the wave vector of the excitation and n is the <strong>condensate</strong> density. For longwavelengths kξ ≪ 1, equation (62) reduces to a l<strong>in</strong>ear phonon spectrum E k ≈ ¯hsk withthe speed of compressional sound s = √ gn/M given by equation (10). In the opposite limitkξ ≫ 1, the spectrum reduces to the free-particle form plus a mean-field Hartree shift fromthe <strong>in</strong>teraction with the background <strong>condensate</strong> E k ≈ (¯h 2 k 2 /2M) + gn.To understand the importance of the sign of the eigenfrequency, it is <strong>in</strong>structive to considerthe case of a <strong>condensate</strong> that moves uniformly with velocity v 0 . As noted <strong>in</strong> connection withequation (24), the <strong>condensate</strong> wave function is (r) = √ ne iq·r , where q = Mv 0 /¯h and the


R158A L Fetter and A A Svidz<strong>in</strong>skychemical potential becomes µ = 1 2 Mv2 0+ gn. The Bogoliubov amplitudes for an excitationwith wave vector k relative to the mov<strong>in</strong>g <strong>condensate</strong> have the form( ) (uk (r) e iq·r u k e ik·r )=(63)v k (r) e −iq·r v k e ik·rwhere the different signs ±iq · r arise from the different phases ±i2q · r <strong>in</strong> the off-diagonalcoupl<strong>in</strong>g terms <strong>in</strong> the Bogoliubov equations (57). The solution with positive norm has theeigenvalueE k (v 0 ) = ¯hk · v 0 + E k (64)as expected from general considerations [107, 108]. In the long-wavelength limit, thisexcitation energy reduces to E k (v 0 ) ≈ ¯hk(v 0 cos θ + s), where θ is the angle between kand v 0 . For v 0 s,the quasiparticle energy becomes negative for certa<strong>in</strong> directions, <strong>in</strong>dicat<strong>in</strong>g the onset of an<strong>in</strong>stability. This behaviour simply reflects the well-known Landau critical velocity for theonset of dissipation, associated with the emission of quasiparticles. It has many analogies withsupersonic flow <strong>in</strong> classical compressible fluids [109] and Cherenkov radiation of photons <strong>in</strong>a dielectric medium [110, 111]. For v 0 >s, the GP description becomes <strong>in</strong>complete becausethe excitation of quasiparticles means that the non<strong>condensate</strong> is no longer negligible.4.3. Quantum-hydrodynamic description of small-amplitude normal modesThe quantum-hydrodynamic forms (16) and (17) of the time-dependent GP equation providea convenient alternative basis for study<strong>in</strong>g the small-amplitude normal modes. The smallperturbations <strong>in</strong> the density n ′ e −iωt and the velocity potential ′ e −iωt obey coupled l<strong>in</strong>earequations [98,112,113] that reduce to equation (33) <strong>in</strong> the TF limit for a static <strong>condensate</strong> [57].A comparison with equations (56a) shows that the quantum-hydrodynamic amplitudesn ′ j = ∗ u j − v j =||(e −iS u j − e iS v j )(65a) ′ j = ¯h¯h2Mi|| 2 (∗ u j + v j ) =2Mi|| (e−iS u j +e iS v j )(65b)are simply l<strong>in</strong>ear comb<strong>in</strong>ations of the Bogoliubov amplitudes u j and v j <strong>in</strong> the presence of thegiven <strong>condensate</strong> solution = e iS ||. The positive normalization condition (59) yields theequivalent quantum-hydrodynamic form∫dV i(n ′ ∗ j ′ j − ′ ∗ j n ′ j ) = ¯h M . (66)For many purposes, the quantum-hydrodynamic modes provide a clearer picture of thedynamical motion.4.4. A s<strong>in</strong>gly quantized vortex <strong>in</strong> an axisymmetric trapEarly numerical studies for small and medium values of the <strong>in</strong>teraction parameter Na/d 0 1exam<strong>in</strong>ed the small-amplitude excitations of a <strong>condensate</strong> with a s<strong>in</strong>gly quantized vortex [114].In particular, the spectrum conta<strong>in</strong>ed an ‘anomalous’ mode with a negative excitation frequencyand positive normalization associated with a large Bogoliubov amplitude u localized <strong>in</strong> thevortex core (see also relevant comments <strong>in</strong> reference [65] concern<strong>in</strong>g the relationship betweenthe sign of the normalization and the sign of the eigenfrequency). The anomalous modecorresponds to a precession of the vortex l<strong>in</strong>e around the z-axis. As seen from the generaldiscussion of the Bogoliubov equations, this anomalous mode <strong>in</strong>dicates the presence of an<strong>in</strong>stability.


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R159S<strong>in</strong>ce the <strong>condensate</strong> wave function has an explicit phase (r) = e iφ |(r ⊥ ,z)|, theBogoliubov amplitudes for an excitation with angular momentum m¯h relative to the vortex<strong>condensate</strong> take the form(um (r)v m (r)) ( e iφ e imφ ) ũ m (r ⊥ ,z)=e −iφ e imφ . (67)ṽ m (r ⊥ ,z)analogous to those <strong>in</strong> equation (63) for a <strong>condensate</strong> <strong>in</strong> uniform motion. Here, the azimuthalquantum number m characterizes the associated density and velocity deformations of the vortexproportional to e imφ (for example, n ′ m =||(ũ m−ṽ m )e imφ , as is clear from equation (65a)). Thenumerical studies [114] found that the anomalous mode has an azimuthal quantum numberm a =−1. Its frequency ω a is negative throughout the relevant range of Na/d 0 1; <strong>in</strong>the non<strong>in</strong>teract<strong>in</strong>g limit, ω a approaches −ω ⊥ , and ω a <strong>in</strong>creases toward 0 from below with<strong>in</strong>creas<strong>in</strong>g Na/d 0 .To understand the particular value m a =−1, it is helpful to recall the non<strong>in</strong>teract<strong>in</strong>g limit,when the negative anomalous mode for the vortex <strong>condensate</strong> signals the <strong>in</strong>stability associatedwith Bose condensation <strong>in</strong> the first excited harmonic oscillator state with excitation energy¯hω ⊥ and unit angular momentum. A particle <strong>in</strong> the <strong>condensate</strong> can make a transition fromthe vortex state back to the true harmonic oscillator ground state, with a change <strong>in</strong> frequency−ω ⊥ and a change <strong>in</strong> angular momentum quantum number −1. More generally, the densityperturbation n ′ a for the anomalous mode with negative frequency −|ω a| is proportional toexp[i(|ω a |t − φ)] and hence precesses <strong>in</strong> the positive sense (namely anticlockwise) at thefrequency |ω a |. Thus the anomalous mode describes the JILA observations of the precessionfrequency of a one-component vortex [37, 85].4.4.1. The near-ideal regime. An explicit perturbation analysis [70,115] of the GP equationfor the <strong>condensate</strong> wave function <strong>in</strong> the weakly <strong>in</strong>teract<strong>in</strong>g limit found the thermodynamiccritical angular velocity c= 1 − 1( )Na Na 2√ + c(2)ω ⊥ 8π d (λ) + ··· (68)z d zwhere the second-order correction depends explicitly on the axial asymmetry λ = ω z /ω ⊥ .Similarly, a perturbation expansion of the Bogoliubov equations <strong>in</strong> the weak-coupl<strong>in</strong>g limitverified the numerical analysis and found the explicit expression for the frequency of theanomalous mode:ω a=−1+ 1( )Na Na 2√ + ω a(2)ω ⊥ 8π d (λ) + ···. (69)z d zIt is evident that c +ω a vanishes at first order, and the detailed analysis shows that the secondordercontribution to the sum is positive for all values of the axial asymmetry parameter λ.The physics of the anomalous mode can be clarified by consider<strong>in</strong>g an axisymmetric<strong>condensate</strong> <strong>in</strong> rotational equilibrium at an angular velocity around the ẑ-axis. In the rotat<strong>in</strong>gframe, the Hamiltonian becomes H − L z , and the Bogoliubov amplitudes have frequenciesω j () = ω j −m j , where ω j is the frequency <strong>in</strong> the nonrotat<strong>in</strong>g frame and m j is the azimuthalquantum number (see equation (67)). For the anomalous mode with m a =−1, the result<strong>in</strong>gfrequency <strong>in</strong> the rotat<strong>in</strong>g frame isω a () = ω a + (70)which is directly analogous to equation (64) for uniform translation. S<strong>in</strong>ce ω a is negative, theanomalous frequency <strong>in</strong> a rotat<strong>in</strong>g frame <strong>in</strong>creases l<strong>in</strong>early toward zero with <strong>in</strong>creas<strong>in</strong>g ; <strong>in</strong>particular, ω a () vanishes at a characteristic rotation frequency ∗ =−ω a =|ω a | (71)


R160A L Fetter and A A Svidz<strong>in</strong>skythat signals the onset of the regime ∗ for which the s<strong>in</strong>gly quantized vortex becomeslocally stable. Equation (69) gives an explicit expression for ∗ <strong>in</strong> the weak-coupl<strong>in</strong>g limit,and detailed comparison with equation (68) <strong>in</strong>dicates that ∗ < c for any axial asymmetryλ (but only because of the second-order contributions). It is natural to identify ∗ with theangular velocity for the onset of local stability with respect to small perturbations; this quantitywas denoted as m <strong>in</strong> connection with the equilibrium energy <strong>in</strong> the TF limit (see figure 5).4.4.2. The Thomas–Fermi regime for a disc-shaped trap. The anomalous negative-frequencymode exists only because the <strong>condensate</strong> conta<strong>in</strong>s a vortex. Hence it cannot be analysed bytreat<strong>in</strong>g the vortex itself as a perturbation. In the TF limit, however, it is possible to useGross’s and Pitaevskii’s [39, 41] solution (19) for a vortex <strong>in</strong> a laterally unbounded fluid asthe basis for a perturbation expansion. A detailed analysis of the Bogoliubov equations foran axisymmetric rotat<strong>in</strong>g flattened trap <strong>in</strong> the TF limit yields the explicit expression for theanomalous mode [116](ω a () = − 3¯hω2 ⊥4µ ln R⊥= − 3 ¯hξ 2 MR⊥2 lnξAs <strong>in</strong> equation (71) for the weak-coupl<strong>in</strong>g limit, equation (72) yields ∗ = 3 ( )¯h R⊥ln2 ξMR 2 ⊥)(R⊥). (72)= m = 3 5 c (73)where the last two equalities follow from (51) and (52). This relation further supports theidentification of ∗ with the metastable rotation frequency m associated with local stabilityof a vortex for small lateral displacements from the centre of the trap. Note that m < c for adisc-shaped <strong>condensate</strong> (<strong>in</strong> the TF limit) (see equations (51) and (52)), similar to the behaviourfor the weak-coupl<strong>in</strong>g regime.4.4.3. Quantum-hydrodynamic analysis of <strong>condensate</strong> normal modes <strong>in</strong> the Thomas–Fermiregime. In addition to the anomalous mode described above, the <strong>condensate</strong> has a sequence ofnormal modes that occur both with and without a vortex. Indeed, one of the early triumphs of thequantum-hydrodynamic description [57] was the detailed agreement between the theoreticalpredictions and the measured frequencies of the lowest few collective normal modes [31]. Foran axisymmetric <strong>condensate</strong>, the normal modes can be classified by their azimuthal quantumnumber m, and modes with ±m are degenerate for a stationary <strong>condensate</strong>.When the <strong>condensate</strong> conta<strong>in</strong>s a vortex, however, the various collective modes areperturbed. In particular, the vortex breaks time-reversal symmetry by impos<strong>in</strong>g a preferredsense of rotation, so modes with ±m are split (this behaviour is analogous to the Zeeman effect<strong>in</strong> which an applied magnetic field splits the magnetic sublevels). In fact, the splitt<strong>in</strong>g of thesedegenerate hydrodynamic modes has been used to detect the presence of a vortex [36,94] andto <strong>in</strong>fer its circulation and angular momentum.In the context of the quantum-hydrodynamic description, the pr<strong>in</strong>cipal effect of the vortexarises through its circulat<strong>in</strong>g velocity field v, which shifts the time derivative ∂ t → ∂ t + ⃗v · ⃗∇.For a normal mode ∝e imφ with azimuthal quantum number m, the perturbation <strong>in</strong> the frequencyhas the form m¯h/Mr⊥ 2 . A detailed analysis shows that the fractional splitt<strong>in</strong>g of the modes is oforder (ω + − ω − )/ω + ∼|m|d⊥ 2 /R2 ⊥, with a numerical coefficient that depends on the particularmode <strong>in</strong> question [64, 113]. Independently, Zambelli and Str<strong>in</strong>gari [117] used sum rules tocalculate the vortex-<strong>in</strong>duced splitt<strong>in</strong>g of the lowest quadrupole mode with m =±2; the twoapproaches yield precisely the same expressions. In the absence of a vortex, the |m| =2 modesimply <strong>in</strong>volves an oscillat<strong>in</strong>g quadrupole distortion, but the vortex-<strong>in</strong>duced splitt<strong>in</strong>g means


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R161that the quadrupole distortion precesses slowly <strong>in</strong> a sense determ<strong>in</strong>ed by the circulation aroundthe vortex. The angular frequency of precession of the eigenaxes of the quadrupole mode isequal to (ω + − ω − )/2|m| =(ω + − ω − )/4 = 7 4 ω ⊥d⊥ 2 /R2 ⊥. Figure 11 shows the differencebetween the two cases (with and without a vortex) for a <strong>condensate</strong> with ≈3.7 × 10 5 87 Rbatoms <strong>in</strong> an elongated trap with ω ⊥ /2π = 171 Hz. In the ENS experiment [36], when onevortex is nucleated at the centre of the <strong>condensate</strong>, the measured frequency splitt<strong>in</strong>g of thequadrupole mode (ω + /2π = 250 Hz) is (ω + − ω − )/2π = 66(±7) Hz. For the experimentalparameters (R ⊥ = 3.8 µm), theory predicts (ω + − ω − )/2π = 7¯h/2πMR⊥ 2 = 56 Hz. Theresult holds <strong>in</strong> the TF limit and is valid with an accuracy of order d⊥ 2 ln(R ⊥/ξ)/R⊥ 2 ∼ 0.15.With this uncerta<strong>in</strong>ty, the theoretical prediction 56(±8) Hz agrees with the experimental value.a) b) c)d) e) f)Figure 11. Transverse oscillations of a stirred <strong>condensate</strong> with 3.7×10 5 atoms <strong>in</strong> an elongated trapwith ω ⊥ /2π = 171 Hz. For (a)–(c), the stirr<strong>in</strong>g frequency /2π = 114 Hz is below the thresholdfor vortex nucleation, whereas for (d)–(f ), the stirr<strong>in</strong>g frequency /2π = 120 Hz has nucleateda vortex (visible at the centre of the <strong>condensate</strong>). The sequences of pictures correspond to timedelays τ = 1, 3, and 5 ms for which the ellipticity <strong>in</strong> the xy-plane is maximum. The fixed axes<strong>in</strong>dicate the excitation basis of the quadrupole mode and the rotat<strong>in</strong>g ones <strong>in</strong>dicate the <strong>condensate</strong>axes. (Taken from reference [36].)One should note that the vortex-<strong>in</strong>duced splitt<strong>in</strong>g of the <strong>condensate</strong> modes is maximumif the vortex is located at the trap centre. If a straight vortex l<strong>in</strong>e is displaced a distanceζ 0 = r 0 /R ⊥ from the z-axis of the TF <strong>condensate</strong>, then the splitt<strong>in</strong>g of the quadrupole mode(m =±2) is given by the expression(1 − 5 4 ζ 2 0[1+ 1 2 ζ 0 4 − ζ 0 6 + 3 ])10 ζ 08 . (74)d⊥2 ω + − ω − = 7ω ⊥R⊥2The splitt<strong>in</strong>g goes to zero if the vortex moves out of the <strong>condensate</strong> (ζ 0 = 1).4.4.4. Numerical analysis for general <strong>in</strong>teraction parameter. García-Ripoll and Pérez-García [104] have performed extensive numerical analyses of the stability of vortices <strong>in</strong>axisymmetric traps with an axial asymmetry parameter λ = ω z /ω ⊥ = 1 (a sphere) and λ = 1 2(one particular cigar-shaped <strong>condensate</strong>). They conclude that a doubly quantized vortex l<strong>in</strong>ehas normal modes with imag<strong>in</strong>ary frequencies and that an external rotation cannot stabilize it.For a s<strong>in</strong>gly quantized vortex <strong>in</strong> a spherical trap, however, they confirm the presence of onenegative-frequency (anomalous) mode with |ω a | < c . For their cigar-shaped <strong>condensate</strong>,


R162A L Fetter and A A Svidz<strong>in</strong>skythey f<strong>in</strong>d additional negative-frequency modes and suggest that such elongated <strong>condensate</strong>sare less stable than spherical or disk-shaped ones. More recent numerical work [85, 118]confirms these f<strong>in</strong>d<strong>in</strong>gs for other geometries, especially that for the ENS experiment [34],where the axial asymmetry is large (ω ⊥ /ω z ≈ R z /R ⊥ ≈ 14). It is expected that a vortex <strong>in</strong>an elongated <strong>condensate</strong> becomes stable only for an external angular velocity m = max|ω a |,where max|ω a | is the absolute value of the most negative of these anomalous modes. For onlymodestly elongated traps, the metastable frequency m exceeds the thermodynamic criticalvalue c ; these results provide an alternative explanation of the ENS observation that the firstvortex appears at an applied rotation ≈70% higher than c . Independently, an analysis of thebend<strong>in</strong>g modes of a <strong>trapped</strong> vortex [119] <strong>in</strong> the TF limit f<strong>in</strong>ds that a vortex <strong>in</strong> a spherical ordisc-shaped <strong>condensate</strong> has only one negative frequency (anomalous) mode, but the numberof such modes <strong>in</strong> an elongated <strong>condensate</strong> <strong>in</strong>creases with the axial asymmetry ratio R z /R ⊥(discussed below <strong>in</strong> section 5.4.2).5. Vortex dynamicsThe preced<strong>in</strong>g sections considered the equilibrium and stability of a vortex <strong>in</strong> a <strong>trapped</strong> Bose<strong>condensate</strong>, us<strong>in</strong>g the stationary GP equation and the Bogoliubov equations that characterizethe small perturbations of the stationary vortex. These approaches are somewhat <strong>in</strong>direct, forthey do not consider the dynamical motion of the vortex core. The present section treats twodifferent methods that address such questions directly.5.1. Time-dependent variational analysisConsider a variational problem for the action ∫ dt L(t) obta<strong>in</strong>ed from the Lagrangian∫ [ ( i¯hL(t) = dV ∗ ∂ ) ]− ∂∗ − ∗ (T + V tr − L z ) − 1 22 ∂t ∂tg||4 . (75)It is easy to verify that the Euler–Lagrange equation for this action is precisely the timedependentGP equation <strong>in</strong> the rotat<strong>in</strong>g frame.If, <strong>in</strong>stead of (r,t), we substitute a trial function that conta<strong>in</strong>s different variationalparameters (for example, the location of the vortex core), the result<strong>in</strong>g time evolution ofthese parameters characterizes the dynamics of the <strong>condensate</strong>. This method is not exact,but it provides an appeal<strong>in</strong>g physical picture. For example, it determ<strong>in</strong>ed the low-energyexcitations of a <strong>trapped</strong> vortex-free <strong>condensate</strong> at zero temperature [120, 121] for generalvalues of the <strong>in</strong>teraction parameter. In the TF limit, this work reproduced the expressionsderived by Str<strong>in</strong>gari [57] based on equation (33).5.1.1. The near-ideal regime. In the near-ideal limit, only the axisymmetric case has beenstudied, and it is natural to start from the non<strong>in</strong>teract<strong>in</strong>g vortex state (36), <strong>in</strong>corporat<strong>in</strong>g smalllateral displacements of the vortex and the centre of mass of the <strong>condensate</strong>, along with aphase that characterizes the velocity field <strong>in</strong>duced by the motion of the <strong>condensate</strong> [122]. Inaddition to the rigid-dipole mode (<strong>in</strong> which the <strong>condensate</strong> and the vortex oscillate togetherat the transverse trap frequency ω ⊥ ), an extra normal mode arises at the anomalous (negative)frequency ω a given <strong>in</strong> equation (69) omitt<strong>in</strong>g the second-order corrections that are beyond thepresent approximation. In this weak-coupl<strong>in</strong>g limit, the result<strong>in</strong>g displacement of the vortexis twice that of the centre of mass, so both must be <strong>in</strong>cluded to obta<strong>in</strong> the correct dynamicalmotion. Detailed analysis confirms the positive normalization and relative displacements foundfrom the Bogoliubov equations for the same axisymmetric trap [115].


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R1635.1.2. The Thomas–Fermi regime for a straight vortex <strong>in</strong> a disc-shaped trap. For a nonaxisymmetrictrap <strong>in</strong> the TF regime, only the nonrotat<strong>in</strong>g case ( = 0) has been analysed,us<strong>in</strong>g the fully anisotropic TF wave function as an appropriate trial state, aga<strong>in</strong> with parametersdescrib<strong>in</strong>g the small displacements of the straight vortex and the centre of mass of the<strong>condensate</strong> [80]. The trial wave function was chosen <strong>in</strong> the form∏(r,t)= B(t)f [r − r 0 (t)] F [r − η 0 (t)] exp [ ix j α j (t) +ixj 2 β j (t) ] . (76)j=x,y,zHere the function f(r) characterizes the vortex l<strong>in</strong>e <strong>in</strong>side the trap, and far away from thevortex core has the approximate form f(r) = e iφ ; the function F(r) is the TF <strong>condensate</strong>density. The time-dependent vector η 0 (t) = (η 0x ,η 0y ,η 0z ) describes the motion of the centreof the <strong>condensate</strong>, while r 0 (t) = (x 0 ,y 0 , 0) describes the motion of the vortex l<strong>in</strong>e <strong>in</strong> thexy-plane. The other variational parameters are the amplitude B(t) and the set α j (t) and β j (t).Substitution of the trial wave function <strong>in</strong>to (75) yields an effective Lagrangian as a function ofthe variational parameters (and their first time derivatives). The result<strong>in</strong>g Lagrangian equationshave a solution that corresponds to the motion of the vortex relative to the <strong>condensate</strong>. For thissolution the vortex motion is described byx 0 = ε 0 R x s<strong>in</strong>(ω a t + φ 0 ) y 0 = ε 0 R y cos(ω a t + φ 0 ) (77)while the displacement of the <strong>condensate</strong> is given byη 0x =− 15ε 0ξ 22R yη 0y =− 15ε 0ξ 22R xlnln( )R⊥ Rxs<strong>in</strong>(ω a t + φ 0 ) (78)ξ R x + R y)Rycos(ω a t + φ 0 ) (79)ξ R x + R y(R⊥whereω a =− 3¯hω ( )xω y4µ ln R⊥=− 3¯h ( )R⊥ln(80)ξ 2MR x R y ξ<strong>in</strong> agreement with that found <strong>in</strong> equation (72). The quantity x0 2/R2 x + y2 0 /R2 y = ε2 0 rema<strong>in</strong>sconstant as the vortex l<strong>in</strong>e follows an elliptic trajectory around the centre of a trap along thel<strong>in</strong>e V tr = constant, and the energy of the system is conserved (as follows from equation (49)).The <strong>condensate</strong> also precesses with the relative phase shift π at the same frequency, butthe amplitude of the <strong>condensate</strong> motion is smaller than that of the vortex l<strong>in</strong>e by a factor∼ξ 2 ln(R ⊥ /|q|ξ)/R x R y .For an axisymmetric TF <strong>condensate</strong> <strong>in</strong> rotational equilibrium at an angular velocity, the Lagrangian (75) provides a more general result for the precession frequency.With the hydrodynamic variables = e iS ||, the first term of the Lagrangian becomes−¯h ∫ dV || 2 ∂S/∂t. S<strong>in</strong>ce the TF <strong>condensate</strong> density vanishes at the surface, the particlecurrent also vanishes there, and it usually suffices to assume a s<strong>in</strong>gle straight vortex displacedlaterally to r 0 (t), with S(r, r 0 ) = arctan[(y − y 0 )/(x − x 0 )] and no image vortex. TheLagrangian becomes∫L = dV Mn(r)ṙ 0 · v 0 (r) − E(r 0 ) + L z (r 0 ) (81)wherev 0 (r) = ¯h M ∇S(r, r 0) =−¯h M ∇ 0S(r, r 0 ) = (κ/2π)ẑ × (r − r 0)(82)|r − r 0 | 2is the circulat<strong>in</strong>g velocity field about the vortex l<strong>in</strong>e. In the special case of a two-dimensional<strong>condensate</strong> with the TF density n(r) = n(0)(1−r⊥ 2 /R2 ⊥) per unit length, equation (81) becomesL 2 = ( ˙φ 0 + )L z2 (r 0 ) − ˙φ 0 L z2 (0) − E 2 (r 0 ) (83)


R164A L Fetter and A A Svidz<strong>in</strong>skywhere φ 0 is the azimuth angle describ<strong>in</strong>g the position of the vortex l<strong>in</strong>e,L z2 (r 0 ) = 1 2 n(0)πR2 ⊥¯h(1 − ζ 0 2 )2 (84)and[ ( )]E 2 (r 0 ) = κ2 Mn(0)2(1 − ζ0 2 8π) ln R⊥+ (1 − ζ0 2 ξ) ln(1 − ζ 0 2 ) − 1+2ζ 02 (85)with ζ 0 = r 0 /R ⊥ (note that 1 n(0) is the mean particle density n per unit length). These2expressions differ from the classical results for a uniform fluid <strong>in</strong> a rotat<strong>in</strong>g cyl<strong>in</strong>der [90,123]because of the parabolic TF density; <strong>in</strong> particular, the TF angular momentum per unit lengthL z2 here is proportional to (1 − ζ0 2)2 , whereas that for a uniform density is proportional to1 − ζ0 2.The Lagrangian dynamical equations show that the vortex precesses at fixed r 0 with theangular frequency˙φ 0 =− + ∂E 2/∂r 0=− − ∂E 2/∂r 0∂L z2 /∂r 0 κMr 0 n(r 0 ) . (86)This result is just that expected from the Magnus force on a straight vortex [124–126]. For smalldisplacements from the centre, the precession frequency <strong>in</strong> a nonrotat<strong>in</strong>g two-dimensional<strong>condensate</strong> reduces to ˙φ 0 ≈ (κ/2πR⊥ 2 ) ln(R ⊥/ξ) ≈ 1 2 c [72], but ˙φ 0 <strong>in</strong>creases with <strong>in</strong>creas<strong>in</strong>gr 0 and eventually diverges near the edge of the <strong>condensate</strong> where the density vanishes.The correspond<strong>in</strong>g results for a three-dimensional disc-shaped TF <strong>condensate</strong> follow fromequations (49) and (81). In particular, the <strong>in</strong>tegration over z means that the total angularmomentum L z3 = N¯h(1 − ζ0 2)5/2 associated with the presence of the vortex differs from thetwo-dimensional result proportional to (1 − ζ0 2)2 . Apart from numerical factors reflect<strong>in</strong>g thethree-dimensional geometry, equation (86) rema<strong>in</strong>s correct. For a straight vortex, it yields m˙φ 0 =− +1 − r0 2 (87)/R2 ⊥where m = 3 2 (¯h/MR2 ⊥ ) ln(R ⊥/ξ) is the metastable frequency (51) for the appearanceof a central vortex <strong>in</strong> a disc-shaped <strong>condensate</strong>. In the special case of a vortex near thecentre (r 0 → 0), this precession frequency reduces to (m<strong>in</strong>us) the correspond<strong>in</strong>g anomalousfrequency ω a () <strong>in</strong> equation (72) for a <strong>condensate</strong> with a s<strong>in</strong>gle central vortex l<strong>in</strong>e. Tounderstand why the precession frequency ˙φ 0 is the negative of the anomalous frequency,recall that the l<strong>in</strong>earized perturbation <strong>in</strong> the density for the anomalous mode is proportional toexp i[m a φ − ω a ()t] = exp i[−φ − ω a ()t] because m a =−1; this latter form shows clearlythat the normal mode propagates around the symmetry axis at an angular frequency −ω a (),with the sense of rotation fixed by the sign of −ω a ().Accord<strong>in</strong>g to (87), for a nonrotat<strong>in</strong>g trap the precession velocity of a displaced vortex<strong>in</strong>creases with the vortex displacement as v = m r 0 /(1 − r0 2/R2 ⊥). It is <strong>in</strong>terest<strong>in</strong>g to estimateat what displacement the vortex velocity becomes supersonic [127]. Assum<strong>in</strong>g that the speedof sound varies radially with the local density asc = c 0√1 − r0 2/R2 ⊥where c 0 = √ µ/M = ω ⊥ R ⊥ / √ 2we obta<strong>in</strong>v/c = ( √ 2 m r 0 /ω ⊥ R ⊥ )(1 − r0 2 /R2 ⊥ )−3/2 .As a result, the vortex velocity becomes supersonic ifr 0>ω ( ) 3/2√ ( ) 3/2⊥√ 1 − r2 02R⊥R ⊥ 2m R⊥2 =1 − r2 03ξ ln(R ⊥ /ξ) R⊥2 . (88)For parameters of JILA experiments [37] R ⊥ /ξ ≈ 33, this gives a critical displacement ofr 0 /R ⊥ ≈ 0.82 where the precession vortex velocity becomes supersonic.


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R1655.2. The method of matched asymptotic expansionsAt zero temperature, the dynamics of a <strong>condensate</strong> <strong>in</strong> a rotat<strong>in</strong>g nonaxisymmetric trap followsfrom the appropriate time-dependent GP equationi¯h ∂ (−¯h2 ∇ 2)=∂t 2M + V tr + g|| 2 − µ() +i¯hΩ · (r × ∇) . (89)A vortex l<strong>in</strong>e <strong>in</strong> the <strong>condensate</strong> will, <strong>in</strong> general, move <strong>in</strong> response to the effect of the nonuniformtrap potential and the external rotation, as well as self-<strong>in</strong>duced effects caused by itsown local curvature. This problem can be solved <strong>in</strong> the case of a large <strong>condensate</strong>, where theTF separation of length scales means that the vortex-core radius ξ is much smaller than the<strong>condensate</strong> radii R j . The relevant mathematics <strong>in</strong>volves the method of matched asymptoticexpansions [128–130].5.2.1. Dynamics of a straight vortex <strong>in</strong> the Thomas–Fermi regime for a disc-shaped trap.As an <strong>in</strong>troduction to these techniques, it is helpful first to concentrate on the case of astraight s<strong>in</strong>gly quantized vortex l<strong>in</strong>e [80], which is applicable to disc-shaped <strong>condensate</strong>swith R z ≪ R ⊥ ; this analysis generalizes two-dimensional results found by Rub<strong>in</strong>ste<strong>in</strong> andPismen [129]. Assume that the vortex is located near the centre of the trap at a transverseposition r ⊥0 (t). In this region, the trap potential does not change significantly on a lengthscale comparable with the vortex-core size ξ. The method of matched asymptotic expansionscompares the solution of equation (89) on two very different length scales:• First, consider the detailed structure of the vortex core. Assume that the vortex moves witha transverse velocity V ⊥ẑ, and transform to a co-mov<strong>in</strong>g frame centred at the vortexcore. Away from the trap centre, the trap potential exerts a force proportional to ∇ ⊥ V trevaluated at the position r ⊥0 (t). The result<strong>in</strong>g steady solution <strong>in</strong>cludes the ‘asymptotic’region |r ⊥ − r ⊥0 |≫ξ.• Second, consider the region far from the vortex (on this scale, the vortex core is effectivelya s<strong>in</strong>gularity). The short-distance behaviour of this latter solution also <strong>in</strong>cludes the regionξ ≪|r ⊥ − r ⊥0 |. The requirement that the two solutions match <strong>in</strong> the overlapp<strong>in</strong>g regionof validity determ<strong>in</strong>es the translational velocity V of the vortex l<strong>in</strong>e.Unfortunately, the details become rather <strong>in</strong>tricate, but the f<strong>in</strong>al answer is elegant andphysical:[V =3¯h ( )]R⊥ 8µln −4Mµ ξ 3¯h(ωx 2 + ω2 y ) (ẑ × ∇ ⊥ V tr )[ (R⊥ln= 3¯h4Mµξ)− 2MR2 ⊥ ](ẑ × ∇ ⊥ V tr ) (90)3¯hwhere R ⊥ for an asymmetric trap is def<strong>in</strong>ed <strong>in</strong> equation (50). This expression has severalnotable features.(a) The motion is along the direction ẑ×∇ ⊥ V tr and hence follows an equipotential l<strong>in</strong>e of V tr .Thus the trajectory conserves energy, which is expected because the GP equation omitsdissipative processes. In the present case of an anisotropic harmonic trap, the trajectoryis elliptical.(b) For a nonrotat<strong>in</strong>g trap ( = 0), the motion is anticlockwise <strong>in</strong> the positive sense at thefrequency given by equation (80), proportional to ω x ω y .


R166A L Fetter and A A Svidz<strong>in</strong>sky(c) With <strong>in</strong>creas<strong>in</strong>g applied rotation , the translational velocity V decreases and vanishesat the special value m = 3¯h(ω2 x + ω2 y ) ( )R⊥ln = 3¯h ( )R⊥8µ ξ 2MR⊥2 ln(91)ξproportional to 1 2 (ω2 x + ω2 y ). This value precisely reproduces equation (51) associatedwith the onset of metastability for small transverse displacements of the vortex from thetrap centre.(d) For > m , the motion is clockwise as seen <strong>in</strong> the rotat<strong>in</strong>g frame. A detailed analysisbased on the normalization of the Bogoliubov amplitudes shows that the positive-normstate has a frequency (compare equation (80))ω a () = 2ω xω yωx 2 + ( − m ). (92)ω2 yNote that this expression differs somewhat from equation (70) because the trap here isanisotropic. The normal-mode frequency is negative and hence unstable for < m ,butit becomes positive and hence stable for > m .This direct analysis of the motion of a straight vortex reproduces the physics of the onset of(static) metastability (51) studied with the GP Hamiltonian and the (dynamic) anomalous mode(73) and (80) studied with the Bogoliubov equations and with the Lagrangian method.5.2.2. Dynamics of a curved vortex <strong>in</strong> the Thomas–Fermi regime. Consider a nonaxisymmetrictrap that rotates with an angular velocity Ω (for convenience, Ω is often takenalong the z-axis). At low temperature <strong>in</strong> a frame rotat<strong>in</strong>g with the same angular velocity, thetrap potential is time <strong>in</strong>dependent, and equation (89) describes the evolution of the <strong>condensate</strong>wave function. In the TF limit, the method of matched asymptotic expansions aga<strong>in</strong> yieldsan approximate solution for the motion of a s<strong>in</strong>gly quantized vortex l<strong>in</strong>e with <strong>in</strong>stantaneousconfiguration r 0 (z, t). Let ˆt be the local tangent to the vortex (def<strong>in</strong>ed with the usual right-handrule), ˆn be the correspond<strong>in</strong>g normal, and ˆb ≡ ˆt ׈n be the b<strong>in</strong>ormal. A generalization ofthe work of Pismen and Rub<strong>in</strong>ste<strong>in</strong> [128,129] eventually yields the explicit expression for thelocal translational velocity of the vortex [119]:V (r 0 ) =− ¯h ( ˆt × ∇Vtr (r 0 )2M g| TF | 2) ( √1+ k ˆb ln ξR 2 ⊥)+ k2+ 2 ∇V tr(r 0 ) × Ω8 ⊥ V tr (r 0 )where k is the local curvature (assumed small, with kξ ≪ 1) and ⊥ is the Laplacian operator<strong>in</strong> the plane perpendicular to Ω.This vector expression holds for general orientation of the gradient of the trap potential,the normal to the vortex l<strong>in</strong>e, and the angular velocity vector. Near the TF boundary of the<strong>condensate</strong>, the denom<strong>in</strong>ator of the first term becomes small, imply<strong>in</strong>g that the numeratorˆt × ∇V tr (r 0 ) must also vanish near the boundary. As a result, the axis of the vortex l<strong>in</strong>e ˆt isparallel to ∇V tr at the surface and hence obeys the <strong>in</strong>tuitive boundary condition that the vortexmust be perpendicular to the <strong>condensate</strong> surface.5.3. Normal modes of a vortex <strong>in</strong> a rotat<strong>in</strong>g two-dimensional TF <strong>condensate</strong>This very general equation (93) applies <strong>in</strong> many different situations [119]. The simplest caseis an <strong>in</strong>itially straight vortex <strong>in</strong> a two-dimensional asymmetric TF <strong>condensate</strong> with Ω = ẑand ω z = 0 (hence no conf<strong>in</strong>ement <strong>in</strong> the z-direction). For small displacements, the x- and(93)


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R167y-coord<strong>in</strong>ates of the vortex core execute harmonic motion ∝ exp[i(κz − ωt)] that can varybetween helical and planar depend<strong>in</strong>g on the relative phase of the x- and y-motion. Thedispersion relation ω κ () depends on the cont<strong>in</strong>uous parameter κ and the rotation frequency, along with the TF radii R x and R y [119]:√( √ )¯hω κ () =± (2 − κ2MR x R 2 Rx 2 − ˜)(2 − κ 2 Ry 2 − 1˜) ln ξy R⊥2 + |κ|2(94)8where4MR 2 x R2 y˜ =¯h(Rx 2 + R2 y ) ln ( ξ√1/R⊥ 2 + |κ|2 /8 ) (95)−1is a dimensionless rotation speed.Of all the various normal modes, a straight vortex l<strong>in</strong>e (κ = 0) has the most negative(anomalous) frequency[ ( )]¯h R⊥ 4µω a () =− ln −2MR x R y ξ ¯h(ωx 2 + ω2 y ) (96)where an analysis similar to that for equation (92) shows that the m<strong>in</strong>us sign corresponds tothe Bogoliubov solution with positive norm. For = 0, the vortex precesses anticlockwiseabout the z-axis <strong>in</strong> the positive sense. With <strong>in</strong>creas<strong>in</strong>g rotation frequency , the precessionfrequency decreases and vanishes at = m , where the metastable rotation frequency <strong>in</strong> twodimensions is m = ¯h(ω2 x + ω2 y )4µln(R⊥ξ)= ¯hMR⊥2 ln(R⊥ξ). (97)As expected, this value is the precession frequency 1 2 c discussed below equation (86)(compare equation (51) for m <strong>in</strong> a three-dimensional disc-shaped TF <strong>condensate</strong>; the differentnumerical coefficient arises from the <strong>in</strong>tegration over the parabolic density <strong>in</strong> the z-direction).More generally, for κ 2 > 0 and a nonaxisymmetric trap (R x > R y ), the oscillationfrequency can be imag<strong>in</strong>ary (and hence unstable) with<strong>in</strong> a range of axial wave numbersdeterm<strong>in</strong>ed by√(2 − ˜)/R x < |κ| ˜ m = 2. In the limit of a uniform unbounded<strong>condensate</strong> (R x ,R y →∞), the general dispersion relation reduces to the familiar one forhelical waves on a long straight vortex l<strong>in</strong>e [7]ω =± ¯h2M κ2 ln(|κ|ξ). (98)Us<strong>in</strong>g this dispersion relation, Barenghi [131] estimated the amplitude of the vortex wavesdue to thermal excitation (the cloud is assumed to rotate at an angular velocity > c , so thevortex is stable). He showed that f<strong>in</strong>ite-temperature effects <strong>in</strong> a Bose <strong>condensate</strong> can distort thevortex state significantly, even at the very low temperatures relevant to the experiments. ForT = 10 −7 K, ¯n ≈ 10 12 –10 13 cm −3 , and R ≈ 5 µm, the amplitude of vortex oscillations canbe 4–14 times the size of the vortex core. At the same time, the thermal excitation of vortexwaves <strong>in</strong> superfluid 4 He is negligible (much smaller than the correspond<strong>in</strong>g vortex-core size).


R168A L Fetter and A A Svidz<strong>in</strong>sky5.4. Normal modes of a vortex <strong>in</strong> a rotat<strong>in</strong>g three-dimensional TF <strong>condensate</strong>Consider a three-dimensional TF <strong>condensate</strong> with ω zz 2 Rz 2 = 2µ/Mω2 z .> 0, conf<strong>in</strong>ed with<strong>in</strong> a TF region5.4.1. General formalism. For a vortex that <strong>in</strong>itially lies along the z-axis, it is straightforwardto f<strong>in</strong>d the pair of coupled equations for the small transverse displacements of the vortex x(z,t)and y(z,t). In particular, we seek solutions of the formx = x(z)s<strong>in</strong>(ωt + ϕ 0 ) y = y(z)cos(ωt + ϕ 0 ) (99)<strong>in</strong> which case the amplitudes x(z)and y(z)describe the vortex shape and obey coupled ord<strong>in</strong>arydifferential equations. Introduc<strong>in</strong>g dimensionless scaled coord<strong>in</strong>ates x → R x x, y → R y y,z → R z z,wef<strong>in</strong>d from equation (93)˜ω(1 − z 2 )x =− d [β(1 − z 2 ) dy ]− y + ˜(1 − z 2 )y (100)dz dz˜ω(1 − z 2 )y =− d [α(1 − z 2 ) dx ]− x + ˜(1 − z 2 )x (101)dz dzwhereα = R2 xR 2 zβ = R2 yR 2 zcharacterize the trap anisotropy and(102)˜ω =2MR xR y¯h ln(R ⊥ /ξ) ω 4MRx ˜ 2 =R2 y¯h(Rx 2 + R2 y ) ln(R ⊥/ξ) (103)are dimensionless angular velocities.These equations (100) and (101) constitute a two-component Sturm–Liouville systemwith natural boundary conditions [132] because the factor 1 − z 2 vanishes at z = ±1.Consequently, the eigenfunctions merely must rema<strong>in</strong> bounded at the surface of the <strong>condensate</strong>.A straightforward generalization of the usual analysis shows that the eigenfunctions obey theorthogonality condition∫ 1−1dz (1 − z 2 )x m y n ∝ δ mn . (104)5.4.2. Special solutions. In the general case of a nonaxisymmetric trap, the result<strong>in</strong>gequations rema<strong>in</strong> coupled, but they separate <strong>in</strong> the particular case of stationary solutions withω = 0. For a nonrotat<strong>in</strong>g trap, such configurations reflect a balance between the effects ofcurvature and the nonuniform trap potential. For example, the small-amplitude stationarysolutions x n (z) rema<strong>in</strong> f<strong>in</strong>ite at the surface z =±1 only for certa<strong>in</strong> special values of the trapanisotropy2α = α n =(105)n(n +1)where n 0 is an <strong>in</strong>teger. The correspond<strong>in</strong>g solutions have the form x n (z) ∝ P n (z),where P n is the familiar Legendre polynomial. The solutions have n nodes and cross thez-axis n times. If α differs from one of these special values (105), there is no stationaryconfiguration. Similarly, the equation for the y-displacement has stationary solutions only ifβ ≡ Ry 2/R2 z = 2/[m(m +1)].


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R169This classification of the solutions accord<strong>in</strong>g to the number of nodes rema<strong>in</strong>s moregenerally valid. In the special case of an axisymmetric <strong>condensate</strong> (α = β), we can considerthe precession frequency ω n of the mode with n nodes as a function of the axial trap anisotropyα. Evidently, the function ω n changes sign at the special value α = α n = 2/[n(n +1)]. Thisobservation allows us to determ<strong>in</strong>e the number of modes with negative frequencies at a fixedvalue of the anisotropy parameter α. Forα 1 (a spherical or disc-shaped <strong>condensate</strong>), onlyone mode has a negative frequency. If 1


R170A L Fetter and A A Svidz<strong>in</strong>skyFor a harmonic transverse external potential ∝r⊥ 2 , the method of matched asymptoticexpansions is valid if the vortex displacement r from the z-axis satisfies the condition r ξ(<strong>in</strong> the vic<strong>in</strong>ity of the vortex core, the trap potential is approximated as a l<strong>in</strong>ear function). For along cigar-shaped <strong>condensate</strong>, the solution for the lowest mode has the form r = r 0 cosh(z/α),where r 0 is the vortex displacement at z = 0. The condition of small vortex displacementimplies that r 0 cosh(1/α) ≪ R ⊥ , while the condition of small vortex curvature kξ ≪ 1implies that r 0 ξ cosh(1/α)/Rz 2α2 ≪ 1. A comb<strong>in</strong>ation of these conditions gives the follow<strong>in</strong>grestriction on the validity of equation (108): exp(1/α) ≪ 2R ⊥ /ξ. For the ENS experiments,1/α ≈ 200 and R ⊥ /ξ ≈ 21, so this condition fails.As mentioned <strong>in</strong> section 4.4.4, the frequency for the onset of metastability m <strong>in</strong>equation (108) can be larger than the thermodynamic critical angular velocity c <strong>in</strong>equation (43). This behaviour is readily understandable because c characterizes the energy ofa straight vortex along the symmetry axis (compare equation (42)), whereas the most unstablenormal-mode amplitude explicitly <strong>in</strong>volves the small-amplitude distortion with no nodes. Fora very elongated <strong>condensate</strong>, the result<strong>in</strong>g vortex dynamics is particularly sensitive to the largecurvature of the <strong>condensate</strong> surface near the two ends of the symmetry axis (<strong>in</strong> contrast to thesmall curvature for the flattened <strong>condensate</strong>).Recent numerical studies [85, 118] of the most negative anomalous modes for a trapgeometry correspond<strong>in</strong>g to the ENS experiments [34, 36] yield values of m that aresignificantly smaller than the prediction given <strong>in</strong> equation (108). Reference [118] mentionsthe possible failure of the TF picture <strong>in</strong> the transverse direction, even though the conventionalTF ratio R ⊥ /ξ is large, at least near the plane z = 0. As confirmation of the validity of theGP equation and the particular role of the anomalous modes, the numerically determ<strong>in</strong>ed [85] m /2π ≈ 0.73ν ⊥ ≈ 124 Hz agrees well with the ENS value obs /2π ≈ 120 Hz for theappearance of the first vortex.For an axisymmetric trap (α = β), we can seek normal-mode solutions <strong>in</strong> the formx(z) = y(z), leav<strong>in</strong>g a s<strong>in</strong>gle equation[ ]˜ω( ˜) − ˜ (1 − z 2 )x =− d [α(1 − z 2 ) dx ]− x (109)dz dzthat depends only on the Doppler-shifted frequency ˜ω( ˜) − ˜ = ˜ω(0). The eigenfunctionsare even or odd functions of z and can be classified accord<strong>in</strong>g to the number of times the vortexcrosses the z-axis (the number of nodes), m = 0, 1, 2,.... Figure 12 shows the dimensionlessfrequency ˜ω(0) as a function of the trap anisotropy α = R⊥ 2 /R2 z for m = 0, 1, and 2. Inagreement with the analytical results, a disc-shaped trap (α 1) has only a s<strong>in</strong>gle mode withnegative frequency ˜ω 0 .For 1


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R1712018161412ω~1086420-2-4-6-8~ω 2~ ω 1ω ~ ~= −Ω c0 1 2 3 4 52 2R ⊥/ R zα=~ω 0Figure 12. Dimensionless frequencies ˜ω ≡˜ω( = 0) for the first three normal modes of a vortex<strong>in</strong> an axisymmetric trap as a function of the axial anisotropy α = R⊥ 2 /R2 z . The lower horizontall<strong>in</strong>e is the negative of the dimensionless thermodynamic critical angular velocity ˜ c = 5. Notethat |˜ω 0 | > ˜ c for α 0 (112)−1−1and x m = x m (z) describes the shape of the mth vortex mode. As we <strong>in</strong>crease the trap rotation,the eigenfrequency is real for ˜


R172A L Fetter and A A Svidz<strong>in</strong>skyfrequency of this mode becomes imag<strong>in</strong>ary <strong>in</strong> the <strong>in</strong>terval | ˜ −|˜ω a ||


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R173units x → R x x, etc, this energy functional becomes∫ [E V (x(z), y(z)) = 2πµR z ξ 2 n(0) dz (1 − x 2 − y 2 − z 2 ) √ 1+α(x ′ ) 2 + β(y ′ ) 2 ln− 2µ(1 − x2 − y 2 − z 2 ) 2 ]¯h(ωx 2 + ω2 y )(R⊥ξ)(116)where n(0) = µ/g is the density at the centre of the vortex-free <strong>condensate</strong>, ξ 2 = ¯h 2 /2Mµ,and the <strong>in</strong>tegration is restricted to the region 1 − x 2 − y 2 − z 2 0. Us<strong>in</strong>g equation (116) onecan obta<strong>in</strong> a simple expression for the angular momentum of the <strong>condensate</strong> <strong>in</strong> the presenceof a curved vortex l<strong>in</strong>e:L z =− ∂E V∂ = 15 8 ¯hN R xR yRx 2 + dz (1 − x 2 − y 2 − z 2 ) 2 (117)R2 ywhere N = 8πR x R y R z n(0)/15 is the total number of particles <strong>in</strong> the <strong>condensate</strong>.The <strong>in</strong>tegration <strong>in</strong> equation (116) is particularly easy for a straight vortex and readilyreproduces equation (49). An expansion for small lateral displacements yields equations (51)and (52) for m and c for a disc-shaped TF <strong>condensate</strong>. In the more general case of arbitrarysmall displacements, equation (116) can be expanded to second order <strong>in</strong> the amplitudes x andy and their derivatives. Use of the dynamical equations that lead to (100) and (101) gives thesimple expression[E V (x(z), y(z)) = 8π ( )]3 µR zξ 2 R⊥n(0) ln − 8 µξ 5 ¯h(ωx 2 + ω2 y )+ 15 ∫ 18 ¯hN dz (1 − z 2 )(xẏ − yẋ) (118)−1The first term of equation (118) reproduces the value of c for a general TF <strong>condensate</strong>, andthe second term becomes a sum over all normal modes of the form (99)[E V (x(z), y(z)) = 8π ( )]3 µR zξ 2 R⊥n(0) ln − 8 µξ 5 ¯h(ωx 2 + ω2 y )+ 158 N ∑ n∫∫ 1¯hω n () dz (1 − z 2 )x n (z)y n (z) (119)−1where the orthogonality condition equation (104) elim<strong>in</strong>ates the cross terms for differentnormal modes. If any one of the normal modes is anomalous (i.e. has negative frequency),then the system is unstable with respect to excitation of that mode. This analysis confirmsthe <strong>in</strong>terpretation of m as the applied rotation frequency at which the frequency of the lastanomalous mode vanishes <strong>in</strong> the rotat<strong>in</strong>g frame. At this applied the location of the vortexl<strong>in</strong>e along the z-axis becomes a local m<strong>in</strong>imum of energy. Note that this conclusion is whollyequivalent to that <strong>in</strong> equation (60) based on the Bogoliubov quasiparticles.One should note that for a cigar-shaped <strong>condensate</strong> with R z 2R ⊥ , there is an <strong>in</strong>tervalof angular velocity of trap rotation when c


R174A L Fetter and A A Svidz<strong>in</strong>sky5.4.4. Precession and tilt<strong>in</strong>g of a straight vortex l<strong>in</strong>e <strong>in</strong> a nearly spherical TF <strong>condensate</strong>.The preced<strong>in</strong>g discussion of vortex dynamics <strong>in</strong> a three-dimensional conf<strong>in</strong>ed <strong>condensate</strong> hasfocused on the small-amplitude displacements from equilibrium. In the special case of aspherical trap, however, the presence of a zero-frequency precess<strong>in</strong>g mode (section 5.4.2)allows a more general analysis of the nonl<strong>in</strong>ear dynamics, which is directly relevant to recentJILA experiments on the evolution of an <strong>in</strong>itially straight vortex <strong>in</strong> a nearly spherical TF<strong>condensate</strong> [94]. In practice, the trap deviates slightly from spherical with R x ≠ R y ≠ R z .For a spherical <strong>condensate</strong>, a motionless straight s<strong>in</strong>gly quantized vortex through the centreof a trap satisfies the general equation (93) for the velocity of a vortex l<strong>in</strong>e because the axis ofthe vortex ˆt lies along ∇V tr . Letx = γ x s y = γ y s z = γ z s (120)specify the axis of the vortex l<strong>in</strong>e, where s is the arc length measured from the trap centre and(γ x , γ y , γ z ) are the direction cos<strong>in</strong>es relative to the pr<strong>in</strong>cipal axes of the anisotropic trap. Forsmall anisotropy, the vortex rema<strong>in</strong>s approximately straight, but the direction cos<strong>in</strong>es becometime dependent. To first order <strong>in</strong> the anisotropy, the curvature k can be omitted <strong>in</strong> equation (93)and | TF | 2 can be approximated by the TF density for a spherical vortex-free <strong>condensate</strong> withTF radius R. Standard perturbation theory yields the nonl<strong>in</strong>ear dynamical equations)(ωz 2 − ω2 y )γ yγ z (121)˙γ x = 5¯h4µ ln ( Rξ˙γ y = 5¯h4µ ln ( Rξ˙γ z = 5¯h4µ ln ( Rξ)(ω 2 x − ω2 z )γ zγ x (122))(ω 2 y − ω2 x )γ xγ y . (123)This set of equations is familiar <strong>in</strong> classical mechanics as Euler’s equations for the torquefreemotion of a rigid body [133–135], where they describe the motion of the angular velocityvector as seen <strong>in</strong> the body-fixed frame. In the present context, this set of three coupled nonl<strong>in</strong>earequations has two first <strong>in</strong>tegrals:γx 2 + γ y 2 + γ z 2 = 1 (124)which verifies that the first-order anisotropy simply rotates the vortex axis; andωx 2 γ x 2 + ω2 y γ y 2 + ω2 z γ z 2 = constant (125)which is the condition of energy conservation.The simplest situation is an axisymmetric trap with ω x = ω y = ω ⊥ , <strong>in</strong> which case thevortex l<strong>in</strong>e precesses uniformly about the z-axis (the symmetry axis) at a fixed polar anglearccos γ z (0) at a frequency [119]ω = 5¯h(ω2 z − ω2 ⊥ ) ( ) 1.96Rγ z (0) ln4µξ= 5¯h2M( 1R 2 z− 1 ) ( 1.96RR⊥2 γ z (0) lnξ)(126)where the numerical factor 1.96 <strong>in</strong>side the logarithm is the same as that discussed belowequation (107). For positive (negative) ω, the precession is anticlockwise (clockwise). Recentexperiments at JILA have observed two recurrences of such precessional motion <strong>in</strong> a slightlyflattened trap with ω z − ω ⊥ ≈ 0.1ω z and a polar tipp<strong>in</strong>g angle of 45 ◦ from the z-axis. In thiscase, equation (126) predicts ω/2π ≈ 0.33 ± 0.03 Hz, <strong>in</strong> an agreement with the observedvalue 0.25 ± 0.02 Hz [94].More generally, for an anisotropic trap (with ω x >ω y >ω z ), the vortex executes closedtrajectories (see figure 13). For <strong>in</strong>itial positions close to the x- and z-axes (the smallest


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R175and largest TF radii), the motion is ‘stable’, rema<strong>in</strong><strong>in</strong>g nearby, but small-amplitude motionabout an <strong>in</strong>itial position close to the y-axis (the <strong>in</strong>termediate TF radius) yields imag<strong>in</strong>aryfrequencies. Thus such trajectories deviate far from the <strong>in</strong>itial neighbourhood, even thoughthey eventually return (this periodic behaviour is familiar from the correspond<strong>in</strong>g solutionsof the Euler equations [133–135]). Reference [119] gives explicit solutions for the result<strong>in</strong>gdynamical motion of a nearly straight vortex <strong>in</strong> a totally anisotropic trap.ZXO .YFigure 13. Typical trajectories of the end of a straight vortex l<strong>in</strong>e (that passes through the <strong>condensate</strong>centre) dur<strong>in</strong>g its motion <strong>in</strong> a slightly nonspherical trap with R x


R176A L Fetter and A A Svidz<strong>in</strong>skyand ρ =〈ˆφ † ˆφ〉). The collective excitation energies E of the system are the eigenvalues of thegeneralized Bogoliubov equations for the coupled amplitudes u(r) and v(r):)(− ¯h22M ∇2 + V tr +2g|| 2 u+2gρ(r) − µ())(v(i¯h ∂ φ −g [ (r) + 2] )( ) ( )u u+−g [ ∗ (r) + ∗2] = E . (128)−i¯h ∂ φv −vEquation (128) is valid at least for temperatures much less than the chemical potential µ whenresonant contributions (the so-called Szepfalusy–Kondor processes) to the self-energies arenot substantial [137]. In addition, we have self-consistency relations for the non<strong>condensate</strong>density ρ(r):ρ(r) = ∑ [ |un (r)| 2 + |v n (r)| 2 ]nexp(E n /k B T)− 1 + |v n(r)| 2 (129)and for the anomalous average (r):(r) =− ∑ [2u n (r)vn ∗(r)]nexp(E n /k B T)− 1 + u n(r)vn ∗ (r) (130)where n denotes quantum numbers specify<strong>in</strong>g the excited states with energies E n (n =0, 1, 2,...). The eigenfunctions u n (r) and v m (r) satisfy the normalization condition:∫ [u∗n (r)u m (r) − vn ∗ (r)v m(r) ] dr = δ nm . (131)Equations (127)–(131) constitute a complete set of the self-consistent equations for theHFB theory. With<strong>in</strong> this theory, the quasiparticle eigenvalues E n <strong>in</strong> equations (128)–(130)must be positive because the <strong>condensate</strong> is def<strong>in</strong>ed to have zero energy. Thus a negativeeigenvalue means a failure of the self-consistency and the associated thermal equilibrium ofthe system. If ρ(r) and (r) are set to zero, we recover the Bogoliubov theory. If we setonly (r) = 0, we obta<strong>in</strong> the Popov approximation. For a vortex-free <strong>condensate</strong> <strong>in</strong> the lowtemperaturelimit, the Popov and Bogoliubov theories give identical excitation spectra [140].The excitation spectrum <strong>in</strong> the HFB theory has an unphysical gap because it does not treatall <strong>condensate</strong>–<strong>condensate</strong> <strong>in</strong>teractions consistently [136]. Gapless modifications of the HFBtheory, the so-called G1 and G2 approximations, are discussed <strong>in</strong> [141, 142]. Normally, thezero-temperature limit of the Popov, G1, and G2 theories should be the Bogoliubov theory(which does not take <strong>in</strong>to account non<strong>condensate</strong> atoms). For a nonrotat<strong>in</strong>g <strong>condensate</strong> witha vortex, however, this is not the case because the vortex is unstable.With<strong>in</strong> the Bogoliubov theory, an isolated vortex <strong>in</strong> a nonrotat<strong>in</strong>g harmonic trap has atleast one normal mode with negative energy. Let us apply the HFB theory for a <strong>condensate</strong>with a vortex. To f<strong>in</strong>d a self-consistent solution for the lowest eigenvalue at low temperatures,one can use a perturbation method analogous to those developed <strong>in</strong> reference [116]. Weconsider a <strong>condensate</strong> <strong>in</strong> an axisymmetric trap that rotates with an angular velocity aroundthe z-axis. We assume that the <strong>condensate</strong> conta<strong>in</strong>s a s<strong>in</strong>gly quantized vortex along the z-axis. For simplicity we consider a disk-shaped <strong>condensate</strong>, so one can omit vortex curvature <strong>in</strong><strong>in</strong>vestigat<strong>in</strong>g the lowest normal mode. The <strong>condensate</strong> wave function has the form = e iφ ||,with = e 2iφ ||, and we can rewrite the generalized Bogoliubov equations as( ) ( ) ( )Hˆu u u0 + ˆV = E(132)v v −vwhereˆ H 0 =(− ¯h22M ∇2 + 1 2 Mω2 z z2 +2g| 0 | 2 − µ())(1 00 1) ( i¯h2)∂φ −g 0+−g0 ∗2 −i¯h ∂ φ(133)


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R177and ˆV <strong>in</strong>cludes the rema<strong>in</strong><strong>in</strong>g part of equation (128). Here, 0 is the wave function for anunbounded <strong>condensate</strong> <strong>in</strong> the xy-plane with the same chemical potential; its excitations obeythe equation( ) ( )Hˆu0 u00 = E . (134)v 0 −v 0Equation (134) has an exact pair of solutions (see reference [116]) with positive norm andenergyE 0 = ¯h. (135)Let us now make the follow<strong>in</strong>g assumption: E 0 ≪ k B T ≪ E 1 ,E 2 ,..., where E 0 is theenergy of the lowest normal mode, which can depend on T . Then the term with n = 0givesthe ma<strong>in</strong> contribution <strong>in</strong> the sum <strong>in</strong> equations (129), (130), and we obta<strong>in</strong>ρ(r) ≈ k BT [|u0 (r)| 2 + |v 0 (r)| 2] (136)E 0(r) ≈− 2k BTu 0 (r)v0 ∗ (r). (137)E 0For a s<strong>in</strong>gly quantized vortex one can derive the expressionρ(r ⊥ = 0,z)≈ 1.44µk ( )BT1 − z2(138)E 0 I 2 gξ 2 Rz2whereI 2 ≈ 16 √ √2πµ 3/2 /3gω z Mis a normalization <strong>in</strong>tegral, and|(r ⊥ = 0,z)|≈0. (139)In first-order perturbation theory, the lowest energy eigenvalue E 0 is def<strong>in</strong>ed by theequation( ) γkB TE 0 = ¯h + E a − 1(140)E 0where E a = (3¯h 2 ω⊥ 2 /4µ) ln(R ⊥/ξ) and γ = 0.077R⊥ 4 /Nξ 4 are positive with N =815 πn(0)R zR⊥ 2 the total number of particles <strong>in</strong> the <strong>condensate</strong>, and µ can be taken as thechemical potential for a nonrotat<strong>in</strong>g trap. Equation (140) has two solutions, one with positiveenergy and one with negative energy that reproduces the previous anomalous mode withE 0 = ¯h − E a as T → 0. The negative solution can be formally omitted, satisfy<strong>in</strong>g therequirement of self-consistency. The positive solution has the formE 0 = 1 [√(Ea − ¯h)22 +4E a γk B T − (E a − ¯h)]. (141)For a nonrotat<strong>in</strong>g trap ( = 0), we f<strong>in</strong>d[√]E 0 = E a1+ 4γk BT− 1 . (142)2 E aIf T → 0, we obta<strong>in</strong> E 0 ≈ γk B T ,soE 0 is proportional to T <strong>in</strong> the low-temperature limit. Infact our method generalizes the Beliaev theory [138] for the vortex state. Recently Pitaevskiiand Str<strong>in</strong>gari actually generalized the Beliaev approach (<strong>in</strong> the density-phase representation)for the <strong>trapped</strong> Thomas–Fermi <strong>condensate</strong> [139].


R178A L Fetter and A A Svidz<strong>in</strong>skyVirtanen et al made numerical calculations of vortex normal modes at f<strong>in</strong>ite T with<strong>in</strong>the Popov, G1, and G2 approximations and demonstrated that for a s<strong>in</strong>gly quantized vortexthere is a self-consistent solution with only positive frequencies <strong>in</strong> the limit T → 0 [143].Their lowest-energy solution corresponds to our equation (142). The vortex mode (142)arises from the presence of quasiparticles (an external p<strong>in</strong>n<strong>in</strong>g potential can also result <strong>in</strong> suchmotion [144]). At low temperatures, the quasiparticles are mostly localized <strong>in</strong> the vortexcoreregion and provide an extra repulsive potential (the term 2gρ(r) <strong>in</strong> equation (128)) thataffects the elementary excitations. At T = 0, the residual localized non<strong>condensate</strong> fractionarises from the <strong>in</strong>teraction between particles; this result follows from equation (138) if wetake E 0 ∝ T at low temperatures. The additional potential has a peak at the vortex core andthe vortex l<strong>in</strong>e precesses around the quasiparticle potential centre with a positive excitationenergy.However, this does not mean that quasiparticles stabilize the vortex <strong>in</strong> a trap. The physicsof the problem is the follow<strong>in</strong>g. At any moment dur<strong>in</strong>g the vortex motion, quasiparticles fillthe vortex core (the relaxation time of quasiparticles is much less than the period of the vortexprecession). The vortex l<strong>in</strong>e participates <strong>in</strong> two motions: first, the vortex precesses around thetrap centre with the frequency ¯hω a =−E a < 0( = 0). The trap potential is responsible forthis unstable mode. The quasiparticles are localized <strong>in</strong> the vortex core and move together withthe vortex; their presence simply slightly changes the chemical potential and slightly decreasesthe normal-mode frequency. In secondary motion, the vortex l<strong>in</strong>e moves around the centre ofmass of the quasiparticles <strong>in</strong> a locally uniform <strong>condensate</strong> (<strong>in</strong> the xy-plane). The amplitudeof this motion is less than ξ and the frequency can be found from equation (142) <strong>in</strong> the limitR x ,R y →∞or E a → 0:¯hω T = √ √γE a k B T = 0.37(µk B Tn 0 R z ξ ln R⊥2 ξ)(143)where n 0 is the density of the vortex-free <strong>condensate</strong> at the vortex location (<strong>in</strong> the plane z = 0).For JILA parameters, γ ≈ 0.3, E a ≈ 1.58 Hz; then for T = 0.8T c we obta<strong>in</strong> ω T ≈ 13.6 Hz.If this mode is thermally excited, its amplitude is given by( ) 6a 1/2 ( )kB T 1/2A = ξ(144)R z ¯hω Twhere a is the scatter<strong>in</strong>g length. For parameters of JILA experiments, A = 0.16ξ. Tak<strong>in</strong>g<strong>in</strong>to account ω T ∝ √ T we obta<strong>in</strong> the follow<strong>in</strong>g temperature dependence: A ∝ T 1/4 . It is<strong>in</strong>terest<strong>in</strong>g to note that the thermal mode (143) exists only <strong>in</strong> the 3D <strong>condensate</strong>; <strong>in</strong> the limitR z =∞, both the mode frequency and the amplitude go to zero.Recent measurements of the lowest vortex modes <strong>in</strong> the JILA experiments are <strong>in</strong> goodquantitative agreement with the solutions of the time-dependent Gross–Pitaevskii equation[37,85,119]. The JILA experiments measure, <strong>in</strong> fact, not only the absolute value, but also thesign of the lowest vortex mode. The negative value of the anomalous-mode frequency meansthat the vortex precesses <strong>in</strong> the same direction as the superfluid flow around the vortex core,which is seen <strong>in</strong> the experiments. An experimental observation of the thermal mode (143)could be the next challeng<strong>in</strong>g problem for future <strong>in</strong>vestigations.6.2. Dissipation and vortex lifetimesIt is valuable to consider dissipation and its role <strong>in</strong> the vortex lifetime. In a nonrotat<strong>in</strong>gtrap, the ground state of the system is a vortex-free <strong>condensate</strong>, so a <strong>condensate</strong> with avortex necessarily constitutes an excited state. In the absence of dissipation, however, the


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R179vortex l<strong>in</strong>e moves along trajectories of constant energy, rema<strong>in</strong><strong>in</strong>g <strong>in</strong>side the <strong>condensate</strong>.The <strong>condensate</strong> with a vortex will be unstable only if there is a mechanism to transfer thesystem to the lower-energy vortex-free state [99]. The dissipative dynamics of a straightvortex due to its <strong>in</strong>teraction with the thermal cloud <strong>in</strong> a <strong>trapped</strong> Bose-condensed gas wasdiscussed by Fedichev and Shlyapnikov [145]. If the vortex l<strong>in</strong>e moves with respect to thenormal component, scatter<strong>in</strong>g of elementary excitations by the vortex produces a friction force,like that <strong>in</strong> superfluid 4 He (see chapter 3 of reference [14]). Such a mechanism can transferenergy and momentum to the thermal cloud. The friction force F can be decomposed <strong>in</strong>tolongitud<strong>in</strong>al and transverse components: F =−Du − D ′ (u ׈n), where u is the velocityof the vortex l<strong>in</strong>e with respect to the normal component, D and D ′ are the longitud<strong>in</strong>al andtransverse friction coefficients, respectively, and ˆn is a local tangent vector to the vortex l<strong>in</strong>e.The transverse friction coefficient is <strong>in</strong>dependent of the scatter<strong>in</strong>g amplitude and is given bythe universal expression D ′ = ¯hρ n /M, where ρ n is the local mass density of the normalcomponent [146]. The longitud<strong>in</strong>al friction coefficient depends on the scatter<strong>in</strong>g process. Inthe limit k B T ≫ µ, one can treat the elementary excitations as s<strong>in</strong>gle particles, with theresult that ρ n ≈ 0.1M 5/2 T 3/2 /¯h 3 and the longitud<strong>in</strong>al friction coefficient is proportional tothe temperature: D ≈ ¯hn(na 3 ) 1/2 T/µ, where n = || 2 is the superfluid density for thevortex-free <strong>condensate</strong> and a is the s-wave scatter<strong>in</strong>g length [145].In the presence of dissipation, the vortex l<strong>in</strong>e moves toward a (local) m<strong>in</strong>imum of theenergy. In a nonrotat<strong>in</strong>g <strong>condensate</strong>, an off-centre vortex precesses around the trap centreand is expected to spiral out to the <strong>condensate</strong> boundary due to the dissipation. Once thevortex reaches the boundary, it presumably decays by emitt<strong>in</strong>g phonons and s<strong>in</strong>gle-particleexcitations. The radial motion of the vortex is governed by the longitud<strong>in</strong>al friction coefficient:v r ≈ Du/¯hn ≈ (na 3 ) 1/2 T u/µ ≪ u, where u is the precessional speed. Us<strong>in</strong>g this expression,one can estimate the characteristic lifetime of the vortex state [145]. At present, no dissipationof the mov<strong>in</strong>g vortex has been observed <strong>in</strong> the JILA experiments [37]. The characteristicdecay time for the dissipative mechanism of Fedichev and Shlyapnikov <strong>in</strong> the JILA conditionsis significantly larger than the lifetime of the <strong>condensate</strong>. The temperature and density are toosmall to see the dissipation.Another factor that can <strong>in</strong>fluence the vortex lifetime is the possibility that a mov<strong>in</strong>g vortexcan emit phonons. It is known that a mov<strong>in</strong>g vortex <strong>in</strong> an <strong>in</strong>f<strong>in</strong>ite compressible fluid emitsphonons, lead<strong>in</strong>g to a slow loss of energy [147]. Recently, Lundh and Ao [125] studied theradiation of sound from a mov<strong>in</strong>g vortex <strong>in</strong> an <strong>in</strong>f<strong>in</strong>ite, uniform system. A homogeneoustwo-dimensional superfluid described by a nonl<strong>in</strong>ear Schröd<strong>in</strong>ger equation is equivalent to(2 + 1)-dimensional electrodynamics, with vortices play<strong>in</strong>g the role of charges and soundcorrespond<strong>in</strong>g to electromagnetic radiation [148, 149]. Thus, a vortex mov<strong>in</strong>g on a circulartrajectory <strong>in</strong> an <strong>in</strong>f<strong>in</strong>ite superfluid radiates sound waves, which are analogous to the cyclotronradiation of an electrical charge mov<strong>in</strong>g along a circular orbit. The power radiated by a vortexwith unit length execut<strong>in</strong>g circular motion with frequency ω at a radius r 0 is given by thefollow<strong>in</strong>g Poynt<strong>in</strong>g vector [125]:P = πQ2 ω 3 r02 (145)4cs2where Q = −¯h √ 2πn/M is the ‘vortex charge’, n is the uniform superfluid density, andc s = √ µ/M is the velocity of sound.In a nonuniform system, such as a two-dimensional or a disk-shaped axisymmetric <strong>trapped</strong><strong>condensate</strong>, an off-centre vortex performs a circular motion around the symmetry axis. If suchmotion excites sound waves (radiates energy), the vortex will move outward toward regionsof lower potential energy, until it eventually escapes from the cloud. In a <strong>trapped</strong> <strong>condensate</strong>,


R180A L Fetter and A A Svidz<strong>in</strong>skyhowever, the excitations all rema<strong>in</strong> conf<strong>in</strong>ed with<strong>in</strong> the <strong>condensate</strong>, and no phonon radiation isexpected. In particular, the wavelength λ of sound that would be emitted exceeds the size R ofthe <strong>condensate</strong>. Indeed, λ ∼ 2πc s /ω and the precession frequency of the straight vortex is ofthe order of ω ∼ ¯h ln(R/ξ)/MR 2 ; as a result, λ/R ∼ (R/ξ) ln(R/ξ) ≫ 1, and the ‘cyclotron’radiation is prohibited.F<strong>in</strong>ally, let us discuss how vortex generation affects the dissipation <strong>in</strong> superfluids.One classic manifestation of superfluidity is that objects travell<strong>in</strong>g below a critical velocitypropagate through a superfluid without dissipation. Accord<strong>in</strong>g to the Landau criterion [107],which relies on the use of Galilean <strong>in</strong>variance, the critical velocity is v L = m<strong>in</strong>[E(p)/p],where E(p) is the energy of an elementary excitation with momentum p. For a homogeneousBose <strong>condensate</strong>, the Bogoliubov spectrum implies a Landau critical velocity equal to thespeed of sound v L = c s . The Landau critical velocity can usually be observed only bymov<strong>in</strong>g microscopic particles through the superfluid. Such motion of microscopic impuritiesthrough a <strong>trapped</strong> gaseous Bose <strong>condensate</strong> was studied recently <strong>in</strong> [150]. As the impuritiestraverse the <strong>condensate</strong>, they dissipate energy by collid<strong>in</strong>g with the stationary <strong>condensate</strong>and radiat<strong>in</strong>g phonons. When the impurity velocity was reduced below the speed of sound,however, the collision probability decreased dramatically, provid<strong>in</strong>g evidence for superfluidity<strong>in</strong> the <strong>condensate</strong>.If a macroscopic object moves through the <strong>condensate</strong>, dissipation can occur due toturbulence and vortex formation <strong>in</strong> the superfluid, even if the object’s velocity is much lowerthan the Landau critical velocity. Recently, dissipation <strong>in</strong> a Bose–E<strong>in</strong>ste<strong>in</strong> condensed gaswas studied by mov<strong>in</strong>g a blue-detuned laser beam through the <strong>condensate</strong> [151, 152]. Thelaser beam repels atoms from its focus and creates a mov<strong>in</strong>g macroscopic ‘hole’ <strong>in</strong> the<strong>condensate</strong>. The observed heat<strong>in</strong>g of the system agrees with the prediction of dissipationwhen the flow field becomes locally supersonic. Numerical simulations of the nonl<strong>in</strong>earSchröd<strong>in</strong>ger equation were used to study the flow field around an object mov<strong>in</strong>g through ahomogeneous <strong>condensate</strong> [28, 124, 153–155]. When the object moves faster than a criticalvelocity v c , these studies show that the superfluid flow becomes unstable aga<strong>in</strong>st the formationof quantized vortex l<strong>in</strong>es, which gives rise to a new dissipative regime. Pairs of vortices withopposite circulation are generated at opposite sides of the object. The rate of the energy transferto the <strong>condensate</strong> by the mov<strong>in</strong>g object <strong>in</strong>creases significantly above this critical velocity forvortex formation. The heat<strong>in</strong>g rate can be expressed as dE/dt = E pair f s , where E pair is theenergy of a vortex pair and f s is the shedd<strong>in</strong>g frequency. The rate of vortex-pair shedd<strong>in</strong>g f sis proportional to v − v c and thus larger when the speed of sound is lower.Other simulations of the GP equation have demonstrated that vortex–antivortex pairs orvortex half-r<strong>in</strong>gs can be generated by superflow around a stationary obstacle [28,154,156,157]or through a small aperture [158]. One might expect similar excitations <strong>in</strong> a rotat<strong>in</strong>g <strong>condensate</strong>.In addition, vortex half-r<strong>in</strong>gs can be nucleated at the <strong>condensate</strong> surface when the localtangential velocity exceeds a critical value.7. Vortex states <strong>in</strong> mixtures and sp<strong>in</strong>or <strong>condensate</strong>sThe advent of multicomponent BECs [159–161] has provided many new possibilities forquantum-mechanical state eng<strong>in</strong>eer<strong>in</strong>g. S<strong>in</strong>ce there is no <strong>in</strong>tr<strong>in</strong>sic difficulty <strong>in</strong> load<strong>in</strong>g andcool<strong>in</strong>g more than one alkali element <strong>in</strong> the same trap, <strong>in</strong>terpenetrat<strong>in</strong>g superfluids can nowbe realized experimentally. B<strong>in</strong>ary mixtures of <strong>condensate</strong>s can consist of different alkalis, ordifferent isotopes, or different hyperf<strong>in</strong>e states of the same alkali atom. Such b<strong>in</strong>ary mixturesof Bose <strong>condensate</strong>s have a great variety of ground states and vortex structures that areexperimentally accessible by vary<strong>in</strong>g the relative particle numbers of different alkalis [6].


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R181In particular, one can move cont<strong>in</strong>uously from regimes of <strong>in</strong>terpenetrat<strong>in</strong>g superfluids to thosewith separated phases. Many alkali b<strong>in</strong>ary mixtures conta<strong>in</strong> a coexistence region, which is theanalogue of 3 He– 4 He <strong>in</strong>terpenetrat<strong>in</strong>g superfluids <strong>in</strong> ultralow-temperature physics [162].7.1. Basic phenomenaMost experiments on Bose–E<strong>in</strong>ste<strong>in</strong> condensation of atomic gases of 87 Rb [1], 7 Li [2], and23 Na [3] have used magnetic traps to condense atoms with a hyperf<strong>in</strong>e sp<strong>in</strong> F = 2 (or F = 1).Such a <strong>condensate</strong> of sp<strong>in</strong>-F bosons constitutes a sp<strong>in</strong>or field〈 ˆψ m (r,t)〉=ζ m (r, t)(r,t) (146)where ˆψ m is the field operator, m labels F z (where −F m F ), is a scalar, and ζ m is anormalized sp<strong>in</strong>or. In magnetic traps, the sp<strong>in</strong>s of the alkali atoms are frozen and maximallyaligned with the local magnetic field B [6]. As a result, ζ is given by the eigenvalue equationˆB · F ζ = Fζ, where F is the hyperf<strong>in</strong>e sp<strong>in</strong> operator and ˆB is a unit vector along B. Thedynamics of 〈 ˆψ m 〉 is therefore completely specified by the scalar field , as<strong>in</strong> 4 He. Thus,even though the alkali atoms carry a sp<strong>in</strong>, they behave <strong>in</strong> magnetic traps like scalar particles.In contrast to the scalar field, however, the sp<strong>in</strong>or field <strong>in</strong> equation (146) possesses a localsp<strong>in</strong>-gauge symmetry: a local gauge change exp[iχ(r,t)]of〈 ˆψ m 〉 can be undone by a localsp<strong>in</strong> rotation exp[−i(χ/F ) ˆB(r,t)· F ]. Because of this symmetry, the effective Hamiltonianof the scalar field is not that of 4 He, but that of a neutral superfluid <strong>in</strong> a velocity field u s . Thevelocity (or gauge field) u s is a direct reflection of the sp<strong>in</strong>-gauge symmetry and it is given byu s =− i¯h M ζ † ∇ζ. (147)The velocity u s can be calculated from the vorticity Ω s of u s , which satisfies the Merm<strong>in</strong>–Horelation [163, 164],Ω s = 1 ( ) ¯h2 ∇ × u s = ɛ αβγ ˆB α ∇ ˆB β × ∇ ˆB γ . (148)2MEquation (148) shows that the spatial variations of B necessary to produce the trapp<strong>in</strong>g potentialwill <strong>in</strong>evitably generate a nonvanish<strong>in</strong>g superfluid velocity [6]u s = (2¯h/M)(1 − B z /B) ∇[arctan(B y /B x )].If B 0 = B 0 ẑ is the magnetic field at the centre of an axisymmetric harmonic trap and ω 0 is themaximum trap frequency, then the sp<strong>in</strong>-gauge effect generates the follow<strong>in</strong>g constant effective‘rotation’ s around the ẑ-axis [6]:Ω s∼−ẑ ¯hω 0(149)ω 0 µ B B 0where µ B is the Bohr magneton. The superfluid velocity u s splits the degeneracy of theharmonic energy levels, breaks the <strong>in</strong>version symmetry of the vortex-nucleation angularvelocity c , and can produce vortex ground states <strong>in</strong> the absence of external rotation if s > c [6]. In current experiments, the sp<strong>in</strong>-gauge effect is small; for example, if ω 0 = 10 Hzand B 0 = 1 G, we obta<strong>in</strong> s /ω 0 ∼ 10 −5 . In oblate traps with ω z ≫ ω ⊥ , however, the sp<strong>in</strong>gaugeeffect can be significant ( s could be comparable with ω ⊥ for large enough values of ω z ).Recently, the MIT group has succeeded <strong>in</strong> trapp<strong>in</strong>g a 23 Na Bose <strong>condensate</strong> by purelyoptical means [160, 161]. In contrast to those <strong>in</strong> a magnetic trap, the sp<strong>in</strong>s of the alkali atoms<strong>in</strong> such an optical trap are essentially free, so the sp<strong>in</strong>or nature of the alkali Bose <strong>condensate</strong>can be fully realized. Specifically, 23 Na atoms possess a hyperf<strong>in</strong>e sp<strong>in</strong>, with F = 1<strong>in</strong>thelower multiplet. All three possible projections of the hyperf<strong>in</strong>e sp<strong>in</strong> can be optically <strong>trapped</strong>


R182A L Fetter and A A Svidz<strong>in</strong>skysimultaneously. Thus the <strong>condensate</strong> is described by a sp<strong>in</strong>-1 sp<strong>in</strong>or. The <strong>in</strong>ternal vortexstructure of a <strong>trapped</strong> sp<strong>in</strong>-1 BEC was <strong>in</strong>vestigated <strong>in</strong> reference [165]. Such vortices and theirstability were also discussed <strong>in</strong> [20,166]. In an optical trap, the ground state of sp<strong>in</strong>-1 bosonssuch as 23 Na, 39 K, and 87 Rb can be either ferromagnetic or ‘polar’, depend<strong>in</strong>g on the scatter<strong>in</strong>glengths <strong>in</strong> different angular momentum channels [20]. The ferromagnetic state also has coreless(or skyrmion) vortices, like textures found <strong>in</strong> superfluid 3 He-A. Because of the wide range ofhyperf<strong>in</strong>e sp<strong>in</strong>s of different alkalis, the optical trap has provided great opportunities to studydifferent sp<strong>in</strong> textures <strong>in</strong> <strong>dilute</strong> quantum gases of atoms with large sp<strong>in</strong>s. This should be afruitful subject for future experiments.Although most of the theoretical effort has concentrated on s<strong>in</strong>gle-<strong>condensate</strong> systems,the first experimental realization of BEC vortices was achieved with a two-species 87 Rb<strong>condensate</strong> [33], follow<strong>in</strong>g the proposal of reference [167]. Several other proposals have beenmade for the dynamical production of a vortex us<strong>in</strong>g the <strong>in</strong>ternal structure of atoms [168–171].The sp<strong>in</strong>-exchange scatter<strong>in</strong>g rate is suppressed for 87 Rb, which makes possible the study ofmagnetically <strong>trapped</strong> multicomponent <strong>condensate</strong>s of these atoms. The two species correspondto two different hyperf<strong>in</strong>e energy levels of 87 Rb, denoted as |1〉 and |2〉; they are separatedby the ground-state hyperf<strong>in</strong>e splitt<strong>in</strong>g. S<strong>in</strong>ce the scatter<strong>in</strong>g lengths are different, the twostates are not equivalent. Typically, the |1〉 ≡|F = 1,m =−1〉 state is <strong>trapped</strong> and cooledto the condensation po<strong>in</strong>t. Once the atoms <strong>in</strong> |1〉 have formed the <strong>condensate</strong> ground state,a two-photon microwave field is applied, <strong>in</strong>duc<strong>in</strong>g transitions between the |1〉 state and the|2〉 ≡|F = 2,m = 1〉 state [33]. As a result, the atoms cycle coherently between the twohyperf<strong>in</strong>e levels with an effective Rabi frequency eff [172]. Two parameters characterize thecoupl<strong>in</strong>g: the detun<strong>in</strong>g and the power. The detun<strong>in</strong>g δ denotes the mismatch of the frequencyof the coupl<strong>in</strong>g electromagnetic field to the frequency difference between the two <strong>in</strong>ternalatomic states. The power is characterized by the Rabi frequency ; it is the rate at whichthe population would oscillate between the two states if δ were zero. When δ is larger than, the population oscillations occur at the effective Rabi frequency eff = √ 2 + δ 2 , whichobviously exceeds .In pr<strong>in</strong>ciple, both states could be cooled simultaneously, with the result that the <strong>condensate</strong>would form <strong>in</strong> a mixture of states. In practice, however, the typical lifetime of atoms <strong>in</strong> the|2〉 state is about 1 s due to <strong>in</strong>elastic sp<strong>in</strong>-exchange collisions, which makes it very difficult toachieve runaway evaporation for this state. In contrast, atoms <strong>in</strong> the |1〉 state have a much longerlifetime of about 75 s [33]. The advantage of us<strong>in</strong>g the |F = 1,m=−1〉 and |F = 2,m= 1〉states is that their magnetic moments are nearly the same, so they can be simultaneouslyconf<strong>in</strong>ed <strong>in</strong> identical and fully overlapp<strong>in</strong>g magnetic trap potentials. Unlike the more familiars<strong>in</strong>gle-component superfluids (see the discussion after equation (153)), where the topologicalconstra<strong>in</strong>ts make it difficult to implant a vortex with<strong>in</strong> an exist<strong>in</strong>g <strong>condensate</strong> <strong>in</strong> a controlledmanner, the coupled two-component <strong>condensate</strong> has a different order parameter and hencedifferent topological constra<strong>in</strong>ts. Indeed, the coupled two-component system allows the directcreation of a |2〉 (or |1〉) state wave function hav<strong>in</strong>g a wide variety of shapes out of a |1〉 (or|2〉) ground-state wave function [167].For example, to form a vortex <strong>in</strong> the two-component system, one should impose aperturbation Hˆ1 that couples the ground state of the system to the vortex state (i.e. the matrixelement of the perturbation operator connect<strong>in</strong>g these two states must be nonzero). The timedependentGP equation describ<strong>in</strong>g the driven, two-component <strong>condensate</strong> is [167]i¯h ∂ ∂t(1) (H ˆ= 0 +V H1 + Hˆ1 + ¯hδ/2 2 ¯h/2¯h/2Hˆ0 +V H2 − Hˆ1 − ¯hδ/2)(1 2)(150)


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R183whereV H1 = U 11 | 1 | 2 + U 12 | 2 | 2 and V H2 = U 21 | 1 | 2 + U 22 | 2 | 2and where Hˆ0 =−(¯h 2 ∇ 2 /2M) + 1 2 Mω2 0 (r2 ⊥ + z2 ) for a spherical trap, M is the atomic mass,ω 0 is the trap frequency, U ij = 4π¯h 2 a ij /M, with a ij the s-wave scatter<strong>in</strong>g lengths for b<strong>in</strong>arycollisions between constituents i and j. Williams and Holland considered the perturbation Hˆ1<strong>in</strong> the follow<strong>in</strong>g form [167]:Hˆ1 = κ[f(r) cos(ωt) + g(r) s<strong>in</strong>(ωt)] (151)where κ is a coupl<strong>in</strong>g coefficient and f(r) and g(r) are prefactors that depend on r. Theexplicit form of Hˆ1 determ<strong>in</strong>es the symmetry of the quantum state be<strong>in</strong>g prepared, so generalf and g can serve to prepare a macroscopic quantum state of arbitrary symmetry. To createa vortex state with one unit of angular momentum, one can take κ = Mω0 2ρ 0, f(r) = x,and g(r) = y <strong>in</strong> cartesian coord<strong>in</strong>ates. This form of perturbation effectively conf<strong>in</strong>es the twohyperf<strong>in</strong>e states <strong>in</strong> separate axially symmetric harmonic oscillator potentials with the sametrap frequency ω 0 . The trap centres are spatially offset <strong>in</strong> the xy-plane by a distance ρ 0 (fromthe centre) and rotate about the symmetry axis at an angular velocity ω. To achieve thisconfiguration experimentally, <strong>in</strong> reference [33] a laser beam was shone <strong>in</strong>to the trap alongthe ẑ-axis, so the cloud sits <strong>in</strong> the middle of the Gaussian beam waist where the gradient ofthe beam <strong>in</strong>tensity is approximately l<strong>in</strong>ear (see figure 14(a)). This arrangement produces aconstant force on the atoms. If the frequency of the laser beam is tuned between the twohyperf<strong>in</strong>e states, the optical dipole force acts <strong>in</strong> opposite directions for each state, displac<strong>in</strong>gthe trap centres for each state. When the beam rotates around the <strong>condensate</strong> at the angularvelocity ω, we obta<strong>in</strong> the desired result.To create a vortex, the angular velocity ω should be close to the value at which a resonanttransfer of population from the nonrotat<strong>in</strong>g <strong>condensate</strong> <strong>in</strong>to the vortex state takes place.Consider the frame co-rotat<strong>in</strong>g with the trap centres at an angular frequency ω. In this frame,the energy of the vortex with one unit of angular momentum is shifted by ¯hω relative to its value<strong>in</strong> the laboratory frame. When this energy shift compensates for both the energy mismatch ¯hδof the <strong>in</strong>ternal coupl<strong>in</strong>g field and the small chemical potential difference between the vortexand the nonrotat<strong>in</strong>g <strong>condensate</strong>, resonant transfer of population takes place (see figure 14(b)).It is obvious that if we change the sign of the detun<strong>in</strong>g δ while keep<strong>in</strong>g the trap rotation fixed, aLaserprofileLaser rotation, ω|2>ω=1 l=0 lBECTwo-photonmicrowave(a)(b)Figure 14. (a) A basic schematic illustration of the technique used to create a vortex. An offresonantlaser provides a rotat<strong>in</strong>g force on the atoms across the <strong>condensate</strong> as a microwave driveof detun<strong>in</strong>g δ is applied. (b) A level diagram show<strong>in</strong>g the microwave transition to very near the|2〉 state, and the modulation due to the laser rotation frequency that couples only to the angularmomentum l = 1 state when ω ≈ δ. (Taken from reference [33].)|1>


R184A L Fetter and A A Svidz<strong>in</strong>skyvortex will be created with opposite circulation. <strong>Vortices</strong> with opposite circulations experienceopposite energy shifts <strong>in</strong> transform<strong>in</strong>g to the rotat<strong>in</strong>g frame and therefore require opposite signsof detun<strong>in</strong>g <strong>in</strong> order to achieve the resonant coupl<strong>in</strong>g.In practice, ω ≫ ω 0 and δ ≫ . The first <strong>in</strong>equality allows the vortex to be generatedrapidly. The ma<strong>in</strong> problem with a slow drive (when ω ≈ ω 0 ) is that the timescale for coupl<strong>in</strong>gto the vortex state is very long, of the order of seconds <strong>in</strong> a trap with ω 0 = 10 Hz. Theweak-coupl<strong>in</strong>g limit, given by the second <strong>in</strong>equality, allows the resonance condition ω ≈ δ toselect energetically the desired state with high fidelity.Figure 15 shows the results of a numerical <strong>in</strong>tegration of equation (150) <strong>in</strong> two dimensions(ω z = 0), with the <strong>condensate</strong> <strong>in</strong>itially <strong>in</strong> the nonrotat<strong>in</strong>g ground state and <strong>in</strong> the <strong>in</strong>ternal state|1〉 [167]. The coupl<strong>in</strong>g drive is turned on at time t = 0, and is turned off at time t = t s bysett<strong>in</strong>g both and ρ 0 to zero. The top and the bottom graphs show the fractional populationand the angular momentum per atom of the |2〉 state as a function of time.The small-amplitude rapid oscillations <strong>in</strong> the top graph correspond to the cycl<strong>in</strong>g between<strong>in</strong>ternal levels at the effective Rabi frequency eff . The gradual rise of this l<strong>in</strong>e reflects coupl<strong>in</strong>gfrom the ground state to the vortex mode caused by the drive Hˆ1 <strong>in</strong> equation (150). Once dur<strong>in</strong>geach Rabi cycle, the angular momentum approaches unity (bottom graph), and, at that time,the |2〉 state wave function approaches a pure vortex mode. By turn<strong>in</strong>g off the coupl<strong>in</strong>g at aprecise time t = t s <strong>in</strong> a given Rabi cycle, the |2〉 state can be prepared to have unit angularmomentum. The maximum possible population transfer to the vortex state us<strong>in</strong>g this schemeobeys a Lorentzian response curve as ω is varied near eff , exhibit<strong>in</strong>g a narrow resonance.Figure 15. Dynamical evolution that can create a vortex. The top graph shows the fractionalpopulation of atoms <strong>in</strong> the |2〉 <strong>in</strong>ternal state. The bottom graph shows the angular momentum ofthe |2〉 state, <strong>in</strong> units of Planck’s constant ¯h. The <strong>in</strong>set shows the amplitude of population transferto the vortex as a function of the trap rotation frequency ω, with = eff − ω. The variousparameters used <strong>in</strong> the calculation are: ω 0 = 10 Hz, δ = 200 Hz, ω = 205.4 Hz, N = 8 × 10 5atoms; M is the mass of the 87 Rb atom; for simulations, the values of scatter<strong>in</strong>g lengths are takento be a 11 = a 22 = a 12 = 5.5 nm, and for t


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R185This situation is shown <strong>in</strong> the <strong>in</strong>set of figure 15, where = eff − ω.In an experiment, it is possible put the <strong>in</strong>itial <strong>condensate</strong> <strong>in</strong>to either the |1〉 or |2〉 state,and then make a vortex <strong>in</strong> the |2〉 or |1〉 state, respectively. The evolution of the vortex can bewatched over timescales from milliseconds to seconds. In reference [33], the vortex was foundto be stable <strong>in</strong> only one of the two possible configurations correspond<strong>in</strong>g to the vortex <strong>in</strong> the|1〉 state, which is the one with the larger scatter<strong>in</strong>g length (with the |2〉 state <strong>in</strong> the core). Theother possibility (the vortex <strong>in</strong> the |2〉 state, which is the one with the lowest self-<strong>in</strong>teractioncoefficient) produces an <strong>in</strong>stability.7.2. Stability theoryWe use the follow<strong>in</strong>g notation for the states: (1, 0) for the state with the vortex <strong>in</strong> |1〉 and (0, 1)for the state with the vortex <strong>in</strong> |2〉. In the JILA experiment [33], the number of particles isthe same for each component (N 1 = N 2 = N) but, <strong>in</strong> general, one could consider any ratiobetween the populations of the different levels. The scatter<strong>in</strong>g lengths for b<strong>in</strong>ary collisionsdepend on the <strong>in</strong>ternal hyperf<strong>in</strong>e level of the atom. For 87 Rb the values of scatter<strong>in</strong>g lengthsare nearly degenerate and <strong>in</strong> the proportion a 11 :a 12 :a 22 = 1.00:0.97:0.94 [173]. Because ofthe relation U 11 >U 12 >U 22 , the experiment is performed <strong>in</strong> a regime <strong>in</strong> which the firstcomponent separates from the second one. Consequently, a favoured configuration has thefirst component spread over the largest part of the space. Numerical simulations show that <strong>in</strong>the equal-population case, N 1 = N 2 = N, and for arbitrary nonl<strong>in</strong>earities, the stationary state(1, 0) is stable while the other state (0, 1) is unstable.The orig<strong>in</strong> of the <strong>in</strong>stability of the state (0, 1) is purely dynamical [174] and can beunderstood with<strong>in</strong> the framework of mean-field theories for the double-<strong>condensate</strong> systemwithout dissipation. Actually, the <strong>in</strong>stability mechanism does not lead to expulsion of thevortex from the <strong>condensate</strong>, but to periodic transfer of the phase s<strong>in</strong>gularity from one speciesto the other. To study the vortex stability, one can start from a pair of coupled Gross–Pitaevskiiequations for the <strong>condensate</strong> wave functions of each species:i¯h ∂ [−¯h2 ∂t ∇ 2]1 =2M + V 1 + U 11 | 1 | 2 + U 12 | 2 | 2 1 (152)i¯h ∂ [−¯h2 ∂t ∇ 22 =2M + V 2 + U 21 | 1 | 2 + U 22 | 2 | 2 ] 2 (153)where V 1 and V 2 are trap potentials for the <strong>condensate</strong> components. These equations are aparticular case of equation (150) when the drive is turned off ( Hˆ1 , , δ = 0). Equations (152)and (153) conserve the number of particles <strong>in</strong> each hyperf<strong>in</strong>e level. However, the angularmomentum of each component is no longer a conserved quantity, and the topological chargeof each species can change through the time evolution. Instead, what is conserved is the totalangular momentum of the system∫∫L z = i¯h d 3 r1 ∗ ∂ φ 1 +i¯h d 3 r2 ∗ ∂ φ 2 . (154)As <strong>in</strong> the JILA experiments, we assume that both potentials are spherically symmetric andhave the formV 1 (r) = V 2 (r) = 1 2 Mω2 0 (r2 ⊥ + z2 ).For stationary configurations <strong>in</strong> which each component has a well-def<strong>in</strong>ed value of the angularmomentum, the time and angular dependence are factored out: i (r ⊥ ,z,φ)= e −iµ it/¯h e iqiφ ψ i (r ⊥ ,z) (155)


R186A L Fetter and A A Svidz<strong>in</strong>skywith i = 1, 2. We focus on three particular configurations, which are the lowest energy stateswith vorticity (q 1 ,q 2 ) = (0, 0), (1, 0), (0, 1). They correspond to the ground state of thedouble <strong>condensate</strong>, and to the s<strong>in</strong>gle-vortex states for the |1〉 and |2〉 species, respectively.L<strong>in</strong>ear stability analysis of the three states gives the follow<strong>in</strong>g results [174]. For the (0, 0)state, the frequencies of all normal modes are positive, as expected for the ground state of thesystem. Among the normal modes of the (1, 0) family, there is a negative eigenvalue, whichmeans that there is a path <strong>in</strong> the configuration space along which the energy decreases (this isjust the analogue of the anomalous mode <strong>in</strong> the one-component system with a vortex). Thispath belongs to a perturbation that takes the vortex out of the <strong>condensate</strong>. As <strong>in</strong> the caseof a s<strong>in</strong>gle-component <strong>condensate</strong>, however, the lifetime of the vortex state is only limitedby the presence of dissipation (without dissipation, the configuration is dynamically stable).F<strong>in</strong>ally, <strong>in</strong> the (0, 1) family, there are normal modes with complex frequencies. The shape ofthe unstable modes is similar to that of the energy-decreas<strong>in</strong>g modes of the (1, 0) family—that is, they are perturbations that push the vortex out of both clouds. The imag<strong>in</strong>ary partof the eigenvalues implies that vortices with unit charge <strong>in</strong> |2〉 are unstable under a genericperturbation of the <strong>in</strong>itial data, whereas those <strong>in</strong> |1〉 can be long-lived. This conclusion isconsistent with the JILA experiments, where a vortex <strong>in</strong> the |2〉 species was found to beunstable [33].Numerical simulations of the vortex behaviour for large perturbations show that the l<strong>in</strong>earlystable state (1, 0) is robust and survives under a wide range of perturbations, suffer<strong>in</strong>g at mosta precession of the vortex core plus changes of the shapes of both components [174]. Thisbehaviour arises <strong>in</strong> both two- and three-dimensional simulations. In contrast, the unstableconfiguration (0, 1) develops a recurrent dynamics. In the first stage, the first component andthe vortex oscillate synchronously (the hole <strong>in</strong> |2〉 p<strong>in</strong>s the peak of |1〉). These oscillationsgrow <strong>in</strong> amplitude, and the vortex spirals out. F<strong>in</strong>ally the first component develops a tail andlater a hole which traps the second component. The hole is a vortex that has been transferredfrom |2〉 to |1〉. Though not completely periodic, this mechanism exhibits some recurrence,and the vortex eventually returns to |2〉. The preced<strong>in</strong>g behaviour persists even for strongperturbations <strong>in</strong> a two-dimensional <strong>condensate</strong>. However, for large perturbations of a threedimensional<strong>condensate</strong>, the dynamics may lead to a turbulent behaviour [174].In figure 16, it is shown how a small <strong>in</strong>itial perturbation makes the phase s<strong>in</strong>gularity <strong>in</strong>|2〉 spiral out of the system while a phase s<strong>in</strong>gularity appears <strong>in</strong> |1〉 and occupies the centre ofthe atomic cloud. This dynamics is recurrent.The preced<strong>in</strong>g results are valid for the equal-population case, N 1 = N 2 . For any ratio ofthe populations N 1 /N 2 and any values of the nonl<strong>in</strong>ear coefficients U ij , the stability conditionsare the follow<strong>in</strong>g [175]:• The configuration (1, 0) is stable if( √ )N 21− 1 > 1 − a 11. (156)N 2 a 12For 87 Rb, the <strong>in</strong>equality (156) is always satisfied, which proves that the configurationwith a vortex <strong>in</strong> |1〉 is always l<strong>in</strong>early stable, as found <strong>in</strong> [33]. Note that the stabilityproperties do not depend on the total number of particles but only on the ratio betweenthe populations.• The stability condition of the configuration (0, 1) is( √ N 2N 1− 1) 2> 1 − a 22a 21. (157)


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R187Figure 16. Evolution of the position of the phase s<strong>in</strong>gularity <strong>in</strong> the xy-plane. Lengths are given<strong>in</strong> units of the trap characteristic length d = √¯h/Mω 0 . (a) Phase s<strong>in</strong>gularity <strong>in</strong> |1〉, (b) phases<strong>in</strong>gularity <strong>in</strong> |2〉. (Taken from reference [175].)This <strong>in</strong>equality fails for a certa<strong>in</strong> range of N 1 /N 2 . For the case of 87 Rb the unstable rangeis N 1 /N 2 ∈ [0.73, 1.49], which means that certa<strong>in</strong> choices of the population imbalanceallow stabilization of the vortex <strong>in</strong> |2〉. These results predict the possibility of stable vortexstates for various multiple-<strong>condensate</strong> systems [175].Energetic considerations show that the extra degree of freedom associated with the secondcomponent allows a more <strong>in</strong>tricate structure for the free-energy surface. As a result, <strong>in</strong> a twocomponentsystem, it is possible to achieve a local m<strong>in</strong>imum <strong>in</strong> the free energy at the centre ofthe trap [126]. The presence of such a m<strong>in</strong>imum implies the existence of a region of energeticstability where the vortex cannot escape and might generate a persistent current.8. Conclusions and outlookIn this paper, we have provided an <strong>in</strong>troductory description of vortices <strong>in</strong> <strong>trapped</strong> Bosecondensedgases. The ma<strong>in</strong> conclusion of our analysis is that the vortex dynamics <strong>in</strong> suchsystems is well described by the time-dependent Gross–Pitaevskii equation (at least for lowtemperatures). The nonuniform nature of the <strong>condensate</strong> results <strong>in</strong> the appearance of anomalousvortex mode(s) with negative frequency and positive norm. Trap rotation shifts the normalmodefrequencies and can stabilize the vortex state. To date, experimental measurements ofvortex dynamics and other properties of vortex states are <strong>in</strong> a good quantitative agreementwith theoretical predictions based on solutions of the GP equation. Deviations from the meanfieldpredictions could arise when the gas parameter ¯n|a| 3 is not very small (semiclassicalcorrections to the mean-field approximation were calculated <strong>in</strong> [176]) or from ‘mesoscopic’


R188A L Fetter and A A Svidz<strong>in</strong>skyeffects associated with the f<strong>in</strong>ite systems. However, there is no experimental evidence for theseeffects so far.We have been able to cover only part of the exist<strong>in</strong>g literature on vortices <strong>in</strong> <strong>trapped</strong><strong>condensate</strong>s. Among important issues that we have not discussed are: different methods ofvortex generation and detection, k<strong>in</strong>etics of vortex nucleation and decay, vortices <strong>in</strong> BECs withattractive <strong>in</strong>teractions and <strong>in</strong> Fermi <strong>condensate</strong>s, other defects <strong>in</strong> BECs (solitons, <strong>in</strong>stantons,vortex solitons, skyrmions, and wave-function and sp<strong>in</strong> monopoles).In the case of superfluid helium, vortex nucleation is associated with p<strong>in</strong>n<strong>in</strong>g of vortex l<strong>in</strong>esat the walls of the conta<strong>in</strong>er. Trapped <strong>condensate</strong>s have no rough surfaces, and the nucleationprocess of quantized vorticity has a different orig<strong>in</strong> [118, 177, 178]. An important question<strong>in</strong> vortex nucleation is that of the role of the thermal component and transverse anisotropy ofmagnetic traps [179, 180].The literature of the past few years conta<strong>in</strong>s many different proposals for the creationof vortices <strong>in</strong> <strong>trapped</strong> BECs, although we considered only a few of them <strong>in</strong> this review. Toillustrate the diversity of different methods, let us cite some other schemes. An experimentalset-up for vortex creation by Berry’s-phase-<strong>in</strong>duced Bose–E<strong>in</strong>ste<strong>in</strong> condensation is proposedby Olshanii and Naraschewski [181]. A related vortex-production scheme employ<strong>in</strong>g theAharonov–Casher effect is discussed by Petrosyan and You [182]. Other proposals suggestthe creation of the vortex state by opto-mechanical stirr<strong>in</strong>g [169]; by a rotat<strong>in</strong>g force [183]; byan adiabatic population transfer of a <strong>condensate</strong> from the ground to the excited Bose-condensedstate via a Raman transition <strong>in</strong>duced by laser light [167,168,170]; by the accidental generationof vortices <strong>in</strong> a quench [184, 185] or <strong>in</strong> self-<strong>in</strong>terference measurements [171].A possible way to create rotat<strong>in</strong>g states from a <strong>trapped</strong> ground-state BEC by us<strong>in</strong>g light<strong>in</strong>ducedforces is proposed by Marzl<strong>in</strong> and Zhang [186]. They show that the dipole potential<strong>in</strong>duced by four travell<strong>in</strong>g-wave laser beams with an appropriate configuration <strong>in</strong> space, phase,and frequency can be used to realize such a system. Vortex states can be <strong>trapped</strong> <strong>in</strong> anevaporative cool<strong>in</strong>g process if the evaporation length is less than the size of the thermallyexcited state [185, 187]. In order to nucleate vortices, the <strong>trapped</strong> gas can be rotated attemperatures above the BEC transition. Recently, it has been suggested that vorticity couldbe impr<strong>in</strong>ted by imag<strong>in</strong>g a BEC through an absorption plate [188]. The method consistsof pass<strong>in</strong>g a far-off-resonant laser pulse through an absorption plate with an azimuthallydependent absorption coefficient, imag<strong>in</strong>g the laser beam onto a BEC, and thus creat<strong>in</strong>gthe correspond<strong>in</strong>g nondissipative Stark-shift potential and <strong>condensate</strong> phase shift. A vortexr<strong>in</strong>g may be formed by translat<strong>in</strong>g one <strong>condensate</strong> through another one [189] (this process isanalogous to r<strong>in</strong>g nucleation by mov<strong>in</strong>g ions <strong>in</strong> superfluid 4 He [190]), or by three-dimensionalsoliton decay [191,192]. Recently the JILA group generated vortex r<strong>in</strong>gs by the decay of darksolitons through the snake <strong>in</strong>stability [193].Many different proposals for the detection of vortices <strong>in</strong> BECs have been mentioned<strong>in</strong> the literature. Some of them are used <strong>in</strong> current experiments. The spatial size of thevortex core <strong>in</strong> the TF regime is too small to be observed; for visualiz<strong>in</strong>g the vortex state,switch<strong>in</strong>g off the trap and lett<strong>in</strong>g the cloud expand ballistically was suggested [84]. Afterexpansion, the size of the vortex core is magnified by approximately the same factor as thesize of the expand<strong>in</strong>g <strong>condensate</strong> [72, 194], so the core becomes observable. Also the vortexstate can be detected by the splitt<strong>in</strong>g of the collective <strong>condensate</strong> modes <strong>in</strong> axisymmetrictraps [64,113,117] or by look<strong>in</strong>g at the phase slip <strong>in</strong> the <strong>in</strong>terference fr<strong>in</strong>ges produced by twoexpand<strong>in</strong>g <strong>condensate</strong>s [72,169]. Dobrek et al [188] proposed an <strong>in</strong>terference method to detectvortices by coherently push<strong>in</strong>g part of the <strong>condensate</strong> with optically <strong>in</strong>duced Bragg scatter<strong>in</strong>g.A detection scheme that reveals the existence of vortex states <strong>in</strong> a cyl<strong>in</strong>drically symmetric trapis discussed by Goldste<strong>in</strong> et al [195]. This scheme relies on the measurement of the second-


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R189order correlation function of the Schröd<strong>in</strong>ger field and yields directly the topological chargeof the vortex state.Also one can detect the vortex state by observ<strong>in</strong>g the off-resonance absorption image ofthe rotational cloud [168]. For a vortex state one should expect a bright ‘hole’ <strong>in</strong> the imagewhich accounts for the vortex core <strong>in</strong> the density distribution. Another possibility is to observethe Doppler frequency shift due to the quantized circular motion of the atoms [168], or byscatter<strong>in</strong>g fast atoms <strong>in</strong> a pure momentum state off a <strong>trapped</strong> atomic cloud [196].Another question that has recently attracted significant theoretical and experimental<strong>in</strong>terest is that of the dynamics and stability of dark solitons and vortex solitons <strong>in</strong> <strong>trapped</strong><strong>condensate</strong>s. Solitary waves (k<strong>in</strong>ks) have been studied <strong>in</strong> many physical contexts [197] andexist <strong>in</strong> different physical, chemical, and biological systems [198]. Recent theoretical studiesdiscuss the dynamics and stability of dark solitons [199–203] (the range of parameters wherethe solitons are dynamically stable has been determ<strong>in</strong>ed <strong>in</strong> [101, 103], while the theory ofdissipative dynamics of a k<strong>in</strong>k <strong>in</strong> f<strong>in</strong>ite-temperature <strong>condensate</strong>s has been developed <strong>in</strong> [204]),as well as suggestions for their creation [170, 188, 205]. Recently, dark solitons <strong>in</strong>side a<strong>condensate</strong> were generated by a phase-impr<strong>in</strong>t<strong>in</strong>g method [206, 207]. Unlike vortices, darksolitons are not topologically stable. At f<strong>in</strong>ite temperature, they exhibit thermodynamic anddynamic (small-amplitude) <strong>in</strong>stabilities. The <strong>in</strong>teraction of the soliton with the thermal cloudcauses dissipation that accelerates the soliton. There is an <strong>in</strong>terest<strong>in</strong>g analogy between solitonsand relativistic particles, <strong>in</strong> which the soliton velocity and speed of sound correspond to theparticle velocity and speed of light <strong>in</strong> vacuum [204]. However, the k<strong>in</strong>ematic mass of thesoliton decreases when its velocity <strong>in</strong>creases. This behaviour is opposite to the case of arelativistic particle, where the k<strong>in</strong>ematic mass <strong>in</strong>creases with velocity, and an <strong>in</strong>f<strong>in</strong>ite forceis required to accelerate the particle beyond the velocity of light. In contrast to the particle,the soliton can reach the velocity of sound. An <strong>in</strong>terest<strong>in</strong>g problem is that of how to createa soliton and a vortex simultaneously (this object is known as a vortex soliton). The vortexsoliton has a topological charge and therefore could be stable.Another challeng<strong>in</strong>g goal for future experiments is the creation of vortex-like states <strong>in</strong>optically conf<strong>in</strong>ed BECs. By relax<strong>in</strong>g the condition of sp<strong>in</strong> polarization imposed by magnetictrapp<strong>in</strong>g, this new method of conf<strong>in</strong>ement permits the study of diverse textures that can beformed by the sp<strong>in</strong>or order parameter, like those <strong>in</strong> superfluid 3 He-A [20]. Also optical trapsallow strong variation of the scatter<strong>in</strong>g length via Feshbach resonances, which provides newpossibilities for manipulat<strong>in</strong>g the <strong>condensate</strong> states.Among other challeng<strong>in</strong>g problems, one should mention that of how to make measurementsof vortex normal modes at higher temperatures, which could establish the connectionbetween the Bogoliubov approximation and self-consistent mean-field theories. Also, it wouldbe <strong>in</strong>terest<strong>in</strong>g to observe vortex dissipation and damp<strong>in</strong>g of vortex normal modes.AcknowledgmentsWe are grateful to B Anderson, E Cornell, J Dalibard, D Feder, M Holland, M L<strong>in</strong>n,G Shlyapnikov and S Str<strong>in</strong>gari for valuable correspondence and discussions. This workbenefited from our participation <strong>in</strong> recent workshops at the Lorentz Centre, Leiden, TheNetherlands, and at the European Centre for Theoretical Studies <strong>in</strong> Nuclear Physics and RelatedAreas, Trento, Italy; we thank H Stoof and S Str<strong>in</strong>gari for organiz<strong>in</strong>g these workshops andfor their hospitality. This research was supported <strong>in</strong> part by the National Science Foundation,Grant No DMR 99-71518, and by Stanford University (AAS).


R190A L Fetter and A A Svidz<strong>in</strong>skyReferences[1] Anderson M H, Ensher J R, Matthews M H, Wieman C E and Cornell E A 1995 Science 269 198[2] Bradley C C, Sackett C A, Tollett J J and Hulet R G 1995 Phys. Rev. Lett. 75 1687[3] Davis K B, Mewes M-O, Andrews M R, van Druten N J, Durfee D S, Kurn D M and Ketterle W 1995 Phys.Rev. Lett. 75 3969[4] Griff<strong>in</strong> A 1993 Excitations <strong>in</strong> Bose-Condensed Liquid (New York: Cambridge University Press)[5] Sokol P 1995 Bose–E<strong>in</strong>ste<strong>in</strong> Condensation ed A Griff<strong>in</strong>, D W Snoke and S Str<strong>in</strong>gari (Cambridge: CambridgeUniversity Press) p 51[6] Ho T-L and Shenoy V B 1996 Phys. Rev. Lett. 77 2595[7] Lifshitz E M and Pitaevskii L P 1980 Statistical Physics part 2, 3rd edn (Oxford: Pergamon) section 29[8] Tilley D R and Tilley J 1986 Superfluidity and Superconductivity 2nd edn (Bristol: Hilger)[9] V<strong>in</strong>en W F 1969 Superconductivity edRDParks (New York: Dekker) ch 20[10] Onsager L 1949 Nuovo Cimento Suppl. 2 6 249Onsager L 1949 Nuovo Cimento Suppl. 2 6 281[11] Feynman R P 1955 Progress <strong>in</strong> Low Temperature Physics vol 1, ed C J Gorter (Amsterdam: North-Holland)p17[12] V<strong>in</strong>en W F 1961 Proc. R. Soc. A 260 218[13] Yarmchuk E J, Gordon M J V and Packard R E 1979 Phys. Rev. Lett. 43 214[14] Donnelly R J 1991 Quantized <strong>Vortices</strong> <strong>in</strong> Helium II (Cambridge: Cambridge University Press)[15] Dolzhanskii F V, Krymov V A and Man<strong>in</strong> D Yu 1990 Usp. Fiz. Nauk 160 1 (Engl. Transl. 1990 Sov. Phys.–Usp.33 495)[16] Sedrakyan D M and Shakhabasyan K M 1991 Usp. Fiz. Nauk 161 3 (Engl. Transl. 1991 Sov. Phys.–Usp. 34555)[17] Fowler G N, Raha S and We<strong>in</strong>er R M 1985 Phys. Rev. C 31 1515[18] Vollhardt D and Wölfle P 1990 The Superfluid Phases of Helium 3 (London: Taylor and Francis) ch 7[19] The term ‘skyrmion’ symbolizes an image of an extended baryon, be<strong>in</strong>g regarded as a topological soliton,made up of bosons but possess<strong>in</strong>g fermion features. Such a soliton was considered by Skyrme <strong>in</strong> his modelwhich describes the low-energy limit of quantum chromodynamics. For a review of the Skyrme model andstrong <strong>in</strong>teractions seeMakhan’kov V G, Rybakov Yu P and Sanyuk V I 1992 Usp. Fiz. Nauk 162 1 (Engl. Transl. 1992 Sov. Phys.–Usp.35 55)[20] Ho T-L 1998 Phys. Rev. Lett. 81 742Ho T-L 1998 Prepr<strong>in</strong>t cond-mat/9803231[21] García-Ripoll J J, Cirac J I, Angl<strong>in</strong> J, Pérez-García V M and Zoller P 2000 Phys. Rev. A 61 053609[22] Krusius M, Hakonen P J and Simola J T 1984 Proc. 17th Int. Conf. on Low Temperature Physics; Physica B+C126 22[23] Volovik G E 1984 Proc. 17th Int. Conf. on Low Temperatures Physics; Physica B+C 126 34For a review of superfluid properties of 3 He-A and 3 He-B, see alsoVolovik G E 1984 Usp. Fiz. Nauk 143 73 (Engl. Transl. 1984 Sov. Phys.–Usp. 27 363)M<strong>in</strong>eev V P 1983 Usp. Fiz. Nauk 139 303 (Engl. Transl. 1983 Sov. Phys.–Usp. 26 160)[24] Fetter A L 1986 Progress <strong>in</strong> Low Temperature Physics vol 10, ed D F Brewer (Amsterdam: North-Holland)p1[25] Salomaa M M and Volovik G E 1987 Rev. Mod. Phys. 59 533[26] Saffman P G 1997 Vortex Dynamics (Cambridge: Cambridge University Press)[27] Jones C A and Roberts P H 1982 J. Phys. A: Math. Gen. 15 2599[28] Frisch T, Pomeau Y and Rica S 1992 Phys. Rev. Lett. 69 1644[29] Koplik J and Lev<strong>in</strong>e H 1993 Phys. Rev. Lett. 71 1375[30] Koplik J and Lev<strong>in</strong>e H 1996 Phys. Rev. Lett. 76 4745[31] Dalfovo F, Giorg<strong>in</strong>i S, Pitaevskii L P and Str<strong>in</strong>gari S 1999 Rev. Mod. Phys. 71 463[32] Baym G and Pethick C J 1996 Phys. Rev. Lett. 76 6[33] Matthews M R, Anderson B P, Haljan P C, Hall D S, Wieman C E and Cornell E A 1999 Phys. Rev. Lett. 832498[34] Madison K W, Chevy F, Wohlleben W and Dalibard J 2000 Phys. Rev. Lett. 84 806[35] Madison K W, Chevy F, Wohlleben W and Dalibard J 2000 J. Mod. Opt. 47 2715[36] Chevy F, Madison K W and Dalibard J 2000 Phys. Rev. Lett. 85 2223[37] Anderson B P, Haljan P C, Wieman C E and Cornell E A 2000 Phys. Rev. Lett. 85 2857[38] Bogoliubov N N 1947 J. Phys. (USSR) 11 23


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R191[39] Gross E P 1961 Nuovo Cimento 20 454[40] Gross E P 1963 J. Math. Phys. 4 195[41] Pitaevskii L P 1961 Zh. Eksp. Teor. Fiz. 40 646 (Engl. Transl. 1961 Sov. Phys.–JETP 13 451)[42] Fetter A L and Walecka J D 1971 Quantum Theory of Many Particle Systems (New York: McGraw-Hill)section 35[43] Fetter A L 1999 Bose–E<strong>in</strong>ste<strong>in</strong> Condensation <strong>in</strong> Atomic Gases (Proc. Int. ‘Enrico Fermi’ School of Physics)ed M Inguscio, S Str<strong>in</strong>gari and CEWieman (Amsterdam: IOS Press) p 201[44] Ties<strong>in</strong>ga E, Williams C W, Julienne P S, Jones K M, Lett P D and Phillips W D 1996 J. Res. Natl Inst. Stand.Technol. 101 505[45] Boesten H M J M, Tsai C C, Gardner J R, He<strong>in</strong>zen D N and Verhaar B J 1997 Phys. Rev. A 55 636[46] Abraham ERI,McAlexander W I, Sackett C A and Hulet R G 1995 Phys. Rev. Lett. 74 1315[47] Dalfovo F and Str<strong>in</strong>gari S 1996 Phys. Rev. A 53 2477[48] Cornish S L, Claussen N R, Roberts J L, Cornell E A and Wieman C E 2000 Phys. Rev. Lett. 85 1795[49] Landau L D and Lifshitz E M 1987 Fluid Mechanics 2nd edn (London: Pergamon) ch 1[50] Fetter A L and Walecka J D 1980 Theoretical Mechanics of Particles and Cont<strong>in</strong>ua (New York: McGraw-Hill)section 48[51] G<strong>in</strong>zburg V L and Pitaevskii L P 1958 Zh. Eksp. Teor. Fiz. 34 1240 (Engl. Transl. 1958 Sov. Phys.–JETP 7 858)[52] Fetter A L 1965 Phys. Rev. 138 A429[53] Fetter A L 1966 Phys. Rev. 151 100[54] Lund F 1991 Phys. Lett. A 159 245[55] Rica S and Tirapegui E 1990 Phys. Rev. Lett. 64 878[56] Rica S and Tirapegui E 1992 Physica D 61 246[57] Str<strong>in</strong>gari S 1996 Phys. Rev. Lett. 77 2360[58] T<strong>in</strong>kham M 1975 Introduction to Superconductivity (New York: McGraw-Hill) ch 5[59] Konotop V V and Pérez-García V M 2000 Phys. Rev. A 62 033610[60] Wilk<strong>in</strong> N K and Gunn J M F 2000 Phys. Rev. Lett. 84 6[61] Moore G and Read N 1991 Nucl. Phys. B 360 362[62] Mottelson B 1999 Phys. Rev. Lett. 83 2695[63] Bertsch G F and Papenbrock T 1999 Phys. Rev. Lett. 83 5412[64] S<strong>in</strong>ha S 1997 Phys. Rev. A 55 4325[65] Rokhsar D S 1997 Phys. Rev. Lett. 79 2164[66] Svidz<strong>in</strong>sky A A and Fetter A L 2000 Physica B 284–288 21[67] Lifshitz E M and Pitaevskii L P 1980 Statistical Physics part 1, 3rd edn (Oxford: Pergamon) section 26[68] Wilk<strong>in</strong> N K, Gunn J M F and Smith R A 1998 Phys. Rev. Lett. 80 2265[69] Butts D A and Rokhsar D S 1999 Nature 397 327[70] L<strong>in</strong>n M and Fetter A L 1999 Phys. Rev. A 60 4910[71] Lundh E, Pethick C J and Smith H 1997 Phys. Rev. A 55 2126[72] Cast<strong>in</strong> Y and Dum R 1999 Eur. Phys. J. D 7 399[73] Shi H and Zheng W 1997 Phys. Rev. A 55 2930[74] Str<strong>in</strong>gari S 1999 Phys. Rev. Lett. 82 4371[75] Lamb H 1945 Hydrodynamics 6th edn (New York: Dover) pp 86–8[76] Fetter A L 1974 J. Low Temp. Phys. 16 533[77] Feder D L, Clark C W and Schneider B I 1999 Phys. Rev. Lett. 82 4956[78] Feder D L, Clark C W and Schneider B I 2000 Phys. Rev. A 61 011601(R)[79] DeConde K and Packard R E 1975 Phys. Rev. Lett. 35 732[80] Svidz<strong>in</strong>sky A A and Fetter A L 2000 Phys. Rev. Lett. 84 5919[81] Dalfovo F, Pitaevskii L P and Str<strong>in</strong>gari S 1996 Phys. Rev. A 54 4213[82] Dalfovo F, Giorg<strong>in</strong>i S, Guilleumas M, Pitaevskii L P and Str<strong>in</strong>gari S 1997 Phys. Rev. A 56 3840[83] Isoshima T and Machida K 1999 Phys. Rev. A 60 3313[84] Lundh E, Pethick C J and Smith H 1998 Phys. Rev. A 58 4816[85] Feder D L, Svidz<strong>in</strong>sky A A, Fetter A L and Clark C W 2001 Phys. Rev. Lett. 86 564[86] Andronikashvili E L, Mamaladze Yu G, Mat<strong>in</strong>yan S G and Tsakadze D S 1961 Usp. Fiz. Nauk 73 3 (Engl.Transl. 1961 Sov. Phys.–Usp. 4 1)[87] Williams G A and Packard R E 1974 Phys. Rev. Lett. 33 280[88] Yarmchuk E J and Packard R E 1982 J. Low Temp. Phys. 46 479[89] Tkachenko V K 1966 Zh. Eksp. Teor. Fiz. 49 1875 (Engl. Transl. 1966 Sov. Phys.–JETP 22 1282)[90] Hess G B 1967 Phys. Rev. 161 189[91] Stauffer D and Fetter A L 1968 Phys. Rev. 168 156


R192A L Fetter and A A Svidz<strong>in</strong>sky[92] Campbell L J and Ziff R M 1979 Phys. Rev. B 20 1886[93] Kavoulakis G M, Mottelson B and Pethick C J 2000 Phys. Rev. A 62 063605[94] Haljan P C, Anderson B P, Codd<strong>in</strong>gton I and Cornell E A 2000 Prepr<strong>in</strong>t cond-mat/0012320[95] Nepomnyashchii Y A 1974 Teor. Mat. Fiz. 20 399[96] Colson W B and Fetter A L 1978 J. Low Temp. Phys. 33 231[97] Fetter A L 1972 Ann. Phys., NY 70 67[98] Fetter A L 1996 Phys. Rev. A 53 4245[99] Pu H, Law C K, Eberly J H and Bigelow N P 1999 Phys. Rev. A 59 1533[100] Aranson I and Ste<strong>in</strong>berg V 1996 Phys. Rev. B 53 75[101] Muryshev A E, van L<strong>in</strong>den van den Heuvell H B and Shlyapnikov G V 1999 Phys. Rev. A 60 R2665[102] Fedichev P O, Muryshev A E and Shlyapnikov G V 1999 Phys. Rev. A 60 3220[103] Feder D L, P<strong>in</strong>dzola M S, Coll<strong>in</strong>s L A, Schneider B I and Clark C W 2000 Phys. Rev. A 62 053606[104] García-Ripoll J J and Pérez-García V M 1999 Phys. Rev. A 60 4864[105] Garay L J, Angl<strong>in</strong> J R, Cirac J I and Zoller P 2001 Phys. Rev. A 63 023611[106] Edwards M, Dodd R J, Clark C W and Burnett K 1996 J. Res. Natl Inst. Stand. Technol. 101 553[107] Landau L D 1941 J. Phys. (USSR) 5 71[108] Lifshitz E M and Pitaevskii L P 1980 Statistical Physics part 2, 3rd edn (Oxford: Pergamon) sections 22 and23[109] Landau L D and Lifshitz E M 1987 Fluid Mechanics 2nd edn (London: Pergamon) ch 9[110] Landau L D, Lifshitz E M and Pitaevskii L P 1984 Electrodynamics of Cont<strong>in</strong>uous Media 2nd edn (Oxford:Pergamon) section 115[111] Jackson J D 1998 Classical Electrodynamics 3rd edn (New York: Wiley) section 13.4[112] Fetter A L and Rokhsar D 1998 Phys. Rev. A 57 1191[113] Svidz<strong>in</strong>sky A A and Fetter A L 1998 Phys. Rev. A 58 3168[114] Dodd R J, Burnett K, Edwards M and Clark C W 1997 Phys. Rev. A 56 587[115] Fetter A L 1998 J. Low Temp. Phys. 113 198[116] Svidz<strong>in</strong>sky A A and Fetter A L, unpublishedSeeSvidz<strong>in</strong>sky A A and Fetter A L 1998 Prepr<strong>in</strong>t cond-mat/9811348[117] Zambelli F and Str<strong>in</strong>gari S 1998 Phys. Rev. Lett. 81 1754[118] García-Ripoll J J and Pérez-García V M 2000 Prepr<strong>in</strong>t cond-mat/0006368[119] Svidz<strong>in</strong>sky A A and Fetter A L 2000 Phys. Rev. A 62 063617[120] Pérez-García V M, Mich<strong>in</strong>el H, Cirac J I, Lewenste<strong>in</strong> M and Zoller P 1996 Phys. Rev. Lett. 77 5320[121] Pérez-García V M, Mich<strong>in</strong>el H, Cirac J I, Lewenste<strong>in</strong> M and Zoller P 1997 Phys. Rev. A 56 1424[122] L<strong>in</strong>n M and Fetter A L 2000 Phys. Rev. A 61 063603[123] Packard R E and Sanders T M Jr 1972 Phys. Rev. A 6 799[124] Jackson B, McCann J F and Adams C S 2000 Phys. Rev. A 61 013604[125] Lundh E and Ao P 2000 Phys. Rev. A 61 063612[126] McGee S A and Holland M J 2000 Prepr<strong>in</strong>t cond-mat/0007143[127] Feder D, private communication[128] Pismen L M and Rub<strong>in</strong>ste<strong>in</strong> J 1991 Physica D 47 353[129] Rub<strong>in</strong>ste<strong>in</strong> B Y and Pismen L M 1994 Physics D 78 1[130] Pismen L M 1999 <strong>Vortices</strong> <strong>in</strong> Nonl<strong>in</strong>ear Fields (Oxford: Clarendon) sections 2.2 and 5.2[131] Barenghi C F 1996 Phys. Rev. A 54 5445[132] Fetter A L and Walecka J D 1980 Theoretical Mechanics of Particles and Cont<strong>in</strong>ua (New York: McGraw-Hill)section 40[133] Landau L D and Lifshitz E M 1960 Mechanics (Oxford: Pergamon) sections 36 and 37[134] Kleppner D and Kolenkow R 1973 An Introduction to Mechanics (New York: McGraw-Hill) section 7.3[135] Fetter A L and Walecka J D 1980 Theoretical Mechanics of Particles and Cont<strong>in</strong>ua (New York: McGraw-Hill)section 28[136] Griff<strong>in</strong> A 1996 Phys. Rev. B 53 9341[137] Fedichev P O and Shlyapnikov G V 1998 Phys. Rev. A 58 3146[138] Beliaev S T 1958 Sov. Phys.–JETP 34 299[139] Pitaevskii L and Str<strong>in</strong>gari S 1998 Phys. Rev. Lett. 81 4541[140] Isoshima T and Machida K 1997 J. Phys. Soc. Japan 66 3502[141] Hutch<strong>in</strong>son DAW,Dodd R J and Burnett K 1998 Phys. Rev. Lett. 81 2198[142] Proukakis N P, Morgan S A, Choi S and Burnett K 1998 Phys. Rev. A 58 2435[143] Virtanen S M M, Simula T P and Salomaa M M 2001 Prepr<strong>in</strong>t cond-mat/0102035


<strong>Vortices</strong> <strong>in</strong> a <strong>trapped</strong> <strong>dilute</strong> Bose–E<strong>in</strong>ste<strong>in</strong> <strong>condensate</strong>R193Virtanen S M M, Simula T P and Salomaa M M 2001 Phys. Rev. Lett. at press[144] Isoshima T and Machida K 1999 Phys. Rev. A 59 2203[145] Fedichev P O and Shlyapnikov G V 1999 Phys. Rev. A 60 R1779[146] Son<strong>in</strong> E B 1997 Phys. Rev. B 55 485[147] Ovch<strong>in</strong>nikov Yu N and Sigal I M 1998 Nonl<strong>in</strong>earity 11 1295[148] Arovas D P and Freire J A 1997 Phys. Rev. B 55 1068[149] Ambegaokar V, Halper<strong>in</strong> B I, Nelson D R and Siggia E D 1980 Phys. Rev. B 21 1806[150] Chikkatur A P, Görlitz A, Stamper-Kurn D M, Inouye S, Gupta S and Ketterle W 2000 Phys. Rev. Lett. 85 483[151] Raman C, Köhl M, Onofrio R, Durfee D S, Kuklewicz C E, Hadzibabic Z and Ketterle W 1999 Phys. Rev. Lett.83 2502[152] Onofrio R, Raman C, Vogels J M, Abo-Shaeer J R, Chikkatur A P and Ketterle W 2000 Phys. Rev. Lett. 852228[153] Huepe C and Brachet M E 1997 C. R. Acad. Sci., Paris 325 195[154] Jackson B, McCann J F and Adams C S 1998 Phys. Rev. Lett. 80 3903[155] Nore C, Huepe C and Brachet M E 2000 Phys. Rev. Lett. 84 2191[156] W<strong>in</strong>iecki T, McCann J F and Adams C S 1999 Phys. Rev. Lett. 82 5186[157] Caradoc-Davies B M, Ballagh R J and Burnett K 1999 Phys. Rev. Lett. 83 895[158] Burkhart S, Bernard M, Avenel O and Varoquaux E 1994 Phys. Rev. Lett. 72 380[159] Myatt C J, Burt E A, Ghrist R W, Cornell E A and Wieman C E 1997 Phys. Rev. Lett. 78 586[160] Stamper-Kurn D M, Andrews M R, Chikkatur A P, Inouye S, Miesner H-J, Stenger J and Ketterle W 1998Phys. Rev. Lett. 80 2027[161] Stenger J, Inouye S, Stamper-Kurn D M, Miesner H-J, Chikkatur A K and Ketterle W 1998 Nature 396 345[162] Wheatley J C 1970 Progress <strong>in</strong> Low Temperature Physics vol 6, ed C J Gorter (Amsterdam: North-Holland)p77[163] Merm<strong>in</strong> N D and Ho T-L 1976 Phys. Rev. Lett. 36 594[164] See, for example,Vollhardt D and Wölfle P 1990 The Superfluid Phases of Helium 3 (London: Taylor and Francis) ch 17, p 210[165] Yip S-K 1999 Phys. Rev. Lett. 83 4677[166] Ohmi T and Machida K 1998 J. Phys. Soc. Japan 67 1822[167] Williams J E and Holland M J 1999 Nature 401 568[168] Marzl<strong>in</strong> K-P, Zhang W and Wright E W 1997 Phys. Rev. Lett. 79 4728[169] Bolda E L and Walls D F 1998 Phys. Lett. A 246 32[170] Dum R, Cirac J I, Lewenste<strong>in</strong> M and Zoller P 1998 Phys. Rev. Lett. 80 2972[171] Ruostekoski J 2000 Phys. Rev. A 61 041603[172] Williams J, Walser R, Cooper J, Cornell E A and Holland M 2000 Phys. Rev. A 61 033612[173] Hall D S, Matthews M R, Ensher J R, Wieman C E and Cornell E A 1998 Phys. Rev. Lett. 81 1539[174] García-Ripoll J J and Pérez-García V M 2000 Phys. Rev. Lett. 84 4264[175] Pérez-García V M and García-Ripoll J J 2000 Phys. Rev. A 62 033601[176] Andersen J and Braaten E 1999 Phys. Rev. A 60 2330[177] Martika<strong>in</strong>en J P, Suom<strong>in</strong>en K A and Sanpera A 2000 Prepr<strong>in</strong>t cond-mat/0005136[178] Dalfovo F and Str<strong>in</strong>gari S 2001 Phys. Rev. A 63 011601[179] Recati A, Zambelli F and Str<strong>in</strong>gari S 2001 Phys. Rev. Lett. 86 377[180] Madison K W, Chevy F, Bret<strong>in</strong> V and Dalibard J 2001 Prepr<strong>in</strong>t cond-mat/0101051[181] Olshanii M and Naraschewski M 1998 Prepr<strong>in</strong>t cond-mat/9811314[182] Petrosyan K G and You L 1999 Phys. Rev. A 59 639[183] Marzl<strong>in</strong> K-P and Zhang W 1998 Phys. Rev. A 57 4761[184] Angl<strong>in</strong> J R and Zurek W H 1999 Phys. Rev. Lett. 83 1707[185] Drummond P D and Corney J F 1999 Phys. Rev. A 60 R2661[186] Marzl<strong>in</strong> K-P and Zhang W 1998 Phys. Rev. A 57 3801[187] Marshall R J, New GHC,Burnett K and Choi S 1999 Phys. Rev. A 59 2085[188] Dobrek L, Gajda M, Lewenste<strong>in</strong> M, Sengstock K, Birkl G and Ertmer W 1999 Phys. Rev. A 60 R3381[189] Jackson B, McCann J F and Adams C S 1999 Phys. Rev. A 60 4882[190] Rayfield G W and Reif F 1964 Phys. Rev. 136 1194[191] Jones C A, Putterman S J and Roberts P H 1986 J. Phys. A: Math. Gen. 19 2991[192] Josserand C and Pomeau Y 1995 Europhys. Lett. 30 43[193] Anderson B P, Haljan P C, Regal C A, Feder D L, Coll<strong>in</strong>s L A, Clark C W and Cornell E A 2000 Prepr<strong>in</strong>tcond-mat/0012444[194] Dalfovo F and Modugno M 2000 Phys. Rev. A 61 023605


R194A L Fetter and A A Svidz<strong>in</strong>sky[195] Goldste<strong>in</strong> E V, Wright E M and Meystre P 1998 Phys. Rev. A 58 576[196] Kuklov A B and Svistunov B V 1999 Phys. Rev. A 60 R769[197] Rajaraman R 1987 Solitons and Instantons (Amsterdam: North-Holland)[198] Kerner B S and Osipov V V 1989 Usp. Fiz. Nauk 157 201 (Engl. Transl. 1989 Sov. Phys.–Usp. 32 101)[199] Zhang W, Walls D F and Sanders B C 1994 Phys. Rev. Lett. 72 60[200] Re<strong>in</strong>hardt W P and Clark C W 1997 J. Phys. B: At. Mol. Phys. 30 L785[201] Morgan S A, Ballagh R J and Burnett K 1997 Phys. Rev. A 55 4338[202] Jackson A D, Kavoulakis G M and Pethick C J 1998 Phys. Rev. A 58 2417[203] Margetis D 1999 J. Math. Phys. 40 5522[204] Fedichev P O, Muryshev A E and Shlyapnikov G V 1999 Phys. Rev. A 60 3220[205] Scott T F, Ballagh R J and Burnett K 1998 J. Phys. B: At. Mol. Phys. 31 L329[206] Burger S, Bongs K, Dettmer S, Ertmer W, Sengstock K, Sanpera A, Shlyapnikov G V and Lewenste<strong>in</strong> M 1999Phys. Rev. Lett. 83 5198[207] Denschlag J, Simsarian J E, Feder D L, Clark C W, Coll<strong>in</strong>s L A, Cubizolles J, Deng L, Hagley E W, Helmerson K,Re<strong>in</strong>hardt W P, Rolston S L, Schneider B I and Phillips W D 2000 Science 287 97

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!